Lipofuscin

From Longevity Wiki

Lipofuscin is a yellow-brown autofluorescent pigment also known as "aging pigment" due to its age-related progressive accumulation. It is a waste product consisting of insoluble granules made of lipids and proteins that accumulate in the lysosomes of cells. Over time, the lysosome becomes clogged and is not able to continue working properly.[1][2][3]

Lipofuscin spots on the upper surface of the hands.

Lipofuscin is proposed as a senescent marker in long-lived, non-dividing cells of different tissues across species. However, it is not 100% specific to senescent cells, as it can accumulate in conditions such as age-related macular degeneration (AMD).[4] Lipofuscin accumulation in the lysosomes cause dysregulation and reduction of its autophagic capacity, generating ROS (reactive oxygen species), elevating lysosomal pH and leading to lysosome leakage.[5] Lipofuscin consists of a non-degradable intralysosomal substance, which forms mainly due to iron-catalyzed oxidation/polymerization of misfolded proteins (~30–70%) and lipid (~20–50%) residues together with metals such as iron, copper, zinc, manganese, and calcium, in a concentration up to 2%.[6][7][8][9][10]

Accumulation of lipofuscin or "aging pigment" is part of normal aging, and should be distinguished from accumulation of ceroid - autofluorescent storage material associated with disease and usually produced under various pathological conditions not necessarily related to aging.[11] Ceroid has been suggested to jeopardize cell performance and viability by inducing membrane fragility, mitochondrial dysfunction, DNA damage, and oxidative stress-induced apoptosis.[12]

Detection of lipofuscin

During the process of aging, lipofuscin accumulates in a nearly linear way in postmitotic senescencent cells (cardiomyocytes, retinal epithelial pigment cells, hepatocytes, neurons and keratinocytes),[13] and has been proposed as a detectable "marker" to estimate aging. This approach is particularly used to determine the age of crabs and other crustaceans either by labor histological fluorescent-microscopy examinations[14] or simply by extractable lipofuscin solvent fluorescence measurements.[15]

Detection of lipofuscin content can be used as a biomarker of old lysosome accumulation, either by its typical autofluorescence properties and fluorescence-based methods, or by selective staining with Sudan black B, which stained lipofuscin granules, allowing for detection in cells, tissues and body fluids.[16][17][18][19]

Relation to aging diseases

Because lipofuscin is a covalently cross-linked aggregate, it cannot be removed from the cytosol by the ubiquitin-proteasome system.[20][21] Furthermore, lipofuscin could belong to Advanced glycation end products (AGEs) deposits.[22]

Isolated lipofuscin aggregates, as shown in vitro, were readily incorporated by fibroblasts and caused cell death at low concentrations (LC50 = 5.0 µg/mL) via a pyroptosis-like pathway. Lipofuscin boosted mitochondrial ROS production and caused lysosomal dysfunction by lysosomal membrane permeabilization leading to reduced lysosome quantity and impaired cathepsin D activity.[23]

Lipofuscin granules accumulation can lead to pathology and accelerate the aging process.[24] The rate of lipofuscin formation has been shown to be negatively correlated with the life expectancy of postmitotic cells, i.e., the higher the rate, the shorter the lifespan of the cell due to decrease of cellular adaptability.[25] Therefore, progressive deposition of lipofuscin might promote the development of age-related pathologies, including macular degeneration, heart failure, and neuro-degenerative diseases.

Dry AMD

One of the diseases associated with the accumulation of lipofuscin is dry age-related macular degeneration (dry AMD) – a disease often diagnosed in people over 70 years of age and a leading cause of rapid vision loss. Dry AMD is a slow-progressing disease in which yellow drusen containing lipofuscin are deposited between the retinal pigment epithelial (RPE) cell layer and Bruch’s membrane.[26] A phototoxic components of lipofuscin such as A2E (Bis-retinoid N-retinyl-N-retinylidene ethanolamine) that induces inflammation and apoptosis in RPE cells,[27] are accumulated with age and mediate damage under blue light exposure.[28][29] It has been reported that iron levels increase in RPE during ageing and this intracellular iron can interact with bisretinoid lipofuscin in RPE to promote cell damage.[30] Therefore, to alleviate the deteriorating effects of lipofuscin on age-related macular degeneration, iron chelation, either independently or in combination with bisretinoid inhibitors could potentially serve as AMD treatments.[31] To protect human RPE cells from oxidative damage, caused by reactive oxygen species generated by the photo-excited lipofuscin, also is able L‐Citrulline, a naturally occurring amino acid with known antioxidant properties[32] and the main active component of Spirulina maxima P-phycocyanin - pigment with anti-inflammatory and antioxidant activities.[33]

The drug Lysoclear is an enzyme developed to enter RPE cells and break down lipofuscin deposits in the lysosomes, a therapeutic approach that proposes to reverse dry age-related macular degeneration and Stargardt's macular degeneration.[34] Phase 1 clinical trials started in 2018.

Zinc deficiency

Zinc-deficient animals showed a greater number of lipofuscin granules.[35] The relationship between zinc deficiency and enhanced lipofuscin accumulation suggest that zinc deficiency may result in the accumulation of substrates for autophagy whereas low zinc does not stimulate autophagy.[36] Autophagy is also inhibited when A2E-treated RPE cells are exposed to blue light.[29] Currently, the only intervention available for the treatment of dry AMD is Age-Related Eye Disease Supplement (AREDS), an oral supplement containing vitamin C, vitamin E, lutein/zeaxanthin, and zinc. It was shown that AREDS can reduce the risk of advanced AMD by about 25% over a 5-year period in patients with intermediate AMD.[37]

Lipofuscin accumulation in aging heart

Lipofuscin granules are found abundantly in myocardial cells.[38][39][40] The myocardial tissues of mice have the ability to eliminate the lipofuscin produced in the cardiomyocytes into the myocardial blood circulation. It is mainly carried out of cardiomyocytes into the myocardial interstitium in the form of small lipofuscin granules, using capsule-like protrusions that are formed on the sarcolemma.[41]

Role of lipofuscin in age-related neurodegeneration

Lipofuscin aggregation represents a risk factor for neurodegeneration.[42]

Progranulin neurons that normally have high levels of progranulin expression are more susceptible to age-related pathology, such as neuronal lipofuscinosis, in GRN−/− mice.[43]

Lipofuscin granules (brown) accumulation in aged (60y) human epidermis according to Rübe et al., & Scherthan (2021).[44]

Lipofuscin-accumulating in skin cells

Lipofuscin is an endogenous photosensitizer that efficiently absorbs ultraviolet radiation and visible light, forming electronic excited states that transfer energy to surrounding molecules. It is assumed, that photosensitized lipofuscin is cytotoxic because of its ability to incorporate redox-active transition metals (Fe+2), resulting in a redox-active surface, able to catalyze the Fenton reaction. Reactive oxygen and nitrogen species (ROS/RNS) generated by photosensitization of lipofuscin leads to DNA damage and strand breaks.[45][46] It was observed that application of vitamin E may reduce the level of lipofuscin in skin biopsies as well as lighten the skin (but not in very old ones).[47][13] Light-induced skin damage can be protected by regulating the ROS-ER stress-autophagy-apoptosis axis with hydrogen sulfide (H2S).[48]

Inhibitors of lipofuscin accumulation

It was suggested that formation of A2E and other toxic lipofuscin bisretinoids, such as A2-DHP-PE (A2-dihydropyridinephosphatidyl-ethanolamine) and atRALdi-PE (all-trans-retinal dimer phosphatidylethanolamine), occurs in the retina in a non-enzymatic manner and can be considered a by-product of a properly functioning visual cycle.[49] The formation of A2E was completely inhibited in total darkness, so, humans with retinal or macular degeneration may slow progression of their disease by limiting exposure to light.[50]

Meclofenoxate

Meclofenoxate is a cholinergic nootropic also known as centrophenoxine. It is an ester of dimethylethanolamine (DMAE) and 4-chlorophenoxyacetic acid (pCPA). Meclofenoxate, as well as DMAE, have been found to increase the dissolution and removal of lipofuscin[51][52][53] leading to lifespan extension of mice.[54][55] As a result of meclophenoxate treatment, a gradual decrease in the myocardial volume occupied by the pigment was noted. After 4-6 weeks of treatment, the pigment bodies were found lodged into the capillary endothelium and the lumen, facilitating the removal of the pigment via blood stream.[56][57]

Remofuscin

Remofuscin, a small molecule belonging to the tetrahydropyridoether class of compounds is able to remove lipofuscin from the RPE by accumulation specifically in RPE pigments and thus stimulation of the exocytosis.[58] Remofuscin formerly known as Soraprazan a potent and reversible selective inhibitor of gastric H,K-ATPase may be a promising drug candidate to manage neurodegenerative diseases related to lipofuscin accumulation.[59] Remofuscin reverses lipofuscin accumulation in aged primary human RPE cells and is non-cytotoxic in aged SD mouse RPE cells in vitro.[60] Mechanism causing lipofuscinolysis may involve the reactive oxygen species generated via the presence of remofuscin. Remofuscin binds to lipofuscin and is a superoxide generator when illuminated with light. Superoxide might help to degrade the polymeric lipofuscin into smaller units which then are transported out of the lysosomes by exocytosis.[60][61] Remofuscin reduces existing levels of lipofuscin in the RPE instead of merely slowing down accumulation of further toxic Vitamin A aggregates.[62][61]

Aging biomarkers were improved in remofuscin-treated Caenorhabditis elegans worms, resulting in a significant (p <0.05) increase in their lifespan.[61] The expression levels of genes related to lysosomes, a nuclear hormone receptor, fatty acid beta-oxidation, and xenobiotic detoxification were increased in remofuscin-treated worms. Moreover, remofuscin failed to extend the lives of C. elegans with loss-of-function mutations of genes related to lysosomes and xenobiotic detoxification, suggesting that these genes are associated with lifespan extension in remofuscin-treated C. elegans.[61]

NMDA receptor antagonists

N-methyl-D-aspartate (NMDA) receptors signaling is a novel mechanism for scavenging N-retinylidene-N-retinylethanolamine (A2E), a component of ocular lipofuscin, in human RPE cells. NMDA receptor antagonists, such as Ro 25-6981, CP-101,606 and AZD6765, degrade lipofuscin via autophagy in human RPE cells.[63] Ro 25-6981 has not yet been approved for clinical use. Among the clinically approved NMDA antagonists, memantine and ifenprodil have been proposed as drug repositioning to remove N-retinylidene-N-retinylethanolamine (A2E), an intracellular lipofuscin component.[64]

ATM inhibition

The increase in lipid peroxidation during oxidative stress increases the content of intra-lysosomal lipofuscins in fibroblasts during senescence.[65] Senescence amelioration in normal aging cells is mediated by the recovered mitochondrial function upon inhibition of a key mediator of DNA damage signaling and repair - Ataxia telangiectasia mutated (ATM).[66][67][68]

ATM inhibitors KU-60019, CP-466722 or antioxidant N-acetyl-cysteine (NAC) significantly reduced lipofuscin accumulation.[69]

See also

  • Ilie, O. D., Ciobica, A., Riga, S., Dhunna, N., McKenna, J., Mavroudis, I., ... & Riga, D. (2020). Mini-review on lipofuscin and aging: focusing on the molecular interface, the biological recycling mechanism, oxidative stress, and the gut-brain axis functionality. Medicina, 56(11), 626. PMID: 33228124 PMCID: PMC7699382 DOI: 10.3390/medicina56110626
  • Nasiri, L., Vaez-Mahdavi, M. R., Hassanpour, H., Ghazanfari, T., Ardestani, S. K., Askari, N., ... & Rahimlou, B. (2023). Increased serum lipofuscin associated with leukocyte telomere shortening in veterans: a possible role for sulfur mustard exposure in delayed-onset accelerated cellular senescence. International Immunopharmacology, 114, 109549. https://doi.org/10.1016/j.intimp.2022.109549 Chronic oxidative stress and continuous inflammatory stimulation in veterans, due to mustard gas poisoning once in 1987, led to cells senescence with increased lipofuscin, and telomere shortening.
  • Nociari, M. M., Lehmann, G. L., Perez Bay, A. E., Radu, R. A., Jiang, Z., Goicochea, S., ... & Rodriguez-Boulan, E. (2014). Beta cyclodextrins bind, stabilize, and remove lipofuscin bisretinoids from retinal pigment epithelium. Proceedings of the National Academy of Sciences, 111(14), E1402-E1408. PMID: 24706818 PMCID: PMC3986126 DOI: 10.1073/pnas.1400530111

References

  1. Strehler, B. L., Mark, D. D., Mildvan, A. S., & Gee, M. V. (1959). Rate and magnitude of age pigment accumulation in the human myocardium. Journal of gerontology, 14(4), 430-439. DOI: 10.1093/geronj/14.4.430
  2. Reichel, W. (1968). Lipofuscin pigment accumulation and distribution in five rat organs as a function of age. Journal of gerontology, 23(2), 145-153. DOI: 10.1093/geronj/23.2.145
  3. Mann, D. M. A., Yates, P. O., & Stamp, J. E. (1978). The relationship between lipofuscin pigment and ageing in the human nervous system. Journal of the Neurological Sciences, 37(1-2), 83-93. DOI: 10.1016/0022-510x(78)90229-0
  4. Georgakopoulou EA, Tsimaratou K, Evangelou K, Fernandez Marcos PJ, Zoumpourlis V, Trougakos IP, Kletsas D, Bartek J, Serrano M, Gorgoulis VG. Specific lipofuscin staining as a novel biomarker to detect replicative and stress-induced senescence. A method applicable in cryo-preserved and archival tissues. Aging (Albany NY). 2013 Jan;5(1):37-50. doi: 10.18632/aging.100527.
  5. Dutta, R. K., Lee, J. N., Maharjan, Y., Park, C., Choe, S. K., Ho, Y. S., ... & Park, R. (2022). Catalase-deficient mice induce aging faster through lysosomal dysfunction. Cell Communication and Signaling, 20(1), 1-22. PMID:36474295 PMC9724376 DOI: 10.1186/s12964-022-00969-2
  6. Höhn, A., Jung, T., Grimm, S., & Grune, T. (2010). Lipofuscin-bound iron is a major intracellular source of oxidants: role in senescent cells. Free Radical Biology and Medicine, 48(8), 1100-1108. PMID: 20116426 DOI: 10.1016/j.freeradbiomed.2010.01.030
  7. Double, K. L., Dedov, V. N., Fedorow, H., Kettle, E., Halliday, G. M., Garner, B., & Brunk, U. T. (2008). The comparative biology of neuromelanin and lipofuscin in the human brain. Cellular and Molecular Life Sciences, 65(11), 1669-1682. PMID: 18278576 Doi:10.1007/s00018-008-7581-9
  8. Terman, A., & Brunk, U. T. (1998). Lipofuscin: mechanisms of formation and increase with age. Apmis, 106(1‐6), 265-276. Doi:10.1111/j.1699-0463.1998.tb01346.x
  9. Marzabadi, M. R., & Løvaas, E. (1996). Spermine prevent iron accumulation and depress lipofuscin accumulation in cultured myocardial cells. Free Radical Biology and Medicine, 21(3), 375-381. DOI:10.1016/0891-5849(96)00038-x
  10. Terman, A., & Brunk, U. T. (2004). Lipofuscin. The international journal of biochemistry & cell biology, 36(8), 1400-1404. Doi:10.1016/j.biocel.2003.08.009
  11. Seehafer, S. S., & Pearce, D. A. (2006). You say lipofuscin, we say ceroid: defining autofluorescent storage material. Neurobiology of aging, 27(4), 576-588. PMID: 16455164 DOI: 10.1016/j.neurobiolaging.2005.12.006
  12. Albaghdadi, M. S., Ikegami, R., Kassab, M. B., Gardecki, J. A., Kunio, M., Chowdhury, M. M., ... & Jaffer, F. A. (2021). Near-infrared autofluorescence in atherosclerosis associates with ceroid and is generated by oxidized lipid-induced oxidative stress. Arteriosclerosis, Thrombosis, and Vascular Biology, 41(7), e385-e398. PMID: 34011166 PMC8222195 DOI: 10.1161/ATVBAHA.120.315612
  13. 13.0 13.1 Georgakopoulou, E. A., Tsimaratou, K., Evangelou, K., Fernandez, M. P., Zoumpourlis, V., Trougakos, I. P., ... & Gorgoulis, V. G. (2013). Specific lipofuscin staining as a novel biomarker to detect replicative and stress-induced senescence. A method applicable in cryo-preserved and archival tissues. Aging (Albany NY), 5(1), 37. PMID: 23449538 PMCID: PMC3616230 DOI: 10.18632/aging.100527
  14. Jung, T., Höhn, A., & Grune, T. (2010). Lipofuscin: detection and quantification by microscopic techniques. Advanced Protocols in Oxidative Stress II, 173-193. PMID: 20072918 DOI: 10.1007/978-1-60761-411-1_13
  15. Pinchuk, A. I., Harvey, H. R., & Eckert, G. L. (2016). Development of biochemical measures of age in the Alaskan red king crab Paralithodes camtschaticus (Anomura): Validation, refinement and initial assessment. Fisheries Research, 183, 92-98.https://doi.org/10.1016/j.fishres.2016.05.019
  16. Evangelou, K., Lougiakis, N., Rizou, S. V., Kotsinas, A., Kletsas, D., Muñoz‐Espín, D., ... & Gorgoulis, V. G. (2017). Robust, universal biomarker assay to detect senescent cells in biological specimens. Aging cell, 16(1), 192-197. PMID: 28165661 PMCID: PMC5242262 DOI: 10.1111/acel.12545
  17. Salmonowicz, H., & Passos, J. F. (2017). Detecting senescence: a new method for an old pigment. Aging Cell, 16(3), 432-434.
  18. Lozano‐Torres, B., Blandez, J. F., García‐Fernández, A., Sancenón, F., & Martínez‐Máñez, R. (2022). Lipofuscin labelling through biorthogonal strain‐promoted azide‐alkyne cycloaddition for the detection of senescent cells. The FEBS Journal. PMID: 35527516 DOI: 10.1111/febs.16477
  19. Evangelou, K., & Gorgoulis, V. G. (2017). Sudan Black B, the specific histochemical stain for lipofuscin: a novel method to detect senescent cells. In Oncogene-Induced Senescence (pp. 111-119). Humana Press, New York, NY. PMID: 27812872 DOI: 10.1007/978-1-4939-6670-7_10
  20. Brunk, U. T., & Terman, A. (2002). Lipofuscin: mechanisms of age-related accumulation and influence on cell function. Free Radical Biology and Medicine, 33(5), 611-619. DOI: 10.1016/s0891-5849(02)00959-0
  21. Höhn, A., & Grune, T. (2013). Lipofuscin: formation, effects and role of macroautophagy. Redox biology, 1(1), 140-144. PMID: 24024146 PMCID: PMC3757681 DOI: 10.1016/j.redox.2013.01.006
  22. Nozynski, J., Zakliczynski, M., Konecka-Mrowka, D., Zakliczynska, H., Pijet, M., Zembala-Nozynska, E., ... & Zembala, M. (2013). Advanced glycation end products and lipofuscin deposits share the same location in cardiocytes of the failing heart. Experimental Gerontology, 48(2), 223-228. PMID: 22982091 DOI: 10.1016/j.exger.2012.09.002
  23. Baldensperger T., Jung T., Heinze T., Schwerdtle T., Höhn A., Grune T. (2024). Age pigment lipofuscin causes oxidative stress, lysosomal dysfunction, and pyroptotic cell death. bioRxiv .03.25.586520; doi: https://doi.org/10.1101/2024.03.25.586520
  24. Feldman, T. B., Dontsov, A. E., Yakovleva, M. A., & Ostrovsky, M. A. (2022). Photobiology of lipofuscin granules in the retinal pigment epithelium cells of the eye: norm, pathology, age. Biophysical Reviews, 1-15. PMID: 36124271 PMCID: PMC9481861 (available on 2023-08-08) DOI: 10.1007/s12551-022-00989-9
  25. Jung, T., Bader, N., & Grune, T. (2007). Lipofuscin: formation, distribution, and metabolic consequences. Annals of the New York Academy of Sciences, 1119(1), 97-111. PMID: 18056959 DOI: 10.1196/annals.1404.008
  26. Jhingan, M., Singh, S. R., Samanta, A., Arora, S., Tucci, D., Amarasekera, S., ... & Chhablani, J. (2021). Drusen ooze: predictor for progression of dry age-related macular degeneration. Graefe's Archive for Clinical and Experimental Ophthalmology, 259(9), 2687-2694. DOI:10.1007/s00417-021-05147-7
  27. Sparrow, J. R., & Boulton, M. (2005). RPE lipofuscin and its role in retinal pathobiology. Experimental eye research, 80(5), 595-606.
  28. Brandstetter, C., Mohr, L. K., Latz, E., Holz, F. G., & Krohne, T. U. (2015). Light induces NLRP3 inflammasome activation in retinal pigment epithelial cells via lipofuscin-mediated photooxidative damage. Journal of Molecular Medicine, 93(8), 905-916. PMID: 25783493 PMC4510924 DOI: 10.1007/s00109-015-1275-1
  29. 29.0 29.1 Jin, H. L., & Jeong, K. W. (2022). Transcriptome Analysis of Long-Term Exposure to Blue Light in Retinal Pigment Epithelial Cells. Biomolecules & therapeutics, 30(3), 291. PMID: 35074938 PMCID: PMC9047491 DOI: 10.4062/biomolther.2021.155
  30. Zhao, T., Guo, X., & Sun, Y. (2021). Iron accumulation and lipid peroxidation in the aging retina: implication of ferroptosis in age-related macular degeneration. Aging and disease, 12(2), 529. PMID: 33815881 PMCID: PMC7990372 DOI: 10.14336/AD.2020.0912
  31. Ueda, K., Kim, H. J., Zhao, J., Song, Y., Dunaief, J. L., & Sparrow, J. R. (2018). Iron promotes oxidative cell death caused by bisretinoids of retina. Proceedings of the National Academy of Sciences, 115(19), 4963-4968. PMID: 29686088 PMCID: PMC5948992 DOI: 10.1073/pnas.1722601115
  32. Hassel, C., Couchet, M., Jacquemot, N., Blavignac, C., Loï, C., Moinard, C., & Cia, D. (2022). Citrulline protects human retinal pigment epithelium from hydrogen peroxide and iron/ascorbate induced damages. Journal of Cellular and Molecular Medicine, 26(10), 2808-2818. PMID: 35460170 PMCID: PMC9097847 DOI: 10.1111/jcmm.17294
  33. Cho, H. M., Jo, Y. D., & Choung, S. Y. (2022). Protective Effects of Spirulina maxima against Blue Light-Induced Retinal Damages in A2E-Laden ARPE-19 Cells and Balb/c Mice. Nutrients, 14(3), 401. PMID: 35276761 PMCID: PMC8840079 DOI: 10.3390/nu14030401
  34. www.ichortherapeutics.com
  35. Julien, S., Biesemeier, A., Kokkinou, D., Eibl, O., & Schraermeyer, U. (2011). Zinc deficiency leads to lipofuscin accumulation in the retinal pigment epithelium of pigmented rats. PLoS One, 6(12), e29245. PMID: 22216222 PMCID: PMC3245262 DOI: 10.1371/journal.pone.0029245
  36. Blasiak, J., Pawlowska, E., Chojnacki, J., Szczepanska, J., Chojnacki, C., & Kaarniranta, K. (2020). Zinc and autophagy in age-related macular degeneration. International Journal of Molecular Sciences, 21(14), 4994. PMID: 32679798 PMCID: PMC7404247 DOI: 10.3390/ijms21144994
  37. Age-Related Eye Disease Study Research Group. (2001). A randomized, placebo-controlled, clinical trial of high-dose supplementation with vitamins C and E, beta carotene, and zinc for age-related macular degeneration and vision loss: AREDS report no. 8. Archives of ophthalmology, 119(10), 1417-1436. PMID: 11594942 PMCID: PMC1462955 DOI: 10.1001/archopht.119.10.1417
  38. Kakimoto, Y., Okada, C., Kawabe, N., Sasaki, A., Tsukamoto, H., Nagao, R., & Osawa, M. (2019). Myocardial lipofuscin accumulation in ageing and sudden cardiac death. Scientific reports, 9(1), 1-8. PMID: 30824797 PMCID: PMC6397159 DOI: 10.1038/s41598-019-40250-0
  39. Li, W. W., Wang, H. J., Tan, Y. Z., Wang, Y. L., Yu, S. N., & Li, Z. H. (2021). Reducing lipofuscin accumulation and cardiomyocytic senescence of aging heart by enhancing autophagy. Experimental Cell Research, 403(1), 112585. PMID: 33811905 DOI: 10.1016/j.yexcr.2021.112585
  40. Wang, L., Xiao, C. Y., Li, J. H., Tang, G. C., & Xiao, S. S. (2022). Transport and Possible Outcome of Lipofuscin in Mouse Myocardium. Advances in Gerontology, 12(3), 247-263.
  41. Wang, L., Xiao, C. Y., Li, J. H., Tang, G. C., & Xiao, S. S. (2020). Observation of the Transport and Removal of Lipofuscin from the Mouse Myocardium using Transmission Electron Microscope. BioRxiv. https://doi.org/10.1101/2020.03.10.985507
  42. Moreno-García, A., Kun, A., Calero, O., Medina, M., & Calero, M. (2018). An overview of the role of lipofuscin in age-related neurodegeneration. Frontiers in Neuroscience, 12, 464. PMID: 30026686 PMCID: PMC6041410 DOI: 10.3389/fnins.2018.00464
  43. Ahmed, Z., Sheng, H., Xu, Y. F., Lin, W. L., Innes, A. E., Gass, J., ... & Lewis, J. (2010). Accelerated lipofuscinosis and ubiquitination in granulin knockout mice suggest a role for progranulin in successful aging. The American journal of pathology, 177(1), 311-324. PMID: 20522652 PMCID: PMC2893674 DOI: 10.2353/ajpath.2010.090915
  44. Rübe, C. E., Bäumert, C., Schuler, N., Isermann, A., Schmal, Z., Glanemann, M., ... & Scherthan, H. (2021). Human skin aging is associated with increased expression of the histone variant H2A. J in the epidermis. npj Aging and Mechanisms of Disease, 7(1), 1-11 PMID:33795696 PMC8016850 DOI:10.1038/s41514-021-00060-z
  45. Tonolli, P. N., Baptista, M. S., & Chiarelli-Neto, O. (2021). Melanin, lipofuscin and the effects of visible light in the skin. Journal of Photochemistry and Photobiology, 7, 100044. https://doi.org/10.1016/j.jpap.2021.100044
  46. Skoczyńska, A., Budzisz, E., Trznadel-Grodzka, E., & Rotsztejn, H. (2017). Melanin and lipofuscin as hallmarks of skin aging. Advances in Dermatology and Allergology/Postępy Dermatologii i Alergologii, 34(2), 97-103. PMID: 28507486 PMCID: PMC5420599 DOI: 10.5114/ada.2017.67070
  47. Monji, A., Morimoto, N., Okuyama, I., Yamashita, N., & Tashiro, N. (1994). Effect of dietary vitamin E on lipofuscin accumulation with age in the rat brain. Brain research, 634(1), 62-68. PMID: 8156392 DOI: 10.1016/0006-8993(94)90258-5
  48. Zhu, S., Li, X., Wu, F., Cao, X., Gou, K., Wang, C., & Lin, C. (2022). Blue light induces skin apoptosis and degeneration through activation of the endoplasmic reticulum stress-autophagy apoptosis axis: Protective role of hydrogen sulfide. Journal of Photochemistry and Photobiology B: Biology, 229, 112426. https://doi.org/10.1016/j.jphotobiol.2022.112426
  49. Sparrow, J. R., Gregory-Roberts, E., Yamamoto, K., Blonska, A., Ghosh, S. K., Ueda, K., & Zhou, J. (2012). The bisretinoids of retinal pigment epithelium. Progress in retinal and eye research, 31(2), 121-135. PMID: 22209824 PMCID: PMC3288746 DOI: 10.1016/j.preteyeres.2011.12.001
  50. Mata, N. L., Weng, J., & Travis, G. H. (2000). Biosynthesis of a major lipofuscin fluorophore in mice and humans with ABCR-mediated retinal and macular degeneration. Proceedings of the National Academy of Sciences, 97(13), 7154-7159. PMID: 10852960 PMCID: PMC16515 DOI: 10.1073/pnas.130110497
  51. Hasan, M., Glees, P., & Spoerri, P. E. (1974). Dissolution and removal of neuronal lipofuscin following dimethylaminoethyl p-chlorophenoxyacetate administration to guinea pigs. Cell and Tissue Research, 150(3), 369-375. PMID: 4367734 DOI: 10.1007/BF00220143
  52. Riga, S., & Riga, D. (1974). Effects of centrophenoxine on the lipofuscin pigments in the nervous system of old rats. Brain Research, 72(2), 265-275. PMID: 4151704 DOI: 10.1016/0006-8993(74)90864-6
  53. Spoerri, P. E., Glees, P., & El Ghazzawi, E. (1974). Accumulation of lipofuscin in the myocardium of senile guinea pigs: dissolution and removal of lipofuscin following dimethylaminoethyl p-chlorophenoxyacetate administration. An electron microscopic study. Mechanisms of Ageing and Development, 3, 311-321. PMID: 4618294 DOI: 10.1016/0047-6374(74)90027-x
  54. Hochschild, R. (1973). Effect of dimethylaminoethanol on the life span of senile male A/J mice. Experimental Gerontology, 8(4), 185-191. https://doi.org/10.1016/0531-5565(73)90025-9
  55. Hochschild, R. (1973). Effect of dimethylaminoethyl p-chlorophenoxyacetate on the life span of male Swiss Webster albino mice. Experimental gerontology, 8(4), 177-183. DOI: 10.1016/0531-5565(73)90024-7
  56. Patro, N. I. S. H. A., Sharma, S. P., & Patro, I. K. (1992). Lipofuscin accumulation in ageing myocardium & its removal by meclophenoxate. The Indian Journal of Medical Research, 96, 192-198. PMID:1512044
  57. Hochschild, R. (1973). Effect of dimethylaminoethanol on the life span of senile male A/J mice. Experimental Gerontology, 8(4), 185-191.
  58. Julien, S., & Schraermeyer, U. (2012). Lipofuscin can be eliminated from the retinal pigment epithelium of monkeys. Neurobiology of aging, 33(10), 2390-2397. PMID: 22244091 DOI: 10.1016/j.neurobiolaging.2011.12.009
  59. Julien‐Schraermeyer, S., Illing, B., Tschulakow, A., Taubitz, T., Guezguez, J., Burnet, M., & Schraermeyer, U. (2020). Penetration, distribution, and elimination of remofuscin/soraprazan in Stargardt mouse eyes following a single intravitreal injection using pharmacokinetics and transmission electron microscopic autoradiography: Implication for the local treatment of Stargardt’s disease and dry age‐related macular degeneration. Pharmacology Research & Perspectives, 8(6), e00683. PMID: 33164337 PMCID: PMC7649431 DOI: 10.1002/prp2.683
  60. 60.0 60.1 Fang, Y., Taubitz, T., Tschulakow, A. V., Heiduschka, P., Szewczyk, G., Burnet, M., ... & Julien-Schraermeyer, S. (2022). Removal of RPE lipofuscin results in rescue from retinal degeneration in a mouse model of advanced Stargardt disease: Role of reactive oxygen species. Free Radical Biology and Medicine, 182, 132-149. PMID: 35219849 DOI: 10.1016/j.freeradbiomed.2022.02.025
  61. 61.0 61.1 61.2 61.3 Oh, M., Yeom, J., Schraermeyer, U., Julien-Schraermeyer, S., & Lim, Y. H. (2022). Remofuscin induces xenobiotic detoxification via a lysosome-to-nucleus signaling pathway to extend the Caenorhabditis elegans lifespan. Scientific reports, 12(1), 1-13. PMID: 35504961 PMCID: PMC9064964 DOI: 10.1038/s41598-022-11325-2
  62. Sears, A. E., Bernstein, P. S., Cideciyan, A. V., Hoyng, C., Issa, P. C., Palczewski, K., ... & Scholl, H. P. (2017). Towards treatment of Stargardt disease: workshop organized and sponsored by the Foundation Fighting Blindness. Translational vision science & technology, 6(5), 6-6.PMID: 28920007 PMCID: PMC5599228 DOI: 10.1167/tvst.6.5.6
  63. Lee, J. R., & Jeong, K. W. (2022). NMDA Receptor Antagonists Degrade Lipofuscin via Autophagy in Human Retinal Pigment Epithelial Cells. Medicina, 58(8), 1129. PMID: 36013596 PMC9415004 DOI: 10.3390/medicina58081129
  64. Lee, J. R., & Jeong, K. W. (2023). N-retinylidene-N-retinylethanolamine degradation in human retinal pigment epithelial cells via memantine-and ifenprodil-mediated autophagy. The Korean Journal of Physiology & Pharmacology: Official Journal of the Korean Physiological Society and the Korean Society of Pharmacology, 27(5), 449. PMID: 37641807 PMC10466070 DOI: 10.4196/kjpp.2023.27.5.449
  65. McHugh, D., & Gil, J. (2018). Senescence and aging: Causes, consequences, and therapeutic avenues. Journal of Cell Biology, 217(1), 65-77.
  66. Khanna, K. K., Lavin, M. F., Jackson, S. P., & Mulhern, T. D. (2001). ATM, a central controller of cellular responses to DNA damage. Cell Death & Differentiation, 8(11), 1052-1065.
  67. Kang, H. T., Park, J. T., Choi, K., Kim, Y., Choi, H. J. C., Jung, C. W., ... & Park, S. C. (2017). Chemical screening identifies ATM as a target for alleviating senescence. Nature chemical biology, 13(6), 616-623. PMID: 28346404 DOI: 10.1038/nchembio.2342
  68. Song, S. B., & Hwang, E. S. (2020). High levels of ROS impair lysosomal acidity and autophagy flux in glucose-deprived fibroblasts by activating ATM and Erk pathways. Biomolecules, 10(5), 761. PMID: 32414146 PMC7277562 DOI: 10.3390/biom10050761
  69. Song, S. B., Shim, W., & Hwang, E. S. (2023). Lipofuscin granule accumulation requires autophagy activation. Molecules and Cells, 46(8), 486-495. PMID: 37438887 PMC10440269 DOI: 10.14348/molcells.2023.0019