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:

Senescent cells promote tissue NAD+ decline during ageing via the activation of CD38+ macrophages

An Author Correction to this article was published on 10 December 2020

This article has been updated

Abstract

Declining tissue nicotinamide adenine dinucleotide (NAD) levels are linked to ageing and its associated diseases. However, the mechanism for this decline is unclear. Here, we show that pro-inflammatory M1-like macrophages, but not naive or M2 macrophages, accumulate in metabolic tissues, including visceral white adipose tissue and liver, during ageing and acute responses to inflammation. These M1-like macrophages express high levels of the NAD-consuming enzyme CD38 and have enhanced CD38-dependent NADase activity, thereby reducing tissue NAD levels. We also find that senescent cells progressively accumulate in visceral white adipose tissue and liver during ageing and that inflammatory cytokines secreted by senescent cells (the senescence-associated secretory phenotype, SASP) induce macrophages to proliferate and express CD38. These results uncover a new causal link among resident tissue macrophages, cellular senescence and tissue NAD decline during ageing and offer novel therapeutic opportunities to maintain NAD levels during ageing.

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: M1 macrophages are characterized by increased NADase activity.
Fig. 2: The NAM-salvage pathway controls NAD levels to regulate macrophage activation and polarization.
Fig. 3: High NADase activity in M1 macrophages is CD38 dependent.
Fig. 4: NAD decline during ageing is associated with increased CD38+ tissue-resident macrophages in eWAT.
Fig. 5: Cytokines secreted by senescent cells promote macrophage CD38 expression and proliferation.
Fig. 6: CD38+ Kupffer cells accumulate in the livers of ageing mice.
Fig. 7: Acute and chronic LPS treatment causes a CD38-dependent decrease in tissue NAD levels.
Fig. 8: Proposed model for how ageing-related inflammation enhances NAD degradation.

Similar content being viewed by others

Data availability

For proteomics data, all files are uploaded to the Center for Computational Mass Spectrometry, and can be downloaded using the following link ftp://massive.ucsd.edu/MSV000083726 (MassIVE ID number: MSV000083726). Data uploads include protein identification and quantification details, spectral library and FASTA file used for analysis. Tabula Muris Senis database is accessible at this link: https://tabula-muris-senis.ds.czbiohub.org. All other data that support the findings of this study are available from the corresponding author upon request. Source data are provided with this paper.

Code availability

Coding used in image analysis is available upon request. Source data are provided with this paper.

Change history

  • 10 December 2020

    An amendment to this paper has been published and can be accessed via a link at the top of the paper.

References

  1. Canto, C., Menzies, K. J. & Auwerx, J. NAD+ metabolism and the control of energy homeostasis: a balancing act between mitochondria and the nucleus. Cell Metab. 22, 31–53 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Yoshino, J., Baur, J. A. & Imai, S. I. NAD+ intermediates: the biology and therapeutic potential of NMN and NR. Cell Metab. 27, 513–528 (2018).

    CAS  PubMed  Google Scholar 

  3. Mitchell, S. J. et al. Nicotinamide improves aspects of healthspan, but not lifespan, in mice. Cell Metab. 27, 667–676 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  5. Bogan, K. L. & Brenner, C. Nicotinic acid, nicotinamide and nicotinamide riboside: a molecular evaluation of NAD+ precursor vitamins in human nutrition. Ann. Rev. Nutr. 28, 115–130 (2008).

    CAS  Google Scholar 

  6. Liu, L. et al. Quantitative analysis of NAD synthesis-breakdown fluxes. Cell Metab. 27, 1067–1080 e1065 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Camacho-Pereira, J. et al. CD38 dictates age-related NAD decline and mitochondrial dysfunction through an SIRT3-dependent mechanism. Cell Metab. 23, 1127–1139 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Jackson, D. G. & Bell, J. I. Isolation of a cDNA encoding the human CD38 (T10) molecule, a cell surface glycoprotein with an unusual discontinuous pattern of expression during lymphocyte differentiation. J. Immunol. 144, 2811–2815 (1990).

    CAS  PubMed  Google Scholar 

  9. Schneider, M. et al. CD38 is expressed on inflammatory cells of the intestine and promotes intestinal inflammation. PLoS ONE 10, e0126007 (2015).

    PubMed  PubMed Central  Google Scholar 

  10. Savarino, A., Bottarel, F., Malavasi, F. & Dianzani, U. Role of CD38 in HIV-1 infection: an epiphenomenon of T-cell activation or an active player in virus/host interactions? AIDS 14, 1079–1089 (2000).

    CAS  PubMed  Google Scholar 

  11. Franceschi, C. et al. Inflamm-aging. An evolutionary perspective on immunosenescence. Ann. N. Y. Acad. Sci. 908, 244–254 (2000).

    CAS  PubMed  Google Scholar 

  12. Franceschi, C. & Campisi, J. Chronic inflammation (inflammaging) and its potential contribution to age-associated diseases. J. Gerontol. A Biol. Sci. Med Sci. 69, S4–S9 (2014).

    PubMed  Google Scholar 

  13. Ganeshan, K. & Chawla, A. Metabolic regulation of immune responses. Annu. Rev. Immunol. 32, 609–634 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Coppe, J. P. et al. Senescence-associated secretory phenotypes reveal cell-nonautonomous functions of oncogenic RAS and the p53 tumor suppressor. PLoS Biol. 6, 2853–2868 (2008).

    CAS  PubMed  Google Scholar 

  15. Murray, P. J. et al. Macrophage activation and polarization: nomenclature and experimental guidelines. Immunity 41, 14–20 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Bieganowski, P. & Brenner, C. Discoveries of nicotinamide riboside as a nutrient and conserved NRK genes establish a Preiss–Handler independent route to NAD+ in fungi and humans. Cell 117, 495–502 (2004).

    CAS  PubMed  Google Scholar 

  17. Ratajczak, J. et al. NRK1 controls nicotinamide mononucleotide and nicotinamide riboside metabolism in mammalian cells. Nat. Commun. 7, 13103 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Cameron, A. M. et al. Inflammatory macrophage dependence on NAD+ salvage is a consequence of reactive oxygen species-mediated DNA damage. Nat. Immunol. 20, 420–432 (2019).

    CAS  PubMed  Google Scholar 

  19. Langston, P. K., Shibata, M. & Horng, T. Metabolism supports macrophage activation. Front. Immunol. 8, 61 (2017).

    PubMed  PubMed Central  Google Scholar 

  20. 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).

    PubMed  PubMed Central  Google Scholar 

  21. Amano, S. U. et al. Local proliferation of macrophages contributes to obesity-associated adipose tissue inflammation. Cell Metab. 19, 162–171 (2014).

    CAS  PubMed  Google Scholar 

  22. Zhang, Z. et al. Mouse macrophage specific knockout of SIRT1 influences macrophage polarization and promotes angiotensin II-induced abdominal aortic aneurysm formation. J. Genet Genomics 45, 25–32 (2018).

    PubMed  Google Scholar 

  23. Cockayne, D. A. et al. Mice deficient for the ecto-nicotinamide adenine dinucleotide glycohydrolase CD38 exhibit altered humoral immune responses. Blood 92, 1324–1333 (1998).

    CAS  PubMed  Google Scholar 

  24. Jablonski, K. A. et al. Novel markers to delineate murine M1 and M2 macrophages. PLoS ONE10, e0145342 (2015).

    PubMed  PubMed Central  Google Scholar 

  25. Shrimp, J. H. et al. Revealing CD38 cellular localization using a cell permeable, mechanism-based fluorescent small-molecule probe. J. Am. Chem. Soc. 136, 5656–5663 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Zhao, Y. J., Lam, C. M. & Lee, H. C. The membrane-bound enzyme CD38 exists in two opposing orientations. Sci. Signal 5, ra67 (2012).

    PubMed  Google Scholar 

  27. Liu, J. et al. Cytosolic interaction of type III human CD38 with CIB1 modulates cellular cyclic ADP-ribose levels. Proc. Natl Acad. Sci. USA 114, 8283–8288 (2017).

  28. Preugschat, F. et al. A pre-steady state and steady state kinetic analysis of the N-ribosyl hydrolase activity of hCD157. Arch. Biochem. Biophys. 564, 156–163 (2014).

    CAS  PubMed  Google Scholar 

  29. Tarrago, M. G. et al. A potent and specific CD38 inhibitor ameliorates age-related metabolic dysfunction by reversing tissue NAD+ decline. Cell Metab. 27, 1081–1095 e1010 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Regdon, Z. et al. LPS protects macrophages from AIF-independent parthanatos by downregulation of PARP1 expression, induction of SOD2 expression, and a metabolic shift to aerobic glycolysis. Free Radic. Biol. Med. 131, 184–196 (2019).

    CAS  PubMed  Google Scholar 

  31. Virag, L., Jaen, R. I., Regdon, Z., Bosca, L. & Prieto, P. Self-defense of macrophages against oxidative injury: fighting for their own survival. Redox Biol. 26, 101261 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Yoshino, J., Mills, K. F., Yoon, M. J. & Imai, S. Nicotinamide mononucleotide, a key NAD+ intermediate, treats the pathophysiology of diet- and age-induced diabetes in mice. Cell Metab. 14, 528–536 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Rabinowitz, S. S. & Gordon, S. Macrosialin, a macrophage-restricted membrane sialoprotein differentially glycosylated in response to inflammatory stimuli. J. Exp. Med. 174, 827–836 (1991).

    CAS  PubMed  Google Scholar 

  34. Xu, X. et al. Obesity activates a program of lysosomal-dependent lipid metabolism in adipose tissue macrophages independently of classic activation. Cell Metab. 18, 816–830 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Cho, K. W., Morris, D. L. & Lumeng, C. N. Flow cytometry analyses of adipose tissue macrophages. Methods Enzymol. 537, 297–314 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Consortium, T. T. M. Single-cell transcriptomics of 20 mouse organs creates a Tabula Muris. Nature 562, 367–372 (2018).

    Google Scholar 

  37. Chakarov, S. et al. Two distinct interstitial macrophage populations coexist across tissues in specific subtissular niches. Science 363, eaau0964 (2019).

  38. Mrdjen, D. et al. High-dimensional single-cell mapping of central nervous system immune cells reveals distinct myeloid subsets in health, aging, and disease. Immunity 48, 380–395 e386 (2018).

    CAS  PubMed  Google Scholar 

  39. Dimri, G. P. et al. A biomarker that identifies senescent human cells in culture and in aging skin in vivo. Proc. Natl Acad. Sci. USA 92, 9363–9367 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Coppe, J. P., Desprez, P. Y., Krtolica, A. & Campisi, J. The senescence-associated secretory phenotype: the dark side of tumor suppression. Annu. Rev. Pathol. 5, 99–118 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Tchkonia, T. et al. Fat tissue, aging, and cellular senescence. Aging Cell 9, 667–684 (2010).

    CAS  PubMed  Google Scholar 

  42. Xu, M. et al. Senolytics improve physical function and increase lifespan in old age. Nat. Med. 24, 1246–1256 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Demaria, M. et al. Cellular senescence promotes adverse effects of chemotherapy and cancer relapse. Cancer Discov. 7, 165–176 (2017).

    CAS  PubMed  Google Scholar 

  44. Pommier, Y., Leo, E., Zhang, H. & Marchand, C. DNA topoisomerases and their poisoning by anticancer and antibacterial drugs. Chem. Biol. 17, 421–433 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Baar, M. P. et al. Targeted apoptosis of senescent cells restores tissue homeostasis in response to chemotoxicity and aging. Cell 169, 132–147 e116 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Schaefer, L. Complexity of danger: the diverse nature of damage-associated molecular patterns. J. Biol. Chem. 289, 35237–35245 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Kratz, M. et al. Metabolic dysfunction drives a mechanistically distinct proinflammatory phenotype in adipose tissue macrophages. Cell Metab. 20, 614–625 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Iqbal, J. & Zaidi, M. TNF regulates cellular NAD+ metabolism in primary macrophages. Biochem. Biophys. Res. Commun. 342, 1312–1318 (2006).

    CAS  PubMed  Google Scholar 

  49. Jenkins, S. J. et al. Local macrophage proliferation, rather than recruitment from the blood, is a signature of TH2 inflammation. Science 332, 1284–1288 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Hashimoto, D. et al. Tissue-resident macrophages self-maintain locally throughout adult life with minimal contribution from circulating monocytes. Immunity 38, 792–804 (2013).

    CAS  PubMed  Google Scholar 

  51. Davies, L. C. et al. Distinct bone marrow-derived and tissue-resident macrophage lineages proliferate at key stages during inflammation. Nat. Commun. 4, 1886 (2013).

    PubMed  Google Scholar 

  52. Ginhoux, F. et al. Fate mapping analysis reveals that adult microglia derive from primitive macrophages. Science 330, 841–845 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Jager, J., Aparicio-Vergara, M. & Aouadi, M. Liver innate immune cells and insulin resistance: the multiple facets of Kupffer cells. J. Intern. Med. 280, 209–220 (2016).

    CAS  PubMed  Google Scholar 

  54. Stahl, E. C., Haschak, M. J., Popovic, B. & Brown, B. N. Macrophages in the aging liver and age-related liver disease. Front. Immunol. 9, 2795 (2018).

    PubMed  PubMed Central  Google Scholar 

  55. The Tabula Muris Consortium. A single-cell transcriptomic atlas characterizes ageing tissues in the mouse. Nature 583, 590–595 (2020).

  56. Ogrodnik, M. et al. Cellular senescence drives age-dependent hepatic steatosis. Nat. Commun. 8, 15691 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Demaria, M. et al. An essential role for senescent cells in optimal wound healing through secretion of PDGF-AA. Dev. Cell 31, 722–733 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Buford, T. W. (Dis)Trust your gut: the gut microbiome in age-related inflammation, health, and disease. Microbiome 5, 80 (2017).

  59. Thevaranjan, N. et al. Age-associated microbial dysbiosis promotes intestinal permeability, systemic inflammation, and macrophage dysfunction. Cell Host Microbe 21, 455–466 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Real, A. M., Hong, S. & Pissios, P. Nicotinamide N-oxidation by CYP2E1 in human liver microsomes. Drug Metab. Dispos. 41, 550–553 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Campisi, J. & Robert, L. Cell senescence: role in aging and age-related diseases. Interdiscip. Top. Gerontol. 39, 45–61 (2014).

    PubMed  PubMed Central  Google Scholar 

  62. Campisi, J. Aging, cellular senescence, and cancer. Annu Rev. Physiol. 75, 685–705 (2013).

    CAS  PubMed  Google Scholar 

  63. Chatterjee, S. et al. CD38–NAD+ axis regulates immunotherapeutic anti-tumor T cell response.Cell Metab. 27, 85–100 (2018).

    CAS  PubMed  Google Scholar 

  64. Lischke, T. et al. CD38 controls the innate immune response against Listeria monocytogenes. Infect. Immun. 81, 4091–4099 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Partida-Sanchez, S. et al. Chemotaxis of mouse bone marrow neutrophils and dendritic cells is controlled by ADP-ribose, the major product generated by the CD38 enzyme reaction. J. Immunol. 179, 7827–7839 (2007).

    CAS  PubMed  Google Scholar 

  66. Ganeshan, K. et al. Energetic trade-offs and hypometabolic states promote disease tolerance. Cell 177, 399–413 e312 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Wei, W., Graeff, R. & Yue, J. Roles and mechanisms of the CD38/cyclic adenosine diphosphate ribose/Ca2+ signaling pathway. World J. Biol. Chem. 5, 58–67 (2014).

    PubMed  PubMed Central  Google Scholar 

  68. Weiss, R. et al. Nicotinamide: a vitamin able to shift macrophage differentiation toward macrophages with restricted inflammatory features. Innate Immun. 21, 813–826 (2015).

    CAS  PubMed  Google Scholar 

  69. Chini, C. C. S. et al. CD38 ecto-enzyme in immune cells is induced during aging and regulates NAD+ and NMN levels. Nat. Metab. 2, 1284–1304 (2020).

    PubMed  PubMed Central  Google Scholar 

  70. Trammell, S. A. & Brenner, C. Targeted, LCMS-based metabolomics for quantitative measurement of NAD+ metabolites. Comput. Struct. Biotechnol. J. 4, e201301012 (2013).

    PubMed  PubMed Central  Google Scholar 

  71. Trammell, S. A. et al. Nicotinamide riboside is uniquely and orally bioavailable in mice and humans. Nat. Commun. 7, 12948 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  72. Schilling, B., Gibson, B. W. & Hunter, C. L. Generation of high-quality SWATH® acquisition data for label-free quantitative proteomics studies using tripleTOF® mass spectrometers. Methods Mol. Biol. 1550, 223–233 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Bruderer, R. et al. Extending the limits of quantitative proteome profiling with data-independent acquisition and application to acetaminophen-treated three-dimensional liver microtissues. Mol. Cell Proteom. 14, 1400–1410 (2015).

    CAS  Google Scholar 

  74. Velarde, M. C., Demaria, M., Melov, S. & Campisi, J. Pleiotropic age-dependent effects of mitochondrial dysfunction on epidermal stem cells. Proc. Natl Acad. Sci. USA 112, 10407–10412 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Wiley, C. D. et al. Mitochondrial dysfunction induces senescence with a distinct secretory phenotype. Cell Metab. 23, 303–314 (2016).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This project was supported by NIH grant R24DK085610 (E.V.), Gladstone Institute intramural funds (E.V.) and Buck Institute intramural funds (E.V. and J.C.). A.J.C. is a recipient of the UC President’s Postdoctoral Fellowship at UCSF and was supported by NIH training grant T32AG000266 (Buck Institute). A.K. was supported by the SENS Research Foundation and NIH grant R01AG051729 (J.C.). B.S. and N.B. were supported by NIH grant U01AG060906 (B.S., principal investigator) and NIH Shared Instrumentation Grant 1S10OD016281 (Buck Institute). M.S.S. and C.B. were supported by NIH grant R01HL147545 and the Roy J. Carver Trust to C.B. We thank M. Walter for help with imaging cells, and P. Dighe for help optimizing Seahorse assay conditions. We thank R. Camarda, V. Byles and D. Powell for reviewing the manuscript and helpful discussions.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization, A.J.C. and E.V.; Methodology, A.J.C. and E.V.; Investigation, A.J.C. (all experiments), A.K. (In vivo experiments, senescent-cell experiments, proteomics and flow cytometry), R.P. (NADase assays and flow cytometry), J.A.L.-D. (in vivo experiments), A.O.P. (single-cell RNA-seq analysis), H.G.K. (flow cytometry), M.S.S. (LC–MS), I.H. (IF imaging and analysis), R.K. (in vivo experiments, IF imaging and analysis), C.D.W (in vivo experiments), H.-S.W. (Seahorse assay), E.G. (Seahorse assay), S.S.I. (RNA analysis), N.B. (proteomics), Q.W. (IF imaging and analysis), I.-J.K. (in vitro experiments), E.S. (in vivo experiments), K.V. (in vivo experiments), K.-O.S. (LC–MS), Y.-M.L. (LC–MS), R.R. (In vivo experiments), I.-B.S. (LC–MS), M.O. (animal housing), B.S. (proteomics), M.S.-K. (IF imaging and analysis), K.I. (provided Cd157 KO and Cd38/Cd157 DKO bone marrow), S.R.Q. (single-cell RNA-seq analysis), J.N. (provided ageing mice), C.B. (LC–MS), J.C. (senescent-cell experiments); writing—original draft, A.J.C. and E.V.; writing—review and editing, all authors; supervision, E.V.; funding acquisition, see Acknowledgements.

Corresponding author

Correspondence to Eric Verdin.

Ethics declarations

Competing interests

All authors have reviewed and approved the manuscript. C.B. is the inventor of intellectual property on the nutritional and therapeutic uses of NR, serves as chief scientific advisor of and holds stock in ChromaDex. E.V. is a scientific cofounder of NAPA Therapeutics. A.J.C., R.P., Q.W., E.S. and K.V. received partial salary support from NAPA Therapeutics. J.C. is a scientific cofounder of Unity Biotechnology. The other authors declare no competing interests.

Additional information

Peer review information Primary Handling Editor: Pooja Jha.

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

Extended data

Extended Data Fig. 1 CD38 expression in human M1 macrophages and analysis of the de novo NAD pathway in BMDMs.

a, mRNA levels of CD38 in human peripheral blood monocytes (PBMC)-derived macrophages treated with recombinant human IL-4 (M2) or LPS (M1) for 18 hours. Representative data from one of three patient samples. (n = 4 independent biological experiments) b, Immunofluorescence of human PBMC derived macrophages stimulated as described above using an anti-human CD38 antibody (Green) and nuclei with DAPI (Blue). Scale bars represents 10μm. Analyzed in PBMCs derived from one patient. c, NADase activity in human PBMC-derived macrophages treated with recombinant human IL-4 (M2) or LPS (M1) for 18 hours. Shown is the mean of two separate experiments from different donors with 2 replicates each. d, Schematic of the de novo NAD synthesis pathway. e, mRNA levels of de novo NAD synthesis pathway enzymes. f, Quantification of tryptophan metabolites measured by LC-MS in M0, M2 and M1 mouse BMDMs activated for 24 hours. ND=not detected. Data shows the mean ± SEM n=3 independent experiments except in A and B. Statistical significance indicated as *P<0.05, **P<0.01, and ***P<0.001; two-sided Student’s t-test.

Source data

Extended Data Fig. 2 Analysis of the role of the NAM-salvage pathway and sirtuins in macrophage activation and polarization.

a, NAD levels quantified by LC-MS in M0, M1 and M2 BMDMs pre-treated or not with 50 nM FK866 and NR for 6 hours prior to stimulation with LPS for an additional 6 hours or IL-4 for 16 hours. b, mRNA levels of M2 genes in BMDMs pre-treated or not with FK866 and NR for 6 hours prior to stimulation with IL-4 for 16 hours. All statistical comparisons are relative to M2 + FK866. c, mRNA levels of M1 genes in BMDMs pre-treated or not with FK866 and NR for 6 hours prior to stimulation with LPS for 6 hours. All statistical comparisons are relative to M1 + FK866. d, Western analysis of Nampt Fl/Fl CreER and Nampt Fl/Fl BMDMs treated with 1 ug/ml tamoxifen. e, mRNA levels of M2 genes in Nampt Fl/Fl CreER and Nampt Fl/Fl BMDMs treated with IL-4 for 16 hours. f, mRNA levels of M1 genes in Nampt Fl/Fl CreER and Nampt Fl/Fl BMDMs treated with LPS for 6 hours. g, mRNA levels of M2 genes in WT BMDMs pretreated with the indicated concentration of the sirtuin inhibitors Ex527 and AGK2 for 30 minutes prior to stimulation with IL-4 for 16 hours. All statistical comparisons are relative to M2. h, mRNA levels of M1 genes in WT BMDMs pretreated with the indicated concentration of the sirtuin inhibitors Ex527 and AGK2 for 30 minutes prior to stimulation with LPS for 6 hours. All statistical comparisons are relative to M1. Data show the mean ± SEM. (n=3 independent experiments). Statistical significance defined as *P<0.05, **P<0.01, and ***P<0.001; two-sided Student’s t-test.

Source data

Extended Data Fig. 3 Heightended NADase activity in M1 macrophages is CD38 dependent and PARP1 independent.

a, Flow cytometry results comparing CD38 surface staining in naive (M0) WT and Cd38 KO BMDMs or BMDMs treated with IL-4 (M2) and LPS (M1) for 16 hours. b-c, NADase activity measured with non-cell permeable εNAD in intact M0, M2, and M1 WT and Cd38 KO BMDMs activated for 16 hours relative to cell number (B) and protein content (C). d, mRNA levels of Cd157 in M0 and M1 WT and Cd38 KO BMDMs treated for 16 hours. e, LC-MS quantification of NR in M0 and M1 WT and Cd38 KO BMDMs treated for 16 hours. f, mRNA levels of anti-oxidant genes in WT and Cd38 KO BMDMs treated with IL-4 (M2) and LPS (M1) for the indicated intervals. g, Western analysis of PARP activity (PARylation) and DNA damage (γH2AX) in WT and Cd38 KO BMDMs treated with IL-4 (M2) and LPS (M1) for the indicated intervals compared to WT MO macrophage treated with 1 mM H2O2 for 10 minutes. h, Western analysis of PARP activity (PARylation) and DNA damage (γH2AX) in WT and Cd38 KO BMDMs treated with IL-4 (M2) and LPS (M1) for 8 hours prior to treatment with 1 mM H2O2 for 10 minutes. Data show the mean ± SEM. (n= at least 3 independent experiments). Statistical significance indicated as *P<0.05, **P<0.01, and ***P<0.001; two-sided Student’s t-test. Unless noted with a bar, statistical comparisons are relative to the appropriate MO WT or Cd38 KO sample of the same genotype.

Source data

Extended Data Fig. 4 CD38+ resident macrophages accumulate in ageing adipose tissue.

a, LC-MS quantification of NAD in eWAT from WT young male mice (6 month) n=7 mice/group, Cd38 KO young male mice (3 month) n=5 mice/group, WT old male mice (25 month) n=10 mice/group, and Cd38 KO old male mice (26 month) n=5 mice/group. NAD concentrations are shown as pmol/mg of tissue. (same WT data from Fig. 4a). b, mRNA levels of Il-1α and IL-10 in eWAT from 6 and 25 month-old WT male mice. (WT young male mice (6 month) n=7 mice/group, WT old male mice (25 month) n=9 mice/group) c, Western analysis of adipose tissue from young (3 Month) and old (19 month) WT male mice to detect PARP activity (PARylation) and DNA damage (γH2AX). Each lane represents one mouse (young n=4 mice/group, old n=4 mice/group). d, mRNA levels of Cd38 in visceral adipose tissue, the stromal vascular fraction, and adipocyte fraction from young (3 month) and old (19 month) WT male mice. (young n=4 mice/group, old n=4 mice/group). e, Flow cytometry gating strategy to identify CD45+ immune cells from the stromal vascular fraction of eWAT. Cells were first gated on forward scatter (FSCA) vs side scatter (SSCA) to discard cell debris and dead or dying cells. Next FSCH (height) vs FSCA (Area) was used to select single cells. Single cells were then gated for auto-fluorescent using the Empty(E) BV421 vs BV711 channels (not used as antibody fluorophores) to discard cells that showed auto-fluorescence in these channels. Then CD45+ cells were selected and analyzed for CD38 and macrophage markers. Flow cytometry gating strategy to identify resident and non-resident macrophages from the stromal vascular fraction of eWAT, showing representative flow plots and histograms for the indicated ages of mice. f, Flow cytometry quantification of CD38- (low) resident macrophages, CD38- non-resident macrophages, and CD38+ (high) non-macrophage immune cells from eWAT of WT male mice for the ages shown. (2 months n=6 mice/group, 6 months n=5 mice/group, 12 months n=5 mice/group, 18 months n=5 mice/group, 25+ months n=7 mice/group) For in vivo experiments, data from individual mice are shown. Statistical significance indicated as *P<0.05, **P<0.01, and ***P<0.001; two-sided Student’s t-test.

Source data

Extended Data Fig. 5 Senescent cell burden is causally linked to increased macrophage CD38 expression.

a, mRNA levels of Il-1α, Cxcl1, and IL-10 in eWAT from 6 month-old WT male mice i.p. injected with Doxo or PBS. (PBS n=8 mice/group, Doxo n=7 mice/group) b, Quantification of CD38-low resident macrophages, and CD38-low non-resident macrophages from eWAT of 6 month-old WT male mice injected with Doxo or PBS. (PBS n=8 mice/group, Doxo n=8 mice/group). c, CD38 mRNA levels in WT and Cd38 KO BMDMs co-cultured (10:1) with non-senescent control mouse dermal fibroblasts (CTRL-MDF) or irradiated senescent MDF (Sen(IR)-MDF) for 24 hours. (n=4 independent biological experiments per condition) d, mRNA levels of Cd38 in WT BMDMs treated with the indicated DAMPs for 16 hours. (n=3 independent biological experiments per condition) e, mRNA levels of inflammatory genes in CTRL-MDF and Sen(IR)-MDF. (n=4 independent biological experiments per condition) f, mRNA levels of Cd38 in BMDMs treated with the indicated concentrations (ng/ml) of recombinant mouse cytokines for 24 hours. (n=3 independent biological experiments per condition) g, Heatmap of significantly upregulated proteins identified by mass spectrometry in conditioned media from CTRL-MDF and Sen(IR)-MDF. (n=4–6 independent biological experiments per condition). For in vivo experiments, data from individual mice are shown. Data show the mean ± SEM. (n= at least 3 independent experiments). Statistical significance indicated as *P<0.05, **P<0.01, and ***P<0.001; two-sided Student’s t-test.

Source data

Extended Data Fig. 6 Single-cell RNA sequencing analysis of inflammatory, NAD consuming, and biosynthetic genes in ageing hepatocytes and liver endothelial cells.

a, Dot-plot and heatmap of the indicated genes in liver hepatocytes based on age. Logarithmic axes, base-10. b, Dot-plot and heatmap of the indicated genes in liver endothelial cells based on age. Logarithmic axes, base-10.

Extended Data Fig. 7 LPS promotes tissue NAD decline via CD38.

a, Representative gating for the splenic leukocyte populations quantified in Fig. 7c, d and Extended Data Fig. 7a. Left six panels show gating for identification of B cells and different myeloid cells, as indicated, as well as gating for CD38-positive cells in all populations. Right six panels show gating for T cell subsets, as indicated. Red arrows indicate sequential gating, pointing from parent plots towards child plots. b, Quantification of immune cell populations and CD38+ immune cells in the spleen of 3 month-old WT male mice i.p. injected with PBS or LPS for 4 weeks, and analyzed by flow cytometry. (PBS n=10 mice/group, LPS n=9 mice/group) c, Western analysis of CD38, CD157, CD68, and NAMPT in eWAT of 3 month-old WT male mice injected with PBS or LPS for 4 weeks and Image J quantification of CD38 protein levels relative to NAMPT. Each lane represents one mouse (PBS n=4 mice/group, LPS n=5 mice/group) d, mRNA levels of NAD consuming enzymes in eWAT from 3 month-old WT male mice injected with PBS or LPS for 4 weeks. (PBS n=4 mice/group, LPS n=5 mice/group) e, mRNA levels of the indicated genes in whole eWAT from 4 month-old WT and Cd38 KO male mice injected with PBS or LPS for 12 hours. (n=10 mice/group) f, Western analysis of eWAT from 4 month-old WT and Cd38 KO male mice injected with PBS or LPS for 12 hours (n=3 mice/group). g, LC-MS quantification of NAD-related metabolites in eWAT from 4 month-old WT and Cd38 KO male mice IP injected with PBS or LPS for 12 hours. (n=10 mice/group) h, LC-MS quantification of NAD-related metabolites in liver from 4 month-old WT and Cd38 KO male mice IP injected with PBS or LPS for 12 hours. (n=10 mice/group) Data from individual mice are shown for in vivo experiments. Data show the mean ± SEM. Statistical significance indicated as *P<0.05, **P<0.01, and ***P<0.001; two-sided Student’s t-test except for 7 g and 7 h one-tailed t-test was used.

Source data

Supplementary information

Source data

Source Data Fig. 1

Statistical Source Data

Source Data Fig. 2

Statistical Source Data

Source Data Fig. 3

Statistical Source Data

Source Data Fig. 3

Unprocessed Western Blots

Source Data Fig. 4

Statistical Source Data

Source Data Fig. 4

Unprocessed Western Blots

Source Data Fig. 5

Statistical Source Data

Source Data Fig. 6

Statistical Source Data

Source Data Fig. 7

Statistical Source Data

Source Data Extended Data Fig. 1

Statistical Source Data

Source Data Extended Data Fig. 2

Statistical Source Data

Source Data Extended Data Fig. 2

Unprocessed Western Blots

Source Data Extended Data Fig. 3

Statistical Source Data

Source Data Extended Data Fig. 3

Unprocessed Western Blots

Source Data Extended Data Fig. 4

Statistical Source Data

Source Data Extended Data Fig. 4

Unprocessed Western Blots

Source Data Extended Data Fig. 5

Statistical Source Data

Source Data Extended Data Fig. 7

Statistical Source Data

Source Data Extended Data Fig. 7

Unprocessed Western Blots

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Covarrubias, A.J., Kale, A., Perrone, R. et al. Senescent cells promote tissue NAD+ decline during ageing via the activation of CD38+ macrophages. Nat Metab 2, 1265–1283 (2020). https://doi.org/10.1038/s42255-020-00305-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s42255-020-00305-3

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