This AOP is licensed under the BY-SA license. This license allows reusers to distribute, remix, adapt, and build upon the material in any medium or format, so long as attribution is given to the creator. The license allows for commercial use. If you remix, adapt, or build upon the material, you must license the modified material under identical terms.

AOP: 3

Title

A descriptive phrase which references both the Molecular Initiating Event and Adverse Outcome.It should take the form “MIE leading to AO”. For example, “Aromatase inhibition leading to reproductive dysfunction” where Aromatase inhibition is the MIE and reproductive dysfunction the AO. In cases where the MIE is unknown or undefined, the earliest known KE in the chain (i.e., furthest upstream) should be used in lieu of the MIE and it should be made clear that the stated event is a KE and not the MIE.  More help

Inhibition of the mitochondrial complex I of nigro-striatal neurons leads to parkinsonian motor deficits

Short name
A name that succinctly summarises the information from the title. This name should not exceed 90 characters. More help
Mitochondrial dysfunction and Neurotoxicity

Graphical Representation

A graphical representation of the AOP.This graphic should list all KEs in sequence, including the MIE (if known) and AO, and the pair-wise relationships (links or KERs) between those KEs. More help
Click to download graphical representation template Explore AOP in a Third Party Tool

Authors

The names and affiliations of the individual(s)/organisation(s) that created/developed the AOP. More help

Anna Bal-Price, European Commission, Joint Research Centre (JRC), Ispra, Italy

Marcel Leist, University of Konstanz, Konstanz, Germany

Stefan Schildknecht, University of Konstanz, Konstanz, Germany

Florianne Tschudi-Monnet, University of Lausanne and SCAHT, Lausanne, Switzerland

Alicia Paini, European Commission, Joint Research Centre (JRC), Ispra, Italy

Andrea Terron, European Food Safety Authority, Parma Italy (corresponding author: Andrea.TERRON@efsa.europa.eu)

Point of Contact

The user responsible for managing the AOP entry in the AOP-KB and controlling write access to the page by defining the contributors as described in the next section.   More help
Andrea Terron   (email point of contact)

Contributors

Users with write access to the AOP page.  Entries in this field are controlled by the Point of Contact. More help
  • Andrea Terron
  • Clemens Wittwehr
  • Anna Price

Coaches

This field is used to identify coaches who supported the development of the AOP.Each coach selected must be a registered author. More help

Status

Provides users with information concerning how actively the AOP page is being developed, what type of use or input the authors feel comfortable with given the current level of development, and whether it is part of the OECD AOP Development Workplan and has been reviewed and/or endorsed. OECD Status - Tracks the level of review/endorsement the AOP has been subjected to. OECD Project Number - Project number is designated and updated by the OECD. SAAOP Status - Status managed and updated by SAAOP curators. More help
Handbook Version OECD status OECD project
v1.0 WPHA/WNT Endorsed 1.33
This AOP was last modified on April 29, 2023 16:02

Revision dates for related pages

Page Revision Date/Time
Inhibition, NADH-ubiquinone oxidoreductase (complex I) March 12, 2018 11:03
Binding of inhibitor, NADH-ubiquinone oxidoreductase (complex I) March 28, 2018 04:51
N/A, Mitochondrial dysfunction 1 March 14, 2024 11:12
Impaired, Proteostasis March 15, 2018 12:46
Neuroinflammation July 15, 2022 09:54
Degeneration of dopaminergic neurons of the nigrostriatal pathway March 15, 2018 12:50
Parkinsonian motor deficits March 12, 2018 12:44
N/A, Mitochondrial dysfunction 1 leads to Degeneration of dopaminergic neurons of the nigrostriatal pathway August 25, 2017 09:28
Binding of inhibitor, NADH-ubiquinone oxidoreductase (complex I) leads to Inhibition, NADH-ubiquinone oxidoreductase (complex I) August 25, 2017 09:35
Inhibition, NADH-ubiquinone oxidoreductase (complex I) leads to N/A, Mitochondrial dysfunction 1 August 25, 2017 09:43
N/A, Mitochondrial dysfunction 1 leads to Impaired, Proteostasis August 25, 2017 08:35
Impaired, Proteostasis leads to Degeneration of dopaminergic neurons of the nigrostriatal pathway August 25, 2017 08:37
Degeneration of dopaminergic neurons of the nigrostriatal pathway leads to Parkinsonian motor deficits August 25, 2017 09:32
Degeneration of dopaminergic neurons of the nigrostriatal pathway leads to Neuroinflammation October 02, 2017 10:30
Neuroinflammation leads to Degeneration of dopaminergic neurons of the nigrostriatal pathway August 25, 2017 08:54
MPP+ December 16, 2016 11:22
Rotenone November 29, 2016 18:42

Abstract

A concise and informative summation of the AOP under development that can stand-alone from the AOP page. The aim is to capture the highlights of the AOP and its potential scientific and regulatory relevance. More help

This Adverse outcome Pathway (AOP) describes the linkage between inhibition of complex I (CI) of the mitochondrial respiratory chain and motor deficit as in parkinsonian disorders. Binding of an inhibitor to complex I has been defined as the molecular initiating event (MIE) that triggers mitochondrial dysfunction, impaired proteostasis, which then cause degeneration of dopaminergic (DA) neurons of the nigro-striatal pathway. Neuroinflammation is triggered early in the neurodegenerative process and exacerbates it significantly. These causatively linked cellular key events result in motor deficit symptoms, typical for parkinsonian disorders, including Parkinson's disease (PD), described in this AOP as an Adverse Outcome (AO). Since the release of dopamine in the striatum by DA neurons of the Substantia Nigra pars compacta (SNpc) is essential for motor control, the key events refer to these two brain structures. The weight-of-evidence supporting the relationship between the described key events is based mainly on effects observed after an exposure to the chemicals rotenone and 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), i.e. two well-known inhibitors of complex I. Data from experiments with these two chemicals reveal a significant concordance in the dose-response relationships between the MIE and AO and within key events (KEs). Also essentiality of the described KEs for this AOP is strong since there is evidence from knock out animal models, engineered cells or replacement therapies that blocking, preventing or attenuating an upstream KE is mitigating the AO. Similarly, there is proved experimental support for the key event relationships (KERs) as multiple studies performed with modulating factors that attenuate (particularly with antioxidants) or augment (e.g. overexpression of viral-mutated α-synuclein) a KE up show that such interference leads to an increase of KE down or the AO. Information from in vitro and in vivo experiments is complemented by human studies in brain tissues from individuals with sporadic Parkinson's disease (Keeney et al., 2006) to support the pathways of toxicity proposed in this AOP.

AOP Development Strategy

Context

Used to provide background information for AOP reviewers and users that is considered helpful in understanding the biology underlying the AOP and the motivation for its development.The background should NOT provide an overview of the AOP, its KEs or KERs, which are captured in more detail below. More help

Strategy

Provides a description of the approaches to the identification, screening and quality assessment of the data relevant to identification of the key events and key event relationships included in the AOP or AOP network.This information is important as a basis to support the objective/envisaged application of the AOP by the regulatory community and to facilitate the reuse of its components.  Suggested content includes a rationale for and description of the scope and focus of the data search and identification strategy/ies including the nature of preliminary scoping and/or expert input, the overall literature screening strategy and more focused literature surveys to identify additional information (including e.g., key search terms, databases and time period searched, any tools used). More help

Summary of the AOP

This section is for information that describes the overall AOP.The information described in section 1 is entered on the upper portion of an AOP page within the AOP-Wiki. This is where some background information may be provided, the structure of the AOP is described, and the KEs and KERs are listed. More help

Events:

Molecular Initiating Events (MIE)
An MIE is a specialised KE that represents the beginning (point of interaction between a prototypical stressor and the biological system) of an AOP. More help
Key Events (KE)
A measurable event within a specific biological level of organisation. More help
Adverse Outcomes (AO)
An AO is a specialized KE that represents the end (an adverse outcome of regulatory significance) of an AOP. More help
Type Event ID Title Short name
MIE 888 Binding of inhibitor, NADH-ubiquinone oxidoreductase (complex I) Binding of inhibitor, NADH-ubiquinone oxidoreductase (complex I)
KE 887 Inhibition, NADH-ubiquinone oxidoreductase (complex I) Inhibition, NADH-ubiquinone oxidoreductase (complex I)
KE 177 N/A, Mitochondrial dysfunction 1 N/A, Mitochondrial dysfunction 1
KE 889 Impaired, Proteostasis Impaired, Proteostasis
KE 188 Neuroinflammation Neuroinflammation
KE 890 Degeneration of dopaminergic neurons of the nigrostriatal pathway Degeneration of dopaminergic neurons of the nigrostriatal pathway
AO 896 Parkinsonian motor deficits Parkinsonian motor deficits

Relationships Between Two Key Events (Including MIEs and AOs)

This table summarizes all of the KERs of the AOP and is populated in the AOP-Wiki as KERs are added to the AOP.Each table entry acts as a link to the individual KER description page. More help

Network View

This network graphic is automatically generated based on the information provided in the MIE(s), KEs, AO(s), KERs and Weight of Evidence (WoE) summary tables. The width of the edges representing the KERs is determined by its WoE confidence level, with thicker lines representing higher degrees of confidence. This network view also shows which KEs are shared with other AOPs. More help

Prototypical Stressors

A structured data field that can be used to identify one or more “prototypical” stressors that act through this AOP. Prototypical stressors are stressors for which responses at multiple key events have been well documented. More help
Name
MPP+
Rotenone

Life Stage Applicability

The life stage for which the AOP is known to be applicable. More help
Life stage Evidence
Adult High

Taxonomic Applicability

Latin or common names of a species or broader taxonomic grouping (e.g., class, order, family) can be selected.In many cases, individual species identified in these structured fields will be those for which the strongest evidence used in constructing the AOP was available. More help
Term Scientific Term Evidence Link
human Homo sapiens High NCBI
rat Rattus norvegicus High NCBI

Sex Applicability

The sex for which the AOP is known to be applicable. More help
Sex Evidence
Mixed High

Overall Assessment of the AOP

Addressess the relevant biological domain of applicability (i.e., in terms of taxa, sex, life stage, etc.) and Weight of Evidence (WoE) for the overall AOP as a basis to consider appropriate regulatory application (e.g., priority setting, testing strategies or risk assessment). More help

Domain of Applicability

Addressess the relevant biological domain(s) of applicability in terms of sex, life-stage, taxa, and other aspects of biological context. More help

This proposed AOP is neither sex-dependent nor associated with certain life stage; however, aged animals may be more sensitive. The relevance of this AOP during the developmental period has not been investigated. In vivo testing has no species restriction. The mouse was the species most commonly used in the experimental models conducted with the chemical stressors; though experimental studies using alternative species have been also performed. (Johnson et al. 2015). However, animal models (rodents in particular) would have limitations as they are poorly representative of the long human life-time as well as of the human long-time exposure to the potential toxicants. Human cell-based models would likely have better predictivity for humans than animal cell models. In this case, toxicokinetics information from in-vivo studies would be essential to test the respective concentrations in-vitro on human cells.

Essentiality of the Key Events

The essentiality of KEs can only be assessed relative to the impact of manipulation of a given KE (e.g., experimentally blocking or exacerbating the event) on the downstream sequence of KEs defined for the AOP. Consequently, evidence supporting essentiality is assembled on the AOP page, rather than on the independent KE pages that are meant to stand-alone as modular units without reference to other KEs in the sequence. The nature of experimental evidence that is relevant to assessing essentiality relates to the impact on downstream KEs and the AO if upstream KEs are prevented or modified. This includes: Direct evidence: directly measured experimental support that blocking or preventing a KE prevents or impacts downstream KEs in the pathway in the expected fashion. Indirect evidence: evidence that modulation or attenuation in the magnitude of impact on a specific KE (increased effect or decreased effect) is associated with corresponding changes (increases or decreases) in the magnitude or frequency of one or more downstream KEs. More help

Essentiality of KEs for this AOP is strong. There is ample evidence from knock out animal models, engineered cells or replacement therapies that blocking, preventing or attenuating an upstream KE is mitigating the AO. In addition, there is experimental support for the KERs as multiple studies performed with modulating factors that attenuate (particularly with antioxidants) or augment (e.g. overexpression of viral-mutated α-synuclein) a KE show that such interference leads to an increase of KE down or the AO.

2 Support for Essentiality of KEs

Defining Question

Are downstream KEs and/or the AO prevented if an upstream KE is blocked ?

High (Strong)

Moderate

Low(Weak)

Direct evidence from specifically designed experimental studies illustrating essentiality for at least one of the important KEs (e.g. stop/reversibility studies, antagonism, knock out models, etc.)

Indirect evidence that sufficient modification of an expected modulating factor attenuates or augments a KE leading to increase in KE down or AO

No or contradictory experimental evidence of the essentiality of any of the KEs

KE1

Inhibition of complex I

STRONG

Rationale: Inactivation of the NADH:Ubiquinone Oxidoreductase Core Subunit S7 (Ndufs 4 gene knockout mice) that produces CI deficiency causes encephalomyopathy, including ataxia and loss of motor skills (Kruse et al., 2008). NDI1-transducted SK-N-MC cells expressing the rotenone-insensitive single subunit NADH dehydrogenase of yeast (NDI1) that acts as a replacement for the entire CI in mammalian cells were completely resistant to 100 nM rotenone-mediated cell death (at 48 hrs of exposure) indicating that rotenone – induced toxicity requires rotenone biding of CI (Sherer et al., 2003). In all rotenone models, mitochondria CI is inhibited at the dose that cause neurodegeneration (Betarbet et al 2000 and 2006).

KE2

Mitochondrial dysfunction

STRONG

Rationale: Many studies showing that antioxidants protect the cells against rotenone or MPTP induced oxidative stress are published (Chen et al. 2015; Lu et al., 2015; Saravanan et al., 2006; Chiu et al., 2015, Sherer et al.2003, Nataraj et al.2015, Wu et al. 1994; Tseng et al. 2014; Li et al. 2010; Kim-Han et al. 2011). This provides (indirect) evidence for essentiality of KE2, if production of reactive oxygen species (ROS) is assumed as direct consequence/sign of mitochondrial dysfunction. Additional evidence comes from experiments with overexpression or activation of antioxidative enzymes (e.g.SOD or ALDH2) , which also prevent rotenone and MPTP/MPP+ induced neurotoxicity (Mudo et al. 2012; Ciu CC et al. 2015). Furthermore, promotion of mitochondrial fusion or blocking of mitochondrial fission prevents or attenuates rotenone and MPTP/MPP+ induced neurotoxicity (Tieu K. et al. 2014).

KE3

Impaired proteostasis

MODERATE

Rationale: Indirect evidence for the role of disturbed alpha-synuclein proteostasis: Lacking of alpha-synuclein expression in mice prevented induction of behavioural symptoms, neuronal degeneration in the nigrostriatal pathway and loss of DA neurons after chronic treatment with MPTP/MPP+ (Fornai et al. 2004; Dauer et al. 2002) . Injection of adeno/lenti-associated virus that expresses wild-type or mutant α-synuclyn into rat, mice or non-human primate SN produced loss of dopaminergic neurons, but the effect is not easily reproduced in transgenic mice overexpressing alpha-synuclein (Kirk, 2002; Klein, 2002; Lo Bianco, 2002; Lauwers, 2003; Kirk, 2003).

Rationale for the role of autophagy: Early dendritic and axonal dystrophy, reduction of striatal dopamine content, and the formation of somatic and dendritic ubiquitinated inclusions in DA neurons were prevented by ablation of Atg7 (an essential autophagy related gene (Friedman et al. 2012)).

Rationale for the role of Ubiquitin Proteosomal System/Autophagic Lysosomal Pathway (UPS/ALP): Protection from DA neuronal death was also observed in multiple experiments through the pharmacological modulation of the UPS, ALP system; however, there are also contradicting data in the literature. (Inden et al. 2007; Fornai et al. 2003; Dehay et al. 2010; Zhu et al. 2007, Fornai et al. 2005).

However, although many lines of evidence exist to support essentiality of impaired proteostasis, a single molecular chain of events cannot be established.

KE4 

Degeneration of DA neurons of nigrostriatal pathway

MODERATE

Receptors for advanced glycated end product (AGEs) can activate NF-kB (a transcription factor involved in the inflammatory response) and they are found on microglia cells and astrocytes. Ablation of receptor for advanced glycated end product (RAGE) proved to be protective against MPTP-induced decreases of TH+ neurons and mitigation of microglia and astrocytes reactivity was observed (Teismann et al. 2012). Inhibition of RAGE, which is upregulated in the striatum following rotenone exposure and in response to neuroinflammation, decreases rotenone-induced apoptosis by suppressing NF(Nuclear Factor)-kB activation, as well as the downstream inflammatory markers TNF-alpha, iNOS and myeloperoxidase (Abdelsalam and Safar, 2015). This showed intermingled links between neuronal injury/death and neuroinflammation. Rotenone-induced neurotoxicity was less pronounced in neuron-enriched cultures than in neuron-glia co-cultures (Gao et al., 2002), suggesting that neuron-glia interactions are critical for rotenone-induced neurodegeneration. In addition, in in vitro systems, a decrease in thyroxine hydrosilase (TH) mRNA expression has been observed to be a sufficient signal to trigger microglial reactivity (Sandström et al., 2017).

KE5

Neuroinflammation

MODERATE

Rationale: Following treatment with Rotenone or MPTP/ MPP+, protection of DA neurons and terminals was observed in vivo and in vitro by inhibiting different feature of neuroinflammation (microglia/astrocyte); however, inhibition was different in different models and considered as an indirect evidence of essentiality (Zhou et al., 2007; Gao et al., 2002 and 2003 and 2015; ; Emmrich et al., 2013; Salama et al., 2012; Chang et al., 2013; Wang et al., 2014; Liu et al., 2012, 2015; Borrajo et al., 2013; Brzozowski et al., 2015; Wang et al., 2006; Chung et al., 2011; Sriram et al., 2014; Feng et al., 2002; Sathe et al., 2012; Khan et al., 2014; Ros-Bernal et al., 2011; Ferger et al., 2004; Chao et al., 2009; Rojo et al., 2010; Qian et al., 2011; Dehmer et al., 2000; Bodea et al., 2014). Mice lacking the type-1 Interferons receptor showed an attenuated pro-inflammatory response and reduced loss of dopaminergic neurons induced by MPTP/MPP+. The neuro-protective potential was also confirmed by treatment with a blocking monoclonal antibody against type-1A IFN receptor (interferon receptor) that increased survival of dopaminergic neurons of TH+ (Main et al., 2016).

KE4 

Degeneration of DA neurons of nigrostriatal pathway

STRONG

Rationale: Clinical and experimental evidences show that the pharmacological replacement of the dopamine (DA) neurofunction by allografting fetal ventral mesencephalic tissues is successfully replacing degenerated DA neurons resulting in the total reversibility of motor deficit in animal model and partial effect is observed in human patient for PD (Widner et al., 1992; Henderson et al., 1991; Lopez-Lozano et al., 1991; Freed et al., 1990; Peschanski et al., 1994; Spencer et al., 1992).

Also, administration of L-DOPA or DA agonists results in an improvement of motor deficits (Calne et al 1970; Fornai et al. 2005). The success of these therapies in man as well as in experimental animal models clearly confirms the causal role of dopamine depletion for PD motor symptoms ( Connolly et al., 2014; Lang et al., 1998; Silva et al., 1997; Cotzias et al., 1969; Uitti et al., 1996; Ferrari-Tonielli et al., 2008; Kelly et al., 1987; Walter et al., 2004; Narabayashi et al., 1984; Matsumoto et al., 1976; De Bie et al., 1999; Uitti et al., 1997; Scott et al., 1998; Moldovan et al., 2015; Deuschl et al., 2006; Fasano et al., 2010; Castrito et al., 2011; Liu et al., 2014; Widner et al., 1992; Henderson et al., 1991; Lopez-Lozano et al., 1991; Freed et al., 1990; Peschanski et al., 1994; Spencer et al., 1992).

Furthermore, experimental evidence from animal models of PD and from in-vitro systems indicate that prevention of apoptosis through ablation of BCL-2 family genes prevents or attenuates neurodegeneration of DA neurons (Offen D et al., 1998; Dietz GPH et al. 2002).

 

Evidence Assessment

Addressess the biological plausibility, empirical support, and quantitative understanding from each KER in an AOP. More help

Concordance of dose-response relationship.

Data from experiments with the stressor compounds rotenone and MPTP (known inhibitors of the mitochondrial Complex I (CI)) reveal a good concordance of the dose-response relationships between the MIE and AO and within KEs. Although the different KEs have been measured using different methodologies, comparison of data from multiple in-vitro/in-vivo studies shows a general agreement in dose-relationship (see table 1 and 2). There is a good consistency when comparing data on KE4 and the AO after exposure to rotenone and MPTP. However, in vivo rodent studies proved that only exposure to low concentrations of rotenone (rat brain concentration between 20-30 nM of rotenone; Betrabet et al., 2000) or MPTP (mice striatum concentration of approximately 12-47 µM MPP+; Fornai et al., 2005; Thomas et al. 2012) after chronic exposure (approximately 5 weeks) reproduced the anatomical, neurochemical behavioural and neuropathological features similar to the ones observed in Parkinson’s disease (PD). Because of the variability of experimental protocols used, a clear no-effect threshold could not be established; nevertheless, these brain concentrations of rotenone (20-30 nM) and MPP+ (approximately 12-47µM) could serve as probabilistic thresholds for chronic exposure that could reproduce features of PD as both concentrations trigger approximately a 50% inhibition of Complex I (see table 3). Generally, a strong response-response relationship is observed within studies. Some exceptions for this rule are observed between KE3/KE5 and KE4, likely because of the all biological complexity associated with these KEs. In this AOP, neuroinflammation was considered to have a direct effect on degeneration of DA neurons. However, it was not clear at which conditions it would become a modulatory factor and for practical reasons was not included in table 1, 2 and 3 but considered in the weight of evidence analysis.

Table1 Dose-response and temporality table for rotenone.

Rotenone Concentration 

KE1aaa

Inhibition of C I

KE2aaa

Mitochondrial dysfunction

KE3aaa

Impaired proteostasis

KE4

Degeneration of DA neurons of nigrostriatal pathway

AO

Parkinsonian motor symptoms

5-10 nM in-vitro

[1]

+

4-72 hours [1]

+

4-72 hours [4]

+

24 hours [3]

-

-

20-30 nM ex-vivo, rat brain concentration

[4-5-2-6]

++

4-72 hours (4-5)

++

4-72 hours [4-5]

++

24 hours [3-2-6]

++a

5 weeks [2-6]

+++aa

5 weeks [2-6]

100 nM in-vitro

[4]

+++

4-72 hours [4]

+++

4-72 hours [4]

+++

24 hours [3]

Corresponding to a concentration above the maximum tolerated dose in-vivo [2-6]

Corresponding to a concentration above the maximum tolerated dose in vivo [2-6]

References: Choi et al. 2008 [1]; Betarbet et al. 2006 [2]; Chou et al. 2010 [3]; Barrientos and Moraes 1999 [4]; Okun et al. 1999 [5]; Betarbet et al. 2000 [6]

-no data available

+: low severity score, ++ intermediate severity score, +++ high severity score

a: 50% of treated animals showed loss of DA neurons in SNpc

aa: All animals affected in KE4 showed impaired motor symptoms

aaa: KE 1, 2 and 3 showed a dose-related severity in the effect and the score ++ was normalized vs. the KE4

Table 2. Dose-Response and Temporality table for MPTP/MPP+

MPTP Administered Dose

MPPBrain Concentration

KE1bb

Inhibition of C I

KE2bb

Mitochondrial dysfunction

KE3b

Impaired proteostasis

KE4

Degeneration of DA neurons of nigrostriatal pathway

AO

Parkinsonian motor symptoms

1 mg/kg sc infusion [1]

-

-

-

+

4 weeks[1]

+aaa

4 weeks [1]

No effect

5 mg/kg sc infusion [1]

-

-

-

++

4 weeks[1]

++aa

4 weeks [1]

+++

4 weeks [1]

20-30 mg/kg sporadic ip. injections (4 times every 2 hours)

[2, 1]

 

47µM [2]^

12µM [1]

+++

4 hrs [2]

+++

4hrs [2]

+++

4 weeks [1]

+++a

1-4 weeks[2,1]

 

+++

4 weeks [1]

References. Fornai et al. 2005 [1]; Thomas et al. 2012 [2]

-no data available

a: approx 50% loss of DA neurons in SNpc

aa: approx 30% loss of DA neurons SN pc

aaa: no loss of DA neurons in SN pc. Reduced level of striata DA

b: for KE3, a dose response effect was observed.

bb: for KE 1 and 2 the severity of the effect was normalized vs. the KE4

^ After single dose MPTP administration, brain concentration was approx. 5.15 µM

Temporal concordance among the MIE, KEs and AO.

There is a strong agreement that loss of DA neurons of the SNpc that project into the putamen is preceded by reduction in DA and degeneration of DA neuronal terminals in the striatum (Bernheimer et al. 1973). The clinical symptoms of a motor deficit are observed when 80% of striatal DA is depleted (Koller et al. 1992) and the sequence of pathological events leading to the adverse outcome has been well-documented (Fujita, et al.2014; O’Malley 2010, Dexter et al. 2013). Temporal concordance (see table 1 and 2) among the KEs can be observed in the experimental models of PD using the chemical stressors rotenone and MPTP (Betarbet 2000 and 2006; Sherer et al. 2003, Fornai et al. 2005). The acute administration of the chemical stressors can trigger a dose-related change from the MIE to impaired proteostasis; however, to trigger KE4 (i.e. degeneration of DA neurons in SNpc with presence of intracytoplasmatic Lewy-like bodies) and motor deficits (AO), proteostasis needs to be disturbed for a minimum period of time (Fornai et al. 2005).

Strength, consistency, and specificity of association of AO and MIE.

Strength and consistency of the association of the AO with the MIE is strong. There is a large body of evidence from in-vitro and in-vivo studies with chemical stressors, showing association between the MIE that triggers an inhibition of CI and the AO (Sherer et al. 2003; Betarbet et al. 2000 and 2006, Fornai et al. 2005). Human data also suggest a link between inhibition of CI and AO (Greenamyre et al. 2001; Schapira et al. 1989; Shults, 2004). Using the two different chemical stressors, rotenone and MPTP, data are consistent and the pattern of activation of the MIE leading of the AO is similar. For rotenone and MPTP, specificity is high; however, there are many inhibitors of the mitochondrial CI without evidence of triggering the AO. When considering  the limited amount of chemical stressors for which empirical data are available for supporting the full sequence of KEs, kinetic and metabolic considerations should be taken into account to demonstrate specificity for these compounds. The issue of specificity was also debated during the external review of this AOP and the following information was added:

The vast majority of empirical support available in the literature is based on complex I inhibitors, such as rotenone and MPTP/MPP+, as well as on studies involving genetic impairment of complex I activity. A relatively wide spectrum of structurally different complex I inhibitors have been described over the course of recent decades. Prominent examples are acetogenins (Nat Prod Rep 2005, 22, 269-303); alkaloids (J Neurochem 1996, 66, 1174-1181); antibiotics (BBA 1998, 1364, 222-235; Eur J Biochem 1994, 219, 691-698; JBC 1970, 245, 1992-1997; Bioorg Med Chem 2003, 11, 4569-4575); pesticides (Biochem Soc Trans 1994, 22, 230-233); quinones (JBC 1971, 246, 2346-2353); or vanilloids (ABB 1989, 270, 573-577). Additional information can be also retreived from Fato et al 2009, Espositi et al. 1993, Lagoa et al. 2011 and Park et al. 2003.

All of these structurally different complex I inhibitors were characterized with isolated mitochondria or with submitochondrial particles. Application of bovine heart mitochondria revealed IC50 values in the range of 20-70 nM for piericidin A, fenpyroximate, rotenone, and phenoxan (Eur. J. Biochem 1994, 219, 691-698). IC50 values in the range of 1-10 nM were detected by application of submitochondrial particles with rotenone, molvizarin, rollinstatin-1 and -2, and piericidin A (Biochem J. 1994, 301, 161-167).

Studies involving neuronal cell cultures or in vivo models are in fact rather rare. A systematic comparison of the IC50 values for complex I inhibition and EC50 values for the reduction of ATP levels; cell death was performed with rat fetal striatal neurons (Exp Neurol 2009, 220, 133-142). Due to the lipophilicity of most of the complex I inhibitors tested, the detected EC50 values were in most cases lower than the IC50 values detected for complex I inhibition. EC50 values detected were: annonacin (60 nM), fenazaquin (45 nM), piericidin A (1.6 nM), rollinstatin- 2 (1 nM), rotenone (8 nM), and squamocin (1 nM).

A systematic investigation involving mesencephalic cultures as well as rats was performed for the complex I inhibitor annonacin, a major acetogenin of soursop, a plant suspected to cause an atypical form of PD in Guadeloupe. Mesencephalic cultures treated for 24 h with annonacin revealed EC50 values of 20 nM (annonacin), 34 nM (rotenone), and 1900 nM (MPP+) (Neurosci 2003, 121(2), 287-296). Intravenous application by minipumps over the course of 28 days indicated a passage of annonacin across the blood-brain barrier, and an energy-dependent loss of ca. 30 % of DA neurons in the substantia nigra (Champi et al.2004)).

Weight of Evidence (WoE).

Biological plausibility, coherence, and consistency of the experimental evidence.

The biological plausibility of this AOP is overall considered strong. When using multiple stressors in different studies and assays, the coherence and consistency of the experimental data is well established. Furthermore, in-vivo and in-vitro studies are also in line with the human evidence from PD patients. In addition, although the mechanistic understanding of parkinsonian disorders (and PD in particular) are not fully clear, the KEs and KERs described in this AOP are considered critical for the development of the disease (Fujita et al. 2015, Shulman et al. 2011, Dexter et al. 2013, Dauer et al. 2003).

1 Support for Biological Plausibility of KERs

Defining Question

High (Strong)

Moderate

Low(Weak)

Is there a mechanistic (i.e. structural or functional) relationship between KEup and KE down consistent with established biological knowledge?

Extensive understanding of the KER based on extensive previous documentation and broad acceptance

The KER is plausible based on analogy to accepted biological relationships, but scientific understanding is not completely established

There is empirical support for a statistical association between KEs but the structural or functional relationship between them is not understood

MIE => KE1

Binding of inhibitor to NADH-ubiquinone oxidoreductase leads of complex I

STRONG

Rationale: As describe in this KER there is an extensive understanding of the functional relationship between binding of an inhibitor to NADH-ubiquinone oxidoreductase (CI) and its inhibition. Different complex I ligands, both naturally occurring, like rotenone (from Derris scandens), piericidin A (from Streptomyces mobaraensis), acetogenins (from various Annonaceae species) and their derivatives, and synthetically manufactured like pyridaben and various piperazin derivatives inhibit the catalytic activity of complex I (Degli Esposti, 1994: Ichimaru et al. 2008; Barrientos and Moraes, 1999; Betarbet et al., 2000).

KE1 => KE2

Inhibition of complex I leads to mitochondrial dysfunction

STRONG

Rationale: There is extensive understanding of the mechanisms explaining how the inhibition of complex I lead to mitochondrial dysfunction (i.e. failure to produce ATP, increase in production of ROS etc). It is well documented that CI inhibition is one of the main sites at which electron leakage to oxygen occurs resulting in oxidative stress (Efremov and Sazanow, 2011; lauren et al. 2010; Greenamyre et al. 2001). These pathological mechanisms are well studied as they are used as readouts for evaluation of mitochondrial dysfunction (Graier et al., 2007; Braun, 2012; Martin, 2011; Correia et al., 2012; Cozzolino et al., 2013

KE2 => KE3

Mitochondrial dysfunction results in impaired proteostasis

MODERATE

Rationale: The weight of evidence supporting the biological plausibility behind the relationship between mitochondrial dysfunction and impaired proteostasis, including the impaired function of UPS and ALP that results in decreased protein degradation and increase protein aggregation is well documented but not fully understood. It is well established that the two main mechanisms that normally remove abnormal proteins (UPS and ALP) rely on physiological mitochondrial function. The role of oxidative stress, due to mitochondrial dysfunction, burdens the proteostasis with oxidized proteins and impairs the chaperone and the degradation systems. This leads to a vicious circle of oxidative stress inducing further mitochondrial impairment (Powers et al., 2009; Zaltieri et al., 2015; McNaught and Jenner, 2001). Therefore, the interaction of mitochondrial dysfunction and UPS /ALP deregulation plays a pivotal role in the pathogenesis of PD (Dagda et al., 2013; Pan et al., 2008; Fornai et al., 2005; Sherer et al., 2002).

KE2 => KE4

Mitochondrial dysfunction leads to the degeneration of dopaminergic neurons of the nigrostriatal pathway

STRONG

Rationale: Mitochondrial are essential for ATP production, ROS management, calcium homeostasis and control of apoptosis. Mitochondrial homeostasis by mitophagy is also an essential process for cellular maintenance (Fujita et al. 2014). Because of their anatomical and physiological characteristics, SNpc DA neurons are considered more vulnerable than other neuronal populations (Sulzer et al. 2013; Surmeier et al.2010). Mechanistic evidence of mutated proteins relate the mitochondrial dysfunction in familial PD with reduced calcium capacity, increased ROS production, increase in mitochondrial membrane permeabilization and increase in cell vulnerability (Koopman et al. 2012; Gandhi et al. 2009). Human studies indicate mitochondrial dysfunction in human idiopathic PD cases in the substantia nigra (Keeney et al., 2006; Parker et al., 1989, 2008; Swerdlow et al., 1996). In addition, systemic application of mitochondrial neurotoxicants such as rotenone or MPTP leads to a preferential loss of nigrostriatal DA neurons (Langston et al., 1983).

KE3 => KE4

Impaired proteostasis leads to degeneration of DA neurons of the nigrostriatal pathway

MODERATE

Rationale: It is well known that impaired proteostasis refers to misfolded and aggregated proteins including alfa-synuclein, deregulated axonal transport of mitochondria and impaired trafficking of cellular organelles. Evidences are linked to PD and experimental PD models as well as from genetic studies (McNaught et al. 2001, 2003; Tieu et al. 2014; Arnold 2011; Rappold et al. 2014). Strong evidence for degeneration of the nigrostriatal pathway comes from the experimental manipulations that directly induce the same disturbances of proteostasis as observed in PD patients (e.g. viral mutated alpha-synuclein expression) or in chronic rotenone/MPTP models trigger degeneration of the nigrostriatal pathway (Kirk et al. 2003; Betarbet et al. 2000 and 2006; Fornai et al. 2005). However, a clear mechanistic proof for the understanding of the exact event triggering cell death is lacking. There is only moderate evidence showing that interventions that correct disturbances of proteostasis after exposure to rotenone would prevent neuronal degeneration and that the disturbances of proteostasis correlate quantitatively under many conditions with the extent of nigrostriatal neuronal degeneration.

KE4 => KE5

Degeneration of DA neurons of the nigrostriatal pathway leads to neuroinflammation

MODERATE

Rationale: The fact that neuronal injury/death can trigger neuroinflammation is supported by evidence in human and experimental models. The evidence that neuroinflammation triggered by neuronal damage can cause neuronal death (vicious circle), is mostly indirect (blockade of any feature of neuroinflammation) or by analogy (Hirsch and Hunot, 2009; Tansey and Goldberg, 2009; Griffin et al., 1998; McGeer and Mc Geer, 1998; Blasko et al., 2004; Cacquevel et al., 2004; Rubio-Perez and Morillas-Ruiz, 2012; Thundyil and Lim, 2014; Barbeito et al., 2010). Neuroinflammation is observed in idiopathic and in genetic human PD as well as in complex I inhibitor exposed humans, non-human primates, and rodent. Components of damaged neurons lead to glial cells activation via Toll-like receptors. Several chemokines and chemokine receptors (fraktalkine, CD200) control the neuron-microglia interactions. Neuroinflammation in response to damaged neurons is not confined to PD, but is common to several neurodegenerative diseases 

KE5 => KE4

Neuroinflammation leads to degeneration of DA neurons of the nigrostriatal pathway

MODERATE

Rationale: The fact that reactive glial cells (microglia and astrocytes) may kill neurons is well accepted. The mechanisms underlying this effect may include the release of cytotoxic signals (e.g. cytokines) or production of ROS and RNS (Chao et al., 1995 ; Brown and Bal-Price, 2003 ; Kraft and Harry, 2011 ; Taetzsch and Block, 2013). However, the responsible mediators differ from model to model. In humans or non-human primates, an inflammatory activation of glial cells is observed years after exposure to complex I inhibitors. Activated microglia and astrocytes form pro-inflammatory cytokines and free radical species, mostly responsible for neuronal damage. Glial reactivity promotes an impairment of blood brain barrier integrity, allowing an infiltration of peripheral leukocytes that exacerbate the neuroinflammatory process and contribute to neurodegeneration.The debris of degenerating neurons causes neuroinflammation, which in turn can trigger neurodegeneration, thus leading to a self-perpetuating vicious cycle.

KE4 => AO

Degeneration of DA neurons of the nigrostriatal pathway leads to parkinsonian motor symptoms

STRONG

Rationale: The mechanistic understanding of the regulatory role of striatal DA in the extrapyramidal motor control system is well established. The loss of DA in the striatum is characteristic of all aetiologies of PD and is not observed in other neurodegenerative diseases (Bernheimer et al. 1973; Reynolds et al. 1986). Characteristic motor symptoms such as bradykinesia, tremor, or rigidity are manifested when more than 80 % of striatal DA is depleted as a consequence of SNpc DA neuronal degeneration (Koller et al. 1992).

Empirical support.

Empirical support is strong. Many studies show evidence for the KERs by showing temporal concordance and dose concordance when using different stressors.

3 Empirical support for KERs

Defining Question

Does the empirical evidence support that a change in the KEup leads to an appropriate change in the KE down? Does KEup occur at lower doses and earlier time points than KE down and is the incidence of KEup higher than that for KE down?

Are inconsistencies in empirical support cross taxa, species and stressors that don’t align with expected pattern of hypothesized AOP?

High (Strong)

Moderate

Low(Weak)

Multiple studies showing dependent change in both exposure to a wide range of specific stressors (extensive evidence for temporal, dose-response and incidence concordance) and no or few critical data gaps or conflicting data.

Demonstrated dependent change in both events following exposure to a small number of specific stressors and some evidence inconsistent with expected pattern that can be explained by factors such as experimental design, technical considerations, differences among laboratories, etc.

Limited or no studies reporting dependent change in both events following exposure to a specific stressor (ie endpoints never measured in the same study or not at all); and/or significant inconsistencies in empirical support across taxa and species that don’t align with expected pattern for hypothesized AOP

MIE => KE1

Binding of inhibitor to NADH-ubiquinone oxidoreductase leads to partial or total inhibition of complex I

STRONG

Rationale: The inhibition of complex I is well documented in a variety of studies using isolated mitochondria or cells as well as in in vivo experiments and in human post mortem PD brains. In many experiments using different inhibitors ie rotenone and MPTP, the observed relationship between the two events was temporal, response and dose concordant (Betarbet et al., 2000 and 2006, Okun et al., 1999, Koopman et al., 2007, Choi et al., 2008, Grivennikova et al., 1997, Barrientos and Moraes 1999).

KE1 => KE2

Inhibition of complex I leads to mitochondrial dysfunction

STRONG

Rationale: There is a large amount of studies showing that the inhibition of CI inhibition results in mitochondrial dysfunctions in a response and dose dependent manner (Barriento and Moraes, 1999).

KE2 => KE3

Mitochondrial dysfunction results in impaired proteostasis

STRONG

Rationale: Based on the existing in vitro and in vivo data it is suggested that mitochondrial dysfunction impairs protein homeostasis (impairment of the UPS and ALP system) through oxidative and nitrosative stress resulting in accumulation of misfolded proteins (including α-synuclein), disruption of microtubule assembly and damaged intracellular transport of proteins and cell organelles. A number of studies performed with chemical stressors showed evidence of temporal, response and dose concordance (Chou et al. 2010; Betarbet et al. 2000 and 2006; Fornai et al. 2005).

KE2 => KE4

Mitochondrial dysfunction directly leads to degeneration of DA neurons of nigrostriatal pathway

STRONG

Rationale: Multiple in vitro studies indicate dose and response-response concordance. As most of the studies were conducted in vitro, the temporal concordance is difficult to establish; however, can be expected based on the well know temporal sequence of the two KEs. (Park et al., 2014; Choi et al., 2014; Marella et al., 2008; Du et al. 2010; Hajieva et al., 2009; Sherer et al., 2003; Sherer et al., 2007; Wen et al. 2011; Swedlow et al., 1996; Jana et al., 2011; Jha et al., 2000; Chinta et al., 2006)

KE3 => KE4a

Impaired proteostasis leads to degeneration of DA neurons of the nigrostriatal pathway

STRONG

Rationale: The empirical support linking impaired proteostasis with degeneration of DA neurons of the nigrostriatal pathway is strong and comes from in-vivo and in-vitro studies performed with different stressor (i.e. Rotenone, MPTP or proteasome inhibitors) and post-mortem human evidences in PD patients supporting a causative link between the two key events. Temporal, effect and dose concordance was established in a number of experiments (Fornai et al. 2005; Fornai et al. 2003; Betabret et al. 2000 and 2006).

KE4a => KE5

Degeneration of DA neurons of nigrostriatal pathway leads to neuroinflammation

MODERATE

Rationale: multiple in vivo and in vitro experiments support the link between degeneration of DA neurons in the nigrostriatal pathway and neuroinflammation. The observation of concomitant presence of reactive microglial and astrocytic cells and degenerated/degenerating DA neurons is also reported in many studies with a good temporal and response concordance. ATP and other damage associated molecular patterns (DAMPs), released from degenerating cells, stimulate P2Y receptors on microglia, leading to their activation. Experimental injection of DAMPs, fraktalkine, or neuromelanin, released by degenerating DA neurons evokes neuroinflammation. Neutralization of DAMPs (e.g. antibodies against HMGB1 or CX3CR1) decreases MPTP-induced neuroinflammation. Toll-like receptor 4 deficient mice display a reduced neuroinflammatory response upon MPTP treatment. Inhibition of RAGE, which is upregulated in striatum upon rotenone exposure, suppresses NF-kB activation and downstream inflammatory markers.

KE5 => KE4b

Neuroinflammation leads to degeneration of DA neurons of nigrostriatal pathway.

MODERATE

Rationale: multiple in vivo and in vitro experiments support the link between neuroinflammation and degeneration of DA neurons in the nigrostriatal pathway. The observation of concomitant presence of reactive microglial and astrocytic cells and degenerated/degenerating DA neurons is also reported in many studies with a good temporal and response concordance. Neuroinflammation has been implicated in dopaminergic neuronal cell death in PD patients (Vivekanantham et al., 2014). LPS injection into the CNS, or applied systemically, evokes glial inflammation and a preferential degeneration of DA neurons. In mouse models with a knockout of either IL-1b, IFN-g, or TNF-a receptors 1 and 2, LPS no longer evokes neuroinflammation and DA neurodegeneration. Experimental interference with CD4+ T cell activation protects from DA neurodegeneration. Transfer of immunosuppressive regulatory T cells protect from DA neurodegeneration. Anti-inflammatory TGF-b1 signaling protects from DA neurodegeneration. Clinical trials indicate a protective influence on DA neuron survival by the antibiotic minocycline blocking microglial reactivity, in association with rasagiline (prevents DA degeneration), and coenzyme Q10/creatine (restoration of cellular ATP).

KE4b => AO

Degeneration of DA neurons of nigrostriatal pathway leads to parkinsonian motor symptoms

STRONG

Rationale: The experimental support linking the degeneration of DA neurons of nigrostriatal pathways with the manifestation of motor symptoms of PD comes from human in vivo observations as well as from monkey, mice and rat in vivo models exposed to an experimental toxin i.e. rotenone and MPTP. Observations in human allow defining correlation between the levels of striatal DA with the onset of motor dysfunction (Lloyd et al. 1975; Hornykiewicz et al. 1986; Bernheimer et al. 1973). Temporal, effect and dose concordance comes from studies performed in multiple animal species following administration of rotenone and MPTP (Bezard et al. 2001; Cannon et al. 2009; Petroske et al. 2001; Alvarez-Fischer et al. 2008; Blesa et al. 2012; Lloyd et a. 1975).

Uncertainties and Inconsistencies.

  • The strength of this AOP is mainly based on MPP+ and rotenone and only limited information on whether other mitochondrial complex I inhibitors also perturb the KEs (specifically degeneration of DA neurons in the SNpc)  or produce a similar AO.
  • Conflicting data exists (Choi et al. 2008) showing that mitochondrial complex I inhibition is not required for DA neuron death induced by rotenone, MPTP/MPP+, or paraquat, challenging the current AOP. The cited research article shows that abolishment of complex I’s activity by inactivation of a gene that codes for a subunit of complex I does not impact the survival of DA neurons in culture. The actions of rotenone, MPTP/MMP+ are independent of complex I. Since some complex I inhibitors also target other complexes, it is possible that impairment of other respiratory complexes may be involved. It was noted that this paper used the approach of genetically deleting an essential chaperone in complex I assembly, and the authors found that absence of complex I activity in this model did not affect the toxicity of rotenone and MPP+. However, the findings have never been confirmed/ continued, neither in the originating laboratory, nor by others. Second, the work did not consider the possibility that some functions of complex I were not affected by the absence of the chaperone (e.g. reverse electron transfer from complex II and III), and that rotenone and MPTP/MPP+ may well cause toxicity by interfering with such residual function (e.g. by favoring channeling of electrons to molecular oxygen). In light of this situation, the publication of Choi et al (2008) should be considered weak in the overall weight of evidence and therefore considered a minor inconsistency. 
  • There is no strict linear relationship between inhibitor binding and reduced mitochondrial function. Low doses of rotenone that inhibit CI activity partially do not alter mitochondrial oxygen consumption. Therefore, bioenergetics defect cannot account alone for rotenone-induced neurodegeneration. Instead, under such conditions, rotenone neurotoxicity may result from oxidative stress (Betarbet et al., 2000). Few studies used human brain cells/human brain mitochondria. Therefore, full quantitative data for humans are not available.
  • It is molecularly unclear how rotenone binding alter CI function, switching it to ROS production. It is still unclear whether the site of superoxide production in CI inhibited mitochondria is complex I itself or not (Singer and Ramsay, 1994).
  • Some studies suggest that rotenone and MPTP/MPP+ may have effects other than CI inhibition, e.g. MPTP and rotenone can induce microtubule disruption (Feng, 2006; Ren et al., 2005; Cappelletti et al., 1999; Cappelletti et al., 2001, Brinkley et al., 1974; Aguilar et al., 2015).
  • There are additional feedback possible between KEs, e.g. ROS production from KE2 may damage CI, this leads to enhancement of KE1.
  • Some KEs e.g. KE 2, 3, 5 pool molecular processes that may need to be evaluated individually at a later stage.
  • The exact molecular link from mitochondrial dysfunction to disturbed proteostasis is still unclear (Malkus et al., 2009; Zaltieri et al., 2015).
  • The role of ATP depletion vs. other features of mitochondrial dysfunction is not clear.
  • The role of a α-synuclein in neuronal degeneration is still unclear as well as the mechanisms leading to its aggregation.
  • It is not clear under which conditions KE3 and KE5 become modulatory factors, and when they are essential. MPTP/MPP+ can induce damage to nigrostriatal neurons without formation of Lewy bodies (Dauer 2003; Forno 1986, 1993). Similarly, discontinuous administration of rotenone, even at high doses, damages the basal ganglia but produce no inclusions (Heikkila et al., 1985; Ferrante et al., 1997, Lapontine 2004). To reproduce the formation of neuronal inclusions, continuous sc infusion of MPTP/MPP+ or rotenone is necessary. Acute intoxication with rotenone seems to spare dopaminergic neurons (Dauer et al., 2003, Ferrante 1997). In addition, in rats chronically infused with rotenone showed a reduction in striatal DARPP-32-positive (dopamine- and cyclic-AMP-regulated phosphoprotein of molecular weight 32,000), cholinergic and NADPH diaphorase-positive neurons (Hoglinger 2003) or in other brain regions. These results would suggest that Rotenone can induce a more widespread neurotoxicity (Aguilar 2015) or the model is not reproducible in all laboratories.
  • Inconsistent effects of MPTP/MPP+ on autophagy (up or down regulation) are reported (Drolet et al., 2004: Dauer et al., 2002). There is conflicting literature on whether increased autophagy would be protective or enhances damage. Similarly, a conflicting literature exists on extent of inhibition or activation of different protein degradation system in PD and a clear threshold of onset is unknown (Malkus et al., 2009; Fornai et al., 2005).
  • The selective vulnerability of the SNpc dopaminergic pathway does not have a molecular explanation.
  • In some instances, the differential vulnerability of various brain regions towards a generalized complex I inhibition found non-dopaminergic lesions, particularly in the striatum, in all animals with nigral lesion, as seen in atypical parkinsonism but not in idiopathic Parkinson's disease (Hoglinger et al., 2003)

  • Priority of the pattern leading to cell death could depend on concentration, time of exposure and species sensitivity; these factors have to be taken into consideration for the interpretation of the study’s result and extrapolation of potential low-dose chronic effect as this AOP refers to long-time exposure.
  • The model of striatal DA loss and its influence on motor output ganglia does not allow to explain specific motor abnormalities observed in PD (e.g. resting tremor vs bradykinesia) (Obeso et al., 2000). Other neurotransmitters (Ach) may play additional roles. Transfer to animal models of PD symptoms is also representing an uncertainties.
  • There are some reports indicating that in subacute rotenone or MPTP models (non-human primates), a significant, sometimes complete, recovery of motor deficits can be observed after termination of toxicant treatment. The role of neuronal plasticity in intoxication recovery and resilience is unclear.
  • This AOP is a linear sequence of KEs. However, mitochondrial dysfunction (and oxidative stress) and impaired proteostasis are influencing each other and this is considered an uncertainties (Malkus et al., 2009).

 

Known Modulating Factors

Modulating factors (MFs) may alter the shape of the response-response function that describes the quantitative relationship between two KES, thus having an impact on the progression of the pathway or the severity of the AO.The evidence supporting the influence of various modulating factors is assembled within the individual KERs. More help

Quantitative Understanding

Optional field to provide quantitative weight of evidence descriptors.  More help

The quantitative understanding of this AOP includes a clear response-response relationship and the identification of a threshold effect. The WoE analysis clearly supports the qualitative AOP as a means to identify and characterize the potential of a chemical to induce DA neuronal loss and the AO. Importantly, both the AO and the KE4 are considered relevant regulatory endpoints for this AOP. The empirical evidence supports existence of a response-response relationship. This response-response is likely triggered by a the brain concentrations of approximately 20-30 nM and 17-47 µM of rotenone and MPTP/MPP+ respectively and both concentrations trigger approx. a 50% inhibition of mitochondrial complex I and this could be considered as a “threshold”. However, a more detailed dose-response analysis for each KE is lacking as well as it is not clear which temporal relationship exists for lower CI inhibitory effects. It is clear from the analysis of the AOP that for the identification of these AOs, the design of the in-vivo studies should be tailored as to a MIE which leads to a long-lasting perturbation of the KEs. This provides the most specific and definite context to trigger neuronal death. To observe KEs relevant for this AOP, new endpoints need to be introduced. Although a dose, response and temporal relationship is evident for most KEs, the quantitative relationship between impaired proteostasis and degeneration of DA neurons has yet to be elucidated. Moving from a qualitative AOP to quantitative AOP would need a clear understanding of effect thresholds and this is still representing a major hurdle for several KEs of this AOP.

Table 3 Summary of quantitative effects at the concentration of rotenone and MPTP triggering the AO.

Concentration

KE1

Inhibition of C I

KE2

Mitochondrial dysfunction

KE3

Impaired proteostasis

KE4

Degeneration of DA neurons of nigrostriatal pathway

AO

Parkinsonian motor symptoms

Rotenone 20-30 nM rat brain concentration

[1-2]

Approx. 53%[4-5]

Approx. 20-53% (decrease in respiration rate)[1-2]

Approx. 20-60% (decrease in UPS (26S) activity) [3]

Neuronal loss (50% of animal affected) [2]

Motor impairment (100% of animals with neuronal loss) [2]

MPP+ 12-47 µM rat brain concentration [4-5]

Approx. 50-75% [5]

Approx. 38% (reduction in phosphorylating respiration) [5]

Approx. 60% (decrease in UPS activity) [4]

Approx. 50% of neuronal loss [4-5]

Motor impairment [4]

[1]; Okun et al., 1999 [2]; Barrientos and Moraes 1999; [3] Borland et al., 2008 [4] Thomas et al., 2012; [5] Betarbet et al., 2000 and 2006.

Summary of the proposed Key Events in this AOP:

Final graph.jpg

Chronic, low level of exposure to environmental chemicals that inhibit complex I could result in mitochondrial dysfunction and oxidative stress, triggering proteasomal dysfunction strongly implicated in parkinsonian disorders, including aggregation/modifications in α-synuclein protein and organelles trafficking. These cellular key events cause DA terminals degeneration in striatum and progressive cell death of DA neurons in SNpc, accompanied by neuroinflammation that potentiates neuronal cell death, finally leading to parkinsonian's motor symptoms. Important to notice that at each step, the effects become regionally restricted such that systemic complex I inhibition eventually results in highly selective degeneration of the nigrostriatal pathway.

Considerations for Potential Applications of the AOP (optional)

Addressess potential applications of an AOP to support regulatory decision-making.This may include, for example, possible utility for test guideline development or refinement, development of integrated testing and assessment approaches, development of (Q)SARs / or chemical profilers to facilitate the grouping of chemicals for subsequent read-across, screening level hazard assessments or even risk assessment. More help
  1. This AOP has been developed in order to evaluate the biological plausibility that the adverse outcome i.e. parkinsonian motor deficits, is linked to a MIE that can be triggered by chemical substances i.e. pesticides and chemicals in general. The relevance of the AOP has been documented by tools compounds known to trigger the described AOP. By means of using a human health outcome that has been shown in epidemiological studies to be association with pesticide exposure, the authors intend to draw attention on this AO in the process of hazard identification. This AOP can be used to support the biological plausibility of this association during the process of evaluation and integration of the epidemiological studies into the risk assessment. It is biologically plausible that a substance triggering the pathway, can be associated with the AO and ultimately with the human health outcome, pending the MoA analysis.
  2. In addition, this AOP can be used to support identification of data gaps that should be explored when a chemical substance is affecting the pathway. Moreover, the AOP provides a scaffold for recommendations on the most adequate study design to investigate the apical endpoints. It is important to note that, although the AO is defined in this AOP as parkinsonian motor deficits, degeneration of DA neurons is already per se an adverse outcome even in situations where it is not leading to parkinsonian motor deficits, and this should be taken into consideration for the regulatory applications of this AOP.
  3. The MIE and KEs identified in this AOP could serve as a basis for assays development that could contribute to an AOP informed-IATA construction which can be applied for different purposes such as: screening and prioritization of chemicals for further testing, hazard characterization or even risk assessment when combined with exposure and ADME information.
  4. This AOP can be used for neurotoxicity assessment, since it is plausible that a compound that binds to the mitochondrial complex I may eventually lead to Parkinsonian motor deficits.

  5. The regulatory applicability of this AOP would be to use experimental findings in model systems representing the MIE and KEs as indicators/alerts for the AO. Risk assessment may be possible if bioavailability at the target cells can be estimated, the toxic concentrations in vitro can be extrapolated to in vivo and exposure scenarios can be simulated.

  6. This AOP can be applied for chemicals that have structural similarities to rotenone or MPTP. However, this AOP may not at the moment be used for chemicals that do not resemble rotenone or MPTP. It is however expected that compounds acting on the same MIE, but belonging to different chemical classes and those that are structurally different, will be tested in the near future in order to substantiate a broader specificity for this AOP. However, it remains evident that chemicals affecting the MIE are potential risk factors for this AO.

References

List of the literature that was cited for this AOP. More help

Abdelsalam RMSafar MM J Neurochem. Neuroprotective effects of vildagliptin in rat rotenone Parkinson's disease model: role of RAGE-NFκB and Nrf2-antioxidant signaling pathways. 2015 Jun;133(5):700-7. doi: 10.1111/jnc.13087. Epub 2015 Mar 26.

Aguilar JS, Kostrzewa RM. Neurotoxin mechanisms and processes relevant to parkinson’s disease: un update. Neurotox Res. DOI 10.1007/s12640-015-9519-y.

Alvarez-Fischer D, Guerreiro S, Hunot S, Saurini F, Marien M, Sokoloff P, Hirsch EC, Hartmann A, Michel PP. Modelling Parkinson-like neurodegeneration via osmotic minipump delivery of MPTP and probenecid. J Neurochem. 2008 Nov;107(3):701-11. doi: 10.1111/j.1471-4159.2008.05651.x. Epub 2008 Sep 16.

Arnold, B., et al. (2011). "Integrating Multiple Aspects of Mitochondrial Dynamics in Neurons: Age-Related Differences and Dynamic Changes in a Chronic Rotenone Model." Neurobiology of Disease 41(1): 189-200.

Barbeito AG, Mesci P, Boillee S. 2010. Motor neuron-immune interactions: the vicious circle of ALS. J Neural Transm 117(8): 981-1000.

Barrientos A., and Moraes C.T. (1999) Titrating the Effects of Mitochondrial Complex I Impairment in the Cell Physiology. Vol. 274, No. 23, pp. 16188–16197.

Bernheimer H, Birkmayer W, Hornykiewicz O, Jellinger K, Seitelberger F. Brain dopamine and the syndromes of Parkinson and Huntington. Clinical, morphological and neurochemical correlations. J Neurol Sci. 1973 Dec;20(4):415-55

Betarbet R, Sherer TB, MacKenzie G, Garcia-Osuna M, Panov AV, Greenamyre JT. 2000. Chronic systemic pesticide exposure reproduces features of Parkinson’s disease. Nat Neurosci. 3:1301–6

Betarbet R, Canet-Aviles RM, Sherer TB, Mastroberardino PG,Mc Lendon C, Kim JH, Lund S, Na HM, taylor G, Bence NF, kopito R, seo BB, Yagi T, Yagi A, Klinfelter G, Cookson MR, Greenmyre JT. 2006. Intersecting pathways to neurodegeneration in Parkinson’s disease: effects of the pesticide rotenone on DJ-1, α-synuclein, and the ubiquitin-proteasome system. Neurobiology disease. (22) 404-20.

Bezard E, Dovero S, Prunier C, Ravenscroft P, Chalon S, Guilloteau D, Crossman AR, Bioulac B, Brotchie JM, Gross CE (2001) Relationship between the appearance of symptoms and the level of nigrostriatal degeneration in a progressive 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-lesioned macaque model of Parkinson's disease. J Neurosci. 21(17):6853-61.

Blasko I, Stampfer-Kountchev M, Robatscher P, Veerhuis R, Eikelenboom P, Grubeck-Loebenstein B. 2004. How chronic inflammation can affect the brain and support the development of Alzheimer's disease in old age: the role of microglia and astrocytes. Aging cell 3(4): 169-176.

Blesa J, Pifl C, Sánchez-González MA, Juri C, García-Cabezas MA, Adánez R, Iglesias E, Collantes M, Peñuelas I, Sánchez-Hernández JJ, Rodríguez-Oroz MC, Avendaño C, Hornykiewicz O, Cavada C, Obeso JA (2012) The nigrostriatal system in the presymptomatic and symptomatic stages in the MPTP monkey model: a PET, histological and biochemical study. Neurobiol Dis. 48(1):79-91.

Bodea LG, Wang Y, Linnartz-Gerlach B, Kopatz J, Sinkkonen L, Musgrove R, et al. 2014. Neurodegeneration by activation of the microglial complement-phagosome pathway. J Neurosci 34(25): 8546-8556.

Borrajo A, Rodriguez-Perez AI, Villar-Cheda B, Guerra MJ, Labandeira-Garcia JL. 2014. Inhibition of the microglial response is essential for the neuroprotective effects of Rho-kinase inhibitors on MPTP-induced dopaminergic cell death. Neuropharmacology 85: 1-8

Brinkley BR, Barham SS, Barranco SC, and Fuller GM. 1974. Rotenone inhibition of spindle microtubule assembly in mammalian cells,” Experimental Cell Research. 85(1)41–46.

Brown GC, Bal-Price A (2003) Inflammatory neurodegeneration mediated by nitric oxide, glutamate, and mitochondria. Mol Neurobiol 27: 325-355

Braun RJ. (2012). Mitochondrion-mediated cell death: dissecting yeast apoptosis for a better understanding of neurodegeneration. Front Oncol 2:182.

Brzozowski MJ, Jenner P, Rose S. 2015. Inhibition of i-NOS but not n-NOS protects rat primary cell cultures against MPP(+)-induced neuronal toxicity. J Neural Transm 122(6): 779-788.

Cacquevel M, Lebeurrier N, Cheenne S, Vivien D. 2004. Cytokines in neuroinflammation and Alzheimer's disease. Curr Drug Targets 5(6): 529-534.

Calne DB, Sandler M (1970) L-Dopa and Parkinsonism. Nature. 226(5240):21-4.

Cannon JR, Tapias V, Na HM, Honick AS, Drolet RE, Greenamyre JT (2009) A highly reproducible rotenone model of Parkinson's disease. Neurobiol Dis. 34(2):279-90.

Cappelletti G, Maggioni MG, Maci R. 1999. Influence of MPP+ on the state of tubulin polymerisation in NGF-differentiated PC12 cells. J Neurosci Res. 56(1):28-35.

Cao S, Theodore S, Standaert DG.2010. Fc gamma receprors are required for NF-kB signaling, microglial activation and dopaminergic neurodegeneration in an AAV-synuclein mouse model of Parkinson's disease. molecular neurodegeneration.5-42.

Cappelletti G, Pedrotti B, Maggioni MG, Maci R. 2001. Microtubule assembly is directly affected by MPP(+)in vitro. Cell Biol Int.25(10):981-4.

Castrioto A, Lozano AM, Poon YY, Lang AE, Fallis M, Moro E. 2011. Ten-year outcome of subthalamic stimulation in Parkinson disease: a blinded evaluation. Arch Neurol. 68(12):1550-6.

Chang CY, Choi DK, Lee DK, Hong YJ, Park EJ. 2013. Resveratrol confers protection against rotenone-induced neurotoxicity by modulating myeloperoxidase levels in glial cells. PLoS One 8(4): e60654.

Champy et al. (2004). Annonacin, a lipophilic inhibitor of mitochondrial complex I, induces nigral and striatal neurodegeneration in rats: possible relevance for atypical Parkinsonism in Guadeloupe. J Neurochem 88: 63-69. 

Chao YX, He BP, Tay SS. 2009. Mesenchymal stem cell transplantation attenuates blood brain barrier damage and neuroinflammation and protects dopaminergic neurons against MPTP toxicity in the substantia nigra in a model of Parkinson's disease. J Neuroimmunol 216(1-2): 39-50.

Chen Y, Zhang DQ, Liao Z, Wang B, Gong S, Wang C, Zhang MZ, Wang GH, Cai H, Liao FF, Xu JP 2015. Anti-oxidant polydatin (piceid) protects against substantia nigral motor degeneration in multiple rodent models of Parkinson's disease. Mol Neurodegener. 2;10(1):4.

Chinta SJ, Andersen JK (2006) Reversible inhibition of mitochondrial complex I activity following chronic dopaminergic glutathione depletion in vitro: implications for Parkinson's disease. Free Radic Biol Med. 41(9):1442-8.

Choi WS., Kruse S.E., Palmiter R, Xia Z., (2008) Mitochondrial complex I inhibition is not required for dopaminergic neuron death induced by rotenone, MPP, or paraquat. PNAS, 105, 39, 15136-15141

Choi BS, Kim H, Lee HJ, Sapkota K, Park SE, Kim S, Kim SJ (2014) Celastrol from 'Thunder God Vine' protects SH-SY5Y cells through the preservation of mitochondrial function and inhibition of p38 MAPK in a rotenone model of Parkinson's disease. Neurochem Res. 39(1):84-96.

Chiu CC, Yeh TH, Lai SC, Wu-Chou YH, Chen CH, Mochly-Rosen D, Huang YC, Chen YJ, Chen CL, Chang YM, Wang HL, Lu CS. 2015. Neuroprotective effects of aldehyde dehydrogenase 2 activation in rotenone-induced cellular and animal models of parkinsonism. Exp Neurol. 263:244-53.

Chou AP, Li S, Fitzmaurice AG, Bronstein JM. 2010. Mechanisms of rotenone-induced proteasome inhibition. NeuroToxicology. 31:367–372. Chung YC, Kim SR, Park JY, Chung ES, Park KW, Won SY, et al. 2011. Fluoxetine prevents MPTP-induced loss of dopaminergic neurons by inhibiting microglial activation. Neuropharmacology 60(6): 963-974.

Correia SC, Santos RX, Perry G, Zhu X, Moreira PI, Smith MA. (2012). Mitochondrial importance in Alzheimer’s, Huntington’s and Parkinson’s diseases. Adv Exp Med Biol 724:205 – 221.

Cotzias GC, Papavasiliou PS, Gellene R. 1969. L-dopa in parkinson's syndrome. N Engl J Med. 281(5):272.

Cozzolino M, Ferri A, Valle C, Carri MT. (2013). Mitochondria and ALS: implications from novel genes and pathways. Mol Cell Neurosci 55:44 – 49.

Dagda RK, Banerjee TD and Janda E. 2013. How Parkinsonian Toxins Dysregulate the Autophagy Machinery. Int. J. Mol. Sci. 14:22163-22189.

Dauer W, Kholodilov N, Vila M, Trillat AC, Goodchild R, Larsen KE, Staal R, Tieu K, Schmitz Y, Yuan CA, Rocha M, Jackson-Lewis V, Hersch S, Sulzer D, Przedborski S, Burke R, Hen R. 2002. Resistance of alpha -synuclein null mice to the parkinsonian neurotoxin MPTP. Proc Natl Acad Sci U S A. 99(22):14524-9.

Dauer W, Kholodilov N, Vila M, Trillat AC, Goodchild R, Larsen KE, Staal R, Tieu K, Schmitz Y, Yuan CA, Rocha M, Jackson-Lewis V, Hersch S, Sulzer D, Przedborski S, Burke R, Hen R. 2002. Resistance of alpha -synuclein null mice to the parkinsonian neurotoxin MPTP. Proc Natl Acad Sci U S A. 99(22):14524-9.

Dauer W, Przerdborski S. 2003. Parkinson’sdisease: Mechanisms and Models.Neuron. 39, 889-9.

De Bie RM, de Haan RJ, Nijssen PC, Rutgers AW, Beute GN, Bosch DA, Haaxma R, Schmand B, Schuurman PR, Staal MJ, Speelman JD. 1999. Unilateral pallidotomy in Parkinson's disease: a randomised, single-blind, multicentre trial. Lancet. 354(9191):1665-9.

Degli Esposti M, Ghelli A. 1994. The mechanism of proton and electron transport in mitochondrial complex I. Biochim Biophys Acta.1187(2):116–120.

Dehay B, Bove J, Rodriguez-Muela N, Perier C, Recasens A, Boya P, Vila M. 2010. Pathogenic lysosomal depletion in Parkinson’s disease. J. Neurosci. 30:12535–12544.

Dehmer T, Lindenau J, Haid S, Dichgans J, Schulz JB. 2000. Deficiency of inducible nitric oxide synthase protects against MPTP toxicity in vivo. J Neurochem 74(5): 2213-2216.

Deuschl G, Schade-Brittinger C, Krack P, Volkmann J, Schäfer H, Bötzel K, Daniels C, Deutschländer A, Dillmann U, Eisner W, Gruber D, Hamel W, Herzog J, Hilker R, Klebe S, Kloss M, Koy J, Krause M, Kupsch A, Lorenz D, Lorenzl S, Mehdorn HM, Moringlane JR, Oertel W, Pinsker MO, Reichmann H, Reuss A, Schneider GH, Schnitzler A, Steude U, Sturm V, Timmermann L, Tronnier V, Trottenberg T, Wojtecki L, Wolf E, Poewe W, Voges J; German Parkinson Study Group, Neurostimulation Section. 2006. A randomized trial of deep-brain stimulation for Parkinson's disease. N Engl J Med. 355(9):896-908.

Dexter D. T., Jenner P.. Parkinson disease: from pathology to molecular disease mechanisms. Free Radical Biology and Medicine 62 (2013) 132-144

Dietz GPH, Stockhausen KV, Dietz B et al. (2008) Membrane-permeable Bcl-xL prevents MPTP-induced dopaminergic neuronal loss in the substantia nigra. J Neurochem 104:757-765. Doi:10.1111/j.1471-4159.2007.05028.

Drolet RE, Behrouz B, Lookingland KJ, Goudreau JL, 2004. Mice lacking α-synuclein have an attenuated loss of striatal dopamine following prolonged chronic MPTP administration. Neurotoxicology. 25(5):761-9.

Drouin-Ouellet J, St-Amour I, Saint-Pierre M, Lamontagne-Prolux J, Kriz J, Barker R, Cicchetti F.2015. Toll-like receptor expression in the blood and brain of patients and a mouse of Parkinson's disease. International Journal of Neuropsychopharmacology. 1-11.

Du T, Li L, Song N, Xie J, Jiang H (2010) Rosmarinic acid antagonized 1-methyl-4-phenylpyridinium (MPP+)-induced neurotoxicity in MES23.5 dopaminergic cells. Int J Toxicol. 29(6):625-33.

Efremov RG, Sazanov LA. Respiratory complex I: 'steam engine' of the cell? Curr Opin Struct Biol. 2011 Aug;21(4):532-40. doi: 10.1016/j.sbi.2011.07.002. Epub 2011 Aug 8. Review.

Efremov RG, Sazanov LA. Structure of the membrane domain of respiratory complex I. Nature. 2011 Aug 7;476(7361):414-20. doi: 10.1038/nature10330.

Emmrich JV, Hornik TC, Neher JJ, Brown GC. 2013. Rotenone induces neuronal death by microglial phagocytosis of neurons. The FEBS journal 280(20): 5030-5038.

Esposti et al. (1993) Complex I and Complex III of mitochondria have common inhibitors acting as ubiquinone antagonists. Biochem Biophys Res Commun 190(3): 1090-6. 

Fasano A, Romito LM, Daniele A, Piano C, Zinno M, Bentivoglio AR, Albanese A. 2010. Motor and cognitive outcome in patients with Parkinson's disease 8 years after subthalamic implants. Brain. 133(9):2664-76.

Fato et al. (2009) Differential effects of mitochondrial complex I inhibitors on production of reactive oxygen species. Biochim Biophys Acta 1787(5): 384-392. 

Feng ZH, Wang TG, Li DD, Fung P, Wilson BC, Liu B, et al. 2002. Cyclooxygenase-2-deficient mice are resistant to 1-methyl-4-phenyl1, 2, 3, 6-tetrahydropyridine-induced damage of dopaminergic neurons in the substantia nigra. Neurosci Lett 329(3): 354-358.

Feng J. Mictrotubule. A common target for parkin and Parkinson's disease toxins. Neuroscientist 2006, 12.469-76.

Ferger B, Leng A, Mura A, Hengerer B, Feldon J. 2004. Genetic ablation of tumor necrosis factor-alpha (TNF-alpha) and pharmacological inhibition of TNF-synthesis attenuates MPTP toxicity in mouse striatum. J Neurochem 89(4): 822-833.

Ferrante RJ, Schulz JB, Kowall NW, Beal MF. 1997. Systematic administration of rotenone produces selective damage in the striatum and globus pallidus, but not in the substantia nigra. Brain Research. (753). 157-2.

Ferrari-Toninelli G, Bonini SA, Cenini G, Maccarinelli G, Grilli M, Uberti D, Memo M. 2008. Dopamine receptor agonists for protection and repair in Parkinson's disease. Curr Top Med Chem. 8(12):1089-99.

Friedman LG, lachenmayer ML, Wang J, He L, Poulose SM, Komatsu M, Holstein GR, Yue Z. 2012. Disrupted autophagy leads to dopaminergic axon and dendrite degeneration and promotes presynaptic accumulation of α-synuclein and LRRK2 in the brain. The Journal of Neuroscience. 32 (22) 7585-93.

Fornai F, Lenzi P, Gesi M, Ferrucci M, Lazzeri G, Busceti C, Ruffoli R, Soldani P, Ruggieri S, Alessandri’ MG, Paparelli A. 2003. Fine structure and mechanisms underlying nigrostriatal inclusions and cell death after proteasome inhibition. The journal of neuroscience. 23 (26) 8955-6.

Fornai F., P. Lenzi, M. Gesi et al., “Methamphetamine produces neuronal inclusions in the nigrostriatal system and in PC12 cells,” Journal of Neurochemistry, vol. 88, no. 1, pp. 114–123, 2004.

Fornai F, Schlüter OM, Lenzi P, Gesi M, Ruffoli R, Ferrucci M, Lazzeri G, Busceti CL, Pontarelli F, Battaglia G, Pellegrini A, Nicoletti F, Ruggieri S, Paparelli A, Südhof TC. 2005. Parkinson-like syndrome induced by continuous MPTP infusion: Convergent roles of the ubiquitinproteasome system and _α-synuclein. PNAS. 102: 3413–3418.

Freed CR, Breeze RE, Rosenberg NL, Schneck SA, Wells TH, Barrett JN, Grafton ST, Huang SC, Eidelberg D, Rottenberg DA. 1990.

Transplantation of human fetal dopamine cells for Parkinson's disease. Results at 1 year. Arch Neurol. 47(5):505-12.

Fujita KA, Ostaszewski M, Matsuoka Y, Ghosh S, Glaab E, Trefois C, Crespo I, Perumal TM, Jurkowski W, Antony PM, Diederich N, Buttini M, Kodama A, Satagopam VP, Eifes S, Del Sol A, Schneider R, Kitano H, Balling R. 2014. Integrating pathways of Parkinson's disease in a molecular interaction map. Mol Neurobiol.49(1):88-102.

Gandhi S, Wood-Kaczmar A, Yao Z, et al. PINK1-associated Parkinson’s disease is caused by neuronal vulnerability to calcium-induced cell death. Molecular Cell. 2009;33:627–638.

Gao HM, Hong JS, Zhang W, Liu B. 2002. Distinct role for microglia in rotenone-induced degeneration of dopaminergic neurons. J Neurosci 22(3): 782-790.

Gao HM, Liu B, Hong JS. 2003. Critical role for microglial NADPH oxidase in rotenone-induced degeneration of dopaminergic neurons. J Neurosci 23(15): 6181-6187.

Gao L, Brenner D, Llorens-Bobadilla E, Saiz-Castro G, Frank T, Wieghofer P, et al. 2015. Infiltration of circulating myeloid cells through CD95L contributes to neurodegeneration in mice. J Exp Med 212(4): 469-480.

Graier WF, Frieden M, Malli R. (2007). Mitochondria and Ca2+ signaling: old guests, new functions. Pflugers Arch 455:375–396.

Greenamyre, J T., Sherer, T.B., Betarbet, R., and Panov A.V. (2001) Critical Review Complex I and Parkinson’s Disease Life, 52: 135–141.

Greenamayre et al. 2010. Lessons from the rotenone model of Parkinson's disease. Trends pharmacol. Sci. 31(4):141-2

Griffin WS, Sheng JG, Royston MC, Gentleman SM, McKenzie JE, Graham DI, et al. 1998. Glial-neuronal interactions in Alzheimer's disease: the potential role of a 'cytokine cycle' in disease progression. Brain Pathol 8(1): 65-72.

Grivennikova, V.G., Maklashina, E.O., E.V. Gavrikova, A.D. Vinogradov (1997) Interaction of the mitochondrial NADH-ubiquinone reductase with rotenone as related to the enzyme active/inactive transition Biochim. Biophys. Acta, 1319 (1997), pp. 223–232

Hajieva P, Mocko JB, Moosmann B, Behl C (2009) Novel imine antioxidants at low nanomolar concentrations protect dopaminergic cells from oxidative neurotoxicity. J Neurochem. 110(1):118-32.

Hoglinger G.U. et al.2003.Chronic systemic complex I inhibition induces a hypokynetic multisystem degeneration in rats. J.neurochem 84:491-502.

Hornykiewicz O, Kish SJ. 1987. Biochemical pathophysiology of parkinson’s disease. In Parkinson’s Disease. M Yahr and K.J. Bergmann, eds (New.York: Raven Press) 19-34.

Jana S, Sinha M, Chanda D, Roy T, Banerjee K, Munshi S, Patro BS, Chakrabarti S (2011) Mitochondrial dysfunction mediated by quinone oxidation products of dopamine: Implications in dopamine cytotoxicity and pathogenesis of Parkinson's disease. Biochim Biophys Acta. 1812(6):663-73.

Jha N, Jurma O, Lalli G, Liu Y, Pettus EH, Greenamyre JT, Liu RM, Forman HJ, Andersen JK (2000) Glutathione depletion in PC12 results in selective inhibition of mitochondrial complex I activity. Implications for Parkinson's disease. J Biol Chem. 275(34):26096-101.

Johnson ME, Bobrovskaya L. 2015. An update on the rotenone models of parkinson’s disease: Their ability to reproduce features of clinical disease and model gene-environment interactions. 946). 101-16.

Heikkila RE, Nicklas WJ, Vyas I, Duvoisin RC. 1985. Dopaminergic toxicity of rotenone and the 1-methyl-4-phenylpyridinium ion after their stereotaxic administration to rats: implication for the mechanism of 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine toxicity. Neurosci Lett. 62(3):389-94.

Henderson BT, Clough CG, Hughes RC, Hitchcock ER, Kenny BG. 1991. Implantation of human fetal ventral mesencephalon to the right caudate nucleus in advanced Parkinson's disease. Arch Neurol. 48(8):822-7.

Hirsch EC, Hunot S. 2009. Neuroinflammation in Parkinson's disease: a target for neuroprotection? Lancet Neurol 8(4): 382-397. Hoglinger GU, Feger J, Annick P, Michel PP, Karine P, Champy P, Ruberg M, Wolfgang WO, Hirsch E. 2003. Chronic systemic complexI inhibition induces a hypokinetic multisystem degeneration in rats. J. Neurochem.. (84) 1-12.

Ichimaru N, Murai M, Kakutani N, Kako J, Ishihara A, Nakagawa Y, Miyoshi H. 2008.. Synthesis and Characterization of New Piperazine-Type Inhibitors for Mitochondrial NADH-Ubiquinone Oxidoreductase (Complex I). Biochemistry. 47(40)10816–10826.

Inden M, Yoshihisa Kitamura, Hiroki Takeuchi, Takashi Yanagida, Kazuyuki Takata, Yuka Kobayashi, Takashi Taniguchi, Kanji Yoshimoto, Masahiko Kaneko, Yasunobu Okuma, Takahiro Taira, Hiroyoshi Ariga and Shun Shimohama. 2007. Neurodegeneration of mouse nigrostriatal dopaminergic system induced by repeated oral administration of rotenone is prevented by 4-phenylbutyrate, a chemical chaperone. Journal of Neurochemistry. 101.(6).1491–4.

Keeney PM,Xie J,Capaldi RA,Bennett JP Jr. (2006) Parkinson's disease brain mitochondrial complex I has oxidatively damaged subunits and is functionally impaired and misassembled. J Neurosci. 10;26(19):5256-64.

Kelly PJ, Ahlskog JE, Goerss SJ, Daube JR, Duffy JR, Kall BA. 1987. Computer-assisted stereotactic ventralis lateralis thalamotomy with microelectrode recording control in patients with Parkinson's disease. Mayo Clin Proc. 62(8):655-64.

Khan MM, Kempuraj D, Zaheer S, Zaheer A. 2014. Glia maturation factor deficiency suppresses 1-methyl-4-phenylpyridinium-induced oxidative stress in astrocytes. J Mol Neurosci 53(4): 590-599.

Kim-Han JS, Dorsey JA, O’Malley KL. 2011. The parkinsonian mimetic MPP+, specifically impairs mitochondrial transport in dopamine axons. The Journal of Neuroscience. 31(19) 7212-1.

Kirk D, Rosenblad C, Burger C, Lundberg C, Johansen TE, Muzyczka N, Mandel R, Bijorklund A. 2002. Parkinson-like neurodegeneration induced by targeted overexpression of α-synuclein in the nigrostriatal system. 22(7) 2780-91.

Kirk D, Annett L, Burger C, Muzyczka N, Mandel R, Bijorklund A. 2003. Nigrostriatal α-synucleinopathy induced by viral vector-mediated overexpression of human α-synuclein: A new primate model of parkinson’s disease. PNAS (100) 2884-9.

Klein RL, King MA, Hamby ME, Meyer EM. 2002. Dopaminergic cell loss induced by human A30P α-synuclein gene transfer to the rat substantia nigra. Hum.Gene.Ther. (13) 605-2.

Koller WC (1992) When does Parkinson's disease begin? Neurology. 42(4 Suppl 4):27-31 Koopman W, Hink M, Verkaart S, Visch H, Smeitink J, Willems P. 2007. Partial complex I inhibition decreases mitochondrial motility and increases matrix protein diffusion as revealed by fluorescence correlation spectroscopy. Biochimica et Biophysica Acta 1767:940-947.

Koopman W, Willems P (2012) Monogenic mitochondrial disorders. New Engl J Med. 22;366(12):1132-41. doi: 10.1056/NEJMra1012478.

Kraft AD, Harry GJ. 2011. Features of microglia and neuroinflammation relevant to environmental exposure and neurotoxicity. International journal of environmental research and public health 8(7): 2980-3018.

Lagoa et al. (2011) Complex I and cytochrome c are molecular targets of flavonoids that inhibit hydrogen peroxide production by mitochondria. Biochimica et Biophys Acta 1807: 1562-1572.

Lang AE, Lozano AM. 1998. Parkinson's disease. Second of two parts. N Engl J Med. 339(16):1130-43.

Langston JW, ballard P, Irwin I. 1983. Chronic parkinsonism in human due to a product of meperidine-analog synthesis. Science. (219) 979-0.

Lapointe N, StHilaire M, martinoli MG, Blanchet J, gould P, Rouillard C, Cicchetti F. 2004. Rotenone induces non-specific central nervous system and systemic toxicity. The FASEB Journal express article 10.1096/fj.03-0677fje

Lauwers E, Debyser Z, Van Drope J, DeStrooper B, Nuttin B. 2003. Neuropathology and neurodegeneration in rodent brain induced by lentiviral vector-mediated overexpression of α-synuclein. Brain pathol. (13) 364-72.

Liu Y, Hu J, Wu J, Zhu C, Hui Y, Han Y, et al. 2012. alpha7 nicotinic acetylcholine receptor-mediated neuroprotection against dopaminergic neuron loss in an MPTP mouse model via inhibition of astrocyte activation. J Neuroinflammation 9: 98.

Liu Y, Li W, Tan C, Liu X, Wang X, Gui Y, Qin L, Deng F, Hu C, Chen L. 2014. Meta-analysis comparing deep brain stimulation of the globus pallidus and subthalamic nucleus to treat advanced Parkinson disease. J Neurosurg. 121(3):709-18.

Liu Y, Zeng X, Hui Y, Zhu C, Wu J, Taylor DH, et al. 2015. Activation of alpha7 nicotinic acetylcholine receptors protects astrocytes against oxidative stress-induced apoptosis: implications for Parkinson's disease. Neuropharmacology 91: 87-96.

Liu W, Kong S, Xie Q, Su J, Li W, Guo H, Li S, Feng X, Su Z, Xu Y, Lai X. Protective effects of apigenin against 1-methyl-4-phenylpyridinium ion induced neurotoxicity in PC12 cells. Int J Mol Med. 2015, 35(3):739-46.

Lloyd KG, Davidson L, Hornykiewicz O (1975) The neurochemistry of Parkinson's disease: effect of L-dopa therapy. J Pharmacol Exp Ther. 195(3):453-64.

Lo Bianco C, Ridet JL, Deglon N, Aebischer P. 2002. Alpha-synucleopathy and selective dopaminergic neuron loss in a rat lentiviral-based model of Parkinson’s disease. Proc.natl.Sci.USA (99)10813-8.

López-Lozano JJ, Bravo G, Abascal J. 1991. Grafting of perfused adrenal medullary tissue into the caudate nucleus of patients with Parkinson's disease. Clinica Puerta de Hierro Neural Transplantation Group. J Neurosurg. 75(2):234-43.

Mangano EN, Litteljohn D, So R, Nelson E, Peters S, Bethune C, et al. 2012. Interferon-gamma plays a role in paraquat-induced neurodegeneration involving oxidative and proinflammatory pathways. Neurobiol Aging 33(7): 1411-1426.

Marella M, Seo BB, Nakamaru-Ogiso E, Greenamyre JT, Matsuno-Yagi A, Yagi T (2008) Protection by the NDI1 gene against neurodegeneration in a rotenone rat model of Parkinson's disease. PLoS One. 3(1):e1433.

Martin LJ. (2011). Mitochondrial pathobiology in ALS. J Bioenerg Biomembr 43:569 – 579.

Matsumoto K, Asano T, Baba T, Miyamoto T, Ohmoto T. 1976. Long-term follow-up results of bilateral thalamotomy for parkinsonism. Appl Neurophysiol. 39(3-4):257-60.

McGeer PL, McGeer EG. 1998. Glial cell reactions in neurodegenerative diseases: Pathophysiology and therapeutic interventions. Alzheimer DisAssocDisord 12 Suppl. 2: S1-S6.

McNaught KS, Jenner P. 2001. Proteasomal function is impaired in substantia nigra in Parkinson’s disease. Neurosci. Lett. 297, 191– 194. McNaught KSC, Olanow W, Halliwell B. 2001. Failure of the ubiquitine-proteasome system in parkinson’s disease. Nature Rev. Neurosci. (2) 589-4.

McNaught KS, Belizaire R, Isacson O, Jenner P, Olanow CW. 2003. Altered proteasomal function in sporadic Parkinson’s disease. Exp. Neurol. 179, 38– 46.

Moldovan AS, Groiss SJ, Elben S, Südmeyer M, Schnitzler A, Wojtecki L. 2015. The treatment of Parkinson's disease with deep brain stimulation: current issues. Neural Regen Res. 10(7):1018-22.

Mudò G, Mäkelä J, Di Liberto V, Tselykh TV, Olivieri M, Piepponen P, Eriksson O, Mälkiä A, Bonomo A, Kairisalo M, Aguirre JA, Korhonen L, Belluardo N, Lindholm D. (2012) Transgenic expression and activation of PGC-1α protect dopaminergic neurons in the MPTP mouse model of Parkinson's disease. Cell Mol Life Sci. 69(7):1153-65.

Narabayashi H, Yokochi F, Nakajima Y. 1984. Levodopa-induced dyskinesia and thalamotomy. J Neurol Neurosurg Psychiatry. 47(8):831-9. Nataraj J, Manivasagam T, Justin Thenmozhi A, Essa MM 2015. Lutein protects dopaminergic neurons against MPTP-induced apoptotic death and motor dysfunction by ameliorating mitochondrial disruption and oxidative stress. Nutr Neurosci. 2015 Mar 2. [Epub ahead of print].

Obeso JA, Rodríguez-Oroz MC, Rodríguez M, Lanciego JL, Artieda J, Gonzalo N, Olanow CW (2000) Pathophysiology of the basal ganglia in Parkinson's disease. Trends Neurosci. 23(10 Suppl):S8-19.

Offen D, Beart PM, Cheung NS et al. (1998) Transegnic mice expressing human Bcl-2 in their neurons are resistant to 6-hydroxydopamine and 1-methyl-4-phenyl-1,2,3,6- tetrahydropyridine neurotoxicity. PNAS 95:5789-5794

O’Malley KL. 2010. The role of axonopathy in Parkinson’s disease. 2010. Experimental Neurobiology. (19). 115-19.

Okun, J.G, Lümmen, P and Brandt U., (1999) Three Classes of Inhibitors Share a Common Binding Domain in Mitochondrial Complex I (NADH:Ubiquinone Oxidoreductase) J. Biol. Chem. 274: 2625-2630. doi:10.1074/jbc.274.5.2625

Pan T, Kondo S, Le W, Jankovic J. 2008. The role of autophagy-lysosome pathway in neurodegeneration associated with Parkinson’s disease. Brain. 131, 1969-1978.

Park C.et al. (2003) Quercetin protects the hydrogen peroxide-induced apoptosis via inhibition of mitochondrial dysfunction in H9c2 cardiomyoblast cells. Biochem Pharmacol 66(7): 1287-1295.

Park SE, Sapkota K, Choi JH, Kim MK, Kim YH, Kim KM, Kim KJ, Oh HN, Kim SJ, Kim S (2014) Rutin from Dendropanax morbifera Leveille protects human dopaminergic cells against rotenone induced cell injury through inhibiting JNK and p38 MAPK signaling. Neurochem Res. 39(4):707-18.

Parker WD Jr, Boyson SJ, Parks JK. 1989. Abnormalities of the electron transport chain in idiopathic Parkinson's disease. Ann Neurol.26(6):719-23.

pasqualiL, Caldarazzo-Ienco Fornai . (2014). MPTP neurotoxicity:actions, mechanisms, and animal modeling of Parkinson's disease. In: Kostrzewa RM (ed) Handbook of neurotoxicity. Springer, Heidelberg, pp237-275.

Peschanski M, Defer G, N'Guyen JP, Ricolfi F, Monfort JC, Remy P, Geny C, Samson Y, Hantraye P, Jeny R. 1994. Bilateral motor improvement and alteration of L-dopa effect in two patients with Parkinson's disease following intrastriatal transplantation of foetal ventral mesencephalon. Brain. 117 ( Pt 3):487-99.

Petroske E, Meredith GE, Callen S, Totterdell S, Lau YS (2001) Mouse model of Parkinsonism: a comparison between subacute MPTP and chronic MPTP/probenecid treatment. Neuroscience. 106(3):589-601.

Powers ET1, Morimoto RI, Dillin A, Kelly JW, Balch WE. 2009.. Biological and Chemical Approaches to Diseases of Proteostasis Deficiency. Ann. Rev. Biochem 78: 959–91.

Purisai MG, McCormack AL, Cumine S, Li J, Isla MZ, Di Monte DA. 2007. Microglial activation as a priming event leading to paraquat-induced dopaminergic cell degeneration. Neurobiol Dis 25(2): 392-400. Parker WD Jr, Parks JK, Swerdlow RH (2008) Complex I deficiency in Parkinson's disease frontal cortex. Brain Res. 1189:215-8.

Qian L, Wu HM, Chen SH, Zhang D, Ali SF, Peterson L, et al. 2011. beta2-adrenergic receptor activation prevents rodent dopaminergic neurotoxicity by inhibiting microglia via a novel signaling pathway. J Immunol 186(7): 4443-4454.

Rappold PM et al.2014. Drp1 inhibition attenuates neurotoxicity and dopamine release deficits in vivo. Nature Communications. 5:5244 doi: 10.1038/ncomms6244.

Ren Y. et al., 2005. Selectivwe vulnerabity of dopaminergic neurons to microtubule depolymerisation. J. Bio. Chem. 280:434105-12. Reynolds GP, Garrett NJ (1986) Striatal dopamine and homovanillic acid in Huntington's disease. J Neural Transm. 65(2):151-5.

Rojo AI, Innamorato NG, Martin-Moreno AM, De Ceballos ML, Yamamoto M, Cuadrado A. 2010. Nrf2 regulates microglial dynamics and neuroinflammation in experimental Parkinson's disease. Glia 58(5): 588-598.

Ros-Bernal F, Hunot S, Herrero MT, Parnadeau S, Corvol JC, Lu L, et al. 2011. Microglial glucocorticoid receptors play a pivotal role in regulating dopaminergic neurodegeneration in parkinsonism. Proc Natl Acad Sci U S A 108(16): 6632-6637.

Rubio-Perez JM, Morillas-Ruiz JM. 2012. A review: inflammatory process in Alzheimer's disease, role of cytokines. ScientificWorldJournal 2012: 756357.

Salama M, Helmy B, El-Gamal M, Reda A, Ellaithy A, Tantawy D, et al. 2013. Role of L-thyroxin in counteracting rotenone induced neurotoxicity in rats. Environmental toxicology and pharmacology 35(2): 270-277.

Sandström JBroyer AZoia DSchilt CGreggio C1, Fournier MDo KQMonnet-Tschudi FPotential mechanisms of development-dependent adverse effects of the herbicide paraquat in 3D rat brain cell cultures.Neurotoxicology. 2017 May;60:116-124. doi: 10.1016/j.neuro.2017.04.010. Epub 2017 Apr 30.

Saravanan KS, Sindhu KM, Senthilkumar KS, Mohanakumar KP. 2006. L-deprenyl protects against rotenone-induced, oxidative stress-mediated dopaminergic neurodegeneration in rats. Neurochem Int.49(1):28-40.

Sathe K, Maetzler W, Lang JD, Mounsey RB, Fleckenstein C, Martin HL, et al. 2012. S100B is increased in Parkinson's disease and ablation protects against MPTP-induced toxicity through the RAGE and TNF-alpha pathway. Brain 135(Pt 11): 3336-3347.

Schapira AH, Cooper JM, Dexter D, Jenner P, Clark JB, and Marsden CD. 1989. Mitochondrial complex I de. ciency in Parkinson’s disease. Lancet. 1,1269.

Scott R, Gregory R, Hines N, Carroll C, Hyman N, Papanasstasiou V, Leather C, Rowe J, Silburn P, Aziz T. 1998. Neuropsychological, neurological and functional outcome following pallidotomy for Parkinson's disease. A consecutive series of eight simultaneous bilateral and twelve unilateral procedures. Brain. 121 ( Pt 4):659-75.

Sherer TB, Betarbet R, Stout AK, Lund S, Baptista M, Panov AV, Cookson MR, Greenamyre JT. 2002. An in vitro model of Parkinson's disease: linking mitochondrial impairment to altered alpha-synuclein metabolism and oxidative damage. J Neurosci. 22(16):7006-15.

Sherer TB, Betarbet R, Testa CM, Seo BB, Richardson JR, Kim JH, et al. 2003. Mechanism of toxicity in rotenone models of Parkinson’s disease. J Neurosci. 23:10756–64.

Sherer TB, Richardson JR, Testa CM, Seo BB, Panov AV, Yagi T, Matsuno-Yagi A, Miller GW, Greenamyre JT (2007) Mechanism of toxicity of pesticides acting at complex I: relevance to environmental etiologies of Parkinson's disease. J Neurochem. 100(6):1469-79.

Shulman JM, DeJager PL, Feany MB. 2011. Parkinson’s disease: Genetics and Pathogenesis. Annu.Rev.Pathol.Mech.Dis. 6:193-2

Shults CW. 2004. Mitochondrial dysfunction and possible treatments in Parkinson’s disease–a review. Mitochondrion 4:641– 648.

Singer TP, Ramsay RR.The reaction sites of rotenone and ubiquinone with mitochondrial NADH dehydrogenase. Biochim Biophys Acta. 1994 Aug 30;1187(2):198-202.

Spencer DD, Robbins RJ, Naftolin F, Marek KL, Vollmer T, Leranth C, Roth RH, Price LH, Gjedde A, Bunney BS. 1992. Unilateral transplantation of human fetal mesencephalic tissue into the caudate nucleus of patients with Parkinson's disease. N Engl J Med. 1992 Nov 26;327(22):1541-8

Sriram K, Matheson JM, Benkovic SA, Miller DB, Luster MI, O'Callaghan JP. 2002. Mice deficient in TNF receptors are protected against dopaminergic neurotoxicity: implications for Parkinson's disease. Faseb J 16(11): 1474-1476.

Sulzer D, Surmeier DJ. 2013. Neuronal vulnerability, pathogenesis, and Parkinson’s disease. Movement Disorders. 28 (6) 715-24.

Surmeier DJ1, Guzman JN, Sanchez-Padilla J, Goldberg JA. 2010. What causes the death of dopaminergic neurons in Parkinson's disease? Prog Brain Res. 2010;183:59-77. doi: 10.1016/S0079-6123(10)83004

Silva MA, Mattern C, Häcker R, Tomaz C, Huston JP, Schwarting RK. 1997. Increased neostriatal dopamine activity after intraperitoneal or intranasal administration of L-DOPA: on the role of benserazide pretreatment. Synapse. 27(4):294-302.

Swerdlow RH, Parks JK, Miller SW, Tuttle JB, Trimmer PA, Sheehan JP, Bennett JP Jr, Davis RE, Parker WD Jr (1996) Origin and functional consequences of the complex I defect in Parkinson's disease. Ann Neurol. 40(4):663-71.

Taetzsch T, Block ML. 2013. Pesticides, microglial NOX2, and Parkinson's disease. J Biochem Mol Toxicol 27(2): 137-149.

Tansey MG, Goldberg MS. 2009. Neuroinflammation in Parkinson's disease: Its role in neuronal death and implications for therapeutic intervention. Neurobiol Dis.

Thomas B., Banerjee R.,Starkova NN., Zhang S., Calingasan NY, Yang L., Wille E., Lorenzo B., Ho D., Beal M., Starkov A. 2012. Mitochondrial permeability transition pore component cyclophilin D distinguishes nigrostriatal dopaminergic death paradigms in the MPTP mouse model of Parkinson's disease. Antioxidants & redox signaling 16 (9) 855-68

Thundyil J, Lim KL. 2014. DAMPs and Neurodegeneration. Ageing research reviews.

Tieu Kim, Imm Jennifer. 2014. Mitochondrial dynamics as potential therapeutic target for Parkinson’s disease? ACNR 14 (1) 6-8.

Tseng YT, Chang FR, Lo YC2. 2014. The Chinese herbal formula Liuwei dihuang protects dopaminergic neurons against Parkinson's toxin through enhancing antioxidative defense and preventing apoptotic death. Phytomedicine. 21(5):724-33.

Uitti RJ, Ahlskog JE. 1996. Comparative Review of Dopamine Receptor Agonists in Parkinson's Disease. C NS Drugs. 5(5):369-88.

Uitti RJ, Wharen RE Jr, Turk MF, Lucas JA, Finton MJ, Graff-Radford NR, Boylan KB, Goerss SJ, Kall BA, Adler CH, Caviness JN, Atkinson EJ. 1997. Unilateral pallidotomy for Parkinson's disease: comparison of outcome in younger versus elderly patients. Neurology. 49(4):1072-7.

Vivekanantham S, Shah S, Dewji R,Dewji A, Khatri C & Ologunde R.2014. Neuroinflammation in Parkinson's disease: role in neurodegeneration and tissue repair. International Journal Of NeuroscienceAccepted. Author Version Posted Online.

Walter BL, Vitek JL. 2004. Surgical treatment for Parkinson's disease. Lancet Neurol. 3(12):719-28.

Widner H, Tetrud J, Rehncrona S, Snow B, Brundin P, Gustavii B, Björklund A, Lindvall O, Langston JW. 1992. Bilateral fetal mesencephalic grafting in two patients with parkinsonism induced by 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP). N Engl J Med. 26;327(22):1556-63.

walker DG, Lue LF, Serrano G, Adler CH, Caviness JN, Sue LI, Beach T.2016. altered expression patterns of inflammation-associated and trophic molecules in substantia nigra and striatum brain samples from Parkinson's disease, incidental Lewy Body disease and normal control cases. Frontiers in Neuroscience. 9:507.

Wang Q, Chu CH, Oyarzabal E, Jiang L, Chen SH, Wilson B, et al. 2014. Subpicomolar diphenyleneiodonium inhibits microglial NADPH oxidase with high specificity and shows great potential as a therapeutic agent for neurodegenerative diseases. Glia 62(12): 2034-2043.

Wang T, Zhang W, Pei Z, Block M, Wilson B, Reece JM, et al. 2006. Reactive microgliosis participates in MPP+-induced dopaminergic neurodegeneration: role of 67 kDa laminin receptor. Faseb J 20(7): 906-915.

Wen Y, Li W, Poteet EC, Xie L, Tan C, Yan LJ, Ju X, Liu R, Qian H, Marvin MA, Goldberg MS, She H, Mao Z, Simpkins JW, Yang SH (2011) Alternative mitochondrial electron transfer as a novel strategy for neuroprotection. J Biol Chem. 286(18):16504-15.

Wu XF, Block ML, Zhang W, Qin L, Wilson B, Zhang WQ, et al. 2005. The role of microglia in paraquat-induced dopaminergic neurotoxicity. Antioxidants & redox signaling 7(5-6): 654-661.

Wu RM, Mohanakumar KP, Murphy DL, Chiueh CC. 1994. Antioxidant mechanism and protection of nigral neurons against MPP+ toxicity by deprenyl (selegiline). Ann N Y Acad Sci. 17;738:214-21.

Yadav S, Gupta SP, Srivastava G, Srivastava PK, Singh MP. 2012. Role of secondary mediators in caffeine-mediated neuroprotection in maneb- and paraquat-induced Parkinson's disease phenotype in the mouse. Neurochem Res 37(4): 875-884.

Zaltieri M, Longhena F, Pizzi M, Missale C, Spano P, Bellucci A. 2015. Mitochondrial Dysfunction and α-Synuclein Synaptic Pathology in Parkinson's Disease: Who's on First? Parkinsons Dis. 2015:108029.

Zhou F, Wu JY, Sun XL, Yao HH, Ding JH, Hu G. 2007. Iptakalim alleviates rotenone-induced degeneration of dopaminergic neurons through inhibiting microglia-mediated neuroinflammation. Neuropsychopharmacology 32(12): 2570-2580.

Zhu JH, Horbinski C, Guo F, Watkins S, Uchiyama Y, Chu CT. 2007.Regulation of autophagy by extracellular signal-regulated protein kinases during 1-methyl-4-phenylpyridinium-induced cell death. Am. J. Pathol. 170:75–86.