AI Chat Paper
Note: Please note that the following content is generated by AMiner AI. SciOpen does not take any responsibility related to this content.
{{lang === 'zh_CN' ? '文章概述' : 'Summary'}}
{{lang === 'en_US' ? '中' : 'Eng'}}
Chat more with AI
PDF (1.2 MB)
Collect
Submit Manuscript AI Chat Paper
Show Outline
Outline
Show full outline
Hide outline
Outline
Show full outline
Hide outline
Review Article | Open Access

Decoding Neurological Mysteries: The Potential Impact of Endogenous Retroviruses on Brain Health

Jiaqi Li1,Liyong Liao1,Xixi Liu1,Yueyan Zhu1Daijing Sun1Chenchun Zhang1Yan Jiang1( )
Institutes of Brain Science, State Key Laboratory of Medical Neurobiology and MOE Frontiers Center for Brain Science, Fudan University, 200032 Shanghai, China

JL, LL, and XL contributed equally to this work

Show Author Information

Abstract

Endogenous retroviruses (ERVs), remnants of ancient viral invasions, make up a significant part of the mammalian genomes. ERVs are typically held in check by complex epigenetic mechanisms, which serve to limit their expansion and potential adverse effects on the genome. However, ERVs can become aberrantly activated in response to stressful challenges, contributing to progression in pathological conditions, including cancer, inflammation, auto-immune disorders, and aging, through various mechanisms. Notably, ERV activation is also detected in the brain and is increasingly recognized as an important factor in neuropsychiatric disorders. In this review, we encapsulate the general understanding of ERVs in both physiological and pathological states and compiled evidence for ERV activation across a spectrum of neuropsychiatric disorders, along with current studies exploring the underlying mechanisms. Despite the accumulating body of evidence, research in this field remains in its infancy and faces substantial challenges. Further studies are warranted to enhance our understanding of ERV activation mechanisms and their roles in neuropsychological conditions, potentially contributing to the development of innovative therapeutic interventions.

References

[1]

Han, M., Perkins, M. H., Novaes, L. S., Xu, T., Chang, H. Advances in transposable elements: From mechanisms to applications in mammalian genomics. Frontiers in Genetics, 2023, 14: 1290146.

[2]

Chuong, E. B., Elde, N. C., Feschotte, C. Regulatory activities of transposable elements: From conflicts to benefits. Nature Reviews Genetics, 2017, 18(2): 71–86.

[3]

Mager, D. L., Stoye, J. P. Mammalian endogenous retroviruses. Microbiology Spectrum, 2015, 3(1): p. MDNA3-0009-2014.

[4]

Duffy, S., Shackelton, L. A., Holmes, E. C. Rates of evolutionary change in viruses: Patterns and determinants. Nature Reviews Genetics, 2008, 9(4): 267–276.

[5]

Zheng, J. L., Wei, Y. T., Han, G. Z. The diversity and evolution of retroviruses: Perspectives from viral “fossils”. Virologica Sinica, 2022, 37(1): 11–18.

[6]

Gifford, R. J., Blomberg, J., Coffin, J. M., Fan, H., Heidmann, T., Mayer, J., Stoye, J., Tristem, M., Johnson, W. E. Nomenclature for endogenous retrovirus (ERV) loci. Retrovirology, 2018, 15: 59.

[7]

Garcia-Montojo, M., Doucet-O’Hare, T., Henderson, L., Nath, A. Human endogenous retrovirus-K (HML-2): A comprehensive review. Critical Reviews in Microbiology, 2018, 44(6): 715–738.

[8]

Giménez-Orenga, K., Oltra, E. Human endogenous retrovirus as therapeutic targets in neurologic disease. Pharmaceuticals, 2021, 14(6): 495.

[9]

Zhu, Y. Y., Sun, D. J., Jakovcevski, M., Jiang, Y. Epigenetic mechanism of SETDB1 in brain: Implications for neuropsychiatric disorders. Translational Psychiatry, 2020, 10: 115.

[10]

Markouli, M., Strepkos, D., Chlamydas, S., Piperi, C. Histone lysine methyltransferase SETDB1 as a novel target for central nervous system diseases. Progress in Neurobiology, 2021, 200: 101968.

[11]

Matsui, T., Leung, D., Miyashita, H., Maksakova, I. A., Miyachi, H., Kimura, H., Tachibana, M., Lorincz, M. C., Shinkai, Y. Proviral silencing in embryonic stem cells requires the histone methyltransferase ESET. Nature, 2010, 464(7290): 927–931.

[12]

Gautam, P., Yu, T., Loh, Y. H. Regulation of ERVs in pluripotent stem cells and reprogramming. Current Opinion in Genetics & Development, 2017, 46: 194–201.

[13]

Wolf, G., Yang, P., Füchtbauer, A. C., Füchtbauer, E. M., Silva, A. M., Park, C., Wu, W., Nielsen, A. L., Pedersen, F. S., MacFarlan, T. S. The KRAB zinc finger protein ZFP809 is required to initiate epigenetic silencing of endogenous retroviruses. Genes & Development, 2015, 29(5): 538–554.

[14]

Wolf, D., Goff, S. P. Embryonic stem cells use ZFP809 to silence retroviral DNAs. Nature, 2009, 458(7242): 1201–1204.

[15]

Jacobs, F. M. J., Greenberg, D., Nguyen, N., Haeussler, M., Ewing, A. D., Katzman, S., Paten, B., Salama, S. R., Haussler, D. An evolutionary arms race between KRAB zinc-finger genes ZNF91/93 and SVA/L1 retrotransposons. Nature, 2014, 516(7530): 242–245.

[16]

Thomas, J. H., Schneider, S. Coevolution of retroelements and tandem zinc finger genes. Genome Research, 2011, 21(11): 1800–1812.

[17]

Imbeault, M., Helleboid, P. Y., Trono, D. KRAB zinc-finger proteins contribute to the evolution of gene regulatory networks. Nature, 2017, 543(7646): 550–554.

[18]

Rowe, H. M., Jakobsson, J., Mesnard, D., Rougemont, J., Reynard, S., Aktas, T., Maillard, P. V., Layard-Liesching, H., Verp, S., Marquis, J. et al. KAP1 controls endogenous retroviruses in embryonic stem cells. Nature, 2010, 463(7278): 237–240.

[19]

Tie, C. H., Fernandes, L., Conde, L., Robbez-Masson, L., Sumner, R. P., Peacock, T., Rodriguez-Plata, M. T., Mickute, G., Gifford, R., Towers, G. J. et al. KAP1 regulates endogenous retroviruses in adult human cells and contributes to innate immune control. EMBO Rep, 2018, 19(10): e45000.

[20]

Li, J. Q., Zheng, S. H., Dong, Y. H., Xu, H., Zhu, Y. Y., Weng, J., Sun, D. J., Wang, S. Y., Xiao, L., Jiang, Y. Histone methyltransferase SETDB1 regulates the development of cortical Htr3a-positive interneurons and mood behaviors. Biological Psychiatry, 2023, 93(3): 279–290.

[21]

Tan, S. L., Nishi, M., Ohtsuka, T., Matsui, T., Takemoto, K., Kamio-Miura, A., Aburatani, H., Shinkai, Y., Kageyama, R. Essential roles of the histone methyltransferase ESET in the epigenetic control of neural progenitor cells during development. Development, 2012, 139(20): 3806–3816.

[22]

Jönsson, M. E., Garza, R., Sharma, Y., Petri, R., Södersten, E., Johansson, J. G., Johansson, P. A., Atacho, D. A., Pircs, K., Madsen, S. et al. Activation of endogenous retroviruses during brain development causes an inflammatory response. EMBO J, 2021, 40(9): e106423.

[23]

Zhao, S., Lu, J. W., Pan, B., Fan, H. T., Byrum, S. D., Xu, C. X., Kim, A., Guo, Y. R., Kanchi, K. L., Gong, W. D. et al. TNRC18 engages H3K9me3 to mediate silencing of endogenous retrotransposons. Nature, 2023, 623(7987): 633–642.

[24]

Yang, bin xia, EL Farran, C., Guo, hong chao, Yu, T., Fang, hai tong, Wang, hao fei, Schlesinger, S., Seah, Y., Goh, G., Neo, S. et al. Systematic identification of factors for provirus silencing in embryonic stem cells. Cell, 2015, 163(1): 230–245.

[25]

Tessier, S., Ferhi, O., Geoffroy, M. C., González-Prieto, R., Canat, A., Quentin, S., Pla, M., Niwa-Kawakita, M., Bercier, P., Rérolle, D. et al. Exploration of nuclear body-enhanced sumoylation reveals that PML represses 2-cell features of embryonic stem cells. Nature Communications, 2022, 13: 5726.

[26]

Rosendorff, A., Sakakibara, S., Lu, S. X., Kieff, E., Xuan, Y., DiBacco, A., Shi, Y. J., Shi, Y., Gill, G. NXP-2 association with SUMO-2 depends on lysines required for transcriptional repression. Proceedings of the National Academy of Sciences of the United States of America, 2006, 103(14): 5308–5313.

[27]

Fukuda, K., Okuda, A., Yusa, K., Shinkai, Y. A CRISPR knockout screen identifies SETDB1-target retroelement silencing factors in embryonic stem cells. Genome Research, 2018, 28(6): 846–858.

[28]

Tchasovnikarova, I. A., Timms, R. T., Matheson, N. J., Wals, K., Antrobus, R., Göttgens, B., Dougan, G., Dawson, M. A., Lehner, P. J. Epigenetic silencing by the HUSH complex mediates position-effect variegation in human cells. Science, 2015, 348(6242): 1481–1485.

[29]

Walsh, C.P., J.R. Chaillet, and T.H. Bestor. Transcription of IAP endogenous retroviruses is constrained by cytosine methylation. Nat Genet, 1998, 20(2): 116–117.

[30]
Lei H. Oh S. P. Okano M. Jüttermann R. Goss K. A. Jaenisch R. Li E. De novo DNA cytosine methyltransferase activities in mouse embryonic stem cells Development 1996 122 10 3195 3205 10.1242/dev.122.10.3195

Lei, H., Oh, S. P., Okano, M., Jüttermann, R., Goss, K. A., Jaenisch, R., Li, E. De novo DNA cytosine methyltransferase activities in mouse embryonic stem cells. Development, 1996, 122(10): 3195–3205.

[31]

Barau, J., Teissandier, A., Zamudio, N., Roy, S., Nalesso, V., Hérault, Y., Guillou, F., Bourc’his, D. The DNA methyltransferase DNMT3C protects male germ cells from transposon activity. Science, 2016, 354(6314): 909–912.

[32]

Min, B., Park, J. S., Jeong, Y. S., Jeon, K., Kang, Y. K. Dnmt1 binds and represses genomic retroelements via DNA methylation in mouse early embryos. Nucleic Acids Research, 2020, 48(15): 8431–8444.

[33]

Chen, T. P., Ueda, Y., Dodge, J. E., Wang, Z. J., Li, E. Establishment and maintenance of genomic methylation patterns in mouse embryonic stem cells by Dnmt3a and Dnmt3b. Molecular and Cellular Biology, 2003, 23(16): 5594–5605.

[34]

Bourc’his, D., Bestor, T. H. Meiotic catastrophe and retrotransposon reactivation in male germ cells lacking Dnmt3L. Nature, 2004, 431(7004): 96–99.

[35]

Ito, S., D’Alessio, A. C., Taranova, O. V., Hong, K., Sowers, L. C., Zhang, Y. Role of Tet proteins in 5mC to 5hmC conversion, ES-cell self-renewal and inner cell mass specification. Nature, 2010, 466(7310): 1129–1133.

[36]

He, Y. F., Li, B. Z., Li, Z., Liu, P., Wang, Y., Tang, Q. Y., Ding, J. P., Jia, Y. Y., Chen, Z. C., Li, L. et al. Tet-mediated formation of 5-carboxylcytosine and its excision by TDG in mammalian DNA. Science, 2011, 333(6047): 1303–1307.

[37]

Leung, D., Du, T. T., Wagner, U., Xie, W., Lee, A. Y., Goyal, P., Li, Y. J., Szulwach, K. E., Jin, P., Lorincz, M. C. et al. Regulation of DNA methylation turnover at LTR retrotransposons and imprinted loci by the histone methyltransferase Setdb1. Proceedings of the National Academy of Sciences of the United States of America, 2014, 111(18): 6690–6695.

[38]

Dong, K. B., Maksakova, I. A., Mohn, F., Leung, D., Appanah, R., Lee, S., Yang, H. W., Lam, L. L., Mager, D. L., Schübeler, D. et al. DNA methylation in ES cells requires the lysine methyltransferase G9a but not its catalytic activity. The EMBO Journal, 2008, 27(20): 2691–2701.

[39]

Gaudet, F., Rideout, W. M. Ⅲ, Meissner, A., Dausman, J., Leonhardt, H., Jaenisch, R. Dnmt1 expression in pre- and postimplantation embryogenesis and the maintenance of IAP silencing. Molecular and Cellular Biology, 2004, 24(4): 1640–1648.

[40]

Kurihara, Y., Kawamura, Y., Uchijima, Y., Amamo, T., Kobayashi, H., Asano, T., Kurihara, H. Maintenance of genomic methylation patterns during preimplantation development requires the somatic form of DNA methyltransferase 1. Developmental Biology, 2008, 313(1): 335–346.

[41]

Sharif, J., Muto, M., Takebayashi, S. I., Suetake, I., Iwamatsu, A., Endo, T. A., Shinga, J., Mizutani-Koseki, Y., Toyoda, T., Okamura, K. et al. The SRA protein Np95 mediates epigenetic inheritance by recruiting Dnmt1 to methylated DNA. Nature, 2007, 450(7171): 908–912.

[42]

Rowe, H. M., Trono, D. Dynamic control of endogenous retroviruses during development. Virology, 2011, 411(2): 273–287.

[43]

Okano, M., Bell, D. W., Haber, D. A., Li, E. DNA methyltransferases Dnmt3a and Dnmt3b are essential for de novo methylation and mammalian development. Cell, 1999, 99(3): 247–257.

[44]

Suetake, I., Shinozaki, F., Miyagawa, J., Takeshima, H., Tajima, S. DNMT3L stimulates the DNA methylation activity of Dnmt3a and Dnmt3b through a direct interaction. Journal of Biological Chemistry, 2004, 279(26): 27816–27823.

[45]

Rowe, H. M., Friedli, M., Offner, S., Verp, S., Mesnard, D., Marquis, J., Aktas, T., Trono, D. De novo DNA methylation of endogenous retroviruses is shaped by KRAB-ZFPs/KAP1 and ESET. Development, 2013, 140(3): 519–529.

[46]

Sharif, J., Endo, T., Nakayama, M., Karimi, M., Shimada, M., Katsuyama, K., Goyal, P., Brind’Amour, J., Sun, M. A., Sun, Z. X. et al. Activation of endogenous retroviruses in Dnmt1–/–ESCs involves disruption of SETDB1-mediated repression by NP95 binding to hemimethylated DNA. Cell Stem Cell, 2016, 19(1): 81–94.

[47]

Groh, S., Schotta, G. Silencing of endogenous retroviruses by heterochromatin. Cellular and Molecular Life Sciences, 2017, 74(11): 2055–2065.

[48]

Wang, Z. Y., Fan, R., Russo, A., Cernilogar, F. M., Nuber, A., Schirge, S., Shcherbakova, I., Dzhilyanova, I., Ugur, E., Anton, T. et al. Dominant role of DNA methylation over H3K9me3 for IAP silencing in endoderm. Nature Communications, 2022, 13: 5447.

[49]

Ohtani, H., Liu, M. M., Zhou, W. D., Liang, G. N., Jones, P. A. Switching roles for DNA and histone methylation depend on evolutionary ages of human endogenous retroviruses. Genome Research, 2018, 28(8): 1147–1157.

[50]

Hutnick, L. K., Golshani, P., Namihira, M., Xue, Z. G., Matynia, A., Yang, X. W., Silva, A. J., Schweizer, F. E., Fan, G. P. DNA hypomethylation restricted to the murine forebrain induces cortical degeneration and impairs postnatal neuronal maturation. Human Molecular Genetics, 2009, 18(15): 2875–2888.

[51]

Bulut-Karslioglu, A., De La Rosa-Velázquez, I., Ramirez, F., Barenboim, M., Onishi-Seebacher, M., Arand, J., Galán, C., Winter, G., Engist, B., Gerle, B. et al. Suv39h-dependent H3K9me3 marks intact retrotransposons and silences LINE elements in mouse embryonic stem cells. Molecular Cell, 2014, 55(2): 277–290.

[52]

Maksakova, I. A., Thompson, P. J., Goyal, P., Jones, S. J., Singh, P. B., Karimi, M. M., Lorincz, M. C. Distinct roles of KAP1, HP1 and G9a/GLP in silencing of the two-cell-specific retrotransposon MERVL in mouse ES cells. Epigenetics & Chromatin, 2013, 6(1): 15.

[53]

Loyola, A., Tagami, H., Bonaldi, T., Roche, D., Quivy, J. P., Imhof, A., Nakatani, Y., Dent, S. Y. R., Almouzni, G. The HP1α–CAF1–SetDB1-containing complex provides H3K9me1 for Suv39-mediated K9me3 in pericentric heterochromatin. EMBO Reports, 2009, 10(7): 769–775.

[54]

Leeb, M., Pasini, D., Novatchkova, M., Jaritz, M., Helin, K., Wutz, A. Polycomb complexes act redundantly to repress genomic repeats and genes. Genes & Development, 2010, 24(3): 265–276.

[55]

Elsässer, S. J., Noh, K. M., Diaz, N., Allis, C. D., Banaszynski, L. A. Histone H3.3 is required for endogenous retroviral element silencing in embryonic stem cells. Nature, 2015, 522(7555): 240–244.

[56]

Sugimoto, J., Sugimoto, M., Bernstein, H., Jinno, Y., Schust, D. A novel human endogenous retroviral protein inhibits cell-cell fusion. Scientific Reports, 2013, 3: 1462.

[57]

Esnault, C., Priet, S., Ribet, D., Vernochet, C., Bruls, T., Lavialle, C., Weissenbach, J., Heidmann, T. A placenta-specific receptor for the fusogenic, endogenous retrovirus-derived, human syncytin-2. Proceedings of the National Academy of Sciences of the United States of America, 2008, 105(45): 17532–17537.

[58]

Frank, J. A., Singh, M., Cullen, H. B., Kirou, R. A., Benkaddour-Boumzaouad, M., Cortes, J. L., Garcia Pérez, J., Coyne, C. B., Feschotte, C. Evolution and antiviral activity of a human protein of retroviral origin. Science, 2022, 378(6618): 422–428.

[59]

Johnson, W. E. Origins and evolutionary consequences of ancient endogenous retroviruses. Nature Reviews Microbiology, 2019, 17(6): 355–370.

[60]

Grow, E. J., Flynn, R. A., Chavez, S. L., Bayless, N. L., Wossidlo, M., Wesche, D. J., Martin, L., Ware, C. B., Blish, C. A., Chang, H. Y. et al. Intrinsic retroviral reactivation in human preimplantation embryos and pluripotent cells. Nature, 2015, 522(7555): 221–225.

[61]

Honda, S., Hatamura, M., Kunimoto, Y., Ikeda, S., Minami, N. Chimeric PRMT6 protein produced by an endogenous retrovirus promoter regulates cell fate decision in mouse preimplantation embryos. Biol Reprod, 2024: ioae002.

[62]

Zhang, W. C., Wu, J., Ward, M., Yang, S., Chuang, Y. A., Xiao, M. F., Li, R. J., Leahy, D., Worley, P. Structural basis of arc binding to synaptic proteins: Implications for cognitive disease. Neuron, 2015, 86(2): 490–500.

[63]

Pastuzyn, E. D., Day, C. E., Kearns, R. B., Kyrke-Smith, M., Taibi, A. V., McCormick, J., Yoder, N., Belnap, D. M., Erlendsson, S., Morado, D. R. et al. The neuronal gene arc encodes a repurposed retrotransposon gag protein that mediates intercellular RNA transfer. Cell, 2018, 172(1–2): 275–288.e18.

[64]

Ashley, J., Cordy, B., Lucia, D., Fradkin, L. G., Budnik, V., Thomson, T. Retrovirus-like gag protein Arc1 binds RNA and traffics across synaptic boutons. Cell, 2018, 172(1–2): 262–274.e11.

[65]

Hendrickson, P. G., Doráis, J. A., Grow, E. J., Whiddon, J. L., Lim, J. W., Wike, C. L., Weaver, B. D., Pflueger, C., Emery, B. R., Wilcox, A. L. et al. Conserved roles of mouse DUX and human DUX4 in activating cleavage-stage genes and MERVL/HERVL retrotransposons. Nature Genetics, 2017, 49(6): 925–934.

[66]

MacFarlan, T. S., Gifford, W. D., Driscoll, S., Lettieri, K., Rowe, H. M., Bonanomi, D., Firth, A., Singer, O., Trono, D., Pfaff, S. L. Embryonic stem cell potency fluctuates with endogenous retrovirus activity. Nature, 2012, 487(7405): 57–63.

[67]

Sakashita, A., Kitano, T., Ishizu, H., Guo, Y. J., Masuda, H., Ariura, M., Murano, K., Siomi, H. Transcription of MERVL retrotransposons is required for preimplantation embryo development. Nature Genetics, 2023, 55(3): 484–495.

[68]

Chuong, E. B., Karim Rumi, M. A., Soares, M. J., Baker, J. C. Endogenous retroviruses function as species-specific enhancer elements in the placenta. Nature Genetics, 2013, 45(3): 325–329.

[69]

Chuong, E. B., Elde, N. C., Feschotte, C. Regulatory evolution of innate immunity through co-option of endogenous retroviruses. Science, 2016, 351(6277): 1083–1087.

[70]

Zhang, Y., Li, T., Preissl, S., et al. Transcriptionally active HERV-H retrotransposons demarcate topologically associating domains in human pluripotent stem cells. Nat Genet, 2019, 51(9): 1380–1388.

[71]

Song, M., Pebworth, M. P., Yang, X. Y., Abnousi, A., Fan, C. X., Wen, J., Rosen, J. D., Choudhary, M. N. K., Cui, X. K., Jones, I. R. et al. Cell-type-specific 3D epigenomes in the developing human cortex. Nature, 2020, 587(7835): 644–649.

[72]

Downey, R. F., Sullivan, F. J., Wang-Johanning, F., Ambs, S., Giles, F. J., Glynn, S. A. Human endogenous retrovirus K and cancer: Innocent bystander or tumorigenic accomplice? International Journal of Cancer, 2015, 137(6): 1249–1257.

[73]

Kassiotis, G. Endogenous retroviruses and the development of cancer. The Journal of Immunology, 2014, 192(4): 1343–1349.

[74]

Büscher, K., Hahn, S., Hofmann, M., Trefzer, U., Özel, M., Sterry, W., Löwer, J., Löwer, R., Kurth, R., Denner, J. Expression of the human endogenous retrovirus-K transmembrane envelope, Rec and Np9 proteins in melanomas and melanoma cell lines. Melanoma Research, 2006, 16(3): 223–234.

[75]

Wang-Johanning, F., Liu, J. S., Rycaj, K., Huang, M., Tsai, K., Rosen, D. G., Chen, D. T., Lu, D. W., Barnhart, K. F., Johanning, G. L. Expression of multiple human endogenous retrovirus surface envelope proteins in ovarian cancer. International Journal of Cancer, 2007, 120(1): 81–90.

[76]

Ejthadi, H. D., Martin, J. H., Junying, J., Roden, D. A., Lahiri, M., Warren, P., Murray, P. G., Nelson, P. N. A novel multiplex RT-PCR system detects human endogenous retrovirus-K in breast cancer. Archives of Virology, 2005, 150(1): 177–184.

[77]

Golan, M., Hizi, A., Resau, J. H., Yaal-Hahoshen, N., Reichman, H., Keydar, I., Tsarfaty, I. Human endogenous retrovirus (HERV-K) reverse transcriptase as a breast cancer prognostic marker. Neoplasia, 2008, 10(6): 521–IN2.

[78]

Depil, S., Roche, C., Dussart, P., Prin, L. Expression of a human endogenous retrovirus, HERV-K, in the blood cells of leukemia patients. Leukemia, 2002, 16(2): 254–259.

[79]

Muster, T., Waltenberger, A., Grassauer, A., Hirschl, S., Caucig, P., Romirer, I., Födinger, D., Seppele, H., Schanab, O., Magin-Lachmann, C. et al. An endogenous retrovirus derived from human melanoma cells. Cancer Research, 2003, 63(24): 8735–8741.

[80]

Büscher, K., Trefzer, U., Hofmann, M., Sterry, W., Kurth, R., Denner, J. Expression of human endogenous retrovirus K in melanomas and melanoma cell lines. Cancer Research, 2005, 65(10): 4172–4180.

[81]

Mao, J., Zhang, Q., Wang, Y., Zhuang, Y., Xu, L., Ma, X., Guan, D., Zhou, J., Liu, J., Wu, X. et al. TERT activates endogenous retroviruses to promote an immunosuppressive tumour microenvironment. EMBO Rep, 2022, 23(4): e52984.

[82]

Contreras-Galindo, R., Kaplan, M. H., Contreras-Galindo, A. C., Gonzalez-Hernandez, M. J., Ferlenghi, I., Giusti, F., Lorenzo, E., Gitlin, S. D., Dosik, M. H., Yamamura, Y. et al. Characterization of human endogenous retroviral elements in the blood of HIV-1-infected individuals. Journal of Virology, 2012, 86(1): 262–276.

[83]

Jones, R. B., Garrison, K. E., Mujib, S., Mihajlovic, V., Aidarus, N., Hunter, D. V., Martin, E., John, V. M., Zhan, W., Faruk, N. F. et al. HERV-K–specific T cells eliminate diverse HIV-1/2 and SIV primary isolates. Journal of Clinical Investigation, 2012, 122(12): 4473–4489.

[84]

Vincendeau, M., Göttesdorfer, I., Schreml, J. M., Wetie, A. G., Mayer, J., Greenwood, A. D., Helfer, M., Kramer, S., Seifarth, W., Hadian, K. et al. Modulation of human endogenous retrovirus (HERV) transcription during persistent and de novo HIV-1 infection. Retrovirology, 2015, 12: 27.

[85]

Srinivasachar Badarinarayan, S., Shcherbakova, I., Langer, S., Koepke, L., Preising, A., Hotter, D., Kirchhoff, F., Sparrer, K. M. J., Schotta, G., Sauter, D. HIV-1 infection activates endogenous retroviral promoters regulating antiviral gene expression. Nucleic Acids Research, 2020, 48(19): 10890–10908.

[86]

Terry, S. N., Manganaro, L., Cuesta-Dominguez, A., Brinzevich, D., Simon, V., Mulder, L. C. F. Expression of HERV-K108 envelope interferes with HIV-1 production. Virology, 2017, 509: 52–59.

[87]

Garrison, K. E., Jones, R. B., Meiklejohn, D. A., Anwar, N., Ndhlovu, L. C., Chapman, J. M., Erickson, A. L., Agrawal, A., Spotts, G., Hecht, F. M. et al. T cell responses to human endogenous retroviruses in HIV-1 infection. PLoS Pathog, 2007, 3(11): e165.

[88]

SenGupta, D., Tandon, R., Vieira, R. G. S., Ndhlovu, L. C., Lown-Hecht, R., Ormsby, C. E., Loh, L., Jones, R. B., Garrison, K. E., Martin, J. N. et al. Strong human endogenous retrovirus-specific T cell responses are associated with control of HIV-1 in chronic infection. Journal of Virology, 2011, 85(14): 6977–6985.

[89]

Sutkowski, N., Conrad, B., Thorley-Lawson, D. A., Huber, B. T. Epstein-barr virus transactivates the human endogenous retrovirus HERV-K18 that encodes a superantigen. Immunity, 2001, 15(4): 579–589.

[90]

Dechaumes, A., Bertin, A., Sane, F., Levet, S., Varghese, J., Charvet, B., Gmyr, V., Kerr-Conte, J., Pierquin, J., Arunkumar, G. et al. Coxsackievirus-B4 infection can induce the expression of human endogenous retrovirus W in primary cells. Microorganisms, 2020, 8(9): E1335.

[91]

Lima-Junior, D. S., Krishnamurthy, S. R., Bouladoux, N., Collins, N., Han, S. J., Chen, E. Y., Constantinides, M. G., Link, V. M., Lim, A. I., Enamorado, M. et al. Endogenous retroviruses promote homeostatic and inflammatory responses to the microbiota. Cell, 2021, 184(14): 3794–3811.e19.

[92]

Manghera, M., Ferguson-Parry, J., Lin, R. T., Douville, R. N. NF-κB and IRF1 induce endogenous retrovirus K expression via interferon-stimulated response elements in its 5’ long terminal repeat. Journal of Virology, 2016, 90(20): 9338–9349.

[93]

Kwon, D. N., Lee, Y. K., Greenhalgh, D. G., Cho, K. Lipopolysaccharide stress induces cell-type specific production of murine leukemia virus type-endogenous retroviral virions in primary lymphoid cells. Journal of General Virology, 2011, 92(2): 292–300.

[94]

Manghera, M., Ferguson, J., Douville, R. ERVK polyprotein processing and reverse transcriptase expression in human cell line models of neurological disease. Viruses, 2015, 7(1): 320–332.

[95]

Herrero, F., Mueller, F. S., Gruchot, J., Küry, P., Weber-Stadlbauer, U., Meyer, U. Susceptibility and resilience to maternal immune activation are associated with differential expression of endogenous retroviral elements. Brain, Behavior, and Immunity, 2023, 107: 201–214.

[96]

Magiorkinis, G., Belshaw, R., Katzourakis, A. ‘There and back again’: Revisiting the pathophysiological roles of human endogenous retroviruses in the post-genomic era. Philosophical Transactions of the Royal Society B: Biological Sciences, 2013, 368(1626): 20120504.

[97]

Herve, C. A., Lugli, E. B., Brand, A., et al. Autoantibodies to human endogenous retrovirus-K are frequently detected in health and disease and react with multiple epitopes. Clin Exp Immunol, 2002, 128(1): 75-82.

[98]

Ejtehadi, H. D. The potential role of human endogenous retrovirus K10 in the pathogenesis of rheumatoid arthritis: A preliminary study. Annals of the Rheumatic Diseases, 2006, 65(5): 612–616.

[99]

Freimanis, G., Hooley, P., Davari Ejtehadi, H., Ali, H. A., Veitch, A., Rylance, P. B., Alawi, A., Axford, J., Nevill, A., Murray, P. G. et al. A role for human endogenous retrovirus-K (HML-2) in rheumatoid arthritis: Investigating mechanisms of pathogenesis. Clinical and Experimental Immunology, 2010, 160(3): 340–347.

[100]

Wang, R. C., Li, H. D., Wu, J. F., Cai, Z. Y., Li, B. Z., Ni, H. X., Qiu, X. F., Chen, H., Liu, W., Yang, Z. H. et al. Gut stem cell necroptosis by genome instability triggers bowel inflammation. Nature, 2020, 580(7803): 386–390.

[101]

Liu, X. Q., Liu, Z. P., Wu, Z. M., Ren, J., Fan, Y. L., Sun, L., Cao, G., Niu, Y. Y., Zhang, B. H., Ji, Q. Z. et al. Resurrection of endogenous retroviruses during aging reinforces senescence. Cell, 2023, 186(2): 287–304.e26.

[102]

Paluvai, H., Di Giorgio, E., Brancolini, C. The histone code of senescence. Cells, 2020, 9(2): 466.

[103]

Di Giorgio, E., Paluvai, H., Dalla, E., Ranzino, L., Renzini, A., Moresi, V., Minisini, M., Picco, R., Brancolini, C. HDAC4 degradation during senescence unleashes an epigenetic program driven by AP-1/p300 at selected enhancers and super-enhancers. Genome Biology, 2021, 22(1): 129.

[104]

Colombo, A. R., Elias, H. K., Ramsingh, G. Senescence induction universally activates transposable element expression. Cell Cycle, 2018, 17(14): 1846–1857.

[105]

Guan, Y. T., Zhang, C., Lyu, G. L., Huang, X. K., Zhang, X. B., Zhuang, T. H., Jia, L. M., Zhang, L. J., Zhang, C., Li, C. et al. Senescence-activated enhancer landscape orchestrates the senescence-associated secretory phenotype in murine fibroblasts. Nucleic Acids Research, 2020, 48(19): 10909–10923.

[106]

De Cecco, M., Criscione, S. W., Peterson, A. L., Neretti, N., Sedivy, J. M., Kreiling, J. A. Transposable elements become active and mobile in the genomes of aging mammalian somatic tissues. Aging, 2013, 5(12): 867–883.

[107]

De Cecco, M., Criscione, S. W., Peckham, E. J., Hillenmeyer, S., Hamm, E. A., Manivannan, J., Peterson, A. L., Kreiling, J. A., Neretti, N., Sedivy, J. M. Genomes of replicatively senescent cells undergo global epigenetic changes leading to gene silencing and activation of transposable elements. Aging Cell, 2013, 12(2): 247–256.

[108]

Patterson, M. N., Scannapieco, A. E., Au, P. H., Dorsey, S., Royer, C. A., Maxwell, P. H. Preferential retrotransposition in aging yeast mother cells is correlated with increased genome instability. DNA Repair (Amst), 2015, 34: 18–27.

[109]

Hu, X., Chris-Anne, M., Colin, S., et al. The remarkable complexity of the brain microbiome in health and disease. BioRxiv, 2023: p. 2023.02. 06.527297.

[110]

Min, B., Jeon, K., Park, J. S., Kang, Y. K. Demethylation and derepression of genomic retroelements in the skeletal muscles of aged mice. Aging Cell, 2019, 18(6): e13042.

[111]

Zhang, H., Li, J. M., Yu, Y., Ren, J., Liu, Q., Bao, Z. S., Sun, S. H., Liu, X. Q., Ma, S., Liu, Z. P. et al. Nuclear lamina erosion-induced resurrection of endogenous retroviruses underlies neuronal aging. Cell Reports, 2023, 42(6): 112593.

[112]

Nevalainen, T., Autio, A., Mishra, B. H., Marttila, S., Jylhä, M., Hurme, M. Aging-associated patterns in the expression of human endogenous retroviruses. PLoS One, 2018, 13(12): e0207407.

[113]

Balestrieri, E., Pica, F., Matteucci, C., Zenobi, R., Sorrentino, R., Argaw-Denboba, A., Cipriani, C., Bucci, I., Sinibaldi-Vallebona, P. Transcriptional activity of human endogenous retroviruses in human peripheral blood mononuclear cells. Biomed Res Int, 2015, 2015: 164529.

[114]

Jintaridth, P., Mutirangura, A. Distinctive patterns of age-dependent hypomethylation in interspersed repetitive sequences. Physiological Genomics, 2010, 41(2): 194–200.

[115]

Yang, S., Liu, C., Jiang, M., Liu, X., Geng, L., Zhang, Y., Sun, S., Wang, K., Yin, J., Ma, S. et al. A single-nucleus transcriptomic atlas of primate liver aging uncovers the pro-senescence role of SREBP2 in hepatocytes. Protein Cell, 2024, 15(2): 98–120.

[116]

Jansz, N., Faulkner, G. J. Endogenous retroviruses in the origins and treatment of cancer. Genome Biology, 2021, 22(1): 147.

[117]

Lamprecht, B., Walter, K., Kreher, S., Kumar, R., Hummel, M., Lenze, D., Köchert, K., Bouhlel, M. A., Richter, J., Soler, E. et al. Derepression of an endogenous long terminal repeat activates the CSF1R proto-oncogene in human lymphoma. Nature Medicine, 2010, 16(5): 571–579.

[118]

Fasching, L., Kapopoulou, A., Sachdeva, R., Petri, R., Jönsson, M. E., Männe, C., Turelli, P., Jern, P., Cammas, F., Trono, D. et al. TRIM28 represses transcription of endogenous retroviruses in neural progenitor cells. Cell Reports, 2015, 10(1): 20–28.

[119]

Lin, H. Y., Cao, X. T. Nuclear innate sensors for nucleic acids in immunity and inflammation. Immunological Reviews, 2020, 297(1): 162–173.

[120]

Chen, R., Ishak, C. A., De Carvalho, D. D. Endogenous retroelements and the viral mimicry response in cancer therapy and cellular homeostasis. Cancer Discovery, 2021, 11(11): 2707–2725.

[121]

Cuellar, T. L., Herzner, A. M., Zhang, X. T., Goyal, Y., Watanabe, C., Friedman, B. A., Janakiraman, V., Durinck, S., Stinson, J., Arnott, D. et al. ---Silencing of retrotransposons by SETDB1 inhibits the interferon response in acute myeloid leukemia--. Journal of Cell Biology, 2017, 216(11): 3535–3549.

[122]

Wu, M., Li, M., Liu, W., Yan, M. B., Li, L., Ding, W. C., Nian, X. M., Dai, W. X., Sun, D., Zhu, Y. Q. et al. Nucleoporin Seh1 maintains Schwann cell homeostasis by regulating genome stability and necroptosis. Cell Reports, 2023, 42(7): 112802.

[123]

Rolland, A., Jouvin-Marche, E., Viret, C., et al. The envelope protein of a human endogenous retrovirus-W family activates innate immunity through CD14/TLR4 and promotes Th1-like responses. J Immunol, 2006, 176(12): 7636-7644.

[124]

Griffin, G. K., Wu, J., Iracheta-Vellve, A., et al. Epigenetic silencing by SETDB1 suppresses tumour intrinsic immunogenicity. Nature, 2021, 595(7866): 309-314.

[125]

Liu, S., Heumüller, S. E., Hossinger, A., Müller, S. A., Buravlova, O., Lichtenthaler, S. F., Denner, P., Vorberg, I. M. Reactivated endogenous retroviruses promote protein aggregate spreading. Nature Communications, 2023, 14: 5034.

[126]

Dai, Y. D., Dias, P., Margosiak, A., Marquardt, K., Bashratyan, R., Hu, W. Y., Haskins, K., Evans, L. H. Endogenous retrovirus Gag antigen and its gene variants are unique autoantigens expressed in the pancreatic islets of non-obese diabetic mice. Immunol Lett, 2020, 223: 62–70.

[127]

Chandrasekaran, S., Espeso-Gil, S., Loh, Y. E., et al. Neuron-specific chromosomal megadomain organization is adaptive to recent retrotransposon expansions. Nat Commun, 2021, 12(1): 7243.

[128]

Morozov, V. A., Dao Thi, V. L., and Denner, J. The transmembrane protein of the human endogenous retrovirus--K (HERV-K) modulates cytokine release and gene expression. PLoS One, 2013, 8(8): e70399.

[129]

Bhat, R. K., Rudnick, W., Antony, J. M., Maingat, F., Ellestad, K. K., Wheatley, B. M., Tönjes, R. R., Power, C. Human endogenous retrovirus-K(Ⅱ) envelope induction protects neurons during HIV/AIDS. PLoS One, 2014, 9(7): e97984.

[130]

Burn, A., Roy, F., Freeman, M., Coffin, J. M. Widespread expression of the ancient HERV-K (HML-2) provirus group in normal human tissues. PLoS Biology, 2022, 20(10): e3001826.

[131]

Wallace, A. D., Wendt, G. A., Barcellos, L. F., de Smith, A. J., Walsh, K. M., Metayer, C., Costello, J. F., Wiemels, J. L., Francis, S. S. To ERV is human: A phenotype-wide scan linking polymorphic human endogenous retrovirus-K insertions to complex phenotypes. Front Genet, 2018, 9: 298.

[132]

Scheltens, P. De Strooper, B. Kivipelto, M. et al., Alzheimer’s disease. The Lancet, 2021, 97(10284): 1577-1590.

[133]

Guo, C., Jeong, H. H., Hsieh, Y. C., et al. Tau Activates Transposable Elements in Alzheimer’s Disease. Cell Reports, 2018, 23(10): 2874-2880.

[134]

Ramirez, P., Zuniga, G., Sun, W., Beckmann, A., Ochoa, E., DeVos, S. L., Hyman, B., Chiu, G., Roy, E. R., Cao, W. et al. Pathogenic tau accelerates aging-associated activation of transposable elements in the mouse central nervous system. Prog Neurobiol, 2022, 208: 102181.

[135]

Dawson, T., Rentia, U., Sanford, J., Cruchaga, C., Kauwe, J. S. K., Crandall, K. A. Locus specific endogenous retroviral expression associated with Alzheimer’s disease. Frontiers in Aging Neuroscience, 2023, 15: 1186470.

[136]

Dembny, P., Newman, A. G., Singh, M., Hinz, M., Szczepek, M., Krüger, C., Adalbert, R., Dzaye, O., Trimbuch, T., Wallach, T. et al. Human endogenous retrovirus HERV-K(HML-2) RNA causes neurodegeneration through Toll-like receptors. JCI Insight, 2020, 5(7): 131093.

[137]

Ochoa, E., Ramirez, P., Gonzalez, E., De Mange, J., Ray, W. J., Bieniek, K. F., Frost, B. Pathogenic tau-induced transposable element-derived dsRNA drives neuroinflammation. Sci Adv, 2023, 9(1): eabq5423.

[138]

Scopa, C., Barnada, S. M., Cicardi, M. E., Singer, M., Trotti, D., Trizzino, M. JUN upregulation drives aberrant transposable element mobilization, associated innate immune response, and impaired neurogenesis in Alzheimer’s disease. Nature Communications, 2023, 14: 8021.

[139]

Sun, W. Y., Samimi, H., Gamez, M., Zare, H., Frost, B. Pathogenic tau-induced PiRNA depletion promotes neuronal death through transposable element dysregulation in neurodegenerative tauopathies. Nature Neuroscience, 2018, 21(8): 1038–1048.

[140]

Brown, R. H., and Al-Chalabi, A., Amyotrophic Lateral Sclerosis. N Engl J Med, 2017, 377(2): 162-172.

[141]

Arru, G., Mameli, G., Deiana, G. A., Rassu, A. L., Piredda, R., Sechi, E., Caggiu, E., Bo, M., Nako, E., Urso, D. et al. Humoral immunity response to human endogenous retroviruses K/W differentiates between amyotrophic lateral sclerosis and other neurological diseases. Eur J Neurol, 2018, 25(8): 1076–e84.

[142]

Douville, R., Liu, J. K., Rothstein, J., Nath, A. Identification of active loci of a human endogenous retrovirus in neurons of patients with amyotrophic lateral sclerosis. Annals of Neurology, 2011, 69(1): 141–151.

[143]

Garcia-Montojo, M., Simula, E. R., Fathi, S., McMahan, C., Ghosal, A., Berry, J. D., Cudkowicz, M., Elkahloun, A., Johnson, K., Norato, G. et al. Antibody response to HML-2 may be protective in amyotrophic lateral sclerosis. Annals of Neurology, 2022, 92(5): 782–792.

[144]

Garson, J. A., Usher, L., Al-Chalabi, A., Huggett, J., Day, E. F., McCormick, A. L. Quantitative analysis of human endogenous retrovirus-K transcripts in postmortem premotor cortex fails to confirm elevated expression of HERV-K RNA in amyotrophic lateral sclerosis. Acta Neuropathologica Communications, 2019, 7(1): 45.

[145]

Mayer, J., Harz, C., Sanchez, L., Pereira, G. C., Maldener, E., Heras, S. R., Ostrow, L. W., Ravits, J., Batra, R., Meese, E. et al. Transcriptional profiling of HERV-K(HML-2) in amyotrophic lateral sclerosis and potential implications for expression of HML-2 proteins. Molecular Neurodegeneration, 2018, 13(1): 39.

[146]

Romano, G., Klima, R., Feiguin, F. TDP-43 prevents retrotransposon activation in the Drosophila motor system through regulation of Dicer-2 activity. BMC Biology, 2020, 18(1): 82.

[147]

Douville, R. N., Nath, A. Human endogenous retrovirus-K and TDP-43 expression bridges ALS and HIV neuropathology. Frontiers in Microbiology, 2017, 8: 1986.

[148]

Simula, E. R., Arru, G., Zarbo, I. R., Solla, P., Sechi, L. A. TDP-43 and HERV-K envelope-specific immunogenic epitopes are recognized in ALS patients. Viruses, 2021, 13(11): 2301.

[149]

Chang, Y. H., Dubnau, J. The Gypsy endogenous retrovirus drives non-cell-autonomous propagation in a drosophila TDP-43 model of neurodegeneration. Current Biology, 2019, 29(19): 3135–3152.e4.

[150]

Krug, L., Chatterjee, N., Borges-Monroy, R., Hearn, S., Liao, W. W., Morrill, K., Prazak, L., Rozhkov, N., Theodorou, D., Hammell, M. et al. Retrotransposon activation contributes to neurodegeneration in a Drosophila TDP-43 model of ALS. PLoS Genet, 2017, 13(3): e1006635.

[151]

Chang, Y. H., Dubnau, J. Endogenous retroviruses and TDP-43 proteinopathy form a sustaining feedback driving intercellular spread of Drosophila neurodegeneration. Nature Communications, 2023, 14: 966.

[152]

Marcus, R. What is multiple sclerosis? JAMA, 2022, 328(20): 2078.

[153]

Gruchot, J., Lewen, I., Dietrich, M., Transgenic expression of the HERV-W envelope protein leads to polarized glial cell populations and a neurodegenerative environment. Proc Natl Acad Sci U, 2023, 120(38): e2308187120.

[154]

Kremer, D., Gruchot, J., Weyers, V., Oldemeier, L., Göttle, P., Healy, L., Ho Jang, J., Kang T Xu, Y., Volsko, C., Dutta, R. et al. pHERV-W envelope protein fuels microglial cell-dependent damage of myelinated axons in multiple sclerosis. Proceedings of the National Academy of Sciences of the United States of America, 2019, 116(30): 15216–15225.

[155]

Mameli, G., Astone, V., Arru, G., Marconi, S., Lovato, L., Serra, C., Sotgiu, S., Bonetti, B., Dolei, A. Brains and peripheral blood mononuclear cells of multiple sclerosis (MS) patients hyperexpress MS-associated retrovirus/HERV-W endogenous retrovirus, but not Human herpesvirus 6. The Journal of General Virology, 2007, 88(Pt 1): 264–274.

[156]

Garcia-Montojo, M., Rodriguez-Martin, E., Ramos-Mozo, P., Ortega-Madueño, I., Dominguez-Mozo, M. I., Arias-Leal, A., García-Martínez, M. Á., Casanova, I., Galan, V., Arroyo, R. et al. Syncytin-1/HERV-W envelope is an early activation marker of leukocytes and is upregulated in multiple sclerosis patients. European Journal of Immunology, 2020, 50(5): 685–694.

[157]

Mameli, G., Cossu, D., Cocco, E., Frau, J., Marrosu, M. G., Niegowska, M., Sechi, L. A. Epitopes of HERV-Wenv induce antigen-specific humoral immunity in multiple sclerosis patients. Journal of Neuroimmunology, 2015, 280: 66–68.

[158]

Alvarez-Lafuente, R., García-Montojo, M., De Las Heras, V., Domínguez-Mozo, M. I., Bartolome, M., Benito-Martin, M. S., Arroyo, R. Herpesviruses and human endogenous retroviral sequences in the cerebrospinal fluid of multiple sclerosis patients. Multiple Sclerosis, 2008, 14(5): 595–601.

[159]

Charvet, B., Reynaud, J. M., Gourru-Lesimple, G., Perron, H., Marche, P. N., Horvat, B. Induction of proinflammatory multiple sclerosis-associated retrovirus envelope protein by human herpesvirus-6A and CD46 receptor engagement. Front Immunol, 2018, 9: 2803.

[160]

Meier, U. C., Cipian, R. C., Karimi, A., Ramasamy, R., Middeldorp, J. M. Cumulative roles for epstein-barr virus, human endogenous retroviruses, and human herpes virus-6 in driving an inflammatory cascade underlying MS pathogenesis. Frontiers in Immunology, 2021, 12: 757302.

[161]

Pérez-Pérez, S., Domínguez-Mozo, M. I., García-Martínez MÁ, García-Frontini, M. C., Villarrubia, N., Costa-Frossard, L., Villar, L. M., Arroyo, R., Álvarez-Lafuente, R. Anti-human herpesvirus 6 A/B antibodies titers correlate with multiple sclerosis-associated retrovirus envelope expression. Front Immunol, 2021, 12: 798003.

[162]

Canli, T. A model of human endogenous retrovirus (HERV) activation in mental health and illness. Medical Hypotheses, 2019, 133: 109404.

[163]

Karlsson, H., Bachmann, S., Schröder, J., McArthur, J., Torrey, E. F., Yolken, R. H. Retroviral RNA identified in the cerebrospinal fluids and brains of individuals with schizophrenia. Proceedings of the National Academy of Sciences of the United States of America, 2001, 98(8): 4634–4639.

[164]

Karlsson, H., Schröder, J., Bachmann, S., Bottmer, C., Yolken, R. H. HERV-W-related RNA detected in plasma from individuals with recent-onset schizophrenia or schizoaffective disorder. Molecular Psychiatry, 2004, 9(1): 12–13.

[165]

Huang, W., Li, S., Hu, Y., Yu, H., Luo, F., Zhang, Q., Zhu, F. Implication of the env gene of the human endogenous retrovirus W family in the expression of BDNF and DRD3 and development of recent-onset schizophrenia. Schizophrenia Bulletin, 2011, 37(5): 988–1000.

[166]

Li, F., Sabunciyan, S., Yolken, R. H., Lee, D., Kim, S., Karlsson, H. Transcription of human endogenous retroviruses in human brain by RNA-seq analysis. PLoS One, 2019, 14(1): e0207353.

[167]

Perron, H., Mekaoui, L., Bernard, C., Veas, F., Stefas, I., Leboyer, M. Endogenous retrovirus type W GAG and envelope protein antigenemia in serum of schizophrenic patients. Biological Psychiatry, 2008, 64(12): 1019–1023.

[168]

Weis, S., Llenos, I. C., Sabunciyan, S., Dulay, J. R., Isler, L., Yolken, R., Perron, H. Reduced expression of human endogenous retrovirus (HERV)-W GAG protein in the cingulate gyrus and hippocampus in schizophrenia, bipolar disorder, and depression. Journal of Neural Transmission, 2007, 114(5): 645–655.

[169]

Tamouza, R., Meyer, U., Foiselle, M., Richard, J. R., Wu, C. L., Boukouaci, W., Le Corvoisier, P., Barrau, C., Lucas, A., Perron, H. et al. Identification of inflammatory subgroups of schizophrenia and bipolar disorder patients with HERV-W ENV antigenemia by unsupervised cluster analysis. Translational Psychiatry, 2021, 11(1): 377.

[170]

Li, X., Wu, X., Li, W., Yan, Q., Zhou, P., Xia, Y., Yao, W., Zhu, F. HERV-W ENV induces innate immune activation and neuronal apoptosis via linc01930/cGAS axis in recent-onset schizophrenia. Int J Mol Sci, 2023, 24(3): 3000.

[171]

Wu, X. L., Yan, Q. J., Liu, L. Z., Xue, X., Yao, W., Li, X. H., Li, W. S., Ding, S., Xia, Y. R., Zhang, D. Y. et al. Domesticated HERV-W env contributes to the activation of the small conductance Ca2+-activated K+ type 2 channels via decreased 5-HT4 receptor in recent-onset schizophrenia. Virologica Sinica, 2023, 38(1): 9–22.

[172]

Li, S., Liu, Z. C., Yin, S. J., Chen, Y. T., Yu, H. L., Zeng, J., Zhang, Q., Zhu, F. Human endogenous retrovirus W family envelope gene activates the small conductance Ca2+-activated K+ channel in human neuroblastoma cells through CREB. Neuroscience, 2013, 247: 164–174.

[173]

Chen, Y. T., Yan, Q. J., Zhou, P., Li, S., Zhu, F. HERV-W env regulates calcium influx via activating TRPC3 channel together with depressing DISC1 in human neuroblastoma cells. Journal of NeuroVirology, 2019, 25(1): 101–113.

[174]

Grube, S., Gerchen, M. F., Adamcio, B., Pardo, L. A., Martin, S., Malzahn, D., Papiol, S., Begemann, M., Ribbe, K., Friedrichs, H. et al. A CAG repeat polymorphism of KCNN3 predicts SK3 channel function and cognitive performance in schizophrenia. EMBO Molecular Medicine, 2011, 3(6): 309–319.

[175]

El-Hassar, L., Simen, A. A., Duque, A., Patel, K. D., Kaczmarek, L. K., Arnsten, A. F. T., Yeckel, M. F. Disrupted in schizophrenia 1 modulates medial prefrontal cortex pyramidal neuron activity through cAMP regulation of transient receptor potential C and small-conductance K+ channels. Biological Psychiatry, 2014, 76(6): 476–485.

[176]

Yao, W., Zhou, P., Yan, Q. J., Wu, X. L., Xia, Y. R., Li, W. S., Li, X. H., Zhu, F. ERVWE1 reduces hippocampal neuron density and impairs dendritic spine morphology through inhibiting Wnt/JNK non-canonical pathway via miR-141-3p in schizophrenia. Viruses, 2023, 15(1): 168.

[177]

Zhang, D., Wu, X., Xue, X., Li, W., Zhou, P., Lv, Z., Zhao, K., Zhu, F. Ancient dormant virus remnant ERVW-1 drives ferroptosis via degradation of GPX4 and SLC3A2 in schizophrenia. Virol Sin, 2024, 39(1): 31–43.

[178]

Frank, O., Giehl, M., Zheng, C., Hehlmann, R., Leib-Mösch, C., Seifarth, W. Human endogenous retrovirus expression profiles in samples from brains of patients with schizophrenia and bipolar disorders. Journal of Virology, 2005, 79(17): 10890–10901.

[179]

Mak, M., Samochowiec, J., Frydecka, D., Pełka-Wysiecka, J., Szmida, E., Karpiński, P., Sąsiadek, M. M., Piotrowski, P., Samochowiec, A., Misiak, B. First-episode schizophrenia is associated with a reduction of HERV-K methylation in peripheral blood. Psychiatry Res, 2019, 271: 459–463.

[180]

Huang, W. J., Liu, Z. C., Wei, W., Wang, G. H., Wu, J. G., Zhu, F. Human endogenous retroviral pol RNA and protein detected and identified in the blood of individuals with schizophrenia. Schizophrenia Research, 2006, 83(2–3): 193–199.

[181]

Suntsova, M., Gogvadze, E. V., Salozhin, S., Gaifullin, N., Eroshkin, F., Dmitriev, S. E., Martynova, N., Kulikov, K., Malakhova, G., Tukhbatova, G. et al. Human-specific endogenous retroviral insert serves as an enhancer for the schizophrenia-linked gene PRODH. Proc Natl Acad Sci USA, 2013, 110(48): 19472–19477.

[182]

Willis, A., Bender, H. U., Steel, G., Valle, D. PRODH variants and risk for schizophrenia. Amino Acids, 2008, 35(4): 673–679.

[183]

Nakamura, A., Okazaki, Y., Sugimoto, J., Oda, T., Jinno, Y. Human endogenous retroviruses with transcriptional potential in the brain. Journal of Human Genetics, 2003, 48(11): 575–581.

[184]

Otowa, T., Tochigi, M., Rogers, M., Umekage, T., Kato, N., Sasaki, T. Insertional polymorphism of endogenous retrovirus HERV-K115 in schizophrenia. Neuroscience Letters, 2006, 408(3): 226–229.

[185]

Diem, O., Schäffner, M., Seifarth, W., Leib-Mösch, C. Influence of antipsychotic drugs on human endogenous retrovirus (HERV) transcription in brain cells. PLoS One, 2012, 7(1): e30054.

[186]

Perron, H., Hamdani, N., Faucard, R., Lajnef, M., Jamain, S., Daban-Huard, C., Sarrazin, S., LeGuen, E., Houenou, J., Delavest, M. et al. Molecular characteristics of Human Endogenous Retrovirus type-W in schizophrenia and bipolar disorder. Transl Psychiatry, 2012, 2: e201.

[187]

Otręba, M., Kośmider, L., Rzepecka-Stojko, A. Antiviral activity of chlorpromazine, fluphenazine, perphenazine, prochlorperazine, and thioridazine towards RNA-viruses. A review. European Journal of Pharmacology, 2020, 887: 173553.

[188]

Andreu, S., Ripa, I., Bello-Morales, R., López-Guerrero, J. A. Valproic acid and its amidic derivatives as new antivirals against alphaherpesviruses. Viruses, 2020, 12(12): E1356.

[189]

Esnault, C., Millet, J., Schwartz, O., Heidmann, T. Dual inhibitory effects of APOBEC family proteins on retrotransposition of mammalian endogenous retroviruses. Nucleic Acids Research, 2006, 34(5): 1522–1531.

[190]

Esnault, C., Priet, S., Ribet, D., Heidmann, O., Heidmann, T. Restriction by APOBEC3 proteins of endogenous retroviruses with an extracellular life cycle: Ex vivo effects and in vivo “traces” on the murine IAPE and human HERV-K elements. Retrovirology, 2008, 5: 75.

[191]

Balestrieri, E., Pitzianti, M., Matteucci, C., D’Agati, E., Sorrentino, R., Baratta, A., Caterina, R., Zenobi, R., Curatolo, P., Garaci, E. et al. Human endogenous retroviruses and ADHD. The World Journal of Biological Psychiatry, 2014, 15(6): 499–504.

[192]

D’Agati, E., Pitzianti, M., Balestrieri, E., Matteucci, C., Sinibaldi Vallebona, P., Pasini, A. First evidence of HERV-H transcriptional activity reduction after methylphenidate treatment in a young boy with ADHD. The New Microbiologica, 2016, 39(3): 237–239.

[193]

Cipriani, C., Pitzianti, M. B., Matteucci, C., D’Agati, E., Miele, M. T., Rapaccini, V., Grelli, S., Curatolo, P., Sinibaldi-Vallebona, P., Pasini, A. et al. The decrease in human endogenous retrovirus-H activity runs in parallel with improvement in ADHD symptoms in patients undergoing methylphenidate therapy. Int J Mol Sci, 2018, 19(11): E3286.

[194]

Balestrieri, E., Cipriani, C., Matteucci, C., et al. Transcriptional activity of human endogenous retrovirus in Albanian children with autism spectrum disorders. New Microbiol, 2016, 39(3): 228-231.

[195]

Balestrieri, E., Arpino, C., Matteucci, C., Sorrentino, R., Pica, F., Alessandrelli, R., Coniglio, A., Curatolo, P., Rezza, G., Macciardi, F. et al. HERVs expression in Autism Spectrum Disorders. PLoS One, 2012, 7(11): e48831.

[196]

Carta, A., Manca, M. A., Scoppola, C., Simula, E. R., Noli, M., Ruberto, S., Conti, M., Zarbo, I. R., Antonucci, R., Sechi, L. A. et al. Antihuman endogenous retrovirus immune response and adaptive dysfunction in autism. Biomedicines, 2022, 10(6): 1365.

[197]

Cipriani, C., Ricceri, L., Matteucci, C., De Felice, A., Tartaglione, A. M., Argaw-Denboba, A., Pica, F., Grelli, S., Calamandrei, G., Sinibaldi Vallebona, P. et al. High expression of Endogenous Retroviruses from intrauterine life to adulthood in two mouse models of Autism Spectrum Disorders. Scientific Reports, 2018, 8: 629.

[198]

Tartaglione, A. M., Cipriani, C., Chiarotti, F., Perrone, B., Balestrieri, E., Matteucci, C., Sinibaldi-Vallebona, P., Calamandrei, G., Ricceri, L. Early behavioral alterations and increased expression of endogenous retroviruses are inherited across generations in mice prenatally exposed to valproic acid. Molecular Neurobiology, 2019, 56(5): 3736–3750.

[199]

Balestrieri, E., Cipriani, C., Matteucci, C., Benvenuto, A., Coniglio, A., Argaw-Denboba, A., Toschi, N., Bucci, I., Miele, M. T., Grelli, S. et al. Children with autism spectrum disorder and their mothers share abnormal expression of selected endogenous retroviruses families and cytokines. Front Immunol, 2019, 10: 2244.

[200]

Cipriani, C., Giudice, M., Petrone, V., Fanelli, M., Minutolo, A., Miele, M. T., Toschi, N., Maracchioni, C., Siracusano, M., Benvenuto, A. et al. Modulation of human endogenous retroviruses and cytokines expression in peripheral blood mononuclear cells from autistic children and their parents. Retrovirology, 2022, 19(1): 26.

[201]

Faucard, R., Madeira, A., Gehin, N., Authier, F. J., Panaite, P. A., Lesage, C., Burgelin, I., Bertel, M., Bernard, C., Curtin, F. et al. Human endogenous retrovirus and neuroinflammation in chronic inflammatory demyelinating polyradiculoneuropathy. EBioMedicine, 2016, 6: 190–198.

[202]

Arru, G., Sechi, E., Mariotto, S., Zarbo, I. R., Ferrari, S., Gajofatto, A., Monaco, S., Deiana, G. A., Bo, M., Sechi, L. A. et al. Antibody response against HERV-W in patients with MOG-IgG associated disorders, multiple sclerosis and NMOSD. J Neuroimmunol, 2020, 338: 577110.

[203]

Jeong, B. H., Lee, Y. J., Carp, R. I., Kim, Y. S. The prevalence of human endogenous retroviruses in cerebrospinal fluids from patients with sporadic Creutzfeldt–Jakob disease. Journal of Clinical Virology, 2010, 47(2): 136–142.

[204]

Rex, C., Nadeau, M. J., Douville, R., Schellenberg, K. Expression of human endogenous retrovirus-K in spinal and bulbar muscular atrophy. Frontiers in Neurology, 2019, 10: 968.

[205]

De Meirleir, K. L., Khaiboullina, S. F., Frémont, M., Hulstaert, J., Rizvanov, A. A., Palotás, A., Lombardi, V. C. Plasmacytoid dendritic cells in the duodenum of individuals diagnosed with myalgic encephalomyelitis are uniquely immunoreactive to antibodies to human endogenous retroviral proteins. In Vivo, 2013, 27(2): 177–187.

[206]

Singh, S., Stafford, P., Schlauch, K. A., Tillett, R. R., Gollery, M., Johnston, S. A., Khaiboullina, S. F., De Meirleir, K. L., Rawat, S., Mijatovic, T. et al. Humoral immunity profiling of subjects with myalgic encephalomyelitis using a random peptide microarray differentiates cases from controls with high specificity and sensitivity. Molecular Neurobiology, 2018, 55(1): 633–641.

[207]

Rodrigues, L. S., da Silva Nali, L. H., Leal, C. O. D., Sabino, E. C., Lacerda, E. M., Kingdon, C. C., Nacul, L., Romano, C. M. HERV-K and HERV-W transcriptional activity in myalgic encephalomyelitis/chronic fatigue syndrome. Autoimmunity Highlights, 2019, 10(1): 12.

[208]

Volcy, K. and Fraser, N.W. DNA damage promotes herpes simplex virus-1 protein expression in a neuroblastoma cell line. J Neurovirol, 2013, 19(1): 57–64.

Stress and Brain
Pages 1-30
Cite this article:
Li J, Liao L, Liu X, et al. Decoding Neurological Mysteries: The Potential Impact of Endogenous Retroviruses on Brain Health. Stress and Brain, 2024, 4(1): 1-30. https://doi.org/10.26599/SAB.2023.9060005

576

Views

128

Downloads

0

Crossref

Altmetrics

Received: 31 December 2023
Accepted: 13 March 2024
Published: 05 March 2024
© The Author(s) 2023

Creative Commons Non Commercial CC BY-NC: This article is distributed under the terms of the Creative Commons Attributtion-NonCommercial 4.0 License (http://www.creativecommons.org/licenses/by-nc/4.0/) which permits non-commercial use, reproduction and distribution of the work without further permission.

Return