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:

N-myristoyltransferase deficiency impairs activation of kinase AMPK and promotes synovial tissue inflammation

Abstract

N-myristoyltransferase (NMT) attaches the fatty acid myristate to the N-terminal glycine of proteins to sort them into soluble and membrane-bound fractions. Function of the energy-sensing AMP-activated protein kinase, AMPK, is myristoylation dependent. In rheumatoid arthritis (RA), pathogenic T cells shift glucose away from adenosine tri-phosphate production toward synthetic and proliferative programs, promoting proliferation, cytokine production, and tissue invasion. We found that RA T cells had a defect in NMT1 function, which prevented AMPK activation and enabled unopposed mTORC1 signaling. Lack of the myristate lipid tail disrupted the lysosomal translocation and activation of AMPK. Instead, myristoylation-incompetent RA T cells hyperactivated the mTORC1 pathway and differentiated into pro-inflammatory TH1 and TH17 helper T cells. In vivo, NMT1 loss caused robust synovial tissue inflammation, whereas forced NMT1 overexpression rescued AMPK activation and suppressed synovitis. Thus, NMT1 has tissue-protective functions by facilitating lysosomal recruitment of AMPK and dampening mTORC1 signaling.

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: N-myristoyltransferase 1 (NMT1) deficiency in T cells from patients with RA.
Fig. 2: NMT1 controls pro-inflammatory T cell differentiation.
Fig. 3: N-myristoylation protects against synovial inflammation.
Fig. 4: Impaired activation of the energy sensor AMPK in RA T cells.
Fig. 5: AMPK activation is myristoylation dependent.
Fig. 6: NMT1 controls lysosomal recruitment of AMPK.
Fig. 7: NMT1 deficiency promotes mTORC1 hyperactivity in RA T cells.
Fig. 8: AMPK activation corrects arthrogenic T cell effector functions in vitro and in vivo.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding author upon request.

References

  1. Rao, D. A. et al. Pathologically expanded peripheral T helper cell subset drives B cells in rheumatoid arthritis. Nature 542, 110–114 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Weyand, C. M. & Goronzy, J. J. T-cell-targeted therapies in rheumatoid arthritis. Nat. Clin. Pract. Rheumatol. 2, 201–210 (2006).

    Article  CAS  PubMed  Google Scholar 

  3. Weyand, C. M. & Goronzy, J. J. Immunometabolism in early and late stages of rheumatoid arthritis. Nat. Rev. Rheumatol. 13, 291–301 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Weyand, C. M., Shen, Y. & Goronzy, J. Redox-sensitive signaling in inflammatory T cells and in autoimmune disease. Free Radic. Biol. Med. 125, 36–43 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Shen, Y. et al. Metabolic control of the scaffold protein TKS5 in tissue-invasive, proinflammatory T cells. Nat. Immunol. 18, 1025–1034 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Yang, Z., Fujii, H., Mohan, S. V., Goronzy, J. J. & Weyand, C. M. Phosphofructokinase deficiency impairs ATP generation, autophagy, and redox balance in rheumatoid arthritis T cells. J. Exp. Med. 210, 2119–2134 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Yang, Z. et al. Restoring oxidant signaling suppresses proarthritogenic T cell effector functions in rheumatoid arthritis. Sci. Transl. Med. 8, 331ra338 (2016).

    Google Scholar 

  8. Tsokos, G. C. Metabolic control of arthritis: switch pathways to treat. Sci. Transl. Med. 8, 331fs338 (2016).

    Article  Google Scholar 

  9. Shao, L. et al. Deficiency of the DNA repair enzyme ATM in rheumatoid arthritis. J. Exp. Med. 206, 1435–1449 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Li, Y. et al. Deficient activity of the nuclease MRE11A induces T cell aging and promotes arthritogenic effector functions in patients with rheumatoid arthritis. Immunity 45, 903–916 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Tsokos, G. C. Fat T cells go to the joint. Nat. Immunol. 18, 955–956 (2017).

    Article  CAS  PubMed  Google Scholar 

  12. Wang, C. W. Lipid droplets, lipophagy, and beyond. Biochim. Biophys. Acta 1861, 793–805 (2016).

    Article  CAS  PubMed  Google Scholar 

  13. Ma, E. H., Poffenberger, M. C., Wong, A. H. & Jones, R. G. The role of AMPK in T cell metabolism and function. Curr. Opin. Immunol. 46, 45–52 (2017).

    Article  CAS  PubMed  Google Scholar 

  14. Gowans, G. J., Hawley, S. A., Ross, F. A. & Hardie, D. G. AMP is a true physiological regulator of AMP-activated protein kinase by both allosteric activation and enhancing net phosphorylation. Cell. Metab. 18, 556–566 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Herzig, S. & Shaw, R. J. AMPK: guardian of metabolism and mitochondrial homeostasis. Nat. Rev. Mol. Cell Biol. 19, 121–135 (2018).

    Article  CAS  PubMed  Google Scholar 

  16. Zhang, C. S. et al. The lysosomal v-ATPase-Ragulator complex is a common activator for AMPK and mTORC1, acting as a switch between catabolism and anabolism. Cell. Metab. 20, 526–540 (2014).

    Article  CAS  PubMed  Google Scholar 

  17. Lin, S. C. & Hardie, D. G. AMPK: sensing glucose as well as cellular energy status. Cell. Metab. 27, 299–313 (2018).

    Article  CAS  PubMed  Google Scholar 

  18. Zhang, C. S. et al. Fructose-1,6-bisphosphate and aldolase mediate glucose sensing by AMPK. Nature 548, 112–116 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Blagih, J. et al. The energy sensor AMPK regulates T cell metabolic adaptation and effector responses in vivo. Immunity 42, 41–54 (2015).

    Article  CAS  PubMed  Google Scholar 

  20. Kim, J., Kundu, M., Viollet, B. & Guan, K. L. AMPK and mTOR regulate autophagy through direct phosphorylation of Ulk1. Nat. Cell Biol. 13, 132–141 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Chi, H. Regulation and function of mTOR signalling in T cell fate decisions. Nat. Rev. Immunol. 12, 325–338 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Delgoffe, G. M. et al. The kinase mTOR regulates the differentiation of helper T cells through the selective activation of signaling by mTORC1 and mTORC2. Nat. Immunol. 12, 295–303 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Sancak, Y. et al. Ragulator-Rag complex targets mTORC1 to the lysosomal surface and is necessary for its activation by amino acids. Cell 141, 290–303 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Hardie, D. G. AMPK and Raptor: matching cell growth to energy supply. Mol. Cell 30, 263–265 (2008).

    Article  CAS  PubMed  Google Scholar 

  25. Oakhill, J. S. et al. beta-Subunit myristoylation is the gatekeeper for initiating metabolic stress sensing by AMP-activated protein kinase (AMPK). Proc. Natl Acad. Sci. USA 107, 19237–19241 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Udenwobele, D. I. et al. Myristoylation: an important protein modification in the immune response. Front. Immunol. 8, 751 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  27. Ducker, C. E., Upson, J. J., French, K. J. & Smith, C. D. Two N-myristoyltransferase isozymes play unique roles in protein myristoylation, proliferation, and apoptosis. Mol. Cancer Res. 3, 463–476 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Yang, S. H. et al. N-myristoyltransferase 1 is essential in early mouse development. J. Biol. Chem. 280, 18990–18995 (2005).

    Article  CAS  PubMed  Google Scholar 

  29. Shrivastav, A. et al. Requirement of N-myristoyltransferase 1 in the development of monocytic lineage. J. Immunol. 180, 1019–1028 (2008).

    Article  CAS  PubMed  Google Scholar 

  30. Weyand, C. M., Yang, Z. & Goronzy, J. J. T-cell aging in rheumatoid arthritis. Curr. Opin. Rheumatol. 26, 93–100 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. 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 

  32. Navarro, M. N. & Cantrell, D. A. Serine-threonine kinases in TCR signaling. Nat. Immunol. 15, 808–814 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Gwinn, D. M. et al. AMPK phosphorylation of raptor mediates a metabolic checkpoint. Mol. Cell 30, 214–226 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Zhang, C. S. et al. Metformin activates AMPK through the lysosomal pathway. Cell. Metab. 24, 521–522 (2016).

    Article  PubMed  Google Scholar 

  35. Fullerton, M. D. et al. Single phosphorylation sites in Acc1 and Acc2 regulate lipid homeostasis and the insulin-sensitizing effects of metformin. Nat. Med. 19, 1649–1654 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Lochner, M., Berod, L. & Sparwasser, T. Fatty acid metabolism in the regulation of T cell function. Trends Immunol. 36, 81–91 (2015).

    Article  CAS  PubMed  Google Scholar 

  37. Dudek, E. et al. N-Myristoyltransferase 1 interacts with calnexin at the endoplasmic reticulum. Biochem. Biophys. Res. Commun. 468, 889–893 (2015).

    Article  CAS  PubMed  Google Scholar 

  38. Ohta, H., Takamune, N., Kishimoto, N., Shoji, S. & Misumi, S. N-Myristoyltransferase 1 enhances human immunodeficiency virus replication through regulation of viral RNA expression level. Biochem. Biophys. Res. Commun. 463, 988–993 (2015).

    Article  CAS  PubMed  Google Scholar 

  39. Kim, S. et al. Blocking myristoylation of SRC inhibits its kinase activity and suppresses prostate cancer progression. Cancer Res. 77, 6950–6962 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Thinon, E. et al. Global profiling of co- and post-translationally N-myristoylated proteomes in human cells. Nat. Commun. 5, 4919 (2014).

    Article  CAS  PubMed  Google Scholar 

  41. Liang, J. et al. Myristoylation confers noncanonical AMPK functions in autophagy selectivity and mitochondrial surveillance. Nat. Commun. 6, 7926 (2015).

    Article  CAS  PubMed  Google Scholar 

  42. Hardie, D. G. AMPK—sensing energy while talking to other signaling pathways. Cell. Metab. 20, 939–952 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Kelly, B., Tannahill, G. M., Murphy, M. P. & O’Neill, L. A. Metformin inhibits the production of reactive oxygen species from NADH:ubiquinone oxidoreductase to limit induction of interleukin-1beta (IL-1beta) and boosts interleukin-10 (IL-10) in lipopolysaccharide (LPS)-activated macrophages. J. Biol. Chem. 290, 20348–20359 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. O’Neill, L. A., Kishton, R. J. & Rathmell, J. A guide to immunometabolism for immunologists. Nat. Rev. Immunol. 16, 553–565 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  45. Perl, A. Review: metabolic control of immune system activation in rheumatic diseases. Arthritis Rheumatol. 69, 2259–2270 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Cool, B. et al. Identification and characterization of a small molecule AMPK activator that treats key components of type 2 diabetes and the metabolic syndrome. Cell. Metab. 3, 403–416 (2006).

    Article  CAS  PubMed  Google Scholar 

  47. Lai, Z. W. et al. Sirolimus in patients with clinically active systemic lupus erythematosus resistant to, or intolerant of, conventional medications: a single-arm, open-label, phase 1/2 trial. Lancet 391, 1186–1196 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Huang, N. & Perl, A. Metabolism as a target for modulation in autoimmune diseases. Trends Immunol. 39, 562–576 (2018).

    Article  CAS  PubMed  Google Scholar 

  49. Wen, Z. et al.The microvascular niche instructs T cells in large vessel vasculitis via the VEGF-Jagged1-Notch pathway. Sci. Transl. Med. 9, pii: eaal3322 (2017).

    Article  Google Scholar 

  50. Fang, Y. et al. Duration of rapamycin treatment has differential effects on metabolism in mice. Cell. Metab. 17, 456–462 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was supported by the National Institutes of Health (grant nos. R01 AR042527, R01 HL 117913, R01 AI108906 and P01 HL129941 to C.M.W. and nos. R01 AI108891, R01 AG045779, U19 AI057266 and I01 BX001669 to J.J.G.).

Author information

Authors and Affiliations

Authors

Contributions

Z.W., C.M.W., and J.J.G. designed the study and analyzed the data. Z.W., K.J., Y.S., Z.Y., Y.L., and B.W. performed experiments. S.S., N.E.R., C.M.W., and J.J.G. were responsible for patient selection, evaluation, and recruitment. L.T. supervised all statistical analyses. C.M.W., Z.W., and J.J.G. wrote the manuscript.

Corresponding author

Correspondence to Cornelia M. Weyand.

Ethics declarations

Competing interests

The authors declare no competing interests.

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 Expression of NMT1 and NMT2.

(a) CD4+CD45RA+ T cells were isolated from patients with rheumatoid arthritis (RA), patients with psoriatic arthritis (PsA) or age-matched healthy individuals and stimulated for 72 h. NMT2 expression was analyzed by flow cytometry. Data (mean ± SEM) from 10 RA patients, 8 PsA patients and 10 healthy individuals. One-way ANOVA and post-ANOVA pair-wise two-group comparisons conducted with Tukey’s method. (b) CD4+CD45RA+ T cells were stimulated for 72 h and NMT1 mRNA expression was measured by RT-PCR. Mean ± SEM from 10 samples in each group. Unpaired Mann-Whitney-Wilcoxon rank test. (c-h) PBMCs were collected from RA patients, PsA patients and healthy individuals. Multicolor flow cytometry was applied to quantify NMT1 expression in CD4+CD45RA+ T cells, CD4+CD45RA T cells, CD8+CD45RA+ T cells, CD8+CD45RA T cells, CD19+ B cells and CD14+ monocytes. Representative histograms and collective MFIs (mean ± SEM) for each of the subpopulations from 11 individuals in each donor cohort. One-way ANOVA and post-ANOVA pair-wise two-group comparisons conducted with Tukey’s method. **p < 0.01. ***p < 0.001

Supplementary Figure 2 Genetic manipulation of NMT1 expression in CD4+ T cells.

(a) CD4+CD45RA+ T cells from RA patients were transfected with a NMT1 expression vector or a control vector. (b) CD4+CD45RA+ T cells from healthy subjects were transfected with NMT1 siRNA or control siRNA. 24 h later, NMT1 protein expression was analyzed with flow cytometry. Representative histograms and collective data from 6 individuals in each group. Paired Mann-Whitney-Wilcoxon rank test. *p < 0.05

Supplementary Figure 3 NMT1 regulates lipid droplet accumulation in T cells.

CD4+CD45RA+ T cells from RA patients and healthy donors were stimulated for 72 h. Neutral lipid droplets were stained with Bodipy 493/503 and analyzed by flow cytometry. (a) Comparison of neutral lipid expression in control and RA T cells. Representative histograms and results from 6 patient-control pairs. Mean ± SEM. **p< 0.01 by paired Mann-Whitney-Wilcoxon rank test. (b) NMT1 activity was restored in T cells from RA patients by transfecting a NMT1-expressing vector. Lipid droplets were detected by staining with Bodipy. Representative histograms are shown. (c) NMT1 activity was inhibited by transfecting T cells from healthy individuals with NMT1 siRNA. Histograms of Bodipy staining for lipid droplets are shown.

Supplementary Figure 4 NMT1 and NMT2 expression in CD4+ T cells is independent of metabolic signals.

Naïve CD4 T cells from 6 healthy donors were stimulated for 72 h in the absence or presence of C75 (20 μM), 3PO (200 nM), ML265 (10 μM), Shikonin (250 nM), pyruvate (1 mM), succinate (1 mM), malic acid (1 mM), A769662 (10 μM), Compound C (1 μM) or rapamycin (10 μM), respectively. Protein expression of NMT1 and NMT2 in CD4 T cells was determined by flow cytometry. One-way ANOVA and post-ANOVA pair-wise two-group comparisons were conducted with Tukey’s method

Supplementary Figure 5 N-myristoylation protects against synovial inflammation.

(a) Scheme of experimental design. NSG mice were engrafted with human synovial tissue. Seven days later, CD4 T cells were FACS sorted from CD4+CD45RO PBMCs of RA patients or healthy individuals, manipulated for NMT1 expression by transfecting with a NMT1 vector, NMT1 siRNA or controls respectively, and added back to the CD4-depleted PBMCs for adoptive transfer into the chimeras. In other experiments, CD45RO PBMCs were transfected. At day 14, synovial tissues were harvested for transcriptome analysis and immunohistochemical staining. (b-e) NMT1 overexpression inhibits synovial inflammation. NSG mice were engrafted with human synovial tissue. Seven days after engraftment, CD45RO PBMCs from RA patients were transfected with a NMT1-expressing or control vector and transferred into the chimeric mice. Seven days later, synovial tissues were harvested for transcriptome analysis and immunostaining. (b) Tissue sections stained with anti-IFN-γ (green) and anti-CD3 (red). Nuclei marked with DAPI. Representative images from 6 grafts. Scale bars 20 μm. (cd) Frequencies of CD3+ T cells and of CD3+IFN-γ+ T cells in tissue sections. (e) Heat map presentation of gene transcripts measured in tissue extracts by qPCR. Data are mean ± SEM from 6 synovial grafts. *p< 0.05 by paired Mann-Whitney-Wilcoxon rank test. (f-i) NMT1 knockdown promotes synovial inflammation. Human synovial tissue was transplanted into NSG mice. After engraftment, CD45RO PBMCs from healthy donors were transfected with NMT1 siRNA or control siRNA and adoptively transferred into the chimeric mice. Tissue transcriptome and immunohistochemical stains were analyzed in synovial explants after 7 days. (f) Tissue-infiltrating human T cells evaluated by dual-color immunostaining of CD3 (red) and IFN-γ (green). Nuclei marked with DAPI. Representative images from 6 grafts. Scale bars 20 μm. (g-h) Frequencies of tissue CD3+ T cells and of CD3+IFN-γ+ T cells. (i) Tissue transcriptome of inflammation-associated genes shown as a heat map. Data are mean ± SEM from 6 synovial grafts. *p< 0.05 by paired Mann-Whitney-Wilcoxon rank test

Supplementary Figure 6 AMPK activity and transcriptional expression in healthy individuals and RA patients.

CD4+CD45RA+ T cells were purified from RA patients or healthy individuals and stimulated for 72 h. (a) AMPK activity was determined by analyzing inhibitory phosphorylation of acetyl-CoA carboxylase (ACC). Protein expression of phospho-ACC was determined by Western blotting. β-actin served as loading control. Representative immunoblots and relative intensities (mean ± SEM) from 5 RA-healthy pairs. Each dot represents the data from one donor. *p < 0.05 by unpaired Mann-Whitney-Wilcoxon rank test. (b) Transcripts for AMPK complex components were analyzed by qPCR. Data are mean ± SEM from 7 RA-healthy pairs.

Supplementary Figure 7 AMPK regulates mTORC1 activation but not its lysosomal localization.

(a) Naïve CD4+ T cells were isolated from healthy individuals or RA patients and stimulated for 72 h. Lysosomes were identified with antibodies to human LAMP1 (red). mTOR was stained with anti-mTOR (green). LAMP1-mTOR co-localization was analyzed by confocal microscopy. Each dot represents one T cell. Scale bars 20 μm. Unpaired Mann-Whitney-Wilcoxon rank test. (b, c) Naïve CD4+ T cells from healthy donors were transfected with AMPKa siRNA or control siRNA and stimulated for 72 h. (b) AMPKα protein expression was analyzed by flow cytometry after 24 hrs. ***p < 0.001 by paired student’s t-test. (c) mTOR expression was quantified in lysosomal fractions and in whole cell lysates. Representative immunoblots from 4 individuals in each group. (d) AMPK-dependent inhibition of mTORC1 activity in T cells. Naïve CD4+ T cells from 6 healthy individuals were stimulated for 72 h in the absence or presence of the AMPK inhibitor Compound C (1 μM). Phospho-S6RP in CD4+ T cells was determined by phosflow cytometry. Paired Mann-Whitney-Wilcoxon rank test. *p < 0.05

Supplementary Figure 8 Anti-inflammatory effects of AMPK activation and mTORC1 inhibition.

(a) The AMPK activator A769662 does not affect Th2 and Treg lineage markers in vivo. NSG mice were implanted with human synovial tissues, reconstituted with CD45RO PBMCs from RA patients, and randomly assigned to control arm (vehicle) or treatment arm (A769662, 30 mg/kg/mouse, twice a day). Synovial grafts were explanted and analyzed for mRNA expression of Th2 and Treg lineage markers (GATA3, IL4, FOXP3) after 7 days of treatment. Data are mean ± SEM from 6 grafts in each group (paired Mann-Whitney-Wilcoxon rank test). (b) The mTORC1 inhibitor rapamycin does not affect Th2 and Treg lineage markers in vivo. NSG mice were implanted with human synovial tissues, reconstituted with CD45RO PBMCs from RA patients, and randomly assigned to control arm (vehicle) and treatment arm (rapamycin, 5 mg/kg/mouse, every other day). Synovial grafts were explanted and analyzed for mRNA expression of Th2 and Treg lineage markers (GATA3, IL4, FOXP3) after 9 days of treatment. Data are mean ± SEM from 6 grafts in each group. Paired Mann-Whitney-Wilcoxon rank test. (c) mTORC1 inhibition abrogates differentiation of RA T cells into pro-inflammatory effector cells. Naïve CD4+ T cells from 6 RA patients were stimulated in the absence or presence of the mTORC1 inhibitor rapamycin (Rapa, 10 μM). Lineage-determining transcription factors and signature cytokines were analyzed. Each dot represents the data from one patient. Paired Mann-Whitney-Wilcoxon rank test. *p< 0.05. (d) Effect of A769662 on CD4+ Treg cell induction in vitro. CD4+CD45RA+ T cells from RA patients were stimulated in the presence of an increasing dose of the AMPK activator A769662 for 4 days and expression of the Treg lineage-determining transcription factor FoxP3 was analyzed by flow cytometry. Data (mean ± SEM) from 6 RA patients. One-way ANOVA and post-ANOVA pair-wise two-group comparisons were conducted with Tukey’s method. ***p < 0.001

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–8, Supplementary Table 1

Reporting Summary

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wen, Z., Jin, K., Shen, Y. et al. N-myristoyltransferase deficiency impairs activation of kinase AMPK and promotes synovial tissue inflammation. Nat Immunol 20, 313–325 (2019). https://doi.org/10.1038/s41590-018-0296-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41590-018-0296-7

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