Next Article in Journal
The Role of Pancreatic Stone Protein (PSP) as a Biomarker of Pregnancy-Related Diseases
Previous Article in Journal
Anatomical and Functional Connectivity of Critical Deep Brain Structures and Their Potential Clinical Application in Brain Stimulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Levodopa-Induced Dyskinesias in Parkinson’s Disease: An Overview on Pathophysiology, Clinical Manifestations, Therapy Management Strategies and Future Directions

by
Lazzaro di Biase
1,2,*,
Pasquale Maria Pecoraro
1,3,
Simona Paola Carbone
1,3,
Maria Letizia Caminiti
1,3 and
Vincenzo Di Lazzaro
1,3
1
Fondazione Policlinico Universitario Campus Bio-Medico, Via Alvaro del Portillo, 200, 00128 Roma, Italy
2
Brain Innovations Lab, Università Campus Bio-Medico di Roma, Via Álvaro del Portillo, 21, 00128 Rome, Italy
3
Unit of Neurology, Neurophysiology, Neurobiology and Psichiatry, Department of Medicine and Surgery, Università Campus Bio-Medico di Roma, Via Alvaro del Portillo, 21, 00128 Roma, Italy
*
Author to whom correspondence should be addressed.
J. Clin. Med. 2023, 12(13), 4427; https://doi.org/10.3390/jcm12134427
Submission received: 21 May 2023 / Revised: 18 June 2023 / Accepted: 27 June 2023 / Published: 30 June 2023
(This article belongs to the Section Clinical Neurology)

Abstract

:
Since its first introduction, levodopa has become the cornerstone for the treatment of Parkinson’s disease and remains the leading therapeutic choice for motor control therapy so far. Unfortunately, the subsequent appearance of abnormal involuntary movements, known as dyskinesias, is a frequent drawback. Despite the deep knowledge of this complication, in terms of clinical phenomenology and the temporal relationship during a levodopa regimen, less is clear about the pathophysiological mechanisms underpinning it. As the disease progresses, specific oscillatory activities of both motor cortical and basal ganglia neurons and variation in levodopa metabolism, in terms of the dopamine receptor stimulation pattern and turnover rate, underlie dyskinesia onset. This review aims to provide a global overview on levodopa-induced dyskinesias, focusing on pathophysiology, clinical manifestations, therapy management strategies and future directions.

1. Introduction

After the description of the pathological correlate of parkinsonian syndrome, characterized by the degeneration of the dopaminergic neurons of the substantia nigra pars compacta (SNc) [1], the main goal of Parkinson’s disease (PD) treatment is to restore dopaminergic transmission [2,3]. Although several options to improve synaptic dopaminergic transmission are available, such as dopamine receptor agonists, COMT inhibitors and MAO-B inhibitors, levodopa still remains the most effective pharmacological treatment for PD [3].
Abnormal involuntary movements, or dyskinesias, along with motor fluctuations, are motor complications of chronic levodopa treatment and neurodegeneration, causing impairment in the quality of life [4]. Even in the first descriptions of levodopa effects in PD patients, in 1967-69, Cotzias and Papavasiliou [5] described the presence of involuntary movements after levodopa administration, during the peak of therapeutic effect [2] in particular in patients with the longest disease duration. The pathophysiology of dyskinesias is still far from being clearly elucidated. The key risk factors for developing such abnormal involuntary movements are levodopa dose and disease duration/severity [6,7]. The non-physiologic intermittent pulsatile stimulation of dopamine (DA) receptors and the subsequent changes in proteins and genes produce alterations in the neuronal firing patterns [8,9,10], which are well-known dyskinesia-contributing factors.
The application of invasive neurophysiological techniques opened the way to closely recording neuronal firing activity [11]. Specific oscillatory activities of both cortical areas and basal ganglia neurons superimposed to dyskinesia appearance have been described, reinforcing the hypothesis of a pivotal role of such oscillations in the genesis of levodopa-induced dyskinesias (LIDs) [12,13,14,15,16,17,18,19]. Recognizing troublesome dyskinesias is crucial for the movement disorder specialist, since their presence is considered a marker of advanced Parkinson’s disease and an indicator of the eligibility for device-aided therapy [20].
In the present review, we aimed to summarize the current state of the art in the field of LIDs. A global overview on LIDs is provided, with a focus on the most recent evidence about clinical aspects, pathophysiology and neurophysiology. Conventional strategies for the therapeutic management of LIDs along with new possible therapeutic approaches are reported.

2. Levodopa-Induced Dyskinesias

2.1. Clinical Risk Factors

The main risk factors for developing LIDs are younger age, female sex (biased by a lower weight), longer disease duration, longer PD duration before the initiation of levodopa, more advanced disease and the dose and duration of levodopa treatment. Studies in the literature agree that levodopa therapy carries a higher risk than DA [6,7]. An RCT, ELLDOPA study, demonstrated how patients receiving a higher levodopa dosage (>300 mg) showed a significantly higher risk of developing LIDs [21]. Abnormal involuntary movements do not occur exclusively with levodopa therapy: dopamine agonists [22,23,24,25,26,27], COMT inhibitors [28] and MAO inhibitors [29] could be correlated with such abnormal movements.

2.2. Genetic Risk Factors

Because LID susceptibility varies greatly from patient to patient, it is possible that genetic factors contribute to the development of LID. Genes involved in dopamine metabolism [30], transport [31] and signaling [32], as well as genes involved in glutamate transmission [33] and synaptic plasticity [34], have been linked to an increased risk of developing LID [35,36].
The following table (Table 1) lists some of the important genes connected to the genetic risk of LID.
A recent systematic review and meta-analysis of genetic factors related to LIDs analyzed 33 studies including a total of 27,092 subjects of different ethnicities. They analyzed 37 genes (22 possibly associated with dyskinesia) and 158 variants (94 possibly related to dyskinesia). The studies reviewed demonstrated inconsistent results, and the meta-analysis failed to demonstrate any association between genetic factors and LID susceptibility [37].
It is crucial to remember that the genetic component of LID is complex and involves the interaction of several genes in addition to the significant influence of non-genetic factors.

2.3. Epidemiology

From the first introduction of levodopa, in the late 1960s, abnormal involuntary movements have been reported in about half of patients after 6 months of treatment [38]. Clinical studies [7,39,40], literature meta-analysis [41] and trials [21,22,23,24,28,42,43,44,45] that investigated the phenomenon of levodopa-induced dyskinesias, focusing on the temporal delay between levodopa-based therapy initiation, the starting time of abnormal involuntary movements and the rate of the occurrence of dyskinesias, are expressed as a scatter plot in Figure 1 and listed in Table 2.
In the study of Ahlskog and Muenter [41], the authors analyzed the cumulative frequency of levodopa-induced dyskinesias through a meta-analysis of the literature during discrete intervals of treatment from 1966 to September 2000, including time-series pre- and post-levodopa availability. The authors concluded that patients with a diagnosis of PD prior to levodopa introduction had a higher frequency of levodopa-induced dyskinesias in the early phase of treatment, with 53.4% of them experiencing involuntary movements just 5–6 months after starting levodopa. This contrasts with modern-era series, in which the occurrence of abnormal involuntary movements was negligible for the first 1–2 years on levodopa. At 4–6 years of follow-up, patients that started antiparkinsonian therapy based on a levodopa regimen still had lower median dyskinesia frequencies (36–39%). Conversely, dyskinesias were present in 90% of PD patients in the only three modern-era series that reported the cumulative dyskinesia frequency over a 9-year period of therapy with levodopa. Of note, in a 5-year, double-blind study, PD patients randomized ab initio to levodopa showed a three-fold probability of exhibiting dyskinesias than subjects initially randomized to Ropirinole, and the median time to dyskinesia appearance was significantly longer for the group that firstly assumed the dopamine agonist. After a 5-year follow-up, 45% of patients randomized to initial therapy with levodopa developed involuntary movements, while, after 10 years on levodopa, dyskinesias occurred in 77.8%. Additionally, the median time to dyskinesia onset was 7 years in subjects originally treated with levodopa [24]. It is worth noting that Rajput and Fenton [39] designed the only reported study so far that highlighted the pattern of the motor complications of long-term levodopa treatment in pathologically confirmed PD. The study concluded that after 14.3 years of levodopa exposure, dyskinesias occurred in 61.9% of patients and represented the earliest and most common adverse effects of levodopa. In the ELLDOPA study [21], early PD patients were randomized to receive different daily carbidopa–levodopa regimens up to 150 mg and 600 mg or assigned to the placebo group for a period of 40 weeks, followed by a withdrawal of 2 weeks. The higher the levodopa dose was administered, the higher the incidence of dyskinesias: abnormal involuntary movements occurred in 16% of patients after taking 600 mg/die of levodopa for 8 months compared to 3% of subjects that were on 150 mg/die. Finally, it is worth citing the Sydney Multicenter Study [45] because it was the only study to deeply define the clinical characteristics of dyskinesias, providing information about their severity. In this study, 149 previously untreated PD patients were randomized to initial blinded treatment with bromocriptine or levodopa. Of the 52 patients surviving at least 15 years, 94% experienced dyskinesias. Moreover, 6 of the 52 subjects (12%) had severe involuntary movements with a Unified Parkinson’s Disease Rating Scale (UPDRS) item 33 of 3. Levodopa-induced dyskinesias started after a mean duration of treatment of 4.2 years for the levodopa group.

2.4. Clinical Manifestations

LIDs are arrhythmic, not stereotyped, casual, discontinuous, involuntary movements that can involve any body district and in particular limbs, the head and the trunk. From a phenomenological point of view, LIDs can appear in different ways as choreic, ballistic, athetoid or dystonic movements. They are present at rest and can be increased by voluntary movements, cognitive activities and emotional stressors.
ON-period dyskinesias are more common than OFF-period dyskinesias, are choreiform or choreoathetoid in nature and can sometimes even be ballistic [46].
Levodopa-induced dyskinesias can be further classified into peak-dose dyskinesias, occurring during the ON phase in coincidence with the peak plasma level of levodopa and affecting mainly the axial musculature and proximal upper limbs. Square-wave dyskinesias are a particularly severe form of ON dyskinesias: this kind of LID occupies the entire ON phase, and the patient does not experience the ON motor state without dyskinetic movements [47].
Another type of dyskinesia, diphasic dyskinesia [48], emerges at lower levodopa plasma levels during the transitions from ON to OFF phases and from OFF to ON phases, involving lower limbs asymmetrically. Diphasic dyskinesias characteristically affect lower limbs with repetitive rapidly alternating dystonic flexion/extension foot movements or leg kicking in a stereotyped pattern, often associated with high-stepping and bizarre gaits [49]. Additionally, it was found that patients who have never had diphasic dyskinesia run the chance of developing it after starting an intestinal infusion of levodopa/carbidopa monotherapy [50].
The last mentioned type occurs only during the OFF phase in the phenomenological form of dystonia [46,51,52,53,54] (Figure 2).

2.5. Objective LID Monitoring

Clinical evaluation remains the gold standard for motor symptom identification and diagnosis in routine clinical practice. However, new technologies, like wearable motion sensor devices, are opening new ways not only for continuous at-home symptom monitoring [55,56] but also for the objective and quantitative description of PD motor symptoms [57,58,59], like tremors [60,61,62,63], bradykinesia [64,65,66], rigidity [66,67,68,69], gait, balance and postural issues [70,71,72,73,74,75], alongside motor complications like motor fluctuations and dyskinesias [76,77,78]. Additionally, among non-wearable sensors, video-based systems represent a reliable solution to assess the features of LIDs [79]. In addition, data science with the development of artificial intelligence and machine learning algorithms will further improve the diagnostic process [80,81,82], motor symptom identification [83,84,85] and the management and optimization of the therapy to avoid motor complications [86] (Figure 3).

2.6. Pathophysiology

Levodopa Pharmacokinetics and Pharmacodynamics

Oral levodopa has a bioavailability of 30% because it is highly metabolized into DA by peripheral Aromatic L-amino acid decarboxylase (AADC) expressed in the gut. Concomitant administration of AADC peripheral inhibitors (AADCI) increases the bioavailability of levodopa up to three times, reducing the required therapeutic dose [87]. Levodopa competes with the transport system of neutral amino acids both in the intestinal mucosa and the blood–brain barrier [87]. Meals with a high protein intake increase the plasmatic concentrations of neutral amino acids, reducing the absorption of levodopa and its therapeutic effect [88]; therefore, the levodopa administration regimen should be adapted to the mealtime in order to enhance its therapeutic effect [89]. In addition, mathematical models showed that not all amino acids compete with levodopa absorption and that a serine-rich diet could even improve the bioavailability by 22% compared with the ante cibum administration [90]. The evolution of pharmacokinetic and pharmacodynamic parameters with disease progression has been widely investigated in the literature [87,91,92,93,94,95,96,97]. Variation in pharmacokinetic values in advanced PD stages is controversial: some studies evidenced no change in levodopa kinetics for advancing disease [87,95,96,98,99], while others demonstrated that levodopa pharmacokinetic parameters may be useful to determine disease severity and the duration of Parkinson’s disease [94,100]. For example, in the work of Adamiak and Kaldonska [94], higher Cmax and AUC values were observed in patients with more advanced Hoehn and Yahr stages. In the same study, the authors further demonstrated a correlation between the age of patients and Tmax, while disease duration was directly related to AUC. Moreover, Nyholm [101] hypothesized that a reduced activity of the levodopa metabolizing enzymes may increase the AUC in patients on longstanding levodopa therapy. This is particularly true for the elderly, probably related to a reduced clearance. Conversely, as demonstrated by Contin and Riva [97], the latency of response and the duration of response to a standard levodopa test dose are significantly shortened with disease progression, while the magnitude of the effect is unchanged [98] or even enhanced [87,95,102,103]. Contin et al. [93,104] highlighted the progressive reduction in levodopa half-life with worsening of the disease over the years and with a negative correlation with the severity of symptoms. The authors proposed the computed half-life of levodopa as an indicator of nigrostriatal dopaminergic functionality and integrity [93,104]. In the work of Triggs and Charles [100], a higher degree of drug receptor occupancy and receptor desensitization indicated an advanced disease stage, while Adamiak and Kaldonska [94] demonstrated a correlation between disease duration and EC50. It has been further postulated that higher EC50 values are underpinned by most advanced disease stages, associated with dyskinesias and motor fluctuations [100,105]. Finally, as highlighted by Adamiak and Kaldonska [94], EC50 itself suggests drug sensitivity, while its changes can express dopamine availability.

2.7. Neurophysiology

Neurophysiology techniques are useful to understand the pathophysiology that underlies LIDs. Non-invasive techniques, like electroencephalogram [106,107,108], let one explore the cortical oscillations related to Parkinson’s disease, while invasive techniques, like local field potential (LFP) recordings through deep brain stimulation (DBS) electrodes, provide a more informative insight into basal ganglia’s pathological oscillations [11].
Indeed, LIDs have been associated with changes in the electrophysiological activity of the motor cortex and basal ganglia circuitry. Both in vivo single-cell recordings and LFPs highlighted the main characteristics of the neuronal firing pattern during LIDs. Single-cell recording using microelectrodes provides information about the frequency and pattern of the discharge of single neurons [109]. The classical model of the basal ganglia function considers LIDs as a result of the over-decreased neuronal firing rates of the globus pallidus internus (GPi), leading to the increased activity of thalamocortical motor circuits [110,111]. These observations have been confirmed in human studies in intra-operatively induced dyskinesias by the administration of apomorphine during pallidotomy in PD patients. A reduction in GPi firing rate was recorded with respect to the OFF state [112,113,114], while no difference was observed between the ON and dyskinetic states [112,113]. The difference in neuronal firing activity concerns not only the frequency but also the pattern of discharge. In ON-state dyskinesias, an increment of burst-like and irregular discharges was observed compared to the OFF state [112,113,114]. Regarding subthalamic nucleus (STN) recording, the mean firing rate of the neurons was not reduced during the ON state without dyskinesias compared to the OFF state. Conversely, it was significantly reduced during LIDs with an increment of spikes in a burst that was absent in the ON state without dyskinesias [113]. The role of the globus pallidus externus (GPe) in LIDs is still unclear. Lozano et al. [115] showed an increment in the firing rate of 50–90%, but further confirmations are needed [115]. The in vivo recording of neuronal oscillatory activity in the GPi, STN and substantia nigra pars reticulata (SNr) is obtained through implanted DBS electrodes. In the STN, peak-dose dyskinesias were associated with an increment in the power of the theta-alpha (4–10 Hz) band with a mean frequency of 8.38 Hz [16,116]. This finding is quite specific and has been confirmed by the increment of the power in the theta-alpha band only when patients exhibited dyskinesia and not during the ON period without dyskinesia. Moreover, in patients with unilateral dyskinesias, this kind of oscillatory activity has been recorded only in the contralateral STN [16]. Patients with diphasic dyskinesias present the same type of oscillatory activity [17]. Peak-dose dyskinesia was associated with theta-alpha activity recorded through electrodes in the dorsal portion of the STN, also known as the motor region [117]. Concerning the GPi, a negative correlation was observed between LFP power in the 8–40 Hz band and the beta bands and LIDs in two patients [118]. LFP oscillations within and coherence between the GPi and STN at low frequencies (<10 Hz) are observed only contralateral to the side of dyskinesia [116]. On the contrary, beta oscillatory activity correlated with the parkinsonian state, rigidity and bradykinesia [119,120]. Recordings of the SNr showed a pathological plasticity and a loss of ability to depotentiate at the output nuclei in patients with dyskinesias [121]. The relationship between hypersynchronization in the gamma frequency band and levodopa-induced dyskinesias has been recently unveiled in the literature [12,13,14,15,16,17,18,19], and the attention on this topic has increased significantly due to its possible clinical and therapeutic implications. For example, Swann and de Hemptinne [12] explored the neuronal activity patterns of dyskinesia by applying a totally implanted multisite brain-recording device in two PD patients treated with DBS, followed over one year. Motor cortex electrocorticography (ECoG) and STN LFP recordings showed that the dyskinetic state was associated with the emergence of a narrowband gamma oscillation both in the motor cortex and STN, pointing out a strong phase coherence. Furthermore, in the study of Halje and Tamte [13], cortical and striatal signals were recorded in a hemiparkinsonian 6-OHDA-induced rat model [122]. Abnormal involuntary movements were neurophysiologically underpinned by a resonant LFP oscillation at 80 Hz in the motor cortex and the striatum of the lesioned hemisphere 10–20 min after receiving intraperitoneal levodopa formulations. Interestingly, this narrowband oscillation was not detectable in the non-dyskinetic animals either ON or OFF levodopa or in the intact hemisphere of any animal [13]. In the same toxin-induced mouse model, Güttler and Altschüler [14] investigated the association of M1 ECoG and motor performance in rats during 21 days of daily treatment with levodopa/benserazide. After levodopa administration, subsequent involuntary movements were accompanied by an increase in cortical narrowband high-gamma oscillations above M1 with an average frequency of 97 Hz. The authors further showed that the gamma power spectrum significantly correlated with the clinical score for abnormal involuntary movements [14]. Narrowband gamma oscillations are not peculiar to levodopa-induced dyskinesias in PD, since such oscillatory activity has been proven also in patients affected by dystonia and myoclonus epilepsy at rest [14,123,124]. A brief overview on the neurophysiological markers of LIDs is schematically depicted in Figure 4.

2.8. Neurotransmitter Systems

Insights for the understanding of LID pathophysiology should arise from recent evidence on neurotransmitter modulatory systems that may influence the dopaminergic transmission under peculiar circumstances. Non-dopaminergic pathways can regulate the dopaminergic transmission of the basal ganglia with direct and indirect mechanisms or can be additionally implied in the metabolism of dopamine after the degeneration of dopaminergic neurons. This topic is particularly demanding to explore, since serotonergic, glutamatergic, noradrenergic, cholinergic, opioid, endocannabinoid and adenosinergic systems have a variable, controversial and only partially explained relationship with LIDs. Recent trends agree on the possible role played by such pathways in LID occurrence and maintenance [47], but a complete understanding of the topic is far from being reached, and further investigations are still needed. The following paragraphs summarize a brief overview on how and when these neurotransmitter pathways interplay in LID pathogenesis. Table 3 depicts an overview on neurotransmitter systems.

2.8.1. Serotonergic System

The serotonergic system originates from the raphe nuclei and, through its cortico-subcortical projections, modulates cognition, vegetative functions and movement [125,126]. As described by Lavoie and Parent [127] and Fox and Chuang [128], the influence on motor control should be attributed to the dense serotonergic innervation of the striatum, SNr and GP. Indeed, serotonergic neurons show biochemical similarities with the dopaminergic ones: they share the vesicular monoamine transporter 2 (VMAT2) and AADC enzyme, which catalyzes the decarboxylation of aromatic amino acids. Through this enzymatic machinery, these neurons can convert also exogenous levodopa into dopamine and subsequently release it in an activity-dependent fashion [129,130,131]. Such possibility to generate dopamine with non-dopaminergic terminals is thought to be the cornerstone of serotonergic influence on LIDs. Additionally, movement regulation and LID induction seem exploited by the serotonergic system through the 5-HT1A, 5-HT1B, 5-HT2A, 5-HT2C and the 5-HT3 receptors. Further evidence proved that removing the forebrain serotonergic innervation by the selective toxin 5,7-dihydroxy-tryptamine (5,7-DHT) almost completely suppressed abnormal involuntary movements [132,133,134]. Finally, it has been extensively demonstrated that increasing the serotonergic tone, by the administration of either selective serotonin reuptake blockers (SSRIs) or the serotonergic precursor 5-hydroxytryptophan, significantly reduced LIDs in hemiparkinsonian rats, without compromising the levodopa therapeutic efficacy [135,136,137]. Keeping in mind that serotonergic and dopaminergic neurons are biochemically related, the contribution of the serotonergic side is beneficial when sufficient dopaminergic terminals are spared, since dopaminergic terminals provide a buffering system for the levodopa-derived dopamine. Conversely, as the disease progresses, the contribution of serotonergic neurons becomes detrimental because serotonergic neurons lack an autoregulatory feedback mechanism for dopamine release. As a consequence, levodopa-derived dopamine is released in an uncontrolled way following levodopa administration.

2.8.2. Glutamatergic System

Glutamate is the major excitatory neurotransmitter in the nervous system, and its transmission depends on three receptor subtypes globally expressed in cortico-subcortical structures: metabotropic receptors coupled to second messenger systems through G-proteins (mGluR), Ionotropic Glutamate N -Methyl-D-Aspartate (NMDA) and Ionotropic Alpha-Amino-3-hydroxy-5-methyl-4-Isoxazolepropionic Acid (AMPA) receptors. As PD progresses, dopaminergic sprouting and reduced DA uptake preserve intrastriatal DA levels [138] with a consequent detrimental glutamatergic control [139,140,141,142]. Dopaminergic decrease and uncontrolled replacement with levodopa alter the glutamatergic transmission within the basal ganglia. Accordingly, corticostriatal glutamatergic activity has been demonstrated to be dramatically increased in PD mouse models [143,144,145]. However, interactions between dopaminergic and glutamatergic systems are complex, since glutamate receptors have a pivotal role in synaptic plasticity, and both postsynaptic changes and the trafficking of the glutamate metabotropic and ionotropic receptors in the synaptic cleft should influence the pathogenesis of LIDs [146]. The three glutamate receptor subtypes are modulated differently throughout the course of PD, and their possible role in LIDs is difficult to establish. For example, group I mGluRs antagonists reduced DA-dependent striatal synaptic plasticity, in terms of both long-term depression (LTD) and long-term potentiation (LTP), raising the possibility of detrimental effects on striatal-dependent motor and cognitive activity [147,148,149]. PD progression additionally influences corticostriatal NMDA-mediated glutamatergic signals, because corticostriatal plasticity depends on either the nigral denervation or differential composition of striatal NMDA receptor subunits [150,151,152]. Differently from mGluR and NMDA receptors, less is clear about the dysregulation of AMPA receptors in PD, and its possible causative role in LIDs is still debatable [142,153,154,155,156]. To conclude, the glutamatergic system is complex and has a plethora of functions that are compromised in a variable fashion in advanced disease stages and appear to be far from completely understood. The literature is controversial on the topic, but it is well established that the chronic levodopa treatment compromises the dopaminergic control on glutamatergic transmission in the long term.

2.8.3. Noradrenergic System

The noradrenergic innervation of the central nervous system (CNS) depends on two main ascending systems. A major source of noradrenaline (NA) is the locus coeruleus (LC), located along the fourth ventricle in the pons, which provides projections to the SNc [157] and striatum [158]. A secondary source of NA is represented by the medullary noradrenergic system, composed of scattered groupings of noradrenergic neurons in the ventrolateral reticular formation and the nucleus of the solitary tract, with primary vegetative functions [159,160,161]. Despite the results of Alachkar and Brotchie [162] and Ribas and Miralles [163], how the noradrenergic system varies throughout the course of PD is a question still open. The NA-synthesizing enzyme DA-β-hydroxylase (DBH) is crucial for NA synthesis and has been exploited for laboratory models: DBH knock-out mice do not produce NA and exhibit both parkinsonism and spontaneous dyskinesias, even if striatal DA is preserved [164]. However, NA loss in PD mouse models has been poorly evaluated in the literature, since in the conventional PD 6-OHDA-induced rat model, the NA transporter (NAT) is blocked prior to 6-OHDA infusion to prevent noradrenergic cell loss. For this reason, only a few studies investigated the effect of additional noradrenergic lesions on LIDs, providing controversial effects on LID severity and duration [165]. For example, direct LC infusions of ibotenic acid reduced LIDs in 6-OHDA-lesioned rats that had previously been rendered dyskinetic [166]. Interestingly, Buck and Ferger [167] demonstrated that exogenous NA, infused in the striatum of hemiparkinsonian rats, elicited LIDs, while Arai and Tomiyama [168] investigated the activity of NAT for DA re-uptake and hypothesized a possible inhibitory role in LIDs. From a receptorial point of view, it is noteworthy that agonists and antagonists of α and β receptors gained attention in recent years due to the possibility to modulate dopaminergic transmission, with variable results. For example, Sommermeyer and Frielingsdorf [169] reduced LIDs with an α-1 receptor antagonist in rodents. Additionally, α-2 receptor antagonists prevented severe/disabling LIDs [170,171,172,173]. Finally, Carpentier and Bonnet [174] demonstrated a reduction in LIDs in humans with propanolol, a β-1/2 receptor antagonist, while β-2 receptor selective antagonists have never been tried for LID management.

2.8.4. Cholinergic System

Nicotine interacts with the nAChRs, ligand-gated ion channels, whose endogenous neurotransmitter is acetylcholine. The most represented nAChR subtypes in regions such as the cortex, hippocampus, thalamus and cerebellum are the α4β2* and α7* nAChRs, while the primary ones in the basal ganglia are the α4β2* and α6β2* subtypes with α7* nAChRs less densely expressed [175,176,177]. An overwhelming body of evidence concluded that nicotine has an established anti-dyskinetic effect, probably mediated by nAChR desensitization/downregulation, with a secondary reduction in striatal dopamine release. In fact, nicotine administration has been proven to alleviate both peak and total LIDs by approximately 60% in MPTP-lesioned nonhuman primates (NHPs) for 30 weeks [178,179,180], and its readministration, after a 10-week washout period, led to an immediate decline in LIDs. Additionally, mice lacking both α4β2* and α6β2* nAChRs had reduced baseline LIDs, suggesting a role as an LID primary regulator for the β2 subtype [181,182,183]. However, it is likely that multiple nAChR populations influence LIDs, including the α4β2* α6β2* and α7 subtypes. Since multiple compensatory changes occur throughout the course of PD, various nAChR subtypes may be differentially related to LIDs during disease progression. The cholinergic system also includes muscarinic receptors mAChRs, numbered from M1 to M5 and coupled to G-proteins [184,185]. These receptors are highly expressed in the striatum [184,185] and do not have an established role in abnormal involuntary movements: atropine had no effect on LIDs [186], while the muscarinic antagonist dicyclomine reduced LIDs in 6-OHDA-lesioned mice.

2.8.5. Opioid System

Endogenous opioid peptides, except endomorphins, share the amino acid enkephalin sequence at the N-terminus, with differing extensions at the C-terminus [187]. Opioid receptors are widely distributed across cortical and subcortical structures of the CNS, such as the basal ganglia, nucleus accumbens and ventral tegmental area. Opioid receptor distribution and activity undergo variable modifications throughout the course of PD [187]. μ-receptor levels are reduced in the striatum and GPi, κ-receptor levels are reduced in the GPe and GPi, while δ-receptor levels are unchanged in the striatum of dyskinetic NHPs. Moreover, signaling is overactive for μ, δ and κ-receptors, respectively, in the striatum, GPi and caudate nucleus and motor cortex of MPTP-lesioned NHPs. Two seminal studies investigated opioid levels in dyskinetic models and found elevated levels of dynorphin B and α-neoendorphin in the dorsolateral striatum [188] and SN [189] of severely dyskinetic rats compared with mildly dyskinetic or non-dyskinetic rats, while dynorphin A has been found elevated in dyskinetic nonhuman primates by Bourdenx and Nilsson [190]. Non-selective opioid receptor antagonists generated ambivalent results for LID control [191,192,193]. Such a lack of clear anti-dyskinetic actions is thought to reflect the interaction of non-subtype-selective ligands with multiple opioid receptors, providing competing pro- and anti-dyskinetic effects. The leading theory in the field considers the blockage of μ and δ receptors as anti-dyskinetic, while the blockade of κ-opioid receptors should promote LIDs [194].

2.8.6. Endocannabinoid System

The endocannabinoid system consists of a family of lipid signaling molecules released on demand from membrane lipid precursors and the relative biochemical enzyme machinery involved in their synthesis and degradation [195,196]. Arachidonoyl ethanolamine (anandamide) [197,198] and 2-arachidonoyl glycerol (2-AG) [199] represent the progenitors of this group of molecules, but the number of new members is rapidly increasing [200]. The main receptors are coupled to G-proteins (CB1 and CB2), while others belong to the transient receptor potential (TRP) family, as well as nuclear peroxisome proliferator-activated receptors (PPAR) [201]. CB1 receptors are represented on GABAergic striatofugal neurons [202,203] and the subthalamic nucleus [204]. The endocannabinoid system has pleiotropic functions in movement: an increased endocannabinoid transmission reduces striatal glutamate release, relieving PD symptoms [143,205], while the activation of CB1 on striatofugal terminals may empower the indirect pathway, amplifying the inhibitory output of the basal ganglia. However, it is a well-known fact that these molecules counterbalance the dopamine-mediated hyperactivity [206,207,208]. LID improvement is attributed to the reduction in the levodopa-induced sensitization of dopamine receptors, normalization of aberrant glutamate release and rebalancing of the maladaptive plasticity in the denervated striatum. In support of this hypothesis, several attempts have shown cannabinoid-mediated improvement in levodopa-induced abnormal involuntary movements in rodent models and NHPs [206,209,210,211,212] and PD patients [213], avoiding global motor suppression [206]. CB1 receptors are also expressed on serotonergic raphe-striatal fibers [214], through which they can act as pseudodopaminergic neurons for abnormal dopamine storage and release, contributing to LID development, as mentioned before (see Section 2.8.1) [132]. To conclude, it is likely that the endocannabinoid system may promote anti-dyskinetic effects, by both dampening the ectopic dopamine release from serotonergic terminals and inhibiting 5-HT release [215,216].

2.8.7. Adenosinergic System

Adenosine, a ubiquitous endogenous nucleoside, is a modulator of neurotransmission exerted by DA, glutamate and acetylcholine and has been implicated in reward-related behavior [217,218,219,220,221,222,223]. Adenosine receptors are widely expressed in the CNS: A1 subtypes are distributed in the cortex, hippocampus and cerebellum, A2A are mainly distributed in the striatum and olfactory bulb, while A2B and A3 subtypes are found at low levels of expression. The activation of adenosine A2A receptors in the striatum regulates dopamine and glutamate release. Adenosine A1 receptors generate functionally interacting complexes in cortical neurons and the basal ganglia [224,225]. A higher number of these receptors has been found in the striatum in dyskinetic patients compared to PD patients without LIDs during autopsy studies. The role of adenosine in LID management has not been primarily investigated so far, and improvement in motor function and OFF time were the primary endpoints in the few studies that tested adenosine A2A receptor antagonists [226,227,228,229,230,231,232,233].
However, specific studies addressing LIDs have not been conducted yet, and a direct anti-dyskinetic effect seems unlikely.

2.9. Imaging Studies

Dopaminergic function can be assessed in vivo using neuroimaging studies [234] with specific ligands for dopamine receptors, the VMAT2, the plasmalemmal DAT [235,236] and postsynaptic DA D1R and D2R receptors. Moreover, the uptake and decarboxylation of levodopa to DA, the storage of DA and the DA turnover could be assessed using the fluorinated analog of levodopa, [18F]fluoro-L-dopa [237], as shown in Figure 5.
Clinical observations of the prevalence of dyskinesias in more advanced disease stages found evidence in a report showing an inverse relationship between [18F]fluoro-L-dopa uptake and dyskinesias [238]. Although presynaptic DA denervation plays a crucial role in the pathogenesis of dyskinesia [239], it is not the only factor involved in the development of dyskinesia, and the pattern of DA receptor stimulation plays a key role too. Patients who developed motor fluctuations showed a greater magnitude but a less sustained decline in [11C]Raclopride binding, with respect to patients with a stable response [240]. Moreover, the relative reduction in [11C]Raclopride binding 1 h after oral levodopa intake advances with disease duration and is more pronounced in patients suffering from LIDs compared to those with a stable response. Moreover, no difference was found between dyskinetic and non-dyskinetic subjects 4 h after levodopa, adding evidence of a more pulsatile pattern of levodopa in subjects with motor complications [240,241]. Similar findings have been observed by other groups [242]. Another marker of advanced disease and a risk factor for dyskinesia onset is the increase in DA turnover. Prolonged scans with [18F]D measure uptake at 90–120 min and reflection uptake, and decarboxylation to fluoroDA and the trapping of fluoroDA in synaptic vesicles also reflect the egress and subsequent metabolism of this trapped radioactivity. The effective distribution volume, which is derived from this reversible tracer model, correlates well with the inverse of the ratio of tracer loss to tracer uptake constants [237], which in turn correlates with classical neurochemical measures of DA turnover [243]. DA turnover measured using this approach is increased early in PD [244], and further increases occur with disease progression [245]. Sossi and de la Fuente-Fernandez [246] investigated the possible age dependency of changes in DA turnover, in terms of DA distribution volume, as a contributing factor to the levodopa-related complications, with the analysis of the plasma input uptake rate constant (Ki) after [18F]D uptake, an indicator of DA synthesis and vesicular storage capacity. The authors concluded that the magnitude of the increase in DA turnover was greater than the magnitude of the decrease in DA synthesis and storage rate in PD patients with younger disease onset, in line with an increased susceptibility of these patients to develop motor complications [247,248,249,250]. DAT downregulation plays a role in the DA turnover; however, serotonergic neurons are able to convert exogenous levodopa into DA too, with a non-regulated release and uptake of DA. The DA system is not the only actor implied in dyskinesia development, since other neurotransmitters play a role too. Piccini and Weeks [251] demonstrated reduced striatal binding of the opioid ligand [11C]Diprenorphine in PD patients with LIDs, presumably reflecting the occupancy of striatal opioid receptors due to increased opioid levels. Studies have demonstrated increased adenosine A2 binding in PD patients with abnormal involuntary movements compared to those dyskinesia-free [252]. Even brain hemodynamics seem to be influenced by levodopa. Hirano and Asanuma [253] used a multimodal PET to image both blood flow (rCBF) and glucose metabolism in the same scanning session: a dissociation between flow and metabolism was found in patients affected by LIDs after the administrations of oral levodopa. Interestingly, during the OFF-medication state, regional glucose metabolism and rCBF matched; conversely, after the administration of oral levodopa, rCBF was greatly enhanced while glucose metabolism was unchanged in the network composed of putamen, pallidum and midbrain pons. Flow/metabolism dissociation was greater in LID patients, supporting the hypothesis of a hemodynamic effect of levodopa. The underlying mechanisms of such evidence have not been solved yet. An increased capillary density in the striatum and midbrain that may drive more blood flow to these regions under the conditions of a high DAergic tone [254,255] is the leading hypothesis in the field so far. Finally, it is worth mentioning the serotonergic contribution to LIDs, since it gained increasing attention in the last years. As mentioned in the “Serotonergic system” paragraph, putaminal serotonergic fibers have the potential to store and release dopamine in a non-physiological manner, promoting LIDs, and this contribution has been further investigated with radioligand tracers. For this purpose, Lee and Seo [256] enrolled 30 patients with PD, classified as dyskinetic, non-dyskinetic and drug-naive and acquired two PET scans and 3T MRI scans for each patient using [(11)C]-3-amino-4-(2-dimethylaminomethylphenylsulfanyl)-benzonitrile (11C-DASB), a ligand of the serotonin transporter, and N-(3-[(18)F]fluoropropyl)-2-carbomethoxy-3-(4-iodophenyl) nortropane (18F-FP-CIT), a DAT radioligand. The 11C-DASB/18F-FP-CIT ratio was computed to estimate serotoninergic fiber innervation relative to dopaminergic fiber availability. The study showed the highest 11C-DASB/18F-FP-CIT ratio in the putamen and pallidum for dyskinetic PD patients, highlighting the pivotal role of the serotonergic innervation in LIDs.

3. Therapeutic Options

Different therapeutic approaches have been proposed in order to face the LID problem [257,258,259]. On the pharmacological side, one possible solution is to avoid the pulsatile administration of levodopa, through a continuous administration of levodopa [260,261,262,263,264] or dopamine agonists like apomorphine [265]. Other pharmacological options include levodopa regimen optimizations in terms of doses and inter-dose timing and the usage of long-release drugs or add-on medications. The aim is to flatten the levodopa fluctuations. However, neurons are responsive to pharmacological stimulation but also to electrical [266], magnetic [267] or ultrasound stimulations [268,269]. Among stimulation techniques, invasive DBS has been used for the treatment of both hyperkinetic [270,271,272,273] and hypokinetic movement disorders, like Parkinson’s disease [274,275,276,277].

3.1. Levodopa Therapy Optimization

The aim of the therapies for LIDs is to reduce the fluctuations of levodopa levels and improve the pharmacokinetics of levodopa by the prevention of dopamine catabolism and the usage of controlled-release levodopa or dopamine agonists [109]. It has been hypothesized for a long time that providing a more continuous and constant levodopa administration could reduce the risk of motor complications, even in early PD patients. This idea was tested in the STRIDE-PD trial, conducted by Stocchi and Rascol [28]: the risk of developing dyskinesias was compared during a 134-week double-blind trial in 747 PD patients randomized to levodopa/carbidopa (LC) or levodopa/carbidopa/entacapone (LCE). It is worth noting that early PD patients on LCE failed to delay the onset of motor complications: the LCE group was associated with a shorter time to onset and an increased frequency of LIDs. For the selection of the correct therapeutic scheme, the pharmacokinetic/pharmacodynamic of each dose of levodopa should be kept in mind to guide the selection (Figure 6).
It is worth noting that the therapeutic approach depends on the type of dyskinesia itself.
For peak-dose and square-wave dyskinesias, the following therapeutic approaches can be considered [47,278,279,280]:
-
Reducing the dose of levodopa and distributing the inter-dose timing.
-
Adding an add-on medication, such as Amantadine, can help to reduce the severity of dyskinesias.
-
For diphasic and square-wave dyskinesias, the following therapeutic approaches can be considered [47,278,279,280]:
-
Adjusting the timing of medication doses: spreading out the doses throughout the day can help to maintain more stable medication levels.
-
Adding an add-on medication, such as Amantadine, can help reduce the severity of dyskinesias.
For OFF dystonia, the following therapeutic approaches can be considered [47,278,279,280]:
-
Adjusting the timing and dosage of medication or increasing the dose of levodopa can help maintain more stable medication levels.
-
Adding an add-on medication, such as an extended-release dopamine agonist, MAO-B inhibitor or COMT inhibitor, can help to reduce the severity of OFF dystonia by potentiating the dopaminergic stimulation.
-
Apomorphine injections or sublingual administration [281] can provide rapid relief from OFF dystonia.
All types of LIDs can benefit from continuous levodopa or Apomorphine administration or deep brain stimulation.
A successful approach to improve pharmacokinetics includes alternative routes of drug delivery to bypass the delayed gastric emptying [282,283]. More constant plasma levodopa concentrations are achieved through a gelified version for the intrajejunal administration of levodopa/carbidopa, in patients with advanced PD [284,285]. The efficacy of this formulation has been confirmed by a randomized controlled trial [260]. It is worth noting that the GLORIA registry, a 24-month, non-interventional and observational registry, conducted by Antonini and Poewe [286], was the study with the largest cohort of PD patients treated with levodopa/carbidopa intestinal gel (LCIG). In particular, in this registry, the authors investigated the impact of such device-aided therapy on advanced PD. Results demonstrated improvements in motor fluctuations and non-motor symptoms, such as sleep, mood and the quality of life. The more frequent adverse effects were weight loss (6.7%), device-related infections (5.9%), device dislocations (4.8%) and polyneuropathy (4.5%).
The evolution of LIDs on LCIG therapy deserves a separate discussion. Unexpectedly, a therapy with LCIG could not completely abort LIDs, but dyskinesias may vary and change their profile throughout the course of the disease, despite an appropriate therapeutic regimen. While a reduction in motor fluctuations is predictable, less is known about how the LID profile is modulated by such therapy. For this purpose, Szász and Constantin [287] investigated dyskinesia features in advanced PD patients before and after 6, 12 and 18 months from LCIG. As expected, motor fluctuations improved, but abnormal involuntary movements changed their pattern. In fact, 18 months after LCIG positioning, severe peak-dose dyskinesias dropped from a mean of 1.61 o 0.04 h, while the mild/moderate ones increased from a mean of 1.97 to 2.79 h. Diphasic dyskinesias were reduced from an average of 4.03 to 1.81 h, and dystonia dropped from a total of 12 to 9 h per day. This work highlights the need for advanced PD patients on LCIG to receive a customized and targeted therapeutic approach [287].
Other approaches to improve pharmacokinetic parameters include the use of extended-release formulations of levodopa [288,289,290], the administration of levodopa gastric retention formulation [282] and the combination of immediate- and extended-release formulations with gastric retention [283].

3.2. Non-Dopaminergic Drugs

Non-levodopa-based therapeutic strategies for dyskinesias rely on the possible role of non-dopaminergic systems, such as the serotonergic ones, for the induction of LIDs [291,292]. The 5-HT1A agonist Buspirone was effective in reducing LIDs after oral administration and reduced levodopa-evoked striatal synaptic dopamine release [291,292]. Another Serotonin 5-HT1A agonist used to manage LIDs is Sarizotan, whose potential feasibility for the scope was not confirmed in a large randomized, placebo-controlled, phase IIb trial [293]. 5-HT1A/1B receptor agonists Eltoprazine and Anpirtoline [294,295] revealed efficacy in reducing LIDs in animal models, while the results were not superior to Amantadine in a translational human study [296]. NMDA antagonists, like Memantine, Remacemide, Dextromethorphan, Milacemide and CP-101.606, were assessed as therapeutic options, without clinical evidence of any anti-dyskinetic effect in experimental models of LIDs [297,298,299,300]. Antiepileptic drugs were tested too with poor results. For example, Gabapentin showed no difference with respect to the placebo in reducing dyskinesias [301]. Moreover, a randomized double-blind, placebo-controlled, parallel-group trial with Levetiracetam failed to show a significant reduction in dyskinesias [302]. Zonisamide at the dosage of 50 mg decreased dyskinesias, but dizziness, apathy and weight loss were side effects reported by the patients [303]. Among antipsychotics, Clozapine, Olanzapine and Quetiapine were the most studied. A large, randomized, placebo-controlled trial with Clozapine showed a reduction in dyskinesias with increased ON time without dyskinesias and no effect on increasing OFF time or increased adverse events [304]. Quetiapine failed to show benefits in dyskinesias compared to placebo [305]. Finally, Olanzapine revealed some anti-dyskinetic effects during a randomized, placebo-controlled, cross-over trial but also increased OFF time [306].

3.3. Deep Brain Stimulation

Another possible therapeutic option is DBS, targeting either the GPi or STN. When the target is the GPi, DBS has a direct effect on LID reduction; conversely, when the STN is chosen, involuntary movements are reduced through an indirect effect on dyskinesias by lowering levodopa thanks to the antiparkinsonian effect of STN stimulation. DBS can lead to new possibilities for LID management, due to the high informative value of basal ganglia oscillatory activity about the motor state and its potential reliability to be real-time adjusted for symptom control [86].

3.4. Closed-Loop Therapy

Predictors of motor performances can be used as feedback to real-time fit and optimize therapeutic regimens in a closed-loop fashion to restore a physiologic dopaminergic stimulation pattern and limit dyskinesias in PD patients [86]. An adaptive closed-loop administration algorithm improves dyskinesias by reducing the fluctuations in dopamine levels and reproducing the normal dopaminergic tone. Biochemical, neurophysiological and wearable sensors are sensing systems expected to provide feedback signals to close this loop. A multiparametric modular sensing system that combines biochemical, neurophysiological and wearable sensor data could adapt the administration of different combinations of antiparkinsonian therapies in real time, as shown in Figure 7 [307,308,309]. Biochemical and neurocomputational models of the levodopa pharmacokinetics and dynamics were shown to predict the motor response with a varying levodopa plasma concentration in both stable and fluctuating PD patients [93,310,311,312,313], and a sigmoid curve was demonstrated to express the relationship between levodopa plasmatic concentration and tapping frequency. Biochemical levodopa sensing approaches include detection in blood [314], sweat [313], skin [315], skeletal muscle [314], subcutaneous tissue samples [316] and electrochemical sensing with amperometry or voltammetry [317]. On the neurophysiological side, synchronized oscillatory rhythms reliably associated with hypokinetic and dyskinetic states provide potentialities as a control signal in closed-loop DBS [12,318,319]. Sensing cortical or subcortical brain activity and real-time symptom correlate extraction during DBS is the most advanced closed-loop strategy: basal ganglia beta activity (13–30 Hz) is related to hypokinetic state and bradykinesia and is, to date, the most promising neurophysiological biomarker for closed-loop DBS in PD [320,321,322], while the basal ganglia activity in the gamma range (25–140 Hz) is the electrophysiological signature of the dyskinetic state [12]. Wearable sensors integrated with machine learning algorithms can convert kinematic motor data into quantitative signals, recognizing and predicting changes in the motor state [323].

4. Discussion

Since its introduction as an antiparkinsonian drug regimen, levodopa has become the leading therapeutic option for motor control in PD [3]. Abnormal involuntary movements are well recognized as a possible complication secondary to levodopa-based therapy, and it is not only movement disorder specialists that have to frequently face this therapy-induced drawback [38]. This review points out the complexity of their neurophysiological correlates and pharmacokinetic and pharmacodynamic models. The literature discussed in depth the temporal relationship between different antiparkinsonian drugs and dyskinesias [7,21,22,23,24,28,39,40,41,42,43,44,45], highlighting the major role played by levodopa over dopamine agonists [22,23,24,25,26,27], COMT inhibitors [28] and MAO inhibitors [29]. Moreover, LIDs are not stereotyped in nature, since they can be different in phenomenology [46,47,48,49]. Such a complexity is reflected in the basal ganglia firing pattern, modified not only in quantity but also in quality [112,113,114]. An exhaustive pathophysiologic explanation of LIDs is an unmet need so far. Imaging studies reinforced the role of dopaminergic transport, progressive presynaptic terminal denervation and the pattern of DA receptor turnover in the advanced disease stages as possible contributing factors but not as the primum movens [239,245,246,247,248,249,250]. Classical therapeutic choices to manage LIDs are limited in efficacy due to the poor feasibility to reach a delicate compromise between excessive and poverty of movement: conventional strategies essentially aim to reduce levodopa fluctuations and stabilize its blood levels to restore a more efficacious and constant dopaminergic tone [47,109,278,279,282,283,284,285]. New therapeutic scenarios will arise from wearable, biochemical and neurophysiological sensing systems, expected to provide feedback signals in a closed-loop system to predict the motor response with a varying levodopa plasma concentration [307,308,309]. The aim of these techniques is to optimize the administration of the currently existing therapeutic options in relation to the real-time motor status.

Author Contributions

Conceptualization, L.d.B.; writing—original draft preparation, L.d.B., P.M.P., M.L.C. and S.P.C.; writing—review and editing, L.d.B. and V.D.L.; supervision, V.D.L. 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.

Data Availability Statement

Not applicable.

Conflicts of Interest

L. di Biase is the scientific director and one of the shareholders of Brain Innovations Srl, a University spinoff of the Campus Bio-Medico University of Rome. The authors declare no conflict of interest.

References

  1. Ehringer, H.; Hornykiewicz, O. Verteilung von Noradrenalin und Dopamin (3-Hydroxytyramin) im Gehirn des Menschen und ihr Verhalten bei Erkrankungen des extrapyramidalen Systems. Klin. Wochenschr. 1960, 38, 1236–1239. [Google Scholar] [CrossRef]
  2. Cotzias, G.C.; Van Woert, M.H.; Schiffer, L.M. Aromatic amino acids and modification of parkinsonism. N. Engl. J. Med. 1967, 276, 374–379. [Google Scholar] [CrossRef] [PubMed]
  3. Connolly, B.S.; Lang, A.E. Pharmacological treatment of Parkinson disease: A review. JAMA 2014, 311, 1670–1683. [Google Scholar] [CrossRef] [PubMed]
  4. Hechtner, M.C.; Vogt, T.; Zollner, Y.; Schroder, S.; Sauer, J.B.; Binder, H.; Singer, S.; Mikolajczyk, R. Quality of life in Parkinson’s disease patients with motor fluctuations and dyskinesias in five European countries. Park. Relat. Disord. 2014, 20, 969–974. [Google Scholar] [CrossRef] [PubMed]
  5. Cotzias, G.C.; Papavasiliou, P.S.; Gellene, R. Modification of Parkinsonism—Chronic treatment with L-dopa. N. Engl. J. Med. 1969, 280, 337–345. [Google Scholar] [CrossRef]
  6. Warren Olanow, C.; Kieburtz, K.; Rascol, O.; Poewe, W.; Schapira, A.H.; Emre, M.; Nissinen, H.; Leinonen, M.; Stocchi, F. Factors predictive of the development of Levodopa-induced dyskinesia and wearing-off in Parkinson’s disease. Mov. Disord. 2013, 28, 1064–1071. [Google Scholar] [CrossRef]
  7. Schrag, A.; Quinn, N. Dyskinesias and motor fluctuations in Parkinson’s disease. A community-based study. Brain A J. Neurol. 2000, 123 Pt 11, 2297–2305. [Google Scholar] [CrossRef] [Green Version]
  8. Colosimo, C.; De Michele, M. Motor fluctuations in Parkinson’s disease: Pathophysiology and treatment. Eur. J. Neurol. 1999, 6, 1–21. [Google Scholar] [CrossRef]
  9. Fabbrini, G.; Mouradian, M.M.; Juncos, J.L.; Schlegel, J.; Mohr, E.; Chase, T.N. Motor fluctuations in Parkinson’s disease: Central pathophysiological mechanisms, Part I. Ann. Neurol. 1988, 24, 366–371. [Google Scholar] [CrossRef]
  10. Mouradian, M.M.; Juncos, J.L.; Fabbrini, G.; Schlegel, J.; Bartko, J.J.; Chase, T.N. Motor fluctuations in Parkinson’s disease: Central pathophysiological mechanisms, Part II. Ann. Neurol. 1988, 24, 372–378. [Google Scholar] [CrossRef]
  11. Assenza, G.; Capone, F.; di Biase, L.; Ferreri, F.; Florio, L.; Guerra, A.; Marano, M.; Paolucci, M.; Ranieri, F.; Salomone, G. Oscillatory activities in neurological disorders of elderly: Biomarkers to target for neuromodulation. Front. Aging Neurosci. 2017, 9, 189, Corrigendum in Front. Aging Neurosci. 2017, 9, 252. [Google Scholar] [CrossRef] [PubMed]
  12. Swann, N.C.; de Hemptinne, C.; Miocinovic, S.; Qasim, S.; Wang, S.S.; Ziman, N.; Ostrem, J.L.; San Luciano, M.; Galifianakis, N.B.; Starr, P.A. Gamma oscillations in the hyperkinetic state detected with chronic human brain recordings in Parkinson’s disease. J. Neurosci. 2016, 36, 6445–6458. [Google Scholar] [CrossRef] [PubMed]
  13. Halje, P.; Tamte, 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] [Green Version]
  14. Güttler, C.; Altschüler, J.; Tanev, K.; Böckmann, S.; Haumesser, J.K.; Nikulin, V.V.; Kühn, A.A.; van Riesen, C. Levodopa-induced dyskinesia are mediated by cortical gamma oscillations in experimental Parkinsonism. Mov. Disord. 2021, 36, 927–937. [Google Scholar] [CrossRef] [PubMed]
  15. Cassidy, M.; Mazzone, P.; Oliviero, A.; Insola, A.; Tonali, P.; Lazzaro, V.D.; Brown, P. Movement-related changes in synchronization in the human basal ganglia. Brain 2002, 125, 1235–1246. [Google Scholar] [CrossRef]
  16. Alonso-Frech, F.; Zamarbide, I.; Alegre, M.; Rodriguez-Oroz, M.C.; Guridi, J.; Manrique, M.; Valencia, M.; Artieda, J.; Obeso, J.A. Slow oscillatory activity and levodopa-induced dyskinesias in Parkinson’s disease. Brain 2006, 129, 1748–1757. [Google Scholar] [CrossRef] [Green Version]
  17. Alegre, M.; López-Azcárate, J.; Alonso-Frech, F.; Rodríguez-Oroz, M.C.; Valencia, M.; Guridi, J.; Artieda, J.; Obeso, J.A. Subthalamic activity during diphasic dyskinesias in Parkinson’s disease. Mov. Disord. 2012, 27, 1178–1181. [Google Scholar] [CrossRef]
  18. Cagnan, H.; Kuhn, A.A.; Brown, P. Co-modulation of finely tuned high-gamma band activity across hemispheres in Parkinson’s disease. Clin. Neurophysiol. 2014, 125, 777–785. [Google Scholar] [CrossRef] [Green Version]
  19. Wiest, C.; Torrecillos, F.; Tinkhauser, G.; Pogosyan, A.; Morgante, F.; Pereira, E.; Tan, H. Finely-tuned gamma oscillations: Spectral characteristics and links to dyskinesia. Exp. Neurol. 2022, 351, 113999. [Google Scholar] [CrossRef]
  20. Antonini, A.; Stoessl, A.J.; Kleinman, L.S.; Skalicky, A.M.; Marshall, T.S.; Sail, K.R.; Onuk, K.; Odin, P.L.A. Developing consensus among movement disorder specialists on clinical indicators for identification and management of advanced Parkinson’s disease: A multi-country Delphi-panel approach. Curr. Med. Res. Opin. 2018, 34, 2063–2073. [Google Scholar] [CrossRef]
  21. Group, P.S. Levodopa and the progression of Parkinson’s disease. New. Engl. J. Med. 2004, 351, 2498–2508. [Google Scholar]
  22. Oertel, W.H.; Wolters, E.; Sampaio, C.; Gimenez-Roldan, S.; Bergamasco, B.; Dujardin, M.; Grosset, D.G.; Arnold, G.; Leenders, K.L.; Hundemer, H.P.; et al. Pergolide versus levodopa monotherapy in early Parkinson’s disease patients: The PELMOPET study. Mov. Disord. 2006, 21, 343–353. [Google Scholar] [CrossRef]
  23. Rascol, O.; Brooks, D.J.; Korczyn, A.D.; De Deyn, P.P.; Clarke, C.E.; Lang, A.E. A five-year study of the incidence of dyskinesia in patients with early Parkinson’s disease who were treated with ropinirole or levodopa. N. Engl. J. Med. 2000, 342, 1484–1491. [Google Scholar] [CrossRef] [PubMed]
  24. Hauser, R.A.; Rascol, O.; Korczyn, A.D.; Jon Stoessl, A.; Watts, R.L.; Poewe, W.; De Deyn, P.P.; Lang, A.E. Ten-year follow-up of Parkinson’s disease patients randomized to initial therapy with ropinirole or levodopa. Mov. Disord. 2007, 22, 2409–2417. [Google Scholar] [CrossRef] [PubMed]
  25. Holloway, R.G.; Shoulson, I.; Fahn, S.; Kieburtz, K.; Lang, A.; Marek, K.; McDermott, M.; Seibyl, J.; Weiner, W.; Musch, B.; et al. Pramipexole vs levodopa as initial treatment for Parkinson disease: A 4-year randomized controlled trial. Arch. Neurol. 2004, 61, 1044–1053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Levin, J.; Krafzcyk, S.; Valkovic, P.; Eggert, T.; Claassen, J.; Botzel, K. Objective measurement of muscle rigidity in Parkinsonian patients treated with subthalamic stimulation. Mov. Disord. 2009, 24, 57–63. [Google Scholar] [CrossRef]
  27. Lees, A.J.; Katzenschlager, R.; Head, J.; Ben-Shlomo, Y. Ten-year follow-up of three different initial treatments in de-novo PD: A randomized trial. Neurology 2001, 57, 1687–1694. [Google Scholar] [CrossRef]
  28. Stocchi, F.; Rascol, O.; Kieburtz, K.; Poewe, W.; Jankovic, J.; Tolosa, E.; Barone, P.; Lang, A.E.; Olanow, C.W. Initiating levodopa/carbidopa therapy with and without entacapone in early Parkinson disease: The STRIDE-PD study. Ann. Neurol. 2010, 68, 18–27. [Google Scholar] [CrossRef]
  29. Shoulson, I.; Oakes, D.; Fahn, S.; Lang, A.; Langston, J.W.; LeWitt, P.; Olanow, C.W.; Penney, J.B.; Tanner, C.; Kieburtz, K.; et al. Impact of sustained deprenyl (selegiline) in levodopa-treated Parkinson’s disease: A randomized placebo-controlled extension of the deprenyl and tocopherol antioxidative therapy of parkinsonism trial. Ann. Neurol. 2002, 51, 604–612. [Google Scholar] [CrossRef]
  30. de Lau, L.M.; Verbaan, D.; Marinus, J.; Heutink, P.; van Hilten, J.J. Catechol-O-methyltransferase Val158Met and the risk of dyskinesias in Parkinson’s disease. Mov. Disord. 2012, 27, 132–135. [Google Scholar] [CrossRef]
  31. Kaiser, R.; Hofer, A.; Grapengiesser, A.; Gasser, T.; Kupsch, A.; Roots, I.; Brockmöller, J. L -dopa-induced adverse effects in PD and dopamine transporter gene polymorphism. Neurology 2003, 60, 1750–1755. [Google Scholar] [CrossRef] [PubMed]
  32. Rieck, M.; Schumacher-Schuh, A.F.; Altmann, V.; Francisconi, C.L.; Fagundes, P.T.; Monte, T.L.; Callegari-Jacques, S.M.; Rieder, C.R.; Hutz, M.H. DRD2 haplotype is associated with dyskinesia induced by levodopa therapy in Parkinson’s disease patients. Pharmacogenomics 2012, 13, 1701–1710. [Google Scholar] [CrossRef] [PubMed]
  33. Ivanova, S.A.; Loonen, A.J.; Pechlivanoglou, P.; Freidin, M.B.; Al Hadithy, A.F.; 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] [Green Version]
  34. Foltynie, T.; Cheeran, B.; Williams-Gray, C.H.; Edwards, M.J.; Schneider, S.A.; Weinberger, D.; Rothwell, J.C.; Barker, R.A.; Bhatia, K.P. BDNF val66met influences time to onset of levodopa induced dyskinesia in Parkinson’s disease. J. Neurol. Neurosurg. Psychiatry 2009, 80, 141–144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Zappia, M.; Annesi, G.; Nicoletti, G.; Arabia, G.; Annesi, F.; Messina, D.; Pugliese, P.; Spadafora, P.; Tarantino, P.; Carrideo, S.; et al. Sex differences in clinical and genetic determinants of levodopa peak-dose dyskinesias in Parkinson disease: An exploratory study. Arch. Neurol. 2005, 62, 601–605. [Google Scholar] [CrossRef] [Green Version]
  36. Strong, J.A.; Dalvi, A.; Revilla, F.J.; Sahay, A.; Samaha, F.J.; Welge, J.A.; Gong, J.; Gartner, M.; Yue, X.; Yu, L. Genotype and smoking history affect risk of levodopa-induced dyskinesias in Parkinson’s disease. Mov. Disord. 2006, 21, 654–659. [Google Scholar] [CrossRef]
  37. Falla, M.; Di Fonzo, A.; Hicks, A.A.; Pramstaller, P.P.; Fabbrini, G. Genetic variants in levodopa-induced dyskinesia (LID): A systematic review and meta-analysis. Park. Relat. Disord. 2021, 84, 52–60. [Google Scholar] [CrossRef]
  38. Duvoisin, R.C. Variations in the ”on-off“ phenomenon. Adv. Neurol. 1974, 5, 339–340. [Google Scholar]
  39. Rajput, A.H.; Fenton, M.E.; Birdi, S.; Macaulay, R.; George, D.; Rozdilsky, B.; Ang, L.C.; Senthilselvan, A.; Hornykiewicz, O. Clinical–pathological study of levodopa complications. Mov. Disord. 2002, 17, 289–296. [Google Scholar] [CrossRef]
  40. Van Gerpen, J.A.; Kumar, N.; Bower, J.H.; Weigand, S.; Ahlskog, J.E. Levodopa-associated dyskinesia risk among Parkinson disease patients in Olmsted County, Minnesota, 1976-1990. Arch. Neurol. 2006, 63, 205–209. [Google Scholar] [CrossRef] [Green Version]
  41. Ahlskog, J.E.; Muenter, M.D. Frequency of levodopa-related dyskinesias and motor fluctuations as estimated from the cumulative literature. Mov. Disord. Off J. Mov. Disord. Soc. 2001, 16, 448–458. [Google Scholar] [CrossRef] [PubMed]
  42. Group, P.S. Impact of deprenyl and tocopherol treatment on Parkinson’s disease in DATATOP patients requiring levodopa. Ann. Neurol. 1996, 39, 37–45. [Google Scholar] [CrossRef] [PubMed]
  43. Block, G.; Liss, C.; Reines, S.; Irr, J.; Nibbelink, D. Comparison of immediate-release and controlled release carbidopa/levodopa in Parkinson’s disease. Eur. Neurol. 1997, 37, 23–27. [Google Scholar] [CrossRef] [PubMed]
  44. Group, P.S.; Group, P.S. Pramipexole vs levodopa as initial treatment for Parkinson disease: A randomized controlled trial. JAMA 2000, 284, 1931–1938. [Google Scholar] [CrossRef] [Green Version]
  45. Hely, M.A.; Morris, J.G.; Reid, W.G.; Trafficante, R. Sydney multicenter study of Parkinson’s disease: Non-L-dopa–responsive problems dominate at 15 years. Mov. Disord. Off J. Mov. Disord. Soc. 2005, 20, 190–199. [Google Scholar] [CrossRef]
  46. Zesiewicz, T.A.; Sullivan, K.L.; Hauser, R.A. Levodopa-induced dyskinesia in Parkinson’s disease: Epidemiology, etiology, and treatment. Curr. Neurol. Neurosci. Rep. 2007, 7, 302–310. [Google Scholar] [CrossRef]
  47. Fox, S.H.; Brotchie, J.M. Levodopa-Induced Dyskinesia in Parkinson’s Disease; Springer: London, UK, 2014. [Google Scholar] [CrossRef]
  48. Hametner, E.; Seppi, K.; Poewe, W. The clinical spectrum of levodopa-induced motor complications. J. Neurol. 2010, 257 (Suppl. 2), S268–S275. [Google Scholar] [CrossRef]
  49. Obeso, J.A.; Grandas, F.; Vaamonde, J.; Luquin, M.R.; Artieda, J.; Lera, G.; Rodriguez, M.E.; Martinez-Lage, J.M. Motor complications associated with chronic levodopa therapy in Parkinson’s disease. Neurology 1989, 39 (Suppl. 2), 11–19. [Google Scholar]
  50. Meloni, M.; Solla, P.; Mascia, M.M.; Marrosu, F.; Cannas, A. Diphasic dyskinesias during levodopa-carbidopa intestinal gel (LCIG) infusion in Parkinson’s disease. Park. Relat. Disord. 2017, 37, 92–96. [Google Scholar] [CrossRef]
  51. Voon, V.; Fernagut, P.-O.; Wickens, J.; Baunez, C.; Rodriguez, M.; Pavon, N.; Juncos, J.L.; Obeso, J.A.; Bezard, E. Chronic dopaminergic stimulation in Parkinson’s disease: From dyskinesias to impulse control disorders. Lancet Neurol. 2009, 8, 1140–1149. [Google Scholar] [CrossRef]
  52. Olanow, C.W.; Calabresi, P.; Obeso, J.A. Continuous Dopaminergic Stimulation as a Treatment for Parkinson’s Disease: Current Status and Future Opportunities. Mov. Disord. 2020, 35, 1731–1744. [Google Scholar] [CrossRef] [PubMed]
  53. di Biase, L.; Di Santo, A.; Caminiti, M.L.; Pecoraro, P.M.; Di Lazzaro, V. Classification of Dystonia. Life 2022, 12, 206. [Google Scholar] [CrossRef]
  54. di Biase, L.; Di Santo, A.; Caminiti, M.L.; Pecoraro, P.M.; Carbone, S.P.; Di Lazzaro, V. Dystonia Diagnosis: Clinical Neurophysiology and Genetics. J. Clin. Med. 2022, 11, 4184. [Google Scholar] [CrossRef] [PubMed]
  55. d’Angelis, O.; Di Biase, L.; Vollero, L.; Merone, M. IoT architecture for continuous long term monitoring: Parkinson’s Disease case study. Internet Things 2022, 20, 100614. [Google Scholar] [CrossRef]
  56. di Biase, L.; Pecoraro, P.M.; Pecoraro, G.; Caminiti, M.L.; Di Lazzaro, V. Markerless radio frequency indoor monitoring for telemedicine: Gait analysis, indoor positioning, fall detection, tremor analysis, vital signs and sleep monitoring. Sensors 2022, 22, 8486. [Google Scholar] [CrossRef] [PubMed]
  57. Sanchez-Ferro, A.; Elshehabi, M.; Godinho, C.; Salkovic, D.; Hobert, M.A.; Domingos, J.; van Uem, J.M.; Ferreira, J.J.; Maetzler, W. New methods for the assessment of Parkinson’s disease (2005 to 2015): A systematic review. Mov. Disord. 2016, 31, 1283–1292. [Google Scholar] [CrossRef] [Green Version]
  58. Hssayeni, M.D.; Jimenez-Shahed, J.; Burack, M.A.; Ghoraani, B. Dyskinesia estimation during activities of daily living using wearable motion sensors and deep recurrent networks. Sci. Rep. 2021, 11, 7865. [Google Scholar] [CrossRef]
  59. Lee, S.I.; Daneault, J.-F.; Golabchi, F.N.; Patel, S.; Paganoni, S.; Shih, L.; Bonato, P. A novel method for assessing the severity of levodopa-induced dyskinesia using wearable sensors. In Proceedings of the 2015 37th Annual International Conference of the IEEE Engineering in Medicine and Biology Society (EMBC), Milano, Italy, 25–29 August 2015; pp. 8087–8090. [Google Scholar]
  60. Deuschl, G.; Krack, P.; Lauk, M.; Timmer, J. Clinical neurophysiology of tremor. J. Clin. Neurophysiol. 1996, 13, 110–121. [Google Scholar] [CrossRef]
  61. Di Pino, G.; Formica, D.; Melgari, J.-M.; Taffoni, F.; Salomone, G.; di Biase, L.; Caimo, E.; Vernieri, F.; Guglielmelli, E. Neurophysiological bases of tremors and accelerometric parameters analysis. In Proceedings of the 2012 4th IEEE RAS & EMBS International Conference on Biomedical Robotics and Biomechatronics (BioRob), Rome, Italy, 24–27 June 2012; pp. 1820–1825. [Google Scholar] [CrossRef]
  62. di Biase, L.; Brittain, J.S.; Shah, S.A.; Pedrosa, D.J.; Cagnan, H.; Mathy, A.; Chen, C.C.; Martin-Rodriguez, J.F.; Mir, P.; Timmerman, L.; et al. Tremor stability index: A new tool for differential diagnosis in tremor syndromes. Brain A J. Neurol. 2017, 140, 1977–1986. [Google Scholar] [CrossRef]
  63. Di Biase, L.; Brittain, J.-S.; Peter, B.; Di Lazzaro, V.; Shah, S.A. Methods and System for Characterising Tremors. U.S. Patent 11, 576, 592 B2, 14 February 2023. [Google Scholar]
  64. Stamatakis, J.; Ambroise, J.; Cremers, J.; Sharei, H.; Delvaux, V.; Macq, B.; Garraux, G. Finger tapping clinimetric score prediction in Parkinson’s disease using low-cost accelerometers. Comput. Intell. Neurosci. 2013, 2013, 717853. [Google Scholar] [CrossRef]
  65. Summa, S.; Tosi, J.; Taffoni, F.; Di Biase, L.; Marano, M.; Rizzo, A.C.; Tombini, M.; Di Pino, G.; Formica, D. Assessing bradykinesia in Parkinson’s disease using gyroscope signals. In Proceedings of the Rehabilitation Robotics (ICORR), London, UK, 17–20 July 2017; pp. 1556–1561. [Google Scholar]
  66. di Biase, L.; Summa, S.; Tosi, J.; Taffoni, F.; Marano, M.; Cascio Rizzo, A.; Vecchio, F.; Formica, D.; Di Lazzaro, V.; Di Pino, G.; et al. Quantitative Analysis of Bradykinesia and Rigidity in Parkinson’s Disease. Front. Neurol. 2018, 9, 121. [Google Scholar] [CrossRef] [Green Version]
  67. Endo, T.; Okuno, R.; Yokoe, M.; Akazawa, K.; Sakoda, S. A novel method for systematic analysis of rigidity in Parkinson’s disease. Mov. Disord. 2009, 24, 2218–2224. [Google Scholar] [CrossRef]
  68. Kwon, Y.; Park, S.H.; Kim, J.W.; Ho, Y.; Jeon, H.M.; Bang, M.J.; Koh, S.B.; Kim, J.H.; Eom, G.M. Quantitative evaluation of parkinsonian rigidity during intra-operative deep brain stimulation. Bio-Med. Mater. Eng. 2014, 24, 2273–2281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Raiano, L.; Di Pino, G.; Di Biase, L.; Tombini, M.; Tagliamonte, N.L.; Formica, D. PDMeter: A Wrist Wearable Device for an at-home Assessment of the Parkinson’s Disease Rigidity. IEEE Trans. Neural Syst. Rehabil. Eng. 2020, 28, 1325–1333. [Google Scholar] [CrossRef] [PubMed]
  70. Moore, S.T.; MacDougall, H.G.; Gracies, J.M.; Cohen, H.S.; Ondo, W.G. Long-term monitoring of gait in Parkinson’s disease. Gait Posture 2007, 26, 200–207. [Google Scholar] [CrossRef] [PubMed]
  71. Schlachetzki, J.C.M.; Barth, J.; Marxreiter, F.; Gossler, J.; Kohl, Z.; Reinfelder, S.; Gassner, H.; Aminian, K.; Eskofier, B.M.; Winkler, J.; et al. Wearable sensors objectively measure gait parameters in Parkinson’s disease. PLoS ONE 2017, 12, e0183989. [Google Scholar] [CrossRef] [PubMed]
  72. Tosi, J.; Summa, S.; Taffoni, F.; Biase, L.d.; Marano, M.; Rizzo, A.C.; Tombini, M.; Schena, E.; Formica, D.; Pino, G.D. Feature Extraction in Sit-to-Stand Task Using M-IMU Sensors and Evaluatiton in Parkinson’s Disease. In Proceedings of the 2018 IEEE International Symposium on Medical Measurements and Applications (MeMeA), Rome, Italy, 11–13 June 2018; pp. 1–6. [Google Scholar] [CrossRef]
  73. Suppa, A.; Kita, A.; Leodori, G.; Zampogna, A.; Nicolini, E.; Lorenzi, P.; Rao, R.; Irrera, F. L-DOPA and freezing of gait in Parkinson’s disease: Objective assessment through a wearable wireless system. Front. Neurol. 2017, 8, 406. [Google Scholar] [CrossRef]
  74. di Biase, L.; Di Santo, A.; Caminiti, M.L.; De Liso, A.; Shah, S.A.; Ricci, L.; Di Lazzaro, V. Gait analysis in Parkinson’s disease: An overview of the most accurate markers for diagnosis and symptoms monitoring. Sensors 2020, 20, 3529. [Google Scholar] [CrossRef]
  75. di Biase, L.; Raiano, L.; Caminiti, M.L.; Pecoraro, P.M.; Di Lazzaro, V. Parkinson’s Disease Wearable Gait Analysis: Kinematic and Dynamic Markers for Diagnosis. Sensors 2022, 22, 8773. [Google Scholar] [CrossRef] [PubMed]
  76. Luis-Martínez, R.; Monje, M.H.G.; Antonini, A.; Sánchez-Ferro, Á.; Mestre, T.A. Technology-Enabled Care: Integrating Multidisciplinary Care in Parkinson’s Disease Through Digital Technology. Front. Neurol. 2020, 11, 575975. [Google Scholar] [CrossRef]
  77. Odin, P.; Chaudhuri, K.R.; Volkmann, J.; Antonini, A.; Storch, A.; Dietrichs, E.; Pirtošek, Z.; Henriksen, T.; Horne, M.; Devos, D.; et al. Viewpoint and practical recommendations from a movement disorder specialist panel on objective measurement in the clinical management of Parkinson’s disease. NPJ Park. Dis. 2018, 4, 14. [Google Scholar] [CrossRef] [PubMed]
  78. Pfister, F.M.; Um, T.T.; Pichler, D.C.; Goschenhofer, J.; Abedinpour, K.; Lang, M.; Endo, S.; Ceballos-Baumann, A.O.; Hirche, S.; Bischl, B. High-resolution motor state detection in Parkinson’s disease using convolutional neural networks. Sci. Rep. 2020, 10, 5860. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Li, M.H.; Mestre, T.A.; Fox, S.H.; Taati, B. Automated assessment of levodopa-induced dyskinesia: Evaluating the responsiveness of video-based features. Park. Relat. Disord. 2018, 53, 42–45. [Google Scholar] [CrossRef]
  80. Alam, M.N.; Garg, A.; Munia, T.T.K.; Fazel-Rezai, R.; Tavakolian, K. Vertical ground reaction force marker for Parkinson’s disease. PLoS ONE 2017, 12, e0175951. [Google Scholar] [CrossRef]
  81. Cavallo, F.; Moschetti, A.; Esposito, D.; Maremmani, C.; Rovini, E. Upper limb motor pre-clinical assessment in Parkinson’s disease using machine learning. Park. Relat. Disord. 2019, 63, 111–116. [Google Scholar] [CrossRef] [Green Version]
  82. Xu, S.; Pan, Z. A novel ensemble of random forest for assisting diagnosis of Parkinson’s disease on small handwritten dynamics dataset. Int. J. Med. Inform. 2020, 144, 104283. [Google Scholar] [CrossRef] [PubMed]
  83. Fisher, J.M.; Hammerla, N.Y.; Ploetz, T.; Andras, P.; Rochester, L.; Walker, R.W. Unsupervised home monitoring of Parkinson’s disease motor symptoms using body-worn accelerometers. Park. Relat. Disord. 2016, 33, 44–50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Ilias, T.; Filip, B.; Radu, C.; Dag, N.; Marina, S.; Mevludin, M. Using measurements from wearable sensors for automatic scoring of Parkinson’s disease motor states: Results from 7 patients. Annu. Int. Conf. IEEE Eng. Med. Biol. Soc. 2017, 2017, 131–134. [Google Scholar] [CrossRef]
  85. di Biase, L.; Raiano, L.; Caminiti, M.L.; Pecoraro, P.M.; Di Lazzaroo, V. Artificial intelligence in Parkinson’s disease—Symptoms identification and monitoring. In Augmenting Neurological Disorder Prediction and Rehabilitation Using Artificial Intelligence; Elsevier: Amsterdam, The Netherlands, 2022; pp. 35–52. [Google Scholar]
  86. di Biase, L.; Tinkhauser, G.; Martin Moraud, E.; Caminiti, M.L.; Pecoraro, P.M.; Di Lazzaro, V. Adaptive, personalized closed-loop therapy for Parkinson’s disease: Biochemical, neurophysiological, and wearable sensing systems. Expert Rev. Neurother. 2021, 21, 1371–1388. [Google Scholar] [CrossRef]
  87. Contin, M.; Martinelli, P. Pharmacokinetics of levodopa. J. Neurol. 2010, 257 (Suppl. 2), S253–S261. [Google Scholar] [CrossRef]
  88. Nutt, J.G.; Fellman, J.H. Pharmacokinetics of levodopa. Clin. Neuropharmacol. 1984, 7, 35–49. [Google Scholar] [CrossRef] [PubMed]
  89. Juncos, J.L.; Fabbrini, G.; Mouradian, M.M.; Serrati, C.; Chase, T.N. Dietary influences on the antiparkinsonian response to levodopa. Arch. Neurol. 1987, 44, 1003–1005. [Google Scholar] [CrossRef]
  90. Guebila, M.B.; Thiele, I. Model-based dietary optimization for late-stage, levodopa-treated, Parkinson’s disease patients. NPJ Syst. Biol. Appl. 2016, 2, 1–8. [Google Scholar] [CrossRef] [Green Version]
  91. Contin, M.; Riva, R.; Martinelli, P.; Albani, F.; Baruzzi, A. Relationship between levodopa concentration, dyskinesias, and motor effect in parkinsonian patients: A 3-year follow-up study. Clin. Neuropharmacol. 1997, 20, 409–418. [Google Scholar] [CrossRef] [PubMed]
  92. Harder, S.; Baas, H. Concentration-response relationship of levodopa in patients at different stages of Parkinson’s disease. Clin. Pharmacol. Ther. 1998, 64, 183–191. [Google Scholar] [CrossRef] [PubMed]
  93. Contin, M.; Riva, R.; Martinelli, P.; Cortelli, P.; Albani, F.; Baruzzi, A. Pharmacodynamic modeling of oral levodopa: Clinical application in Parkinson’s disease. Neurology 1993, 43, 367. [Google Scholar] [CrossRef]
  94. Adamiak, U.; Kaldonska, M.; Klodowska-Duda, G.; Wyska, E.; Safranow, K.; Bialecka, M.; Gawronska-Szklarz, B. Pharmacokinetic-pharmacodynamic modeling of levodopa in patients with advanced Parkinson disease. Clin. Neuropharmacol. 2010, 33, 135–141. [Google Scholar] [CrossRef]
  95. Fabbrini, G.; Juncos, J.; Mouradian, M.; Serrati, C.; Chase, T. Levodopa pharmacokinetic mechanisms and motor fluctuations in Parkinson’s disease. Ann. Neurol. Off J. Am. Neurol. Assoc. Child Neurol. Soc. 1987, 21, 370–376. [Google Scholar] [CrossRef]
  96. Gancher, S.T.; Nutt, J.G.; Woodward, W.R. Peripheral pharmacokinetics of levodopa in untreated, stable, and fluctuating parkinsonian patients. Neurology 1987, 37, 940. [Google Scholar] [CrossRef]
  97. Contin, M.; Riva, R.; Martinelli, P.; Triggs, E.; Albani, F.; Baruzzi, A. Rate of motor response to oral levodopa and the clinical progression of Parkinson’s disease. Neurology 1996, 46, 1055–1058. [Google Scholar] [CrossRef]
  98. Contin, M.; Riva, R.; Martinelli, P.; Cortelli, P.; Albani, F.; Baruzzi, A. Longitudinal monitoring of the levodopa concentration-effect relationship in Parkinson’s disease. Neurology 1994, 44, 1287. [Google Scholar] [CrossRef] [PubMed]
  99. Contin, M.; Riva, R.; Martinelli, P.; Albani, F.; Avoni, P.; Baruzzi, A. Levodopa therapy monitoring in patients with Parkinson disease: A kinetic-dynamic approach. Ther. Drug Monit. 2001, 23, 621–629. [Google Scholar] [CrossRef] [PubMed]
  100. Triggs, E.; Charles, B.; Contin, M.; Martinelli, P.; Cortelli, P.; Riva, R.; Albani, F.; Baruzzi, A. Population pharmacokinetics and pharmacodynamics of oral levodopa in parkinsonian patients. Eur. J. Clin. Pharmacol. 1996, 51, 59–67. [Google Scholar] [CrossRef] [PubMed]
  101. Nyholm, D. Pharmacokinetic optimisation in the treatment of Parkinson’s disease. Clin. Pharmacokinet. 2006, 45, 109–136. [Google Scholar] [CrossRef]
  102. Nutt, J.G.; Carter, J.H.; Lea, E.S.; Sexton, G.J. Evolution of the response to levodopa during the first 4 years of therapy. Ann. Neurol. Off J. Am. Neurol. Assoc. Child Neurol. Soc. 2002, 51, 686–693. [Google Scholar] [CrossRef]
  103. Hughes, A.J.; Frankel, J.P.; Kempster, P.A.; Stern, G.M.; Lees, A.J. Motor response to levodopa in patients with parkinsonian motor fluctuations: A follow-up study over three years. J. Neurol. Neurosurg. Psychiatry 1994, 57, 430–434. [Google Scholar] [CrossRef] [Green Version]
  104. Contin, M.; Riva, R.; Martinelli, P.; Cortelli, P.; Albani, F.; Baruzzi, A. A levodopa kinetic-dynamic study of the rate of progression in Parkinson’s disease. Neurology 1998, 51, 1075–1080. [Google Scholar] [CrossRef]
  105. Baas, H.; Zehrden, F.; Selzer, R.; Kohnen, R.; Loetsch, J.; Harder, S. Pharmacokinetic-pharmacodynamic relationship of levodopa with and without tolcapone in patients with Parkinson’s disease. Clin. Pharmacokinet. 2001, 40, 383–393. [Google Scholar] [CrossRef]
  106. Richter, U.; Halje, P.; Petersson, P. Mechanisms underlying cortical resonant states: Implications for levodopa-induced dyskinesia. Rev. Neurosci. 2013, 24, 415–429. [Google Scholar] [CrossRef]
  107. Melgari, J.-M.; Curcio, G.; Mastrolilli, F.; Salomone, G.; Trotta, L.; Tombini, M.; Di Biase, L.; Scrascia, F.; Fini, R.; Fabrizio, E. Alpha and beta EEG power reflects L-dopa acute administration in parkinsonian patients. Front. Aging Neurosci. 2014, 6, 302. [Google Scholar] [CrossRef]
  108. di Biase, L.; Ricci, L.; Caminiti, M.L.; Pecoraro, P.M.; Carbone, S.P.; Di Lazzaro, V. Quantitative High Density EEG Brain Connectivity Evaluation in Parkinson’s Disease: The Phase Locking Value (PLV). J. Clin. Med. 2023, 12, 1450. [Google Scholar] [CrossRef] [PubMed]
  109. Bastide, M.F.; Meissner, W.G.; Picconi, B.; Fasano, S.; Fernagut, P.O.; Feyder, M.; Francardo, V.; Alcacer, C.; Ding, Y.; Brambilla, R.; et al. Pathophysiology of L-dopa-induced motor and non-motor complications in Parkinson’s disease. Prog. Neurobiol. 2015, 132, 96–168. [Google Scholar] [CrossRef] [PubMed]
  110. Albin, R.L.; Young, A.B.; Penney, J.B. The functional anatomy of basal ganglia disorders. Trends Neurosci. 1989, 12, 366–375. [Google Scholar] [CrossRef] [PubMed]
  111. DeLong, M.R. Primate models of movement disorders of basal ganglia origin. Trends Neurosci. 1990, 13, 281–285. [Google Scholar] [CrossRef] [PubMed]
  112. Lee, J.I.; Verhagen Metman, L.; Ohara, S.; Dougherty, P.M.; Kim, J.H.; Lenz, F.A. Internal pallidal neuronal activity during mild drug-related dyskinesias in Parkinson’s disease: Decreased firing rates and altered firing patterns. J. Neurophysiol. 2007, 97, 2627–2641. [Google Scholar] [CrossRef] [Green Version]
  113. Levy, R.; Dostrovsky, J.O.; Lang, A.E.; Sime, E.; Hutchison, W.D.; Lozano, A.M. Effects of apomorphine on subthalamic nucleus and globus pallidus internus neurons in patients with Parkinson’s disease. J. Neurophysiol. 2001, 86, 249–260. [Google Scholar] [CrossRef]
  114. Merello, M.; Balej, J.; Delfino, M.; Cammarota, A.; Betti, O.; Leiguarda, R. Apomorphine induces changes in GPi spontaneous outflow in patients with Parkinson’s disease. Mov. Disord. 1999, 14, 45–49. [Google Scholar] [CrossRef]
  115. Lozano, A.M.; Lang, A.E.; Levy, R.; Hutchison, W.; Dostrovsky, J. Neuronal recordings in Parkinson’s disease patients with dyskinesias induced by apomorphine. Ann. Neurol. 2000, 47 (Suppl. 1), S141–S146. [Google Scholar]
  116. Foffani, G.; Ardolino, G.; Meda, B.; Egidi, M.; Rampini, P.; Caputo, E.; Baselli, G.; Priori, A. Altered subthalamo-pallidal synchronisation in parkinsonian dyskinesias. J. Neurol. Neurosurg. Psychiatry 2005, 76, 426–428. [Google Scholar] [CrossRef] [Green Version]
  117. Rodriguez-Oroz, M.C.; Lopez-Azcarate, J.; Garcia-Garcia, D.; Alegre, M.; Toledo, J.; Valencia, M.; Guridi, J.; Artieda, J.; Obeso, J.A. Involvement of the subthalamic nucleus in impulse control disorders associated with Parkinson’s disease. Brain A J. Neurol. 2011, 134 Pt 1, 36–49. [Google Scholar] [CrossRef] [Green Version]
  118. Silberstein, P.; Oliviero, A.; Di Lazzaro, V.; Insola, A.; Mazzone, P.; Brown, P. Oscillatory pallidal local field potential activity inversely correlates with limb dyskinesias in Parkinson’s disease. Exp. Neurol. 2005, 194, 523–529. [Google Scholar] [CrossRef] [PubMed]
  119. Little, S.; Pogosyan, A.; Kuhn, A.A.; Brown, P. beta band stability over time correlates with Parkinsonian rigidity and bradykinesia. Exp. Neurol. 2012, 236, 383–388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Lopez-Azcarate, J.; Tainta, M.; Rodriguez-Oroz, M.C.; Valencia, M.; Gonzalez, R.; Guridi, J.; Iriarte, J.; Obeso, J.A.; Artieda, J.; Alegre, M. Coupling between beta and high-frequency activity in the human subthalamic nucleus may be a pathophysiological mechanism in Parkinson’s disease. J. Neurosci. 2010, 30, 6667–6677. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Prescott, I.A.; Liu, L.D.; Dostrovsky, J.O.; Hodaie, M.; Lozano, A.M.; Hutchison, W.D. Lack of depotentiation at basal ganglia output neurons in PD patients with levodopa-induced dyskinesia. Neurobiol. Dis. 2014, 71, 24–33. [Google Scholar] [CrossRef] [PubMed]
  122. Nadjar, A.; Gerfen, C.R.; Bezard, E. Priming for l-dopa-induced dyskinesia in Parkinson’s disease: A feature inherent to the treatment or the disease? Prog. Neurobiol. 2009, 87, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Miocinovic, S.; Swann, N.C.; de Hemptinne, C.; Miller, A.; Ostrem, J.L.; Starr, P.A. Cortical gamma oscillations in isolated dystonia. Park. Relat. Disord. 2018, 49, 104–105. [Google Scholar] [CrossRef]
  124. Kempf, F.; Brücke, C.; Salih, F.; Trottenberg, T.; Kupsch, A.; Schneider, G.H.; Doyle Gaynor, L.M.; Hoffmann, K.T.; Vesper, J.; Wöhrle, J. Gamma activity and reactivity in human thalamic local field potentials. Eur. J. Neurosci. 2009, 29, 943–953. [Google Scholar] [CrossRef]
  125. Benarroch, E.E. Serotonergic modulation of basal ganglia circuits: Complexity and therapeutic opportunities. Neurology 2009, 73, 880–886. [Google Scholar] [CrossRef]
  126. Berger, M.; Gray, J.A.; Roth, B.L. The expanded biology of serotonin. Annu. Rev. Med. 2009, 60, 355–366. [Google Scholar] [CrossRef] [Green Version]
  127. Lavoie, B.; Parent, A. Immunohistochemical study of the serotoninergic innervation of the basal ganglia in the squirrel monkey. J. Comp. Neurol. 1990, 299, 1–16. [Google Scholar] [CrossRef]
  128. Fox, S.H.; Chuang, R.; Brotchie, J.M. Serotonin and Parkinson’s disease: On movement, mood, and madness. Mov. Disord. Off J. Mov. Disord. Soc. 2009, 24, 1255–1266. [Google Scholar] [CrossRef] [PubMed]
  129. Arai, R.; Karasawa, N.; Geffard, M.; Nagatsu, I. L-DOPA is converted to dopamine in serotonergic fibers of the striatum of the rat: A double-labeling immunofluorescence study. Neurosci. Lett. 1995, 195, 195–198. [Google Scholar] [CrossRef] [PubMed]
  130. Arai, R.; Karasawa, N.; Geffard, M.; Nagatsu, T.; Nagatsu, I. Immunohistochemical evidence that central serotonin neurons produce dopamine from exogenousl-DOPA in the rat, with reference to the involvement of aromaticl-amino acid decarboxylase. Brain Res. 1994, 667, 295–299. [Google Scholar] [CrossRef] [PubMed]
  131. Tanaka, H.; Kannari, K.; Maeda, T.; Tomiyama, M.; Suda, T.; Matsunaga, M. Role of serotonergic neurons in L-DOPA-derived extracellular dopamine in the striatum of 6-OHDA-lesioned rats. Neuroreport 1999, 10, 631–634. [Google Scholar] [CrossRef]
  132. 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]
  133. Eskow, K.L.; Dupre, K.B.; Barnum, C.J.; Dickinson, S.O.; Park, J.Y.; Bishop, C. The role of the dorsal raphe nucleus in the development, expression, and treatment of L-dopa-induced dyskinesia in hemiparkinsonian rats. Synapse 2009, 63, 610–620. [Google Scholar] [CrossRef] [Green Version]
  134. 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]
  135. 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]
  136. Conti, M.M.; Ostock, C.Y.; Lindenbach, D.; Goldenberg, A.A.; Kampton, E.; Dell’isola, R.; Katzman, A.C.; Bishop, C. Effects of prolonged selective serotonin reuptake inhibition on the development and expression of L-DOPA-induced dyskinesia in hemi-parkinsonian rats. Neuropharmacology 2014, 77, 1–8. [Google Scholar] [CrossRef] [Green Version]
  137. 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] [Green Version]
  138. Lee, J.; Zhu, W.-M.; Stanic, D.; Finkelstein, D.I.; Horne, M.H.; Henderson, J.; Lawrence, A.J.; O’Connor, L.; Tomas, D.; Drago, J. Sprouting of dopamine terminals and altered dopamine release and uptake in Parkinsonian dyskinaesia. Brain 2008, 131, 1574–1587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Ahmed, I.; Bose, S.K.; Pavese, N.; Ramlackhansingh, A.; Turkheimer, F.; Hotton, G.; Hammers, A.; Brooks, D.J. Glutamate NMDA receptor dysregulation in Parkinson’s disease with dyskinesias. Brain 2011, 134, 979–986. [Google Scholar] [CrossRef] [PubMed]
  140. Calabresi, P.; Mercuri, N.; Sancesario, G.; Bernardi, G. Electrophysiology of dopamine-denervated striatal neurons. Implications for Parkinson’s disease. Brain A J. Neurol. 1993, 116, 433–452. [Google Scholar]
  141. Hallett, P.J.; Dunah, A.; Ravenscroft, P.; Zhou, S.; Bezard, E.; Crossman, A.R.; Brotchie, J.M.; Standaert, D.G. Alterations of striatal NMDA receptor subunits associated with the development of dyskinesia in the MPTP-lesioned primate model of Parkinson’s disease. Neuropharmacology 2005, 48, 503–516. [Google Scholar] [CrossRef]
  142. Picconi, B.; Gardoni, F.; Centonze, D.; Mauceri, D.; Cenci, M.A.; Bernardi, G.; Calabresi, P.; Di Luca, M. Abnormal Ca2+-calmodulin-dependent protein kinase II function mediates synaptic and motor deficits in experimental parkinsonism. J. Neurosci. 2004, 24, 5283–5291. [Google Scholar] [CrossRef]
  143. Gubellini, P.; Picconi, B.; Bari, M.; Battista, N.; Calabresi, P.; Centonze, D.; Bernardi, G.; Finazzi-Agro, A.; Maccarrone, M. Experimental parkinsonism alters endocannabinoid degradation: Implications for striatal glutamatergic transmission. J. Neurosci. 2002, 22, 6900–6907. [Google Scholar] [CrossRef] [Green Version]
  144. Calabresi, P.; Pisani, A.; Mercuri, N.B.; Bernardi, G. Heterogeneity of metabotropic glutamate receptors in the striatum: Electrophysiological evidence. Eur. J. Neurosci. 1993, 5, 1370–1377. [Google Scholar] [CrossRef]
  145. Greenamyre, J.T. Glutamatergic influences on the basal ganglia. Clin. Neuropharmacol. 2001, 24, 65–70. [Google Scholar] [CrossRef]
  146. Duty, S. Targeting glutamate receptors to tackle the pathogenesis, clinical symptoms and levodopa-induced dyskinesia associated with Parkinson’s disease. CNS Drugs 2012, 26, 1017–1032. [Google Scholar] [CrossRef]
  147. Gubellini, P.; Saulle, E.; Centonze, D.; Bonsi, P.; Pisani, A.; Bernardi, G.; Conquet, F.; Calabresi, P. Selective involvement of mGlu1 receptors in corticostriatal LTD. Neuropharmacology 2001, 40, 839–846. [Google Scholar] [CrossRef]
  148. Gubellini, P.; Saulle, E.; Centonze, D.; Costa, C.; Tropepi, D.; Bernardi, G.; Conquet, F.; Calabresi, P. Corticostriatal LTP requires combined mGluR1 and mGluR5 activation. Neuropharmacology 2003, 44, 8–16. [Google Scholar] [CrossRef] [PubMed]
  149. Rüdiger, M.; Maria, F.; Rodney, C.; Adele, G.; Kari, J.; Carmela, M.; Richard, S.; Karl, S.; Pier, S.; Stuck, B.E. (International Commission on Non-Ionizing Radiation Protection). ICNIRP Guidelines for Limiting Exposure to Electric Fields Induced by Movement of the Human Body in a Static Magnetic Field and by Time-Varying Magnetic Fields below 1 Hz. Health Phys. 2014, 106, 418–425. [Google Scholar]
  150. Sgambato-Faure, V.; Cenci, M.A. Glutamatergic mechanisms in the dyskinesias induced by pharmacological dopamine replacement and deep brain stimulation for the treatment of Parkinson’s disease. Prog. Neurobiol. 2012, 96, 69–86. [Google Scholar] [CrossRef] [PubMed]
  151. Picconi, B.; Centonze, D.; Håkansson, K.; Bernardi, G.; Greengard, P.; Fisone, G.; Cenci, M.A.; Calabresi, P. Loss of bidirectional striatal synaptic plasticity in L-DOPA–induced dyskinesia. Nat. Neurosci. 2003, 6, 501–506. [Google Scholar] [CrossRef] [PubMed]
  152. Vastagh, C.; Gardoni, F.; Bagetta, V.; Stanic, J.; Zianni, E.; Giampà, C.; Picconi, B.; Calabresi, P.; Di Luca, M. N-methyl-D-aspartate (NMDA) receptor composition modulates dendritic spine morphology in striatal medium spiny neurons. J. Biol. Chem. 2012, 287, 18103–18114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Lee, C.-Y.; Lee, C.-H.; Shih, C.-C.; Liou, H.-H. Paraquat inhibits postsynaptic AMPA receptors on dopaminergic neurons in the substantia nigra pars compacta. Biochem. Pharmacol. 2008, 76, 1155–1164. [Google Scholar] [CrossRef]
  154. Bernard, V.; Gardiol, A.; Faucheux, B.; Bloch, B.; Agid, Y.; Hirsch, E.C. Expression of glutamate receptors in the human and rat basal ganglia: Effect of the dopaminergic denervation on AMPA receptor gene expression in the striatopallidal complex in Parkinson’s disease and rat with 6-OHDA lesion. J. Comp. Neurol. 1996, 368, 553–568. [Google Scholar] [CrossRef]
  155. Betarbet, R.; Porter, R.H.; Greenamyre, J.T. GluR1 glutamate receptor subunit is regulated differentially in the primate basal ganglia following nigrostriatal dopamine denervation. J. Neurochem. 2000, 74, 1166–1174. [Google Scholar] [CrossRef]
  156. Santini, E.; Valjent, E.; Usiello, A.; Carta, M.; Borgkvist, A.; Girault, J.-A.; Hervé, D.; Greengard, P.; Fisone, G. Critical involvement of cAMP/DARPP-32 and extracellular signal-regulated protein kinase signaling in L-DOPA-induced dyskinesia. J. Neurosci. 2007, 27, 6995–7005. [Google Scholar] [CrossRef] [Green Version]
  157. Berridge, C.W.; Waterhouse, B.D. The locus coeruleus–noradrenergic system: Modulation of behavioral state and state-dependent cognitive processes. Brain Res. Rev. 2003, 42, 33–84. [Google Scholar] [CrossRef]
  158. Mason, S.T.; Fibiger, H.C. Regional topography within noradrenergic locus coeruleus as revealed by retrograde transport of horseradish peroxidase. J. Comp. Neurol. 1979, 187, 703–724. [Google Scholar] [CrossRef] [PubMed]
  159. Gaspar, P. Anatomy of the Noradrenergic Pathways in the Primate Brain Alteration in Parkinson’s Disease. In Noradrenergic Mechanisms in Parkinson’s Disease; CRC Press: Boca Raton, FL, USA, 1994; pp. 74–88. [Google Scholar]
  160. Margules, D.; Lewis, M.J.; Dragovich, J.A.; Margules, A.S. Hypothalamic norepinephrine: Circadian rhythms and the control of feeding behavior. Science 1972, 178, 640–643. [Google Scholar] [CrossRef] [PubMed]
  161. Pfaus, J.G. Pathways of sexual desire. J. Sex. Med. 2009, 6, 1506–1533. [Google Scholar] [CrossRef]
  162. Alachkar, A.; Brotchie, J.M.; Jones, O.T. Changes in the mRNA levels of α 2A and α 2C adrenergic receptors in rat models of parkinson’s disease and L-DOPA-induced dyskinesia. J. Mol. Neurosci. 2012, 46, 145–152. [Google Scholar] [CrossRef] [PubMed]
  163. Ribas, C.; Miralles, A.; Busquets, X.; García-Sevilla, J.A. Brain α2-adrenoceptors in monoamine-depleted rats: Increased receptor density, G coupling proteins, receptor turnover and receptor mRNA. Br. J. Pharmacol. 2001, 132, 1467–1476. [Google Scholar] [CrossRef] [Green Version]
  164. Rommelfanger, K.; Edwards, G.; Freeman, K.; Liles, L.; Miller, G.; Weinshenker, D. Norepinephrine loss produces more profound motor deficits than MPTP treatment in mice. Proc. Natl. Acad. Sci. USA 2007, 104, 13804–13809. [Google Scholar] [CrossRef] [Green Version]
  165. Perez, V.; Marín, C.; Rubio, A.; Aguilar, E.; Barbanoj, M.; Kulisevsky, J. Effect of the additional noradrenergic neurodegeneration to 6-OHDA-lesioned rats in levodopa-induced dyskinesias and in cognitive disturbances. J. Neural. Transm. 2009, 116, 1257–1266. [Google Scholar] [CrossRef]
  166. Miguelez, C.; Aristieta, A.; Cenci, M.A.; Ugedo, L. The locus coeruleus is directly implicated in L-DOPA-induced dyskinesia in parkinsonian rats: An electrophysiological and behavioural study. PLoS ONE 2011, 6, e24679. [Google Scholar] [CrossRef]
  167. Buck, K.; Ferger, B. Comparison of intrastriatal administration of noradrenaline and l-DOPA on dyskinetic movements: A bilateral reverse in vivo microdialysis study in 6-hydroxydopamine-lesioned rats. Neuroscience 2009, 159, 16–20. [Google Scholar] [CrossRef]
  168. Arai, A.; Tomiyama, M.; Kannari, K.; Kimura, T.; Suzuki, C.; Watanabe, M.; Kawarabayashi, T.; Shen, H.; Shoji, M. Reuptake of L-DOPA-derived extracellular DA in the striatum of a rodent model of Parkinson’s disease via norepinephrine transporter. Synapse 2008, 62, 632–635. [Google Scholar] [CrossRef]
  169. Sommermeyer, H.; Frielingsdorf, J.; Knorr, A. Effects of prazosin on the dopaminergic neurotransmission in rat brain. Eur. J. Pharmacol. 1995, 276, 267–270. [Google Scholar] [CrossRef]
  170. Henry, B.; Fox, S.H.; Peggs, D.; Crossman, A.R.; Brotchie, J.M. The α2-adrenergic receptor antagonist idazoxan reduces dyskinesia and enhances anti-parkinsonian actions of L-dopa in the MPTP-lesioned primate model of Parkinson’s disease. Mov. Disord. Off J. Mov. Disord. Soc. 1999, 14, 744–753. [Google Scholar] [CrossRef]
  171. Savola, J.M.; Hill, M.; Engstrom, M.; Merivuori, H.; Wurster, S.; McGuire, S.G.; Fox, S.H.; Crossman, A.R.; Brotchie, J.M. Fipamezole (JP-1730) is a potent α2 adrenergic receptor antagonist that reduces levodopa-induced dyskinesia in the MPTP-lesioned primate model of Parkinson’s disease. Mov. Disord. Off J. Mov. Disord. Soc. 2003, 18, 872–883. [Google Scholar] [CrossRef] [PubMed]
  172. Fox, S.H.; Henry, B.; Hill, M.P.; Peggs, D.; Crossman, A.R.; Brotchie, J.M. Neural mechanisms underlying peak-dose dyskinesia induced by levodopa and apomorphine are distinct: Evidence from the effects of the alpha2 adrenoceptor antagonist idazoxan. Mov. Disord. Off J. Mov. Disord. Soc. 2001, 16, 642–650. [Google Scholar] [CrossRef] [PubMed]
  173. Johnston, T.H.; Fox, S.H.; Piggott, M.J.; Savola, J.M.; Brotchie, J.M. The α2 adrenergic antagonist fipamezole improves quality of levodopa action in Parkinsonian primates. Mov. Disord. 2010, 25, 2084–2093. [Google Scholar] [CrossRef] [PubMed]
  174. Carpentier, A.; Bonnet, A.; Vidailhet, M.; Agid, Y. Improvement of levodopa-induced dyskinesia by propranolol in Parkinson’s disease. Neurology 1996, 46, 1548–1551. [Google Scholar] [CrossRef] [PubMed]
  175. Millar, N.S.; Gotti, C. Diversity of vertebrate nicotinic acetylcholine receptors. Neuropharmacology 2009, 56, 237–246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Albuquerque, E.X.; Pereira, E.F.; Alkondon, M.; Rogers, S.W. Mammalian nicotinic acetylcholine receptors: From structure to function. Physiol. Rev. 2009, 89, 73–120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Quik, M.; Wonnacott, S. α6β2* and α4β2* nicotinic acetylcholine receptors as drug targets for Parkinson’s disease. Pharmacol. Rev. 2011, 63, 938–966. [Google Scholar] [CrossRef] [Green Version]
  178. Quik, M.; Cox, H.; Parameswaran, N.; O’Leary, K.; Langston, J.W.; Di Monte, D. Nicotine reduces levodopa-induced dyskinesias in lesioned monkeys. Ann. Neurol. Off J. Am. Neurol. Assoc. Child Neurol. Soc. 2007, 62, 588–596. [Google Scholar] [CrossRef]
  179. Quik, M.; Mallela, A.; Ly, J.; Zhang, D. Nicotine reduces established levodopa-induced dyskinesias in a monkey model of Parkinson’s disease. Mov. Disord. 2013, 28, 1398–1406. [Google Scholar] [CrossRef] [PubMed]
  180. Quik, M.; Mallela, A.; Chin, M.; McIntosh, J.M.; Perez, X.A.; Bordia, T. Nicotine-mediated improvement in L-dopa-induced dyskinesias in MPTP-lesioned monkeys is dependent on dopamine nerve terminal function. Neurobiol. Dis. 2013, 50, 30–41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Huang, L.Z.; Grady, S.R.; Quik, M. Nicotine reduces L-DOPA-induced dyskinesias by acting at β2* nicotinic receptors. J. Pharmacol. Exp. Ther. 2011, 338, 932–941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Quik, M.; Park, K.M.; Hrachova, M.; Mallela, A.; Huang, L.Z.; McIntosh, J.M.; Grady, S.R. Role for α6 nicotinic receptors in l-dopa-induced dyskinesias in parkinsonian mice. Neuropharmacology 2012, 63, 450–459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Quik, M.; Campos, C.; Grady, S.R. Multiple CNS nicotinic receptors mediate L-dopa-induced dyskinesias: Studies with parkinsonian nicotinic receptor knockout mice. Biochem. Pharmacol. 2013, 86, 1153–1162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Caulfield, M.P.; Birdsall, N.J. International Union of Pharmacology. XVII. Classification of muscarinic acetylcholine receptors. Pharmacol. Rev. 1998, 50, 279–290. [Google Scholar]
  185. Langmead, C.J.; Watson, J.; Reavill, C. Muscarinic acetylcholine receptors as CNS drug targets. Pharmacol. Ther. 2008, 117, 232–243. [Google Scholar] [CrossRef]
  186. Gomez-Mancilla, B.; Bedard, P.J. Effect of nondopaminergic drugs on L-dopa-induced dyskinesias in MPTP-treated monkeys. Clin. Neuropharmacol. 1993, 16, 418–427. [Google Scholar] [CrossRef]
  187. Huot, P.; Johnston, T.H.; Koprich, J.B.; Fox, S.H.; Brotchie, J.M. The pharmacology of L-DOPA-induced dyskinesia in Parkinson’s disease. Pharmacol. Rev. 2013, 65, 171–222. [Google Scholar] [CrossRef]
  188. Hanrieder, J.; Ljungdahl, A.; Fälth, M.; Mammo, S.E.; Bergquist, J.; Andersson, M. L-DOPA-induced dyskinesia is associated with regional increase of striatal dynorphin peptides as elucidated by imaging mass spectrometry. Mol. Cell. Proteom. 2011, 10. [Google Scholar] [CrossRef] [Green Version]
  189. Ljungdahl, A.; Hanrieder, J.; Fälth, M.; Bergquist, J.; Andersson, M. Imaging mass spectrometry reveals elevated nigral levels of dynorphin neuropeptides in L-DOPA-induced dyskinesia in rat model of Parkinson’s disease. PLoS ONE 2011, 6, e25653. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Bourdenx, M.; Nilsson, A.; Wadensten, H.; Fälth, M.; Li, Q.; Crossman, A.R.; Andrén, P.E.; Bezard, E. Abnormal structure-specific peptide transmission and processing in a primate model of Parkinson’s disease and L-DOPA-induced dyskinesia. Neurobiol. Dis. 2014, 62, 307–312. [Google Scholar] [CrossRef] [PubMed]
  191. Samadi, P.; Grégoire, L.; Bédard, P.J. Opioid antagonists increase the dyskinetic response to dopaminergic agents in parkinsonian monkeys: Interaction between dopamine and opioid systems. Neuropharmacology 2003, 45, 954–963. [Google Scholar] [CrossRef] [PubMed]
  192. Samadi, P.; Grégoire, L.; Tahar, A.H.; Di Paolo, T.; Rouillard, C.; Bédard, P.J. Naltrexone in the short-term decreases antiparkinsonian response to l-Dopa and in the long-term increases dyskinesias in drug-naive parkinsonian monkeys. Neuropharmacology 2005, 49, 165–173. [Google Scholar] [CrossRef]
  193. Samadi, P.; Grégoire, L.; Rouillard, C.; Bédard, P.J. Dyskinesias occur in response to saline and naltrexone alone after priming with combination of dopaminergic agents and naltrexone in the MPTP parkinsonian monkeys. Neurobiol. Dis. 2005, 19, 266–272. [Google Scholar] [CrossRef]
  194. Ikeda, K.; Yoshikawa, S.; Kurokawa, T.; Yuzawa, N.; Nakao, K.; Mochizuki, H. TRK-820, a selective kappa opioid receptor agonist, could effectively ameliorate L-DOPA-induced dyskinesia symptoms in a rat model of Parkinson’s disease. Eur. J. Pharmacol. 2009, 620, 42–48. [Google Scholar] [CrossRef]
  195. Kano, M.; Ohno-Shosaku, T.; Hashimotodani, Y.; Uchigashima, M.; Watanabe, M. Endocannabinoid-mediated control of synaptic transmission. Physiol. Rev. 2009, 89, 309–380. [Google Scholar] [CrossRef]
  196. O’sullivan, S. Cannabinoids go nuclear: Evidence for activation of peroxisome proliferator-activated receptors. Br. J. Pharmacol. 2007, 152, 576–582. [Google Scholar] [CrossRef] [Green Version]
  197. Devane, W.A.; Hanuš, L.; Breuer, A.; Pertwee, R.G.; Stevenson, L.A.; Griffin, G.; Gibson, D.; Mandelbaum, A.; Etinger, A.; Mechoulam, R. Isolation and structure of a brain constituent that binds to the cannabinoid receptor. Science 1992, 258, 1946–1949. [Google Scholar] [CrossRef]
  198. Di Marzo, V.; Fontana, A.; Cadas, H.; Schinelli, S.; Cimino, G.; Schwartz, J.-C.; Piomelli, D. Formation and inactivation of endogenous cannabinoid anandamide in central neurons. Nature 1994, 372, 686–691. [Google Scholar] [CrossRef] [Green Version]
  199. Mechoulam, R.; Ben-Shabat, S.; Hanus, L.; Ligumsky, M.; Kaminski, N.E.; Schatz, A.R.; Gopher, A.; Almog, S.; Martin, B.R.; Compton, D.R. Identification of an endogenous 2-monoglyceride, present in canine gut, that binds to cannabinoid receptors. Biochem. Pharmacol. 1995, 50, 83–90. [Google Scholar] [CrossRef] [PubMed]
  200. López-Moreno, J.A.; González-Cuevas, G.; Moreno, G.; Navarro, M. The pharmacology of the endocannabinoid system: Functional and structural interactions with other neurotransmitter systems and their repercussions in behavioral addiction. Addict. Biol. 2008, 13, 160–187. [Google Scholar] [CrossRef] [PubMed]
  201. Console-Bram, L.; Marcu, J.; Abood, M.E. Cannabinoid receptors: Nomenclature and pharmacological principles. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2012, 38, 4–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Wallmichrath, I.; Szabo, B. Cannabinoids inhibit striatonigral GABAergic neurotransmission in the mouse. Neuroscience 2002, 113, 671–682. [Google Scholar] [CrossRef]
  203. Martín, A.B.; Fernandez-Espejo, E.; Ferrer, B.; Gorriti, M.A.; Bilbao, A.; Navarro, M.; Rodriguez de Fonseca, F.; Moratalla, R. Expression and function of CB1 receptor in the rat striatum: Localization and effects on D1 and D2 dopamine receptor-mediated motor behaviors. Neuropsychopharmacology 2008, 33, 1667–1679. [Google Scholar] [CrossRef] [Green Version]
  204. Morera-Herreras, T.; Ruiz-Ortega, J.; Gomez-Urquijo, S.; Ugedo, L. Involvement of subthalamic nucleus in the stimulatory effect of Δ9-tetrahydrocannabinol on dopaminergic neurons. Neuroscience 2008, 151, 817–823. [Google Scholar] [CrossRef]
  205. Gerdeman, G.L.; Ronesi, J.; Lovinger, D.M. Postsynaptic endocannabinoid release is critical to long-term depression in the striatum. Nat. Neurosci. 2002, 5, 446–451. [Google Scholar] [CrossRef]
  206. Morgese, M.G.; Cassano, T.; Cuomo, V.; Giuffrida, A. Anti-dyskinetic effects of cannabinoids in a rat model of Parkinson’s disease: Role of CB1 and TRPV1 receptors. Exp. Neurol. 2007, 208, 110–119. [Google Scholar] [CrossRef] [Green Version]
  207. Giuffrida, A.; Parsons, L.; Kerr, T.; De Fonseca, F.R.; Navarro, M.; Piomelli, D. Dopamine activation of endogenous cannabinoid signaling in dorsal striatum. Nat. Neurosci. 1999, 2, 358–363. [Google Scholar] [CrossRef]
  208. Masserano, J.; Karoum, F.; Wyatt, R. SR 141716A, a CB1 cannabinoid receptor antagonist, potentiates the locomotor stimulant effects of amphetamine and apomorphine. Behav. Pharmacol. 1999, 10, 429–432. [Google Scholar] [CrossRef]
  209. Kreitzer, A.C.; Malenka, R.C. Endocannabinoid-mediated rescue of striatal LTD and motor deficits in Parkinson’s disease models. Nature 2007, 445, 643–647. [Google Scholar] [CrossRef] [PubMed]
  210. Fox, S.H.; Henry, B.; Hill, M.; Crossman, A.; Brotchie, J. Stimulation of cannabinoid receptors reduces levodopa-induced dyskinesia in the MPTP-lesioned nonhuman primate model of Parkinson’s disease. Mov. Disord. Off J. Mov. Disord. Soc. 2002, 17, 1180–1187. [Google Scholar] [CrossRef] [PubMed]
  211. Martinez, A.; Macheda, T.; Morgese, M.G.; Trabace, L.; Giuffrida, A. The cannabinoid agonist WIN55212-2 decreases L-DOPA-induced PKA activation and dyskinetic behavior in 6-OHDA-treated rats. Neurosci. Res. 2012, 72, 236–242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Walsh, S.; Gorman, A.M.; Finn, D.P.; Dowd, E. The effects of cannabinoid drugs on abnormal involuntary movements in dyskinetic and non-dyskinetic 6-hydroxydopamine lesioned rats. Brain Res. 2010, 1363, 40–48. [Google Scholar] [CrossRef]
  213. Sieradzan, K.; Fox, S.; Hill, M.; Dick, J.; Crossman, A.; Brotchie, J. Cannabinoids reduce levodopa-induced dyskinesia in Parkinson’s disease: A pilot study. Neurology 2001, 57, 2108–2111. [Google Scholar] [CrossRef] [PubMed]
  214. Hermann, H.; Marsicano, G.; Lutz, B. Coexpression of the cannabinoid receptor type 1 with dopamine and serotonin receptors in distinct neuronal subpopulations of the adult mouse forebrain. Neuroscience 2002, 109, 451–460. [Google Scholar] [CrossRef]
  215. Mathur, B.N.; Capik, N.A.; Alvarez, V.A.; Lovinger, D.M. Serotonin induces long-term depression at corticostriatal synapses. J. Neurosci. 2011, 31, 7402–7411. [Google Scholar] [CrossRef] [Green Version]
  216. Nakazi, M.; Bauer, U.; Nickel, T.; Kathmann, M.; Schlicker, E. Inhibition of serotonin release in the mouse brain via presynaptic cannabinoid CB 1 receptors. Naunyn-Schmiedebergs Arch. Pharmacol. 2000, 361, 19–24. [Google Scholar]
  217. Linden, J. Molecular approach to adenosine receptors: Receptor-mediated mechanisms of tissue protection. Annu. Rev. Pharmacol. Toxicol. 2001, 41, 775–787. [Google Scholar] [CrossRef]
  218. Cunha, R.A. Neuroprotection by adenosine in the brain: From A 1 receptor activation to A 2A receptor blockade. Purinergic Signal. 2005, 1, 111–134. [Google Scholar] [CrossRef] [Green Version]
  219. Gomes, C.V.; Kaster, M.P.; Tomé, A.R.; Agostinho, P.M.; Cunha, R.A. Adenosine receptors and brain diseases: Neuroprotection and neurodegeneration. Biochim. Biophys. Acta 2011, 1808, 1380–1399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. V Lopes, L.; M Sebastiao, A.; A Ribeiro, J. Adenosine and related drugs in brain diseases: Present and future in clinical trials. Curr. Top. Med. Chem. 2011, 11, 1087–1101. [Google Scholar] [CrossRef]
  221. Borea, P.A.; Gessi, S.; Merighi, S.; Varani, K. Adenosine as a multi-signalling guardian angel in human diseases: When, where and how does it exert its protective effects? Trends Pharmacol. Sci. 2016, 37, 419–434. [Google Scholar] [CrossRef] [PubMed]
  222. Burnstock, G. Purinergic signalling: Therapeutic developments. Front. Pharmacol. 2017, 8, 661. [Google Scholar] [CrossRef] [Green Version]
  223. Jacobson, K.A.; Merighi, S.; Varani, K.; Borea, P.A.; Baraldi, S.; Aghazadeh Tabrizi, M.; Romagnoli, R.; Baraldi, P.G.; Ciancetta, A.; Tosh, D.K. A3 adenosine receptors as modulators of inflammation: From medicinal chemistry to therapy. Med. Res. Rev. 2018, 38, 1031–1072. [Google Scholar] [CrossRef] [PubMed]
  224. Fuxe, K.; Ferré, S.; Zoli, M.; Agnati, L.F. Integrated events in central dopamine transmission as analyzed at multiple levels. Evidence for intramembrane adenosine A2A/dopamine D2 and adenosine A1/dopamine D1 receptor interactions in the basal ganglia. Brain Res. Rev. 1998, 26, 258–273. [Google Scholar] [CrossRef]
  225. Fuxe, K.; Ferré, S.; Genedani, S.; Franco, R.; Agnati, L.F. Adenosine receptor–dopamine receptor interactions in the basal ganglia and their relevance for brain function. Physiol. Behav. 2007, 92, 210–217. [Google Scholar] [CrossRef]
  226. Hauser, R.A.; Cantillon, M.; Pourcher, E.; Micheli, F.; Mok, V.; Onofrj, M.; Huyck, S.; Wolski, K. Preladenant in patients with Parkinson’s disease and motor fluctuations: A phase 2, double-blind, randomised trial. Lancet Neurol. 2011, 10, 221–229. [Google Scholar] [CrossRef]
  227. Pourcher, E.; Hauser, R.A. Adenosinergic Receptor Antagonists: Clinical Experience in Parkinson’s Disease. In The Adenosinergic System: A Non-Dopaminergic Target in Parkinson’s Disease; Springer: Cham, Switzerland, 2015; pp. 291–307. [Google Scholar]
  228. LeWitt, P.A.; Guttman, M.; Tetrud, J.W.; Tuite, P.J.; Mori, A.; Chaikin, P.; Sussman, N.M.; Group, U.S. Adenosine A2A receptor antagonist istradefylline (KW-6002) reduces “off” time in Parkinson’s disease: A double-blind, randomized, multicenter clinical trial (6002-US-005). Ann. Neurol. 2008, 63, 295–302. [Google Scholar] [CrossRef]
  229. Hauser, R.A.; Shulman, L.M.; Trugman, J.M.; Roberts, J.W.; Mori, A.; Ballerini, R.; Sussman, N.M. Study of istradefylline in patients with Parkinson’s disease on levodopa with motor fluctuations. Mov. Disord. Off J. Mov. Disord. Soc. 2008, 23, 2177–2185. [Google Scholar] [CrossRef]
  230. Stacy, M.; Silver, D.; Mendis, T.; Sutton, J.; Mori, A.; Chaikin, P.; Sussman, N. A 12-week, placebo-controlled study (6002-US-006) of istradefylline in Parkinson disease. Neurology 2008, 70, 2233–2240. [Google Scholar] [CrossRef] [PubMed]
  231. Mizuno, Y.; Kondo, T.; Group, J.I.S. Adenosine A2A receptor antagonist istradefylline reduces daily OFF time in Parkinson’s disease. Mov. Disord. 2013, 28, 1138–1141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Hauser, R.A.; Olanow, C.W.; Kieburtz, K.D.; Pourcher, E.; Docu-Axelerad, A.; Lew, M.; Kozyolkin, O.; Neale, A.; Resburg, C.; Meya, U. Tozadenant (SYN115) in patients with Parkinson’s disease who have motor fluctuations on levodopa: A phase 2b, double-blind, randomised trial. Lancet Neurol. 2014, 13, 767–776. [Google Scholar] [CrossRef]
  233. Therapies, B. Biotie’s tozadenant (SYN115) meets primary and multiple secondary endpoints in phase 2b study in Parkinson’s disease. Press. Release Dec. 2012, 11. [Google Scholar]
  234. Politis, M. Neuroimaging in Parkinson disease: From research setting to clinical practice. Nat. Rev. Neurol. 2014, 10, 708–722. [Google Scholar] [CrossRef]
  235. Au, W.L.; Adams, J.R.; Troiano, A.R.; Stoessl, A.J. Parkinson’s disease: In vivo assessment of disease progression using positron emission tomography. Brain Res. Mol. Brain Res. 2005, 134, 24–33. [Google Scholar] [CrossRef] [PubMed]
  236. Brooks, D.J.; Frey, K.A.; Marek, K.L.; Oakes, D.; Paty, D.; Prentice, R.; Shults, C.W.; Stoessl, A.J. Assessment of neuroimaging techniques as biomarkers of the progression of Parkinson’s disease. Exp. Neurol. 2003, 184 (Suppl. 1), S68–S79. [Google Scholar] [CrossRef]
  237. Sossi, V.; Doudet, D.J.; Holden, J.E. A reversible tracer analysis approach to the study of effective dopamine turnover. J. Cereb. Blood Flow. Metab. 2001, 21, 469–476. [Google Scholar] [CrossRef] [Green Version]
  238. Linazasoro, G.; Antonini, A.; Maguire, R.P.; Leenders, K.L. Pharmacological and PET studies in patient’s with Parkinson’s disease and a short duration-motor response: Implications in the pathophysiology of motor complications. J. Neural Transm. 2004, 111, 497–509. [Google Scholar] [CrossRef]
  239. Hong, J.Y.; Oh, J.S.; Lee, I.; Sunwoo, M.K.; Ham, J.H.; Lee, J.E.; Sohn, Y.H.; Kim, J.S.; Lee, P.H. Presynaptic dopamine depletion predicts levodopa-induced dyskinesia in de novo Parkinson disease. Neurology 2014, 82, 1597–1604. [Google Scholar] [CrossRef]
  240. de la Fuente-Fernández, R.; Lu, J.Q.; Sossi, V.; Jivan, S.; Schulzer, M.; Holden, J.E.; Lee, C.S.; Ruth, T.J.; Calne, D.B.; Stoessl, A.J. Biochemical variations in the synaptic level of dopamine precede motor fluctuations in Parkinson’s disease: PET evidence of increased dopamine turnover. Ann. Neurol. 2001, 49, 298–303. [Google Scholar] [CrossRef] [PubMed]
  241. Stoessl, A.J. Central pharmacokinetics of levodopa: Lessons from imaging studies. Mov. Disord. 2015, 30, 73–79. [Google Scholar] [CrossRef] [PubMed]
  242. Pavese, N.; Evans, A.H.; Tai, Y.F.; Hotton, G.; Brooks, D.J.; Lees, A.J.; Piccini, P. Clinical correlates of levodopa-induced dopamine release in Parkinson disease: A PET study. Neurology 2006, 67, 1612–1617. [Google Scholar] [CrossRef] [PubMed]
  243. Doudet, D.J.; Chan, G.L.; Holden, J.E.; McGeer, E.G.; Aigner, T.A.; Wyatt, R.J.; Ruth, T.J. 6-[18F]Fluoro-L-DOPA PET studies of the turnover of dopamine in MPTP-induced parkinsonism in monkeys. Synapse 1998, 29, 225–232. [Google Scholar] [CrossRef]
  244. Sossi, V.; de La Fuente-Fernandez, R.; Holden, J.E.; Doudet, D.J.; McKenzie, J.; Stoessl, A.J.; Ruth, T.J. Increase in dopamine turnover occurs early in Parkinson’s disease: Evidence from a new modeling approach to PET 18 F-fluorodopa data. J. Cereb. Blood Flow. Metab. 2002, 22, 232–239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Sossi, V.; de la Fuente-Fernandez, R.; Holden, J.E.; Schulzer, M.; Ruth, T.J.; Stoessl, J. Changes of dopamine turnover in the progression of Parkinson’s disease as measured by positron emission tomography: Their relation to disease-compensatory mechanisms. J. Cereb. Blood Flow. Metab. 2004, 24, 869–876. [Google Scholar] [CrossRef] [Green Version]
  246. Sossi, V.; de la Fuente-Fernandez, R.; Schulzer, M.; Adams, J.; Stoessl, J. Age-related differences in levodopa dynamics in Parkinson’s: Implications for motor complications. Brain 2006, 129, 1050–1058. [Google Scholar] [CrossRef] [Green Version]
  247. Golbe, L.I. Young-onset Parkinson’s disease: A clinical review. Neurology 1991, 41 Pt 1, 168–173. [Google Scholar] [CrossRef]
  248. Grandas, F.; Galiano, M.L.; Tabernero, C. Risk factors for levodopa-induced dyskinesias in Parkinson’s disease. J. Neurol. 1999, 246, 1127–1133. [Google Scholar] [CrossRef]
  249. Quinn, N.; Critchley, P.; Marsden, C.D. Young onset Parkinson’s disease. Mov. Disord. 1987, 2, 73–91. [Google Scholar] [CrossRef]
  250. Kumar, N.; Van Gerpen, J.A.; Bower, J.H.; Ahlskog, J.E. Levodopa-dyskinesia incidence by age of Parkinson’s disease onset. Mov. Disord. 2005, 20, 342–344. [Google Scholar] [CrossRef]
  251. Piccini, P.; Weeks, R.A.; Brooks, D.J. Alterations in opioid receptor binding in Parkinson’s disease patients with levodopa-induced dyskinesias. Ann. Neurol. 1997, 42, 720–726. [Google Scholar] [CrossRef] [PubMed]
  252. Mishina, M.; Ishiwata, K.; Naganawa, M.; Kimura, Y.; Kitamura, S.; Suzuki, M.; Hashimoto, M.; Ishibashi, K.; Oda, K.; Sakata, M.; et al. Adenosine A(2A) receptors measured with [C]TMSX PET in the striata of Parkinson’s disease patients. PLoS ONE 2011, 6, e17338. [Google Scholar] [CrossRef] [Green Version]
  253. Hirano, S.; Asanuma, K.; Ma, Y.; Tang, C.; Feigin, A.; Dhawan, V.; Carbon, M.; Eidelberg, D. Dissociation of metabolic and neurovascular responses to levodopa in the treatment of Parkinson’s disease. J. Neurosci. Off J. Soc. Neurosci. 2008, 28, 4201–4209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Faucheux, B.A.; Bonnet, A.M.; Agid, Y.; Hirsch, E.C. Blood vessels change in the mesencephalon of patients with Parkinson’s disease. Lancet 1999, 353, 981–982. [Google Scholar] [CrossRef] [PubMed]
  255. Ohlin, K.E.; Francardo, V.; Lindgren, H.S.; Sillivan, S.E.; O’Sullivan, S.S.; Luksik, A.S.; Vassoler, F.M.; Lees, A.J.; Konradi, C.; Cenci, M.A. Vascular endothelial growth factor is upregulated by L-dopa in the parkinsonian brain: Implications for the development of dyskinesia. Brain 2011, 134, 2339–2357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Lee, J.-Y.; Seo, S.; Lee, J.S.; Kim, H.-J.; Kim, Y.K.; Jeon, B.S. Putaminal serotonergic innervation: Monitoring dyskinesia risk in Parkinson disease. Neurology 2015, 85, 853–860. [Google Scholar] [CrossRef]
  257. Stocchi, F.; Tagliati, M.; Olanow, C.W. Treatment of levodopa-induced motor complications. Mov. Disord. 2008, 23 (Suppl. 3), S599–S612. [Google Scholar] [CrossRef]
  258. Obeso, J.A.; Olanow, C.W.; Nutt, J.G. Levodopa motor complications in Parkinson’s disease. Trends Neurosci. 2000, 23, S2–S7. [Google Scholar] [CrossRef]
  259. Del Sorbo, F.; Albanese, A. Levodopa-induced dyskinesias and their management. J. Neurol. 2008, 255 (Suppl. 4), 32–41. [Google Scholar] [CrossRef]
  260. 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]
  261. Salomone, G.; Marano, M.; di Biase, L.; Melgari, J.M.; Di Lazzaro, V. Dopamine dysregulation syndrome and punding in levodopa-carbidopa intestinal gel (LCIG) infusion: A serious but preventable complication. Park. Relat. Disord. 2015, 21, 1124–1125. [Google Scholar] [CrossRef]
  262. Melgari, J.M.; Salomone, G.; di Biase, L.; Marano, M.; Scrascia, F.; Di Lazzaro, V. Dyskinesias during levodopa-carbidopa intestinal gel (LCIG) infusion: Management inclinical practice. Park. Relat. Disord. 2015, 21, 327–328. [Google Scholar] [CrossRef] [PubMed]
  263. Marano, M.; Naranian, T.; di Biase, L.; Di Santo, A.; Poon, Y.Y.; Arca, R.; Cossu, G.; Marano, P.; Di Lazzaro, V.; Fasano, A. Complex dyskinesias in Parkinson patients on levodopa/carbidopa intestinal gel. Park. Relat. Disord. 2019, 69, 140–146. [Google Scholar] [CrossRef] [PubMed]
  264. Marano, M.; Pizzicannella, M.; di Biase, L.; Rea, R.; Di Santo, A.; Martino, M.; Andrisani, G.; Pandolfi, M.; Di Matteo, F.M.; Di Lazzaro, V. Jejunal pulling syndrome: A peculiar LCIG complication. Park. Relat. Disord. 2018, 52, 113–114. [Google Scholar] [CrossRef] [PubMed]
  265. Manson, A.J.; Turner, K.; Lees, A.J. Apomorphine monotherapy in the treatment of refractory motor complications of Parkinson’s disease: Long-term follow-up study of 64 patients. Mov. Disord. 2002, 17, 1235–1241. [Google Scholar] [CrossRef]
  266. Di Lazzaro, V.; Rothwell, J.C. Corticospinal activity evoked and modulated by non-invasive stimulation of the intact human motor cortex. J. Physiol. 2014, 592, 4115–4128. [Google Scholar] [CrossRef]
  267. Lefaucheur, J.P.; Aleman, A.; Baeken, C.; Benninger, D.H.; Brunelin, J.; Di Lazzaro, V.; Filipović, S.R.; Grefkes, C.; Hasan, A.; Hummel, F.C.; et al. Evidence-based guidelines on the therapeutic use of repetitive transcranial magnetic stimulation (rTMS): An update (2014-2018). Clin. Neurophysiol. 2020, 131, 474–528. [Google Scholar] [CrossRef]
  268. di Biase, L.; Falato, E.; Di Lazzaro, V. Transcranial Focused Ultrasound (tFUS) and Transcranial Unfocused Ultrasound (tUS) Neuromodulation: From Theoretical Principles to Stimulation Practices. Front. Neurol. 2019, 10, 549. [Google Scholar] [CrossRef] [Green Version]
  269. di Biase, L.; Falato, E.; Caminiti, M.L.; Pecoraro, P.M.; Narducci, F.; Di Lazzaro, V. Focused Ultrasound (FUS) for Chronic Pain Management: Approved and Potential Applications. Neurol. Res. Int. 2021, 2021, 8438498. [Google Scholar] [CrossRef]
  270. di Biase, L.; Munhoz, R.P. Deep brain stimulation for the treatment of hyperkinetic movement disorders. Expert Rev. Neurother. 2016, 16, 1067–1078. [Google Scholar] [CrossRef] [PubMed]
  271. Kern, D.S.; Picillo, M.; Thompson, J.A.; Sammartino, F.; di Biase, L.; Munhoz, R.P.; Fasano, A. Interleaving Stimulation in Parkinson’s Disease, Tremor, and Dystonia. Stereotact. Funct. Neurosurg. 2018, 96, 379–391. [Google Scholar] [CrossRef] [PubMed]
  272. Sandoe, C.; Krishna, V.; Basha, D.; Sammartino, F.; Tatsch, J.; Picillo, M.; di Biase, L.; Poon, Y.Y.; Hamani, C.; Reddy, D.; et al. Predictors of deep brain stimulation outcome in tremor patients. Brain Stimul. 2018, 11, 592–599. [Google Scholar] [CrossRef] [PubMed]
  273. Krack, P.; Volkmann, J.; Tinkhauser, G.; Deuschl, G. Deep Brain Stimulation in Movement Disorders: From Experimental Surgery to Evidence-Based Therapy. Mov. Disord. 2019, 34, 1795–1810. [Google Scholar] [CrossRef] [PubMed]
  274. Rizzone, M.; Fasano, A.; Daniele, A.; Zibetti, M.; Merola, A.; Rizzi, L.; Piano, C.; Piccininni, C.; Romito, L.; Lopiano, L. Long-term outcome of subthalamic nucleus DBS in Parkinson’s disease: From the advanced phase towards the late stage of the disease? Park. Relat. Disord. 2014, 20, 376–381. [Google Scholar] [CrossRef]
  275. di Biase, L.; Fasano, A. Low-frequency deep brain stimulation for Parkinson’s disease: Great expectation or false hope? Mov. Disord. 2016, 31, 962–967. [Google Scholar] [CrossRef]
  276. Tinkhauser, G.; Pogosyan, A.; Debove, I.; Nowacki, A.; Shah, S.A.; Seidel, K.; Tan, H.; Brittain, J.S.; Petermann, K.; di Biase, L.; et al. Directional local field potentials: A tool to optimize deep brain stimulation. Mov. Disord. 2018, 33, 159–164. [Google Scholar] [CrossRef]
  277. di Biase, L.; Piano, C.; Bove, F.; Ricci, L.; Caminiti, M.L.; Stefani, A.; Viselli, F.; Modugno, N.; Cerroni, R.; Calabresi, P. Intraoperative Local Field Potential Beta Power and Three-Dimensional Neuroimaging Mapping Predict Long-Term Clinical Response to Deep Brain Stimulation in Parkinson Disease: A Retrospective Study. Neuromodulation Technol. Neural Interface 2023. [Google Scholar] [CrossRef]
  278. Espay, A.J.; Morgante, F.; Merola, A.; Fasano, A.; Marsili, L.; Fox, S.H.; Bezard, E.; Picconi, B.; Calabresi, P.; Lang, A.E. Levodopa-induced dyskinesia in Parkinson disease: Current and evolving concepts. Ann. Neurol. 2018, 84, 797–811. [Google Scholar] [CrossRef] [Green Version]
  279. Kwon, D.K.; Kwatra, M.; Wang, J.; Ko, H.S. Levodopa-Induced Dyskinesia in Parkinson’s Disease: Pathogenesis and Emerging Treatment Strategies. Cells 2022, 11, 3736. [Google Scholar] [CrossRef]
  280. Vijayakumar, D.; Jankovic, J. Drug-Induced Dyskinesia, Part 1: Treatment of Levodopa-Induced Dyskinesia. Drugs 2016, 76, 759–777. [Google Scholar] [CrossRef]
  281. Hislop, J.; Margolesky, J.; Shpiner, D.S. Sublingual apomorphine in treatment of Parkinson’s disease: A review. Int. J. Neurosci. 2022, 1–7. [Google Scholar] [CrossRef]
  282. LeWitt, P.A. Levodopa therapy for Parkinson’s disease: Pharmacokinetics and pharmacodynamics. Mov. Disord. 2015, 30, 64–72. [Google Scholar] [CrossRef]
  283. Poewe, W.; Antonini, A. Novel formulations and modes of delivery of levodopa. Mov. Disord. Off. J. Mov. Disord. Soc. 2015, 30, 114–120. [Google Scholar] [CrossRef] [PubMed]
  284. Antonini, A.; Odin, P.; Opiano, L.; Tomantschger, V.; Pacchetti, C.; Pickut, B.; Gasser, U.E.; Calandrella, D.; Mancini, F.; Zibetti, M.; et al. Effect and safety of duodenal levodopa infusion in advanced Parkinson’s disease: A retrospective multicenter outcome assessment in patient routine care. J. Neural Transm. 2013, 120, 1553–1558. [Google Scholar] [CrossRef] [PubMed]
  285. Antonini, A.; Isaias, I.U.; Canesi, M.; Zibetti, M.; Mancini, F.; Manfredi, L.; Dal Fante, M.; Lopiano, L.; Pezzoli, G. Duodenal levodopa infusion for advanced Parkinson’s disease: 12-month treatment outcome. Mov. Disord. 2007, 22, 1145–1149. [Google Scholar] [CrossRef] [PubMed]
  286. Antonini, A.; Poewe, W.; Chaudhuri, K.R.; Jech, R.; Pickut, B.; Pirtošek, Z.; Szasz, J.; Valldeoriola, F.; Winkler, C.; Bergmann, L. Levodopa-carbidopa intestinal gel in advanced Parkinson’s: Final results of the GLORIA registry. Park. Relat. Disord. 2017, 45, 13–20. [Google Scholar] [CrossRef] [Green Version]
  287. Szász, J.A.; Constantin, V.A.; Orbán-Kis, K.; Bancu, L.A.; Ciorba, M.; Mihály, I.; Nagy, E.E.; Szász, R.M.; Kelemen, K.; Simu, M.A. Management challenges of severe, complex dyskinesia. Data from a large cohort of patients treated with Levodopa-carbidopa intestinal gel for advanced Parkinson’s disease. Brain Sci. 2021, 11, 826. [Google Scholar] [CrossRef]
  288. Hauser, R.A.; Hsu, A.; Kell, S.; Espay, A.J.; Sethi, K.; Stacy, M.; Ondo, W.; O’Connell, M.; Gupta, S.; investigators, I.A.-P. Extended-release carbidopa-levodopa (IPX066) compared with immediate-release carbidopa-levodopa in patients with Parkinson’s disease and motor fluctuations: A phase 3 randomised, double-blind trial. Lancet Neurol. 2013, 12, 346–356. [Google Scholar] [CrossRef]
  289. 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. Park. Relat. Disord. 2014, 20, 1335–1340. [Google Scholar] [CrossRef] [Green Version]
  290. Lewitt, P.A.; Ellenbogen, A.; Chen, D.; Lal, R.; McGuire, K.; Zomorodi, K.; Luo, W.; Huff, F.J. Actively transported levodopa prodrug XP21279: A study in patients with Parkinson disease who experience motor fluctuations. Clin. Neuropharmacol. 2012, 35, 103–110. [Google Scholar] [CrossRef] [PubMed]
  291. Bonifati, V.; Fabrizio, E.; Cipriani, R.; Vanacore, N.; Meco, G. Buspirone in levodopa-induced dyskinesias. Clin. Neuropharmacol. 1994, 17, 73–82. [Google Scholar] [CrossRef] [PubMed]
  292. Politis, M.; Wu, K.; Loane, C.; Brooks, D.J.; Kiferle, L.; Turkheimer, F.E.; Bain, P.; Molloy, S.; Piccini, P. Serotonergic mechanisms responsible for levodopa-induced dyskinesias in Parkinson’s disease patients. J. Clin. Investig. 2014, 124, 1340–1349. [Google Scholar] [CrossRef] [PubMed]
  293. Goetz, C.G.; Damier, P.; Hicking, C.; Laska, E.; Muller, T.; Olanow, C.W.; Rascol, O.; Russ, H. Sarizotan as a treatment for dyskinesias in Parkinson’s disease: A double-blind placebo-controlled trial. Mov. Disord. 2007, 22, 179–186. [Google Scholar] [CrossRef] [PubMed]
  294. Bezard, E.; Munoz, A.; Tronci, E.; Pioli, E.Y.; Li, Q.; Porras, G.; Bjorklund, A.; Carta, M. Anti-dyskinetic effect of anpirtoline in animal models of L-DOPA-induced dyskinesia. Neurosci. Res. 2013, 77, 242–246. [Google Scholar] [CrossRef] [PubMed]
  295. Bezard, E.; Tronci, E.; Pioli, E.Y.; Li, Q.; Porras, G.; Bjorklund, A.; Carta, M. Study of the antidyskinetic effect of eltoprazine in animal models of levodopa-induced dyskinesia. Mov. Disord. 2013, 28, 1088–1096. [Google Scholar] [CrossRef]
  296. 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] [Green Version]
  297. Tronci, E.; Fidalgo, C.; Zianni, E.; Collu, M.; Stancampiano, R.; Morelli, M.; Gardoni, F.; Carta, M. Effect of memantine on L-DOPA-induced dyskinesia in the 6-OHDA-lesioned rat model of Parkinson’s disease. Neuroscience 2014, 265, 245–252. [Google Scholar] [CrossRef]
  298. Clarke, C.E.; Cooper, J.A.; Holdich, T.A.H.; Study, G. A Randomized, Double-blind, Placebo-controlled, Ascending-dose Tolerability and Safety Study of Remacemide as Adjuvant Therapy in Parkinson’s Disease with Response Fluctuations. Clin. Neuropharmacol. 2001, 24, 133–138. [Google Scholar] [CrossRef]
  299. Giuffra, M.E.; Sethy, V.H.; Davis, T.L.; Mouradian, M.M.; Chase, T.N. Milacemide therapy for Parkinson’s disease. Mov. Disord. 1993, 8, 47–50. [Google Scholar] [CrossRef]
  300. Nutt, J.G.; Gunzler, S.A.; Kirchhoff, T.; Hogarth, P.; Weaver, J.L.; Krams, M.; Jamerson, B.; Menniti, F.S.; Landen, J.W. Effects of a NR2B selective NMDA glutamate antagonist, CP-101,606, on dyskinesia and parkinsonism. Mov. Disord. 2008, 23, 1860–1866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  301. Van Blercom, N.; Lasa, A.; Verger, K.; Masramón, X.; Sastre, V.M.; Linazasoro, G. Effects of Gabapentin on the Motor Response to Levodopa: A Double-Blind, Placebo-Controlled, Crossover Study in Patients With Complicated Parkinson Disease. Clin. Neuropharmacol. 2004, 27, 124–128. [Google Scholar] [CrossRef] [PubMed]
  302. Wolz, M.; Löhle, M.; Strecker, K.; Schwanebeck, U.; Schneider, C.; Reichmann, H.; Grählert, X.; Schwarz, J.; Storch, A. Levetiracetam for levodopa-induced dyskinesia in Parkinson’s disease: A randomized, double-blind, placebo-controlled trial. J. Neural Transm. 2010, 117, 1279–1286. [Google Scholar] [CrossRef] [PubMed]
  303. Iwata, Y.; Irie, S.; Uchida, H.; Suzuki, T.; Watanabe, K.; Iwashita, S.; Mimura, M. Effects of zonisamide on tardive dyskinesia: A preliminary open-label trial. J. Neurol. Sci. 2012, 315, 137–140. [Google Scholar] [CrossRef]
  304. Durif, F.; Debilly, B.; Galitzky, M.; Morand, D.; Viallet, F.; Borg, M.; Thobois, S.; Broussolle, E.; Rascol, O. Clozapine improves dyskinesias in Parkinson disease: A double-blind, placebo-controlled study. Neurology 2004, 62, 381–388. [Google Scholar] [CrossRef]
  305. Katzenschlager, R.; Manson, A.; Evans, A.; Watt, H.; Lees, A. Low dose quetiapine for drug induced dyskinesias in Parkinson’s disease: A double blind cross over study. J. Neurol. Neurosurg. Psychiatry 2004, 75, 295–297. [Google Scholar]
  306. Manson, A.; Schrag, A.; Lees, A. Low-dose olanzapine for levodopa induced dyskinesias. Neurology 2000, 55, 795–799. [Google Scholar] [CrossRef]
  307. di Biase, L. Method and Device for the Objective Characterization of Symptoms of Parkinson’s Disease. WO Patent WO2022054112A1, 17 March 2022. [Google Scholar]
  308. di Biase, L. Method for the Management of Oral Therapy in Parkinson’s Disease. WO Patent WO2022054113A1, 17 March 2022. [Google Scholar]
  309. di Biase, L. Adaptive Method and System for a Personalized Daily Infusion Therapy of Parkinson’s Disease. WO Patent WO2022054111A1, 17 March 2022. [Google Scholar]
  310. Baston, C.; Contin, M.; Calandra Buonaura, G.; Cortelli, P.; Ursino, M. A mathematical model of levodopa medication effect on basal ganglia in parkinson’s Disease: An application to the alternate finger tapping task. Front. Hum. Neurosci. 2016, 10, 280. [Google Scholar] [CrossRef]
  311. Ursino, M.; Magosso, E.; Lopane, G.; Calandra-Buonaura, G.; Cortelli, P.; Contin, M. Mathematical modeling and parameter estimation of levodopa motor response in patients with parkinson disease. PLoS ONE 2020, 15, e0229729. [Google Scholar] [CrossRef] [Green Version]
  312. Chan, P.L.; Nutt, J.G.; Holford, N.H. Modeling the short-and long-duration responses to exogenous levodopa and to endogenous levodopa production in Parkinson’s disease. J. Pharmacokinet. Pharmacodyn. 2004, 31, 243–268. [Google Scholar] [CrossRef]
  313. Ursino, M.; Baston, C. Aberrant learning in Parkinson’s disease: A neurocomputational study on bradykinesia. Eur. J. Neurosci. 2018, 47, 1563–1582. [Google Scholar] [CrossRef] [PubMed]
  314. Deleu, D.; Sarre, S.; Ebinger, G.; Michotte, Y. Simultaneous monitoring of levodopa, dopamine and their metabolites in skeletal muscle and subcutaneous tissue in different pharmacological conditions using microdialysis. J. Pharm. Biomed. Anal. 1993, 11, 577–585. [Google Scholar] [CrossRef] [PubMed]
  315. Goud, K.Y.; Moonla, C.; Mishra, R.K.; Yu, C.; Narayan, R.; Litvan, I.; Wang, J. Wearable electrochemical microneedle sensor for continuous monitoring of levodopa: Toward Parkinson management. ACS Sens. 2019, 4, 2196–2204. [Google Scholar] [CrossRef] [PubMed]
  316. Dizdar, N.; Kullman, A.; Norlander, B.; Olsson, J.-E.; Kagedal, B. Human pharmacokinetics of L-3, 4-dihydroxyphenylalanine studied with microdialysis. Clin. Chem. 1999, 45, 1813–1820. [Google Scholar] [CrossRef] [Green Version]
  317. Van Gompel, J.J.; Chang, S.-Y.; Goerss, S.J.; Kim, I.Y.; Kimble, C.; Bennet, K.E.; Lee, K.H. Development of intraoperative electrochemical detection: Wireless instantaneous neurochemical concentration sensor for deep brain stimulation feedback. Neurosurg. Focus 2010, 29, E6. [Google Scholar] [CrossRef] [Green Version]
  318. Rosin, B.; Slovik, M.; Mitelman, R.; Rivlin-Etzion, M.; Haber, S.N.; Israel, Z.; Vaadia, E.; Bergman, H. Closed-loop deep brain stimulation is superior in ameliorating parkinsonism. Neuron 2011, 72, 370–384. [Google Scholar] [CrossRef] [Green Version]
  319. Little, S.; Pogosyan, A.; Neal, S.; Zavala, B.; Zrinzo, L.; Hariz, M.; Foltynie, T.; Limousin, P.; Ashkan, K.; FitzGerald, J. Adaptive deep brain stimulation in advanced Parkinson disease. Ann. Neurol. 2013, 74, 449–457. [Google Scholar] [CrossRef] [Green Version]
  320. Brown, P.; Oliviero, A.; Mazzone, P.; Insola, A.; Tonali, P.; Di Lazzaro, V. Dopamine dependency of oscillations between subthalamic nucleus and pallidum in Parkinson’s disease. J. Neurosci. 2001, 21, 1033–1038. [Google Scholar] [CrossRef] [Green Version]
  321. Hammond, C.; Bergman, H.; Brown, P. Pathological synchronization in Parkinson’s disease: Networks, models and treatments. Trends Neurosci. 2007, 30, 357–364. [Google Scholar] [CrossRef]
  322. Neumann, W.J.; Degen, K.; Schneider, G.H.; Brücke, C.; Huebl, J.; Brown, P.; Kühn, A.A. Subthalamic synchronized oscillatory activity correlates with motor impairment in patients with Parkinson’s disease. Mov. Disord. 2016, 31, 1748–1751. [Google Scholar] [CrossRef] [Green Version]
  323. Godinho, C.; Domingos, J.; Cunha, G.; Santos, A.T.; Fernandes, R.M.; Abreu, D.; Gonçalves, N.; Matthews, H.; Isaacs, T.; Duffen, J. A systematic review of the characteristics and validity of monitoring technologies to assess Parkinson’s disease. J. Neuroeng. Rehabil. 2016, 13, 24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Relationship between the incidence of levodopa-induced dyskinesias and levodopa exposure time. Data in the graph derive from the mean levodopa exposure time from each study reported in Table 2 expressed in years, the blue line represents the trend line for the reported data.
Figure 1. Relationship between the incidence of levodopa-induced dyskinesias and levodopa exposure time. Data in the graph derive from the mean levodopa exposure time from each study reported in Table 2 expressed in years, the blue line represents the trend line for the reported data.
Jcm 12 04427 g001
Figure 2. Levodopa-induced dyskinesia types. Legend: red line = plasma levodopa concentration [L-dopa] over time; lower dashed line = levodopa concentration threshold for OFF–ON transition; upper dashed line = levodopa concentration threshold for peak-dose dyskinesia transition; orange boxes = OFF dystonia periods; blue boxes = diphasic dyskinesia periods; yellow box = square-wave dyskinesia period; green box = peak-dose dyskinesia period.
Figure 2. Levodopa-induced dyskinesia types. Legend: red line = plasma levodopa concentration [L-dopa] over time; lower dashed line = levodopa concentration threshold for OFF–ON transition; upper dashed line = levodopa concentration threshold for peak-dose dyskinesia transition; orange boxes = OFF dystonia periods; blue boxes = diphasic dyskinesia periods; yellow box = square-wave dyskinesia period; green box = peak-dose dyskinesia period.
Jcm 12 04427 g002
Figure 3. LID and ON/OFF motor status recognition trough convolutional neural network. (Figure drawn by Dr. Pfister, reproduced, under the terms of the Creative Commons Attribution 4.0 License, from [78]).
Figure 3. LID and ON/OFF motor status recognition trough convolutional neural network. (Figure drawn by Dr. Pfister, reproduced, under the terms of the Creative Commons Attribution 4.0 License, from [78]).
Jcm 12 04427 g003
Figure 4. LID neurophysiology findings at cortical and basal ganglia level (for whole description and references, see the corresponding paragraph: 2.7. Neurophysiology).
Figure 4. LID neurophysiology findings at cortical and basal ganglia level (for whole description and references, see the corresponding paragraph: 2.7. Neurophysiology).
Jcm 12 04427 g004
Figure 5. Nuclear neuroimaging ligands for Parkinson’s disease studies. Legend: DAT: Dopamine Transporter; PET: Positron Emission Tomography; SPECT: Single Photon Emission Computed Tomography; VMAT-2: Vesicular Monoamine Transporter 2. (The figure was partly generated using Servier Medical Art, provided by Servier, licensed under a Creative Commons Attribution 3.0 unported license).
Figure 5. Nuclear neuroimaging ligands for Parkinson’s disease studies. Legend: DAT: Dopamine Transporter; PET: Positron Emission Tomography; SPECT: Single Photon Emission Computed Tomography; VMAT-2: Vesicular Monoamine Transporter 2. (The figure was partly generated using Servier Medical Art, provided by Servier, licensed under a Creative Commons Attribution 3.0 unported license).
Jcm 12 04427 g005
Figure 6. Upper graph: pharmacokinetic of 4 single doses of levodopa without the computation of the previous doses (Doses 1 to 4). Lower graph: real pharmacokinetic of the same 4 single doses of levodopa describing the net effect of the therapeutic scheme (the net dosing scheme is the result of the combination of Doses 1 to 4). Legend: lower dashed line = levodopa concentration threshold for OFF–ON transition; upper dashed line = levodopa concentration threshold for peak-dose dyskinesia transition.
Figure 6. Upper graph: pharmacokinetic of 4 single doses of levodopa without the computation of the previous doses (Doses 1 to 4). Lower graph: real pharmacokinetic of the same 4 single doses of levodopa describing the net effect of the therapeutic scheme (the net dosing scheme is the result of the combination of Doses 1 to 4). Legend: lower dashed line = levodopa concentration threshold for OFF–ON transition; upper dashed line = levodopa concentration threshold for peak-dose dyskinesia transition.
Jcm 12 04427 g006
Figure 7. Adaptive closed-loop therapy system for Parkinson’s disease patients. The four therapies, alone or in combination (deep brain stimulation, oral therapy, intestinal infusion therapy and subcutaneous infusion therapy) can be administered in adaptive controlled closed-loop way, through a multiparametric (LFP, biochemical and kinematic monitoring) sensing system. (Reproduced, under the terms of the Creative Commons Attribution 4.0 License, from [86]).
Figure 7. Adaptive closed-loop therapy system for Parkinson’s disease patients. The four therapies, alone or in combination (deep brain stimulation, oral therapy, intestinal infusion therapy and subcutaneous infusion therapy) can be administered in adaptive controlled closed-loop way, through a multiparametric (LFP, biochemical and kinematic monitoring) sensing system. (Reproduced, under the terms of the Creative Commons Attribution 4.0 License, from [86]).
Jcm 12 04427 g007
Table 1. Genetic risk factor for LIDs.
Table 1. Genetic risk factor for LIDs.
GeneFunctionGene VariantRef.
COMTInvolved in the metabolism of dopamineVal158Met[30]
DRD2Encodes the D2 subtype of the dopamine receptorANKK1 TTCTA haplotype[32]
BDNFInvolved in neuronal survival and synaptic plasticityVal66Met[34]
SLC6A3Dopamine transporter gene (DAT)40-bp VNTR[31]
GRIN2AEncodes a subunit of the NMDA receptor, involved in glutamate transmissionrs7192557 and rs8057394[33]
Legend: BDNF, brain-derived neurotrophic factor gene; COMT, catechol-O-methyltransferase; DRD2, dopamine receptor D2 gene; GRIN2A, N-methyl- d -aspartate receptor subunit 2A; SLC6A3, dopamine transporter gene.
Table 2. Relationship between the incidence of levodopa-induced dyskinesias and levodopa exposure time.
Table 2. Relationship between the incidence of levodopa-induced dyskinesias and levodopa exposure time.
Type of StudyN° PatientsLevodopa
Exposure
% Dyskinetic PatientsRef.
Meta-analysis335 (pre-levodopa era)3–6 w23.3[41]
606 (pre-levodopa era)2–4 m34.5
606 (pre-levodopa era)5–6 m53.4
2645 (pre-levodopa era)7–12 m54.8
982 (pre-levodopa era)1–2 y71.9
297(pre-levodopa era)2.5–3.5 y56.7
4327–12 m7
575 1–2 y28.7
7472.5–3.5 y26.9
1599 4–6 y36.2
514 9–15+ y87.8
Prospective, double-blind, randomized clinical trial885 y45[24]
2710 y77.8
Clinico-pathological 426.4 y31[39]
14.3 y61.9
Community-based87<5 y11[7]
6–9 y32
>10 y89
Community-based1265 y30[40]
10 y59
DATATOP35220.5 m30[42]
FIRST187 on IR 5 y 20.6[43]
193 on CR5 y21.8
056-study45 5 y45[23]
CALM-PD study1313 y 54[44]
PELMOPET study903 y26[22]
ELLDOPA study92 (150 mg/die)8 m3[21]
88 (300 mg/die)8 m2
91 (600 mg/die)8 m16
STRIDE-PD study372 (Ldopa)2 y32[28]
373 (Ldopa + entecapone)38
SIDNEY study5215 y94% (12% severe dyskinesias)[45]
Legend: w, weeks; m, months; y, years.
Table 3. Summary table of the main characteristics of neurotransmitter systems involved in LIDs.
Table 3. Summary table of the main characteristics of neurotransmitter systems involved in LIDs.
NeurotransmitterReceptorTherapeutic TargetEffect on LIDs
Serotonergic systemSerotonin5-HT1A
5-HT1B
5-HT1A and 5-HT1B Receptor Agonists- Density of serotonergic terminals in the striatum directly correlates with the severity of LIDs
- Serotonergic neurons convert exogenous levodopa into dopamine and release it without autoregulatory feedback
5-HT2A
5-HT2C
5-HT2A Receptors Antagonists
5-HT3
-
Not established
-
SERT inhibition
Glutamatergic systemGlutamatemGluRMGluR antagonist- Altered trafficking
- Hyperactive
NMDANMDA receptor antagonist (GluN2A/B subunit)- Altered trafficking
- Alteration of subunit composition
- Supersensitivity in the putamen following long-term levodopa
AMPAAMPA receptor antagonist- Altered trafficking
- Increased index of rectification (IR) of AMPA current in striatal medium spiny neurons
- Increased activity of Ca2+ -permeable AMPAR due to hyperphosphorylation
of GluR1 subunit
Noradrenergic systemNoradrenalineα-1/2α receptor antagonist- NA loss causes parkinsonism and spontaneous dyskinesias in DBH knock-out mice
- NA infusion in the striatum promotes LID in hemiparkinsonian rats
- NAT activity should re-uptake DA and reduce LIDs
- Controversial evidence
β-1/2β receptor antagonist
Cholinergic systemAcetylcholinenAChR (α4β2* and α6β2* subtypes)β2* nAChR agonistβ2 subtype reduces LIDs, but nAChR vary over the course of PD
mAChR (m1 to m5)Variable results with muscarinic antagonistsNot established
Opioid systemEnkephalinδδ—receptor selective antagonist- Elevated levels of dynorphin B, α-neoendorphin and Dynorphin A in the dorsolateral striatum and SN
- μ and δ receptors promote LIDs
- κ receptor reduces LIDs
β-endorphin
Endomorphin
μμ—receptor selective antagonist
Dynorphin A
Dynorphin B
α-neoendorphin
β-neoendorphin
κκ—receptor selective agonist
Endocannabinoid systemAnandamide
2-AG
CB1/2CB-receptor agonist- The stimulation of the CB1 receptors reduces LIDs by:
● Desensitization of DA receptors
● Normalizing aberrant glutamate release
- Net anti-dyskinetic effect
- CB1 receptors can also promote LIDs by dopamine synthesis in serotonergic raphe-striatal fibers
TRPNot establishedNot established
PPARNot establishedNot established
Adenosinergic systemAdenosineA2A- A2A receptor antagonist- Not clearly established, but the activation of this receptor in the striatum regulates amplification of dopamine and glutamate release
- A direct anti-dyskinetic effect seems unlikely
A2BNot establishedNot established (poorly expressed in CNS)
A3Not establishedNot established (poorly expressed in CNS)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

di Biase, L.; Pecoraro, P.M.; Carbone, S.P.; Caminiti, M.L.; Di Lazzaro, V. Levodopa-Induced Dyskinesias in Parkinson’s Disease: An Overview on Pathophysiology, Clinical Manifestations, Therapy Management Strategies and Future Directions. J. Clin. Med. 2023, 12, 4427. https://doi.org/10.3390/jcm12134427

AMA Style

di Biase L, Pecoraro PM, Carbone SP, Caminiti ML, Di Lazzaro V. Levodopa-Induced Dyskinesias in Parkinson’s Disease: An Overview on Pathophysiology, Clinical Manifestations, Therapy Management Strategies and Future Directions. Journal of Clinical Medicine. 2023; 12(13):4427. https://doi.org/10.3390/jcm12134427

Chicago/Turabian Style

di Biase, Lazzaro, Pasquale Maria Pecoraro, Simona Paola Carbone, Maria Letizia Caminiti, and Vincenzo Di Lazzaro. 2023. "Levodopa-Induced Dyskinesias in Parkinson’s Disease: An Overview on Pathophysiology, Clinical Manifestations, Therapy Management Strategies and Future Directions" Journal of Clinical Medicine 12, no. 13: 4427. https://doi.org/10.3390/jcm12134427

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