Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Nov 20.
Published in final edited form as: Free Radic Biol Med. 2019 Jun 12;144:279–292. doi: 10.1016/j.freeradbiomed.2019.06.007

Interrogating Parkinson’s Disease Associated Redox Targets: Potential Application of CRISPR Editing

MA Artyukhova a, YY Tyurina b, CT Chu c, TM Zharikova a,e, H Bayir b,d, VE Kagan b,f,g,h,i, PS Timashev a,j,k
PMCID: PMC6832799  NIHMSID: NIHMS1532591  PMID: 31201850

Abstract

Loss of dopaminergic neurons in the substantia nigra is one of the pathogenic hallmarks of Parkinson’s disease, yet the underlying molecular mechanisms remain enigmatic. While aberrant redox metabolism strongly associated with iron dysregulation and accumulation of dysfunctional mitochondria is considered as one of the major contributors to neurodegeneration and death of dopaminergic cells, the specific anomalies in the molecular machinery and pathways leading to the PD development and progression have not been identified. The high efficiency and relative simplicity of a new genome editing tool, CRISPR/Cas9, make its applications attractive for deciphering molecular changes driving PD-related impairments of redox metabolism and lipid peroxidation in relation to mishandling of iron, aggregation and oligomerization of alpha-synuclein and mitochondrial injury as well as in mechanisms of mitophagy and programs of regulated cell death (apoptosis and ferroptosis). These insights into the mechanisms of PD pathology may be used for the identification of new targets for therapeutic interventions and innovative approaches to genome editing, including CRISPR/Cas9.

Keywords: Parkinson’s disease, CRISPR/Cas 9, Ferroptosis, Iron, Iron homeostasis, Lipid peroxidation, Mitophagy, Mitochondrial dysfunction, ROS

Graphical Abstract

graphic file with name nihms-1532591-f0007.jpg


“The art of medicine consists in amusing the patient while nature cures the disease.”

Voltaire

1. Introduction

Parkinson’s disease (PD) – the condition first described by James Parkinson at the beginning of the XIX century – still remains one of the most debilitating diseases worldwide. PD is a progressive neurodegenerative disorder involving the loss of dopaminergic (DA) neurons in the substantia nigra (SN) of the midbrain [1]. PD is characterized by severe motor deficits such as bradykinesia, postural instability and tremor. Most patients with PD exhibit impaired cognition, especially deficits in attention, executive functioning, visual-spatial processing and varying degree of dementia [2, 3]. Not only does PD affects the patient’s quality of life, but as populations age, patients with PD are expected to impose an increasing social and economic burden [4].

Morphological hallmarks of PD include loss of 50-70% of DA neurons, and the presence of Lewy Bodies (LBs) enriched in abnormal α-synuclein (α-syn) filaments [5, 6]. Whether these LBs contribute to disease pathogenesis or serve as markers of underlying cellular dysfunction or adaptation remains controversial.

Most cases of PD are sporadic and approximately 10% are familial in etiology [7]. Familial PD is a genetically inherited form of PD accompanied by a variety of disturbances suggesting that mutations in one or more genes can lead to the synthesis of abnormal proteins associated with increased aggregation and/or changes in enzymatic activity. On the contrary, sporadic PD neither has apparent genetic linkage nor clear idea on the contribution of genetic risk variants and environmental risk factors to this form of PD [8]. Among many possible genetic variants potentially contributing to the sporadic PD, in this review we will focus on those associated with aberrant redox mechanisms and leading to triggering of cell death programs, including apoptosis and ferroptosis. The emphasis will be on the enzymatic pathways of selective phospholipid peroxidation as mechanisms of apoptotic and ferroptotic cell death programs in the context of the SN iron dishomeostasis that lead to the disturbed redox balance and H2O2 (ROS) accumulation feeding the enzymatic lipid peroxidation. We will also discuss the role of lipid peroxidation as a source of lipid mediators (oxidized PUFA) that affect inflammation and abnormalities associated with the PD progression [913].

Currently, symptomatic drug therapy is the main strategy for PD. No mechanism based therapeutic approaches have been developed that may prevent, control or reduce the pathological manifestations of PD [14, 15]. Moreover, the symptomatic therapeutic modalities used are associated with a variety of side effects. In the field of cell replacement therapies, a few studies showed the potential of generation of DA neurons from human embryonic stem cells (hESCs) followed by implantation of these cells in animal models of PD [1619]. The initial results showed the increase in dopamine levels in the brain of experimental animals [16, 17]. However, this approach has a lot of unsolved problems, including risk of brain tumors, phenotype instability of hESC-derived DA neurons, immunologic responses, not to mention ethical concerns and the necessity to examine the safety and efficacy of the treatment for PD patients.

Based on the possible pathogenic role of α-syn and the demonstrated capacity of microglia to degrade abnormal intracellular α-syn filaments and prevent damage of DA neurons [7], vaccines against α-syn might be a good therapeutic strategy. However, no studies developing this approach have been published so far.

There is an obvious urgency in mechanistic studies of a new type aimed at integrative understanding of genetic and environmental factors triggering multiple aberrant metabolic pathways and erroneous interactions of various macromolecules during overall PD pathogenesis. In this regard, promising opportunities may be associated with CRISPR/Cas9 system – a new tool developed during the last five years – which enables quick and precise genome editing in almost any living creature [2022]. Several recent publications indicate that CRISPR/Cas9 has a potential to broaden and accelerate the fundamental research aimed at understanding of pathogenic mechanisms of neurodegenerative diseases and lead to new therapies, especially for PD [2325].

This new technology can be applied for genomewide screening experiments to simultaneously assess large numbers of genes and identify previously unknown PD-associated genetic variants that enable subsequent generation of adequate cellular and animal models effectively applicable for drug discovery. There are several excellent recent reviews considering different applications of CRISPR/Cas9 technology in relation to PD [2629]. This review is focused on the aspects of direct reprogramming of host neuronal and immune cells in the context of targeting the redox mechanisms of lipid signaling and neuro-inflammation associated with genetic programs of regulated cell death (Table 1).

Table 1. Genes implicated in Parkinson’s disease impairment of redox metabolism and lipid peroxidation.

Here we reviewed most promising candidate genes and processes that highlight PD pathology: mitophagy, iron homeostasis, ferroptotic and apoptotic cell death programs.

Gene Protein Function
ACSL4   Acyl-CoA Synthetase Long-Chain Family Member 4   Preferably converts arachidonic acid into fatty acyl-CoA esters, and thereby plays a key role in ferroptosis
ALOXs   Lipoxygenases   Involved in peroxidation of polyunsaturated fatty acids in ferroptosis
APP Amyloid precursor protein Regulator of synapse formation, neural plasticity and iron export
FBXO7 F-box only protein 7 Mediates the ubiquitination and subsequent proteasomal degradation of target proteins; regulation of mitophagy
GCLC/GCLM   Glutamate-cysteine ligase   Biosynthesis of glutathione
GPX4   Glutathione peroxidase 4   Reduces membrane phospholipid hydroperoxides to suppress ferroptosis
GSS   Glutathione synthetase   Biosynthesis of glutathione
IREB2 Iron-responsive element-binding protein 2 Regulator of TFRC, FPN, DMT1 translational activity
MAPT tau Microtubule-associated protein; regulates APP trafficking in neurons
NME4 (NDPK-D) Nucleoside diphosphate kinase D Synthesis of nucleoside triphosphates other than ATP; regulator of mitophagy
PARK2 Parkin Ubiquitin ligase, regulator of mitophagy
PARK7 Nucleic acid deglycase DJ-1 Redox-sensitive chaperone and a sensor for oxidative stress. Required for mitochondrial morphology and function
PINK1 PTEN-induced putative kinase 1 Kinase, phosphorylates Parkin and ubiquitin to regulate mitophagy
SLC40A1 Solute carrier family 40 member 1 (Ferroportin) Transmembrane protein that transports iron from the inside of a cell; promotes ferroptosis
SNCA α-synuclein Regulator of dopamine biosynthesis, release and transport; suppressor of apoptosis

2.1. Overview

In the early 2000s, Clustered Regularly Interspaced Short Palindromic Repeats (CRISPR) has been found in the genomes of different prokaryotes, providing them with immunity against viral infections [3033]. Since then, three types and twelve subtypes of CRISPR/Cas systems have been discovered [34]. The best characterized is the type II CRISPR system [3538] that consists of the endonuclease Cas9, bacterial CRISPR RNA (crRNA) and trans-activating crRNA (tracrRNA) (Fig. 1). Later these naturally occurring bacterial RNAs were fused together to generate the single guide RNA molecule (gRNA) [37]. The target activity of CRISPR/Cas9 systems is provided by protospacer adjacent motif (PAM) – a sequence within an invading DNA that helps bacteria to recognize pathogenic genetic material from its own [39, 40]. Only if the spacer sequence is fully homologous to PAM the CRISPR/Cas9 system will target viral or plasmid genetic material by introducing double-stranded (ds)DNA breaks in the invading DNA [41]. These discoveries have led to the realization that the CRISPR/Cas9 system can be used as a new tool for genome editing in a variety of organisms. For this, the DNA sequence of interest is selected and a dsDNA break is introduced in a selected region by simply adding designed gRNA, Cas9 protein and your repair template (a plasmid) if necessary (Fig. 1). The technology offers new tools for a broad spectrum of genetic manipulations, including introduction of point mutations [42, 43], deletion and insertion of short and large DNA fragments into an organism’s genome [44, 45], epigenome editing [46] and CRISPR-based activation and inactivation of different genes [47, 48]. A number of genome editing technologies preceded CRISPR/Cas9 system: zinc-finger nucleases (ZFNs) [49, 50] and transcription activator-like effector nucleases (TALENs) [5052]. However, CRISPR/Cas9 genome editing technology possesses a number of advantages over previous approaches. First, CRISPR/Cas9 system is more precise as it utilizes specific gRNAs fully complementary through Watson-Crick base pairing to the PAM and targeted DNA sequence. It makes CRISPR/Cas9 technology more specific and easier to design. Second, CRISPR/Cas9 technology saves a lot of research time. Modern gRNA designing tools allow a researcher to create a perfect guide RNA for a genomic locus of interest within an hour. The simple principle of CRISPR/Cas9 genome editing and an easy to perform procedure allows generation of genetically/epigenetically modified cells within a few weeks. Third, an important point about CRISPR is its versatility. By changing gRNA sequence one can target almost any sequence of any living creature from algae to humans.

Fig. 1. CRISPR/Cas9 genome editing system.

Fig. 1.

The CRISPR/Cas9 two-component system is composed of Cas9 endonuclease and a single guide RNA (gRNA) molecule. Endonuclease Cas9 generates site-specific double-strand breaks in the genomic locus of interest by direct pairing of gRNA molecule with the target DNA. The targeting specificity relies on the presence of PAM sequence (protospacer adjacent motif) in the DNA sequence of interest. Upon Cas9 action double-stranded breaks are introduced into the DNA, which in mammalian cells are usually repaired either by non-homologous end joining (NHEJ), an error-prone mechanism that generates random insertion and deletion mutations at the targeted locus (used for knock-outs), or by homology-directed repair (HDR), for which an exogenous repair template is provided.

2.2. Application of CRISPR/Cas9 genome editing to Parkinson’s disease

Difficulties in deciphering the PD pathogenic mechanisms are due, to a large extent, to the multiplicity of mutations with Mendelian autosomal dominant (AD), SNCA, LRRK2, VPS35, HTRA2, EIF4G1, GBA, or autosomal recessive (AR) inheritance – PARK2 (encoding for Parkin protein), PINK1, PARK7 (known as DJ-1), ATP13A2, PLA2G6, FBXO7 [5361]. With the familial forms of PD accounting only for 10% cases [14], and the rest of them considered sporadic [58], it is difficult to sort through many genetic risk variants [62, 63] to determine if they could be causally linked to PD pathogenesis. Application of CRISPR/Cas9 and CRISPR for screening and identification of new therapeutic redox targets may be quite valuable.

3. The role of mitochondria in Parkinson’s disease

3.1. Parkin/Pink1-dependent mitophagy

Mitochondria are unique power-houses of cells with multiple additional signaling and metabolic functions [64]. As the major electron transporting redox organelle, mitochondria can also generate reactive oxygen species (ROS) essential for the dynamic regulation of their own organization via fission, fusion and mitophagy [6466]. Discoordination of mitochondrial redox machinery may lead to their damage which, in turn, triggers self-elimination through selective autophagy, known as mitophagy. Mitophagy has been observed in degenerating SN neurons of patients with PD or the related Lewy body dementia (LBD) [67]. Inefficient mitophagy or excessive or prolonged activation of mitophagy may contribute to neurodegeneration [11, 68]. In response to different physiological stimuli, various signaling cascades can be activated to trigger specific types of mitophagy [69]. Impairment of this essential cellular process perturbs mitochondrial function and leads to accumulation of damaged mitochondria within the cells and furthermore promotes cell and tissue damage [68]. Mitochondrial damage and excessive ROS production have also been considered as causative pathogenic factors in animals with PD-associated mutations [8, 70].

PTEN-induced putative kinase 1(PINK1), a cytosolic E3 ubiquitin ligase (Parkin), and F-box only protein 7 (Fbxo7) are essential for proper mitochondrial functioning (Fig. 2). Their disruption can affect mitochondrial structure, function and sensitize cells to pro-oxidant factors (Table 1) [7173].

Fig. 2. Mitophagy in neurons.

Fig. 2.

Left: following stress, PINK1 is stabilized on the mitochondrial outer membrane (MOM), promoting Parkin recruitment to the mitochondria. Parkin ubiquitinates several MOM proteins. Adaptor proteins (p62, OPTIN) recognize phosphorylated mitochondrial proteins and initiate autophagosome formation through binding with LC3, thus promoting mitophagy. During apoptosis mitochondrial inner membrane (MIM) phospholipid cardiolipin (CL) interacts with cytochrome c (cyt c) and forms a peroxidase complex that catalyzes CL oxidation. The cyt c/CL complex is capable of promoting covalent hetero-oligomerization of α-syn with cyt c into high molecular weight aggregates that occur in Lewy neuritis of PD patients. Right: another protein NDPK-D (NME4) plays a key role in mitophagy initiation. In healthy mitochondria, NME4 interacts with MIM and OPA1. In this topology, NME4 has NDP kinase (phosphotransfer) activity. Depolarization of mitochondrion initiates redistribution of CL and its externalization to the mitochondrial surface; hexameric NME4 forms a complex with CL and acts like a rotary machine transferring CL from MIM to MOM through the intermembrane space (IMS) interacting simultaneously with MIM and MOM. Binding to CL suppresses NDP kinase activity, but facilitates the intermembrane transfer of the phospholipid.

Importantly, PINK1 and Parkin act in a common pathway to control turnover of mitochondria via ubiquitin-dependent mitophagy [71, 74]. PINKI-Parkin pathway regulates mitochondrial dynamics, biogenesis, transport and recruitment of autophagic machinery, to ensure proper elimination of defective organelles [68, 74, 75]. Normally, PINK1 is imported into mitochondria to reach the mitochondrial inner membrane (MIM) where it is cleaved by several proteases and degraded by the ubiquitin-proteasome system [74]. The import of PINK1 depends on mitochondrial membrane potential. Following the loss of mitochondrial membrane potential, active PINK1 starts to accumulate on the mitochondrial outer membrane (MOM), promoting Parkin translocation to the damaged mitochondria [76, 77]. Upon phosphorylation, Parkin is activated to initiate ubiquitination of numerous MOM proteins, induce recruitment of ubiquitin-binding mitophagy receptors to promote capture of damaged mitochondria by the autophagosome [68, 78, 79].

In flies, mutations in PINK1 and Parkin homologues cause mitochondrial dysfunction associated with PD, such as loss of DA neurons, mitochondrial enlargement and disintegration, muscle degeneration and shortened lifespan [71, 80]. However, PINK1 and Parkin knockout mice failed to reproduce PD related conditions observed in human patients [81]. Moreover, PINK1- and Parkin-independent pathways of mitophagy occur in neuronal cultures and in vivo [8285]. Recently, significant phenotypic differences between several PINK1 knockout animal models and human PD patients have been documented [8688]. Yang et al. observed marked neuronal loss in CRISPR-modified PINK1 knockout rhesus monkey that was not reproduced in studies on PINK1 deficient mice and pigs - the difference possibly associated with the primate-specific expression and function of PINK1 [89]. It was assumed that PINK1 functions may be diverse and the severity and complexity of PINK1 depleted phenotypes can depend on the type and the number of mutations in the single gene. Furthermore, the mosaicism of CRISPR/Cas9-mediated mutations can cause different degrees of PINK1 loss [89].

CRISPR/Cas9 technology can be also used in searches of new genes, transcription factors or other regulators of Parkin/PINK1 expression in cells. For example, Potting et al. applied genome-wide CRISPR/Cas9 knockout screening and discovered 53 positive and negative regulators of PARKIN abundance in the context of transcriptional repression and mitophagy [90].

Another protein, DJ-1 (Table 1), has been suggested to play a key role in regulating oxidative stress, ROS formation, mitochondrial function and autophagy [91, 92]. Mutations in the DJ-1 gene (PARK7) have been linked to autosomal-recessive early on-set PD [53]. DJ-1 is mainly localized to the cytoplasm, however, under normal conditions only a small portion of DJ-1 is present in mitochondria. Upon decrease of mitochondrial membrane potential, DJ-1 translocates into mitochondria promoting clearance of damaged organelle by mitophagy [93]. Hao et al. reported that DJ-1 knockout leads to mitochondrial dysfunction in age-dependent manner in both Drosophila and mouse. DJ-1 knockout flies manifest similar phenotypes as PINK1 and Parkin mutants: male sterility, shortened lifespan, and reduced climbing ability [94]. Moreover, DJ-1 up-regulation can rescue PINK1 deficiency in mutant flies and protect neurons against oxidative stress-induced cell death in a rat model of PD [94, 95]. While DJ-1 and PINK1/Parkin may participate in two parallel pathways whose function critically impacts mitochondrial activity the precise relationships between these three genes remain enigmatic [94]. DJ-1 has been also shown to reduce α-syn aggregation and toxicity [96]. It has been demonstrated that DJ-1 can directly interact with monomeric α-syn in vitro and in cells and reduce its dimerization. In contrast, DJ-1 mutants failed to antagonize α-syn dimerization and abrogate the protective effects of wild-type DJ-1 [97].

Fbxo7 plays a role in modulating of mitochondrial homeostasis [98]. The PD-associated mutations in this gene may result in mitochondrial dysfunction through impairment of mitochondrial respiration or accumulation of defective mitochondria via impaired mitophagy [73]. In response to mitochondrial depolarization Fbxo7 can promote mitophagy through direct interaction with Parkin and Pink1 [98]. Under stress-induced conditions cytosolic Fbxo7 relocates to the mitochondria. Moreover, Fbxo7 participates in Parkin translocation to the mitochondria via direct physical interaction. Importantly, the loss of Fbxo7 expression results in a significant inhibition of Parkin recruitment to depolarized mitochondria and subsequent autophagic clearance of the dysfunctional organelle [98, 99].

Interestingly, overexpression of Fbxo7 in Drosophila Parkin mutants significantly rescued the Parkin depleted phenotypes, including loss of DA neurons, muscle degeneration, and mitochondrial disruption, but failed to rescue the effects of PINK1 silencing in mammalian cells [98]. These observations are compatible with recent reports that PINK1 plays additional roles in neuronal health through activation of cytosolic signaling pathways [100].

Several attempts to use CRISPR/Cas9 technology to model and treat PD-related conditions associated with impairment of Parkin/PINK1-dependent mitophagy have been made. However, knockout of Parkin and PINK1 in animal models of PD did not reproduce PD related behavior and pathological alterations observed in PD patients [81, 8689]. These results suggest that PD pathology is much more complex and depends on the proper execution of many other mechanisms. Mutations and improper functioning of Parkin/PINK1 can launch mechanisms that compensate for the loss of their function and stabilize mitochondrial dynamics. Moreover, execution of mitophagy can occur independently from Parkin/PINK1 pathway through other mechanisms. One of many possible applications of CRISPR can be the unearthing such mechanisms and deciphering their role in PD. First, meta-analysis of previous research of such mechanisms and pathways should be performed with subsequent testing in vitro by generation of CRISPR gRNA libraries and their use in screening with CRISPR/Cas9. Interestingly, Burchell et al. showed that Fbxo7 overexpression rescued Parkin depletion - but not PINK1 silencing [98]. This suggests that other protein(s) can be involved in Parkin/PINK1-driven mitophagy and maintenance of mitochondrial functions or that PINK1 has independent functions regulating neuronal health [100102]. Hence, revealing the respective genes and their protein products may be another challenging employment of CRISPR/Cas9 screening system.

3.2. α-synuclein

Missense mutations and duplications or triplications of the SNCA gene, which encodes for α-syn (Table 1), have been linked to PD with autosomal dominant inheritance and to sporadic forms of PD as well [57, 60, 61]. Elevated expression levels of endogenous α-syn induce mitochondrial dysfunction, characterized by impaired respiration and mitochondrial membrane depolarization [12]. Aggregation of endogenous α-syn monomers generates beta sheet-rich oligomers that localize to the mitochondria in close proximity to several mitochondrial proteins including ATP synthase. Oligomeric complexes are able to impair complex I dependent respiration, oxidize ATP synthase beta subunit, induce production of ROS that subsequently stimulates peroxidation of lipids in the MIM and neuronal death [12, 103, 104]. These pro-oxidant events increase the probability of permeability transition pore (PTP) opening, leading to equilibration of ionic gradients, swelling of the matrix, cristae unfolding, rupture of the MOM ultimately triggering neuronal death [12]. Inhibition of oligomer induced oxidation prevents the pathological induction of PTP [105]. Furthermore, suppression of lipid peroxidation by bis-allylic deuterated polyunsaturated fatty acids (D-PUFAs) protects glio-neuronal cell cultures from death [104]. α-Syn can also exert its pro-oxidant effects via interactions with a mitochondrial intermembrane space hemoprotein, cytochrome c (cyt c) (Fig. 2) [79]. This interaction is realized with the participation of a mitochondria-specific phospholipid, cardiolipin (CL). The latter forms a complex with cyt c that displays a peroxidase activity and oxidizes CL. Notably, α-syn can bind with cyt c/CL to yield a triple complex retaining a high pro-oxidant potential capable of damaging mitochondria. The presence of CL in the cyt c-CL complex is required to mediate cyt c interaction with α-syn. This unexpected pro-oxidant catalytic activity of α-syn represents a source of persistent oxidative stress in affected DA neurons that leads to chronic neurodegeneration. Interestingly, α-syn-cyt c aggregates were detected in SN DA neurons of rotenone-treated rats and LBs of PD patients [106].

In spite of these strong circumstantial pieces of evidence for a pro-oxidant role of role α-syn in mitochondrial and, possibly, extra-mitochondrial compartments, the intricacies of its redox partnerships still remain enigmatic. CRISPR/Cas9 genome editing technology can be utilized for direct modification of a SNCA gene transcriptional activity, thus, leading to stable and long-term effect on gene expression. α-Syn transcriptional activity is regulated through wide range of genetic and epigenetic mechanisms, particularly DNA methylation patterns. It was shown that increased α-syn expression levels can occur due to demethylation of CpG dinucleotides at SNCA intron 1 [107]. Interestingly, brains of PD patients show hypomethylation at SNCA intron 1 and therefore, high levels of α-syn. Kantor et al. developed a system for target editing of SNCA methylation profiles within intron 1. This group established a lentiviral vector, harboring gRNA, an engineered nuclease null form Cas9 (dCas9) fused with the catalytic domain of DNA-methyltransferase 3A (DNMT3A) to target hypomethylated CpG islands in the SNCA intron 1 region. The system was successfully applied to human induced pluripotent stem cell (hiPSC)-derived DA neurons from PD patients with the SNCA gene triplication. The study showed downregulation of SNCA mRNA levels, 25% decreased protein abundance in hiPSC-derived DA neurons compared to controls without lentiviral construct. Moreover, CRISPR/dCas9 epigenome editing influenced PD-related phenotype. Results demonstrated increased cellular viability and mitochondrial functioning by alleviating the susceptibility to oxidative stress [24]. While mutations in SNCA gene are likely causative to PD progression, the exact contribution of α-syn to PD development is not fully understood.

3.3. Cardiolipin-driven mitophagy

Dependently on the context of mitochondrial injury, mitophagy can also occur independently of Parkin/Pink1 [108110]. One of the pathways triggering elimination of dysfunctional mitochondria is realized via signaling by a mitochondria-specific phospholipid, CL. Normally CL is synthesized and localized to the MIM, particularly its matrix leaflet. However, in depolarized mitochondria, CL is redistributed to the MOM [84]. This process includes CL transfers through the MIM, intermembrane space (IMS) and culminates in CL externalization on the mitochondrial surface where it is recognized by the microtubule associated protein 1 light chain 3 (LC3). The latter targets the injured mitochondria to autophagosomes and lysosomal degradation [111]. The transfer of CL through the IMS occurs with the assistance of the hexameric nucleoside diphosphate kinase D (NME4, NDPK-D, or NM23-H4) (Table 1), encoded by the NME4 gene (Fig. 2). RNAi knockdown of endogenous NDPK-D decreased CCCP-induced CL externalization and mitochondrial degradation [112]. Externalization of CL to the mitochondrial surface and enhanced mitophagy is triggered by a number of mitochondrial protonophoric uncouplers (such as CCCP) as well as an inhibitor of mitochondrial complex I, rotenone which is also known as a chemical PD inducer in experimental animals [83, 113].

Apart from CL binding and further externalization to MOM NDPK-D interacts with OPA1, a MIM-located dynamin-related GTPase, responsible for MIM fusion and mitochondrial quality control [114]. OPA1/NDPK-D interaction helps to fuel OPA1 with GTP to maintain OPA1 functions. Interestingly, OPA1 is a negative regulator of NME4/NDPK-D-supported CL transfer. Loss of NDP kinase activity induced by mitophagic stimuli weakens NME4/NDPK-D interaction with OPA1, thus releasing NME4/NDPK-D to crosslink MIM and MOM, facilitating CL transfer, resulting in the mitochondrial fragmentation observed during mitophagy [111, 114].

CL has the capacity to buffer synucleinopathy in SNCA mutated neurons (containing A53T and E46K mutations) by withdrawing α-syn monomers out of potentially pathological oligomers and fibrils [11]. α-Syn has been reported to bind to both mitochondrial membranes and mitochondria associated endoplasmic reticulum (ER) membranes [115, 116]. CL translocates to the MOM where it can facilitate α-syn refolding from aggregated β-sheet forms back to monomers comprising α-helices, effectively buffering synucleinopathy. The observed effects of CL on refolding of α-syn fibrils may be unique to CL-containing mitochondrial and mitochondria associated membranes. The increased abundance and prolonged CL exposure on the MOM in SNCA-mutants needed to refold mutant α-syn alter membrane dynamics and may initiate the depolarization of mitochondrial membranes, leading to increased LC3 recruitment to the MOM. Mutated α-syn had a significantly reduced ability to competitively inhibit LC3 binding to CL, as far as they compete for the same binding site on CL, thus increasing mitochondrial stress and triggering excessive mitophagy. Furthermore, co-culture of SNCA-mutant neurons with their isogenic controls results in transmission of mitochondrial pathology in non-mutant control neurons [11]. Thus, CL has two effects on α-syn: facilitating the formation of cyt c/CL-α-syn triple complexes with peroxidase activity capable of catalyzing prooxidant reactions, including those of covalent oligomerization of α-syn with cyt c and CL itself. These oligomers may become components of LBs relevant to the PD pathogenesis. Independently of this, CL by itself can suppress the formation of physical aggregates of α-syn thus counteracting the development of PD inducing reactions. One can imagine that CRISPR/Cas9 manipulated forms of α-syn as well as of cyt c may be instrumental in getting better insights in these conflicting roles of CL in PD development.

Intracellular quality control mechanisms are increasingly recognized as key factors in aging and NME4/NDPK-D-dependent mitophagy can also be an active player in this process [111]. Up to date no mechanistic studies on CL, NDPK-D in PD pathology have been conducted. The role of NDPK-D in CL redistribution and mitophageal signaling has been documented in animal models and cells treated with mitochondrial respiratory inhibitors (rotenone) or uncouplers (CCCP). However these results have not been confirmed on model animals that reproduce PD phenotype. The exact role of CL in modulation of α-syn pathology in relation to PD has not been established. It is also possible that abnormal α-syn itself may affect NME4/NDPK-D activity and facilitate CL translocation from MIM to MOM, thus promoting mitophagy.

CRISPR/Cas9 is a good tool to test whether NME4/NDPK-D activity is the only mechanism that promotes CL transgression to the MOM. Furthermore, creation of transgenic animal and cell models carrying various combinations of mutations in NDPK-D and α-syn may solve the puzzle of involvement of these genes and their protein products in PD pathogenesis.

3.4. ROS

Mitochondria synthesize ATP but also leak electrons to generate ROS, that are both cellular signaling molecules and damaging oxidants [117]. The generation and extensive accumulation of ROS in the mitochondria leads to the formation of mitochondrial PTP and mitochondrial membrane hyperpolarization. The damaged mitochondria then release the pro-apoptotic protein cyt c into the cytoplasm, where it activates nuclear fragmentation and caspase-3-dependent apoptosis [118]. In neurodegenerative diseases, ROS can cause damage to macromolecules such as proteins, lipids, polysaccharides, or nucleic acids in neurons. Abundance of fatty acids prone to peroxidation; high intracellular concentrations of transition metals (especially iron), capable of catalyzing the formation of reactive hydroxyl radicals; low levels of antioxidants; reduced capability to regenerate are the intrinsic properties of neurons, that make them highly vulnerable to the harmful effects of ROS. Furthermore, oxidative damage to proteins, lipids and DNA has been observed in post-mortem brain samples from PD patients [119].

3.5. Lipid mediators

Chronic pro-inflammatory immune activity is being increasingly associated with neurodegenerative disorders such as PD [14]. Oxidation of free polyunsaturated fatty acids (PUFAs) produces lipid mediators that exert diversified effects on the tissue homeostasis in health and disease [120]. Free PUFAs are normally esterified into cellular (phospho)lipids, but can be hydrolyzed and released by Ca-dependent phospholipases A2 (PLA2) and act as substrates for oxidation reactions catalyzed by several enzymes, especially lipoxygenases (LOXs), cyclooxygenases (COXs), cytochrome P450 isoforms and peroxidases [121]. A plethora of lipid mediators, including thromboxanes, pro-inflammatory prostaglandins and leukotrienes, regulators with both anti-inflammatory and proresolution activities, such as the lipoxins, resolvins and protectins are formed [122].

Recently a new pathway for selective generation of lipid mediators from a mitochondrial-specific phospholipid, CL, has been reported. This novel pathway localizes to mitochondria, and begins with the direct cyt c-catalyzed oxidation of PUFA-CL with subsequent hydrolysis of oxidation products, oxidized CL (CLox), by two types of Ca2+-independent phospholipase A2 (iPLA2) specific towards oxidatively modified phospholipids. The presence of sufficient amounts of oxidizing equivalents (H2O2 or lipid hydroperoxides) is essential for the peroxidase activity of cyt c/CL complexes and CL oxidation [120]. The brain has an unprecedented variety of CL species that can be utilized for the biosynthesis of diversified lipid mediators via cyt c-catalyzed oxidation processes [123]. Tyurina et al. demonstrated that this Ca2+-independent pathway, that triggers the oxidation of polyunsaturated CLs and accumulation of their hydrolysis products (oxidized linoleic, arachidonic acids and mono-lyso CLs), is activated in vivo after acute tissue injury [120].

Mitochondria are an active player in PD pathology. Long term exposure of mitochondria (especially those located in DA neurons of SN) to pro-oxidant conditions would cause CL externalization leading to its binding with cyt c to a peroxidase complex. Disrupted electron transport will inevitably generate superoxide anions that will dismutate directly or via Mn-SOD catalyzed reaction to yield H2O2. The latter can feed the peroxidase reaction of cyt c/CL complex to directly oxidize mitochondrial phospholipids, particularly CL. Among the oxidation products, there will be signals of apoptotic death (Fig. 3) as well as other oxygen-containing derivatives that can be hydrolyzed by Ca2+-independent PLA2γ (iPLA2γ) present in mitochondria (Fig. 3). These reactions will yield a plethora of pro- and anti-inflammatory lipid mediators – eicosanoids and docosanoids [120]. It is believed – but not experimentally proven – that these reactions may contribute to PD associated pathogenic processes.

Fig. 3. Schema showing a mitochondrial pathway for generation of lipid mediators.

Fig. 3.

Enzymatic oxidation of cardiolipin (CL) by cytochrome c (cyt c) in mitochondria during apoptosis and its hydrolysis by iPLA yields mono-lyso CL as well as a diversified series of pro-inflammatory and pro-resolving lipids mediators.

CRISPR/Cas9 protocol can be employed to test how Parkin, PINK1, Fbxo7, α-syn, CL, PLA2 and iPLA2 over-expression/downregulation/knockout affect the process of mitophagy, integrity of DA neurons and contribute to PD pathology.

4. Iron accumulation

Iron is essential for life as a part of many enzymes of oxidative metabolism. Yet free ionic iron or “loosely-bound” small molecular complexes of iron may be potentially toxic because of their very high redox activity, particularly with regards to their involvement in the poorly controlled production of ROS in oxygen-rich environments. Understanding mechanisms of iron metabolism in the brain is of great importance as iron may be involved in the pathogenesis of several neurodegenerative disorders.

Strictly controlled levels and activity of iron and ROS are crucial for normal cell and organismal function, although the aberrant accumulation of iron and excessive ROS production is recognized in a number of chronic degenerative conditions [124, 125]. In eukaryotic organisms, excessive mitochondrial iron-dependent production of ROS may result in cell death and damage to mitochondrial (mt)DNA by interfering with essential metabolic pathways [126].

Iron accumulation occurs in the brain tissues of aging animals, including humans, in areas primary responsible for the execution of motor and cognitive functions, particularly SN, globus pallidus and dentate nucleus [127]. The mechanisms underlying extensive iron accumulation in these structures likely involve changed balance between all major elements of the complex system fulfilling import, export, storage, redistribution and precise chaperon-mediated delivery of iron to the particular protein clients in neurons [128]. However, the exact knowledge about the role and contribution of individual components of this sophisticated machinery to the overall PD related iron dishomeostasis is incomplete [128]. Iron is believed to play a key role in PD pathogenesis. Excessive iron accumulation in SN of PD patients promotes death of DA neurons [9, 129, 130]. High sensitivity of these cells to various damaging agents can be related to the presence of dopamine, the major neurotransmitter released by DA neurons. Importantly, interaction between iron and dopamine can promote the production of various toxic metabolic intermediates and end-products (e.g. dopamine-o-quinone, amino-chrome, 5,6-indolequinone) that can be involved in redox-cycling reactions and cause mitochondrial dysfunction and α-syn oligomerization in DA neurons [131137]. Notably, in vitro studies implicated elevated iron levels in triggering α-syn aggregation to toxic fibrils that are present in LBs [137, 138].

Tau protein associates with amyloid precursor protein (APP) (Table 1) and regulates its trafficking in neurons where APP stabilizes surface presentation of transmembrane iron exporter, ferroportin (FPN) (Table 1), and facilitates iron efflux from neurons (Fig. 4) [139141]. Tau deficiency prevents APP from being transported to the neuronal surface that further leads to increase in intracellular APP levels in tau-deficient neurons, thus preventing it from interaction with FPN and impairs iron trafficking in DA neurons [141]. Iron accumulation within the neurons, especially DA neurons, involved in PD pathology, leads to neuronal death observed in in vivo and in vitro PD models and PD patients [140]. High amounts of insoluble tau levels with simultaneous reduction of soluble tau proteins is observed in familial and sporadic forms of PD even in individuals without dementia [140142]. In a recent study, Lei et al. reported that tau-knockout mice develop iron accumulation and SN neuronal loss, with concomitant cognitive deficits and Parkinsonism. Importantly, changes in tau levels and iron accumulation were specific to the SN and were absent from any other brain regions [141]. Interestingly, accumulation of pathological neurofibrillary tangles of the microtubule-associated protein tau has been shown to contribute to PD pathology [143]. Moreover, tau protein encoded by MAPT gene (Table 1) was shown to be a risk factor for late-onset sporadic PD, possibly due to its involvement in regulation of axonal transport, synaptic function, DNA stabilization and protection from heat damage and oxidative stress. In 2017 Hallmann et al. used CRISPR/Cas9 genome editing in stem cell model of fronto-temporal dementia (FTD) to repair FTD-associated mutation in the MAPT gene. For the FTD stem cell model researchers utilized hiPSC-derived neural progenitor cells (NPCs) from patients carrying N279K MAPT mutation. It has been demonstrated that FTD NPCs-differentiated astrocytes with mutation in the MAPT gene expressed increased levels of 4R-TAU isoforms and showed changes in the transcriptome profiles as well as higher vulnerability to oxidative stress compared to the control group [144].

Fig. 4. Iron homeostasis in the brain.

Fig. 4.

Iron uptake in neurons occur via transferrin receptor 1 (TfR1), it is followed by the internalization of Tf/TfR1 complex inside the cell and iron dissociation from Tf. Divalent metal transporter 1 (DMT1) facilitates iron transfer into the cytosol, where cytosolic iron chaperones of PCBP family, PCBP1 and PCBP2, transfers iron to the appropriate enzymes and proteins. Excess iron can be stored inside the cavity of extracellular protein, ferritin, or exported by transmembrane iron carrier, ferroportin (FPN). The amyloid precursor protein (APP) that is transported to the cell membrane by microtubule-associated protein tau is essential for FPN stabilization at the neuronal surface; therefore facilitating iron efflux from neuronal cells. Pathological accumulation of iron due to impairment in iron homeostasis can provoke iron-dependent oxidations of DNA, proteins and lipids that threaten cell health and integrity. Treatment with iron chelators might reduce excessive iron accumulation and restore iron balance in cells.

Ceruloplasmin (Cp) participates in iron transport by oxidizing Fe2+ to Fe3+ resented by FPN. Dysfunction of Cp is widely observed in PD-affected SN and suggests it may contribute to the pathogenic pro-oxidant iron accumulation [145]. Cp is also essential for transferrin (Tf) dependent delivery of iron into neuronal cells through that specific binding to the Tf receptors (TfR1) (Table 1) on the cell surface, which triggers endocytosis of Tf-TfR1 complex [146]. Endosomal iron can then be transported across the membrane to the cytosol by endosomal divalent metal transporter (DMT1) (Fig. 4) [147]. Tf is required for both iron uptake and efflux; when the cell requires iron for the proper functioning Tf binds to TfR1, promoting iron import into the cell, alternatively Tf removes iron from the FPN/Cp complex when the cell is iron replete and redistributes excess iron throughout the body [148]. Depletion of Tf in serum and SN has been described in PD as a contributor to iron accumulation in the brain. Furthermore, Tf is mis-compartmentalized in the mitochondria of SN neurons in PD that could further contribute to aberrant regulation of Tf in the brain [148].

In the cytosol, iron may be transported by binding to metallochaperones that deliver iron to target apoproteins [127]. One of these chaperones, poly(rC)-binding protein 1 (PCBP1), transfers iron to ferritin, an intracellular protein that detoxifies excess Fe2+ by oxidizing and storing the resultant Fe3+ inside its cavity [127, 149]. PCBP1-depleted cells show reduction in ferritin iron loading and simultaneous increase in cytosolic iron levels [150]. Another protein from the PCBP family, PCBP2, acquires ferrous iron from DMT 1 and transfers it to the appropriate enzymes and proteins in iron-PCBP1/PCBP2 complex. Importantly, PCBP2 facilitates iron export by binding to FPN [151]. PCBP1 and PCBP2 may function as iron chaperones for enzymes containing mono- and dinuclear iron centers, especially the family of non-heme iron-containing dioxygenases such as LOXs – enzymes involved in regulation of ferroptotic cell death program (see below). Depletion of one or both PCBPs corresponds with impaired iron delivery to the target cytosolic proteins [152].

Iron homeostasis in all mammalian cells is regulated by cytosolic iron response proteins (IRP½). In neurons IRP2 is the main sensor of labile intracellular iron, it binds to transcripts of ferritin, TfR1 and other target genes to control the expression of iron metabolizing proteins at the post-transcriptional level [9, 153]. In normal DA neurons and other cells under iron-depleted conditions iron-regulatory elements (IREs) (Table 1) located on the 3’-UTR of DMT1 and TfR1 mRNAs stabilize these mRNAs by IRP binding followed by enhanced translation of proteins. Alternatively, when intracellular iron levels are elevated, IRPs bind to IREs in the 5’-UTR of DMT 1, ferritin, APP and α-syn mRNAs, thus leading to repression of translation from these mRNAs [9]. Jiang et al., showed that impairment of IREs regulation through IRPs binding promoted DMT1 upregulation in the SN of 6-hydroxydopamine (6-OHDA)-induced PD rats. Overexpression of DMT1 caused increased iron influx, resulting in damaged mitochondrial function and increased ROS generation [154].

Iron chelators might reduce harmful iron-dependent oxidations of DNA, proteins and lipids [155]. For instance, clioquinol specifically reduces iron levels in the SN and prevents the onset of both motor dysfunction and cognitive decline in tau-deficient mice [141]. Unfortunately, treatment with iron chelators such as clioquinol, deferiprone or deferasirox could also remove iron from tissues unaffected by the disease or promote other side effects. Among the side-effects of iron-chelating compounds are nephrotoxicity, reduction of striatal dopamine, and neutropenia or agranulocytosis [148].

CRISPR/Cas9-based drug screening technologies may contribute to the development of new small molecule regulators that will target pathological iron accumulation in PD more selectively and effectively. Importantly, CRISPR/ Cas9 drug screening strategy can be applied for personalized treatment of PD patients. Based on unique array of genetic variants and specific disturbances of every PD patient, appropriate treatment strategies and medicines can be designed targeting the specific disease drivers in a given patient.

Overall, dysregulated iron homeostasis in neurons and glia of the SN is an established feature of PD. Iron homeostasis is regulated through multiple pathways and cellular mechanisms. APP and tau protein regulate FPN functioning and stabilize FPN in the cellular membrane to support iron export from the cell. Moreover, many other proteins, Tf, TfR1, DMT1, PCBPs and ferritin, are important for proper iron metabolism inside the cell. Interestingly, translational levels of these iron homeostatic proteins are regulated by intracellular iron levels and specific regulatory elements IRPs and IREs. The impairments in these mechanisms can lead to dysregulation of DMT 1, FPN, Tf, ferritin or APP translational activity and iron accumulation. CRISPR/Cas9 epigenome editing technology can be used for validation of the effects of up- and down-regulation of transcriptional and translational activity of iron homeostatic genes and their protein products in DA neurons. However, dysregulation of iron levels can also occur due to mutations in the reported genes, which can influence normal protein functioning and underlie the abnormalities in iron homeostasis. Disturbances in some of these proteins are essential and sufficient for development of PD phenotype, while impairments in others are just the consequences. However, which ones are really involved in PD is not clear. CRISPR/Cas9 multiple screening approaches can assist in separating the causative iron homeostatic genes associated with the PD pathology from those that are not directly related to the PD pathogenesis.

5. Programmed Cell Death

5.1. Apoptosis

Apoptotic death of DA neurons has been implicated in PD [156]. In the context of this review, we would like to focus on the mechanisms of intrinsic apoptosis because its early stages take place in mitochondria, strongly depend on changes in redox environment and include CL oxidation as a required stage of the execution of the entire program [157]. As has been described above, mitochondrial injury triggers trans-migration of unoxidized CL from the matrix side of the MIM to the mitochondrial surface thus signaling mitophagy and elimination of dysfunctional mitochondria. These pro-survival pathways may be effective in maintaining mitochondrial health within the cells. However, if the injurious factors continue to progress and the cell cannot cope with the intensity of the impact via mitophagy, the next initiated program is apoptotic elimination of the “high risk” cell continuously producing massive ROS. This stage is characterized by the formation of cyt c/CL complexes with a strong peroxidase activity towards bound CLs [10]. Within these complexes, CL induces substantial conformational changes and mobility of the protein such that hexa-coordinated iron loses one of the coordination bonds (Fe-S/Met80) and converts into a penta-coordinated state capable of binding small molecule oxidants, such as H2O2. The latter may be provided by the disrupted electron transport chain whereby cyt c interacting with CL obtains a highly negative redox potential precluding the acceptance of electrons from the respiratory complex III [158]. As a result, the flow of electron may be re-directed towards other acceptors, among them molecular oxygen to yield superoxide anion-radical. Spontaneous or MnSOD-catalyzed dismutation of the latter produces H2O2 that feeds the catalytic peroxidase cycle of cyt c/CL complexes and perpetuated CL peroxidation.

Several small molecule regulators such as mitochondria-targeted GS-nitroxides and imidazole-substituted fatty acids have been employed to effectively suppress the peroxidase CL oxidation and apoptosis [159]. However, their effectiveness in the context of PD has not been assessed. Because PD pathogenesis has been strongly associated with apoptosis and accumulation of specific CL peroxidation products (Fig. 5), it seems reasonable to suggest that optimized combination of small molecules inhibitors may be effective in anti-PD therapy. However, such a development may require detailed genetic studies unequivocally identifying all the participating protein-based mechanisms. These include above mentioned transmembrane transporters of CL to the mitochondrial surface, electron transporters generating and providing H2O2, kinases participating in phosphorylation of cyt c thus determining its affinity towards interactions with CL, protein mechanisms of CL biosynthesis and remodeling and numerous mitochondrial proteins capable of binding CL hence “out-compete” cyt c. Obviously, this extensive and time-consuming research may benefit from the utilization of the effective CRISPR/Cas9 technology.

Fig. 5. Cardiolipin oxidation products in substantia nigra.

Fig. 5.

Figure shows accumulation of specific cardiolipin oxidation (CLox) products in substantia nigra (SN) using rat rotenone model of PD [164], Full MS spectra of CL (left panel) and CLox (right panel) from SN of rats exposed to rotenone (10 days after treatment). Insert: Significantly higher levels of CLox molecular species were detected in SN of rotenone treated rats and associated with key pathological features of PD such as bradykinesia, postural instability/gait disturbances and rigidity.

5.2. Ferroptosis – Overview

Ferroptosis is a form of iron-dependent, oxidative cell death with specific role of lipid hydroperoxides accumulating in conditions of GPX4 insufficiency [160, 161]. Ferroptosis can be inhibited by lipophilic radical scavengers (eg, ferrostatin-1, different homologues of vitamin E, and Trolox), by inhibitors of LOX (eg, baicalein) and by iron chelators (deferoxamine) [161, 162]. Ferroptosis has been implicated in the pathological cell death associated with degenerative diseases, including PD [163]. The sensitivity to ferroptosis is strongly dependent on the access to oxidizable PUFA and their esterification in specific phospholipids, iron homeostasis, as well as on the sufficiency of the thiols.

Glutathione peroxidase 4 (GPX4) (Table 1) catalyzes the reduction of lipid peroxides to lipid alcohols. Inactivation of GPX4 through depletion of reduced glutathione (GSH) with erastin, or with a direct GPX4 inhibitor RSL3, ultimately induces intensive lipid peroxidation that causes cell death [165]. Hydroperoxy derivatives of PUFA-PEs also cause ferroptotic death when added to cells with inactivated GPX4 [166]. GSH, a cofactor of GPX4, is an essential intracellular antioxidant synthesized from glutamate, cysteine and glycine by cytosolic glutamate-cysteine ligase (GCL) and glutathione synthetase (GSS) (Fig. 6a) (Table 1). Due to the overwhelming importance of several amino acids for biological activity of GPX4, amino acid metabolism is tightly linked to the regulation of ferroptosis [167].

Fig. 6. Ferroptotic cell death in neurons.

Fig. 6.

a. Ferroptosis is triggered by accumulation of lipid hydroperoxides (PL-OOH). Glutathione peroxidase 4 (GPX4) catalyzes the reduction of lipid peroxides to lipid alcohols (PL-OH). Inactivation of GPX4 through depletion of reduced glutathione (GSH) with erastin, or with the direct GPX4 inhibitor RSL3, ultimately induces intensive lipid peroxidation that causes ferroptotic cell death. GSH, a cofactor of GPX4, is an essential intracellular antioxidant synthesized from glutamate (that is exchanged for cystine by system xc-), cysteine and glycine by cytosolic glutamate-cysteine ligase (GCL) and glutathione synthetase (GSS). Lipoxygenases (LOXs), nonheme, iron-containing proteins, normally utilizes free PUFAs as their substrates.

b. LOX substrate specificity is switched from free PUFAs to PUFA-containing polyunsaturated phosphatidylethanolamines (PEs). Phosphatidylethanolamine-binding protein 1 (PEBP1) complexes with 15-LOXs and changes their substrate competence to produce hydroperoxy-PE (PE-OOH). Inadequate reduction of PE-OOH due to dysfunction of GPX4 leads to ferroptosis.

Cysteine availability limits the biosynthesis of glutathione, furthermore, some cells can bypass the requirement for cystine import via the cystine/glutamate antiporter system xc- by converting methionine to cysteine through transsulfuration pathway [165]. Glutamate and glutamine are also important for ferroptotic cell death regulation [168]. Glutamate is exchanged for cystine by system xc-, so high extracellular concentrations of glutamate inhibit system xc- and induce ferroptosis [160]. Thus, accumulation of extracellular glutamate could serve as a natural trigger for inducing ferroptosis in physiological contexts. Free PUFAs are substrates for synthesis of lipid signaling mediators, yet they must be esterified into membrane phospholipids and undergo oxidation to become ferroptotic signals [166]. Previous studies suggest that phosphatidyl-ethanolamines (PEs), containing arachidonic acid (AA) or its elongation product, an isomer of docosatetraenoic acid, adrenic acid (AdA), are key phospholipids that undergo oxidation and provoke ferroptotic cell death [166, 169].

Acyl-CoA synthetase long-chain family member 4 (ACSL4) (Table 1) is involved in the biosynthesis and remodeling of PUFA-PEs in cellular membranes and, furthermore, promotes cell sensitivity to ferroptosis. Loss of function mutations in ACSL4 deplete the substrates for lipid peroxidation and increase resistance to ferroptosis [169, 170]. Alternatively, cells that are supplemented with arachidonic acid or other PUFAs are sensitized to ferroptosis [170].

LOXs (Table 1), non-heme, iron-containing proteins, utilize free PUFAs as their substrates (Fig. 6a) [171]. Interestingly, LOX substrate specificity can be switched from free PUFAs to PUFA-containing PEs (Fig. 6b). The mechanism that alters the substrate specificity of LOX remains to be enigmatic; however for 15-lipoxygenases (15-LOXs) the requirement in Phosphatidylethanolamine-Binding Protein 1 (PEBP1) has been demonstrated.PEBP1 forms complexes with 15-LO1 and 15LO2, two isoforms of 15LOX and changes their substrate competence to produce hydroperoxy-PE [172]. Notably, increased levels of lipid peroxidation products have been consistently observed in the SN of PD patients and in the brains of PD animal models [173, 174].

Several studies reported potential links between ferroptosis and the pathogenesis of neurodegenerative diseases. Inhibitors of ferroptosis, ferrostatins and liprox-statins, were found to be effective protectors against neurodegeneration in models of PD, Huntington’s, and Alzheimer’s disease models [163, 175, 176]. In line with the key role of GPX4 in ferroptosis, insufficiency of the enzyme – due to chemical inactivation, deletion or loss of function mutations - cause ferroptotic cell death characterized by extensive lipid peroxidation. Therefore, CRISPR/Cas9 genome/epigenome editing aimed at the restoration of the GPX4 activity might be beneficial. Another application of the CRISPR/Cas9 genome editing technology may be promising in studies of 15-LOX/PEBP1 interactions in determining selectivity and specificity in PUFA-PE oxidation and switching the substrate specificity from free to esterified PUFA. Oxidation of PUFA-PE by 15-LOXs is considered as the key mechanism that generates lipid signals of ferroptotic cell death. There are, however, studies that implicate other isoforms of LOXs (3-5-, 12-) in triggering ferroptotic death [177179]. CRISPR/Cas9 silencing or activation of respective LOX isoforms may give more quantitative description of their role and significance in lipid peroxidation, execution of ferroptosis and involvement in PD pathology.

Empirical research has established many potential metabolic abnormalities that may represent the specific key mechanisms of PD pathogenesis. However, the diversity of these findings and the lack in understanding the connections between them slow down the progress in the development of specific treatments. Several papers in a recent special issue of Science (December 14, 2018) “Illuminating the Brain” provided new insights into an integrative functional genomic/transcriptomics analysis, a powerful approach to gain comprehensive information on the entire the genome as well as epigenomic and transcriptomic features of the healthy and diseased (mostly psychiatric conditions) of human brain across various regions and cell types [180182]. The success of this approach in several neuropsychiatric conditions (e.g. schizophrenia, autism spectrum disorder (ASD), bipolar disorder) assures that that it will also be applied to PD patients. Simultaneously, new advancements based on CRISPR/Cas9 genome/epigenome editing protocols are being rapidly developed. As in sporadic PD, individual genomic variations are highly specific, patient-derived iPSCs may provide important personalized insights into the genetic basis of PD heritability, facilitate creation of patient-specific experimental cell models, drug discovery and develop new strategies for effective treatment [27, 183186].

Among many potentially important targets for CRISPR/Cas9 based research we focused in this brief review on aberrant redox mechanisms and pathways related to redox metabolism, mitochondria and regulated cell death programs that have been identified as strongly associated with the PD pathogenesis, yet without definitive conclusions leading to design of new therapies. With the appreciation of the significant role of the post-translational modifications of proteins and lipid protein interactions, we are optimistic that the power and relative simplicity of the CRISPR/Cas9 genome editing and screening protocols will generate new information on metabolic impairments essential for PD pathogenesis. This enthusiasm is supported by the fact that employment of genome-editing technology has already generated new animal models in which the introduction of custom-made modifications into the genome replicated key features of PD [26].

Highlights:

  • CRISPR/Cas9 genome editing in Parkinson disease

  • Redox lipidomics of signaling in cell death programs

  • Pathogenic impairments of iron regulation

  • New therapies of neurodegeneration

  • Cardiolipin signals of mitochondrial injury

7. Acknowledgements

This review article is supported by Russian academic excellence project “5-100” and by the Russian Science Foundation under grant # 17-15-01487 as well as NIH US grants (P01HL114453, U19AI068021,NS076511, NS061817, AG026389, NS101628-co-funded by NINDS and NIA).

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • [1].Hornykiewicz O, [Dopamine (3-hydroxytyramine) in the central nervous system and its relation to the Parkinson syndrome in man], Deutsche medizinische Wochenschrift 87 (1962) 1807–10. [DOI] [PubMed] [Google Scholar]
  • [2].Goedert M, Spillantini MG, Del Tredici K, Braak H, 100 years of Lewy pathology, Nature reviews. Neurology 9(1) (2013) 13–24. [DOI] [PubMed] [Google Scholar]
  • [3].Irwin DJ, Lee VM, Trojanowski JQ, Parkinson’s disease dementia: convergence of alpha-synuclein, tau and amyloid-beta pathologies, Nature reviews. Neuroscience 14(9) (2013) 626–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [4].de Lau LM, Breteler MM, Epidemiology of Parkinson’s disease, The Lancet. Neurology 5(6) (2006) 525–35. [DOI] [PubMed] [Google Scholar]
  • [5].Hoang QQ, Pathway for Parkinson disease, Proceedings of the National Academy of Sciences of the United States of America 111(7) (2014) 2402–3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [6].Spillantini MG, Schmidt ML, Lee VM, Trojanowski JQ, Jakes R, Goedert M, Alpha-synuclein in Lewy bodies, Nature 388(6645) (1997) 839–40. [DOI] [PubMed] [Google Scholar]
  • [7].Hickman S, Izzy S, Sen P, Morsett L, El Khoury J, Microglia in neurodegeneration, Nature neuroscience 21(10) (2018) 1359–1369. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [8].Dauer W, Przedborski S, Parkinson’s disease: mechanisms and models, Neuron 39(6) (2003) 889–909. [DOI] [PubMed] [Google Scholar]
  • [9].Masaldan S, Bush AI, Devos D, Rolland AS, Moreau C, Striking while the iron is hot: Iron metabolism and ferroptosis in neurodegeneration, Free radical biology & medicine (2018). [DOI] [PubMed] [Google Scholar]
  • [10].Kagan VE, Tyurin VA, Jiang J, Tyurina YY, Ritov VB, Amoscato AA, Osipov AN, Belikova NA, Kapralov AA, Kini V, Vlasova II, Zhao Q, Zou M, Di P, Svistunenko DA, Kurnikov IV, Borisenko GG, Cytochrome c acts as a cardiolipin oxygenase required for release of proapoptotic factors, Nature chemical biology 1 (4) (2005) 223–32. [DOI] [PubMed] [Google Scholar]
  • [11].Ryan T, Bamm VV, Stykel MG, Coackley CL, Humphries KM, Jamieson-Williams R, Ambasudhan R, Mosser DD, Lipton SA, Harauz G, Ryan SD, Cardiolipin exposure on the outer mitochondrial membrane modulates alpha-synuclein, Nature communications 9(1) (2018) 817. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [12].Ludtmann MHR, Angelova PR, Horrocks MH, Choi ML, Rodrigues M, Baev AY, Berezhnov AV, Yao Z, Little D, Banushi B, Al-Menhali AS, Ranasinghe RT, Whiten DR, Yapom R, Dolt KS, Devine MJ, Gissen P, Kunath T, Jaganjac M, Pavlov EV, Klenerman D, Abramov AY, Gandhi S, alpha-synuclein oligomers interact with ATP synthase and open the permeability transition pore in Parkinson’s disease, Nature communications 9(1) (2018) 2293. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [13].Basil MC, Levy BD, Specialized pro-resolving mediators: endogenous regulators of infection and inflammation, Nature reviews. Immunology 16(1) (2016) 51–67. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [14].Abeliovich A, Gitler AD, Defects in trafficking bridge Parkinson’s disease pathology and genetics, Nature 539(7628) (2016) 207–216. [DOI] [PubMed] [Google Scholar]
  • [15].Connolly BS, Lang AE, Pharmacological treatment of Parkinson disease: a review, Jama 311(16) (2014) 1670–83. [DOI] [PubMed] [Google Scholar]
  • [16].Grealish S, Diguet E, Kirkeby A, Mattsson B, Heuer A, Bramoulle Y, Van Camp N, Perrier AL, Hantraye P, Bjorklund A, Parmar M, Human ESC-derived dopamine neurons show similar preclinical efficacy and potency to fetal neurons when grafted in a rat model of Parkinson’s disease, Cell stem cell 15(5) (2014) 653–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [17].Kriks S, Shim JW, Piao J, Ganat YM, Wakeman DR, Xie Z, Carrillo-Reid L, Auyeung G, Antonacci C, Buch A, Yang L, Beal MF, Surmeier DJ, Kordower JH, Tabar V, Studer L, Dopamine neurons derived from human ES cells efficiently engraft in animal models of Parkinson’s disease, Nature 480(7378) (2011) 547–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [18].Perrier AL, Tabar V, Barberi T, Rubio ME, Bruses J, Topf N, Harrison NL, Studer L, Derivation of midbrain dopamine neurons from human embryonic stem cells, Proceedings of the National Academy of Sciences of the United States of America 101(34) (2004) 12543–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [19].Wang YK, Zhu WW, Wu MH, Wu YH, Liu ZX, Liang LM, Sheng C, Hao J, Wang L, Li W, Zhou Q, Hu BY, Human Clinical-Grade Parthenogenetic ESC-Derived Dopaminergic Neurons Recover Locomotive Defects of Nonhuman Primate Models of Parkinson’s Disease, Stem cell reports 11(1) (2018) 171–182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [20].Adli M, The CRISPR tool kit for genome editing and beyond, Nature communications 9(1) (2018) 1911. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [21].Hsu PD, Lander ES, Zhang F, Development and applications of CRISPR-Cas9 for genome engineering, Cell 157(6) (2014) 1262–78. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [22].Sander JD, Joung JK, CRISPR-Cas systems for editing, regulating and targeting genomes, Nature biotechnology 32(4) (2014) 347–55. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [23].Heman-Ackah SM, Manzano R, Hoozemans JJM, Scheper W, Flynn R, Haerty W, Cowley SA, Bassett AR, Wood MJA, Alpha-synuclein induces the unfolded protein response in Parkinson’s disease SNCA triplication iPSC-derived neurons, Human molecular genetics 26(22) (2017) 4441–4450. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [24].Kantor B, Tagliafierro L, Gu J, Zamora ME, Ilich E, Grenier C, Huang ZY, Murphy S, Chiba-Falek O, Downregulation of SNCA Expression by Targeted Editing of DNA Methylation: A Potential Strategy for Precision Therapy in PD, Molecular therapy : the journal of the American Society of Gene Therapy 26(11) (2018) 2638–2649. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [25].Kim S, Yun SP, Lee S, Umanah GE, Bandaru VVR, Yin X, Rhee P, Karuppagounder SS, Kwon SH, Lee H, Mao X, Kim D, Pandey A, Lee G, Dawson VL, Dawson TM, Ko HS, GBA1 deficiency negatively affects physiological alpha-synuclein tetramers and related multimers, Proceedings of the National Academy of Sciences of the United States of America 115(4) (2018) 798–803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [26].Cota-Coronado JA, Sandoval-Avila S, Gaytan-Davila YP, Diaz NF, Vega-Ruiz B, Padilla-Camberos E, Diaz-Martinez NE, New transgenic models of Parkinson’s disease using genome editing technology, Neurologia (2017). [DOI] [PubMed] [Google Scholar]
  • [27].Calatayud C, Carola G, Consiglio A, Raya A, Modeling the genetic complexity of Parkinson’s disease by targeted genome edition in iPS cells, Current opinion in genetics & development 46 (2017) 123–131. [DOI] [PubMed] [Google Scholar]
  • [28].Bonthron DT, Foulkes WD, Genetics meets pathology - an increasingly important relationship, The Journal of pathology 241 (2) (2017) 119–122. [DOI] [PubMed] [Google Scholar]
  • [29].Tagliafierro L, Chiba-Falek O, Up-regulation of SNCA gene expression: implications to synucleinopathies, Neurogenetics 17(3) (2016) 145–57. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [30].Mojica FJ, Diez-Villasenor C, Garcia-Martinez J, Soria E, Intervening sequences of regularly spaced prokaryotic repeats derive from foreign genetic elements, Journal of molecular evolution 60(2) (2005) 174–82. [DOI] [PubMed] [Google Scholar]
  • [31].Bolotin A, Quinquis B, Sorokin A, Ehrlich SD, Clustered regularly interspaced short palindrome repeats (CRISPRs) have spacers of extrachromosomal origin, Microbiology 151 (Pt 8) (2005) 2551–61. [DOI] [PubMed] [Google Scholar]
  • [32].Jansen R, Embden JD, Gaastra W, Schouls LM, Identification of genes that are associated with DNA repeats in prokaryotes, Molecular microbiology 43(6) (2002) 1565–75. [DOI] [PubMed] [Google Scholar]
  • [33].Mojica FJ, Diez-Villasenor C, Soria E, Juez G, Biological significance of a family of regularly spaced repeats in the genomes of Archaea, Bacteria and mitochondria, Molecular microbiology 36(1) (2000) 244–6. [DOI] [PubMed] [Google Scholar]
  • [34].Makarova KS, Haft DH, Barrangou R, Brouns SJ, Charpentier E, Horvath P, Moineau S, Mojica FJ, Wolf YI, Yakunin AF, van der Oost J, Koonin EV, Evolution and classification of the CRISPR-Cas systems, Nature reviews. Microbiology 9(6) (2011) 467–77. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [35].Taketani Y, Kitamoto K, Sakisaka T, Kimakura M, Toyono T, Yamagami S, Amano S, Kuroda M, Moore T, Usui T, Ouchi Y, Repair of the TGFBI gene in human corneal keratocytes derived from a granular corneal dystrophy patient via CRISPR/Cas9-induced homology-directed repair, Scientific reports 7(1) (2017) 16713. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [36].Leidy-Davis T, Cheng K, Goodwin LO, Morgan JL, Juan WC, Roca X, Ong ST, Bergstrom DE, Viable Mice with Extensive Gene Humanization (25-kbp) Created Using Embryonic Stem Cell/Blastocyst and CRISPR/Zygote Injection Approaches, Scientific reports 8(1) (2018) 15028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [37].Jinek M, Chylinski K, Fonfara I, Hauer M, Doudna JA, Charpentier E, A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity, Science 337(6096) (2012) 816–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [38].Sapranauskas R, Gasiunas G, Fremaux C, Barrangou R, Horvath P, Siksnys V, The Streptococcus thermophilus CRISPR/Cas system provides immunity in Escherichia coli, Nucleic acids research 39(21) (2011) 9275–82. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [39].Mojica FJ, Diez-Villasenor C, Garcia-Martinez J, Almendros C, Short motif sequences determine the targets of the prokaryotic CRISPR defence system, Microbiology 155(Pt 3) (2009) 733–40. [DOI] [PubMed] [Google Scholar]
  • [40].Deveau H, Barrangou R, Garneau JE, Labonte J, Fremaux C, Boyaval P, Romero DA, Horvath P, Moineau S, Phage response to CRISPR-encoded resistance in Streptococcus thermophilus, Journal of bacteriology 190(4) (2008) 1390–400. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [41].Garneau JE, Dupuis ME, Villion M, Romero DA, Barrangou R, Boyaval P, Fremaux C, Horvath P, Magadan AH, Moineau S, The CRISPR/Cas bacterial immune system cleaves bacteriophage and plasmid DNA, Nature 468(7320) (2010) 67–71. [DOI] [PubMed] [Google Scholar]
  • [42].Bialk P, Sansbury B, Rivera-Torres N, Bloh K, Man D, Kmiec EB, Analyses of point mutation repair and allelic heterogeneity generated by CRISpR/Cas9 and single-stranded DNA oligonucleotides, Scientific reports 6 (2016) 32681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [43].Jiang W, Liu L, Chang Q, Xing F, Ma Z, Fang Z, Zhou J, Fu L, Wang H, Huang X, Chen X, Li Y, Li S, Production of Wilson Disease Model Rabbits with Homology-Directed Precision Point Mutations in the ATP7B Gene Using the CRISPR/Cas9 System, Scientific reports 8(1) (2018) 1332. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [44].Yan S, Tu Z, Liu Z, Fan N, Yang H, Yang S, Yang W, Zhao Y, Ouyang Z, Lai C, Yang H, Li L, Liu Q, Shi H, Xu G, Zhao H, Wei H, Pei Z, Li S, Lai L, Li XJ, A Huntingtin Knockin Pig Model Recapitulates Features of Selective Neurodegeneration in Huntington’s Disease, Cell 173(4) (2018) 989–1002 e13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [45].Shinmyo Y, Tanaka S, Tsunoda S, Hosomichi K, Tajima A, Kawasaki H, CRISPR/Cas9-mediated gene knockout in the mouse brain using in utero electroporation, Scientific reports 6 (2016) 20611. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [46].Kwon DY, Zhao YT, Lamonica JM, Zhou Z, Locus-specific histone deacetylation using a synthetic CRISPR-Cas9-based HDAC, Nature communications 8 (2017) 15315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [47].Zheng Y, Shen W, Zhang J, Yang B, Liu YN, Qi H, Yu X, Lu SY, Chen Y, Xu YZ, Li Y, Gage FH, Mi S, Yao J, CRISPR interference-based specific and efficient gene inactivation in the brain, Nature neuroscience 21(3) (2018) 447–454. [DOI] [PubMed] [Google Scholar]
  • [48].Zhou H, Liu J, Zhou C, Gao N, Rao Z, Li H, Hu X, Li C, Yao X, Shen X, Sun Y, Wei Y, Liu F, Ying W, Zhang J, Tang C, Zhang X, Xu H, Shi L, Cheng L, Huang P, Yang H, In vivo simultaneous transcriptional activation of multiple genes in the brain using CRISPR-dCas9-activator transgenic mice, Nature neuroscience 21(3) (2018) 440–446. [DOI] [PubMed] [Google Scholar]
  • [49].Miller JC, Holmes MC, Wang J, Guschin DY, Lee YL, Rupniewski I, Beausejour CM, Waite AJ, Wang NS, Kim KA, Gregory PD, Pabo CO, Rebar EJ, An improved zinc-finger nuclease architecture for highly specific genome editing, Nature biotechnology 25(7) (2007) 778–85. [DOI] [PubMed] [Google Scholar]
  • [50].Wood AJ, Lo TW, Zeitler B, Pickle CS, Ralston EJ, Lee AH, Amora R, Miller JC, Leung E, Meng X, Zhang L, Rebar EJ, Gregory PD, Urnov FD, Meyer BJ, Targeted genome editing across species using ZFNs and TAlEns , Science 333(6040) (2011) 307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [51].Reyon D, Tsai SQ, Khayter C, Foden JA, Sander JD, Joung JK, FLASH assembly of TALENs for high-throughput genome editing, Nature biotechnology 30(5) (2012) 460–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [52].Hockemeyer D, Wang H, Kiani S, Lai CS, Gao Q, Cassady JP, Cost GJ, Zhang L, Santiago Y, Miller JC, Zeitler B, Cherone JM, Meng X, Hinkley SJ, Rebar EJ, Gregory PD, Urnov FD, Jaenisch R, Genetic engineering of human pluripotent cells using TALE nucleases, Nature biotechnology 29(8) (2011) 731–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [53].Bonifati V, Rizzu P, van Baren MJ, Schaap O, Breedveld GJ, Krieger D, Dekker MC, Squitieri F, Ibanez P, Joosse M, van Dongen JW, Vanacore N, van Swieten JC, Brice A, Meco G, van Duijn CM, Oostra BA, Heutink P, Mutations in the DJ-1 gene associated with autosomal recessive early-onset parkinsonism, Science 299(5604) (2003) 256–9. [DOI] [PubMed] [Google Scholar]
  • [54].Dehay B, Ramirez A, Martinez-Vicente M, Perier C, Canron MH, Doudnikoff E, Vital A, Vila M, Klein C, Bezard E, Loss of P-type ATPase ATP13A2/PARK9 function induces general lysosomal deficiency and leads to Parkinson disease neurodegeneration, Proceedings of the National Academy of Sciences of the United States of America 109(24) (2012) 9611–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [55].Di Fonzo A, Dekker MC, Montagna P, Baruzzi A, Yonova EH, Correia Guedes L, Szczerbinska A, Zhao T, Dubbel-Hulsman LO, Wouters CH, de Graaff E, Oyen WJ, Simons EJ, Breedveld GJ, Oostra AA, Horstink MW, Bonifati V, FBXO7 mutations cause autosomal recessive, early-onset parkinsonian-pyramidal syndrome, Neurology 72(3) (2009) 240–5. [DOI] [PubMed] [Google Scholar]
  • [56].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 392(6676) (1998) 605–8. [DOI] [PubMed] [Google Scholar]
  • [57].Kruger R, Kuhn W, Muller T, Woitalla D, Graeber M, Kosel S, Przuntek H, Epplen JT, Schols L, Riess O, Ala30Pro mutation in the gene encoding alpha-synuclein in Parkinson’s disease, Nature genetics 18(2) (1998) 106–8. [DOI] [PubMed] [Google Scholar]
  • [58].Klein C, Westenberger A, Genetics of Parkinson’s disease, Cold Spring Harbor perspectives in medicine 2(1) (2012) a008888. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [59].Mandemakers W, Snellinx A, O’Neill MJ, de Strooper B, LRRK2 expression is enriched in the striosomal compartment of mouse striatum, Neurobiology of disease 48(3) (2012) 582–93. [DOI] [PubMed] [Google Scholar]
  • [60].Polymeropoulos MH, Lavedan C, Leroy E, Ide SE, Dehejia A, Dutra A, Pike B, Root H, Rubenstein J, Boyer R, Stenroos ES, Chandrasekharappa S, Athanassiadou A, Papapetropoulos T, Johnson WG, Lazzarini AM, Duvoisin RC, Di Iorio G, Golbe LI, Nussbaum RL, Mutation in the alpha-synuclein gene identified in families with Parkinson’s disease, Science 276(5321) (1997) 2045–7. [DOI] [PubMed] [Google Scholar]
  • [61].Przedborski S, The two-century journey of Parkinson disease research, Nature reviews. Neuroscience 18(4) (2017) 251–259. [DOI] [PubMed] [Google Scholar]
  • [62].Satake W, Nakabayashi Y, Mizuta I, Hirota Y, Ito C, Kubo M, Kawaguchi T, Tsunoda T, Watanabe M, Takeda A, Tomiyama H, Nakashima K, Hasegawa K, Obata F, Yoshikawa T, Kawakami H, Sakoda S, Yamamoto M, Hattori N, Murata M, Nakamura Y, Toda T, Genome-wide association study identifies common variants at four loci as genetic risk factors for Parkinson’s disease, Nature genetics 41 (12) (2009) 1303–7. [DOI] [PubMed] [Google Scholar]
  • [63].Simon-Sanchez J, Schulte C, Bras JM, Sharma M, Gibbs JR, Berg D, Paisan-Ruiz C, Lichtner P, Scholz SW, Hernandez DG, Kruger R, Federoff M, Klein C, Goate A, Perlmutter J, Bonin M, Nalls MA, Illig T, Gieger C, Houlden H, Steffens M, Okun MS, Racette BA, Cookson MR, Foote KD, Fernandez HH, Traynor BJ, Schreiber S, Arepalli S, Zonozi R, Gwinn K, van der Brug M, Lopez G, Chanock SJ, Schatzkin A, Park Y, Hollenbeck A, Gao J, Huang X, Wood NW, Lorenz D, Deuschl G, Chen H, Riess O, Hardy JA, Singleton AB, Gasser T, Genome-wide association study reveals genetic risk underlying Parkinson’s disease, Nature genetics 41 (12) (2009) 1308–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [64].Friedman JR, Nunnari J, Mitochondrial form and function, Nature 505(7483) (2014) 335–43. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [65].Youle RJ, van der Bliek AM, Mitochondrial fission, fusion, and stress, Science 337(6098) (2012) 1062–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [66].Quiros PM, Mottis A, Auwerx J, Mitonuclear communication in homeostasis and stress, Nature reviews. Molecular cell biology 17(4) (2016) 213–26. [DOI] [PubMed] [Google Scholar]
  • [67].Zhu JH, Guo F, Shelburne J, Watkins S, Chu CT, Localization of phosphorylated ERK/MAP kinases to mitochondria and autophagosomes in Lewy body diseases, Brain pathology 13(4) (2003) 473–81. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [68].Palikaras K, Lionaki E, Tavernarakis N, Mechanisms of mitophagy in cellular homeostasis, physiology and pathology, Nature cell biology 20(9) (2018) 1013–1022. [DOI] [PubMed] [Google Scholar]
  • [69].Ashrafi G, Schwarz TL, The pathways of mitophagy for quality control and clearance of mitochondria, Cell death and differentiation 20(1) (2013) 31–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [70].Ryan SD, Dolatabadi N, Chan SF, Zhang X, Akhtar MW, Parker J, Soldner F, Sunico CR, Nagar S, Talantova M, Lee B, Lopez K, Nutter A, Shan B, Molokanova E, Zhang Y, Han X, Nakamura T, Masliah E, Yates JR 3rd, Nakanishi N, Andreyev AY, Okamoto S, Jaenisch R, Ambasudhan R, Lipton SA, Isogenic human iPSC Parkinson’s model shows nitrosative stress-induced dysfunction in MEF2-PGC1alpha transcription, Cell 155(6) (2013) 1351–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [71].Clark IE, Dodson MW, Jiang C, Cao JH, Huh JR, Seol JH, Yoo SJ, Hay BA, Guo M, Drosophila pink1 is required for mitochondrial function and interacts genetically with parkin, Nature 441(7097) (2006) 1162–6. [DOI] [PubMed] [Google Scholar]
  • [72].Valente EM, Abou-Sleiman PM, Caputo V, Muqit MM, Harvey K, Gispert S, Ali Z, Del Turco D, Bentivoglio AR, Healy DG, Albanese A, Nussbaum R, Gonzalez-Maldonado R, Deller T, Salvi S, Cortelli P, Gilks WP, Latchman DS, Harvey RJ, Dallapiccola B, Auburger G, Wood NW, Hereditary early-onset Parkinson’s disease caused by mutations in PINK1, Science 304(5674) (2004) 1158–60. [DOI] [PubMed] [Google Scholar]
  • [73].Delgado-Camprubi M, Esteras N, Soutar MP, Plun-Favreau H, Abramov AY, Deficiency of Parkinson’s disease-related gene Fbxo7 is associated with impaired mitochondrial metabolism by PARP activation, Cell death and differentiation 24(1) (2017) 120–131. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [74].Pickles S, Vigie P, Youle RJ, Mitophagy and Quality Control Mechanisms in Mitochondrial Maintenance, Current biology : CB 28(4) (2018) R170–R185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [75].Harper JW, Ordureau A, Heo JM, Building and decoding ubiquitin chains for mitophagy, Nature reviews. Molecular cell biology 19(2) (2018) 93–108. [DOI] [PubMed] [Google Scholar]
  • [76].Hasson SA, Kane LA, Yamano K, Huang CH, Sliter DA, Buehler E, Wang C, Heman-Ackah SM, Hessa T, Guha R, Martin SE, Youle RJ, High-content genome-wide RNAi screens identify regulators of parkin upstream of mitophagy, Nature 504(7479) (2013) 291–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [77].Matsuda N, Sato S, Shiba K, Okatsu K, Saisho K, Gautier CA, Sou YS, Saiki S, Kawajiri S, Sato F, Kimura M, Komatsu M, Hattori N, Tanaka K, PINK1 stabilized by mitochondrial depolarization recruits Parkin to damaged mitochondria and activates latent Parkin for mitophagy, The Journal of cell biology 189(2) (2010) 211–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [78].Kane LA, Lazarou M, Fogel AI, Li Y, Yamano K, Sarraf SA, Banerjee S, Youle RJ, PINK1 phosphorylates ubiquitin to activate Parkin E3 ubiquitin ligase activity, The Journal of cell biology 205(2) (2014) 143–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [79].Aguirre JD, Dunkerley KM, Mercier P, Shaw GS, Structure of phosphorylated UBL domain and insights into PINK1-orchestrated parkin activation, Proceedings of the National Academy of Sciences of the United States of America 114(2) (2017) 298–303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [80].Yang Y, Gehrke S, Imai Y, Huang Z, Ouyang Y, Wang JW, Yang L, Beal MF, Vogel H, Lu B, Mitochondrial pathology and muscle and dopaminergic neuron degeneration caused by inactivation of Drosophila Pink1 is rescued by Parkin, Proceedings of the National Academy of Sciences of the United States of America 103(28) (2006) 10793–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [81].Blesa J, Przedborski S, Parkinson’s disease: animal models and dopaminergic cell vulnerability, Frontiers in neuroanatomy 8 (2014) 155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [82].Chu CT, Mechanisms of selective autophagy and mitophagy: Implications for neurodegenerative diseases, Neurobiology of disease 122 (2019) 23–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [83].Chu CT, Ji J, Dagda RK, Jiang JF, Tyurina YY, Kapralov AA, Tyurin VA, Yanamala N, Shrivastava IH, Mohammadyani D, Wang KZQ, Zhu J, Klein-Seetharaman J, Balasubramanian K, Amoscato AA, Borisenko G, Huang Z, Gusdon AM, Cheikhi A, Steer EK, Wang R, Baty C, Watkins S, Bahar I, Bayir H, Kagan VE, Cardiolipin externalization to the outer mitochondrial membrane acts as an elimination signal for mitophagy in neuronal cells, Nature cell biology 15(10) (2013) 1197–1205. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [84].McWilliams TG, Prescott AR, Montava-Garriga L, Ball G, Singh F, Barini E, Muqit MMK, Brooks SP, Ganley IG, Basal Mitophagy Occurs Independently of PINK1 in Mouse Tissues of High Metabolic Demand, Cell metabolism 27(2) (2018) 439–449 e5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [85].Lee JJ, Sanchez-Martinez A, Zarate AM, Beninca C, Mayor U, Clague MJ, Whitworth AJ, Basal mitophagy is widespread in Drosophila but minimally affected by loss of Pink1 or parkin, The Journal of cell biology 217(5) (2018) 1613–1622. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [86].Dawson TM, Ko HS, Dawson VL, Genetic animal models of Parkinson’s disease, Neuron 66(5) (2010) 646–61. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [87].Kitada T, Pisani A, Porter DR, Yamaguchi H, Tscherter A, Martella G, Bonsi P, Zhang C, Pothos EN, Shen J, Impaired dopamine release and synaptic plasticity in the striatum of PINK1-deficient mice, Proceedings of the National Academy of Sciences of the United States of America 104(27) (2007) 11441–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [88].Zhou X, Xin J, Fan N, Zou Q, Huang J, Ouyang Z, Zhao Y, Zhao B, Liu Z, Lai S, Yi X, Guo L, Esteban MA, Zeng Y, Yang H, Lai L, Generation of CRISPR/Cas9-mediated gene-targeted pigs via somatic cell nuclear transfer, Cellular and molecular life sciences : CMLS 72(6) (2015) 1175–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [89].Yang W, Liu Y, Tu Z, Xiao C, Yan S, Ma X, Guo X, Chen X, Yin P, Yang Z, Yang S, Jiang T, Li S, Qin C, Li XJ, CRISPR/Cas9-mediated PINK1 deletion leads to neurodegeneration in rhesus monkeys, Cell research 29(4) (2019) 334–336. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [90].Potting C, Crochemore C, Moretti F, Nigsch F, Schmidt I, Manneville C, Carbone W, Knehr J, DeJesus R, Lindeman A, Maher R, Russ A, McAllister G, Reece-Hoyes JS, Hoffman GR, Roma G, Muller M, Sailer AW, Helliwell SB, Genome-wide CRISPR screen for PARKIN regulators reveals transcriptional repression as a determinant of mitophagy, Proceedings of the National Academy of Sciences of the United States of America 115(2) (2018) E180–E189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [91].Cookson MR, DJ-1, PINK1, and their effects on mitochondrial pathways, Movement disorders : official journal of the Movement Disorder Society 25 Suppl 1 (2010) S44–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [92].Lopert P, Patel M, Brain mitochondria from DJ-1 knockout mice show increased respiration-dependent hydrogen peroxide consumption, Redox biology 2 (2014) 667–72. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [93].Ariga H, Takahashi-Niki K, Kato I, Maita H, Niki T, Iguchi-Ariga SM, Neuroprotective function of DJ-1 in Parkinson’s disease, Oxidative medicine and cellular longevity 2013 (2013) 683920. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [94].Hao LY, Giasson BI, Bonini NM, DJ-1 is critical for mitochondrial function and rescues PINK1 loss of function, Proceedings of the National Academy of Sciences of the United States of America 107(21) (2010) 9747–52. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [95].Inden M, Taira T, Kitamura Y, Yanagida T, Tsuchiya D, Takata K, Yanagisawa D, Nishimura K, Taniguchi T, Kiso Y, Yoshimoto K, Agatsuma T, Koide-Yoshida S, Iguchi-Ariga SM, Shimohama S, Ariga H, PARK7 DJ-1 protects against degeneration of nigral dopaminergic neurons in Parkinson’s disease rat model, Neurobiology of disease 24(1) (2006) 144–58. [DOI] [PubMed] [Google Scholar]
  • [96].Zhou W, Freed CR, DJ-1 up-regulates glutathione synthesis during oxidative stress and inhibits A53T alpha-synuclein toxicity, The Journal of biological chemistry 280(52) (2005) 43150–8. [DOI] [PubMed] [Google Scholar]
  • [97].Zondler L, Miller-Fleming L, Repici M, Goncalves S, Tenreiro S, Rosado-Ramos R, Betzer C, Straatman KR, Jensen PH, Giorgini F, Outeiro TF, DJ-1 interactions with alpha-synuclein attenuate aggregation and cellular toxicity in models of Parkinson’s disease, Cell death & disease 5 (2014) e1350. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [98].Burchell VS, Nelson DE, Sanchez-Martinez A, Delgado-Camprubi M, Ivatt RM, Pogson JH, Randle SJ, Wray S, Lewis PA, Houlden H, Abramov AY, Hardy J, Wood NW, Whitworth AJ, Laman H, Plun-Favreau H, The Parkinson’s disease-linked proteins Fbxo7 and Parkin interact to mediate mitophagy, Nature neuroscience 16(9) (2013) 1257–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [99].Narendra D, Tanaka A, Suen DF, Youle RJ, Parkin is recruited selectively to impaired mitochondria and promotes their autophagy, The Journal of cell biology 183(5) (2008) 795–803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [100].Wang ESKZQ, Otero PA, Bateman NW, Cheng MH, Scott AL, Wu C, Bahar I, Shih Y-T, Hsueh Y-P, Chu CT, PINK1 Interacts with VCP/p97 and Activates PKA to Promote nSfL1 C/p47 Phosphorylation and Dendritic Arborization in Neurons, eneuro 5(6) (2018) ENEURO.0466-18.2018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [101].Verma M, Wills Z, Chu CT, Excitatory Dendritic Mitochondrial Calcium Toxicity: Implications for Parkinson’s and Other Neurodegenerative Diseases, Frontiers in neuroscience 12 (2018) 523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [102].Tsai PI, Lin CH, Hsieh CH, Papakyrikos AM, Kim MJ, Napolioni V, Schoor C, Couthouis J, Wu RM, Wszolek ZK, Winter D, Greicius MD, Ross OA, Wang X, PINK1 Phosphorylates MIC60/Mitofilin to Control Structural Plasticity of Mitochondrial Crista Junctions, Molecular cell 69(5) (2018) 744–756 e6. [DOI] [PubMed] [Google Scholar]
  • [103].Cremades N, Cohen SI, Deas E, Abramov AY, Chen AY, Orte A, Sandal M, Clarke RW, Dunne P, Aprile FA, Bertoncini CW, Wood NW, Knowles TP, Dobson CM, Klenerman D, Direct observation of the interconversion of normal and toxic forms of alpha-synuclein, Cell 149(5) (2012) 1048–59. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [104].Angelova PR, Horrocks MH, Klenerman D, Gandhi S, Abramov AY, Shchepinov MS, Lipid peroxidation is essential for alpha-synuclein-induced cell death, Journal of neurochemistry 133(4) (2015) 582–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [105].Brenner C, Moulin M, Physiological roles of the permeability transition pore, Circulation research 111(9) (2012) 1237–47. [DOI] [PubMed] [Google Scholar]
  • [106].Bayir H, Kapralov AA, Jiang J, Huang Z, Tyurina YY, Tyurin VA, Zhao Q, Belikova NA, Vlasova II, Maeda A, Zhu J, Na HM, Mastroberardino PG, Sparvero LJ, Amoscato AA, Chu CT, Greenamyre JT, Kagan VE, Peroxidase mechanism of lipid-dependent cross-linking of synuclein with cytochrome C: protection against apoptosis versus delayed oxidative stress in Parkinson disease, The Journal of biological chemistry 284(23) (2009) 15951–69. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [107].Jowaed A, Schmitt I, Kaut O, Wullner U, Methylation regulates alpha-synuclein expression and is decreased in Parkinson’s disease patients’ brains, The Journal of neuroscience : the official journal of the Society for Neuroscience 30(18) (2010) 6355–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [108].Boissan M, Montagnac G, Shen Q, Griparic L, Guitton J, Romao M, Sauvonnet N, Lagache T, Lascu I, Raposo G, Desbourdes C, Schlattner U, Lacombe ML, Polo S, van der Bliek AM, Roux A, Chavrier P, Membrane trafficking. Nucleoside diphosphate kinases fuel dynamin superfamily proteins with GTP for membrane remodeling, Science 344(6191) (2014) 1510–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [109].Francois-Moutal L, Maniti O, Marcillat O, Granjon T, New insights into lipid-Nucleoside Diphosphate Kinase-D interaction mechanism: protein structural changes and membrane reorganisation, Biochimica et biophysica acta 1828(2) (2013) 906–15. [DOI] [PubMed] [Google Scholar]
  • [110].Anand R, Wai T, Baker MJ, Kladt N, Schauss AC, Rugarli E, Langer T, The i-AAA protease YME1L and OMA1 cleave OPA1 to balance mitochondrial fusion and fission, The Journal of cell biology 204(6) (2014) 919–29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [111].Schlattner U, Tokarska-Schlattner M, Epand RM, Boissan M, Lacombe ML, Kagan VE, NME4/nucleoside diphosphate kinase D in cardiolipin signaling and mitophagy, Laboratory investigation; a journal of technical methods and pathology 98(2) (2018) 228–232. [DOI] [PubMed] [Google Scholar]
  • [112].Kagan VE, Jiang J, Huang Z, Tyurina YY, Desbourdes C, Cottet-Rousselle C, Dar HH, Verma M, Tyurin VA, Kapralov AA, Cheikhi A, Mao G, Stolz D, St Croix CM, Watkins S, Shen Z, Li Y, Greenberg ML, Tokarska-Schlattner M, Boissan M, Lacombe ML, Epand RM, Chu CT, Mallampalli RK, Bayir H, Schlattner U, NdPk-D (NM23-H4)-mediated externalization of cardiolipin enables elimination of depolarized mitochondria by mitophagy, Cell death and differentiation 23(7) (2016) 1140–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [113].Chu CT, Bayir H, Kagan VE, LC3 binds externalized cardiolipin on injured mitochondria to signal mitophagy in neurons: implications for Parkinson disease, Autophagy 10(2) (2014) 376–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [114].Schlattner U, Tokarska-Schlattner M, Ramirez S, Tyurina YY, Amoscato AA, Mohammadyani D, Huang Z, Jiang J, Yanamala N, Seffouh A, Boissan M, Epand RF, Epand RM, Klein-Seetharaman J, Lacombe ML, Kagan VE, Dual function of mitochondrial Nm23-H4 protein in phosphotransfer and intermembrane lipid transfer: a cardiolipin-dependent switch, The Journal of biological chemistry 288(1) (2013) 111–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [115].Nakamura K, Nemani VM, Azarbal F, Skibinski G, Levy JM, Egami K, Munishkina L, Zhang J, Gardner B, Wakabayashi J, Sesaki H, Cheng Y, Finkbeiner S, Nussbaum RL, Masliah E, Edwards RH, Direct membrane association drives mitochondrial fission by the Parkinson disease-associated protein alpha-synuclein, The Journal of biological chemistry 286(23) (2011) 20710–26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [116].Guardia-Laguarta C, Area-Gomez E, Rub C, Liu Y, Magrane J, Becker A, Voos W, Schon EA, Przedborski S, alpha-Synuclein is localized to mitochondria-associated ER membranes, The Journal of neuroscience : the official journal of the Society for Neuroscience 34(1) (2014) 249–59. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [117].Sena LA, Chandel NS, Physiological roles of mitochondrial reactive oxygen species, Molecular cell 48(2) (2012) 158–67. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [118].Shen XL, Song N, Du XX, Li Y, Xie JX, Jiang H, Nesfatin-1 protects dopaminergic neurons against MPP(+)/MPTP-induced neurotoxicity through the C-Raf-ERK½-dependent anti-apoptotic pathway, Scientific reports 7 (2017) 40961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [119].Franco-Iborra S, Vila M, Perier C, Mitochondrial Quality Control in Neurodegenerative Diseases: Focus on Parkinson’s Disease and Huntington’s Disease, Frontiers in neuroscience 12 (2018) 342. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [120].Tyurina YY, Poloyac SM, Tyurin VA, Kapralov AA, Jiang J, Anthonymuthu TS, Kapralova VI, Vikulina AS, Jung MY, Epperly MW, Mohammadyani D, Klein-Seetharaman J, Jackson TC, Kochanek PM, Pitt BR, Greenberger JS, Vladimirov YA, Bayir H, Kagan VE, A mitochondrial pathway for biosynthesis of lipid mediators, Nature chemistry 6(6) (2014) 542–52. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [121].Rouzer CA, Marnett LJ, Endocannabinoid oxygenation by cyclooxygenases, lipoxygenases, and cytochromes P450: cross-talk between the eicosanoid and endocannabinoid signaling pathways, Chemical reviews 111 (10) (2011) 5899–921. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [122].Serhan CN, Chiang N, Van Dyke TE, Resolving inflammation: dual anti-inflammatory and pro-resolution lipid mediators, Nature reviews. Immunology 8(5) (2008) 349–61. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [123].Bayir H, Tyurin VA, Tyurina YY, Viner R, Ritov V, Amoscato AA, Zhao Q, Zhang XJ, Janesko-Feldman KL, Alexander H, Basova LV, Clark RS, Kochanek PM, Kagan VE, Selective early cardiolipin peroxidation after traumatic brain injury: an oxidative lipidomics analysis, Annals of neurology 62(2) (2007) 154–69. [DOI] [PubMed] [Google Scholar]
  • [124].Dixon SJ, Stockwell BR, The role of iron and reactive oxygen species in cell death, Nature chemical biology 10(1) (2014) 9–17. [DOI] [PubMed] [Google Scholar]
  • [125].Belaidi AA, Bush AI, Iron neurochemistry in Alzheimer’s disease and Parkinson’s disease: targets for therapeutics, Journal of neurochemistry 139 Suppl 1 (2016) 179–197. [DOI] [PubMed] [Google Scholar]
  • [126].Esposito LA, Melov S, Panov A, Cottrell BA, Wallace DC, Mitochondrial disease in mouse results in increased oxidative stress, Proceedings of the National Academy of Sciences of the United States of America 96(9) (1999) 4820–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [127].Rouault TA, Iron metabolism in the CNS: implications for neurodegenerative diseases, Nature reviews. Neuroscience 14(8) (2013) 551–64. [DOI] [PubMed] [Google Scholar]
  • [128].Jiang H, Wang J, Rogers J, Xie J, Brain Iron Metabolism Dysfunction in Parkinson’s Disease, Molecular neurobiology 54(4) (2017) 3078–3101. [DOI] [PubMed] [Google Scholar]
  • [129].Noble W, Hanger DP, Miller CC, Lovestone S, The importance of tau phosphorylation for neurodegenerative diseases, Frontiers in neurology 4 (2013) 83. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [130].Hamza TH, Zabetian CP, Tenesa A, Laederach A, Montimurro J, Yearout D, Kay DM, Doheny KF, Paschall J, Pugh E, Kusel VI, Collura R, Roberts J, Griffith A, Samii A, Scott WK, Nutt J, Factor SA, Payami H, Common genetic variation in the HLA region is associated with late-onset sporadic Parkinson’s disease, Nature genetics 42(9) (2010) 781–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [131].Aguirre P, Urrutia P, Tapia V, Villa M, Paris I, Segura-Aguilar J, Nunez MT, The dopamine metabolite aminochrome inhibits mitochondrial complex I and modifies the expression of iron transporters DMT1 and FPN1, Biometals : an international journal on the role of metal ions in biology, biochemistry, and medicine 25(4) (2012) 795–803. [DOI] [PubMed] [Google Scholar]
  • [132].Dibenedetto D, Rossetti G, Caliandro R, Carloni P, A molecular dynamics simulation-based interpretation of nuclear magnetic resonance multidimensional heteronuclear spectra of alpha-synuclein.dopamine adducts, Biochemistry 52(38) (2013) 6672–83. [DOI] [PubMed] [Google Scholar]
  • [133].LaVoie MJ, Ostaszewski BL, Weihofen A, Schlossmacher MG, Selkoe DJ, Dopamine covalently modifies and functionally inactivates parkin, Nature medicine 11(11) (2005) 1214–21. [DOI] [PubMed] [Google Scholar]
  • [134].Munoz P, Huenchuguala S, Paris I, Segura-Aguilar J, Dopamine oxidation and autophagy, Parkinson’s disease 2012 (2012) 920953. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [135].Norris EH, Giasson BI, Hodara R, Xu S, Trojanowski JQ, Ischiropoulos H, Lee VM, Reversible inhibition of alpha-synuclein fibrillization by dopaminochrome-mediated conformational alterations, The Journal of biological chemistry 280(22) (2005) 21212–9. [DOI] [PubMed] [Google Scholar]
  • [136].Zucca FA, Segura-Aguilar J, Ferrari E, Munoz P, Paris I, Sulzer D, Sarna T, Casella L, Zecca L, Interactions of iron, dopamine and neuromelanin pathways in brain aging and Parkinson’s disease, Progress in neurobiology 155 (2017) 96–119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [137].Hare DJ, Double KL, Iron and dopamine: a toxic couple, Brain : a journal of neurology 139(Pt 4) (2016) 1026–35. [DOI] [PubMed] [Google Scholar]
  • [138].Li WJ, Jiang H, Song N, Xie JX, Dose- and time-dependent alpha-synuclein aggregation induced by ferric iron in SK-N-SH cells, Neuroscience bulletin 26(3) (2010) 205–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [139].Duce JA, Tsatsanis A, Cater MA, James SA, Robb E, Wikhe K, Leong SL, Perez K, Johanssen T, Greenough MA, Cho HH, Galatis D, Moir RD, Masters CL, McLean C, Tanzi RE, Cappai R, Barnham KJ, Ciccotosto GD, Rogers JT, Bush AI, Iron-export ferroxidase activity of beta-amyloid precursor protein is inhibited by zinc in Alzheimer’s disease, Cell 142(6) (2010) 857–67. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [140].Stankowski JN, Dawson VL, Dawson TM, Ironing out tau’s role in parkinsonism, Nature medicine 18(2) (2012) 197–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [141].Lei P, Ayton S, Finkelstein DI, Spoerri L, Ciccotosto GD, Wright DK, Wong BX, Adlard PA, Cherny RA, Lam LQ, Roberts BR, Volitakis H, Egan GF, McLean CA, Cappai R, Duce JA, Bush AI, Tau deficiency induces parkinsonism with dementia by impairing APP-mediated iron export, Nature medicine 18(2) (2012) 291–5. [DOI] [PubMed] [Google Scholar]
  • [142].Compta Y, Parkkinen L, O’Sullivan SS, Vandrovcova J, Holton JL, Collins C, Lashley T, Kallis C, Williams DR, de Silva R, Lees AJ, Revesz T, Lewy- and Alzheimer-type pathologies in Parkinson’s disease dementia: which is more important?, Brain : a journal of neurology 134(Pt 5) (2011) 1493–1505. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [143].C. International Parkinson Disease Genomics, Nalls MA, Plagnol V, Hernandez DG, Sharma M, Sheerin UM, Saad M, Simon-Sanchez J, Schulte C, Lesage S, Sveinbjornsdottir S, Stefansson K, Martinez M, Hardy J, Heutink P, Brice A, Gasser T, Singleton AB, Wood NW, Imputation of sequence variants for identification of genetic risks for Parkinson’s disease: a meta-analysis of genomewide association studies, Lancet 377(9766) (2011) 641–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [144].Hallmann AL, Arauzo-Bravo MJ, Mavrommatis L, Ehrlich M, Ropke A, Brockhaus J, Missler M, Sterneckert J, Scholer HR, Kuhlmann T, Zaehres H, Hargus G, Astrocyte pathology in a human neural stem cell model of frontotemporal dementia caused by mutant TAU protein, Scientific reports 7 (2017) 42991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [145].Ayton S, Lei P, Duce JA, Wong BX, Sedjahtera A, Adlard PA, Bush AI, Finkelstein DI, Ceruloplasmin dysfunction and therapeutic potential for Parkinson disease, Annals of neurology 73(4) (2013) 554–9. [DOI] [PubMed] [Google Scholar]
  • [146].Ohgami RS, Campagna DR, Greer EL, Antiochos B, McDonald A, Chen J, Sharp JJ, Fujiwara Y, Barker JE, Fleming MD, Identification of a ferrireductase required for efficient transferrin-dependent iron uptake in erythroid cells, Nature genetics 37(11) (2005) 1264–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [147].Stoyanovsky DA, Tyurina YY, Shrivastava I, Bahar I, Tyurin VA, Protchenko O, Jadhav S, Bolevich SB, Kozlov AV, Vladimirov YA, Shvedova AA, Philpott CC, Bayir H, Kagan VE, Iron catalysis of lipid peroxidation in ferroptosis: Regulated enzymatic or random free radical reaction?, Free radical biology & medicine (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [148].Ayton S, Lei P, McLean C, Bush AI, Finkelstein DI, Transferrin protects against Parkinsonian neurotoxicity and is deficient in Parkinson’s substantia nigra, Signal transduction and targeted therapy 1 (2016) 16015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [149].Honarmand Ebrahimi K, Bill E, Hagedoorn PL, Hagen WR, The catalytic center of ferritin regulates iron storage via Fe(II)-Fe(III) displacement, Nature chemical biology 8(11) (2012) 941–8. [DOI] [PubMed] [Google Scholar]
  • [150].Shi H, Bencze KZ, Stemmler TL, Philpott CC, A cytosolic iron chaperone that delivers iron to ferritin, Science 320(5880) (2008) 1207–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [151].Yanatori I, Yasui Y, Tabuchi M, Kishi F, Chaperone protein involved in transmembrane transport of iron, The Biochemical journal 462(1) (2014) 25–37. [DOI] [PubMed] [Google Scholar]
  • [152].Frey AG, Nandal A, Park JH, Smith PM, Yabe T, Ryu MS, Ghosh MC, Lee J, Rouault TA, Park MH, Philpott CC, Iron chaperones PCBP1 and PCBP2 mediate the metallation of the dinuclear iron enzyme deoxyhypusine hydroxylase, Proceedings of the National Academy of Sciences of the United States of America 111 (22) (2014) 8031–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [153].Meyron-Holtz EG, Ghosh MC, Iwai K, LaVaute T, Brazzolotto X, Berger UV, Land W, Ollivierre-Wilson H, Grinberg A, Love P, Rouault TA, Genetic ablations of iron regulatory proteins 1 and 2 reveal why iron regulatory protein 2 dominates iron homeostasis, The EMBO journal 23(2) (2004) 386–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [154].Jiang H, Song N, Xu H, Zhang S, Wang J, Xie J, Up-regulation of divalent metal transporter 1 in 6-hydroxydopamine intoxication is IRE/IRP dependent, Cell research 20(3) (2010) 345–56. [DOI] [PubMed] [Google Scholar]
  • [155].Boddaert N, Le Quan Sang KH, Rotig A, Leroy-Willig A, Gallet S, Brunelle F, Sidi D, Thalabard JC, Munnich A, Cabantchik ZI, Selective iron chelation in Friedreich ataxia: biologic and clinical implications, Blood 110(1) (2007) 401–8. [DOI] [PubMed] [Google Scholar]
  • [156].Anglade P, Vyas S, Javoy-Agid F, Herrero MT, Michel PP, Marquez J, Mouatt-Prigent A, Ruberg M, Hirsch EC, Agid Y, Apoptosis and autophagy in nigral neurons of patients with Parkinson’s disease, Histology and histopathology 12(1) (1997) 25–31. [PubMed] [Google Scholar]
  • [157].Ji J, Kline AE, Amoscato A, Samhan-Arias AK, Sparvero LJ, Tyurin VA, Tyurina YY, Fink B, Manole MD, Puccio AM, Okonkwo DO, Cheng JP, Alexander H, Clark RS, Kochanek PM, Wipf P, Kagan VE, Bayir H, Lipidomics identifies cardiolipin oxidation as a mitochondrial target for redox therapy of brain injury, Nature neuroscience 15(10) (2012) 1407–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [158].Basova LV, Kurnikov IV, Wang L, Ritov VB, Belikova NA, Vlasova II, Pacheco AA, Winnica DE, Peterson J, Bayir H, Waldeck DH, Kagan VE, Cardiolipin switch in mitochondria: shutting off the reduction of cytochrome c and turning on the peroxidase activity, Biochemistry 46(11) (2007) 3423–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [159].Atkinson J, Kapralov AA, Yanamala N, Tyurina YY, Amoscato AA, Pearce L, Peterson J, Huang Z, Jiang J, Samhan-Arias AK, Maeda A, Feng W, Wasserloos K, Belikova NA, Tyurin VA, Wang H, Fletcher J, Wang Y, Vlasova II, Klein-Seetharaman J, Stoyanovsky DA, Bayir H, Pitt BR, Epperly MW, Greenberger JS, Kagan VE, A mitochondria-targeted inhibitor of cytochrome c peroxidase mitigates radiation-induced death, Nature communications 2 (2011) 497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [160].Dixon SJ, Lemberg KM, Lamprecht MR, Skouta R, Zaitsev EM, Gleason CE, Patel DN, Bauer AJ, Cantley AM, Yang WS, Morrison B 3rd, Stockwell BR, Ferroptosis: an iron-dependent form of nonapoptotic cell death, Cell 149(5) (2012) 1060–72. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [161].Galluzzi L, Vitale I, Aaronson SA, Abrams JM, Adam D, Agostinis P, Alnemri ES, Altucci L, Amelio I, Andrews DW, Annicchiarico-Petruzzelli M, Antonov AV, Arama E, Baehrecke EH, Barlev NA, Bazan NG, Bernassola F, Bertrand MJM, Bianchi K, Blagosklonny MV, Blomgren K, Borner C, Boya P, Brenner C, Campanella M, Candi E, Carmona-Gutierrez D, Cecconi F, Chan FK, Chandel NS, Cheng EH, Chipuk JE, Cidlowski JA, Ciechanover A, Cohen GM, Conrad M, Cubillos-Ruiz JR, Czabotar PE, D’Angiolella V, Dawson TM, Dawson VL, De Laurenzi V, De Maria R, Debatin KM, DeBerardinis RJ, Deshmukh M, Di Daniele N, Di Virgilio F, Dixit VM, Dixon SJ, Duckett CS, Dynlacht BD, El-Deiry WS, Elrod JW, Fimia GM, Fulda S, Garcia-Saez AJ, Garg AD, Garrido C, Gavathiotis E, Golstein P, Gottlieb E, Green DR, Greene LA, Gronemeyer H, Gross A, Hajnoczky G, Hardwick JM, Harris IS, Hengartner MO, Hetz C, Ichijo H, Jaattela M, Joseph B, Jost PJ, Juin PP, Kaiser WJ, Karin M, Kaufmann T, Kepp O, Kimchi A, Kitsis RN, Klionsky DJ, Knight RA, Kumar S, Lee SW, Lemasters JJ, Levine B, Linkermann A, Lipton SA, Lockshin RA, Lopez-Otin C, Lowe SW, Luedde T, Lugli E, MacFarlane M, Madeo F, Malewicz M, Malorni W, Manic G, Marine JC, Martin SJ, Martinou JC, Medema JP, Mehlen P, Meier P, Melino S, Miao EA, Molkentin JD, Moll UM, Munoz-Pinedo C, Nagata S, Nunez G, Oberst A, Oren M, Overholtzer M, Pagano M, Panaretakis T, Pasparakis M, Penninger JM, Pereira DM, Pervaiz S, Peter ME, Piacentini M, Pinton P, Prehn JHM, Puthalakath H, Rabinovich GA, Rehm M, Rizzuto R, Rodrigues CMP, Rubinsztein DC, Rudel T, Ryan KM, Sayan E, Scorrano L, Shao F, Shi Y, Silke J, Simon HU, Sistigu A, Stockwell BR, Strasser A, Szabadkai G, Tait SWG, Tang D, Tavernarakis N, Thorburn A, Tsujimoto Y, Turk B, Vanden Berghe T, Vandenabeele P, Vander Heiden MG, Villunger A, Virgin GW, Vousden KH, Vucic D, Wagner EF, Walczak H, Wallach D, Wang Y, Wells JA, Wood W, Yuan J, Zakeri Z, Zhivotovsky B, Zitvogel L, Melino G, Kroemer G, Molecular mechanisms of cell death: recommendations of the Nomenclature Committee on Cell Death 2018, Cell death and differentiation 25(3) (2018) 486–541. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [162].Ito K, Eguchi Y, Imagawa Y, Akai S, Mochizuki H, Tsujimoto Y, MPP+ induces necrostatin-1- and ferrostatin-1-sensitive necrotic death of neuronal SH-SY5Y cells, Cell death discovery 3 (2017) 17013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [163].Do Van B, Gouel F, Jonneaux A, Timmerman K, Gele P, Petrault M, Bastide M, Laloux C, Moreau C, Bordet R, Devos D, Devedjian JC, Ferroptosis, a newly characterized form of cell death in Parkinson’s disease that is regulated by PKC, Neurobiology of disease 94 (2016) 169–78. [DOI] [PubMed] [Google Scholar]
  • [164].Tyurina YY, Polimova AM, Maciel E, Tyurin VA, Kapralova VI, Winnica DE, Vikulina AS, Domingues MR, McCoy J, Sanders LH, Bayir H, Greenamyre JT, Kagan VE, LC/MS analysis of cardiolipins in substantia nigra and plasma of rotenone-treated rats: Implication for mitochondrial dysfunction in Parkinson’s disease, Free radical research 49(5) (2015) 681–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [165].Stockwell BR, Friedmann Angeli JP, Bayir H, Bush AI, Conrad M, Dixon SJ, Fulda S, Gascon S, Hatzios SK, Kagan VE, Noel K, Jiang X, Linkermann A, Murphy ME, Overholtzer M, Oyagi A, Pagnussat GC, Park J, Ran Q, Rosenfeld CS, Salnikow K, Tang D, Torti FM, Torti SV, Toyokuni S, Woerpel KA, Zhang DD, Ferroptosis: A Regulated Cell Death Nexus Linking Metabolism, Redox Biology, and Disease, Cell 171(2) (2017) 273–285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [166].Kagan VE, Mao G, Qu F, Angeli JP, Doll S, Croix CS, Dar HH, Liu B, Tyurin VA, Ritov VB, Kapralov AA, Amoscato AA, Jiang J, Anthonymuthu T, Mohammadyani D, Yang Q, Proneth B, Klein-Seetharaman J, Watkins S, Bahar I, Greenberger J, Mallampalli RK, Stockwell BR, Tyurina YY, Conrad M, Bayir H, Oxidized arachidonic and adrenic PEs navigate cells to ferroptosis, Nature chemical biology 13(1) (2017) 81–90. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [167].Angeli JPF, Shah R, Pratt DA, Conrad M, Ferroptosis Inhibition: Mechanisms and Opportunities, Trends in pharmacological sciences 38(5) (2017) 489–498. [DOI] [PubMed] [Google Scholar]
  • [168].Gao M, Monian P, Quadri N, Ramasamy R, Jiang X, Glutaminolysis and Transferrin Regulate Ferroptosis, Molecular cell 59(2) (2015) 298–308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [169].Doll S, Proneth B, Tyurina YY, Panzilius E, Kobayashi S, Ingold I, Irmler M, Beckers J, Aichler M, Walch A, Prokisch H, Trumbach D, Mao G, Qu F, Bayir H, Fullekrug J, Scheel CH, Wurst W, Schick JA, Kagan VE, Angeli JP, Conrad M, ACSL4 dictates ferroptosis sensitivity by shaping cellular lipid composition, Nature chemical biology 13(1) (2017) 91–98. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [170].Yuan H, Li X, Zhang X, Kang R, Tang D, Identification of ACSL4 as a biomarker and contributor of ferroptosis, Biochemical and biophysical research communications 478(3) (2016) 1338–43. [DOI] [PubMed] [Google Scholar]
  • [171].Yang WS, Stockwell BR, Ferroptosis: Death by Lipid Peroxidation, Trends in cell biology 26(3) (2016) 165–176. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [172].Wenzel SE, Tyurina YY, Zhao J, St Croix CM, Dar HH, Mao G, Tyurin VA, Anthonymuthu TS, Kapralov AA, Amoscato AA, Mikulska-Ruminska K, Shrivastava IH, Kenny EM, Yang Q, Rosenbaum JC, Sparvero LJ, Emlet DR, Wen X, Minami Y, Qu F, Watkins SC, Holman TR, VanDemark AP, Kellum JA, Bahar I, Bayir H, Kagan VE, PEBP1 Wardens Ferroptosis by Enabling Lipoxygenase Generation of Lipid Death Signals, Cell 171 (3) (2017) 628–641 e26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [173].Dexter DT, Carter CJ, Wells FR, Javoy-Agid F, Agid Y, Lees A, Jenner P, Marsden CD, Basal lipid peroxidation in substantia nigra is increased in Parkinson’s disease, Journal of neurochemistry 52(2) (1989) 381–9. [DOI] [PubMed] [Google Scholar]
  • [174].Garcia E, Limon D, Perez-De La Cruz V, Giordano M, Diaz-Munoz M, Maldonado PD, Herrera-Mundo MN, Pedraza-Chaverri J, Santamaria A, Lipid peroxidation, mitochondrial dysfunction and neurochemical and behavioural deficits in different neurotoxic models: protective role of S-allylcysteine, Free radical research 42(10) (2008) 892–902. [DOI] [PubMed] [Google Scholar]
  • [175].Hambright WS, Fonseca RS, Chen L, Na R, Ran Q, Ablation of ferroptosis regulator glutathione peroxidase 4 in forebrain neurons promotes cognitive impairment and neurodegeneration, Redox biology 12 (2017) 8–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [176].Chen L, Hambright WS, Na R, Ran Q, Ablation of the Ferroptosis Inhibitor Glutathione Peroxidase 4 in Neurons Results in Rapid Motor Neuron Degeneration and Paralysis, The Journal of biological chemistry 290(47) (2015) 28097–106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [177].Shah R, Shchepinov MS, Pratt DA, Resolving the Role of Lipoxygenases in the Initiation and Execution of Ferroptosis, ACS central science 4(3) (2018) 387–396. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [178].Shintoku R, Takigawa Y, Yamada K, Kubota C, Yoshimoto Y, Takeuchi T, Koshiishi I, Torii S, Lipoxygenase-mediated generation of lipid peroxides enhances ferroptosis induced by erastin and RSL3, Cancer science 108(11) (2017) 2187–2194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [179].Yang WS, Kim KJ, Gaschler MM, Patel M, Shchepinov MS, Stockwell BR, Peroxidation of polyunsaturated fatty acids by lipoxygenases drives ferroptosis, Proceedings of the National Academy of Sciences of the United States of America 113(34) (2016) E4966–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [180].Gandal MJ, Zhang P, Hadjimichael E, Walker RL, Chen C, Liu S, Won H, van Bakel H, Varghese M, Wang Y, Shieh AW, Haney J, Parhami S, Belmont J, Kim M, Moran Losada P, Khan Z, Mleczko J, Xia Y, Dai R, Wang D, Yang YT, Xu M, Fish K, Hof PR, Warrell I, Fitzgerald D, White K, Jaffe AE, Psych EC, Peters MA, Gerstein M, Liu C, Iakoucheva LM, Pinto D, Geschwind DH, Transcriptome-wide isoform-level dysregulation in ASD, schizophrenia, and bipolar disorder, Science 362(6420) (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [181].Li M, Santpere G, Imamura Kawasawa Y, Evgrafov OV, Gulden FO, Pochareddy S, Sunkin SM, Li Z, Shin Y, Zhu Y, Sousa AMM, Werling DM, Kitchen RR, Kang HJ, Pletikos M, Choi J, Muchnik S, Xu X, Wang D, Lorente-Galdos B, Liu S, Giusti-Rodriguez P, Won H, de Leeuw CA, Pardinas AF, BrainSpan C, Psych EC, Psych EDS, Hu M, Jin F, Li Y, Owen MJ, O’Donovan MC, Walters JTR, Posthuma D, Levitt P, Weinberger DR, Hyde TM, Kleinman JE, Geschwind DH, Hawrylycz MJ, State MW, Sanders SJ, Sullivan PF, Gerstein MB, Lein ES, Knowles JA, Sestan N, Reimers MA, Integrative functional genomic analysis of human brain development and neuropsychiatric risks, Science 362(6420) (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [182].Wang D, Liu S, Warrell J, Won H, Shi X, Navarro FCP, Clarke D, Gu M, Emani P, Yang YT, Xu M, Gandal MJ, Lou S, Zhang J, Park JJ, Yan C, Rhie SK, Manakongtreecheep K, Zhou H, Nathan A, Peters M, Mattei E, Fitzgerald D, Brunetti T, Moore J, Jiang Y, Girdhar K, Hoffman GE, Kalayci S, Gumus ZH, Crawford GE, Psych EC, Roussos P, Akbarian S, Jaffe AE, White KP, Weng Z, Sestan N, Geschwind DH, Knowles JA, Gerstein MB, Comprehensive functional genomic resource and integrative model for the human brain, Science 362(6420) (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [183].di Domenico A, Carola G, Calatayud C, Pons-Espinal M, Munoz JP, Richaud-Patin Y, Fernandez-Carasa I, Gut M, Faella A, Parameswaran J, Soriano J, Ferrer I, Tolosa E, Zorzano A, Cuervo AM, Raya A, Consiglio A, Patient-Specific iPSC-Derived Astrocytes Contribute to Non-Cell-Autonomous Neurodegeneration in Parkinson’s Disease, Stem cell reports (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [184].Ishikawa T, Imamura K, Kondo T, Koshiba Y, Hara S, Ichinose H, Furujo M, Kinoshita M, Oeda T, Takahashi J, Takahashi R, Inoue H, Genetic and pharmacological correction of aberrant dopamine synthesis using patient iPSCs with BH4 metabolism disorders, Human molecular genetics 25(23) (2016) 5188–5197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [185].Inoue N, Ogura S, Kasai A, Nakazawa T, Ikeda K, Higashi S, Isotani A, Baba K, Mochizuki H, Fujimura H, Ago Y, Hayata-Takano A, Seiriki K, Shintani Y, Shintani N, Hashimoto H, Knockdown of the mitochondria-localized protein p13 protects against experimental parkinsonism, EMBO reports 19(3) (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [186].Jennings CG, Landman R, Zhou Y, Sharma J, Hyman J, Movshon JA, Qiu Z, Roberts AC, Roe AW, Wang X, Zhou H, Wang L, Zhang D, Desimone R, Feng G, Opportunities and challenges in modeling human brain disorders in transgenic primates, Nature neuroscience 19(9) (2016) 1123–30. [DOI] [PubMed] [Google Scholar]

RESOURCES