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:

Disruption of sugar nucleotide clearance is a therapeutic vulnerability of cancer cells

Abstract

Identifying metabolic steps that are specifically required for the survival of cancer cells but are dispensable in normal cells remains a challenge1. Here we report a therapeutic vulnerability in a sugar nucleotide biosynthetic pathway that can be exploited in cancer cells with only a limited impact on normal cells. A systematic examination of conditionally essential metabolic enzymes revealed that UXS1, a Golgi enzyme that converts one sugar nucleotide (UDP-glucuronic acid, UDPGA) to another (UDP-xylose), is essential only in cells that express high levels of the enzyme immediately upstream of it, UGDH. This conditional relationship exists because UXS1 is required to prevent excess accumulation of UDPGA, which is produced by UGDH. UXS1 not only clears away UDPGA but also limits its production through negative feedback on UGDH. Excess UDPGA disrupts Golgi morphology and function, which impedes the trafficking of surface receptors such as EGFR to the plasma membrane and diminishes the signalling capacity of cells. UGDH expression is elevated in several cancers, including lung adenocarcinoma, and is further enhanced during chemoresistant selection. As a result, these cancer cells are selectively dependent on UXS1 for UDPGA detoxification, revealing a potential weakness in tumours with high levels of UGDH.

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: UXS1 is essential for cell survival in a manner that correlates with cell-line-specific expression of the upstream enzyme UGDH.
Fig. 2: UXS1 is required to detoxify the UDP-glucuronic acid produced by UGDH.
Fig. 3: UDPGA accumulation causes Golgi dysfunction that leads to EGFR inactivation.
Fig. 4: UXS1 is a cancer-selective therapeutic target because of elevated UGDH expression in cancer subsets and chemoresistant cancer cells.

Similar content being viewed by others

Data availability

The transcriptomic data included in this study have been deposited to the Gene Expression Omnibus (GEO) (accession number: GSE240344). Uncropped western blots are provided in the Supplementary Information. Additional raw data and materials that support the findings of the study are available from the corresponding author on request. Source data are provided with this paper.

References

  1. Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: the next generation. Cell 144, 646–674 (2011).

    Article  CAS  PubMed  Google Scholar 

  2. Kim, D. et al. SHMT2 drives glioma cell survival in ischaemia but imposes a dependence on glycine clearance. Nature 520, 363–367 (2015).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  3. Carlisle, A. E. et al. Selenium detoxification is required for cancer-cell survival. Nat. Metab. 2, 603–611 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Spears, M. E. et al. De novo sphingolipid biosynthesis necessitates detoxification in cancer cells. Cell Rep. 40, 111415 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Lee, N., Spears, M. E., Carlisle, A. E. & Kim, D. Endogenous toxic metabolites and implications in cancer therapy. Oncogene 39, 5709–5720 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Meyers, R. M. et al. Computational correction of copy number effect improves specificity of CRISPR–Cas9 essentiality screens in cancer cells. Nat. Genet. 49, 1779–1784 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Barretina, J. et al. The Cancer Cell Line Encyclopedia enables predictive modelling of anticancer drug sensitivity. Nature 483, 603–607 (2012).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  8. Axelrod, J., Kalckar, H. M., Maxwell, E. S. & Strominger, J. L. Enzymatic formation of uridine diphosphoglucuronic acid. J. Biol. Chem. 224, 79–90 (1957).

    Article  CAS  PubMed  Google Scholar 

  9. Prydz, K. & Dalen, K. T. Synthesis and sorting of proteoglycans. J. Cell Sci. 113, 193–205 (2000).

    Article  CAS  PubMed  Google Scholar 

  10. Mulder, G. J. Glucuronidation and its role in regulation of biological activity of drugs. Annu. Rev. Pharmacol. Toxicol. 32, 25–49 (1992).

    Article  CAS  PubMed  Google Scholar 

  11. Moriarity, J. L. et al. UDP-glucuronate decarboxylase, a key enzyme in proteoglycan synthesis: cloning, characterization, and localization. J. Biol. Chem. 277, 16968–16975 (2002).

    Article  CAS  PubMed  Google Scholar 

  12. Kakizaki, I. et al. A novel mechanism for the inhibition of hyaluronan biosynthesis by 4-methylumbelliferone. J. Biol. Chem. 279, 33281–33289 (2004).

    Article  CAS  PubMed  Google Scholar 

  13. Kultti, A. et al. 4-Methylumbelliferone inhibits hyaluronan synthesis by depletion of cellular UDP-glucuronic acid and downregulation of hyaluronan synthase 2 and 3. Exp. Cell. Res. 315, 1914–1923 (2009).

    Article  CAS  PubMed  Google Scholar 

  14. Clarkin, C. E., Allen, S., Wheeler-Jones, C. P., Bastow, E. R. & Pitsillides, A. A. Reduced chondrogenic matrix accumulation by 4-methylumbelliferone reveals the potential for selective targeting of UDP-glucose dehydrogenase. Matrix Biol. 30, 163–168 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Kjellén, L. & Lindahl, U. Proteoglycans: structures and interactions. Annu. Rev. Biochem. 60, 443–475 (1991).

    Article  PubMed  Google Scholar 

  16. Winkler, J., Abisoye-Ogunniyan, A., Metcalf, K. J. & Werb, Z. Concepts of extracellular matrix remodelling in tumour progression and metastasis. Nat. Commun. 11, 5120 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  17. Vitale, D. et al. Proteoglycans and glycosaminoglycans as regulators of cancer stem cell function and therapeutic resistance. FEBS J. 286, 2870–2882 (2019).

    Article  CAS  PubMed  Google Scholar 

  18. Neufeld, E. F. & Hall, C. W. Inhibition of UDP-D-glucose dehydrogenase by UDP-D-xylose: a possible regulatory mechanism. Biochem. Biophys. Res. Commun. 19, 456–461 (1965).

    Article  CAS  PubMed  Google Scholar 

  19. Gainey, P. A. & Phelps, C. F. Interactions of uridine diphosphate glucose dehydrogenase with the inhibitor uridine diphosphate xylose. Biochem. J 145, 129–134 (1975).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Beattie, N. R., Keul, N. D., Sidlo, A. M. & Wood, Z. A. Allostery and hysteresis are coupled in human UDP-glucose dehydrogenase. Biochemistry 56, 202–211 (2017).

    Article  CAS  PubMed  Google Scholar 

  21. Reiling, J. H. et al. A CREB3–ARF4 signalling pathway mediates the response to Golgi stress and susceptibility to pathogens. Nat. Cell Biol. 15, 1473–1485 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Ignashkova, T. I. et al. Cell survival and protein secretion associated with Golgi integrity in response to Golgi stress-inducing agents. Traffic 18, 530–544 (2017).

    Article  CAS  PubMed  Google Scholar 

  23. Hiraoka, S. et al. Nucleotide-sugar transporter SLC35D1 is critical to chondroitin sulfate synthesis in cartilage and skeletal development in mouse and human. Nat. Med. 13, 1363–1367 (2007).

    Article  CAS  PubMed  Google Scholar 

  24. Stanley, P. Golgi glycosylation. Cold Spring Harb. Perspect. Biol. 3, a005199 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  25. Varki, A. Factors controlling the glycosylation potential of the Golgi apparatus. Trends Cell Biol. 8, 34–40 (1998).

    Article  CAS  PubMed  Google Scholar 

  26. Tarentino, A. L., Gomez, C. M. & Plummer, T. H. Jr Deglycosylation of asparagine-linked glycans by peptide:N-glycosidase F. Biochemistry 24, 4665–4671 (1985).

    Article  CAS  PubMed  Google Scholar 

  27. Unal, E. S., Zhao, R., Qiu, A. & Goldman, I. D. N-linked glycosylation and its impact on the electrophoretic mobility and function of the human proton-coupled folate transporter (HsPCFT). Biochim. Biophys. Acta 1778, 1407–1414 (2008).

    Article  PubMed  Google Scholar 

  28. Contessa, J. N., Bhojani, M. S., Freeze, H. H., Rehemtulla, A. & Lawrence, T. S. Inhibition of N-linked glycosylation disrupts receptor tyrosine kinase signaling in tumor cells. Cancer Res. 68, 3803–3809 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Ornitz, D. M. & Itoh, N. The fibroblast growth factor signaling pathway. Wiley Interdiscip. Rev. Dev. Biol. 4, 215–266 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Hakuno, F. & Takahashi, S.-I. IGF1 receptor signaling pathways. J. Mol. Endocrinol. 61, T69–T86 (2018).

    Article  CAS  PubMed  Google Scholar 

  31. Ouhtit, A., Rizeq, B., Saleh, H. A., Rahman, M. M. & Zayed, H. Novel CD44-downstream signaling pathways mediating breast tumor invasion. Int. J. Biol. Sci. 14, 1782–1790 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Normanno, N. et al. Epidermal growth factor receptor (EGFR) signaling in cancer. Gene 366, 2–16 (2006).

    Article  CAS  PubMed  Google Scholar 

  33. Jayaprakash, N. G. & Surolia, A. Role of glycosylation in nucleating protein folding and stability. Biochem. J. 474, 2333–2347 (2017).

    Article  CAS  PubMed  Google Scholar 

  34. Lis, H. & Sharon, N. Protein glycosylation. Structural and functional aspects. Eur. J. Biochem. 218, 1–27 (1993).

    Article  CAS  PubMed  Google Scholar 

  35. Taylor, E. S., Pol-Fachin, L., Lins, R. D. & Lower, S. K. Conformational stability of the epidermal growth factor (EGF) receptor as influenced by glycosylation, dimerization and EGF hormone binding. Proteins 85, 561–570 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Saha, S. et al. UDP-glucose 6-dehydrogenase expression as a predictor of survival in patients with pulmonary adenocarcinoma. Int. J. Surg. Oncol. 5, e85 (2020).

    Google Scholar 

  37. Hagiuda, D. et al. Clinicopathological and prognostic significance of nuclear UGDH localization in lung adenocarcinoma. Biomed. Res. 40, 17–27 (2019).

    Article  CAS  PubMed  Google Scholar 

  38. Arnold, J. M. et al. UDP-glucose 6-dehydrogenase regulates hyaluronic acid production and promotes breast cancer progression. Oncogene 39, 3089–3101 (2020).

    Article  CAS  PubMed  Google Scholar 

  39. Teoh, S. T., Ogrodzinski, M. P. & Lunt, S. Y. UDP-glucose 6-dehydrogenase knockout impairs migration and decreases in vivo metastatic ability of breast cancer cells. Cancer Lett. 492, 21–30 (2020).

    Article  CAS  PubMed  Google Scholar 

  40. Lin, L.-H. et al. Targeting UDP-glucose dehydrogenase inhibits ovarian cancer growth and metastasis. J. Cell. Mol. Med. 24, 11883–11902 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Tang, T. et al. A mouse knockout library for secreted and transmembrane proteins. Nat. Biotechnol. 28, 749–755 (2010).

    Article  CAS  PubMed  Google Scholar 

  42. Vitale, D. L. et al. Initial identification of UDP-glucose dehydrogenase as a prognostic marker in breast cancer patients, which facilitates epirubicin resistance and regulates hyaluronan synthesis in MDA-MB-231 cells. Biomolecules 11, 246 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Ahmed, N. et al. Unique proteome signature of post-chemotherapy ovarian cancer ascites-derived tumor cells. Sci. Rep. 6, 30061 (2016).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  44. Paul, D. et al. Global proteomic profiling identifies etoposide chemoresistance markers in non-small cell lung carcinoma. J. Proteomics 138, 95–105 (2016).

    Article  CAS  PubMed  Google Scholar 

  45. Radominska-Pandya, A., Bratton, S. M., Redinbo, M. R. & Miley, M. J. The crystal structure of human UDP-glucuronosyltransferase 2B7 C-terminal end is the first mammalian UGT target to be revealed: the significance for human UGTs from both the 1A and 2B families. Drug Metab. Rev. 42, 133–144 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Mackenzie, P. I. et al. Regulation of UDP glucuronosyltransferase genes. Curr. Drug Metab. 4, 249–257 (2003).

    Article  CAS  PubMed  Google Scholar 

  47. Nagar, S. & Blanchard, R. L. Pharmacogenetics of uridine diphosphoglucuronosyltransferase (UGT) 1A family members and its role in patient response to irinotecan. Drug Metab. Rev. 38, 393–409 (2006).

    Article  CAS  PubMed  Google Scholar 

  48. Corsello, S. M. et al. Discovering the anticancer potential of non-oncology drugs by systematic viability profiling. Nat. Cancer 1, 235–248 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. 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  ADS  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Barger, C. J., Branick, C., Chee, L. & Karpf, A. R. Pan-cancer analyses reveal genomic features of FOXM1 overexpression in cancer. Cancers 11, 251 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Nonnenmacher, Y. et al. Analysis of mitochondrial metabolism in situ: Combining stable isotope labeling with selective permeabilization. Metab. Eng. 43, 147–155 (2017).

    Article  CAS  PubMed  Google Scholar 

  53. Vercesi, A. E., Bernardes, C. F., Hoffmann, M. E., Gadelha, F. R. & Docampo, R. Digitonin permeabilization does not affect mitochondrial function and allows the determination of the mitochondrial membrane potential of Trypanosoma cruzi in situ. J. Biol. Chem. 266, 14431–14434 (1991).

    Article  CAS  PubMed  Google Scholar 

  54. Sachs, N. et al. A living biobank of breast cancer organoids captures disease heterogeneity. Cell 172, 373–386 (2018).

    Article  CAS  PubMed  Google Scholar 

  55. 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 

Download references

Acknowledgements

We thank M. Walker, R. Cummings, C. Gao, K. Reddig and M. Pacold for advice, discussion and other assistance and D. Sabatini, S. Wolfe, W. Xue, M. Green and M. Lee for materials, including vectors and cell lines. This work was supported by the Suh Kyungbae Foundation Young Investigator Award and National Institutes of Health grant R01CA269711 to D.K. and by a Mogam Fellowship to N.L. The electron-microscopy work was supported by award numbers S10OD025113-01 and S10OD021580 from the National Center for Research Resources. O.P. and A.J.M.W. were supported by NIH grants GM122502 and DK068429. C.J. was funded by an Edward Mallinckrodt, Jr. Foundation Award and NIH grant R01AA029124. H.L.G. was supported by NIH grant R50CA221780 and A.M.M. by NIH grant R01CA218805. J.B.S. was funded by a Worcester Foundation grant and a Smith Family Foundation grant.

Author information

Authors and Affiliations

Authors

Contributions

M.B.D. and D.K. conceived the project and designed the research. M.B.D. (with assistance described below) performed experiments involving metabolic gene knockouts, knockouts with concomitant overexpression and combination knockouts, examining as readouts of cell viability, cell cycle analyses, organelle morphology, protein expression and size shift, protein localization, xenograft growth and mouse survival, glycosaminoglycan levels and hyaluronic acid levels. N.L. assisted in carrying out and analysing cell viability and xenograft growth and mouse survival-based experiments. M.S. assisted with the cell-viability-based experiments. K.S. provided tumour samples and assisted in experimental design and in analyses of experiments determining protein levels in patient-derived tumour tissues and normal tissues. R.L. and L.J.Z. analysed the RNA-seq data. O.P. performed GC–MS analyses and A.J.M.W. assisted in experimental design and in analyses. H.L.G. produced chemoresistant organoids and, with A.M.M., assisted with experimental design using them. C.A. carried out N-glycan profiling and quantitative analysis. T.T. and J.B.S. advised with the design and carried out analyses of the 13C labelling experiments with assistance from C.J. M.B.D. and D.K. wrote the manuscript with consultation from all authors.

Corresponding author

Correspondence to Dohoon Kim.

Ethics declarations

Competing interests

M.B.D. and D.K. are listed as authors on a patent application filed by University of Massachusetts Chan Medical School on targeting UXS1 in cancer therapy. All other authors declare no competing interests.

Peer review

Peer review information

Nature thanks Frederic Bard, Joshua Rabinowitz and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

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 UXS1 is required in a subset of cancer cells, and validation of CRISPR/cas9 reagents for UXS1.

(a) Ranked list of the standard deviation of CERES gene-scores of all metabolic genes across 572 solid cancer cell lines, identifying UXS1 as one of the top enzymes having variable dependency. (b) UXS1 gene-score across cancer cell lines using pan-cancer dependency dataset6 showing that only a subset of cancer cell lines (gene-scores below −0.5) is dependent on UXS1 for survival. A lower gene score means that a gene is more likely to be essential in a given cell line. A score of 0 is equivalent to a non-essential gene, whereas a score of −1 corresponds to the median of all common essential genes. (c) UXS1 immunoblots showing KO of UXS1 in HT1080 and H838 cell lines 9 days post transduction. (d) Schematic of UXS1 g2 and UXS1 CRISPR resistant g2 sequence, demonstrating the strategy of using 1 silent mutation in the PAM sequence (yellow highlight) and 2 silent mutations in the seed sequence (green highlight) to prevent CRISPR/cas9 targeting. (e) Relative viability of DLD1 cells either overexpressing blank vector or g2-resistant UXS1 then subjected to UXS1 KO with g1 or g2, demonstrating that UXS1 KO toxicity is on target (CTRL = 1.0) (n = 3 per group). (f) Immunoblot of UXS1 levels in A549 and (g) DLD1 cells overexpressing blank or UXS1-g2 resistant vectors when subjected to CRISPR/Cas9 mediated KO of UXS1 (h) TLCv2 system51; LentiCRISPR v2 was modified into an all-in-one dox inducible system. The addition of doxycycline induces Cas9-2A-eGFP. The U6 promoter drives constitutive sgRNA expression. UXS1 iKO and control iKO was prepared by cloning UXS1-g2 and CTRL (non-targeting guide) into the TLCv2 system, respectively. (i) Relative viability of UXS1 iKO and control iKO cells with and without doxycycline induction (control iKO without doxycycline = 1.0) (n = 3 per group). Induction consisted of 100 ng/ml doxycycline treatment for 48 h; cells were cultured for 4 more days before measuring viability. (j) UXS1 immunoblots showing KO of UXS1 in CCD18Lu and CCD8Lu cells, 9 days post-transduction. Data are shown as mean ± s.d. except in b where box plot shows median (centre) with interquartile range of 25% to 75%, minima and maxima. e, i, Two-tailed unpaired Student’s t-test. P values from left to right in e: p = 0.12, p = 0.002, i: p = 0.663, p = 5.4E-06 (n.s.: not significant, **p < 0.01, ****p < 0.0001). d and h was created using Biorender (https://biorender.com/).

Source Data

Extended Data Fig. 2 Additional data for UXS1 loss led cell cycle defects and cell death rescue.

(a) Gating strategy for cell cycle analysis. Cells were first gated by forward (FSC-A) and side scatters (SSC-A) to remove cell debris, then by forward scatter area (FSC-A) and height (FSC-H) to select for single cells before plotting histograms for cell cycle profiles. This example gating strategy is shown in control iKO cells. (b) Representative cell cycle profiles for control iKO cells and (c) UXS1 iKO cells 5 days after doxycycline induction. The cell cycle phases were determined by fitting a univariate cell cycle model using the Watson pragmatic algorithm (d) Relative viability of DLD1 cells subjected to UXS1 KO then treated with z-vad, Necrostain, or Ferrostatin at the indicated concentrations (Untreated = 1.0) (n = 3 per group). Data are mean shown as ± s.d. d, Two-tailed unpaired Student’s t-test. P values from left to right in d: p = 0.03, p = 0.002, p = 0.032 (*p < 0.05, **p < 0.01, ***p < 0.001).

Source Data

Extended Data Fig. 3 Supporting data that UDPGA is toxic.

(a) Relative viability of DLD1 cells subjected to preemptive CTRL (gray) or UGDH (faucet; orange) KO, followed by CTRL or UXS1 (drain) KO. Values are relative to cells subjected to CTRL KO then CTRL KO ( = 1.0) (n = 3 per group). (b) Immunoblot of UXS1 and UGDH protein levels under the KO combinations shown in a. (c) GC-MS quantification of UDPGA in cell-lines subjected to CTRL (gray) or UXS1 KO (pink) at 8 days of transduction (n = 3 per group). (d) Relative viability of A549 UXS1 iKO cells grown in media containing increasing glucose concentrations. Values are relative to uninduced cells grown in media containing 25 mM glucose (=1.0) (n = 3 per condition). (e) GC-MS quantification of UDPGA (blue; left Y axis) and relative viability (red; right Y axis) of A549 cells treated with increasing concentrations of UDPGA, showing dose-dependent increase in UDPGA and accompanying decrease in viability (n = 3 per group). (f) Viability of H838 cells overexpressing either blank vector (gray) or UGDH (red) then subjected to CTRL or UXS1 KO. Values are relative to blank vector overexpressing, CTRL KO cells (=1.0) (n = 3 per group). (g) Immunoblot of UXS1 and UGDH protein levels under the OE/KO combinations shown in c; LE indicates low exposure, HE indicates high exposure. (h) Relative UDPGA produced [M + 6 UDPGA / (M + 6 UDP-glucose/ total UDP-glucose)] in control, UGDH KO, and UXS1 KO A549 cells (n = 3 per group). (i) Labelled UDP-glucose (M + 6) in control and UXS1 KO A549 cells (<1 h) (n = 3 per group). (j) Unlabelled UDP-glucose (M + 0) in control and UXS1 KO A549 cells (<1 h) (n = 3 per group). (k) Total UDP-glucose (M + 6 + M + 0) in control and UXS1 KO A549 cells (<1 h) (n = 3 per group). (l) Fractional labeling of UDP-glucose [M + 6 UDP-glucose/ total UDP-glucose] of control and UXS1 KO cells (<1 h) (n = 3 per group). Data are shown as mean ± s.d. a, c-f, i, Two-tailed unpaired Student’s t-test. P values from left to right in a: p = 0.002, p = 0.0001. p = 7.9E-06, c: p = 0.52, p = 0.006, p = 0.013, d: p = 0.0006, p = 0.015, e: (UDPGA) p = 0.016, p = 0.035, p = 0.0001, (viability) p = 0.0007, p = 0.0008, p = 3.14E-05 f: p = 0.001, p = 0.011, p = 0.001 (n.s.: not significant, *p < 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001).

Source Data

Extended Data Fig. 4 Downstream proteoglycan impact of UGDH and UXS1 KO.

(a(Methods) from UGDH KO, UXS1 KO, and UGDH-UXS1 DKO A549 cells and in (b) DLD1 cells (n = 3 per group). (c) Hyaluronic acid levels quantified via competitive ELISA (Methods) from the conditioned media obtained from UGDH KO, UXS1 KO, and UGDH-UXS1 DKO A549 cells and in (d) DLD1 cells (n = 3 per group). Data are shown as mean ± s.d. a-d, Two-tailed unpaired Student’s t-test. P values from left to right in a: p = 0.024, p = 0.019, p = 0.028, p = 0.012, p = 0.031, b: p = 0.337, p = 0.145, p = 0.117, p = 0.098, p = 0.137, c: p = 0.479, p = 0.19, p = 0.015, p = 0.018, p = 0.027, d: p = 0.492, p = 0.71, p = 0.004, p = 0.002, p = 0.004 (n.s.: not significant,*p < 0.05, **p < 0.01).

Source Data

Extended Data Fig. 5 Additional details of Golgi stress, altered glycosylation, and signaling/cell cycle changes following UXS1 ablation.

(a) ARF4 immunoblots in UGDH high (UXS1 (A549) iKO and H1944) and low (HT1080) cells, 7 days post UXS1 KO, showing that ARF4 is only induced only in UGDH high cells following UXS1 ablation. (b) Changes in Oligomannose glycans (c) Paucimannose glycans (d) Sialylated glycans and (e) glycans with poly-LacNac extensions in A549 cells subjected to CTRL or UXS1 KO (n = 2 per group). (f) O-glycan profiling analysis performed on A549 cells subjected to CTRL or UXS1 KO at 7d post-transduction (n = 2 per group). (g) Immunoblots of phospho-histone H3 and actin from A549 UXS1 iKO and control iKO cells at indicated timepoints post doxycycline induction. Nocodazole-treated cells are used as a positive control for enrichment for cells in M phase. (h) Dot plot of total Golgi area per cell, from A549 UXS1 iKO cells induced via doxycycline, then labeled using GM130 at indicated days post-induction. No serum (orange) cells were serum starved for 12 h before labeling, which results in G0 cell cycle arrest and synchronization. p values from left to right are p = 0.48, p = 0.21, p = 0.15, p = 0.044. (i) GSEA enrichment plot showing that EGFR signaling and Cell Cycle genesets are highly depleted in UXS1 KO A549 cells compared to CTRL KO (n = 50 cells per group). (j) Same findings in another UXS1 sensitive cell line (H460). Cell Cycle genes are the most depleted geneset for ‘C2; canonical pathways’ (as shown in Fig. 3a), while EGFR Signaling genes are the most depleted geneset for the ‘C2; chemical and genetic perturbations’ ontology scheme. Data are shown as mean ± s.d. except in h where box plot shows median (centre) with interquartile range of 25% to 75%, minima and maxima. h, Two-tailed unpaired Student’s t-test. (n.s.: not significant, *p < 0.05).

Source Data

Extended Data Fig. 6 Additional details of golgi and ER morphology, and EGFR localization after UXS1 ablation.

(a) Immunofluorescent images of UXS1 iKO or control iKO cells, 7 days post doxycycline induction, immunostained with Giantin (medial-golgi, green), (c) TGN46 (trans-golgi, green), or (e) Calnexin (ER, red); scale bar=20 µm. (b) Dot plot of total golgi area per cell from uninduced and induced UXS1 iKO cells at 7 days postinduction, as labeled with Giantin, (d) TGN-46, or (f) Calnexin (n = 50 cells for each group). (g) Immunoblot for ER stress markers IRE1α, CHOP, and BiP in A549 and DLD1 UXS1 KO cells 7 days post transduction. Tunicamycin treated cells are positive control for ER stress. (h) Dot plot of total golgi area per cell, labeled using GM130, from control, UXS1 KO, and UGDH-UXS1 DKO (A549) cells, 7 days post UXS1 transduction; n = 50 cells per group. (i) Quantification of relative actin-normalized EGFR expression (immunoblots from 3 independent experiments) in induced compared to uninduced UXS1 iKO A549 cells, 5d postinduction. In each experiment, the uninduced EGFR = 1.0. (j) Gating strategy for EGFR cell surface quantitation analysis. Cells were first gated by forward (FSC-A) and side scatters (SSC-A) to remove cell debris, then by forward scatter area (FSC-A) and height (FSC-H) to select for single cells before plotting histograms for FITC channel. Example gating strategy is shown for UXS1 iKO cells. (k) Quantitation of cell surface EGFR labeling via 3 independent flow cytometry experiments. Data are shown as mean ± s.d. except in b, d, f, h where box plot shows median (centre) with interquartile range of 25% to 75%, minima and maxima. b, d, f, h, i, k, Two-tailed unpaired Student’s t-test. P values from left to right in b: p = 3.59E-10 in d: p = 2.94E-10 in f: p = 0.95 in h: p = 1.96E-09, p = 0.285 in i: p = 0.011 in k: p = 0.001 (n.s.: not significant, *p < 0.05, ***p < 0.001, ****p < 0.0001).

Source Data

Extended Data Fig. 7 UDPGA accumulation occurs in cytosol and overexpression of SLC35D1 does not impact toxicity of UXS1 KO.

(a) Viability of A549 and DLD1 cells overexpressing either blank vector (gray) or SLC35D1 (dark green) then subjected to CTRL or UXS1 KO. Values are relative to blank vector overexpressing, CTRL KO cells (=1.0) (n = 3 per group). (b) Immunoblot of UXS1 and SLC35D1 protein levels under the OE/KO combinations shown in a. (c) GC-MS UDPGA quantitation of cytoplasmic and organellar fractions of either induced or uninduced A549 UXS1 iKO cells (n = 3 per group). (d) Immunoblots of proteins located in nucleus (PARP1), ER (Calnexin), Golgi (GM130), mitochondria (NDUFS3), and cytoplasm (PSPH and tubulin) in cytosolic and organellar fractions. Cytochrome c is used as a control for intact organelles (mitochondria). Data are shown as mean ± s.d. a, c, Two-tailed unpaired Student’s t-test. P values: from left to right in a: p = 0.391, p = 0.379, p = 0.792, p = 0.106 in c: p = 0.018, p = 0.002 (n.s.: not significant, *p < 0.05, **p < 0.01).

Source Data

Extended Data Fig. 8 Additional in-vivo data.

(a) Schematic of in-vivo xenograft experiments shown in Fig. 4a–f, p. A549/H460/HT1080 UXS1 iKO cells were injected subcutaneously into 6 weeks old nude female mice and allowed for tumors to establish. Once the mean tumor was 200–250mm3, the mice were divided into groups for control (uninduced) and dox chow induction of UXS1 KO. A549 xenograft experiment had additional cisplatin arms. The animals were euthanized when tumor volume reached 2000mm3. (b) Immunoblots for UXS1 and actin of H460 and HT1080 UXS1 iKO cells in culture with and without doxycycline induction. (c) Relative viability of H460 UXS1 iKO and HT1080 UXS1 iKO cells with and without doxycycline induction (without doxycycline = 1.0). Induction consisted of 100 ng/ml doxycycline treatment for 48 h; cells were cultured for 4 more days before measuring viability. (d) Immunoblot for UXS1 and actin in six individual tumors each from the control and UXS1 iKO group of HT1080 UXS1 iKO xenograft experiment. (e) Weight of mice in grams from xenograft experiments of A549, (f) H460, and (g) HT1080 from control and UXS1 iKO arms of the experiment. (h) LC-MS Quantification of UDPGA from endpoint tumors from another A549 iKO cohort where tumors were allowed to form for 17 days, induced for KO, then euthanized at 19 days postinduction (p = 0.003). (i) Immunoblot for ARF4 and actin from four individual tumors in panel h. (j) Immunoblot for UGDH and actin from organs harvested from control and cisplatin-treated mice (representative pair shown) treated with the same regimen of cisplatin as in Fig. 4n (4 mg/kg iv 1x weekly, 3 weeks). (k) Quantitation of relative, actin normalized UGDH band intensities from control and cisplatin-treated mice (n = 4 each) with the same regimen. Values are relative to cisplatin-treated kidney in each mouse set at 1.0. Data are shown as mean ± s.d. c, h, k, Two-tailed unpaired Student’s t-test. P values from left to right in c: p = 1.34E-08, p = 0.44 h: p = 0.003 in k: p = 0.966, p = 0.978, p = 0.478, p = 0.471. (n.s.: not significant, **p < 0.01, ****p < 0.0001). a was created using Biorender (https://biorender.com/).

Source Data

Extended Data Fig. 9 UGDH is elevated in cancer subsets.

(a) Immunoblots for UGDH and actin from 11 breast cancer patient tissue samples and 8 normal breast tissues; cancer tissues are labeled in red and normal tissues are labeled in blue. Numbers indicate deidentified patient code, and tumors and normals sharing the same number are adjacent samples from the same patient. (b) Quantification of relative, actin normalized UGDH band intensities from panel a (p = 0.023). (c) Expression profiles of UGDH in 10 types of normal and tumor tissues. The plot is obtained from GEPIA255, a web gene expression profiling tool that plots normalized mRNA-seq data from patient tumor tissues and normal tissues obtained from TCGA and GTEx. All tumor types shown have over 2-fold UGDH expression in tumors than normal tissues (q value < 0.01). q values have been determined by two-way ANOVA and adjusted for false discovery rate. The dotted line indicates mean UGDH TPM of liver normal samples, highest among other normal organs. (d) Top dot plot showing viability following UXS1 KO (y axis) and bottom dot plot showing corresponding UGDH protein expression level (y axis) for 19 cancer and 9 normal cell lines of varying lineage (x axis). Cancer lines are divided between High (red) and Low(gray) UGDH expressing lines based on arbitrary cut-off 1.25 units; normal lines are labeled as blue dots. (e) Immunoblots of UGDH and actin for cancer cell lines and normal cells (non-transformed primary or immortalized cells) shown in Fig. 4h and panel d; UXS1 KO sensitive cell lines are labeled in red and insensitive lines are labeled in gray and normal lines are labeled in blue. (f) Immunoblots for UGDH and actin from 6 normal liver, 6 normal kidney and, 5 normal colon patient tissue samples compared with 2 representative (UGDH high) lung cancer patient samples; cancer tissues are labeled in red and normal tissues are labeled in blue. Data shown in b is shown as mean ± s.d. b, Two-tailed unpaired Student’s t-test.

Source Data

Extended Data Fig. 10 Additional chemo-resistance data and RNA interference data for UXS1.

(a) Viability of parental (chemo-sensitive) and chemo-resistant Patient-derived organoid (PDO) TPN-02 subjected to varying doses of cisplatin. (n = 3 per group). (b) Immunoblot for UGDH and actin from parental and chemo-resistant PDO TPN-02. (c) Relative viabilities of parental and cisplatin-resistant TPN-02 subjected to UXS1 KO, indicating that cisplatin-resistant organoids are more sensitive to UXS1 loss. Values are relative to parental cells CTRL-KO ( = 1.0) (n = 3 per group). (d) Sigmoidal drug response curves for MDAMB231 parental (gray) and cisplatin-resistant (violet) cells in response to cisplatin treatment and MDAMB231 parental (gray) and paclitaxel-resistant (orange) counterparts in response to paclitaxel (n = 3 per group). (e) Immunoblot for UGDH and actin from MDAMB231 parental and drug-resistant counterparts. (f) Relative viabilities of MDAMB231 cisplatin (violet) and paclitaxel (orange) resistant clones subjected to UXS1 KO, indicating that MDAMB231 resistant clones are more sensitive to UXS1 loss. Values are relative to CTRL-KO for each clone (=1.0) (n = 3 per group). (g) Sigmoidal drug response curves for H2170 parental (gray) and cisplatin-resistant (violet) cells in response to Cisplatin treatment and H2170 parental (gray) and paclitaxel-resistant (orange) cells in response to paclitaxel (n = 3 per group). (h) Immunoblot for UGDH and actin from H2170 parental and drug-resistant clones. (i) Relative viabilities of H2170 cisplatin (violet) and paclitaxel (orange) resistant clones subjected to UXS1 KO, indicating that H2170 resistant clones are more sensitive to UXS1 loss. Values are relative to CTRL-KO for each clone (=1.0) (n = 3 per group). (j) Immunoblots for glycoproteins FGFR1, FGFR4, CD44 and UXS1 in A549 cells subjected to RNAi. UXS1 blots indicate that shUXS1-2 and shUXS1-3, but not shUXS1-1, result in effective knockdown of UXS1, # denotes that shUXS1-1 is a poor performing shRNA. (k) Relative viabilities of A549 cells subjected to RNA interference (RNAi) mediated knock-down of UXS1, showing that the two guides which effectively knock down UXS1 result in toxicity (n = 2 per group). (l) Dot plot of total golgi area per cell, labeled using GM130, from A549 cells subjected to shGFP or shUXS1-2, 7 days post-transduction; n = 50 cells per group. Data are shown as mean ± s.d. except in l where box plot shows median (centre) with interquartile range of 25% to 75%, minima and maxima. a, c, f, i, l, Two-tailed unpaired Student’s t-test. P values from left to right in a p = 0.005, p = 0.007, p = 4.12E-05 in c p = 0.006, p = 1.13E-05 f p = 0.002, p = 6.07E-05, p = 2.47E-05, p = 0.021 in i p = 0.006, p = 0.019, p = 2.4E-5, p = 0.021 in l p = 4.8E-20. (*p < 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001).

Source Data

Supplementary information

Supplementary Information

This file contains Supplementary Discussion, Supplementary Tables 1 and 2 and Supplementary Figs. 1–12.

Reporting Summary

Peer Review File

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

Doshi, M.B., Lee, N., Tseyang, T. et al. Disruption of sugar nucleotide clearance is a therapeutic vulnerability of cancer cells. Nature 623, 625–632 (2023). https://doi.org/10.1038/s41586-023-06676-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-023-06676-3

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