Open access peer-reviewed chapter

Parkinson’s Disease: Insights from Drosophila Model

Written By

Mohamad Ayajuddin, Abhik Das, Limamanen Phom, Priyanka Modi, Rahul Chaurasia, Zevelou Koza, Abuno Thepa, Nukshimenla Jamir, Pukhrambam Rajesh Singh, Sentinungla Longkumer, Pardeshi Lal and Sarat Chandra Yenisetti

Submitted: 03 October 2017 Reviewed: 27 October 2017 Published: 28 February 2018

DOI: 10.5772/intechopen.72021

From the Edited Volume

Drosophila melanogaster - Model for Recent Advances in Genetics and Therapeutics

Edited by Farzana Khan Perveen

Chapter metrics overview

1,644 Chapter Downloads

View Full Metrics

Abstract

Parkinson’s disease (PD) is a medical condition that has been known since ancient times. It is the second most common neurodegenerative disorder affecting approximately 1% of the population over 50 years. It is characterized by both motor and non-motor symptoms. Most of PD cases are sporadic while 5–10% cases are familial. Environment factors such as exposure to pesticides, herbicides and other heavy metals are expected to be the main cause of sporadic form of the disease. Mutation of the susceptible genes such as SNCA, PINK1, PARKIN, DJ1, and others are considered to be the main cause of the familial form of disease. Drosophila offers many advantages for studying human neurodegenerative diseases and their underlying molecular and cellular pathology. Shorter life span; large number of progeny; conserved molecular mechanism(s) among fly, mice and human; availability of many techniques, and tools to manipulate gene expression makes drosophila a potential model system to understand the pathology associated with PD and to unravel underlying molecular mechanism(s) responsible for dopaminergic neurodegeneration in PD—understanding of which will be of potential assistance to develop therapeutic strategies to PD. In the present review, we made an effort to discuss the contribution of fly model to understand pathophysiology of PD, in understanding the biological functions of genes implicated in PD; to understand the gene-environment interaction in PD; and validation of clues that are generated through genome-wide association studies (GWAS) in human through fly; further to screen and develop potential therapeutic molecules for PD. In nutshell, fly has been a great model system which has immensely contributed to the biomedical research relating to understand and addressing the pathology of human neurological diseases in general and PD in particular.

Keywords

  • dopamine
  • Drosophila
  • Parkinson’s disease
  • mitochondrial dysfunction
  • neuroprotective therapeutics
  • pathophysiology and translational research

1. Introduction

Parkinson’s disease (PD) is a medical condition that has been known about since ancient times in Indian and Chinese civilization [1, 2]. It is referred to as Kampavata in the ancient Indian medical system of Ayurveda (“kampa” means tremor in Sanskrit). An Egyptian papyrus from the twelfth century BC mentions a king drooling with age [3]. In Western medical literature, the tremor symptom was described by the physician Galen in 175 AD [4]. In 1817, James Parkinson wrote an essay on “shaking palsy” based on six cases that he had observed in his own practice and on walks around his neighborhood. The essay was intended to encourage others to study the disease. This established the disease as a recognized medical condition. He termed this medical condition as “shaking palsy or paralysis agitans”. He published a detailed medical essay entitled “An Essay on the Shaking Palsy” where he described shaking palsy as “involuntary tremulous motion, with lessened muscular power, in parts not in action and even when supported; with a propensity to bend the trunk forwards, and to pass from a walking to a running pace: the senses and intellects being uninjured” [5]. The term “Parkinson’s disease” was coined in 1865 by William Sanders and later popularized by French neurologist Jean Martin Charcot [6]. Charcot and colleagues described the clinical symptoms of this disease, noting two prototypes: the tremorous and the rigid or akinetic form. They described the detailed arthritic changes, dysautonomia, and pain that can accompany Parkinson’s disease. He recognized that PD patients are not markedly weak and do not necessarily have tremor [7]. All these observations instigated the curiosity among the clinicians to understand this condition better with an aim to improve the patient’s quality of life.

Advertisement

2. Pathophysiology

Parkinson’s disease is the second most common neurodegenerative disorder affecting approximately 1% of the population over 50 years [8]. A central pathological hallmark of PD is the selective loss of dopamine (DA) neurons in the substantia nigra pars compacta (SN). These dopaminergic neurons are required for proper motor function, and their deficiency manifests its characteristic features: bradykinesia, tremors, and rigidity [9]. A second neuropathological hallmark of PD is the Lewy body (LB), which is a cytoplasmic spherical proteinaceous inclusion. LBs have been reported to contain various proteins including α-synuclein, ubiquitin, parkin, and neurofilaments [10]. The mechanisms by which α-synuclein and other proteins aggregate to form Lewy pathology are uncertain, but may involve oxidative modifications and/or cross-linking. Although, the neurodegeneration of PD was considered to be confined to dopaminergic cell loss in the SN, cell loss, and neuropathology is found to occur in other parts as well including the locus coeruleus, raphe, nucleus basalis of Meynert, dorsal motor nucleus of the vagus, cerebral cortex, olfactory bulb, and autonomic nervous system [11]. Several non-motor symptoms such as sleep disturbances, constipation, cognitive decline, depression, fear, anxiety, bladder problems, weight changes, fatigue and loss of energy, hypotension, and sexual problems can be dominant and debilitating in a sizeable number of patients, affecting the quality of their life [12]. Till date, treatments address only the symptoms but they fail to stop the progression of the disease and PD patients continue to experience a higher mortality rate compared to the general population [13].

2.1. Sporadic Parkinson’s disease

A sporadic PD has unknown cause with implication of environmental influence coupled with genetic factors. The pathology of PD therefore may be multifactorial involving gene and environment interactions. Studies indicate role of neurotoxicants or neuroprotective compounds in pathogenesis of nigrostriatal degeneration, supporting the concept of association between the environment and PD [14]. Additionally, the identification of the mutated α-synuclein (SNCA) gene that cause familial PD [15] as a risk factor for sporadic disease [16] provides a genetic background for the disease. Studies suggest that rural people, well water consumption, pesticide use, and occupations like rural farming, mining, and welding have an increased risk of PD [17, 18]. Epidemiological studies suggest association of PD with environmental toxic factors, primarily the mitochondrial complex I inhibitors such as rotenone [19]. Some other findings suggest that exposure to pesticide such as bipyridyl, organochlorine, and carbamate derivatives could contribute to PD [20].

2.2. 1-Methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)

Exploring the contribution of environmental exposure markedly advanced our understanding of the mechanisms involved in the development of PD. Initial evidence came from findings that subjects exposed to 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) developed PD-like symptoms [21]. MPTP was accidentally discovered in a synthesis process that went wrong, and, although it may have caused some problems in certain circles, today it represents the most important and most frequently used parkinsonian toxin applied in animal models. 1-Methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) is a representative strong neurotoxin that has been recognized from several young drug addicts from Northern California developed severe parkinsonism [22]. Since then, environmental exposure to pesticides [23], polychlorinated biphenyls [24], organic solvents [25], metals [26], and air pollutants [27] has been proposed to increase risk for PD.

2.3. Mechanism of MPTP neurotoxicity

Though the exact mechanism regarding the mode of MPTP toxicity is not known, it has been postulated that MPP+ entry into dopaminergic neurons is dependent on selective uptake by dopamine transporter localizing it and interfering into mitochondrial activity. MPTP is not toxic per se, but becomes toxic once it is converted to 1-methyl-4-phenyl-2,3-dihydropyridinium (MPDP+) by action of monoamine oxidase B (MAO-B) in glial cells and serotoninergic neurons [28] followed by oxidation into 1-methyl-4-phenylpyridinium (MPP+), which is a highly toxic compound [29]. Then the dopamine transporter (DAT) carries it to dopaminergic neurons leading to its accumulation in cytoplasm and into synaptic vesicles by the vesicular monoamine transporter (VMAT). The driving force of mitochondrial membrane potential lets MPP+ enter these organelles, where it blocks complex I [30]. This leads to abnormally increased concentrations of the toxin to interfere with mitochondrial respiration by blocking the mitochondrial oxidation. Thus it results in impairment of ATP synthesis and involving in the generation of oxidative stress.

2.4. Paraquat

Paraquat (1,1-dimethyl-4,4′-bipyridinium dichloride) is a quaternary nitrogen herbicide widely used for broadleaf weed control. It is a fast-acting, non-selective compound which destroys tissues of green plants on contact and by translocation with the plant. Significant damage to the brain was seen in individuals who died from paraquat intoxication [31]. For many years, experimental studies using paraquat were focusing on its effects on lung, liver, and kidney, probably because the toxicity induced by this herbicide in these organs is responsible for death after acute exposure [32]. Epidemiological studies in agricultural communities have suggested an increased risk for PD due to paraquat use, raising the possibility that paraquat could be an environmental parkinsonian toxin. This chemical causes extensive oxidative stress in mitochondria of the cell, resulting in the perturbation of biochemical processes, cell death, multi-organ failure, and neurodegenerative diseases [33].

It is still vague about how the molecular mechanism of paraquat leads to cell death. However, studies have shown that paraquat can trigger the sequential phosphorylation of c-Jun N-terminal kinase (JNK), c-Jun, and the activation of caspase-3 consequently leading to neuronal death both in vitro and in vivo [34], suggesting that JNK signaling pathway contributes to paraquat-induced neurodegeneration.

2.5. Rotenone

Rotenone is a commonly used natural pesticide prepared from the roots of certain tropical plants, such as Derris elliptica. It is a classical high-affinity-specific inhibitor of mitochondrial complex I. This lipophilic compound crosses the blood-brain barrier rapidly and accumulates in subcellular organelles such as mitochondria where it impairs oxidative phosphorylation by inhibiting complex I of the electron transport chain [35]. Postmortem studies implicated mitochondrial impairments [36], and epidemiological studies suggested an association with environmental toxins, in particular mitochondrial complex I inhibitors such as rotenone [37]. In vitro, rotenone has been shown to produce cell apoptosis, accumulation, and aggregation of α-synuclein and ubiquitin, oxidative damage, and endoplasmic reticulum stress [38, 39]. In a study in post mortem idiopathic PD brain, the substantia nigra is seen to comprise of a strong inhibition of complex I activity [40] suggesting this could be the cause of degeneration of dopaminergic neurons.

Advertisement

3. Familial/genetic Parkinson’s disease

Till date, 15 known genes and 21 loci have been identified for familial PD. Some of the genes are discussed further.

3.1. SNCA

In a study involving a large Italian family, Polymeropoulos and colleagues identified the missense mutations in the SNCA gene. Through a traditional linkage approach, they managed to track the underlying genetic injury to an area located in the long arm of human chromosome number 4 [41]. This discovery has been a FRAME shift in the genetic research of PD. A separate study showed that the α-synuclein protein is the main constituent of the Lewy body which is the pathological hallmark of PD [42]. These two vital research findings brought about a link between sporadic and familial forms of PD.

Five different missense mutations in SNCA have been implicated in PD namely A53T, A30P, E46K, H50Q, and G51D mutation. Clinically, early age onset of parkinsonism with a positive initial response to levodopa treatment are seen among patients with missense mutations but later on the disease progresses rapidly with dementia as a common feature. Eventually, appearance of cognitive decline, hallucinations, and fluctuations of consciousness in patients becomes clear. Histopathological study reveals an abundant LB pathology [43].

3.2. LRRK2

An association between apparent autosomal-dominant parkinsonism and chromosome number 12 was suggested by findings on a study involving a large Japanese family [44]. Later it was established that mutations in the gene LRRK2 lead to the basic genetic cause of chromosome number 12 linked PD [45]. The most frequent LRRK2 mutation is G2019S, detected in approximately 1% of sporadic and about 3–6% of familial PD cases [46]. R1441G is the second most common mutation after G2019S [47]. Most of the LRRK2 cases described demonstrate LB in the brainstem accompanied by loss of neurons in the SN. Only a minority of cases exhibit neurofibrillary tangle pathology, glial cytoplasmic inclusions, or neuronal nigral loss without LB [45].

3.3. PARKIN

An uncommon syndrome characterized by early dystonia and problems from levodopa treatment, osteotendinous hyperreflexia and comparatively slow motor progression was described in Japan in 1973, which is now known to be an autosomal-recessive juvenile parkinsonism (AR-JP) [48]. Mutations in parkin gene was identified as a cause of this condition [49]. AR-JP maps to the long arm of chromosome number 6 and linked to the markers D6S305 and D6S253 [50]. It was found that the former is deleted in an AR-JP Japanese patient [51]. By positional cloning within this microdeletion, Kitada and colleagues isolated a cDNA clone of 2960 bp with a 1395 bp open reading frame which code a 465 amino acid protein consisting of an N-terminal ubiquitin-like domain and RING domain having two RING finger motifs. The gene spans more than 500 kb and has 12 exons of which 5 exons (3–7) are found to be deleted in the patient. Also four other AR-JP unrelated patients have a deletion that affect only the exon 4. A 4.5 kb transcript expressed in various human tissues abundantly in the brain, including the substantia nigra, is shorter in brain tissue from one of the groups of exon 4 deleted patients. Therefore inferring that the mutations in this newly identified gene is responsible for the pathogenesis of AR-JP. In a number of families, PD is related with heterozygous parkin mutations through an apparently dominant way of transmission, suggesting that the carriers of a sole parkin mutation might be at risk to develop the PD [52].

3.4. DJ1

A homozygous deletion and a missense mutation in the Daisuke-Junko-1 (DJ-1) gene as a cause of autosomal-recessive early onset PD was identified [53]. A number of novel DJ-1 mutations have been discovered in patients with early onset PD. However, these mutations are rare and can be found in only ̴ about 1% of early onset PD cases [54]. At the clinical level, the phenotype of DJ-1 subjects is the same as that of parkin and PINK1-related parkinsonism, with age at onset (AAO) usually around the mid-30s, good response to levodopa treatment, slow disease progression, and often focal dystonia such as blepharospasm [55]. However, the neuropathology of DJ-1-linked parkinsonism still remains unidentified.

3.5. PINK1

Mutations in the phosphate and tensin homolog-induced Putative kinase 1 (PINK1) were initially identified in a Sicilian family with autosomal-recessive parkinsonism [56]. In most of PINK1 mutations, the type of mutation seen is missense. However, mutations of copy number, genetic, and exonic rearrangements have been described [57]. In both the cases of familial and sporadic PD, investigation of mutation has recognized homozygous and compound heterozygous type PINK1 mutations. This raised the potential role on a single PINK1 heterozygous mutation to be a possible risk factor for PD [58]. In a number of families, PINK1 mutations have been associated with early onset PD and PINK1 mutations association in sporadic cases is about 2–4%. In a clinical phenotype, this type of parkinsonism is quite comparable to those seen among patients with parkin and DJ-1 mutations. They display progressive levodopa-responsive disease gradually [57].

3.6. Vacuolar protein sorting-associated protein 35 (VPS35)

Mutation in VPS35 causes monogenic form of PD was described by using exome sequencing in two separate studies with identification of p.D620N mutation in VPS35 among the members of a Swiss kindred with a late onset, autosomal-dominant PD [59]. An independent study published the identification of the p.D620N mutation in a large multi-generation Austrian PD family and in two additional families screened for VPS35 mutations [60]. The VPS35-linked families reportedly fulfill the London Brain Bank criteria for PD, but there are fairly limited clinical and pathological details on these cases.

3.7. ATPase (P-type) 13A2 (ATP13A2)

In a neuronal P-type ATPase gene, ATP13A2, a loss of function was initially described in a consanguineous Jordanian family [61]. Clinically, the subjects showed a very early onset of the disease accompanied by rigid-akinetic phenotype with reduced resting tremor, pyramidal syndrome, progressive cognitive impairment, vertical gaze palsy, mini myoclonus, insomnia, and levodopa responsive [61]. Mutations of this gene mapping on chromosome 1p36 cause PD in atypical form which is known as Kufor-Rakeb syndrome [62]. Clinical phenotype of this early onset pallido-pyramidal syndrome varies in severity but only a handful of cases and families have been reported [63, 64].

3.8. PLA2G6

Homozygous mutations in phospholipase A2 gene (PLA2G6) located on chromosome 22q13.1 was reported in three patients of two inbred Pakistani families. The phenotypes were associated with cognitive and psychiatric problems and dystonic features, pyramidal syndrome [65]. In an Asian group with early onset recessive parkinsonism caused by compound heterozygous mutations in PLA2G6, the phenotype reported were frontotemporal lobar atrophy and dementia [66]. The PLA2G6 gene encodes a protein of phospholipase A2 group VI (PLA2G6), which act as catalysis of fatty acids elimination from phospholipids and involved in maintaining membrane phospholipids homeostasis [67].

3.9. UCH-L1

Missense mutations of gene coding for the ubiquitin carboxyl-terminal hydrolase L1 (UCH-L1), an ubiquitin recycling enzyme located on chromosome 4p14 was reported in a German family with an autosomal-dominant transmission PD [68]. The phenotype of affected individuals was consistent with that of idiopathic PD. An epidemiological study suggested an association between the UCH-L1 gene S18Y variant and PD [69]. Overexpression of UCH-L1 gene upregulated aggresomes formation through dysfunction of proteasome system [70].

3.10. HtrA2

Various studies indicate risk factor for parkinsonism due to loss of function of the gene encoding Omi/high temperature requiring A2 mitochondrial protein (HtrA2) in German [71] and Belgians PD patients [72]. The serine protease Omi/HtrA2 is released from mitochondria and promotes apoptosis [73] and mutations of Omi/HtrA2 gene affect its protease activity linked to mitochondrial dysfunction [71, 72]. Although it acts independently of parkin, Omi/HtrA2 functions in the PINK1/parkin pathway downstream of PINK1 [74]. These findings were not confirmed in Omi/HtrA2 knockout mutants in contrast to PINK1 or parkin null mutants [75].

3.11. EIF4G1

Mutations in the eukaryotic translation initiation factor 4-gamma (EIF4G1) was identified as a risk factor in a study of a northern French family with autosomal-dominant late onset parkinsonism on the chromosome 3q26-q28. Genomic analysis identified a heterozygous mutation in EIF4G1 which was confirmed subsequently in 2 PD patients and 2 Lewy body disease patients among 225 more patients [76]. Further, in all the affected members of another multiplex unrelated family, a pathogenic mutation was detected including in one unaffected 86-year-old family member suggesting an incomplete penetrance [77].

Advertisement

4. Animal models of Parkinson’s disease

4.1. Criteria for modeling PD in animals

Being a neurodegenerative disorder, the prominent hall mark of PD is progressive loss of dopaminergic neurons in the Substantia niagra parse compacta (SNpc) [78]. Together with DA neurons, there is a loss of cholinergic neurons, serotonergic neurons, and nor adrenergic neurons in the brain [79]. The prominent biomarker being aggregation of Lewy bodies in intracytoplasmic space [80]. A combination of all these features shows the hallmark motor and non-motor symptom of PD.

A suitable model for PD should show histopathologically characterizable progressive loss of DA neurons together with other neurons and significant reduction in DA level. Moreover, the onset of the disease should be in adulthood this should manifest in such a way that it would mimic the PD-affected human motor symptom such as bradykinesia, rigidity, postural instability, and resting tremor, with motor features being responsive to L-DOPA or any anti-PD drug therapy. Even though non human primate and mouse has been the traditional model of PD, because of low cost of maintenance, shorter life cycle, and defined neuropathological profile is making zebrafish and Drosophila are among the emerging and more interesting model of PD [81].

4.2. Advantages and limitation of modeling PD using drosophila

4.2.1. Advantages

Drosophila offers many advantages for studying human neurodegenerative diseases and their underlying molecular and cellular pathology. Benefits include a faster time frame due to the shorter life span of the fly (10–14 day from pupae to adult), large number of progeny, availability of many techniques, and tools to manipulate gene expression [82]. Also Drosophila has been studied for longer than any other model out there which makes its anatomy and phenotypes very well known to experimental biologists [83]. A well developed CNS and prominent number of DA neurons [84] combined with well characterized behaviors which are conserved among strains in 90% cases [85] makes fly very cost effective and efficient model. Genetically, it has been estimated that nearly 75% of disease-related genes in humans have functional orthologs in the fly [86]. While overall similarity at nucleotide or protein level is 40% but in conserved domain it is 80–90% or higher [87].

4.2.2. Disadvantages

The major difference being an invertebrate there will be some prominent difference in physiology with human (e.g., brain anatomy, cardio vascular system, and respiratory system) which relating complex motor behavior with human a difficult task [88]. For CNS-related studies, there is also an issue on blood-brain permeability difference [89]. The metabolic differences are also to be considered when studying drug efficacy and toxin-induced disease phenotype.

4.3. Relevance of study of PD in flies

4.3.1. Drosophila mimicking human PD symptoms

Although the physiological difference between human and flies are very prominent core pathology observed in human PD can be produced in a very accurate extent by toxin-induced or transgenic modification. It is as accurate as area specific and age-dependent DA neuron loss as observed in PD condition and hallmark PD biomarker the LB and LN inclusions are visible in transgenic system [90]. Drosophila also performs complex behavior such as mating, aggression, conditioning to fear, learning and motor behaviors such as flying, climbing, and walking [85] which like human are affected by the PD onset and progression. These multitudes of behavior are very much helpful in characterizing different aspects of PD.

4.4. Physiological attributes of fly brain

Drosophila model of PD has two principle phenotypes: the specific loss of DAergic neurons with aging brain and defects in motor behavior. In fly brain, the DA neurons are distributed as a group of clusters throughout the brain and project their effect on different behavioral patterns of the fly by different functional areas of the adult brain. Target areas include:

  • The mushroom bodies: involved in memory formation and motivation.

  • The Central Complex: controls the motor behavior.

The phenomenon of different part of brain controlling different behavior pattern highly resembles the mammalian brain ( Figure 1 ) [84].

Figure 1.

(A and C) Schematic representation of an adult fly brain with the distribution of DA neurons grouped in clusters and arranged with bilateral symmetry (image adapted from Botella et al. [91]). (B and D) Confocal Z-stack of TH > mCD8::GFP brain; anti-nc82 immunoreactivity together with GFP labeling reveals dopaminergic neurons in the anterior and posterior brain (image adapted from White et al. [84]).

4.5. Genetic similarity between Drosophila and human

Drsophila shares 61% homology with human genetically. In fact all the familial or sporadic genes reported so far in human associated with parkinsonism are available in Drosophila as a homolog. But there is no homolog for the gene α-synuclein which produces Lewy bodies and Lewy neuritis at the extracellular matrix of brain a hallmark biomarker under PD condition in mammalian brain. Nevertheless α-synuclein transgenic models respond very well under PD conditions and recapitulate the PD phenotypes. Given ( Table 1 ) are the list of parkinsonian genes and their homologs in flies.

Gene/protein Inheritance Fly homolog Protein function
α-Synuclein AD None Pre-synaptic protein
Parkin AR Parkin/CG10523 E3 ubiquitin ligase
UCH-L1 unclear Uch/CG4265 E3 ubiquitin hydrolase/ligase
PINK1 AR Pink1/CG4523 Mitochondrial kinase
DJ-1 AR DJ-1a/CG6646
DJ-1b/CG1349
Redox sensor/Chaperone
LRRK2 AD lrrk2/CG5483 Kinase/GTPase
HtrA2 AD HtrA2/CG8486 Mitochondrial pro-apoptotic protease
GBA unclear CG33090 Lysosomal enzyme
POLG unclear tamas/CG8987 Mitochondrial DNA polymerase
Tau unclear tau/CG31057 Microtubule stabilization

Table 1.

Showing parkinsonian genes and their fly homologs [92].

Notes: UCH-L1 = ubiquitin carboxyl-terminal esterase L1; PINK1 = PTEN-induced putative kinase 1; LRRK2 = leucine-rich repeat kinase 2; HtrA2 = high temperature requirement protein A2; GBA = glucocerebrosidase; POLG = polymerase gamma; AD = autosomal dominant; AR = autosomal recessive.

Advertisement

5. Gene-environment interaction studies in PD

The identification of PD symptoms subsequent to ingestion of MPTP led to the idea that environmental factors could be related to the causes of pathogenesis of the disease. When ingested, MPTP is metabolized to neurotoxin MPP+ which was originally developed as an herbicide, cyperquat. The chemical structure of MPP+ is similar to that of the widely used herbicide paraquat. This finding suggested that exposure to environmental factors such as pesticide, herbicides may contribute to human sporadic PD. Over the years, environmental factors, including pesticides and herbicides, metals, tobacco and caffeine, head injuries, etc. have been largely considered as possible PD risk factors.

Over the last few decades either through genetic or toxin challenges many animal models have been developed to study PD. Using Drosophila melanogaster as a model has proved to be of great value and has contributed significantly toward understanding the mechanism underlying PD pathogenesis. Table 2 illustrates the interaction studies between environmental toxins and PD genes using Drosophila model.

Gene/protein Toxin(s) Observations Reference
α-Synuclein Rotenone Drosophila larvae expressing mutant human α-synuclein showed significant loss of DA neurons and deficit in locomotion [93]
DJ-1 Paraquat
Rotenone
DJ-1 KO flies selectively sensitive to oxidative stress induced by toxins [94]
[95]
LRRK2 Rotenone hLRRk2 transgenic flies showed loss of DA neurons and sensitivity to environmental toxin [96]
Parkin Rotenone Transgenic flies expressing parkin mutation show age-dependent DA neuron degeneration and locomotor deficits [97]
PINK1 Rotenone Flies expressing PINK1 mutations show increased rates of mortality in the mutants [98]

Table 2.

Gene-environment interaction studies in drosophila model of PD.

Environmental toxins such as paraquat (herbicide) increases the risk factors for PD and this susceptibility is influenced by the genetic background. Fly model exposed to paraquat shows neurodegenerative symptoms induced by oxidative stress which are similar with most behavioral characters of PD. As seen from epidemiological studies, male flies are more susceptibility to paraquat toxicity than female flies. Drosophila mutant for dopamine regulating genes show variable susceptibility to paraquat such as mutations which increase DA pathway function helps in reducing paraquat neurotoxicity while loss of function mutations increases susceptibility to paraquat by decreasing dopamine levels. The loss of function mutation in negative regulator of DA production (Catecholamines-up (Catsup)) acts by delaying the onset of PD symptoms and loss of DA neurons and confers protection against paraquat exposure [99]. Drosophila DJ-1 mutants developed motor deficits when exposed to paraquat [100]. Loss of function of DJ-1β mutants confers resistance to paraquat and lowers mortality of DA neuron and overexpression of DJ-1α in DA neurons, protects against paraquat toxicity [95]. Acute paraquat exposure in Drosophila showed elevated levels of oxidative stress markers and mitochondrial dysfunction [101]. Role of ubiquitin proteosome system (UPS) to sporadic PD is not very clear though it is a potential target for PD risk associated with pesticide [102]. Data from epidemiological studies show that paraquat in addition with maneb or ziram increases PD risk [103]. Drosophila knockdown of E1 ligase when exposed to paraquat + maneb showed significant DA neuron loss thus imply synergistic effects of the pesticides on risk for PD [104].

Rotenone exposure suppresses proteasomal activity through ATP depletion thus inhibiting mitochondrial function [105]. Parkin through its E3 ligase function offers neuroprotection against neuronal insults including rotenone [106]. Rotenone alters parkin solubility increasing intracellular aggregation and S-nitrosylation of parkin [107]. Loss of parkin increases rotenone-induced DA cell death in mice [108]. Drosophila shows negative geotaxis characteristics and rotenone exposure has shown to cause mortality and locomotor defects affecting the climbing ability of flies [109]. Rotenone also inhibits learning and memory functions in fly and the damage caused shows severe effect than those in α-synuclein A30P mutant [110]. Exposure of adult Drosophila to sublethal doses of rotenone causes concentration-dependent locomotor deficiencies, specific dopaminergic neuronal loss, and reduction in the DA levels in flies [111]. Non-motor symptoms such as circadian rhythms in Drosophila are also altered following exposure to rotenone [112]. All these studies highlight the utility of drosophila model to understand the gene-environment interaction in PD.

5.1. Drosophila PD models and associated tools

Drosophila is proved to be one of the important genetic models to study the disease mechanism in PD. Even though with limitations, the fly model enables rapid and elaborate genetic study, providing in depth cell and molecular studies which offers a unique look into the mechanisms and pathways underlying PD pathogenesis [113]. Currently 14 genes for PD have been identified of which 12 genes have homology with Drosphila. Loss of function and gain of function analysis using fly model would provide insights into the mechanism of action of these genes associated with PD ( Table 3 ).

Gene/symbol Drosophila homolog (Over)Expression construct (Over)Expression construct with point mutation Loss of function mutants
SNCA/ PARK1 No YES YES NO
PARK2 encoding parkin/ PARK2 Parkin YES YES YES
UCHL1/PARK5 Uch NO NO NO
PINK1/PARK6 Pink1 YES YES YES
PARK7 encoding DJ1/PARK7 Dj-1α and Dj-1β YES NO YES
LRRK2/PARK8 Lrrk YES YES YES
ATP13A2/PARK9 CG32000 YES NO NO
HTRA2/PARK13 Htra2 YES YES YES
PLA2G6/PARK14 iPLA2-VIA NO NO YES
FBOX7/PARK15 No homolog YES YES NO
VPS35/PARK17 Vps35 YES Yes YES
EIG4G1/PARK18 eIF4G No NO NO
DNAJC6/PARK19 auxillin YES NO YES
SYNJ1/PARK20 Synj YES YES NO

Table 3.

Drosophila model of Parkinson’s disease (genetic) (adapted from Vanhauwaert and Verstreken [113]).

5.2. Mitochondrial genetics of PD: insights from drosophila model

Drosophila model has made major contribution in the research area of mitochondrial genetics. The early hints of PINK1/parkin on mitochondrial homeostasis came from studies using Drosophila model [98, 114, 115]. The most compelling evidence for a mitochondrial etiology of PD was derived from the study of genes mediating familial forms of the disease in fly model [116, 117]. Mutations in Pink1 (PARK6), which encodes a serine-threonine kinase localized to mitochondria and parkin (PARK2), which encodes a RING finger-containing E3 ubiquitin ligase have been found in recessively inherited and sporadic PD cases [56, 118]. Drosophila PINK1 and parkin function in the same genetic pathway, with pink1 acting upstream of parkin, to regulate mitochondrial integrity in testes, muscle, and dopaminergic neurons [98, 115, 119]. Flies lacking PINK1 or parkin function are viable and show muscle degeneration and TUNEL staining, indicative of cell death [98, 115, 119]. Parkin and PINK1 mutant flies show locomotor defects, abnormal wing position, and dented thoraces [98, 115, 119]. Mitochondrial defects seen in parkin or PINK1 mutant flies are majorly found in the muscle cells though other cell types like DA neurons in the fly brain also show mitochondrial defects which suggests that loss of PINK1 or parkin would result in systemic mitochondrial defects but not extended to all tissues with similar extent [98, 115, 119]. PINK1 and parkin mutant flies show very similar phenotypes suggesting that these genes act together in protecting mitochondria from damage. Drosophila expressing wild-type parkin in a PINK1 mutant partially rescues PINK1-associated phenotypes. Alternatively, PINK1 wild type expression in a parkin mutant does not rescue the parkin-associated phenotypes which shows that parkin acts downstream of PINK1 [98, 115]. Parkin is also associated with degradation of mitochondria by autophagy and proteasomal degradation of mitochondrial components [113, 120].

Mitochondrial DNA mutations were first associated with different sporadic or maternally inherited neuromuscular disorders [121]. These disorders were characterized by either the accumulation of multiple mtDNA deletions in post-mitotic tissues [122] or tissue-specific mtDNA depletion [123], and a genetic defect affecting nuclear genes involved in mtDNA replication and maintenance [124]. Mitochondrial DNA often exists in a state of heteroplasmy (cells carrying mtDNA of different genotypes), in which mutant mtDNA co-exists in cells with wild-type mtDNA. Pathology is seen when the frequency of such a mutation reaches a threshold [125]. It accumulates throughout life and is thought to contribute to diseases of aging that include neurodegeneration, metabolic disorders, cancer, heart disease, and sarcopenia [126, 127]. A new investigation of mtDNA in the dopaminergic neurons [128] expanded the previous results showing a prevalent deletion in single neurons on a background of multiple mtDNA deletions [129]. It showed that complex I and complex II are most consistently affected in single neurons, which also displayed a reduced mtDNA copy number [128]. Stimulation of autophagy, activation of the PINK1/parkin pathway, or decreased levels of mitofusin results in a selective decrease in lethal mtDNA deletion [130]. Similar to flies lacking parkin, the flies with mtDNA deletions display striking mitochondrial abnormalities such as disrupted cristae resulting in reduced ATP levels leading to apoptosis and subsequently neurodegeneration.

These dynamic processes regulate mitochondrial function by enabling mitochondrial recruitment to critical subcellular compartments, content exchange between mitochondria, mitochondrial shape control, mitochondrial communication with the cytosol, and mitochondrial quality control making the mitochondria readily adapt to changes in cellular requirements. When mitochondrial dynamics is disrupted, cellular dysfunction follows. The view of mitochondrial dynamics has expanded from a curious phenomenon into an integral cell biological process influencing many cellular functions and survival [131].

5.3. Mitochondrial dynamics: fusion and fission

Mitochondria undergo frequent fission and fusion events contributing to the filamentous and interconnected morphology of the organelles and serve crucial physiological functions [132]. The first mitofusin was discovered in mutant Drosophila characterized by male sterility [133]. During cell division, mitochondrial fission plays an important role for proper inheritance of mitochondria to the daughter cells. Moreover, organelle fission appears to be important to meet the energy demands of cells at particular subcellular locations. This is true especially for neurons, which heavily depend on energy supply, where mitochondria can travel via microtubule-associated motor proteins to serve specialized neuronal functions such as synaptic transmission. Furthermore, an interconnected network of contiguous mitochondrial organelles seems to be required for the intramitochondrial exchange of metabolic substrates as well as the maintenance of respiratory capacity and mitochondrial membrane potential in hypoxic cellular regions [134].

Mitochondrial fission and fusion events are regulated and coordinated by evolutionarily highly conserved molecular machineries. The molecules regulating mitochondrial dynamics include the homologous GTPases Mitofusin 1 (Mfn1) and Mitofusin 2 (Mfn2), which mediate fusion of the mitochondrial outer membrane, and Optic atrophy 1 (Opa1), a GTPase required for fusion of the inner membrane. Mitochondrial fission, conversely, requires Dynamin-related protein 1 (Drp1), which is also a GTPase [131, 135]. The Drosophila melanogaster genome encodes two homologs of Mfn, one being Fuzzy onion (Fzo) [133]. The expression of Fzo is restricted to the testes, and mutations in fzo causes mitochondrial fusion defects in testes and male sterility [133]. The second Drosophila Mfn homolog is a largely uncharacterized protein known as mitochondrial assembly regulatory factor (Marf), which is expressed in germline and somatic cells [136]. The Drosophila genome also encodes single homologs of opa1 [137] and drp1 [138], both of which have been shown to function in mitochondrial dynamics in flies.

5.4. Genome-wide association studies (GWAS) and genetic screens

GWAS are large-scale population-based genotyping studies that were designed to capture common genetic risk loci and searches for small variations, called single nucleotide polymorphisms or SNPs, which occur more frequently in people with particular disease than in people without the disease. These approaches lead to the identification of new disease causing genes, new biological pathways to explain disease origin or progression, and potential therapeutic targets which is much more precise than the corresponding information from linkage-based studies.

Maraganore and colleagues conducted the first GWAS for PD. It had limited sample sizes (few hundred patients) and patient-control series were sometimes mismatched, but they remain in the genetic history of PD and suggested the low heritability of PD generating a large amount of genetic data into public domain to be further examined and completed by other researchers [139]. Over the last two decades, human GWAS have begun to reveal the genetic risk factors for countless common disorders with complex genetic etiologies including most of the major causes of morbidity and mortality in the developed world [140].

Involvement of multiple genes and pathways complicate the identifying and developing effective therapeutic strategies in different geographical and ethnically divergent populations. Here lies the necessity and importance of GWAS relating to complex neurological disorders such as PD and cancers. Mutations in five genes have been identified to contribute to Mendelian forms of PD risk in fewer than 5% of those with PD which suggest that additional genes too contribute to disease risk [141]. In 2009, two GWAS papers provided unequivocal evidence for an association of the MAPT locus and SNCA variations with sporadic PD. Additionally, both papers implicated variants close to LRRK2 and at two new loci on chromosome 1 (1q32) and chromosome 4, close to the bone marrow stromal cell antigen 1 (BST1) gene [142]. Shortly after these findings, Pankratz et al. conducted the first GWAS in familial PD, confirming the previous discoveries and providing preliminary evidence for an association of a new locus containing the genes cyclin G association kinase (GAK) and diacylglycerol kinase theta (DGKQ) with PD [142, 143]. A meta-analysis on more than seven million polymorphisms originating either from GWAS datasets and/or from smaller scale PD association studies was performed, where 10 loci showed genome-wide significant association with PD risk (BST1, CCDCC2/HIP1R, DGKQ/GAK, GBA, LRRK2, MAPT, MCCC1/LAMP3, SNCA, STK39, and SYT11/RAB25) and novel evidence for genome-wide significant association with polymorphism in STGA8 was found [144]. To date, the largest GWAS was performed in 2014 carrying out a meta-analysis in all existing European-ancestry PD GWAS data with 13,708 PD cases, 95,282 controls and a replication study using genotyping array called ‘Neuro X’ in an independent data set identified 26 independent SNPs which showed genome-wide risk for PD [145].The first papers about the potential impact of risk loci on age at onset (AAO) in PD were published in 2015. The results using polygenic score analysis showed that patients with an early AAO had a significantly higher polygenic score when compared to those with late AAO [146]. GWAS showed a genetic association with PD in the HLA region (chromosome 6p21.3), which was designated PARK18 and the common variant associated with late onset sporadic PD was rs3129882 in intron 1 of HLA-DRA [147] GWAS sporadic PD was performed and sterol regulatory element binding transcription factor 1 (SREBF1) was identified as risk factor for PD [148]. Later, an unbiased approach on genome-wide RNAi screen was performed in Drosophila cells and validated in Hela cells to identify genes regulating the PINK1/parkin pathway, which act in a common genetic pathway in mediating the autophagic degradation of mitochondria (mitophagy) and found 20 genes that have a conserved function in promoting translocation and degradation of depolarized mitochondria. But most notable genes involved were (SREBF1) Fbox and WD40 domain protein 7 (FBXW7), and other components of the sterol regulatory element binding protein (SREBP) lipogenesis pathway indicating a role of lipids in mitochondrial homeostasis, which further showed that this pathway regulate mitophagy and also share a common mechanistic link between autosomal-recessive and sporadic PD [149].

The loci currently associated with PD account for only a very small amount (3–5%) of the expected heritability of PD, suggesting that additional heritable factors (genetic or epigenetic) also play a role in transforming susceptibility to PD. While current evidence suggests that common genetic variants play a role in the etiology of typical PD. GWAS by their inherent design may not be able to detect rare variants [143, 150]. Most SNPs detected by GWAS are not likely causal variants for disease risk but rather informative markers hence it is often not productive to study their direct functional consequences. Also cases selected for GWAS may not be particularly enhanced with genetic susceptibility alleles, moreover the effect sizes identified for most variants are reduced. Thus multiple approaches including linkage analysis, sequencing, and sibpair analysis would be needed to discover additional variants/causative genes and susceptibility loci. Large-scale genome and exome sequencing in conjunction with denser genotyping in large cohorts may help to identify the loci that contribute to the “missing heritability” previously unnoticed by earlier generation technologies [151].

Advertisement

6. Utilizing Drosophila to understand human GWAS signals

6.1. GAL4-UAS system

In 1993, Brand and Perrimon developed the GAL4-UAS system for precise spatial and temporal patterns directing gene expression in Drosophila and has been considered as a powerful research tool. A bipartite approach in which a transcriptional activator, the GAL4 gene binds to specific cis-enhancer elements, upstream activation sequence (UAS) leading to activation of the adjacent gene, and thousands of GAL4 driver lines available from individual labs and public stock collections allow expression of desired target genes, typically cDNA transgenes under control of upstream activating sequence (UAS) sites [152]. Findings showed that several familial forms of parkinsonism result from increased gene dosage of α-syn, based on this expression levels of the α-syn transgene was augmented by generating an α-syn expression construct bearing sequence alterations designed to improve the translational efficiency of this cDNA in Drosophila and further maximized α-syn protein expression in the fly brain by making use of flies bearing two copies each of the UAS-α-syn transgene and the TH-GAL4 driver; these control approximately doubled the abundance of α-syn protein relative to flies bearing a single copy of each of these transgenes generating a more robust Drosophila model for studying synucleinopathies [153].

6.2. RNA interference (RNAi)

RNAi is an RNA-dependent gene-silencing process that is regulated by the RNA-induced silencing complex (RISC) and triggered by short double-stranded RNA (dsRNA) molecules.

Efficient silencing of gene expression by dsRNA was first discovered by Fire and Mello [154]. RNAi silencing of a specific target gene relies on the ability of small interfering RNAs (siRNAs), long double-stranded RNAs (dsRNAs), or short hairpin RNAs (shRNAs) to target mRNA molecules for degradation [155, 156, 157]. Drosophila in vivo RNAi techniques screen both the whole genome and subsets of genes. A total of 10,689 different genes (78% of the Drosophila genome) were assayed that affect susceptibility to intestinal Serratia marcescens infection [158]. Of these, 8.3% (885 genes) were defined as hits; the majority 89.3% (790 genes) were susceptibility candidates, and 95 genes (10.7% of hits) were negative regulators. A total of 78 and 56 genes were found to function only in the gut and hemocytes, respectively, and 79 functioned in both. A primary screened of 6923 UAS-IR strains for genes involved in the glycosylation of a neural glycoprotein and identified 171candidates [159]. These were further confirmed by knock down experiments, using in-silico analysis and a secondary set of UAS-IR strains that targeted regions distinct from those of the primary strains. A total of 2970 genes were knocked down by neuron-specific RNAi in search for genes involved in the formation, growth, and maintenance of the neuromuscular junction (NMJ) [160]. Knockdown of 158 genes in post-mitotic neurons led to abnormalities in the neuromuscular system. Genome-wide small interfering RNA (siRNA) screening yielded gene candidates involved in characterization of TOMM7 required for stabilizing PINK1 on the outer mitochondrial membrane following mitochondrial damage. Also, HSPA1L (HSP70 family member) and BAG4 found to have mutually opposing roles in the regulation of parkin translocation. RNAi screens revealed that SIAH3 localize to mitochondria, inhibits PINK1 accumulation after mitochondrial insult, and reducing parkin translocation. Overall, screens provide a rich resource to understand mitochondrial quality control [161].

Using this inducible RNAi technique, large-scale screens for various biological processes have been performed successfully in Drosophila, proving RNAi-based in vivo screen adequate and efficient. However, RNAi-based screens have relatively high levels of false positives and negatives. To validate the screening results, experimental, and computational analyses have been proposed which will increase the accuracy of RNAi-based screen results.

6.3. CRISPR/Cas9

Clustered regularly interspaced short palindromic repeat/CRISPR-associated protein 9 (CRISPR/Cas9) was first applied in mammalian cells in 2013 which has been used as an essential tool in biotechnology [162]. The CRISPR/Cas9 system is a novel genome modification method in which gRNA direct the nuclease Cas9 to selected sequences of genomic DNA, and Cas9 cuts both strands at a specific location. Non-homologous end joining (NHEJ) or homology-directed repair (HDR) repairs the genomic DNA resulting in mutations that can interrupt the open reading frame and cause gene inactivation. For example, loss in function of parkin and Pink1 genes causes PD, CRISPR/Cas9-mediated mutations can mimic knockout of the parkin and/or Pink1 gene [163]. Thus, when both alleles are mutated by CIRSPR/Cas9, the complete loss of parkin or Pink1 will mimic the genetic mutations in PD patients. Indeed, CRISPR/Cas9 has been used to generate pig models of PD by targeting the genes for parkin, Pink1, and DJ1 [164]. In addition to genome editing in germline cells, CRISPR/Cas9 can efficiently target genes in somatic tissues, such as neurons in the brain [165, 166]. In PD patients, continuous loss of dopaminergic neurons in the substantia nigra is a fundamental pathological feature. Thus, gRNAs and Cas9 can be delivered to the substantia nigra of animal brains by a viral system to examine the effect of parkin or Pink1 loss in adult brains. This approach is especially useful for investigating the age-related neuropathology in PD. Also, by gaining toxicity of mutant proteins Cas9-mediated knock-in mutations within the genome can develop animal models of those neurodegenerative diseases.

For transgenic PD animal models that express mutant a-syn, CRISPR/Cas9 can be designed to reduce the expression of mutant genes via NHEJ, which can lead to gene inactivation, in dopaminergic neurons. Besides, replacing the mutant gene by CRISPR/Cas9 via HDR with normal DNA sequences can also lead to the genetic correction of DNA mutations in PD animal models. Even though efficiency of such gene replacement is low at present, the rapid development of CRISPR/Cas9 system offers a promising attempt to produce knock-in models of human diseases [167].

6.4. Deciphering the pathways of therapeutic molecules: role of drosophila model

The first published study about compound treatments in a drosophila PD model reported the effects of drugs commonly used for treating PD on the locomotor phenotype of α-synuclein expressing flies and showed that some of them were able to suppress that phenotype [168]. Subsequently, and given the ability of increased chaperone activity to counteract α-synuclein toxicity [169], the effect of geldanamycin (GA), an antibiotic able to interfere with Hsp90 activity and activate stress response, was assayed over α-synuclein expressing flies [170, 171]. Notably, feeding these flies with GA protected DA neurons against α-synuclein-induced degeneration, and this protection was driven by an increase in Hsp70 levels [171]. Inhibitors of the histone deacetylase SIRT2 also showed a protective effect against α-synuclein toxicity [172]. Other studies have been also performed in several drosophila PD models to look for potentially therapeutic compounds directed to reduce oxidative stress damage. As explained previously, the study of α-synuclein toxicity in flies led to the identification of Phase II detoxification pathway as a possible target for therapeutic treatment [173]. In fact, feeding α-synuclein-expressing flies or Drosophila parkin mutants with pharmacological inducers of that pathway like sulforaphane or allyl disulfide suppresses the neuronal loss of both PD models [173]. Besides, it has been shown that dietary supplementation with S-methyl-L-cysteine (SMLC) inhibits the locomotor and circadian rhythm defects caused by ectopic expression of human α-synuclein in drosophila [174]. SMLC participates in the catalytic antioxidant mechanism involving methionine sulfoxide reductase A (MSRA), one of the enzymes that catalyze the oxidation of the amino acid methionine to methionine sulfoxide, a reversible reaction that has been postulated to act protecting cells from oxidative damage. Furthermore, grape extract supplementation has been shown to recover locomotor ability and lifespan in α-synuclein-expressing flies. It is known that grape extracts contain several polyphenols, compounds with antioxidant properties [175]. Other drosophila PD models in which treatments with antioxidant compounds have been shown to be beneficial are those involving the DJ-1α and DJ-1β genes [176, 177]. Compounds with antioxidant and anti-inflammatory properties such as celastrol and minocycline conferred potent DA neuroprotection in RNAi DJ-1α mutants [176]. We have also recently demonstrated that chronic treatments with antioxidant compounds are able to modify the lifespan phenotype of DJ-1β mutant flies, thus suggesting that oxidative stress plays a causal role in such phenotype [177]. It is known that rapamycin is a small molecule inhibitor of TOR signaling that has been shown to lead to 4E–BP hypophosphorylation in vitro and in vivo [178, 179]. Notably administration of rapamycin was able to suppress all pathologic phenotypes in parkin and PINK1 mutants. Moreover, this suppression was found to be 4E-BP dependent, since the administration of rapamycin to parkin and Thor or PINK1 and Thor double mutants was completely unable to suppress these phenotypes [180]. Since 4E-BP activity can be manipulated by small molecule inhibitors such as rapamycin, this pathway represents a viable therapeutic target for PD treatment. Moreover, it has been recently suggested that parkin mutants, apart from the described phenotypes, also present altered zinc homeostasis. This is supported by the fact that dietary zinc supplementation in the form of zinc chloride increased lifespan as well as the percentage of parkin mutant flies reaching adulthood while this supplemented diet was deleterious to control flies [181]. Since most PD cases are sporadic and could be associated to different environmental agents, it is also essential the use of toxin-induced drosophila PD models to assay the beneficial effects of candidate compounds. Polyphenol administration was also found to exert a beneficial effect on flies exposed to paraquat and iron, protecting, rescuing, and restoring the impaired locomotor activity caused by exposure to those agents [182]. Other antioxidant compounds such as melatonin have also been found to rescue locomotor deficits and DA neurodegeneration in flies exposed to rotenone [19].

6.4.1. LRRK2 kinase inhibitors

Several LRRK2 kinase inhibitors, including CZC-25146, GW5074, and sorafenib, have been tested in rodents, as well as in Caenorhabditis elegans and drosophila models, and have been shown to protect against LRRK 2 (G2019S)-induced neurodegeneration [183]. These findings indicate that increased kinase activity of LRRK2 is neurotoxic and that inhibition of LRRK2 activity can have a disease-modifying effect.

6.4.2. Molecular chaperones

6.4.2.1. HSF-1 modulators

Endogenous molecular chaperone function can be modulated pharmacologically with compounds that augment endogenous chaperone levels. Several HSF-1 modulators including celastrol and carbenoxolone can trigger HSF-1 activation, leading to downstream induction of Hsp70 expression [184]. Celastrol has been demonstrated to be effective against protein aggregation and toxicity in various neurodegenerative disease models, including dopaminergic neuroprotection in a Drosophila model of PD [176]. Carbenoxolone has demonstrated the ability to attenuate 𝛼-synuclein and ubiquitin aggregation in vitro and in vivo [185, 186]. Thus, it may have potential as a chaperone-mediated therapeutic option for PD. Hsp90 Inhibitors a naturally occurring small molecule antibiotic, geldanamycin (GA), inhibits the interaction between Hsp90 and HSF-1, leading to increased Hsp70 expression [187]. In vitro cell studies have demonstrated the capability of this compound to decrease 𝛼-synuclein aggregation and reduce cell toxicity [188], and its neuroprotective effects have been shown in Drosophila and MPTP mouse models of PD [169, 189]. Other analogues of GA include 17-AAG and 17-DMAG, which similarly prevent 𝛼-synuclein aggregation and toxicity, but are more potent and less toxic than GA [190]. Moreover, 17-AAG has poor permeability of the BBB, limiting its pharmacological usage for neurodegenerative diseases [185]. Consequently, compound library screening for small molecule Hsp90 inhibitors with improved pharmacokinetics, including BBB permeability, have led to the identification of SNX compounds [185]. These compounds are associated with an increase in Hsp70 activity in the brain and a reduction in 𝛼-synuclein oligomerization and toxicity in vitro [190].

Insights regarding identification of pathways through which different therapeutic molecules confer neuroprotection is briefed in Table 4 .

Pathway/process Compound treatment * Drosophila model Modified phenotype/s References
Oxidative stress Sulforaphane and allyl disulfide *
S-methyl-Lcysteine
Polyphenols
α-tocopherol
SOD
Melatonin
Bacopa monieri
leaf extract
Parkin
α-synuclein
α-synuclein
α-synuclein
Paraquat and iron
DJ-1β
PINK1
PINK1
DJ-1β
Paraquat
Rotenone
Paraquat
DA neuron number
DA neuron number
Locomotor activity
Lifespan, locomotor activity
Locomotor activity
Lifespan
Ommatidial degeneration
Ommatidial degeneration
Lifespan
Locomotor activity
Locomotor activity, DA neuron
Oxidative markers levels
[173]
[173]
[174]
[175]
[182]
[177]
[191]
[191]
[177]
[99]
[99]
[192]
Oxidative stress/inflammatory process Minocycline *
Celastrol
DJ-1α DA neuron number, dopamine Levels
DA neuron number, dopamine levels, locomotor activity, and
survival rate under oxidativestress condition
[176]
TOR signaling Rapamycin* parkin/PINK1 Thoracic indentations, locomotors activity, DA neuron number, and muscle integrity [180]
Removal of excess or toxic protein forms Geldanamycin* α-synuclein DA neuron number [170]
[171]
Zinc homeostasis Zinc chloride* parkin Lifespan, locomotor activity, and percentage of adulthood
survivors
[181]
Chaperone therapies (HSF-1 modulators)
Trigger HSF-1 activation
Induces downstream Hsp70 expression
Celastrol
Carbenoxolone
α-synuclein
α-synuclein
dopaminergic neuroprotection [184]
Hsp90 inhibitors
Inhibits the interaction between Hsp90 and HSF-1, leading to increased Hsp70 expression and activity
Geldanamycin
17-AAG
17-DMAG
SNX-2112
α-synuclein decrease α- synuclein aggregation and reduce cell toxicity [187]
[190]
mTOR-dependent pathways/AMPK Metformin
AICAR
Drosophila melanogaster
mutated for LRRK2
Reduced cell death [193]
mTORC1 Rapamycin and Rp analogues
(CCI-779, RAD001 and AP23573)
Drosophila melanogaster
mutated for PINK-1 and
Parkin
Reduced mitochondrial
Dysfunction
[180]
mTor-independent pathways/unknown Spermidine α-synuclein Reduced motor dysfunction,
increased lifespan; Reduced
neuronal cell loss
[194]
LRKK2 kinase inhibitors GW5074, and sorafenib α-synuclein Protect again DA neuron degeneration
locomotor activity
[183]
Histone Deacetylase inhibitors α-synuclein Protect again DA
locomotor activity
[195]
Antitumor agents Geldanamycin α-synuclein Protect again DA
Mobilized the stress response and increase levels of chaperon HSP70
[195]

Table 4.

Listing of pathways through which neuroprotective compounds confer neuroprotection: lessons from drosophila model of PD.

All treatments were administered as dietary complement.


Advertisement

7. Conclusion

Bench to bedside: role of drosophila in translational research.

Bench to bedside is a term used to describe the process by which the results of research done in the laboratory are directly used to develop new ways to treat patients. Taking advantage of studies from animal models such as drosophila certain pharmacotherapies and non-pharmacotherapies have been developed which are in different stages in clinical trials to validate their efficacy, safety, and tolerability. Pharmacotherapies include adenosine A2A receptor antagonists [196], glutamate AMPA receptor antagonists [197], monoamine oxidase inhibitors [198], anti-apoptotic agents [199], and antioxidants [200]. Non-pharmacotherapies also offer alternative approaches for treatment of the disease which include the use of viral vector gene therapy [201], stem cell transplants [202], and microRNAs [203]. Nevertheless, additional trials enrolling larger numbers of PD patients are still needed to better understand the neuroprotective effects of these therapies.

Advertisement

Acknowledgments

This work is partly supported by Department of Biotechnology (DBT), India grant numbers (BT/405/NE/U-Excel/2013 and BT/PR16868/NER/95/328/2015) and DST-SERB (Department of Science and Technology-Science and Engineering Research Board), India grant number (EMR/2016/002375).

Advertisement

Conflict of interest

Authors declares no conflict of interest.

References

  1. 1. Manyam BV. Paralysis agitans and levodopa in “Ayurveda”: Ancient Indian medical treatise. Movement Disorders. 1990;5:47-48
  2. 2. Zhang ZX, Dong ZH, Roman GC. Early descriptions of Parkinson’s disease in ancient China. Archives of Neurology. 2006;63:782-784
  3. 3. Garcia-Ruiz PJ. Prehistoria de la enfermedad de Parkinson (Prehistory of Parkinson's disease). Neurologia (in Spanish). 2004;19(10):735-737
  4. 4. Available from: http://www.parkinsons.org/parkinsons-history.html [Accessed: Oct 2, 2017]
  5. 5. Parkinson J. An Essay on the Shaking Palsy. Needly and Jones, London: Whittingham and Rowland for Sherwood; 1817
  6. 6. Lees AJ. Unresolved issues relating to the shaking palsy on the celebration of James Parkinson's 250th birthday. Movement Disorders. 2007;22(17):S327-S334
  7. 7. Charcot J-M. De la paralysie agitante. In: Oeuvres Comple’tes (t 1) Lec¸ons sur les maladies du syste’me nerveux. Paris: A Delahaye; 1877. pp. 155-188. [In English: Charcot, J-M. On Parkinson’s disease. In: Lectures on Diseases of the Nervous System Delivered at the Salpeˆtrie’re (transl. Sigerson G). London: New Sydenham Society; 1872. pp. 129-156]
  8. 8. Modi P, Ayajuddin M, Phom L, et al. Understanding pathophysiology of sporadic Parkinson’s disease in Drosophila model: Potential opportunities and notable limitations. In: Jolanta Dorszewska and Wojciech Kozubski, editors. Challenges in Parkinson's Disease. European Union: In-tech Croatia; 2016. pp. 217-244
  9. 9. Postuma RB, Berg D, Stern M, et al. MDS clinical diagnostic criteria for Parkinson's disease. Movement Disorders. 2015;30(12):1591-1601
  10. 10. Jellinger KA. Formation and development of Lewy pathology: A critical update. Journal of Neurology. 2009;256(3):270-279
  11. 11. Hornykiewicz O, Kish SJ. Biochemical pathophysiology of Parkinson’s disease. Advances in Neurology. 1987;45:19-34
  12. 12. Pfeiffer RF. Non-motor symptoms in Parkinson's disease. Parkinsonism & Related Disorders. 2016;22:S119-S122
  13. 13. Levy G, Schupf N, Tang M, et al. Combined effect of age and severity on the risk of dementia in Parkinson's disease. Annals of Neurology. 2002;51(6):722-729
  14. 14. Di Monte DA, Lavasani M, Manning Bog AB. Environmental factors in Parkinson's disease. Neurotoxicology. 2002;23:487-502
  15. 15. Polymeropoulos MH, Lavedan C, Leroy E, et al. Mutation in the alpha-synuclein gene identified in families with Parkinson's disease. Science. 1997;276:2045-2047
  16. 16. Simon-Sanchez J, Schulte C, Bras J. Genome-wide association study reveals genetic risk underlying Parkinson’s disease. Nature Genetics. 2009;41:1308-1312
  17. 17. Elbaz A, Clavel J, Rathouz PJ, et al. Professional exposure to pesticides and Parkinson disease. Annals of Neurology. 2009;66:494-504
  18. 18. Ritz BR, Manthripragada AD, Costello S, et al. Dopamine transporter genetic variants and pesticides in Parkinson's disease. Environmental Health Perspectives 2009;117:964-969
  19. 19. Coulom H, Birman S. Chronic exposure to rotenone models sporadic Parkinson’s disease in Drosophila melanogaster. The Journal of Neuroscience. 2004;24(48):10993-10998
  20. 20. Liou HH, Tsai MC, Chen CJ, et al. Environmental risk factors and Parkinson's disease – A case-control study in Taiwan. Neurology. 1997;48(6):1583-1588
  21. 21. Ballard PA, Tetrud JW, Langston JW. Permanent human parkinsonism due to 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP): Seven cases. Neurology. 1985;35(7):949-956
  22. 22. Langston JW, Ballard P, Tetrud JW, Irwin I. Chronic parkinsonism in humans due to a product of meperidine-analog synthesis. Science. 1983;219:979-980
  23. 23. Priyadarshi A, Khuder SA, Schaub EA, et al. A meta-analysis of Parkinson’s disease and exposures to pesticides. Neurotoxicology. 2000;21:435-440
  24. 24. Steenland K, Hein MJ, Cassinelli RT, et al. Polychlorinated biphenyls and neurodegenerative disease mortality in an occupational cohort. Epidemiology. 2006;17(1):8-13
  25. 25. Goldman SM, Quinlan PJ, Ross GW, et al. Solvent exposures and Parkinson disease risk in twins. Annals of Neurology. 2012;71(6):776-784
  26. 26. Gorell JM, Johnson CC, Rybicki BA, et al. Occupational exposures to metals as risk factors for Parkinson's disease. Neurology. 1997;48(3):650-658
  27. 27. Finkelstein MM, Jerrett M. A study of the relationships between Parkinson's disease and markers of traffic-derived and environmental manganese air pollution in two Canadian cities. Environmental Research. 2007;104(3):420-432
  28. 28. Vitalis T, Fouquet C, Alvarez C, et al. Developmental expression of monoamine oxidases A and B in the central and peripheral nervous systems of the mouse. The Journal of Comparative Neurology. 2002;442(4):331-347
  29. 29. Singer TP, Ramsay RR, McKeown K, et al. Mechanism of the neurotoxicity of 1-methyl-4-phenylpyridinium (MPP+), the toxic bioactivation product of 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP). Toxicology. 1988;49(1):17-23
  30. 30. Ramsay RR, Singer TP. Energy-dependent uptake of N-methyl-4-phenylpyridinium, the neurotoxic metabolite of 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine, by mitochondria. The Journal of Biological Chemistry. 1986;261(17):7585-7587
  31. 31. Grant H, Lantos PL, Parkinson C. Cerebral damage in paraquat poisoning. Histopathology. 1980;4:185-195
  32. 32. Bové J, Prou D, Perier C, et al. Toxin-induced models of Parkinson's disease. NeuroRx. 2005;2(3):484-494
  33. 33. Mohammadi-Bardbori A,, Ghazi-Khansari M. Alternative electron acceptors: Proposed mechanism of paraquat mitochondrial toxicity. Environmental Toxicology and Pharmacology. 2008;26(1):1-5
  34. 34. Peng J, Ou Mao X, Stevenson FF, et al. The herbicide paraquat induces dopaminergic nigral apoptosis through sustained activation of the JNK pathway. Journal of Biological Chemistry. 2004;279(31):32626-32632
  35. 35. Sherer TB, Betarbet R, Testa CM, et al. Mechanism of toxicity in rotenone models of Parkinson's disease. The Journal of Neuroscience. 2003;23(34):10756-10764
  36. 36. Schapira AH, Cooper JM, Dexter D, et al. Mitochondrial complex I deficiency in Parkinson's disease. Journal of Neurochemistry. 1990;54(3):823-827
  37. 37. Di Monte DA: The environment and Parkinson's disease:is the nigrostriatal system preferentially targeted by neurotoxins? Lancet Neurol. 2003;2(9):531-538
  38. 38. Ryu EJ, Harding HP, Angelastro JM, et al. Endoplasmic reticulum stress and the unfolded protein response in cellular models of Parkinson's disease. Journal of Neuroscience. 2002;22:10690-10698
  39. 39. Sherer TB, Betarbet R, Stout AK, et al. An in vitro model of Parkinson’s disease: Linking mitochondrial impairment to altered a-synuclein metabolism and oxidative damage. Journal of Neuroscience. 2002;22:7006-7015
  40. 40. Schapira A, Jenner P. Etiology and pathogenesis of Parkinson’s disease. Movement Disorders. 2011;26:1049-1055
  41. 41. Polymeropoulos MH, Higgins JJ, Golbe LI, et al. Mapping of a gene for Parkinson’s disease to chromosome 4q21–q23. Science. 1996;274:1197-1199
  42. 42. Spillantini MG, Crowther RA, Jakes R, et al. Alpha-Synuclein in filamentous inclusions of Lewy bodies from Parkinson’s disease and dementia with lewy bodies. Proceedings of the National Academy of Sciences of the United States of America. 1998;95:6469-6473
  43. 43. Sheerin UM, Houlden H, Wood NW. Advances in the genetics of Parkinson’s disease: A guide for the clinician. Movement Disorders Clinical Practice. 2014;1:3-13
  44. 44. Funayama M, Hasegawa K, Kowa H, et al. A new locus for Parkinson’s disease (PARK8) maps to chromosome 12p11.2–q13.1. Annals of Neurology. 2002;51:296-301
  45. 45. Zimprich A, Biskup S, Leitner P, et al. Mutations in LRRK2 cause autosomal-dominant parkinsonism with pleomorphic pathology. Neuron. 2004;44:601-607
  46. 46. Healy DG, Falchi M, O’Sullivan SS, et al. Phenotype, genotype, and worldwide genetic penetrance of LRRK2-associated Parkinson’s disease: A case-control study. Lancet Neurology. 2008;7:583-590
  47. 47. Di Fonzo A, Tassorelli C, De Mari M, et al. Comprehensive analysis of the LRRK2 gene in sixty families with Parkinson’s disease. European Journal of Human Genetics. 2006;14:322-331
  48. 48. Khan NL, Brooks DJ, Pavese N, et al. Progression of nigrostriatal dysfunction in a parkin kindred: An [18F]dopa PET and clinical study. Brain. 2002;125:2248-2256
  49. 49. Kitada T, Asakawa S, Hattori N, et al. Mutations in the parkin gene cause autosomal recessive juvenile parkinsonism. Nature. 1998;392:605-608
  50. 50. Matsumine H, Saito M, Shimoda-Matsubayashi S, et al. Localization of a gene for an autosomal recessive form of juvenile Parkinsonism to chromosome 6q25.2-27. American Journal of Human Genetics. 1997;60(3):588-596
  51. 51. Matsumine H, Yamamura Y, Hattori N, et al. A microdeletion of D6S305 in a family of autosomal recessive juvenile parkinsonism (PARK2). Genomics. 1998;49(1):143-146
  52. 52. Lohmann E, Periquet M, Bonifati V, et al. How much phenotypic variation can be attributed to parkin genotype? Annals of Neurology. 2003;54:176-185
  53. 53. Bonifati V, Rizzu P, Van Baren MJ, et al. Mutations in the DJ-1 Gene associated with autosomal recessive early-onset parkinsonism. Science. 2003;299:256-259
  54. 54. Abou-Sleiman PM, Healy DG, Quinn N, et al. The role of pathogenic DJ-1 mutations in Parkinson’ s disease. Annals of Neurology 2003;54:283-286
  55. 55. Massano J, and Bhatia KP. Clinical approach to Parkinson’s disease: Features, diagnosis, and principles of management. Cold Spring Harbor Perspectives in Medicine 2012;2(6):a008870. DOI: 10.1101/cshperspect.a008870
  56. 56. Valente EM, Abou-Sleiman PM, Caputo V, et al. Hereditary early-onset Parkinson’s disease caused by mutations in PINK1. Science. 2004;304:1158-1160
  57. 57. Samaranch L, Lorenzo-Betancor O, Arbelo JM, et al. PINK1-linked parkinsonism is associated with Lewy body pathology. Brain. 2010;133:1128-1142
  58. 58. Nuytemans K, Theuns J, Cruts M, et al. Genetic etiology of Parkinson disease associated with mutations in the SNCA, PARK2, PINK1, PARK7, and LRRK2 genes: A mutation update. Human Mutation. 2010;31:763-780
  59. 59. Vilariño-Güell C, Wider C, Ross OA, et al. VPS35 mutations in Parkinson disease. American Journal of Human Genetics. 2011;89(1):162-167
  60. 60. Zimprich A, Benet-Pagès A, Struhal W, et al. A mutation in VPS35, encoding a subunit of the retromer complex, causes late-onset Parkinson disease. American Journal of Human Genetics. 2011;89:168-175
  61. 61. Najim al-Din AS, Wriekat A, Mubaidin A, et al. Pallido-pyramidal degeneration, supranuclear upgaze paresis and dementia: Kufor-Rakeb syndrome. Acta Neurologica Scandinavica. 1994;89:347-352
  62. 62. Hampshire DJ, Roberts E, Crow Y, et al. Kufor-Rakeb syndrome, pallido-pyramidal degeneration with supranuclear up gaze paresis and dementia, maps to 1p36. Journal of Medical Genetics. 2001;38:680-682
  63. 63. Ning YP, Kanai K, Tomiyama H, et al. PARK9-linked parkinsonism in eastern Asia: Mutation detection in ATP13A2 and clinical phenotype. Neurology. 2008;70:1491-1493
  64. 64. Schneider SA, Coro Paisan-Ruiz, Niall P et al.: ATP13A2 mutations (PARK9) cause neurodegeneration with brain iron accumulation. Movement Disorders, The Official Journal of MDS. 2010;25:979-984
  65. 65. Paisan-Ruiz C, Bhatia KP, Li A, et al. Characterization of PLA2G6 as a locus for dystonia-parkinsonism. Annals of Neurology. 2009;65:19-23
  66. 66. Yoshino H, Tomiyama H, Tachibana N, et al. Phenotypic spectrum of patients with PLA2G6 mutation and PARK14-linked parkinsonism. Neurology. 2010;75:1356-1361
  67. 67. Baburina I, Jackowski S. Cellular responses to excess phospholipid. The Journal of Biological Chemistry. 1999;274:9400-9408
  68. 68. Leroy E, Boyer R, Auburger G, et al. The ubiquitin pathway in Parkinson's disease. Nature. 1998;395:451-452
  69. 69. Ragland M, Hutter C, Zabetian C, et al. Association between the ubiquitin carboxyl-terminal esterase L1 gene (UCHL1) S18Y variant and Parkinson's disease: A HuGE review and meta analysis. American Journal of Epidemiology. 2009;170:1344-1357
  70. 70. Ardley HC, Scott GB, Rose SA, et al. UCH-L1 aggresome formation in response to proteasome impairment indicates a role in inclusion formation in Parkinson's disease. Journal of Neurochemistry. 2004;90:379-391
  71. 71. Strauss KM, Martins LM, Plun-Favreau H, et al. Loss of function mutations in the gene encoding Omi/HtrA2 in Parkinson’s disease. Human Molecular Genetics. 2005;14:2099-2111
  72. 72. Bogaerts V, Nuytemans K, Reumers J, et al. Genetic variability in the mitochondrial serine protease HTRA2 contributes to risk for Parkinson disease. Human Mutation. 2008;29:832-840
  73. 73. Suzuki Y, Imai Y, Nakayama H, et al. A serine protease, HtrA2, is released from the mitochondria and interacts with XIAP, inducing cell death. Molecular Cell. 2001;8:613-621
  74. 74. Whitworth AJ, Lee JR, Ho VM, et al. Rhomboid-7 and HtrA2/Omi act in a common pathway with the Parkinson’s disease factors Pink1 and parkin. Disease Models & Mechanisms. 2008;1:168-174
  75. 75. Yun J, Cao JH, Dodson MW, et al. Loss of function analysis suggests that Omi/HtrA2 is not an essential component of the PINK1/PARKIN pathway in vivo. The Journal of Neuroscience. 2008;283:14500-14510
  76. 76. Chartier-Harlin MC, Dachsel J, Hulihan M, et al. EIF4G1 mutations in familial parkinsonism. Parkinsonism & Related Disorders. 2009;15:145-146
  77. 77. Nuytemans K, Bademci G, Inchausti V, et al. Whole exome sequencing of rare variants in EIF4G1 and VPS35 in Parkinson disease. Neurology. 2013;80:982-989
  78. 78. Wooten GF. Functional anatomical and behavioral consequences of dopamine receptor stimulation. Annals of the New York Academy of Sciences. 1997;835:153-156
  79. 79. Jellinger KA. The pathology of Parkinson’s disease. Advances in Neurology. 2001;86:55-72
  80. 80. Lewy FH. Paralysis agitans. Pathologische anatomie. In: Lewandowski M, editor. Handbuch der Neurologie. Berlin: Springer-Verlag; 1912. pp. 920-933
  81. 81. Pienaar IS, Gotz J, Feany MB. Parkinson’s disease: Insights from non-traditional model organisms. Progress in Neurobiology. 2010;92:558-571
  82. 82. Bilen J, Bonini NM. Drosophila as a model for human neurodegenerative disease. Annual Review of Genetics. 2005;39:153-171
  83. 83. Venken KJ, Bellen HJ. Emerging technologies for gene manipulation in Drosophila melanogaster. Nature Reviews. Genetics. 2005;6:167-178
  84. 84. White KE, Humphrey DM, Hirth F. The dopaminergic system in the aging brain of Drosophila Frontier in Neuroscience. 2010;4(205):1-12
  85. 85. Lessing D, Bonini NM. Maintaining the brain: Insight into human Neurodegeneration from Drosophila mutants. Nature Reviews. Genetics. 2009;10:359-370
  86. 86. Lloyd TE, Taylor JP. Flightless flies: Drosophila models of neuromuscular disease. Annals of the New York Academy of Sciences. 2010;1184:1-20
  87. 87. Pandey UB, Nichols CD. Human disease models in Drosophila melanogaster and the role of the fly in therapeutic drug discovery. Pharmacological Reviews. 2011;63:411-436
  88. 88. Prüßing K, Voigt A, Schulz JB. Drosophila melanogaster as a model organism for Alzheimer’s disease. Molecular Neurodegeneration. 2013;8:35-45
  89. 89. Mayer F, Mayer N, Chinn L, Pinsonneault RL et al. Evolutionary conservation of vertebrate blood-brain barrier chemoprotective mechanisms in Drosophila. The Journal of Neuroscience. 2009;29:3538-3550
  90. 90. Feany MB, Bender WW. A Drosophila model of Parkinson’s disease. Nature. 2000;404:394-398
  91. 91. Botella JA, Bayersdorfer F, Gmeiner F et al. Modelling Parkinson’s disease in Drosophila. NeuroMolecular Medicine. 2009;11:268-280
  92. 92. Hirth F. Drosophila melanogaster in the study of human neurodegeneration. CNS & Neurological Disorders – Drug Targets. 2010;9:504-523
  93. 93. Varga SJ, Qi C, Podolsky E, et al. A new Drosophila model to study the interaction between genetic and environmental factors in Parkinson’s disease. Brain Research 2014;1583:277-286
  94. 94. Meulener M, Whitworth AJ, Armstrong-Gold CE, et al. Drosophila DJ-1 mutants are selectively sensitive to environmental toxins associated with Parkinson's disease. Current Biology. 2005;15:1572-1577
  95. 95. Menzies FM, Yenisetti SC, Min KT. Roles of Drosophila DJ-1 in survival of dopaminergic neurons and oxidative stress. Current Biology. 2005;15:1578-1582
  96. 96. Venderova K, Kabbach G, Abdel-Messih E, et al. Leucine-rich repeat kinase 2 interacts with parkin, DJ-1 and PINK-1 in a Drosophila Melanogaster model of Parkinson's disease. Human Molecular Genetics. 2009;18:4390-4404
  97. 97. Wang C, Lu R, Ouyang X, et al. Drosophila overexpressing parkin R275W mutant exhibits dopaminergic neuron degeneration and mitochondrial abnormalities. The Journal of Neuroscience. 2007;27(32):8563-8570
  98. 98. Clark IE, Dodson MW, Jiang C, et al. Drosophila pink1 is required for mitochondrial function and interacts genetically with parkin. Nature. 2006;441(7097):1162-1166
  99. 99. Chaudhuri A, Bowling K, Funderburk C, et al. Interaction of genetic and environmental factors in a Drosophila parkinsonism model. The Journal of Neuroscience. 2007;27(10):2457-2467
  100. 100. Park J, Kim SY, Cha GH, et al. Drosophila DJ-1 mutants show oxidative stress-sensitive locomotive dysfunction. Gene. 2005;361:133-139
  101. 101. Hosamani R, Muralidhara. Acute exposure of Drosophila melanogaster to Paraquat causes oxidative stress and mitochondrial dysfunction. Archives of Insect Biochemistry and Physiology. 2013;83(1):25-40
  102. 102. Ebrahimi-Fakhari D, Wahlster L, McLean P. Protein degradation pathways in Parkinson’s disease: Curse or blessing. Acta Neuropathologica. 2012;124:153-172
  103. 103. Wang A, Costello S, Cockburn M, et al. Parkinson’s disease risk from ambient exposure to pesticides. European Journal of Epidemiology. 2011;26(7):547-555
  104. 104. Martin CA, Barajas A, Lawless G, et al. Synergistic effects on dopamine cell death in a Drosophila model of chronic toxic exposure. Neurotoxicology. 2014;44:344-351
  105. 105. Sun F, Kanthasamy A, Anantharam V, et al. Envionmental neurotoxic chemicals-induced ubiquitin proteosome system dysfunction in the pathogenesis and progression of parkinson’s disease. Pharmacology & Therapeutics. 2007;114:327-344
  106. 106. Wang C, Ko HS, Thomas B, et al. Stress-induced alterations in parkin solubility promote parkin aggregation and compromise parkin's protective function. Human Molecular Genetics. 2005;14(24):3885-3897
  107. 107. Yao D, Gu Z, Nakamura T, et al. Nitrosative stress linked to sporadic Parkinson's disease: S-nitrosylation of parkin regulates its E3 ubiquitin ligase activity. Proceedings of the National Academy of Sciences of the United States of America. 2004;101(29):10810-10814
  108. 108. Casarejos MJ, Menendez J, Solano RM, et al. Susceptibility to rotenone isincreased in neurons from parkin null mice and is reduced by minocycline. Journal of Neurochemistry. 2006;97(4):934-946
  109. 109. Liao J, Morin LW, Ahmad ST. Methods to characterize spontaneous and startle induced locomotion in a rotenone induced parkinson’s disease model of Drosophila. Journal of Visualized Experiments. 2014;90:1-6
  110. 110. Shu YU, Meng J, Ying X, et al. Rotenone induces more serious learning and memory impairment that α-synuclein A30P does in Drosophila. Journal of Shanghai Jiaotong University. 2011;15(3):229-234
  111. 111. Khatri DK, Juvekar AR. Abrogation of locator impairment in a rotenone-induced Drosophila melanogaster and zebrafish model of parkinson’s disease by ellagic acid and curcumin. International Journal of Nutrition, Pharmacology, Neurological Diseases. 2016;6(2):90-96
  112. 112. Subramanian P, Prasanna V, Jayapalan JJ, et al. Role of Bacopa monnieri in the temporal regulation of oxidative stress in clock mutant (cryb) of Drosophila melanogaster. Journal of Insect Physiology. 2014;65:37-44
  113. 113. Vanhauwaert R, Verstreken P: Flies with Parkinson's disease. Experimental Neurology. 2015;274(Pt A):42-51
  114. 114. Greene JC, Whitworth AJ, Kuo I, et al. Mitochondrial pathology and apoptotic muscle degeneration in Drosophila parkin mutants. Proceedings of the National Academy of Sciences of the United States of America. 2003;100(7):4078-4083
  115. 115. Park J, Lee SB, Lee S, et al. Mitochondrial dysfunction in Drosophila PINK1 mutants is complemented by parkin. Nature. 2006;441(7097):1157-1161
  116. 116. Dodson MW, Guo M. Pink1, Parkin, DJ-1 and mitochondrial dysfunction in Parkinson’s disease. Current Opinion in Neurobiology. 2007;17:331-337
  117. 117. Wood-Kaczmar A, Gandhi S, Wood NW. Understanding the molecular causes of Parkinson’s disease. Trends in Molecular Medicine. 2006;12:521-528
  118. 118. Bonifati V, Rohé CF, Breedveld GJ, et al. Early-onset parkinsonism associated with PINK1 mutations: Frequency, genotypes, and phenotypes. Neurology. 2005;65(1):87-95
  119. 119. Yang Y, Gehrke S, Imai Y, et al. 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 USA. 2006;103:10793-10798
  120. 120. Vincow ES, Merrihew G, Thomas RE, et al. The PINK1-parkin pathway promotes both mitophagy and selective respiratory chain turnover in vivo. Proceedings of the National Academy of Sciences of the United States of America. 2013;110(16):6400-6405
  121. 121. Holt IJ, Harding AE, Petty RK, Morgan-Hughes JA. A new mitochondrial disease associated with mitochondrial DNA heteroplasmy. American Journal of Human Genetics. 1990;46:428-433
  122. 122. Zeviani M, Servidei S, Gellera C, et al. An autosomal dominant disorder with multiple deletions of mitochondrial DNA starting at the D-loop region. Nature. 1989;339:309-311
  123. 123. Moraes CT, Shanske S, Tritschler HJ, et al. mtDNA depletion with variable tissue expression: A novel genetic abnormality in mitochondrial diseases. American Journal of Human Genetics. 1991;48:492-501
  124. 124. Zeviani M. Nucleus-driven mutations of human mitochondrial DNA. Journal of Inherited Metabolic Disease. 1992;15:456-471
  125. 125. Lightowlers RN, Taylor RW, Turnbull DM. Mutations causing mitochondrial disease: What is new and what challenges remain? Science. 2015;349:1494-1499
  126. 126. Payne BA, Chinnery PF. Mitochondrial dysfunction in aging: Much progress but many unresolved questions. Biochimica et Biophysica Acta. 1847;2015:1347-1353
  127. 127. Lane RK, Hilsabeck T, Rea SL. The role of mitochondrial dysfunction in age-related diseases. Biochimica et Biophysica Acta. 1847;2015:1387-1400
  128. 128. Grunewald A, Rygiel KA, Hepplewhite PD, et al. Mitochondrial DNA depletion in respiratory chain-deficient Parkinson disease neurons. Annals of Neurology. 2016;79:366-378
  129. 129. Bender A, Krishnan KJ, Morris CM, et al. High levels of mitochondrial DNA deletions in substantia nigra neurons in aging and Parkinson disease. Nature Genetics. 2006;38:515-517
  130. 130. Kandul NP, Zhang T, Hay BA, Guo M. Selective removal of deletion-bearing mitochondrial DNA in heteroplasmic Drosophila. Nature Communications. 2016;7:13100
  131. 131. Detmer SA, Chan DC. Functions and dysfunctions of mitochondrial dynamics. Nature Reviews. Molecular Cell Biology. 2007;8:870-879
  132. 132. Chen H, Chomyn A, Chan DC. Disruption of fusion results in mitochondrial heterogeneity and dysfunction. The Journal of Biological Chemistry. 2005;280:26185-26192
  133. 133. Hales KG, Fuller MT. Developmentally regulated mitochondrial fusion mediated by a conserved, novel, predicted GTPase. Cell. 1997;90:121-129
  134. 134. Skulachev VP. Mitochondrial filaments and clusters as intracellular power-transmitting cables. Trends in Biochemical Sciences. 2001;26:23-29
  135. 135. Hoppins S, Lackner L, Nunnari J. The machines that divide and fuse mitochondria. Annual Review of Biochemistry. 2007;76:751-780
  136. 136. Hwa JJ, Hiller MA, Fuller MT, Santel A. Differential expression of the Drosophila mitofusin genes fuzzy onions (fzo) and dmfn. Mechanisms of Development. 2002;116:213-216
  137. 137. Yarosh W, Monserrate J, Tong JJ, et al. The molecular mechanisms of OPA1-mediated optic atrophy in Drosophila model and prospects for antioxidant treatment. PLoS Genet. 2008;4:e6
  138. 138. Verstreken P, Ly CV, Venken KJ, et al. Synaptic mitochondria are critical for mobilization of reserve pool vesicles at Drosophila neuromuscular junctions. Neuron. 2005;47(3):365-378
  139. 139. Maraganore DM, de Andrade M, Lesnick TG, et al. High-resolution whole-genome association study of Parkinson disease. American Journal of Human Genetics. 2005;77:685-693
  140. 140. Hardy J, Singleton A. Genomewide association studies and humandisease. The New England Journal of Medicine. 2009;360:1759-1768
  141. 141. Pankratz N, Foroud T. Genetics of Parkinson disease. Genetics in Medicine. 2007;9:801-811
  142. 142. Satake W, Nakabayashi Y, Mizuta I, et al. Genome-wide association study identifies common variants at four loci as genetic risk factors for Parkinson’s disease. Nature Genetics. 2009;41:1303-1307
  143. 143. Pankratz N, Wilk JB, Latourelle JC, et al. Genome wide association study for susceptibility genes contributing to familial Parkinson disease. Human Genetics. 2009;124:593-605
  144. 144. Lill CM, Roehr JT, McQueen MB, et al. Comprehensive research synopsis and systematic meta-analyses in Parkinson’s disease genetics: The PDgene database. PLoS Genetics. 2012;8:4-13
  145. 145. Nalls MA, Pankratz N, Lill CM, et al. Large-scale meta-analysis of genome-wide association data identifies six new risk loci for Parkinson’s disease. Nature Genetics. 2014;056:1-7
  146. 146. Escott-Price V, Nalls MA, Morris HR, et al. Polygenic risk of Parkinson disease is correlated with disease age at onset. Annals of Neurology. 2015;77:582-591
  147. 147. Hamza TH, Zabetian CP, Tenesa A, et al. Common genetic variation in the HLA region is associated with late-onset sporadic Parkinson’s disease. Nature Genetics. 2010;42:781-785
  148. 148. Do CB et al. Web-based genome-wide association study identifies two novel loci and a substantial genetic component for Parkinson’s disease. PLoS Genetics. 2011;7(6):e1002141
  149. 149. Ivatt RM, Sanchez-Martinez A, Godena VK, et al. Genome-wide RNAi screen identifies the Parkinson disease GWAS risk locus SREBF1 as a regulator of mitophagy. Proceedings of the National Academy of Sciences of USA. 2014;111:8494-8499
  150. 150. Bodmer W, Bonilla C. Common and rare variants in multifactorial susceptibility to common diseases. Nature Genetics. 2008;40:695-701
  151. 151. Keller MF, Saad M, Bras J et al. Using genome-wide complex trait analysis to quantify “missing heritability” in Parkinson’s disease. Human Molecular Genetics 2012;21:4996-5009
  152. 152. Brand AH, Perrimon N. Raf acts downstream of the EGF receptor to determine dorsoventral polarity during Drosophila oogenesis. Genes & Development. 1994;8:629-639
  153. 153. Singleton AB, Farrer M, Johnson J, et al. Alpha-Synuclein locus triplication causes Parkinson's disease. Science 2003;302:841
  154. 154. Fire A, Xu S, Montgomery MK, et al. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature. 1998;391:806-811
  155. 155. Meister G, Tuschl T. Mechanisms of gene silencing by double-stranded RNA. Nature. 2004;431:343-349
  156. 156. Silva J, Chang K, Hannon GJ, Rivas FV. RNA-interference-based functional genomics in mammalian cells: Reverse genetics coming of age. Oncogene. 2004;23:8401-8409
  157. 157. Ghildiyal M, Zamore PD. Small silencing RNAs: An expanding universe. Nature Reviews. Genetics. 2009;10:94-108
  158. 158. Cronin SJ, Nehme NT, Limmer S, et al. Genome-wide RNAi screen identifies genes involved in intestinal pathogenic bacterial infection. Science. 2009;325:340-343
  159. 159. Yamamoto-Hino M, Kanie Y, Awano W, et al. Identification of genes required for neural-specific glycosylation using functional genomics. PLoS Genetics. 2010;6:e1001254
  160. 160. Valakh V, Naylor SA, Berns DS, DiAntonio A. A large-scale RNAi screen identifies functional classes of genes shaping synaptic development and maintenance. Developmental Biology. 2012;366:163-171
  161. 161. Samuel AH, Lesley AK, Koji Y, et al. High-content genome-wide RNAi screens identifyregulators of parkin upstream of mitophagy. Nature. 2013;504:291-295
  162. 162. Cong L, Ran FA, Cox D, et al. Multiplex genome engineering using CRISPR/Cas systems. Science. 2013;339:819-823
  163. 163. Chen BH, Gilbert LA, Cimini BA, et al. Dynamic imaging of genomic loci in living human cells by an optimized CRISPR/cas system. Cell. 2014;155:1479-1491
  164. 164. Wang X, Cao C, Huang J, et al. One-generation of triple gene-targeted pigs using CRISPR/Cas9system. Scientific Reports. 2016;6:20620
  165. 165. Heidenreich M, Zhang F. Applications of CRISPR-Cas systems in neuroscience. Nature Reviews. Neuroscience. 2016;17:36-44
  166. 166. Walters BJ, Azam AB, Gillon CJ, et al. Advanced in vivo use of CRISPR/Cas9 and anti-sense DNA inhibition for gene manipulation in the brain. Frontiers in Genetics. 2016;6:362
  167. 167. Yang W, Tu Z, Sun Q, Li XJ. CRISPR/Cas9: Implications for Modeling and therapy of neurodegenerative diseases. Frontiers in Molecular Neuroscience. 2016;9:30
  168. 168. Pendleton RG, Parvez F, Sayed M, et al: Effects of pharmacological agents upon a transgenic model of Parkinson’s disease in Drosophila melanogaster. Journal of Pharmacology and Experimental Therapeutics. 2002;300(1):91-96
  169. 169. Auluck PK, Chan HYE, Trojanowski JQ, et al. Chaperone suppression of α- synuclein toxicity in a Drosophila model for Parkinson’s disease. Science. 2002;295(5556):865-868
  170. 170. Auluck PK, Meulener MC, Bonini NM. Mechanisms of suppression of α-synuclein neurotoxicity by geldanamycin in Drosophila. Journal of Biological Chemistry. 2005;280(4):2873-2878
  171. 171. Auluck PK, Bonini NM. Pharmacological prevention of Parkinson disease in Drosophila. Nature Medicine. 2002;8(11):1185-1186
  172. 172. Outeiro TF, Kontopoulos E, Altmann SM, et al. Sirtuin 2 inhibitors rescue α-synuclein-mediated toxicity in models of Parkinson’s disease. Science. 2007;317(5837):516-519
  173. 173. Trinh K, Moore K, Wes PD, et al. Induction of the phase II detoxification pathway suppresses neuron loss in Drosophila models of Parkinson’s disease. Journal of Neuroscience. 2008;28(2):465-472
  174. 174. Wassef R, Haenold R, Hansel A, et al. Methionine sulfoxide reductase A and a dietary supplement S-methyl-L-cysteine prevent Parkinson’s-like symptoms. Journal of Neuroscience. 2007;27(47):12808-12816
  175. 175. Long J, Gao H, Sun L, et al. Grape extract protects mitochondria from oxidative damage and improves Locomotor dysfunction and extends lifespan in a Drosophila Parkinson’s disease model. Rejuvenation Research. 2009;12(5):321-331
  176. 176. Faust K, Gehrke S, Yang Y, et al. Neuroprotective effects of compounds with antioxidant and anti-inflammatory properties in a Drosophila model of Parkinson’s disease. BMC Neuroscience. 2009;10:109
  177. 177. Lavara-Culebras E, Muñoz-Soriano V, Gómez-Pastor R, et al. Effects of pharmacological agents on the lifespan phenotype of Drosophila DJ-1β mutants. Gene. 2010;462(1-2):26-33
  178. 178. Brunn GJ, Hudson CC, Sekuli’c A, et al. Phosphorylation of the translational repressor PHAS-I by the mammalian target of rapamycin. Science. 1997;277(5322):99-101
  179. 179. Burnett PE, Barrow RK, Cohen NA, et al. RAFT1 phosphorylation of the translational regulators p70 S6 kinase and 4E-BP1. Proceedings of the National Academy of Sciences of the United States of America. 1998;95(4):14321437
  180. 180. Tain LS, Mortiboys H, Tao RN, et al. Rapamycin activation of 4E-BP prevents parkinsonian dopaminergic neuron loss. Nature Neuroscience. 2009;12(9):1129-1135
  181. 181. Saini N, Schaffner W. Zinc supplement greatly improves the condition of parkin mutant Drosophila. Biological Chemistry. 2010;391(5):513-518
  182. 182. Jimenez-Del-Rio M, Guzman-Martinez C, Velez-Pardo C. The effects of polyphenols on survival and locomotor activity in Drosophila melanogaster exposed to iron and paraquat. Neurochemical Research. 2010;35(2):227-238
  183. 183. Liu Z, Hamamichi S, Lee BD et al: Inhibitors of LRRK2 kinase attenuate neurodegeneration and Parkinson-like phenotypes in Caenorhabditis elegans and Drosophila Parkinson's disease models. Human Molecular Genetics. 2011;20:3933-3942
  184. 184. Neef DW, Jaeger AM, Thiele DJ. Heat shock transcription factor 1 as a therapeutic target in neurodegenerative diseases. Nature Reviews Drug Discovery. 2011;10(12):930-944
  185. 185. Thakur P, Nehru B. Long-termheat shock proteins (HSPs) induction by carbenoxolone improves hallmark features of Parkinson’s disease in a rotenone-based model. Neuropharmacology. 2014;79:190-200
  186. 186. Kilpatrick K, Novoa JA, Hancock T, et al. Chemical induction of Hsp70 reduces 𝛼-synuclein aggregation in neuroglioma cells. ACS Chemical Biology. 2013;8(7):1460-1468
  187. 187. Chiosis G, Tao H. Purine-scaffold Hsp90 inhibitors. IDrugs: The Investigational Drugs Journal. 2006;9(11):778-782
  188. 188. McLean PJ, Klucken J, Shin Y, et al. Geldanamycin induces Hsp70 and prevents 𝛼-synuclein aggregation and toxicity in vitro. Biochemical and Biophysical Research Communications. 2004;321(3):665-669
  189. 189. Shen HY, He JC, Wang Y, et al. Geldanamycin induces heat shock protein 70 and protects against MPTP-induced dopaminergic neurotoxicity in mice. Journal of Biological Chemistry. 2005;280(48):39962-39969
  190. 190. Putcha P, Danzer KM, Kranich LR et al. Brain-permeable small-molecule inhibitors of Hsp90 prevent 𝛼-synuclein oligomer formation and rescue 𝛼-synuclein-induced toxicity. Journal of Pharmacology and Experimental Therapeutics. 2010;332(3):849-857
  191. 191. Wang D, Qian L, Xiong H, et al. Antioxidants protect PINK1-dependent dopaminergic neurons in Drosophila. Proceedings of the National Academy of Sciences of the United States of America. 2006;103(36):13520-13525
  192. 192. Cooper AA, Gitler AD, Cashikar A, et al. α-synuclein blocks ER-Golgi traffic and Rab1 rescues neuron loss in Parkinson’s models. Science. 2006;313(5785):324-328
  193. 193. Ng CH, Guan MS, Koh C, et al. AMP kinase activation mitigates dopaminergic dysfunction and mitochondrial abnormalities in Drosophila models of Parkinson’s disease. The Journal of Neuroscience. 2012;32(143):11-17
  194. 194. Buttner S, Broeskamp F, Sommer C, et al. Spermidine protects against alpha-synuclein neurotoxicity. Cell Cycle. 2014;13(3903):8. DOI: 10.4161/15384101.2014.973309
  195. 195. Kontopoulos E, Parvin JD, Feany MB. Alpha-synuclein acts in the nucleus to inhibit histone acetylation and promote neurotoxicity. Human Molecular Genetics. 2006;15(20):3012-3023
  196. 196. Pinna A. Adenosine A2A receptor antagonists in Parkinson's disease: Progress in clinical trials from the newly approved istradefylline to drugs in early development and those already discontinued. CNS Drugs. 2014;28(5):455-475
  197. 197. Stocchi F, Rascol O, Destee A, et al. AFQ056 in Parkinson patients with levodopa-induced dyskinesia: 13-week, randomized, dose-finding study. Movement Disorders. 2013;28(13):1838-1846
  198. 198. Borgohain R, Szasz J, Stanzione P, et al. For the study 016 investigators. Randomized trial of safinamide add-on to levodopa in Parkinson's disease with motor fluctuations. Movement Disorders 2014;29(2):229-237
  199. 199. Sureda FX, Junyent F, Verdaguer E, et al. Antiapoptotic drugs: A therapautic strategy for the prevention of neurodegenerative diseases. Current Pharmaceutical Design. 2011;17(3):230-245
  200. 200. Holmay MJ, Terpstra M, Coles LD, et al. N-Acetylcysteine boosts brain and blood glutathione in Gaucher and Parkinson diseases. Clinical Neuropharmacology. 2013;36(4):103-106
  201. 201. Palfi S, Gurruchaga JM, Ralph GS, et al. Long-term safety and tolerability of ProSavin, a lentiviral vector-based gene therapy for Parkinson's disease: A dose escalation, open-label, phase 1/2 trial. Lancet. 2014;383(9923):1138-1146
  202. 202. Nishimura K, Takahashi J. Therapeutic application of stem cell technology toward the treatment of Parkinson's disease. Biological & Pharmaceutical Bulletin. 2013;36(2):171-175
  203. 203. Harraz MM, Dawson TM, Dawson VL. MicroRNAs in Parkinson's disease. Journal of Chemical Neuroanatomy. 2011;42(2):127-130

Written By

Mohamad Ayajuddin, Abhik Das, Limamanen Phom, Priyanka Modi, Rahul Chaurasia, Zevelou Koza, Abuno Thepa, Nukshimenla Jamir, Pukhrambam Rajesh Singh, Sentinungla Longkumer, Pardeshi Lal and Sarat Chandra Yenisetti

Submitted: 03 October 2017 Reviewed: 27 October 2017 Published: 28 February 2018