1932

Abstract

According to the endosymbiotic theory, most of the DNA of the original bacterial endosymbiont has been lost or transferred to the nucleus, leaving a much smaller (∼16 kb in mammals), circular molecule that is the present-day mitochondrial DNA (mtDNA). The ability of mtDNA to escape mitochondria and integrate into the nuclear genome was discovered in budding yeast, along with genes that regulate this process. Mitochondria have emerged as key regulators of innate immunity, and it is now recognized that mtDNA released into the cytoplasm, outside of the cell, or into circulation activates multiple innate immune signaling pathways. Here, we first review the mechanisms through which mtDNA is released into the cytoplasm, including several inducible mitochondrial pores and defective mitophagy or autophagy. Next, we cover how the different forms of released mtDNA activate specific innate immune nucleic acid sensors and inflammasomes. Finally, we discuss how intracellular and extracellular mtDNA release, including circulating cell-free mtDNA that promotes systemic inflammation, are implicated in human diseases, bacterial and viral infections, senescence and aging.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-biochem-032620-104401
2023-06-20
2024-04-28
Loading full text...

Full text loading...

/deliver/fulltext/biochem/92/1/annurev-biochem-032620-104401.html?itemId=/content/journals/10.1146/annurev-biochem-032620-104401&mimeType=html&fmt=ahah

Literature Cited

  1. 1.
    Thorsness PE, Fox TD. 1990. Escape of DNA from mitochondria to the nucleus in Saccharomyces cerevisiae. Nature 346:376–79
    [Google Scholar]
  2. 2.
    Thorsness PE, Fox TD. 1993. Nuclear mutations in Saccharomyces cerevisiae that affect the escape of DNA from mitochondria to the nucleus. Genetics 134:21–28
    [Google Scholar]
  3. 3.
    Sprenger H-G, MacVicar T, Bahat A, Fiedler KU, Hermans S et al. 2021. Cellular pyrimidine imbalance triggers mitochondrial DNA–dependent innate immunity. Nat. Metab. 3:636–50
    [Google Scholar]
  4. 4.
    Hancock-Cerutti W, Wu Z, Xu P, Yadavalli N, Leonzino M et al. 2022. ER-lysosome lipid transfer protein VPS13C/PARK23 prevents aberrant mtDNA-dependent STING signaling. J. Cell Biol. 221:7e202106046
    [Google Scholar]
  5. 5.
    Sliter DA, Martinez J, Hao L, Chen X, Sun N et al. 2018. Parkin and PINK1 mitigate STING-induced inflammation. Nature 561:258–62
    [Google Scholar]
  6. 6.
    Matsui H, Ito J, Matsui N, Uechi T, Onodera O, Kakita A. 2021. Cytosolic dsDNA of mitochondrial origin induces cytotoxicity and neurodegeneration in cellular and zebrafish models of Parkinson's disease. Nat. Commun. 12:3101
    [Google Scholar]
  7. 7.
    Weindel CG, Bell SL, Vail KJ, West KO, Patrick KL, Watson RO 2020. LRRK2 maintains mitochondrial homeostasis and regulates innate immune responses to Mycobacterium tuberculosis. eLife 9:e51071
    [Google Scholar]
  8. 8.
    West AP, Shadel GS. 2017. Mitochondrial DNA in innate immune responses and inflammatory pathology. Nat. Rev. Immunol. 17:363–75
    [Google Scholar]
  9. 9.
    Shadel GS, Clayton DA. 1997. Mitochondrial DNA maintenance in vertebrates. Annu. Rev. Biochem. 66:409–35
    [Google Scholar]
  10. 10.
    Falkenberg M, Gustafsson CM. 2020. Mammalian mitochondrial DNA replication and mechanisms of deletion formation. Crit. Rev. Biochem. Mol. Biol. 55:509–24
    [Google Scholar]
  11. 11.
    Garrido N, Griparic L, Jokitalo E, Wartiovaara J, van der Bliek AM, Spelbrink JN. 2003. Composition and dynamics of human mitochondrial nucleoids. Mol. Biol. Cell 14:1583–96
    [Google Scholar]
  12. 12.
    Kukat C, Wurm CA, Spåhr H, Falkenberg M, Larsson N-G, Jakobs S 2011. Super-resolution microscopy reveals that mammalian mitochondrial nucleoids have a uniform size and frequently contain a single copy of mtDNA. PNAS 108:13534–39
    [Google Scholar]
  13. 13.
    Kukat C, Davies KM, Wurm CA, Spåhr H, Bonekamp NA et al. 2015. Cross-strand binding of TFAM to a single mtDNA molecule forms the mitochondrial nucleoid. PNAS 112:11288–93
    [Google Scholar]
  14. 14.
    Ngo HB, Lovely GA, Phillips R, Chan DC. 2014. Distinct structural features of TFAM drive mitochondrial DNA packaging versus transcriptional activation. Nat. Commun. 5:3077
    [Google Scholar]
  15. 15.
    Kazak L, Reyes A, Holt IJ 2012. Minimizing the damage: Repair pathways keep mitochondrial DNA intact. Nat. Rev. Mol. Cell Biol. 13:659–71
    [Google Scholar]
  16. 16.
    Bonawitz ND, Clayton DA, Shadel GS. 2006. Initiation and beyond: multiple functions of the human mitochondrial transcription machinery. Mol. Cell 24:813–25
    [Google Scholar]
  17. 17.
    Tan BG, Mutti CD, Shi Y, Xie X, Zhu X et al. 2022. The human mitochondrial genome contains a second light strand promoter. Mol. Cell 82:3646–60.e9
    [Google Scholar]
  18. 18.
    Nicholls TJ, Gustafsson CM. 2018. Separating and segregating the human mitochondrial genome. Trends Biochem. Sci. 43:869–81
    [Google Scholar]
  19. 19.
    Nicholls TJ, Nadalutti CA, Motori E, Sommerville EW, Gorman GS et al. 2018. Topoisomerase 3α is required for decatenation and segregation of human mtDNA. Mol. Cell 69:9–23.e6
    [Google Scholar]
  20. 20.
    Wu Z, Sainz AG, Shadel GS. 2021. Mitochondrial DNA: cellular genotoxic stress sentinel. Trends Biochem. Sci. 46:812–21
    [Google Scholar]
  21. 21.
    Bruni F, Lightowlers RN, Chrzanowska-Lightowlers ZM. 2017. Human mitochondrial nucleases. FEBS J. 284:1767–77
    [Google Scholar]
  22. 22.
    Dhir A, Dhir S, Borowski LS, Jimenez L, Teitell M et al. 2018. Mitochondrial double-stranded RNA triggers antiviral signalling in humans. Nature 560:238–42
    [Google Scholar]
  23. 23.
    Tigano M, Vargas DC, Tremblay-Belzile S, Fu Y, Sfeir A. 2021. Nuclear sensing of breaks in mitochondrial DNA enhances immune surveillance. Nature 591:477–81
    [Google Scholar]
  24. 24.
    McArthur K, Whitehead LW, Heddleston JM, Li L, Padman BS et al. 2018. BAK/BAX macropores facilitate mitochondrial herniation and mtDNA efflux during apoptosis. Science 359:eaao6047
    [Google Scholar]
  25. 25.
    Riley JS, Quarato G, Cloix C, Lopez J, O'Prey J et al. 2018. Mitochondrial inner membrane permeabilisation enables mtDNA release during apoptosis. EMBO J. 37:e99238
    [Google Scholar]
  26. 26.
    Chen L, Dong J, Liao S, Wang S, Wu Z et al. 2022. Loss of Sam50 in hepatocytes induces cardiolipin-dependent mitochondrial membrane remodeling to trigger mtDNA release and liver injury. Hepatology 76:51389–408
    [Google Scholar]
  27. 27.
    Maekawa H, Inoue T, Ouchi H, Jao T-M, Inoue R et al. 2019. Mitochondrial damage causes inflammation via cGAS-STING signaling in acute kidney injury. Cell Rep. 29:1261–73.e6
    [Google Scholar]
  28. 28.
    Zhang Y-F, Zhou L, Mao H-Q, Yang F-H, Chen Z, Zhang L 2021. Mitochondrial DNA leakage exacerbates odontoblast inflammation through gasdermin D-mediated pyroptosis. Cell Death Discov. 7:381
    [Google Scholar]
  29. 29.
    Wang L-Q, Liu T, Yang S, Sun L, Zhao Z-Y et al. 2021. Perfluoroalkyl substance pollutants activate the innate immune system through the AIM2 inflammasome. Nat. Commun. 12:2915
    [Google Scholar]
  30. 30.
    Cosentino K, Hertlein V, Jenner A, Dellmann T, Gojkovic M et al. 2022. The interplay between BAX and BAK tunes apoptotic pore growth to control mitochondrial-DNA-mediated inflammation. Mol. Cell 82:933–49.e9
    [Google Scholar]
  31. 31.
    Kim J, Gupta R, Blanco LP, Yang S, Shteinfer-Kuzmine A et al. 2019. VDAC oligomers form mitochondrial pores to release mtDNA fragments and promote lupus-like disease. Science 366:1531–36
    [Google Scholar]
  32. 32.
    García N, Chávez E. 2007. Mitochondrial DNA fragments released through the permeability transition pore correspond to specific gene size. Life Sci. 81:1160–66
    [Google Scholar]
  33. 33.
    Collins LV, Hajizadeh S, Holme E, Jonsson I-M, Tarkowski A. 2004. Endogenously oxidized mitochondrial DNA induces in vivo and in vitro inflammatory responses. J. Leukocyte Biol. 75:995–1000
    [Google Scholar]
  34. 34.
    Xian H, Watari K, Sanchez-Lopez E, Offenberger J, Onyuru J et al. 2022. Oxidized DNA fragments exit mitochondria via mPTP- and VDAC-dependent channels to activate NLRP3 inflammasome and interferon signaling. Immunity 55:81370–85.e8
    [Google Scholar]
  35. 35.
    He W-R, Cao L-B, Yang Y-L, Hua D, Hu M-M, Shu H-B. 2021. VRK2 is involved in the innate antiviral response by promoting mitostress-induced mtDNA release. Cell. Mol. Immunol. 18:1186–96
    [Google Scholar]
  36. 36.
    Luzwick JW, Dombi E, Boisvert RA, Roy S, Park S et al. 2021. MRE11-dependent instability in mitochondrial DNA fork protection activates a cGAS immune signaling pathway. Sci. Adv. 7:eabf9441
    [Google Scholar]
  37. 37.
    Magnani L, Colantuoni M, Mortellaro A. 2022. Gasdermins: new therapeutic targets in host defense, inflammatory diseases, and cancer. Front. Immunol. 13:898298
    [Google Scholar]
  38. 38.
    Huang LS, Hong Z, Wu W, Xiong S, Zhong M et al. 2020. mtDNA activates cGAS signaling and suppresses the YAP-mediated endothelial cell proliferation program to promote inflammatory injury. Immunity 52:475–86.e5
    [Google Scholar]
  39. 39.
    Torre-Minguela C, Gómez AI, Couillin I, Pelegrín P. 2021. Gasdermins mediate cellular release of mitochondrial DNA during pyroptosis and apoptosis. FASEB J. 35:8e21757
    [Google Scholar]
  40. 40.
    Rodríguez-Nuevo A, Díaz-Ramos A, Noguera E, Díaz-Sáez F, Duran X et al. 2018. Mitochondrial DNA and TLR9 drive muscle inflammation upon Opa1 deficiency. EMBO J. 37:e96553
    [Google Scholar]
  41. 41.
    Bueno M, Zank D, Buendia-Roldán I, Fiedler K, Mays BG et al. 2019. PINK1 attenuates mtDNA release in alveolar epithelial cells and TLR9 mediated profibrotic responses. PLOS ONE 14:e0218003
    [Google Scholar]
  42. 42.
    Newman LE, Tadepalle N, Novak SW, Schiavon CR, Rojas GR et al. 2022. Endosomal removal and disposal of dysfunctional, immunostimulatory mitochondrial DNA. bioRxiv 2022.10.12.511955. https://doi.org/10.1101/2022.10.12.511955
  43. 43.
    Oka T, Hikoso S, Yamaguchi O, Taneike M, Takeda T et al. 2012. Mitochondrial DNA that escapes from autophagy causes inflammation and heart failure. Nature 485:251–55
    [Google Scholar]
  44. 44.
    Nakahira K, Haspel JA, Rathinam VAK, Lee S-J, Dolinay T et al. 2011. Autophagy proteins regulate innate immune responses by inhibiting the release of mitochondrial DNA mediated by the NALP3 inflammasome. Nat. Immunol. 12:222–30
    [Google Scholar]
  45. 45.
    Sen A, Kallabis S, Gaedke F, Jüngst C, Boix J et al. 2022. Mitochondrial membrane proteins and VPS35 orchestrate selective removal of mtDNA. Nat. Commun. 13:6704
    [Google Scholar]
  46. 46.
    De Gaetano A, Solodka K, Zanini G, Selleri V, Mattioli AV et al. 2021. Molecular mechanisms of mtDNA-mediated inflammation. Cells 10:2898
    [Google Scholar]
  47. 47.
    Riley JS, Tait SW. 2020. Mitochondrial DNA in inflammation and immunity. EMBO Rep. 21:e49799
    [Google Scholar]
  48. 48.
    Zhang Q, Raoof M, Chen Y, Sumi Y, Sursal T et al. 2010. Circulating mitochondrial DAMPs cause inflammatory responses to injury. Nature 464:104–7
    [Google Scholar]
  49. 49.
    Zhang Q, Itagaki K, Hauser CJ. 2010. Mitochondrial DNA is released by shock and activates neutrophils via p38 map kinase. Shock 34:155–59
    [Google Scholar]
  50. 50.
    Caielli S, Athale S, Domic B, Murat E, Chandra M et al. 2016. Oxidized mitochondrial nucleoids released by neutrophils drive type I interferon production in human lupus. J. Exp. Med. 213:697–713
    [Google Scholar]
  51. 51.
    Lood C, Blanco LP, Purmalek MM, Carmona-Rivera C, De Ravin SS et al. 2016. Neutrophil extracellular traps enriched in oxidized mitochondrial DNA are interferogenic and contribute to lupus-like disease. Nat. Med. 22:146–53
    [Google Scholar]
  52. 52.
    Wang H, Li T, Chen S, Gu Y, Ye S. 2015. Neutrophil extracellular trap mitochondrial DNA and its autoantibody in systemic lupus erythematosus and a proof-of-concept trial of metformin. Arthritis Rheumatol. 67:3190–200
    [Google Scholar]
  53. 53.
    Todkar K, Chikhi L, Desjardins V, El-Mortada F, Pépin G, Germain M. 2021. Selective packaging of mitochondrial proteins into extracellular vesicles prevents the release of mitochondrial DAMPs. Nat. Commun. 12:1971
    [Google Scholar]
  54. 54.
    König T, Nolte H, Aaltonen MJ, Tatsuta T, Krols M et al. 2021. MIROs and DRP1 drive mitochondrial-derived vesicle biogenesis and promote quality control. Nat. Cell Biol. 23:1271–86
    [Google Scholar]
  55. 55.
    Jing R, Hu Z-K, Lin F, He S, Zhang S-S et al. 2020. Mitophagy-mediated mtDNA release aggravates stretching-induced inflammation and lung epithelial cell injury via the TLR9/MyD88/NF-κB pathway. Front. Cell Dev. Biol. 8:819
    [Google Scholar]
  56. 56.
    Rabas N, Palmer S, Mitchell L, Ismail S, Gohlke A et al. 2021. PINK1 drives production of mtDNA-containing extracellular vesicles to promote invasiveness. J. Cell Biol. 220:12e202006049
    [Google Scholar]
  57. 57.
    Man SM, Kanneganti T-D. 2016. Converging roles of caspases in inflammasome activation, cell death and innate immunity. Nat. Rev. Immunol. 16:7–21
    [Google Scholar]
  58. 58.
    West AP, Shadel GS, Ghosh S. 2011. Mitochondria in innate immune responses. Nat. Rev. Immunol. 11:389–402
    [Google Scholar]
  59. 59.
    Gao D, Wu J, Wu Y-T, Du F, Aroh C et al. 2013. Cyclic GMP-AMP synthase is an innate immune sensor of HIV and other retroviruses. Science 341:903–6
    [Google Scholar]
  60. 60.
    Ishikawa H, Barber GN. 2008. STING is an endoplasmic reticulum adaptor that facilitates innate immune signalling. Nature 455:674–78
    [Google Scholar]
  61. 61.
    Zhong B, Yang Y, Li S, Wang YY, Li Y et al. 2008. The adaptor protein MITA links virus-sensing receptors to IRF3 transcription factor activation. Immunity 29:538–50
    [Google Scholar]
  62. 62.
    Ritchie C, Carozza JA, Li L. 2022. Biochemistry, cell biology, and pathophysiology of the innate immune cGAS–cGAMP–STING pathway. Annu. Rev. Biochem. 91:599–628
    [Google Scholar]
  63. 63.
    Rongvaux A, Jackson R, Harman CCD, Li T, West AP et al. 2014. Apoptotic caspases prevent the induction of type I interferons by mitochondrial DNA. Cell 159:1563–77
    [Google Scholar]
  64. 64.
    White MJ, McArthur K, Metcalf D, Lane RM, Cambier JC et al. 2014. Apoptotic caspases suppress mtDNA-induced STING-mediated type I IFN production. Cell 159:1549–62
    [Google Scholar]
  65. 65.
    West AP, Khoury-Hanold W, Staron M, Tal MC, Pineda CM et al. 2015. Mitochondrial DNA stress primes the antiviral innate immune response. Nature 520:553–57
    [Google Scholar]
  66. 66.
    Dou Z, Ghosh K, Vizioli MG, Zhu J, Sen P et al. 2017. Cytoplasmic chromatin triggers inflammation in senescence and cancer. Nature 550:402–6
    [Google Scholar]
  67. 67.
    Glück S, Guey B, Gulen MF, Wolter K, Kang TW et al. 2017. Innate immune sensing of cytosolic chromatin fragments through cGAS promotes senescence. Nat. Cell Biol. 19:1061–70
    [Google Scholar]
  68. 68.
    Larsson N-G, Wang J, Wilhelmsson H, Oldfors A, Rustin P et al. 1998. Mitochondrial transcription factor A is necessary for mtDNA maintance and embryogenesis in mice. Nat. Genet. 18:231–36
    [Google Scholar]
  69. 69.
    Andreeva L, Hiller B, Kostrewa D, Lässig C, de Oliveira Mann CC et al. 2017. cGAS senses long and HMGB/TFAM-bound U-turn DNA by forming protein–DNA ladders. Nature 549:394–98
    [Google Scholar]
  70. 70.
    Isidoro Cobo TNT, Mangalhara KC, Lana A, Yeang C, Han C et al. 2022. DNMT3A and TET2 restrain mitochondrial DNA-mediated interferon signaling in macrophages. Immunity 55:81386–401.e10
    [Google Scholar]
  71. 71.
    Chung KW, Dhillon P, Huang S, Sheng X, Shrestha R et al. 2019. Mitochondrial damage and activation of the STING pathway lead to renal inflammation and fibrosis. Cell Metab. 30:784–99.e5
    [Google Scholar]
  72. 72.
    Hu M, Zhou M, Bao X, Pan D, Jiao M et al. 2021. ATM inhibition enhances cancer immunotherapy by promoting mtDNA leakage and cGAS/STING activation. J. Clin. Investig. 131:3e139333
    [Google Scholar]
  73. 73.
    Al Khatib I, Deng J, Symes SA, Zhang H, Huang S et al. 2022. Activation of the cGAS-STING innate immune response in cells with deficient mitochondrial topoisomerase TOP1MT. bioRxiv 2022.03.08.483326. https://www.biorxiv.org/content/10.1101/2022.03.08.483326v1
  74. 74.
    Torres-Odio S, Lei Y, Gispert S, Maletzko A, Key J et al. 2021. Loss of mitochondrial protease CLPP activates type I IFN responses through the mitochondrial DNA–cGAS–STING signaling axis. J. Immunol. 206:1890–900
    [Google Scholar]
  75. 75.
    Zhong Z, Liang S, Sanchez-Lopez E, He F, Shalapour S et al. 2018. New mitochondrial DNA synthesis enables NLRP3 inflammasome activation. Nature 560:198–203
    [Google Scholar]
  76. 76.
    Lepelley A, Della Mina E, Van Nieuwenhove E, Waumans L, Fraitag S et al. 2021. Enhanced cGAS-STING–dependent interferon signaling associated with mutations in ATAD3A. J. Exp. Med. 218:10e20201560
    [Google Scholar]
  77. 77.
    Peralta S, Goffart S, Williams SL, Diaz F, Garcia S et al. 2018. ATAD3 controls mitochondrial cristae structure, influencing mtDNA replication and cholesterol levels in muscle. J. Cell Sci. 131:jcs217075
    [Google Scholar]
  78. 78.
    Gerhold JM, Cansiz-Arda Ş, Lõhmus M, Engberg O, Reyes A et al. 2015. Human mitochondrial DNA-protein complexes attach to a cholesterol-rich membrane structure. Sci. Rep. 5:15292
    [Google Scholar]
  79. 79.
    He J, Mao C-C, Reyes A, Sembongi H, Di Re M et al. 2007. The AAA+ protein ATAD3 has displacement loop binding properties and is involved in mitochondrial nucleoid organization. J. Cell Biol. 176:141–46
    [Google Scholar]
  80. 80.
    Lewis SC, Uchiyama LF, Nunnari J. 2016. ER-mitochondria contacts couple mtDNA synthesis with mitochondrial division in human cells. Science 353:aaf5549
    [Google Scholar]
  81. 81.
    Bronner DN, Abuaita BH, Chen X, Fitzgerald KA, Nuñez G et al. 2015. Endoplasmic reticulum stress activates the inflammasome via NLRP3- and caspase-2-driven mitochondrial damage. Immunity 43:451–62
    [Google Scholar]
  82. 82.
    Günther C, Rösing S, Ullrich F, Meisterfeld S, Schmidt F et al. 2022. Chronic ER stress promotes cGAS/mtDNA-induced autoimmunity via ATF6 in myotonic dystrophy type 2. Preprint, Res. Sq. https://assets.researchsquare.com/files/rs-1784722/v1_covered.pdf?c=1656426386
  83. 83.
    Ichim G, Lopez J, Ahmed SU, Muthalagu N, Giampazolias E et al. 2015. Limited mitochondrial permeabilization causes DNA damage and genomic instability in the absence of cell death. Mol. Cell 57:860–72
    [Google Scholar]
  84. 84.
    Desai R, Frazier AE, Durigon R, Patel H, Jones AW et al. 2017. ATAD3 gene cluster deletions cause cerebellar dysfunction associated with altered mitochondrial DNA and cholesterol metabolism. Brain 140:1595–610
    [Google Scholar]
  85. 85.
    Dang EV, McDonald JG, Russell DW, Cyster JG. 2017. Oxysterol restraint of cholesterol synthesis prevents AIM2 inflammasome activation. Cell 171:1057–71.e11
    [Google Scholar]
  86. 86.
    York AG, Williams KJ, Argus JP, Zhou QD, Brar G et al. 2015. Limiting cholesterol biosynthetic flux spontaneously engages type I IFN signaling. Cell 163:1716–29
    [Google Scholar]
  87. 87.
    Blanc M, Hsieh WY, Robertson KA, Watterson S, Shui G et al. 2011. Host defense against viral infection involves interferon mediated down-regulation of sterol biosynthesis. PLOS Biol. 9:e1000598
    [Google Scholar]
  88. 88.
    Chen W, Li S, Yu H, Liu X, Huang L et al. 2016. ER adaptor SCAP translocates and recruits IRF3 to perinuclear microsome induced by cytosolic microbial DNAs. PLOS Pathog. 12:e1005462
    [Google Scholar]
  89. 89.
    Chu T-T, Tu X, Yang K, Wu J, Repa JJ, Yan N 2021. Tonic prime-boost of STING signalling mediates Niemann–Pick disease type C. Nature 596:570–75
    [Google Scholar]
  90. 90.
    Field CS, Baixauli F, Kyle RL, Puleston DJ, Cameron AM et al. 2020. Mitochondrial integrity regulated by lipid metabolism is a cell-intrinsic checkpoint for Treg suppressive function. Cell Metab. 31:422–37.e5
    [Google Scholar]
  91. 91.
    Yuan L, Mao Y, Luo W, Wu W, Xu H et al. 2017. Palmitic acid dysregulates the Hippo–YAP pathway and inhibits angiogenesis by inducing mitochondrial damage and activating the cytosolic DNA sensor cGAS–STING–IRF3 signaling mechanism. J. Biol. Chem. 292:15002–15
    [Google Scholar]
  92. 92.
    Hu H, Zhao R, He Q, Cui C, Song J et al. 2022. cGAS-STING mediates cytoplasmic mitochondrial-DNA-induced inflammatory signal transduction during accelerated senescence of pancreatic β-cells induced by metabolic stress. FASEB J. 36:e22266
    [Google Scholar]
  93. 93.
    Ma XM, Geng K, Law BY-K, Wang P, Pu YL et al. 2022. Lipotoxicity-induced mtDNA release promotes diabetic cardiomyopathy by activating the cGAS-STING pathway in obesity-related diabetes. Cell Biol. Toxicol. https://doi.org/10.1007/s10565-021-09692-z
    [Google Scholar]
  94. 94.
    Yan M, Li Y, Luo Q, Zeng W, Shao X et al. 2022. Mitochondrial damage and activation of the cytosolic DNA sensor cGAS–STING pathway lead to cardiac pyroptosis and hypertrophy in diabetic cardiomyopathy mice. Cell Death Discov. 8:258
    [Google Scholar]
  95. 95.
    Bai J, Cervantes C, Liu J, He S, Zhou H et al. 2017. DsbA-L prevents obesity-induced inflammation and insulin resistance by suppressing the mtDNA release-activated cGAS-cGAMP-STING pathway. PNAS 114:12196–201
    [Google Scholar]
  96. 96.
    Bai J, Cervantes C, He S, He J, Plasko GR et al. 2020. Mitochondrial stress-activated cGAS-STING pathway inhibits thermogenic program and contributes to overnutrition-induced obesity in mice. Commun. Biol. 3:257
    [Google Scholar]
  97. 97.
    Fermaintt CS, Takahashi-Ruiz L, Liang H, Mooberry SL, Risinger AL. 2021. Eribulin activates the cGAS-STING pathway via the cytoplasmic accumulation of mitochondrial DNA. Mol. Pharmacol. 100:309–18
    [Google Scholar]
  98. 98.
    Yan X, Yao C, Fang C, Han M, Gong C et al. 2022. Rocaglamide promotes the infiltration and antitumor immunity of NK cells by activating cGAS-STING signaling in non-small cell lung cancer. Int. J. Biol. Sci. 18:585–98
    [Google Scholar]
  99. 99.
    Carroll EC, Jin L, Mori A, Muñoz-Wolf N, Oleszycka E et al. 2016. The vaccine adjuvant chitosan promotes cellular immunity via DNA sensor cGAS-STING-dependent induction of type I interferons. Immunity 44:597–608
    [Google Scholar]
  100. 100.
    Sabnam S, Rizwan H, Pal S, Pal A. 2020. CEES-induced ROS accumulation regulates mitochondrial complications and inflammatory response in keratinocytes. Chem. Biol. Interact. 321:109031
    [Google Scholar]
  101. 101.
    Zeng X, Li X, Zhang Y, Cao C, Zhou Q. 2022. IL6 induces mtDNA leakage to affect the immune escape of endometrial carcinoma via cGAS-STING. J. Immunol. Res. 2022:3815853
    [Google Scholar]
  102. 102.
    Aarreberg LD, Esser-Nobis K, Driscoll C, Shuvarikov A, Roby JA, Gale M. 2019. Interleukin-1β induces mtDNA release to activate innate immune signaling via cGAS-STING. Mol. Cell 74:801–15.e6
    [Google Scholar]
  103. 103.
    Willemsen J, Neuhoff MT, Hoyler T, Noir E, Tessier C et al. 2021. TNF leads to mtDNA release and cGAS/STING-dependent interferon responses that support inflammatory arthritis. Cell Rep. 37:109977
    [Google Scholar]
  104. 104.
    Xu MM, Pu Y, Han D, Shi Y, Cao X et al. 2017. Dendritic cells but not macrophages sense tumor mitochondrial DNA for cross-priming through signal regulatory protein α signaling. Immunity 47:363–73.e5
    [Google Scholar]
  105. 105.
    Zhang Q, Wei J, Liu Z, Huang X, Sun M et al. 2022. STING signaling sensing of DRP1-dependent mtDNA release in Kupffer cells contributes to lipopolysaccharide-induced liver injury in mice. Redox. Biol. 54:102367
    [Google Scholar]
  106. 106.
    Kerur N, Fukuda S, Banerjee D, Kim Y, Fu D et al. 2018. cGAS drives noncanonical-inflammasome activation in age-related macular degeneration. Nat. Med. 24:50–61
    [Google Scholar]
  107. 107.
    Swanson KV, Deng M, Ting JP. 2019. The NLRP3 inflammasome: molecular activation and regulation to therapeutics. Nat. Rev. Immunol. 19:477–89
    [Google Scholar]
  108. 108.
    Yabal M, Calleja DJ, Simpson DS, Lawlor KE. 2019. Stressing out the mitochondria: mechanistic insights into NLRP3 inflammasome activation. J. Leukocyte Biol. 105:377–99
    [Google Scholar]
  109. 109.
    Shimada K, Crother TR, Karlin J, Dagvadorj J, Chiba N et al. 2012. Oxidized mitochondrial DNA activates the NLRP3 inflammasome during apoptosis. Immunity 36:401–14
    [Google Scholar]
  110. 110.
    Tumurkhuu G, Shimada K, Dagvadorj J, Crother TR, Zhang W et al. 2016. Ogg1-dependent DNA repair regulates NLRP3 inflammasome and prevents atherosclerosis. Circ. Res. 119:e76–90
    [Google Scholar]
  111. 111.
    Jin Y, Liu Y, Xu L, Xu J, Xiong Y et al. 2022. Novel role for caspase 1 inhibitor VX765 in suppressing NLRP3 inflammasome assembly and atherosclerosis via promoting mitophagy and efferocytosis. Cell Death. Dis. 13:512
    [Google Scholar]
  112. 112.
    Zhong Z, Umemura A, Sanchez-Lopez E, Liang S, Shalapour S et al. 2016. NF-κB restricts inflammasome activation via elimination of damaged mitochondria. Cell 164:896–910
    [Google Scholar]
  113. 113.
    Wu KKL, Long K, Lin H, Siu PMF, Hoo RLC et al. 2021. The APPL1-Rab5 axis restricts NLRP3 inflammasome activation through early endosomal-dependent mitophagy in macrophages. Nat. Commun. 12:6637
    [Google Scholar]
  114. 114.
    Guo W, Liu W, Chen Z, Gu Y, Peng S et al. 2017. Tyrosine phosphatase SHP2 negatively regulates NLRP3 inflammasome activation via ANT1-dependent mitochondrial homeostasis. Nat. Commun. 8:2168
    [Google Scholar]
  115. 115.
    Huang Y, Zhou JH, Zhang H, Canfran-Duque A, Singh AK et al. 2022. Brown adipose TRX2 deficiency activates mtDNA-NLRP3 to impair thermogenesis and protect against diet-induced insulin resistance. J. Clin. Investig. 132:9e148852
    [Google Scholar]
  116. 116.
    Gupta P, Sharma G, Lahiri A, Barthwal MK. 2022. FOXO3a acetylation regulates PINK1, mitophagy, inflammasome activation in murine palmitate-conditioned and diabetic macrophages. J. Leukocyte Biol. 111:611–27
    [Google Scholar]
  117. 117.
    Peng J, Wang H, Gong Z, Li X, He L et al. 2020. Idebenone attenuates cerebral inflammatory injury in ischemia and reperfusion via dampening NLRP3 inflammasome activity. Mol. Immunol. 123:74–87
    [Google Scholar]
  118. 118.
    Wu Y, Hao C, Liu X, Han G, Yin J et al. 2020. MitoQ protects against liver injury induced by severe burn plus delayed resuscitation by suppressing the mtDNA-NLRP3 axis. Int. Immunopharmacol. 80:106189
    [Google Scholar]
  119. 119.
    Moriyama M, Nagai M, Maruzuru Y, Koshiba T, Kawaguchi Y, Ichinohe T 2020. Influenza virus-induced oxidized DNA activates inflammasomes. iScience 23:101270
    [Google Scholar]
  120. 120.
    Li S, Li H, Zhang YL, Xin QL, Guan ZQ et al. 2020. SFTSV infection induces BAK/BAX-dependent mitochondrial DNA release to trigger NLRP3 inflammasome activation. Cell Rep. 30:4370–85.e7
    [Google Scholar]
  121. 121.
    Hua K-F, Chou J-C, Ka S-M, Tasi Y-L, Chen A et al. 2015. Cyclooxygenase-2 regulates NLRP3 inflammasome-derived IL-1β production. J. Cell. Physiol. 230:863–74
    [Google Scholar]
  122. 122.
    Guo H, Liu H, Jian Z, Cui H, Fang J et al. 2019. Nickel induces inflammatory activation via NF-κB, MAPKs, IRF3 and NLRP3 inflammasome signaling pathways in macrophages. Aging 11:11659–72
    [Google Scholar]
  123. 123.
    Li Y, Shen Y, Jin K, Wen Z, Cao W et al. 2019. The DNA repair nuclease MRE11A functions as a mitochondrial protector and prevents T cell pyroptosis and tissue inflammation. Cell Metab. 30:477–92.e6
    [Google Scholar]
  124. 124.
    Ren J-D, Wu X-B, Jiang R, Hao D-P, Liu Y. 2016. Molecular hydrogen inhibits lipopolysaccharide-triggered NLRP3 inflammasome activation in macrophages by targeting the mitochondrial reactive oxygen species. Biochim. Biophys. Acta Mol. Cell Res. 1863:50–55
    [Google Scholar]
  125. 125.
    Jung SS, Moon JS, Xu JF, Ifedigbo E, Ryter SW et al. 2015. Carbon monoxide negatively regulates NLRP3 inflammasome activation in macrophages. Am. J. Physiol. Lung Cell. Mol. Physiol. 308:L1058–67
    [Google Scholar]
  126. 126.
    Li S, Liang F, Kwan K, Tang Y, Wang X et al. 2018. Identification of ethyl pyruvate as a NLRP3 inflammasome inhibitor that preserves mitochondrial integrity. Mol. Med. 24:8
    [Google Scholar]
  127. 127.
    Liu X, Zhou W, Zhang X, Lu P, Du Q et al. 2016. Dimethyl fumarate ameliorates dextran sulfate sodium-induced murine experimental colitis by activating Nrf2 and suppressing NLRP3 inflammasome activation. Biochem. Pharmacol. 112:37–49
    [Google Scholar]
  128. 128.
    Xian H, Liu Y, Rundberg Nilsson A, Gatchalian R, Crother TR et al. 2021. Metformin inhibition of mitochondrial ATP and DNA synthesis abrogates NLRP3 inflammasome activation and pulmonary inflammation. Immunity 54:1463–77.e11
    [Google Scholar]
  129. 129.
    Lu B, Kwan K, Levine YA, Olofsson PS, Yang H et al. 2014. α7 nicotinic acetylcholine receptor signaling inhibits inflammasome activation by preventing mitochondrial DNA release. Mol. Med. 20:350–58
    [Google Scholar]
  130. 130.
    Lo Y-H, Huang Y-W, Wu Y-H, Tsai C-S, Lin Y-C et al. 2013. Selective inhibition of the NLRP3 inflammasome by targeting to promyelocytic leukemia protein in mouse and human. Blood 121:3185–94
    [Google Scholar]
  131. 131.
    Jin T, Perry A, Smith P, Jiang J, Xiao TS. 2013. Structure of the absent in melanoma 2 (AIM2) pyrin domain provides insights into the mechanisms of AIM2 autoinhibition and inflammasome assembly. J. Biol. Chem. 288:13225–35
    [Google Scholar]
  132. 132.
    Jin T, Perry A, Jiang J, Smith P, Curry JA et al. 2012. Structures of the HIN domain:DNA complexes reveal ligand binding and activation mechanisms of the AIM2 inflammasome and IFI16 receptor. Immunity 36:561–71
    [Google Scholar]
  133. 133.
    Dombrowski Y, Peric M, Koglin S, Kaymakanov N, Schmezer V et al. 2012. Honey bee (Apis mellifera) venom induces AIM2 inflammasome activation in human keratinocytes. Allergy 67:1400–7
    [Google Scholar]
  134. 134.
    Devi TD, Babu M, Mäkinen P, Kaikkonen MU, Heinaniemi M et al. 2017. Aggravated postinfarct heart failure in type 2 diabetes is associated with impaired mitophagy and exaggerated inflammasome activation. Am. J. Pathol. 187:2659–73
    [Google Scholar]
  135. 135.
    Wang Y, Chen C, Chen J, Sang T, Peng H et al. 2022. Overexpression of NAG-1/GDF15 prevents hepatic steatosis through inhibiting oxidative stress-mediated dsDNA release and AIM2 inflammasome activation. Redox Biol. 52:102322
    [Google Scholar]
  136. 136.
    Li W, Li Y, Siraj S, Jin H, Fan Y et al. 2019. FUN14 domain-containing 1–mediated mitophagy suppresses hepatocarcinogenesis by inhibition of inflammasome activation in mice. Hepatology 69:604–21
    [Google Scholar]
  137. 137.
    Sridevi Gurubaran I, Hytti M, Kaarniranta K, Kauppinen A 2022. Epoxomicin, a selective proteasome inhibitor, activates AIM2 inflammasome in human retinal pigment epithelium cells. Antioxidants 11:1288
    [Google Scholar]
  138. 138.
    Qi M, Dai D, Liu J, Li Z, Liang P et al. 2020. AIM2 promotes the development of non-small cell lung cancer by modulating mitochondrial dynamics. Oncogene 39:2707–23
    [Google Scholar]
  139. 139.
    Kawasaki T, Kawai T. 2014. Toll-like receptor signaling pathways. Front. Immunol. 5:461
    [Google Scholar]
  140. 140.
    Okabe Y, Kawane K, Akira S, Taniguchi T, Nagata S. 2005. Toll-like receptor–independent gene induction program activated by mammalian DNA escaped from apoptotic DNA degradation. J. Exp. Med. 202:1333–39
    [Google Scholar]
  141. 141.
    Chan MP, Onji M, Fukui R, Kawane K, Shibata T et al. 2015. DNase II-dependent DNA digestion is required for DNA sensing by TLR9. Nat. Commun. 6:5853
    [Google Scholar]
  142. 142.
    Sasai M, Linehan MM, Iwasaki A. 2010. Bifurcation of Toll-like receptor 9 signaling by adaptor protein 3. Science 329:1530–34
    [Google Scholar]
  143. 143.
    Julian MW, Shao G, VanGundy ZC, Papenfuss TL, Crouser ED. 2013. Mitochondrial transcription factor A, an endogenous danger signal, promotes TNFα release via RAGE- and TLR9-responsive plasmacytoid dendritic cells. PLOS ONE 8:e72354
    [Google Scholar]
  144. 144.
    Liu Y, Yan W, Tohme S, Chen M, Fu Y et al. 2015. Hypoxia induced HMGB1 and mitochondrial DNA interactions mediate tumor growth in hepatocellular carcinoma through Toll-like receptor 9. J. Hepatol. 63:114–21
    [Google Scholar]
  145. 145.
    Zhang Z, Meng P, Han Y, Shen C, Li B et al. 2015. Mitochondrial DNA-LL-37 complex promotes atherosclerosis by escaping from autophagic recognition. Immunity 43:1137–47
    [Google Scholar]
  146. 146.
    Sansone P, Savini C, Kurelac I, Chang Q, Amato LB et al. 2017. Packaging and transfer of mitochondrial DNA via exosomes regulate escape from dormancy in hormonal therapy-resistant breast cancer. PNAS 114:E9066–75
    [Google Scholar]
  147. 147.
    Takaoka A, Wang Z, Choi MK, Yanai H, Negishi H et al. 2007. DAI (DLM-1/ZBP1) is a cytosolic DNA sensor and an activator of innate immune response. Nature 448:501–5
    [Google Scholar]
  148. 148.
    Chiang DC, Li Y, Ng SK. 2021. The role of the Z-DNA binding domain in innate immunity and stress granules. Front. Immunol. 11:625504
    [Google Scholar]
  149. 149.
    Szczesny B, Marcatti M, Ahmad A, Montalbano M, Brunyánszki A et al. 2018. Mitochondrial DNA damage and subsequent activation of Z-DNA binding protein 1 links oxidative stress to inflammation in epithelial cells. Sci. Rep. 8:914
    [Google Scholar]
  150. 150.
    Saada J, McAuley RJ, Marcatti M, Tang TZ, Motamedi M, Szczesny B. 2022. Oxidative stress induces Z-DNA-binding protein 1–dependent activation of microglia via mtDNA released from retinal pigment epithelial cells. J. Biol. Chem. 298:101523
    [Google Scholar]
  151. 151.
    Baik JY, Liu Z, Jiao D, Kwon H-J, Yan J et al. 2021. ZBP1 not RIPK1 mediates tumor necroptosis in breast cancer. Nat. Commun. 12:2666
    [Google Scholar]
  152. 152.
    Chen D, Tong J, Yang L, Wei L, Stolz DB et al. 2018. PUMA amplifies necroptosis signaling by activating cytosolic DNA sensors. PNAS 115:3930–35
    [Google Scholar]
  153. 153.
    Rai P, Janardhan KS, Meacham J, Madenspacher JH, Lin WC et al. 2021. IRGM1 links mitochondrial quality control to autoimmunity. Nat. Immunol. 22:312–21
    [Google Scholar]
  154. 154.
    Caielli S, Cardenas J, de Jesus AA, Baisch J, Walters L et al. 2021. Erythroid mitochondrial retention triggers myeloid-dependent type I interferon in human SLE. Cell 184:4464–79.e19
    [Google Scholar]
  155. 155.
    Saffran HA, Pare JM, Corcoran JA, Weller SK, Smiley JR. 2007. Herpes simplex virus eliminates host mitochondrial DNA. EMBO Rep. 8:188–93
    [Google Scholar]
  156. 156.
    Duguay BA, Smiley JR. 2013. Mitochondrial nucleases ENDOG and EXOG participate in mitochondrial DNA depletion initiated by herpes simplex virus 1 UL12.5. J. Virol. 87:11787–97
    [Google Scholar]
  157. 157.
    Moriyama M, Koshiba T, Ichinohe T. 2019. Influenza A virus M2 protein triggers mitochondrial DNA-mediated antiviral immune responses. Nat. Commun. 10:4624
    [Google Scholar]
  158. 158.
    Aguirre S, Luthra P, Sanchez-Aparicio MT, Maestre AM, Patel J et al. 2017. Dengue virus NS2B protein targets cGAS for degradation and prevents mitochondrial DNA sensing during infection. Nat. Microbiol. 2:17037
    [Google Scholar]
  159. 159.
    Lai J-H, Wang M-Y, Huang C-Y, Wu C-H, Hung L-F et al. 2018. Infection with the dengue RNA virus activates TLR9 signaling in human dendritic cells. EMBO Rep. 19:e46182
    [Google Scholar]
  160. 160.
    Liu X-N, Li L-W, Gao F, Jiang Y-F, Yuan W-Z et al. 2022. cGAS restricts PRRSV replication by sensing the mtDNA to increase the cGAMP activity. Front. Immunol. 13:887054
    [Google Scholar]
  161. 161.
    Zheng Y, Liu Q, Wu Y, Ma L, Zhang Z et al. 2018. Zika virus elicits inflammation to evade antiviral response by cleaving cGAS via NS1-caspase-1 axis. EMBO J. 37:e99347
    [Google Scholar]
  162. 162.
    Sato H, Hoshi M, Ikeda F, Fujiyuki T, Yoneda M, Kai C 2021. Downregulation of mitochondrial biogenesis by virus infection triggers antiviral responses by cyclic GMP-AMP synthase. PLOS Pathog. 17:e1009841
    [Google Scholar]
  163. 163.
    Domizio JD, Gulen MF, Saidoune F, Thacker VV, Yatim A et al. 2022. The cGAS–STING pathway drives type I IFN immunopathology in COVID-19. Nature 603:145–51
    [Google Scholar]
  164. 164.
    Costa TJ, Potje SR, Fraga-Silva TFC, da Silva-Neto JA, Barros PR et al. 2022. Mitochondrial DNA and TLR9 activation contribute to SARS-CoV-2-induced endothelial cell damage. Vasc. Pharmacol. 142:106946
    [Google Scholar]
  165. 165.
    Scozzi D, Cano M, Ma L, Zhou D, Zhu JH et al. 2021. Circulating mitochondrial DNA is an early indicator of severe illness and mortality from COVID-19. JCI Insight 6:4e143299
    [Google Scholar]
  166. 166.
    Seth RB, Sun L, Ea C-K, Chen ZJ. 2005. Identification and characterization of MAVS, a mitochondrial antiviral signaling protein that activates NF-κB and IRF3. Cell 122:669–82
    [Google Scholar]
  167. 167.
    Jabir MS, Hopkins L, Ritchie ND, Ullah I, Bayes HK et al. 2015. Mitochondrial damage contributes to Pseudomonas aeruginosa activation of the inflammasome and is downregulated by autophagy. Autophagy 11:166–82
    [Google Scholar]
  168. 168.
    Wang B, Zhou C, Wu Q, Lin P, Pu Q et al. 2022. cGAS modulates cytokine secretion and bacterial burdens by altering the release of mitochondrial DNA in pseudomonas pulmonary infection. Immunology 166:408–23
    [Google Scholar]
  169. 169.
    Qin S, Lin P, Wu Q, Pu Q, Zhou C et al. 2020. Small-molecule inhibitor of 8-oxoguanine DNA glycosylase 1 regulates inflammatory responses during Pseudomonas aeruginosa infection. J. Immunol. 205:2231–42
    [Google Scholar]
  170. 170.
    Xue Y, Du M, Zhu M-J. 2017. Quercetin suppresses NLRP3 inflammasome activation in epithelial cells triggered by Escherichia coli O157:H7. Free Radical Biol. Med. 108:760–69
    [Google Scholar]
  171. 171.
    Xu L, Li M, Yang Y, Zhang C, Xie Z et al. 2022. Salmonella induces the cGAS-STING-dependent type I interferon response in murine macrophages by triggering mtDNA release. mBio 13:e03632-21
    [Google Scholar]
  172. 172.
    Wiens KE, Ernst JD. 2016. The mechanism for type I interferon induction by Mycobacterium tuberculosis is bacterial strain-dependent. PLOS Pathog. 12:e1005809
    [Google Scholar]
  173. 173.
    Weindel CG, Martinez EL, Zhao X, Mabry CJ, Bell SL et al. 2022. Mitochondrial ROS promotes susceptibility to infection via gasdermin D-mediated necroptosis. Cell 185:3214–31.e23
    [Google Scholar]
  174. 174.
    Kim B-R, Kim B-J, Kook Y-H, Kim B-J. 2020. Mycobacterium abscessus infection leads to enhanced production of type 1 interferon and NLRP3 inflammasome activation in murine macrophages via mitochondrial oxidative stress. PLOS Pathog. 16:e1008294
    [Google Scholar]
  175. 175.
    Jang JY, Blum A, Liu J, Finkel T. 2018. The role of mitochondria in aging. J. Clin. Investig. 128:3662–70
    [Google Scholar]
  176. 176.
    Franceschi C, Garagnani P, Parini P, Giuliani C, Santoro A. 2018. Inflammaging: a new immune–metabolic viewpoint for age-related diseases. Nat. Rev. Endocrinol. 14:576–90
    [Google Scholar]
  177. 177.
    Shadel GS, Horvath TL. 2015. Mitochondrial ROS signaling in organismal homeostasis. Cell 163:560–69
    [Google Scholar]
  178. 178.
    Pinti M, Cevenini E, Nasi M, De Biasi S, Salvioli S et al. 2014. Circulating mitochondrial DNA increases with age and is a familiar trait: implications for “inflamm-aging.”. Eur. J. Immunol. 44:1552–62
    [Google Scholar]
  179. 179.
    Moon JS, Goeminne LJE, Kim JT, Tian JW, Kim SH et al. 2020. Growth differentiation factor 15 protects against the aging-mediated systemic inflammatory response in humans and mice. Aging Cell 19:e13195
    [Google Scholar]
  180. 180.
    Verschoor CP, Loukov D, Naidoo A, Puchta A, Johnstone J et al. 2015. Circulating TNF and mitochondrial DNA are major determinants of neutrophil phenotype in the advanced-age, frail elderly. Mol. Immunol. 65:148–56
    [Google Scholar]
  181. 181.
    Lazo S, Noren Hooten N, Green J, Eitan E, Mode NA et al. 2021. Mitochondrial DNA in extracellular vesicles declines with age. Aging Cell 20:e13283
    [Google Scholar]
  182. 182.
    Chocron ES, Munkácsy E, Pickering AM. 2019. Cause or casualty: the role of mitochondrial DNA in aging and age-associated disease. Biochim. Biophys. Acta Mol. Basis Dis. 1865:285–97
    [Google Scholar]
  183. 183.
    Kaarniranta K, Uusitalo H, Blasiak J, Felszeghy S, Kannan R et al. 2020. Mechanisms of mitochondrial dysfunction and their impact on age-related macular degeneration. Prog. Retin. Eye Res. 79:100858
    [Google Scholar]
  184. 184.
    Fukuda S, Narendran S, Varshney A, Nagasaka Y, Wang SB et al. 2021. Alu complementary DNA is enriched in atrophic macular degeneration and triggers retinal pigmented epithelium toxicity via cytosolic innate immunity. Sci. Adv. 7:eabj3658
    [Google Scholar]
  185. 185.
    Jauhari A, Baranov SV, Suofu Y, Kim J, Singh T et al. 2020. Melatonin inhibits cytosolic mitochondrial DNA–induced neuroinflammatory signaling in accelerated aging and neurodegeneration. J. Clin. Investig. 130:3124–36
    [Google Scholar]
  186. 186.
    Zhong W, Rao Z, Xu J, Sun Y, Hu H et al. 2022. Defective mitophagy in aged macrophages promotes mitochondrial DNA cytosolic leakage to activate STING signaling during liver sterile inflammation. Aging Cell 21:e13622
    [Google Scholar]
  187. 187.
    Liu J, Chen H, Lin X, Yi J, Ye W et al. 2022. Age-related activation of cyclic GMP–AMP synthase–stimulator of interferon genes signaling in the auditory system is associated with presbycusis in C57BL/6J male mice. Neuroscience 481:73–84
    [Google Scholar]
  188. 188.
    Lei Y, Guerra Martinez C, Torres-Odio S, Bell SL, Birdwell CE et al. 2021. Elevated type I interferon responses potentiate metabolic dysfunction, inflammation, and accelerated aging in mtDNA mutator mice. Sci Adv 7:eabe7548
    [Google Scholar]
  189. 189.
    Wiley CD, Velarde MC, Lecot P, Liu S, Sarnoski EA et al. 2016. Mitochondrial dysfunction induces senescence with a distinct secretory phenotype. Cell Metab. 23:303–14
    [Google Scholar]
  190. 190.
    Correia-Melo C, Marques FD, Anderson R, Hewitt G, Hewitt R et al. 2016. Mitochondria are required for pro-ageing features of the senescent phenotype. EMBO J. 35:724–42
    [Google Scholar]
  191. 191.
    Vizioli MG, Liu T, Miller KN, Robertson NA, Gilroy K et al. 2020. Mitochondria-to-nucleus retrograde signaling drives formation of cytoplasmic chromatin and inflammation in senescence. Genes Dev. 34:428–45
    [Google Scholar]
  192. 192.
    Martini H, Passos JF 2023. Cellular senescence: All roads lead to mitochondria. FEBS J. 290:1186–202
    [Google Scholar]
  193. 193.
    Glück S, Ablasser A. 2019. Innate immunosensing of DNA in cellular senescence. Curr. Opin. Immunol. 56:31–36
    [Google Scholar]
  194. 194.
    Lv N, Zhao Y, Liu X, Ye L, Liang Z et al. 2022. Dysfunctional telomeres through mitostress-induced cGAS/STING activation to aggravate immune senescence and viral pneumonia. Aging Cell 21:e13594
    [Google Scholar]
  195. 195.
    Schuliga M, Read J, Blokland KEC, Waters DW, Burgess J et al. 2020. Self DNA perpetuates IPF lung fibroblast senescence in a cGAS-dependent manner. Clin. Sci. 134:889–905
    [Google Scholar]
  196. 196.
    Schuliga M, Kanwal A, Read J, Blokland KEC, Burgess JK et al. 2021. A cGAS-dependent response links DNA damage and senescence in alveolar epithelial cells: a potential drug target in IPF. Am. J. Physiol. Lung Cell. Mol. Physiol. 321:L859–71
    [Google Scholar]
  197. 197.
    Passos J, Chapman J, Salmonowicz H, Victorelli S, Martini H et al. 2022. Sub-lethal apoptotic stress enables mtDNA release during senescence and drives the SASP. Preprint, Res. Sq.. https://assets.researchsquare.com/files/rs-1247316/v1_covered.pdf?c=1646941050
  198. 198.
    Tran M, Reddy PH. 2021. Defective autophagy and mitophagy in aging and Alzheimer's disease. Front. Neurosci. 14:612757
    [Google Scholar]
  199. 199.
    Hajizadeh S, DeGroot J, TeKoppele JM, Tarkowski A, Collins LV. 2003. Extracellular mitochondrial DNA and oxidatively damaged DNA in synovial fluid of patients with rheumatoid arthritis. Arthritis Res. Ther. 5:R234
    [Google Scholar]
  200. 200.
    Zhang B, Asadi S, Weng Z, Sismanopoulos N, Theoharides TC. 2012. Stimulated human mast cells secrete mitochondrial components that have autocrine and paracrine inflammatory actions. PLOS ONE 7:e49767
    [Google Scholar]
  201. 201.
    Poli C, Augusto JF, Dauvé J, Adam C, Preisser L et al. 2017. IL-26 confers proinflammatory properties to extracellular DNA. J. Immunol. 198:3650–61
    [Google Scholar]
  202. 202.
    Bryant JD, Lei Y, VanPortfliet JJ, Winters AD, West AP. 2022. Assessing mitochondrial DNA release into the cytosol and subsequent activation of innate immune-related pathways in mammalian cells. Curr. Protocols 2:e372
    [Google Scholar]
  203. 203.
    Hepokoski ML, Odish M, Lam MT, Coufal NG, Rolfsen ML et al. 2022. Absolute quantification of plasma mitochondrial DNA by droplet digital PCR marks COVID-19 severity over time during intensive care unit admissions. Am. J. Physiol. Lung Cell. Mol. Physiol. 323:L84–92
    [Google Scholar]
  204. 204.
    Ban-Ishihara R, Ishihara T, Sasaki N, Mihara K, Ishihara N. 2013. Dynamics of nucleoid structure regulated by mitochondrial fission contributes to cristae reformation and release of cytochrome c. PNAS 110:11863–68
    [Google Scholar]
  205. 205.
    Mussil B, Suspène R, Caval V, Durandy A, Wain-Hobson S, Vartanian J-P. 2019. Genotoxic stress increases cytoplasmic mitochondrial DNA editing by human APOBEC3 mutator enzymes at a single cell level. Sci. Rep. 9:3109
    [Google Scholar]
  206. 206.
    Anindya R. 2022. Cytoplasmic DNA in cancer cells: several pathways that potentially limit DNase2 and TREX1 activities. Biochim. Biophys. Acta Mol. Cell Res. 1869:119278
    [Google Scholar]
  207. 207.
    Van Der Burgh R, Nijhuis L, Pervolaraki K, Compeer EB, Jongeneel LH et al. 2014. Defects in mitochondrial clearance predispose human monocytes to interleukin-1β hypersecretion. J. Biol. Chem. 289:5000–12
    [Google Scholar]
  208. 208.
    Gkirtzimanaki K, Kabrani E, Nikoleri D, Polyzos A, Blanas A et al. 2018. IFNα impairs autophagic degradation of mtDNA promoting autoreactivity of SLE monocytes in a STING-dependent fashion. Cell Rep 25:921–33.e5
    [Google Scholar]
  209. 209.
    Kuck JL, Obiako BO, Gorodnya OM, Pastukh VM, Kua J et al. 2015. Mitochondrial DNA damage-associated molecular patterns mediate a feed-forward cycle of bacteria-induced vascular injury in perfused rat lungs. Am. J. Physiol. Lung Cell. Mol. Physiol. 308:L1078–85
    [Google Scholar]
  210. 210.
    Yao X, Carlson D, Sun Y, Ma L, Wolf SE et al. 2015. Mitochondrial ROS induces cardiac inflammation via a pathway through mtDNA damage in a pneumonia-related sepsis model. PLOS ONE 10:e0139416
    [Google Scholar]
  211. 211.
    Tumburu L, Ghosh-Choudhary S, Seifuddin FT, Barbu EA, Yang S et al. 2021. Circulating mitochondrial DNA is a proinflammatory DAMP in sickle cell disease. Blood 137:3116–26
    [Google Scholar]
  212. 212.
    Wu Z, Oeck S, West AP, Mangalhara KC, Sainz AG et al. 2019. Mitochondrial DNA stress signaling protects the nuclear genome. Nat. Metab. 1:1209–18
    [Google Scholar]
  213. 213.
    Ding Z, Liu S, Wang X, Khaidakov M, Dai Y, Mehta JL. 2013. Oxidant stress in mitochondrial DNA damage, autophagy and inflammation in atherosclerosis. Sci. Rep. 3:1077
    [Google Scholar]
  214. 214.
    Cobo I, Tanaka TN, Mangalhara KC, Lana A, Yeang C et al. 2022. DNMT3A and TET2 restrain mitochondrial DNA-mediated interferon signaling in macrophages. Immunity 55:81386–401.e10
    [Google Scholar]
  215. 215.
    Nie S, Lu J, Wang L, Gao M 2020. Pro-inflammatory role of cell-free mitochondrial DNA in cardiovascular diseases. IUBMB Life 72:1879–90
    [Google Scholar]
  216. 216.
    Cotticelli MG, Xia S, Truitt R, Doliba NM, Rozo AV et al. 2023. Acute frataxin knockdown in induced pluripotent stem cell-derived cardiomyocytes activates a type I interferon response. Dis. Model Mech. 16:5dmm049497
    [Google Scholar]
  217. 217.
    McCarthy CG, Wenceslau CF, Goulopoulou S, Ogbi S, Baban B et al. 2015. Circulating mitochondrial DNA and Toll-like receptor 9 are associated with vascular dysfunction in spontaneously hypertensive rats. Cardiovasc. Res. 107:119–30
    [Google Scholar]
  218. 218.
    Guo MM-H, Huang Y-H, Wang F-S, Chang L-S, Chen K-D, Kuo H-C. 2022. CD36 is associated with the development of coronary artery lesions in patients with Kawasaki disease. Front. Immunol. 13:790095
    [Google Scholar]
  219. 219.
    Boyapati RK, Dorward DA, Tamborska A, Kalla R, Ventham NT et al. 2018. Mitochondrial DNA is a pro-inflammatory damage-associated molecular pattern released during active IBD. Inflamm. Bowel Dis. 24:2113–22
    [Google Scholar]
  220. 220.
    Marques PE, Amaral SS, Pires DA, Nogueira LL, Soriani FM et al. 2012. Chemokines and mitochondrial products activate neutrophils to amplify organ injury during mouse acute liver failure. Hepatology 56:1971–82
    [Google Scholar]
  221. 221.
    Garcia-Martinez I, Santoro N, Chen Y, Hoque R, Ouyang X et al. 2016. Hepatocyte mitochondrial DNA drives nonalcoholic steatohepatitis by activation of TLR9. J. Clin. Investig. 126:859–64
    [Google Scholar]
  222. 222.
    An P, Wei L-L, Zhao S, Sverdlov DY, Vaid KA et al. 2020. Hepatocyte mitochondria-derived danger signals directly activate hepatic stellate cells and drive progression of liver fibrosis. Nat. Commun. 11:2362
    [Google Scholar]
  223. 223.
    Bae JH, Jo SI, Kim SJ, Lee JM, Jeong JH et al. 2019. Circulating cell-free mtDNA contributes to AIM2 inflammasome-mediated chronic inflammation in patients with type 2 diabetes. Cells 8:328
    [Google Scholar]
  224. 224.
    Yu C-H, Davidson S, Harapas CR, Hilton JB, Mlodzianoski MJ et al. 2020. TDP-43 triggers mitochondrial DNA release via mPTP to activate cGAS/STING in ALS. Cell 183:636–49.e18
    [Google Scholar]
  225. 225.
    Long G, Gong R, Wang Q, Zhang D, Huang C. 2022. Role of released mitochondrial DNA in acute lung injury. Front. Immunol. 13:973089
    [Google Scholar]
  226. 226.
    Tsuji N, Tsuji T, Ohashi N, Kato A, Fujigaki Y, Yasuda H. 2016. Role of mitochondrial DNA in septic AKI via Toll-like receptor 9. J. Am. Soc. Nephrol. 27:2009–20
    [Google Scholar]
  227. 227.
    Liu J, Jia Z, Gong W. 2021. Circulating mitochondrial DNA stimulates innate immune signaling pathways to mediate acute kidney injury. Front. Immunol. 12:680648
    [Google Scholar]
  228. 228.
    Hepokoski M, Wang J, Li K, Li Y, Gupta P et al. 2021. Altered lung metabolism and mitochondrial DAMPs in lung injury due to acute kidney injury. Am. J. Physiol. Lung Cell. Mol. Physiol. 320:L821–31
    [Google Scholar]
  229. 229.
    Comish PB, Liu M-M, Huebinger R, Carlson D, Kang R, Tang D. 2022. The cGAS-STING pathway connects mitochondrial damage to inflammation in burn-induced acute lung injury in rat. Burns 48:168–75
    [Google Scholar]
  230. 230.
    Yamanouchi S, Kudo D, Yamada M, Miyagawa N, Furukawa H, Kushimoto S. 2013. Plasma mitochondrial DNA levels in patients with trauma and severe sepsis: time course and the association with clinical status. J. Crit. Care 28:1027–31
    [Google Scholar]
  231. 231.
    Simmons JD, Lee YL, Mulekar S, Kuck JL, Brevard SB et al. 2013. Elevated levels of plasma mitochondrial DNA DAMPs are linked to clinical outcome in severely injured human subjects. Ann. Surg. 258:591–6; discussion 96–8
    [Google Scholar]
  232. 232.
    Jeon H, Lee J, Lee S, Kang SK, Park SJ et al. 2019. Extracellular vesicles from KSHV-infected cells stimulate antiviral immune response through mitochondrial DNA. Front. Immunol. 10:876
    [Google Scholar]
/content/journals/10.1146/annurev-biochem-032620-104401
Loading
/content/journals/10.1146/annurev-biochem-032620-104401
Loading

Data & Media loading...

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error