Skip to content
Publicly Available Published by De Gruyter October 30, 2019

Mitochondrial dysfunction and energy deprivation in the mechanism of neurodegeneration

Nörodejenerasyon mekanizmasında mitokondriyal fonksiyon bozukluğu ve enerji yoksunluğu
  • Andrey Y. Abramov ORCID logo EMAIL logo and Plamena R. Angelova EMAIL logo

Abstract

Energy-producing organelles mitochondria are involved in a number of cellular functions. Deregulation of mitochondrial function due to mutations or effects of mitochondrial toxins is proven to be a trigger for diverse pathologies, including neurodegenerative disorders. Despite the extensive research done in the last decades, the mechanisms by which mitochondrial dysfunction leads to neuronal deregulation and cell death have not yet been fully elucidated. Brain cells are specifically dependent on mitochondria due to their high energy demands to maintain neuronal ion gradients and signal transduction, and also, to mediate neuronal health through the processes of mitochondrial calcium homeostasis, mitophagy, mitochondrial reactive oxygen species production and mitochondrial dynamics. Some of these processes have been independently implicated in the mechanism of neuronal loss in neurodegeneration. Moreover, it is increasingly recognised that these processes are interdependent and interact within the mitochondria to ensure proper neuronal function and survival.

Öz

Enerji üreten organeller mitokondriler birçok hücresel fonksiyonda rol oynar. Mitokondriyal toksinlerin mutasyonlara veya etkilerine bağlı olarak mitokondriyal fonksiyonun düzensizleştirilmesinin, nörodejeneratif hastalıklar dahil olmak üzere çeşitli patolojiler için bir tetikleyici olduğu kanıtlanmıştır. Son yıllarda yapılan kapsamlı araştırmalara rağmen, mitokondriyal disfonksiyonun nöronal düzensizliğe ve hücre ölümüne yol açtığı mekanizmalar henüz tam olarak açıklanamamıştır. Beyin hücreleri, nöronal iyon gradyanlarını ve sinyal iletimini sürdürmek için yüksek enerji taleplerinden ve ayrıca mitokondriyal kalsiyum homeostazisi, mitofajisi, mitokondriyal reaktif oksijen türlerinin üretimi ve mitokondriyal dinamiği süreçleri yoluyla nöronal sağlığa aracılık etmek için yüksek enerji taleplerinden dolayı mitokondriye spesifik olarak bağımlıdır. Bu işlemlerin bazıları bağımsız olarak nörodejenerasyondaki nöronal kayıp mekanizmasında yer almıştır. Üstelik, giderek artan bir şekilde, uygun nöronal fonksiyon ve sağkalım sağlamak için bu işlemlerin birbirine bağımlı olduğu ve mitokondri içinde etkileşime girdiği kabul edilmektedir.

Introduction

The brain is an organ which consumes up to 10 times more oxygen and 20% of the total glucose than any other parts of the body. In regard to where most of the oxygen is used for mitochondrial respiration and for ATP production, neurons could be considered as the types of cells with high rate of energy production and the one with the highest rate of energy consumption. The majority of ATP in mammalian cells is produced in the processes of glycolysis and oxidative phosphorylation in mitochondria. Most of the mitochondrial reviews on brain pathology introduce these organelles as the “energy plants” of the cells. Clearly, mitochondrial dysfunction directly leads to a decrease in ATP level and consequently to brain cell death. As well, this should be expectable for brain cells that produce more than 90% of their ATP in the process of oxidative phosphorylation [1]. Further to that, the implication of glycolysis in ATP production for neurons has been suggested to be almost none. Despite some controversy regarding the absence of glycolysis in neurons, it is clear that dysfunction in mitochondrial respiration and oxidative phosphorylation should lead to changes in ATP level and, possibly, to energy deprivation and death in neurons.

Mitochondria are organelles with complex functionality that includes number of interdependent processes, loss of function of which can lead to cell death. The major function of mitochondria is ATP production in the F0-F1-ATP synthase. This enzyme uses the electrochemical transmembrane potential (ΔΨm) generated by the electron transport chain, three of four complexes of which are working as a proton pump. All major mitochondrial processes are linked to the membrane potential such as: ATP production, mitochondrial Ca2+ uptake (Ca2+ is entering mitochondria via calcium uniporter via an electrogenic mechanism), reactive oxygen species production (ROS), and the initiation of the process of programmed cell death in response to loss of ΔΨm. Importantly, fast drop in mitochondrial membrane potential caused by the opening of the permeability transition pore (PTP) could be triggered by mitochondrial calcium overload and ROS production [2]. This renders the deregulation of mitochondrial processes to be especially important in the mechanism of neuronal loss in neurodegeneration. Mitochondrial pathology has been associated with a wide range of neurodegenerative diseases. In primary mitochondrial diseases, diseases caused by mutations in mitochondrial DNA or nuclear DNA encoding mitochondrial proteins, perturbation in mitochondrial function alone is sufficient and necessary to trigger neuronal death [3]. It is less clear whether the mitochondrial dysfunction seen in the sporadic late onset neurodegenerative diseases is necessary for pathogenesis or a bystander effect of disease. In this review we explore the different mechanisms by which mitochondrial function leads to progressive neuronal death in different forms of neurodegeneration and study the evidence for its importance in causing neuronal death.

Pathologies induced by ATP deprivation

Changes in mitochondrial bioenergetics due to damage to the respiratory chain may be caused by mitochondrial DNA mutations or by mutations in nuclear genes encoding proteins of the Electron Transport Chain (ETC) of mitochondria. Mutations in the mtDNA encoding complex I leads to Leber’s hereditary optic neuropathy. Mutations in nuclear genes encoding complex I or complex II result in Leigh’s syndrome. Complex V mutations cause ataxia and retinitis pigmentosa. Mutations in tRNA genes result in MELAS (mitochondrial myopathy, encephalopathy, lactic acidosis, and stroke-like episodes), or MERRF (myoclonic epilepsy with ragged red fibers). Mitochondrial diseases manifest themselves with wide ranging neurological symptoms such as loss of vision, deafness, ataxia, seizures, external ophthalmoplegia, and cognitive impairment, but also muscle weakness [4], [5]. However, mutations directly affecting mitochondrial energy production (mostly mtDNA mutations) comprise a relatively small number of neurodegenerative diseases. Moreover, some human neurons with mitochondrial mutations have a higher ΔΨm due to consumption of ATP in F0-F1-ATPase to compensate lack of activity ETC by glycolysis [6]. However, despite the high glycolytic activity, stimulation of the cells with these mutations with higher functional activity induces energy deprivation [7].

Mitochondrial uncoupling (transfer of the protons from one site of membrane to another not through F0-F1-ATP without ATP synthesis) is important process which helps to maintain thermoregulation in mammalian [8]. It acts through specific family of mitochondrial uncoupling proteins (UCP). However, upregulation of UCP or UCP deregulation could lead to a lower ATP level and further to the development of pathology [9], [10]. In the cells with familial forms of neurodegenerative disorders, such as motoneuron disease (VCP) or parkinsonism (HTRA2 deficiency) profound mitochondrial uncoupling lead to the lack of ATP and neuronal loss [11], [12]. Importantly, uncoupling could be induced by abnormal function of the mitochondrial proteins – F0-F1-ATP or ATP/ADP exchanger [11], [12], [13].

Oxidative damage and energy deprivation

Mitochondria produce reactive oxygen species (ROS) as a byproduct of ETC and also in some enzymes of TCA cycle or Monoamine oxidase [14]. It makes this organelle the major candidate for a trigger but also a target of oxidative stress. Oxidation of the proteins, DNA, lipid peroxidation can be a stimulus for cellular dysfunction and induction of cell death by itself [14], [15]. One of the important features of mitochondrial ROS or free radicals from different sources is the fact that they are one of the major triggers for opening of the permeability transition pore (PTP), [16]. Enhanced level of ROS in combination with calcium overload shown to be involved in cell death in Parkinson’s disease [17], [18], Alzheimer’s disease [19], [20], [21], and number of other neurodegenerative disorders [22], [23] (Figure 1).

Figure 1: The role of mitochondria in cellular mechanism of neurodegeneration.Mitochondria interact with various cellular processes and lead to cell death.
Figure 1:

The role of mitochondria in cellular mechanism of neurodegeneration.

Mitochondria interact with various cellular processes and lead to cell death.

However, ROS production and oxidative stress can have direct and indirect effects on mitochondrial metabolism that cause energy deprivation in cells. Thus, oxidation of DNA activates the DNA-repairing enzyme Poly-(ADP-ribose) polymerase (PARP) and the consumption of NAD, respectively. It reduces mitochondrial NADH followed by the decrease in ATP production and as result neuronal loss is triggered (Figure 1). This mechanism was shown for Alzheimer’s disease [24], [25], Parkinson’s disease [26] and excitotoxicity [27], [28].

Oxidative stress in Parkinson’s disease could be the reason for inhibition of the glucose transporter that leads to NADH deprivation and inhibition of mitochondrial respiration and further to ATP deprivation [29], [30], [31]. In familial form of frontotemporal dementia inhibition of the TCA cycle limits the NADH-a substrate availability for the electron transport chain, inhibits mitochondrial respiration, utilization of ATP in glycolysis for the maintenance of ΔΨm and overproduces ROS [23]. It is interesting that in another form of tauopathy-progressive supranuclear palsy (PSP), the mechanism of mitochondrial pathology is similar, but the most important is that ROS overproduction and ATP depletion in mesenchymal stem cells with this pathology lead to a reduced ability of this type of cells for differentiation [32].

It should be noted that production of GSH upon increased oxidative stress in the cell is also occurring at an energy cost (in the form of NADPH) and it could also be a substrate limiting factor for the mitochondrial ATP production and thus induces neurodegeneration [33], [34] (Figure 1).

Cellular quality control

Autophagy is a process whereby cellular components are degraded by merging into the autophagosomes. Autophagosomes fuse with lysosomes, which contain hydrolytic enzymes that break down cellular components. Selective autophagic degradation of mitochondria, termed mitophagy, is necessary for the sustained turnover of mitochondria, the adjustment of mitochondrion numbers to changing metabolic demands, or the removal of damaged mitochondria [35].

Functional mitochondrial alterations, such as loss of mitochondrial potential or permeabilization of the outer mitochondria membrane (OMM), trigger mitophagy, probably in an attempt to limit potential deleterious effects associated with damaged mitochondria such as excessive ROS production or enhanced release of mitochondrial pro-apoptotic factors [36].

Most mitophagy studies in mammalian cells have focused on PTEN-induced putative kinase protein 1 (PINK1)/Parkin-dependent mitochondrial degradation by autophagy. Parkin and PINK1 are encoded by the PARK2 and PARK6 genes, respectively; both are responsible for familial Parkinson’s disease (PD) and have been reported to be associated with mitophagy [37], [38]. PINK1 is expressed in the cytoplasm and constitutively translocates into the mitochondrial inner membrane (IMM) where it is promptly degraded by the mitochondrial inner membrane rhomboid protease presenilin-associated rhomboid-like protein (PARL) [30], [38], [39], [40]. When mitochondria lose their membrane potential, PINK1 can target to the mitochondria, but cannot translocate across the mitochondrial outer membrane; therefore, it accumulates there. Accordingly, only depolarized mitochondria are marked by PINK1 accumulation. Parkin translocates to mitochondria in a PINK1-dependent manner [37], [38], [41], [42]. Further, Parkin triggers the ubiquitination of many mitochondrial proteins such as mitochondrial assembly regulatory factor (MARF) in flies or mitofusin 1, mitofusin 2, and voltage-dependent anion channel 1 (VDAC1) in mammalian cells [43], [44], [45]. In PD, overall autophagic degradation, including mitophagy, seems to be impaired. Indeed, in experimental PD models and post mortem PD brain samples, abnormal mitochondria readily accumulate in the cytosol of affected neurons [46], [47], indicating that it cannot be efficiently degraded through mitophagy. Accumulation of dysfunctional mitochondria may contribute to dopaminergic cell death by an increased production of ROS and an enhanced release of mitochondrial apoptogenic factors [48], [49], [50], [51]. Defective autophagy in PD originates, at least in part [51], from a pathogenic reduction in the amount of functional lysosomes [46], [52], [53]. Lysosomal breakdown in PD appears secondary to the abnormal permeabilization of lysosomal membranes by mitochondrially driven oxidative attack [46], leading to a vicious cycle in which increased ROS production from dysfunctional mitochondria contributes to defective autophagy by oxidatively damaging lysosomal membranes, thereby resulting in a further accumulation of altered mitochondria which cannot be degraded through mitophagy. Supporting a pathogenic role for decreased autophagy/mitophagy in PD, pharmacological restoration of lysosomal-mediated degradation by rapamycin is able to reduce the cytosolic accumulation of undegraded autophagosomes (which contain abnormal mitochondria) and attenuate dopaminergic neurodegeneration in MPTP-treated mice [46], [54]. Besides a general impairment of autophagic degradation, specific defects in mitophagy may also occur in PD.

For instance, PD-linked mutations in PINK1 and Parkin have been shown to disrupt the coordinated normal regulatory role of these molecules at promoting autophagic degradation of dysfunctional mitochondria, thereby leading to defective mitophagy [38], [44], [55].

Parkin has been reported to be recruited to mitochondria, which is followed by a stimulation of mitochondrial autophagy. Relocation of Parkin to mitochondria induced by a collapse of ΔΨm relies on PINK1 expression and that overexpression of WT but not of mutated PINK1 causes Parkin translocation to mitochondria, even in cells with normal ΔΨm. Co-overexpression of Parkin and PINK1 collapses the normal tubular mitochondrial network into mitochondrial aggregates and/or large perinuclear clusters, many of which are surrounded by autophagic vacuoles [55].

It should be noted that some specific pathways of mitophagy and autophagy could be corrected by other ones. Thus, acidification of the cytosol induces autophagy and PINK1-dependent and independent mitophagy which potentially can activate mitophagy in PINK1 or PARKIN form of pathology [56].

However, pathological mutation in p62 – one of the pathways of autophagy lead to motoneuron disease. It induces mitochondrial dysfunction due to substrate deprivation and ATP level decrease. Importantly, it can be restored by the use of Nrf2 activator [57]. Because Nrf2 controls production of mitochondrial substrates, one could assume that it also could be cell protective in PINK1 model of Parkinson’s disease [58], [59], [60]. It should be noted that changes in autophagy or mitophagy are less likely to be the initial trigger of neurodegeneration, but mitochondrial dysfunction of non-removed damaged mitochondria or misfolded proteins could induce pathological cascades which are usually described for neurodegenerative disorders [61].

Mitochondrial Ca2+ and neurodegeneration

Mitochondrial calcium overload is one of the major triggers for cell death in neurodegeneration and many other pathologies [62]. Pathology in mitochondrial transport, and, specifically in enzyme responsible for calcium efflux from mitochondria could lead to neurodegeneration in Parkinson’s disease [29], [63]. These conditions could potentially be corrected by the limitation but not by the complete inhibition of calcium uptake by mitochondria with pharmacological compounds [64]. However, neurons spend ~80–90% of their total ATP for the maintenance of calcium homeostasis. Consider this, any activation of calcium signal in the cells with mitochondrial dysfunction or inhibition of mitochondria with environmental toxins may lead to a cell collapse due to ATP deprivation [7], [65], [66] (Figure 1).

The role of mitochondria is central to the viability and vitality of the organism. Proper functioning of these organelles is essential for the compatibility to life of every cell and any dysfunction of these organelles will result in the reduction in the efficiency of mitochondria to produce ATP. Neurons are heavily dependent on the OXPHOS for the production of their ATP and any disturbance in this process, both ways, would have detrimental effects for the cell, as OXPHOS plays a central role as well in other processes as calcium buffering, ROS production, signaling and programmed cell death. Thus, any unbalances in the process of OXPHOS would render neurons extremely susceptible to the development of a disease condition, e.g. neurodegeneration. Therefore, substances, able to balance out this process hold a high therapeutic potential in the field of brain diseases.

  1. Conflict of interest: Authors have no conflict of interest.

References

1. Bergman O, Ben-Shachar D. Mitochondrial oxidative phosphorylation system (OXPHOS) deficits in schizophrenia: possible interactions with cellular processes. Can J Psychiatry 2016;61:457–69.10.1177/0706743716648290Search in Google Scholar PubMed PubMed Central

2. Bernardi P. Mitochondrial transport of cations: channels, exchangers, and permeability transition. Physiol Rev 1999;79:1127–55.10.1152/physrev.1999.79.4.1127Search in Google Scholar PubMed

3. Wallace DC. Mitochondrial diseases in man and mouse. Science 1999;283:1482–8.10.1126/science.283.5407.1482Search in Google Scholar PubMed

4. Schon EA, Manfredi G. Neuronal degeneration and mitochondrial dysfunction. J Clin Invest 2003;111:303–12.10.1172/JCI200317741Search in Google Scholar

5. Gorman GS, Chinnery PF, DiMauro S, Hirano M, Koga Y, McFarland R, et al. Mitochondrial diseases. Nat Rev Dis Primers 2016;2:16080.10.1093/med/9780199658602.003.0011Search in Google Scholar

6. Abramov AY, Smulders-Srinivasan TK, Kirby DM, Acin-Perez R, Enriquez JA, Lightowlers RN, et al. Mechanism of neurodegeneration of neurons with mitochondrial DNA mutations. Brain 2010;133:797–807.10.1093/brain/awq015Search in Google Scholar PubMed PubMed Central

7. Kovac S, Preza E, Houlden H, Walker MC, Abramov AY. Impaired bioenergetics in mutant mitochondrial DNA determines cell fate during seizure-like activity. Mol Neurobiol 2018;56:321–34.10.1007/s12035-018-1078-9Search in Google Scholar PubMed

8. Himms-Hagen J. Cellular thermogenesis. Annu Rev Physiol 1976;38:315–51.10.1146/annurev.ph.38.030176.001531Search in Google Scholar PubMed

9. Weydt P, Pineda VV, Torrence AE, Libby RT, Satterfield TF, Lazarowski ER, et al. Thermoregulatory and metabolic defects in Huntington’s disease transgenic mice implicate PGC-1alpha in Huntington’s disease neurodegeneration. Cell Metab 2006;4:349–62.10.1016/j.cmet.2006.10.004Search in Google Scholar PubMed

10. Marcuzzo S, Isaia D, Bonanno S, Malacarne C, Cavalcante P, Zacheo A, et al. FM19G11-loaded gold nanoparticles enhance the proliferation and self-renewal of ependymal stem progenitor cells derived from ALS mice. Cells 2019;8:279.10.3390/cells8030279Search in Google Scholar PubMed PubMed Central

11. Plun-Favreau H, Burchell VS, Holmstrom KM, Yao Z, Deas E, Cain K, et al. HtrA2 deficiency causes mitochondrial uncoupling through the F(1)F(0)-ATP synthase and consequent ATP depletion. Cell Death Dis 2012;3:e335.10.1038/cddis.2012.77Search in Google Scholar PubMed PubMed Central

12. Bartolome F, Wu HC, Burchell VS, Preza E, Wray S, Mahoney CJ, et al. Pathogenic VCP mutations induce mitochondrial uncoupling and reduced ATP levels. Neuron 2013;78:57–64.10.1016/j.neuron.2013.02.028Search in Google Scholar PubMed PubMed Central

13. Ludtmann MH, Arber C, Bartolome F, de Vicente M, Preza E, Carro E, et al. Mutations in valosin-containing protein (VCP) decrease ADP/ATP translocation across the mitochondrial membrane and impair energy metabolism in human neurons. J Biol Chem 2017;292:8907–17.10.1074/jbc.M116.762898Search in Google Scholar PubMed PubMed Central

14. Angelova PR, Abramov AY. Functional role of mitochondrial reactive oxygen species in physiology. Free Radic Biol Med 2016;100:81–5.10.1016/j.freeradbiomed.2016.06.005Search in Google Scholar PubMed

15. Abramov AY, Scorziello A, Duchen MR. Three distinct mechanisms generate oxygen free radicals in neurons and contribute to cell death during anoxia and reoxygenation. J Neurosci 2007;27:1129–38.10.1523/JNEUROSCI.4468-06.2007Search in Google Scholar PubMed PubMed Central

16. Bernardi P, Rasola A, Forte M, Lippe G. The mitochondrial permeability transition pore: channel formation by F-ATP synthase, integration in signal transduction, and role in pathophysiology. Physiol Rev 2015;95:1111–55.10.1152/physrev.00001.2015Search in Google Scholar PubMed PubMed Central

17. Gandhi S, Vaarmann A, Yao Z, Duchen MR, Wood NW, Abramov AY. Dopamine induced neurodegeneration in a PINK1 model of Parkinson’s disease. PLoS One 2012;7:e37564.10.1371/journal.pone.0037564Search in Google Scholar PubMed PubMed Central

18. Ludtmann MH, Angelova PR, Horrocks MH, Choi ML, Rodrigues M, Baev AY, et al. α-Synuclein oligomers interact with ATP synthase and open the permeability transition pore in Parkinson’s disease. Nat Commun 2018;9:2293.10.1038/s41467-018-04422-2Search in Google Scholar PubMed PubMed Central

19. Park L, Anrather J, Forster C, Kazama K, Carlson GA, Iadecola C. Abeta-induced vascular oxidative stress and attenuation of functional hyperemia in mouse somatosensory cortex. J Cereb Blood Flow Metab 2004;24:334–42.10.1097/01.WCB.0000105800.49957.1ESearch in Google Scholar PubMed

20. Shevtsova EF, Vinogradova DV, Kireeva EG, Reddy VP, Aliev G, Bachurin SO. Dimebon attenuates the Abeta-induced mitochondrial permeabilization. Curr Alzheimer Res 2014;11:422–9.10.2174/1567205011666140505094808Search in Google Scholar PubMed

21. Abramov AY, Canevari L, Duchen MR. Beta-amyloid peptides induce mitochondrial dysfunction and oxidative stress in astrocytes and death of neurons through activation of NADPH oxidase. J Neurosci 2004;24:565–75.10.1523/JNEUROSCI.4042-03.2004Search in Google Scholar PubMed PubMed Central

22. Cozzolino M, Carri MT. Mitochondrial dysfunction in ALS. Prog Neurobiol 2012;97:54–66.10.1016/j.pneurobio.2011.06.003Search in Google Scholar PubMed

23. Esteras N, Rohrer JD, Hardy J, Wray S, Abramov AY. Mitochondrial hyperpolarization in iPSC-derived neurons from patients of FTDP-17 with 10+16 MAPT mutation leads to oxidative stress and neurodegeneration. Redox Biol 2017;12:410–22.10.1016/j.redox.2017.03.008Search in Google Scholar PubMed PubMed Central

24. Love S, Barber R, Wilcock GK. Increased poly(ADP-ribosyl)ation of nuclear proteins in Alzheimer’s disease. Brain 1999; 122(Pt 2):247–53.10.1093/brain/122.2.247Search in Google Scholar PubMed

25. Abeti R, Abramov AY, Duchen MR. Beta-amyloid activates PARP causing astrocytic metabolic failure and neuronal death. Brain 2011;134:1658–72.10.1093/brain/awr104Search in Google Scholar PubMed

26. Delgado-Camprubi M, Esteras N, Soutar MP, Plun-Favreau H, Abramov AY. Deficiency of Parkinson’s disease-related gene Fbxo7 is associated with impaired mitochondrial metabolism by PARP activation. Cell Death Differ 2016;24:120–31.10.1038/cdd.2016.104Search in Google Scholar PubMed PubMed Central

27. Duan Y, Gross RA, Sheu SS. Ca2+-dependent generation of mitochondrial reactive oxygen species serves as a signal for poly(ADP-ribose) polymerase-1 activation during glutamate excitotoxicity. J Physiol 2007;585:741–58.10.1113/jphysiol.2007.145409Search in Google Scholar PubMed PubMed Central

28. Abramov AY, Duchen MR. Mechanisms underlying the loss of mitochondrial membrane potential in glutamate excitotoxicity. Biochim Biophys Acta 2008;1777:953–64.10.1016/j.bbabio.2008.04.017Search in Google Scholar PubMed

29. Gandhi S, Wood-Kaczmar A, Yao Z, Plun-Favreau H, Deas E, Klupsch K, et al. PINK1-associated Parkinson’s disease is caused by neuronal vulnerability to calcium-induced cell death. Mol Cell 2009;33:627–38.10.1016/j.molcel.2009.02.013Search in Google Scholar PubMed PubMed Central

30. Deas E, Plun-Favreau H, Gandhi S, Desmond H, Kjaer S, Loh SH, et al. PINK1 cleavage at position A103 by the mitochondrial protease PARL. Hum Mol Genet 2010;20:867–79.10.1093/hmg/ddq526Search in Google Scholar PubMed PubMed Central

31. Yao Z, Gandhi S, Burchell VS, Plun-Favreau H, Wood NW, Abramov AY. Cell metabolism affects selective vulnerability in PINK1-associated Parkinson’s disease. J Cell Sci 2011;124:4194–202.10.1242/jcs.088260Search in Google Scholar

32. Angelova PR, Barilani M, Lovejoy C, Dossena M, Vigano M, Seresini A, et al. Mitochondrial dysfunction in Parkinsonian mesenchymal stem cells impairs differentiation. Redox Biol 2018;14:474–84.10.1016/j.redox.2017.10.016Search in Google Scholar

33. Camandola S, Mattson MP. Aberrant subcellular neuronal calcium regulation in aging and Alzheimer’s disease. Biochim Biophys Acta 2011;1813:965–73.10.1016/j.bbamcr.2010.10.005Search in Google Scholar

34. Esteras N, Dinkova-Kostova AT, Abramov AY. Nrf2 activation in the treatment of neurodegenerative diseases: a focus on its role in mitochondrial bioenergetics and function. Biol Chem 2016;397:383–400.10.1515/hsz-2015-0295Search in Google Scholar

35. Youle RJ, Narendra DP. Mechanisms of mitophagy. Nat Rev Mol Cell Biol 2010;12:9–14.10.1038/nrm3028Search in Google Scholar

36. Tait SW, Green DR. Mitochondria and cell death: outer membrane permeabilization and beyond. Nat Rev Mol Cell Biol 2010;11:621–32.10.1038/nrm2952Search in Google Scholar

37. Narendra D, Tanaka A, Suen DF, Youle RJ. Parkin is recruited selectively to impaired mitochondria and promotes their autophagy. J Cell Biol 2008;183:795–803.10.1083/jcb.200809125Search in Google Scholar

38. Narendra DP, Jin SM, Tanaka A, Suen DF, Gautier CA, Shen J, et al. PINK1 is selectively stabilized on impaired mitochondria to activate Parkin. PLoS Biol 2010;8:e1000298.10.1371/journal.pbio.1000298Search in Google Scholar

39. Abramov AY, Zamaraeva MV, Hagelgans AI, Azimov RR, Krasilnikov OV. Influence of plant terpenoids on the permeability of mitochondria and lipid bilayers. Biochim Biophys Acta 2001;1512:98–110.10.1016/S0005-2736(01)00307-8Search in Google Scholar

40. Shi G, Lee JR, Grimes DA, Racacho L, Ye D, Yang H, et al. Functional alteration of PARL contributes to mitochondrial dysregulation in Parkinson’s disease. Hum Mol Genet 2011;20:1966–74.10.1093/hmg/ddr077Search in Google Scholar PubMed

41. Matsuda N, Sato S, Shiba K, Okatsu K, Saisho K, Gautier CA, et al. PINK1 stabilized by mitochondrial depolarization recruits Parkin to damaged mitochondria and activates latent Parkin for mitophagy. J Cell Biol 2010;189:211–21.10.1083/jcb.200910140Search in Google Scholar PubMed PubMed Central

42. Vives-Bauza C, Zhou C, Huang Y, Cui M, de Vries RL, Kim J, et al. PINK1-dependent recruitment of Parkin to mitochondria in mitophagy. Proc Natl Acad Sci USA 2009;107:378–83.10.1073/pnas.0911187107Search in Google Scholar PubMed PubMed Central

43. Gegg ME, Cooper JM, Chau KY, Rojo M, Schapira AH, Taanman JW. Mitofusin 1 and mitofusin 2 are ubiquitinated in a PINK1/parkin-dependent manner upon induction of mitophagy. Hum Mol Genet 2010;19:4861–70.10.1093/hmg/ddq419Search in Google Scholar PubMed PubMed Central

44. Geisler S, Holmstrom KM, Treis A, Skujat D, Weber SS, Fiesel FC, et al. The PINK1/Parkin-mediated mitophagy is compromised by PD-associated mutations. Autophagy 2010;6:871–8.10.4161/auto.6.7.13286Search in Google Scholar PubMed

45. Tanaka A. Parkin-mediated selective mitochondrial autophagy, mitophagy: Parkin purges damaged organelles from the vital mitochondrial network. FEBS Lett 2010;584:1386–92.10.1016/j.febslet.2010.02.060Search in Google Scholar PubMed PubMed Central

46. Dehay B, Bove J, Rodriguez-Muela N, Perier C, Recasens A, Boya P, et al. Pathogenic lysosomal depletion in Parkinson’s disease. J Neurosci 2010;30:12535–44.10.1523/JNEUROSCI.1920-10.2010Search in Google Scholar PubMed PubMed Central

47. Vila M, Bove J, Dehay B, Rodriguez-Muela N, Boya P. Lysosomal membrane permeabilization in Parkinson disease. Autophagy 2011;7:98–100.10.4161/auto.7.1.13933Search in Google Scholar PubMed

48. Vila M, Jackson-Lewis V, Vukosavic S, Djaldetti R, Liberatore G, Offen D, et al. Bax ablation prevents dopaminergic neurodegeneration in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine mouse model of Parkinson’s disease. Proc Natl Acad Sci USA 2001;98:2837–42.10.1073/pnas.051633998Search in Google Scholar PubMed PubMed Central

49. Vila M, Przedborski S. Targeting programmed cell death in neurodegenerative diseases. Nat Rev Neurosci 2003;4:365–75.10.1038/nrn1100Search in Google Scholar PubMed

50. Perier C, Tieu K, Guegan C, Caspersen C, Jackson-Lewis V, Carelli V, et al. Complex I deficiency primes Bax-dependent neuronal apoptosis through mitochondrial oxidative damage. Proc Natl Acad Sci U S A 2005;102:19126–31.10.1073/pnas.0508215102Search in Google Scholar PubMed PubMed Central

51. Perier C, Bove J, Wu DC, Dehay B, Choi DK, Jackson-Lewis V, et al. Two molecular pathways initiate mitochondria-dependent dopaminergic neurodegeneration in experimental Parkinson’s disease. Proc Natl Acad Sci USA 2007;104:8161–6.10.1073/pnas.0609874104Search in Google Scholar PubMed PubMed Central

52. Martinez-Vicente M, Talloczy Z, Kaushik S, Massey AC, Mazzulli J, Mosharov EV, et al. Dopamine-modified alpha-synuclein blocks chaperone-mediated autophagy. J Clin Invest 2008;118:777–88.10.1172/JCI32806Search in Google Scholar

53. Chu Y, Dodiya H, Aebischer P, Olanow CW, Kordower JH. Alterations in lysosomal and proteasomal markers in Parkinson’s disease: relationship to alpha-synuclein inclusions. Neurobiol Dis 2009;35:385–98.10.1016/j.nbd.2009.05.023Search in Google Scholar PubMed

54. Bove J, Martinez-Vicente M, Vila M. Fighting neurodegeneration with rapamycin: mechanistic insights. Nat Rev Neurosci 2011;12:437–52.10.1038/nrn3068Search in Google Scholar PubMed

55. Vives-Bauza C, de Vries RL, Tocilescu M, Przedborski S. PINK1/Parkin direct mitochondria to autophagy. Autophagy 2010;6:315–6.10.4161/auto.6.2.11199Search in Google Scholar PubMed

56. Berezhnov AV, Soutar MP, Fedotova EI, Frolova MS, Plun-Favreau H, Zinchenko VP, et al. Intracellular pH modulates autophagy and mitophagy. J Biol Chem 2016;291:8701–8.10.1074/jbc.M115.691774Search in Google Scholar PubMed PubMed Central

57. Bartolome F, Esteras N, Martin-Requero A, Boutoleau-Bretonniere C, Vercelletto M, Gabelle A, et al. Pathogenic p62/SQSTM1 mutations impair energy metabolism through limitation of mitochondrial substrates. Sci Rep 2017;7:1666.10.1038/s41598-017-01678-4Search in Google Scholar PubMed PubMed Central

58. Holmstrom KM, Baird L, Zhang Y, Hargreaves I, Chalasani A, Land JM, et al. Nrf2 impacts cellular bioenergetics by controlling substrate availability for mitochondrial respiration. Biol Open 2013;2:761–70.10.1242/bio.20134853Search in Google Scholar PubMed PubMed Central

59. Ludtmann MH, Angelova PR, Zhang Y, Abramov AY, Dinkova-Kostova AT. Nrf2 affects the efficiency of mitochondrial fatty acid oxidation. Biochem J 2014;457:415–24.10.1042/BJ20130863Search in Google Scholar PubMed PubMed Central

60. Dinkova-Kostova AT, Kostov RV, Canning P. Keap1, the cysteine-based mammalian intracellular sensor for electrophiles and oxidants. Arch Biochem Biophys 2017;617:84–93.10.1016/j.abb.2016.08.005Search in Google Scholar PubMed PubMed Central

61. Abramov AY, Berezhnov AV, Fedotova EI, Zinchenko VP,Dolgacheva LP. Interaction of misfolded proteins and mitochondria in neurodegenerative disorders. Biochem Soc Trans 2017;45:1025–33.10.1042/BST20170024Search in Google Scholar PubMed

62. Abeti R, Abramov AY. Mitochondrial Ca(2+) in neurodegenerative disorders. Pharmacol Res 2015;99:377–81.10.1016/j.phrs.2015.05.007Search in Google Scholar PubMed

63. Ludtmann MH, Kostic M, Horne A, Gandhi S, Sekler I, Abramov AY. LRRK2 deficiency induced mitochondrial Ca(2+) efflux inhibition can be rescued by Na(+)/Ca(2+)/Li(+) exchanger upregulation. Cell Death Dis 2019;10:265.10.1038/s41419-019-1469-5Search in Google Scholar PubMed PubMed Central

64. Angelova PR, Vinogradova D, Neganova ME, Serkova TP, Sokolov VV, Bachurin SO, et al. Pharmacological sequestration of mitochondrial calcium uptake protects neurons against glutamate excitotoxicity. Mol Neurobiol 2018;56:2244–55.10.1007/s12035-018-1204-8Search in Google Scholar PubMed PubMed Central

65. Domijan AM, Abramov AY. Fumonisin B1 inhibits mitochondrial respiration and deregulates calcium homeostasis–implication to mechanism of cell toxicity. Int J Biochem Cell Biol 2011;43:897–904.10.1016/j.biocel.2011.03.003Search in Google Scholar PubMed

66. Little D, Luft C, Mosaku O, Lorvellec M, Yao Z, Paillusson S, et al. A single cell high content assay detects mitochondrial dysfunction in iPSC-derived neurons with mutations in SNCA. Sci Rep 2018;8:9033.10.1038/s41598-018-27058-0Search in Google Scholar PubMed PubMed Central

Received: 2019-06-08
Accepted: 2019-09-24
Published Online: 2019-10-30
Published in Print: 2019-12-18

©2019 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 23.4.2024 from https://www.degruyter.com/document/doi/10.1515/tjb-2019-0255/html
Scroll to top button