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:

Inflammatory macrophage dependence on NAD+ salvage is a consequence of reactive oxygen species–mediated DNA damage

Abstract

The adoption of Warburg metabolism is critical for the activation of macrophages in response to lipopolysaccharide. Macrophages stimulated with lipopolysaccharide increase their expression of nicotinamide phosphoribosyltransferase (NAMPT), a key enzyme in NAD+ salvage, and loss of NAMPT activity alters their inflammatory potential. However, the events that lead to the cells' becoming dependent on NAD+ salvage remain poorly defined. We found that depletion of NAD+ and increased expression of NAMPT occurred rapidly after inflammatory activation and coincided with DNA damage caused by reactive oxygen species (ROS). ROS produced by complex III of the mitochondrial electron-transport chain were required for macrophage activation. DNA damage was associated with activation of poly(ADP-ribose) polymerase, which led to consumption of NAD+. In this setting, increased NAMPT expression allowed the maintenance of NAD+ pools sufficient for glyceraldehyde-3-phosphate dehydrogenase activity and Warburg metabolism. Our findings provide an integrated explanation for the dependence of inflammatory macrophages on the NAD+ salvage pathway.

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: NAD+ is depleted in inflammatory macrophages, and NAD+ salvage is required for viability.
Fig. 2: NAD+ salvage regulates the core metabolism of inflammatory macrophages through changes in GAPDH activity.
Fig. 3: NAMPT loss of function modulates inflammatory macrophage activation.
Fig. 4: NAMPT inhibition reduces in vivo inflammation, modulating inflammatory macrophage metabolism.
Fig. 5: Nicotinamide mononucleotide addition rescues the inhibitory effects of NAMPT loss of function on inflammatory macrophage activation.
Fig. 6: γ+LPS exposure induces rapid NAD+ depletion and DNA damage accumulation.
Fig. 7: γ+LPS–induced mitochondrial ROS drives oxidative DNA damage.
Fig. 8: Mitochondrial ROS is produced by complex III and required for inflammatory macrophage polarization.

Similar content being viewed by others

Data availability

All data generated or analyzed during this study are included in this published article (and its Supplementary information) and the data that support the findings of this study are available from the corresponding author upon reasonable request. Next-generation sequencing data can be accessed at Gene Expression Omnibus under accession code GSE123596.

References

  1. Wynn, T. A., Chawla, A. & Pollard, J. W. Macrophage biology in development, homeostasis and disease. Nature 496, 445–455 (2013).

    Article  CAS  Google Scholar 

  2. Huang, S. C. et al. Cell-intrinsic lysosomal lipolysis is essential for alternative activation of macrophages. Nat. Immunol. 15, 846–855 (2014).

    Article  CAS  Google Scholar 

  3. Jha, A. K. et al. Network integration of parallel metabolic and transcriptional data reveals metabolic modules that regulate macrophage polarization. Immunity 42, 419–430 (2015).

    Article  CAS  Google Scholar 

  4. Mills, E. L. et al. Succinate dehydrogenase supports metabolic repurposing of mitochondria to drive inflammatory macrophages. Cell 167, 457–470.e413 (2016).

    Article  CAS  Google Scholar 

  5. Tannahill, G. M. et al. Succinate is an inflammatory signal that induces IL-1β through HIF-1α. Nature 496, 238–242 (2013).

    Article  CAS  Google Scholar 

  6. Krawczyk, C. M. et al. Toll-like receptor-induced changes in glycolytic metabolism regulate dendritic cell activation. Blood 115, 4742–4749 (2010).

    Article  CAS  Google Scholar 

  7. O’Neill, L. A. & Pearce, E. J. Immunometabolism governs dendritic cell and macrophage function. J. Exp. Med. 213, 15–23 (2016).

    Article  Google Scholar 

  8. Locasale, J. W. & Cantley, L. C. Metabolic flux and the regulation of mammalian cell growth. Cell Metab. 14, 443–451 (2011).

    Article  CAS  Google Scholar 

  9. Verdin, E. NAD+ in aging, metabolism, and neurodegeneration. Science 350, 1208–1213 (2015).

    Article  CAS  Google Scholar 

  10. Busso, N. et al. Pharmacological inhibition of nicotinamide phosphoribosyltransferase/visfatin enzymatic activity identifies a new inflammatory pathway linked to NAD. PLoS One 3, e2267 (2008).

    Article  Google Scholar 

  11. Halvorsen, B. et al. Increased expression of NAMPT in PBMC from patients with acute coronary syndrome and in inflammatory M1 macrophages. Atherosclerosis 243, 204–210 (2015).

    Article  CAS  Google Scholar 

  12. Liu, T. F., Yoza, B. K., El Gazzar, M., Vachharajani, V. T. & McCall, C. E. NAD+-dependent SIRT1 deacetylase participates in epigenetic reprogramming during endotoxin tolerance. J. Biol. Chem. 286, 9856–9864 (2011).

    Article  CAS  Google Scholar 

  13. Schilling, E. et al. Inhibition of nicotinamide phosphoribosyltransferase modifies LPS-induced inflammatory responses of human monocytes. Innate Immun. 18, 518–530 (2012).

    Article  CAS  Google Scholar 

  14. Galli, M., Van Gool, F., Rongvaux, A., Andris, F. & Leo, O. The nicotinamide phosphoribosyltransferase: a molecular link between metabolism, inflammation, and cancer. Cancer Res. 70, 8–11 (2010).

    Article  CAS  Google Scholar 

  15. Gerner, R. R. et al. NAD metabolism fuels human and mouse intestinal inflammation. Gut 67, 1813–1823 (2018).

    Article  CAS  Google Scholar 

  16. Hasmann, M. & Schemainda, I. FK866, a highly specific noncompetitive inhibitor of nicotinamide phosphoribosyltransferase, represents a novel mechanism for induction of tumor cell apoptosis. Cancer Res. 63, 7436–7442 (2003).

    CAS  PubMed  Google Scholar 

  17. Ronchi, J. A. et al. A spontaneous mutation in the nicotinamide nucleotide transhydrogenase gene of C57BL/6J mice results in mitochondrial redox abnormalities. Free. Radic. Biol. Med. 63, 446–456 (2013).

    Article  CAS  Google Scholar 

  18. Everts, B. et al. Commitment to glycolysis sustains survival of NO-producing inflammatory dendritic cells. Blood 120, 1422–1431 (2012).

    Article  CAS  Google Scholar 

  19. Tan, B. et al. Inhibition of nicotinamide phosphoribosyltransferase (NAMPT), an enzyme essential for NAD+ biosynthesis, leads to altered carbohydrate metabolism in cancer cells. J. Biol. Chem. 290, 15812–15824 (2015).

    Article  CAS  Google Scholar 

  20. Liberti, M. V. et al. A Predictive Model for Selective Targeting of the Warburg Effect through GAPDH Inhibition with a Natural Product. Cell Metab. 26, 648–659.e648 (2017).

    Article  CAS  Google Scholar 

  21. Barth, M. W., Hendrzak, J. A., Melnicoff, M. J. & Morahan, P. S. Review of the macrophage disappearance reaction. J. Leukoc. Biol. 57, 361–367 (1995).

    Article  CAS  Google Scholar 

  22. Dockery, L. E., Gunderson, C. C. & Moore, K. N. Rucaparib: the past, present, and future of a newly approved PARP inhibitor for ovarian cancer. Onco. Target Ther. 10, 3029–3037 (2017).

    Article  CAS  Google Scholar 

  23. Morales, A. J. et al. A type I IFN-dependent DNA damage response regulates the genetic program and inflammasome activation in macrophages. Elife 6, e24655 (2017).

    Article  Google Scholar 

  24. Zingarelli, B., O’Connor, M., Wong, H., Salzman, A. L. & Szabo, C. Peroxynitrite-mediated DNA strand breakage activates poly-adenosine diphosphate ribosyl synthetase and causes cellular energy depletion in macrophages stimulated with bacterial lipopolysaccharide. J. Immunol. 156, 350–358 (1996).

    CAS  PubMed  Google Scholar 

  25. Hayes, J. D. & Dinkova-Kostova, A. T. The Nrf2 regulatory network provides an interface between redox and intermediary metabolism. Trends Biochem. Sci. 39, 199–218 (2014).

    Article  CAS  Google Scholar 

  26. Nunes, P., Demaurex, N. & Dinauer, M. C. Regulation of the NADPH oxidase and associated ion fluxes during phagocytosis. Traffic 14, 1118–1131 (2013).

    CAS  PubMed  Google Scholar 

  27. Samuni, Y., Goldstein, S., Dean, O. M. & Berk, M. The chemistry and biological activities of N-acetylcysteine. Biochim. Biophys. Acta 1830, 4117–4129 (2013).

    Article  CAS  Google Scholar 

  28. Ezeriņa, D., Takano, Y., Hanaoka, K., Urano, Y. & Dick, T. P. N-Acetyl cysteine functions as a fast-acting antioxidant by triggering intracellular H2S and sulfane sulfur production. Cell Chem. Biol. 25, 447–459.e444 (2018).

    Article  Google Scholar 

  29. Murphy, M. P. How mitochondria produce reactive oxygen species. Biochem. J. 417, 1–13 (2009).

    Article  CAS  Google Scholar 

  30. Bleier, L. & Drose, S. Superoxide generation by complex III: from mechanistic rationales to functional consequences. Biochim. Biophys. Acta 1827, 1320–1331 (2013).

    Article  CAS  Google Scholar 

  31. Conradt, P., Dittmar, K. E., Schliephacke, H. & Trowitzsch-Kienast, W. Myxothiazol: a reversible blocker of the cell cycle. J. Antibiot. (Tokyo) 42, 1158–1162 (1989).

    Article  CAS  Google Scholar 

  32. Matalonga, J. et al. The nuclear receptor LXR limits bacterial infection of host macrophages through a mechanism that impacts cellular NAD metabolism. Cell Rep. 18, 1241–1255 (2017).

    Article  CAS  Google Scholar 

  33. Van Gool, F. et al. Intracellular NAD levels regulate tumor necrosis factor protein synthesis in a sirtuin-dependent manner. Nat. Med. 15, 206–210 (2009).

    Article  Google Scholar 

  34. Sampath, D., Zabka, T. S., Misner, D. L., O’Brien, T. & Dragovich, P. S. Inhibition of nicotinamide phosphoribosyltransferase (NAMPT) as a therapeutic strategy in cancer. Pharmacol. Ther. 151, 16–31 (2015).

    Article  CAS  Google Scholar 

  35. Tan, B. et al. Pharmacological inhibition of nicotinamide phosphoribosyltransferase (NAMPT), an enzyme essential for NAD + biosynthesis, in human cancer cells: metabolic basis and potential clinical implications. J. Biol. Chem. 288, 3500–3511 (2013).

    Article  CAS  Google Scholar 

  36. Jackson, S. P. & Bartek, J. The DNA-damage response in human biology and disease. Nature 461, 1071–1078 (2009).

    Article  CAS  Google Scholar 

  37. Pereira-Lopes, S. et al. NBS1 is required for macrophage homeostasis and functional activity in mice. Blood 126, 2502–2510 (2015).

    Article  CAS  Google Scholar 

  38. Herrtwich, L. et al. DNA damage signaling instructs polyploid macrophage fate in granulomas. Cell 167, 1264–1280 e1218 (2016).

    Article  CAS  Google Scholar 

  39. Singh, I. et al. High mobility group protein-mediated transcription requires DNA damage marker γ-H2AX. Cell Res. 25, 837–850 (2015).

    Article  CAS  Google Scholar 

  40. Weintz, G. et al. The phosphoproteome of toll-like receptor-activated macrophages. Mol. Syst. Biol. 6, 371 (2010).

    Article  Google Scholar 

  41. Leung, A., Todorova, T., Ando, Y. & Chang, P. Poly(ADP-ribose) regulates post-transcriptional gene regulation in the cytoplasm. RNA Biol. 9, 542–548 (2012).

    Article  CAS  Google Scholar 

  42. Mills, E. L. et al. Itaconate is an anti-inflammatory metabolite that activates Nrf2 via alkylation of KEAP1. Nature 556, 113–117 (2018).

    Article  CAS  Google Scholar 

  43. Palsson-McDermott, E. M. & O’Neill, L. A. The Warburg effect then and now: from cancer to inflammatory diseases. Bioessays 35, 965–973 (2013).

    Article  CAS  Google Scholar 

  44. Garaude, J. et al. Mitochondrial respiratory-chain adaptations in macrophages contribute to antibacterial host defense. Nat. Immunol. 17, 1037–1045 (2016).

    Article  CAS  Google Scholar 

  45. Nakahira, K. et al. Autophagy proteins regulate innate immune responses by inhibiting the release of mitochondrial DNA mediated by the NALP3 inflammasome. Nat. Immunol. 12, 222–230 (2011).

    Article  CAS  Google Scholar 

  46. West, A. P. et al. TLR signalling augments macrophage bactericidal activity through mitochondrial ROS. Nature 472, 476–480 (2011).

    Article  CAS  Google Scholar 

  47. Park, J. et al. Mitochondrial ROS govern the LPS-induced pro-inflammatory response in microglia cells by regulating MAPK and NF-κB pathways. Neurosci. Lett. 584, 191–196 (2015).

    Article  CAS  Google Scholar 

  48. Bulua, A. C. et al. Mitochondrial reactive oxygen species promote production of proinflammatory cytokines and are elevated in TNFR1-associated periodic syndrome (TRAPS). J. Exp. Med. 208, 519–533 (2011).

    Article  CAS  Google Scholar 

  49. Afgan, E. et al. The Galaxy platform for accessible, reproducible and collaborative biomedical analyses: 2016 update. Nucleic Acids Res. 44, W3–W10 (2016).

    Article  CAS  Google Scholar 

  50. Ramirez, F. et al. deepTools2: a next generation web server for deep-sequencing data analysis. Nucleic Acids Res. 44, W160–W165 (2016).

    Article  CAS  Google Scholar 

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

    Google Scholar 

  52. Liao, Y., Smyth, G. K. & Shi, W. featureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics 30, 923–930 (2014).

    Article  CAS  Google Scholar 

  53. 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  Google Scholar 

  54. Lu, W., Wang, L., Chen, L., Hui, S. & Rabinowitz, J. D. Extraction and quantitation of nicotinamide adenine dinucleotide redox cofactors. Antioxid. Redox Signal 28, 167–179 (2018).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank members of the Pearce laboratories for support and discussions, and B. Sleckman and A. Morales at Cornell Weill and E. Latz and M. Lauterbach at University of Bonn for their generous advice during these studies. We also thank D. Braas and the UCLA Metabolomics Centre and the DeepSequencing facility at the Max Planck Institute of Immunobiology and Epigenetics for their technical support. This work was funded by National Institutes of Health grants AI 110481 (E.J.P.) and CA18125 (E.L.P.), a Capes Humboldt Research Fellowship (A.C.), a Sir Henry Wellcome Fellowship awarded by The Wellcome Trust (D.J.P.), and the Max Planck Society (E.J.P and E.L.P).

Author information

Authors and Affiliations

Authors

Contributions

A.M.C., A.C., B.K., E.L.P. and E.J.P. designed experiments and provided conceptual input. A.M.C., A.C., L.J.F., C.S.F., D.J.P., R.L.K., A.E.P, F.H. and J.M.B., performed experiments and developed methodologies. A.M.C., A.C., L.J.F., D.E.S., E.L.P. and E.J.P. analyzed data. A.M.C. and E.J.P. wrote the manuscript.

Corresponding author

Correspondence to Edward J. Pearce.

Ethics declarations

Competing interests

E.J.P. and E.L.P. are founders of Rheos Medicines, and E.L.P. is a member of the Scientific Advisory Board for ImmunoMet Therapeutics.

Additional information

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

Integrated supplementary information

Supplementary Figure 1 Inflammatory macrophages are depleted of NAD+, and NAD+ salvage is required for viability.

(a) NAD+ in M0, M(γ+LPS), M(LPS) and M(IL-4) polarized for 18 h, quantified by NAD cycling assay (n = 3 biologically independent samples, representative of three independent experiments). (b) NADH, NADP+, NADPH in M0, M(γ+LPS), M(LPS) and M(IL-4) polarized for 18 h, as quantified by LC-MS (n = 3 biologically independent samples, representative of three independent experiments). (c) Expression of NAMPT mRNA in macrophages polarized as indicated for 18h, normalized to mRNA encoding HPRT and presented relative to M0 control cells, set as 1 (n = 3 biologically independent samples, representative of three independent experiments). (d) RNA-seq analysis of expressed enzymes in the Preiss Handler Pathway or de novo NAD+ synthesis pathway in M0, M(γ+LPS), M(LPS) and M(IL-4) polarized for 18 h. Statistically significant (adjusted p value < 0.1) upregulated (> 2 fold) genes denoted by * in M(γ+LPS) polarized cells compared to M0 (n = 3 biologically independent samples). (e) NADH levels of M0, M(γ+LPS), M(LPS) and M(IL-4) cultured for 18 h in the presence or absence of 50 nM FK866, relative to M0 (n = 3 biologically independent samples, representative of two independent experiments). (f) Viability of M(γ+LPS) treated with increasing FK866 concentrations from 20 nM – 1600 nM as determined by flow cytometry, presented as percentage of vehicle treated control (n = 3 biologically independent samples, representative of three independent experiments). (g) NAD+ of bone marrow macrophages derived from C57Bl/6N mice. M0, M(γ+LPS) +/- 50 nM FK866, M(LPS) +/- 50nM FK866 and M(IL-4) cultured for 18 h, and NAD+ quantified by LC-MS (n = 3 biologically independent samples, representative of two independent experiments). Error bars are mean ± SEM. Data were analyzed by one-way ANOVA with Tukey’s multiple comparison test. *p < 0.05, ***p < 0.001 and ****p < 000.1.

Supplementary Figure 2 Inhibition of NAMPT affects glycolytic metabolism of inflammatory macrophages due to diminished limiting of GAPDH activity by NAD+.

(a-f) M(γ+LPS) or M(LPS) polarized for 18 h with or without FK866 and subsequently analyzed. (a,b) ECAR after glucose injection (a: n = 5 technical replicates, b: n = 4 technical replicates, both representative of over ten independent experiments). (c,d) Lactate production (n = 3 biologically independent samples, representative of two independent experiments). (e,f) basal OCR (e: n = 4 technical replicates, f: n = 4 technical replicates, both are representative of over ten independent experiments. (g-j) M(γ+LPS) or M(LPS) polarized for 18 h with vehicle control, FK866, GPP78 or STF118804 and subsequently analyzed. Real time changes in the ECAR (g) and ECAR after glucose injection (h) (n = 5 technical replicates, representative of three independent experiments). NAD+ as quantified by LC-MS (i) and ATP levels (j) (n = 4 biologically independent samples, representative of two independent experiments). (k, l) Volcano plots showing differentially expressed genes in M(γ+LPS) (k) or M(LPS) (l) untreated cells compared to FK866 treated, highlighting key genes in the glycolytic pathway as obtained from the Kyoto Encyclopedia of genes and genomes (pathway identifier: 00010) (n = 3 biologically independent samples). (m) LC-MS quantitation of glycolytic metabolites in control M(LPS) or M(LPS) + FK866 (n = 4 biologically independent samples, representative of three independent experiments). (n-q) Macrophages were polarized for 18h with or without 1 μM heptelidic acid (HA). (n) Basal ECAR, and basal OCR of M(γ+LPS) (n = 4 technical replicates, representative of two independent experiments). (o) ATP quantification of M(γ+LPS) (n = 3 biologically independent samples, representative of two independent experiments). (p) Basal ECAR, and basal OCR of M(LPS) (n = 4 technical replicates, representative of two independent experiments). (q) ATP quantification of M(LPS) (n = 3 biologically independent samples, representative of two independent experiments). (r) Western Blot showing NAMPT and β-actin expression in macrophages transduced with empty vector (EV) shRNA or Nampt shRNA. The blot was cropped to show relevant bands, and is representative of four independent experiments with similar results. (s) Viability of macrophages transduced with EV shRNA or Nampt shRNA, polarized for 18 h with γ+LPS and assessed by flow cytometry (n = 3 biologically independent samples, representative of three independent experiments). Error bars are mean ± SEM. Data were analyzed in c,d,m,o,q,s by unpaired, two-sided Student’s t-test and in h-j by one-way ANOVA with Tukey’s multiple comparison test. *p < 0.05, ***p < 0.001 and ****p < 000.1

Supplementary Figure 3 Modulation of inflammatory macrophage phenotype by an NAMPT inhibitor is linked to changes in glycolytic metabolism.

(a) Example flow cytometric gating strategy used to identify live, CD11b+, F4/80+ macrophages. (b) Intracellular staining for TNF production in M0, M(γ+LPS) and M(γ+LPS) + FK866 as assessed by flow cytometry after 8 h of polarization, representative flow cytomtery plots with indicated percentage of TNF+ cells and SEM (n = 3 biologically independent samples, representative of two independent experiments). (c) Heatmap of antigen presentation and pro-inflammatory gene expression in macrophages polarized for 18 h as indicated. Statistically significant (adjusted p value < 0.1) up- or down-regulated (>2 fold) genes denoted by * in M(γ+LPS) compared to M(γ+LPS) + FK (n = 3 biologically independent samples). (d,e) M(γ+LPS) polarized with vehicle control or 50 nM FK866, GPP78 or STF118804 for 18h and assessed by flow cytometry. (d) Viability relative to control treated M(γ+LPS) (n = 4 biologically independent samples, representative of three independent experiments). (e) CD80, NOS2, and MHC class II expression (n = 3 biologically independent samples, representative of three independent experiments). (f) Expression of NOS2 and production of IL-6 and TNFα of M(γ+LPS) polarized for 18h with or without 1 μM HA and measured by flow cytometry and ELISA, respectively (n = 3 biologically independent samples, representative of two independent experiments). (g) Expression of markers RELMα, CD206 and CD301 in M(IL-4) polarized for 18h with or without FK866 treatment, analyzed by flow cytometry (n = 3 biologically independent samples, representative of three independent experiments). Error bars are mean ± SEM. Data were analyzed by one-way ANOVA with Tukey’s multiple comparison test. *p < 0.05, **p < 0.01, ***p < 0.001 and ****p < 000.1.

Supplementary Figure 4 Inhibition of NAMPT affects macrophage metabolism but does not modulate the frequency of immune cells in peritoneal lavage fluid or adipose tissue.

Frequency of eosinophils (a), neutrophils (b), and dendritic cells (c) in peritoneal lavage 12 hours after i.p. administration of PBS, FK866, LPS or LPS and FK866 (n = 5 mice, representative of two independent experiments). (d) Basal ECAR and basal OCR of peritoneal macrophages isolated from mice given i.p. PBS, FK866, LPS or LPS and FK866 (n = 5 technical replicates derived from 5 pooled biological samples, representative of three independent experiments). (e) Basal ECAR and basal OCR of macrophages isolated from epididymal adipose tissue 12 hours after i.p. administration of PBS, FK866, LPS or LPS and FK866 (n = 5 technical replicates derived from 5 pooled biological samples, representative of two independent experiments). Frequency of macrophages (f) eosinophils (g), neutrophils (h), and dendritic cells (o) in epididymal adipose tissue and basal ECAR and basal OCR of macrophages isolated from epididymal adipose tissue 12 hours after i.p. administration of PBS, FK866, LPS or LPS and FK866 (n = 5 mice, representative of two independent experiments). Error bars are mean ± SEM. Data were analyzed by one-way ANOVA with Tukey’s multiple comparison test. *p < 0.05, **p < 0.01, ***p < 0.001 and ****p < 000.1.

Supplementary Figure 5 The addition of exogenous pyruvate does not ‘rescue’ the metabolic or phenotypic effects of NAMPT inhibition in inflammatory macrophages.

Real-time changes in the ECAR (a) and ECAR after glucose injection (b) measured using Seahorse, of M(γ+LPS) untreated or treated with 50 nM FK866, 5 mM pyruvate or both FK and pryuvate for 18 h (n = 6 technical replicates, data are representative of two independent experiments). NAD+ (c), and F1,6BP and 2PG/3PG (d) levels measured by LC-MS in macrophages polarized to M(γ+LPS) untreated or treated with 50 nM FK866, 5 mM pryuvate or both FK and pyruvate for 18 h (n = 4 biologically independent samples, representative of one experiment). Viability (e) and expression of markers CD80, NOS2 and MHC class II (f) M(γ+LPS) with or without FK866, pyruvate or both FK and pyruvate for 18 h analyzed by flow cytometry (n = four biologically independent samples, representative of two independent experiments). Error bars are mean ± SEM. Data were analyzed by one-way ANOVA with Tukey’s multiple comparison test. *p < 0.05 and ****p < 000.1.

Supplementary Figure 6 Depletion of NAD+ is associated with PARP expression and DNA damage.

(a) NADH, NADP+, NADPH in M0, M(γ+LPS), M(LPS) and M(IL-4) polarized for 18 h, as quantified by LC-MS (n = 3 biologically independent samples, representative of three concordant experiments). (b) Expression of NAMPT mRNA in macrophages polarized as indicated for 1 h, normalized to mRNA encoding HPRT and presented relative to M0, set as 1 (n = 3 biologically independent samples, representative of two concordant experiments). (c) NAMPT and β-actin (loading control) expression as shown by immunoblot in M0, M(γ+LPS), M(LPS) and M(IL-4) polarized for 1 h. The blot was cropped to show relevant bands, and is representative of three independent experiments with similar results. (d) CD38 expression of M0, M(γ+LPS), M(LPS) and M(IL-4) polarized for 1 h or 18 h, analysed by flow cytometry (n = 3 biologically independent samples, representative of three concordant experiments). Heatmaps of PARP expression at 1 h (e) and 18 h (f), and (g) genes annotated with the Gene Ontology term ‘cellular response to DNA damage stimulus’, at 1 h, in M0, M(γ+LPS), M(LPS) and M(IL-4). Statistically significant (adjusted p value < 0.1) up or down-regulated (> 2 fold) genes denoted by * in M(γ+LPS) polarized cells compared to M0 (n = 3 biologically independent samples). (h,i) C57Bl/6N M0, M(γ+LPS), M(LPS) and M(IL-4) polarized for 1 h. (h) NAD+ measured by LCMS (n = 3 biologically independent samples, representative of three concordant experiments). (i) γ-H2AX by flow cytometry (n = 3 biologically independent samples, representative of two concordant experiments). (j) Viability of M(γ+LPS) polarized with increasing concentrations of Rucaparib, assessed by flow cytometry at 18 h (n = 3 biologically independent samples, representative of five concordant experiments). (k,l) M0 and M(γ+LPS) treated with 10 μM Ruc or vehicle control for 18 h and subsequently analyzed. (k) Expression of CD80, NOS2 and MHC class II measured by flow cytometry (n = 3 biologically independent samples, representative of five concordant experiments). (l) IL-6 and TNF production by ELISA (n = 4 biologically independent samples, representative of three concordant experiments). (m-o) M0, M(LPS) or M(IL-4) were treated with or without 10 μM Ruc for 4 h, and γ-H2AX expression was assessed by flow cytometry. (n = 3 biologically independent samples, representative of two independent experiments). (p) ECAR after glucose injection of M(γ+LPS) polarized with or without 50 nM FK866, 10 μM Ruc or both inhibitors in combination (n = 6 technical replicates, representative of five concordant experiments). Error bars are mean ± SEM. Data were analyzed in a,b,d,h-k by one-way ANOVA with Tukey’s multiple comparison test and in l-o by unpaired, two-tailed students t-test. *p < 0.05, **p < 0.01, ***p < 0.001 and ****p < 000.1.

Supplementary Figure 7 LPS-induced mitochondrial ROS drives oxidative DNA damage.

(a) Nitrite production by M0, M(γ+LPS), M(LPS) and M(IL-4) polarized for 1 h or 18 h. Samples for which readings were non-detectable indicated as n.d. (n = 3 biologically independent experiments, representative of two independent experiments).(b) NRF2 and β-actin (loading control) expression in M0, M(γ+LPS), M(LPS) and M(IL-4) polarized for 1 h or 18 h as indicated. The blot was cropped to show relevant bands, and is representative of two independent experiments with similar results. Real-time changes in OCR (c), basal OCR (d) and SRC (e) of macrophages unstimulated, or stimulated with LPS for 0.5-4 h as indicated (n = 5 technical replicates, representative of three independent experiments). Macrophages were treated with 10 mM NAC (f, g) or MitoTempol (MT; h) for 1 h, then polarized with LPS. Subsequently MitoSox staining (f) and γ-H2AX expression (g,h) were assessed by flow cytometry (n = 3 biologically independent samples, representative of three (g,h) or four (f) independent experiments). (i) 8OHG immunoflouresence of cells treated with etoposide (positive control), or stained with an IgG control or secondary antibody only, and analysed by confocal microscopy. Scale bar: 5 μm, white arrows indicate nuclear co-localization of 8OHG staining, insert shows an enlargement of nuclei with 8OHG staining (data are representative of three independent experiments. Each experiment was performed with 3 biologically independent samples with 10 images collected per sample, per condition). Error bars are mean ± SEM. Data were analyzed by one-way ANOVA with Tukey’s multiple comparison test. *p < 0.05, **p < 0.01, ***p < 0.001 and ****p < 000.1.

Supplementary Figure 8 Inhibition of complex III and ROS scavenging affect inflammatory macrophage function.

(a) Viability of M(γ+LPS) polarized in the presence of CI inhibitor (Rotenone), CIII Qi inhibitor (Antimycin A) or a Complex III Qo inhibitor (Myx) for 1 h then harvested and assessed by flow cytometry. Viability presented as percentage control M(γ+LPS) (n = 3 biologically independent samples, representative of five independent experiments). (b) NAD+ and NADH quantification by LCMS in M(γ+LPS) polarized for 1 h with or without Myx (n = 3 biologically independent samples, representative of three independent experiments). (c) Viability of M(γ+LPS) at 18 h, after treatment with Myx or vehicle control for the first hour of polarization. Viability shown relative to vehicle control M(γ+LPS) (n = 3 biologically independent samples, representative of three independent experiments). (d,e) M(γ+LPS) were polarized with or without 10 mM NAC for 18 h then analyzed. (d) CD80, NOS2, MHC class II expression (n = 4 biologically independent samples, representative of three independent experiments). (e) IL-6 and TNF production (n = 3 biologically independent samples, representative of three independent experiments). (f,g) Macrophages were pre-treated for 2 h with NAC then stimulated with γ+LPS, 1 h later cells were washed and fresh γ+LPS added. 18 h later cells were analysed. (f) CD80, NOS2, MHC class II expression (n = 4 biologically independent samples, representative of two independent experiments). (g) IL-6 and TNF production (n = 4 biologically independent samples, representative of two independent experiments). Error bars are mean ± SEM. Data were analyzed by unpaired, two-tailed students t-test. *p < 0.05, **p < 0.01, ***p < 0.001 and ****p < 000.1.

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cameron, A.M., Castoldi, A., Sanin, D.E. et al. Inflammatory macrophage dependence on NAD+ salvage is a consequence of reactive oxygen species–mediated DNA damage. Nat Immunol 20, 420–432 (2019). https://doi.org/10.1038/s41590-019-0336-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41590-019-0336-y

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing