Mitochondria as central hub of the immune system

Cristiane Naffah de Souza Breda, Gustavo Gastão Davanzo, Paulo José Basso, Niels Olsen Saraiva Câmara, Pedro Manoel Mendes Moraes-Vieira, Cristiane Naffah de Souza Breda, Gustavo Gastão Davanzo, Paulo José Basso, Niels Olsen Saraiva Câmara, Pedro Manoel Mendes Moraes-Vieira

Abstract

Nearly 130 years after the first insights into the existence of mitochondria, new rolesassociated with these organelles continue to emerge. As essential hubs that dictate cell fate, mitochondria integrate cell physiology, signaling pathways and metabolism. Thus, recent research has focused on understanding how these multifaceted functions can be used to improve inflammatory responses and prevent cellular dysfunction. Here, we describe the role of mitochondria on the development and function of immune cells, highlighting metabolic aspects and pointing out some metabolic- independent features of mitochondria that sustain cell function.

Keywords: And immune cells; Cell fate; Immunometabolism; Mitochondrial function.

Copyright © 2019 The Authors. Published by Elsevier B.V. All rights reserved.

Figures

Fig. 1
Fig. 1
Immature and mature neutrophils have different metabolic requirements. A. In the presence of normal glucose levels, immature neutrophils undergo autophagy to a proper maturation through FAO associated with OXPHOS. In the same way, during a low glucose levels context (tumoral environment), immature neutrophils prioritize FAO associated with OXPHOS to increase NADPH levels and consequently oxidants production. Excess of oxidants leads to T cell inhibition, contributing to tumor growth. B. Mature neutrophils prioritize aerobic glycolysis and generate NADPH from PPP. NADPH is the Nox2 substrate to produce oxidants inside the phagosome and crucial to eliminate phagocyted pathogens. Oxidants from Nox2 are also important to NETs formation from genomic DNA while mtROS are important to mitochondrial NETs production. Intracellular Ca2+ levels stimulate mitochondrial ATP production by the mTOR pathway. The ATP acts in an autocrine way by binding the P2Y2 receptors in the neutrophils stimulating intracellular Ca2+ mobilization and chemotaxis. Mitochondrial fission also contributes to neutrophil chemotaxis.
Fig. 2
Fig. 2
Metabolism of macrophages. A. Under LPS and IFN-γ stimuli, macrophages preferentially use glucose to produce lactate. In the mitochondria, citrate can be exported and used to produce fatty acids and itaconate. The mitochondria increase the production of oxidants and succinate, which are capable to stabilize the transcription factor HIF-1α in the cytosol, triggering the transcription of HIF-1α-target genes. The M1 macrophage is characterized by mitochondria fission. B. Metabolism of M2 macrophages. Alternatively activated macrophages (M2) use glucose and glutamine to feed the TCA cycle and produce ATP by oxidative phosphorylation. These cells are known to have elongated mitochondria, and consequently more efficient energy production. The use of fatty acid oxidation and glycolysis by M2 macrophages are under review in the literature.
Fig. 3
Fig. 3
Metabolism of dendritic cells and T lymphocytes. Inflammatory signals drive glycolysis, increasing glucose uptake and glycogenosis in DCs. The conversion from glucose to pyruvate increases citrate levels in the mitochondria, which can be exported to the cytosol. The PPP generates NADPH, used in the FAS. FAS are used to generate membranes to the endoplasmic reticulum and Golgi complex, in an NADPH-dependent manner, both necessary to increase protein synthesis. Dendritic cells process antigens and present them using the MHC at the cell surface where they can be recognized by the TCR. This event engages ROS production, which facilitates the activation of nuclear factor of activated T cell (NFAT) activation. This transcription factor induces IL-2, a T cell growth cytokine.
Fig. 4
Fig. 4
Essential metabolism and mitochondrial regulation on B cells (A) and Antibody-Secreting Cells (B). A. B cells are activated through T cell-independent or -dependent mechanisms. B cell activation increases both aerobic glycolysis and OXPHOS rates, enhancing the expression of the glucose transporter Glut1 and mitochondrial mass. Part of the TCA cycle fueling is used to de novo lipid biosynthesis via citrate transport from the mitochondria into the cytoplasm. B cell activation promotes both oxidants and antioxidants production, which the balance regulates B cell survival, effector functions and cell fate. Higher mtROS levels are important for sustaining both mitochondrial activity and mass, activating the CSR. On the other hand, lower mtROS levels were directly associated with plasma cell differentiation. Moreover, B cell activation also increases the mitochondria number and change their anatomy from an elongated to a rounded shape. Recently, CpG-stimulated TLR9 in B cells causes mtDNA releasing into the extracellular environment and its associated in stimulating anti-viral responses via type I IFN production. B. ASCs have high metabolic demand in order to produce antibodies. Part of the glucose is used to glycosylate antibodies and intracellular membrane expansion. Higher antioxidant activity is associated with increased IgM secretion by ASCs. Different subtypes of plasma cells show different metabolic profiles. LLPCs show increased glucose uptake, CD98 amino acid transporter expression and autophagosome mass compared to SLPCs. a.a: amino acids. ASC: Antibody-Secreting Cell, BCR, B Cell Receptor, CSR, Class-Switch Recombination, FAS, Fatty Acid Synthesis, GSH, Glutathione, LLPCs, Long-Lived Plasma Cells, OXPHOS, Oxidative Phosphorylation, PPP, Pentose Phosphate Pathway, SLPCs, Short-Lived Plasma Cells, TCR, T Cell Receptor, TLR, Toll-Like Receptor, TCA, Tricarboxylic Acid Cycle, xCT, Cystine/glutamate transporter.

References

    1. Alwarawrah Y., Kiernan K., MacIver N.J. Changes in nutritional status impact immune cell metabolism and function. Front. Immunol. 2018;9:1055.
    1. Andersen C.J., Murphy K.E., Fernandez M.L. Impact of obesity and metabolic syndrome on immunity. Adv. Nutr. 2016;7(1):66–75.
    1. Chang C.H., Qiu J., O'Sullivan D., Buck M.D., Noguchi T., Curtis J.D. Metabolic competition in the tumor microenvironment is a driver of cancer progression. Cell. 2015;162(6):1229–1241.
    1. Choi B.S., Martinez-Falero I.C., Corset C., Munder M., Modolell M., Müller I. Differential impact of L-arginine deprivation on the activation and effector functions of T cells and macrophages. J. Leukoc. Biol. 2009;85(2):268–277.
    1. Pearce E.L., Pearce E.J. Metabolic pathways in immune cell activation and quiescence. Immunity. 2013;38(4):633–643.
    1. O'Rourke B. From bioblasts to mitochondria: ever expanding roles of mitochondria in cell physiology. Front. Physiol. 2010;1:7.
    1. Weinberg S.E., Sena L.A., Chandel N.S. Mitochondria in the regulation of innate and adaptive immunity. Immunity. 2015;42(3):406–417.
    1. Angajala A., Lim S., Phillips J.B., Kim J.H., Yates C., You Z. Diverse roles of mitochondria in immune responses: novel insights into immuno-metabolism. Front. Immunol. 2018;9
    1. Wai T., Langer T. Mitochondrial dynamics and metabolic regulation. Trends Endocrinol. Metabol. 2016;27(2):105–117.
    1. Hamanaka R.B., Chandel N.S. Mitochondrial reactive oxygen species regulate cellular signaling and dictate biological outcomes. Trends Biochem. Sci. 2010;35(9):505–513.
    1. Zorov D.B., Juhaszova M., Sollott S.J. Mitochondrial ROS-induced ROS release: an update and review. Biochim. Biophys. Acta. 2006;1757(5–6):509–517.
    1. Droge W. Free radicals in the physiological control of cell function. Physiol. Rev. 2002;82(1):47–95.
    1. Balaban R.S., Nemoto S., Finkel T. Mitochondria, oxidants, and aging. Cell. 2005;120(4):483–495.
    1. Sena L.A., Chandel N.S. Physiological roles of mitochondrial reactive oxygen species. Mol. Cell. 2012;48(2):158–167.
    1. Rashida Gnanaprakasam J.N., Wu R., Wang R. Metabolic reprogramming in modulating T cell reactive oxygen species generation and antioxidant capacity. Front. Immunol. 2018;9:1075.
    1. Kamiński M.M., Röth D., Krammer P.H., Gülow K. Mitochondria as oxidative signaling organelles in T-cell activation: physiological role and pathological implications. Arch Immunol Ther Exp (Warsz) 2013;61(5):367–384.
    1. Lunt S.Y., Vander Heiden M.G. Aerobic glycolysis: meeting the metabolic requirements of cell proliferation. Annu. Rev. Cell Dev. Biol. 2011;27:441–464.
    1. Mookerjee S.A., Goncalves R.L.S., Gerencser A.A., Nicholls D.G., Brand M.D. The contributions of respiration and glycolysis to extracellular acid production. Biochim. Biophys. Acta. 2015;1847(2):171–181.
    1. Gladden L.B. Lactate metabolism: a new paradigm for the third millennium. J. Physiol. 2004;558(Pt 1):5–30.
    1. W O. Versuche an überlebendem carcinomgewebe. In: M S., editor. Klin Wochenschr. 1923. pp. 776–777.
    1. Love D.C., Hanover J.A. The hexosamine signaling pathway: deciphering the "O-GlcNAc code. Sci. STKE. 2005;2005(312):re13.
    1. Chiaradonna F., Ricciardiello F., Palorini R. The nutrient-sensing hexosamine biosynthetic pathway as the hub of cancer metabolic rewiring. Cells. 2018;7(6)
    1. de Jesus T., Shukla S., Ramakrishnan P. Too sweet to resist: control of immune cell function by O-GlcNAcylation. Cell. Immunol. 2018;333:85–92.
    1. Hart G.W., Slawson C., Ramirez-Correa G., Lagerlof O. Cross talk between O-GlcNAcylation and phosphorylation: roles in signaling, transcription, and chronic disease. Annu. Rev. Biochem. 2011;80:825–858.
    1. Duan G., Walther D. The roles of post-translational modifications in the context of protein interaction networks. PLoS Comput. Biol. 2015;11(2)
    1. Yang X., Qian K. Protein O-GlcNAcylation: emerging mechanisms and functions. Nat. Rev. Mol. Cell Biol. 2017;18(7):452–465.
    1. Moremen K.W., Tiemeyer M., Nairn A.V. Vertebrate protein glycosylation: diversity, synthesis and function. Nat. Rev. Mol. Cell Biol. 2012;13(7):448–462.
    1. Lyons J.J., Milner J.D., Rosenzweig S.D. Glycans instructing immunity: the emerging role of altered glycosylation in clinical immunology. Front Pediatr. 2015;3:54.
    1. Torres C.R., Hart G.W. Topography and polypeptide distribution of terminal N-acetylglucosamine residues on the surfaces of intact lymphocytes. Evidence for O-linked GlcNAc. J. Biol. Chem. 1984;259(5):3308–3317.
    1. Hanover J.A. Glycan-dependent signaling: O-linked N-acetylglucosamine. FASEB J. 2001;15(11):1865–1876.
    1. Hart G.W., Housley M.P., Slawson C. Cycling of O-linked beta-N-acetylglucosamine on nucleocytoplasmic proteins. Nature. 2007;446(7139):1017–1022.
    1. Lefebvre T., Dehennaut V., Guinez C., Olivier S., Drougat L., Mir A.M. Dysregulation of the nutrient/stress sensor O-GlcNAcylation is involved in the etiology of cardiovascular disorders, type-2 diabetes and Alzheimer's disease. Biochim. Biophys. Acta. 2010;1800(2):67–79.
    1. Chen L.B. Mitochondrial membrane potential in living cells. Annu. Rev. Cell Biol. 1988;4:155–181.
    1. Schultz B.E., Chan S.I. Structures and proton-pumping strategies of mitochondrial respiratory enzymes. Annu. Rev. Biophys. Biomol. Struct. 2001;30:23–65.
    1. Mills E., O'Neill L.A. Succinate: a metabolic signal in inflammation. Trends Cell Biol. 2014;24(5):313–320.
    1. Williams N.C., O'Neill L.A.J. A role for the krebs cycle intermediate citrate in metabolic reprogramming in innate immunity and inflammation. Front. Immunol. 2018;9:141.
    1. Murphy M.P., O'Neill L.A.J. Krebs cycle reimagined: the emerging roles of succinate and itaconate as signal transducers. Cell. 2018;174(4):780–784.
    1. O'Neill L.A., Kishton R.J., Rathmell J. A guide to immunometabolism for immunologists. Nat. Rev. Immunol. 2016;16(9):553–565.
    1. Pearce E.J., Everts B. Dendritic cell metabolism. Nat. Rev. Immunol. 2015;15(1):18–29.
    1. Everts B., Amiel E., Huang S.C., Smith A.M., Chang C.H., Lam W.Y. TLR-driven early glycolytic reprogramming via the kinases TBK1-IKKvarepsilon supports the anabolic demands of dendritic cell activation. Nat. Immunol. 2014;15(4):323–332.
    1. Tannahill G.M., Curtis A.M., Adamik J., Palsson-McDermott E.M., McGettrick A.F., Goel G. Succinate is an inflammatory signal that induces IL-1beta through HIF-1alpha. Nature. 2013;496(7444):238–242.
    1. Lampropoulou V., Sergushichev A., Bambouskova M., Nair S., Vincent E.E., Loginicheva E. Itaconate links inhibition of succinate dehydrogenase with macrophage metabolic remodeling and regulation of inflammation. Cell Metabol. 2016;24(1):158–166.
    1. Tol V.A. Aspects of long-chain acyl-COA metabolism. Mol. Cell. Biochem. 1975;7(1):19–31.
    1. McGarry J.D., Brown N.F. The mitochondrial carnitine palmitoyltransferase system. From concept to molecular analysis. Eur. J. Biochem. 1997;244(1):1–14.
    1. Nomura M., Liu J., Rovira, Gonzalez-Hurtado E., Lee J., Wolfgang M.J. Fatty acid oxidation in macrophage polarization. Nat. Immunol. 2016;17(3):216–217.
    1. Patsoukis N., Bardhan K., Chatterjee P., Sari D., Liu B., Bell L.N. PD-1 alters T-cell metabolic reprogramming by inhibiting glycolysis and promoting lipolysis and fatty acid oxidation. Nat. Commun. 2015;6:6692.
    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(6):3299–3303.
    1. Pearce E.L., Walsh M.C., Cejas P.J., Harms G.M., Shen H., Wang L.S. Enhancing CD8 T cell memory by modulating fatty acid metabolism. Nature. 2009;460(7251):103–107.
    1. Divakaruni A.S., Hsieh W.Y., Minarrieta L., Duong T.N., Kim K.K.O., Desousa B.R. Etomoxir inhibits macrophage polarization by disrupting CoA homeostasis. Cell Metabol. 2018;28(3):490–503. e7.
    1. Raud B., Roy D.G., Divakaruni A.S., Tarasenko T.N., Franke R., Ma E.H. Etomoxir actions on regulatory and memory T cells are independent of cpt1a-mediated fatty acid oxidation. Cell Metabol. 2018;28(3):504–515. e7.
    1. Yao C.H., Liu G.Y., Wang R., Moon S.H., Gross R.W., Patti G.J. Identifying off-target effects of etomoxir reveals that carnitine palmitoyltransferase I is essential for cancer cell proliferation independent of β-oxidation. PLoS Biol. 2018;16(3)
    1. Curi R., Newsholme P., Pithon-Curi T.C., Pires-de-Melo M., Garcia C., Homem-de-Bittencourt Júnior P.I. Metabolic fate of glutamine in lymphocytes, macrophages and neutrophils. Braz. J. Med. Biol. Res. 1999;32(1):15–21.
    1. Wood T. Physiological functions of the pentose phosphate pathway. Cell Biochem. Funct. 1986;4(4):241–247.
    1. Perner A., Nielsen S.E., Rask-Madsen J. High glucose impairs superoxide production from isolated blood neutrophils. Intensive Care Med. 2003;29(4):642–645.
    1. Gray G.R., Stamatoyannopoulos G., Naiman S.C., Kliman M.R., Klebanoff S.J., Austin T. Neutrophil dysfunction, chronic granulomatous disease, and non-spherocytic haemolytic anaemia caused by complete deficiency of glucose-6-phosphate dehydrogenase. Lancet. 1973;2(7828):530–534.
    1. Stanton R.C. Glucose-6-phosphate dehydrogenase, NADPH, and cell survival. IUBMB Life. 2012;64(5):362–369.
    1. McCommis K.S., Baines C.P. The role of VDAC in cell death: friend or foe? Biochim. Biophys. Acta. 2012;1818(6):1444–1450.
    1. Seth R.B., Sun L., Ea C.K., Chen Z.J. Identification and characterization of MAVS, a mitochondrial antiviral signaling protein that activates NF-kappaB and IRF 3. Cell. 2005;122(5):669–682.
    1. Rojo M., Legros F., Chateau D., Lombès A. Membrane topology and mitochondrial targeting of mitofusins, ubiquitous mammalian homologs of the transmembrane GTPase Fzo. J. Cell Sci. 2002;115(Pt 8):1663–1674.
    1. Fritz S., Rapaport D., Klanner E., Neupert W., Westermann B. Connection of the mitochondrial outer and inner membranes by Fzo1 is critical for organellar fusion. J. Cell Biol. 2001;152(4):683–692.
    1. Galonek H.L., Hardwick J.M. Upgrading the BCL-2 network. Nat. Cell Biol. 2006;8(12):1317–1319.
    1. Youle R.J. Cell biology. Cellular demolition and the rules of engagement. Science. 2007;315(5813):776–777.
    1. Chan D.C. Mitochondria: dynamic organelles in disease, aging, and development. Cell. 2006;125(7):1241–1252.
    1. Walther D.M., Rapaport D. Biogenesis of mitochondrial outer membrane proteins. Biochim. Biophys. Acta. 2009;1793(1):42–51.
    1. Mazunin I.O., Levitskii S.A., Patrushev M.V., Kamenski P.A. Mitochondrial matrix processes. Biochemistry (Mosc.) 2015;80(11):1418–1428.
    1. Gurung P., Lukens J.R., Kanneganti T.D. Mitochondria: diversity in the regulation of the NLRP3 inflammasome. Trends Mol. Med. 2015;21(3):193–201.
    1. Schroder K., Tschopp J. The inflammasomes. Cell. 2010;140(6):821–832.
    1. Broz P., Dixit V.M. Inflammasomes: mechanism of assembly, regulation and signalling. Nat. Rev. Immunol. 2016;16(7):407–420.
    1. Artyomov M., Sergushichev A., Schilling J.D. Integrating immunometabolism and macrophage diversity. Semin. Immunol. 2016;28(5):417–424.
    1. Ingelsson B., Soderberg D., Strid T., Soderberg A., Bergh A.C., Loitto V. Lymphocytes eject interferogenic mitochondrial DNA webs in response to CpG and non-CpG oligodeoxynucleotides of class C. Proc. Natl. Acad. Sci. U. S. A. 2018;115(3):E478–E487.
    1. Yousefi S., Mihalache C., Kozlowski E., Schmid I., Simon H.U. Viable neutrophils release mitochondrial DNA to form neutrophil extracellular traps. Cell Death Differ. 2009;16(11):1438–1444.
    1. Dorn G.W. Evolving concepts of mitochondrial dynamics. Annu. Rev. Physiol. 2018;81:1–17.
    1. Gao Z., Li Y., Wang F., Huang T., Fan K., Zhang Y. Mitochondrial dynamics controls anti-tumour innate immunity by regulating CHIP-IRF1 axis stability. Nat. Commun. 2017;8(1):1805.
    1. Buck M.D., O'Sullivan D., Klein Geltink R.I., Curtis J.D., Chang C.H., Sanin D.E. Mitochondrial dynamics controls T cell fate through metabolic programming. Cell. 2016;166(1):63–76.
    1. Eisner V., Picard M., Hajnoczky G. Mitochondrial dynamics in adaptive and maladaptive cellular stress responses. Nat. Cell Biol. 2018;20(7):755–765.
    1. Tilokani L., Nagashima S., Paupe V., Prudent J. Mitochondrial dynamics: overview of molecular mechanisms. Essays Biochem. 2018;62(3):341–360.
    1. Garaude J., Acín-Pérez R., Martínez-Cano S., Enamorado M., Ugolini M., Nistal-Villán E. Mitochondrial respiratory-chain adaptations in macrophages contribute to antibacterial host defense. Nat. Immunol. 2016;17(9):1037–1045.
    1. Lam G.Y., Huang J., Brumell J.H. The many roles of NOX2 NADPH oxidase-derived ROS in immunity. Semin. Immunopathol. 2010;32(4):415–430.
    1. DeLeo F.R., Allen L.A., Apicella M., Nauseef W.M. NADPH oxidase activation and assembly during phagocytosis. J. Immunol. 1999;163(12):6732–6740.
    1. Nauseef W.M. Assembly of the phagocyte NADPH oxidase. Histochem. Cell Biol. 2004;122(4):277–291.
    1. Schrader M., Costello J., Godinho L.F., Islinger M. Peroxisome-mitochondria interplay and disease. J. Inherit. Metab. Dis. 2015;38(4):681–702.
    1. Pascual-Ahuir A., Manzanares-Estreder S., Proft M. Pro- and antioxidant functions of the peroxisome-mitochondria connection and its impact on aging and disease. Oxid Med Cell Longev. 2017;2017:9860841.
    1. De Duve C., Baudhuin P. Peroxisomes (microbodies and related particles) Physiol. Rev. 1966;46(2):323–357.
    1. Fransen M., Nordgren M., Wang B., Apanasets O. Role of peroxisomes in ROS/RNS-metabolism: implications for human disease. Biochim. Biophys. Acta. 2012;1822(9):1363–1373.
    1. Andreyev A.Y., Kushnareva Y.E., Starkov A.A. Mitochondrial metabolism of reactive oxygen species. Biochemistry (Mosc.) 2005;70(2):200–214.
    1. Muller F. The nature and mechanism of superoxide production by the electron transport chain: its relevance to aging. J Am Aging Assoc. 2000;23(4):227–253.
    1. Turrens J.F. Mitochondrial formation of reactive oxygen species. J. Physiol. 2003;552(Pt 2):335–344.
    1. Muller F.L., Liu Y., Van Remmen H. Complex III releases superoxide to both sides of the inner mitochondrial membrane. J. Biol. Chem. 2004;279(47):49064–49073.
    1. Thomas C., Mackey M.M., Diaz A.A., Cox D.P. Hydroxyl radical is produced via the Fenton reaction in submitochondrial particles under oxidative stress: implications for diseases associated with iron accumulation. Redox Rep. 2009;14(3):102–108.
    1. Liochev S.I. The mechanism of "Fenton-like" reactions and their importance for biological systems. A biologist's view. Met. Ions Biol. Syst. 1999;36:1–39.
    1. Hayyan M., Hashim M.A., AlNashef I.M. Superoxide ion: generation and chemical implications. Chem. Rev. 2016;116(5):3029–3085.
    1. Sies H. Strategies of antioxidant defense. Eur. J. Biochem. 1993;215(2):213–219.
    1. Paiva C.N., Bozza M.T. Are reactive oxygen species always detrimental to pathogens? Antioxidants Redox Signal. 2014;20(6):1000–1037.
    1. Hurst J.K. What really happens in the neutrophil phagosome? Free Radic. Biol. Med. 2012;53(3):508–520.
    1. Pattison D.I., Davies M.J., Hawkins C.L. Reactions and reactivity of myeloperoxidase-derived oxidants: differential biological effects of hypochlorous and hypothiocyanous acids. Free Radic. Res. 2012;46(8):975–995.
    1. Aratani Y. Myeloperoxidase: its role for host defense, inflammation, and neutrophil function. Arch. Biochem. Biophys. 2018;640:47–52.
    1. Marí M., Morales A., Colell A., García-Ruiz C., Fernández-Checa J.C. Mitochondrial glutathione, a key survival antioxidant. Antioxidants Redox Signal. 2009;11(11):2685–2700.
    1. Orrenius S., Gogvadze V., Zhivotovsky B. Mitochondrial oxidative stress: implications for cell death. Annu. Rev. Pharmacol. Toxicol. 2007;47:143–183.
    1. Hwang C., Sinskey A.J., Lodish H.F. Oxidized redox state of glutathione in the endoplasmic reticulum. Science. 1992;257(5076):1496–1502.
    1. Griffith O.W., Meister A. Origin and turnover of mitochondrial glutathione. Proc. Natl. Acad. Sci. U. S. A. 1985;82(14):4668–4672.
    1. Mak T.W., Grusdat M., Duncan G.S., Dostert C., Nonnenmacher Y., Cox M. Glutathione primes T cell metabolism for inflammation. Immunity. 2017;46(6):1089–1090.
    1. Marí M., Colell A., Morales A., von Montfort C., Garcia-Ruiz C., Fernández-Checa J.C. Redox control of liver function in health and disease. Antioxidants Redox Signal. 2010;12(11):1295–1331.
    1. Valko M., Leibfritz D., Moncol J., Cronin M.T., Mazur M., Telser J. Free radicals and antioxidants in normal physiological functions and human disease. Int. J. Biochem. Cell Biol. 2007;39(1):44–84.
    1. Moncada S., Palmer R.M., Higgs E.A. Nitric oxide: physiology, pathophysiology, and pharmacology. Pharmacol. Rev. 1991;43(2):109–142.
    1. Alderton W.K., Cooper C.E., Knowles R.G. Nitric oxide synthases: structure, function and inhibition. Biochem. J. 2001;357(Pt 3):593–615.
    1. Bogdan C. Nitric oxide synthase in innate and adaptive immunity: an update. Trends Immunol. 2015;36(3):161–178.
    1. Szabó C. Multiple pathways of peroxynitrite cytotoxicity. Toxicol. Lett. 2003;140–141:105–112.
    1. Szabó C., Ischiropoulos H., Radi R. Peroxynitrite: biochemistry, pathophysiology and development of therapeutics. Nat. Rev. Drug Discov. 2007;6(8):662–680.
    1. Radi R., Cassina A., Hodara R., Quijano C., Castro L. Peroxynitrite reactions and formation in mitochondria. Free Radic. Biol. Med. 2002;33(11):1451–1464.
    1. Knoops B., Goemaere J., Van der Eecken V., Declercq J.P. Peroxiredoxin 5: structure, mechanism, and function of the mammalian atypical 2-Cys peroxiredoxin. Antioxidants Redox Signal. 2011;15(3):817–829.
    1. Park J., Choi H., Kim B., Chae U., Lee D.G., Lee S.R. Peroxiredoxin 5 (Prx5) decreases LPS-induced microglial activation through regulation of Ca. Free Radic. Biol. Med. 2016;99:392–404.
    1. Knoops B., Argyropoulou V., Becker S., Ferté L., Kuznetsova O. Multiple roles of peroxiredoxins in inflammation. Mol. Cells. 2016;39(1):60–64.
    1. Tal M.C., Sasai M., Lee H.K., Yordy B., Shadel G.S., Iwasaki A. Absence of autophagy results in reactive oxygen species-dependent amplification of RLR signaling. Proc. Natl. Acad. Sci. U. S. A. 2009;106(8):2770–2775.
    1. Sadatomi D., Nakashioya K., Mamiya S., Honda S., Kameyama Y., Yamamura Y. Mitochondrial function is required for extracellular ATP-induced NLRP3 inflammasome activation. J. Biochem. 2017;161(6):503–512.
    1. Sandhir R., Halder A., Sunkaria A. Mitochondria as a centrally positioned hub in the innate immune response. Biochim Biophys Acta Mol Basis Dis. 2017;1863(5):1090–1097.
    1. van Horssen J., van Schaik P., Witte M. In press: Inflammation and mitochondrial dysfunction: a vicious circle in neurodegenerative disorders? Neurosci. Lett. 2017
    1. Hsu C.C., Tseng L.M., Lee H.C. Role of mitochondrial dysfunction in cancer progression. Exp. Biol. Med. 2016;241:1281–1295.
    1. Ashrafi G., Schwarz T.L. The pathways of mitophagy for quality control and clearance of mitochondria. Cell Death Differ. 2013;20(1):31–42.
    1. Wallace D.C. A mitochondrial paradigm of metabolic and degenerative diseases, aging, and cancer: a dawn for evolutionary medicine. Annu. Rev. Genet. 2005;39:359–407.
    1. Ježek J., Cooper K.F., Strich R. Reactive oxygen species and mitochondrial dynamics: the yin and yang of mitochondrial dysfunction and cancer progression. Antioxidants (Basel) 2018;7(1)
    1. Jin S.M., Youle R.J. PINK1- and Parkin-mediated mitophagy at a glance. J. Cell Sci. 2012;125(Pt 4):795–799.
    1. Poole A.C., Thomas R.E., Yu S., Vincow E.S., Pallanck L. The mitochondrial fusion-promoting factor mitofusin is a substrate of the PINK1/parkin pathway. PLoS One. 2010;5(4)
    1. Tanaka A., Cleland M.M., Xu S., Narendra D.P., Suen D.F., Karbowski M. Proteasome and p97 mediate mitophagy and degradation of mitofusins induced by Parkin. J. Cell Biol. 2010;191(7):1367–1380.
    1. Ziviani E., Tao R.N., Whitworth A.J. Drosophila parkin requires PINK1 for mitochondrial translocation and ubiquitinates mitofusin. Proc. Natl. Acad. Sci. U. S. A. 2010;107(11):5018–5023.
    1. Chan N.C., Salazar A.M., Pham A.H., Sweredoski M.J., Kolawa N.J., Graham R.L. Broad activation of the ubiquitin-proteasome system by Parkin is critical for mitophagy. Hum. Mol. Genet. 2011;20(9):1726–1737.
    1. Soubannier V., McLelland G.L., Zunino R., Braschi E., Rippstein P., Fon E.A. A vesicular transport pathway shuttles cargo from mitochondria to lysosomes. Curr. Biol. 2012;22(2):135–141.
    1. Nathan C. Neutrophils and immunity: challenges and opportunities. Nat. Rev. Immunol. 2006;6(3):173–182.
    1. Rosales C., Demaurex N., Lowell C.A., Uribe-Querol E. Neutrophils: their role in innate and adaptive immunity. J Immunol Res. 2016;2016:1469780.
    1. Mayadas T.N., Cullere X., Lowell C.A. The multifaceted functions of neutrophils. Annu. Rev. Pathol. 2014;9:181–218.
    1. Sbarra A.J., Karnovsky M.L. The biochemical basis of phagocytosis. I. Metabolic changes during the ingestion of particles by polymorphonuclear leukocytes. J. Biol. Chem. 1959;234(6):1355–1362.
    1. Borregaard N., Herlin T. Energy metabolism of human neutrophils during phagocytosis. J. Clin. Investig. 1982;70(3):550–557.
    1. Valentine W.N., Beck W.S. Biochemical studies on leucocytes. I. Phosphatase activity in health, leucocytosis, and myelocytic leucemia. J. Lab. Clin. Med. 1951;38(1):39–55.
    1. Maianski N.A., Geissler J., Srinivasula S.M., Alnemri E.S., Roos D., Kuijpers T.W. Functional characterization of mitochondria in neutrophils: a role restricted to apoptosis. Cell Death Differ. 2004;11(2):143–153.
    1. Riffelmacher T., Clarke A., Richter F.C., Stranks A., Pandey S., Danielli S. Autophagy-dependent generation of free fatty acids is critical for normal neutrophil differentiation. Immunity. 2017;47(3):466–480. e5.
    1. Rožman S., Yousefi S., Oberson K., Kaufmann T., Benarafa C., Simon H.U. The generation of neutrophils in the bone marrow is controlled by autophagy. Cell Death Differ. 2015;22(3):445–456.
    1. Skendros P., Mitroulis I., Ritis K. Autophagy in neutrophils: from granulopoiesis to neutrophil extracellular traps. Front Cell Dev Biol. 2018;6:109.
    1. Rice C.M., Davies L.C., Subleski J.J., Maio N., Gonzalez-Cotto M., Andrews C. Tumour-elicited neutrophils engage mitochondrial metabolism to circumvent nutrient limitations and maintain immune suppression. Nat. Commun. 2018;9(1):5099.
    1. Winterbourn C.C., Hampton M.B., Livesey J.H., Kettle A.J. Modeling the reactions of superoxide and myeloperoxidase in the neutrophil phagosome: implications for microbial killing. J. Biol. Chem. 2006;281(52):39860–39869.
    1. Brinkmann V., Reichard U., Goosmann C., Fauler B., Uhlemann Y., Weiss D.S. Neutrophil extracellular traps kill bacteria. Science. 2004;303(5663):1532–1535.
    1. Steinberg B.E., Grinstein S. Unconventional roles of the NADPH oxidase: signaling, ion homeostasis, and cell death. Sci. STKE. 2007;2007(379):pe11.
    1. Douda D.N., Yip L., Khan M.A., Grasemann H., Palaniyar N. Akt is essential to induce NADPH-dependent NETosis and to switch the neutrophil death to apoptosis. Blood. 2014;123(4):597–600.
    1. Douda D.N., Khan M.A., Grasemann H., Palaniyar N. SK3 channel and mitochondrial ROS mediate NADPH oxidase-independent NETosis induced by calcium influx. Proc. Natl. Acad. Sci. U. S. A. 2015;112(9):2817–2822.
    1. Yipp B.G., Kubes P. NETosis: how vital is it? Blood. 2013;122(16):2784–2794.
    1. Pilsczek F.H., Salina D., Poon K.K., Fahey C., Yipp B.G., Sibley C.D. A novel mechanism of rapid nuclear neutrophil extracellular trap formation in response to Staphylococcus aureus. J. Immunol. 2010;185(12):7413–7425.
    1. Parker H., Dragunow M., Hampton M.B., Kettle A.J., Winterbourn C.C. Requirements for NADPH oxidase and myeloperoxidase in neutrophil extracellular trap formation differ depending on the stimulus. J. Leukoc. Biol. 2012;92(4):841–849.
    1. Papayannopoulos V., Metzler K.D., Hakkim A., Zychlinsky A. Neutrophil elastase and myeloperoxidase regulate the formation of neutrophil extracellular traps. J. Cell Biol. 2010;191(3):677–691.
    1. Wynn J.L., Scumpia P.O., Delano M.J., O'Malley K.A., Ungaro R., Abouhamze A. Increased mortality and altered immunity in neonatal sepsis produced by generalized peritonitis. Shock. 2007;28(6):675–683.
    1. Adam-Vizi V., Starkov A.A. Calcium and mitochondrial reactive oxygen species generation: how to read the facts. J. Alzheimer's Dis. 2010;20(Suppl 2):S413–S426.
    1. Amini P., Stojkov D., Felser A., Jackson C.B., Courage C., Schaller A. Neutrophil extracellular trap formation requires OPA1-dependent glycolytic ATP production. Nat. Commun. 2018;9(1):2958.
    1. Behnen M., Möller S., Brozek A., Klinger M., Laskay T. Extracellular acidification inhibits the ROS-dependent formation of neutrophil extracellular traps. Front. Immunol. 2017;8:184.
    1. Rodríguez-Espinosa O., Rojas-Espinosa O., Moreno-Altamirano M.M., López-Villegas E.O., Sánchez-García F.J. Metabolic requirements for neutrophil extracellular traps formation. Immunology. 2015;145(2):213–224.
    1. Bao Y., Ledderose C., Seier T., Graf A.F., Brix B., Chong E. Mitochondria regulate neutrophil activation by generating ATP for autocrine purinergic signaling. J. Biol. Chem. 2014;289(39):26794–26803.
    1. Bao Y., Ledderose C., Graf A.F., Brix B., Birsak T., Lee A. mTOR and differential activation of mitochondria orchestrate neutrophil chemotaxis. J. Cell Biol. 2015;210(7):1153–1164.
    1. Zhou W., Cao L., Jeffries J., Zhu X., Staiger C.J., Deng Q. Neutrophil-specific knockout demonstrates a role for mitochondria in regulating neutrophil motility in zebrafish. Dis Model Mech. 2018;11(3)
    1. Zheng X., Chen M., Meng X., Chu X., Cai C., Zou F. Phosphorylation of dynamin-related protein 1 at Ser616 regulates mitochondrial fission and is involved in mitochondrial calcium uniporter-mediated neutrophil polarization and chemotaxis. Mol. Immunol. 2017;87:23–32.
    1. Epelman S., Lavine K.J., Randolph G.J. Origin and functions of tissue macrophages. Immunity. 2014;41(1):21–35.
    1. Shi C., Pamer E.G. Monocyte recruitment during infection and inflammation. Nat. Rev. Immunol. 2011;11(11):762–774.
    1. Correa-da-Silva F., Pereira J.A.S., de Aguiar C.F., de Moraes-Vieira P.M.M. Mitoimmunity-when mitochondria dictates macrophage function. Cell Biol. Int. 2018;42(6):651–655.
    1. Castoldi A., Naffah de Souza C., Camara N.O., Moraes-Vieira P.M. The macrophage switch in obesity development. Front. Immunol. 2015;6:637.
    1. Braga T.T., Agudelo J.S., Camara N.O. Macrophages during the fibrotic process: M2 as friend and foe. Front. Immunol. 2015;6:602.
    1. West A.P., Brodsky I.E., Rahner C., Woo D.K., Erdjument-Bromage H., Tempst P. TLR signalling augments macrophage bactericidal activity through mitochondrial ROS. Nature. 2011;472(7344):476–480.
    1. Freemerman A.J., Johnson A.R., Sacks G.N., Milner J.J., Kirk E.L., Troester M.A. Metabolic reprogramming of macrophages: glucose transporter 1 (GLUT1)-mediated glucose metabolism drives a proinflammatory phenotype. J. Biol. Chem. 2014;289(11):7884–7896.
    1. Fukuzumi M., Shinomiya H., Shimizu Y., Ohishi K., Utsumi S. Endotoxin-induced enhancement of glucose influx into murine peritoneal macrophages via GLUT1. Infect. Immun. 1996;64(1):108–112.
    1. Krawczyk C.M., Holowka T., Sun J., Blagih J., Amiel E., DeBerardinis R.J. Toll-like receptor-induced changes in glycolytic metabolism regulate dendritic cell activation. Blood. 2010;115(23):4742–4749.
    1. Diskin C., Pålsson-McDermott E.M. Metabolic modulation in macrophage effector function. Front. Immunol. 2018;9
    1. Jha A.K., Huang S.C., Sergushichev A., Lampropoulou V., Ivanova Y., Loginicheva E. Network integration of parallel metabolic and transcriptional data reveals metabolic modules that regulate macrophage polarization. Immunity. 2015;42(3):419–430.
    1. Mills E.L., Ryan D.G., Prag H.A., Dikovskaya D., Menon D., Zaslona Z. Itaconate is an anti-inflammatory metabolite that activates Nrf2 via alkylation of KEAP1. Nature. 2018;556(7699):113.
    1. McFadden B.A., Purohit S. Itaconate, an isocitrate lyase-directed inhibitor in Pseudomonas indigofera. J. Bacteriol. 1977;131(1):136–144.
    1. Mills E.L., Ryan D.G., Prag H.A., Dikovskaya D., Menon D., Zaslona Z. Itaconate is an anti-inflammatory metabolite that activates Nrf2 via alkylation of KEAP1. Nature. 2018;556(7699):113–117.
    1. Mills E.L., Kelly B., Logan A., Costa A.S.H., Varma M., Bryant C.E. Succinate dehydrogenase supports metabolic repurposing of mitochondria to drive inflammatory macrophages. Cell. 2016;167(2):457–470. e13.
    1. Van den Bossche J., Baardman J., Otto N.A., van der Velden S., Neele A.E., van den Berg S.M. Mitochondrial dysfunction prevents repolarization of inflammatory macrophages. Cell Rep. 2016;17(3):684–696.
    1. Krzyszczyk P., Schloss R., Palmer A., Berthiaume F. The role of macrophages in acute and chronic wound healing and interventions to promote pro-wound healing phenotypes. Front. Physiol. 2018;9
    1. Huang S.C., Smith A.M., Everts B., Colonna M., Pearce E.L., Schilling J.D. Metabolic reprogramming mediated by the mTORC2-IRF4 signaling Axis is essential for macrophage alternative activation. Immunity. 2016;45(4):817–830.
    1. Vats D., Mukundan L., Odegaard J.I., Zhang L., Smith K.L., Morel C.R. Oxidative metabolism and PGC-1β attenuate macrophage-mediated inflammation. Cell Metabol. 2006;4(1):13–24.
    1. Huang S.C., Everts B., Ivanova Y., O'Sullivan D., Nascimento M., Smith A.M. Cell-intrinsic lysosomal lipolysis is essential for alternative activation of macrophages. Nat. Immunol. 2014;15(9):846–855.
    1. Namgaladze D., Brune B. Fatty acid oxidation is dispensable for human macrophage IL-4-induced polarization. Biochim. Biophys. Acta. 2014;1841(9):1329–1335.
    1. Covarrubias A.J., Aksoylar H.I., Yu J., Snyder N.W., Worth A.J., Iyer S.S. Akt-mTORC1 signaling regulates Acly to integrate metabolic input to control of macrophage activation. Elife. 2016;5
    1. Wang F., Zhang S., Vuckovic I., Jeon R., Lerman A., Folmes C.D. Glycolytic stimulation is not a requirement for M2 macrophage differentiation. Cell Metabol. 2018;28(3):463–475. e4.
    1. Wang Y., Subramanian M., Yurdagul A., Jr., Barbosa-Lorenzi V.C., Cai B., de Juan-Sanz J. Mitochondrial fission promotes the continued clearance of apoptotic cells by macrophages. Cell. 2017;171(2):331–345. e22.
    1. Escoll P., Song O.R., Viana F., Steiner B., Lagache T., Olivo-Marin J.C. Legionella pneumophila modulates mitochondrial dynamics to trigger metabolic repurposing of infected macrophages. Cell Host Microbe. 2017;22(3):302–316. e7.
    1. Colegio O.R., Chu N.Q., Szabo A.L., Chu T., Rhebergen A.M., Jairam V. Functional polarization of tumour-associated macrophages by tumour-derived lactic acid. Nature. 2014;513(7519):559–563.
    1. Olefsky J.M., Glass C.K. Macrophages, inflammation, and insulin resistance. Annu. Rev. Physiol. 2010;72:219–246.
    1. Kratz M., Coats B.R., Hisert K.B., Hagman D., Mutskov V., Peris E. Metabolic dysfunction drives a mechanistically distinct proinflammatory phenotype in adipose tissue macrophages. Cell Metabol. 2014;20(4):614–625.
    1. Serbulea V., Upchurch C.M., Schappe M.S., Voigt P., DeWeese D.E., Desai B.N. Macrophage phenotype and bioenergetics are controlled by oxidized phospholipids identified in lean and obese adipose tissue. Proc. Natl. Acad. Sci. U. S. A. 2018;115(27) E6254-e63.
    1. Biswas S.K., Chittezhath M., Shalova I.N., Lim J.Y. Macrophage polarization and plasticity in health and disease. Immunol. Res. 2012;53(1–3):11–24.
    1. Moraes-Vieira P.M., Yore M.M., Dwyer P.M., Syed I., Aryal P., Kahn B.B. RBP4 activates antigen-presenting cells, leading to adipose tissue inflammation and systemic insulin resistance. Cell Metabol. 2014;19(3):512–526.
    1. Merad M., Sathe P., Helft J., Miller J., Mortha A. The dendritic cell lineage: ontogeny and function of dendritic cells and their subsets in the steady state and the inflamed setting. Annu. Rev. Immunol. 2013;31:563–604.
    1. Jantsch J., Chakravortty D., Turza N., Prechtel A.T., Buchholz B., Gerlach R.G. Hypoxia and hypoxia-inducible factor-1 alpha modulate lipopolysaccharide-induced dendritic cell activation and function. J. Immunol. 2008;180(7):4697–4705.
    1. Palmieri F. The mitochondrial transporter family (SLC25): physiological and pathological implications. Pflügers Archiv. 2004;447(5):689–709.
    1. Everts B., Amiel E., Huang S.C., Smith A.M., Chang C.H., Lam W.Y. TLR-driven early glycolytic reprogramming via the kinases TBK1-IKKϵ supports the anabolic demands of dendritic cell activation. Nat. Immunol. 2014;15(4):323–332.
    1. Everts B., Amiel E., van der Windt G.J., Freitas T.C., Chott R., Yarasheski K.E. Commitment to glycolysis sustains survival of NO-producing inflammatory dendritic cells. Blood. 2012;120(7):1422–1431.
    1. Thwe P.M., Pelgrom L., Cooper R., Beauchamp S., Reisz J.A., D'Alessandro A. Cell-intrinsic glycogen metabolism supports early glycolytic reprogramming required for dendritic cell immune responses. Cell Metabol. 2017;26(3):558–567. e5.
    1. Zaccagnino P., Saltarella M., Maiorano S., Gaballo A., Santoro G., Nico B. An active mitochondrial biogenesis occurs during dendritic cell differentiation. Int. J. Biochem. Cell Biol. 2012;44(11):1962–1969.
    1. Del Prete A., Zaccagnino P., Di Paola M., Saltarella M., Oliveros Celis C., Nico B. Role of mitochondria and reactive oxygen species in dendritic cell differentiation and functions. Free Radic. Biol. Med. 2008;44(7):1443–1451.
    1. Sheng K.C., Pietersz G.A., Tang C.K., Ramsland P.A., Apostolopoulos V. Reactive oxygen species level defines two functionally distinctive stages of inflammatory dendritic cell development from mouse bone marrow. J. Immunol. 2010;184(6):2863–2872.
    1. Svajger U., Obermajer N., Jeras M. Dendritic cells treated with resveratrol during differentiation from monocytes gain substantial tolerogenic properties upon activation. Immunology. 2010;129(4):525–535.
    1. Malinarich F., Duan K., Hamid R.A., Bijin A., Lin W.X., Poidinger M. High mitochondrial respiration and glycolytic capacity represent a metabolic phenotype of human tolerogenic dendritic cells. J. Immunol. 2015;194(11):5174–5186.
    1. Oberkampf M., Guillerey C., Mouries J., Rosenbaum P., Fayolle C., Bobard A. Mitochondrial reactive oxygen species regulate the induction of CD8(+) T cells by plasmacytoid dendritic cells. Nat. Commun. 2018;9(1):2241.
    1. Ryu S.W., Han E.C., Yoon J., Choi C. The mitochondrial fusion-related proteins Mfn2 and OPA1 are transcriptionally induced during differentiation of bone marrow progenitors to immature dendritic cells. Mol. Cells. 2015;38:89–94.
    1. Swainson L., Kinet S., Manel N., Battini J.L., Sitbon M., Taylor N. Glucose transporter 1 expression identifies a population of cycling CD4+CD8+ human thymocytes with high CXCR4-induced chemotaxis. Proc. Natl. Acad. Sci. U. S. A. 2005;102:12867–12872.
    1. MacIver N.J., Michalek R.D., Rathmell J.C. Metabolic regulation of T lymphocytes. Annu. Rev. Immunol. 2013;31:259–283.
    1. Menk A.V., Scharping N.E., Moreci R.S., Zeng X., Guy C., Salvatore S. Early TCR signaling induces rapid aerobic glycolysis enabling distinct acute T cell effector functions. Cell Rep. 2018;22(6):1509–1521.
    1. Shi L.Z., Wang R., Huang G., Vogel P., Neale G., Green D.R. HIF1alpha-dependent glycolytic pathway orchestrates a metabolic checkpoint for the differentiation of TH17 and Treg cells. J. Exp. Med. 2011;208(7):1367–1376.
    1. Frauwirth K.A., Riley J.L., Harris M.H., Parry R.V., Rathmell J.C., Plas D.R. The CD28 signaling pathway regulates glucose metabolism. Immunity. 2002;16(6):769–777.
    1. Wang R., Dillon C.P., Shi L.Z., Milasta S., Carter R., Finkelstein D. The transcription factor Myc controls metabolic reprogramming upon T lymphocyte activation. Immunity. 2011;35(6):871–882.
    1. Vander Heiden M.G., Cantley L.C., Thompson C.B. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science. 2009;324(5930):1029–1033.
    1. Johnson M.O., Wolf M.M., Madden M.Z., Andrejeva G., Sugiura A., Contreras D.C. Distinct regulation of Th17 and Th1 cell differentiation by glutaminase-dependent metabolism. Cell. 2018;175(7):1780–1795. e19.
    1. Kuwahara M., Izumoto M., Honda H., Inoue K., Imai Y., Suzuki J. 2017. Glutamine Metabolism Regulates Th2 Cell Differentiation via the α-ketoglutarate-dependent Demethylation of Histone H3K27.
    1. Chang C.H., Curtis J.D., Maggi L.B., Jr., Faubert B., Villarino A.V., O'Sullivan D. Posttranscriptional control of T cell effector function by aerobic glycolysis. Cell. 2013;153(6):1239–1251.
    1. Sena L.A., Li S., Jairaman A., Prakriya M., Ezponda T., Hildeman D.A. Mitochondria are required for antigen-specific T cell activation through reactive oxygen species signaling. Immunity. 2013;38(2):225–236.
    1. Contento R.L., Campello S., Trovato A.E., Magrini E., Anselmi F., Viola A. Adhesion shapes T cells for prompt and sustained T-cell receptor signalling. EMBO J. 2010;29(23):4035–4047.
    1. Quintana A., Schwindling C., Wenning A.S., Becherer U., Rettig J., Schwarz E.C. T cell activation requires mitochondrial translocation to the immunological synapse. Proc. Natl. Acad. Sci. U. S. A. 2007;104(36):14418–14423.
    1. Berod L., Friedrich C., Nandan A., Freitag J., Hagemann S., Harmrolfs K. De novo fatty acid synthesis controls the fate between regulatory T and T helper 17 cells. Nat. Med. 2014;20(11):1327–1333.
    1. Endo Y., Asou H.K., Matsugae N., Hirahara K., Shinoda K., Tumes D.J. Obesity drives Th17 cell differentiation by inducing the lipid metabolic kinase. ACC1. Cell Rep. 2015;12(6):1042–1055.
    1. Raha S., Raud B., Oberdorfer L., Castro C.N., Schreder A., Freitag J. Disruption of de novo fatty acid synthesis via acetyl-CoA carboxylase 1 inhibition prevents acute graft-versus-host disease. Eur. J. Immunol. 2016;46(9):2233–2238.
    1. Lee J., Walsh M.C., Hoehn K.L., James D.E., Wherry E.J., Choi Y. Regulator of fatty acid metabolism, acetyl coenzyme a carboxylase 1, controls T cell immunity. J. Immunol. 2014;192(7):3190–3199.
    1. Devadas S., Zaritskaya L., Rhee S.G., Oberley L., Williams M.S. Discrete generation of superoxide and hydrogen peroxide by T cell receptor stimulation: selective regulation of mitogen-activated protein kinase activation and fas ligand expression. J. Exp. Med. 2002;195(1):59–70.
    1. Kaminski M., Kiessling M., Suss D., Krammer P.H., Gulow K. Novel role for mitochondria: protein kinase Ctheta-dependent oxidative signaling organelles in activation-induced T-cell death. Mol. Cell. Biol. 2007;27(10):3625–3639.
    1. Kaminski M.M., Roth D., Sass S., Sauer S.W., Krammer P.H., Gulow K. Manganese superoxide dismutase: a regulator of T cell activation-induced oxidative signaling and cell death. Biochim. Biophys. Acta. 2012;1823(5):1041–1052.
    1. Kaminski M.M., Sauer S.W., Klemke C.D., Suss D., Okun J.G., Krammer P.H. Mitochondrial reactive oxygen species control T cell activation by regulating IL-2 and IL-4 expression: mechanism of ciprofloxacin-mediated immunosuppression. J. Immunol. 2010;184(9):4827–4841.
    1. Yi J.S., Holbrook B.C., Michalek R.D., Laniewski N.G., Grayson J.M. Electron transport complex I is required for CD8+ T cell function. J. Immunol. 2006;177(2):852–862.
    1. Los M., Schenk H., Hexel K., Baeuerle P.A., Droge W., Schulze-Osthoff K. IL-2 gene expression and NF-kappa B activation through CD28 requires reactive oxygen production by 5-lipoxygenase. EMBO J. 1995;14(15):3731–3740.
    1. Jackson S.H., Devadas S., Kwon J., Pinto L.A., Williams M.S. T cells express a phagocyte-type NADPH oxidase that is activated after T cell receptor stimulation. Nat. Immunol. 2004;5(8):818–827.
    1. Kwon J., Shatynski K.E., Chen H., Morand S., de Deken X., Miot F. The nonphagocytic NADPH oxidase Duox1 mediates a positive feedback loop during T cell receptor signaling. Sci. Signal. 2010;3(133):ra59.
    1. Kaminski M.M., Sauer S.W., Kaminski M., Opp S., Ruppert T., Grigaravicius P. T cell activation is driven by an ADP-dependent glucokinase linking enhanced glycolysis with mitochondrial reactive oxygen species generation. Cell Rep. 2012;2(5):1300–1315.
    1. McKinstry K.K., Strutt T.M., Swain S.L. Regulation of CD4+ T-cell contraction during pathogen challenge. Immunol. Rev. 2010;236:110–124.
    1. van der Windt G.J., Everts B., Chang C.H., Curtis J.D., Freitas T.C., Amiel E. Mitochondrial respiratory capacity is a critical regulator of CD8+ T cell memory development. Immunity. 2012;36(1):68–78.
    1. Taub D.D., Hesdorffer C.S., Ferrucci L., Madara K., Schwartz J.B., Goetzl E.J. Distinct energy requirements for human memory CD4 T-cell homeostatic functions. FASEB J. 2013;27(1):342–349.
    1. O'Sullivan D., van der Windt G.J., Huang S.C., Curtis J.D., Chang C.H., Buck M.D. Memory CD8(+) T cells use cell-intrinsic lipolysis to support the metabolic programming necessary for development. Immunity. 2014;41(1):75–88.
    1. D'Souza A.D., Parikh N., Kaech S.M., Shadel G.S. Convergence of multiple signaling pathways is required to coordinately up-regulate mtDNA and mitochondrial biogenesis during T cell activation. Mitochondrion. 2007;7(6):374–385.
    1. Boothby M.R., Hodges E., Thomas J.W. Molecular regulation of peripheral B cells and their progeny in immunity. Genes Dev. 2019;33(1–2):26–48.
    1. Browne E.P. Regulation of B-cell responses by Toll-like receptors. Immunology. 2012;136(4):370–379.
    1. Chen X., Jensen P.E. The role of B lymphocytes as antigen-presenting cells. Arch Immunol Ther Exp (Warsz) 2008;56(2):77–83.
    1. Carter M.J., Mitchell R.M., Meyer Sauteur P.M., Kelly D.F., Truck J. The antibody-secreting cell response to infection: kinetics and clinical applications. Front. Immunol. 2017;8:630.
    1. Schroeder H.W., Jr., Cavacini L. Structure and function of immunoglobulins. J. Allergy Clin. Immunol. 2010;125(2 Suppl 2):S41–S52.
    1. Tsui C., Martinez-Martin N., Gaya M., Maldonado P., Llorian M., Legrave N.M. Protein kinase C-beta dictates B cell fate by regulating mitochondrial remodeling, metabolic reprogramming, and heme biosynthesis. Immunity. 2018;48(6):1144–11459 e5.
    1. Mendoza P., Martinez-Martin N., Bovolenta E.R., Reyes-Garau D., Hernansanz-Agustin P., Delgado P. R-Ras2 is required for germinal center formation to aid B cells during energetically demanding processes. Sci. Signal. 2018;11(532)
    1. Baumgarth N. The double life of a B-1 cell: self-reactivity selects for protective effector functions. Nat. Rev. Immunol. 2011;11(1):34–46.
    1. Montecino-Rodriguez E., Dorshkind K. B-1 B cell development in the fetus and adult. Immunity. 2012;36(1):13–21.
    1. Clarke A.J., Riffelmacher T., Braas D., Cornall R.J., Simon A.K. B1a B cells require autophagy for metabolic homeostasis and self-renewal. J. Exp. Med. 2018;215(2):399–413.
    1. Pillai S., Cariappa A. The follicular versus marginal zone B lymphocyte cell fate decision. Nat. Rev. Immunol. 2009;9(11):767–777.
    1. Stein M., Dutting S., Mougiakakos D., Bosl M., Fritsch K., Reimer D. A defined metabolic state in pre B cells governs B-cell development and is counterbalanced by Swiprosin-2/EFhd1. Cell Death Differ. 2017;24(7):1239–1252.
    1. Zouali M., Richard Y. Marginal zone B-cells, a gatekeeper of innate immunity. Front. Immunol. 2011;2:63.
    1. Nutt S.L., Hodgkin P.D., Tarlinton D.M., Corcoran L.M. The generation of antibody-secreting plasma cells. Nat. Rev. Immunol. 2015;15(3):160–171.
    1. Jellusova J., Cato M.H., Apgar J.R., Ramezani-Rad P., Leung C.R., Chen C. Gsk3 is a metabolic checkpoint regulator in B cells. Nat. Immunol. 2017;18(3):303–312.
    1. Cho S.H., Raybuck A.L., Stengel K., Wei M., Beck T.C., Volanakis E. Germinal centre hypoxia and regulation of antibody qualities by a hypoxia response system. Nature. 2016;537(7619):234–238.
    1. Jang K.J., Mano H., Aoki K., Hayashi T., Muto A., Nambu Y. Mitochondrial function provides instructive signals for activation-induced B-cell fates. Nat. Commun. 2015;6:6750.
    1. Caro-Maldonado A., Wang R., Nichols A.G., Kuraoka M., Milasta S., Sun L.D. Metabolic reprogramming is required for antibody production that is suppressed in anergic but exaggerated in chronically BAFF-exposed B cells. J. Immunol. 2014;192(8):3626–3636.
    1. Dufort F.J., Gumina M.R., Ta N.L., Tao Y., Heyse S.A., Scott D.A. Glucose-dependent de novo lipogenesis in B lymphocytes: a requirement for atp-citrate lyase in lipopolysaccharide-induced differentiation. J. Biol. Chem. 2014;289(10):7011–7024.
    1. Waters L.R., Ahsan F.M., Wolf D.M., Shirihai O., Teitell M.A. Initial B cell activation induces metabolic reprogramming and mitochondrial remodeling. iScience. 2018;5:99–109.
    1. Kunisawa J., Sugiura Y., Wake T., Nagatake T., Suzuki H., Nagasawa R. Mode of bioenergetic metabolism during B cell differentiation in the intestine determines the distinct requirement for vitamin B1. Cell Rep. 2015;13(1):122–131.
    1. Irish J.M., Czerwinski D.K., Nolan G.P., Levy R. Kinetics of B cell receptor signaling in human B cell subsets mapped by phosphospecific flow cytometry. J. Immunol. 2006;177(3):1581–1589.
    1. Polikowsky H.G., Wogsland C.E., Diggins K.E., Huse K., Irish J.M. Cutting edge: redox signaling hypersensitivity distinguishes human germinal center B cells. J. Immunol. 2015;195(4):1364–1367.
    1. Wheeler M.L., Defranco A.L. Prolonged production of reactive oxygen species in response to B cell receptor stimulation promotes B cell activation and proliferation. J. Immunol. 2012;189(9):4405–4416.
    1. Richards S.M., Clark E.A. BCR-induced superoxide negatively regulates B-cell proliferation and T-cell-independent type 2 Ab responses. Eur. J. Immunol. 2009;39(12):3395–3403.
    1. Sugamata R., Donko A., Murakami Y., Boudreau H.E., Qi C.F., Kwon J. Duox1 regulates primary B cell function under the influence of IL-4 through BCR-mediated generation of hydrogen peroxide. J. Immunol. 2019;202(2):428–440.
    1. Vene R., Delfino L., Castellani P., Balza E., Bertolotti M., Sitia R. Redox remodeling allows and controls B-cell activation and differentiation. Antioxidants Redox Signal. 2010;13(8):1145–1155.
    1. Price M.J., Patterson D.G., Scharer C.D., Boss J.M. Progressive upregulation of oxidative metabolism facilitates plasmablast differentiation to a T-independent antigen. Cell Rep. 2018;23(11):3152–3159.
    1. Lam W.Y., Becker A.M., Kennerly K.M., Wong R., Curtis J.D., Llufrio E.M. Mitochondrial pyruvate import promotes long-term survival of antibody-secreting plasma cells. Immunity. 2016;45(1):60–73.
    1. Lam W.Y., Jash A., Yao C.H., D'Souza L., Wong R., Nunley R.M. Metabolic and transcriptional modules independently diversify plasma cell lifespan and function. Cell Rep. 2018;24(9):2479–24792 e6.
    1. Watanabe-Matsui M., Muto A., Matsui T., Itoh-Nakadai A., Nakajima O., Murayama K. Heme regulates B-cell differentiation, antibody class switch, and heme oxygenase-1 expression in B cells as a ligand of Bach2. Blood. 2011;117(20):5438–5448.

Source: PubMed

3
Abonnere