Next Article in Journal
Different Impact of Beta-Blockers on Long-Term Mortality in Heart Failure Patients with and without Chronic Obstructive Pulmonary Disease
Next Article in Special Issue
Blood Neurofilament Light Chain in Parkinson’s Disease: Comparability between Parkinson’s Progression Markers Initiative (PPMI) and Asian Cohorts
Previous Article in Journal
First-Line Anaplastic Lymphoma Kinase (ALK) Inhibitors for ALK-Positive Lung Cancer in Asian Populations: Systematic Review and Network Meta-Analysis
Previous Article in Special Issue
Evolution of Quality of Life in Persons with Parkinson’s Disease: A Prospective Cohort Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Current Knowledge on the Background, Pathophysiology and Treatment of Levodopa-Induced Dyskinesia—Literature Review

by
Michał Hutny
1,*,
Jagoda Hofman
1,
Aleksandra Klimkowicz-Mrowiec
2 and
Agnieszka Gorzkowska
3
1
Students’ Scientific Society, Department of Neurorehabilitation, Faculty of Medical Sciences in Katowice, Medical University of Silesia, 40-752 Katowice, Poland
2
Department of Internal Medicine and Gerontology, Faculty of Medicine, Medical College, Jagiellonian University, 30-688 Kraków, Poland
3
Department of Neurorehabilitation, Faculty of Medical Sciences, School of Medicine, Medical University of Silesia, 40-752 Katowice, Poland
*
Author to whom correspondence should be addressed.
J. Clin. Med. 2021, 10(19), 4377; https://doi.org/10.3390/jcm10194377
Submission received: 3 August 2021 / Revised: 2 September 2021 / Accepted: 22 September 2021 / Published: 25 September 2021

Abstract

:
Levodopa remains the primary drug for controlling motor symptoms in Parkinson’s disease through the whole course, but over time, complications develop in the form of dyskinesias, which gradually become more frequent and severe. These abnormal, involuntary, hyperkinetic movements are mainly characteristic of the ON phase and are triggered by excess exogenous levodopa. They may also occur during the OFF phase, or in both phases. Over the past 10 years, the issue of levodopa-induced dyskinesia has been the subject of research into both the substrate of this pathology and potential remedial strategies. The purpose of the present study was to review the results of recent research on the background and treatment of dyskinesia. To this end, databases were reviewed using a search strategy that included both relevant keywords related to the topic and appropriate filters to limit results to English language literature published since 2010. Based on the selected papers, the current state of knowledge on the morphological, functional, genetic and clinical features of levodopa-induced dyskinesia, as well as pharmacological, genetic treatment and other therapies such as deep brain stimulation, are described.

1. Introduction

Treatment of Parkinson’s disease (PD) is inseparable from the use of levodopa, l-3,4-dihydroxyphenylalanine (L-DOPA). Physiologically, it is a precursor molecule in the synthesis of catecholamines that act as neurotransmitters in nervous system—epinephrine, norepinephrine (NE) and dopamine (DA). Decreased DA levels, associated with a degenerative process in the substantia nigra (SN), are the trigger for the cardinal motor features of PD, such as bradykinesia and rigidity [1]. The first symptoms of PD appear when DA levels in the striatum are reduced by 80%, with a corresponding neuronal DA loss of 60% [2]. During this period, treatment is in “the honeymoon phase” [1]. L-DOPA administration alleviates motor symptoms of PD, such as bradykinesia and rigidity, without the presence of abnormal involuntary movements (AIMs) [1]. As the disease progresses, DA biosynthesis and storage decreases. When the dose of L-DOPA entering the brain across the blood–brain barrier exceeds the dopamine storage capacity of dopaminergic neurons, there is an increased dopaminergic activity, leading to hyperkinetic movements—levodopa-induced dyskinesias (LIDs) [1]. They occur at the peak dose, with an increase or decrease in the drug concentration, or both, which is referred to as biphasic LID. As the duration of L-DOPA therapy increases, the level of DA required to induce LID decreases [3].
Motor complications in PD patients develop after varying lengths of therapy. The prevalence of LIDs varies between the sources [4,5,6]; however, about 30% of PD patients develop dyskinesia in the first 5 to 6 years of L-DOPA treatment [5,6]. This fact naturally raises the question, what are the determinants of their occurrence and severity? Another important question is whether there are any methods by which PD patients with LID can be helped to improve their quality of life. This article attempts to answer these questions by reviewing the literature published on this topic in the last decade.

2. Materials and Methods

Database searches (PubMed, PubMed Central, Medline) were conducted over a 4-month period, from November 2020 to February 2021. The search strategy was based on creating input keywords consisting of “levodopa-induced dyskinesia” and the following related terms: “genetic”, “therapeutic”, “background”, “pathogenesis”, “therapy”, “treatment”, “deep brain stimulation”, “dopamine receptor”, “dopaminergic activity” and “signal”. The filters applied to the results limited them to English language papers published since 2010. Reviews, systematic reviews, personal communication, letters to the editor, conference material and case reports were not included in the study. The online Mendelian Inheritance in Men database was used as a reference for gene loci and terminology.

3. Results

The findings were divided into sections on background (structural and functional; genetic; clinical) and treatment, divided into preclinical studies (Section 3.5) and clinical trials (Section 3.6). The criteria were the purpose of the studies—studies describing the involvement of neuroreceptors, genetic and clinical factors in the development of dyskinesias and the corresponding anatomical and functional changes in brain tissue were classified in the background section, while studies focusing on measuring the efficacy and safety of therapeutic strategies in controlling dyskinesias were classified in the treatment section.
It must be noted, however, that the studies presented below were performed on different subjects. Dyskinesia in animal models may be very different from dyskinesia in humans, which limits the translation of the results in animal studies to practice in humans. The animal subjects that resemble human PD and the LID phenotype the most are primates lesioned with methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP). They present symptoms typical for human PD—rigidity, postural tremor, akinesia—which can be controlled using L-DOPA treatment [7,8]. MPTP primates also experience non-motor symptoms of PD, such as cognitive deficits and sleep disturbances [9]. A limitation of this study model is that the subjects may develop a different severity of parkinsonism (mild vs. advanced) depending on the response to MPTP lesioning [8]. Another issue, which concerns not only MTPT primates, but also 6-hydroxydopamine (6-OHDA)-lesioned rodent models, is the fact that lesioning is an acute induction of pathology in animal, which contrasts with the progressive and chronic nature of the neurodegenerative pathology that is PD [10]. The other most popular animal models are mentioned before 6-OHDA lesioned rodent, including rats and mice. Due to obvious differences in morphology between humans and rodents, dyskinetic movements in the latter case must be measured using different tests. Dyskinesia in a rodent is usually measured with the AIMs scale, measuring forelimb (hyperkinetic movements of contralateral forelimb), axial (twisted posture), orolingual (empty jaw movements) and locomotive (circular movements) dyskinesia [11]. L-Dopa treatment in these animals is reflected by the development of dyskinesia [12]. The topography of lesioning in a PD model rodent is also an important issue, as the medial forebrain bundle lesioning results in creating the subjects with severe and evenly distributed DA denervation, whereas lesioning of the striatum leads to a heterogeneous distribution of DA denervation [13]. Bearing all the above in mind, the readers are advised to use Table S1, (see Supplementary Materials: “Table S1: Summary of reviewed articles concerning the studied subjects and their parkinsonian/dyskinetic phenotypes.”), which clearly describes each enrolled study concerning the type of studied subjects and their “parkinsonian” and “dyskinetic” phenotype.

3.1. Structural and Functional Background

3.1.1. Grey and White Matter

From the very beginning of PD, magnetic resonance imaging (MRI) shows anatomical changes in the brain. These involve white matter hyperintensities (WMHs), and grey matter changes in the size of striatum. Both of these phenomena were examined for the significance of their intensity to the development of LID.
In animal model studies comparing severely dyskinetic 6-OHDA-lesioned rats with saline treated rats, it was shown that the striatum size as well as the asymmetry between the sides increased with the increasing duration of L-DOPA treatment, with a predominance on the lesioned side. The degree of asymmetry between sites in the size of the dorsal striatum components, putamen and caudate nucleus, at the onset of PD, was found to be a predictor of the development of LID, but also of the response to L-DOPA treatment [3]. The microscopic changes, although evident, did not appear to be directly related to striatal growth or the axial, limb and orolingual AIMs scores, measured together using established criteria of amplitude and severity [14].
The primary motor cortex (M1) in animals has also been shown to undergo changes associated with PD progression. In the rats that developed LID, an increase in the dendritic spike density was observed that was significantly correlated with the severity of LID. The dendrites of tested neurons mainly received glutamatergic excitatory signals; therefore, the increased density of dendritic spines was thought to reflect the development of hypersensitivity to glutamatergic input in these neurons. These results suggest a role for M1 neurons in the development of LID [15]. In humans, although M1 volume was significantly negatively correlated with the development and severity of diphasic LID, no such relationship could be proven for density; therefore, only volume can be considered as a potential marker of diphasic dyskinesia in patients [16].
Lesioning the striatum of rats decreases the dopaminergic innervation in M1 ipsilateral to the lesion. It is represented by the loss of tyrosine hydroxylase fibres, whose density in the lesioned side is 75% lower than in the side contralateral to the lesion. Lesioning also influences the levels of M1 monoamines and their metabolites. The dopamine and norepinephrine levels are decreased, but administration of L-DOPA restores the level of DA back to the regular values [17].
A retrospective analysis of 484 patients showed an association between WMHs and LID in patients with PD. The results showed that periventricular, lobar, basal ganglia and infratentorial hyperintensities were associated with the occurrence of dyskinesia. Although such a result provides evidence of an association between lesions and LID, it also indicates that any WMHs’ location is not dominant in influencing the development of LID [18].

3.1.2. Basal Ganglia

Neurons of basal ganglia: the internal globus pallidus (GPi) and subthalamic nucleus (STN) present an altered firing rate in dyskinetic patients. Both grouped discharges, as well as low frequency firing, are significantly more frequently found in this group [19]. In MPTP parkinsonian non-human primates, chronic treatment with L-DOPA alters the activity of basal ganglia without changing the striatal DA release [20]. It also increases interactions between the STN and the entopeduncular nucleus, which, in rodents, is a structure analogous to human GPi [21]. Therefore, it is possible that changes in the firing pattern of basal neurons contribute to the development of LID [19,20]. Although the electrophysiological parameters of STN are not associated with the degree of severity of AIMs, the attenuation of rodent STN rebalances the activity of D1R and D2R signalling pathways, which results in the decreased severity of LID [22].

3.1.3. Cerebellum

Cerebellar circuits are specifically addressed in studies of LID development. The functional connectivity (FC) of the cerebellum and other parts of the brain at rest was studied in groups of PD patients without LID, PD patients with LID and controls. The functional connectivity between the motor cerebellum and posterior cortical and cerebellum in PD patients is higher than in non-PD patients. The presence of LIDs is positively correlated with an increase in the FC between lobule VIIIb and the left inferior frontal gyrus [23]. Cerebellar involvement in dyskinesia is also associated with red nucleus (RN) iron levels. This parameter is elevated in the PD patients with dyskinesia, whereas it remains unchanged in the PD patients without dyskinesias. As the RN is involved in cerebellar circuits, it is thought that the iron level in RN reflects an increased activity of cerebellar pathways. Their increased activity may be a compensating mechanism in response to the development of LID [24].

3.1.4. Circulation and Angiogenesis

Local cerebral blood flow (CBF) is increased under hypercapnia in areas of vascular proliferation. D1R activity stimulates vasodilatation and angiogenesis [25]. Abnormally elevated levels of the putaminal cerebral glucose metabolism rate (CMR) are observed during an OFF phase in both dyskinetic and non-dyskinetic patients. These high CMR values decrease after the administration of L-DOPA. It also causes an increase in putaminal CBF values in 75% of dyskinetic PD patients and in 50% of non-dyskinetic PD patients. The difference between the two groups of patients, although it exists, is not significant enough; therefore, neither CBF nor CMR can be considered as an indicator of changes leading to the development of LID in patients with PD. On the other hand, the parameter that significantly differentiates dyskinetic and non-dyskinetic patients is the resting activity of the primary somatosensory cortex and M1. In addition, the topography of dyskinetic symptoms reflects cortical areas that show abnormal CBF and CMR baseline values [26]. Although there is no significant difference in the CBF and CMR values, these parameters can be further used to determine the putaminal hyper-perfusion/hypometabolism index (PHI). It compares the ON-phase putaminal perfusion and glucose metabolism with values measured for the thalamus, which is a structure similar in size to the putamen that is not affected by D1R activity. The PHI reflects abnormal activity of the putamen after the administration of L-DOPA. The presence of LID in PD patients is associated with higher PHI values. It is possible that this parameter can be used as a marker to assess the risk of developing LID in patients with PD [27].

3.1.5. Alterations in Cerebrospinal Fluid

Cerebrospinal fluid (CSF) analysis is one of the basic neurological examinations providing a very useful insight into the CNS state. Significant differences in the DA and NA turnover rate in CSF are observed between dyskinetic and non-dyskinetic PD patients (higher in the dyskinetic group). In PD patients not treated with L-DOPA, a decrease in the CSF concentrations of the following metabolites: dihydroxyphenylacetic acid (DOPAC) and homovanillic acid, is observed. Physiologically, DA concentrations increase as the age-dependent DOPAC/DA ratio decreases. These observations suggest DA in CSF, as well as the concentrations of their metabolites, are potential prognostic markers for the development of LID [28], although they need confirmation in future studies.

3.1.6. Metabolic Changes

A DA deficiency results in an impaired metabolism of neurotransmitters—mainly glutamate (Glu), glutamine and aspartate in the following specific brain regions: midbrain and right cortex. Parkinsonian and dyskinetic rats show lower levels of synaptophysin compared to control animals, which may indicate reduced synaptic plasticity. Synaptophysin is a glycoprotein of presynaptic vesicle membranes, responsible for the docking and fusion of neurotransmitter-carrying vesicles. Its levels were shown to decrease with age and to be reduced in neurological disorders such as PD and Alzheimer’s disease [29]. Lower levels of synaptophysin and postsynaptic density protein 95 (PDS95) are correlated with reduced dendritic spine density [30], which impairs synaptic plasticity [31]. Changes in astrocyte activity, resulting from elevated myo-inositol levels, may also contribute to the development of dyskinesia [32].

3.2. Genetic Background

The latest technologies offered by genetics and molecular science have been used to study the genetic determinants of dyskinesias in Parkinson’s disease. Single nucleotide polymorphism (SNP) in the GALNT14 (2p23.1; MIM 608225) gene (rs144125291) [33], as well as in the gene for the GRIN2A subunit of the NMDA receptor (rs7192557 and rs8057394) [34] are significantly correlated with LID [33]. GALNT14 encodes information concerning one of the Golgi proteins involved in glycosylation. It may, therefore, be involved in the basal levels of neuroinflammation in the brain [33]. The MAOB gene polymorphism (MIM 309860)—rs1799836, G allele—is associated with an increased risk of LID development, when compared to the A allele [35]. The levels of long non-coding RNA (lncRNA) molecules, such as NONRATT023402, are altered in dyskinetic rats. The regulatory targets of this lncRNA molecule are the Gsto2 (10q25.1; MIM 612314) and Ptger3 (1p31.1; MIM 176806) genes, which are associated with PD and LID [36].
As the DA levels decrease during PD, gene expression is altered in the spiny projection neurons (SPNs) of the direct (dSPNs) and indirect (iSPNs) pathways. These changes are particularly evident in the genes regulating MAPK signalling [37]. The difference between dSPNs and iSPNs in the level of expression of Ap-1, ERK and cyclic AMP-responsive element binding protein-dependent genes is clear, with more changes occurring in dSPNs. Cellular homeostatic mechanisms counteract these changes, but without an effect on Ap-1-dependent gene expression [37]. The genes associated with the development of AIMs—Sstr2 (17q25.1; MIM 182452), Sstr4 (20p11.21; MIM 182454) and Kcnn3 (1q21.3; MIM 602983)—are overexpressed in dSPNs [37].
Chronic L-DOPA treatment leads to changes in the DNA methylation activity of parkinsonian rats by increasing the expression levels of genes for DNA demethylases, Tet3 (2p13.1; MIM 613555) and Gadd45b (19p13.3; MIM 604948). These epigenetic changes correlate with the downregulation of other genes’ expression, as a result of prolonged L-DOPA administration. They are also associated with an increased expression of genes involved in synaptic plasticity [38]. Bromodomain and extraterminal family proteins bind to acetylated histones, thus playing an effector role in transcriptional regulation. Their dysfunction has been associated with the reorganisation of dorsal striatal chromatin and with changes in corticostriatal plasticity [39].

3.3. Alterations of Neurotransmission

A dopamine transporter (DAT) is a presynaptic membrane molecule responsible for transporting DA to neuronal cells. The lower activity of DAT membrane expression is thought to reflect the depletion of dopaminergic neurons [40]. The expression level can be measured using single-photon emission computed tomography, where the quantity of DATs is reflected by the specific binding ratio (SBR) of a radiotracer [41]. Using the same technique, it is possible to examine changes to the pre- and post-synaptic membranes of dopaminergic neurons. It was shown that the uptake of a tracer for the DAT decreased significantly more in PD patients than in healthy controls. The condition of the post-synaptic membrane, evaluated using a radiotracer for D2R and D3R, did not differ between the PD patients and the healthy controls [42]. Measurement of the putaminal SBRs of the radiotracer for the DAT in a 4-year follow-up study showed a significant reduction in signal in the PD patients who developed dyskinesia [43].
Using a functional MRI technique, it is possible to investigate which connections between the cortex and striatum are modulated by chronic L-DOPA treatment and evaluate the effect of the dopaminergic modulation of connections on the severity of LID in patients with established dyskinesia, as well as in non-dyskinetic PD patients [44]. An acute dose of L-DOPA triggers changes in dopaminergic modulation (via feedback connections from the putamen to the M1 and presupplementary area (preSMA)), although the extent of the provoked response differed significantly between the patients with and without LID. Increased signalling between the putamen and M1 is thought to account for the inadequate ‘Go’ response observed in NoGo trials [45]. No-Go activity in the preSMA has also been shown to be a predictive measure for the severity of dyskinesia. Given the overactivity of putamen in LID, this structure may be a key factor in the development of dyskinesia [46].
Dyskinesia can be induced without administration of L-DOPA—by the reactivation of striatal neurons. As an animal study shows, inhibition of their activity thus leads to a lack of dyskinetic response to L-DOPA. Neurons in striatum are mainly SPNs, but the γ-aminobutyric acid (GABA)-ergic fast spiking interneurons (FSIs) and iSPNs are also present in this brain region [47]. The activity of dSPNs decreases with dopamine depletion in the midbrain, resulting in disproportion in the activity of dSPNs and iSPNs [48,49]. After the administration of L-DOPA, the activity of dSPNs increases again, while the activity of iSPNs decreases, leading to a restoration of balance [48,49]. In LID, the effect of L-DOPA on the activity of both SPN populations is enhanced [48,49]. The activation of dSPNs alone is sufficient to induce LID, as demonstrated by laser light provocation. Both selective D1R and D2R agonists activate a comparable number of striatal neurons, and both induce LID, although selective D1 agonists trigger more vigorous changes in the glow rate and more severe dyskinetic symptoms [48]. Interestingly, the antiparkinsonian effect of L-DOPA is also mediated to a greater extent by D1R than D2R [50]. The activity patterns of SPNs are altered in the PD and LIDs of MPTP parkinsonian primates. Dopamine depletion is associated with an increased burst spike frequency or pause loss. Due to the complexity of interactions between dopaminergic and glutamatergic systems, the exact mechanism leading to these changes remains unclear. As mentioned earlier, dopaminergic signalling provokes a response that is different for the SNP subpopulations, resulting in an imbalance of their activity [51].
The striatum and cortex of a parkinsonian brain are bidirectionally connected in rats [52]. As the disease enters an advanced stage, these connections change their characteristics [53]. It is unclear whether the connections become unidirectional [54] or whether the observed imbalances are due to the increased strength of one direction relative to the other [52]. Dyskinetic events are associated with a frequency peak of 80 Hz, in both an animal model and humans [52,54,55]. Long-term potentiation (LTP) and long-term depression (LTD) are mechanisms that contribute to synaptic plasticity. Long-term potentiation occurs in indirect pathways in parkinsonian mice and in direct pathways during dyskinesias, whereas LTD occurs in direct pathways in parkinsonian mice and in indirect pathways during dyskinesias [54]. Thus, dyskinesia is not triggered by impaired LTP induction, but results from a reduced excitability difference between dSPNs and iSPNs during the ON-phase [54,55,56].

3.3.1. Inhibitors of Monoamine Oxidase B and Catechol-O-Methyltransferase

Both of these enzymes are one of the key components of DA metabolism. Their inhibitors are widely used as adjuvant therapy in the treatment of PD with L-DOPA [57]. Identification of the SNPs of genes for catechol-O-methyltransferase [58] and monoamine oxidase B (MAO-B) [59] was thought to be useful for estimating the activity of these enzymes, which, in turn, influences DA metabolism and, indirectly, dyskinetic symptoms [58,59].

3.3.2. Dopaminergic Pathways and Dopamine Receptors

The results of the studies on rodent parkinsonian model animals showed that three distinct signalling pathways—PKA/DARPP-32, ERK and mTORC1—play an important role in the mechanism of LID development [60,61]. They are all activated by a common intracellular cascade that is triggered in striatonigral SPNs expressing D1Rs [60,61]. Long-term L-DOPA administration increases the activity in these pathways, leading to changes in striatal bidirectional synaptic plasticity [61]. The activation of one pathway may affect the activity of the other pathways [61]. DARPP-32 signalling modulates the ERK and mTORC1 pathways [60,61]
Recent studies have closely examined the effect of D1R activation on ERK1/2 through interaction with the tyrosine phosphatase Shp-2 [62]. Not only has Shp-2 activation been shown to be a key factor for ERK1/2 and mTOR activation via phosphorylation and, consequently, for LID expression, but it has also been demonstrated that long-term stimulation with intermittent L-DOPA or D1R agonists leads to Shp-2 phosphorylation. This effect can be counteracted by D1R antagonists, consequently influencing the levels of the D1R/Shp-2 pathway end products—p-mTOR and p-ERK1/2 [62]. Rapamycin was shown to selectively affect mTORC1 in the striatum without affecting either the PKA/DARPP-32 and ERK pathways or the expression of the α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor (AMPA) and NMDA receptor subunits [61].
In lesioned animal models, the potential of D1R to associate with PKA and ERK1/2 was increased, and this change occurred concomitantly with the onset of LID [63]. However, within 2 weeks of treatment with L-DOPA, D1R was not able to activate the PKA and ERK1/2 pathways. These observations indicate a remodelling of the direct striatal pathways that results from the oversensitivity of dSPNs [63]. The rate of D1R-dependent ERK1/2 phosphorylation is modulated by mGluR5, which alters the mGluR5/PLC/PKC pathway through the modulation of calcium/dependent responses. However, it does not influence the PKA activity [64]. The hypersensitivity of the D1R/cAMP/PKA pathway is associated with the elevation of the Gαolf protein level. This protein is the major striatal stimulating G-protein, whose activity leads to increased cAMP levels. Interestingly, despite the decreased Gαolf protein levels in mutant animals, the D1R/cAMP/PKA pathway was found to be hypersensitive in these subjects. Therefore, the activity of Gαolf may not be essential for the hypersensitivity of this pathway [65]. Chronic L-DOPA administration is associated with increased levels of Gαolf protein in striatonigral SPNs of parkinsonian mice, and a decreased level of this protein in striatopallidal SPNs [66].
Knocking-out casein kinase 2 (CK2) is also associated with reduced Golf levels in striatonigral SPNs [67]. The knockout (KO) of CK2 in striatonigral neurons in model animals results in a decreased severity of dyskinesia. This also correlates with lower pERK levels in dSPNs and cholinergic aspiny interneurons (ChIs). When CK2 is knocked-out in striatopallidal neurons, the severity of dyskinesia and the pERK levels in dSPNs and ChIs increase. However, in the latter case, the LID can be attenuated by coadministration of caffeine with L-DOPA [67].
An increased expression of Nurr1 transcription factor in striatal neurons of dyskinetic rats is associated with the increased cortically evoked firing rate and the increased spine density of SNPs [68]. Both downstream D1R/D2R activation pathways, PKA and ERK1/2, are influenced by the activity of leucine-rich repeat kinase 2 (LRRK2). The interactions between these pathways and LRRK2 are not yet well understood, but their further investigation may result in a new therapeutic strategy to modulate the downstream activity of dopaminergic receptors, leading to better control of LID [69]. The dopaminergic system in the midbrain is also regulated by the activity of vesicular glutamate transporter type 3 in a circadian-dependant matter [70].
The surface distribution of dopaminergic receptors is dictated by anchoring protein-PSD95 [71]. Its levels are increased in dopamine-depleted brains, causing lower diffusion of D1R in the membrane. The association between DA loss and increased levels of PSD95 might be involved in the development of dyskinesia in non-human parkinsonian primates [72]. The activity of parkin, an E3 ubiquitin-protein ligase, is impaired not only in genetic parkinsonism, but also in sporadic PD patients. In animal models, the loss of function of this enzyme leads to the sequestration of ubiquitinated proteins. It is associated with D1R-mediated decreased proteasome catalytic activity, which is reflected by the severity of AIMs [73].
Ca2+/calmodulin-dependent protein kinase IIα (CaMKIIα) binds to the intracellular domain of D2R. The Gαi domain of D2R is also localised in this region [74]. This may indicate that CaMKIIα affects D2R activity in the adenylyl cyclase signalling pathway. In parkinsonian rats, the level of interaction between CaMKIIα and D2R is elevated. Decreasing its level leads to the attenuation of LID with an efficacy similar to D2R agonists [75].
An increased D3R response is observed in animals with severe dyskinesia, which is associated with an increased GABA release. This differs significantly from animals with lower AIMs scores. D3R can affect the ERK pathway independently of D1R, thus adding up to the effect of D1R modulation [76]. The deletion of D3R is associated with decreased levels of FosB, ERK and H3 activity. It also causes an attenuation of dyskinesia, with preservation of the therapeutic effect of L-DOPA [77]. The main locus of D3R, whose levels are elevated in dyskinetic patients, is the GP [78]. A lack of D5 signalling in D5R animals is associated with the decreased therapeutic effect of L-DOPA and the increased expression of LID. As LID increases, LTD plasticity is reduced in these animals. Decreased acetylcholine (ACh) release also negatively affects the activation of M1 postsynaptic receptors in dSPNs [79].

3.3.3. GTPases of Ras Family

Raf/Mek/ERK is a pathway that activates the ERK signalling cascade. It is stimulated by Ras GTPases, whose activity is increased by DA and glutamate receptors. Decreasing the striatal levels of Ras-guanine nucleotide-releasing factor (Ras-GFR1) leads to inhibition of the Ras/ERK pathway in parkinsonian mice. Ras-GFR1 affects the Ras-ERK pathway only in dSPNs, without affecting the cholinergic interneurons of ChIs [80]. It, therefore, affects specific downstream D1R pathways, consequently inhibiting the development of hypersensitivity to DA stimulation in the neurons. This is associated with LID developing less, although dopaminergic stimulation retains its therapeutic effect on parkinsonian motor symptoms [80,81].
The administration of drugs that modify the MEK-ERK core components further potentiates the aforementioned antidyskinetic effect [82]. Eliminating Ras-GRF1 decreases ERK, PKA and mTOR signalling intensity [83]. Consequently, the Ras-GRF1 levels in striatonigral SPNs also affect LTP-dependent synaptic plasticity. Such a relationship was not observed in striatopallidal SPNs, which may be due to the involvement of DARPP-32 or Ras-GRF2 activity [84]. The Ras-ERK pathway in mice can be overactive due to the overexpression of Ras-GRF1, or due to the reduced gene expression of the Ras inhibitor (Nf+/−). Interestingly, in both cases, the severity of dyskinesia does not exceed normal levels. However, the administration of lovastatin, a Ras inhibitor, improves the control of dyskinetic symptoms only in the group with normal Nf+/− gene expression [85].

3.3.4. FosB Transcription Factor

Chronic treatment with L-DOPA leads to an increased level of the truncated splice variant of FosB (ΔFosB). Among others, this transcription factor regulates the genes responsible for plasticity in striatal neurons. Elevated levels of ΔFosB are associated with a loss of FosB function, which results in the early development of dyskinesia and unstable firing rate changes during the ON phase in L-DOPA-treated animals [86]. MSH1 histone kinase activity also increases with ΔFosB levels, consistent with MSK1′s target, histone H3. H3 is localised on c-fos-associated nucleosomes, and its phosphorylation results in an increased expression of this gene [81]. In SPNs expressing D1R, MSK1 also decreases the Polycomb-group-proteins-dependent gene repression in parkinsonian mice [87].
The level of c-Fos expression in M1 neurons differs between sides in an animal model of unilateral MFB injury after acute (p = 0.003) or chronic (p = 0.001) L-DOPA treatment, although it is elevated when compared to the control. It is also positively correlated with the total AIMs score (p = 0.13). Another immediate-early gene product, ARC, which is associated with synaptic plasticity, showed no such correlation [17]. Both of these genes are overexpressed in the dorsolateral bed nucleus of the terminal stria (dlBST) in dyskinetic rats. This structure is not included in the basal ganglia, and as such, has not been implicated in the development of LID. Interestingly, inhibition of ∆FosB expression in dlBST results in a reduced AIMs score [88]. The neuronal NO synthase inhibitor 7-nitroindazole also downregulates FosB/ΔFosB expression and reduces dyskinetic symptoms in a rodent model [89].

3.3.5. Glutamatergic Signalling

Neurotransmission systems, other than those that are dopaminergic, also affect SPNs and contribute to the development and expression of LID. One of these is the glutamatergic system. As the duration of L-DOPA therapy increases, NMDAR activity and its composition changes [90]. Interactions of Rph3A protein with the GluN2A subunit of NMDARs are increased in parkinsonian dyskinetic rats. Altering the PSD95 function leads to improved AIMs scores in animals. This may be a promising target for translation to humans as an antidyskinetic therapy, despite the potential post-synaptic interactions involving PSD95 [91]. This protein also contributes to the overactivity of NMDAR in dyskinetic animals. It modulates interactions between Fyn kinase and the NR2B subunit [92]. AMPA antagonists directly acting on the glutamate receptors in M1 are effective in reducing the severity of dyskinesia. NMDA antagonists did not alter dyskinesia in parkinsonian rats, although simultaneous action on NMDA and AMPA receptors increased the effect of AMPA antagonism [93].

3.3.6. Cholinergic Signalling

The density of nicotine cholinergic receptors (nAChRs) in the striatum of dyskinetic PD patients remains normal, in contrast to the reduced levels of nAChRs in non-dyskinetic patients [94]. This indicates that cholinergic activity is preserved in LID, which may be a mechanism of overcoming DA loss in striatal neurons. Cholinergic activity induces DA release from otherwise serotoninergic neurons. The regulatory mechanisms of dopaminergic transmission do not affect the above DA release, as evidenced by the lack of significant differences in the measurement of DAT binding in the striatum between dyskinetic and non-dyskinetic patients [94].
The metabotropic muscarinic cholinergic receptors (mAChRs) of PD dyskinetic rats, specifically M1R and M4R, are involved in LID expression. Interestingly, M4R has a bidirectional function—endogenous ACh inhibits LID and striato-nigral SPN activation via M4R stimulation, whereas during LID, it facilitates its effect on postsynaptic M4 mAChRs. The exogenous stimulation of mAChRs in M4, presumably at presynaptic dopaminergic terminals or ChIs, also potentiates an inhibitory effect on LID [95]. The activation of M4Rs by endogenous ACh potentiates LTD and inhibits LTP in the dSPN glutamatergic synapses of animal models. Thus, the function of M4Rs is similar to that of D2R in iSPN. The exogenous stimulation of M4R increases synaptic plasticity in dSPNs and thereby attenuates LID [96].

3.3.7. Serotonergic Signalling

Serotonergic neurons are capable of converting 5-HT into DA, storing the produced DA, and then releasing it. This function is referred to as the presynaptic serotonergic LID mechanism [97,98]. Surgical or pharmacological lysis of serotonin raphe projections in dyskinetic rats has been shown to block LID, providing evidence for a key role of dopamine released from serotonergic neurons in the pathogenesis of LID [99]. The observed greater severity of dyskinesia in animals with a large number of serotonergic endings in the striatum is not due to DA release per se but rather to an abnormal activation pattern (with a predominance of D1R over D2R) and lack of an adequate DA gradient [100]. A study in rodent and primate models of PD, as well as post-mortem studies of brain tissue examination, have shown significant upregulation of the 5-HT transporter (SERT) in the striatum of subjects experiencing LID. The expression of dyskinesia was correlated with a radiotracer binding to SERT [97]. Brain neurotrophic factor (BDNF) overexpression, which can be caused by L-DOPA treatment, has been shown to correlate with axonal growth of serotonin neurons in the striatum and putamen of dyskinetic rats, which was further associated with the increased activity of the cAMP/PKA and mTORC pathways [101]. In the DA-depleted striata of PD model monkeys, SERT+ axons were significantly denser in the sensorimotor part of the striatum. The synaptic incidence was also significantly increased in the area of dopaminergic denervation [102]. The above results suggest that synaptic plasticity and proliferation of serotonergic neurons, which follow the dopaminergic denervation and further L-DOPA treatment, may be important factors in the development of motor symptoms in PD, as well as in LID [97,101,102]. In dyskinetic rats, selective serotonin reuptake inhibitors (SSRIs) such as citalopram, paroxetine and fluoxetine show a dose-dependent antidyskinetic effect when administered together with L-DOPA. In contrast, no such effect was observed in animals treated with apomorphine, a direct DA agonist. This indicates a presynaptic mechanism of antidyskinetic action [103,104]. L-DOPA administration does not affect 5-HT receptor activity, although it reduces the effect of SSRI on the neuron firing rate in the dorsal raphe nucleus (DRN) [105].

3.4. Clinical and Environmental Background

Certain non-modifiable factors and clinical features are associated with an earlier onset or higher severity of LID. Female gender is an independent risk factor for developing LID. Patients whose tremor predominates over rigidity and akinesia have been shown to have a lower risk of LID [3], but these data are inconclusive. In another study, patients who initially had a tremor had a significantly higher risk of LID during follow-up [106]. The course of treatment also influences the characteristics of LID in PD. A higher mean dose of L-DOPA during the first 6 months of therapy [6], as well as greater severity of motor symptoms [107], are significantly associated with an early onset of LID. These observations are mutually consistent—patients who initially have more severe parkinsonian motor symptoms require higher doses of L-DOPA to restore their normal motor function [3]. Moreover, the need for more rapid escalation of the drug dose is a direct reflection of the rapid degeneration of the dopaminergic system and correlates with an earlier onset of LID [106]. It has been pointed out, however, that it is not clear whether L-DOPA administration is a contributing factor to the development of LID itself, or rather just a trigger [3]. Nonetheless, a less-pulsatile administration of L-DOPA, e.g., 6× doses per day, has been shown in clinical trials to reduce the risk of LID development in both PD patients who have never previously initiated L-DOPA treatment and in non-dyskinetic patients treated with L-DOPA [108]. A continuous supply of L-DOPA via an enteral levodopa-carbidopa gel reduces fluctuations in DA levels. In patients with dyskinetic symptoms lasting more than 4 h/day, a reduction in LID duration, severity and accompanying pain was observed after the administration of a levodopa-carbidopa enteral gel. In patients with a duration of dyskinetic episodes less than 4 h/day, the duration of dyskinesias increased, albeit with a significant reduction in severity and pain [109].
Adenosine A2A receptor antagonists may reduce the severity of dyskinesia, which is associated with greater receptor availability in PD patients who develop LID [110]. Caffein, as an A2A antagonist, may be a factor delaying the onset of LID [111]. The N-of-1 study on nicotine use demonstrated a significant effect of nicotine on the occurrence of LID compared to a placebo [112].

3.5. Treatment-Preclinical Studies

3.5.1. Modification of L-DOPA Delivery and Release

As LIDs are motor symptoms caused by pulsatile changes in DA concentration due to intermittent L-dopa administration, the obvious strategy to prevent them is the continuous delivery and release of the drug (continuous dopamine stimulation; CDS)—this way, theoretically, the DA concentration will reach a stable level at which the therapeutic, antiparkinsonian function is active without LID side effects.
L-DOPA/benserazide poly(lactic-glucosic acid) (LBM) microspheres are a novel therapeutic option for the treatment of PD. In animals, LBM showed a dose-dependent antiparkinsonian effect and did not significantly increase AIMs scores. In control animals treated with L-DOPA, the D1R/Shp-2/ERK1/2 pathway activity increased with duration and onset of dyskinesia, which was not observed in LBM-treated animals [113]. These results are consistent with a previous study that also demonstrated the efficacy of LBM therapy in reducing LID’s frequency [114]. Abnormal PKA/DARPP-32 signalling leads to the increased phosphorylation of tau protein, which in turn activates SPNs and results in LID. PKA/DARPP-32 and tau protein phosphorylation activity was increased in the mice with parkinsonism treated with standard i.p. L-DOPA/benserazide (LS) therapy. A dose-dependent inhibition of PKA/DARPP-32 and tau protein phosphorylation was observed in the LBM-treated animals [115].
Striatal PKA-related glutamatergic signalling and glutamatergic receptors (AMPA) may play an important role in the expression and occurrence of LIDs. During chronic repeated L-DOPA treatment, the AMPA receptor subunit, GluR1, becomes over-phosphorylated at the serine-831 and serine-845 position. Such changes (elevated levels of pGluR1S831 and pGluR1S845) are not found in LBM-treated animals [116].
An alternative method of L-DOPA delivery is chitosan-coated nanoliposomes (CCN). ERK1/2, DARPP-32 and FosB/ΔFosB expression and LID induction are significantly less elevated in parkinsonian rats treated with CCN, compared to the same animal model treated with LS. However, these parameters in PD rats treated with CCN are higher than in a control group of saline-treated animals [117].

3.5.2. Amantadine Counteracts LID

Amantadine, an antiparkinsonian drug, is used to reduce LID. In a study conducted in an animal model, the administration of amantadine decreased the severity of LID episodes. Interestingly, GABA release in SN, which was increased after L-DOPA treatment and during AIMs, increased significantly less when L-DOPA was administered concomitantly with amantadine. This suggested that the mechanism of the antidyskinetic effect of amantadine may be related to an effect on nigral GABA levels [118]. Recent studies have shown that amantadine inhibits dyskinesia by blocking the inward-rectifying type 2 K+ channel in SPNs rather than NMDA receptors [57].

3.5.3. Serotonin in the Treatment of LID

The serotoninergic system is largely present in the basal ganglia. Chronic L-DOPA treatment and a decreased DA concentration leads to its alterations [119]. It is speculated that drugs that modify 5-HT activity may have an inhibitory effect on LID. However, they require careful administration because an overdose of serotonin agonists leads to 5-HT syndrome, the symptomatology of which is similar (tremor and rigidity) to PD [120]. The 5-HT1A partial agonist buspirone inhibits LID in rats by influencing the 5-HT1A receptors, although it does not alter STN electrical activity [121]. It also regulates GABA and Glu release and burst activity in pars reticular of SN (SNr) [122].
The 5-HT1A agonist 8-hydroxy-2-dipropylaminotetralin (8-OH-DPAT) significantly reduces LID, but this benefit is outweighed by the 5-HT syndrome that is caused by 8-OH-DPAT [123]. Nonetheless, 8-OH-DPAT delays LID development [123] and reduces elevated cortical gamma activity that was caused by D1 and D2 agonism [124]. The efficacy of SSRI (citalopram) in inhibiting dyskinesia in rats is similar to 8-OH-DPAT and buspirone, but it does not induce significant 5-HT syndrome [125]. LID inhibition by another 5-HT1A agonist, BMY-14802, increases adequately with the dose (20 > 10 > 5 mg/kg). The antidyskinetic effect of BMY-14802 was observed not only in LID induced by L-DOPA administration, but also in LID induced by the administration of D1 and D2 agonists. However, this effect can be reversed by using 5-HT1A receptor antagonists in L-DOPA-induced LID [126].
The 5-HT3 antagonism was another serotoninergic system checkpoint that was investigated for its potential to reduce LID. In rats, the first significant differences in the AIMs scores were observed after 9 days of treatment with ondansetron (5-HT3 antagonist) and L-DOPA, compared with animals treated with L-DOPA alone. This difference steadily increases. Interestingly, ondansetron administration was discontinued on day 23, resulting in significant differences between the L-DOPA and ondansetron and the control groups on day 30. These results suggest the potential of 5-HT3 antagonists, such as ondansetron, in controlling LID in the treatment of PD, although further research is necessary in order to understand this correlation [127].
The serotonin precursor molecule, 5-hydroxy-tryptophan (5-HTP), has also been studied to control LID. The use of 5-HTP increases the cytoplasmic levels of serotonin, which, by competing with DA in serotonergic neurons, decreases the amount of DA released by these neurons [128,129]. Dyskinetic rats with 6-OHDA damage were administered different doses of 5-HTP, L-DOPA, L-DOPA and 5-HTP or L-DOPA and 5-HTP and a 5-HT antagonist. After evaluating the motor functions of the test animals, 5-HTP was shown to have antidyskinetic potential. It also inhibited the development of dyskinesias in parkinsonian rats treated from the beginning with the combination of L-DOPA and 5-HTP. The control of parkinsonism symptoms by L-DOPA was not worsened when 5-HTP was used [128].
Finasteride, a 5-α-reductase inhibitor, is involved in neurosteroid synthesis [130]. This is a group of molecules involved in the physiology of nervous tissue and the pathophysiology of its diseases, including PD [131]. Neurosteroids interact with the dorsal raphe nucleus and influence serotonergic activity [132], also by regulating the DA release from serotonergic neurons [130]. Since the latter is an important factor in the development of LID, finasteride has been studied in dyskinetic rats for its antidyskinetic potential. The results of this study suggest that it may be effective in controlling LID in both acute and chronic treatment, alone or in combination with L-DOPA [130].

3.5.4. β-Adrenoceptor Blockade—Propranolol

To date, the modulation of LID by adrenergic compounds has emphasised the α-adrenergic receptor (αAR), despite the high density of β-adrenergic receptors (βAR) in the striatum of PD patients [133]. The optical isomer (−) of propranolol affects βARs in the dorsal striatum and is the isomer responsible for reducing LID, with a two-fold longer antidyskinetic effect compared with (±) compounds [134]. Although propranolol alone can reduce the motor activity of rats, which is a highly undesirable effect in the treatment of PD, there is a broad doses spectrum in which this reduction does not occur, with a concomitant significant effect on LID. Supportive therapy with propranolol normalises aberrant signalling in the striatum of rats [134]. Propranolol could not attenuate LID induced by D1 and D2 receptors agonists, indicating that propranolol acts through a presynaptic mechanism [135]. Due to the association between the reduction in striatal DA, propranolol administration and the reduction in LID reduction, it was suggested that the antidyskinetic mechanism of propranolol may be related to the reduction in striatal DA levels. It may be caused by the secondary anticholinergic effect of propranolol. An inhibition of βARs results in a decrease in intracellular cAMP, which leads to a decrease in the spontaneous firing rate of ChIs, resulting in a decrease in ACh levels. Striatal DA release is regulated by ACh, and therefore, the effect of propranolol on ACh levels could possibly restore the balance between ACh and DA in LID, which should be examined in the further studies [136].

3.5.5. Nicotinic Receptor and Its Agonists

The acetylcholine receptors (AChR) of nigrostratial neurons modulate neurotransmission, including dopaminergic signalling [137]. Nicotine, a direct agonist of nAChR, has been shown to significantly reduce AIMs scores after 8 (p < 0.01) and 12 (p < 0.05) weeks of administration in rats. Although acute nicotine administration stimulates nAChR, thereby increasing dopamine release [138], intermittent nicotine treatment desensitises nAChR. It also decreases dopamine release, which may explain its antidyskinetic effect—in later stages of PD, DAT levels are reduced, resulting in a less efficient clearance of excess DA [139]. Interestingly, a more recent experiment showed no significant improvement in dyskinesia in female mice after a 10-week treatment of nicotine, as well as in the group treated for 5 days with α7 nAChR partial agonist (product name: AZD0328). Brain-derived neurotrophic factor levels in the striatum, but not in the prefrontal cortex, were shown to be significantly correlated with the development of LID. Nicotine failed to significantly reduce high BDNF levels in the striatum but succeeded at the prefrontal cortex level. It was suggested that the reason for the lack of modulation of LID by the drugs tested is due to the condition of the animals—their nigrostriatal damage was almost complete, indicating that the treatment of LID with nicotine requires a less damaged nigrostriatal pathway [140].

3.5.6. Ionotropic Receptors of Glu

The previously described increase in glutamatergic transmission in the development of PD and LID indicates that its inhibition appears to be an attractive therapeutic approach for the treatment of LID. As the disease progresses, an increase in the NMDA subunits, NMDA-R1 and NMDA-R2, acting as an ion channel and regulatory unit, respectively, becomes apparent [90]. Studies on Compound Formula Rehmannia showed its inhibitory effect on NMDA-R1 and NMDA-R2 expression. It also increased the expression of GABA-RB1, which becomes down-regulated during LID. This effect was correlated with the decrease in LID in test animals [141]. Memantine, a non-competitive NMDAR antagonist, exhibits a similar mechanism through which it attenuates LID in rat models of PD [142].

3.5.7. Histamine H2 Receptor Antagonism

Given the widespread distribution of H2 receptors in the basal ganglia involved in the development of LID, both in their input and output regions [143], it is conceivable that H2 receptors may be targets for antidyskinetic therapy. Chronic treatment with ranitidine reduces LID expression in 6-OHDA rats treated with L-DOPA. A suggested mechanism of action of ranitidine to reduce dyskinesias involves the inhibition of the PKA pathway in the SPNs [144]. Ranitidine inhibits striatal Glu efflux, as well as GABA release in SNr, thus affecting the pre- and postsynaptic mechanisms of LID development. The pre-treatment of rats with ranitidine reduces AIMs scores, without reducing the efficacy of anti-PD treatment [145].

3.5.8. Opioid Modulators

As opioids affect neurotransmission in the basal ganglia and the intensity of opioid signalling is significantly altered during PD [146], drugs that modulate this transmission system appear to be an effective alternative to standard antidyskinetic therapy. Treatment with a combination of L-DOPA and nalbuphine (κ receptor agonist and μ receptor antagonist) reduce the expression of ΔFosB in the striatum of MPTP-lesioned parkinsonian primates, as well as other LID markers, such as prodynorphin, dynorphin A, Cdk5, Thr34-phosphorylation of DARPP-32, to normal values. Drug administration did not cause any AEs, and even the sedative effect present in nalbuphine monotherapy was not induced by L-DOPA administration [147]. The significant results of this experiment may be due to the dualistic nature of the treatment strategy used—the highly selective µ-opioid receptor antagonism did not improve dyskinesia in the tested 6-OHDA rats [148]. Only μ receptor agonists appeared to reduce the incidence of LID in MPTP parkinsonian primates when used alone [149].
Nociceptin/orphanin FQ opioid peptide (NOP) receptors, despite some functional similarities, differ from opioid receptors—they are not activated by the most standard opioid receptor ligands. Although both NOP agonists, AT-390 and AT-403, exhibited dose-dependent overall hypolocomotor effects in parkinsonian rats, AT-403 produced a significant, albeit mild, improvement in dyskinesia [150].

3.5.9. cAMP and cGMP Signalling–Inhibitors of Phosphodiesterase 10A

Phosphodiesterase 10A (PDE10A) is an enzyme expressed exclusively in brain tissue, including the SPNs expressing D1R and D2R [151]. Hydrolysing cyclic nucleotides cAMP and cGMP affects DA signalling in SPNs, showing an effect similar to D2R antagonists. In PD patients, PDE10A expression is significantly reduced in the striatum and GP. Due to the presence of PDE10A in both populations of SNPs, it is possible that its theoretical antidyskinetic effect could exceed that shown by D2R antagonists [152]. In MPTP parkinsonian primates, a novel PDE10A inhibitor (product name: MR1916) showed high penetration into the brain and antidyskinetic effects greater than amantadine over the entire dose range—not only with acute but also with chronic administration. It also did not cause significant AEs [153].

3.5.10. Genetic Treatment

Another therapeutic approach to LID focuses on modifying the expression of LID-related genes. Considering the latest techniques used in molecular sciences, this area appears to be a very promising and effective strategy, and genes related to the dopaminergic system and dopaminergic pathways seem to be the first choice. In parkinsonian rats, interesting results were obtained by the overexpression of DA autoreceptor D2Rs in the DRN, which is involved in the previously described uncontrolled uptake and release of L-DOPA in serotonergic neurons. The DRN, in which D2Rs were overexpressed, was insensitive to high, LID-inducing doses of L-DOPA and dopaminergic agonists. This demonstrated that genetic control of serotonergic DRN neurons could be used to reduce dyskinesia in studied animals [154]. Inhibition of the p11 gene (1q21.3; MIM 114085) in the dorsal striatum of parkinsonian mice produced surprising results. Its product, S100A10, is known to activate the serotonin 5-HT1B receptor, which, in turn, is associated with decreased LID. Thus, downregulating p11 expression should rather increase LID, and, in fact, it inhibited dyskinesia more effectively than the direct pharmacological activation of 5-HT1B. These results suggested that the effect of p11 on LID development may function through mechanisms not involving 5-HT1B [155]. The transplantation of dopaminergic neurons has also been investigated for the effect of L-DOPA on graft survival. Both allogeneic and xenogeneic transplants were viable but showed no significant improvement of motor functions in the test animals. Furthermore, post-transplantation L-DOPA treatment induced an immune response in xenogeneic grafts [156]. It has been shown that ERK hypersensitivity, secondary to striatal damage in rats, can be reduced by overexpression of G protein-coupled receptor kinase (GRK) (two-fold). This also resulted in decreased Akt activity and decreased ΔFosB levels in the striatum. Increasing GRK expression with lentivirus may be a potential strategy to reduce dyskinesia, although this strategy requires further examination [157]. A direct target of GRK, a G protein-coupled receptor, is also β-arrestin2, a protein scaffolding various intracellular molecules [158]. In rats with β-arrestin2 overexpression, a significant reduction in LID efficiency for L-DOPA was observed, without affecting the therapeutic efficacy of L-DOPA. Knocking out β-arrestin2 had the opposite effect of increasing dyskinesia [159,160]. Knocking out adenylyl cyclase 5 (AC5) appeared to be another option to modulate downstream dopaminergic pathways. A significant attenuation of LID was observed in AC5 KO mice, coupled with decreased activation of the PKA pathway, resulting in a decreased ERK level, as well as lower FosB/ΔFosB expression [161].
Knocking out the α5 subunit of nAChR in 6-OHDA-lesioned mice leads to the attenuation of dyskinesia in these animals. Compared with wild-type individuals, KO mice showed less dopaminergic denervation in SN, as manifested by reduced rotational movements after amphetamine administration. DAT activity was also lower in modified individuals, resulting in less 6-OHDA neurotoxicity [162].
As mentioned earlier, CaMKIIα is involved in LID expression [75]. Modulating its function through inhibitors such as KN-93 has been shown to reduce LID in rats. Although it did not show a significant antiparkinsonian effect, it did reduce pGluR1S845 levels and Gad1 (2q31.1; MIM 605363) and Nur77 (12q13.13; MIM 139139) gene expression, all of which were increased after 3 weeks of L-DOPA treatment [163].
CaV1.3 calcium channels are present in striatal SPNs [164]. The administration of nimodipine (CaV1.3 antagonist) showed a reduction in LID in studies in dyskinetic animals. Recently the effects of CaV1.3 antagonism on LID expression have been studied more directly, by silencing the genes responsible for striatal CaV1.3 channels at the mRNA level. Even partial silencing of CaV1.3 in the striatum led to the complete prevention of LID. The method presented here was designed to specifically target the striatum, as otherwise, silencing CaV1.3 in other tissues could lead to serious cardiac, neurological and psychological consequences [165].
Peroxisome proliferator-activated receptors (PPAR) have been shown to be involved in the antidyskinetic effect of URB597 (an inhibitor of endocannabinoid catabolism) and capsazapine. Moreover, the direct activation of PPARγ led to a decrease in LID in parkinsonian rats, which was also correlated with an increase in ERK phosphorylation [166].

3.6. Treatment-Clinical Trials

3.6.1. Modification of L-DOPA Delivery and Release

As mentioned in the clinical background section, one of the CDS methods is the use of levodopa-carbidopa intestinal gel (LCIG) [109]. In patients with advanced PD, the standard of care is a 16-hour long infusion of LCIG, whose superiority over oral L-DOPA in improving the life quality and motor functions in these patients has been repeatedly described in various populations of observational studies and clinical trials (phase three) [167,168,169,170,171,172,173]. It has been shown to reduce dyskinesia by more than 40% [168]. Some recent studies have suggested that, in certain situations, such as sleep disturbances, the 24-hour LCIG infusion may be more beneficial in terms of sleep quality [174] and the improvement of daily time with dyskinesia (phase three) [175]. Previous studies have shown that approximately 30% of patients have experienced serious AEs associated with LCIG therapy [168,172]. Less severe AEs, such as abdominal pain, occurred in up to 87% of patients [168]. Although, in general, LCIG treatment had similar safety levels as standard oral L-DOPA treatment; it also carries a risk of complications associated with gastrointestinal tube insertion guidance [168,169,172,176,177,178]. It was also observed that initiating LCIG monotherapy bears a risk of developing diphasic dyskinesia in patients who have not experienced this type of dyskinesia before. This problem might be resolved by increased morning and continuous LCIG flow or by using ER L-DOPA or DA agonists during bedtime [179]. Despite the presence of the above AEs, LCIG can also be successfully used in older, advanced PD patients [180], although there is a significant risk of therapy discontinuation due to the presence of AEs [98].
Apomorphine was the first DR agonist introduced for the treatment of PD [181]. Continuous subcutaneous apomorphine injection (CSAI) is an apomorphine delivery method that has been used for nearly 40 years in many countries [182] as a monotherapy or adjunct therapy to standard oral L-DOPA therapy. The efficacy of CSAI in controlling dyskinesias varies between studies. It is generally considered a likely option in patients with advanced PD [182,183,184,185,186] and has been described as a transitional option between the standard treatment and DBS, providing better control of dyskinesia [181,187]. However, it has sometimes been pointed out, that the increase in LEDD due to combined treatment with oral L-DOPA and CSAI led to a lack of initial improvement in dyskinesias and was a reason for discontinuing therapy [181,187,188]. In contrast, this problem was not observed with CSAI used in monotherapy, and the positive effect on dyskinesias was beneficial [185].
An ER L-DOPA/carbidopa (CD-LD) (product name: IPX066) prolongs the duration of the ON phase without dyskinesia, shortens the duration of the OFF phase, and does not increase the severity of LIDs in patients studied during phase three clinical trials, compared with the treatment effects of immediate release (IR) CD-LD, and CD-LD with entacapone [189,190]. Concomitant treatment with CD-LD and DA agonists, MAO-B inhibitors, or amantadine does not significantly affect treatment outcomes, although small changes in treatment efficacy have been observed [190].

3.6.2. Amantadine Counteracts LID

Oral immediate-release amantadine is widely used in early PD as an alternative to L-DOPA treatment. In a meta-analysis, the anti-dyskinetic effect of amantadine was proven, also noting its vision-related AEs [191]. It is possible that amantadine may delay and consequently reduce the incidence of LID [192]. Despite its long half-life, it is usually administered in two daily doses, which, due to its high concentrations at night, may lead to sleep disturbances [193].
Extended-release amantadine (ERA) capsules (product name: ADS-5102) significantly (27%; p = 0.005) reduce dyskinesia and increase the duration of drug efficacy without dyskinesias in the experimental group compared to the control group (phase three trials) [194,195]. ERA has also been shown to improve MDS-UPDRS (phase two and three trials) [195,196]. The adversary effects (AEs) of long-term ERA therapy were dose-dependent and most commonly included falls, sleep disorders and hallucinations [196]. In most cases, they were not caused by the drug itself, but were due to the severity of the PD course [197]. NMDA-receptor antagonism caused hallucinations, and cholinergic antagonism caused symptoms such as dry mouth, constipation and nausea [194,197,198]

3.6.3. D3 Selective Agonism

Another proposed solution to control LID is to affect D1R activity by modulating D3R activity, either with D3 partial agonists or with D3 antagonists. Both receptors are co-expressed in dSPN and together are able to form a heteromeric D1R-D3R complex [77,199]. Overactivity of D3R, which is observed in dyskinesia [76], increases the membrane binding of D1R, consequently increasing the activity of downstream D1R pathways [200]. Studies in animal models have shown that partial D3 agonists are able to reduce the incidence of LID, which may be due to the restoration of the balance between D1R binding to the membrane and its internalisation [200]. Partial D3 agonists also did not reduce the efficacy of L-DOPA in reducing Parkinsonian symptoms, as observed with D3R antagonists [77,199,201]. Pramipexole is a non-ergoline agonist that prefers D3 receptors. Pramipexol monotherapy (achieved by gradually replacing previously used drugs with pramipexole) led to significant decrease in Core Assessment Program for Surgical Intervention Therapies scores compared with baseline values after 4 weeks of the treatment and was maintained at this level. Patients treated with pramipexole had a significant improvement of MDS-UPDRS part III (improvement in motor symptoms) and quality-of-life after 2 weeks of therapy. The only observed AEs were increased dyskinesia and the prolonged duration of the OFF phase. D3 agonism is a potential candidate for reducing LID in clinical practice because it combines the efficacy of PD treatment, LID control and safety of use. This may be due to the synergic effect of D1 and D3. The agonists preferring D3 may provide stimulation levels to overcome the hypokinetic symptoms of PD, but also low enough to not induce LID [202].
A new drug (product name: IRL790), a selective D3 receptor antagonist, has been proposed [203]. Dyskinetic primates show increased D3R expression in the striatum [201]; therefore, D3 receptor antagonism may be a potential strategy for reducing dyskinesia. D3 receptor antagonists, partial agonists and D3R deletion attenuated the occurrence of LID [77]. During the phase 1b clinical trial, treatment with IRL790 caused a significant number of AEs, with 90.1% of the patients taking D3-selective antagonists experiencing at least one AE, which was less than the AEs in the placebo group. The most common AE in the IRL790 group was worsening parkinsonism. Significantly more AEs were observed during dose titration of the drug than during the dose-size stabilisation phase. Treatment of efficacy was measured using the Unified Dyskinesia Rating Scale, showing an 11.5 point median reduction and an 8.2 point mean reduction in the intervention group compared to the placebo. The drug has been shown to be able to significantly reduce LID, although the concomitant presence of AEs must also be taken into account [203].

3.6.4. MAO-B Inhibition

Safinamide is relatively new drug whose proven mechanism of action is MAO-B inhibition [204]. It has been studied as a potential adjuvant therapy to standard L-DOPA treatment in PD, albeit in a limited population [204,205,206]. The results of these phase III studies did not allow for a firm conclusion on the efficacy of safinamide as an add-on therapy to L-DOPA in patients with advanced PD [204,205,206], neither did studies in a rodent model [207]. AEs associated with safinamide treatment such as cataracts, asthenia, fever, falls, back pain, dyskinesias, worsening of PD, headaches and insomnia have been observed in more than 10% of patients [204,205].

3.6.5. Serotonin in the Treatment of LID

The results of a phase I/IIa trial showed that treatment with 5-HT1A/B agonist, eltoprazine, at a 5.0 and 7.5 mg dose, significantly reduced Clinical Dyskinesia Rating Scale scores, although a significant reduction in the maximum LID severity was observed only at the 5.0 mg dose. No significant changes in MDS-UPDRS-III scale scores were observed. While the above therapeutic results are not as spectacular as those previously reported in animal models, such a difference was to be expected given the differences in the complexity of the substrate of LID in humans and rodents [208].
The antidyskinetic potential of 5-HTP was also studied in a small group of twelve patients with parkinsonism in a phase IIa clinical trial. The efficacy of the drug was measured using UPDRS Part III and IV scores, as well as UDysRS and patient self-report diaries. The results of a 16-week follow-up after treatment with 50 mg of 5-HTP showed significant improvements in all the study scores compared to baseline and the placebo group. Despite the promising results of this study, its limitations must be taken into account, and larger group studies are needed to adequately assess the antidyskinetic potential of 5-HTP in patients with parkinsonism [209].

3.6.6. Histamine H2 Receptor Antagonism

Although the results of ranitidine in animal models appear promising, a phase IIa clinical trial with famotidine in a human model showed no significant differences between the treatment effect of famotidine (doses: 80, 120, 160 mg/day) and a placebo. No significant AEs were observed in this study, demonstrating the safety of famotidine. The lack of clinically significant results in this experiment illustrates that the complexity of LID problems in humans cannot always be adequately represented in animal models, and that optimistic results from studies in rats or non-human primates may not necessarily translate to human experiments [210].

3.6.7. Invasive and Non-Invasive Brain Stimulation

Deep brain stimulation (DBS) is an invasive treatment option for advanced PD, in which the therapeutic window is increasingly narrow and pharmacological control of the disease is increasingly difficult. DBS can control both motor symptoms of parkinsonism and dyskinesias [211]. Its primary target is the STN, although the efficacy of stimulation of this area in controlling LID and motor symptoms of PD has recently been compared with the stimulation of GPi [211,212]. Stimulation of the STN is widely used in clinical practice, and as shown in recent studies, is more effective in reducing the equivalent daily dose of L-DOPA (LEDD) compared to GPi-DBS [211,213]. Although the DBS-STN is usually used to treat advanced PD, it was also studied in a phase four clinical trial in PD patients with mild motor complications and a short (1.5 years) history of dyskinesia [214]. The use of DBS-STN resulted in a better treatment outcome. However, AEs were also more frequent than in the pharmacological-only treatment group [214]. Some phase three trials have suggested that GPi stimulation is associated with a more significant reduction in dyskinesia than STN-DBS, but the significance of this difference has not been proven for treatment durations longer than 12 months [212,215]. The results of retrospective analysis showed that STN-DBS improved the MDS-UPDRS IV score in 57–61.9% of patients, depending on the group, with the improvement exceeding the direct effect of LEDD reduction on LID. This effect increased with time after treatment (69% vs. 77% at 1 year vs. 2 years after treatment, respectively) [216]. Interestingly, stimulation of the area above the STN (including the zona incerta) had better efficacy than stimulation of the STN alone (18 vs. 15 patients; p = 0.048) [217].
Repetitive preSMA transcranial magnetic stimulation (rTMS) has been studied as an alternative, non-invasive stimulatory intervention meant to control LID [218]. As described previously, preSMA activity is increased in dyskinetic PD patients, which may reflect pathological processes in the development of LID or attempts to compensate the pathology [46]. The results of this study showed that preSMA overactivity is one of the factors contributing to the development of dyskinesia, and that rTMS exhibits an antidyskinetic effect proportional to the strength of the applied magnetic field [218]. This study confirmed the results of previous studies showing the efficacy of rTMS in neurodegenerative diseases, including PD. Low-frequency rTMS has been shown to reduce dyskinesia in PD patients, and high-frequency rTMS has been shown to improve their motor function [219].

4. Conclusions

Structural changes in the brain that become apparent as an increase in asymmetry in the striatum size between sides, with a predominance of the degenerate side, leads to the development of LID. Functional changes in the basal ganglia appear to be central in the development of LID, particularly low frequency firing. Bidirectional plasticity in the cortico-striatal pathway is altered. SPNs of the direct and indirect pathway show an altered frequency and pattern of firing in LID. The excitability of dSPNs and iSPNs in the ON-phase is impaired. Dopaminergic downstream pathways—PKA/DARPP-32, ERK and mTORC1—are activated as part of a common cascade, beginning in striatonigral D1Rs SPNs. With the duration of L-DOPA treatment, the activity of these pathways increases. D3R, located mainly in the GP, shows an additive effect on LID, and its deletion reduces the activation of FosB, ERK and H3 histone. Risk factors for the development of dyskinesias include a higher mean L-DOPA dose size over 6 months, a higher dose escalation rate, earlier onset of PD, greater severity of motor symptoms and female gender. New routes of administration of L-DOPA (LBM, CCN) have been introduced, as well as ER compounds LD and CD, and amantadine, with very satisfactory results in lowering LID in advanced PD patients, whereas the standard oral immediate-response compounds of amantadine or apomorphine are useful as adjunct therapy for early (amantadine) or advanced (apomorphine) PD. Selective D3R agonists, such as pramipexole, may be considered as a moderate alternative to L-DOPA. Both 5-HT-1A and 5-HT-1A/B agonists showed dose-dependent antidyskinetic effects. SSRI also reduced LID, without the risk of inducing significant 5-HT syndrome. Propranolol is thought to affect dyskinesia by decreasing the firing rate in ChIs, leading to a decrease in ACh levels and consequently to a decrease in DA release. Although in the clinical practice STN-DBS is the only stimulation applied in the advanced stage of PD, the literature suggests that DBS can be applied to two regions, STN and GPi. Additional stimulation of areas above the STN may have better results than stimulation of the STN alone.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/jcm10194377/s1, Table S1: Summary of reviewed articles concerning the studied subjects and their parkinsonian/dyskinetic phenotypes.

Author Contributions

Conceptualisation, M.H.; methodology, M.H., J.H., A.K.-M. and A.G.; software, M.H., J.H., A.K.-M. and A.G.; validation, M.H., J.H., A.K.-M. and A.G.; formal analysis, A.K.-M. and A.G.; investigation, M.H., J.H., A.K.-M. and A.G.; resources, M.H., J.H., A.K.-M. and A.G.; data curation, M.H., J.H., A.K.-M. and A.G.; writing—original draft preparation, M.H., J.H., A.K.-M. and A.G.; writing—review and editing, M.H., J.H., A.K.-M. and A.G.; supervision, A.K.-M. and A.G.; project administration, A.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Albin, R.L.; Leventhal, D.K. The missing, the short, and the long: L-dopa responses and dopamine actions. Ann. Neurol. 2018, 82, 4–19. [Google Scholar] [CrossRef]
  2. Ockleford, C.; Adriaanse, P.; Berny, P.; Brock, T.; Duquesne, S.; Grilli, S.; Hernandez-Jerez, A.F.; Hougaard Bennekou, S.; Klein, M.; Kuhl, T.; et al. Investigation into experimental toxicological properties of plant protection products having a potential link to Parkinson’s disease and childhood leukaemia. EFSA J. 2017, 15, e04691. [Google Scholar] [PubMed] [Green Version]
  3. Eusebi, P.; Romoli, M.; Paoletti, F.P.; Tambasco, N.; Calabresi, P.; Parnetti, L. Risk factors of levodopa-induced dyskinesia in Parkinson’s disease: Results from the PPMI cohort. NPJ Parkinsons Dis. 2018, 4, 33. [Google Scholar] [CrossRef]
  4. Calabresi, P.; Standaert, D.G. Dystonia and levodopa-induced dyskinesias in Parkinson’s disease: Is there a connection? Neurobiol. Dis. 2019, 132, 104579. [Google Scholar] [CrossRef] [PubMed]
  5. Bjornestad, A.; Forsaa, E.B.; Pedersen, K.F.; Tysnes, O.-B.; Larsen, J.P.; Alves, G. Risk and course of motor complications in a population-based incident Parkinson’s disease cohort. Parkinsonism Relat. Disord. 2016, 22, 48–53. [Google Scholar] [CrossRef]
  6. Turcano, P.; Mielke, M.M.; Bower, J.H.; Parisi, J.E.; Cutsforth-Gregory, J.K.; Ahlskog, J.E.; Savica, R. Levodopa-induced dyskinesia in Parkinson disease: A population-based cohort study. Neurology 2018, 91, e2238–e2243. [Google Scholar] [CrossRef] [PubMed]
  7. Meissner, W.; Prunier, C.; Guilloteau, D.; Chalon, S.; Gross, C.E.; Bezard, E. Time-course of nigrostriatal degeneration in a progressive MPTP-lesioned macaque model of Parkinson’s disease. Mol. Neurobiol. 2003, 28, 209–218. [Google Scholar] [CrossRef]
  8. Potts, L.F.; Wu, H.; Singh, A.; Marcilla, I.; Luquin, M.R.; Papa, S.M. Modeling Parkinson’s disease in monkeys for translational studies, a critical analysis. Exp. Neurol. 2014, 256, 133–143. [Google Scholar] [CrossRef] [Green Version]
  9. Choudhury, G.R.; Daadi, M.M. Charting the onset of Parkinson-like motor and non-motor symptoms in nonhuman primate model of Parkinson’s disease. PLoS ONE 2018, 13, e0202770. [Google Scholar] [CrossRef]
  10. Burns, R.S.; Chiueh, C.C.; Markey, S.P.; Ebert, M.H.; Jacobowitz, D.M.; Kopin, I.J. A primate model of parkinsonism: Selective destruction of dopaminergic neurons in the pars compacta of the substantia nigra by N-methyl-4-phenyl-1,2,3,6-tetrahydropyridine. Proc. Natl. Acad. Sci. USA 1983, 80, 4546–4550. [Google Scholar] [CrossRef] [Green Version]
  11. Lundblad, M.; Picconi, B.; Lindgren, H.; Cenci, M.A. A model of L-DOPA-induced dyskinesia in 6-hydroxydopamine lesioned mice: Relation to motor and cellular parameters of nigrostriatal function. Neurobiol. Dis. 2004, 16, 110–123. [Google Scholar] [CrossRef]
  12. Picconi, B.; Paillé, V.; Ghiglieri, V.; Bagetta, V.; Barone, I.; Lindgren, H.S.; Bernardi, G.; Cenci, M.A.; Calabresi, P. l-DOPA dosage is critically involved in dyskinesia via loss of synaptic depotentiation. Neurobol. Dis. 2008, 29, 327–335. [Google Scholar] [CrossRef]
  13. Francardo, V.; Recchia, A.; Popovic, N.; Andersson, D.; Nissbrandt, H.; Cenci, M.A. Impact of the lesion procedure on the profiles of motor impairment and molecular responsiveness to L-DOPA in the 6-hydroxydopamine mouse model of Parkinson’s disease. Neurobiol. Dis. 2011, 42, 327–340. [Google Scholar] [CrossRef] [Green Version]
  14. Fletcher, E.J.R.; Finlay, C.J.; Lopez, A.A.; Crum, W.R.; Vernon, A.C.; Duty, S. Neuroanatomical and Microglial alterations in the striatum of levodopa-treated, dyskinetic hemi-Parkinsonian rats. Front. Neurosci. 2020, 14, 567222. [Google Scholar] [CrossRef]
  15. Ueno, T.; Nishijima, H.; Ueno, S.; Tomiyama, M. Spine enlargement of pyramidal tract-type neurons in the motor cortex of a rat model of levodopa-induced dyskinesia. Front. Neurosci. 2017, 11, 206. [Google Scholar] [CrossRef] [Green Version]
  16. Zhi, Y.; Wang, M.; Yuan, Y.-S.; Shen, Y.-T.; Ma, K.-W.; Gan, C.-T.; Si, Q.-Q.; Wang, L.-N.; Cao, S.-W.; Zhang, K.-Z. The increased gray matter volumes of precentral gyri in Parkinson’s disease patients with diphasic dyskinesia. Aging 2019, 11, 9661–9671. [Google Scholar] [CrossRef] [PubMed]
  17. Lindenbach, D.; Conti, M.M.; Ostock, C.Y.; Dupre, K.B.; Bishop, C. Alterations in primary motor cortex neurotransmission and gene expression in hemi-Parkinsonian rats with drug-induced dyskinesia. Neuroscience 2015, 310, 12–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Chung, S.J.; Yoo, H.S.; Lee, Y.H.; Jung, J.H.; Baik, K.; Ye, B.S.; Sohn, Y.H.; Lee, P.H. White matter hyperintensities and risk of levodopa-induced dyskinesia in Parkinson’s disease. Ann. Clin. Transl. Neurol. 2020, 7, 229–238. [Google Scholar] [CrossRef] [PubMed]
  19. Li, X.; Zhuang, P.; Li, Y. Altered neuronal firing pattern of the basal ganglia nucleus plays a role in levodopa-induced dyskinesia in patients with Parkinson’s disease. Front. Hum. Neurosci. 2015, 9, 630. [Google Scholar] [CrossRef] [Green Version]
  20. Porras, G.; De Deurwaerdere, P.; Li, Q.; Marti, M.; Morgenstern, R.; Sohr, R.; Bezard, E.; Morari, M.; Meissner, W.G. L-dopa-induced dyskinesia: Beyond an excessive dopamine tone in the striatum. Sci. Rep. 2014, 4, 3730. [Google Scholar] [CrossRef] [Green Version]
  21. Aristeta, A.; Azkona, G.; Sagarduy, A.; Miguelez, C.; Ruiz-Ortega, J.Á.; Sanchez-Pernaute, R.; Ugedo, L. The role of the subthalamic nucleus in L-DOPA induced dyskinesia in 6-hydroxydopamine lesioned rats. PLoS ONE 2012, 7, e42652. [Google Scholar] [CrossRef]
  22. Aristieta, A.; Ruiz-Ortega, J.A.; Morera-Herreras, T.; Miguelez, C.; Ugedo, L. Acute L-DOPA administration reverses changes in firing pattern and low frequency oscillatory activity in the entopeduncular nucleus from long term L-DOPA treated 6-OHDA-lesioned rats. Exp. Neurol. 2019, 322, 113036. [Google Scholar] [CrossRef]
  23. Yoo, H.S.; Choi, Y.H.; Chung, S.J.; Lee, Y.H.; Ye, B.S.; Sohn, Y.H.; Lee, J.M.; Lee, P.H. Cerebellar connectivity in Parkinson’s disease with levodopa-induced dyskinesia. Ann. Clin. Transl. Neurol. 2019, 6, 2251–2260. [Google Scholar] [CrossRef] [Green Version]
  24. Lewis, M.M.; Du, G.; Kidacki, M.; Patel, N.; Shaffer, M.L.; Mailman, R.B.; Huang, X. Higher iron in the red nucleus marks Parkinson’s dyskinesia. Neurobiol. Aging 2013, 34, 1497–1503. [Google Scholar] [CrossRef] [Green Version]
  25. Jourdain, V.A.; Schindlbeck, K.A.; Tang, C.C.; Niethammer, M.; Choi, Y.Y.; Markowitz, D.; Nazem, A.; Nardi, D.; Carras, N.; Feigin, A. Increased putamen hypercapnic vasoreactivity in levodopa-induced dyskinesia. JCI Insight 2017, 2, e96411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Jourdain, V.A.; Tang, C.C.; Holtbernd, F.; Dresel, C.; Choi, Y.Y.; Ma, Y.; Dhawan, V.; Eidelberg, D. Flow-metabolism dissociation in the pathogenesis of levodopa-induced dyskinesia. JCI Insight 2016, 1, e86615. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Aljuaid, M.; Booth, S.; Hobson, D.E.; Borys, A.; Williams, K.; Katako, A.; Ryner, L.; Goertzen, A.L.; Ko, J.H. Blood flow and glucose metabolism dissociation in the putamen is predictive of levodopa induced dyskinesia in Parkinson’s disease patients. Front. Neurol. 2019, 10, 1217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Andersen, A.D.; Blaabjerg, M.; Binzer, M.; Kamal, A.; Thagesen, H.; Kjaer, T.W.; Stenager, E.; Gramsbergen, J.B.P. Cerebrospinal fluid levels of catecholamines and its metabolites in Parkinson’s disease: Effect of l-DOPA treatment and changes in levodopa-induced dyskinesia. J. Neurochem. 2017, 141, 614–625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Rao, J.S.; Kellom, M.; Kim, H.-W.; Rapoport, S.I. Neuroinflammation and synaptic loss. Neurochem. Res. 2012, 37, 903–910. [Google Scholar] [CrossRef] [PubMed]
  30. Liu, Y.; Zhang, Y.; Zheng, X.; Fang, T.; Yang, X.; Luo, X.; Guo, A.; Newell, K.A.; Huang, X.F.; Yu, Y. Galantamine improves cognition, hippocampal inflammation, and synaptic plasticity impairments induced by lipopolysaccharide in mice. J. Neuroinflamm. 2018, 15, 112. [Google Scholar] [CrossRef] [Green Version]
  31. Gipson, C.D.; Olive, M.F. Structural and functional plasticity of dendritic spines—Root or result of behavior? Genes Brain Behav. 2017, 16, 101–117. [Google Scholar] [CrossRef] [Green Version]
  32. Yang, C.; Zhang, T.; Wang, W.; Xiang, Y.; Huang, Q.; Xie, C.; Zhao, L.; Zheng, H.; Yang, Y.; Gao, H. Brain-region specific metabolic abnormalities in Parkinson’s disease and levodopa-induced dyskinesia. Front. Aging Neurosci. 2020, 12, 75. [Google Scholar] [CrossRef]
  33. Ryu, H.S.; Park, K.W.; Choi, N.; Kim, J.; Park, Y.M.; Jo, S.; Kim, M.J.; Kim, Y.J.; Kim, J.; Kim, K.; et al. Genomic analysis identifies new loci associated with motor complications in Parkinson’s disease. Front. Neurol. 2020, 11, 570. [Google Scholar] [CrossRef]
  34. Ivanova, S.A.; Loonen, A.J.M.; Pechlivanoglou, P.; Freidin, M.B.; Al Hadithy, A.F.Y.; Rudikov, E.V.; Zhukova, I.A.; Govorin, N.V.; Sorokina, V.A.; Fedorenko, O.Y.; et al. NMDA receptor genotypes associated with the vulnerability to develop dyskinesia. Transl. Psychiatry 2012, 2, e67. [Google Scholar] [CrossRef]
  35. Kakinuma, S.; Beppu, M.; Sawai, S.; Nakayama, A.; Hirano, S.; Yamanaka, Y.; Yamamoto, T.; Masafumi, C.; Aisihaer, X.; Aersilan, A.; et al. Monoamine oxidase B rs1799836 G allele polymorphism is a risk factor for early development of levodopa-induced dyskinesia in Parkinson’s disease. eNeurologicalSci 2020, 19, 100239. [Google Scholar] [CrossRef]
  36. Han, C.-L.; Liu, Y.-P.; Sui, Y.P.; Chen, N.; Du, T.T.; Jiang, Y.; Guo, C.J.; Wang, K.L.; Wang, Q.; Fan, S.Y.; et al. Integrated transcriptome expression profiling reveals a novel lncRNA associated with l-DOPA-induced dyskinesia in a rat model of Parkinson’s disease. Aging 2020, 12, 718–739. [Google Scholar] [CrossRef]
  37. Heiman, M.; Heilbut, A.; Francardo, V.; Kulicke, R.; Fenster, R.J.; Kolaczyk, E.D.; Mesirov, J.P.; Surmeier, D.J.; Cenci, M.A.; Greengard, P. Molecular adaptations of striatal spiny projection neurons during levodopa-induced dyskinesia. Proc. Natl. Acad. Sci. USA 2014, 111, 4578–4583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Figge, D.A.; Eskow Jaunarajs, K.L.; Standaert, D.G. Dynamic DNA methylation regulates levodopa-induced dyskinesia. J. Neurosci. 2016, 36, 6514–6524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Figge, D.A.; Standaert, D.G. Dysregulation of BET proteins in levodopa-induced dyskinesia. Neurobiol. Dis. 2017, 102, 125–132. [Google Scholar] [CrossRef] [PubMed]
  40. Suwjin, S.R.; van Boheemen, C.J.M.; de Haan, R.J.; Tissingh, G.; Booij, J.; de Bie, R.M. The diagnostic accuracy of dopamine transporter SPECT imaging to detect nigrostriatal cell loss in patients with Parkinson’s disease or clinically uncertain parkinsonism: A systematic review. EJNMMI Res. 2015, 5, 12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Palermo, G.; Giannoni, S.; Frosini, D.; Morganti, R.; Volterrani, D.; Bonuccelli, U.; Pavese, N.; Ceravolo, R. Dopamine transporter, age, and motor complications in Parkinson’s disease: A clinical and single-photon emission computed tomography study. Mov. Disord. 2020, 35, 1028–1036. [Google Scholar] [CrossRef]
  42. Jakobson Mo, S.; Linder, J.; Forsgren, L.; Holmberg, H.; Larsson, A.; Riklund, K. Pre- and postsynaptic dopamine SPECT in Idiopathic Parkinsonian diseases: A follow-up study. Biomed. Res. Int. 2013, 2013, 143532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Jeong, E.H.; Sunwoo, M.K.; Song, Y.S. Serial I-123-FP-CIT SPECT image Findings of Parkinson’s disease patients with levodopa-induced dyskinesia. Front. Neurol. 2018, 9, 1133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Cerasa, A.; Pugliese, P.; Messina, D.; Morelli, M.; Gioia, M.C.; Salsone, M.; Novellino, F.; Nicoletti, G.; Arabia, G.; Quattrone, A. Prefrontal alterations in Parkinson’s disease with levodopa-induced dyskinesia during fMRI motor task. Mov. Disord. 2012, 27, 364–371. [Google Scholar] [CrossRef] [PubMed]
  45. Herz, D.M.; Haagensen, B.N.; Christensen, M.S.; Madsen, K.H.; Rowe, J.B.; Løkkegaard, A.; Siebner, H.R. Abnormal dopaminergic modulation of striato-cortical networks underlies levodopa-induced dyskinesias in humans. Brain 2015, 138, 1658–1666. [Google Scholar] [CrossRef] [Green Version]
  46. Herz, D.M.; Haagensen, B.N.; Christensen, M.S.; Madsen, K.H.; Rowe, J.B.; Løkkegaard, A.; Siebner, H.R. The acute brain response to levodopa heralds dyskinesias in Parkinson disease. Ann. Neurol. 2014, 75, 829–836. [Google Scholar] [CrossRef] [Green Version]
  47. Girasole, A.E.; Lum, M.Y.; Nathaniel, D.; Bair-Marshall, C.J.; Guenthner, C.J.; Luo, L.; Kreitzer, A.C.; Nelson, A.B. A subpopulation of striatal neurons mediates levodopa-induced dyskinesia. Neuron 2018, 97, 787–795.e6. [Google Scholar] [CrossRef]
  48. Ryan, M.B.; Bair-Marshall, C.; Nelson, A.B. Aberrant striatal activity in parkinsonism and levodopa-induced dyskinesia. Cell Rep. 2018, 23, 3438–3446.e5. [Google Scholar] [CrossRef]
  49. Parker, J.G.; Marshall, J.D.; Ahanonu, B.; Wu, Y.W.; Kim, T.H.; Grewe, B.F.; Zhang, Y.; Li, J.Z.; Ding, J.B.; Ehlers, M.D.; et al. Diametric neural ensemble dynamics in parkinsonian and dyskinetic states. Nature 2019, 557, 177–182. [Google Scholar] [CrossRef]
  50. Li, L.; Zhou, F.M. Parallel dopamine D1 receptor activity dependence of l-Dopa-induced normal movement and dyskinesia in mice. Neuroscience 2013, 236, 66–76. [Google Scholar] [CrossRef] [Green Version]
  51. Singh, A.; Liang, L.; Kaneoke, Y.; Cao, X.; Papa, S.M. Dopamine regulates distinctively the activity patterns of striatal output neurons in advanced parkinsonian primates. J. Neurophysiol. 2015, 113, 1533–1544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Belić, J.J.; Halje, P.; Richter, U.; Petersson, P.; Kotaleski, J.H. Untangling cortico-striatal connectivity and cross-frequency coupling in L-DOPA-induced dyskinesia. Front. Syst. Neurosci. 2016, 10, 26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Alberico, S.L.; Kim, Y.-C.; Lence, T.; Narayanan, N.S. Axial levodopa-induced dyskinesias and neuronal activity in the dorsal striatum. Neuroscience 2017, 343, 240–249. [Google Scholar] [CrossRef] [Green Version]
  54. Thiele, S.L.; Chen, B.; Lo, C.; Gertler, T.S.; Warre, R.; Surmeier, J.D.; Brotchie, J.M.; Nash, J.E. Selective loss of bi-directional synaptic plasticity in the direct and indirect striatal output pathways accompanies generation of parkinsonism and l-DOPA induced dyskinesia in mouse models. Neurobiol. Dis. 2014, 71, 334–344. [Google Scholar] [CrossRef] [Green Version]
  55. Halje, P.; Tamtè, M.; Richter, U.; Mohammed, M.; Cenci, M.A.; Petersson, P. Levodopa-induced dyskinesia is strongly associated with resonant cortical oscillations. J. Neurosci. 2012, 32, 16541–16551. [Google Scholar] [CrossRef] [PubMed]
  56. Perez, X.A.; Zhang, D.; Bordia, T.; Quik, M. Striatal D1 medium spiny neuron activation induces dyskinesias in parkinsonian mice. Mov. Disord. 2017, 32, 538–548. [Google Scholar] [CrossRef] [Green Version]
  57. Shen, W.; Ren, W.; Zhai, S.; Yang, B.; Vanoye, C.G.; Mitra, A.; George, A.L., Jr.; Surmeier, D.J. Striatal Kir2 K+ channel inhibition mediates the antidyskinetic effects of amantadine. J. Clin. Investig. 2020, 130, 2593–2601. [Google Scholar] [CrossRef]
  58. Okada, M.; Nakao, R.; Hosoi, R.; Zhang, M.-R.; Fukumura, T.; Suzuki, K.; Inoue, O. Microdialysis with radiometric monitoring of -[β-11C]DOPA to assess dopaminergic metabolism: Effect of inhibitors of -amino acid decarboxylase, monoamine oxidase, and catechol-O-methyltransferase on rat striatal dialysate. J. Cereb. Blood Flow Metab. 2011, 31, 124–131. [Google Scholar] [CrossRef] [Green Version]
  59. Loonen, A.J.M.; Ivanova, S.A.; Pozhidaev, I.V.; Freidin, M.B.; Zhukova, I.A.; Osmanova, D.Z.; Zhukova, N.G.; Mironova, Y.A.; Tiguntsev, V.V.; Fedorenko, O.Y.; et al. Polymorphisms of catechol-O-Methyl Transferase (COMT) gene in vulnerability to levodopa-induced dyskinesia. J. Pharm. Pharm. Sci. 2018, 21, 340. [Google Scholar]
  60. Torkaman-Boutorabi, A.; Shahidi, G.A.; Choopani, S.; Rezvani, M.; Pourkosary, K.; Golkar, M.; Zarrindast, M.-R. The catechol-O-methyltransferase and monoamine oxidase B polymorphisms and levodopa therapy in the Iranian patients with sporadic Parkinson’s disease. Acta Neurobiol. Exp. 2012, 72, 272–282. [Google Scholar]
  61. Santini, E.; Feyder, M.; Gangarossa, G.; Bateup, H.S.; Greengard, P.; Fisone, G. Dopamine- and cAMP-regulated Phosphoprotein of 32-kDa (DARPP-32)-dependent Activation of Extracellular Signal-regulated Kinase (ERK) and Mammalian Target of Rapamycin Complex 1 (mTORC1) Signaling in experimental Parkinsonism. J. Biol. Chem. 2012, 287, 27806–27812. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Calabrese, V.; Di Maio, A.; Marino, G.; Cardinale, A.; Natale, G.; De Rosa, A.; Campanelli, F.; Mancini, M.; Napolitano, F.; Avallone, L.; et al. Rapamycin, by inhibiting mTORC1 signaling, prevents the loss of striatal bidirectional synaptic plasticity in a rat model of L-DOPA-induced dyskinesia. Front. Aging Neurosci. 2020, 12, 230. [Google Scholar] [CrossRef]
  63. Wu, N.; Wan, Y.; Song, L.; Qi, C.; Liu, Z.; Gan, J. The abnormal activation of D1R/Shp-2 complex involved in levodopa-induced dyskinesia in 6-hydroxydopamine-lesioned Parkinson’s rats. Neuropsychiatr. Dis. Treat. 2018, 14, 1779–1786. [Google Scholar] [CrossRef] [Green Version]
  64. Jones-Tabah, J.; Mohammad, H.; Hadj-Youssef, S.; Kim, L.E.H.; Martin, R.D.; Benaliouad, F.; Tanny, J.C.; Clarke, P.B.S.; Hébert, T.E. Dopamine D1 receptor signalling in dyskinetic Parkinsonian rats revealed by fiber photometry using FRET-based biosensors. Sci. Rep. 2020, 10, 14426. [Google Scholar] [CrossRef]
  65. Fieblinger, J.; Sebastianutto, I.; Alcacer, C.; Bimpisidis, Z.; Maslava, N.; Sandberg, S.; Engblom, D.; Cenci, M.A. Mechanisms of dopamine D1 receptor-mediated ERK1/2 activation in the Parkinsonian striatum and their modulation by metabotropic glutamate receptor type 5. J. Neurosci. 2014, 34, 4728–4740. [Google Scholar] [CrossRef] [Green Version]
  66. Alcacer, C.; Santini, E.; Valjent, E.; Gaven, F.; Girault, J.A.; Hervé, D. Gαolf mutation allows parsing the role of cAMP-dependent and extracellular signal-regulated kinase-dependent signaling in l-3,4-dihydroxyphenylalanine-induced dyskinesia. J. Neurosci. 2012, 32, 5900–5910. [Google Scholar] [CrossRef]
  67. Goto, S. Striatal Gαolf/cAMP signal-dependent mechanism to generate levodopa-induced dyskinesia in Parkinson’s disease. Front. Cell Neurosci. 2017, 11, 364. [Google Scholar] [CrossRef]
  68. Södersten, E.; Feyder, M.; Lerdrup, M.; Gomes, A.-L.; Kryh, H.; Spigolon, G.; Caboche, J.; Fisone, G.; Hansen, K. Dopamine signaling leads to loss of polycomb repression and aberrant gene activation in experimental Parkinsonism. PLoS Genet. 2014, 10, e1004574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Sellnow, R.C.; Steece-Collier, K.; Altwal, F.; Sandoval, I.M.; Kordower, J.H.; Collier, T.J.; Sortwell, C.E.; West, A.R.; Manfredsson, F.P. Striatal Nurr1 facilitates the dyskinetic state and exacerbates levodopa-induced dyskinesia in a rat model of Parkinson’s disease. J. Neurosci. 2020, 40, 3675–3691. [Google Scholar] [CrossRef] [PubMed]
  70. Stanic, J.; Mellone, M.; Cirnaru, M.D.; Perez-Carrion, M.; Zianni, E.; Di Luca, M.; Gardoni, F.; Piccoli, G. LRRK2 phosphorylation level correlates with abnormal motor behaviour in an experimental model of levodopa-induced dyskinesias. Mol. Brain. 2016, 9, 53. [Google Scholar] [CrossRef] [Green Version]
  71. Divitio, C.B.; Steece-Collier, K.; Case, D.T.; Williams, S.P.; Stancati, J.A.; Zhi, L.; Rubio, M.E.; Sortwell, C.E.; Collier, T.J.; Sulzer, D.; et al. Loss of VGLUT3 produces circadian-dependent hyperdopaminergia and ameliorates motor dysfunction and l-Dopa-mediated dyskinesias in a model of Parkinson’s disease. J. Neurosci. 2015, 35, 14983–14999. [Google Scholar] [CrossRef] [Green Version]
  72. Zhang, J.; Saur, T.; Duke, A.N.; Grant, S.G.N.; Platt, D.M.; Rowlett, J.K.; Isacson, O.; Yao, W.-D. Motor impairments, striatal degeneration, and altered dopamine-glutamate interplay in mice lacking PSD-95. J. Neurogenet. 2014, 28, 98–111. [Google Scholar] [CrossRef] [Green Version]
  73. Porras, G.; Berthet, A.; Dehay, B.; Li, Q.; Ladepeche, L.; Normand, E.; Dovero, S.; Martinez, A.; Doudnikoff, E.; Martin-Négrier, M.L.; et al. PSD-95 expression controls l-DOPA dyskinesia through dopamine D1 receptor trafficking. J. Clin. Investig. 2012, 122, 3977–3989. [Google Scholar] [CrossRef] [Green Version]
  74. Berthet, A.; Bezard, E.; Porras, G.; Fasano, S.; Barroso-Chinea, P.; Dehay, B.; Martinez, A.; Thiolat, M.L.; Nosten-Bertrand, M.; Giros, B.; et al. l-DOPA impairs proteasome activity in parkinsonism through D1 dopamine receptor. J. Neurosci. 2012, 32, 681–691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Lyons, D.; de Jaeger, X.; Rosen, L.G.; Ahmad, T.; Lauzon, N.M.; Zunder, J.; Coolen, L.M.; Rushlow, W.; Laviolette, S.R. Opiate exposure and withdrawal induces a molecular memory switch in the basolateral amygdala between ERK1/2 and CaMKIIα-dependent signaling substrates. J. Neurosci. 2013, 33, 14693–14704. [Google Scholar] [CrossRef] [Green Version]
  76. Zhang, S.F.; Xie, C.L.; Wang, Q.; Liu, Z.G. Interactions of CaMKII with dopamine D2 receptors: Roles in levodopa-induced dyskinesia in 6-hydroxydopamine lesioned Parkinson’s rats. Sci. Rep. 2014, 4, 6811. [Google Scholar] [CrossRef] [PubMed]
  77. Albarrán-Bravo, S.; Ávalos-Fuentes, J.A.; Cortés, H.; Rodriguez-Sánchez, M.; Leyva-García, N.; Rangel-Barajas, C.; Erlij, D.; Florán, B. Severity of dyskinesia and D3R signaling changes induced by L-DOPA treatment of hemiparkinsonian rats are features inherent to the treated subjects. Biomolecules 2019, 9, 431. [Google Scholar] [CrossRef] [Green Version]
  78. Solís, O.; Garcia-Montes, J.R.; González-Granillo, A.; Xu, M.; Moratalla, R. Dopamine D3 receptor modulates l-DOPA-induced dyskinesia by targeting D1 receptor-mediated striatal signaling. Cereb. Cortex 2017, 27, 435–446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Payer, D.E.; Guttman, M.; Kish, S.J.; Tong, J.; Adams, J.R.; Rusjan, P.; Houle, S.; Furukawa, Y.; Wilson, A.A.; Boileau, I. D3 dopamine receptor-preferring [11C]PHNO PET imaging in Parkinson patients with dyskinesia. Neurology 2016, 86, 224–230. [Google Scholar] [CrossRef] [Green Version]
  80. Castello, J.; Cortés, M.; Malave, L.; Kottmann, A.; Sibley, D.R.; Friedman, E.; Rebholz, H. The dopamine D5 receptor contributes to activation of cholinergic interneurons during L-DOPA induced dyskinesia. Sci. Rep. 2020, 10, 2542. [Google Scholar] [CrossRef] [Green Version]
  81. Bido, S.; Solari, N.; Indrigo, M.; D’Antoni, A.; Brambilla, R.; Morari, M.; Fasano, S. Differential involvement of Ras-GRF1 and Ras-GRF2 in L-DOPA-induced dyskinesia. Ann. Clin. Transl. Neurol. 2015, 2, 662–678. [Google Scholar] [CrossRef] [PubMed]
  82. Fasano, S.; Bezard, E.; D’Antoni, A.; Francardo, V.; Indrigo, M.; Qin, L.; Doveró, S.; Cerovic, M.; Cenci, M.A.; Brambilla, R. Inhibition of Ras-guanine nucleotide-releasing factor 1 (Ras-GRF1) signaling in the striatum reverts motor symptoms associated with l-dopa–induced dyskinesia. Proc. Natl. Acad. Sci. USA 2010, 107, 21824–21829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Eshraghi, M.; Ramírez-Jarquín, U.N.; Shahani, N.; Nuzzo, T.; De Rosa, A.; Swarnkar, S.; Galli, N.; Rivera, O.; Tsaprailis, G.; Scharager-Tapia, C.; et al. RasGRP1 is a causal factor in the development of l-DOPA–induced dyskinesia in Parkinson’s disease. Neuroscience 2020, 6, eeaz7001. [Google Scholar]
  84. Cerovic, M.; Bagetta, V.; Pendolino, V.; Ghiglieri, V.; Fasano, S.; Morella, I.; Hardingham, N.; Heuer, A.; Papale, A.; Marchisella, F.; et al. Derangement of Ras-guanine nucleotide-releasing factor 1 (Ras-GRF1) and extracellular signal-regulated kinase (ERK) dependent striatal plasticity in L-DOPA-induced dyskinesia. Biol. Psychiatry 2015, 77, 106–115. [Google Scholar] [CrossRef] [PubMed]
  85. Ruiz-DeDiego, I.; Fasano, S.; Solís, O.; Garcia-Montes, J.R.; Brea, J.; Loza, M.I.; Brambilla, R.; Moratalla, R. Genetic enhancement of Ras-ERK pathway does not aggravate L-DOPA-induced dyskinesia in mice but prevents the decrease induced by lovastatin. Sci. Rep. 2018, 8, 15381. [Google Scholar] [CrossRef]
  86. Beck, G.; Singh, A.; Zhang, J.; Potts, L.F.; Woo, J.-M.; Park, E.S.; Mochizuki, H.; Mouradian, M.M.; Papa, S.M. Role of striatal ΔFosB in l-Dopa–induced dyskinesias of parkinsonian nonhuman primates. Proc. Natl. Acad. Sci. USA 2019, 116, 18664–18672. [Google Scholar] [CrossRef] [Green Version]
  87. Cortés, M.; Malave, L.; Castello, J.; Flajolet, M.; Cenci, M.A.; Friedman, E.; Rebholz, H. CK2 oppositely modulates l-DOPA-induced dyskinesia via striatal projection neurons expressing D1 or D2 receptors. J. Neurosci. 2017, 37, 11930–11946. [Google Scholar] [CrossRef]
  88. Bastide, M.F.; Glangetas, C.; Doudnikoff, E.; Li, Q.; Bourdenx, M.; Fernagut, P.O.; Dumont, É.C.; Georges, F.; Bézard, E. Involvement of the bed nucleus of the stria terminalis in L-Dopa induced dyskinesia. Sci. Rep. 2017, 7, 2348. [Google Scholar] [CrossRef] [Green Version]
  89. Padovan-Neto, F.E.; Ferreira, N.R.; de Oliveira-Tavaras, D.; de Aguiar, D.; da Silva, C.A.; Raisman-Vozari, R.; Del Bel, E. Anti-dyskinetic effect of the neuronal nitric oxide synthase inhibitor is linked to decrease of FosB/DeltaFosB expression. Neurosci. Lett. 2013, 541, 126–131. [Google Scholar] [CrossRef] [Green Version]
  90. Singh, A.; Jenkins, M.A.; Burke, K.J., Jr.; Beck, G.; Jenkins, A.; Scimemi, A.; Traynelis, S.F.; Papa, S.M. Glutamatergic tuning of hyperactive striatal projection neurons controls the motor response to dopamine replacement in Parkinsonian primates. Cell Rep. 2018, 22, 941–952. [Google Scholar] [CrossRef] [Green Version]
  91. Stanic, J.; Mellone, M.; Napolitano, F.; Racca, C.; Zianni, E.; Minocci, D.; Ghiglieri, V.; Thiolat, M.L.; Li, Q.; Longhi, A.; et al. Rabphilin 3A: A novel target for the treatment of levodopa-induced dyskinesias. Neurobiol. Dis. 2017, 108, 54–64. [Google Scholar] [CrossRef]
  92. Ba, M.; Kong, M.; Ma, G. Postsynaptic density protein 95-regulated NR2B tyrosine phosphorylation and interactions of Fyn with NR2B in levodopa-induced dyskinesia rat models. Drug Des. Dev. Ther. 2015, 9, 199–206. [Google Scholar] [CrossRef] [Green Version]
  93. Lindenbach, D.; Conti, M.M.; Ostock, C.Y.; George, J.A.; Goldenberg, A.A.; Melikhov-Sosin, M.; Nuss, E.E.; Bishop, C. The role of primary motor cortex (M1) glutamate and GABA signaling in l-DOPA-induced dyskinesia in Parkinsonian rats. J. Neurosci. 2016, 36, 9873–9887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Brumberg, J.; Küsters, S.; Al-Momani, E.; Marotta, G.; Cosgrove, K.P.; van Dyck, C.H.; Herrmann, K.; Homola, G.A.; Pezzoli, G.; Buck, A.K.; et al. Cholinergic activity and levodopa-induced dyskinesia: A multitracer molecular imaging study. Ann. Clin. Transl. Neurol. 2017, 4, 632–639. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Brugnoli, A.; Pisanò, C.A.; Morari, M. Striatal and nigral muscarinic type 1 and type 4 receptors modulate levodopa-induced dyskinesia and striato-nigral pathway activation in 6-hydroxydopamine hemilesioned rats. Neurobiol. Dis. 2020, 144, 105044. [Google Scholar] [CrossRef]
  96. Shen, W.; Plotkin, J.L.; Francardo, V.; Ko, W.K.; Xie, Z.; Li, Q.; Fieblinger, T.; Wess, J.; Neubig, R.R.; Lindsley, C.W.; et al. M4 muscarinic receptor signaling ameliorates striatal plasticity deficits in models of L-DOPA-induced dyskinesia. Neuron 2015, 88, 762–763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Sahin, G.; Thompson, L.H.; Lavisse, S.; Ozgur, M.; Rbah-Vidal, L.; Dollé, F.; Hantraye, P.; Kirik, D. Differential dopamine receptor occupancy underlies L-DOPA-induced dyskinesia in a rat model of Parkinson’s disease. PLoS ONE 2014, 9, e90759. [Google Scholar]
  98. Rylander, D.; Parent, M.; O’Sullivan, S.S.; Dovero, S.; Lees, A.J.; Bezard, E.; Descarries, L.; Cenci, M.A. Maladaptive plasticity of serotonin axon terminals in levodopa-induced dyskinesia. Ann. Neurol. 2010, 68, 619–628. [Google Scholar] [CrossRef]
  99. Carta, M.; Carlsson, T.; Kirik, D.; Björklund, A. Dopamine released from 5-HT terminals is the cause of L-DOPA-induced dyskinesia in parkinsonian rats. Brain 2007, 130, 1819–1833. [Google Scholar] [CrossRef] [Green Version]
  100. Tronci, E.; Napolitano, F.; Muñoz, A.; Fidalgo, C.; Rossi, F.; Björklund, A.; Usiello, A.; Carta, M. BDNF over-expression induces striatal serotonin fiber sprouting and increases the susceptibility to l-DOPA-induced dyskinesia in 6-OHDA-lesioned rats. Exp. Neurol. 2017, 297, 73–81. [Google Scholar] [CrossRef]
  101. Gagnon, D.; Gregoire, L.; Di Paolo, T.; Parent, M. Serotonin hyperinnervation of the striatum with high synaptic incidence in Parkinsonian monkeys. Brain Struct. Funct. 2016, 221, 3675–3691. [Google Scholar] [CrossRef]
  102. Inden, M.; Abe, M.; Minamino, H.; Takata, K.; Yoshimoto, K.; Tooyama, I.; Kitamura, Y. Effect of selective serotonin reuptake inhibitors via 5-HT1A receptors on L-DOPA-induced rotational behavior in a hemiparkinsonian rat model. J. Pharmacol. Sci. 2012, 119, 10–19. [Google Scholar] [CrossRef] [Green Version]
  103. Bishop, C.; George, J.A.; Buchta, W.; Goldenberg, A.A.; Mohamed, M.; Dickinson, S.O.; Eissa, S.; Eskow Jaunarajs, K.L. Serotonin transporter inhibition attenuates l-DOPA-induced dyskinesia without compromising l-DOPA efficacy in hemi-parkinsonian rats. Eur. J. Neurosci. 2012, 36, 2839–2848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Miguelez, C.; Navailles, S.; De Deurwaerdère, P.; Ugedo, L. The acute and long-term L-DOPA effects are independent from changes in the activity of dorsal raphe serotonergic neurons in 6-OHDA lesioned rats. Br. J. Pharmacol. 2016, 173, 2135–2146. [Google Scholar] [CrossRef] [PubMed]
  105. Santos-Lobato, B.L.; Schumacher-Schuh, A.F.; Rieder, C.R.M.; Hutz, M.H.; Borges, V.; Ferraz, H.B.; Mata, I.F.; Zabetian, C.P.; Tumas, V. Diagnostic prediction model for levodopa-induced dyskinesia in Parkinson’s disease. Arq. Neuropsiquiatr. 2020, 78, 206–216. [Google Scholar] [CrossRef]
  106. Hong, J.Y.; Sunwoo, M.K.; Yoon, J.H.; Kang, S.Y.; Sohn, Y.H.; Lee, P.H.; Kim, S.H. Rapid drug increase and early onset of levodopa-induced dyskinesia in Parkinson’s disease. PLoS ONE 2020, 15, e0237472. [Google Scholar] [CrossRef] [PubMed]
  107. Lin, M.M.; Laureno, R. Less pulsatile levodopa therapy (6 doses daily) is associated with a reduced incidence of dyskinesia. J. Mov. Disord. 2019, 12, 37–42. [Google Scholar] [CrossRef] [PubMed]
  108. Poewe, W.; Chaudhuri, K.R.; Bergmann, L.; Antonini, A. Levodopa–carbidopa intestinal gel in a subgroup of patients with dyskinesia at baseline from the GLORIA Registry. Neurodegener. Dis. Manag. 2019, 9, 39–46. [Google Scholar] [CrossRef] [PubMed]
  109. Ramlackhansingh, A.F.; Bose, S.K.; Ahmed, I.; Turkheimer, F.E.; Pavese, N.; Brooks, D.J. Adenosine 2A receptor availability in dyskinetic and nondyskinetic patients with Parkinson disease. Neurology 2011, 76, 1811–1816. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Wills, A.-M.A.; Eberly, S.; Tennis, M.; Lang, A.E.; Messing, S.; Togasaki, D.; Tanner, C.M.; Kamp, C.; Chen, J.F.; Oakes, D.; et al. Caffeine consumption and risk of dyskinesia in CALM-PD. Mov. Disord. 2013, 28, 380–383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Riggare, S.; Unruh, K.T.; Sturr, J.; Domingos, J.; Stamford, J.A.; Svenningsson, P.; Hägglund, M. Patient-driven N-of-1 in Parkinson’s disease; lessons learned from a placebo-controlled study of the effect of nicotine on dyskinesia. Methods Inf. Med. 2017, 56, e123–e128. [Google Scholar]
  112. Popa, L.C.; Leucota, D.-C.; Tohanean, N.; Popa, S.-L.; Perju-Dumbrava, L. Intrajejunal vs oral levodopa-carbidopa therapy in Parkinson disease. Medicine 2020, 99, 46. [Google Scholar] [CrossRef]
  113. Wan, Y.; Wu, N.; Song, l.; Wang, X.; Liu, Z.; Yuan, W.; Gan, J. Levodopa/benserazide loaded microspheres alleviate L-dopa induced dyskinesia through preventing the over-expression of D1R/Shp-2/ERK1/2 signaling pathway in a rat model of Parkinson’s disease. Front. Aging Neurosci. 2017, 9, 331. [Google Scholar] [CrossRef]
  114. Yang, X.; Zheng, R.; Cai, Y.; Liao, M.; Yuan, W.; Liu, Z. Controlled-release levodopa methyl ester/benserazide-loaded nanoparticles ameliorate levodopa-induced dyskinesia in rats. Int. J. Nanomed. 2012, 7, 2077–2086. [Google Scholar]
  115. Xie, C.I.; Wang, W.W.; Zhang, S.-F.; Yuan, M.-L.; Che, J.-Y.; Gan, J.; Song, L.; Yuan, W.-E.; Liu, Z.-G. Levodopa/benserazide microsphere (LBM) prevents L-dopa induced dyskinesia by inactivation of the DR1/PKA/P-tau pathway in 6-OHDA-lesioned Parkinson’s rats. Sci. Rep. 2014, 4, 7506. [Google Scholar] [CrossRef] [Green Version]
  116. Yang, X.; Chen, Y.; Hong, X.; Wu, N.; Song, L.; Yuan, W.; Liu, Z. Levodopa/benserazide microspheres reduced levodopa-induced dyskinesia by downregulating phosphorylated GluR1 expression in 6-OHDA-lesioned rats. Drug Des. Dev. Ther. 2012, 6, 341–347. [Google Scholar]
  117. Cao, X.; Hou, D.; Wang, L.; Li, S.; Sun, S.; Ping, Q.; Xu, Y. Effects and molecular mechanism of chitosan-coated levodopa nanoliposomes on behavior of dyskinesia rats. Biol. Res. 2016, 49, 32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Bido, S.; Marti, M.; Morari, M. Amantadine attenuates levodopa-induced dyskinesia in mice and rats preventing the accompanying rise in nigral GABA levels. J. Neurochem. 2011, 118, 1043–1055. [Google Scholar] [CrossRef]
  119. Stansley, B.L.; Yamamoto, B.K. Chronic l-dopa decreases serotonin neurons in a subregion of the dorsal raphe nucleus. J. Pharmacol. Exp. Ther. 2014, 351, 440–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Hernandez, M.; Walsch, M.; Stead, T.; Quinones, A.; Ganti, L. Serotonin syndrome in the emergency department. Cureus 2019, 11, e6307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Sagarduy, A.; Llorente, J.; Miguelez, C.; Morera-Herreras, T.; Ruiz-Ortega, J.A.; Ugedo, L. Buspirone requires the intact nigrostriatal pathway to reduce the activity of the subthalamic nucleus via 5-HT1A receptors. Exp. Neurol. 2016, 277, 35–45. [Google Scholar] [CrossRef] [Green Version]
  122. Vegas-Suárez, S.; Pisanò, C.A.; Requejo, C.; Bengoetxea, H.; Laufente, J.V.; Morari, M.; Miguelez, C.; Ugedo, L. 6-Hydroxydopamine lesion and levodopa treatment modify the effect of buspirone in the substantia nigra pars reticulata. Br. J. Pharmacol. 2020, 177, 3957–3974. [Google Scholar] [CrossRef]
  123. Nahimi, A.; Høltzermann, M.; Landau, A.M.; Simonsen, M.; Jakobsen, S.; Alstrup, A.K.O.; Vang, K.; Møller, A.; Wegener, G.; Gjedde, A.; et al. Serotonergic modulation of receptor occupancy in rats treated with L-DOPA after unilateral 6-OHDA lesioning. J. Neurochem. 2012, 120, 806–817. [Google Scholar] [CrossRef] [Green Version]
  124. Dupre, K.B.; Cruz, A.V.; McCoy, A.J.; Delaville, C.; Gerber, C.M.; Eyring, K.W.; Walters, J.R. Effects of L-dopa priming on cortical high beta and high gamma oscillatory activity in a rodent model of Parkinson’s disease. Neurobiol. Dis. 2016, 86, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Lindenbach, D.; Palumbo, N.; Ostock, C.Y.; Vilceus, N.; Conti, M.M.; Bishop, C. Side effect profile of 5-HT treatments for Parkinson’s disease and L-DOPA-induced dyskinesia in rats. Br. J. Pharmacol. 2014, 172, 119–130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Bhide, N.; Lindenbach, D.; Surrena, M.A.; Goldenberg, A.A.; Bishop, C.; Berger, S.P.; Paquette, M.A. The effects of BMY-14802 against L-DOPA- and dopamine agonist-induced dyskinesia in the hemiparkinsonian rat. Psychopharmacology 2013, 227, 533–544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Aboulghasemi, N.; Jahromy, M.H.; Ghasemi, A. Anti-dyskinetic efficacy of 5-HT3 receptor antagonist in the hemi-parkinsonian rat model. IBRO Rep. 2019, 6, 40–44. [Google Scholar] [CrossRef] [PubMed]
  128. Tronci, E.; Lisci, C.; Stancampiano, R.; Fidalgo, C.; Collu, M.; Devoto, P.; Carta, M. 5-Hydroxy-tryptophan for the treatment of l-DOPA-induced dyskinesia in the rat Parkinson’s disease model. Neurobiol. Dis. 2013, 60, 108–114. [Google Scholar] [CrossRef] [PubMed]
  129. Maffei, M.E. 5-Hydroxytryptophan (5-HTP): Natural occurrence, analysis, biosynthesis, biotechnology, physiology and toxicology. Int. J. Mol. Sci. 2021, 22, 181. [Google Scholar] [CrossRef] [PubMed]
  130. Frau, R.; Savoia, P.; Fanni, S.; Fiorentini, C.; Fidalgo, C.; Tronci, E.; Stancampiano, R.; Meloni, M.; Cannas, A.; Marrosu, F.; et al. The 5-alpha reductase inhibitor finasteride reduces dyskinesia in a rat model of Parkinson’s disease. Exp. Neurol. 2017, 291, 1–7. [Google Scholar] [CrossRef] [PubMed]
  131. Yilmaz, C.; Karali, K.; Fodelianaki, G.; Gravanis, A.; Chavakis, T.; Charalampopoulos, I.; Alexaki, V.I. Neurosteroids as regulators of neuroinflammation. Front. Neuroendocrinol. 2019, 55, 100788. [Google Scholar] [CrossRef]
  132. Robichaud, M.; Debonnel, G. Allopregnanolone and ganaxolone increase the firing activity of dorsal raphe nucleus serotonergic neurons in female rats. Int. J. Neuropsychopharmacol. 2006, 9, 191–200. [Google Scholar] [CrossRef] [Green Version]
  133. Hara, M.; Fukui, R.; Hieda, E.; Kuroiwa, M.; Bateup, H.S.; Kano, T.; Greengard, P.; Nishi, A. Role of adrenoceptors in the regulation of dopamine/DARPP-32 signaling in neostriatal neurons. J. Neurochem. 2010, 113, 1046–1059. [Google Scholar] [CrossRef] [Green Version]
  134. Lindenbach, D.; Ostock, C.Y.; Eskow Jaunarajs, K.L.; Dupre, K.B.; Barnum, C.J.; Bhide, N.; Bishop, C. Behavioral and cellular modulation of l-DOPA-induced dyskinesia by β-adrenoceptor blockade in the 6-hydroxydopamine-lesioned rat. J. Pharmacol. Exp. Ther. 2011, 337, 755–765. [Google Scholar] [CrossRef] [Green Version]
  135. Bhide, N.; Lindenbach, D.; Barnum, C.J.; George, J.A.; Surrena, M.A.; Bishop, C. Effects of the beta-adrenergic receptor antagonist Propranolol on dyskinesia and L-DOPA-induced striatal DA efflux in the hemi-parkinsonian rat. J. Neurochem. 2015, 134, 222–232. [Google Scholar] [CrossRef]
  136. Shi, Z.; Barnford, I.J.; McKinley, J.W.; Devi, S.P.S.; Vahedipour, A.; Bamford, N.S. Propranolol relieves L-dopa-induced dyskinesia in Parkinsonian mice. Brain Sci. 2020, 10, 903. [Google Scholar] [CrossRef]
  137. Crans, R.A.J.; Wouters, E.; Valle-León, M.; Taura, J.; Massari, C.M.; Fernández-Dueñas, V.; Stove, C.P.; Ciruela, F. Striatal dopamine D2-muscarinic acetylcholine M1 receptor—Receptor interaction in a model of movement disorders. Front. Pharmacol. 2020, 11, 194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Seppä, T.; Ruotsalainen, M.; Laakso, I.; Tuominen, R.; Ahtee, L. Effect of acute nicotine administration on striatal dopamine output and metabolism in rats kept at different ambient temperatures. Br. J. Pharmacol. 2000, 130, 1147–1155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Bordia, T.; McIntosh, J.M.; Quik, M. The nicotine-mediated decline in l-dopa-induced dyskinesias is associated with a decrease in striatal dopamine release. J. Neurochem. 2013, 125, 291–302. [Google Scholar] [CrossRef] [PubMed]
  140. Leino, S.; Kohtala, S.; Rantamäki, T.; Koski, S.K.; Rannanpää, S.; Salminen, O. Dyskinesia and brain-derived neurotrophic factor levels after long-term levodopa and nicotinic receptor agonist treatments in female mice with near-total unilateral dopaminergic denervation. BMC Neurosci. 2018, 19, 77. [Google Scholar] [CrossRef] [PubMed]
  141. Teng, L.; Hong, F.; Zhang, C.; He, J.; Wang, H. Compound Formula Rehmannia alleviates levodopa-induced dyskinesia in Parkinson’s disease. Neural Regen. Res. 2014, 9, 407–412. [Google Scholar] [PubMed]
  142. Ogawa, M.; Zhou, Y.; Tsuji, R.; Kasahara, J.; Goto, S. Intrastriatal memantine infusion dampens levodopa-induced dyskinesia and motor deficits in a mouse model of hemiparkinsonism. Front. Neurol. 2019, 10, 1258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Lim, S.A.O.; Xia, R.; Ding, Y.; Won, L.; Ray, W.J.; Hitchcock, S.A.; McGehee, D.S.; Kang, U.J. Enhanced histamine H2 excitation of striatal cholinergic interneurons in l-DOPA-induced dyskinesia. Neurobiol. Dis. 2015, 76, 67–76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Cui, G.; Yang, X.; Wang, X.; Zhang, Z.; Yue, X.; Shi, H.; Shen, X. Ranitidine reduced levodopa-induced dyskinesia in a rat model of Parkinson’s disease. Neuropsychiatr. Dis. Treat. 2014, 10, 39–46. [Google Scholar]
  145. Shi, H.; Yang, X.; Zhao, H.; Zhang, S.; Zu, J.; Zhang, W.; Shen, X.; Cui, G.; Hua, F.; Yan, C. Ranitidine reduced levodopa-induced dyskinesia by remodeling neurochemical changes in hemiparkinsonian model of rats. Neuropsychiatr. Dis. Treat. 2015, 11, 1331–1337. [Google Scholar]
  146. Johansson, P.A.; Andersson, M.; Andersson, K.E.; Cenci, M.A. Alterations in cortical and basal ganglia levels of opioid receptor binding in a rat model of l-DOPA-induced dyskinesia. Neurobiol. Dis. 2001, 8, 220–239. [Google Scholar] [CrossRef] [Green Version]
  147. Potts, L.F.; Park, E.S.; Woo, J.-M.; Dyavar Shetty, B.L.; Singh, A.; Braithwaite, S.P.; Voronkov, M.; Papa, S.M.; Mouradian, M.M. Dual κ-agonist/μ-antagonist opioid receptor modulation reduces levodopa-induced dyskinesia and corrects dysregulated striatal changes in the nonhuman primate model of Parkinson disease. Ann. Neurol. 2015, 77, 930–941. [Google Scholar] [CrossRef]
  148. Bartlett, M.J.; So, L.Y.; Szabò, L.; Skinner, D.P.; Parent, K.L.; Heien, M.L.; Vanderah, T.W.; Polt, R.; Sherman, S.J.; Falk, T. Highly-selective µ-opioid receptor antagonism does not block L-DOPA-induced dyskinesia in a rodent model. BMC Res. Notes 2020, 13, 149. [Google Scholar] [CrossRef]
  149. Bezard, E.; Li, Q.; Hulme, H.; Fridjonsdottir, E.; Nilsson, A.; Pioli, E.; Andren, P.E.; Crossman, A.R. µ opioid receptor agonism for L-DOPA-induced dyskinesia in Parkinson’s disease. J. Neurosci. 2020, 40, 6812–6819. [Google Scholar] [CrossRef] [PubMed]
  150. Arcuri, L.; Novello, S.; Frassineti, M.; Mercatelli, D.; Pisanò, C.A.; Morella, I.; Fasano, S.; Journigan, B.V.; Meyer, M.E.; Polgar, W.E.; et al. Anti-Parkinsonian and anti-dyskinetic profiles of two novel potent and selective nociceptin/orphanin FQ receptor agonists. Br. J. Pharmacol. 2018, 175, 782–796. [Google Scholar] [CrossRef] [Green Version]
  151. Megens, A.A.H.P.; Hendrickx, H.M.R.; Mahieu, M.M.A.; Wellens, A.L.Y.; de Boer, P.; Vanhoof, G. PDE10A inhibitors stimulate or suppress motor behavior dependent on the relative activation state of the direct and indirect striatal output pathways. Pharmacol. Res. Perspect. 2014, 2, e00057. [Google Scholar] [CrossRef] [Green Version]
  152. Niccolini, F.; Foltynie, T.; Marques, T.R.; Muhlert, N.; Tziortzi, A.C.; Searle, G.E.; Natesan, S.; Kapur, S.; Rabiner, E.A.; Gunn, R.N.; et al. Loss of phosphodiesterase 10A expression is associated with progression and severity in Parkinson’s disease. Brain 2015, 138, 3003–3015. [Google Scholar] [CrossRef] [Green Version]
  153. Beck, G.; Maehara, S.; Chang, P.Y.; Papa, S.M. A selective phosphodiesterase 10A inhibitor reduces L-dopa-induced dyskinesias in Parkinsonian monkeys. Mov. Disord. 2018, 33, 805–814. [Google Scholar] [CrossRef]
  154. Sellnow, R.C.; Newman, J.H.; Chambers, N.; West, A.R.; Steece-Collier, K.; Sandoval, I.M.; Benskey, M.J.; Bishop, C.; Manfredsson, F.P. Regulation of dopamine neurotransmission from serotonergic neurons by ectopic expression of the dopamine D2 autoreceptor blocks levodopa-induced dyskinesia. Acta Neuropathol. Commun. 2019, 7, 8. [Google Scholar] [CrossRef] [PubMed]
  155. Marongiu, R.; Arango-Lievano, M.; Francardo, V.; Morgenstern, P.; Zhang, X.; Cenci, M.A.; Svenningsson, P.; Greengard, P.; Kaplitt, M.G. Gene therapy blockade of dorsal striatal p11 improves motor function and dyskinesia in parkinsonian mice. Proc. Natl. Acad. Sci. USA 2016, 113, 1423–1428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Breger, L.S.; Kienle, K.; Smith, G.A.; Dunnett, S.B.; Lane, E.L. Influence of chronic L-DOPA treatment on immune response following allogeneic and xenogeneic graft in a rat model of Parkinson’s disease. Brain Behav. Immun. 2017, 61, 155–164. [Google Scholar] [CrossRef] [PubMed]
  157. Ahmed, M.R.; Bychkov, E.; Kook, S.; Zurkovsky, L.; Dalby, K.N.; Gurevich, E.V. Overexpression of GRK6 rescues l-DOPA-induced signaling abnormalities in the dopamine-depleted striatum of hemiparkinsonian rats. Exp. Neurol. 2015, 266, 42–54. [Google Scholar] [CrossRef] [Green Version]
  158. Şensoy, Ö. The single nucleotide β -arrestin2 variant, A248T, resembles dynamical properties of activated arrestin. Turk. J. Chem. 2020, 44, 409–420. [Google Scholar] [CrossRef] [PubMed]
  159. Zhang, X.R.; Zhang, Z.-R.; Chen, S.-Y.; Wang, W.-W.; Wang, X.-S.; He, J.-C.; Xie, C.-L. β-arrestin2 alleviates L-dopa–induced dyskinesia via lower D1R activity in Parkinson’s rats. Aging 2019, 11, 12315–12327. [Google Scholar] [CrossRef] [PubMed]
  160. Urs, N.M.; Bido, S.; Peterson, S.M.; Daigle, T.L.; Bass, C.E.; Gainetdinov, R.R.; Bezard, E.; Caron, M.G. Targeting β-arrestin2 in the treatment of l-DOPA–induced dyskinesia in Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2015, 112, E2517–E2526. [Google Scholar] [CrossRef] [Green Version]
  161. Park, H.-Y.; Kang, Y.-M.; Kang, Y.; Park, T.-S.; Ryu, Y.-K.; Hwang, J.-H.; Kim, Y.-H.; Chung, B.-H.; Nam, K.-H.; Kim, M.-R. Inhibition of adenylyl cyclase type 5 prevents l-DOPA-induced dyskinesia in an animal model of Parkinson’s disease. J. Neurosci. 2014, 34, 11744–11753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Leino, S.; Koski, S.K.; Hänninen, R.; Tapanainen, T.; Rannanpää, S.; Salminen, O. Attenuated dopaminergic neurodegeneration and motor dysfunction in hemiparkinsonian mice lacking the α5 nicotinic acetylcholine receptor subunit. Neuropharmacology 2018, 138, 371–380. [Google Scholar] [CrossRef]
  163. Yang, X.; Wu, N.; Song, L.; Liu, Z. Intrastriatal injections of KN-93 ameliorates levodopa-induced dyskinesia in a rat model of Parkinson’s disease. Neuropsychiatr. Dis. Treat. 2013, 9, 1213–1220. [Google Scholar]
  164. Plotkin, J.L.; Shen, W.; Rafalovich, I.; Sebel, L.E.; Day, M.; Chan, C.S.; Surmeier, D.J. Regulation of dendritic calcium release in striatal spiny projection neurons. J. Neurophysiol. 2013, 110, 2325–2336. [Google Scholar] [CrossRef] [Green Version]
  165. Steece-Collier, K.; Stancat, J.A.; Collier, N.J.; Sandoval, I.M.; Mercado, N.M.; Sortwell, C.E.; Collier, T.J.; Manfredsson, F.P. Genetic silencing of striatal CaV1.3 prevents and ameliorates levodopa dyskinesia. Mov. Disord. 2019, 34, 697–707. [Google Scholar] [CrossRef]
  166. Martinez, A.A.; Morgese, M.G.; Pisanu, A.; Macheda, T.; Paquette, M.A.; Seillier, A.; Cassano, T.; Carta, A.R.; Giuffrida, A. Activation of PPAR gamma receptors reduces levodopa-induced dyskinesias in 6-OHDA-lesioned rats. Neurobiol. Dis. 2015, 74, 295–304. [Google Scholar] [CrossRef] [Green Version]
  167. Fernandez, H.H.; Vanagunas, A.; Odin, P.; Espay, A.J.; Hauser, R.A.; Standaert, D.G.; Chatamra, K.; Benesh, J.; Pritchett, Y.; Hass, S.L.; et al. Levodopa-carbidopa intestinal gel in advanced Parkinson’s disease open-label study: Interim results. Parkinsonism Relat. Disord. 2013, 19, 339–345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Poewe, W.; Bergmann, L.; Kukreja, P.; Robieson, W.Z.; Antonini, A. Levodopa-carbidopa intestinal gel monotherapy: GLORIA registry demographics, efficacy, and safety. J. Parkinsons Dis. 2019, 9, 531–541. [Google Scholar] [CrossRef] [Green Version]
  169. Băjenaru, O.; Ene, A.; Popescu, B.O.; Szász, J.A.; Sabău, M.; Mureşan, D.F.; Perju-Dumbrava, L.; Popescu, C.D.; Constantinescu, A.; Buraga, I.; et al. The effect of levodopa-carbidopa intestinal gel infusion long-term therapy on motor complications in advanced Parkinson’s disease: A multicenter Romanian experience. J. Neural Transm. 2016, 123, 407–414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Lopiano, L.; Modugno, N.; Marano, P.; Sensi, M.; Meco, G.; Cannas, A.; Gusmaroli, G.; Tamma, F.; Mancini, F.; Quatrale, R.; et al. Motor outcomes in patients with advanced Parkinson’s disease treated with levodopa/carbidopa intestinal gel in Italy: An interim analysis from the GREENFIELD observational study. Neurol. Sci. 2016, 37, 1785–1792. [Google Scholar] [CrossRef] [PubMed]
  171. Lopiano, L.; Modugno, N.; Marano, P.; Sensi, M.; Meco, G.; Solla, P.; Gusmaroli, G.; Tamma, F.; Mancini, F.; Quatrale, R.; et al. Motor and non-motor outcomes in patients with advanced Parkinson’s disease treated with levodopa/carbidopa intestinal gel: Final results of the GREENFIELD observational study. J. Neurol. 2019, 266, 2164–2176. [Google Scholar] [CrossRef] [Green Version]
  172. Fernandez, H.H.; Standaert, D.G.; Hauser, R.A.; Lang, A.E.; Fung, V.S.C.; Klostermann, F.; Lew, M.F.; Odin, P.; Steiger, M.; Yakupov, E.Z.; et al. Levodopa-carbidopa intestinal gel in advanced Parkinson’s disease: Final 12-month, open-label results. Mov. Disord. 2015, 30, 500–509. [Google Scholar] [CrossRef] [PubMed]
  173. Thakkar, S.; Fung, V.S.C.; Merola, A.; Rollins, M.; Soileau, M.J.; Kovács, N. 24-hour levodopa-carbidopa intestinal gel: Clinical experience and practical recommendations. CNS Drugs 2021, 35, 137–149. [Google Scholar] [CrossRef] [PubMed]
  174. Cruse, B.; Morales-Briceño, H.; Chang, F.C.F.; Mahant, N.; Ha, A.D.; Kim, S.D.; Wolfe, N.; Kwan, V.; Tsui, D.S.; Griffith, J.M.; et al. 24-hour levodopa-carbidopa intestinal gel may reduce troublesome dyskinesia in advanced Parkinson’s disease. NPJ Parkinsons Dis. 2018, 4, 34. [Google Scholar] [CrossRef] [PubMed]
  175. Olanow, C.W.; Kieburtz, K.; Odin, P.; Espay, A.J.; Standaert, D.G.; Fernandez, H.H.; Vanagunas, A.; Othman, A.A.; Widnell, K.L.; Robieson, W.Z.; et al. Continuous intrajejunal infusion of levodopa-carbidopa intestinal gel for patients with advanced Parkinson’s disease: A randomised, controlled, double-blind, double-dummy study. Lancet Neurol. 2014, 13, 141–149. [Google Scholar] [CrossRef] [Green Version]
  176. Fabbri, M.; Zibetti, M.; Beccaria, L.; Merola, A.; Romagnolo, A.; Montanaro, E.; Ferreira, J.J.; Palermo, S.; Lopiano, L. Levodopa/carbidopa intestinal gel infusion and weight loss in Parkinson’s disease. Eur. J. Neurol. 2019, 26, 490–496. [Google Scholar] [CrossRef] [Green Version]
  177. Antonini, A.; Yegin, A.; Preda, C.; Bergmann, L.; Poewe, W.; GLORIA Study Investigators and Coordinators. Global long-term study on motor and non-motor symptoms and safety of levodopa-carbidopa intestinal gel in routine care of advanced Parkinson’s disease patients; 12-month interim outcomes. Parkinsonism Relat. Disord. 2015, 21, 231–235. [Google Scholar] [CrossRef]
  178. Morgante, F.; Oppo, V.; Fabbri, M.; Olivola, E.; Sorbera, C.; De Micco, R.; Ielo, G.C.; Colucci, F.; Bonvegna, S.; Novelli, A.; et al. Levodopa-carbidopa intrajejunal infusion in Parkinson’s disease: Untangling the role of age. J. Neurol. 2021, 268, 1728–1737. [Google Scholar] [CrossRef]
  179. Meloni, M.; Solla, P.; Mascia, M.M.; Marrosu, F.; Cannas, A. Diphasic dyskinesias during levodopa-carbidopa intestinal gel (LCIG) infusion in Parkinson’s disease. Parkinsonism Relat. Disord. 2017, 37, 92–96. [Google Scholar] [CrossRef]
  180. Santos García, D.; Martínez Castrillo, J.C.; Puente Périz, V.; Seoane Urgorri, A.; Fernández Díez, S.; Benita León, V.; Udaeta Baldivieso, B.; Campolongo Perillo, A.; Mariscal Pérez, N. Clinical management of patients with advanced Parkinson’s disease treated with continuous intestinal infusion of levodopa/carbidopa. Neurodegener. Dis. Manag. 2016, 6, 187–202. [Google Scholar] [CrossRef] [Green Version]
  181. García Ruiz, P.J.; Sesar Ignacio, A.; Ares Pensado, B.; Castro García, A.; Alonso Frech, F.; Alvarez López, M.; Arbelo González, J.; Baiges Octavio, J.; Burguera Hernández, J.A.; Calopa Garriga, M.; et al. Efficacy of long-term continuous subcutaneous apomorphine infusion in advanced Parkinson’s disease with motor fluctuations: A multicenter study. Mov. Disord. 2008, 23, 1130–1136. [Google Scholar] [CrossRef]
  182. Houvenghel, J.-F.; Drapier, S.; Duprez, J.; Robert, G.H.; Riou, A.; Drapier, D.; Sauleau, P.; Vérin, M. Effects of continuous subcutaneous apomorphine infusion in Parkinson’s disease without cognitive impairment on motor, cognitive, psychiatric symptoms and quality of life. J. Neurol. Sci. 2018, 395, 113–118. [Google Scholar] [CrossRef]
  183. Katzenschlager, R.; Poewe, W.; Rascol, O.; Trenkwalder, C.; Deuschl, G.; Chaudhuri, K.R.; Henriksen, T.; van Laar, T.; Lockhart, D.; Staines, H.; et al. Long-term safety and efficacy of apomorphine infusion in Parkinson’s disease patients with persistent motor fluctuations: Results of the open-label phase of the TOLEDO study. Parkinsonism Relat. Disord. 2021, 83, 79–85. [Google Scholar] [CrossRef]
  184. Henriksen, T.; Staines, H. Continuous subcutaneous apomorphine infusion in Parkinson’s Disease: A single-center, long-term follow-up study of the causes for discontinuation. J. Pers. Med. 2021, 11, 525. [Google Scholar] [CrossRef]
  185. Papuć, E.; Trzciniecka, O.; Rejdak, K. Continuous subcutaneous apomorphine monotherapy in Parkinson’s disease. Ann. Agric. Environ. Med. 2019, 26, 133–137. [Google Scholar] [CrossRef] [PubMed]
  186. Drapier, S.; Eusebio, A.; Degos, B.; Vérin, M.; Durif, F.; Azulay, J.P.; Viallet, F.; Rouaud, T.; Moreau, C.; Defebvre, L.; et al. Quality of life in Parkinson’s disease improved by apomorphine pump: The OPTIPUMP cohort study. J. Neurol. 2016, 263, 1111–1119. [Google Scholar] [CrossRef]
  187. Olivola, E.; Fasano, A.; Varanese, S.; Lena, F.; Santilli, M.; Femiano, C.; Centonze, D.; Modugno, N. Continuous subcutaneous apomorphine infusion in Parkinson’s disease: Causes of discontinuation and subsequent treatment strategies. Neurol. Sci. 2019, 40, 1917–1923. [Google Scholar] [CrossRef]
  188. Meira, B.; Degos, B.; Corsetti, E.; Doulazmi, M.; Berthelot, E.; Virbel-Fleischman, C.; Dodet, P.; Méneret, A.; Mariani, L.L.; Delorme, C.; et al. Long-term effect of apomorphine infusion in advanced Parkinson’s disease: A real-life study. NPJ Parkinsons Dis. 2021, 7, 50. [Google Scholar] [CrossRef]
  189. LeWitt, P.A.; Metman, L.V.; Rubens, R.; Khanna, S.; Kell, S.; Gupta, S. Effect of concomitant medications on the safety and efficacy of extended-release carbidopa-levodopa (IPX066) in patients with advanced Parkinson disease: A post hoc analysis. Clin. Neuropharmacol. 2018, 41, 47–55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Stocchi, F.; Hsu, A.; Khanna, S.; Ellenbogen, A.; Mahler, A.; Liang, G.; Dillmann, U.; Rubens, R.; Kell, S.; Gupta, S. Comparison of IPX066 with carbidopa–levodopa plus entacapone in advanced PD patients. Parkinsonism Relat. Disord. 2014, 20, 1335–1340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Kong, M.; Ba, M.; Ren, C.; Yu, L.; Dong, S.; Yu, G.; Liang, H. An updated meta-analysis of amantadine for treating dyskinesia in Parkinson’s disease. Oncotarget 2017, 8, 57316–57326. [Google Scholar] [CrossRef] [Green Version]
  192. Kim, A.; Kim, Y.E.; Yun, J.Y.; Kim, H.-J.; Yang, H.-J.; Lee, W.-W.; Shin, C.W.; Park, H.; Jung, Y.J.; Kim, A.; et al. Amantadine and the risk of dyskinesia in patients with early Parkinson’s disease: An open-label, pragmatic trial. J. Mov. Disord. 2018, 11, 65–71. [Google Scholar] [CrossRef] [PubMed]
  193. Hauser, R.A.; Walsh, R.R.; Pahwa, R.; Chernick, D.; Formella, A.E. Amantadine ER (Gocovri®) significantly increases ON time without any dyskinesia: Pooled analyses from pivotal trials in Parkinson’s disease. Front. Neurol. 2021, 12, 645706. [Google Scholar] [CrossRef] [PubMed]
  194. Pahwa, R.; Tanner, C.M.; Hauser, R.A.; Sethi, K.; Isaacson, S.; Truong, D.; Struck, L.; Ruby, A.E.; McClure, N.L.; Went, G.T. Amantadine extended release for levodopa-induced dyskinesia in Parkinson’s disease (EASED Study). Mov. Disord. 2015, 30, 788–795. [Google Scholar] [CrossRef] [PubMed]
  195. Oertel, W.; Eggert, K.; Pahwa, R.; Tanner, C.M.; Hauser, R.A.; Trenkwalder, C.; Ehret, R.; Azulay, J.P.; Isaacson, S.; Felt, L.; et al. Randomized, placebo-controlled trial of ADS-5102 (amantadine) extended-release capsules for levodopa-induced dyskinesia in Parkinson’s disease (EASE LID 3). Mov. Disord. 2017, 32, 1701–1709. [Google Scholar] [CrossRef]
  196. Tanner, C.M.; Pahwa, R.; Hauser, R.A.; Oertel, W.H.; Isaacson, S.H.; Jankovic, J.; Johnson, R.; Chernick, D.; Hubble, J. EASE LID 2: A 2-year open-label trial of gocovri (amantadine) extended release for dyskinesia in Parkinson’s disease. J. Parkinsons Dis. 2020, 10, 543–558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Hauser, R.A.; Pahwa, R.; Tanner, C.M.; Oertel, W.; Isaacson, S.H.; Johnson, R.; Felt, L.; Stempien, M.J. ADS-5102 (amantadine) extended-release capsules for levodopa-induced dyskinesia in Parkinson’s disease (EASE LID 2 study): Interim results of an open-label safety study. J. Parkinsons Dis. 2017, 7, 511–522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Pahwa, R.; Tanner, C.M.; Hauser, R.A.; Isaacson, S.H.; Nausieda, P.A.; Truong, D.D.; Agarwal, P.; Hull, K.L.; Lyons, K.E.; Johnson, R.; et al. ADS-5102 (amantadine) extended-release capsules for levodopa-induced dyskinesia in Parkinson disease (EASE LID study). JAMA Neurol. 2017, 74, 941–949. [Google Scholar] [CrossRef]
  199. Xu, W.; Wang, X.; Tocker, A.M.; Huang, P.; Reith, M.E.A.; Liu-Chen, L.-Y.; Smith, A.B., 3rd; Kortagere, S. Functional characterization of a novel series of biased signaling dopamine D3 receptor agonists. ACS Chem. Neurosci. 2017, 8, 486–500. [Google Scholar] [CrossRef] [PubMed]
  200. Berthet, A.; Porras, G.; Duodnikoff, E.; Stark, H.; Cador, M.; Bezard, E.; Bloch, B. Pharmacological analysis demonstrates dramatic alteration of D1 dopamine receptor neuronal distribution in the rat analog of l-DOPA-induced dyskinesia. J. Neurosci. 2009, 29, 4829–4835. [Google Scholar] [CrossRef]
  201. Visanji, N.P.; Fox, S.H.; Johnston, T.; Reyes, G.; Millan, M.J.; Brotchie, J.M. Dopamine D3 receptor stimulation underlies the development of L-DOPA-induced dyskinesia in animal models of Parkinson’s disease. Neurobiol. Dis. 2009, 35, 184–192. [Google Scholar] [CrossRef] [PubMed]
  202. Utsumi, H.; Okuma, Y.; Kano, O.; Suzuki, Y.; Iijima, M.; Tomimitsu, H.; Hashida, H.; Kubo, S.-I.; Suzuki, M.; Nanri, K.; et al. Evaluation of the Efficacy of pramipexole for treating levodopa-induced dyskinesia in patients with Parkinson’s disease. Intern. Med. 2013, 52, 325–332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Svenningsson, P.; Johansson, A.; Nyholm, D.; Tsitsi, P.; Hansson, F.; Sonesson, C.; Tedroff, J. Safety and tolerability of IRL790 in Parkinson’s disease with levodopa-induced dyskinesia—A phase 1b trial. NPJ Parkinsons Dis. 2018, 4, 35. [Google Scholar] [CrossRef]
  204. Borgohain, R.; Szasz, J.; Stanzione, P.; Meshram, C.; Bhatt, M.; Chirilineau, D.; Stocchi, F.; Lucini, V.; Giuliani, R.; Forrest, E.; et al. Randomized trial of safinamide add-on to levodopa in Parkinson’s disease with motor fluctuations. Mov. Disord. 2014, 29, 229–237. [Google Scholar] [CrossRef] [PubMed]
  205. Borgohain, R.; Szasz, J.; Stanzione, P.; Meshram, C.; Bhatt, M.H.; Chirilineau, D.; Stocchi, F.; Lucini, V.; Giuliani, R.; Forrest, E.; et al. Two-year, randomized, controlled study of safinamide as add-on to levodopa in mid to late Parkinson’s disease. Mov. Disord. 2014, 29, 1273–1280. [Google Scholar] [CrossRef]
  206. Cattaneo, C.; Jost, W.H.; Bonizzoni, E. Long-term efficacy of safinamide on symptoms severity and quality of life in fluctuating Parkinson’s disease patients. J. Parkinsons Dis. 2020, 10, 89–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Gardoni, F.; Morari, M.; Kulisevsky, J.; Brugnolii, A.; Novello, S.; Pisanò, C.A.; Caccia, C.; Mellone, M.; Melloni, E.; Padoani, G.; et al. Safinamide modulates striatal glutamatergic signaling in a rat model of levodopa-induced dyskinesia. J. Pharmacol. Exp. Ther. 2018, 367, 442–451. [Google Scholar] [CrossRef] [Green Version]
  208. Svenningsson, P.; Rosenblad, C.; Af Edholm Arvidsson, K.; Wictorin, K.; Keywood, C.; Shankar, B.; Lowe, D.A.; Björklund, A.; Widner, H. Eltoprazine counteracts l-DOPA-induced dyskinesias in Parkinson’s disease: A dose-finding study. Brain 2015, 138, 963–973. [Google Scholar] [CrossRef]
  209. Meloni, M.; Puligheddu, M.; Sanna, F.; Cannas, A.; Farris, R.; Tronci, E.; Figorilli, M.; Defazio, G.; Carta, M. Efficacy and safety of 5-hydroxytryptophan on levodopa-induced motor complications in Parkinson’s disease: A preliminary finding. J. Neurol. Sci. 2020, 415, 116869. [Google Scholar] [CrossRef]
  210. Mestre, T.A.; Shah, B.B.; Connolly, B.S.; de Aquino, C.; Al Dhakeel, A.; Walsh, R.; Ghate, T.; Lui, L.P.; Fox, S.H. Famotidine, a histamine H2 receptor antagonist, does not reduce levodopa-induced dyskinesia in Parkinson’s disease: A proof-of-concept study. Mov. Disord. Clin. Pract. 2014, 1, 219–224. [Google Scholar] [CrossRef]
  211. Fan, S.-Y.; Wang, K.-L.; Hu, W.; Eisinger, R.S.; Han, A.; Han, C.-L.; Wang, Q.; Michitomo, S.; Zhang, J.-G.; Weng, F.; et al. Pallidal versus subthalamic nucleus deep brain stimulation for levodopa-induced dyskinesia. Ann. Clin. Transl. Neurol. 2020, 7, 59–68. [Google Scholar] [CrossRef]
  212. Liu, Y.; Li, F.; Luo, H.; He, Q.; Chen, L.; Cheng, Y.; Zhang, W.; Xie, Z. Improvement of deep brain stimulation in dyskinesia in Parkinson’s disease: A meta-analysis. Front. Neurol. 2019, 10, 151. [Google Scholar] [CrossRef] [Green Version]
  213. Ryu, H.-S.; Kim, M.-S.; You, S.; Kim, M.-J.; Kim, Y.-J.; Kim, J.; Kim, K.; Chung, S.J. Comparison of pallidal and subthalamic deep brain stimulation in Parkinson’s disease: Therapeutic and adverse effects. J. Mov. Disord. 2017, 10, 80–86. [Google Scholar] [CrossRef] [Green Version]
  214. Schuepbach, W.M.M.; Rau, J.; Knudsen, K.; Volkmann, J.; Krack, P.; Timmermann, L.; Hälbig, T.D.; Hesekamp, H.; Navarro, S.M.; Meier, N.; et al. Neurostimulation for Parkinson’s disease with early motor complications. N. Engl. J. Med. 2013, 368, 610–622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Weaver, F.M.; Follett, K.A.; Stern, M.; Luo, P.; Harris, C.L.; Hur, K.; Marks, W.J., Jr.; Rothlind, J.; Sagher, O.; Moy, C.; et al. Randomized trial of deep brain stimulation for Parkinson disease. Thirty-six-month outcomes. Neurology 2012, 79, 55–65. [Google Scholar] [CrossRef] [Green Version]
  216. Sobstyl, M.; Zabek, M.; Zaczynski, A.; Gorecki, W.; Mossakowski, Z.; Brzuszkiewicz-Kuzmicka, G. Unilateral subthalamic nucleus stimulation in the treatment of asymmetric Parkinson’s disease with early motor complications. Turk. Neurosurg. 2017, 27, 294–300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Kim, J.H.; Chang, W.S.; Jung, H.H.; Chang, J.W. Effect of subthalamic deep brain stimulation on levodopa-induced dyskinesia in Parkinson’s disease. Yonsei Med. J. 2015, 56, 1316–1321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Lohse, A.; Meder, D.; Nielsen, S.; Lund, A.E.; Herz, D.M.; Løkkegaard, A.; Siebner, H.R. Low-frequency transcranial stimulation of pre-supplementary motor area alleviates levodopa-induced dyskinesia in Parkinson’s disease: A randomized cross-over trial. Brain Commun. 2020, 2, fcaa147. [Google Scholar] [CrossRef]
  219. Ni, Z.; Chen, R. Transcranial magnetic stimulation to understand pathophysiology and as potential treatment for neurodegenerative diseases. Transl. Neurodegener. 2015, 4, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hutny, M.; Hofman, J.; Klimkowicz-Mrowiec, A.; Gorzkowska, A. Current Knowledge on the Background, Pathophysiology and Treatment of Levodopa-Induced Dyskinesia—Literature Review. J. Clin. Med. 2021, 10, 4377. https://doi.org/10.3390/jcm10194377

AMA Style

Hutny M, Hofman J, Klimkowicz-Mrowiec A, Gorzkowska A. Current Knowledge on the Background, Pathophysiology and Treatment of Levodopa-Induced Dyskinesia—Literature Review. Journal of Clinical Medicine. 2021; 10(19):4377. https://doi.org/10.3390/jcm10194377

Chicago/Turabian Style

Hutny, Michał, Jagoda Hofman, Aleksandra Klimkowicz-Mrowiec, and Agnieszka Gorzkowska. 2021. "Current Knowledge on the Background, Pathophysiology and Treatment of Levodopa-Induced Dyskinesia—Literature Review" Journal of Clinical Medicine 10, no. 19: 4377. https://doi.org/10.3390/jcm10194377

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop