Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 2882 word(s) 2882 2022-01-26 04:07:42 |
2 format correction -55 word(s) 2827 2022-02-11 07:59:38 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Kim, J. Peroxisomal Stress Response and Inter-Organelle Communication. Encyclopedia. Available online: https://encyclopedia.pub/entry/19356 (accessed on 16 June 2024).
Kim J. Peroxisomal Stress Response and Inter-Organelle Communication. Encyclopedia. Available at: https://encyclopedia.pub/entry/19356. Accessed June 16, 2024.
Kim, Jinoh. "Peroxisomal Stress Response and Inter-Organelle Communication" Encyclopedia, https://encyclopedia.pub/entry/19356 (accessed June 16, 2024).
Kim, J. (2022, February 11). Peroxisomal Stress Response and Inter-Organelle Communication. In Encyclopedia. https://encyclopedia.pub/entry/19356
Kim, Jinoh. "Peroxisomal Stress Response and Inter-Organelle Communication." Encyclopedia. Web. 11 February, 2022.
Peroxisomal Stress Response and Inter-Organelle Communication
Edit

Peroxisomes are single membrane-bound organelles found in all eukaryotic cells and organisms, from yeast to plants and mammals. They are key regulators of cellular and metabolic homeostasis. These organelles play important roles in redox metabolism, the oxidation of very-long-chain fatty acids (VLCFAs), and the biosynthesis of ether phospholipids. Given the essential role of peroxisomes in cellular homeostasis, peroxisomal dysfunction has been linked to various pathological conditions, tissue functional decline, and aging. In the past few decades, a variety of cellular signaling and metabolic changes have been reported to be associated with defective peroxisomes, suggesting that many cellular processes and functions depend on peroxisomes. Peroxisomes communicate with other subcellular organelles, such as the nucleus, mitochondria, endoplasmic reticulum (ER), and lysosomes. These inter-organelle communications are highly linked to the key mechanisms by which cells surveil defective peroxisomes and mount adaptive responses to protect them from damages. 

peroxisome reactive oxygen species acetyl-CoA plasmalogen ER stress mitochondrial dysfunction apoptosis pexophagy

1. Introduction

Eukaryotic cells contain a variety of membrane-enclosed organelles. Among them, peroxisomes are ubiquitous single membrane-bound organelles that play essential roles in cellular metabolism. Peroxisomes are the major site for both the production and degradation of hydrogen peroxide (H2O2) due to their high content of peroxisomal oxidases and antioxidant enzymes, such as catalase. Peroxisomes are also critical for the biosynthesis of ether phospholipids, cholesterol, bile acids, and polyunsaturated fatty acids (PUFAs). The importance of peroxisomes is evidenced by peroxisomal biogenesis disorders (PBDs), rare genetic diseases caused by mutations in any of the 13 different PEX genes encoding peroxins responsible for the import of membrane or matrix proteins to peroxisomes [1]. Zellweger syndrome (ZS) is the most severe PBD, and patients with ZS show severe brain, liver, kidney, and bone malfunctions [2].
Given the essential roles of peroxisomes in cellular and metabolic homeostasis [3][4], it is not surprising to find that peroxisomal dysfunction is strongly linked to the loss of redox homeostasis, dysregulated lipid metabolism, mitochondrial dysfunction, altered gene expression, and elevated ER stress and cell death. More recently, it has become clear that peroxisome deficiency can also alter functions of specific subcellular organelles [5][6][7], such as mitochondria, endoplasmic reticulum (ER), and lysosome. Peroxisomal stress signals can also be transduced from peroxisomes to the nucleus, as peroxisomal dysfunction often induces global transcriptional changes [8][9]. Thus, similar to other organelle stress pathways (e.g., ER stress), there are specific peroxisomal stress response pathways that allow eukaryotic cells to cope with malfunctioning peroxisomes [10].
Peroxisomes can communicate with other organelles through membrane contact sites, peroxisome-derived metabolites, as well as retrograde signaling. However, it remains largely unclear how peroxisomal stress response pathways are activated and how they protect cells from damages. Understanding peroxisomal stress response pathways and how peroxisomes communicate with other organelles are important areas of peroxisome research because of their vital role in cellular homeostasis, aging, and aging-related diseases [11][12][13][14].

2. Peroxisome Biogenesis and Peroxisomal Import Machinery

Peroxisomes are single membrane-bound organelles found in all eukaryotic cells and organisms, from yeast to plants and mammals [15]. Under normal conditions, the peroxisome has a half-life of 1.5 to 2 days [16], suggesting that peroxisome biogenesis and degradation are very dynamic processes. Peroxisomal biogenesis is mostly regulated by peroxins encoded by PEX genes. To date, more than 30 conserved peroxins have been identified in yeast, plants, and many animal species [17]. Peroxisomes can be formed by de novo biogenesis from the ER and mitochondria [17][18][19]. Pre-peroxisomal vesicles originating from the ER contain the peroxisomal membrane protein (PMP) PEX16, which binds to the PMP receptor PEX3. PEX3 functions as a docking receptor for the cytosolic chaperone and PMP receptor PEX19 [20][21]. Together, these peroxins recruit other PMPs containing membrane peroxisome targeting signals (PTS) to the peroxisomal membrane during the peroxisome maturation process.
Peroxisomal matrix proteins are transported into the organelle after synthesis in the cytosol [22]. PTSs are essential for the correct sorting of matrix proteins [23]. Most peroxisomal matrix proteins contain a PTS type 1 (PTS1) sequence at the C-terminus consisting of an S-K-L motif or a conservative variant [24][25]. A small set of matrix proteins contains a PTS type 2 (PTS2) sequence, a nonapeptide with the consensus sequence (R/K)-(L/I/V)-X5-(H/Q)-(L/A) (X = any amino acid) that is localized near or at the N-terminus of the protein [26][27]. PTS1 proteins are transported to the peroxisome by the shuttling receptor PEX5, whereas PTS2 proteins are delivered to the organelle by PEX5 and PEX7 [23]. The cargo-loaded receptor proteins bind to the peroxisomal membrane proteins PEX13 and PEX14 [28][29]. After docking, the cargo proteins are translocated across the peroxisomal membrane and imported into the organelle lumen, while receptor protein PEX5 is then mono- or polyubiquitinated by E3 ubiquitin ligases PEX2, PEX10, and PEX12 [23][30][31]. PEX5 monoubiquitination is required for its recycling to the cytosol, while polyubiquitinated PEX5 can be targeted to the proteasome for degradation [32][33][34]. Cytosolic AAA-ATPases PEX1 and PEX6 form a heterohexameric ring as a trimer of PEX1/PEX6 heterodimers, which is anchored to the peroxisome membrane by PEX26 in higher organisms [35][36][37][38][39]. The PEX1-PEX6-PEX26 complex is responsible for releasing ubiquitinated PEX5 from the peroxisomal membrane to the cytosol [33][34][35]. In the cytosol, PEX5 can be deubiquitinated by deubiquitylating (DUB) enzymes (Ubp15 in yeast [40] and USP9X in mammals [41]) for the next cycle of peroxisomal import.
Mature peroxisomes divide and grow by elongation, membrane constriction, and fission [22][42][43][44][45]. In mammals, the peroxisomal membrane protein PEX11β plays a critical role in peroxisomal proliferation [42][46][47]. Exogenous expression of PEX11β in mammalian cells enhances peroxisomal proliferation [42], whereas the genetic defects of human PEX11β decrease peroxisomal abundance [47]. Although the molecular mechanisms underlying the modulation of PEX11β activity are not fully understood, docosahexaenoic acid (DHA) is known to play an important role in the initiation of peroxisomal elongation and fission [48]. Itoyama et al. [48] demonstrated that DHA, a peroxisomal β-oxidation metabolite, induced peroxisomal division by augmenting the hyper-oligomerization of Pex11pβ and the formation of Pex11pβ-enriched regions on elongated peroxisomes. The function of Pex11β in peroxisome proliferation depends on the homo-oligomerization through its N-terminal domain [46][49]. In addition, Pex11β interacts with the membrane via the N-terminal amphipathic helix to deform the membrane into elongated tubular peroxisomes [50][51][52]. Subsequently, Pex11β forms a ternary fission machinery complex with mitochondrial fission factor (Mff) and dynamin-like protein 1 (DLP1) at the constricted membrane region of elongated peroxisomes, promoting the scission of the membrane [51][53].

3. Key Metabolic Functions of Peroxisome

Peroxisomes are multifunctional organelles that play essential roles in several key metabolic pathways, including the β-oxidation of very-long-chain fatty acids (VLCFAs, >C22), the α-oxidation of branched-chain fatty acids, the oxidation of D-amino acids and polyamines, and the synthesis of ether phospholipids, bile acids, and DHA [2][17]. During these metabolic processes (e.g., VLCFA β-oxidation), reactive oxygen species (ROS) are produced as byproducts and subsequently neutralized by several antioxidant enzymes such as catalase [54][55][56]. About 50 different enzymes are involved in peroxisomal metabolic functions [2][57], and some of them are shared with other subcellular compartments, including mitochondria [57]. Accordingly, cooperation with other organelles, including mitochondria, ER, lipid droplets, and lysosomes, is necessary for proper peroxisomal metabolic functions.
Most fatty acids (FAs) are catabolized by β-oxidation, which removes two carbons from their carboxyl terminus [58]. In all eukaryotic cells, fatty acid oxidation occurs in both mitochondria and peroxisomes. FAs (<C20) are oxidized in mitochondria and converted to CO2 and H2O through the mitochondrial citric acid cycle and oxidative phosphorylation. However, VLCFAs cannot be oxidized in mitochondria. Instead, VLCFAs translocate across the peroxisomal membrane via ATP-binding cassette (ABC) subfamily D half-transporters (ABCD) and are broken down to acetyl-CoA, propionyl-CoA, or medium-chain acyl-CoAs. These different acyl-CoAs can be further converted into acylcarnitines in the peroxisome and translocate to mitochondria through the mitochondrial carnitine/acylcarnitine translocase (CACT). Acylcarnitines are then converted back to acyl-CoAs in mitochondria and enter the citric acid cycle [59].
Ether phospholipids are particular classes of phospholipids containing a glycerol backbone with an ether or vinyl-ether bond at the sn-1 position [60]. Plasmalogen is one type of ether phospholipid containing a vinyl-ether bond that is widely found in the brain, heart, and white blood cells [17]. The biosynthesis of plasmalogen begins in peroxisomes [17][60][61]. Dihydroxyacetone (DHAP), a glycolysis intermediate, is used as a building block of the glycerol backbone of plasmalogen. The peroxisomal matrix enzyme glyceronephosphate O-acyltransferase (GNPAT) transfers the acyl chain of acyl-CoA to DHAP to generate acyl-DHAP. Then, alkyl dihydroxyacetone phosphate synthase (AGPS) exchanges the acyl chain for an alkyl group to form alkyl-DHAP. Recently, Honsho et al. [62] showed that acyl/alkyl-DHAP reductase (ADHAPR/DHRS7B) localizes in peroxisomes and the ER as a type I integral membrane protein in HeLa cells. The authors demonstrated that ER-localized ADHAPR reduces alkyl-DHAP, while peroxisomal ADHAPR preferentially reduces acyl-DHAP before the subsequent synthesis of alkyl-DHAP catalyzed by alkyl-DHAP synthase (ADAPS) [62]. The remaining steps of plasmalogen synthesis take place in the ER [63][64].
Peroxisomes are also important for ROS metabolism. The name “peroxisome” was first introduced by De Duve in 1965 to define an organelle containing several enzymes that produce or break down hydrogen peroxide (H2O2) [4][65]. Peroxisomes are responsible for up to 20% of the cell’s oxygen consumption and 35% of the H2O2 production in mammalian tissues [66]. Peroxisomes produce different ROS species such as superoxide, hydroxyl radicals, and hydrogen peroxide as byproducts of metabolic processes by multiple oxidases present in the organelle. Acyl-CoA oxidases (ACOX) are the most abundant ROS-generating enzymes in the peroxisome [59]. D-amino acid oxidase (DAO), D-aspartate oxidase (DDO), L-pipecolic acid oxidase (PIPOX), L-α-hydroxy acid oxidases (HAO), polyamine oxidase (PAOX), and xanthine oxidase (XDH) are also important H2O2 producers in human peroxisomes. To protect the organelle from oxidative damage, human peroxisomes have several antioxidant enzymes including catalase (CAT), copper/zinc-containing superoxide dismutase (SOD1), peroxiredoxin 5 (PRDX5), glutathione S-transferase kappa 1 (GSTK1), microsomal GST1 (MGST1), and epoxide hydrolase 2 (EPHX2) [67].

4. Cellular Responses to Peroxisomal Dysfunction

The metabolic functions of peroxisomes are critical for cellular homeostasis, and the impairment of proper peroxisomal function elicits a variety of cellular responses (Figure 1).
Figure 1. Cellular responses to peroxisomal dysfunction. Impaired peroxisomal function elicits cellular responses such as the impaired ROS homeostasis, dysregulated lipid metabolism, mitochondrial dysfunction, altered gene expression, pexophagy, elevated ER stress, and apoptosis.

5. Peroxisome-Organelle Communication

Peroxisomes are multifunctional and dynamic organelles that actively contribute to cellular signaling, cell fate, immunity and inflammation, and aging [6]. To accomplish these cellular activities, peroxisomes interact and coordinate with other subcellular organelles, such as mitochondria, ER, lysosomes, and lipid droplets [6][59][68]. The cellular changes that accompany peroxisomal dysfunction are strongly associated with peroxisome-organelle crosstalk. Here, the current knowledge on peroxisome-organelle communication involving membrane contact sites, peroxisome-derived signaling metabolites, and retrograde signaling is summarized (Figure 2 and Figure 3).
Figure 2. Peroxisome-organelle membrane contact sites and their roles in organelle functions. Peroxisomes’ contact with mitochondria involves (1) peroxisomal Pex11 and mitochondrial Mdm34, (2) peroxisomal Pex34 and an unknown mitochondrial protein, (3) mitofusin Fzo1 on both peroxisome and mitochondria, and (4) peroxisomal ACBD2/ECI2 and mitochondrial TOMM20. The interaction between peroxisomes and the ER involves (1) peroxisomal ACBD5 and resident ER protein VAPB and (2) peroxisomal ABCD3 and ER protein ATF6α upon Ceapin treatment. In mammalian cells, the contact between peroxisomes and lysosomes involves phospholipid PI(4,5)P2 on the peroxisomal membrane and lysosomal Syt7. Abbreviations: Pex11, peroxin 11; Mdm34, mitochondrial distribution and morphology protein 34; Pex34, peroxin 34; Fzo1, mitofusin; ACBD2/ECI2, acyl-coenzyme A-binding domain or enoyl-CoA-δ isomerase 2; TOMM20, translocase of outer mitochondrial membrane 20; ACBD5, acyl-CoA binding domain-containing protein 5; VAPB, vesicle-associated membrane protein-associated protein B/C; ABCD3, ATP binding cassette subfamily D member 3; ATF6, activating transcription factor 6; PI(4,5)P2, phosphatidylinositol 4,5-bisphosphate; Syt7, synaptotagmin VII.
Figure 3. Metabolic and molecular signaling pathways mediate peroxisome-organelle crosstalk in response to peroxisomal dysfunction. (1) Peroxisome-nucleus: Peroxisomal dysfunction activates nuclear gene transcription through the transcription factors PPARα, AP-1, or Hnf4. The accumulation of long-chain fatty acids (LCFAs) activates PPARα to upregulate LONP2 and CAT expression to restore peroxisomal function (Blue arrow). Elevated ROS activates JNK signaling and AP-1 to induce the transcription of the pro-inflammatory cytokine IL-6. Peroxisomal dysfunction activates Hnf4 through unknown mechanisms to induce the expression of lipase 3 (Lip3), which leads to excessive free fatty acids (FFA) in the cytoplasm. FFAs hyper-activate Hnf4 via a positive feedback loop. (2) Peroxisome-mitochondrion: Peroxisomal dysfunction causes mitochondrial swelling, mitochondrial ROS generation, mitochondrial DNA (mtDNA) damage, disruption of oxidative phosphorylation (OXPHOS), and decreased inner membrane potential. Excessive ROS and elevated VLCFA are the potential signaling molecules that mediate the peroxisome-mitochondrion crosstalk. Peroxisomal dysfunction decreases plasmalogen, resulting in the inhibition of mitochondrial fission. (3) Peroxisome-ER: Peroxisomal dysfunction activates PERK signaling through unknown mechanisms to increase the phosphorylation of eIF2α and ATF4 induction. (4) Peroxisome-lysosome: Peroxisomal dysfunction due to the loss of Acox1 decreases peroxisome-derived acetyl-CoA, resulting in Raptor acetylation and lysosome localization of mTOR, leading to ULK1 activation and elevated lipophagy. Abbreviations: PPARα, peroxisome proliferator activated receptor alpha; AP-1, activator protein-1; Hnf4, hepatocyte nuclear factor 4; LONP2, lon peptidase 2; CAT, catalase; upd3, unpaired 3; IL-6, interleukin 6; Lip3, lipase 3; ROS, reactive oxygen species; JNK, Jun N-terminal kinase; VLCFAs, very-long-chain fatty acids; PERK, protein kinase RNA-like endoplasmic reticulum kinase; eIF2α, eukaryotic translation initiation factor 2a, ATF4, activating transcription factor 4; mTOR, mechanistic target of rapamycin kinase; ULK1, Unc-51-like autophagy-activating kinase 1.

6. Peroxisomal Dysfunction in Aging and Aging-Related Diseases

Several lines of evidence suggest that peroxisomal dysfunction is an underappreciated cause of aging [11][12][13][14]. Peroxisomal protein import is known to be compromised in aged tissues [11][12][13][14], and almost all peroxisomal proteins are downregulated during aging [8][13]. In senescent human fibroblasts, catalase import was significantly reduced, leading to the accumulation of H2O2 and the further disruption of peroxisome import [11]. Conversely, the overexpression of Pex5 has been shown to restore peroxisomal import function and preserve cardiac health in aged oenocytes [14]. Importantly, Sebastiani et al. [69] studied the proteomic signatures of centenarians and identified that proteins encoding components of peroxisomes were significantly downregulated in centenarians’ serum compared with younger individuals. In addition, the bile acid metabolism pathway and cholesterol homeostasis pathway were also downregulated in centenarians. These studies provide strong evidence for a role of impaired peroxisomal homeostasis in aging.
As peroxisomal activities decline with age [70], it has been suggested that peroxisomal dysfunctions might be associated with the pathogenesis of age-related neurodegenerative disease, including Alzheimer’s disease [71], Parkinson’s disease [72], and amyotrophic lateral sclerosis [73]), as well as age-related metabolic disease, such as cardiovascular disease [74], obesity [75], diabetes [76], and nonalcoholic liver disease [77]. These age-related pathologies are likely due to increased oxidative stress [78][79] and decreased plasmalogen synthesis [80]
Alzheimer’s disease (AD) is the most common neurodegenerative disease in the elderly. It is characterized by the progressive loss of synapses and neurons especially in the frontal cortex and hippocampus accompanied by significant memory loss and cognitive impairment [81][82]. The pathological hallmarks of AD are the accumulation of the extracellular beta-amyloid (Aβ) plaques and intracellular fibrillary deposits of hyperphosphorylated tau protein (neurofibrillary tangles). Oxidative stress is thought to be a primary culprit in AD pathogenesis [83][84]. Accordingly, the hippocampus of AD mice was found to exhibit high levels of markers of lipid peroxidation and DNA/RNA oxidation [83]. Several lines of evidence suggest a link between peroxisomal dysfunction and AD pathogenesis. The accumulation of VLCFAs and decreased plasmalogen levels were observed in the cortical regions of AD patients [85]. Also, the levels of DHA in the brain and liver tissue of AD patients were reduced compared to those in control individuals [86]. Impaired peroxisomal function might also contribute to alterations of peroxisomal metabolites in AD pathology. Shi et al. [87] showed that the inhibition of peroxisomal β-oxidation increased the level of VLCFAs and Aβ contents in rat cerebral cortex. On the other hand, treatment with peroxisome proliferators attenuated AD-related pathology and improved spatial memory in transgenic mice models of AD [88]. Although peroxisomal dysfunction is associated with AD pathogenesis, whether decreased peroxisomal activity is the primary cause, bystander, or consequence in AD pathology remains to be determined [89].
Cardiovascular diseases (CVDs) are the major cause of death in elderly populations [90], and the prevalence of CVDs has been shown to increase with age [91]. Aging plays a critical role in the impairment of the cardiovascular system and increases the risk of CVDs [92]. Peroxisomal dysfunction has been implicated in CVDs. For instance, cardiac-specific overexpression of catalase, predominantly located in peroxisomes [93], was found to prevent the progression to overt heart failure [74]. In addition, catalase activity is significantly decreased in failing myocardium [94]. Refsum’s disease is a peroxisomal disorder caused by the impaired α-oxidation of branched-chain fatty acids, and patients with Refsum’s disease develop cardiac arrhythmia and heart failure later in life [95][96][97]
Aging is accompanied by changes in the distribution and composition of adipose tissue, particularly an increase in fat deposition within the white adipose tissue (abdominal obesity), which is a major contributor to insulin resistance and metabolic syndrome [98][99][100]. As life expectancy increases, the prevalence of obesity has also risen steadily among the elderly [100]. A recent study demonstrated that peroxisomes play an important role in adipose dysfunction associated with obesity [75]. The authors found that obese mice exhibited downregulation of peroxisomal genes in white adipose tissue, and the knockdown of Pex5 increased ROS levels and inflammation in adipocytes. Accordingly, catalase knockout mice also showed accelerated obesity compared to control mice. These phenotypes were attenuated by improving peroxisomal biogenesis through treatment with the PPARα agonist fenofibrate [75]. Additionally, catalase-deficient patients frequently show diabetic pathology, presumably due to the accumulation of oxidative damage in pancreatic β-cells [101].

References

  1. Klouwer, F.C.; Berendse, K.; Ferdinandusse, S.; Wanders, R.J.; Engelen, M. Zellweger spectrum disorders: Clinical overview and management approach. Orphanet J. Rare. Dis. 2015, 10, 151.
  2. Waterham, H.R.; Ferdinandusse, S.; Wanders, R.J. Human disorders of peroxisome metabolism and biogenesis. Biochim. Biophys. Acta 2016, 1863, 922–933.
  3. He, A.; Dean, J.M.; Lodhi, I.J. Peroxisomes as cellular adaptors to metabolic and environmental stress. Trends Cell Biol. 2021, 31, 656–670.
  4. Wanders, R.J. Metabolic functions of peroxisomes in health and disease. Biochimie 2014, 98, 36–44.
  5. Schrader, M.; Kamoshita, M.; Islinger, M. Organelle interplay-peroxisome interactions in health and disease. J. Inherit. Metab. Dis. 2020, 43, 71–89.
  6. Lin, T.K.; Lin, K.J.; Lin, K.L.; Liou, C.W.; Chen, S.D.; Chuang, Y.C.; Wang, P.W.; Chuang, J.H.; Wang, T.J. When Friendship Turns Sour: Effective Communication Between Mitochondria and Intracellular Organelles in Parkinson’s Disease. Front. Cell Dev. Biol. 2020, 8, 607392.
  7. Jin, Y.; Strunk, B.S.; Weisman, L.S. Close encounters of the lysosome-peroxisome kind. Cell 2015, 161, 197–198.
  8. Huang, K.; Chen, W.; Zhu, F.; Li, P.; Li, P.W.-L.; Kapahi, P.; Bai, H. RiboTag translatomic profiling of Drosophila oenocytes under aging and induced oxidative stress. BMC Genom. 2019, 20, 50.
  9. Huang, K.; Kim, J.; Vo, P.; Miao, T.; Bai, H. Peroxisome import stress impairs ribosome biogenesis and induces integrative stress response through eIF2α phosphorylation. bioRxiv 2021.
  10. Rackles, E.; Witting, M.; Forné, I.; Zhang, X.; Zacherl, J.; Schrott, S.; Fischer, C.; Ewbank, J.J.; Osman, C.; Imhof, A.; et al. Reduced peroxisomal import triggers peroxisomal retrograde signaling. Cell Rep. 2021, 34, 108653.
  11. Legakis, J.E.; Koepke, J.I.; Jedeszko, C.; Barlaskar, F.; Terlecky, L.J.; Edwards, H.J.; Walton, P.A.; Terlecky, S.R. Peroxisome senescence in human fibroblasts. Mol. Biol. Cell 2002, 13, 4243–4255.
  12. Giordano, C.R.; Terlecky, S.R. Peroxisomes, cell senescence, and rates of aging. Biochim. Biophys. Acta Mol. Basis Dis. 2012, 1822, 1358–1362.
  13. Narayan, V.; Ly, T.; Pourkarimi, E.; Murillo, A.B.; Gartner, A.; Lamond, A.; Kenyon, C. Deep Proteome Analysis Identifies Age-Related Processes in C. elegans. Cell Syst. 2016, 3, 144–159.
  14. Huang, K.; Miao, T.; Chang, K.; Kim, J.; Kang, P.; Jiang, Q.; Simmonds, A.J.; Di Cara, F.; Bai, H. Impaired peroxisomal import in Drosophila oenocytes causes cardiac dysfunction by inducing upd3 as a peroxikine. Nat. Commun. 2020, 11, 2943.
  15. Islinger, M.; Cardoso, M.; Schrader, M. Be different—The diversity of peroxisomes in the animal kingdom. Biochim. et Biophys. Acta 2010, 1803, 881–897.
  16. Huybrechts, S.J.; Van Veldhoven, P.P.; Brees, C.; Mannaerts, G.P.; Los, G.V.; Fransen, M. Peroxisome Dynamics in Cultured Mammalian Cells. Traffic 2009, 10, 1722–1733.
  17. Lodhi, I.J.; Semenkovich, C.F. Peroxisomes: A Nexus for Lipid Metabolism and Cellular Signaling. Cell Metab. 2014, 19, 380–392.
  18. Sugiura, A.; Mattie, S.; Prudent, J.; McBride, H.M. Newly born peroxisomes are a hybrid of mitochondrial and ER-derived pre-peroxisomes. Nature 2017, 542, 251–254.
  19. Van der Zand, A.; Gent, J.; Braakman, I.; Tabak, H.F. Biochemically distinct vesicles from the endoplasmic reticulum fuse to form peroxisomes. Cell 2012, 149, 397–409.
  20. Matsuzaki, T.; Fujiki, Y. The peroxisomal membrane protein import receptor Pex3p is directly transported to peroxisomes by a novel Pex19p- and Pex16p-dependent pathway. J. Cell Biol. 2008, 183, 1275–1286.
  21. Fang, Y.; Morrell, J.C.; Jones, J.M.; Gould, S.J. PEX3 functions as a PEX19 docking factor in the import of class I peroxisomal membrane proteins. J. Cell Biol. 2004, 164, 863–875.
  22. Lazarow, P.B.; Fujiki, Y. Biogenesis of Peroxisomes. Annu. Rev. Cell Biol. 1985, 1, 489–530.
  23. Francisco, T.; Rodrigues, T.; Dias, A.F.; Barros-Barbosa, A.; Bicho, D.; Azevedo, J.E. Protein transport into peroxisomes: Knowns and unknowns. BioEssays 2017, 39, 39.
  24. Platta, H.; Erdmann, R. The peroxisomal protein import machinery. FEBS Lett. 2007, 581, 2811–2819.
  25. Gould, S.J.; Keller, G.-A.; Hosken, N.; Wilkinson, J.; Subramani, S. A conserved tripeptide sorts proteins to peroxisomes. J. Cell Biol. 1989, 108, 1657–1664.
  26. Lazarow, P.B. Chapter 3.1.7. The import receptor Pex7p and the PTS2 targeting sequence. Biochim. Biophys. Acta Bioenerg. 2006, 1763, 1599–1604.
  27. Swinkels, B.W.; Gould, S.J.; Bodnar, A.G.; Rachubinski, R.A.; Subramani, S. A novel, cleavable peroxisomal targeting signal at the amino-terminus of the rat 3-ketoacyl-CoA thiolase. EMBO J. 1991, 10, 3255–3262.
  28. Meinecke, M.; Cizmowski, C.; Schliebs, W.; Krüger, V.; Beck, S.; Wagner, R.H.; Erdmann, R. The peroxisomal importomer constitutes a large and highly dynamic pore. Nat. Cell Biol. 2010, 12, 273–277.
  29. Rucktäschel, R.; Girzalsky, W.; Erdmann, R. Protein import machineries of peroxisomes. Biochim. Biophys. Acta Biomembr. 2011, 1808, 892–900.
  30. Platta, H.W.; El Magraoui, F.; Baumer, B.E.; Schlee, D.; Girzalsky, W.; Erdmann, R. Pex2 and Pex12 Function as Protein-Ubiquitin Ligases in Peroxisomal Protein Import. Mol. Cell. Biol. 2009, 29, 5505–5516.
  31. Platta, H.W.; Brinkmeier, R.; Reidick, C.; Galiani, S.; Clausen, M.P.; Eggeling, C. Regulation of peroxisomal matrix protein import by ubiquitination. Biochim. Biophys. Acta Bioenerg. 2016, 1863, 838–849.
  32. Thoms, S.; Erdmann, R. Peroxisomal matrix protein receptor ubiquitination and recycling. Biochim. Biophys. Acta Bioenerg. 2006, 1763, 1620–1628.
  33. Carvalho, A.F.; Pinto, M.P.; Grou, C.P.; Alencastre, I.S.; Fransen, M.; Sa-Miranda, C.; Azevedo, J.E. Ubiquitination of Mammalian Pex5p, the Peroxisomal Import Receptor. J. Biol. Chem. 2007, 282, 31267–31272.
  34. Grou, C.P.; Carvalho, A.F.; Pinto, M.P.; Alencastre, I.S.; Rodrigues, T.A.; Freitas, M.O.; Francisco, T.; Sá-Miranda, C.; Azevedo, J.E. The peroxisomal protein import machinery—A case report of transient ubiquitination with a new flavor. Cell. Mol. Life Sci. 2008, 66, 254–262.
  35. Matsumoto, N.; Tamura, S.; Fujiki, Y. The pathogenic peroxin Pex26p recruits the Pex1p–Pex6p AAA ATPase complexes to peroxisomes. Nat. Cell Biol. 2003, 5, 454–460.
  36. Blok, N.B.; Tan, D.; Wang, R.Y.; Penczek, P.A.; Baker, D.; DiMaio, F.; Rapoport, T.A.; Walz, T. Unique double-ring structure of the peroxisomal Pex1/Pex6 ATPase complex revealed by cryo-electron microscopy. Proc. Natl. Acad. Sci. USA 2015, 112, E4017–E4025.
  37. Ciniawsky, S.; Grimm, I.; Saffian, D.; Girzalsky, W.; Erdmann, R.; Wendler, P. Molecular snapshots of the Pex1/6 AAA+ complex in action. Nat. Commun. 2015, 6, 7331.
  38. Tan, D.; Blok, N.B.; Rapoport, T.A.; Walz, T. Structures of the double-ring AAA ATPase Pex1-Pex6 involved in peroxisome biogenesis. FEBS J. 2015, 283, 986–992.
  39. Gardner, B.M.; Chowdhury, S.; Lander, G.C.; Martin, A. The Pex1/Pex6 complex is a heterohexameric AAA+ motor with alternating and highly coordinated subunits. J. Mol. Biol. 2015, 427, 1375–1388.
  40. Debelyy, M.; Platta, H.; Saffian, D.; Hensel, A.; Thoms, S.; Meyer, H.E.; Warscheid, B.; Girzalsky, W.; Erdmann, R. Ubp15p, a Ubiquitin Hydrolase Associated with the Peroxisomal Export Machinery. J. Biol. Chem. 2011, 286, 28223–28234.
  41. Grou, C.P.; Francisco, T.; Rodrigues, T.A.; Freitas, M.O.; Pinto, M.P.; Carvalho, A.F.; Domingues, P.; Wood, S.A.; Rodríguez-Borges, J.E.; Sá-Miranda, C.; et al. Identification of Ubiquitin-specific Protease 9X (USP9X) as a Deubiquitinase Acting on Ubiquitin-Peroxin 5 (PEX5) Thioester Conjugate. J. Biol. Chem. 2012, 287, 12815–12827.
  42. Schrader, M.; Reuber, B.E.; Morrell, J.C.; Jimenez-Sanchez, G.; Obie, C.; Stroh, T.A.; Valle, D.; Schroer, T.A.; Gould, S.J. Expression of PEX11beta mediates peroxisome proliferation in the absence of extracellular stimuli. J. Biol. Chem. 1998, 273, 29607–29614.
  43. Schrader, M.; Fahimi, H.D. Growth and Division of Peroxisomes. Adv. Appl. Microbiol. 2006, 255, 237–290.
  44. Fagarasanu, A.; Fagarasanu, M.; Rachubinski, R.A. Maintaining peroxisome populations: A story of division and inheritance. Annu. Rev. Cell Dev. Biol. 2007, 23, 321–344.
  45. Thoms, S.; Erdmann, R. Dynamin-related proteins and Pex11 proteins in peroxisome division and proliferation. FEBS J. 2005, 272, 5169–5181.
  46. Okumoto, K.; Tamura, S.; Honsho, M.; Fujiki, Y. Peroxisome: Metabolic Functions and Biogenesis. Adv. Exp. Med. Biol. 2020, 1299, 3–17.
  47. Ebberink, M.S.; Koster, J.; Visser, G.; van Spronsen, F.; Stolte-Dijkstra, I.; Smit, G.P.; Fock, J.M.; Kemp, S.; Wanders, R.J.; Waterham, H.R. A novel defect of peroxisome division due to a homozygous non-sense mutation in the PEX11β gene. J. Med. Genet. 2012, 49, 307–313.
  48. Itoyama, A.; Honsho, M.; Abe, Y.; Moser, A.; Yoshida, Y.; Fujiki, Y. Docosahexaenoic acid mediates peroxisomal elongation, a prerequisite for peroxisome division. J. Cell Sci. 2012, 125, 589–602.
  49. Kobayashi, S.; Tanaka, A.; Fujiki, Y. Fis1, DLP1, and Pex11p coordinately regulate peroxisome morphogenesis. Exp. Cell Res. 2007, 313, 1675–1686.
  50. Opaliński, Ł.; Kiel, J.A.; Williams, C.; Veenhuis, M.; Van Der Klei, I.J. Membrane curvature during peroxisome fission requires Pex11. EMBO J. 2011, 30, 5–16.
  51. Yoshida, Y.; Niwa, H.; Honsho, M.; Itoyama, A.; Fujiki, Y. Pex11mediates peroxisomal proliferation by promoting deformation of the lipid membrane. Biol. Open 2015, 4, 710–721.
  52. Su, J.; Thomas, A.S.; Grabietz, T.; Landgraf, C.; Volkmer, R.; Marrink, S.J.; Williams, C.; Melo, M.N. The N-terminal amphipathic helix of Pex11p self-interacts to induce membrane remodelling during peroxisome fission. Biochim. Biophys. Acta Biomembr. 2018, 1860, 1292–1300.
  53. Itoyama, A.; Michiyuki, S.; Honsho, M.; Yamamoto, T.; Moser, A.; Yoshida, Y.; Fujiki, Y. Mff functions with Pex11pβ and DLP1 in peroxisomal fission. Biol. Open 2013, 2, 998–1006.
  54. Bonekamp, N.A.; Völkl, A.; Fahimi, H.D.; Schrader, M. Reactive oxygen species and peroxisomes: Struggling for balance. BioFactors 2009, 35, 346–355.
  55. Islinger, M.; Li, K.W.; Seitz, J.; Völkl, A.; Lüers, G.H. Hitchhiking of Cu/Zn Superoxide Dismutase to Peroxisomes–Evidence for a Natural Piggyback Import Mechanism in Mammals. Traffic 2009, 10, 1711–1721.
  56. del Río, L.A.; Corpas, F.J.; Sandalio, L.M.; Palma, J.M.; Gómez, M.; Barroso, J.B. Reactive oxygen species, antioxidant systems and nitric oxide in peroxisomes. J. Exp. Bot. 2002, 53, 1255–1272.
  57. Wanders, R.J.; Waterham, H.R. Biochemistry of Mammalian Peroxisomes Revisited. Annu. Rev. Biochem. 2006, 75, 295–332.
  58. Cipolla, C.M.; Lodhi, I.J. Peroxisomal Dysfunction in Age-Related Diseases. Trends Endocrinol. Metab. 2017, 28, 297–308.
  59. Wanders, R.J.A.; Waterham, H.R.; Ferdinandusse, S. Metabolic Interplay between Peroxisomes and Other Subcellular Organelles Including Mitochondria and the Endoplasmic Reticulum. Front. Cell Dev. Biol. 2016, 3, 83.
  60. Nagan, N.; Zoeller, R.A. Plasmalogens: Biosynthesis and functions. Prog. Lipid Res. 2001, 40, 199–229.
  61. Dean, J.M.; Lodhi, I.J. Structural and functional roles of ether lipids. Protein Cell 2018, 9, 196–206.
  62. Honsho, M.; Tanaka, M.; Zoeller, R.A.; Fujiki, Y. Distinct Functions of Acyl/Alkyl Dihydroxyacetonephosphate Reductase in Peroxisomes and Endoplasmic Reticulum. Front. Cell Dev. Biol. 2020, 8, 855.
  63. Gallego-García, A.; Monera-Girona, A.J.; Pajares-Martínez, E.; Bastida-Martínez, E.; Pérez-Castaño, R.; Iniesta, A.A.; Fontes, M.; Padmanabhan, S.; Elías-Arnanz, M. A bacterial light response reveals an orphan desaturase for human plasmalogen synthesis. Science 2019, 366, 128–132.
  64. Werner, E.R.; Keller, M.A.; Sailer, S.; Lackner, K.; Koch, J.; Hermann, M.; Coassin, S.; Golderer, G.; Werner-Felmayer, G.; Zo-eller, R.A.; et al. The TMEM189 gene encodes plasmanylethanolamine desaturase which introduces the characteristic vinyl ether double bond into plasmalogens. Proc. Natl. Acad. Sci. USA 2020, 117, 7792–7798.
  65. De Duve, C.; Baudhuin, P. Peroxisomes (microbodies and related particles). Physiol. Rev. 1966, 46, 323–357.
  66. Boveris, A.; Oshino, N.; Chance, B. The cellular production of hydrogen peroxide. Biochem. J. 1972, 128, 617–630.
  67. Fransen, M.; Nordgren, M.; Wang, B.; Apanasets, O. Role of peroxisomes in ROS/RNS-metabolism: Implications for human disease. Biochim. et Biophys. Acta Mol. Basis Dis. 2011, 1822, 1363–1373.
  68. Fransen, M.; Lismont, C.; Walton, P. The Peroxisome-Mitochondria Connection: How and Why? Int. J. Mol. Sci. 2017, 18, 1126.
  69. Sebastiani, P.; Federico, A.; Morris, M.; Gurinovich, A.; Tanaka, T.; Chandler, K.B.; Andersen, S.L.; Denis, G.; Costello, C.E.; Ferrucci, L.; et al. Protein signatures of centenarians and their offspring suggest centenarians age slower than other humans. Aging Cell 2021, 20, e13290.
  70. Titorenko, V.I.; Terlecky, S.R. Peroxisome Metabolism and Cellular Aging. Traffic 2010, 12, 252–259.
  71. Lizard, G.; Rouaud, O.; Demarquoy, J.; Cherkaoui-Malki, M.; Iuliano, L. Potential roles of peroxisomes in Alzheimer’s disease and in dementia of the Alzheimer’s type. J. Alzheimer Dis. 2012, 29, 241–254.
  72. Jo, D.S.; Park, N.Y.; Cho, D.-H. Peroxisome quality control and dysregulated lipid metabolism in neurodegenerative diseases. Exp. Mol. Med. 2020, 52, 1486–1495.
  73. Pharaoh, G.; Sataranatarajan, K.; Street, K.; Hill, S.; Gregston, J.; Ahn, B.; Kinter, C.; Kinter, M.; Van Remmen, H. Metabolic and Stress Response Changes Precede Disease Onset in the Spinal Cord of Mutant SOD1 ALS Mice. Front. Neurosci. 2019, 13, 487.
  74. Qin, F.; Lennon-Edwards, S.; Lancel, S.; Biolo, A.; Siwik, D.A.; Pimentel, D.R.; Dorn, G.W.; Kang, Y.J.; Colucci, W.S. Cardiac-specific overexpression of catalase identifies hydrogen peroxide-dependent and -independent phases of myocardial remodeling and prevents the progression to overt heart failure in G(alpha)q-overexpressing transgenic mice. Circ. Heart Fail. 2010, 3, 306–313.
  75. Piao, L.; Dorotea, D.; Jiang, S.; Koh, E.H.; Oh, G.T.; Ha, H. Impaired Peroxisomal Fitness in Obese Mice, a Vicious Cycle Exacerbating Adipocyte Dysfunction via Oxidative Stress. Antioxid. Redox Signal. 2019, 31, 1339–1351.
  76. Hwang, I.; Lee, J.; Huh, J.Y.; Park, J.; Lee, H.B.; Ho, Y.-S.; Ha, H. Catalase Deficiency Accelerates Diabetic Renal Injury Through Peroxisomal Dysfunction. Diabetes 2012, 61, 728–738.
  77. Francque, S.; Szabo, G.; Abdelmalek, M.F.; Byrne, C.D.; Cusi, K.; Dufour, J.-F.; Roden, M.; Sacks, F.; Tacke, F. Nonalcoholic steatohepatitis: The role of peroxisome proliferator-activated receptors. Nat. Rev. Gastroenterol. Hepatol. 2021, 18, 24–39.
  78. Marcus, D.L.; Thomas, C.; Rodriguez, C.; Simberkoff, K.; Tsai, J.S.; Strafaci, J.A.; Freedman, M.L. Increased peroxidation and reduced antioxidant enzyme activity in Alzheimer’s disease. Exp. Neurol. 1998, 150, 40–44.
  79. Sheikh, F.G.; Pahan, K.; Khan, M.; Barbosa, E.; Singh, I. Abnormality in catalase import into peroxisomes leads to severe neurological disorder. Proc. Natl. Acad. Sci. USA 1998, 95, 2961–2966.
  80. Goodenowe, D.B.; Cook, L.L.; Liu, J.; Lu, Y.; Jayasinghe, D.A.; Ahiahonu, P.W.; Heath, D.; Yamazaki, Y.; Flax, J.; Krenitsky, K.F.; et al. Peripheral ethanolamine plasmalogen deficiency: A logical causative factor in Alzheimer’s disease and dementia. J. Lipid Res. 2007, 48, 2485–2498.
  81. Graham, W.V.; Bonito-Oliva, A.; Sakmar, T.P. Update on Alzheimer’s Disease Therapy and Prevention Strategies. Annu. Rev. Med. 2017, 68, 413–430.
  82. Cimini, A.; Moreno, S.; D’Amelio, M.; Cristiano, L.; D’Angelo, B.; Falone, S.; Benedetti, E.; Carrara, P.; Fanelli, F.; Cecconi, F.; et al. Early biochemical and morphological modifications in the brain of a transgenic mouse model of Alzheimer’s disease: A role for peroxisomes. J. Alzheimer Dis. 2009, 18, 935–952.
  83. Fanelli, F.; Sepe, S.; D’Amelio, M.; Bernardi, C.; Cristiano, L.; Cimini, A.; Cecconi, F.; Ceru’, M.P.; Moreno, S. Age-dependent roles of peroxisomes in the hippocampus of a transgenic mouse model of Alzheimer’s disease. Mol. Neurodegener. 2013, 8, 8.
  84. Nunomura, A.; Tamaoki, T.; Motohashi, N.; Nakamura, M.; McKeel, D.W., Jr.; Tabaton, M.; Lee, H.-G.; Smith, M.A.; Perry, G.; Zhu, X. The Earliest Stage of Cognitive Impairment in Transition from Normal Aging to Alzheimer Disease Is Marked by Prominent RNA Oxidation in Vulnerable Neurons. J. Neuropathol. Exp. Neurol. 2012, 71, 233–241.
  85. Kou, J.; Kovacs, G.G.; Höftberger, R.; Kulik, W.; Brodde, A.; Forss-Petter, S.; Hönigschnabl, S.; Gleiss, A.; Brügger, B.; Wanders, R.; et al. Peroxisomal alterations in Alzheimer’s disease. Acta Neuropathol. 2011, 122, 271–283.
  86. Astarita, G.; Jung, K.M.; Berchtold, N.C.; Nguyen, V.Q.; Gillen, D.L.; Head, E.; Cotman, C.W.; Piomelli, D. Deficient liver biosynthesis of docosahexaenoic acid correlates with cognitive impairment in Alzheimer’s disease. PLoS ONE 2010, 5, e12538.
  87. Shi, R.; Zhang, Y.; Shi, Y.; Shi, S.; Jiang, L. Inhibition of peroxisomal β-oxidation by thioridazine increases the amount of VLCFAs and Aβ generation in the rat brain. Neurosci. Lett. 2012, 528, 6–10.
  88. Inestrosa, N.C.; Carvajal, F.J.; Zolezzi, J.M.; Tapia-Rojas, C.; Serrano, F.; Karmelic, D.; Toledo, E.M.; Toro, A.; Toro, J.; Santos, M.J. Peroxisome proliferators reduce spatial memory impairment, synaptic failure, and neurodegeneration in brains of a double transgenic mice model of Alzheimer’s disease. J. Alzheimer Dis. 2013, 33, 941–959.
  89. Berger, J.; Dorninger, F.; Forss-Petter, S.; Kunze, M. Peroxisomes in brain development and function. Biochim. Biophys. Acta 2016, 1863, 934–955.
  90. Dai, D.-F.; Chen, T.; Johnson, S.; Szeto, H.; Rabinovitch, P.S. Cardiac Aging: From Molecular Mechanisms to Significance in Human Health and Disease. Antioxid. Redox Signal. 2012, 16, 1492–1526.
  91. Yazdanyar, A.; Newman, A.B. The Burden of Cardiovascular Disease in the Elderly: Morbidity, Mortality, and Costs. Clin. Geriatr. Med. 2009, 25, 563–577.
  92. Rodgers, J.L.; Jones, J.; Bolleddu, S.I.; Vanthenapalli, S.; Rodgers, L.E.; Shah, K.; Karia, K.; Panguluri, S.K. Cardiovascular Risks Associated with Gender and Aging. J. Cardiovasc. Dev. Dis. 2019, 6, 19.
  93. Mittler, R. Oxidative stress, antioxidants and stress tolerance. Trends Plant Sci. 2002, 7, 405–410.
  94. Bäumer, A.T.; Flesch, M.; Wang, X.; Shen, Q.; Feuerstein, G.Z.; Böhm, M. Antioxidative Enzymes in Human Hearts with Idiopathic Dilated Cardiomyopathy. J. Mol. Cell. Cardiol. 2000, 32, 121–130.
  95. Colasante, C.; Chen, J.; Ahlemeyer, B.; Baumgart-Vogt, E. Peroxisomes in cardiomyocytes and the peroxisome/peroxisome proliferator-activated receptor-loop. Thromb. Haemost. 2015, 113, 452–463.
  96. Koh, J.T.; Choi, H.H.; Ahn, K.Y.; Kim, J.U.; Kim, J.H.; Chun, J.-Y.; Baik, Y.H.; Kim, K.K. Cardiac Characteristics of Transgenic Mice Overexpressing Refsum Disease Gene-Associated Protein within the Heart. Biochem. Biophys. Res. Commun. 2001, 286, 1107–1116.
  97. Wanders, R.J.; Komen, J.C. Peroxisomes, Refsum’s disease and the alpha- and omega-oxidation of phytanic acid. Biochem. Soc. Trans. 2007, 35, 865–869.
  98. Miard, S.; Picard, F. Obesity and aging have divergent genomic fingerprints. Int. J. Obes. 2008, 32, 1873–1874.
  99. Barzilai, N.; Huffman, D.M.; Muzumdar, R.H.; Bartke, A. The Critical Role of Metabolic Pathways in Aging. Diabetes 2012, 61, 1315–1322.
  100. Jura, M.; Kozak, L. Obesity and related consequences to ageing. AGE 2016, 38, 23.
  101. Goth, L.; Eaton, J.W. Hereditary catalase deficiencies and increased risk of diabetes. Lancet 2000, 356, 1820–1821.
More
Information
Subjects: Cell Biology
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 811
Revisions: 2 times (View History)
Update Date: 15 Feb 2022
1000/1000
Video Production Service