Immunometabolic profiling of T cells from patients with relapsing-remitting multiple sclerosis reveals an impairment in glycolysis and mitochondrial respiration

Claudia La Rocca, Fortunata Carbone, Veronica De Rosa, Alessandra Colamatteo, Mario Galgani, Francesco Perna, Roberta Lanzillo, Vincenzo Brescia Morra, Giuseppe Orefice, Ilaria Cerillo, Ciro Florio, Giorgia Teresa Maniscalco, Marco Salvetti, Diego Centonze, Antonio Uccelli, Salvatore Longobardi, Andrea Visconti, Giuseppe Matarese, Claudia La Rocca, Fortunata Carbone, Veronica De Rosa, Alessandra Colamatteo, Mario Galgani, Francesco Perna, Roberta Lanzillo, Vincenzo Brescia Morra, Giuseppe Orefice, Ilaria Cerillo, Ciro Florio, Giorgia Teresa Maniscalco, Marco Salvetti, Diego Centonze, Antonio Uccelli, Salvatore Longobardi, Andrea Visconti, Giuseppe Matarese

Abstract

Background: Metabolic reprogramming is shaped to support specific cell functions since cellular metabolism controls the final outcome of immune response. Multiple sclerosis (MS) is an autoimmune disease resulting from loss of immune tolerance against central nervous system (CNS) myelin. Metabolic alterations of T cells occurring during MS are not yet well understood and their studies could have relevance in the comprehension of the pathogenetic events leading to loss of immune tolerance to self and to develop novel therapeutic strategies aimed at limiting MS progression.

Methods and results: In this report, we observed that extracellular acidification rate (ECAR) and oxygen consumption rate (OCR), indicators of glycolysis and oxidative phosphorylation, respectively, were impaired during T cell activation in naïve-to-treatment relapsing remitting (RR)MS patients when compared with healthy controls. These results were also corroborated at biochemical level by a reduced expression of the glycolitic enzymes aldolase, enolase 1, hexokinase I, and by reduction of Krebs cycle enzymes dihydrolipoamide-S-acetyl transferase (DLAT) and dihydrolipoamide-S-succinyl transferase (DLST). Treatment of RRMS patients with interferon beta-1a (IFN beta-1a) was able to restore T cell glycolysis and mitochondrial respiration as well as the amount of the metabolic enzymes to a level comparable to that of healthy controls. These changes associated with an up-regulation of the glucose transporter-1 (GLUT-1), a key element in intracellular transport of glucose.

Conclusions: Our data suggest that T cells from RRMS patients display a reduced engagement of glycolysis and mitochondrial respiration, reversible upon IFN beta-1a treatment, thus suggesting an involvement of an altered metabolism in the pathogenesis of MS.

Keywords: Adipocytokines; IFN beta-1a; Metabolism; Multiple sclerosis.

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

Figures

Fig. 1
Fig. 1
Peripheral immunometabolic signature of RRMS patients before and after IFN beta-1a treatment. The graphs show the number per mm3 of several peripheral blood immune cell subpopulations and the serum concentration of several cytokines implicated in the control of metabolism and immune system functions in healthy controls (n = 24), naїve-to-treatment (n = 15) and IFN beta-1a treated (n = 15) RRMS patients. Data are shown as mean ± s.e.m. Comparisons were evaluated using non-parametric one-way ANOVA test (Kruskal-Wallis test) and Dunn's post-hoc test *P ≤ 0.05, **P ≤ 0.001, ***P ≤ 0.0001.
Fig. 2
Fig. 2
Impaired engagement of glycolysis and mitochondrial respiration of T cells from RRMS is reversed by IFN beta-1a treatment. The graphs represent the metabolic profile of PBMCs isolated from healthy controls, naїve-to-treatment and IFN beta-1a treated RRMS patients. (A) Kinetic profile of ECAR in PBMCs stimulated with OKT3 for 12 h. Healthy controls n = 25, naїve-to-treatment n = 11 and IFN beta-1a treated n = 17 RRMS patients. The data are shown as mean ± s.e.m. ECAR was measured in real time, under basal conditions and in response to glucose, oligomycin and 2-DG. Indices of glycolytic pathway activation, calculated from PBMCs ECAR profile: (B) basal, (C) maximal glycolysis and (D) glycolytic capacity. Data are expressed as mean ± s.e.m. (E) Kinetic profile of OCR in PBMCs stimulated with OKT3 for 12 h. Healthy controls n = 23, naїve-to-treatment n = 11 and IFN beta-1a treated n = 17 RRMS patients. The data are shown as mean ± s.e.m. OCR was measured in real time, under basal conditions and in response to oligomycin, FCCP, Antimycin A and Rotenone. Indices of mitochondrial respiratory function, calculated from PBMCs OCR profile: (F) basal, (G) maximal respiration and (H) spare capacity in PBMCs stimulated with OKT3 for 12 h. Data are expressed as mean ± s.e.m. Comparisons were evaluated using non-parametric one-way ANOVA test (Kruskal-Wallis test) and Dunn's post-hoc test *P ≤ 0.05, **P ≤ 0.001, ***P ≤ 0.0001.
Fig. 3
Fig. 3
Biochemical pathways of CD4+ T cells from RRMS patients before and after IFN beta-1a treatment. (A) Immunoblot for aldolase, hexokinase I, enolase 1, Glut-1, DLAT and DLST on CD4+ T cells from healthy control, naїve-to-treatment and IFN beta-1a treated RRMS patient upon 12 h anti-CD3/CD28 stimulation. Total ERK 1/2 served as a loading control. One representative out of at least three independent experiments. (B) The graphs show the relative densitometric quantitation of aldolase, hexokinase I, enolase 1, Glut-1, DLAT and DLST normalized on total ERK1/2 in healthy controls, naїve -to-treatment and IFN beta-1a treated RRMS patients. One representative experiment is shown, from one individual for each condition; in total we runned al least 3 healthy controls, 3 naive-to-treatment and 3 RRMS IFN beta-1a treated patients. We scanned 3 times at least three films with different exposures from each subject, and averaged values were utilized as densitometry to reduce variations among samples. Comparisons were evaluated using non-parametric one-way ANOVA test (Kruskal-Wallis test) and Dunn's post-hoc test *P ≤ 0.05, **P ≤ 0.001, ***P ≤ 0.0001.

References

    1. Compston A., Coles A. Multiple sclerosis. Lancet. 2008;372:1502–1517.
    1. Costantino C.M., Baecher-Allan C., Hafler D.A. Multiple sclerosis and regulatory T cells. J Clin Immunol. 2008;28:697–706.
    1. Vukmanovic-Stejic M., Zhang Y., Cook J.E., Fletcher J.M., McQuaid A., Masters J.E. Human CD4 + CD25hi Foxp3 + regulatory T cells are derived by rapid turnover of memory populations in vivo. J Clin Invest. 2006;116:2423–2433.
    1. Vukmanovic-Stejic M., Agius E., Booth N., Dunne P.J., Lacy K.E., Reed J.R. The kinetics of CD4 + Foxp3 + T cell accumulation during a human cutaneous antigen-specific memory response in vivo. J Clin Invest. 2008;118:3639–3650.
    1. De Rosa V., Galgani M., Porcellini A., Colamatteo A., Santopaolo M., Zuchegna C. Glycolysis controls the induction of human regulatory T cells by modulating the expression of FOXP3 exon 2 splicing variants. Nat Immunol. 2015;16:1174–1184.
    1. Carrieri P.B., Carbone F., Perna F., Bruzzese D., La Rocca C., Galgani M. Longitudinal assessment of immuno-metabolic parameters in multiple sclerosis patients during treatment with glatiramer acetate. Metabolism. 2015;64:1112–1121.
    1. Cantarini L., Pucino V., Vitale A., Talarico R., Lucherini O.M., Magnotti F. Immunometabolic biomarkers of inflammation in Behçet's disease: relationship with epidemiological profile, disease activity and therapeutic regimens. Clin Exp Immunol. 2016;184:197–207.
    1. Galgani M., Nugnes R., Bruzzese D., Perna F., De Rosa V., Procaccini C. Meta-immunological profiling of children with type 1 diabetes identifies new biomarkers to monitor disease progression. Diabetes. 2013;62:2481–2491.
    1. Markowitz C.E. Interferon-beta: mechanism of action and dosing issues. Neurology. 2007;68(24 Suppl. 4):S8–11. [Review]
    1. Pearce E.L., Poffenberger M.C., Chang C.H., Jones R.G. Fueling immunity: insights into metabolism and lymphocyte function. Science. 2013;342:1242454.
    1. Michalek R.D., Gerriets V.A., Jacobs S.R., Macintyre A.N., MacIver N.J., Mason E.F. Cutting edge: distinct glycolytic and lipid oxidative metabolic programs are essential for effector and regulatory CD4+ T cell subsets. J Immunol. 2011;186:3299–3303.
    1. Bettencourt I.A., Powell J.D. Targeting metabolism as a novel therapeutic approach to autoimmunity, inflammation, and transplantation. J Immunol. 2017;198:999–1005.
    1. Okuda Y., Okuda M., Apatoff B.R., Posnett D.N. The activation of memory CD4(+) T cells and CD8(+) T cells in patients with multiple sclerosis. J Neurol Sci. 2005;235:11–17.
    1. Praksova P., Stourac P., Bednarik J., Vlckova E., Mikulkova Z., Michalek J. Immunoregulatory T cells in multiple sclerosis and the effect of interferon beta and glatiramer acetate treatment on T cell subpopulations. J Neurol Sci. 2012;319:18–23.
    1. Matarese G., Carrieri P.B., La Cava A., Perna F., Sanna V., De Rosa V. Leptin increase in multiple sclerosis associates with reduced number of CD4(+)CD25 + regulatory T cells. Proc Natl Acad Sci U S A. 2005;102:5150–5155.
    1. Zastrow O., Seidel B., Kiess W., Thiery J., Keller E., Böttner A. The soluble leptin receptor is crucial for leptin action: evidence from clinical and experimental data. Int J Obes Relat Metab Disord. 2003;27:1472–1478.
    1. LaceyDL TimmsE, TanHL Kelley M.J., Dunstan C.R., Burgess T. Osteoprotegerin ligand is a cytokine that regulates osteoclast differentiation and activation. Cell. 1998;93:165–176.
    1. Browner W.S., Lui L.Y., Cummings S.R. Associations of serum osteoprotegerin levels with diabetes, stroke, bone density, fractures, and mortality in elderly women. J Clin Endocrinol Metab. 2001;86:631–637.
    1. Hofbauer L.C., Cepok S., Hemmer B. Osteoprotegerin is highly expressed in the spinal cord and cerebrospinal fluid. Acta Neuropathol. 2004;107:575–577.
    1. Gerritse K., Laman J.D., Noelle R.J., Aruffo A., Ledbetter J.A., Boersma W.J. CD40-CD40 Ligand interactions in experimental allergic encephalomyelitis and multiple sclerosis. Proc Natl Acad Sci U S A. 1996;93:2499–2504.
    1. Munroe M.E., Arbiser J.L., Bishop G.A. Honokiol, a natural plant product, inhibits inflammatory signals and alleviates inflammatory arthritis. J Immunol. 2007;179:753–763.
    1. Sosvorova L., Kanceva R., Vcelak J., Kancheva L., Mohapl M., Starka L. The comparison of selected cerebrospinal fluid and serum cytokine levels in patients with multiple sclerosis and normal pressure hydrocephalus. Neuro Endocrinol Lett. 2015;36:564–571.
    1. Huang W.X., Huang P., Hillert J. Systemic upregulation of CD40 and CD40 ligand mRNA expression in multiple sclerosis. Mult Scler. 2000;6:61–65.
    1. Carbone F., De Rosa V., Carrieri P.B., Montella S., Bruzzese D., Porcellini A. Regulatory T cell proliferative potential is impaired in human autoimmune disease. Nat Med. 2014;20:69–74.
    1. Procaccini C., Carbone F., Di Silvestre D., Brambilla F., De Rosa V., Galgani M. The proteomic landscape of human ex vivo regulatory and conventional T cells reveals specific metabolic requirements. Immunity. 2016;44:406–421.
    1. Gerriets V.A., Kishton R.J., Johnson M.O., Cohen S., Siska P.J., Nichols A.G. Foxp3 And toll-like receptor signaling balance T(reg) cell anabolic metabolism for suppression. Nat Immunol. 2016;17:1459–1466.
    1. 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. 2013;210:2119–2134.
    1. De Riccardis L., Rizzello A., Ferramosca A., Urso E., De Robertis F., Danieli A. Bioenergetics profile of CD4(+) T cells in relapsing remitting multiple sclerosis subjects. J Biotechnol. 2015;202:31–39.
    1. Burke J.D., Platanias L.C., Fish E.N. Beta interferon regulation of glucose metabolism is PI3K/Akt dependent and important for antiviral activity against coxsackievirus B3. J Virol. 2014;88:3485–3495.
    1. Negrotto L., Farez M.F., Correale J. Immunologic effects of metformin and pioglitazone treatment on metabolic syndrome and multiple sclerosis. JAMA Neurol. 2016;73:520–528.
    1. Choi I.Y., Piccio L., Childress P., Bollman B., Ghosh A., Brandhorst S. A diet mimicking fasting promotes regeneration and reduces autoimmunity and multiple sclerosis symptoms. Cell Rep. 2016;15:2136–2146.
    1. Choi I.Y., Lee C., Longo V.D. Nutrition and fasting mimicking diets in the prevention and treatment of autoimmune diseases and immunosenescence. Mol Cell Endocrinol. 2017;7207:30055-2.
    1. Feinstein D.L., Galea E., Gavrilyuk V., Brosnan C.F., Whitacre C.C., Dumitrescu-Ozimek L. Peroxisome proliferator-activated receptor-gamma agonists prevent experimental autoimmune encephalomyelitis. Ann Neurol. 2002;51:694–702.
    1. Sun Y., Tian T., Gao J., Liu X., Hou H., Cao R. Metformin ameliorates the development of experimental autoimmune encephalomyelitis by regulating T helper 17 and regulatory T cells in mice. J Neuroimmunol. 2016;292:58–67.
    1. Mokry L.E., Ross S., Timpson N.J., Sawcer S., Davey Smith G., Richards J.B. Obesity and multiple sclerosis: a Mendelian randomization study. PLoS Med. 2016;13:e1002053.
    1. Chitnis T., Grag J., Weinstock-Guttman B., Belman A., Olsen C., Misra M. Distinct effects of obesity and puberty on risk and age at onset of pediatric MS. Ann Clin Transl Neurol. 2016;4:897–907.
    1. Hedström A.K., Lima Bomfim I., Barcellos L., Gianfrancesco M., Schaefer C., Kockum I. Interaction between adolescent obesity and HLA risk genes in the etiology of multiple sclerosis. Neurology. 2014;82:865–872.
    1. Hedström A.K., Olsson T., Alfredsson L. Body mass index during adolescence, rather than childhood, is critical in determining MS risk. Mult Scler. 2016;22:878–883.

Source: PubMed

3
Subscribe