Obesity-Induced Metabolic Stress Leads to Biased Effector Memory CD4+ T Cell Differentiation via PI3K p110δ-Akt-Mediated Signals

Claudio Mauro, Joanne Smith, Danilo Cucchi, David Coe, Hongmei Fu, Fabrizia Bonacina, Andrea Baragetti, Gaia Cermenati, Donatella Caruso, Nico Mitro, Alberico L Catapano, Enrico Ammirati, Maria P Longhi, Klaus Okkenhaug, Giuseppe D Norata, Federica M Marelli-Berg, Claudio Mauro, Joanne Smith, Danilo Cucchi, David Coe, Hongmei Fu, Fabrizia Bonacina, Andrea Baragetti, Gaia Cermenati, Donatella Caruso, Nico Mitro, Alberico L Catapano, Enrico Ammirati, Maria P Longhi, Klaus Okkenhaug, Giuseppe D Norata, Federica M Marelli-Berg

Abstract

Low-grade systemic inflammation associated to obesity leads to cardiovascular complications, caused partly by infiltration of adipose and vascular tissue by effector T cells. The signals leading to T cell differentiation and tissue infiltration during obesity are poorly understood. We tested whether saturated fatty acid-induced metabolic stress affects differentiation and trafficking patterns of CD4+ T cells. Memory CD4+ T cells primed in high-fat diet-fed donors preferentially migrated to non-lymphoid, inflammatory sites, independent of the metabolic status of the hosts. This was due to biased CD4+ T cell differentiation into CD44hi-CCR7lo-CD62Llo-CXCR3+-LFA1+ effector memory-like T cells upon priming in high-fat diet-fed animals. Similar phenotype was observed in obese subjects in a cohort of free-living people. This developmental bias was independent of any crosstalk between CD4+ T cells and dendritic cells and was mediated via direct exposure of CD4+ T cells to palmitate, leading to increased activation of a PI3K p110δ-Akt-dependent pathway upon priming.

Keywords: Akt; CD4; T lymphocyte; differentiation; effector memory; high-fat diet; inflammation; obesity; palmitate; saturated fatty acid.

Copyright © 2017 The Author(s). Published by Elsevier Inc. All rights reserved.

Figures

Graphical abstract
Graphical abstract
Figure 1
Figure 1
Priming following HFD Induces Differentiation of a CD4+ T Cell Population that Readily Migrates to Inflamed, Non-lymphoid Tissues Independent of the Metabolic Status of the Host In vivo-primed CD4+ T cells isolated from pooled lymph nodes of HFD or CD Marilyn female donor mice, fluorescently labeled with CFSE or PKH26, respectively, in different tissue compartments (as described) of HFD or CD C57BL/6 male recipient mice 48 hr post-i.v. injection of donor cells. The male recipients were previously i.p. injected with zymosan (1 mg/mouse) and IFNγ (600 U/mouse). (A) Dot plots and quantification of donor cells in the blood circulation of CD recipient mice; 2 hr, 24 hr, and 48 hr post-i.v. injection of donor cells. (B–D) Dot plots and quantification of the localization of donor cells, fluorescently labeled with CFSE (HFD) or PKH26 (CD) in the peritoneal cavity (B), mesenteric lymph nodes (C), and spleen (D) of HFD or CD recipient mice. (E) Survival curve of C57BL/6 male skin grafts on HFD or CD Marilyn female recipient mice up to 28 days post-transplantation. (A–D) n = 4–6 independent recipients. (E) n = 7 HFD, n = 9 CD independent recipients. (A–D) The values denote mean ± SEM. ∗p < 0.05; ∗∗p < 0.01; and ∗∗∗p < 0.001. Mantel-Cox test, ∗p < 0.05.
Figure 2
Figure 2
HFD-Induced Metabolic Stress Promotes the Differentiation of Primed CD4+ T Cells to a CXCR3+-LFA1+ Effector Memory-like Phenotype in Mice (A–F) Cell surface staining of CCR7 (A), CD62L (B), CXCR3 (C), LFA1 (D), CD25 (E), and CD44 (F) in in vivo-primed CD4+ T cells isolated from pooled lymph nodes of the HFD or CD Marilyn female donor mice used in Figures 1A–1D (absolute numbers). (G) Ex vivo chemotaxis (3 hr, 6 hr, and 24 hr time points) of in vivo-primed CD4+ T cells isolated from pooled lymph nodes of the HFD or CD Marilyn female donor mice used in Figures 1A–1D in response to CXCL10 (300 ng/mL), CCL19/21 (200 ng/mL of each chemokine), or media only. (H and I) Intracellular staining of IFNγ (H) and IL4 (I) in in vivo-primed CD4+ T cells isolated from mesenteric lymph nodes of HFD or CD mice. (A–F) n = 3–6 independent donors. (G) n = 3 independent donors (each donor was tested in duplicate). (H and I) n = 5 independent mice. The values denote mean ± SEM. ∗p < 0.05; ∗∗p < 0.01; and ∗∗∗p < 0.001.
Figure 3
Figure 3
Obesity and Body Fat Distribution in Humans Positively Associates with an Increase in CXCR3+ Effector Memory T Cells (A–D) Number of total lymphocytes (A) and % of helper, naïve, and effector memory T cells (B–D) purified from the peripheral blood of lean, overweight, and obese subjects stratified according to BMI. (E) % of effector memory T cells purified from the peripheral blood of obese subjects stratified according to BMI either + (H) T cells with android/gynoid ratio. (A) n = 1,172. (B–D and F–H) n = 187. (E) n = 21 BMI ∗p < 0.05.
Figure 4
Figure 4
Antigen Presentation by DC from Metabolically Stressed Donors Does Not Affect T Cell Differentiation to a Pro-inflammatory Effector Memory-like Phenotype (A) Dot plots and quantification of in vitro proliferation of CFSE-labeled CD4+ T cells isolated from pooled lymph nodes of HFD or CD Marilyn female mice incubated for 3 days with CD11c+ DC isolated from the spleen of HFD or CD C57BL/6 male mice. The undivided and fourth division populations are quantified by dilution of the CFSE-label. (B) Cell surface staining of CD44, CCR7, CD62L, and CXCR3 in the undivided population of CFSE-labeled CD4+ T cells shown in (A). (C) Cell surface staining of CD40, CD80, and CD86 in CD11c+ DC used in (A). (A and B) n = 3 independent mice (each mouse was tested in triplicate). (C) n = 3 independent mice. The values denote mean ± SEM. ∗∗p < 0.01 and ∗∗∗p < 0.001.
Figure 5
Figure 5
Direct Exposure of CD4+ T Cells to Saturated FA Causes Enhanced Activation of a PI3K p110δ-Akt-Dependent Pathway upon Priming (A) Levels and densitometric quantification of p-Akt (S473), Akt, p-S6 (S235/236), and β-actin protein expression in in vivo-primed CD4+ T cells isolated from mesenteric lymph nodes of CD and HFD mice i.p. injected with the PI3K inhibitor IC87114 or left untreated. (B) Levels and densitometric quantification of p-Akt (S473) and Akt protein expression in CD4+ T cells purified from the peripheral blood of lean and obese human subjects. (C) Membrane fluidity index in in vivo-primed CD4+ T cells isolated from mesenteric lymph nodes of CD and HFD mice, calculated as the ratio of oleic acid (C18:1) to linoleic acid (C18:2). (D) Cell surface staining of CTxB in in vitro activated (Act) CD4+ T cells isolated from the lymph nodes of CD and HFD mice. The quantification of aggregation of CTxB is measured by bright detail intensity of the signal. (E–H) Cell surface staining of CD44 with CCR7 (E), CD62L (F), CXCR3 (G), and LFA1 (H) in in vivo-primed CD4+ T cells isolated from mesenteric lymph nodes of PED and PCD mice. (I and J) Protein levels and densitometric quantification of p-Akt (S473), Akt, p-S6 (S235/236), and S6 protein expression (I), and gene expression of CCR7 and CD62L (J) in CD4+ T cells isolated from pooled lymph nodes of mice and cultured overnight with IL-7 (1 ng/mL) in the presence or absence of palmitic, linoleic, or stearic acid (200 μM), followed by activation with plate bound anti-CD3 (0.5 μg/mL) and anti-CD28 (2.5 μg/mL) for the indicated time points. (K and L) Levels of CCR7 and CD62L gene expression (K) and CCR7 protein (L) in CD4+ T cells isolated from pooled lymph nodes of mice, then activated with plate bound anti-CD3 (0.5 μg/mL) and anti-CD28 (2.5 μg/mL) for the indicated time points in the presence or absence of the Akt activator SC79 (500 nM). (A and B) Each lane shows data from independent mouse/human samples. (C–H) n = 3–6 independent mice. (I) Representative data of n = 3 independent mice. (J–L) 3–6 pooled mice. The values denote mean ± SEM. ∗p < 0.05; ∗∗p < 0.01; and ∗∗∗p < 0.001.
Figure 6
Figure 6
The PI3K, Akt Pathway Is Key for the Differentiation of CXCR3+ Effector Memory T Cells (A–D) Cell surface staining of CD44 with CCR7 (A), CD62L (B), CXCR3 (C), and LFA1 (D) in in vivo-primed CD4+ T cells isolated from mesenteric lymph nodes of HFD P110δD910A or WT C57BL/6 mice. (E) Levels and densitometric quantification of p-Akt (S473), Akt, p-S6 (S235/236), and S6 protein expression in in vivo-primed CD4+ T cells from (A)–(D). (F) Ex vivo chemotaxis (3 hr, 6 hr, and 24 hr time points) of in vivo-primed CD4+ T cells isolated from mesenteric lymph nodes of HFD or CD OT-II P110δD910A mice in response to CXCL10 (300 ng/mL), CCL19/21 (200 ng/mL of each chemokine), or media only. (G–J) Cell surface staining of CD44 with CCR7 (G), CD62L (H), CXCR3 (I), and LFA1 (J) in in vivo-primed CD4+ T cells isolated from mesenteric lymph nodes of CD and HFD mice i.p. injected with the PI3K inhibitor IC87114, the Akt activator SC79, or left untreated. (A–E and G–J) n = 5 independent mice. (F) n = 4 independent mice (each mouse was tested in duplicate). The values denote mean ± SEM. ∗p < 0.05; ∗∗p < 0.01; and ∗∗∗p < 0.001.
Figure 7
Figure 7
Inhibition of FAO Prevents Enhanced Effector Memory Differentiation in Saturated Fatty Acid-Enriched Diet (A) OCR of naive CD4+ T cells isolated from pooled lymph nodes of mice and cultured overnight with IL-7 (1 ng/mL) in the presence or absence of palmitic or stearic acid (200 μM). (B–E) Cell surface staining of CD44 with CCR7 (B), CD62L (C), CXCR3 (D), and LFA1 (E) in in vivo-primed CD4+ T cells isolated from mesenteric lymph nodes of PED or PCD mice i.p. injected with etomoxir or left untreated. (A) Six pooled mice tested in 6–10 replicates per treatment. (B–E) n = 5 independent mice. The values denote mean ± SEM. ∗p < 0.05 and ∗∗p < 0.01.

References

    1. Ammirati E., Cianflone D., Banfi M., Vecchio V., Palini A., De Metrio M., Marenzi G., Panciroli C., Tumminello G., Anzuini A. Circulating CD4+CD25hiCD127lo regulatory T-Cell levels do not reflect the extent or severity of carotid and coronary atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 2010;30:1832–1841.
    1. Ammirati E., Cianflone D., Vecchio V., Banfi M., Vermi A.C., De Metrio M., Grigore L., Pellegatta F., Pirillo A., Garlaschelli K. Effector memory T cells are associated with atherosclerosis in humans and animal models. J. Am. Heart Assoc. 2012;1:27–41.
    1. Baragetti A., Norata G.D., Sarcina C., Rastelli F., Grigore L., Garlaschelli K., Uboldi P., Baragetti I., Pozzi C., Catapano A.L. High density lipoprotein cholesterol levels are an independent predictor of the progression of chronic kidney disease. J. Intern. Med. 2013;274:252–262.
    1. Bjørndal B., Burri L., Staalesen V., Skorve J., Berge R.K. Different adipose depots: their role in the development of metabolic syndrome and mitochondrial response to hypolipidemic agents. J. Obes. 2011;2011:490650.
    1. Chen Y., Demir Y., Valujskikh A., Heeger P.S. The male minor transplantation antigen preferentially activates recipient CD4+ T cells through the indirect presentation pathway in vivo. J. Immunol. 2003;171:6510–6518.
    1. Chen Y., Tian J., Tian X., Tang X., Rui K., Tong J., Lu L., Xu H., Wang S. Adipose tissue dendritic cells enhances inflammation by prompting the generation of Th17 cells. PLoS ONE. 2014;9:e92450.
    1. Cildir G., Akıncılar S.C., Tergaonkar V. Chronic adipose tissue inflammation: all immune cells on the stage. Trends Mol. Med. 2013;19:487–500.
    1. Cole M.A., Murray A.J., Cochlin L.E., Heather L.C., McAleese S., Knight N.S., Sutton E., Jamil A.A., Parassol N., Clarke K. A high fat diet increases mitochondrial fatty acid oxidation and uncoupling to decrease efficiency in rat heart. Basic Res. Cardiol. 2011;106:447–457.
    1. Eguchi K., Manabe I., Oishi-Tanaka Y., Ohsugi M., Kono N., Ogata F., Yagi N., Ohto U., Kimoto M., Miyake K. Saturated fatty acid and TLR signaling link β cell dysfunction and islet inflammation. Cell Metab. 2012;15:518–533.
    1. Endo Y., Asou H.K., Matsugae N., Hirahara K., Shinoda K., Tumes D.J., Tokuyama H., Yokote K., Nakayama T. Obesity drives Th17 cell differentiation by inducing the lipid metabolic kinase, ACC1. Cell Rep. 2015;12:1042–1055.
    1. Faulds M.H., Dahlman-Wright K. Metabolic diseases and cancer risk. Curr. Opin. Oncol. 2012;24:58–61.
    1. Finlay D.K., Rosenzweig E., Sinclair L.V., Feijoo-Carnero C., Hukelmann J.L., Rolf J., Panteleyev A.A., Okkenhaug K., Cantrell D.A. PDK1 regulation of mTOR and hypoxia-inducible factor 1 integrate metabolism and migration of CD8+ T cells. J. Exp. Med. 2012;209:2441–2453.
    1. Fu H., Kishore M., Gittens B., Wang G., Coe D., Komarowska I., Infante E., Ridley A.J., Cooper D., Perretti M., Marelli-Berg F.M. Self-recognition of the endothelium enables regulatory T-cell trafficking and defines the kinetics of immune regulation. Nat. Commun. 2014;5:3436.
    1. Gerriets V.A., Rathmell J.C. Metabolic pathways in T cell fate and function. Trends Immunol. 2012;33:168–173.
    1. Gupta S., Pablo A.M., Jiang Xc., Wang N., Tall A.R., Schindler C. IFN-gamma potentiates atherosclerosis in ApoE knock-out mice. J. Clin. Invest. 1997;99:2752–2761.
    1. Haghikia A., Jörg S., Duscha A., Berg J., Manzel A., Waschbisch A., Hammer A., Lee D.H., May C., Wilck N. Dietary fatty acids directly impact central nervous system autoimmunity via the small intestine. Immunity. 2015;43:817–829.
    1. Hamann A., Klugewitz K., Austrup F., Jablonski-Westrich D. Activation induces rapid and profound alterations in the trafficking of T cells. Eur. J. Immunol. 2000;30:3207–3218.
    1. Karlsson E.A., Sheridan P.A., Beck M.A. Diet-induced obesity impairs the T cell memory response to influenza virus infection. J. Immunol. 2010;184:3127–3133.
    1. Lantz O., Grandjean I., Matzinger P., Di Santo J.P. Gamma chain required for naïve CD4+ T cell survival but not for antigen proliferation. Nat. Immunol. 2000;1:54–58.
    1. Larbi A., Franceschi C., Mazzatti D., Solana R., Wikby A., Pawelec G. Aging of the immune system as a prognostic factor for human longevity. Physiology (Bethesda) 2008;23:64–74.
    1. Lindholm C.R., Ertel R.L., Bauwens J.D., Schmuck E.G., Mulligan J.D., Saupe K.W. A high-fat diet decreases AMPK activity in multiple tissues in the absence of hyperglycemia or systemic inflammation in rats. J. Physiol. Biochem. 2013;69:165–175.
    1. Liu Y., Wan Q., Guan Q., Gao L., Zhao J. High-fat diet feeding impairs both the expression and activity of AMPKa in rats’ skeletal muscle. Biochem. Biophys. Res. Commun. 2006;339:701–707.
    1. Macintyre A.N., Finlay D., Preston G., Sinclair L.V., Waugh C.M., Tamas P., Feijoo C., Okkenhaug K., Cantrell D.A. Protein kinase B controls transcriptional programs that direct cytotoxic T cell fate but is dispensable for T cell metabolism. Immunity. 2011;34:224–236.
    1. Mathis D., Shoelson S.E. Immunometabolism: an emerging frontier. Nat. Rev. Immunol. 2011;11:81.
    1. Mauro C., Marelli-Berg F.M. T cell immunity and cardiovascular metabolic disorders: does metabolism fuel inflammation? Front. Immunol. 2012;3:173.
    1. Mehta M.M., Chandel N.S. Targeting metabolism for lupus therapy. Sci. Transl. Med. 2015;7:274fs5.
    1. Molinero L.L., Yin D., Lei Y.M., Chen L., Wang Y., Chong A.S., Alegre M.L. High-fat diet-induced obesity enhances allograft rejection. Transplantation. 2016;100:1015–1021.
    1. Moreira J.B., Wohlwend M., Alves M.N., Wisløff U., Bye A. A small molecule activator of AKT does not reduce ischemic injury of the rat heart. J. Transl. Med. 2015;13:76.
    1. Norata G.D., Garlaschelli K., Ongari M., Raselli S., Grigore L., Catapano A.L. Effects of fractalkine receptor variants on common carotid artery intima-media thickness. Stroke. 2006;37:1558–1561.
    1. Norata G.D., Raselli S., Grigore L., Garlaschelli K., Dozio E., Magni P., Catapano A.L. Leptin:adiponectin ratio is an independent predictor of intima media thickness of the common carotid artery. Stroke. 2007;38:2844–2846.
    1. Norata G.D., Caligiuri G., Chavakis T., Matarese G., Netea M.G., Nicoletti A., O’Neill L.A., Marelli-Berg F.M. The cellular and molecular basis of translational immunometabolism. Immunity. 2015;43:421–434.
    1. O’Sullivan D., van der Windt G.J., Huang S.C., Curtis J.D., Chang C.H., Buck M.D., Qiu J., Smith A.M., Lam W.Y., DiPlato L.M. Memory CD8(+) T cells use cell-intrinsic lipolysis to support the metabolic programming necessary for development. Immunity. 2014;41:75–88.
    1. Okkenhaug K., Bilancio A., Farjot G., Priddle H., Sancho S., Peskett E., Pearce W., Meek S.E., Salpekar A., Waterfield M.D. Impaired B and T cell antigen receptor signaling in p110delta PI 3-kinase mutant mice. Science. 2002;297:1031–1034.
    1. Reynolds C.M., McGillicuddy F.C., Harford K.A., Finucane O.M., Mills K.H., Roche H.M. Dietary saturated fatty acids prime the NLRP3 inflammasome via TLR4 in dendritic cells-implications for diet-induced insulin resistance. Mol. Nutr. Food Res. 2012;56:1212–1222.
    1. Robertson N.J., Chai J.G., Millrain M., Scott D., Hashim F., Manktelow E., Lemonnier F., Simpson E., Dyson J. Natural regulation of immunity to minor histocompatibility antigens. J. Immunol. 2007;178:3558–3565.
    1. Rocha V.Z., Folco E.J., Sukhova G., Shimizu K., Gotsman I., Vernon A.H., Libby P. Interferon-gamma, a Th1 cytokine, regulates fat inflammation: a role for adaptive immunity in obesity. Circ. Res. 2008;103:467–476.
    1. Rocha V.Z., Folco E.J., Ozdemir C., Sheikine Y., Christen T., Sukhova G.K., Tang E.H., Bittencourt M.S., Santos R.D., Luster A.D. CXCR3 controls T-cell accumulation in fat inflammation. Arterioscler. Thromb. Vasc. Biol. 2014;34:1374–1381.
    1. Schrauwen P., Wagenmakers A.J., van Marken Lichtenbelt W.D., Saris W.H., Westerterp K.R. Increase in fat oxidation on a high-fat diet is accompanied by an increase in triglyceride-derived fatty acid oxidation. Diabetes. 2000;49:640–646.
    1. Shamshiev A.T., Ampenberger F., Ernst B., Rohrer L., Marsland B.J., Kopf M. Dyslipidemia inhibits Toll-like receptor-induced activation of CD8alpha-negative dendritic cells and protective Th1 type immunity. J. Exp. Med. 2007;204:441–452.
    1. Shaw A.C., Joshi S., Greenwood H., Panda A., Lord J.M. Aging of the innate immune system. Curr. Opin. Immunol. 2010;22:507–513.
    1. Shriver L.P., Manchester M. Inhibition of fatty acid metabolism ameliorates disease activity in an animal model of multiple sclerosis. Sci. Rep. 2011;1:79.
    1. Swamy M., Beck-Garcia K., Beck-Garcia E., Hartl F.A., Morath A., Yousefi O.S., Dopfer E.P., Molnár E., Schulze A.K., Blanco R. A cholesterol-based allostery model of T cell receptor phosphorylation. Immunity. 2016;44:1091–1101.
    1. Veillard N.R., Steffens S., Pelli G., Lu B., Kwak B.R., Gerard C., Charo I.F., Mach F. Differential influence of chemokine receptors CCR2 and CXCR3 in development of atherosclerosis in vivo. Circulation. 2005;112:870–878.
    1. Viola A., Schroeder S., Sakakibara Y., Lanzavecchia A. T lymphocyte costimulation mediated by reorganization of membrane microdomains. Science. 1999;283:680–682.
    1. Wen H., Gris D., Lei Y., Jha S., Zhang L., Huang M.T., Brickey W.J., Ting J.P. Fatty acid-induced NLRP3-ASC inflammasome activation interferes with insulin signaling. Nat. Immunol. 2011;12:408–415.
    1. Yin Y., Choi S.C., Xu Z., Perry D.J., Seay H., Croker B.P., Sobel E.S., Brusko T.M., Morel L. Normalization of CD4+ T cell metabolism reverses lupus. Sci. Transl. Med. 2015;7:274ra18.
    1. Ying H., Fu H., Rose M.L., McCormack A.M., Sarathchandra P., Okkenhaug K., Marelli-Berg F.M. Genetic or pharmaceutical blockade of phosphoinositide 3-kinase p110δ prevents chronic rejection of heart allografts. PLoS ONE. 2012;7:e32892.

Source: PubMed

3
订阅