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:

The kinase p38 activated by the metabolic regulator AMPK and scaffold TAB1 drives the senescence of human T cells

Subjects

Abstract

In T lymphocytes, the mitogen-activated protein kinase (MAPK) p38 regulates pleiotropic functions and is activated by canonical MAPK signaling or the alternative activation pathway downstream of the T cell antigen receptor (TCR). Here we found that senescent human T cells lacked the canonical and alternative pathways for the activation of p38 but spontaneously engaged the metabolic master regulator AMPK to trigger recruitment of p38 to the scaffold protein TAB1, which caused autophosphorylation of p38. Signaling via this pathway inhibited telomerase activity, T cell proliferation and the expression of key components of the TCR signalosome. Our findings identify a previously unrecognized mode for the activation of p38 in T cells driven by intracellular changes such as low-nutrient and DNA-damage signaling (an 'intrasensory' pathway). The proliferative defect of senescent T cells was reversed by blockade of AMPK-TAB1–dependent activation of p38.

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

Figure 1: Spontaneous activation of p38 in the absence of upstream canonical and alternative pathways in senescent T cells.
Figure 2: Spontaneous activation of p38 triggered by AMPK and mediated by TAB1 in senescent T cells.
Figure 3: DNA-damage and low-nutrient signals converge at AMPK, which drives TAB1-dependent activation of p38 in T cells.
Figure 4: AMPK-triggered recruitment of p38 to TAB1 causes autophosphorylation of p38.
Figure 5: Silencing of the expression of AMPKα or TAB1 restores telomerase and proliferation in senescent T cells.
Figure 6: Activation of p38 by a specific AMPK agonist reproduces senescent features in nonsenescent T cells.

Similar content being viewed by others

References

  1. Huang, Y. & Wange, R.L. T cell receptor signaling: beyond complex complexes. J. Biol. Chem. 279, 28827–28830 (2004).

    Article  CAS  PubMed  Google Scholar 

  2. Akbar, A.N., Beverley, P.C. & Salmon, M. Will telomere erosion lead to a loss of T-cell memory? Nat. Rev. Immunol. 4, 737–743 (2004).

    Article  CAS  PubMed  Google Scholar 

  3. Fletcher, J.M. et al. Cytomegalovirus-specific CD4+ T cells in healthy carriers are continuously driven to replicative exhaustion. J. Immunol. 175, 8218–8225 (2005).

    Article  CAS  PubMed  Google Scholar 

  4. Henson, S.M. et al. KLRG1 signaling induces defective Akt (ser473) phosphorylation and proliferative dysfunction of highly differentiated CD8+ T cells. Blood 113, 6619–6628 (2009).

    Article  CAS  PubMed  Google Scholar 

  5. Libri, V. et al. Cytomegalovirus infection induces the accumulation of short-lived, multifunctional CD4+CD45RA+CD27+ T cells: the potential involvement of interleukin-7 in this process. Immunology 132, 326–339 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Warrington, K.J., Vallejo, A.N., Weyand, C.M. & Goronzy, J.J. CD28 loss in senescent CD4+ T cells: reversal by interleukin-12 stimulation. Blood 101, 3543–3549 (2003).

    Article  CAS  PubMed  Google Scholar 

  7. Vallejo, A.N. CD28 extinction in human T cells: altered functions and the program of T-cell senescence. Immunol. Rev. 205, 158–169 (2005).

    Article  CAS  PubMed  Google Scholar 

  8. Weng, N.P., Akbar, A.N. & Goronzy, J. CD28 T cells: their role in the age-associated decline of immune function. Trends Immunol. 30, 306–312 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Di Mitri, D. et al. Reversible senescence in human CD4+CD45RA+CD27 memory T cells. J. Immunol. 187, 2093–2100 (2011).

    Article  CAS  PubMed  Google Scholar 

  10. Ashwell, J.D. The many paths to p38 mitogen-activated protein kinase activation in the immune system. Nat. Rev. Immunol. 6, 532–540 (2006).

    Article  CAS  PubMed  Google Scholar 

  11. Rudd, C.E. MAPK p38: alternative and nonstressful in T cells. Nat. Immunol. 6, 368–370 (2005).

    Article  CAS  PubMed  Google Scholar 

  12. Chang, L. & Karin, M. Mammalian MAP kinase signalling cascades. Nature 410, 37–40 (2001).

    Article  CAS  PubMed  Google Scholar 

  13. Dodeller, F. & Schulze-Koops, H. The p38 mitogen-activated protein kinase signaling cascade in CD4 T cells. Arthritis Res. Ther. 8, 205 (2006).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  14. Salvador, J.M. et al. Alternative p38 activation pathway mediated by T cell receptor-proximal tyrosine kinases. Nat. Immunol. 6, 390–395 (2005).

    Article  CAS  PubMed  Google Scholar 

  15. Plunkett, F.J. et al. The loss of telomerase activity in highly differentiated CD8+CD28CD27 T cells is associated with decreased Akt (Ser473) phosphorylation. J. Immunol. 178, 7710–7719 (2007).

    Article  CAS  PubMed  Google Scholar 

  16. Yan, M. et al. Activation of stress-activated protein kinase by MEKK1 phosphorylation of its activator SEK1. Nature 372, 798–800 (1994).

    Article  CAS  PubMed  Google Scholar 

  17. Jirmanova, L., Sarma, D.N., Jankovic, D., Mittelstadt, P.R. & Ashwell, J.D. Genetic disruption of p38α Tyr323 phosphorylation prevents T-cell receptor-mediated p38α activation and impairs interferon-γ production. Blood 113, 2229–2237 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Round, J.L. et al. Scaffold protein Dlgh1 coordinates alternative p38 kinase activation, directing T cell receptor signals toward NFAT but not NF-κB transcription factors. Nat. Immunol. 8, 154–161 (2007).

    Article  CAS  PubMed  Google Scholar 

  19. Rincon, M. & Davis, R.J. Choreography of MAGUKs during T cell activation. Nat. Immunol. 8, 126–127 (2007).

    Article  CAS  PubMed  Google Scholar 

  20. Ge, B. et al. MAPKK-independent activation of p38α mediated by TAB1-dependent autophosphorylation of p38α. Science 295, 1291–1294 (2002).

    Article  CAS  PubMed  Google Scholar 

  21. Ge, B. et al. TAB1β (transforming growth factor-β-activated protein kinase 1-binding protein 1β), a novel splicing variant of TAB1 that interacts with p38α but not TAK1. J. Biol. Chem. 278, 2286–2293 (2003).

    Article  CAS  PubMed  Google Scholar 

  22. Li, J., Miller, E.J., Ninomiya-Tsuji, J., Russell, R.R. III & Young, L.H. AMP-activated protein kinase activates p38 mitogen-activated protein kinase by increasing recruitment of p38 MAPK to TAB1 in the ischemic heart. Circ. Res. 97, 872–879 (2005).

    Article  CAS  PubMed  Google Scholar 

  23. Hardie, D.G. AMP-activated protein kinase: an energy sensor that regulates all aspects of cell function. Genes Dev. 25, 1895–1908 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Rolf, J. et al. AMPKα1: a glucose sensor that controls CD8 T-cell memory. Eur. J. Immunol. 43, 889–896 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Finlay, D. & Cantrell, D.A. Metabolism, migration and memory in cytotoxic T cells. Nat. Rev. Immunol. 11, 109–117 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Sanders, M.J. et al. Defining the mechanism of activation of AMP-activated protein kinase by the small molecule A-769662, a member of the thienopyridone family. J. Biol. Chem. 282, 32539–32548 (2007).

    Article  CAS  PubMed  Google Scholar 

  27. d'Adda di Fagagna, F. Living on a break: cellular senescence as a DNA-damage response. Nat. Rev. Cancer 8, 512–522 (2008).

    Article  CAS  PubMed  Google Scholar 

  28. Mondal, A.M. et al. p53 isoforms regulate aging- and tumor-associated replicative senescence in T lymphocytes. J. Clin. Invest. 123, 5247–5257 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Passos, J.F. & Von Zglinicki, T. Oxygen free radicals in cell senescence: are they signal transducers? Free Radic. Res. 40, 1277–1283 (2006).

    Article  CAS  PubMed  Google Scholar 

  30. Blattler, S.M., Rencurel, F., Kaufmann, M.R. & Meyer, U.A. In the regulation of cytochrome P450 genes, phenobarbital targets LKB1 for necessary activation of AMP-activated protein kinase. Proc. Natl. Acad. Sci. USA 104, 1045–1050 (2007).

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  31. Hodes, R.J., Hathcock, K.S. & Weng, N.P. Telomeres in T and B cells. Nat. Rev. Immunol. 2, 699–706 (2002).

    Article  CAS  PubMed  Google Scholar 

  32. Reed, J.R. et al. Telomere erosion in memory T cells induced by telomerase inhibition at the site of antigenic challenge in vivo. J. Exp. Med. 199, 1433–1443 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Lanna, A. et al. IFN-alpha inhibits telomerase in human CD8+ T Cells by both hTERT downregulation and Induction of p38 MAPK signalling. J. Immunol. (2013).

  34. Akbar, A.N. & Vukmanovic-Stejic, M. Telomerase in T lymphocytes: use it and lose it? J. Immunol. 178, 6689–6694 (2007).

    Article  CAS  PubMed  Google Scholar 

  35. Akbar, A.N. & Henson, S.M. Are senescence and exhaustion intertwined or unrelated processes that compromise immunity? Nat. Rev. Immunol. 11, 289–295 (2011).

    Article  CAS  PubMed  Google Scholar 

  36. Chang, C.H. et al. Posttranscriptional control of T cell effector function by aerobic glycolysis. Cell 153, 1239–1251 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Shibuya, H. et al. TAB1: an activator of the TAK1 MAPKKK in TGF-β signal transduction. Science 272, 1179–1182 (1996).

    Article  CAS  PubMed  Google Scholar 

  38. Jacquet, S. et al. The relationship between p38 mitogen-activated protein kinase and AMP-activated protein kinase during myocardial ischemia. Cardiovasc. Res. 76, 465–472 (2007).

    Article  CAS  PubMed  Google Scholar 

  39. Jaswal, J.S., Gandhi, M., Finegan, B.A., Dyck, J.R. & Clanachan, A.S. p38 mitogen-activated protein kinase mediates adenosine-induced alterations in myocardial glucose utilization via 5′-AMP-activated protein kinase. Am. J. Physiol. Heart Circ. Physiol. 292, H1978–H1985 (2007).

    Article  CAS  PubMed  Google Scholar 

  40. Alam, S.M. et al. Counter-regulation of T cell effector function by differentially activated p38. J. Exp. Med. 211, 1257–1270 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. De Nicola, G.F. et al. Mechanism and consequence of the autoactivation of p38α mitogen-activated protein kinase promoted by TAB1. Nat. Struct. Mol. Biol. 20, 1182–1190 (2013).

    Article  CAS  Google Scholar 

  42. Lee, J.C. et al. Inhibition of p38 MAP kinase as a therapeutic strategy. Immunopharmacology 47, 185–201 (2000).

    Article  CAS  PubMed  Google Scholar 

  43. Tivey, H.S., Brook, A.J., Rokicki, M.J., Kipling, D. & Davis, T. p38 (MAPK) stress signalling in replicative senescence in fibroblasts from progeroid and genomic instability syndromes. Biogerontology 14, 47–62 (2013).

    Article  CAS  PubMed  Google Scholar 

  44. Bernet, J.D. et al. p38 MAPK signaling underlies a cell-autonomous loss of stem cell self-renewal in skeletal muscle of aged mice. Nat. Med. 20, 265–271 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Escors, D. et al. Targeting dendritic cell signaling to regulate the response to immunization. Blood 111, 3050–3061 (2008).

    Article  CAS  PubMed  Google Scholar 

  46. Tangeman, L., Wyatt, C.N. & Brown, T.L. Knockdown of AMP-activated protein kinase α1 and α2 catalytic subunits. J RNAi Gene Silencing 8, 470–478 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Sarbassov, D.D., Guertin, D.A., Ali, S.M. & Sabatini, D.M. Phosphorylationand regulation of Akt/PKB by the rictor-mTOR complex. Science 307, 1098–1101 (2005).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank S.S. Marelli for discussions. Supported by the Medical Research Council (A.L.), the Biotechnology and Biological Science Research Council (BB/J006750/1 to A.N.A. and S.M.H.) and the Instituto de Salud Carlos III, Spain (D.E.).

Author information

Authors and Affiliations

Authors

Contributions

A.L. conceived of and performed the study, analyzed and interpreted the data and wrote the paper; S.M.H. performed experiments; D.E. provided lentiviral tools; A.N.A. wrote the paper and provided overall direction; and all authors read and approved the final version of the manuscript.

Corresponding author

Correspondence to Arne N Akbar.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Loss of the TCR signalosome by highly differentiated T cells.

(a) Immunoblot of total Lat, SLP-76 and PLC-γ 1 expression in freshly isolated human CD4+ CD27/CD28 defined subsets. GAPDH was used as a loading control. (b) The relative expression of Lat, SLP-76 and PLC-γ 1 vs. total GAPDH in the CD4+ CD27/CD28 defined subsets from 3 different individuals. (c) Representative overlay and (d) pooled Fluo-4 AM data from 3 independent experiments showing intracellular calcium influx after αCD3 activation (10 μg/mL, 5) in CD27+ CD28+ and CD27- CD28- CD4+ T cells, as measured by flow-cytometry. In (b) a one-way analysis of variance (ANOVA) for repeated-measures with a Bonferroni post-test correction. In (d) a paired Student’s t test. *p< 0,05, **p<0.01, and ***p< 0.001. Error bars depict s.e.m

Supplementary Figure 2 Confirmation of the knockdown of AMPKα and TAB1 by lentiviral vector in senescent CD27CD28 CD4+ T cells.

(a) Extended map and (b) schematic representation of the p-Siren HIV lentiviral vector system used for gene knock-down. The shRNA of interest was cloned downstream the U6-promoter between the BamHI and EcoRI restriction sites. Measurement of (c) AMPKα and (d) TAB1 expression by both immunoblot (top) and quantitative PCR (bottom) in highly differentiated CD27- CD28- CD4+ T cells transduced with either shAMPKα or shTAB1, as indicated. An irrelevant scrambled short hairpin (shCtrl) expressing vector was used as internal control. Immunoblots are representative of 3 independent experiments. Samples for quantitative PCR were analyzed from 4 different donors and normalized against housekeeping GAPDH. Knockdown efficiency was evaluated 96 hours post-transduction (day 7). **p<0.01 and ***p< 0.001 values were assessed by a paired Student’s t test. Errors bars depict s.e.m.

Supplementary Figure 3 Enhanced levels of ROS endogenously activate ATM, AMPK and p38 in senescent CD27CD28 CD4+ T cells.

(a) Intracellular ROS levels by Dihydroethidium (DHE) in CD27+ CD28+ and CD27- CD28- CD4+ T cells directly ex vivo, as measured by flow cytometry. Data are representative of 4 independent experiments. (b) Representative overlays and (c) pooled phospho-flow data from 3 different donors showing the effect of the ROS scavenger superoxide-dismutase (SOD; 100 U, 60’) on AMPK (Thr172), p38 (Thr180,Tyr182) and ATM (Ser1921) phosphorylation in senescent CD27- CD28- CD4+ T cells. All *p< 0,05, **p<0.01, and ***p< 0.001 values were calculated by a paired Student’s t test. Error bars depict s.e.m

Supplementary Figure 4 Silencing p38 restores telomerase and proliferative activities in senescent human CD27CD28 CD4+ T cells.

(a) Experimental design. Purified human senescent CD27- CD28- CD4+ T cells were activated by αCD3 plus rh-IL-2, and then transduced 48 hours later with lentiviral particles. A second hit transduction was performed 24 hours later. Following transduction, medium containing rh-IL-2 was replaced every 2-3 days. CD27- CD28- CD4+ T cells were maintained in activation medium and analyzed for functional readouts at the indicated time points (see text). For long-term cultures, transduced cells were subjected to a number of two subsequent rounds of activation by αCD3 plus rh-IL2, every 10 days. (b) Measurement of p38 expression by both immunoblot (left) and quantitative PCR (right) in CD27- CD28- CD4+ T cells, transduced as indicated. Immunoblots are representative of 2 independent experiments. Samples for quantitative PCR were analyzed from 3 different donors and normalized against housekeeping GAPDH. Knockdown efficiency was evaluated 96 hours post-transduction (day 7). (c) Telomerase activity by TRAP assay in CD27- CD28- CD4+ T cells transduced as indicated. Experiments were performed from 3 separate donors, 96 hours post-transduction. Proliferative activity by either (d) 3H thymidine incorporation or (e) dye dilution assay in CD27- CD28- CD4+ T cells, transduced as indicated. Experiments were performed from 3 separate donors, 96 hours after transduction. (f) The pooled results of 3 separate experiments performed as in (e). (g) Replicative lifespan by cumulative PDs of long-term cultured CD4+ CD27- CD28- T cells, transduced as indicated. Cumulative PDs were assessed by absolute cell number enumeration from 3 separate donors. All *p< 0,05, **p<0.01, and ***p< 0.001 values were calculated by a paired Student’s t test. Error bars depict s.e.m.

Supplementary Figure 5 p38-mediated modulation of cell-cycle machinery in senescent human CD27CD28 CD4+ T cells.

Representative overlay and pooled phospho-flow data (n=3) showing the effect of the p38 inhibitor BIRB (500 nM) on (a,b) p27, (c,d) D1 cyclin and (e,f) under-phosphorylated Rb expression in CD27- CD28- CD4+ T cells, activated by CD3/rh-IL-2 for 36 hours. (g) Rb inactivation and (h) p53 activation were assessed by measuring the ratio of the inhibitory phosphorylated Rb (Ser807,Ser811) to under-phosphorylated Rb and phosphorylation of p53 at residue Ser46, respectively. All data are from 3 independent experiments. All **p<0.01 values were calculated by a paired Student’s t test. Error bars depict s.e.m

Supplementary Figure 6 AMPK activation inhibits human T cell telomerase and proliferation via p38.

Measurement of (a) telomerase activity by TRAP assay and (b) proliferation by thymidine incoporation in relatively undifferentiated CD27+ CD28+ CD4+ T cells activated by αCD3/CD28 antibodies and transduced with either shCtrl or shP38, in the presence of DMSO control or the AMPK agonist A-769662 (150 μM). (c) Representative Facs plot and (d) pooled data of Ki67 expression by CD27+ CD28+ CD4+ T cells activated by CD3/CD28 in the presence or absence of the p38 inhibitor BIRB 796 (500 nM), under glucose starvation. Activated CD27+ CD28+ CD4+ T cells grown in the presence of glucose [20mM] served as positive proliferation control. (e) Telomerase activity by TRAP assay in glucose starved CD27+ CD28+ CD4+T cells activated for 36 hours in the presence or absence of BIRB 796. In (a,b) *p< 0,05, **p<0.01, and ***p< 0.001 values were calculated by a one-way analysis of variance (ANOVA) for repeated-measures with a Bonferroni post-test correction; in (d,e) *p< 0,05 was calculated by a paired Student’s t test. Error bars depict s.e.m.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–6 (PDF 2199 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lanna, A., Henson, S., Escors, D. et al. The kinase p38 activated by the metabolic regulator AMPK and scaffold TAB1 drives the senescence of human T cells. Nat Immunol 15, 965–972 (2014). https://doi.org/10.1038/ni.2981

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ni.2981

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