Diverse Roles of Mitochondria in Immune Responses: Novel Insights Into Immuno-Metabolism

Anusha Angajala, Sangbin Lim, Joshua B Phillips, Jin-Hwan Kim, Clayton Yates, Zongbing You, Ming Tan, Anusha Angajala, Sangbin Lim, Joshua B Phillips, Jin-Hwan Kim, Clayton Yates, Zongbing You, Ming Tan

Abstract

Lack of immune system cells or impairment in differentiation of immune cells is the basis for many chronic diseases. Metabolic changes could be the root cause for this immune cell impairment. These changes could be a result of altered transcription, cytokine production from surrounding cells, and changes in metabolic pathways. Immunity and mitochondria are interlinked with each other. An important feature of mitochondria is it can regulate activation, differentiation, and survival of immune cells. In addition, it can also release signals such as mitochondrial DNA (mtDNA) and mitochondrial ROS (mtROS) to regulate transcription of immune cells. From current literature, we found that mitochondria can regulate immunity in different ways. First, alterations in metabolic pathways (TCA cycle, oxidative phosphorylation, and FAO) and mitochondria induced transcriptional changes can lead to entirely different outcomes in immune cells. For example, M1 macrophages exhibit a broken TCA cycle and have a pro-inflammatory role. By contrast, M2 macrophages undergo β-oxidation to produce anti-inflammatory responses. In addition, amino acid metabolism, especially arginine, glutamine, serine, glycine, and tryptophan, is critical for T cell differentiation and macrophage polarization. Second, mitochondria can activate the inflammatory response. For instance, mitochondrial antiviral signaling and NLRP3 can be activated by mitochondria. Third, mitochondrial mass and mobility can be influenced by fission and fusion. Fission and fusion can influence immune functions. Finally, mitochondria are placed near the endoplasmic reticulum (ER) in immune cells. Therefore, mitochondria and ER junction signaling can also influence immune cell metabolism. Mitochondrial machinery such as metabolic pathways, amino acid metabolism, antioxidant systems, mitochondrial dynamics, mtDNA, mitophagy, and mtROS are crucial for immune functions. Here, we have demonstrated how mitochondria coordinate to alter immune responses and how changes in mitochondrial machinery contribute to alterations in immune responses. A better understanding of the molecular components of mitochondria is necessary. This can help in the development of safe and effective immune therapy or prevention of chronic diseases. In this review, we have presented an updated prospective of the mitochondrial machinery that drives various immune responses.

Keywords: TCA cycle; fatty acid oxidation; memory T cell; oxidative phosphorylation; regulatory T cell.

Figures

Figure 1
Figure 1
Mitochondria roles in immune responses: mitochondrial components are involved in immune functions. (1) Metabolic pathways such as TCA cycle, oxidative phosphorylation (OXPHOS), and fatty acid oxidation (FAO) are important for macrophage polarization and T cell differentiation. (2) Amino acid metabolism (Tryptophan metabolism is shown here.) also contributes to mitochondrial immune functions. Tryptophan depletion can cause immune tolerance. NAD+ is synthesized from tryptophan. NAD+ contributes to SIRT3-mediated mitochondrial biogenesis. In addition, NAD+ is required for metabolic pathways and CD4+ T cell differentiation. (3) Mitochondrial ROS (mtROS) control immune cell transcription, metabolism, and NLRP3 mediated inflammation. Antioxidants such as glutathione balance mtROS. SIRT3 is involved in regulating many mitochondrial proteins such as IDH2 (TCA cycle), MnSOD [reactive oxygen species (ROS) balance], and FOXO3. SIRT3 inhibits mtROS. (4) Mitophagy is crucial for removing damaged mitochondria. Damaged mitochondria are a source of mtROS so mitophagy balances mtROS. Pink1, localized on the mitochondrial outer membrane, binds to parkin and initiates mitophagy. Parkin mutations can increase the susceptibility to the intracellular bacteria Mycobacterium leprae and Salmonella enterica. Decreased mitophagy results in increased ROS which further increases the susceptibility to infections. Hepatitis B and C viruses use mitophagy to their benefit. These viruses protect themselves from mitochondria induced apoptosis by activating mitophagy. (5) Mitochondrial antiviral signaling (MAVS) in the mitochondria membrane initiates the inflammatory response. Upon viral infections, viral RNA is sensed by RIG-I. RIG-I activates MAVS (located on the outer membrane of mitochondria). MAVS signaling can activate NLRP3 inflammasome and NF-κB/IRF transcription. (6) NLRP3 inflammasome contributes to the activation of caspase-1 and leads to NF-κB signaling and production of IL-1β and IL-18. (7) Mitochondrial DNA (mtDNA) can be released from mitochondria into the cytosol and activate the NLRP3 inflammasome and production of IL-1β and IL-18. Also, mtROS can induce mtDNA mutation. (8) Mitochondrial dynamics, fission (OPA-1) and fusion (Drp1), is associated with activated T cells and memory T cells, respectively. (9) ER and mitochondria junction: in CD8+ memory T cells, mitochondria are placed near the ER. HK-1 (hexokinase-1, an important enzyme in glycolysis) in conjunction with pAKT, mammalian target of rapamycin complex 2 (mTORC2), and glycogen synthase kinase 3β (GSK-3β) mediated signaling cascades regulate CD8+ memory T cell metabolism. (10) Uncoupling protein 2 (UCP2), a mitochondrial membrane protein, allows protons to enter the mitochondrial matrix, thereby decreasing mitochondrial membrane potential which further decreases mtROS.
Figure 2
Figure 2
Metabolic regulation in macrophages and T cells: (A) metabolism during macrophage polarization: polarization of monocytes to M1 or M2 phenotypes is controlled by the cytokines produced by other immune cells. M1 macrophages are activated by IFN-γ produced by Th1 cells. M1 macrophages tend to store surplus FA as triacyclglycerols and cholesteryl esters in lipid droplets, and they exhibit higher aerobic glycolysis and lower oxidative phosphorylation (OXPHOS). Nitric oxide production is higher in M1. Uncoupling protein 2 (UCP2) expression is decreased in M1 macrophages. Contrarily, M2 macrophages are activated by IL-4 or IL-13 to regulate anti-inflammation and promote Th2 response and tissue repair. M2 macrophages adopt a metabolic program dominated by fatty acid-fueled OXPHOS and channel FA toward re-esterification and β-oxidation. Silencing UCP2 impairs M2 macrophage activation by IL-4. High adenosine monophosphate-activated protein kinase (AMPK) and low NO is the reason for high OXPHOS in M2 macrophages. (B) Metabolism during T cells differentiation: naïve T cells are dependent on OXPHOS as their primary metabolic pathway. By contrast, activated T cells exhibit higher glycolysis than OXPHOS. After differentiation, Th1, Th2, and Th17 have higher glycolysis than OXPHOS and high mTORC1 activity. Memory T cells and regulatory T cells undergo AMPK-dependent FAO and have variable mTORC1.
Figure 3
Figure 3
Metabolism in M1 macrophages. (A) M1 macrophage metabolic regulation: M1 macrophages are prevalent in obese adipose tissue. Glucose uptake is increased in M1 macrophages. Importantly, the TCA cycle exhibits two breaks. The first break involves the enzyme isocitrate dehydrogenase (IDH) which results in increased levels of citrate and itaconic acid. Citrate feeds fatty acid (FA) synthesis for prostaglandin (PG) and nitric oxide (NO) production while itaconic acid has anti-bacterial properties. The second break happens with the enzyme succinate dehydrogenase (SDH) which causes increased succinate levels. Succinate stabilizes HIF-1α which binds to the interleukin (IL)-1β promoter boosting IL-1β production and inflammation. Increased flow through the aspartate arginosuccinate shunt (AASS) replenishes the TCA cycle which further increases citrate levels and feeds the urea cycle which contributes to NO production. Glutamine is converted into glutamate by glutamate synthase (GS). Glutamate can further be converted into α-ketoglutarate (αKG). Low αKG/succinate ratio strengthens M1 macrophage activation. Glutamine-synthetase inhibition skews M2-polarized macrophages toward the M1-like phenotype characterized by reduced intracellular glutamine and increased succinate with enhanced glucose flux through glycolysis which is partly related to HIF-1α activation. (B) M1 macrophage transcriptional regulation: in M1 macrophages, the TCA cycle is broken. As a result, citrate is converted to free fatty acids (FFAs). FFAs activate macrophage IKK and JNK1 signaling molecules to induce M1 polarization. Activation of IKK and JNK1 in macrophages produces IL-1β and TNFα which can initiate inflammatory response. Adiponectin is released from adipose tissue and is involved in the breakdown of fatty acids. Adiponectin signaling inhibits M1 programming.
Figure 4
Figure 4
Metabolism in M2 macrophages. (A) M2 macrophage metabolic regulation: M2 macrophages adopt a metabolic program fueled by fatty acids and oxidative phosphorylation (OXPHOS). M2 macrophages are seen in lean adipose tissue and lipid uptake is increased in M2 macrophages (8). Both OXPHOS and FAO are important for the anti-inflammatory function of M2 macrophages. In M2 macrophages, glycolysis is impaired due to low HIF-1α activity and low expression of the active 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase 1 (PFKFB1). Also, high α-KG/succinate ratio promotes the M2 phenotype. This could be due to the high flow of the TCA cycle or glutamine metabolism. M2 macrophages will channel FA toward re-esterification and β-oxidation. β-oxidation is channeled through carinitine palmitoyl transferase 1 (CPT1) (9). The high rates of OXPHOS in M2 macrophages are caused by low nitric oxide (NO) production and high adenosine monophosphate-activated protein kinase (AMPK) activity. (B) M2 macrophage transcriptional regulation: omega-3 polyunsaturated fatty acids (PUFA) undergo enzymatic transformation by 12/15-LO (lipoxygenase) and other enzymes in the production of specialized pro-resolving lipid mediators (SPMs) and intermediate metabolites that are potent PPAR ligands. M2-inducing cytokines IL-4 and IL-13 promote STAT6 phosphorylation to induce transcription of PPARs (γ and δ) and their coactivator PPARγ-coactivator-1β (PGC-1β). M2 exhibit enhanced PPAR-γ activity. PPAR-γ forms a heterodimer with RXR to induce OXHOS regulating proteins. PGC-1β promotes the STAT6 transcription complex and expression of M2 proteins such as arginase-1 and the pattern-recognition/endocytic receptor CD206. The STAT6 transcription complex further increases transcription of TCA cycle enzymes such as citrate synthase, CD36 (fatty acid transporter), ATP synthase, 12/15 lipoxygenase (converts omega-3 PUFA to SPMs), and LCAD (long chain acyl dehydrogenase, an enzyme for β-oxidation). Adiponectin activates PPARs in M2 macrophages and consequently increases β-oxidation.
Figure 5
Figure 5
Tryptophan metabolism mediated immune cell response. L-tryptophan (Trp) can only be obtained in the human body from dietary sources. It is catabolized by the enzyme indoleamine-2,3-dioxygenase-1 (IDO-1) to L-kynurenine (Kyn). Interferon-γ produced by Th1 cells stimulates IDO-1 production in antigen-presenting cells (APCs). Tumor cells are also known to have increased production of IDO-1. Increases in IDO-1 cause conversion of Trp to Kyn resulting in lower Trp and higher Kyn in plasma. Kyn is a ligand for aryl hydrocarbon receptor (AhR). Mitochondrial AhR (MitoAhR) is localized in the inner membrane space of mitochondria and is also affected by Kyn. (1) Immune tolerance by kynurenine: increased Kyn causes immune tolerance by inhibiting proliferation of both T cells and NK cells by inhibiting maturation of dendritic cells and increasing proliferation of immune suppresser cells (Treg and MDSC). Tryptophan depletion also causes apoptosis of T cells due to nutrient deficiency. (2) NAD+ mediated immune function: in addition, Trp is one of the sources for NAD+ biosynthesis. NAD+ is an important metabolite for metabolic processes [glycolysis, TCA cycle, and oxidative phosphorylation (OXPHOS)] and production of NADPH. During oxidative burst in macrophages, NAPPH is oxidized in the presence of NADPH oxidase to produce superoxide which is utilized to kill pathogens. NAD+ contributes to CD4+ T cell differentiation and SIRT3-mediated mitochondrial biogenesis. (3) mTOR inhibitor rapamycin can be used to treat systemic lupus erythematosus (SLE) and mitochondrial disease: T cell dysfunction is associated with SLE. In SLE, an increase in kynurenine and kynurenine mediated mTOR stimulation was observed. In addition, accumulation of reactive oxygen species (ROS) producing mitochondria and decreased Drp1 are associated with SLE. Rapamycin can increase Drp1 by inhibiting mTORC1. (4) Rapamycin inhibits mitochondrial ROS (mtROS): rapamycin inhibits mtROS through several mechanisms. Complex I in mitochondria is a source of ROS generation. Rapamycin can inhibit mtROS production by inhibiting complex I formation. In addition to this, rapamycin can inhibit mtROS by inducing autophagy and activating p62 and the NRF2 pathway.
Figure 6
Figure 6
Mitochondria related signal cascades to activate immunity. Mitochondrial antiviral signaling (MAVS) is activated by the viral RNA sensor RIG-I and promotes inflammatory and immune gene expression via transcription factor NF-κB and IRFs. MAVS receptors can be found in the mitochondrial outer membrane. Mitochondrial ROS (mtROS) are also able to activate MAVS independently of RNA. The NLRP3 inflammasome also signals from the outer mitochondrial membrane. Like MAVS, NLRP3 also responds to mtROS and can cause mitochondrial damage that promotes the generation of additional ROS. NLRP3 drives the production of IL-1β and IL-18. Mitochondrial DNA can also activate NLRP3. Finally, MAVS promotes oligomerization of NLRP3 at the mitochondria. Therefore, two major innate immune pathways (RIG-I/MAVS and NLRP3) are all dependent on mitochondria.
Figure 7
Figure 7
Mitochondrial fission and fusion contribute to immune responses. Mitochondrial fission and fusion controls mitochondrial mass. Nutrient deprivation induces an increase in mitochondrial fusion and suppression of mitophagy. On the other hand, prolonged DNA damage leads to mitochondrial fission. Fusion (tubular mitochondria) is controlled by two dynamin-like GTPases for the outer membrane proteins (mitofusin MFN1 and MFN2) and inner membrane proteins (OPA1). Fission (fragmented mitochondria) is coordinated by four outer membrane proteins (Drp1, Mff, Mid49, and mid52). Fusion and fission of mitochondria contribute to immune response. (1) Memory T cells have increased fusion generating more oxidative phosphorylation (OXPHOS) and FAO. (2) Clearance of apoptotic cells by phagocytes is known as efferocytosis. Resting macrophages have displayed tubular mitochondria. (3) Fission contributes to mitochondria mobility. During antigen-specific T cell and NK cell activation, mitochondria fission increased, and mitochondria move toward the immune synapse (IS) and facilitate ATP production at the IS. (4) Uptake of apoptotic cells by macrophages during efferocytosis requires fission and increased Drp1. (5) In mouse-derived bone marrow macrophages, knocking down Drp1 initiates NLRP3 inflammasomal activation, and inducing mitochondrial fission attenuated NLRP3 inflammasomal assembly and activation. (6) Activated effector T cells have increased mitochondrial fission and exhibit higher rates of glycolysis.
Figure 8
Figure 8
Reactive oxygen species (ROS)-mediated cell signaling in T cells: low/moderate levels of ROS promote mTOR, Myc, and nuclear factor of activated T cell (NFAT) activation which is associated with increased glycolysis and inflammatory response in T cells. ROS levels are regulated by the glutathione pathway (GSH) and NRF2–Keap1–Cul3 trimeric complex. Abrogation of GSH and the trimeric complex, which may cause high levels of ROS, impairs the inflammatory response of T cells. In addition, high ROS and high adenosine monophosphate-activated protein kinase (AMPK) levels inhibit mTOR. It is reported that naïve T cells, CD4+ regulatory T cells, and chronically stimulated T cells depend on OXPHOS as their primary metabolism method while CD8+ effector and CD4+ Th1 cells mainly depend on glycolysis. This metabolic switching is controlled by ROS.
Figure 9
Figure 9
Antioxidants neutralize mitochondrial ROS (mtROS). (1) Antioxidants in mitochondria: superoxide (O2−) is released from complex I (into the matrix) and III (into both the matrix and intermembrane space). This is converted to H2O2 in the presence of superoxide dismutase (SOD). Mn-SOD (manganese dismutase) is localized in mitochondria matrix and Cu/Zn-SOD (in the intermembrane space and cytosol). SIRT3 is involved in deacetylation of MnSOD. Catalase converts H2O2 to H2O. Catalase is also localized in mitochondria (MCAT). Glutathione converts OH− to H2O to reduce the cellular damage. Glutathione (GSH) is oxidized and reduced to glutathione disulfide (GSSG) in a reversible reaction. NADPH from the pentose phosphate pathway reduces GSSG to form GSH in presence of glutathione peroxidase (Gpx). (2) Antioxidants play an important role in neutralizing mtROS. Uncoupling protein 2 (UCP2) (mitochondrial uncoupler protein), which helps to shuttle H+ ions from the intermembrane space to the mitochondria matrix, decreases the membrane potential to inhibit excessive reactive oxygen species (ROS) production. Excessive ROS production can impact ATP synthesis, cytokine production, mitochondrial DNA (mtDNA) mutation, and post-translational modification of cellular proteins by S-nitosylation or glutathionylation.

References

    1. Grimm S. The ER-mitochondria interface: the social network of cell death. Biochim Biophys Acta (2012) 1823(2):327–34.10.1016/j.bbamcr.2011.11.018
    1. Zhou R, Yazdi AS, Menu P, Tschopp J. A role for mitochondria in NLRP3 inflammasome activation. Nature (2011) 469(7329):221–5.10.1038/nature09663
    1. Bantug GR, Fischer M, Grählert J, Balmer ML, Unterstab G, Develioglu L, et al. Mitochondria-endoplasmic reticulum contact sites function as immunometabolic hubs that orchestrate the rapid recall response of memory CD8(+) T cells. Immunity (2018) 48(3):542–555.e6.10.1016/j.immuni.2018.02.012
    1. Pearce EL, Pearce EJ. Metabolic pathways in immune cell activation and quiescence. Immunity (2013) 38(4):633–43.10.1016/j.immuni.2013.04.005
    1. Carr EL, Kelman A, Wu GS, Gopaul R, Senkevitch E, Aghvanyan A, et al. Glutamine uptake and metabolism are coordinately regulated by ERK/MAPK during T lymphocyte activation. J Immunol (2010) 185(2):1037–44.10.4049/jimmunol.0903586
    1. Le A, Lane AN, Hamaker M, Bose S, Gouw A, Barbi J, et al. Glucose-independent glutamine metabolism via TCA cycling for proliferation and survival in B cells. Cell Metab (2012) 15(1):110–21.10.1016/j.cmet.2011.12.009
    1. Wynn TA, Chawla A, Pollard JW. Macrophage biology in development, homeostasis and disease. Nature (2013) 496(7446):445–55.10.1038/nature12034
    1. He C, Carter AB. The metabolic prospective and redox regulation of macrophage polarization. J Clin Cell Immunol (2015) 6(6):371.10.4172/2155-9899.1000371
    1. De Simone R, Ajmone-Cat MA, Pandolfi M, Bernardo A, De Nuccio C, Minghetti L, et al. The mitochondrial uncoupling protein-2 is a master regulator of both M1 and M2 microglial responses. J Neurochem (2015) 135(1):147–56.10.1111/jnc.13244
    1. Mills EL, O’Neill LA. Reprogramming mitochondrial metabolism in macrophages as an anti-inflammatory signal. Eur J Immunol (2016) 46(1):13–21.10.1002/eji.201445427
    1. Shapiro H, Lutaty A, Ariel A. Macrophages, meta-inflammation, and immuno-metabolism. ScientificWorldJournal (2011) 11:2509–29.10.1100/2011/397971
    1. Phan K, Ng W, Lu VM, McDonald KL, Fairhall J, Reddy R, et al. Association between IDH1 and IDH2 mutations and preoperative seizures in patients with low-grade versus high-grade glioma: a systematic review and meta-analysis. World Neurosurg (2018) 111:e539–45.10.1016/j.wneu.2017.12.112
    1. Kong MJ, Han SJ, Kim JI, Park JW, Park KM. Mitochondrial NADP(+)-dependent isocitrate dehydrogenase deficiency increases cisplatin-induced oxidative damage in the kidney tubule cells. Cell Death Dis (2018) 9(5):488.10.1038/s41419-018-0537-6
    1. Jha AK, Huang SC, Sergushichev A, Lampropoulou V, Ivanova Y, Loginicheva E, et al. Network integration of parallel metabolic and transcriptional data reveals metabolic modules that regulate macrophage polarization. Immunity (2015) 42(3):419–30.10.1016/j.immuni.2015.02.005
    1. Tannahill GM, Curtis AM, Adamik J, Palsson-McDermott EM, McGettrick AF, Goel G, et al. Succinate is an inflammatory signal that induces IL-1beta through HIF-1alpha. Nature (2013) 496(7444):238–42.10.1038/nature11986
    1. Tan Z, Xie N, Cui H, Moellering DR, Abraham E, Thannickal VJ, et al. Pyruvate dehydrogenase kinase 1 participates in macrophage polarization via regulating glucose metabolism. J Immunol (2015) 194(12):6082–9.10.4049/jimmunol.1402469
    1. Coe DJ, Kishore M, Marelli-Berg F. Metabolic regulation of regulatory T cell development and function. Front Immunol (2014) 5:590.10.3389/fimmu.2014.00590
    1. Sena LA, Li S, Jairaman A, Prakriya M, Ezponda T, Hildeman DA, et al. Mitochondria are required for antigen-specific T cell activation through reactive oxygen species signaling. Immunity (2013) 38(2):225–36.10.1016/j.immuni.2012.10.020
    1. Arora R, Schmitt D, Karanam B, Tan M, Yates C, Dean-Colomb W. Inhibition of the Warburg effect with a natural compound reveals a novel measurement for determining the metastatic potential of breast cancers. Oncotarget (2015) 6(2):662–78.10.18632/oncotarget.2689
    1. Fischer K, Hoffmann P, Voelkl S, Meidenbauer N, Ammer J, Edinger M, et al. Inhibitory effect of tumor cell-derived lactic acid on human T cells. Blood (2007) 109(9):3812–9.10.1182/blood-2006-07-035972
    1. Colegio OR, Chu NQ, Szabo AL, Chu T, Rhebergen AM, Jairam V, et al. Functional polarization of tumour-associated macrophages by tumour-derived lactic acid. Nature (2014) 513(7519):559–63.10.1038/nature13490
    1. Buck MD, Sowell RT, Kaech SM, Pearce EL. Metabolic instruction of immunity. Cell (2017) 169(4):570–86.10.1016/j.cell.2017.04.004
    1. Selleri S, Bifsha P, Civini S, Pacelli C, Dieng MM, Lemieux W, et al. Human mesenchymal stromal cell-secreted lactate induces M2-macrophage differentiation by metabolic reprogramming. Oncotarget (2016) 7(21):30193–210.10.18632/oncotarget.8623
    1. Marino A, Menghini R, Fabrizi M, Casagrande V, Mavilio M, Stoehr R, et al. ITCH deficiency protects from diet-induced obesity. Diabetes (2014) 63(2):550–61.10.2337/db13-0802
    1. Adamson S, Leitinger N. Phenotypic modulation of macrophages in response to plaque lipids. Curr Opin Lipidol (2011) 22(5):335–42.10.1097/MOL.0b013e32834a97e4
    1. Rombaldova M, Janovska P, Kopecky J, Kuda O. Omega-3 fatty acids promote fatty acid utilization and production of pro-resolving lipid mediators in alternatively activated adipose tissue macrophages. Biochem Biophys Res Commun (2017) 490(3):1080–5.10.1016/j.bbrc.2017.06.170
    1. Ayers S, Switnicki MP, Angajala A, Lammel J, Arumanayagam AS, Webb P. Genome-wide binding patterns of thyroid hormone receptor beta. PLoS One (2014) 9(2):e81186.10.1371/journal.pone.0081186
    1. Lammel Lindemann JA, Angajala A, Engler DA, Webb P, Ayers SD. Thyroid hormone induction of human cholesterol 7 alpha-hydroxylase (Cyp7a1) in vitro. Mol Cell Endocrinol (2014) 388(1–2):32–40.10.1016/j.mce.2014.02.003
    1. Zubair H, Azim S, Srivastava SK, Ahmad A, Bhardwaj A, Aslam Khan M, et al. Glucose metabolism reprogrammed by overexpression of ikkepsilon promotes pancreatic tumor growth. Cancer Res (2016) 76(24):7254–64.10.1158/0008-5472.CAN-16-1666
    1. Vats D, Mukundan L, Odegaard JI, Zhang L, Smith KL, Morel CR, et al. Oxidative metabolism and PGC-1beta attenuate macrophage-mediated inflammation. Cell Metab (2006) 4(1):13–24.10.1016/j.cmet.2006.08.006
    1. Pearce EL, Walsh MC, Cejas PJ, Harms GM, Shen H, Wang LS, et al. Enhancing CD8 T-cell memory by modulating fatty acid metabolism. Nature (2009) 460(7251):103–7.10.1038/nature08097
    1. O’Sullivan D, van der Windt GJ, Huang SC, Curtis JD, Chang CH, Buck MD, et al. Memory CD8(+) T cells use cell-intrinsic lipolysis to support the metabolic programming necessary for development. Immunity (2014) 41(1):75–88.10.1016/j.immuni.2014.06.005
    1. Cui G, Staron MM, Gray SM, Ho PC, Amezquita RA, Wu J, et al. IL-7-induced glycerol transport and TAG synthesis promotes memory CD8+ T cell longevity. Cell (2015) 161(4):750–61.10.1016/j.cell.2015.03.021
    1. Michalek RD, Gerriets VA, Jacobs SR, Macintyre AN, MacIver NJ, Mason EF, et al. Cutting edge: distinct glycolytic and lipid oxidative metabolic programs are essential for effector and regulatory CD4+ T cell subsets. J Immunol (2011) 186(6):3299–303.10.4049/jimmunol.1003613
    1. Wu D, Sanin DE, Everts B, Chen Q, Qiu J, Buck MD, et al. Type 1 interferons induce changes in core metabolism that are critical for immune function. Immunity (2016) 44(6):1325–36.10.1016/j.immuni.2016.06.006
    1. Geiger R, Rieckmann JC, Wolf T, Basso C, Feng Y, Fuhrer T, et al. L-arginine modulates T cell metabolism and enhances survival and anti-tumor activity. Cell (2016) 167(3):829–842.e13.10.1016/j.cell.2016.09.031
    1. Ron-Harel N, Santos D, Ghergurovich JM, Sage PT, Reddy A, Lovitch SB, et al. Mitochondrial biogenesis and proteome remodeling promote one-carbon metabolism for T cell activation. Cell Metab (2016) 24(1):104–17.10.1016/j.cmet.2016.06.007
    1. Gostner JM, Becker K, Kofler H, Strasser B, Fuchs D. Tryptophan metabolism in allergic disorders. Int Arch Allergy Immunol (2016) 169(4):203–15.10.1159/000445500
    1. Soliman H, Mediavilla-Varela M, Antonia S. Indoleamine 2,3-dioxygenase: is it an immune suppressor? Cancer J (2010) 16(4):354–9.10.1097/PPO.0b013e3181eb3343
    1. Strong MJ, Xu G, Coco J, Baribault C, Vinay DS, Lacey MR, et al. Differences in gastric carcinoma microenvironment stratify according to EBV infection intensity: implications for possible immune adjuvant therapy. PLoS Pathog (2013) 9(5):e1003341.10.1371/journal.ppat.1003341
    1. Ma N, Guo P, Zhang J, He T, Kim SW, Zhang G, et al. Nutrients mediate intestinal bacteria-mucosal immune crosstalk. Front Immunol (2018) 9:5.10.3389/fimmu.2018.00005
    1. Vera JF, Brenner MK, Dotti G. Immunotherapy of human cancers using gene modified T lymphocytes. Curr Gene Ther (2009) 9(5):396–408.10.2174/156652309789753338
    1. Jürgens B, Raberger J, Fuchs D, Heitger A. Indoleamine 2,3-dioxygenase in human hematopoietic stem cell transplantation. Int J Tryptophan Res (2010) 3:77–90.10.4137/IJTR.S4076
    1. Rodriguez Cetina Biefer H, Vasudevan A, Elkhal A. Aspects of tryptophan and nicotinamide adenine dinucleotide in immunity: a new twist in an old tale. Int J Tryptophan Res (2017) 10:1178646917713491.10.1177/1178646917713491
    1. Hwang HJ, Dornbos P, Steidemann M, Dunivin TK, Rizzo M, LaPres JJ. Mitochondrial-targeted aryl hydrocarbon receptor and the impact of 2,3,7,8-tetrachlorodibenzo-p-dioxin on cellular respiration and the mitochondrial proteome. Toxicol Appl Pharmacol (2016) 304:121–32.10.1016/j.taap.2016.04.005
    1. Tullius SG, Biefer HR, Li S, Trachtenberg AJ, Edtinger K, Quante M, et al. NAD+ protects against EAE by regulating CD4+ T-cell differentiation. Nat Commun (2014) 5:5101.10.1038/ncomms6101
    1. Duan W. Sirtuins: from metabolic regulation to brain aging. Front Aging Neurosci (2013) 5:36.10.3389/fnagi.2013.00036
    1. Wakino S, Hasegawa K, Itoh H. Sirtuin and metabolic kidney disease. Kidney Int (2015) 88(4):691–8.10.1038/ki.2015.157
    1. Perl A, Hanczko R, Lai ZW, Oaks Z, Kelly R, Borsuk R, et al. Comprehensive metabolome analyses reveal N-acetylcysteine-responsive accumulation of kynurenine in systemic lupus erythematosus: implications for activation of the mechanistic target of rapamycin. Metabolomics (2015) 11(5):1157–74.10.1007/s11306-015-0772-0
    1. Lai ZW, Kelly R, Winans T, Marchena I, Shadakshari A, Yu J, 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 (2018) 391(10126):1186–96.10.1016/S0140-6736(18)30485-9
    1. Khan NA, Nikkanen J, Yatsuga S, Jackson C, Wang L, Pradhan S, et al. mTORC1 regulates mitochondrial integrated stress response and mitochondrial myopathy progression. Cell Metab (2017) 26(2):419–428.e5.10.1016/j.cmet.2017.07.007
    1. Miwa S, Jow H, Baty K, Johnson A, Czapiewski R, Saretzki G, et al. Low abundance of the matrix arm of complex I in mitochondria predicts longevity in mice. Nat Commun (2014) 5:3837.10.1038/ncomms4837
    1. Mimaki M, Wang X, McKenzie M, Thorburn DR, Ryan MT. Understanding mitochondrial complex I assembly in health and disease. Biochim Biophys Acta (2012) 1817(6):851–62.10.1016/j.bbabio.2011.08.010
    1. Ko JH, Yoon SO, Lee HJ, Oh JY. Rapamycin regulates macrophage activation by inhibiting NLRP3 inflammasome-p38 MAPK-NFkappaB pathways in autophagy- and p62-dependent manners. Oncotarget (2017) 8(25):40817–31.10.18632/oncotarget.17256
    1. Liu PS, Wang H, Li X, Chao T, Teav T, Christen S, et al. alpha-ketoglutarate orchestrates macrophage activation through metabolic and epigenetic reprogramming. Nat Immunol (2017) 18(9):985–94.10.1038/ni.3796
    1. Palmieri EM, Menga A, Martín-Pérez R, Quinto A, Riera-Domingo C, De Tullio G, et al. Pharmacologic or genetic targeting of glutamine synthetase skews macrophages toward an M1-like phenotype and inhibits tumor metastasis. Cell Rep (2017) 20(7):1654–66.10.1016/j.celrep.2017.07.054
    1. He L, Weber KJ, Schilling JD. Glutamine modulates macrophage lipotoxicity. Nutrients (2016) 8(4):215.10.3390/nu8040215
    1. He C, Ryan AJ, Murthy S, Carter AB. Accelerated development of pulmonary fibrosis via Cu,Zn-superoxide dismutase-induced alternative activation of macrophages. J Biol Chem (2013) 288(28):20745–57.10.1074/jbc.M112.410720
    1. West AP, Brodsky IE, Rahner C, Woo DK, Erdjument-Bromage H, Tempst P, et al. TLR signalling augments macrophage bactericidal activity through mitochondrial ROS. Nature (2011) 472(7344):476–80.10.1038/nature09973
    1. Rousset S, Emre Y, Join-Lambert O, Hurtaud C, Ricquier D, Cassard-Doulcier AM. The uncoupling protein 2 modulates the cytokine balance in innate immunity. Cytokine (2006) 35(3–4):135–42.10.1016/j.cyto.2006.07.012
    1. Dostert C, Pétrilli V, Van Bruggen R, Steele C, Mossman BT, Tschopp J. Innate immune activation through Nalp3 inflammasome sensing of asbestos and silica. Science (2008) 320(5876):674–7.10.1126/science.1156995
    1. Nakahira K, Haspel JA, Rathinam VA, Lee SJ, Dolinay T, Lam HC, et al. Autophagy proteins regulate innate immune responses by inhibiting the release of mitochondrial DNA mediated by the NALP3 inflammasome. Nat Immunol (2011) 12(3):222–30.10.1038/ni.1980
    1. Li N, Ragheb K, Lawler G, Sturgis J, Rajwa B, Melendez JA, et al. Mitochondrial complex I inhibitor rotenone induces apoptosis through enhancing mitochondrial reactive oxygen species production. J Biol Chem (2003) 278(10):8516–25.10.1074/jbc.M210432200
    1. Kamiński MM, Sauer SW, Kamiński M, Opp S, Ruppert T, Grigaravičius P, et al. T cell activation is driven by an ADP-dependent glucokinase linking enhanced glycolysis with mitochondrial reactive oxygen species generation. Cell Rep (2012) 2(5):1300–15.10.1016/j.celrep.2012.10.009
    1. Watanabe-Matsui M, Muto A, Matsui T, Itoh-Nakadai A, Nakajima O, Murayama K, et al. 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):5438–48.10.1182/blood-2010-07-296483
    1. Jang KJ, Mano H, Aoki K, Hayashi T, Muto A, Nambu Y, et al. Mitochondrial function provides instructive signals for activation-induced B-cell fates. Nat Commun (2015) 6:6750.10.1038/ncomms7750
    1. Narendra DP, Jin SM, Tanaka A, Suen DF, Gautier CA, Shen J, et al. PINK1 is selectively stabilized on impaired mitochondria to activate Parkin. PLoS Biol (2010) 8(1):e1000298.10.1371/journal.pbio.1000298
    1. Kirienko NV, Ausubel FM, Ruvkun G. Mitophagy confers resistance to siderophore-mediated killing by Pseudomonas aeruginosa. Proc Natl Acad Sci U S A (2015) 112(6):1821–6.10.1073/pnas.1424954112
    1. Mira MT, Alcaïs A, Nguyen VT, Moraes MO, Di Flumeri C, Vu HT, et al. Susceptibility to leprosy is associated with PARK2 and PACRG. Nature (2004) 427(6975):636–40.10.1038/nature02326
    1. Malhotra D, Darvishi K, Lohra M, Kumar H, Grover C, Sood S, et al. Association study of major risk single nucleotide polymorphisms in the common regulatory region of PARK2 and PACRG genes with leprosy in an Indian population. Eur J Hum Genet (2006) 14(4):438–42.10.1038/sj.ejhg.5201563
    1. Ali S, Vollaard AM, Widjaja S, Surjadi C, van de Vosse E, van Dissel JT. PARK2/PACRG polymorphisms and susceptibility to typhoid and paratyphoid fever. Clin Exp Immunol (2006) 144(3):425–31.10.1111/j.1365-2249.2006.03087.x
    1. Manzanillo PS, Ayres JS, Watson RO, Collins AC, Souza G, Rae CS, et al. The ubiquitin ligase parkin mediates resistance to intracellular pathogens. Nature (2013) 501(7468):512–6.10.1038/nature12566
    1. Pua HH, Guo J, Komatsu M, He YW. Autophagy is essential for mitochondrial clearance in mature T lymphocytes. J Immunol (2009) 182(7):4046–55.10.4049/jimmunol.0801143
    1. Willinger T, Flavell RA. Canonical autophagy dependent on the class III phosphoinositide-3 kinase Vps34 is required for naive T-cell homeostasis. Proc Natl Acad Sci U S A (2012) 109(22):8670–5.10.1073/pnas.1205305109
    1. Kim SJ, Syed GH, Siddiqui A. Hepatitis C virus induces the mitochondrial translocation of Parkin and subsequent mitophagy. PLoS Pathog (2013) 9(3):e1003285.10.1371/journal.ppat.1003285
    1. Kim SJ, Khan M, Quan J, Till A, Subramani S, Siddiqui A. Hepatitis B virus disrupts mitochondrial dynamics: induces fission and mitophagy to attenuate apoptosis. PLoS Pathog (2013) 9(12):e1003722.10.1371/journal.ppat.1003722
    1. Shao BZ, Xu ZQ, Han BZ, Su DF, Liu C. NLRP3 inflammasome and its inhibitors: a review. Front Pharmacol (2015) 6:262.10.3389/fphar.2015.00262
    1. Shimada K, Crother TR, Karlin J, Dagvadorj J, Chiba N, Chen S, et al. Oxidized mitochondrial DNA activates the NLRP3 inflammasome during apoptosis. Immunity (2012) 36(3):401–14.10.1016/j.immuni.2012.01.009
    1. Allam R, Lawlor KE, Yu EC, Mildenhall AL, Moujalled DM, Lewis RS, et al. Mitochondrial apoptosis is dispensable for NLRP3 inflammasome activation but non-apoptotic caspase-8 is required for inflammasome priming. EMBO Rep (2014) 15(9):982–90.10.15252/embr.201438463
    1. Zhou QF, Xu SM, Wang HQ, Xing LM, Fu R, Shao ZH. [Single nucleotide polymorphism of mitochondrial DNA D-LOOP region in peripheral blood lymphocytes of immuno-related pancytopenia patients]. Zhongguo Shi Yan Xue Ye Xue Za Zhi (2017) 25(1):186–91.10.7534/j.issn.1009-2137.2017.01.033
    1. Seth RB, Sun L, Ea CK, Chen ZJ. Identification and characterization of MAVS, a mitochondrial antiviral signaling protein that activates NF-kappaB and IRF 3. Cell (2005) 122(5):669–82.10.1016/j.cell.2005.08.012
    1. Buskiewicz IA, Montgomery T, Yasewicz EC, Huber SA, Murphy MP, Hartley RC, et al. Reactive oxygen species induce virus-independent MAVS oligomerization in systemic lupus erythematosus. Sci Signal (2016) 9(456):ra115.10.1126/scisignal.aaf1933
    1. Park S, Juliana C, Hong S, Datta P, Hwang I, Fernandes-Alnemri T, et al. The mitochondrial antiviral protein MAVS associates with NLRP3 and regulates its inflammasome activity. J Immunol (2013) 191(8):4358–66.10.4049/jimmunol.1301170
    1. Hee JS, Cresswell P. Viperin interaction with mitochondrial antiviral signaling protein (MAVS) limits viperin-mediated inhibition of the interferon response in macrophages. PLoS One (2017) 12(2):e0172236.10.1371/journal.pone.0172236
    1. Chao T, Wang H, Ho PC. Mitochondrial control and guidance of cellular activities of T cells. Front Immunol (2017) 8:473.10.3389/fimmu.2017.00473
    1. Westermann B. Molecular machinery of mitochondrial fusion and fission. J Biol Chem (2008) 283(20):13501–5.10.1074/jbc.R800011200
    1. Zorzano A, Sebastián D, Segalés J, Palacín M. The molecular machinery of mitochondrial fusion and fission: an opportunity for drug discovery? Curr Opin Drug Discov Devel (2009) 12(5):597–606.
    1. Khan M, Syed GH, Kim SJ, Siddiqui A. Mitochondrial dynamics and viral infections: a close nexus. Biochim Biophys Acta (2015) 1853(10 Pt B):2822–33.10.1016/j.bbamcr.2014.12.040
    1. Quintana A, Hoth M. Mitochondrial dynamics and their impact on T cell function. Cell Calcium (2012) 52(1):57–63.10.1016/j.ceca.2012.02.005
    1. Baixauli F, Martín-Cófreces NB, Morlino G, Carrasco YR, Calabia-Linares C, Veiga E, et al. The mitochondrial fission factor dynamin-related protein 1 modulates T-cell receptor signalling at the immune synapse. EMBO J (2011) 30(7):1238–50.10.1038/emboj.2011.25
    1. Abarca-Rojano E, Muñiz-Hernández S, Moreno-Altamirano MM, Mondragón-Flores R, Enriquez-Rincón F, Sánchez-García FJ. Re-organization of mitochondria at the NK cell immune synapse. Immunol Lett (2009) 122(1):18–25.10.1016/j.imlet.2008.10.008
    1. Junger WG. Immune cell regulation by autocrine purinergic signalling. Nat Rev Immunol (2011) 11(3):201–12.10.1038/nri2938
    1. Singaravelu K, Nelson C, Bakowski D, de Brito OM, Ng SW, Di Capite J, et al. Mitofusin 2 regulates STIM1 migration from the Ca2+ store to the plasma membrane in cells with depolarized mitochondria. J Biol Chem (2011) 286(14):12189–201.10.1074/jbc.M110.174029
    1. Wang Y, Subramanian M, Yurdagul A, Jr, Barbosa-Lorenzi VC, Cai B, de Juan-Sanz J, et al. Mitochondrial fission promotes the continued clearance of apoptotic cells by macrophages. Cell (2017) 171(2):331–345.e22.10.1016/j.cell.2017.08.041
    1. Cid-Castro C, Hernandez-Espinosa DR, Moran J. ROS as regulators of mitochondrial dynamics in neurons. Cell Mol Neurobiol (2018) 38(5):995–1007.10.1007/s10571-018-0584-7
    1. Zheng Q, Zhao Y, Guo J, Zhao S, Fei C, Xiao C, et al. Iron overload promotes mitochondrial fragmentation in mesenchymal stromal cells from myelodysplastic syndrome patients through activation of the AMPK/MFF/Drp1 pathway. Cell Death Dis (2018) 9(5):515.10.1038/s41419-018-0552-7
    1. Tsushima K, Bugger H, Wende AR, Soto J, Jenson GA, Tor AR, et al. Mitochondrial reactive oxygen species in lipotoxic hearts induce post-translational modifications of AKAP121, DRP1, and OPA1 that promote mitochondrial fission. Circ Res (2018) 122(1):58–73.10.1161/CIRCRESAHA.117.311307
    1. Li N, Wang H, Jiang C, Zhang M. Renal ischemia/reperfusion-induced mitophagy protects against renal dysfunction via Drp1-dependent-pathway. Exp Cell Res (2018) 369(1):27–33.10.1016/j.yexcr.2018.04.025
    1. Caza TN, Fernandez DR, Talaber G, Oaks Z, Haas M, Madaio MP, et al. HRES-1/Rab4-mediated depletion of Drp1 impairs mitochondrial homeostasis and represents a target for treatment in SLE. Ann Rheum Dis (2014) 73(10):1888–97.10.1136/annrheumdis-2013-203794
    1. Oaks Z, Winans T, Caza T, Fernandez D, Liu Y, Landas SK, et al. Mitochondrial dysfunction in the liver and antiphospholipid antibody production precede disease onset and respond to rapamycin in lupus-prone mice. Arthritis Rheumatol (2016) 68(11):2728–39.10.1002/art.39791
    1. Buck MD, O’Sullivan D, Klein Geltink RI, Curtis JD, Chang CH, Sanin DE, et al. Mitochondrial dynamics controls T cell fate through metabolic programming. Cell (2016) 166(1):63–76.10.1016/j.cell.2016.05.035
    1. Kang YJ, Bang BR, Han KH, Hong L, Shim EJ, Ma J, et al. Regulation of NKT cell-mediated immune responses to tumours and liver inflammation by mitochondrial PGAM5-Drp1 signalling. Nat Commun (2015) 6:8371.10.1038/ncomms9371
    1. Liu P, Xie Q, Wei T, Chen Y, Chen H, Shen W. Activation of the NLRP3 inflammasome induces vascular dysfunction in obese OLETF rats. Biochem Biophys Res Commun (2015) 468(1–2):319–25.10.1016/j.bbrc.2015.10.105
    1. Park S, Won JH, Hwang I, Hong S, Lee HK, Yu JW. Defective mitochondrial fission augments NLRP3 inflammasome activation. Sci Rep (2015) 5:15489.10.1038/srep15489
    1. Zapico SC, Ubelaker DH. mtDNA mutations and their role in aging, diseases and forensic sciences. Aging Dis (2013) 4(6):364–80.10.14336/AD.2013.0400364
    1. Ribas V, Garcia-Ruiz C, Fernandez-Checa JC. Glutathione and mitochondria. Front Pharmacol (2014) 5:151.10.3389/fphar.2014.00151
    1. Acosta MJ, Vazquez Fonseca L, Desbats MA, Cerqua C, Zordan R, Trevisson E, et al. Coenzyme Q biosynthesis in health and disease. Biochim Biophys Acta (2016) 1857(8):1079–85.10.1016/j.bbabio.2016.03.036
    1. Quijano C, Trujillo M, Castro L, Trostchansky A. Interplay between oxidant species and energy metabolism. Redox Biol (2016) 8:28–42.10.1016/j.redox.2015.11.010
    1. Fraternale A, Brundu S, Magnani M. Glutathione and glutathione derivatives in immunotherapy. Biol Chem (2017) 398(2):261–75.10.1515/hsz-2016-0202
    1. Mensurado S, Rei M, Lança T, Ioannou M, Gonçalves-Sousa N, Kubo H, et al. Tumor-associated neutrophils suppress pro-tumoral IL-17+ gammadelta T cells through induction of oxidative stress. PLoS Biol (2018) 16(5):e2004990.10.1371/journal.pbio.2004990
    1. Allen M, Bailey C, Cahatol I, Dodge L, Yim J, Kassissa C, et al. Mechanisms of control of Mycobacterium tuberculosis by NK cells: role of glutathione. Front Immunol (2015) 6:508.10.3389/fimmu.2015.00508
    1. Valdivia A, Ly J, Gonzalez L, Hussain P, Saing T, Islamoglu H, et al. Restoring cytokine balance in HIV-positive individuals with low CD4 T cell counts. AIDS Res Hum Retroviruses (2017) 33(9):905–18.10.1089/AID.2016.0303
    1. Short JD, Downs K, Tavakoli S, Asmis R. Protein thiol redox signaling in monocytes and macrophages. Antioxid Redox Signal (2016) 25(15):816–35.10.1089/ars.2016.6697
    1. Allen EM, Mieyal JJ. Protein-thiol oxidation and cell death: regulatory role of glutaredoxins. Antioxid Redox Signal (2012) 17(12):1748–63.10.1089/ars.2012.4644
    1. Yarosz EL, Chang CH. The role of reactive oxygen species in regulating T cell-mediated immunity and disease. Immune Netw (2018) 18(1):e14.10.4110/in.2018.18.e14
    1. Gerriets VA, Rathmell JC. Metabolic pathways in T cell fate and function. Trends Immunol (2012) 33(4):168–73.10.1016/j.it.2012.01.010
    1. Beckermann KE, Dudzinski SO, Rathmell JC. Dysfunctional T cell metabolism in the tumor microenvironment. Cytokine Growth Factor Rev (2017) 35:7–14.10.1016/j.cytogfr.2017.04.003
    1. Franchina DG, Dostert C, Brenner D. Reactive oxygen species: involvement in T cell signaling and metabolism. Trends Immunol (2018) 39(6):489–502.10.1016/j.it.2018.01.005
    1. Fracchia KM, Pai CY, Walsh CM. Modulation of T cell metabolism and function through calcium signaling. Front Immunol (2013) 4:324.10.3389/fimmu.2013.00324
    1. Kouidhi S, Elgaaied AB, Chouaib S. Impact of metabolism on T-cell differentiation and function and cross talk with tumor microenvironment. Front Immunol (2017) 8:270.10.3389/fimmu.2017.00270
    1. Previte DM, O’Connor EC, Novak EA, Martins CP, Mollen KP, Piganelli JD. Reactive oxygen species are required for driving efficient and sustained aerobic glycolysis during CD4+ T cell activation. PLoS One (2017) 12(4):e0175549.10.1371/journal.pone.0175549
    1. Sivitz WI, Yorek MA. Mitochondrial dysfunction in diabetes: from molecular mechanisms to functional significance and therapeutic opportunities. Antioxid Redox Signal (2010) 12(4):537–77.10.1089/ars.2009.2531
    1. Perl A, Gergely P, Jr, Puskas F, Banki K. Metabolic switches of T-cell activation and apoptosis. Antioxid Redox Signal (2002) 4(3):427–43.10.1089/15230860260196227
    1. Laskowski M, Augustynek B, Kulawiak B, Koprowski P, Bednarczyk P, Jarmuszkiewicz W, et al. What do we not know about mitochondrial potassium channels? Biochim Biophys Acta (2016) 1857(8):1247–57.10.1016/j.bbabio.2016.03.007
    1. Gergely P, Jr, Grossman C, Niland B, Puskas F, Neupane H, Allam F, et al. Mitochondrial hyperpolarization and ATP depletion in patients with systemic lupus erythematosus. Arthritis Rheum (2002) 46(1):175–90.10.1002/1529-0131(200201)46:1<175::AID-ART10015>;2-H
    1. Perl A, Gergely P, Jr, Nagy G, Koncz A, Banki K. Mitochondrial hyperpolarization: a checkpoint of T-cell life, death and autoimmunity. Trends Immunol (2004) 25(7):360–7.10.1016/j.it.2004.05.001
    1. Fernandez D, Perl A. Metabolic control of T cell activation and death in SLE. Autoimmun Rev (2009) 8(3):184–9.10.1016/j.autrev.2008.07.041
    1. Esparza-Molto PB, Cuezva JM. The role of mitochondrial H(+)-ATP synthase in cancer. Front Oncol (2018) 8:53.10.3389/fonc.2018.00053
    1. Vojtísková A, Jesina P, Kalous M, Kaplanová V, Houstek J, Tesarová M, et al. Mitochondrial membrane potential and ATP production in primary disorders of ATP synthase. Toxicol Mech Methods (2004) 14(1–2):7–11.10.1080/15376520490257347
    1. Barbi de Moura M, Vincent G, Fayewicz SL, Bateman NW, Hood BL, Sun M, et al. Mitochondrial respiration – an important therapeutic target in melanoma. PLoS One (2012) 7(8):e40690.10.1371/journal.pone.0040690
    1. Park DW, Zmijewski JW. Mitochondrial dysfunction and immune cell metabolism in sepsis. Infect Chemother (2017) 49(1):10–21.10.3947/ic.2017.49.1.10
    1. Sheu SS, Nauduri D, Anders MW. Targeting antioxidants to mitochondria: a new therapeutic direction. Biochim Biophys Acta (2006) 1762(2):256–65.10.1016/j.bbadis.2005.10.007
    1. D’Souza GG, Wagle MA, Saxena V, Shah A. Approaches for targeting mitochondria in cancer therapy. Biochim Biophys Acta (2011) 1807(6):689–96.10.1016/j.bbabio.2010.08.008
    1. Adlam VJ, Harrison JC, Porteous CM, James AM, Smith RA, Murphy MP, et al. Targeting an antioxidant to mitochondria decreases cardiac ischemia-reperfusion injury. FASEB J (2005) 19(9):1088–95.10.1096/fj.05-3718com
    1. Dare AJ, Bolton EA, Pettigrew GJ, Bradley JA, Saeb-Parsy K, Murphy MP. Protection against renal ischemia-reperfusion injury in vivo by the mitochondria targeted antioxidant MitoQ. Redox Biol (2015) 5:163–8.10.1016/j.redox.2015.04.008
    1. Reddy PH, Tripathi R, Troung Q, Tirumala K, Reddy TP, Anekonda V, et al. Abnormal mitochondrial dynamics and synaptic degeneration as early events in Alzheimer’s disease: implications to mitochondria-targeted antioxidant therapeutics. Biochim Biophys Acta (2012) 1822(5):639–49.10.1016/j.bbadis.2011.10.011
    1. Wu X, Cheng J, Wang X. Dietary antioxidants: potential anticancer agents. Nutr Cancer (2017) 69(4):521–33.10.1080/01635581.2017.1299872
    1. Kim HK, Han J. Mitochondria-targeted antioxidants for the treatment of cardiovascular disorders. Adv Exp Med Biol (2017) 982:621–46.10.1007/978-3-319-55330-6_32
    1. Kalyanaraman B, Cheng G, Hardy M, Ouari O, Lopez M, Joseph J, et al. A review of the basics of mitochondrial bioenergetics, metabolism, and related signaling pathways in cancer cells: therapeutic targeting of tumor mitochondria with lipophilic cationic compounds. Redox Biol (2018) 14:316–27.10.1016/j.redox.2017.09.020
    1. Mailloux RJ, Treberg JR. Protein S-glutathionlyation links energy metabolism to redox signaling in mitochondria. Redox Biol (2016) 8:110–8.10.1016/j.redox.2015.12.010
    1. Chen CL, Chen J, Rawale S, Varadharaj S, Kaumaya PP, Zweier JL, et al. Protein tyrosine nitration of the flavin subunit is associated with oxidative modification of mitochondrial complex II in the post-ischemic myocardium. J Biol Chem (2008) 283(41):27991–8003.10.1074/jbc.M802691200
    1. Espinosa-Diez C, Miguel V, Mennerich D, Kietzmann T, Sánchez-Pérez P, Cadenas S, et al. Antioxidant responses and cellular adjustments to oxidative stress. Redox Biol (2015) 6:183–97.10.1016/j.redox.2015.07.008
    1. Chouchani ET, Methner C, Nadtochiy SM, Logan A, Pell VR, Ding S, et al. Cardioprotection by S-nitrosation of a cysteine switch on mitochondrial complex I. Nat Med (2013) 19(6):753–9.10.1038/nm.3212
    1. Bhatt KN, Butler J. Myocardial energetics and heart failure: a review of recent therapeutic trials. Curr Heart Fail Rep (2018) 15(3):191–7.10.1007/s11897-018-0386-8
    1. Shanmugam MK, Shen H, Tang FR, Arfuso F, Rajesh M, Wang L, et al. Potential role of genipin in cancer therapy. Pharmacol Res (2018) 133:195–200.10.1016/j.phrs.2018.05.007
    1. Quryshi N, Norwood Toro LE, Ait-Aissa K, Kong A, Beyer AM. Chemotherapeutic-induced cardiovascular dysfunction: physiological effects, early detection-the role of telomerase to counteract mitochondrial defects and oxidative stress. Int J Mol Sci (2018) 19(3):E797.10.3390/ijms19030797
    1. Dashdorj A, Jyothi KR, Lim S, Jo A, Nguyen MN, Ha J, et al. Mitochondria-targeted antioxidant MitoQ ameliorates experimental mouse colitis by suppressing NLRP3 inflammasome-mediated inflammatory cytokines. BMC Med (2013) 11:178.10.1186/1741-7015-11-178
    1. Cordero MD, Alcocer-Gómez E, Culic O, Carrión AM, de Miguel M, Díaz-Parrado E, et al. NLRP3 inflammasome is activated in fibromyalgia: the effect of coenzyme Q10. Antioxid Redox Signal (2014) 20(8):1169–80.10.1089/ars.2013.5198
    1. Morton SU, Prabhu SP, Lidov HGW, Shi J, Anselm I, Brownstein CA, et al. AIFM1 mutation presenting with fatal encephalomyopathy and mitochondrial disease in an infant. Cold Spring Harb Mol Case Stud (2017) 3(2):a001560.10.1101/mcs.a001560
    1. Fukumoto K, Nguyen TB, Chiba S, Sakata-Yanagimoto M. Review of the biologic and clinical significance of genetic mutations in angioimmunoblastic T-cell lymphoma. Cancer Sci (2018) 109(3):490–6.10.1111/cas.13393
    1. Moshtaghi A, Vaziri H, Sariri R, Shaigan H. Polymorphism of MnSOD (Val16Ala) gene in pregnancies with blighted ovum: a case-control study. Int J Reprod Biomed (Yazd) (2017) 15(8):503–8.10.29252/ijrm.15.8.503

Source: PubMed

3
Iratkozz fel