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:

Ornithine aminotransferase supports polyamine synthesis in pancreatic cancer

Abstract

There is a need to develop effective therapies for pancreatic ductal adenocarcinoma (PDA), a highly lethal malignancy with increasing incidence1 and poor prognosis2. Although targeting tumour metabolism has been the focus of intense investigation for more than a decade, tumour metabolic plasticity and high risk of toxicity have limited this anticancer strategy3,4. Here we use genetic and pharmacological approaches in human and mouse in vitro and in vivo models to show that PDA has a distinct dependence on de novo ornithine synthesis from glutamine. We find that this process, which is mediated through ornithine aminotransferase (OAT), supports polyamine synthesis and is required for tumour growth. This directional OAT activity is usually largely restricted to infancy and contrasts with the reliance of most adult normal tissues and other cancer types on arginine-derived ornithine for polyamine synthesis5,6. This dependency associates with arginine depletion in the PDA tumour microenvironment and is driven by mutant KRAS. Activated KRAS induces the expression of OAT and polyamine synthesis enzymes, leading to alterations in the transcriptome and open chromatin landscape in PDA tumour cells. The distinct dependence of PDA, but not normal tissue, on OAT-mediated de novo ornithine synthesis provides an attractive therapeutic window for treating patients with pancreatic cancer with minimal toxicity.

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: PDA uses glutamine for DNS and polyamine synthesis.
Fig. 2: OAT is required for PDA polyamine synthesis and tumour growth.
Fig. 3: KRAS promotes DNS and polyamine synthesis in PDA.
Fig. 4: OAT inhibition suppresses PDA and alters its transcriptome similar to ODC1.

Similar content being viewed by others

Data availability

All data supporting the findings in this study are available within the article, its supplementary information and from the corresponding author upon reasonable request. The RNA-seq and ATAC-seq data have been deposited into the Gene Expression Omnibus of the National Center for Biotechnology Information and are accessible through GEO Series accession number GSE193411. Box plots in Extended Data Fig. 4a were generated using GEPIA56 (http://gepia.cancer-pku.cn/help.html), which uses a combination of public datasets from both The Cancer Genome Atlas and the Genotype–Tissue project for expression analyses. In Extended Data Fig. 9a,b, bar plots and data were generated using TRANSFAC database (v.1.9) (https://genexplain.com/transfac-2-0/). For gel source data, see Supplementary Fig. 1Source data are provided with this paper.

References

  1. Rahib, L., Wehner, M. R., Matrisian, L. M. & Nead, K. T. Estimated projection of US cancer incidence and death to 2040. JAMA Netw. Open 4, e214708 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  2. Siegel, R. L., Miller, K. D., Wagle, N. S. & Jemal, A. Cancer statistics, 2023. CA Cancer J. Clin. 73, 17–48 (2023).

    Article  PubMed  Google Scholar 

  3. Encarnacion-Rosado, J. & Kimmelman, A. C. Harnessing metabolic dependencies in pancreatic cancers. Nat. Rev. Gastroenterol. Hepatol. 18, 482–492 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  4. Martinez-Reyes, I. & Chandel, N. S. Cancer metabolism: looking forward. Nat. Rev. Cancer 21, 669–680 (2021).

    Article  CAS  PubMed  Google Scholar 

  5. Ginguay, A., Cynober, L., Curis, E. & Nicolis, I. Ornithine aminotransferase, an important glutamate-metabolizing enzyme at the crossroads of multiple metabolic pathways. Biology (Basel) 6, 18 (2017).

    PubMed  Google Scholar 

  6. Casero, R. A. Jr, Murray Stewart, T. & Pegg, A. E. Polyamine metabolism and cancer: treatments, challenges and opportunities. Nat. Rev. Cancer 18, 681–695 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Zaytouni, T. et al. Critical role for arginase 2 in obesity-associated pancreatic cancer. Nat. Commun. 8, 242 (2017).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  8. Ying, H. et al. Oncogenic Kras maintains pancreatic tumors through regulation of anabolic glucose metabolism. Cell 149, 656–670 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Son, J. et al. Glutamine supports pancreatic cancer growth through a KRAS-regulated metabolic pathway. Nature 496, 101–105 (2013).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  10. Vande Voorde, J. et al. Improving the metabolic fidelity of cancer models with a physiological cell culture medium. Sci. Adv. 5, eaau7314 (2019).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  11. Hensley, C. T. et al. Metabolic heterogeneity in human lung tumors. Cell 164, 681–694 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Sullivan, M. R. et al. Quantification of microenvironmental metabolites in murine cancers reveals determinants of tumor nutrient availability. eLife 8, e44235 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Tsai, P. Y. et al. Adaptation of pancreatic cancer cells to nutrient deprivation is reversible and requires glutamine synthetase stabilization by mTORC1. Proc. Natl Acad. Sci. USA https://doi.org/10.1073/pnas.2003014118 (2021).

  14. Apiz-Saab, J. J. et al. Pancreatic tumors activate arginine biosynthesis to adapt to myeloid-driven amino acid stress. Preprint at bioRxiv https://doi.org/10.1101/2022.06.21.497008 (2022).

  15. Menjivar, R. E. et al. Arginase 1 is a key driver of immune suppression in pancreatic cancer. eLife 12, e80721 (2023).

    Article  PubMed  Google Scholar 

  16. Miska, J. et al. Polyamines drive myeloid cell survival by buffering intracellular pH to promote immunosuppression in glioblastoma. Sci. Adv. 7, eabc8929 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  17. Gitto, S. B. et al. Difluoromethylornithine combined with a polyamine transport inhibitor is effective against gemcitabine resistant pancreatic cancer. Mol. Pharm. 15, 369–376 (2018).

    Article  CAS  PubMed  Google Scholar 

  18. Nilsson, J. A. et al. Targeting ornithine decarboxylase in Myc-induced lymphomagenesis prevents tumor formation. Cancer Cell 7, 433–444 (2005).

    Article  CAS  PubMed  Google Scholar 

  19. Meyskens, F. L. Jr & Gerner, E. W. Development of difluoromethylornithine (DFMO) as a chemoprevention agent. Clin. Cancer Res. 5, 945–951 (1999).

    CAS  PubMed  Google Scholar 

  20. Holbert, C. E., Cullen, M. T., Casero, R. A. Jr & Stewart, T. M. Polyamines in cancer: integrating organismal metabolism and antitumour immunity. Nat. Rev. Cancer 22, 467–480 (2022).

    Article  CAS  PubMed  Google Scholar 

  21. Daune, G., Gerhart, F. & Seiler, N. 5-Fluoromethylornithine, an irreversible and specific inhibitor of l-ornithine:2-oxo-acid aminotransferase. Biochem. J. 253, 481–488 (1988).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Seiler, N., Sarhan, S. & Knodgen, B. Inhibition of ornithine aminotransferase by 5-fluoromethylornithine: protection against acute thioacetamide intoxication by elevated tissue ornithine levels. Pharmacol. Toxicol. 70, 373–380 (1992).

    Article  CAS  PubMed  Google Scholar 

  23. Zielonka, M. et al. Pharmacologic rescue of hyperammonemia-induced toxicity in zebrafish by inhibition of ornithine aminotransferase. PLoS ONE 13, e0203707 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  24. Burns, M. R., Graminski, G. F., Weeks, R. S., Chen, Y. & O’Brien, T. G. Lipophilic lysine–spermine conjugates are potent polyamine transport inhibitors for use in combination with a polyamine biosynthesis inhibitor. J. Med. Chem. 52, 1983–1993 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Samal, K. et al. AMXT-1501, a novel polyamine transport inhibitor, synergizes with DFMO in inhibiting neuroblastoma cell proliferation by targeting both ornithine decarboxylase and polyamine transport. Int. J. Cancer 133, 1323–1333 (2013).

    Article  CAS  PubMed  Google Scholar 

  26. Puleston, D. J. et al. Polyamine metabolism is a central determinant of helper T cell lineage fidelity. Cell 184, 4186–4202.e20 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Wagner, A. et al. Metabolic modeling of single Th17 cells reveals regulators of autoimmunity. Cell 184, 4168–4185.e21 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Alexander, E. T., Minton, A., Peters, M. C., Phanstiel, O. T. & Gilmour, S. K. A novel polyamine blockade therapy activates an anti-tumor immune response. Oncotarget 8, 84140–84152 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  29. Hayes, C. S. et al. Polyamine-blocking therapy reverses immunosuppression in the tumor microenvironment. Cancer Immunol. Res. 2, 274–285 (2014).

    Article  ADS  CAS  PubMed  Google Scholar 

  30. Park, M. H., Nishimura, K., Zanelli, C. F. & Valentini, S. R. Functional significance of eIF5A and its hypusine modification in eukaryotes. Amino Acids 38, 491–500 (2010).

    Article  CAS  PubMed  Google Scholar 

  31. Hardbower, D. M. et al. Ornithine decarboxylase regulates M1 macrophage activation and mucosal inflammation via histone modifications. Proc. Natl Acad. Sci. USA 114, E751–E760 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Pasini, A., Caldarera, C. M. & Giordano, E. Chromatin remodeling by polyamines and polyamine analogs. Amino Acids 46, 595–603 (2014).

    Article  CAS  PubMed  Google Scholar 

  33. Perera, R. M. & Bardeesy, N. Pancreatic cancer metabolism: breaking it down to build it back up. Cancer Discov. 5, 1247–1261 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Croghan, M. K., Aickin, M. G. & Meyskens, F. L. Dose-related α-difluoromethylornithine ototoxicity. Am. J. Clin. Oncol. 14, 331–335 (1991).

    Article  CAS  PubMed  Google Scholar 

  35. Romijn, J. C., Verkoelen, C. F. & Splinter, T. A. Problems of pharmacokinetic studies on α-difluoromethylornithine in mice. Cancer Chemother. Pharmacol. 19, 30–34 (1987).

    Article  CAS  PubMed  Google Scholar 

  36. Iwami, K., Wang, J. Y., Jain, R., McCormack, S. & Johnson, L. R. Intestinal ornithine decarboxylase: half-life and regulation by putrescine. Am. J. Physiol. 258, G308–G315 (1990).

    CAS  PubMed  Google Scholar 

  37. Kobayashi, K., Morris, H. P. & Katunuma, N. Studies on the turnover rates of ornithine aminotransferase in Morris hepatoma 44 and host liver. J. Biochem. 80, 1085–1089 (1976).

    Article  CAS  PubMed  Google Scholar 

  38. Simell, O. & Takki, K. Raised plasma-ornithine and gyrate atrophy of the choroid and retina. Lancet 1, 1031–1033 (1973).

    Article  CAS  PubMed  Google Scholar 

  39. Bolkenius, F. N., Knodgen, B. & Seiler, N. dl-canaline and 5-fluoromethylornithine. Comparison of two inactivators of ornithine aminotransferase. Biochem. J. 268, 409–414 (1990).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Halmekyto, M., Alhonen, L., Alakuijala, L. & Janne, J. Transgenic mice over-producing putrescine in their tissues do not convert the diamine into higher polyamines. Biochem. J. 291, 505–508 (1993).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Zigmond, E. et al. Suppression of hepatocellular carcinoma by inhibition of overexpressed ornithine aminotransferase. ACS Med. Chem. Lett. 6, 840–844 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Mendoza, M. C., Er, E. E. & Blenis, J. The Ras–ERK and PI3K–mTOR pathways: cross-talk and compensation. Trends Biochem. Sci. 36, 320–328 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Bodoky, G. et al. A phase II open-label randomized study to assess the efficacy and safety of selumetinib (AZD6244 [ARRY-142886]) versus capecitabine in patients with advanced or metastatic pancreatic cancer who have failed first-line gemcitabine therapy. Invest. New Drugs 30, 1216–1223 (2012).

    Article  CAS  PubMed  Google Scholar 

  44. Kenney, C. et al. Phase II study of selumetinib, an orally active inhibitor of MEK1 and MEK2 kinases, in KRASG12R-mutant pancreatic ductal adenocarcinoma. Invest. New Drugs 39, 821–828 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Hingorani, S. R. et al. Preinvasive and invasive ductal pancreatic cancer and its early detection in the mouse. Cancer Cell 4, 437–450 (2003).

    Article  CAS  PubMed  Google Scholar 

  46. Jonkers, J. et al. Synergistic tumor suppressor activity of BRCA2 and p53 in a conditional mouse model for breast cancer. Nat. Genet. 29, 418–425 (2001).

    Article  CAS  PubMed  Google Scholar 

  47. Sastra, S. A. & Olive, K. P. Quantification of murine pancreatic tumors by high-resolution ultrasound. Methods Mol. Biol. 980, 249–266 (2013).

    Article  CAS  PubMed  Google Scholar 

  48. Wapnir, I. L., Wartenberg, D. E. & Greco, R. S. Three dimensional staging of breast cancer. Breast Cancer Res. Treat. 41, 15–19 (1996).

    Article  CAS  PubMed  Google Scholar 

  49. Samuels, Y. et al. Mutant PIK3CA promotes cell growth and invasion of human cancer cells. Cancer Cell 7, 561–573 (2005).

    Article  CAS  PubMed  Google Scholar 

  50. Debnath, J., Muthuswamy, S. K. & Brugge, J. S. Morphogenesis and oncogenesis of MCF-10A mammary epithelial acini grown in three-dimensional basement membrane cultures. Methods 30, 256–268 (2003).

    Article  CAS  PubMed  Google Scholar 

  51. Siddiqui, K. M. & Chopra, D. P. Primary and long term epithelial cell cultures from human fetal normal colonic mucosa. In Vitro 20, 859–868 (1984).

    Article  CAS  PubMed  Google Scholar 

  52. Marin-Valencia, I. et al. Analysis of tumor metabolism reveals mitochondrial glucose oxidation in genetically diverse human glioblastomas in the mouse brain in vivo. Cell Metab. 15, 827–837 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Mascanfroni, I. D. et al. Metabolic control of type 1 regulatory T cell differentiation by AHR and HIF1-α. Nat. Med. 21, 638–646 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Chong, J., Wishart, D. S. & Xia, J. Using MetaboAnalyst 4.0 for comprehensive and integrative metabolomics data analysis. Curr. Protoc. Bioinformatics 68, e86 (2019).

    Article  PubMed  Google Scholar 

  55. Bookout, A. L. & Mangelsdorf, D. J. Quantitative real-time PCR protocol for analysis of nuclear receptor signaling pathways. Nucl. Recept. Signal. 1, e012 (2003).

    Article  PubMed  PubMed Central  Google Scholar 

  56. Tang, Z. et al. GEPIA: a web server for cancer and normal gene expression profiling and interactive analyses. Nucleic Acids Res. 45, W98–W102 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  58. Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  59. 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 

  60. 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  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  61. The R Development Core Team. R: A language and Environment for Statistical Computing (R Foundation for Statistical Computing, 2016).

  62. Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Feng, J., Liu, T., Qin, B., Zhang, Y. & Liu, X. S. Identifying ChIP-seq enrichment using MACS. Nat. Protoc. 7, 1728–1740 (2012).

    Article  CAS  PubMed  Google Scholar 

  64. Consortium, E. P. An integrated encyclopedia of DNA elements in the human genome. Nature 489, 57–74 (2012).

    Article  ADS  Google Scholar 

  65. 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 

  66. Ramirez, F., Dundar, F., Diehl, S., Gruning, B. A. & Manke, T. deepTools: a flexible platform for exploring deep-sequencing data. Nucleic Acids Res. 42, W187–191 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Roehle, K. et al. cIAP1/2 antagonism eliminates MHC class I-negative tumors through T cell-dependent reprogramming of mononuclear phagocytes. Sci. Transl Med. 13, eabf5058 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank members of the N.Y.K. Laboratory for helpful comments and discussions and M.-S. Tsao (Princess Margaret Hospital, Toronto, Canada) for the HPDE cells. This work was supported by the NIH/National Cancer Institute (grant 1R01 CA270110-01 to N.Y.K.), an American Cancer Society Research Scholar Grant (RSG-17-070-01-TBG to N.Y.K.) and by funds from Boston Children’s Hospital. S.K.D., A.M.B. and M.H. were funded by NIH 1R01 AI158488-01 and the Hale Center for Pancreatic Cancer Research.

Author information

Authors and Affiliations

Authors

Contributions

M.-S.L. performed all experiments. I.N. and A.A. assisted M.-S.L. with mouse work. T.Z. provided early data for glutamine nitrogen flux into DNS and in vivo data with ARG2 knockdown or Arg2 KO (Extended Data Fig. 7k,l). L.L. performed the first in vivo infusion (Fig. 1e,f and Extended Data Fig. 2h,i) with assistance from M.-S.L. C.D., assisted by L.D., analysed and quantified all metabolite data, except those presented in Extended Data Figs. 1d,e and 7k,l, which were completed by D.S.H. C.B.C. supervised and discussed the metabolite quantification analyses. U.J. designed and supervised analyses for RNA-seq and ATAC-seq, which were analysed by A.L. with help from G.Y. S.K.D., A.M.B. and M.T.H. designed, conducted and analysed the immune profiling assays (Extended Data Fig. 11). Human whole tissue sections and tissue microarrays were provided by M.M.-K. with help from M.E.M. and J.L.B., with pathological assessment and scoring completed by M.M.-K. N.Y.K. conceived, designed and supervised the study and obtained funding. M.-S.L. and N.Y.K. wrote the manuscript with feedback from all authors.

Corresponding author

Correspondence to Nada Y. Kalaany.

Ethics declarations

Competing interests

S.K.D. received research funding unrelated to this project from Eli Lilly and Company, Novartis Pharmaceuticals, Genocea and Bristol-Myers Squibb and is a founder, science advisory board member and equity holder in Kojin. M.M.-K. has served as a compensated consultant for AstraZeneca, Bristol-Myers Squibb, Sanofi and Janssen Oncology and has received loyalties from Elsevier, all unrelated to the current work.

Peer review

Peer review information

Nature thanks Saverio Tardito 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 figures and tables

Extended Data Fig. 1 PDA favors the use of glutamine over arginine for polyamine synthesis.

a, Schematic depicting all 3 pathways leading to synthesis of the polyamine precursor ornithine: De novo ornithine synthesis (DNS) via OAT in red, urea cycle via ARG2 in yellow and creatine synthesis via GATM in purple. ARG, arginase; ASL, argininosuccinate lyase; ASS1, argininosuccinate synthase 1; CPS1, carbamoyl-phosphate synthase 1; GAMT, guanidinoacetate N-methyltransferase; GATM, glycine amidinotransferase; GLS, glutaminase; GSA, glutamate-γ-semialdehyde; α-KG, α-ketoglutarate; NOS, nitric oxide synthase; OAA, oxaloacetate; OAT, ornithine aminotransferase; ODC1, ornithine decarboxylase 1; OTC, ornithine transcarbamoylase; P5C, pyrroline-5-carboxylate; P5CS, pyrroline-5-carboxylate synthase (product of ALDH18A1 gene or aldehyde dehydrogenase 18 family member A1); P5CDH, pyrroline-5-carboxylate dehydrogenase (product of ALDH4A1 gene or aldehyde dehydrogenase 4 family member A1); PRODH, proline dehydrogenase 1; PYCR, pyrroline-5-carboxylate reductase; SMS, spermine synthase; SRM, spermidine synthase. b, c, Schematics tracing the fates of 15N-(amine) of glutamine (b) or all 4 nitrogens of 15N4-arginine (c) into ornithine and polyamine synthesis (left), or the urea cycle (right). Circles in White: 12C; in Red (b) or Green (c): 15N; in Gray: 14N; in Black: 14N when the urea cycle is off, as in PDA cells in vitro, but 15N when the urea cycle is on, as in PDA tumors in vivo7. Thicker arrows indicate enhanced flux into DNS and polyamine synthesis (b) or into generation of argininosuccinate, urea, ornithine and polyamines (c). d, Percent 15N-labeled metabolites in AsPC-1 cells fed 15N-(amine)Gln for 24 h. n = 6 biological replicates. e, Percent 15N-labeled metabolites in AsPC-1 cells fed 15N4-Arg for 24 h. Consistent with urea cycle inactivity, only 15N4-Arg-derived citrulline (M+3), the result of nitrogen oxide synthase (NOS) activity, but not citrulline (M+2), product of ornithine transcarbamoylase (OTC) in the urea cycle, was detected (see schematic in c). Furthermore, arginine-derived 15N-argininosuccinate (As, M+4) but not (M+2) was detected, indicating reverse argininosuccinate lyase (ASL) reaction10, rather than transfer of 15N from citrulline via argininosuccinate synthetase (ASS1) in the urea cycle (see schematic in c). n = 4 biological replicates. f,g, Percent labeled 15N-ornithine (f) and 15N-putrescine (g) in 29 cancer cell lines representing 5 cancer types (PDA; BRCA: breast carcinoma; LUAD, COAD and PRAD: adenocarcinomas of the lung, colon and prostate, respectively) with tissue-matched normal cell lines, indicated by arrowheads, that were fed 15N4-Arg for 24 h. n = 4 biological replicates per cell line. h,i, Relative abundance of 15N-labeled ornithine (h) or 15N-labeled putrescine (i) normalized to 15N-labeled glutamate in cell lines fed 15N-(amine)Gln for 24 h, as described in Fig. 1b, c. n = 4 biological replicates. M+1 and M+2 indicate a mass shift of 1 or 2 nitrogens, respectively. Data represent the mean ± s.d. p-values were obtained by one-way ANOVA, followed by Tukey test. Stars indicate statistical significance between each cancer cell line and its tissue-matched normal cell line/s. Data are representative of six (d), three (e) or two (fi) independent experiments.

Source Data

Extended Data Fig. 2 Enhanced de novo ornithine synthesis is a distinct feature of PDA.

a, Doubling times of cell lines in Fig. 1b, c. n = 8 biological replicates. b, Percent labeled 15N-proline in 15N-(amine)Gln-fed cells from Fig. 1b, c. n = 4 biological replicates. c, Percent labeled 15N-argininosuccinate (M+4) in 15N4-Arg-fed cells described in Extended Data Fig. 1f, g. n = 4 biological replicates. d,e, Percent 15N-labeled metabolites in AsPC-1 cells fed for 24 h, either 650 μM 15N-(amine)Gln (d) or 64 μM 15N4-Arg (e) in the presence of 64 μM unlabeled arginine (d) or 650 μM unlabeled glutamine (e). These amino acid concentrations reflect levels found in human plasma. n = 4 biological replicates. f,g, Percent labeled 15N-ornithine and 15N-putrescine in 11 cancer cell lines representing 3 cancer types (PDA; BRCA: breast carcinoma; LUAD, lung adenocarcinoma) with tissue-matched normal cell lines (arrowheads), that were fed 15N-(amine)Gln (f) or 15N4-Arg (g) and maintained for 24 h in plasma glutamine and arginine levels as described in d,e. n = 4 biological replicates. h,i, 15N enrichment in plasma glutamine (h) and percent 15N-labeled glutamine in normal pancreas or PDA tumors (i) derived from tumor-bearing iKrasG12D mice and non-tumor-bearing iKrasWT mice treated with Dox (2g l−1 drinking water) for 3 weeks prior to infusion with 15N-(amine)Gln for 1, 2 or 3 h as described in Fig. 1e–g. n = 4 mice per group. j, 15N enrichment in plasma arginine of iKras mice described in Fig. 1g, that were treated with Dox for 3 weeks prior to infusion with 15N4-Arg for 3 h. n = 4 mice per group. k,l, Percent 15N-labeled arginine (k) and relative abundance of total ornithine and putrescine (l) in normal pancreas or PDA tumors derived from either control or tumor-bearing mice described in Fig. 1g, that were infused with 15N4-Arg for 3 h. n = 4 mice per group. Data represent the mean ± s.d. p-values were obtained by one-way (b,c,f,g) or two-way (hj) ANOVA, followed by Tukey test, or unpaired two-tailed t-test (k,l). In b,f,g, statistical significance is for each cancer cell line vs. its tissue-matched normal cell line/s. In ag, data are representative of two independent experiments.

Source Data

Extended Data Fig. 3 Polyamines are enriched in PDA cells and their tumor microenvironment.

a, Principal component (PC) analysis of the abundance of 263 polar metabolites in plasma or tumor interstitial fluid (TIF) or normal interstitial fluid (NIF) from PDA tumors or normal pancreas of iKrasG12D or iKrasWT mice, respectively, described in Fig. 1h, i. Data were obtained by Metaboanalyst 4.0. n = 6 biological replicates. b, Top 10 enriched metabolic pathways in TIF compared to plasma (top); or TIF compared to NIF (bottom) based on abundance of metabolites in a. FDR false discovery rate. c, Volcano plot illustrating the fold change (log2-transformed) in metabolite levels between TIF and plasma of tumor-bearing iKrasG12D mice described in a. n = 6 biological replicates. Pink dots indicate significantly altered metabolites (> 1.5-fold; p < 0.01). d, Heatmaps listing in descending order of statistical significance (p < 0.05 by unpaired two-tailed t-test), metabolites in arginine metabolism from TIF vs. plasma of iKrasG12D mice (left) or TIF vs. NIF (right) described in a. n = 6 biological replicates. e,f, Relative live cell number (e) percent dead cells (f) in 4 KRAS-mutant and 1 non-KRAS-mutant (BxPC-3) PDA cell lines as well as normal HPDE cells grown in TIF arginine levels (2 µM). n = 8 biological replicates. g, Principal component (PC) analysis of the abundance of intracellular polar metabolites (263) in PDA tumors of iKrasG12D mice vs. normal pancreas of iKrasWT mice from a. n = 6 biological replicates. h, Top 10 enriched metabolic pathways in PDA tumors compared to normal pancreas of mice from g. i, Heatmap listing in descending order of statistical significance (p < 0.05 by unpaired two-tailed t-test), metabolites involved in arginine metabolism from PDA tumors and normal pancreas described in g. n = 6 biological replicates. j, Relative abundance of intracellular ornithine, putrescine, and spermidine in PDA tumors or normal pancreas described in g. n = 6 biological replicates. Data in ad and gi were obtained by Metaboanalyst 4.0. In d,i, Red indicates higher level and blue lower level, relative to the median. Data in e,f,j represent the mean ± s.d. p-values were obtained by two-way ANOVA followed by Tukey test (e,f) or by unpaired two-tailed t-test (j). In e,f, statistical significance is for days 3, 5 or 7 vs. day1. Data are representative of two independent experiments.

Source Data

Extended Data Fig. 4 OAT is enriched in human and mouse PDA tumors.

a, Gene expression analysis showing higher mRNA levels of OAT but not ALDH18A1 (aldehyde dehydrogenase 18 family member A1), ARG2 or ODC1 in human tumors compared to normal tissues. Data were derived from TCGA and GTEx datasets and represent 5 cancer types including pancreatic adenocarcinoma (PAAD), breast carcinoma (BRCA), lung, colon and prostate adenocarcinomas (LUAD, COAD and PRAD, respectively). TPM, transcript per million. T, tumor and N, normal tissue. Number of tissue samples n is indicated at bottom of panel. Box plots represent the interquartile range of data with the middle line being the median and whiskers spanning the minimum and maximum values. Plots and statistics were generated by GEPIA56 software. *p < 0.01 (one-way ANOVA). b, mRNA levels of genes in a quantified by qPCR, in 20 human cancer cell lines and 3 tissue-matched normal cell lines from the panel described in Fig. 1b, c. Ct indicates cycle threshold for each gene in the normal cell lines and is inversely proportional to mRNA levels. Data represent the mean of 3 technical replicates and are representative of two independent experiments. c, Representative immunohistochemical staining of OAT in PDA tumors and normal pancreas of tumor-bearing iKrasG12D or non-tumor-bearing iKrasWT mice treated with Dox (2g l−1 drinking water) for 3 weeks. Scale bar: 100 μm. d, Representative immunohistochemical OAT staining in human PDA, pancreatic intraepithelial neoplasia (PanIn) and adjacent normal cells in whole tissue sections of resected tumors from 4 different patients (Cases 1-4). Scale bar: 100 μm. In c,d, framed top corner insets represent a 5-fold magnification of the area delineated by a dashed square; data are representative of 2 independent experiments. e, H scores of OAT staining in PDA tumors and patient-matched normal pancreatic tissue performed on tissue microarrays (TMAs) of resected tumors from n = 31 patients. Data represent the mean ± s.d and p = 0.0003 was obtained by paired two-tailed t-test.

Extended Data Fig. 5 OAT is required for PDA growth.

a, Levels of ODC1, OAT or ARG2 proteins in AsPC-1 cells with knockdown of ODC1, OAT, ARG2 or GATM (left) and mRNA levels of GATM (right) in AsPC-1 cells with GATM knockdown as compared to Scramble (Scr). 2 hairpins per gene were used. b, Percent 15N-labeled metabolites in proline synthesis, urea cycle and creatine synthesis pathways in AsPC-1 cells with knockdown of ODC1, OAT, ARG2 or GATM (described in Fig. 2b, c), that were fed 15N4-Arg for 24 h. n = 4 biological replicates. c, Schematic demonstrating reversal of the OAT reaction upon ODC1 loss, accompanied by a compensatory increase in ARG2 and GATM activities to re-generate ornithine as demonstrated in b. d,e, Proliferation of AsPC-1 cells with knockdown of ODC1, OAT, ARG2, GATM or Scr (d) or those with knockdown of ODC1, OAT, or Scr, that were grown in the presence or absence of 10 μM putrescine (e). n = 8 biological replicates. f, Levels of ODC1 and OAT proteins in MIA PaCa-2 cells harboring ODC1, OAT or Scr knockdown. Arrowhead indicates non-specific band detected by ODC1 antibody (ab97395). g,h, Relative abundance of 15N-labeled putrescine (g) or total putrescine or ornithine (h) in cells from f, that were fed 15N-(amine)Gln or 15N4-Arg for 24 h. n = 4 biological replicates. i, mRNA levels of ODC1 and OAT genes in SUIT-2 cells with control Scr, ODC1, or OAT knockdown. j, Proliferation of MIA PaCa-2 and SUIT-2 cells with either ODC1, OAT, or Scr knockdown that were grown in the presence or absence of 10 μM putrescine. n = 8 biological replicates. Data represent the mean ± s.d. in a,b,gi, or mean ± s.e.m. in d,e,j. p-values were obtained by one-way ANOVA (a,b,gi) or two-way ANOVA (d,e,j), followed by Tukey test. In a and f, β-Actin was used as a loading control. Data are representative of three (a,d,e,f,i,j) or two (b,g,h) independent experiments.

Source Data

Extended Data Fig. 6 Silencing of OAT decreases polyamine pools in PDA cells and suppresses proliferation.

a, Levels of ODC1 and OAT proteins in AsPC-1 cells with control Scramble (Scr), ODC1 or OAT knockdown overexpressing either GFP control or respective rescue cDNAs for ODC1 or OAT. b,c, Relative abundance of 15N-labeled (b) or total (c) ornithine and putrescine in AsPC-1 cells described in a, that were fed 15N-(amine)Gln for 24 h. n = 4 biological replicates. d, Proliferation of AsPC-1 cells in a. n = 8 biological replicates. e, Levels of ODC1 and OAT proteins in HPDE cells with control Scr, ODC1 or OAT knockdown. f, Proliferation of HPDE cells from e. n = 8 biological replicates. g, Representative ultrasound images of orthotopic xenografts 3 weeks post-injection of AsPC-1 cells bearing knockdown of Scr, ODC1 or OAT into the pancreas of Rag1−/− mice, as described in Fig. 2f. T, Tumor; S, Spleen; K, Kidney. h, Volumes by ultrasound, of orthotopic human PDA tumors from g and Fig. 2f. n = 9 except for shODC1 #2, n = 8 mice per group. i, Levels of ODC1 and OAT proteins in 2 iKras cell lines with control Scr, Odc1 or Oat knockdown (2 hairpins per gene). j,k, Relative abundance of 15N-labeled (j) or total (k) ornithine and putrescine in iKras cells from i. n = 4 biological replicates. l, Proliferation of iKras cells from i, grown in the presence or absence of 10 μM putrescine. n = 8 biological replicates. Data represent the mean ± s.d. (b,c,h,j,k) or mean ± s.e.m. (d,f,l). p-values were obtained by one-way (b,c,j,k) or two-way (d,f,h,l) ANOVA, followed by Tukey test. In d,l, statistical significance is for each condition vs. control “shScr + GFP” (d) or each gene knockdown vs. Scr control in the presence or absence of putrescine (l). In e,i, Arrowhead indicates non-specific band detected by ODC1 antibody (ab97395). In a,e,i, β-Actin was used as a loading control. Data are representative of two (ac, ik) or three (d,e,f,l) independent experiments.

Source Data

Extended Data Fig. 7 OAT maintains in vivo tumor polyamine pools supporting PDA growth.

a, Generation of mouse orthotopic PDA transplants lacking Odc1 or Oat. iKras clonal cell lines with CRISPR/Cas9 knockout of Odc1 (clones #3 and #10) or Oat (clones #10 and #11) using single guide (sg) RNAs each, or sgControl (sgCtrl) were injected (5 x 105 cells) into the pancreas of non-Cre-expressing iKras mice. All mice were treated with Dox (2g l−1 drinking water) and monitored for tumor growth by ultrasound over 3 weeks, prior to subjecting them to a 3-hour infusion with 15N-(amine)Gln. b, Levels of ODC1 and OAT proteins in iKras cells from a. c,d, Relative abundance of 15N-labeled (c) or total (d) ornithine and putrescine in iKras cells from a that were fed for 24 h 15N-(amine)Gln or 15N4-Arg. n = 4 biological replicates. e, Proliferation of iKras cells described in a. n = 8 biological replicates. fh, Volumes (ultrasound) of orthotopic iKras tumors with control (Ctrl), Odc1 or Oat knockout from a, injected into non-tumor-bearing mice (f), with representative ultrasound images (g), and weights (h), 3 weeks post-injection of iKras cells. n = 9 (sgCtrl); n = 10 (sgOdc1 #10; sgOat #10); n = 6 (sgOdc1 #3; sgOat #11). T, Tumor; S, Spleen. i, 15N enrichment in plasma glutamine of tumor-bearing mice in fh, that were infused with 15N-(amine)Gln over 3 h, 3 weeks post-tumor cell injection (related to Fig. 2g). n = 4 mice per group. j, Relative abundance of indicated metabolites in orthotopic tumors derived from iKras cells with knockout for Odc1, Oat, or control, described in a. n = 6 mice per group. k, Relative abundance of putrescine and spermidine in human orthotopic PDA tumors derived from AsPC-1 cells with ARG2 or Scramble (Scr) knockdown, that were injected (105 cells) into the pancreas of Rag1−/− mice and grown for 6 weeks. n = 7 mice per group. l, Relative abundance of the indicated metabolites in mouse orthotopic transplant tumors derived from KPC cells (see Methods) expressing or lacking Arg2 (Arg2+/+ or Arg2−/−, respectively) that were injected (2.5 x 105 cells) into the pancreas of mice of the same strain and grown for 2 weeks. n = 6 mice per group. In jl, Box plots represent medians ± 10-90 percentile and whiskers span minimum and maximum values. m, Levels of OAT and ODC1 proteins in iKras cells with control (sgCtrl) or Odc1 or Oat knockout (sgOdc1 or sgOat) that are overexpressing GFP control or the respective gene cDNA (Odc1 or Oat). n,o, Relative abundance of 15N-labeled (n) or total (o) ornithine and putrescine levels in iKras cells described in m, that were fed 15N-(amine)Gln for 24 h. n = 4 biological replicates. p, Proliferation of iKras cells described in m. n = 8 biological replicates. q, Volumes of syngeneic orthotopic tumor transplants derived from iKras cells described in m that were grown and quantified by ultrasound, 2 and 3 weeks post-cell injection (related to Fig. 2h). n = 8 mice (sgCtrl + GFP); n = 7 (sgCtrl + Oat); n = 9 (sgOat #10 + GFP; sgOat #10 + Oat). r, Metabolite abundance in TIF or plasma of mice related to Fig. 2i. n = 8 except sgCtrl + Oat (7). In i,j, mr, sgOdc1 clone #10 and sgOat clone #10 from ah were used. Data represent the mean ± s.d. (c,d,f,h,i,n,o,q,r) or mean ± s.e.m. (e,p). p-values were obtained by one-way (c,d,h,j,n,o) or two-way (e,f,i,p,q,r) ANOVA followed by Tukey test, or by unpaired two-tailed t-test (k,l). In e, np, statistical significance is for each condition vs. control knockout “sgCtrl” (e) or each condition vs. “sgCtrl + GFP” (np). In b,m, Arrowhead indicates non-specific band detected by ODC1 antibody (ab97395) and β-Actin was used as a loading control. Data are representative of two (bd, mo) or three (e,p) independent experiments.

Source Data

Extended Data Fig. 8 Oncogenic KRAS induces the expression of OAT and polyamine synthesis genes.

a, mRNA levels of Srm, Sms and Arg2 in iKras cell lines #1 and #2 maintained in Dox (1μg ml−1) for 24 h prior to Dox deprivation for 24, 48, or 72 h. n = 3 biological replicates. b,c, mRNA levels of KRAS, OAT, ODC1, SRM, SMS and ARG2 in human PDA AsPC-1 (b) or MIA PaCa-2 (c) cells with Dox-inducible knockdown of GFP (Tet on-shGFP) or KRAS (Tet on-shKRAS hairpins #1 and #2) that were cultured in Dox (1μg ml−1) for 24, 48 or 72 h. In b, n = 3 and in c, n =2 biological replicates. d, mRNA levels of ornithine and polyamine synthesis genes (OAT, ODC1, SRM, SMS and ARG2) in AsPC-1 and MIA PaCa-2 cells treated with vehicle control DMSO or inhibitors of: PI3K (BKM 120, 150 nM); AKT (MK2206, 200 nM); MEK (AZD6244, 50 nM); mTORC1 (Rapamycin, 20 nM) for 72 h. n = 3 biological replicates. Data represent the mean ± s.d. (a,b,d) of biological replicates from 3 independent experiments or the mean of biological replicates from 2 independent experiments (c). p-values were obtained by two-way (a,b) or one-way (d) ANOVA followed by Tukey test. Statistical significance is for Off Dox vs. On Dox (a) or for shKRAS vs. shGFP (b) at each indicated time, or for each inhibitor vs. DMSO (d). e, Levels of proteins in cells from d. β-Actin was used as loading control and data represent two independent experiments.

Extended Data Fig. 9 Transcription factor KLF6 mediates KRAS-driven polyamine synthesis in PDA.

a, b, Putative transcription factor (TF) binding sites identified and scored by FMatch tool in the promoter regions of all 4 ornithine and polyamine synthesis genes responsive to KRASG12D, i.e. OAT, ODC1, SRM and SMS. Predicted TFs are listed in descending order of statistical significance (p-value cut-off of 0.01), with matrix score similarity ranging from 0-1, where 1 is a perfect match. Eight are conserved in human (a) and mouse (b), out of which six (in blue) have binding sites present in the OAT promoter. c, mRNA levels of Myc, Oat, Odc1, Srm, Sms and Arg2 in 2 iKras cell lines with Scramble control (Scr) or Myc knockdown (2 distinct hairpins). n = 2 biological replicates. d, mRNA levels validating knockdown (2 or 3 hairpins per gene) of each of the 6 conserved TFs with binding sites for OAT (blue in a,b), and assessing their effects on expression of Oat, Odc1, Srm, Sms and Arg2 in iKras #1 cell line. n = 3 biological replicates. e, mRNA levels of ornithine and polyamine synthesis genes (Oat, Odc1, Srm, Sms and Arg2) in iKras #2 cells with Scr or Klf6 knockdown (2 distinct hairpins). n = 3 biological replicates. f, Levels of OAT, ODC1, SRM, SMS proteins in iKras cells with Klf6 or Scramble (Scr) knockdown, with or without Dox (72 h). β-Actin was used as loading control. g, Fold change in mRNA levels of Myc and each of the 6 TFs from d in iKras #1 cell line deprived of Dox for 48 h. n = 3 biological replicates. h, Relative abundance of total ornithine and putrescine in iKras#1 cells described in Fig. 3h. Cells were fed 15N-(amine)Gln or 15N4-Arg 24 h before harvest. n = 4 biological replicates. In ce, g,h, data represent the mean ± s.d. p-values were obtained by one-way (d,e) or two-way (h) ANOVA followed by Tukey test, or paired two-tailed t-test (g). In d,e,g, statistical significance is for each TF knockdown vs. control Scramble (d,e), or for Off Dox vs. On Dox (g). Data represent the mean of biological replicates from 2 independent experiments (c) or mean ± s.d. of biological replicates from 3 independent experiments (d,e,g). In f,h, data are representative of 2 independent experiments.

Source Data

Extended Data Fig. 10 5-FMO suppresses polyamine synthesis and PDA growth similar to DFMO but without off-target effects.

a, Relative abundance of 15N-labeled ornithine and putrescine in iKras cells #1 treated for 72 h with inhibitors of OAT (5-FMO) or ODC1 (DFMO) and fed 15N-(amine)Gln or 15N4-Arg 24 h before harvest. n = 4 biological replicates. b,c, Relative abundance of total ornithine and putrescine in AsPC-1 (b, related to Fig. 4a) or iKras (c) cells treated as in a. n = 4 biological replicates. d, Percent growth of human PDA cells AsPC-1 and MIA PaCa-2 with knockdown of OAT #1 or Scramble (Scr) that were treated with 5-FMO (0, 0.001, 0.025, 0.05, 0.1, 0.25, 0.5, 0.75, 1, 2.5 mM) for 7 days. e, Percent growth over 7 days of AsPC-1 cells treated with 5-FMO as in d, in the presence or absence of putrescine (10 μM). f, Percent growth of iKras #1 cells with control (Ctrl) or Oat knockout, that were treated with 5-FMO as in d. g,h, Percent growth of human PDA cells with Scramble or ODC1 #1 knockdown (g) or mouse iKras #1 cells with Ctrl or Odc1 knockout (h) that were treated with DFMO (0, 0.001, 0.025, 0.05, 0.1, 0.25, 0.5, 0.75, 1, 2.5 mM) for 7 days. In f,h, sgOdc1 clone #10 and sgOat clone #10 from Extended Data Fig. 7a were used. In dh, n = 8 biological replicates and arrowheads indicate concentrations used in ac. i, Relative abundance of 13C-labeled (M+4) and total putrescine in AsPC-1 cells with knockdown of control Scramble (Scr), ODC1 or OAT that were treated with vehicle control (water) or 10 μM 13C4-putrescine for 1 h. n = 4 biological replicates. j, Relative abundance of 13C-labeled (M+4) and total putrescine in AsPC-1 cells that were pre-treated (or non-pre-treated) for 15 min with 1 μM AMXT-1501, then fed 10 μM 13C4-putrescine for 0, 50, 100, 150, 200 and 240 min while continuing the same AMXT-1501 treatment (0 or 1 μM) but in absence or presence of 5-FMO (0 or 100 μM). n = 4 biological replicates. In i,j, data represent the mean of n = 4 biological replicates from two independent experiments. k, Proliferation of AsPC-1 cells that were treated or not treated with 1 μM AMXT-1501 and/or 100 μM 5-FMO for 7 days in the presence or absence of 10 μM putrescine. Data represent the mean of n = 32 biological replicates from four independent experiments with n = 8 per experiment. l, 15N enrichment in plasma glutamine of tumor-bearing male iKrasG12D mice from Fig. 4f that were treated with control saline or 5-FMO (10 mg kg−1 and 30 mg kg−1) for 14 days, prior to infusing them for 3 h with 15N-(amine)Gln, as described in Fig. 4g. n = 4 mice per group. m, Relative abundance of total ornithine and putrescine in tumors of 5-FMO treated male iKras mice, related to Fig. 4g. n = 4. n,o, Weights of livers (n) and growth curves (o) of male or female iKrasG12D mice described in Fig. 4f. Male mice: n = 11 per dose; Female mice: n = 10 for saline, n = 9 for 10mg kg−1 and n = 11 for 30mg kg−1. p, Proliferation fold increase in number over 7 days of AsPC-1 cells treated with gemcitabine at 0, 0.1, 0.2, 0.4, 0.6, 0.8, 1, 2 μM in the presence or absence of 5-FMO (0, 100 or 500 μM). n = 8 biological replicates. Data represent the mean ± s.d. (ac, in, p) or mean ± s.e.m. (dh, o). p-values were obtained by one-way (ac, m, n) or two-way (il, o, p) ANOVA followed by Tukey test or paired two-tailed t-test (dh). In i, indicated p-values are for each knockdown in the presence vs. absence of putrescine at the indicated times; † and ‡ indicate significance of p < 0.0001 for each gene knockdown vs. Scr knockdown in absence (†) of putrescine; or in presence (‡) of putrescine, except for 13C4-putrescine in shOAT #1 and total putrescine in shOAT #2, where p = 0.0003. In j, significance is for each drug condition vs. vehicle control (black line). In p, † (p = 0.0006) and ‡ (p < 0.0001) indicate statistical significance between each gemcitabine-treated group and non-treated control in the absence of 5-FMO (grey bars). Data are representative of two (ac, g,h), three (df,p) independent experiments.

Source Data

Extended Data Fig. 11 Immune profiling reveals no major differences in immune subsets upon genetic or pharmacological inhibition of OAT.

a, Analysis of select immune populations in orthotopic tumors derived from iKras cells with knockout of Odc1, Oat, or control (Ctrl) that were injected (5 x 105 cells) into the pancreas of non-Cre-expressing iKras mice and grown for 3 weeks. bf, Immune profiling of PDA tumors from male and female iKrasG12D mice that were treated with Dox (2g l−1 drinking water) for 7 days, then daily intraperitoneally injected with control saline or 5-FMO (10 mg kg−1 and 30 mg kg−1) for 14 days while under continuous treatment with Dox. Single cell infiltrates from tumors and draining lymph nodes were analyzed by flow cytometry. Analysis of select immune populations (b), ratio of Tregs to CD8+ T cells and percent PD1+ CD8+ T of total CD8+ T cells (c) in PDA tumors. Analysis of select immune populations in pancreatic draining lymph nodes (dLN, d). Representative flow plots showing Tregs (CTLA-4+ and FOXP3+) in PDA tumor (e) and pancreatic draining lymph node (f) derived from saline-treated iKrasG12D mice. For tumor samples, values indicate the average of 2 replicate samples obtained from different regions of the tumor mass. n = 9-10 mice per group using both sexes (male: n = 5 per group; female: n = 4 for saline and 10 mg kg−1 5-FMO, n = 5 for 30 mg kg−1 5-FMO). In ad, Data represent the mean ± s.d. p-values were obtained by one-way ANOVA followed by Tukey test. For gating details, analysis of select immune populations gated out of CD45+ live, single cells: CD11b+/− = CD45+; MoMDSCs = CD11b+ Gr1low; GrMDSCs = CD11b+ Gr1high; macrophages = CD11b+ SiglecF Gr1; eosinophils = CD11b+ SiglecF+; CD103+/MHCIIhigh; DCs = CD11b CD4 CD8 CD11c+ MHCII+; CD4 T cells = CD11b CD4+; CD8 T cells = CD11b CD8+; Tregs = CD11b CD4+ FOXP3+ CTLA4+; B cells = CD11b CD4 CD8 B220+. All positive gates were set based on clearly separated populations except in the case of Gr-1 staining, which was divided into tertiles based on mean fluorescence intensity. Unstained cells were used to define negative gates. Forward scatter and side scatter were used to identify cell-sized objects. Fragments with FSC less than 10% of the sample median were excluded from analysis. Cell doublets were excluded based on deviation from the diagonal of FSC area vs FSC height.

Extended Data Fig. 12 OAT silencing results in transcriptional and epigenetic changes similar to those of ODC1 knockdown.

a, Venn diagrams showing a higher overlapping number of differentially expressed genes (q < 0.05) in AsPC-1 cells with knockdown of OAT or ODC1 (608 down-regulated and 599 up-regulated), as compared to ARG2 and ODC1 or ARG2 and OAT. Numbers reflect differential expression common to 2 hairpins per gene. b, ATAC-Seq data showing significant changes (q < 0.05) in chromatin access at enhancers located < 20 kb from transcription start sites in AsPC-1 cells with knockdown of ODC1, compared to control Scramble (Scr), OAT, or ARG2 (2 hairpins per gene). n = 2 biological replicates per hairpin. Significant changes detected include losses and gains of chromatin accessibility near 175 and 135 enhancers, respectively. c, Correlation of changes in gene expression and nearby (< 25 kb) chromatin accessibility for 403 genes that were consistently up-regulated upon ODC1 and OAT silencing in AsPC-1 cells (cluster V, Fig. 4h). Each dot represents a gene. Gene-linked enhancers (< 25 kb) show gains in accessibility upon knockdown of ODC1 or OAT, but not ARG2. d, List of top 18 pathways in descending significance (FDR 0.003-0.30) that were differentially altered at the transcriptional level (either negative or positive enrichment) upon ODC1 knockdown compared to Scr control in AsPC-1 PDA cells described in a, b. Negatively enriched pathways related to growth factors, cytokines and response to starvation are highlighted in yellow and those involved in cell shape, migration, differentiation and ion transport are highlighted in blue. Asterisks indicate representative pathways illustrated in e and Fig. 4j. e, Heatmaps of genes from representative pathways (blue) in d displaying negative enrichment upon knockdown of ODC1 compared to Scr control (n = 3), with corresponding GSEA plots. NES normalized enrichment scores, FDR false discovery rate, Nom. Nominal. f, List of differentially expressed genes from top 18 negatively enriched pathways in d, that are part of Cluster I in Fig. 4h and are indicated with dark blue dots in the correlation plots of Fig. 4i. Genes in Quadrant (Q) III of Fig. 4i plots associate with decreased chromatin accessibility, and are more numerous in cells with knockdown of ODC1 (37/58) or OAT (44/58), than ARG2 (22/58). g, Relative mRNA levels (by qPCR) of 7 representative genes in AsPC-1 cells with knockdown of ODC1, OAT, or ARG2 (2 hairpins per gene) compared to control Scr, that were maintained for 72 h in the presence or absence of 10 μM Putrescine. Genes were randomly chosen from Q III described in f and Fig. 4i, and display common decreases in expression upon OAT or ODC1 knockdown but not ARG2, along with concordant decreases in chromatin access. Data represent the mean ± s.d. of 3 biological replicates from 3 independent experiments. p-values were obtained by two-way ANOVA followed by Tukey test. Statistical significance is for knockdown of ODC1, OAT or ARG2 vs. Scr under similar putrescine condition, unless otherwise indicated by lines.

Supplementary information

Supplementary Fig. 1

Uncropped immunoblots for Fig. 3 and Extended Data Figs. 5–9.

Reporting Summary

Supplementary Tables

Supplementary Table 1: Antibodies. Supplementary Table 2: Chemicals and other reagents. Supplementary Table 3: Genetic alterations in cancer cell lines. Supplementary Table 4: shRNA and sgRNA sequences. Supplementary Table 5: qPCR primer sequences.

Source data

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lee, MS., Dennis, C., Naqvi, I. et al. Ornithine aminotransferase supports polyamine synthesis in pancreatic cancer. Nature 616, 339–347 (2023). https://doi.org/10.1038/s41586-023-05891-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-023-05891-2

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: Translational Research

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

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