Next Article in Journal
Regeneration of Articular Cartilage in Lizard Knee from Resident Stem/Progenitor Cells
Next Article in Special Issue
The Role of Mitochondrial DNA in Mediating Alveolar Epithelial Cell Apoptosis and Pulmonary Fibrosis
Previous Article in Journal
Biogas Production from Sugarcane Waste: Assessment on Kinetic Challenges for Process Designing
Previous Article in Special Issue
Mitochondrial Transcription Factor A and Mitochondrial Genome as Molecular Targets for Cisplatin-Based Cancer Chemotherapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mitochondria: A Therapeutic Target for Parkinson’s Disease?

1
Department of Neurological Surgery, Case Western Reserve University, Cleveland, OH 44106, USA
2
Department of Physiology, Case Western Reserve University, Cleveland, OH 44106, USA
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2015, 16(9), 20704-20730; https://doi.org/10.3390/ijms160920704
Submission received: 10 June 2015 / Revised: 14 August 2015 / Accepted: 20 August 2015 / Published: 1 September 2015
(This article belongs to the Special Issue Mitochondrial Dysfunction in Ageing and Diseases)

Abstract

:
Parkinson’s disease (PD) is one of the most common neurodegenerative disorders. The exact causes of neuronal damage are unknown, but mounting evidence indicates that mitochondrial-mediated pathways contribute to the underlying mechanisms of dopaminergic neuronal cell death both in PD patients and in PD animal models. Mitochondria are organized in a highly dynamic tubular network that is continuously reshaped by opposing processes of fusion and fission. Defects in either fusion or fission, leading to mitochondrial fragmentation, limit mitochondrial motility, decrease energy production and increase oxidative stress, thereby promoting cell dysfunction and death. Thus, the regulation of mitochondrial dynamics processes, such as fusion, fission and mitophagy, represents important mechanisms controlling neuronal cell fate. In this review, we summarize some of the recent evidence supporting that impairment of mitochondrial dynamics, mitophagy and mitochondrial import occurs in cellular and animal PD models and disruption of these processes is a contributing mechanism to cell death in dopaminergic neurons. We also summarize mitochondria-targeting therapeutics in models of PD, proposing that modulation of mitochondrial impairment might be beneficial for drug development toward treatment of PD.

1. Introduction

Parkinson’s disease (PD) is the second most common neurodegenerative disorder after Alzheimer’s disease, affecting over 1% of the population older than 60 years of age. Clinically, it is diagnosed primarily based on motor abnormalities including bradykinesia, resting tremor, and cogwheel rigidity [1]. A key characteristic of pathology in PD is the degeneration of the nigrostriatal (NS) dopaminergic pathway which is one of the most important dopamine (DA) pathways in the brain and contains about 80% of the total brain DA. Despite a large number of studies on the pathogenesis of PD, there is still inconclusive evidence about why dopaminergic neurons are selectively degenerated. Currently, there is no effective restorative treatment available for PD, only symptomatic treatment is available.
Among a number of proposed mechanisms involved in PD pathogenesis, mitochondrial dysfunction has been repeatedly implicated as the cause of the death of DA neurons in PD [2,3,4,5]. Mitochondria are critical for many cellular functions, such as intermediary metabolism [6,7], redox signaling [8], calcium homeostasis [9,10,11], cell proliferation [12,13], development [14,15] and cell death [16,17,18]. Mitochondrial dysfunction is mainly characterized by the generation of reactive oxygen species (ROS), a defect in mitochondrial electron transport complex enzyme activities, ATP depletion, caspase 3 release and depletion of mitochondrial DNA. In this review, we summarize evidence on the critical involvement of mitochondria in both genetic mutation and environmental toxin-induced PD. We propose a causal role for mitochondrial dysfunction in the development of PD, because (1) neurotoxins causing parkinsonism, such as 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), rotenone, paraquat, induce dopaminergic neuronal death through direct inhibition of mitochondrial complex I activity; (2) mutant proteins from PD-related genes associate with mitochondria where they elicit diverse mitochondrial dysregulation and subsequently cause neuronal degeneration; (3) therapeutic agents that target mitochondrial protein or inhibit mitochondrial damage can reduce neuropathological phenotypes of PD in animal models and cells from PD patients.

2. Mitochondrial Dysfunction in Parkinson’s Disease

Aberrant mitochondrial function is one of the major cytopathologies in PD and has been widely accepted as a central pathogenic mechanism underlying PD pathogenesis. Chronic systemic administration of rotenone, a specific complex I inhibitor and a pesticide, results in neuropathologic and behavioral changes in rats that are similar to human PD [19,20]. 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), a meperidine analog found to cause parkinsonism in humans, exerts its toxic effects through metabolism to 1-methyl-4-phenylpyridinium (MPP+), another complex I inhibitor [19,21]. These compounds have long been used in animal models of PD. Furthermore a number of familial forms of PD are associated with mutations in genes encoding both mitochondrially targeted proteins and proteins involved in mitochondrial function and/or oxidative stress responses, including mutations in PINK-1, DJ-1, parkin, and leucine-rich repeat kinase 2 (LRRK2) [22]. Interestingly, one study reported that the size of mitochondria in dopaminergic neurons in the substantia nigra (SN; susceptible to PD degeneration) is smaller than in neighboring non-dopaminergic neurons or in dopaminergic neurons of the ventral tegmental area which is more resistant in PD, suggesting a basis for the increased vulnerability of SN neurons to subtle changes in mitochondrial maintenance and function [23]. In addition, increased oxidative stress due to mitochondrial compromise in PD model animals has been proposed to contribute to the degeneration of dopaminergic neurons [23].
Consistent with the evidence from basic science, clinical studies also showed that mitochondrial damage plays a predominant role in the development of PD in patients. Mild deficiency in mitochondrial respiratory electron transport chain NADH dehydrogenase (Complex I) activity has been reported in the substantia nigra [24] as well as platelets [25,26] and lymphocytes [27,28] in PD patients, suggesting a systemic inhibition of complex I activity in PD patients. Mitochondrial dysfunction could lead to increased oxidative stress. Indeed oxidative damage to lipids, proteins and DNA has been detected in brain tissue from PD patients [29,30]. A recent proteomic analysis of mitochondria-enriched fractions from post-mortem PD substantia nigra revealed differential expression of multiple mitochondrial proteins in PD brains as compared with control brains [23]. In further support for a “mitochondrial genetics” hypothesis for PD pathophysiology, Bender et al. [31] reported higher levels of mitochondrial DNA deletions in nigral neurons from PD patients. Moreover, both Bender et al. [31] and Kraytsberg et al. [32] reported higher levels of mitochondrial DNA deletions in nigral neurons of aged humans with sharp elevations starting shortly before age 70. This correlates with the known risk factor of age in PD. It is possible that there is an accumulation of mitochondrial dysfunction and of reactive oxygen species (ROS) damage during aging which needs to reach a critical threshold for cellular dysfunction and degeneration to be observed. How a systemic dysregulation in mitochondrial function or oxidative damage leads to tissue or cell type specific vulnerability still remains to be elucidated.
Taken together, although studies over many years on PD indicate an important role of mitochondria in PD-associated pathology, the process by which the mitochondria become dysfunctional in PD and whether correction of mitochondrial defects could provide neuroprotection in PD remain to be determined.

3. Environmental Toxins that Influence Mitochondrial Function

Some toxins used to model DA loss in PD, such as MPTP and rotenone, impair respiratory chain function by inhibiting complex I [33,34,35,36]. These complex I inhibitors replicate some of the key motor features of PD and lead to DA neuronal loss. Intravenous injection of the compound MPTP by drug addicts caused a condition that closely resembles the anatomic and clinical features of PD [37,38]. Multiple models have been developed in the laboratory in which the chronic infusion of the pesticide rotenone or combination of herbicides and pesticides [39] lead to a pattern of cell death and DA loss similar to that of PD. The precise role of environmental toxins in the cause of PD remains to be defined, but these data support the hypothesis that environmental toxins could introduce mitochondrial dysfunction and lead to parkinsonism in human. Indeed, epidemiological studies showed that the prevalence of sporadic PD is higher among farming communities [40]. Exposure to pesticides or herbicides elicited a three-to fourfold increased risk of developing PD [41]. All of these data suggest an environmental contribution to the etiology of sporadic PD.

4. Genetic Factors Associated with PD

In a twin study, Tanner et al. showed that for early onset cases, monozygotic concordance was twice that of dizygotic concordance, suggesting that genetic factors are important in the early onset PD [42]. Many examples of familial parkinsonism have also been reported (See review [43]). Mutations of several genes have been linked to familial PD and parkinsonian syndromes [44,45,46].

4.1. Mitochondrial DNA Mutations and Deletions in PD

Mitochondrial DNA (mtDNA) encodes 13 subunits of respiratory chain proteins, including seven complex I, one complex III, three complex IV, and two complex V subunits. Until now, there has been no PD-associated genetic mutations in mtDNA reported [47]. However, mtDNA deletions have been observed in individual dopaminergic neurons dissected from postmortem human substantia nigra tissue [31]. In addition, mutations in the gene encoding human mtDNA polymerase subunit γ (POLG) leads to clinical parkinsonism associated with multiple mtDNA deletions. [48,49]. Furthermore, proliferator-activated coactivator-1 α (PGC-1 α), one of the key regulators of mitochondrial biogenesis, was found to be decreased in PD patients through a genome-wide association study (GWAS) [50]. Directed deletion of transcriptional factor A (TFAM) in mouse DA neurons using the Cre-LoxP system, termed as MitoPark mice, causes marked deletion of mtDNA, severe impairment of oxidative phosphorylation and slowly progressive motor deficits in the DA system that mimic human parkinsonism as well as altered response to L-3,4-dihydroxyphenylalanine (L-DOPA) treatment [51,52,53]. The MitoPark mice of PD provided direct evidence that mitochondrial dysfunction in DA neurons can causes PD-related phenotypes. Consistently increased level of mtDNA deletions in the striatum of PD patients have been reported [54] and mtDNA deletions were significantly higher in neurons with impaired cytochrome oxidase activity [31,32]. These findings support a mitochondrial genetic contribution in PD.

4.2. Nuclear Gene Mutations Affecting Mitochondrial Function

Many of the PD susceptible genes identified are related to mitochondrial function (Table 1). These genes and their potential contribution to PD have previously been extensively reviewed [43]. PD linked genes that affect mitochondrial function include, but are not limited to, α-synuclein, Parkin, Phosphatase and tensin homologue (PTEN)-induced putative kinase 1 (PINK1), DJ-1 and LRRK2. In this review we will focus on their involvement and effects on dysregulation of mitochondrial dynamics, mitophagy and mitochondrial redox, and mitochondrial protein import.
Alpha-synuclein: α-synuclein is aggregation-prone protein which attains an increased propensity to aggregate because of the presence of its hydrophobic non-amyloid β component domain. Missense mutations in the alpha-synuclein gene are associated with autosomal dominant PD [55]. The fibrillar form of α-synucleinis a major component of Lewy bodies and has been demonstrated to trigger neurotoxicity in PD [56,57].
Parkin: Parkin is a RING finger containing ubiquitin E3 ligase. It is known to mediate poly-ubiquitination of its substrates for proteasomal degradation. Mutations in the parkin gene cause early onset juvenile autosomal recessive PD, and Parkin mutations are the most common cause of young onset PD. The loss of parkin E3 ligase activity results in accumulation of its substrates leading to neurotoxicity in autosomal recessive PD [58,59].
PINK1: Phosphatase and tensin homologue (PTEN)-induced putative kinase 1 (PINK1) is a serine/threonine kinase localized in mitochondria. Mutations in PINK1 are associated with a rare form of autosomal recessive PD. PINK mutations result in the loss of PINK1 function which leads to aberrant phosphorylation of its substrates to cause PD [60,61].
Table 1. Animal models of Parkinson’s disease (PD)-related genes affect mitochondrial function.
Table 1. Animal models of Parkinson’s disease (PD)-related genes affect mitochondrial function.
Animal ModelsGenetic Manipulation in AnimalsMotor PhenotypesPD Pathology and Mitochondrial FunctionReferences
Alpha-Synuclein transgenic micehA53T alpha-Synuclein in mice; mPrP promoter Severe leading to paralysis and premature deathLewy body-like inclusion in older mice; mitochondrial dysfunction; no dopaminergic neuronal loss [62,63,64]
hA30P alpha-synuclein in mice; mThy-1 promoterSevere leading to paralysis Lewy body-like inclusion; sensorimotor neuronal loss in brain stem[65]
Alpha-synuclein overexpression in mice (Thy1 promoter)Progressive declines in spontaneous and motor activityNo DA neuronal degeneration, mitochondrial dysregulation[66,67]
hA53T alpha-synuclein expressing in SN DA neurons of miceBody weight loss; normal locomotion activity Progressive DA neuronal loss; aberrant mitochondrial inclusion[68]
LRRK2 transgenic or knock-in miceLRRK2 R1441G mice; BAC promoterRearing activity decrease in older miceDA neurite degeneration; Tau phosphorylation increase; no DA neuronal degeneration; mitochondrial dysfunction[69,70]
LRRK2 G2019S knock-in miceAbsentAbnormal mitochondrial morphology; mitochondrial dysfunction; no DA neuronal loss[71]
Parkin−/− miceParkin germline inactivationConflicting: either absent or subtle motor movement disturbanceAbsent[72,73]
DJ1−/− miceDJ1 germline inactivationAge-dependent declines in locomotor activityImpaired DA update; no DA neuronal degeneration [74]
PINK1−/− micePINK1 germline inactivationAge-dependent declines in spontaneous activityImpaired DA update; no DA neuronal degeneration; mitochondrial abnormalities[75,76]
Mito-Park miceDAT riven cre; loxed p TFAMBegins at 3–4 months; declines in spontaneous and rearing activityAbnormal mitochondrial aggregates; DA reduction in the striatum; progressive DA neuronal degeneration [51]
Double-mutant miceA53T alpha-synuclein overexpression in Parkin−/− miceAbsentAltered mitochondrial structure and morphology; no DA neuronal loss[77]
DJ-1: DJ-1 belongs to the peptidase C56 family of proteins. Wild-type DJ-1 can serve as a chaperone, protease, regulator of transcription and autophagy through redox regulation. DJ-1 seems to be cytoprotective specifically under conditions related to oxidative stress. Its protective action is the result of the modification of cysteine residues on DJ-1 to cysteine-sulfinic and cysteine-sulfonic acids under oxidative stress [78]. DJ-1 mutations associated with PD are rare and account for 1%–2% of autosomal recessive early-onset PD [79].
LRRK2: Leucine-rich repeat kinase 2 (LRRK2) is a protein encoded by the PARK8 locus. It has a conserved serine-threonine kinase mitogen-activated protein kinase kinase kinase (MAPKKK) domain, a member of the Roc (Ras of complex) GTPase family [80,81,82]. To date, there are over 50 variants identified in PD patients. The mutation G2019S (Gly2019 to Ser) that takes place in the MAPKKK domain has been recognized as the most common cause of dominant familial PD and accounts for up to 2% of sporadic PD cases [83]. The G2019S mutant augments the kinase activity of LRRK2, which is associated with increased toxicity in dopaminergic neurons [84].

4.3. Mitochondrial Dynamics Impairment in Parkinson’s Disease

Mitochondria are organized in a highly dynamic tubular network that is continuously reshaped by opposing processes of fusion and fission [85]. Mitochondrial fission and fusion were first observed in yeast, and since have been observed in all mammalian cells [86]. A delicate balance is maintained between fusion and fission to ensure the normal function of mitochondria. Specifically the fusion process is important for mitochondrial interactions and communication, and fission facilitates the segregation of mitochondria into daughter cells and enhances mitochondrial renewal and distribution along cytoskeletal tracks. Fusion and fission events enable the proper exchanging and mixing of mitochondrial membranes and contents. This dynamic process controls not only mitochondrial morphology, but also the subcellular location and function of mitochondria. Defects in either fusion or fission limit mitochondrial motility, decrease energy production and increase oxidative stress, thereby promoting cell dysfunction and death [87,88]. The two opposing processes, fusion and fission, are controlled by evolutionarily conserved large GTPases that belong to the dynamin family of proteins. In mammalian cells, mitochondrial fusion is regulated by mitofusin-1 and -2 (MFN1/2) and optic atrophy 1 (OPA1), whereas mitochondrial fission is controlled by the dynamin-1-related protein, (Drp1) [89,90] and its mitochondrial adaptors such as Fis1, Mff and MiD49/51 [91,92,93]. Drp1 is primarily found in the cytosol, but it translocates from the cytosol to the mitochondrial surface in response to various cellular stimuli to regulate mitochondrial morphology [1]. At the mitochondrial surface, Drp1 is thought to wrap around the mitochondria to induce fission using its GTPase activity [94].
In terms of PD, no mutations in typical mitochondria fission and fusion genes have yet been identified in PD patients. However, increasing evidence from both toxin models and genetic mutations in PD animal models supports the hypothesis that mitochondrial dynamic regulation and dysfunction are involved in PD. Evidence from toxin-induced PD models support a role for mitochondrial fission/fusion in the pathogenesis of PD. The Parkinsonian neurotoxins, 6-hydroxy dopamine (6-OHDA), rotenone, and MPP+, all induce mitochondrial fragmentation, leading to dopaminergic cell death in neuronal cultures [91,95,96]. Inhibition of pro-fission Drp1 or overexpression of pro-fusion protein mitofusin-1 (Mfn1) using genetic techniques prevents both neurotoxin-induced mitochondrial fission and neuronal cell death [97,98,99]. Loss of Mfn2 or conditional knockout of neuronal Mfn2 in mice has recently been reported to result in age-dependent motor deficits, followed by the loss of dopaminergic terminals in the striatum [100,101], suggesting a role for Mfn2 in parkinsonism. Also, Drp1 is required for synaptic formation [102] and lack of Drp1 leads to an impairment of brain development in mice [103,104]. Recently it has been reported that the Drp1 is critical for targeting mitochondria to the terminal synapses of dopaminergic neurons and deletion of Drp1 gene in dopaminergic neurons rapidly eliminates DA terminals in the caudate-putamen and causes cell bodies in the midbrain to degenerate and lose α-synuclein [105]. Taken together, all these support that molecular machinery which maintains the balance of fusion and fission dynamics in the cells might contribute to the pathogenesis of PD.
In addition, several genes, whose mutated forms are associated with familial PD, affect mitochondrial dynamics: these include PINK1, Parkin, LRRK2 and DJ-1 [45,94,96,106,107,108,109,110]. Among these genes, the role of PINK1/Parkin pathway in regulation of mitochondrial dynamics seems to be opposite from LRRK2 and DJ-1; mutations in PINK1/Parkin lead to mitochondrial fusion [44,111] whereas mutations in LRRK2 or DJ-1 promote mitochondrial fission [44,112]. Fibroblasts from PD patients carrying PINK1 or Parkin mutations exhibited a more fragmented mitochondrial network, showing mitochondrial dysfunction [113,114]. The mitochondrial network could also be reduced by the depletion of Drp1, or overexpression of OPA1 or Mfn2 [115,116,117,118]. Deficiency in DJ-1 in cell lines, cultured neurons and lymphoblasts derived from DJ-1-deficient patients displayed aberrant mitochondrial morphology [119]. Further, in double-mutant mice in which alpha-synuclein mutant is overexpressed and parkin is ablated, severe genotype-, age- and region-dependent mitochondrial morphological alterations were found in neuronal somata. The number of structurally altered mitochondria was significantly increased in the SN of these double-mutants mice [77]. These studies further support the involvement of vulnerable genes in mitochondrial dynamics regulation in PD.
Wild-type LRRK2 interacts and colocalizes with several key regulators of mitochondrial fusion/fission, suggesting that it might have multiple regulatory roles [120]. Furthermore, mutant LRRK2 G2019S, the most common mutation in the population of familial PD patients, has been recently demonstrated to interact with Drp1 and to promote mitochondrial fragmentation, leading to mitochondrial dysfunction and neuronal abnormalities [121,122]. This fragmentation can be reduced by expression of the dominant-negative Drp1K38A or overexpression of the fusion protein Mfn2 [121,122]. iPS cells derived from PD patients carrying the G2019S mutation show excessive mitochondrial fission, aberrant autophagy and neuronal damage in DA neurons differentiated in vitro [123]. More importantly, treatment with P110, a selective peptide inhibitor of Drp1 recently developed in Qi’s group [124,125], reduced mitochondrial fragmentation and damage, and corrected excessive autophagy. In this study it was also shown that G2019S mutated LRRK2 protein primarily phosphorylates Drp1 at T595 resulting in aberrant mitochondrial fragmentation [123].
Together, these findings suggest that impairment of mitochondrial dynamics might contribute to the pathogenesis and progression of PD.

5. Mitophagy/Autophagy Impairment in Parkinson’s Disease

Autophagy is a process of cellular degradation in which cargos are degraded by autophagosomes fused with lysosomes [126]. In addition to non-selective cargos, autophagosomes can degrade protein aggregates or damaged mitochondria. Mitochondria-associated autophagy (mitophagy) is regulated by autophagy receptors that preferentially bind ubiquitylated mitochondria and subsequently recruit the autophagosome protein light chain 3 (LC3) through their LC3-interaction region (LIR) motif [127]. LC3 in mammals, also known as Atg8 in yeast, plays crucial roles in both autophagosome membrane biogenesis and cargo recognition [128,129]. In yeast, Atg32 functions as a receptor on mitochondria to initiate mitophagy through its interaction with Atg8 [130,131]. In mammals, FUNDC1 [132], p62 [133], BNIP3 [134], and AMBRA1 [135], have been recognized as receptors for mitophagy, all of which bind to LC3 via the LIR motif [136]. The implication of these receptors in mitophagy regulation, however, seems to be dependent of experimental conditions. Whether and how these receptors cooperatively regulate mitophagy remains to be determined.
A number of lines of evidence suggest a critical role for defective autophagy/mitophagy in neurodegeneration in PD. Cultured cells exposed to parkinsonian neurotoxins such as MPP+, rotenone, or 6-OHDA showed an increased number of autophagosomes and associated neuronal cell death [125,137,138]. In recent years, PINK1-depenent activation of parkin has been recognized as a major pathway of mitophagy [139]. In mammalian cells, parkin is recruited to depolarized mitochondria, which are subsequently eliminated by autophagy. Such parkin recruitment to mitochondria depends on PINK1 accumulation on mitochondria and therefore PINK1 is a key molecule in the signal transduction of mitophagy [140]. The failure of PINK1/parkin-mediated mitophagic process leads to accumulation of damaged mitochondria, which results in an increase in ROS and cell death [141,142]. Because these studies were all conducted in cells with overexpression of PINK1 and Parkin and because Parkin at endogenous levels fail to mediate mitophagy in PD patient cells [143], the matter of whether these proteins, at the endogenous levels, cooperatively regulate mitophagy remains to be validated. Knock-down of DJ-1, another PD-related gene, resulted in decreased mitochondrial membrane potential, increased reactive oxygen species, excessive mitochondrial fragmentation and impaired autophagy [74,119,144]. Interestingly, overexpression of PINK1 and parkin can rescue mitochondrial fragmentation and dysfunction induced by the depletion of DJ-1 [145,146]. Given that wild-type and mutant DJ-1 can interact with PINK1 [147] and that Parkin, as an E3 ligase on mitochondria, catalyzes the ubiquitination of DJ-1 [148], PINK1, Parkin and DJ-1 may be operative in the pathway of PINK1/parkin-mediated mitophagy.
Pathogenic LRRK2-mediated autophagy has been observed in a variety of cell cultures [149,150], in neurons derived from patient-induced pluripotent stem cells [151] and in animal models in which a mutant form of LRRK2 is expressed [152]. Either knock-down of LRRK2 by siRNA or treatment with a LRRK2 kinase inhibitor caused an increase in autophagic fluxin-cultured cells [153,154]. Overexpression of LRRK2, especially mutant forms, seems to suppress autophagy [155]. In contrast, studies from other two groups show that mutant forms of LRRK2 may induce autophagy via an ERK1/2-dependent pathway [149,156]. Although the detailed mechanisms by which LRRK2 and its mutants mediate autophagy are not clear, LRRK2 may disrupt the balance of autophagy through damaging lysosome-related calcium storage and cargo degradation [157,158]. Further, our group recently reported excessive mitophagy in a variety of cells expressing the LRRK2 G2019S mutant, which was accompanied by mitochondrial depolarization, recruitment of p62 to the mitochondria, increased LC3II levels and lysosomal activity as well as death of dopaminergic neurons [123,159]. We showed that the LRRK2 G2019S mutant caused excessive mitophagyby phosphorylating its substrate including fission protein Drp1 and mitochondrial outer membrane protein Bcl2 [123,159]. Thus, the pathway that LRRK2 mediation of mitophagy might be different from those occurred during autophagy.
Taken together, a large body of studies indicates that the different PD-related genes contribute to the pathogenesis of PD at the intersection of mitochondrial dysfunction and autophagy. The loss of mitochondrial membrane potential, which is associated with mitochondrial dysfunction, seems to be a common signal for mitochondria to be degraded via mitophagy. Moreover, occurrence of mitochondrial fragmentation due to impairment of mitochondrial fusion and fission often precedes the induction of mitophagy. Thus, it is possible that mitophagy may be a multistep process starting with the degradation of profusion/fission proteins, resulting in an imbalance of mitochondrial dynamics and the subsequent clearance of mitochondria [86]. However, the factors that regulate these processes involved in mitochondrial “quality control”, including mitochondrial dynamics, mitochondria-associated degradation and mitophagy in PD, remain to be determined.

6. Mitochondrial Redox Signaling in Parkinson’s Disease

Decreased Complex I activity in the SN of PD patients and animal models has been repeatedly observed. The defect in complex I results in impairment of electron transport and causes ROS accumulation in mitochondria which lead to neuronal degeneration. Neurotoxins causing parkinsonism, MPP+ are selectively taken up into dopaminergic neurons in which it inhibits Complex I activity [160]. Rotenone also inhibits Complex I by impairing oxidative phosphorylation [161]. These studies demonstrate a contribution from ROS to the pathogenesis of dopamine neuronal loss in PD. The genetic PD-linked proteins play a significant role in this process. Alpha-synuclein mutant A53T can enter mitochondria where it binds to the Complex I subunit to inhibit Complex I activity, producing ROS [162,163]. Functional studies showed that alpha-synuclein associated with mitochondria induces cytochrome c release, increased calcium and ROS levels resulting in dopaminergic neuronal death [164]. In addition, expression of mutants of PD-related genes Parkin, PINK1, DJ-1 and LRRK2 in cultured cells all increased ROS. This evidence has been well-summarized in other reviews [29,165]. However, it is uncertain if such elevated ROS are directly caused by these mutants or through indirect cellular effects.
Besides ROS, reactive nitrogen species (RNS) mediating nitrosative stress is also implicated in SN neuronal loss in PD [166]. RNS are generated by the reaction of superoxide with nitric oxide (NO), which results in the production of peroxynitrite. NO inhibits several enzymes including complexes I and IV of the mitochondrial electron transport chain, which in turn lead to ROS generation [167,168]. Increased expression of iNOS and nNOS were observed in basal ganglia of postmortem brain of PD patients [169]. In the mouse MPTP model, there was a significant upregulation of iNOS associated with the gliosis in the SN [170]. Inhibition of nNOS has been reported to protect against neurotoxicity in MPTP-induced PD animal model [171,172]. These observations suggest that NO and its metabolite peroxynitriteare implicated in the pathogenesis of PD.

7. Mitochondrial Protein Import in Parkinson’s Disease

Mitochondria possess their own DNA and translational machinery; however, there are only a small number of mitochondrial proteins encoded by mtDNA that are synthesized within the organelle. The majority of mitochondrial proteins are nuclear-encoded and have to be imported into the organelle. The translocase of the outer mitochondrial membrane (TOM complex) plays central roles in controlling protein entry to mitochondria [173]. The TOM complex is the main entry portal for most mitochondrial proteins that are synthesized in cytoplasm. The TOM complex contains seven subunits including TOM40, TOM22, two proprotein receptors, TOM20, and TOM70, and three smaller proteins, TOM5, TOM6 and TOM7. In general, imported proteins bind to one of these receptors. With the assistance of TOM22 and TOM5, the imported proteins pass the TOM 40 channel [174]. TOM complexes also need to cooperate with TIM (Translocase of the Inner Membrane) 23 complexes to import matrix-targeted proteins [175]. Thus, the TOM and the TIM23 complex direct the translocation of oxidative phosphorylation and metabolite transporter proteins to the inner membrane.
A number of studies have reported that PD-associated genes, alpha-synuclein, PINK1 and Parkin can interact with the TOM complex, disrupting the mitochondrial protein import. Alpha-synuclein can enter mitochondria where it mainly localizes at the inner membrane. The import of alpha-synuclein to mitochondria is through the TOM complex [176,177]. In postmortem brain samples of PD and in brain tissues of mice overexpressing alpha-synuclein, TOM 40 levels changed together with the levels of alpha-synuclein [177]. Expression with either wild-type alpha-synuclein or mutant A53T in cultured cells resulted in the loss of TOM40, whereas over-expression of TOM40 prevented the cellular damage caused by expression of either alpha-synuclein wild-type or its mutant A53T [177]. Thus, alpha-synuclein mutants may impair mitochondrial function by suppression of TOM40-dependent mitochondrial protein import pathways. PINK1 is also imported into the mitochondria via the TOM complex channel [178,179] and then is degraded in a membrane potential-dependent manner. Under mitochondrial depolarization PINK1, interacting with the TOM complex, recruits and activates Parkin, leading to degradation of mitochondrial outer membrane proteins and ultimately mitophagy [178]. Thus, pathogenic mutations in PINK1 and parkin may disrupt this pathway, resulting in the accumulation of dysfunctional mitochondria.

8. Potential Therapeutics Targeting Mitochondria for Treatment of PD

Given the importance of mitochondrial dysfunction in the pathogenesis of PD, therapeutics targeting mitochondria have been studied to prevent or treat PD. Although the cellular and pathological phenotypes from neurotoxin-induced and genetic mutant-associated PD models are different, the outcomes of mitochondrial dysfunction, including mitochondrial dynamic impairments, increased ROS, and impaired bioenergetics, seem to be common pathways. Thus, the mechanistic information on mitochondrial dysfunction in PD models provides potential targets for the development of therapeutic approaches for treatment of PD.
Here, we summarize reported therapeutic agents that reduce PD pathology in models of PD (see Table 2). We categorize these agents into (1) modulating PD-related genetic mutants; (2) modulating mitochondrial proteins; and (3) modulating the consequences of mitochondrial dysfunction. In Table 2, we list the neuroprotective effects of these agents especially in either animal models of PD or neuron derived patient-iPS cells.
Modulating PD-related genetic mutants: LRRK2 mutants are the most common genetic mutants in both familial and sporadic PD. Modulation of the LRRK2 kinase domain has been attractive for the development of therapeutics of PD. LRRK2 inhibitors, PF-06447475 and GW5074, have been shown to increase dopaminergic neuronal survival in primary neuronal cultures and neurons- derived from patient iPS cells [180,181]. Increasing glyoxalase activity of DJ-1 by supplying D-lactate and glycolate rescues the requirement for DJ-1 in maintenance of mitochondrial potential, increases cytocatalytic rate of DJ-1, and reduces neuronal death caused by paraquat and down-regulation of PINK1 [182]. However, thus far, these reagents are only tested in cultured cells so far. Whether treatment with these pharmacological agents protect neurons from PD in animal models remains to be determined.
Table 2. Therapeutic agents that target mitochondria for treatment of Parkinson’s disease
Table 2. Therapeutic agents that target mitochondria for treatment of Parkinson’s disease
CategoryAgentMolecular ActionPD ModeTherapeutic Effects
Modulating PD-related genesPF-06447475LRRK2 kinase inhibitorTransgenic rat with LRRK2 G2019SReduce behavioral and neuropathological phenotypes
Reduce inflammation [183]
GW5074LRRK2 kinase inhibitorDA neurons derived from PD patient iPS cellsSuppression of ROS
Improve mitochondrial respiration
Increase DA neuronal survival [180]
FX2149LRRK2 GTPase inhibitormouse inflammation modelReduce neuroinflammation
Inhibit microglial activity [184]
Modulating mitochondrial proteinAlda-1ALDH2 activatorRotenone- and MPTP-induced animal modelsImprove mitochondrial membrane potential
Inhibit mitochondrial ROS
Reduce dopaminergic cell death [185]
TRO40303inhibitor of mitochondrial transition poreMice expressing alha-synucleinUpregulate mitochondrial proteins
Increase TH expression [186]
P110Peptide inhibitor of Drp1DA neurons from LRRK2 G2019S PD patient iPS cellsImprove mitochondrial membrane potential
Inhibit mitochondrial ROS
Increase mitochondrial integrity
Reduce autophagy
Improve DA neuronal morphology and survival [123]
Mdivi-1Inhibitor of mitochondrial fragmentationMPTP-induced mouse PD modelImprove mitochondrial morphology
Improve mouse behavioral outcome
PINK1−/− mouse modelReduce DA neuronal loss in SN
Restore dopamine level [187]
Modulating mitochondrial dysfunctionQ18-OH-quinoline-based iron chelatorMPTP-induced mouse PD modelReduce DA neuronal degeneration in SN
Decrease mitochondrial iron pool [188]
RapamycinmTOR inhibitor6-OHDA-induced rat PD modelInhibit oxidative stress
Inhibit mitochondrial apoptosis [189]
EdaravoneROS scavengerRotenone-induced rat PD modelInhibit mitochondrial apoptosis
Reduce ROS [190]
MelatoninAntioxidantRotenone-induced rat PD modelSuppress calcium level
6-OHDA-induced rat PD modelInhibit mitochondrial ROS
Enhance complex I activity [191,192]
QuercetinBioflavonoidRotenone-induced rat PD modelInhibit mitochondrial ROS generation
Inhibit p53 level
Inhibit nuclear translocation of NF-kappaB
Inhibit mitochondrial apoptosis [193]
CNB-001Curcumin derivativeMPTP-induced mouse PD modelImprove mitochondrial morphology
Inhibit mitochondrial apoptotic pathway
Improve mitochondrial membrane potential [194]
Alpha-Lipoic acidAntioxidant Rotenone-induced rat PD modelIncrease mitochondrial complex I activity
Inhibit ROS generation
Increase mitochondrial biogenesis
Increase glutathione [195]
LycopeneChemical carotene Rotenone-induced rat PD modelInhibit mitochondrial apoptotic pathway
Increase SOD activity
Increase glutathione
Inhibit lipid peroxidation [196]
Modulating mitochondrial proteins: Aldehyde dehydrogenase 2 (ALDH2), located in mitochondrial matrix, functions as a cellular protector against oxidative stress by detoxification of cytotoxic aldehydes. Alda-1 is a small molecule that enhances ALDH enzyme activity and protects against oxidative toxicity [197]. Treatment with Alda-1 can reduce rotenone-induced apoptosis in both SH-SY5Y cells and primary dopaminergic neurons. Moreover, intraperitoneal administration of Alda-1 can improve mitochondrial membrane potential, inhibit mitochondrial ROS and reduce death of tyrosine hydroxylase (TH)-positive dopaminergic neurons in rotenone- or MPTP-induced PD animal models [185]. Cholesterol oximes such as olesoxime and TRO40303 are small molecules that interact with the mitochondrial outer membrane protein VDAC and limit opening of the mitochondrial transition pore in response to oxidative stress [198]. Olesoxime can protect differentiated SHSY-5Y cells from cell death, and reduce neurite retraction and cytoplasmic shrinkage induced by alpha-synuclein overexpression [199]. Low dose TRO40303 upregulates a number of mitochondrial proteins including Drp1 and VDAC and enhances expression of tyrosine hydroxylase in mice overexpressing alpha-synuclein [186]. As mentioned above, our group has developed a peptide inhibitor P110 that selectively blocks the protein-protein interactions between Drp1 and its mitochondrial adaptor Fis1 [125]. Treatment with P110 significantly reduced mitochondrial fragmentation, decreased mitochondrial ROS and improved mitochondrial integrity in dopaminergic neurons exposed to MPP+ in neurons expressing the LRRK2 G2019S mutation, and in dopaminergic neurons derived from LRRK2 G2019S patient-iPS cells [123,125]. Importantly, we showed that treatment with P110 had minor effects on mitochondrial dynamics and neuronal survival under physiological conditions [123,125]. In addition, Mdivi-1, an inhibitor of mitochondrial fragmentation, was reported to reduce behavioral and neuropathological phenotypes in an MPTP-induced PD mouse model, in addition to its protective effects in dopaminergic neurons exposed to neurotoxins [187].
Modulating the consequences of mitochondrial dysfunction: Modulation of downstream mitochondrial dysfunction may also provide therapeutic opportunities, in both sporadic and familial PD. In the past decades, a number of natural products and small molecules have been reported to protect against neuropathology associated with PD, at least in part through protecting mitochondria. Treatment with these agents has been shown to reduce mitochondrial ROS, inhibit mitochondrial apoptotic pathways, and increase mitochondrial complex I activity. As a consequence, treatment with these agents can reduce neuronal degeneration in PD in culture and in animals. These agents have been extensively reviewed [200,201,202,203]. Here, we only list those that have significant protection in animal models of PD (Table 2).

9. Concluding Remarks

Accumulating evidence supports the hypothesis that mitochondrial abnormalities and dysfunction could critically influence neuronal degeneration in both sporadic and faimilial PD. Dysregulation of mitochondrial dynamics and mitophagy have been centrally implicated in the neuropathology of PD. This evidence supports the idea that mitochondrial damage might be a primary cause initiating the progression of PD. Thus, targeting mitochondria may offer the opportunities for drug development to treat neurodegenerative diseases such as parkinsonism.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Duvoisin, R.C. Overview of Parkinson’s disease. Ann. N. Y. Acad. Sci. 1992, 648, 187–193. [Google Scholar] [CrossRef] [PubMed]
  2. Dauer, W.; Przedborski, S. Parkinson’s disease: Mechanisms and models. Neuron 2003, 39, 889–909. [Google Scholar] [CrossRef]
  3. Dawson, T.M.; Dawson, V.L. Molecular pathways of neurodegeneration in Parkinson’s disease. Science 2003, 302, 819–822. [Google Scholar] [CrossRef] [PubMed]
  4. Yao, Z.; Wood, N.W. Cell death pathways in Parkinson’s disease: Role of mitochondria. Antioxid. Redox Signal. 2009, 11, 2135–2149. [Google Scholar] [CrossRef] [PubMed]
  5. Ellis, C.E.; Murphy, E.J.; Mitchell, D.C.; Golovko, M.Y.; Scaglia, F.; Barcelo-Coblijn, G.C.; Nussbaum, R.L. Mitochondrial lipid abnormality and electron transport chain impairment in mice lacking alpha-synuclein. Mol. Cell. Biol. 2005, 25, 10190–10201. [Google Scholar] [CrossRef] [PubMed]
  6. Herrmann, J.M.; Longen, S.; Weckbecker, D.; Depuydt, M. Biogenesis of mitochondrial proteins. Adv. Exp. Med. Biol. 2012, 748, 41–64. [Google Scholar] [PubMed]
  7. Sas, K.; Robotka, H.; Toldi, J.; Vecsei, L. Mitochondria, metabolic disturbances, oxidative stress and the kynurenine system, with focus on neurodegenerative disorders. J. Neurol. Sci. 2007, 257, 221–239. [Google Scholar] [CrossRef] [PubMed]
  8. Daiber, A. Redox signaling (cross-talk) from and to mitochondria involves mitochondrial pores and reactive oxygen species. Biochim. Biophys. Acta 2010, 1797, 897–906. [Google Scholar] [CrossRef] [PubMed]
  9. Bononi, A.; Missiroli, S.; Poletti, F.; Suski, J.M.; Agnoletto, C.; Bonora, M.; de Marchi, E.; Giorgi, C.; Marchi, S.; Patergnani, S.; et al. Mitochondria-associated membranes (MAMs) as hotspot Ca2+ signaling units. Adv. Exp. Med. Biol. 2012, 740, 411–437. [Google Scholar] [PubMed]
  10. Rizzuto, R.; de Stefani, D.; Raffaello, A.; Mammucari, C. Mitochondria as sensors and regulators of calcium signalling. Nat. Rev. Mol. Cell Biol. 2012, 13, 566–578. [Google Scholar] [CrossRef] [PubMed]
  11. Cali, T.; Ottolini, D.; Brini, M. Mitochondrial Ca2+ as a key regulator of mitochondrial activities. Adv. Exp. Med. Biol. 2012, 942, 53–73. [Google Scholar] [PubMed]
  12. Osteryoung, K.W.; Nunnari, J. The division of endosymbiotic organelles. Science 2003, 302, 1698–704. [Google Scholar] [CrossRef] [PubMed]
  13. Antico Arciuch, V.G.; Elguero, M.E.; Poderoso, J.J.; Carreras, M.C. Mitochondrial regulation of cell cycle and proliferation. Antioxid. Redox Signal. 2012, 16, 1150–1180. [Google Scholar] [CrossRef] [PubMed]
  14. Moyes, C.D.; Hood, D.A. Origins and consequences of mitochondrial variation in vertebrate muscle. Annu. Rev. Physiol. 2003, 65, 177–201. [Google Scholar] [CrossRef] [PubMed]
  15. Schatten, H.; Prather, R.S.; Sun, Q.Y. The significance of mitochondria for embryo development in cloned farm animals. Mitochondrion 2005, 5, 303–321. [Google Scholar] [CrossRef] [PubMed]
  16. Giorgi, C.; Baldassari, F.; Bononi, A.; Bonora, M.; de Marchi, E.; Marchi, S.; Missiroli, S.; Patergnani, S.; Rimessi, A.; Suski, J.M.; et al. Mitochondrial Ca2+ and apoptosis. Cell Calcium 2012, 52, 36–43. [Google Scholar] [CrossRef] [PubMed]
  17. Rasola, A.; Bernardi, P. Mitochondrial permeability transition in Ca2+-dependent apoptosis and necrosis. Cell Calcium 2011, 50, 222–233. [Google Scholar] [CrossRef] [PubMed]
  18. Wang, C.; Youle, R.J. The role of mitochondria in apoptosis*. Annu. Rev. Genet. 2009, 43, 95–118. [Google Scholar] [CrossRef] [PubMed]
  19. Burbach, J.P.; Smits, S.; Smidt, M.P. Transcription factors in the development of midbrain dopamine neurons. Ann. N. Y. Acad. Sci. 2003, 991, 61–68. [Google Scholar] [CrossRef] [PubMed]
  20. Wallen, A.; Perlmann, T. Transcriptional control of dopamine neuron development. Ann. N. Y. Acad. Sci. 2003, 991, 48–60. [Google Scholar] [CrossRef] [PubMed]
  21. Dreyer, S.D.; Zhou, G.; Baldini, A.; Winterpacht, A.; Zabel, B.; Cole, W.; Johnson, R.L.; Lee, B. Mutations in LMX1B cause abnormal skeletal patterning and renal dysplasia in nail patella syndrome. Nat. Genet. 1998, 19, 47–50. [Google Scholar] [CrossRef] [PubMed]
  22. Smidt, M.P.; van Schaick, H.S.; Lanctot, C.; Tremblay, J.J.; Cox, J.J.; van der Kleij, A.A.; Wolterink, G.; Drouin, J.; Burbach, J.P. A homeodomain gene Ptx3 has highly restricted brain expression in mesencephalic dopaminergic neurons. Proc. Natl Acad. Sci. USA 1997, 94, 13305–13310. [Google Scholar] [CrossRef] [PubMed]
  23. Smidt, M.P.; Asbreuk, C.H.; Cox, J.J.; Chen, H.; Johnson, R.L.; Burbach, J.P. A second independent pathway for development of mesencephalic dopaminergic neurons requires Lmx1b. Nat. Neurosci. 2000, 3, 337–341. [Google Scholar] [PubMed]
  24. Mann, V.M.; Cooper, J.M.; Daniel, S.E.; Srai, K.; Jenner, P.; Marsden, C.D.; Schapira, A.H. Complex I, iron, and ferritin in Parkinson’s disease substantia nigra. Ann. Neurol. 1994, 36, 876–881. [Google Scholar] [CrossRef] [PubMed]
  25. Blandini, F.; Nappi, G.; Greenamyre, J.T. Quantitative study of mitochondrial complex I in platelets of parkinsonian patients. Mov. Disord. 1998, 13, 11–15. [Google Scholar] [CrossRef] [PubMed]
  26. Haas, R.H.; Nasirian, F.; Nakano, K.; Ward, D.; Pay, M.; Hill, R.; Shults, C.W. Low platelet mitochondrial complex I and complex II/III activity in early untreated Parkinson’s disease. Ann. Neurol. 1995, 37, 714–722. [Google Scholar] [CrossRef] [PubMed]
  27. Barroso, N.; Campos, Y.; Huertas, R.; Esteban, J.; Molina, J.A.; Alonso, A.; Gutierrez-Rivas, E.; Arenas, J. Respiratory chain enzyme activities in lymphocytes from untreated patients with Parkinson disease. Clin. Chem. 1993, 39, 667–69. [Google Scholar] [PubMed]
  28. Yoshino, H.; Nakagawa-Hattori, Y.; Kondo, T.; Mizuno, Y. Mitochondrial complex I and II activities of lymphocytes and platelets in Parkinson’s disease. J. Neural Transm Parkinsons Dis. Dement. Sect. 1992, 4, 27–34. [Google Scholar] [CrossRef]
  29. Dias, V.; Junn, E.; Mouradian, M.M. The role of oxidative stress in Parkinson’s disease. J. Parkinsons Dis. 2013, 3, 461–491. [Google Scholar] [PubMed]
  30. Hwang, O. Role of oxidative stress in Parkinson’s disease. Exp. Neurobiol. 2013, 22, 11–17. [Google Scholar] [CrossRef] [PubMed]
  31. Bender, A.; Krishnan, K.J.; Morris, C.M.; Taylor, G.A.; Reeve, A.K.; Perry, R.H.; Jaros, E.; Hersheson, J.S.; Betts, J.; Klopstock, T.; et al. High levels of mitochondrial DNA deletions in substantia nigra neurons in aging and Parkinson disease. Nat. Genet. 2006, 38, 515–517. [Google Scholar] [CrossRef] [PubMed]
  32. Kraytsberg, Y.; Kudryavtseva, E.; McKee, A.C.; Geula, C.; Kowall, N.W.; Khrapko, K. Mitochondrial DNA deletions are abundant and cause functional impairment in aged human substantia nigra neurons. Nat. Genet. 2006, 38, 518–520. [Google Scholar] [CrossRef] [PubMed]
  33. Langston, J.W.; Ballard, P.; Tetrud, J.W.; Irwin, I. Chronic Parkinsonism in humans due to a product of meperidine-analog synthesis. Science 1983, 219, 979–980. [Google Scholar] [CrossRef] [PubMed]
  34. Betarbet, R.; Sherer, T.B.; MacKenzie, G.; Garcia-Osuna, M.; Panov, A.V.; Greenamyre, J.T. Chronic systemic pesticide exposure reproduces features of Parkinson’s disease. Nat. Neurosci. 2000, 3, 1301–1306. [Google Scholar] [PubMed]
  35. Testa, C.M.; Sherer, T.B.; Greenamyre, J.T. Rotenone induces oxidative stress and dopaminergic neuron damage in organotypic substantia nigra cultures. Brain Res. Mol. Brain Res. 2005, 134, 109–118. [Google Scholar] [CrossRef] [PubMed]
  36. Smeyne, R.J.; Jackson-Lewis, V. The MPTP model of Parkinson’s disease. Brain Res. Mol. Brain Res. 2005, 134, 57–66. [Google Scholar] [CrossRef] [PubMed]
  37. Langston, J.W.; Ballard, P.A., Jr. Parkinson’s disease in a chemist working with 1-methyl-4-phenyl-1,2,5,6-tetrahydropyridine. N. Engl. J. Med. 1983, 309, 310. [Google Scholar] [PubMed]
  38. Kopin, I.J. Toxins and Parkinson’s disease: MPTP parkinsonism in humans and animals. Adv. Neurol. 1987, 45, 137–144. [Google Scholar] [PubMed]
  39. Barbeau, A. Parkinson’s disease: Clinical features and etiopahthology. In Amsterdam Handbook of Clinical Neurology; Viken, P.J., Bruyn, G.W., Klawans, H.L., Eds.; Elsevier Science Publishers: Amsterdam, The Netherlands, 1986; pp. 87–108. [Google Scholar]
  40. Jenner, P.; Olanow, C.W. Understanding cell death in Parkinson’s disease. Ann. Neurol. 1998, 44 (Suppl. 1), S72–S84. [Google Scholar] [CrossRef] [PubMed]
  41. Nunes, I.; Tovmasian, L.T.; Silva, R.M.; Burke, R.E.; Goff, S.P. Pitx3 is required for development of substantia nigra dopaminergic neurons. Proc. Natl. Acad. Sci. USA 2003, 100, 4245–4250. [Google Scholar] [CrossRef] [PubMed]
  42. Van den Munckhof, P.; Luk, K.C.; Ste-Marie, L.; Montgomery, J.; Blanchet, P.J.; Sadikot, A.F.; Drouin, J. Pitx3 is required for motor activity and for survival of a subset of midbrain dopaminergic neurons. Development 2003, 130, 2535–2542. [Google Scholar] [CrossRef] [PubMed]
  43. Klein, C.; Westenberger, A. Genetics of Parkinson’s disease. Cold Spring Harb. Perspect. Med. 2012, 2, a008888. [Google Scholar] [CrossRef] [PubMed]
  44. Zetterstrom, R.H.; Solomin, L.; Jansson, L.; Hoffer, B.J.; Olson, L.; Perlmann, T. Dopamine neuron agenesis in Nurr1-deficient mice. Science 1997, 276, 248–250. [Google Scholar] [CrossRef] [PubMed]
  45. Wallen, A.; Zetterstrom, R.H.; Solomin, L.; Arvidsson, M.; Olson, L.; Perlmann, T. Fate of mesencephalic AHD2-expressing dopamine progenitor cells in NURR1 mutant mice. Exp. Cell Res. 1999, 253, 737–46. [Google Scholar] [CrossRef] [PubMed]
  46. Smidt, M.P.; Smits, S.M.; Burbach, J.P. Molecular mechanisms underlying midbrain dopamine neuron development and function. Eur. J. Pharmacol. 2003, 480, 75–88. [Google Scholar] [CrossRef] [PubMed]
  47. Franco-Iborra, S.; Vila, M.; Perier, C. The Parkinson Disease Mitochondrial Hypothesis: Where Are We at? Neuroscientist 2015. [Google Scholar] [CrossRef] [PubMed]
  48. Luoma, P.; Melberg, A.; Rinne, J.O.; Kaukonen, J.A.; Nupponen, N.N.; Chalmers, R.M.; Oldfors, A.; Rautakorpi, I.; Peltonen, L.; Majamaa, K.; et al. Parkinsonism, premature menopause, and mitochondrial DNA polymerase gamma mutations: Clinical and molecular genetic study. Lancet 2004, 364, 875–882. [Google Scholar] [CrossRef]
  49. Henchcliffe, C.; Beal, M.F. Mitochondrial biology and oxidative stress in Parkinson disease pathogenesis. Nat. Clin. Pract. Neurol. 2008, 4, 600–609. [Google Scholar] [CrossRef] [PubMed]
  50. Zheng, B.; Liao, Z.; Locascio, J.J.; Lesniak, K.A.; Roderick, S.S.; Watt, M.L.; Eklund, A.C.; Zhang-James, Y.; Kim, P.D.; Hauser, M.A.; et al. PGC-1α, a potential therapeutic target for early intervention in Parkinson’s disease. Sci. Transl Med. 2010, 2, 52ra73. [Google Scholar] [PubMed]
  51. Ekstrand, M.I.; Terzioglu, M.; Galter, D.; Zhu, S.; Hofstetter, C.; Lindqvist, E.; Thams, S.; Bergstrand, A.; Hansson, F.S.; Trifunovic, A.; et al. Progressive parkinsonism in mice with respiratory-chain-deficient dopamine neurons. Proc. Natl. Acad. Sci. USA 2007, 104, 1325–1330. [Google Scholar] [CrossRef] [PubMed]
  52. Ekstrand, M.I.; Galter, D. The MitoPark Mouse—An animal model of Parkinson’s disease with impaired respiratory chain function in dopamine neurons. Parkinsonism Relat. Disord. 2009, 15 (Suppl. 3), S185–S188. [Google Scholar] [CrossRef]
  53. Gellhaar, S.; Marcellino, D.; Abrams, M.B.; Galter, D. Chronic L-DOPA induces hyperactivity, normalization of gait and dyskinetic behavior in MitoPark mice. Genes Brain Behav. 2015, 14, 260–270. [Google Scholar] [CrossRef] [PubMed]
  54. Ikebe, S.; Tanaka, M.; Ohno, K.; Sato, W.; Hattori, K.; Kondo, T.; Mizuno, Y.; Ozawa, T. Increase of deleted mitochondrial DNA in the striatum in Parkinson’s disease and senescence. Biochem. Biophys. Res. Commun. 1990, 170, 1044–1048. [Google Scholar] [CrossRef]
  55. Lee, V.M.; Trojanowski, J.Q. Mechanisms of Parkinson’s disease linked to pathological alpha-synuclein: New targets for drug discovery. Neuron 2006, 52, 33–38. [Google Scholar] [CrossRef] [PubMed]
  56. Peelaerts, W.; Bousset, L.; van der Perren, A.; Moskalyuk, A.; Pulizzi, R.; Giugliano, M.; van den Haute, C.; Melki, R.; Baekelandt, V. Synuclein strains cause distinct synucleinopathies after local and systemic administration. Nature 2015, 522, 340–344. [Google Scholar] [CrossRef] [PubMed]
  57. Selkoe, D.; Dettmer, U.; Luth, E.; Kim, N.; Newman, A.; Bartels, T. Defining the native state of alpha-synuclein. Neurodegener. Dis. 2014, 13, 114–117. [Google Scholar] [CrossRef] [PubMed]
  58. Riess, O.; Jakes, R.; Kruger, R. Genetic dissection of familial Parkinson’s disease. Mol. Med. Today 1998, 4, 438–444. [Google Scholar] [CrossRef]
  59. De Silva, H.R.; Khan, N.L.; Wood, N.W. The genetics of Parkinson’s disease. Curr. Opin. Genet. Dev. 2000, 10, 292–298. [Google Scholar] [CrossRef]
  60. Tan, J.M.; Dawson, T.M. Parkin blushed by PINK1. Neuron 2006, 50, 527–529. [Google Scholar] [CrossRef] [PubMed]
  61. Kubo, S.; Hattori, N.; Mizuno, Y. Recessive Parkinson’s disease. Mov. Disord. 2006, 21, 885–893. [Google Scholar] [CrossRef] [PubMed]
  62. Giasson, B.I.; Duda, J.E.; Quinn, S.M.; Zhang, B.; Trojanowski, J.Q.; Lee, V.M. Neuronal alpha-synucleinopathy with severe movement disorder in mice expressing A53T human alpha-synuclein. Neuron 2002, 34, 521–533. [Google Scholar] [CrossRef]
  63. Lee, M.K.; Stirling, W.; Xu, Y.; Xu, X.; Qui, D.; Mandir, A.S.; Dawson, T.M.; Copeland, N.G.; Jenkins, N.A.; Price, D.L. Human alpha-synuclein-harboring familial Parkinson’s disease-linked Ala-53 —> Thr mutation causes neurodegenerative disease with alpha-synuclein aggregation in transgenic mice. Proc. Natl. Acad. Sci. USA 2002, 99, 8968–8973. [Google Scholar] [CrossRef] [PubMed]
  64. Gispert, S.; Del Turco, D.; Garrett, L.; Chen, A.; Bernard, D.J.; Hamm-Clement, J.; Korf, H.W.; Deller, T.; Braak, H.; Auburger, G.; et al. Transgenic mice expressing mutant A53T human alpha-synuclein show neuronal dysfunction in the absence of aggregate formation. Mol. Cell. Neurosci. 2003, 24, 419–429. [Google Scholar] [CrossRef]
  65. Neumann, M.; Kahle, P.J.; Giasson, B.I.; Ozmen, L.; Borroni, E.; Spooren, W.; Muller, V.; Odoy, S.; Fujiwara, H.; Hasegawa, M.; et al. Misfolded proteinase K-resistant hyperphosphorylated alpha-synuclein in aged transgenic mice with locomotor deterioration and in human alpha-synucleinopathies. J. Clin. Investig. 2002, 110, 1429–11439. [Google Scholar] [CrossRef] [PubMed]
  66. Rockenstein, E.; Mallory, M.; Hashimoto, M.; Song, D.; Shults, C.W.; Lang, I.; Masliah, E. Differential neuropathological alterations in transgenic mice expressing alpha-synuclein from the platelet-derived growth factor and Thy-1 promoters. J. Neurosci. Res. 2002, 68, 568–578. [Google Scholar] [CrossRef] [PubMed]
  67. Fleming, S.M.; Salcedo, J.; Fernagut, P.O.; Rockenstein, E.; Masliah, E.; Levine, M.S.; Chesselet, M.F. Early and progressive sensorimotor anomalies in mice overexpressing wild-type human alpha-synuclein. J. Neurosci. 2004, 24, 9434–9440. [Google Scholar] [CrossRef] [PubMed]
  68. Chen, L.; Xie, Z.; Turkson, S.; Zhuang, X. A53T human alpha-synuclein overexpression in transgenic mice induces pervasive mitochondria macroautophagy defects preceding dopamine neuron degeneration. J. Neurosci. 2015, 35, 890–905. [Google Scholar] [CrossRef] [PubMed]
  69. Li, Y.; Liu, W.; Oo, T.F.; Wang, L.; Tang, Y.; Jackson-Lewis, V.; Zhou, C.; Geghman, K.; Bogdanov, M.; Przedborski, S.; et al. Mutant LRRK2(R1441G) BAC transgenic mice recapitulate cardinal features of Parkinson’s disease. Nat. Neurosci. 2009, 12, 826–828. [Google Scholar] [CrossRef] [PubMed]
  70. Dranka, B.P.; Gifford, A.; McAllister, D.; Zielonka, J.; Joseph, J.; O'Hara, C.L.; Stucky, C.L.; Kanthasamy, A.G.; Kalyanaraman, B. A novel mitochondrially-targeted apocynin derivative prevents hyposmia and loss of motor function in the leucine-rich repeat kinase 2 (LRRK2(R1441G)) transgenic mouse model of Parkinson’s disease. Neurosci. Lett. 2014, 583, 159–164. [Google Scholar] [CrossRef] [PubMed]
  71. Yue, M.; Hinkle, K.M.; Davies, P.; Trushina, E.; Fiesel, F.C.; Christenson, T.A.; Schroeder, A.S.; Zhang, L.; Bowles, E.; Behrouz, B.; et al. Progressive dopaminergic alterations and mitochondrial abnormalities in LRRK2 G2019S knock-in mice. Neurobiol. Dis. 2015, 78, 172–195. [Google Scholar] [CrossRef] [PubMed]
  72. Goldberg, M.S.; Fleming, S.M.; Palacino, J.J.; Cepeda, C.; Lam, H.A.; Bhatnagar, A.; Meloni, E.G.; Wu, N.; Ackerson, L.C.; Klapstein, G.J.; et al. Parkin-deficient mice exhibit nigrostriatal deficits but not loss of dopaminergic neurons. J. Biol. Chem. 2003, 278, 43628–43635. [Google Scholar] [CrossRef] [PubMed]
  73. Perez, F.A.; Palmiter, R.D. Parkin-deficient mice are not a robust model of parkinsonism. Proc. Natl. Acad. Sci. USA 2005, 102, 2174–2179. [Google Scholar] [CrossRef] [PubMed]
  74. Lopert, P.; Patel, M. Brain mitochondria from DJ-1 knockout mice show increased respiration-dependent hydrogen peroxide consumption. Redox Biol. 2014, 2, 667–672. [Google Scholar] [CrossRef] [PubMed]
  75. Akundi, R.S.; Huang, Z.; Eason, J.; Pandya, J.D.; Zhi, L.; Cass, W.A.; Sullivan, P.G.; Bueler, H. Increased mitochondrial calcium sensitivity and abnormal expression of innate immunity genes precede dopaminergic defects in Pink1-deficient mice. PLoS ONE 2011, 6, e16038. [Google Scholar] [CrossRef] [PubMed]
  76. Gispert, S.; Ricciardi, F.; Kurz, A.; Azizov, M.; Hoepken, H.H.; Becker, D.; Voos, W.; Leuner, K.; Muller, W.E.; Kudin, A.P.; et al. Parkinson phenotype in aged PINK1-deficient mice is accompanied by progressive mitochondrial dysfunction in absence of neurodegeneration. PLoS ONE 2009, 4, e5777. [Google Scholar] [CrossRef] [PubMed]
  77. Stichel, C.C.; Zhu, X.R.; Bader, V.; Linnartz, B.; Schmidt, S.; Lubbert, H. Mono- and double-mutant mouse models of Parkinson’s disease display severe mitochondrial damage. Hum. Mol. Genet. 2007, 16, 2377–2393. [Google Scholar] [CrossRef] [PubMed]
  78. Bonifati, V.; Oostra, B.A.; Heutink, P. Linking DJ-1 to neurodegeneration offers novel insights for understanding the pathogenesis of Parkinson’s disease. J. Mol. Med. (Berl.) 2004, 82, 163–174. [Google Scholar] [CrossRef] [PubMed]
  79. Abou-Sleiman, P.M.; Healy, D.G.; Wood, N.W. Causes of Parkinson’s disease: Genetics of DJ-1. Cell Tissue Res. 2004, 318, 185–188. [Google Scholar] [CrossRef] [PubMed]
  80. Zimprich, A.; Biskup, S.; Leitner, P.; Lichtner, P.; Farrer, M.; Lincoln, S.; Kachergus, J.; Hulihan, M.; Uitti, R.J.; Calne, D.B.; et al. Mutations in LRRK2 cause autosomal-dominant parkinsonism with pleomorphic pathology. Neuron 2004, 44, 601–607. [Google Scholar] [CrossRef] [PubMed]
  81. Paisan-Ruiz, C.; Jain, S.; Evans, E.W.; Gilks, W.P.; Simon, J.; van der Brug, M.; de Munain, A.L.; Aparicio, S.; Gil, A.M.; Khan, N.; et al. Cloning of the gene containing mutations that cause PARK8-linked Parkinson’s disease. Neuron 2004, 44, 595–600. [Google Scholar] [CrossRef] [PubMed]
  82. West, A.B.; Moore, D.J.; Choi, C.; Andrabi, S.A.; Li, X.; Dikeman, D.; Biskup, S.; Zhang, Z.; Lim, K.L.; Dawson, V.L.; et al. Parkinson’s disease-associated mutations in LRRK2 link enhanced GTP-binding and kinase activities to neuronal toxicity. Hum. Mol. Genet. 2007, 16, 223–232. [Google Scholar] [CrossRef] [PubMed]
  83. Cookson, M.R. The role of leucine-rich repeat kinase 2 (LRRK2) in Parkinson’s disease. Nat. Rev. Neurosci. 2010, 11, 791–797. [Google Scholar] [CrossRef] [PubMed]
  84. West, A.B.; Moore, D.J.; Biskup, S.; Bugayenko, A.; Smith, W.W.; Ross, C.A.; Dawson, V.L.; Dawson, T.M. Parkinson’s disease-associated mutations in leucine-rich repeat kinase 2 augment kinase activity. Proc. Natl. Acad. Sci. USA 2005, 102, 16842–16847. [Google Scholar] [CrossRef] [PubMed]
  85. Simon, H.H.; Saueressig, H.; Wurst, W.; Goulding, M.D.; O'Leary, D.D. Fate of midbrain dopaminergic neurons controlled by the engrailed genes. J. Neurosci. 2001, 21, 3126–3134. [Google Scholar] [PubMed]
  86. Yu-chin Su, X.Q. Impairment of mitochondrial dynamics: A target for the treatment of neurological disorders? Future Med. 2013, 8, 333–346. [Google Scholar]
  87. Tanner, C.M.; Goldman, S.M. Epidemiology of Parkinson’s disease. Neurol. Clin. 1996, 14, 317–35. [Google Scholar] [CrossRef]
  88. Caradoc-Davies, T.H.; Weatherall, M.; Dixon, G.S.; Caradoc-Davies, G.; Hantz, P. Is the prevalence of Parkinson’s disease in New Zealand really changing? Acta Neurol. Scand. 1992, 86, 40–44. [Google Scholar] [CrossRef] [PubMed]
  89. Morgante, L.; Rocca, W.A.; di Rosa, A.E.; de Domenico, P.; Grigoletto, F.; Meneghini, F.; Reggio, A.; Savettieri, G.; Castiglione, M.G.; Patti, F.; et al. Prevalence of Parkinson’s disease and other types of parkinsonism: A door-to-door survey in three Sicilian municipalities. The Sicilian Neuro-Epidemiologic Study (SNES) Group. Neurology 1992, 42, 1901–1907. [Google Scholar] [PubMed]
  90. Mutch, W.J.; Dingwall-Fordyce, I.; Downie, A.W.; Paterson, J.G.; Roy, S.K. Parkinson’s disease in a Scottish city. Br. Med. J. (Clin. Res. Ed.) 1986, 292, 534–536. [Google Scholar] [CrossRef]
  91. Polymeropoulos, M.H.; Higgins, J.J.; Golbe, L.I.; Johnson, W.G.; Ide, S.E.; di Iorio, G.; Sanges, G.; Stenroos, E.S.; Pho, L.T.; Schaffer, A.A.; et al. Mapping of a gene for Parkinson’s disease to chromosome 4q21-q23. Science 1996, 274, 1197–1199. [Google Scholar] [CrossRef] [PubMed]
  92. Dekker, M.C.; Bonifati, V.; van Duijn, C.M. Parkinson’s disease: Piecing together a genetic jigsaw. Brain 2003, 126, 1722–1733. [Google Scholar] [CrossRef] [PubMed]
  93. Langston, J.W. Epidemiology versus genetics in Parkinson’s disease: Progress in resolving an age-old debate. Ann. Neurol. 1998, 44 (Suppl. 1), S45–S52. [Google Scholar] [CrossRef] [PubMed]
  94. Bowers, W.J.; Howard, D.F.; Federoff, H.J. Gene therapeutic strategies for neuroprotection: Implications for Parkinson’s disease. Exp. Neurol. 1997, 144, 58–68. [Google Scholar] [CrossRef] [PubMed]
  95. Kitada, T.; Asakawa, S.; Hattori, N.; Matsumine, H.; Yamamura, Y.; Minoshima, S.; Yokochi, M.; Mizuno, Y.; Shimizu, N. Mutations in the parkin gene cause autosomal recessive juvenile parkinsonism. Nature 1998, 392, 605–608. [Google Scholar] [PubMed]
  96. Lincoln, S.; Vaughan, J.; Wood, N.; Baker, M.; Adamson, J.; Gwinn-Hardy, K.; Lynch, T.; Hardy, J.; Farrer, M. Low frequency of pathogenic mutations in the ubiquitin carboxy-terminal hydrolase gene in familial Parkinson’s disease. Neuroreport 1999, 10, 427–429. [Google Scholar] [CrossRef] [PubMed]
  97. Golbe, L.I.; Di Iorio, G.; Bonavita, V.; Miller, D.C.; Duvoisin, R.C. A large kindred with autosomal dominant Parkinson’s disease. Ann. Neurol. 1990, 27, 276–282. [Google Scholar] [CrossRef] [PubMed]
  98. Kruger, R.; Kuhn, W.; Muller, T.; Woitalla, D.; Graeber, M.; Kosel, S.; Przuntek, H.; Epplen, J.T.; Schols, L.; Riess, O. Ala30Pro mutation in the gene encoding alpha-synuclein in Parkinson’s disease. Nat. Genet. 1998, 18, 106–108. [Google Scholar] [CrossRef] [PubMed]
  99. Polymeropoulos, M.H.; Lavedan, C.; Leroy, E.; Ide, S.E.; Dehejia, A.; Dutra, A.; Pike, B.; Root, H.; Rubenstein, J.; Boyer, R.; et al. Mutation in the alpha-synuclein gene identified in families with Parkinson’s disease. Science 1997, 276, 2045–2047. [Google Scholar] [CrossRef] [PubMed]
  100. Pham, A.H.; Meng, S.; Chu, Q.N.; Chan, D.C. Loss of Mfn2 results in progressive, retrograde degeneration of dopaminergic neurons in the nigrostriatal circuit. Hum. Mol. Genet. 2012, 21, 4817–4826. [Google Scholar] [CrossRef] [PubMed]
  101. Lee, S.; Sterky, F.H.; Mourier, A.; Terzioglu, M.; Cullheim, S.; Olson, L.; Larsson, N.G. Mitofusin 2 is necessary for striatal axonal projections of midbrain dopamine neurons. Hum. Mol. Genet. 2012, 21, 4827–35. [Google Scholar] [CrossRef] [PubMed]
  102. Li, H.; Chen, Y.; Jones, A.F.; Sanger, R.H.; Collis, L.P.; Flannery, R.; McNay, E.C.; Yu, T.; Schwarzenbacher, R.; Bossy, B.; et al. Bcl-xL induces Drp1-dependent synapse formation in cultured hippocampal neurons. Proc. Natl. Acad. Sci. USA 2008, 105, 2169–2174. [Google Scholar] [CrossRef] [PubMed]
  103. Ishihara, N.; Nomura, M.; Jofuku, A.; Kato, H.; Suzuki, S.O.; Masuda, K.; Otera, H.; Nakanishi, Y.; Nonaka, I.; Goto, Y.; et al. Mitochondrial fission factor Drp1 is essential for embryonic development and synapse formation in mice. Nat. Cell. Biol. 2009, 11, 958–966. [Google Scholar] [CrossRef] [PubMed]
  104. Wakabayashi, J.; Zhang, Z.; Wakabayashi, N.; Tamura, Y.; Fukaya, M.; Kensler, T.W.; Iijima, M.; Sesaki, H. The dynamin-related GTPase Drp1 is required for embryonic and brain development in mice. J. Cell Biol. 2009, 186, 805–816. [Google Scholar] [CrossRef] [PubMed]
  105. Berthet, A.; Margolis, E.B.; Zhang, J.; Hsieh, I.; Zhang, J.; Hnasko, T.S.; Ahmad, J.; Edwards, R.H.; Sesaki, H.; Huang, E.J.; et al. Loss of mitochondrial fission depletes axonal mitochondria in midbrain dopamine neurons. J. Neurosci. 2014, 34, 14304–14317. [Google Scholar] [CrossRef] [PubMed]
  106. Zarranz, J.J.; Alegre, J.; Gomez-Esteban, J.C.; Lezcano, E.; Ros, R.; Ampuero, I.; Vidal, L.; Hoenicka, J.; Rodriguez, O.; Atares, B.; et al. The new mutation, E46K, of alpha-synuclein causes Parkinson and Lewy body dementia. Ann. Neurol. 2004, 55, 164–173. [Google Scholar] [CrossRef] [PubMed]
  107. Bonifati, V.; Rizzu, P.; Squitieri, F.; Krieger, E.; Vanacore, N.; van Swieten, J.C.; Brice, A.; van Duijn, C.M.; Oostra, B.; Meco, G.; et al. DJ-1(PARK7), a novel gene for autosomal recessive, early onset parkinsonism. Neurol. Sci. 2003, 24, 159–160. [Google Scholar] [CrossRef] [PubMed]
  108. Iwatsubo, T.; Ito, G.; Takatori, S.; Hannno, Y.; Kuwahara, T. Pathogenesis of Parkinson’s disease: Implications from familial Parkinson’s disease. Rinsho Shinkeigaku 2005, 45, 899–901. [Google Scholar] [PubMed]
  109. Bialecka, M.; Hui, S.; Klodowska-Duda, G.; Opala, G.; Tan, E.K.; Drozdzik, M. Analysis of LRRK 2 G 2019 S and I 2020 T mutations in Parkinson’s disease. Neurosci. Lett. 2005, 390, 1–3. [Google Scholar] [CrossRef] [PubMed]
  110. Smith, R.G. The aging process: Where are the drug opportunities? Curr. Opin. Chem. Biol. 2000, 4, 371–376. [Google Scholar] [CrossRef]
  111. Le, W.; Conneely, O.M.; He, Y.; Jankovic, J.; Appel, S.H. Reduced Nurr1 expression increases the vulnerability of mesencephalic dopamine neurons to MPTP-induced injury. J. Neurochem. 1999, 73, 2218–2221. [Google Scholar] [PubMed]
  112. Xiao, Q.; Castillo, S.O.; Nikodem, V.M. Distribution of messenger RNAs for the orphan nuclear receptors Nurr1 and Nur77 (NGFI-B) in adult rat brain using in situ hybridization. Neuroscience 1996, 75, 221–230. [Google Scholar] [CrossRef]
  113. Rakovic, A.; Grunewald, A.; Seibler, P.; Ramirez, A.; Kock, N.; Orolicki, S.; Lohmann, K.; Klein, C. Effect of endogenous mutant and wild-type PINK1 on Parkin in fibroblasts from Parkinson disease patients. Hum. Mol. Genet. 2010, 19, 3124–3137. [Google Scholar] [CrossRef] [PubMed]
  114. Rakovic, A.; Grunewald, A.; Kottwitz, J.; Bruggemann, N.; Pramstaller, P.P.; Lohmann, K.; Klein, C. Mutations in PINK1 and Parkin Impair Ubiquitination of Mitofusins in Human Fibroblasts. PLoS ONE 2011, 6, e16746. [Google Scholar] [CrossRef] [PubMed]
  115. Jin, H.J.; Li, C.G. Tanshinone IIA and Cryptotanshinone Prevent Mitochondrial Dysfunction in Hypoxia-Induced H9c2 Cells: Association to Mitochondrial ROS, Intracellular Nitric Oxide, and Calcium Levels. Evid. Based Complement. Alternat. Med. 2013, 2013, 610694. [Google Scholar] [CrossRef] [PubMed]
  116. Poole, A.C.; Thomas, R.E.; Andrews, L.A.; McBride, H.M.; Whitworth, A.J.; Pallanck, L.J. The PINK1/Parkin pathway regulates mitochondrial morphology. Proc. Natl. Acad. Sci. USA 2008, 105, 1638–1643. [Google Scholar] [CrossRef] [PubMed]
  117. Yang, Y.; Ouyang, Y.; Yang, L.; Beal, M.F.; McQuibban, A.; Vogel, H.; Lu, B. Pink1 regulates mitochondrial dynamics through interaction with the fission/fusion machinery. Proc. Natl. Acad. Sci. USA 2008, 105, 7070–7075. [Google Scholar] [CrossRef] [PubMed]
  118. Yu, W.; Sun, Y.; Guo, S.; Lu, B. The PINK1/Parkin pathway regulates mitochondrial dynamics and function in mammalian hippocampal and dopaminergic neurons. Hum. Mol. Genet. 2011, 20, 3227–3240. [Google Scholar] [CrossRef] [PubMed]
  119. Wang, X.; Petrie, T.G.; Liu, Y.; Liu, J.; Fujioka, H.; Zhu, X. Parkinson’s disease-associated DJ-1 mutations impair mitochondrial dynamics and cause mitochondrial dysfunction. J. Neurochem. 2012, 121, 830–839. [Google Scholar] [CrossRef] [PubMed]
  120. Ryan, B.J.; Hoek, S.; Fon, E.A.; Wade-Martins, R. Mitochondrial dysfunction and mitophagy in Parkinson’s: From familial to sporadic disease. Trends Biochem. Sci. 2015, 40, 200–210. [Google Scholar] [CrossRef] [PubMed]
  121. Niu, J.; Yu, M.; Wang, C.; Xu, Z. Leucine-rich repeat kinase 2 disturbs mitochondrial dynamics via Dynamin-like protein. J. Neurochem. 2012, 122, 650–658. [Google Scholar] [CrossRef] [PubMed]
  122. Wang, X.; Yan, M.H.; Fujioka, H.; Liu, J.; Wilson-Delfosse, A.; Chen, S.G.; Perry, G.; Casadesus, G.; Zhu, X. LRRK2 regulates mitochondrial dynamics and function through direct interaction with DLP1. Hum. Mol. Genet. 2012, 21, 1931–1944. [Google Scholar] [CrossRef] [PubMed]
  123. Su, Y.C.; Qi, X. Inhibition of excessive mitochondrial fission reduced aberrant autophagy and neuronal damage caused by LRRK2 G2019S mutation. Hum. Mol. Genet. 2013, 22, 4545–4561. [Google Scholar] [CrossRef] [PubMed]
  124. Guo, X.; Disatnik, M.H.; Monbureau, M.; Shamloo, M.; Mochly-Rosen, D.; Qi, X. Inhibition of mitochondrial fragmentation diminishes Huntington’s disease-associated neurodegeneration. J. Clin. Investig. 2013, 123, 5371–5388. [Google Scholar] [CrossRef] [PubMed]
  125. Qi, X.; Qvit, N.; Su, Y.C.; Mochly-Rosen, D. A novel Drp1 inhibitor diminishes aberrant mitochondrial fission and neurotoxicity. J. Cell Sci. 2013, 126, 789–802. [Google Scholar] [CrossRef] [PubMed]
  126. Yamamoto, A.; Yue, Z. Autophagy and its normal and pathogenic states in the brain. Annu. Rev. Neurosci. 2014, 37, 55–78. [Google Scholar] [CrossRef] [PubMed]
  127. Redmann, M.; Dodson, M.; Boyer-Guittaut, M.; Darley-Usmar, V.; Zhang, J. Mitophagy mechanisms and role in human diseases. Int. J. Biochem. Cell Biol. 2014, 53, 127–33. [Google Scholar] [CrossRef] [PubMed]
  128. Fass, E.; Amar, N.; Elazar, Z. Identification of essential residues for the C-terminal cleavage of the mammalian LC3: A lesson from yeast Atg8. Autophagy 2007, 3, 48–50. [Google Scholar] [CrossRef] [PubMed]
  129. Kabeya, Y.; Mizushima, N.; Ueno, T.; Yamamoto, A.; Kirisako, T.; Noda, T.; Kominami, E.; Ohsumi, Y.; Yoshimori, T. LC3, a mammalian homologue of yeast Apg8p, is localized in autophagosome membranes after processing. EMBO J. 2000, 19, 5720–5728. [Google Scholar] [CrossRef] [PubMed]
  130. Kanki, T.; Wang, K.; Cao, Y.; Baba, M.; Klionsky, D.J. Atg32 is a mitochondrial protein that confers selectivity during mitophagy. Dev. Cell 2009, 17, 98–109. [Google Scholar] [CrossRef] [PubMed]
  131. Okamoto, K.; Kondo-Okamoto, N.; Ohsumi, Y. Mitochondria-anchored receptor Atg32 mediates degradation of mitochondria via selective autophagy. Dev. Cell 2009, 17, 87–97. [Google Scholar] [CrossRef] [PubMed]
  132. Liu, L.; Feng, D.; Chen, G.; Chen, M.; Zheng, Q.; Song, P.; Ma, Q.; Zhu, C.; Wang, R.; Qi, W.; et al. Mitochondrial outer-membrane protein FUNDC1 mediates hypoxia-induced mitophagy in mammalian cells. Nat. Cell Biol. 2012, 14, 177–185. [Google Scholar] [CrossRef] [PubMed]
  133. Narendra, D.; Kane, L.A.; Hauser, D.N.; Fearnley, I.M.; Youle, R.J. p62/SQSTM1 is required for Parkin-induced mitochondrial clustering but not mitophagy; VDAC1 is dispensable for both. Autophagy 2010, 6, 1090–1106. [Google Scholar] [CrossRef] [PubMed]
  134. Shi, R.Y.; Zhu, S.H.; Li, V.; Gibson, S.B.; Xu, X.S.; Kong, J.M. BNIP3 interacting with LC3 triggers excessive mitophagy in delayed neuronal death in stroke. CNS Neurosci. Ther. 2014, 20, 1045–1055. [Google Scholar] [CrossRef] [PubMed]
  135. Strappazzon, F.; Nazio, F.; Corrado, M.; Cianfanelli, V.; Romagnoli, A.; Fimia, G.M.; Campello, S.; Nardacci, R.; Piacentini, M.; Campanella, M.; et al. AMBRA1 is able to induce mitophagy via LC3 binding, regardless of PARKIN and p62/SQSTM1. Cell. Death Differ. 2015, 22, 419–432. [Google Scholar] [CrossRef] [PubMed]
  136. Johansen, T.; Lamark, T. Selective autophagy mediated by autophagic adapter proteins. Autophagy 2011, 7, 279–296. [Google Scholar] [CrossRef] [PubMed]
  137. Wu, F.; Xu, H.D.; Guan, J.J.; Hou, Y.S.; Gu, J.H.; Zhen, X.C.; Qin, Z.H. Rotenone impairs autophagic flux and lysosomal functions in Parkinson’s disease. Neuroscience 2015, 284, 900–911. [Google Scholar] [CrossRef] [PubMed]
  138. Arsikin, K.; Kravic-Stevovic, T.; Jovanovic, M.; Ristic, B.; Tovilovic, G.; Zogovic, N.; Bumbasirevic, V.; Trajkovic, V.; Harhaji-Trajkovic, L. Autophagy-dependent and -independent involvement of AMP-activated protein kinase in 6-hydroxydopamine toxicity to SH-SY5Y neuroblastoma cells. Biochim. Biophys. Acta 2012, 1822, 1826–1836. [Google Scholar] [CrossRef] [PubMed]
  139. Youle, R.J.; Narendra, D.P. Mechanisms of mitophagy. Nat. Rev. Mol. Cell Biol. 2011, 12, 9–14. [Google Scholar] [CrossRef] [PubMed]
  140. Narendra, D.P.; Youle, R.J. Targeting mitochondrial dysfunction: Role for PINK1 and Parkin in mitochondrial quality control. Antioxid. Redox Signal. 2011, 14, 1929–1938. [Google Scholar] [CrossRef] [PubMed]
  141. Ashrafi, G.; Schlehe, J.S.; LaVoie, M.J.; Schwarz, T.L. Mitophagy of damaged mitochondria occurs locally in distal neuronal axons and requires PINK1 and Parkin. J. Cell Biol. 2014, 206, 655–670. [Google Scholar] [CrossRef] [PubMed]
  142. Geisler, S.; Holmstrom, K.M.; Skujat, D.; Fiesel, F.C.; Rothfuss, O.C.; Kahle, P.J.; Springer, W. PINK1/Parkin-mediated mitophagy is dependent on VDAC1 and p62/SQSTM1. Nat. Cell Biol. 2010, 12, 119–131. [Google Scholar] [CrossRef] [PubMed]
  143. Rakovic, A.; Shurkewitsch, K.; Seibler, P.; Grunewald, A.; Zanon, A.; Hagenah, J.; Krainc, D.; Klein, C. Phosphatase and tensin homolog (PTEN)-induced putative kinase 1 (PINK1)-dependent ubiquitination of endogenous Parkin attenuates mitophagy: Study in human primary fibroblasts and induced pluripotent stem cell-derived neurons. J. Biol. Chem. 2013, 288, 2223–2237. [Google Scholar] [CrossRef] [PubMed]
  144. McCoy, M.K.; Cookson, M.R. DJ-1 regulation of mitochondrial function and autophagy through oxidative stress. Autophagy 2011, 7, 531–532. [Google Scholar] [CrossRef] [PubMed]
  145. Thomas, K.J.; McCoy, M.K.; Blackinton, J.; Beilina, A.; van der Brug, M.; Sandebring, A.; Miller, D.; Maric, D.; Cedazo-Minguez, A.; Cookson, M.R. DJ-1 acts in parallel to the PINK1/parkin pathway to control mitochondrial function and autophagy. Hum. Mol. Genet. 2011, 20, 40–50. [Google Scholar] [CrossRef] [PubMed]
  146. Irrcher, I.; Aleyasin, H.; Seifert, E.L.; Hewitt, S.J.; Chhabra, S.; Phillips, M.; Lutz, A.K.; Rousseaux, M.W.; Bevilacqua, L.; Jahani-Asl, A.; et al. Loss of the Parkinson’s disease-linked gene DJ-1 perturbs mitochondrial dynamics. Hum. Mol. Genet. 2010, 19, 3734–46. [Google Scholar] [CrossRef] [PubMed]
  147. Tang, B.; Xiong, H.; Sun, P.; Zhang, Y.; Wang, D.; Hu, Z.; Zhu, Z.; Ma, H.; Pan, Q.; Xia, J.H.; et al. Association of PINK1 and DJ-1 confers digenic inheritance of early-onset Parkinson’s disease. Hum. Mol. Genet. 2006, 15, 1816–1825. [Google Scholar] [CrossRef] [PubMed]
  148. Xiong, H.; Wang, D.; Chen, L.; Choo, Y.S.; Ma, H.; Tang, C.; Xia, K.; Jiang, W.; Ronai, Z.; Zhuang, X.; et al. Parkin, PINK1, and DJ-1 form a ubiquitin E3 ligase complex promoting unfolded protein degradation. J. Clin. Investig. 2009, 119, 650–660. [Google Scholar] [CrossRef] [PubMed]
  149. Plowey, E.D.; Cherra, S.J., 3rd; Liu, Y.J.; Chu, C.T. Role of autophagy in G2019S-LRRK2-associated neurite shortening in differentiated SH-SY5Y cells. J. Neurochem. 2008, 105, 1048–1056. [Google Scholar] [CrossRef] [PubMed]
  150. Yakhine-Diop, S.M.; Bravo-San Pedro, J.M.; Gomez-Sanchez, R.; Pizarro-Estrella, E.; Rodriguez-Arribas, M.; Climent, V.; Aiastui, A.; Lopez de Munain, A.; Fuentes, J.M.; Gonzalez-Polo, R.A. G2019S LRRK2 mutant fibroblasts from Parkinson’s disease patients show increased sensitivity to neurotoxin 1-methyl-4-phenylpyridinium dependent of autophagy. Toxicology 2014, 324, 1–9. [Google Scholar] [CrossRef] [PubMed]
  151. Sanchez-Danes, A.; Richaud-Patin, Y.; Carballo-Carbajal, I.; Jimenez-Delgado, S.; Caig, C.; Mora, S.; di Guglielmo, C.; Ezquerra, M.; Patel, B.; Giralt, A.; et al. Disease-specific phenotypes in dopamine neurons from human iPS-based models of genetic and sporadic Parkinson’s disease. EMBO Mol. Med. 2012, 4, 380–395. [Google Scholar] [CrossRef] [PubMed]
  152. Lachenmayer, M.L.; Yue, Z. Genetic animal models for evaluating the role of autophagy in etiopathogenesis of Parkinson disease. Autophagy 2012, 8, 1837–1838. [Google Scholar] [CrossRef] [PubMed]
  153. Alegre-Abarrategui, J.; Christian, H.; Lufino, M.M.; Mutihac, R.; Venda, L.L.; Ansorge, O.; Wade-Martins, R. LRRK2 regulates autophagic activity and localizes to specific membrane microdomains in a novel human genomic reporter cellular model. Hum. Mol. Genet. 2009, 18, 4022–4034. [Google Scholar] [CrossRef] [PubMed]
  154. Saez-Atienzar, S.; Bonet-Ponce, L.; Blesa, J.R.; Romero, F.J.; Murphy, M.P.; Jordan, J.; Galindo, M.F. The LRRK2 inhibitor GSK2578215A induces protective autophagy in SH-SY5Y cells: Involvement of Drp-1-mediated mitochondrial fission and mitochondrial-derived ROS signaling. Cell Death Dis. 2014, 5, e1368. [Google Scholar] [CrossRef] [PubMed]
  155. Gomez-Suaga, P.; Luzon-Toro, B.; Churamani, D.; Zhang, L.; Bloor-Young, D.; Patel, S.; Woodman, P.G.; Churchill, G.C.; Hilfiker, S. Leucine-rich repeat kinase 2 regulates autophagy through a calcium-dependent pathway involving NAADP. Hum. Mol. Genet. 2012, 21, 511–525. [Google Scholar] [CrossRef] [PubMed]
  156. Bravo-San Pedro, J.M.; Niso-Santano, M.; Gomez-Sanchez, R.; Pizarro-Estrella, E.; Aiastui-Pujana, A.; Gorostidi, A.; Climent, V.; Lopez de Maturana, R.; Sanchez-Pernaute, R.; Lopez de Munain, A.; et al. The LRRK2 G2019S mutant exacerbates basal autophagy through activation of the MEK/ERK pathway. Cell. Mol. Life Sci. 2013, 70, 121–136. [Google Scholar] [CrossRef] [PubMed]
  157. Gomez-Suaga, P.; Hilfiker, S. LRRK2 as a modulator of lysosomal calcium homeostasis with downstream effects on autophagy. Autophagy 2012, 8, 692–693. [Google Scholar] [CrossRef] [PubMed]
  158. Orenstein, S.J.; Kuo, S.H.; Tasset, I.; Arias, E.; Koga, H.; Fernandez-Carasa, I.; Cortes, E.; Honig, L.S.; Dauer, W.; Consiglio, A.; et al. Interplay of LRRK2 with chaperone-mediated autophagy. Nat. Neurosci. 2013, 16, 394–406. [Google Scholar] [CrossRef] [PubMed]
  159. Su, Y.C.; Guo, X.; Qi, X. Threonine 56 phosphorylation of Bcl-2 is required for LRRK2 G2019S-induced mitochondrial depolarization and autophagy. Biochim. Biophys. Acta 2015, 1852, 12–21. [Google Scholar] [CrossRef] [PubMed]
  160. Ramsay, R.R.; Salach, J.I.; Singer, T.P. Uptake of the neurotoxin 1-methyl-4-phenylpyridine (MPP+) by mitochondria and its relation to the inhibition of the mitochondrial oxidation of NAD+-linked substrates by MPP+. Biochem. Biophys. Res. Commun. 1986, 134, 743–748. [Google Scholar] [CrossRef]
  161. Marey-Semper, I.; Gelman, M.; Levi-Strauss, M. The high sensitivity to rotenone of striatal dopamine uptake suggests the existence of a constitutive metabolic deficiency in dopaminergic neurons from the substantia nigra. Eur. J. Neurosci. 1993, 5, 1029–1034. [Google Scholar] [CrossRef] [PubMed]
  162. Chinta, S.J.; Mallajosyula, J.K.; Rane, A.; Andersen, J.K. Mitochondrial alpha-synuclein accumulation impairs complex I function in dopaminergic neurons and results in increased mitophagy in vivo. Neurosci. Lett. 2010, 486, 235–239. [Google Scholar] [CrossRef] [PubMed]
  163. Devi, L.; Raghavendran, V.; Prabhu, B.M.; Avadhani, N.G.; Anandatheerthavarada, H.K. Mitochondrial import and accumulation of alpha-synuclein impair complex I in human dopaminergic neuronal cultures and Parkinson disease brain. J. Biol. Chem. 2008, 283, 9089–9100. [Google Scholar] [CrossRef] [PubMed]
  164. Martin, L.J.; Pan, Y.; Price, A.C.; Sterling, W.; Copeland, N.G.; Jenkins, N.A.; Price, D.L.; Lee, M.K. Parkinson’s disease alpha-synuclein transgenic mice develop neuronal mitochondrial degeneration and cell death. J. Neurosci. 2006, 26, 41–50. [Google Scholar] [CrossRef] [PubMed]
  165. Zuo, L.; Motherwell, M.S. The impact of reactive oxygen species and genetic mitochondrial mutations in Parkinson’s disease. Gene 2013, 532, 18–23. [Google Scholar] [CrossRef] [PubMed]
  166. Nakamura, T.; Prikhodko, O.A.; Pirie, E.; Nagar, S.; Akhtar, M.W.; Oh, C.K.; McKercher, S.R.; Ambasudhan, R.; Okamoto, S.I.; Lipton, S.A. Aberrant protein S-nitrosylation contributes to the pathophysiology of neurodegenerative diseases. Neurobiol. Dis. 2015. [Google Scholar] [CrossRef] [PubMed]
  167. Van Muiswinkel, F.L.; Steinbusch, H.W.; Drukarch, B.; de Vente, J. Identification of NO-producing and -receptive cells in mesencephalic transplants in a rat model of Parkinson’s disease: A study using NADPH-d enzyme- and NOSc/cGMP immunocytochemistry. Ann. N. Y. Acad. Sci. 1994, 738, 289–304. [Google Scholar] [CrossRef] [PubMed]
  168. Gu, Z.; Nakamura, T.; Lipton, S.A. Redox reactions induced by nitrosative stress mediate protein misfolding and mitochondrial dysfunction in neurodegenerative diseases. Mol. Neurobiol. 2010, 41, 55–72. [Google Scholar] [CrossRef] [PubMed]
  169. Levecque, C.; Elbaz, A.; Clavel, J.; Richard, F.; Vidal, J.S.; Amouyel, P.; Tzourio, C.; Alperovitch, A.; Chartier-Harlin, M.C. Association between Parkinson’s disease and polymorphisms in the nNOS and iNOS genes in a community-based case-control study. Hum. Mol. Genet. 2003, 12, 79–86. [Google Scholar] [CrossRef] [PubMed]
  170. Joniec, I.; Ciesielska, A.; Kurkowska-Jastrzebska, I.; Przybylkowski, A.; Czlonkowska, A.; Czlonkowski, A. Age- and sex-differences in the nitric oxide synthase expression and dopamine concentration in the murine model of Parkinson’s disease induced by 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine. Brain Res. 2009, 1261, 7–19. [Google Scholar] [CrossRef] [PubMed]
  171. Watanabe, Y.; Kato, H.; Araki, T. Protective action of neuronal nitric oxide synthase inhibitor in the MPTP mouse model of Parkinson’s disease. Metab. Brain Dis. 2008, 23, 51–69. [Google Scholar] [CrossRef] [PubMed]
  172. Castagnoli, K.; Palmer, S.; Castagnoli, N., Jr. Neuroprotection by (R)-deprenyl and 7-nitroindazole in the MPTP C57BL/6 mouse model of neurotoxicity. Neurobiology (Bp) 1999, 7, 135–149. [Google Scholar] [PubMed]
  173. Dolezal, P.; Likic, V.; Tachezy, J.; Lithgow, T. Evolution of the molecular machines for protein import into mitochondria. Science 2006, 313, 314–318. [Google Scholar] [CrossRef] [PubMed]
  174. De Marcos-Lousa, C.; Sideris, D.P.; Tokatlidis, K. Translocation of mitochondrial inner-membrane proteins: Conformation matters. Trends Biochem. Sci. 2006, 31, 259–267. [Google Scholar] [CrossRef] [PubMed]
  175. Baker, M.J.; Frazier, A.E.; Gulbis, J.M.; Ryan, M.T. Mitochondrial protein-import machinery: Correlating structure with function. Trends Cell. Biol. 2007, 17, 456–464. [Google Scholar] [CrossRef] [PubMed]
  176. Gottschalk, W.K.; Lutz, M.W.; He, Y.T.; Saunders, A.M.; Burns, D.K.; Roses, A.D.; Chiba-Falek, O. The Broad Impact of TOM40 on Neurodegenerative Diseases in Aging. J. Parkinsons Dis. Alzheimers Dis. 2014, 1, 12. [Google Scholar] [PubMed]
  177. Bender, A.; Desplats, P.; Spencer, B.; Rockenstein, E.; Adame, A.; Elstner, M.; Laub, C.; Mueller, S.; Koob, A.O.; Mante, M.; et al. TOM40 mediates mitochondrial dysfunction induced by alpha-synuclein accumulation in Parkinson’s disease. PLoS ONE 2013, 8, e62277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Okatsu, K.; Kimura, M.; Oka, T.; Tanaka, K.; Matsuda, N. Unconventional PINK1 localization to the outer membrane of depolarized mitochondria drives Parkin recruitment. J. Cell Sci. 2015, 128, 964–978. [Google Scholar] [CrossRef] [PubMed]
  179. Kato, H.; Lu, Q.; Rapaport, D.; Kozjak-Pavlovic, V. Tom70 is essential for PINK1 import into mitochondria. PLoS ONE 2013, 8, e58435. [Google Scholar] [CrossRef] [PubMed]
  180. Cooper, O.; Seo, H.; Andrabi, S.; Guardia-Laguarta, C.; Graziotto, J.; Sundberg, M.; McLean, J.R.; Carrillo-Reid, L.; Xie, Z.; Osborn, T.; et al. Pharmacological rescue of mitochondrial deficits in iPSC-derived neural cells from patients with familial Parkinson’s disease. Sci. Transl. Med. 2012, 4, 141ra90. [Google Scholar] [CrossRef] [PubMed]
  181. Reinhardt, P.; Schmid, B.; Burbulla, L.F.; Schondorf, D.C.; Wagner, L.; Glatza, M.; Hoing, S.; Hargus, G.; Heck, S.A.; Dhingra, A.; et al. Genetic correction of a LRRK2 mutation in human iPSCs links parkinsonian neurodegeneration to ERK-dependent changes in gene expression. Cell. Stem Cell. 2013, 12, 354–367. [Google Scholar] [CrossRef] [PubMed]
  182. Toyoda, Y.; Erkut, C.; Pan-Montojo, F.; Boland, S.; Stewart, M.P.; Muller, D.J.; Wurst, W.; Hyman, A.A.; Kurzchalia, T.V. Products of the Parkinson’s disease-related glyoxalase DJ-1, D-lactate and glycolate, support mitochondrial membrane potential and neuronal survival. Biol. Open 2014, 3, 777–784. [Google Scholar] [CrossRef] [PubMed]
  183. Daher, J.P.; Abdelmotilib, H.A.; Hu, X.; Volpicelli-Daley, L.A.; Moehle, M.S.; Fraser, K.B.; Needle, E.; Chen, Y.; Steyn, S.J.; Galatsis, P.; et al. LRRK2 Pharmacological Inhibition Abates alpha-Synuclein Induced Neurodegeneration. J. Biol. Chem. 2015, 290, 19433–19444. [Google Scholar] [CrossRef] [PubMed]
  184. Li, T.; He, X.; Thomas, J.M.; Yang, D.; Zhong, S.; Xue, F.; Smith, W.W. A novel GTP-binding inhibitor, FX2149, attenuates LRRK2 toxicity in Parkinson’s disease models. PLoS ONE 2015, 10, e0122461. [Google Scholar] [CrossRef] [PubMed]
  185. Chiu, C.C.; Yeh, T.H.; Lai, S.C.; Wu-Chou, Y.H.; Chen, C.H.; Mochly-Rosen, D.; Huang, Y.C.; Chen, Y.J.; Chen, C.L.; Chang, Y.M.; et al. Neuroprotective effects of aldehyde dehydrogenase 2 activation in rotenone-induced cellular and animal models of parkinsonism. Exp. Neurol. 2015, 263, 244–253. [Google Scholar] [CrossRef] [PubMed]
  186. Richter, F.; Gao, F.; Medvedeva, V.; Lee, P.; Bove, N.; Fleming, S.M.; Michaud, M.; Lemesre, V.; Patassini, S.; de La Rosa, K.; et al. Chronic administration of cholesterol oximes in mice increases transcription of cytoprotective genes and improves transcriptome alterations induced by alpha-synuclein overexpression in nigrostriatal dopaminergic neurons. Neurobiol. Dis. 2014, 69, 263–275. [Google Scholar] [CrossRef] [PubMed]
  187. Rappold, P.M.; Cui, M.; Grima, J.C.; Fan, R.Z.; de Mesy-Bentley, K.L.; Chen, L.; Zhuang, X.; Bowers, W.J.; Tieu, K. Drp1 inhibition attenuates neurotoxicity and dopamine release deficits in vivo. Nat. Commun. 2014, 5, 5244. [Google Scholar] [CrossRef] [PubMed]
  188. Mena, N.P.; Garcia-Beltran, O.; Lourido, F.; Urrutia, P.J.; Mena, R.; Castro-Castillo, V.; Cassels, B.K.; Nunez, M.T. The novel mitochondrial iron chelator 5-((methylamino)methyl)-8-hydroxyquinoline protects against mitochondrial-induced oxidative damage and neuronal death. Biochem. Biophys. Res. Commun. 2015, 463, 787–792. [Google Scholar] [CrossRef] [PubMed]
  189. Jiang, J.; Zuo, Y.; Gu, Z. Rapamycin protects the mitochondria against oxidative stress and apoptosis in a rat model of Parkinson’s disease. Int. J. Mol. Med. 2013, 31, 825–832. [Google Scholar] [PubMed]
  190. Xiong, N.; Xiong, J.; Khare, G.; Chen, C.; Huang, J.; Zhao, Y.; Zhang, Z.; Qiao, X.; Feng, Y.; Reesaul, H.; et al. Edaravone guards dopamine neurons in a rotenone model for Parkinson’s disease. PLoS ONE 2011, 6, e20677. [Google Scholar] [CrossRef] [PubMed]
  191. Dabbeni-Sala, F.; di Santo, S.; Franceschini, D.; Skaper, S.D.; Giusti, P. Melatonin protects against 6-OHDA-induced neurotoxicity in rats: A role for mitochondrial complex I activity. FASEB J. 2001, 15, 164–170. [Google Scholar] [CrossRef] [PubMed]
  192. Zaitone, S.A.; Hammad, L.N.; Farag, N.E. Antioxidant potential of melatonin enhances the response to L-dopa in 1-methyl 4-phenyl 1,2,3,6-tetrahydropyridine-parkinsonian mice. Pharmacol. Rep. 2013, 65, 1213–1226. [Google Scholar] [CrossRef]
  193. Karuppagounder, S.S.; Madathil, S.K.; Pandey, M.; Haobam, R.; Rajamma, U.; Mohanakumar, K.P. Quercetin up-regulates mitochondrial complex-I activity to protect against programmed cell death in rotenone model of Parkinson’s disease in rats. Neuroscience 2013, 236, 136–148. [Google Scholar] [CrossRef] [PubMed]
  194. Jayaraj, R.L.; Elangovan, N.; Dhanalakshmi, C.; Manivasagam, T.; Essa, M.M. CNB-001, a novel pyrazole derivative mitigates motor impairments associated with neurodegeneration via suppression of neuroinflammatory and apoptotic response in experimental Parkinson’s disease mice. Chem. Biol. Interact. 2014, 220, 149–157. [Google Scholar] [CrossRef] [PubMed]
  195. Abdin, A.A.; Sarhan, N.I. Intervention of mitochondrial dysfunction-oxidative stress-dependent apoptosis as a possible neuroprotective mechanism of alpha-lipoic acid against rotenone-induced parkinsonism and L-dopa toxicity. Neurosci. Res. 2011, 71, 387–395. [Google Scholar] [CrossRef] [PubMed]
  196. Kaur, H.; Chauhan, S.; Sandhir, R. Protective effect of lycopene on oxidative stress and cognitive decline in rotenone induced model of Parkinson’s disease. Neurochem. Res. 2011, 36, 1435–1443. [Google Scholar] [CrossRef] [PubMed]
  197. Chen, C.H.; Budas, G.R.; Churchill, E.N.; Disatnik, M.H.; Hurley, T.D.; Mochly-Rosen, D. Activation of aldehyde dehydrogenase-2 reduces ischemic damage to the heart. Science 2008, 321, 1493–1145. [Google Scholar] [CrossRef] [PubMed]
  198. Bordet, T.; Buisson, B.; Michaud, M.; Drouot, C.; Galea, P.; Delaage, P.; Akentieva, N.P.; Evers, A.S.; Covey, D.F.; Ostuni, M.A.; et al. Identification and characterization of cholest-4-en-3-one, oxime (TRO19622), a novel drug candidate for amyotrophic lateral sclerosis. J. Pharmacol. Exp. Ther. 2007, 322, 709–720. [Google Scholar] [CrossRef] [PubMed]
  199. Gouarne, C.; Tracz, J.; Paoli, M.G.; Deluca, V.; Seimandi, M.; Tardif, G.; Xilouri, M.; Stefanis, L.; Bordet, T.; Pruss, R.M. Protective role of olesoxime against wild-type alpha-synuclein-induced toxicity in human neuronally differentiated SHSY-5Y cells. Br. J. Pharmacol. 2015, 172, 235–245. [Google Scholar] [CrossRef] [PubMed]
  200. Fernandez-Moriano, C.; Gonzalez-Burgos, E.; Gomez-Serranillos, M.P. Mitochondria-Targeted Protective Compounds in Parkinson’s and Alzheimer’s Diseases. Oxid. Med. Cell. Longev. 2015, 2015, 408927. [Google Scholar] [CrossRef] [PubMed]
  201. Valadas, J.S.; Vos, M.; Verstreken, P. Therapeutic strategies in Parkinson’s disease: What we have learned from animal models. Ann. N. Y. Acad. Sci. 2015, 1338, 16–37. [Google Scholar] [CrossRef] [PubMed]
  202. Procaccio, V.; Bris, C.; de la Barca, J.M.C.; Oca, F.; Chevrollier, A.; Amati-Bonneau, P.; Bonneau, D.; Reynier, P. Perspectives of drug-based neuroprotection targeting mitochondria. Rev. Neurol. (Paris) 2014, 170, 390–400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Yadav, A.; Agarwal, S.; Tiwari, S.K.; Chaturvedi, R.K. Mitochondria: Prospective targets for neuroprotection in Parkinson’s disease. Curr. Pharm. Des. 2014, 20, 5558–5573. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Luo, Y.; Hoffer, A.; Hoffer, B.; Qi, X. Mitochondria: A Therapeutic Target for Parkinson’s Disease? Int. J. Mol. Sci. 2015, 16, 20704-20730. https://doi.org/10.3390/ijms160920704

AMA Style

Luo Y, Hoffer A, Hoffer B, Qi X. Mitochondria: A Therapeutic Target for Parkinson’s Disease? International Journal of Molecular Sciences. 2015; 16(9):20704-20730. https://doi.org/10.3390/ijms160920704

Chicago/Turabian Style

Luo, Yu, Alan Hoffer, Barry Hoffer, and Xin Qi. 2015. "Mitochondria: A Therapeutic Target for Parkinson’s Disease?" International Journal of Molecular Sciences 16, no. 9: 20704-20730. https://doi.org/10.3390/ijms160920704

Article Metrics

Back to TopTop