Skip to main content

ORIGINAL RESEARCH article

Front. Neuroanat., 20 October 2010
Volume 4 - 2010 | https://doi.org/10.3389/fnana.2010.00140

Vulnerability of mesostriatal dopaminergic neurons in Parkinson’s disease

  • 1 Department of Anatomy, Faculty of Medicine, University of La Laguna, Tenerife, Spain
  • 2 Spanish Network in Neurodegenerative Diseases (CIBERNED), Madrid, Spain

The term vulnerability was first associated with the midbrain dopaminergic neurons 85 years ago, before they were identified as monoaminergic neurons, when Foix and Nicolesco (1925) reported the loss of neuromelanin containing neurons in the midbrain of patients with post-encephalitic Parkinson’s disease (PD). A few years later, Hassler (1938) showed that degeneration is more intense in the ventral tier of the substantia nigra compacta than in its dorsal tier and the ventral tegmental area (VTA), outlining the concept of differential vulnerability of midbrain dopaminergic (DA-) neurons. Nowadays, we know that other neuronal groups degenerate in PD, but the massive loss of nigral DA-cells is its pathological hallmark, having a pivotal position in the pathophysiology of the disease as it is responsible for the motor symptoms. Data from humans as well as cellular and animal models indicate that DA-cell degeneration is a complex process, probably precipitated by the convergence of different risk factors, mediated by oxidative stress, and involving pathogenic factors arising within the DA-neuron (intrinsic factors), and from its environment and distant interconnected brain regions (extrinsic factors). In light of current data, intrinsic factors seem to be preferentially involved in the first steps of the degenerative process, and extrinsic factors in its progression. A controversial issue is the relative weight of the impairment of common cell functions, such as energy metabolism and proteostasis, and specific dopaminergic functions, such as pacemaking activity and DA handling, in the pathogenesis of DA-cell degeneration. Here we will review the current knowledge about the relevance of these factors at the beginning and during the progression of PD, and in the differential vulnerability of midbrain DA-cells.

Introduction

The level of interest that mesostriatal dopaminergic (DA-) neurons have attained within the brain motor circuits in the last 30 years is mostly because of the relevance of this cell group in the pathophysiology of Parkinson’s disease (PD). PD is a neurodegenerative disorder characterized by neuronal loss and the presence of intraneuronal inclusions known as Lewy bodies (LBs) in different brain centers (Braak et al., 2003), which in its sporadic form affects more than 1% of people over the age of 60 (Nussbaum and Ellis, 2003; de Lau and Breteler, 2006). Besides midbrain DA-cells, this neuronal loss affects the locus coeruleus, dorsal nucleus of the vagus, hypothalamus, nucleus basalis of Meynert, olfactory bulb, and the peripheral and enteric systems, although they do not all display the same susceptibility to degeneration and clinical impact (Pearce et al., 1995; Forno, 1996). Post-mortem studies show that at the end-stage of the disease, the cell loss in non-dopaminergic nuclei ranges between 30 and 50%, while in the SN it reaches 80% (Victorof et al., 1996; Zarow et al., 2003; Thannickal et al., 2007), with most DA-cells dying during the preclinical period (Greffard et al., 2006). So, the massive loss of SN DA-cells is considered the hallmark of PD, and responsible for most, if not all, motor symptoms characterizing parkinsonian patients (Hornykiewicz, 2006). Nowadays, basic research is mainly focused on elucidating the cause and cellular mechanisms underlying DA-cell degeneration. Despite these efforts, the etiology and pathogenesis of PD remain unclear. Current data point to it being more of a heterogeneous entity, arising from different causes and pathogenic mechanisms, and converging in a relatively uniform phenotype. A number of studies support the relevance of the impairment of common cellular functions, such as mitochondrial metabolism and protein degradation, and pathogenic mechanisms, such as inflammation and excitotoxicity, while others highlight the role of metabolic and functional aspects specific to DA-cells in the degenerative process. However, the relative weight of each of these factors, and the reason why midbrain DA-cells are particularly susceptible remain unknown. Our knowledge is based on data from different sources, including patients with familial forms of PD, whose clinical and pathological features in many cases differ from those in sporadic PD, and from animal and cellular models based on the manipulation of cellular functions probably involved in the pathogenesis of PD. The aim of this review is to summarize and discuss the current data in favor of and against the relevance of these factors in the sequence of events leading to DA-cell degeneration.

Common Cellular Factors and Pathogenic Mechanisms

Mitochondrial Dysfunction

Mitochondria are organelles where a number of important cellular functions occur, including oxidative energy metabolism, calcium homeostasis, and control of cell death (Kroemer and Blomgren, 2007; Celsi et al., 2009; Galluzzi et al., 2009). Furthermore, deficient mitochondrial metabolism may generate reactive oxygen species with catastrophic consequences for the cell. These facts make mitochondrial dysfunction an attractive candidate for playing a key role in neurodegeneration and aging.

The relationship between mitochondrial dysfunction and DA-cell degeneration was reported for the first time by Schapira et al. (1990) who found reduced activity of mitochondrial complex I (NADH-quinone oxidoereductase) in the substantia nigra of parkinsonian patients. The enzymatic deficiency was later confirmed in the cerebral cortex (Keeney et al., 2006; Parker et al., 2008), and in blood platelets (Krige et al., 1992; Haas et al., 1995; Schapira, 2008), suggesting that a number of patients can undergo a systemic decrease in mitochondrial complex I activity. In addition, cells transfected with mitochondrial DNA from platelets of PD patients show complex I deficiency and other features of PD brains (Swerdlow et al., 1996; Borland et al., 2009).

The involvement of mitochondrial dysfunction was also supported by the finding in the early 80s of an acute and persistent parkinsonism in humans after intravenous injection of an illicit drug which contained 1-methyl-4-phenyl 1,2,3,6-tetrahydropyridine (MPTP) (Langston et al., 1983). MPTP is converted by the monoamine oxidase enzyme of glial cells into its active metabolite MPP+ and taken up by DA-cells through the dopamine transporter (DAT). Once inside the cell, MPP+ can exert its neurotoxic effect by means of different mechanisms, including the inhibition of the mitochondrial complex I (Nicklas et al., 1985; Di Monte et al., 1986a; Mizuno et al., 1988). Taking advantage of this effect, MPTP has been widely used for inducing models of PD (DA-cell degeneration) in mice and monkeys (Przedborski and Jackson-Lewis, 1998; Blum et al., 2001).

The structural similarity between MPP+ and the herbicide paraquat raised the possibility of a relationship between exposure to environmental toxins, impairment of oxidative phosphorylation and PD. This idea was supported by epidemiological studies reporting a positive association between the incidence of PD and exposure to pesticides (Semchuk et al., 1992; Liou et al., 1997; Dhillon et al., 2008), although others could not confirm such a relationship (Firestone et al., 2005). Experimental data show that paraquat, subchronically administrated in mice, induces midbrain DA-cell loss (McCormack et al., 2002) and α-synuclein accumulation (Manning-Bog et al., 2003), but its cytotoxic effect is due to the production of reactive oxygen species at cytosolic level, with potential indirect effects on mitochondria, but not to the direct inhibition of the mitochondrial complex I (Di Monte et al., 1986b; Miller, 2007). Rotenone is another pesticide widely used to induce an animal model of PD (Betarbet et al., 2000; Bove et al., 2005). Its adequacy to reproduce PD features has been a matter of controversy (Cicchetti et al., 2009). Some authors argue a high variability in the lesion pattern, including brain regions unaffected in PD as well as the liver and gastrointestinal tract, suggesting that it acts as a systemic unspecific toxin (Hoglinger et al., 2003; Lapointe et al., 2004). In contrast, others maintain that, by using the appropriate administration schedule, rotenone only produces PD specific brain lesions, including dopaminergic degeneration, alpha-synuclein-positive inclusions, and impairment of the ubiquitin–proteasome system (UPS) (Sherer et al., 2003; Cannon et al., 2009). Because of its high lipophilicy, rotenone can cross all cell membranes and accumulates in cellular organelles. Its main cytotoxic effect comes from the interaction with the ND1 and PSST subunits of the mitochondrial complex I, resulting in the loss of its enzymatic activity (Schuler and Casida, 2001). Interestingly, although the complex I impairment is evident at the systemic level, degeneration only affects nigrostriatal neurons (Betarbet et al., 2000), supporting the idea that a systemic mitochondrial dysfunction together with the special susceptibility of midbrain DA-cells may be the origin of DA-cell degeneration in sporadic PD.

Convincing evidence for the involvement of mitochondrial dysfunction in the pathogenesis of PD also arises from the discovery of mutations in genes which encode the Parkin, PINK1, DJ-1, α-synuclein, and LRRK2 proteins.

Mutations in the Parkin gene (PARK2) were first associated with PD in a Japanese family (Kitada et al., 1998). Today, they are recognized as the most frequent cause of juvenile-onset parkinsonism, although later-onset cases have also been reported (Lohmann et al., 2003). Parkin is a ubiquitously expressed protein, present in different subcellular fractions (Shimura et al., 1999), with several domains for protein–protein interactions and E3 ubiquitin protein ligase activity (Zhang et al., 2000). Parkin-associated endothelin-receptor (Pael-R) is one of the substrates of parkin ligase activity which has been shown to be accumulated in the brain of patients with PARK2 mutations. When this receptor is overexpressed in cells it tends to become unfolded and insoluble, and induces cell death (Imai et al., 2001). Parkin has also been recognized as a mitochondrial protection factor. Its overexpression in cells prevents the mitochondrial swelling induced by ceramide (Darios et al., 2003), upregulates the expression of complex I subunits and reduces the accumulation of reactive oxygen species in mitochondria (Kuroda et al., 2006). While its deficiency leads to a decrease in the levels of complex I and IV subunits and mitochondrial respiratory capacity in mice (Palacino et al., 2004).

PTEN-induced putative kinase 1 (PINK1) is a serine/threonine kinase with autophosphorylation capacity which is transcribed in the nucleus, and imported to the mitochondria (Silvestri et al., 2005). Its mutations are associated with hereditary early-onset PD, although heterozygous mutations have also been found in sporadic early-onset forms (Valente et al., 2004). PINK1 regulates the activity and phosphorylation of TRAP1 (TNF-receptor-associated protein 1), a protein which acts as a chaperone to prevent protein misfolding (Pridgeon et al., 2007), HtrA2/Omi, which acts as a protease involved in the degradation of misfolded proteins (Alnemri, 2007), and Parkin. PINK1 regulates the mitochondrial localization of Parkin through its phosphorylation (Kim et al., 2008). Neurons from PINK1 knockout mice develop mitochondrial dysfunction, including reduced membrane potential, increased production of oxygen radicals and sensitivity to apoptosis (Wood-Kaczmar et al., 2008). Furthermore, the striatum of PINK1 knockout mice shows an increase in the number of large mitochondria and impairment of complex I and II activity (Gautier et al., 2008). So, mitochondrial dysfunction induced by PINK1 mutations may in part be due to defects in Parkin functions.

DJ-1 is a 189 amino acid/20 kDa multifunctional protein with antioxidant and transcription modulatory activity (Takahashi et al., 2001; Taira et al., 2004). Under normal conditions, DJ-1 preferentially localizes in the cytosol, but under conditions of oxidative stress it translocates to the mitochondria and nucleus (Zhang et al., 2005). This redistribution ability has been related to its neuroprotective capacity (Canet-Aviles et al., 2004; Ashley et al., 2009). Mutations in its encoding gene (PARK7) are a rare cause of recessively inherited PD (Bonifati et al., 2003). DJ-1 deficient mice show motor abnormalities and changes in DA handing, but not spontaneous DA-cell degeneration (Goldberg et al., 2005). However, they become more susceptible to MPTP than their wild-type congeners (Kim et al., 2005).

Alpha-synuclein is a presynaptic protein whose relevance in PD comes from the fact that it is present in LBs, and because mutations and triplications of its gene locus (PARK1) are a cause of autosomal dominant PD (Polymeropoulos et al., 1997; Singleton et al., 2003). The mechanism by which α-synuclein exerts its toxicity has been related to the tendency of the mutated and over-expressed protein to aggregate in form of fibrils (Conway et al., 2000; Fredenburg et al., 2007). Oligomeric intermediates of the α-synuclein fibrillation pathway, called protofibrils, can form annular pores in cell membranes, including those of mitochondria, leading to their permeabilization. This phenomenon may be enhanced in DA-cells because cytosolic DA interacts with α-synuclein slowing the conversion of protofibrils to fibrils (Conway et al., 2000; Rochet et al., 2004). Overexpression of mutant and wild-type α-synuclein can induce damage at different cell levels, including proteasome, endoplasmic reticulum, and mitochondria (Hsu et al., 2000). Mitochondrial impairment includes oxidation of mitochondria-associated proteins (Poon et al., 2005), mitochondrial DNA damage and reduced complex IV activity (Martin et al., 2006). In this respect, higher α-synuclein levels have been found in mitochondria from substantia nigra of PD patients than in those from healthy individuals (Devi et al., 2008).

Leucine-rich repeat kinase 2 (LRRK2) is a widely distributed protein with a multidomain structure which includes a serine/threonine kinase domain (Giasson et al., 2006). Mutations in the gene encoding LRRK2 (PARK8) are a common cause of autosomal dominant PD, and are also present in about 1% of patients with sporadic PD (Paisan-Ruiz et al., 2004; Di Fonzo et al., 2005). The toxicity of several of these mutations appears to be associated with an increase in its kinase activity (Gloeckner et al., 2006). Although the involvement of mitochondria in the pathogenesis of LRRK2 mutant-associated PD remains unclear, the facts that LRRK2 partially colocalizes with mitochondrial markers (Biskup et al., 2006), and that overexpression of mutant LRRK2 proteins induces Apaf1 dependent apoptosis (Iaccarino et al., 2007), suggest that LRRK2 mutations can compromise mitochondrial function.

In sum, a large body of evidence supports the idea that mitochondrial dysfunction plays a key role in the pathogenesis of DA-cell degeneration. However, this idea is challenged by certain facts. For example, diseases associated with mitochondrial DNA mutations are typically multisystemic, involving heart, skeletal muscle, and brain, but not specifically DA-cells. Furthermore, DA-cell degeneration is not relevant in most neurodegenerative diseases associated with mutations in the nuclear genes which encode proteins that are targeted to mitochondria (Schon and Manfredi, 2003). Thus, although mitochondria impairment may be involved in the degenerative process in many cases of PD, other factors must contribute to making DA-cells particularly vulnerable in this disorder.

Protein Homeostasis

Cells control the quantity and quality of intracellular proteins through the balance between the rate of their synthesis, maturation and degradation (Shah and Di Napoli, 2007). Folding is part of this process converting newly synthesized proteins into functional molecules. The endoplasmic reticulum contains most of the machinery responsible for maintaining an appropriate protein folding through a coordinated signaling program known as unfolded protein response (UPR) (Hosoi and Ozawa, 2009). The response of the endoplasmic reticulum includes translational attenuation, induction of chaperones, and degradation of misfolded proteins. The two principal pathways of intracellular protein degradation are the UPS, responsible for the bulk of the turnover of short-lived cytosolic proteins, and the autophagy–lysosome pathway (ALP), involved in the degradation of long-lived stable proteins as well as in the recycling of organelles (Jellinger, 2009). The capacity of this proteostasis network is enough to maintain protein homeostasis under normal conditions and under short-term stressful circumstances. However, this seems to be overwhelmed by chronic stress as probably occurs in aging, cancer, or metabolic and neurodegenerative diseases. In these cases, proteins lose their normal configuration and activity, and become misfolded with a propensity to aggregate. The structural instability of some mutated or postranslationally modified proteins makes them particularly inclined to aggregate to each other as well as to other normal proteins. These protein aggregates may interfere with critical intracellular processes including UPR function, and induce cytotoxicity.

A number of findings indicate that these cellular functions are compromised in DA-cells in PD. For example, proteinaceous aggregations, containing ubiquitin, and α-synuclein and known as Lewy bodies (LBs), are abundant in the SN and other brain centers affected by degeneration in sporadic PD (Braak et al., 2003). The enzymatic activity and expression of some α-subunits of the proteasome and the proteasomal activator PA700 have been found to be low in the SN of PD patients, while in unaffected brain regions they are high, suggesting that neurons in these regions, but not those in the SN, can activate compensatory responses against the proteolytic stress (McNaught et al., 2001, 2003). Even in healthy subjects, the levels of the proteasome activator PA28 in the SN pars compacta (SNc) have been reported to be lower than in other brain regions (McNaught et al., 2002a, 2003). In addition, a recent immunohistochemistry study showed that under normal conditions, PA28 and PA700 expression are very low in the rat SNc, and that different stressful stimuli induce PA28/PA700 overexpression in a variety of cells but not in midbrain DA-neurons (McNaught et al., 2010). Therefore, it is possible that the UPS is constitutively weak in midbrain DA-cells, contributing to their particular vulnerability in sporadic PD.

The relevance of proteolytic stress is also supported by data arising from familial PD. As mentioned above, although the mechanism by which Parkin mutations induce degeneration is not well known, it has been mainly related to a defect in its ubiquitin ligase activity. Parkin interacts with the 26S proteosome through its ubiquitin ligase domain facilitating the recognition of ubiquitinated substrates by the proteasomal activator PA700 (Pickart, 2004). Parkin acts in conjunction with other enzymes to ubiquitinate a variety of substrates, such as Pael-R, which have been found to be accumulated and non-ubiquinated in LBs (Murakami et al., 2004). In addition, Parkin prevents cell death induced by overexpression of Pael-R (Imai et al., 2001). Therefore, Parkin mutations could impair ubiquitination and degradation of substrate proteins which might be aggregated and cause cytotoxicity.

Ubiquitin C-terminal hydrolase L1 (UCH-L1) is a 230 amino acid/26 kDa protein only expressed in neurons, which acts as a deubiquinating enzyme removing ubiquitin from protein adducts before their entry in the proteosome (Leroy et al., 1998). A missense mutation in the UCH-L1 gene (PARK5) has been found in two German siblings with PD, and is considered today to be a rare cause of familiar PD (Olanow and McNaught, 2006). It has been reported that the in vitro deubiquinating activity of mutated UCH-L1 form is low (Nishikawa et al., 2003), and that inhibition of the ubiquitin hydrolase activity leads to the formation of LB-like inclusions and DA-cell degeneration in cell cultures (McNaught et al., 2002b). So, it is possible that, in these patients, deubiquitination failure blocks the proteasomal processing of proteins, leads to the formation of aggregates and interferes with the supply of ubiquitin monomers to additional unwanted proteins (Olanow and McNaught, 2006).

The protective role of DJ-1 has been mostly associated to its ability to translocate to the mitochondria and nucleus where it can exert antioxidant and transcriptional modulatory actions (Zhang et al., 2005). However, its molecular structure and in vitro properties indicate that it can also work as a molecular chaperone and protease (Abou-Sleiman et al., 2003; Lee et al., 2003). In addition, such proteolytic activity is lost when it is mutated (Olzmann et al., 2004; Shendelman et al., 2004). So, mutations in DJ-1 can compromise UPS function and promote protein aggregation.

Under physiological conditions, α-synuclein is monomeric and unfolded, and preferentially degraded by the proteasome in an ubiquitin-independent manner (Bennett et al., 1999). But in high concentration or as a mutated form, it resists proteasomal degradation (Stefanis et al., 2001; Tanaka et al., 2001), becomes misfolded and oligomeric, and tends to self-aggregate and aggregate to other proteins (Bennett, 2005). As mentioned above, the toxicity of α-synuclein has been related to the ability of α-synuclein intermediate aggregates to form ring-like pores in cellular membranes, increasing their permeability (Lashuel et al., 2002). In vitro studies reveal that α-synuclein aggregates can also bind to the proteasomal protein S6’, a subunit of the 19S cap of the 26S proteasome, resulting in the inhibition of its function (Snyder et al., 2003).

On the other hand, similar to other protein complexes and membrane proteins too large in size to pass through the proteasome barrel, α-synuclein oligomers and aggregates can also be cleared by the ALP (Pan et al., 2008). The principal pathways for α-synuclein clearance through lysosomes are the macroautophagy, which is also involved in the recycling of damaged organelles, and the chaperone-mediated autophagy (CMA) (Cuervo et al., 2004). In CMA, α-synuclein couples to the chaperone Hsc70 forming a protein–molecular chaperone complex that binds to the lysosomal membrane (Mak et al., 2010). The A30P and A53T α-synuclein mutants, which are associated with familial PD, remain firmly attached to the chaperone, blocking their lysosomal translocation and degradation, and the binding of other CMA protein substrates (Cuervo et al., 2004; Pan et al., 2008). In addition, missense mutations of the gene encoding lysosomal ATPase 13A2 (PARK9), which are associated with an atypical form of juvenile-onset PD, lead to deficient autophagy activity with α-synuclein aggregation (Ramirez et al., 2006).

In spite of the unequivocal evidence for the impairment of proteostasis in the pathogenesis of DA-cell degeneration, it is still not clear whether degeneration starts with this phenomenon, whether the proteolytic stress does or does not play a determinant role in the relative specificity of DA-cell loss in PD, and whether LB-like proteinaceous inclusions are toxic entities promoting neuron death or whether they represent protective mechanisms by which neurons sequester unwanted proteins (Harrower et al., 2005). Based on the regional distribution of LBs, Braak et al. (2003) proposed a neuropathological staging scheme according to which PD starts in non-dopaminergic structures of the medulla oblongata and progresses rostrally, affecting nigral DA-cells later on in the course of the disease. However, this idea contrasts with quantitative studies reporting that most SN cell loss and striatal denervation occurs during the 5–10 year preclinical period (Gibb and Lees, 1988; Greffard et al., 2006), suggesting that cell loss and α-synuclein deposition are not parallel processes (Burke et al., 2008). It should also be noted that LBs and α-synuclein immunoreactive inclusions are present in different brain regions in different neurodegenerative diseases encompassed by the term LB disorders, including PD with dementia and dementia with LBs (Lippa et al., 2007), and in normal aging (Hindle, 2010); that the density of LBs in neurodegenerative diseases does not correlate with the intensity of cell degeneration, and that LBs are virtually absent in juvenile-onset PD associated with Parkin mutations (Olanow and McNaught, 2006).

Neuroinflammation

The participation of inflammatory reaction in the degeneration of midbrain DA-cells was suggested for the first time by the report of an increased number of microglial cells immunoreactive for the major histocompatibility complex (MHC) antigen II in the SN of PD patients (McGeer et al., 1988). The presence of activated microglia was thereafter confirmed using other markers of microglial activation (Mogi et al., 1994a; Banati et al., 1998; Imamura et al., 2003). The fact that DA-cells express the receptors for some of the cytokines found in the SN of PD patients, such as the tumor necrosis factor (TNF) α, interleukin (IL)-1β and interferon γ (Boka et al., 1994), and that the expression of some of these, e.g., TNFα receptor 1, increases in PD (Mogi et al., 2000), indicates that DA-cells may be sensitive to the direct effect of cytokines. However, cytokines can also affect DA-cells by inducing enzymes such as nitric oxide synthase, cyclo-oxygenase 2, and NADPH-oxidase which produce toxic reactive species (Hunot et al., 1996; Knott et al., 2000).

Interestingly, an analysis of cerebrospinal fluid in PD patients showed an increase in TNFα, IL-1β, and IL-6 concentration (Mogi et al., 1994b; Blum-Degen et al., 1995), and an analysis of peripheral blood showed an increase in activated CD4(+) CD25(+) lymphocytes (Bas et al., 2001). Furthermore, the population of circulating CD3+ CD4 bright+ CD8 dull+ lymphocytes is significantly greater in PD patients than in age-matched control subjects (Hisanaga et al., 2001). Since the count of these lymphocytes usually increases after a viral infection, the authors proposed that postinfectious immune abnormalities could be associated with the pathogenesis of PD.

Animal and cellular models of PD have reproduced most cellular and molecular inflammatory events observed in the brain of PD patients, but have also provided additional insight for understanding the role of inflammation in DA-cell degeneration. For example, microglial activation has been found in the SN of mice and monkeys after peripheral injection of MPTP (Liberatore et al., 1999; Hurley et al., 2003; McGeer et al., 2003). The addition of α-synuclein or neuromelanin, two products released by degenerating DA-cells, to microglial cultures induces activation of proinflammatory pathways, and the injection of α-synuclein protofibrils in the rat SN induces microglial activation and loss of adjacent neurons (Wilms et al., 2003, 2009). Furthermore, the administration of steroidal and non-steroidal anti-inflammatory drugs reduce the toxicity of 6-OHDA and MPTP in rodents (Kurkowska-Jastrzebska et al., 2002; Di Matteo et al., 2006; Hirsch and Hunot, 2009). Taken together, these findings suggest that damaged DA-cells might upregulate and/or release products which activate microglial cells, and these lead to a local neuroinflammatory reaction which helps to perpetuate degeneration. Interestingly, recent in vivo and in vitro studies show that DA-degeneration induces an increase in the activity of angiotensin converting enzyme and angiotensin II (AII) levels, that the addition of AII to midbrain cell cultures increases 6-OHDA toxicity, and that angiotensin type-1 receptor antagonists inhibit both DA-cell degeneration and NADPH-oxidase and early microglial activation (Muñoz et al., 2006; Rodriguez-Pallares et al., 2008; Joglar et al., 2009). So, the brain angiotensin system may play a key role in the inflammatory process in PD. The fact that inflammatory changes are also present in other neurodegenerative disorders, such as Alzheimer’s disease, Huntington’s disease and amyotrophic lateral sclerosis (McGeer and McGeer, 2004), suggests a common phenomenon in the progression of neuronal degeneration.

Glutamate Excitotoxicity

Glutamate is the major excitatory neurotransmitter in the mammalian brain, and is also a potent neurotoxin when it reaches high extracellular concentrations. The levels of extracellular glutamate are regulated by neuronal and glial transporters, and by its metabolism and recycling via glutamine synthetase in astrocytes, which prevent the extracellular glutamate concentration from rising to neurotoxic values (Plaitakis and Shashidharan, 2000; Dervan et al., 2004). Glutamatergic actions are exerted through two families of glutamate receptors, ionotropic and metabotropic. According to their agonist selectivity, ionotropic receptors have been classified into: N-methyl-d-aspartate (NMDA), α-amino-3-hydroxyl-5-methyl-4-isoxazole-propionate (AMPA), and kainic acid (KA) receptors (Burnashev et al., 1992; Hollmann and Heinemann, 1994; Geiger et al., 1995). Metabotropic receptors are coupled by G-proteins to different secondary messenger systems, and have also been classified into different groups (Groups I–III; Conn and Pin, 1997). Extracellular glutamate comes from two different sources, the neuronal Ca2+-dependent pool, and the glial Ca2+-independent pool. Under basal conditions, the glial pool provides 60% of extracellular glutamate, although at present its functional relevance is controversial. After K+-induced depolarization, the neuronal pool augments from 30% to 60% of the total extracellular glutamate (Xue et al., 1996; Baker et al., 2002; Cavelier and Attwell, 2005). Current data show that the glutamatergic input to the SN comes from the pedunculopontine tegmental nucleus (PPN), and particularly from the subthalamic nucleus (STN) (Charara et al., 1996; Iribe et al., 1999; Forster and Blaha, 2003). Curiously, hodological and electron microscopic studies indicate that the bulk of subthalamic afferents to the SN reach the pars reticulata and form asymmetrical synapses with TH-negative (presumably GABAergic) dendrites, with those making contact with TH-positive being more than 10% (Kita and Kitai, 1987). So, depending on the target cell and the receptor type, glutamate can directly excite DA-cells, indirectly inhibit them by exciting GABAergic neurons, and directly inhibit or excite them by activating metabotropic receptors on DA-cells (Lee and Tepper, 2009).

Glutamate signaling may be affected by alterations in its homeostasis, which leads to an increase in the extracellular concentration, or by alterations in the energy metabolism of its target cells, which lead to a decrease in the activation threshold for their receptors, particularly those of the NMDA type (Novelli et al., 1988). Experimental data indicate that glutamate might contribute to DA-cell degeneration by means of both mechanisms. In a model of PD in mice, the loss of striatal DA-terminals is accompanied by an increase in the size and number of astrocytes, but the density of transporters per astrocyte is low, suggesting that the astrocyte function is compromised (Dervan et al., 2004). On the other hand, NMDA antagonists reduce the loss of DA-neurons due to the inhibition of the cellular metabolism in midbrain cell culture (Zeevalk et al., 1998), and the loss of DA-cells induced by MPP+ and MPTP in rodents and in non-human primates (Turski and Lachowicz, 1991). STN activity has been found to be enhanced in both the experimental models of PD (Bergman et al., 1994) and in the PD patients (Benazzouz et al., 2002), and its modulation results in an improvement of motor symptoms (Limousin et al., 1998; Garcia et al., 2005). Hyperactivity of STN has even been detected before the appearance of clinical signs in MPTP-treated monkeys, suggesting that it may be a compensatory mechanism for the maintenance of striatal dopaminergic homeostasis (Bezard et al., 1999a), but it could also contribute to the progression of DA-cell loss from the first stages of the disease.

Aging

The relationship between aging and DA-cell degeneration is based on several facts. PD in its sporadic form is an age-related degenerative disorder, with a mean onset age of 55, and the incidence increases from 20/100.000 to 120/100.000 at age 70 (Dauer and Przedborski, 2003). Lewy bodies, the pathological hallmarks of PD (Braak et al., 2003), are also present in normal aged brains (Saito et al., 2004; Hindle, 2010). Oxidative stress is currently considered to play a central role in both aging (Harman, 1956; Squier, 2001) and PD (Dauer and Przedborski, 2003; Jenner, 2003). Mitochondrial DNA deletions occurring during normal aging, probably as a consequence of oxidative stress, are particularly abundant in midbrain DA-cells (Bender et al., 2006; Kraytsberg et al., 2006). In addition, morphological studies in monkeys (Pakkenberg et al., 1995; Gerhardt et al., 2002) and humans (McGeer et al., 1977; Fearnley and Lees, 1991; Pakkenberg et al., 1995; Cabello et al., 2002; Gerhardt et al., 2002) have reported a decrease in the number of midbrain DA-cells with age, suggesting that midbrain DA-neurons undergo a low intensity degenerative process during normal aging. Based on these findings, it has been proposed that everyone should get PD if they live long enough (Chan et al., 2007; Surmeier, 2007). However, this idea is challenged by other studies performed on monkeys (Irwin et al., 1994; McCormack et al., 2004; Collier et al., 2007) and humans (Kubis et al., 2000) according to which the number of midbrain DA-cells does not change during normal aging, suggesting that the age-related loss of striatal DA levels (Kish et al., 1992; Haycock et al., 2003) is due to a progressive chemical and functional decline rather than to degeneration. Furthermore, biochemical studies performed on humans show that the regional pattern of striatal DA loss during normal aging substantially differs from that typically observed in PD (Kish et al., 1992). Consistent with these data, we have recently observed that the loss of DA in aged rats is higher in the ventral striatum than in the dorsal striatum (Cruz-Muros et al., 2007). In addition, two different age-related processes were detected. One is characterized by DOPA decarboxylase decrease involving both the nigrostriatal and the mesolimbic compartments, and is responsible for the moderate DA decrease in the dorsal striatum. The other process is characterized by axonal degeneration and aggregation of α-synuclein in their original somata, restricted to mesolimbic regions and responsible for the decline of tyrosine hydroxylase activity and the greater decrease in DA levels in this compartment. Therefore, although both the nigrostriatal and the mesolimbic systems are vulnerable to aging, in contrast to what occurs in PD, the mesolimbic system seems to be more vulnerable to normal aging than the nigrostriatal system. Furthermore, although several age-related aspects, such as oxidative stress or impairment of the energy metabolism, make aging the most important risk factor for PD, anatomical and biochemical features of normal aging in the mesostriatal system seem to be more probably related to age-associated behavioral and mood disturbances than to parkinsonian motor signs.

Specific Dopaminergic Factors

The aforementioned studies indicate that intrinsic (mitochondrial dysfunction and proteostasis impairment), extrinsic (neuroinflammation and excitotoxicity) and age-related factors participate in the degenerative process of DA-cells in PD. However, a question of critical importance remains to be solved: why are midbrain DA-neurons, and especially some of these neurons, particularly vulnerable to degeneration? It is known that in PD, DA-cells in the SN are more vulnerable than those in the VTA, and that within the SN, neurons lying in its ventrolateral and caudal region (SNcv) are more vulnerable than those in the rostromedial and dorsal region (SNrm) (Hassler, 1938; German et al., 1989; Damier et al., 1999). However, there is no evidence for substantial differences in the ER, mitochondria or other structural organelles between DA-cells and non-DA cells, or between midbrain DA-cells that show different susceptibility to degeneration. So, differential vulnerability should be linked to physiological, neurochemical, and/or metabolic aspects which are specific to DA-cells and also differ from one DA-subpopulation to another, making some of them more vulnerable or resistant.

Physiological Profile of DA-Neurons

A characteristic of midbrain DA-cells is that, unlike most neurons in the brain, they display two electrophysiological patterns. One phasic pattern in the form of bursts of action potentials with a mean frequency of 20 Hz and triggered by excitatory inputs from different sources (Schultz, 2002; Omelchenko and Sesack, 2005), and the other tonic pattern, with a mean frequency of 2–4 Hz and which is generated in absence of afferents (Sulzer and Schmitz, 2007). This intrinsic pacemaking activity is believed to be important in maintaining basal DA levels in the striatum, and may be detected from the first postnatal days in mice. During the first weeks, this is driven by sodium and hyperpolarization-activated and cyclic nucleotide-gated cation (Na+/HCN) channels. Thereafter, the expression of L-type Ca2+ channels increases in SN DA-cells, and the pacemaking is driven by Ca2+ currents. Adult VTA DA-cells also have L-type Ca2+ channels, but as opposed to SN DA-cells, their pacemaking continues to be driven by sodium channels. The L-type Ca2+ channels used by SN DA-cells have a Cav 1.3 pore-forming subunit which allows them to open at a relatively hyperpolarized potential (Striessnig et al., 2006; Chan et al., 2007). While in most cells, Ca2+ channel opening is a rare and brief event, the sustained used of Cav 1.3 Ca2+ channels by SN DA-cells results in a large Ca2+ influx which requires a high metabolic cost to be handled. Cytosolic Ca2+ should be pumped back across the plasma membrane against a concentration gradient or rapidly sequestered. The two organelles most directly involved in Ca2+ handling are ER and mitochondrion. Cytosolic Ca2+ is moved into ER by high-affinity ATP-dependent transporters. Thereafter, it is released to the cytosol to be used in modulatory functions or pumped to the extracellular space (Berridge, 2002; Choi et al., 2006). Ca2+ released from the ER can also enter mitochondria through apposition points between both organelles (Rizzuto and Pozzan, 2006). This ER-mitochondrial system contributes to Ca2+ buffering but also to ATP production (McCormack et al., 1990). However, it may be overburdened by the sustained use of Cav1.3 Ca2+ channels. Both depletion and increase in intraluminal ER Ca2+ concentration induce ER stress and compromise proteostasis (LaFerla, 2002; Paschen and Mengesdorf, 2005). Furthermore, the high demand of ATP needed to move Ca2 out of the cell increases the electron transport chain activity leading to the production of reactive oxygen species (Sulzer and Schmitz, 2007; Chan et al., 2009). The relevance of the Ca2+-dependence of pacemaking activity in the vulnerability of SN DA-cells is based on the finding that blockade of Cav1.3 Ca2+ channels in adult SN DA-cells induces a reversion of the pacemaking pattern to that of adult VTA and juvenile SN DA-cells, also protecting them against rotenone and MPTP (Chan et al., 2007). Reinforcing the role Ca2+ homeostasis in the differential vulnerability of midbrain DA-cells, DA-cells in the dorsomedial region of SN and VTA express calbindin-D28k (Gerfen et al., 1987; McRitchie and Halliday, 1995; Liang et al., 1996; McRitchie et al., 1996). The ability of this protein to buffer intracellular Ca2+, together with its localization in DA-cell subpopulations resistant to degeneration, have suggested that calbindin-D28k confers neuroprotection (Gerfen et al., 1985; Yamada et al., 1990; German et al., 1992). However, some findings contrast with this hypothesis. For example, no differences have been observed between the neurotoxic effect of MPTP in midbrain-DA cells of calbindin-D28k-deficient mice and their wild-type littermates (Airaksinen et al., 1997). Furthermore, DA-cells in the lateral region of the SN which express calbindin-D28k are more sensitive to the neurotoxic effect of 6-OHDA than other SN DA-cells which do not express calbindin-D28k (Rodriguez et al., 2001).

Metabolic Phenotype of DA-Neurons

Another peculiarity of DA-cells implicated in their vulnerability concerns DA itself. Under normal conditions, most DA acting as a neurotransmitter is contained inside synaptic vesicles of DA-cells (Jonsson, 1971). When DA concentration increases in the cytosol, DA is rapidly metabolized via monoamino oxidase or by autooxidation. Both enzymatic and non-enzymatic catabolisms of DA are important sources of reactive oxygen species which have been implicated in DA-cell degeneration. Through its enzymatic metabolism, DA is converted into 3,4-dihydroxyphenylacetaldehyde and 3,4-dihydroxyphenylacetic acid, also generating hydrogen peroxide. This may be neutralized by antioxidant enzymes, or, in the presence of iron, generate high toxic hydroxyl radicals. Through autooxidation, DA is converted into reactive quinones and superoxide anion. This can be converted into hydrogen peroxide by superoxide dismutase, or react with nitric oxide to generate the highly toxic radical peroxynitrite (Barzilai et al., 2001). Reactive quinones may be oxidized to cyclized aminochromes, which are thereafter polymerized to form neuromelanin, or interact with cysteine residues of different molecules inhibiting their activity (Miyazaki and Asanuma, 2009). On the basis of this interaction, the formation of quinones by DA oxidation has gained interest in the pathogenesis of PD. There are studies supporting the notion that DA-quinone, or some of its metabolic products, can inhibit the ubiquitin ligase activity of parkin (LaVoie et al., 2005), the permeability of mitochondrial pores (Berman and Hasting, 1999) and the proteasomal function (Zafar et al., 2006), as well as induce microglial activation (Kuhn et al., 2006).

Neuromelanin is another product from the DA metabolism that has been implicated in the pathogenesis of PD. Neuromelanin is a complex polymeric molecule which probably arises from both the enzymatic (Hasting, 1995; Sanchez-Ferrer et al., 1995) and non-enzymatic (Fornsted et al., 1986; Sulzer et al., 2000) metabolism of catecholamines, and is particularly abundant in DA-cells in the SN and noradrenergic cells in the locus coeruleus of humans, being responsible of their dark color (Fasano et al., 2006). Neuromelanin was first proposed as playing a role in PD on the basis of morphological studies reporting a correlation between neuromelanin concentration and DA-cell degeneration (Mann and Yates, 1983; Hirsch et al., 1988). However, this idea was challenged by the report that DA-cells in the dorsal tier of the SN contain higher levels of neuromelanin than those in the ventral tier which are more vulnerable to degeneration (Gibb, 1992). Experimental studies performed during the last two decades indicate that neuromelanin may play a dual role, protecting from or contributing to DA-cell degeneration, depending on the environment (Ben-Shachar et al., 1991; Sulzer et al., 2000). On the one hand, neuromelanin can bind MPP+ (D’Amato et al., 1986) and paraquat (Lindquist et al., 1988), reducing their neurotoxicity, and accumulate dopaminergic drugs, regulating their intracellular concentration (Salazar et al., 1978). Neuromelanin also has the capability to sequester redox-active ions, including iron, thereby reducing the formation of free hydroxyl radicals (Enochs et al., 1994; Zecca et al., 2003). However, under oxidative conditions, neuromelanin can interact with hydrogen peroxide, becoming a source of free radicals and releasing toxic metals which accelerate cell death (Swartz et al., 1992; Shima et al., 1997). In addition, as mentioned above (Wilms et al., 2003), neuromelanin released from degenerated DA-cells can induce microglial activation, also perpetuating the degenerative process.

It should be also noted that the cytosolic levels of DA are closely related to two transport processes: DA uptake from the extracellular space, and once inside the cell, DA packing into small synaptic vesicles. DA uptake is performed by the DAT, a 12-transmembrane domain glycoprotein with three N-glycosylation sites in the second extracellular loop (Horn, 1990; Giros and Caron, 1993; Uhl et al., 1994). DA vesicular storage is performed by the vesicular monoamine transporter type 2 (VMAT2), also present in other monoaminergic cells (Miller et al., 1999). DAT may contribute to the vulnerability of DA-cells serving as an entrance for DA and its metabolites, while VMAT2 may serve as a neuroprotective factor by sequestering DA into vesicles, and preventing interaction with their catabolic enzymes. So, the functional balance between both transporters may affect the differential vulnerability between different midbrain DA-cell subpopulations. This idea is supported by the fact that VMAT2 deficient mice are more susceptible to MPTP than their homologous wild-type (Takahashi et al., 1997). Furthermore, they display spontaneous and progressive DA-cell degeneration (Caudle et al., 2007), and non-motor parkinsonian symptoms (Taylor et al., 2009), suggesting that VMAT2 may also be involved in the degeneration of other monoaminergic cells in PD. Studies carried out in our laboratory show that during normal aging, in parallel with the decline in vesicular DA uptake, VMAT2 is deglycosylated and moved from the vesicle membrane to the soluble compartment (Cruz-Muros et al., 2008), probably contributing to the increase of the vulnerability of DA-cells with age.

On the other hand, the damaging role of DAT is supported by the fact that its expression and functional integrity are required for making DA-cells susceptible to DA analog neurotoxins (Kopin, 1992; Gainetdinov et al., 1997; Bezard et al., 1999b), and by the report of an anatomical correlation between the distribution of DA-neurons expressing high DAT mRNA levels and those showing high vulnerability to degeneration in PD and animal models of PD (Cerruti et al., 1993; Hurd et al., 1994; Uhl et al., 1994). However, this correlation does not apply in some DA-cell groups. DA-cells in the parabrachialis pigmentosus region of the A10 cell group, which contain higher levels of DATmRNA than those in the dorsal tier of SN in monkeys (Haber et al., 1995) and rats (Shimada et al., 1992) are also more resistant to MPTP (Varastet et al., 1994) and 6-OHDA (Rodriguez et al., 2001) respectively. In addition, a comparative analysis of DAT and VMAT2 mRNA expressions in the ventral midbrain of rats revealed no differences between both mRNA expression patterns, with the highest levels in the SNrm, followed by the SNcv, and the lowest ones in VTA (Gonzalez-Hernandez et al., 2004). So, the involvement of DA transporters in the differential vulnerability of DA-cells could be due to aspects other than differences in their mRNA levels. Given that DAT activity depends on its glycosylation status and membrane expression, we then explored a possible relationship between DAT glycosylation and function and the differential vulnerability of DA-cells (Afonso-Oramas et al., 2009). Glycosylated-DAT expression, DA uptake and DAT Vmax were significantly higher in nigrostriatal neuron terminals than in those of mesolimbic neurons. These findings are consistent with those reporting that the number of 3H-DA uptake sites in the dSt is higher than in the vSt (Missale et al., 1985; Marshall et al., 1990; Cass et al., 1992), and correlate regional differences in DA uptake with the glycosylation status of DAT in both striatal compartments. Differences in DAT regulation may contribute to defining the pattern of DA time signaling and functional specialization in the dSt and vSt (Wickens et al., 2003). Interestingly, the expression pattern of glycosylated-DAT in the human midbrain and striatum showed a close anatomical relationship with DA-degeneration in parkinsonian patients. In addition, this relationship was confirmed in rodent and monkey models of PD, and in HEK-cells expressing the wild-type and a partially deglycosylated DAT form.

Concluding Remarks

Nearly two centuries after PD was first described by James Parkinson, its etiopathogenesis is still unknown. There is a tendency to consider PD as a multicausal disorder resulting from the convergence of environmental-, genetic-, and age-related factors. In recent years, mitochondrial dysfunction and proteostasis impairment have been proposed as having a leading role in the degenerative process. However, this idea contrasts with the selectivity of neuronal populations affected in PD. In spite of the large body of evidence supporting the participation of mitochondria and ER in the pathogenesis of PD, it is difficult to accept that a failure in functions common to all cells is the main factor responsible for degeneration of particular neuronal groups. Ca2+-dependent pacemaking and DA handling are two oxidative stress generating activities specific to midbrain DA-cells. The reliance on Cav1.3 Ca2+ channels to drive pacemaking and the pattern of DAT maturation and DA uptake also explain the differential susceptibility between nigrostriatal and mesolimbic neurons. Morphological and molecular studies performed in the last 20 years, with the aid of laser capture microdissection and more recently microarray analysis, show differences in the expression levels of more than 100 genes between vulnerable (nigrostriatal) and resistant (mesolimbic) DA-cells (Chung et al., 2005; Greene et al., 2005). These genes belong to different categories, including Ca2+ buffering proteins, mitochondrial proteins, L-type Ca2+ and inward rectifier K+ channels, axon guidance and transcription factors, etc. (for review see Gonzalez-Hernandez et al., 2009). But the function and relevance of most of these in the differential vulnerability of DA-cells is still unknown.

Finally, we should not forget that, to a varying degree, other neuronal populations, such as noradrenergic neurons in the locus coeruleus, and cholinergic neurons in the dorsal nucleus of the vagus are also affected in PD, and that Ca2+-dependent pacemaking and DA handling cannot explain their vulnerability. It would be interesting to elucidate whether there is a vulnerability factor common to all cell populations involved in PD, or whether independent factors determine the start of the degenerative process in each neuronal group for their posterior convergence in a common pathogenic route.

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

This work was partially supported by Ministerio de Ciencia e Innovación (grant no. BFU2007-66561) and CIBERNED (ISCIII). Josmar Salas-Hernandez and Javier Castro-Hernandez were supported by Fundación Canaria de Invetigación y Salud (FUNCIS).

References

Abou-Sleiman, P. M., Healy, D. G., Quinn, N., Lees, A. J., and Wood, N. W. (2003). The role of pathogenic DJ-1 mutations in Parkinson’s disease. Ann. Neurol. 54, 283–286.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Afonso-Oramas, D., Cruz-Muros, I., Alvarez de la Rosa, D., Abreu, P., Giraldez, T., Castro-Hernandez, J., Salas-Hernandez, J., Lanciego, J. L., Rodriguez, M., and Gonzalez-Hernandez, T. (2009). Dopamine transporter glycosylation correlates with the vulnerability of midbrain dopaminergic cells in Parkinson’s disease. Neurobiol. Dis. 36, 494–508.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Airaksinen, M. S., Thoenen, H., and Meyer, M. (1997). Vulnerability of midbrain dopaminergic neurons in calbindin-D28k-deficient mice: lack of evidence for a neuroprotective role of endogenous calbindin in MPTP-treated and weaver mice. Eur. J. Neurosci. 9, 120–127.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Alnemri, E. S. (2007). HtrA2 and Parkinson’s disease: think PINK? Nat. Cell Biol. 9, 1227–1229.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Ashley, A. K., Hanneman, W. H., Katoh, T., Moreno, J. A., Pollack, A., Tjalkens, R. B., and Legare, M. E. (2009). Analysis of targeted mutation in DJ-1 on cellular function in primary astrocytes. Toxicol. Lett. 184, 186–191.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Baker, D. A., Xi, Z. X., Shen, H., Swanson, C. J., and Kalivas, P. W. (2002). The origin and neuronal function of in vivo nonsynaptic glutamate. J. Neurosci. 22, 9134–9141.

Pubmed Abstract | Pubmed Full Text

Banati, R. B., Daniel, S. E., and Blunt, S. B. (1998). Glial pathology but absence of apoptotic nigral neurons in long-standing Parkinson’s disease. Mov. Disord. 13, 221–227.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Barzilai, A., Melalmed, E., and Shirvan, A. (2001). Is there a rationale for neuroprotection against dopamine toxicity in Parkinson’s disease. Cell. Mol. Neurobiol. 21, 215–235.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Bas, J., Calopa, M., Mestre, M., Mollevi, D. G., Cutillas, B., Ambrosio, S., and Buendia, E. (2001). Lymphocyte populations in Parkinson’s disease and in rat models of parkinsonism. J. Neuroimmunol. 113, 146–152.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Ben-Shachar, D., Riederer, P., and Youdim, M. B. (1991). Iron-melanin interaction and lipid peroxidation: implications for Parkinson’s disease. J. Neurochem. 57, 1609–1614.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Benazzouz, A., Breit, S., Koudsie, A., Pollak, P., Krack, P., and Benabid, A. L. (2002). Intraoperative microrecordings of the subthalamic nucleus in Parkinson’s disease. Mov. Disord. 17 (Suppl. 3), S145–S149.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Bender, A., Krishnan, K. J., Morris, C. M., Taylor, G. A., Reeve, A. K., Perry, R. H., Jaros, E., Hersheson, J. S., Betts, J., Klopstock, T., Taylor, R. W., and Turnbull, D. M. (2006). High levels of mitochondrial DNA deletions in substantia nigra neurons in aging and Parkinson disease. Nat. Genet. 38, 515–517.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Bennett, M. C. (2005). The role of alpha-synuclein in neurodegenerative diseases. Pharmacol. Ther. 105, 311–331.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Bennett, M. C., Bishop, J. F., Leng, Y., Chock, P. B., Chase, T. N., and Mouradian, M. M. (1999). Degradation of alpha-synuclein by proteasome. J. Biol. Chem. 274, 33855–33858.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Bergman, H., Wichmann, T., Karmon, B., and DeLong, M. R. (1994). The primate subthalamic nucleus. II. Neuronal activity in the MPTP model of parkinsonism. J. Neurophysiol. 72, 507–520.

Pubmed Abstract | Pubmed Full Text

Berman, S. B., and Hasting, T. G. (1999). Dopamine oxidation alters mitochondrial respiration and induces permeability transition in brain mitochondria: implications for Parkinson’s disease. J. Neurochem. 73, 1127–1137.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Berridge, M. J. (2002). The endoplasmic reticulum: a multifunctional signaling organelle. Cell Calcium 32, 235–249.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Betarbet, R., Sherer, T. B., MacKenzie, G., Garcia-Osuna, M., Panov, A. V., and Greenamyre, J. T. (2000). Chronic systemic pesticide exposure reproduces features of Parkinson’s disease. Nat. Neurosci. 3, 1301–1306.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Bezard, E., Boraud, T., Bioulac, B., and Gross, C. E. (1999a). Involvement of the subthalamic nucleus in glutamatergic compensatory mechanisms. Eur. J. Neurosci. 11, 2167–2170.

CrossRef Full Text

Bezard, E., Gross, C. E., Fournier, M. C., Dovero, S., Bloch, B., and Jaber, M. (1999b). Absence of MPTP-induced neuronal death in mice lacking the dopamine transporter. Exp. Neurol. 155, 268–273.

CrossRef Full Text

Biskup, S., Moore, D. J., Celsi, F., Higashi, S., West, A. B., Andrabi, S. A., Kurkinen, K., Yu, S. W., Savitt, J. M., Waldvogel, H. J., Faull, R. L., Emson, P. C., Torp, R., Ottersen, O. P., Dawson, T. M., and Dawson, V. L. (2006). Localization of LRRK2 to membranous and vesicular structures in mammalian brain. Ann. Neurol. 60, 557–569.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Blum, D., Torch, S., Lambeng, N., Nissou, M., Benabid, A. L., Sadoul, R., and Verna, J. M. (2001). Molecular pathways involved in the neurotoxicity of 6-OHDA, dopamine and MPTP: contribution to the apoptotic theory in Parkinson’s disease. Prog. Neurobiol. 65, 135–172.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Blum-Degen, D., Muller, T., Kuhn, W., Gerlach, M., Przuntek, H., and Riederer, P. (1995). Interleukin-1 beta and interleukin-6 are elevated in the cerebrospinal fluid of Alzheimer’s and de novo Parkinson’s disease patients. Neurosci. Lett. 202, 17–20.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Boka, G., Anglade, P., Wallach, D., Javoy-Agid, F., Agid, Y., and Hirsch, E. C. (1994). Immunocytochemical analysis of tumor necrosis factor and its receptors in Parkinson’s disease. Neurosci. Lett. 172, 151–154.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Bonifati, V., Rizzu, P., van Baren, M. J., Schaap, O., Breedveld, G. J., Krieger, E., Dekker, M. C., Squitieri, F., Ibanez, P., Joosse, M., van Dongen, J. W., Vanacore, N., van Swieten, J. C., Brice, A., Meco, G., van Duijn, C. M., Oostra, B. A., and Heutink, P. (2003). Mutations in the DJ-1 gene associated with autosomal recessive early-onset parkinsonism. Science 299, 256–259.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Borland, M. K., Mohanakumar, K. P., Rubinstein, J. D., Keeney, P. M., Xie, J., Capaldi, R., Dunham, L. D., Trimmer, P. A., and Bennett, J. P. Jr. (2009). Relationships among molecular genetic and respiratory properties of Parkinson’s disease cybrid cells show similarities to Parkinson’s brain tissues. Biochim. Biophys. Acta 1792, 68–74.

Pubmed Abstract | Pubmed Full Text

Bove, J., Prou, D., Perier, C., and Przedborski, S. (2005). Toxin-induced models of Parkinson’s disease. NeuroRx 2, 484–494.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Braak, H., Del Tredici, K., Rub, U., de Vos, R. A., Jansen Steur, E. N., and Braak, E. (2003). Staging of brain pathology related to sporadic Parkinson’s disease. Neurobiol. Aging 24, 197–211.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Burke, R. E., Dauer, W. T., and Vonsattel, J. P. (2008). A critical evaluation of the Braak staging scheme for Parkinson’s disease. Ann. Neurol. 64, 485–491.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Burnashev, N., Monyer, H., Seeburg, P. H., and Sakmann, B. (1992). Divalent ion permeability of AMPA receptor channels is dominated by the edited form of a single subunit. Neuron 8, 189–198.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Cabello, C. R., Thune, J. J., Pakkenberg, H., and Pakkenberg, B. (2002). Ageing of substantia nigra in humans: cell loss may be compensated by hypertrophy. Neuropathol. Appl. Neurobiol. 28, 283–291.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Canet-Aviles, R. M., Wilson, M. A., Miller, D. W., Ahmad, R., McLendon, C., Bandyopadhyay, S., Baptista, M. J., Ringe, D., Petsko, G. A., and Cookson, M. R. (2004). The Parkinson’s disease protein DJ-1 is neuroprotective due to cysteine-sulfinic acid-driven mitochondrial localization. Proc. Natl. Acad. Sci. U.S.A. 101, 9103–9108.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Cannon, J. R., Tapias, V., Na, H. M., Honick, A. S., Drolet, R. E., and Greenamyre, J. T. (2009). A highly reproducible rotenone model of Parkinson’s disease. Neurobiol. Dis. 34, 279–290.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Cass, W. A., Gerhardt, G. A., Mayfield, R. D., Curella, P., and Zahniser, N. R. (1992). Differences in dopamine clearance and diffusion in rat striatum and nucleus accumbens following systemic cocaine administration. J. Neurochem. 59, 259–266.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Caudle, W. M., Richardson, J. R., Wang, M. Z., Taylor, T. N., Guillot, T. S., McCormack, A. L., Colebrooke, R. E., Di Monte, D. A., Emson, P. C., and Miller, G. W. (2007). Reduced vesicular storage of dopamine causes progressive nigrostriatal neurodegeneration. J. Neurosci. 27, 8138–8148.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Cavelier, P., and Attwell, D. (2005). Tonic release of glutamate by a DIDS-sensitive mechanism in rat hippocampal slices. J. Physiol. 564, 397–410.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Celsi, F., Pizzo, P., Brini, M., Leo, S., Fotino, C., Pinton, P., and Rizzuto, R. (2009). Mitochondria, calcium and cell death: a deadly triad in neurodegeneration. Biochim. Biophys. Acta 1787, 335–344.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Cerruti, C., Walther, D. M., Kuhar, M. J., and Uhl, G. R. (1993). Dopamine transporter mRNA expression is intense in rat midbrain neurons and modest outside midbrain. Brain Res. Mol. Brain Res. 18, 181–186.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Chan, C. S., Gertler, T. S., and Surmeier, D. J. (2009). Calcium homeostasis, selective vulnerability and Parkinson’s disease. Trends Neurosci. 32, 249–256.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Chan, C. S., Guzman, J. N., Ilijic, E., Mercer, J. N., Rick, C., Tkatch, T., Meredith, G. E., and Surmeier, D. J. (2007). ‘Rejuvenation’ protects neurons in mouse models of Parkinson’s disease. Nature 447, 1081–1086.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Charara, A., Smith, Y., and Parent, A. (1996). Glutamatergic inputs from the pedunculopontine nucleus to midbrain dopaminergic neurons in primates: phaseolus vulgaris-leucoagglutinin anterograde labeling combined with postembedding glutamate and GABA immunohistochemistry. J. Comp. Neurol. 364, 254–266.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Choi, Y. M., Kim, S. H., Chung, S., Uhm, D. Y., and Park, M. K. (2006). Regional interaction of endoplasmic reticulum Ca2+ signals between soma and dendrites through rapid luminal Ca2+ diffusion. J. Neurosci. 26, 12127–12136.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Chung, C. Y., Seo, H., Sonntag, K. C., Brooks, A., Lin, L., and Isacson, O. (2005). Cell type-specific gene expression of midbrain dopaminergic neurons reveals molecules involved in their vulnerability and protection. Hum. Mol. Genet. 14, 1709–1725.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Cicchetti, F., Drouin-Ouellet, J., and Gross, R. E. (2009). Environmental toxins and Parkinson’s disease: what have we learned from pesticide-induced animal models? Trends Pharmacol. Sci. 30, 475–483.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Collier, T. J., Lipton, J., Daley, B. F., Palfi, S., Chu, Y., Sortwell, C., Bakay, R. A., Sladek, J. R., Jr., and Kordower, J. H. (2007). Aging-related changes in the nigrostriatal dopamine system and the response to MPTP in nonhuman primates: diminished compensatory mechanisms as a prelude to parkinsonism. Neurobiol. Dis. 26, 56–65.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Conn, P. J., and Pin, J. P. (1997). Pharmacology and functions of metabotropic glutamate receptors. Annu. Rev. Pharmacol. Toxicol. 37, 205–237.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Conway, K. A., Lee, S. J., Rochet, J. C., Ding, T. T., Williamson, R. E., and Lansbury, P. T. Jr. (2000). Acceleration of oligomerization, not fibrillization, is a shared property of both alpha-synuclein mutations linked to early-onset Parkinson’s disease: implications for pathogenesis and therapy. Proc. Natl. Acad. Sci. U.S.A. 97, 571–576.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Cruz-Muros, I., Afonso-Oramas, D., Abreu, P., Barroso-Chinea, P., Rodriguez, M., Gonzalez, M. C., and Hernandez, T. G. (2007). Aging of the rat mesostriatal system: differences between the nigrostriatal and the mesolimbic compartments. Exp. Neurol. 204, 147–161.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Cruz-Muros, I., Afonso-Oramas, D., Abreu, P., Rodriguez, M., Gonzalez, M. C., and Gonzalez-Hernandez, T. (2008). Deglycosylation and subcellular redistribution of VMAT2 in the mesostriatal system during normal aging. Neurobiol. Aging 29, 1702–1711.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Cuervo, A. M., Stefanis, L., Fredenburg, R., Lansbury, P. T., and Sulzer, D. (2004). Impaired degradation of mutant alpha-synuclein by chaperone-mediated autophagy. Science 305, 1292–1295.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

D’Amato, R. J., Lipman, Z. P., and Snyder, S. H. (1986). Selectivity of the parkinsonian neurotoxin MPTP: toxic metabolite MPP+ binds to neuromelanin. Science 231, 987–989.

CrossRef Full Text

Damier, P., Hirsch, E. C., Agid, Y., and Graybiel, A. M. (1999). The substantia nigra of the human brain. II. Patterns of loss of dopamine-containing neurons in Parkinson’s disease. Brain 122(Pt 8), 1437–1448.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Darios, F., Corti, O., Lucking, C. B., Hampe, C., Muriel, M. P., Abbas, N., Gu, W. J., Hirsch, E. C., Rooney, T., Ruberg, M., and Brice, A. (2003). Parkin prevents mitochondrial swelling and cytochrome c release in mitochondria-dependent cell death. Hum. Mol. Genet. 12, 517–526.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Dauer, W., and Przedborski, S. (2003). Parkinson’s disease: mechanisms and models. Neuron 39, 889–909.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

de Lau, L. M., and Breteler, M. M. (2006). Epidemiology of Parkinson’s disease. Lancet Neurol. 5, 525–535.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Dervan, A. G., Meshul, C. K., Beales, M., McBean, G. J., Moore, C., Totterdell, S., Snyder, A. K., and Meredith, G. E. (2004). Astroglial plasticity and glutamate function in a chronic mouse model of Parkinson’s disease. Exp. Neurol. 190, 145–156.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Devi, L., Raghavendran, V., Prabhu, B. M., Avadhani, N. G., and Anandatheerthavarada, H. K. (2008). Mitochondrial import and accumulation of alpha-synuclein impair complex I in human dopaminergic neuronal cultures and Parkinson disease brain. J. Biol. Chem. 283, 9089–9100.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Dhillon, A. S., Tarbutton, G. L., Levin, J. L., Plotkin, G. M., Lowry, L. K., Nalbone, J. T., and Shepherd, S. (2008). Pesticide/environmental exposures and Parkinson’s disease in East Texas. J. Agromedicine 13, 37–48.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Di Fonzo, A., Rohe, C. F., Ferreira, J., Chien, H. F., Vacca, L., Stocchi, F., Guedes, L., Fabrizio, E., Manfredi, M., Vanacore, N., Goldwurm, S., Breedveld, G., Sampaio, C., Meco, G., Barbosa, E., Oostra, B. A., and Bonifati, V. (2005). A frequent LRRK2 gene mutation associated with autosomal dominant Parkinson’s disease. Lancet 365, 412–415.

Pubmed Abstract | Pubmed Full Text

Di Matteo, V., Pierucci, M., Di Giovanni, G., Di Santo, A., Poggi, A., Benigno, A., and Esposito, E. (2006). Aspirin protects striatal dopaminergic neurons from neurotoxin-induced degeneration: an in vivo microdialysis study. Brain Res. 1095, 167–177.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Di Monte, D., Jewell, S. A., Ekstrom, G., Sandy, M. S., and Smith, M. T. (1986a). 1-Methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) and 1-methyl-4-phenylpyridine (MPP+) cause rapid ATP depletion in isolated hepatocytes. Biochem. Biophys. Res. Commun. 137, 310–315.

CrossRef Full Text

Di Monte, D., Sandy, M. S., Ekstrom, G., and Smith, M. T. (1986b). Comparative studies on the mechanisms of paraquat and 1-methyl-4-phenylpyridine (MPP+) cytotoxicity. Biochem. Biophys. Res. Commun. 137, 303–309.

CrossRef Full Text

Enochs, W. S., Sarna, T., Zecca, L., Riley, P. A., and Swartz, H. M. (1994). The roles of neuromelanin, binding of metal ions, and oxidative cytotoxicity in the pathogenesis of Parkinson’s disease: a hypothesis. J. Neural. Transm. Park. Dis. Dement. Sect. 7, 83–100.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Fasano, M., Bergamasco, B., and Lopiano, L. (2006). Modifications of iron-neuromelanin system in Parkinson’s disease. J. Neurochem. 96, 909–916.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Fearnley, J. M., and Lees, A. J. (1991). Ageing and Parkinson’s disease: substantia nigra regional selectivity. Brain 114 (Pt 5), 2283–2301.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Firestone, J. A., Smith-Weller, T., Franklin, G., Swanson, P., Longstreth, W. T. Jr., and Checkoway, H. (2005). Pesticides and risk of Parkinson disease: a population-based case-control study. Arch. Neurol. 62, 91–95.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Foix, C., and Nicolesco, J. (1925). Anatomie cérébrale. Les noyaux gris centraux et la región Mésencéphalo-sous-optique., Suivi d’un apéndice sur l’anatomie pathologique de la maladie de Parkinson. Paris: Masson et Cie, pp. 508–538.

Forno, L. S. (1996). Neuropathology of Parkinson’s disease. J. Neuropathol. Exp. Neurol. 55, 259–272.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Fornsted, B., Rosengren, E., and Carlsson, A. (1986). Occurrence and distribution of 5-S-cysteinyl derivates of dopamine, dopa and dopac in brains of eight mammalian species. Neuropharmacology 25, 451–454.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Forster, G. L., and Blaha, C. D. (2003). Pedunculopontine tegmental stimulation evokes striatal dopamine efflux by activation of acetylcholine and glutamate receptors in the midbrain and pons of the rat. Eur. J. Neurosci. 17, 751–762.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Fredenburg, R. A., Rospigliosi, C., Meray, R. K., Kessler, J. C., Lashuel, H. A., Eliezer, D., and Lansbury, P. T. Jr. (2007). The impact of the E46K mutation on the properties of alpha-synuclein in its monomeric and oligomeric states. Biochemistry 46, 7107–7118.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Gainetdinov, R. R., Fumagalli, F., Jones, S. R., and Caron, M. G. (1997). Dopamine transporter is required for in vivo MPTP neurotoxicity: evidence from mice lacking the transporter. J. Neurochem. 69, 1322–1325.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Galluzzi, L., Blomgren, K., and Kroemer, G. (2009). Mitochondrial membrane permeabilization in neuronal injury. Nat. Rev. Neurosci. 10, 481–494.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Garcia, L., D’Alessandro, G., Fernagut, P. O., Bioulac, B., and Hammond, C. (2005). Impact of high-frequency stimulation parameters on the pattern of discharge of subthalamic neurons. J. Neurophysiol. 94, 3662–3669.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Gautier, C. A., Kitada, T., and Shen, J. (2008). Loss of PINK1 causes mitochondrial functional defects and increased sensitivity to oxidative stress. Proc. Natl. Acad. Sci. U.S.A. 105, 11364–11369.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Geiger, J. R., Melcher, T., Koh, D. S., Sakmann, B., Seeburg, P. H., Jonas, P., and Monyer, H. (1995). Relative abundance of subunit mRNAs determines gating and Ca2+ permeability of AMPA receptors in principal neurons and interneurons in rat CNS. Neuron 15, 193–204.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Gerfen, C. R., Baimbridge, K. G., and Miller, J. J. (1985). The neostriatal mosaic: compartmental distribution of calcium-binding protein and parvalbumin in the basal ganglia of the rat and monkey. Proc. Natl. Acad. Sci. U.S.A. 82, 8780–8784.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Gerfen, C. R., Herkenham, M., and Thibault, J. (1987). The neostriatal mosaic: II. Patch- and matrix-directed mesostriatal dopaminergic and non-dopaminergic systems. J. Neurosci. 7, 3915–3934.

Pubmed Abstract | Pubmed Full Text

Gerhardt, G. A., Cass, W. A., Yi, A., Zhang, Z., and Gash, D. M. (2002). Changes in somatodendritic but not terminal dopamine regulation in aged rhesus monkeys. J. Neurochem. 80, 168–177.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

German, D. C., Manaye, K., Smith, W. K., Woodward, D. J., and Saper, C. B. (1989). Midbrain dopaminergic cell loss in Parkinson’s disease: computer visualization. Ann. Neurol. 26, 507–514.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

German, D. C., Manaye, K. F., Sonsalla, P. K., and Brooks, B. A. (1992). Midbrain dopaminergic cell loss in Parkinson’s disease and MPTP-induced parkinsonism: sparing of calbindin-D28k-containing cells. Ann. N.Y. Acad. Sci. 648, 42–62.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Giasson, B. I., Covy, J. P., Bonini, N. M., Hurtig, H. I., Farrer, M. J., Trojanowski, J. Q., and Van Deerlin, V. M. (2006). Biochemical and pathological characterization of Lrrk2. Ann. Neurol. 59, 315–322.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Gibb, W. (1992). Melanin, tyrosine hydroxylase, calbindin and substance P in the human midbrain and substantia nigra in relation to nigrastriatal projections and differential neuron susceptibility in Parkinson’s disease. Brain Res. 581, 283–291.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Gibb, W. R., and Lees, A. J. (1988). A comparison of clinical and pathological features of young- and old-onset Parkinson’s disease. Neurology 38, 1402–1406.

Pubmed Abstract | Pubmed Full Text

Giros, B., and Caron, M. G. (1993). Molecular characterization of the dopamine transporter. Trends Pharmacol. Sci. 14, 43–49.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Gloeckner, C. J., Kinkl, N., Schumacher, A., Braun, R. J., O’Neill, E., Meitinger, T., Kolch, W., Prokisch, H., and Ueffing, M. (2006). The Parkinson disease causing LRRK2 mutation I2020T is associated with increased kinase activity. Hum. Mol. Genet. 15, 223–232.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Goldberg, M. S., Pisani, A., Haburcak, M., Vortherms, T. A., Kitada, T., Costa, C., Tong, Y., Martella, G., Tscherter, A., Martins, A., Bernardi, G., Roth, B. L., Pothos, E. N., Calabresi, P., and Shen, J. (2005). Nigrostriatal dopaminergic deficits and hypokinesia caused by inactivation of the familial Parkinsonism-linked gene DJ-1. Neuron 45, 489–496.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Gonzalez-Hernandez, T., Afonso-Oramas, D., and Cruz-Muros, I. (2009). Phenotype, compartmental organization and differential vulnerability of nigral dopaminergic neurons. J. Neural. Transm. Suppl. 21–37.

Gonzalez-Hernandez, T., Barroso-Chinea, P., De La Cruz Muros, I., Del Mar Perez-Delgado, M., and Rodriguez, M. (2004). Expression of dopamine and vesicular monoamine transporters and differential vulnerability of mesostriatal dopaminergic neurons. J. Comp. Neurol. 479, 198–215.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Greene, J. G., Dingledine, R., and Greenamyre, J. T. (2005). Gene expression profiling of rat midbrain dopamine neurons: implications for selective vulnerability in parkinsonism. Neurobiol. Dis. 18, 19–31.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Greffard, S., Verny, M., Bonnet, A. M., Beinis, J. Y., Gallinari, C., Meaume, S., Piette, F., Hauw, J. J., and Duyckaerts, C. (2006). Motor score of the unified Parkinson disease rating scale as a good predictor of Lewy body-associated neuronal loss in the substantia nigra. Arch. Neurol. 63, 584–588.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Haas, R. H., Nasirian, F., Nakano, K., Ward, D., Pay, M., Hill, R., and Shults, C. W. (1995). Low platelet mitochondrial complex I and complex II/III activity in early untreated Parkinson’s disease. Ann. Neurol. 37, 714–722.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Haber, S. N., Ryoo, H., Cox, C., and Lu, W. (1995). Subsets of midbrain dopaminergic neurons in monkeys are distinguished by different levels of mRNA for the dopamine transporter: comparison with the mRNA for the D2 receptor, tyrosine hydroxylase and calbindin immunoreactivity. J. Comp. Neurol. 362, 400–410.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Harman, D. (1956). Aging: a theory based on free radical and radiation chemistry. J. Gerontol. 11, 298–300.

Pubmed Abstract | Pubmed Full Text

Harrower, T. P., Michell, A. W., and Barker, R. A. (2005). Lewy bodies in Parkinson’s disease: protectors or perpetrators? Exp. Neurol. 195, 1–6.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Hassler, R. (1938). Zur pathologie der paralysis agitans und des postencephalitschen Parkinsonismus. J. Psychol. Neurol. 48, 387–476.

Hasting, T. G. (1995). Enzymatic oxidation of dopamine: the role of prostaglandin H synthase. J. Neurochem. 64, 919–924.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Haycock, J. W., Becker, L., Ang, L., Furukawa, Y., Hornykiewicz, O., and Kish, S. J. (2003). Marked disparity between age-related changes in dopamine and other presynaptic dopaminergic markers in human striatum. J. Neurochem. 87, 574–585.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Hindle, J. V. (2010). Ageing, neurodegeneration and Parkinson’s disease. Age Ageing 39, 156–161.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Hirsch, E., Graybiel, A., and Agid, Y. (1988). Melanized dopamine neurons are differentially susceptible to degeneration in Parkinson’s disease. Nature 334, 345–348.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Hirsch, E. C., and Hunot, S. (2009). Neuroinflammation in Parkinson’s disease: a target for neuroprotection? Lancet Neurol. 8, 382–397.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Hisanaga, K., Asagi, M., Itoyama, Y., and Iwasaki, Y. (2001). Increase in peripheral CD4 bright+ CD8 +dull T cells in Parkinson disease. Arch. Neurol. 58, 1580–1583.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Hoglinger, G. U., Feger, J., Prigent, A., Michel, P. P., Parain, K., Champy, P., Ruberg, M., Oertel, W. H., and Hirsch, E. C. (2003). Chronic systemic complex I inhibition induces a hypokinetic multisystem degeneration in rats. J. Neurochem. 84, 491–502.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Hollmann, M., and Heinemann, S. (1994). Cloned glutamate receptors. Annu. Rev. Neurosci. 17, 31–108.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Horn, A. S. (1990). Dopamine uptake: a review of progress in the last decade. Prog. Neurobiol. 34, 387–400.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Hornykiewicz, O. (2006). The discovery of dopamine deficiency in the parkinsonian brain. J. Neural Transm. Suppl. 70, 9–15.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Hosoi, T., and Ozawa, K. (2009). Endoplasmic reticulum stress in disease: mechanisms and therapeutic opportunities. Clin. Sci. (Lond) 118, 19–29.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Hsu, L. J., Sagara, Y., Arroyo, A., Rockenstein, E., Sisk, A., Mallory, M., Wong, J., Takenouchi, T., Hashimoto, M., and Masliah, E. (2000). alpha-synuclein promotes mitochondrial deficit and oxidative stress. Am. J. Pathol. 157, 401–410.

Pubmed Abstract | Pubmed Full Text

Hunot, S., Boissiere, F., Faucheux, B., Brugg, B., Mouatt-Prigent, A., Agid, Y., and Hirsch, E. C. (1996). Nitric oxide synthase and neuronal vulnerability in Parkinson’s disease. Neuroscience 72, 355–363.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Hurd, Y. L., Pristupa, Z. B., Herman, M. M., Niznik, H. B., and Kleinman, J. E. (1994). The dopamine transporter and dopamine D2 receptor messenger RNAs are differentially expressed in limbic- and motor-related subpopulations of human mesencephalic neurons. Neuroscience 63, 357–362.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Hurley, S. D., O’Banion, M. K., Song, D. D., Arana, F. S., Olschowka, J. A., and Haber, S. N. (2003). Microglial response is poorly correlated with neurodegeneration following chronic, low-dose MPTP administration in monkeys. Exp. Neurol. 184, 659–668.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Iaccarino, C., Crosio, C., Vitale, C., Sanna, G., Carri, M. T., and Barone, P. (2007). Apoptotic mechanisms in mutant LRRK2-mediated cell death. Hum. Mol. Genet. 16, 1319–1326.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Imai, Y., Soda, M., Inoue, H., Hattori, N., Mizuno, Y., and Takahashi, R. (2001). An unfolded putative transmembrane polypeptide, which can lead to endoplasmic reticulum stress, is a substrate of Parkin. Cell 105, 891–902.

Pubmed Abstract | Pubmed Full Text

Imamura, K., Hishikawa, N., Sawada, M., Nagatsu, T., Yoshida, M., and Hashizume, Y. (2003). Distribution of major histocompatibility complex class II-positive microglia and cytokine profile of Parkinson’s disease brains. Acta Neuropathol. 106, 518–526.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Iribe, Y., Moore, K., Pang, K. C., and Tepper, J. M. (1999). Subthalamic stimulation-induced synaptic responses in substantia nigra pars compacta dopaminergic neurons in vitro. J. Neurophysiol. 82, 925–933.

Pubmed Abstract | Pubmed Full Text

Irwin, I., DeLanney, L. E., McNeill, T., Chan, P., Forno, L. S., Murphy, G. M., Jr., Di Monte, D. A., Sandy, M. S., and Langston, J. W. (1994). Aging and the nigrostriatal dopamine system: a non-human primate study. Neurodegeneration 3, 251–265.

Pubmed Abstract | Pubmed Full Text

Jellinger, K. A. (2009). Recent advances in our understanding of neurodegeneration. J. Neural. Transm. 116, 1111–1162.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Jenner, P. (2003). Oxidative stress in Parkinson’s disease. Ann. Neurol. 53 (Suppl. 3), S26–S36; discussion S28–S36.

Pubmed Abstract | Pubmed Full Text

Joglar, B., Rodriguez-Pallares, J., Rodriguez-Perez, A. I., Rey, P., Guerra, M. J., and Labandeira-Garcia. J. L. (2009). The inflammatory response in the MPTP model of Parkinson’s disease is mediated by brain angiotensin: relevance to progression of the disease. J. Neurochem. 109, 656–669.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Jonsson, G. (1971). Quantitation of fluorescence of biogenic monoamines. Prog. Histochem. Cytochem. 2, 244–299.

Keeney, P. M., Xie, J., Capaldi, R. A., and Bennett, J. P. Jr. (2006). Parkinson’s disease brain mitochondrial complex I has oxidatively damaged subunits and is functionally impaired and misassembled. J. Neurosci. 26, 5256–5264.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kim, R. H., Smith, P. D., Aleyasin, H., Hayley, S., Mount, M. P., Pownall, S., Wakeham, A., You-Ten, A. J., Kalia, S. K., Horne, P., Westaway, D., Lozano, A. M., Anisman, H., Park, D. S., and Mak, T. W. (2005). Hypersensitivity of DJ-1-deficient mice to 1-methyl-4-phenyl-1,2,3,6-tetrahydropyrindine (MPTP) and oxidative stress. Proc. Natl. Acad. Sci. U.S.A. 102, 5215–5220.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kim, Y., Park, J., Kim, S., Song, S., Kwon, S. K., Lee, S. H., Kitada, T., Kim, J. M., and Chung, J. (2008). PINK1 controls mitochondrial localization of Parkin through direct phosphorylation. Biochem. Biophys. Res. Commun. 377, 975–980.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kish, S. J., Shannak, K., Rajput, A., Deck, J. H., and Hornykiewicz, O. (1992). Aging produces a specific pattern of striatal dopamine loss: implications for the etiology of idiopathic Parkinson’s disease. J. Neurochem. 58, 642–648.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kita, H., and Kitai, S. T. (1987). Efferent projections of the subthalamic nucleus in the rat: light and electron microscopic analysis with the PHA-L method. J. Comp. Neurol. 260, 435–452.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kitada, T., Asakawa, S., Hattori, N., Matsumine, H., Yamamura, Y., Minoshima, S., Yokochi, M., Mizuno, Y., and Shimizu, N. (1998). Mutations in the parkin gene cause autosomal recessive juvenile parkinsonism. Nature 392, 605–608.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Knott, C., Stern, G., and Wilkin, G. P. (2000). Inflammatory regulators in Parkinson’s disease: iNOS, lipocortin-1, and cyclooxygenases-1 and -2. Mol. Cell. Neurosci. 16, 724–739.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kopin, I. J. (1992). Features of the dopaminergic neurotoxin MPTP. Ann. N. Y. Acad. Sci. 648, 96–104.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kraytsberg, Y., Kudryavtseva, E., McKee, A. C., Geula, C., Kowall, N. W., and Khrapko, K. (2006). Mitochondrial DNA deletions are abundant and cause functional impairment in aged human substantia nigra neurons. Nat. Genet. 38, 518–520.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Krige, D., Carroll, M. T., Cooper, J. M., Marsden, C. D., and Schapira, A. H. (1992). Platelet mitochondrial function in Parkinson’s disease. The Royal Kings and Queens Parkinson Disease Research Group. Ann. Neurol. 32, 782–788.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kroemer, G., and Blomgren, K. (2007). Mitochondrial cell death control in familial Parkinson disease. PLoS Biol. 5, e206. doi: 10.1371/journal.pbio.0050206.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kubis, N., Faucheux, B. A., Ransmayr, G., Damier, P., Duyckaerts, C., Henin, D., Forette, B., Le Charpentier, Y., Hauw, J. J., Agid, Y., and Hirsch, E. C. (2000). Preservation of midbrain catecholaminergic neurons in very old human subjects. Brain 123 (Pt 2), 366–373.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kuhn, D. M., Francescutti-Verbeem, D. M., and Thomas, D. M. (2006). Dopamine quinones activate microglia and induce a neurotoxic gene expression profile: relationship to methamphetamine-induced nerve ending damage. Ann. N.Y. Acad. Sci. 1074, 31–41.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kurkowska-Jastrzebska, I., Babiuch, M., Joniec, I., Przybylkowski, A., Czlonkowski, A., and Czlonkowska, A. (2002). Indomethacin protects against neurodegeneration caused by MPTP intoxication in mice. Int. Immunopharmacol. 2, 1213–1218.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kuroda, Y., Mitsui, T., Kunishige, M., and Matsumoto, T. (2006). Parkin affects mitochondrial function and apoptosis in neuronal and myogenic cells. Biochem. Biophys. Res. Commun. 348, 787–793.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

LaFerla, F. M. (2002). Calcium dyshomeostasis and intracellular signalling in Alzheimer’s disease. Nat. Rev. Neurosci. 3, 862–872.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Langston, J. W., Ballard, P., Tetrud, J. W., and Irwin, I. (1983). Chronic Parkinsonism in humans due to a product of meperidine-analog synthesis. Science 219, 979–980.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Lapointe, N., St-Hilaire, M., Martinoli, M. G., Blanchet, J., Gould, P., Rouillard, C., and Cicchetti, F. (2004). Rotenone induces non-specific central nervous system and systemic toxicity. FASEB J. 18, 717–719.

Pubmed Abstract | Pubmed Full Text

Lashuel, H. A., Petre, B. M., Wall, J., Simon, M., Nowak, R. J., Walz, T., and Lansbury, P. T. Jr. (2002). Alpha-synuclein, especially the Parkinson’s disease-associated mutants, forms pore-like annular and tubular protofibrils. J. Mol. Biol. 322, 1089–1102.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

LaVoie, M. J., Ostaszewski, B. L., Weihofen, A., Schlossmacher, M. G., and Selkoe, D. J. (2005). Dopamine covalently modifies and functionally inactivates parkin. Nat. Med. 11, 1214–1221.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Lee, C. R., and Tepper, J. M. (2009). Basal ganglia control of substantia nigra dopaminergic neurons. J. Neural. Transm. Suppl. 71–90.

Lee, S. J., Kim, S. J., Kim, I. K., Ko, J., Jeong, C. S., Kim, G. H., Park, C., Kang, S. O., Suh, P. G., Lee, H. S., and Cha, S. S. (2003). Crystal structures of human DJ-1 and Escherichia coli Hsp31, which share an evolutionarily conserved domain. J. Biol. Chem. 278, 44552–44559.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Leroy, E., Boyer, R., Auburger, G., Leube, B., Ulm, G., Mezey, E., Harta, G., Brownstein, M. J., Jonnalagada, S., Chernova, T., Dehejia, A., Lavedan, C., Gasser, T., Steinbach, P. J., Wilkinson, K. D., and Polymeropoulos, M. H. (1998). The ubiquitin pathway in Parkinson’s disease. Nature 395, 451–452.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Liang, C. L., Sinton, C. M., and German, D. C. (1996). Midbrain dopaminergic neurons in the mouse: co-localization with Calbindin-D28K and calretinin. Neuroscience 75, 523–533.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Liberatore, G. T., Jackson-Lewis, V., Vukosavic, S., Mandir, A. S., Vila, M., McAuliffe, W. G., Dawson, V. L., Dawson, T. M., and Przedborski, S. (1999). Inducible nitric oxide synthase stimulates dopaminergic neurodegeneration in the MPTP model of Parkinson disease. Nat. Med. 5, 1403–1409.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Limousin, P., Krack, P., Pollak, P., Benazzouz, A., Ardouin, C., Hoffmann, D., and Benabid, A. L. (1998). Electrical stimulation of the subthalamic nucleus in advanced Parkinson’s disease. N. Engl. J. Med. 339, 1105–1111.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Lindquist, N. G., Larsson, B. S., and Lydén-Sokolowski, A. (1988). Autoradiography of [14C] paraquat or [14C] diquat in frogs and mice: accumulation in neuromelanin. Neurosci. Lett. 93, 1–6.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Liou, H. H., Tsai, M. C., Chen, C. J., Jeng, J. S., Chang, Y. C., Chen, S. Y., and Chen, R. C. (1997). Environmental risk factors and Parkinson’s disease: a case-control study in Taiwan. Neurology 48, 1583–1588.

Pubmed Abstract | Pubmed Full Text

Lippa, C. F., Duda, J. E., Grossman, M., Hurtig, H. I., Aarsland, D., Boeve, B. F., Brooks, D. J., Dickson, D. W., Dubois, B., Emre, M., Fahn, S., Farmer, J. M., Galasko, D., Galvin, J. E., Goetz, C. G., Growdon, J. H., Gwinn-Hardy, K. A., Hardy, J., Heutink, P., Iwatsubo, T., Kosaka, K., Lee, V. M., Leverenz, J. B., Masliah, E., McKeith, I. G., Nussbaum, R. L., Olanow, C. W., Ravina, B. M., Singleton, A. B., Tanner, C. M., Trojanowski, J. Q., and Wszolek, Z. K. (2007). DLB and PDD boundary issues: diagnosis, treatment, molecular pathology, and biomarkers. Neurology 68, 812–819.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Lohmann, E., Periquet, M., Bonifati, V., Wood, N. W., De Michele, G., Bonnet, A. M., Fraix, V., Broussolle, E., Horstink, M. W., Vidailhet, M., Verpillat, P., Gasser, T., Nicholl, D., Teive, H., Raskin, S., Rascol, O., Destee, A., Ruberg, M., Gasparini, F., Meco, G., Agid, Y., Durr, A., and Brice, A. (2003). How much phenotypic variation can be attributed to parkin genotype? Ann. Neurol. 54, 176–185.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Mak, S. K., McCormack, A. L., Manning-Bog, A. B., Cuervo, A. M., and Di Monte, D. A. (2010). Lysosomal degradation of alpha-synuclein in vivo. J. Biol. Chem. 285, 13621–13629.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Mann, D., and Yates, P. (1983). Possible role of neuromelanin in the pathogenesis of Parkinson’s disease. Mech. Ageing Dev. 21, 193–203.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Manning-Bog, A. B., McCormack, A. L., Purisai, M. G., Bolin, L. M., and Di Monte, D. A. (2003). Alpha-synuclein overexpression protects against paraquat-induced neurodegeneration. J. Neurosci. 23, 3095–3099.

Pubmed Abstract | Pubmed Full Text

Marshall, J. F., O’Dell, S. J., Navarrete, R., and Rosenstein, A. J. (1990). Dopamine high-affinity transport site topography in rat brain: major differences between dorsal and ventral striatum. Neuroscience 37, 11–21.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Martin, L. J., Pan, Y., Price, A. C., Sterling, W., Copeland, N. G., Jenkins, N. A., Price, D. L., and Lee, M. K. (2006). Parkinson’s disease alpha-synuclein transgenic mice develop neuronal mitochondrial degeneration and cell death. J. Neurosci. 26, 41–50.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

McCormack, A. L., Di Monte, D. A., Delfani, K., Irwin, I., DeLanney, L. E., Langston, W. J., and Janson, A. M. (2004). Aging of the nigrostriatal system in the squirrel monkey. J. Comp. Neurol. 471, 387–395.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

McCormack, A. L., Thiruchelvam, M., Manning-Bog, A. B., Thiffault, C., Langston, J. W., Cory-Slechta, D. A., and Di Monte, D. A. (2002). Environmental risk factors and Parkinson’s disease: selective degeneration of nigral dopaminergic neurons caused by the herbicide paraquat. Neurobiol. Dis. 10, 119–127.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

McCormack, J. G., Halestrap, A. P., and Denton, R. M. (1990). Role of calcium ions in regulation of mammalian intramitochondrial metabolism. Physiol. Rev. 70, 391–425.

Pubmed Abstract | Pubmed Full Text

McGeer, P. L., Itagaki, S., Akiyama, H., and McGeer, E. G. (1988). Rate of cell death in parkinsonism indicates active neuropathological process. Ann. Neurol. 24, 574–576.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

McGeer, P. L., and McGeer, E. G. (2004). Inflammation and the degenerative diseases of aging. Ann. N. Y. Acad. Sci. 1035, 104–116.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

McGeer, P. L., McGeer, E. G., and Suzuki, J. S. (1977). Aging and extrapyramidal function. Arch. Neurol. 34, 33–35.

Pubmed Abstract | Pubmed Full Text

McGeer, P. L., Schwab, C., Parent, A., and Doudet, D. (2003). Presence of reactive microglia in monkey substantia nigra years after 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine administration. Ann. Neurol. 54, 599–604.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

McNaught, K. S., Belizaire, R., Isacson, O., Jenner, P., and Olanow, C. W. (2003). Altered proteasomal function in sporadic Parkinson’s disease. Exp. Neurol. 179, 38–46.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

McNaught, K. S., Belizaire, R., Jenner, P., Olanow, C. W., and Isacson, O. (2002a). Selective loss of 20S proteasome alpha-subunits in the substantia nigra pars compacta in Parkinson’s disease. Neurosci. Lett. 326, 155–158.

CrossRef Full Text

McNaught, K. S., Jnobaptiste, R., Jackson, T., and Jengelley, T. A. (2010). The pattern of neuronal loss and survival may reflect differential expression of proteasome activators in Parkinson’s disease. Synapse 64, 241–250.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

McNaught, K. S., Olanow, C. W., Halliwell, B., Isacson, O., and Jenner, P. (2001). Failure of the ubiquitin-proteasome system in Parkinson’s disease. Nat. Rev. Neurosci. 2, 589–594.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

McNaught, K. S., Shashidharan, P., Perl, D. P., Jenner, P., and Olanow, C. W. (2002b). Aggresome-related biogenesis of Lewy bodies. Eur. J. Neurosci. 16, 2136–2148.

CrossRef Full Text

McRitchie, D. A., and Halliday, G. M. (1995). Calbindin D28k-containing neurons are restricted to the medial substantia nigra in humans. Neuroscience 65, 87–91.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

McRitchie, D. A., Hardman, C. D., and Halliday, G. M. (1996). Cytoarchitectural distribution of calcium binding proteins in midbrain dopaminergic regions of rats and humans. J. Comp. Neurol. 364, 121–150.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Miller, G. W. (2007). Paraquat: the red herring of Parkinson’s disease research. Toxicol. Sci. 100, 1–2.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Miller, G. W., Gainetdinov, R. R., Levey, A. I., and Caron, M. G. (1999). Dopamine transporters and neuronal injury. Trends Pharmacol. Sci. 20, 424–429.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Missale, C., Castelletti, L., Govoni, S., Spano, P. F., Trabucchi, M., and Hanbauer, I. (1985). Dopamine uptake is differentially regulated in rat striatum and nucleus accumbens. J. Neurochem. 45, 51–56.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Miyazaki, I., and Asanuma, M. (2009). Approaches to prevent dopamine quinone-induced neurotoxicity. Neurochem. Res. 34, 698–706.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Mizuno, Y., Sone, N., Suzuki, K., and Saitoh, T. (1988). Studies on the toxicity of 1-methyl-4-phenylpyridinium ion (MPP+) against mitochondria of mouse brain. J. Neurol. Sci. 86, 97–110.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Mogi, M., Harada, M., Kondo, T., Riederer, P., Inagaki, H., Minami, M., and Nagatsu, T. (1994a). Interleukin-1 beta, interleukin-6, epidermal growth factor and transforming growth factor-alpha are elevated in the brain from parkinsonian patients. Neurosci. Lett. 180, 147–150.

CrossRef Full Text

Mogi, M., Harada, M., Riederer, P., Narabayashi, H., Fujita, K., and Nagatsu, T. (1994b). Tumor necrosis factor-alpha (TNF-alpha) increases both in the brain and in the cerebrospinal fluid from parkinsonian patients. Neurosci. Lett. 165, 208–210.

CrossRef Full Text

Mogi, M., Togari, A., Kondo, T., Mizuno, Y., Komure, O., Kuno, S., Ichinose, H., and Nagatsu, T. (2000). Caspase activities and tumor necrosis factor receptor R1 (p55) level are elevated in the substantia nigra from parkinsonian brain. J. Neural. Transm. 107, 335–341.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Muñoz, A., Rey, P., Guerra, M. J., Mendez-Alvarez, E., Soto-Otero, R., and Labandeira-Garcia, J. L. (2006). Reduction of dopaminergic degeneration and oxidative stress by inhibition of angiotensin converting enzyme in a MPTP model of parkinsonism. Neuropharmacology 51, 112–120.

CrossRef Full Text

Murakami, T., Shoji, M., Imai, Y., Inoue, H., Kawarabayashi, T., Matsubara, E., Harigaya, Y., Sasaki, A., Takahashi, R., and Abe, K. (2004). Pael-R is accumulated in Lewy bodies of Parkinson’s disease. Ann. Neurol. 55, 439–442.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Nicklas, W. J., Vyas, I., and Heikkila, R. E. (1985). Inhibition of NADH-linked oxidation in brain mitochondria by 1-methyl-4-phenyl-pyridine, a metabolite of the neurotoxin, 1-methyl-4-phenyl-1,2,5,6-tetrahydropyridine. Life Sci. 36, 2503–2508.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Nishikawa, K., Li, H., Kawamura, R., Osaka, H., Wang, Y. L., Hara, Y., Hirokawa, T., Manago, Y., Amano, T., Noda, M., Aoki, S., and Wada, K. (2003). Alterations of structure and hydrolase activity of parkinsonism-associated human ubiquitin carboxyl-terminal hydrolase L1 variants. Biochem. Biophys. Res. Commun. 304, 176–183.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Novelli, A., Reilly, J. A., Lysko, P. G., and Henneberry, R. C. (1988). Glutamate becomes neurotoxic via the N-methyl-d-aspartate receptor when intracellular energy levels are reduced. Brain Res. 451, 205–212.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Nussbaum, R. L., and Ellis, C. E. (2003). Alzheimer’s disease and Parkinson’s disease. N. Engl. J. Med. 348, 1356–1364.

Pubmed Abstract | Pubmed Full Text

Olanow, C. W., and McNaught, K. S. (2006). Ubiquitin-proteasome system and Parkinson’s disease. Mov. Disord. 21, 1806–1823.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Olzmann, J. A., Brown, K., Wilkinson, K. D., Rees, H. D., Huai, Q., Ke, H., Levey, A. I., Li, L., and Chin, L. S. (2004). Familial Parkinson’s disease-associated L166P mutation disrupts DJ-1 protein folding and function. J. Biol. Chem. 279, 8506–8515.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Omelchenko, N., and Sesack, S. R. (2005). Laterodorsal tegmental projections to identified cell populations in the rat ventral tegmental area. J. Comp. Neurol. 483, 217–235.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Paisan-Ruiz, C., Jain, S., Evans, E. W., Gilks, W. P., Simon, J., van der Brug, M., Lopez de Munain, A., Aparicio, S., Gil, A. M., Khan, N., Johnson, J., Martinez, J. R., Nicholl, D., Carrera, I. M., Pena, A. S., de Silva, R., Lees, A., Marti-Masso, J. F., Perez-Tur, J., Wood, N. W., and Singleton, A. B. (2004). Cloning of the gene containing mutations that cause PARK8-linked Parkinson’s disease. Neuron 44, 595–600.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Pakkenberg, H., Andersen, B. B., Burns, R. S., and Pakkenberg, B. (1995). A stereological study of substantia nigra in young and old rhesus monkeys. Brain Res. 693, 201–206.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Palacino, J. J., Sagi, D., Goldberg, M. S., Krauss, S., Motz, C., Wacker, M., Klose, J., and Shen, J. (2004). Mitochondrial dysfunction and oxidative damage in parkin-deficient mice. J. Biol. Chem. 279, 18614–18622.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

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

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Parker, W. D. Jr., Parks, J. K., and Swerdlow, R. H. (2008). Complex I deficiency in Parkinson’s disease frontal cortex. Brain Res. 1189, 215–218.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Paschen, W., and Mengesdorf, T. (2005). Endoplasmic reticulum stress response and neurodegeneration. Cell Calcium 38, 409–415.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Pearce, R. K., Hawkes, C. H., and Daniel, S. E. (1995). The anterior olfactory nucleus in Parkinson’s disease. Mov. Disord. 10, 283–287.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Pickart, C. M. (2004). Back to the future with ubiquitin. Cell 116, 181–190.

Pubmed Abstract | Pubmed Full Text

Plaitakis, A., and Shashidharan, P. (2000). Glutamate transport and metabolism in dopaminergic neurons of substantia nigra: implications for the pathogenesis of Parkinson’s disease. J. Neurol. 247 (Suppl. 2), II25–II35.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Polymeropoulos, M. H., Lavedan, C., Leroy, E., Ide, S. E., Dehejia, A., Dutra, A., Pike, B., Root, H., Rubenstein, J., Boyer, R., Stenroos, E. S., Chandrasekharappa, S., Athanassiadou, A., Papapetropoulos, T., Johnson, W. G., Lazzarini, A. M., Duvoisin, R. C., Di Iorio, G., Golbe, L. I., and Nussbaum, R. L. (1997). Mutation in the alpha-synuclein gene identified in families with Parkinson’s disease. Science 276, 2045–2047.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Poon, H. F., Frasier, M., Shreve, N., Calabrese, V., Wolozin, B., and Butterfield, D. A. (2005). Mitochondrial associated metabolic proteins are selectively oxidized in A30P alpha-synuclein transgenic mice--a model of familial Parkinson’s disease. Neurobiol. Dis. 18, 492–498.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Pridgeon, J. W., Olzmann, J. A., Chin, L. S., and Li, L. (2007). PINK1 protects against oxidative stress by phosphorylating mitochondrial chaperone TRAP1. PLoS Biol. 5, e172. doi: 10.1371/journal.pbio.0050172.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Przedborski, S., and Jackson-Lewis, V. (1998). Mechanisms of MPTP toxicity. Mov. Disord. 13 (Suppl. 1), 35–38.

Pubmed Abstract | Pubmed Full Text

Ramirez, A., Heimbach, A., Grundemann, J., Stiller, B., Hampshire, D., Cid, L. P., Goebel, I., Mubaidin, A. F., Wriekat, A. L., Roeper, J., Al-Din, A., Hillmer, A. M., Karsak, M., Liss, B., Woods, C. G., Behrens, M. I., and Kubisch, C. (2006). Hereditary parkinsonism with dementia is caused by mutations in ATP13A2, encoding a lysosomal type 5 P-type ATPase. Nat. Genet. 38, 1184–1191.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Rizzuto, R., and Pozzan, T. (2006). Microdomains of intracellular Ca2+: molecular determinants and functional consequences. Physiol. Rev. 86, 369–408.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Rochet, J. C., Outeiro, T. F., Conway, K. A., Ding, T. T., Volles, M. J., Lashuel, H. A., Bieganski, R. M., Lindquist, S. L., and Lansbury, P. T. (2004). Interactions among alpha-synuclein, dopamine, and biomembranes: some clues for understanding neurodegeneration in Parkinson’s disease. J. Mol. Neurosci. 23, 23–34.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Rodriguez, M., Barroso-Chinea, P., Abdala, P., Obeso, J., and Gonzalez-Hernandez, T. (2001). Dopamine cell degeneration induced by intraventricular administration of 6-hydroxydopamine in the rat: similarities with cell loss in parkinson’s disease. Exp. Neurol. 169, 163–181.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Rodriguez-Pallares, J., Rey, P., Parga, J. A., Muñoz, A., Guerra, M. J., and Labandeira-Garcia, J. L. (2008).Brain angiotensin enhances dopaminergic cell death via microglial activation and NADPH-derived ROS. Neurobiol. Dis. 31, 58–73.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Saito, Y., Ruberu, N. N., Sawabe, M., Arai, T., Kazama, H., Hosoi, T., Yamanouchi, H., and Murayama, S. (2004). Lewy body-related alpha-synucleinopathy in aging. J. Neuropathol. Exp. Neurol. 63, 742–749.

Pubmed Abstract | Pubmed Full Text

Salazar, M., Sokoloski, T. D., and Patil, P. N. (1978). Binding of dopaminergic drugs by the neuromelanin of the substantia nigra, synthetic melanins and melanin granules. Fed. Proc. 37, 2403–2407.

Pubmed Abstract | Pubmed Full Text

Sanchez-Ferrer, A., Rodriguez-Lopez, J. N., Garcia-Canovas, F., and Garcia-Carmona, F. (1995). Tyrosinase: a comprehensive review of its mechanism. Biochim. Biophys. Acta. 1247, 1–11.

Pubmed Abstract | Pubmed Full Text

Schapira, A. H. (2008). Mitochondria in the aetiology and pathogenesis of Parkinson’s disease. Lancet Neurol. 7, 97–109.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Schapira, A. H., Cooper, J. M., Dexter, D., Clark, J. B., Jenner, P., and Marsden, C. D. (1990). Mitochondrial complex I deficiency in Parkinson’s disease. J. Neurochem. 54, 823–827.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Schon, E. A., and Manfredi, G. (2003). Neuronal degeneration and mitochondrial dysfunction. J. Clin. Invest. 111, 303–312.

Pubmed Abstract | Pubmed Full Text

Schuler, F., and Casida, J. E. (2001). The insecticide target in the PSST subunit of complex I. Pest. Manag. Sci. 57, 932–940.

CrossRef Full Text

Schultz, W. (2002). Getting formal with dopamine and reward. Neuron 36, 241–263.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Semchuk, K. M., Love, E. J., and Lee, R. G. (1992). Parkinson’s disease and exposure to agricultural work and pesticide chemicals. Neurology 42, 1328–1335.

Pubmed Abstract | Pubmed Full Text

Shah, I. M., and Di Napoli, M. (2007). The ubiquitin-proteasome system and proteasome inhibitors in central nervous system diseases. Cardiovasc. Hematol. Disord. Drug Targets 7, 250–273.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Shendelman, S., Jonason, A., Martinat, C., Leete, T., and Abeliovich, A. (2004). DJ-1 is a redox-dependent molecular chaperone that inhibits alpha-synuclein aggregate formation. PLoS Biol. 2, e362. doi: 10.1371/journal.pone.0001884.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Sherer, T. B., Kim, J. H., Betarbet, R., and Greenamyre, J. T. (2003). Subcutaneous rotenone exposure causes highly selective dopaminergic degeneration and alpha-synuclein aggregation. Exp. Neurol. 179, 9–16.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Shima, T., Sarna, T., Swartz, H. M., Stroppolo, A., Gerbasi, R., and Zecca, L. (1997). Binding of iron to neuromelanin of human substantia nigra and synthetic melanin: an electron paramagnetic resonance spetroscopy study. Free Radic. Biol. Med. 23, 110–119.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Shimada, S., Kitayama, S., Walther, D., and Uhl, G. (1992). Dopamine transporter mRNA: dense expression in ventral midbrain neurons. Brain Res. Mol. Brain Res. 13, 359–362.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Shimura, H., Hattori, N., Kubo, S., Yoshikawa, M., Kitada, T., Matsumine, H., Asakawa, S., Minoshima, S., Yamamura, Y., Shimizu, N., and Mizuno, Y. (1999). Immunohistochemical and subcellular localization of Parkin protein: absence of protein in autosomal recessive juvenile parkinsonism patients. Ann. Neurol. 45, 668–672.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Silvestri, L., Caputo, V., Bellacchio, E., Atorino, L., Dallapiccola, B., Valente, E. M., and Casari, G. (2005). Mitochondrial import and enzymatic activity of PINK1 mutants associated to recessive parkinsonism. Hum. Mol. Genet. 14, 3477–3492.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Singleton, A. B., Farrer, M., Johnson, J., Singleton, A., Hague, S., Kachergus, J., Hulihan, M., Peuralinna, T., Dutra, A., Nussbaum, R., Lincoln, S., Crawley, A., Hanson, M., Maraganore, D., Adler, C., Cookson, M. R., Muenter, M., Baptista, M., Miller, D., Blancato, J., Hardy, J., and Gwinn-Hardy, K. (2003). alpha-Synuclein locus triplication causes Parkinson’s disease. Science 302, 841.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Snyder, H., Mensah, K., Theisler, C., Lee, J., Matouschek, A., and Wolozin, B. (2003). Aggregated and monomeric alpha-synuclein bind to the S6’ proteasomal protein and inhibit proteasomal function. J. Biol. Chem. 278, 11753–11759.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Squier, T. C. (2001). Oxidative stress and protein aggregation during biological aging. Exp. Gerontol. 36, 1539–1550.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Stefanis, L., Larsen, K. E., Rideout, H. J., Sulzer, D., and Greene, L. A. (2001). Expression of A53T mutant but not wild-type alpha-synuclein in PC12 cells induces alterations of the ubiquitin-dependent degradation system, loss of dopamine release, and autophagic cell death. J. Neurosci. 21, 9549–9560.

Pubmed Abstract | Pubmed Full Text

Striessnig, J., Koschak, A., Sinnegger-Brauns, M. J., Hetzenauer, A., Nguyen, N. K., Busquet, P., Pelster, G., and Singewald, N. (2006). Role of voltage-gated L-type Ca2+ channel isoforms for brain function. Biochem. Soc. Trans. 34, 903–909.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Sulzer, D., Bogulavsky, J., Larsen, K. E., Behr, G., Karatekin, E., Kleinman, M. H., Turro, N., Krantz, D., Edwards, R. H., Greene, L. A., and Zecca, L. (2000). Neuromelanin biosynthesis is driven by excess cytosolic catecholamines not accumulated by synaptic vesicles. Proc. Natl. Acad. Sci. U.S.A. 97, 11869–11874.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Sulzer, D., and Schmitz, Y. (2007). Parkinson’s disease: return of an old prime suspect. Neuron 55, 8–10.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Surmeier, D. J. (2007). Calcium, ageing, and neuronal vulnerability in Parkinson’s disease. Lancet Neurol. 6, 933–938.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Swartz, H. M., Sarna, T., and Zecca, L. (1992). Modulation by neuromelanin of the availability and reactivity of metal ions. Ann. Neurobiol. 32 Suppl. S69–S75.

CrossRef Full Text

Swerdlow, R. H., Parks, J. K., Miller, S. W., Tuttle, J. B., Trimmer, P. A., Sheehan, J. P., Bennett, J. P. Jr., Davis, R. E., and Parker, W. D. Jr. (1996). Origin and functional consequences of the complex I defect in Parkinson’s disease. Ann. Neurol. 40, 663–671.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Taira, T., Saito, Y., Niki, T., Iguchi-Ariga, S. M., Takahashi, K., and Ariga, H. (2004). DJ-1 has a role in antioxidative stress to prevent cell death. EMBO Rep. 5, 213–218.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Takahashi, K., Taira, T., Niki, T., Seino, C., Iguchi-Ariga, S. M., and Ariga, H. (2001). DJ-1 positively regulates the androgen receptor by impairing the binding of PIASx alpha to the receptor. J. Biol. Chem. 276, 37556–37563.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Takahashi, N., Miner, L. L., Sora, I., Ujike, H., Revay, R. S., Kostic, V., Jackson-Lewis, V., Przedborski, S., and Uhl, G. R. (1997). VMAT2 knockout mice: heterozygotes display reduced amphetamine-conditioned reward, enhanced amphetamine locomotion, and enhanced MPTP toxicity. Proc. Natl. Acad. Sci. U.S.A. 94, 9938–9943.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Tanaka, Y., Engelender, S., Igarashi, S., Rao, R. K., Wanner, T., Tanzi, R. E., Sawa, A., V, L. D., Dawson, T. M., and Ross, C. A. (2001). Inducible expression of mutant alpha-synuclein decreases proteasome activity and increases sensitivity to mitochondria-dependent apoptosis. Hum. Mol. Genet. 10, 919–926.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Taylor, T. N., Caudle, W. M., Shepherd, K. R., Noorian, A., Jackson, C. R., Iuvone, P. M., Weinshenker, D., Greene, J. G., and Miller, G. W. (2009). Nonmotor symptoms of Parkinson’s disease revealed in an animal model with reduced monoamine storage capacity. J. Neurosci. 29, 8103–8113.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Thannickal, T. C., Lai, Y. Y., and Siegel, J. M. (2007).Hypocretin (orexin) cell loss in Parkinson’s disease. Brain 130, 1586–1595.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Turski, W. A., and Lachowicz, L. (1991). The possibility of participation of different protein kinases in the regulation of degradation of Glp6[125I]Tyr8SP6-11 with different endopeptidases in subcellular fractions of rat brain. Int. J. Biochem. 23, 317–321.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Uhl, G. R., Walther, D., Mash, D., Faucheux, B., and Javoy-Agid, F. (1994). Dopamine transporter messenger RNA in Parkinson’s disease and control substantia nigra neurons. Ann. Neurol. 35, 494–498.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Valente, E. M., Salvi, S., Ialongo, T., Marongiu, R., Elia, A. E., Caputo, V., Romito, L., Albanese, A., Dallapiccola, B., and Bentivoglio, A. R. (2004). PINK1 mutations are associated with sporadic early-onset parkinsonism. Ann. Neurol. 56, 336–341.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Varastet, M., Riche, D., Maziere, M., and Hantraye, P. (1994). Chronic MPTP treatment reproduces in baboons the differential vulnerability of mesencephalic dopaminergic neurons observed in Parkinson’s disease. Neuroscience 63, 47–56.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Victorof, J., Zarow, C., Mack, W. J., Hsu, E., and Chui, H. C. (1996). Physical aggression is associated with preservation of substantia nigra pars compacta in Alzheimer disease. Arch. Neurol. 53, 428–434.

Pubmed Abstract | Pubmed Full Text

Wickens, J. R., Reynolds, J. N., and Hyland, B. I. (2003). Neural mechanisms of reward-related motor learning. Curr. Opin. Neurobiol. 13, 685–690.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Wilms, H., Rosenstiel, P., Romero-Ramos, M., Arlt, A., Schafer, H., Seegert, D., Kahle, P. J., Odoy, S., Claasen, J. H., Holzknecht, C., Brandenburg, L. O., Deuschl, G., Schreiber, S., Kirik, D., and Lucius, R. (2009). Suppression of MAP kinases inhibits microglial activation and attenuates neuronal cell death induced by alpha-synuclein protofibrils. Int. J. Immunopathol. Pharmacol. 22, 897–909.

Pubmed Abstract | Pubmed Full Text

Wilms, H., Rosenstiel, P., Sievers, J., Deuschl, G., Zecca, L., and Lucius, R. (2003). Activation of microglia by human neuromelanin is NF-kappaB dependent and involves p38 mitogen-activated protein kinase: implications for Parkinson’s disease. FASEB J. 17, 500–502.

Pubmed Abstract | Pubmed Full Text

Wood-Kaczmar, A., Gandhi, S., Yao, Z., Abramov, A. Y., Miljan, E. A., Keen, G., Stanyer, L., Hargreaves, I., Klupsch, K., Deas, E., Downward, J., Mansfield, L., Jat, P., Taylor, J., Heales, S., Duchen, M. R., Latchman, D., Tabrizi, S. J., and Wood, N. W. (2008). PINK1 is necessary for long term survival and mitochondrial function in human dopaminergic neurons. PLoS One 3, e2455. doi: 10.1371/journal.pone.0002455.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Xue, C. J., Ng, J. P., Li, Y., and Wolf, M. E. (1996). Acute and repeated systemic amphetamine administration: effects on extracellular glutamate, aspartate, and serine levels in rat ventral tegmental area and nucleus accumbens. J. Neurochem. 67, 352–363.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Yamada, T., McGeer, P. L., Baimbridge, K. G., and McGeer, E. G. (1990). Relative sparing in Parkinson’s disease of substantia nigra dopamine neurons containing calbindin-D28K. Brain Res. 526, 303–307.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Zafar, S. K., Siegel, D., and Ross, D. (2006). A potential role for cyclized quinones derived from dopamine, DOPA, and 3, 4-dihydroxyphenylacetic acid in proteasomal inhibition. Mol. Pharmacol. 70, 1079–1086.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Zarow, C., Lyness, S. A., Mortimer, J. A., and Chui, H. C. (2003). Neuronal loss is greater in the locus coeruleus than nucleus basalis and substantia nigra in Alzheimer and Parkinson diseases. Arch. Neurol. 60, 337–341.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Zeevalk, G. D., Bernard, L. P., and Nicklas, W. J. (1998). Role of oxidative stress and the glutathione system in loss of dopamine neurons due to impairment of energy metabolism. J. Neurochem. 70, 1421–1430.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Zecca, L., Zucca, F. A., Wilms, H., and Sulzer, D. (2003). Neuromelanin of the substantia niga: a neuronal black hole with protective and toxic characteristics. Trends Neurosci. 26, 578–580.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Zhang, L., Shimoji, M., Thomas, B., Moore, D. J., Yu, S. W., Marupudi, N. I., Torp, R., Torgner, I. A., Ottersen, O. P., Dawson, T. M., and Dawson, V. L. (2005). Mitochondrial localization of the Parkinson’s disease related protein DJ-1: implications for pathogenesis. Hum. Mol. Genet. 14, 2063–2073.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Zhang, Y., Gao, J., Chung, K. K., Huang, H., Dawson, V. L., and Dawson, T. M. (2000). Parkin functions as an E2-dependent ubiquitin-protein ligase and promotes the degradation of the synaptic vesicle-associated protein, CDCrel-1. Proc. Natl. Acad. Sci. U.S.A. 97, 13354–13359.

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Keywords: Parkinson’s disease, neurodegeneration, aging, differential vulnerability, nigrostriatal, mesolimbic, mitochondrion, proteostasis

Citation: González-Hernández T, Cruz-Muros I, Afonso-Oramas D, Salas-Hernández J and Castro-Hernández J (2010) Vulnerability of mesostriatal dopaminergic neurons in Parkinson’s disease. Front. Neuroanat. 4:140. doi: 10.3389/fnana.2010.00140

Received: 07 July 2010; Accepted: 24 September 2010;
Published online: 20 October 2010

Edited by:

Jose L. Lanciego, University of Navarra, Spain

Reviewed by:

Thomas Wichman, Emory University, USA
Jose L. Labandeira-Garcia, Universidade de Santiago de Compostela, Spain

Copyright: © 2010 González-Hernández, Cruz-Muros, Afonso-Oramas, Salas-Hernández and Castro-Hernández. This is an open-access article subject to an exclusive license agreement between the authors and the Frontiers Research Foundation, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are credited.

*Correspondence: Tomás González-Hernández, Department of Anatomy, Faculty of Medicine, University of La Laguna, 38207 La Laguna, Tenerife, Spain. e-mail: tgonhern@ull.es

Download