CNS Drugs DOI 10.1007/s40263-014-0205-z

REVIEW ARTICLE

Pharmacological Strategies for the Management of LevodopaInduced Dyskinesia in Patients with Parkinson’s Disease Eva Schaeffer • Andrea Pilotto • Daniela Berg

Ó Springer International Publishing Switzerland 2014

Abstract L-Dopa-induced dyskinesias (LID) are the most common adverse effects of long-term dopaminergic therapy in Parkinson’s disease (PD). However, the exact mechanisms underlying dyskinesia are still unclear. For a long time, nigrostriatal degeneration and pulsatile stimulation of striatal postsynaptic receptors have been highlighted as the key factors for the development of LID. In recent years, PD models have revealed a wide range of non-dopaminergic neurotransmitter systems involved in pre- and postsynaptic changes and thereby contributing to the pathophysiology of LID. In the current review, we focus on therapeutic LID targets, mainly based on agents acting on dopaminergic, glutamatergic, serotoninergic, adrenergic, and cholinergic systems. Despite a large number of clinical trials, currently only amantadine and, to a lesser extent, clozapine are being used as effective strategies in the treatment of LID in clinical settings. Thus, in the second part of the article, we review the placebo-controlled trials on LID treatment in order to disentangle the changing scenario of drug development. Promising results include the extension of L-dopa action without inducing LID of the novel monoamine oxidase B- and glutamate-release inhibitor safinamide; however, this had no obvious effect on existing LID. Others, like the metabotropic E. Schaeffer  A. Pilotto  D. Berg (&) Department of Neurodegeneration, Hertie Institute for Clinical Brain Research, University of Tuebingen, Hoppe Seyler-Strasse 3, 72076 Tu¨bingen, Germany e-mail: [email protected] E. Schaeffer  D. Berg German Center of Neurodegenerative Diseases (DZNE), Tu¨bingen, Germany A. Pilotto Neurology Unit, Centre for Ageing Brain and Neurodegenerative Disorders, University of Brescia, Brescia, Italy

glutamate-receptor antagonist AFQ056, showed promising results in some of the studies; however, confirmation is still lacking. Thus, to date, strategies of continuous dopaminergic stimulation seem the most promising to prevent or ameliorate LID. The success of future therapeutic strategies once moderate to severe LID occur will depend on the translation from preclinical experimental models into clinical practice in a bidirectional process.

Key Points The pathophysiology of L-dopa-induced dyskinesias (LID) involves complex pre- and postsynaptic mechanisms, including a wide range of modulations in dopaminergic and non-dopaminergic transmitter systems. Despite great research efforts and a large numbers of new drugs proposed, only a few trials have shown promising results. These need to be replicated in larger cohorts. For clinical application, amantadine and clozapine are the only effective drugs for LID reduction that are currently available, besides the adjustment of dopaminergic therapy.

1 Introduction L-3,4-Dihydroxyphenylalanine

(L-Dopa), the initial ‘gold standard therapy’ for Parkinson’s disease (PD), is still one of the most effective and widely used therapeutic options in

E. Schaeffer et al.

the treatment of this neurodegenerative disorder. However, its use is still limited by the development of motor fluctuations and L-dopa-induced dyskinesia (LID) [1], which manifests in the majority of patients after varying lengths of treatment. Nearly 40 % of PD patients develop LID after 4–6 years of L-dopa treatment [2]. In recent decades, considerable efforts have been made to understand the underlying pathological mechanisms to either prevent or at least alleviate LID. This review gives an overview on the most important pathomechanisms involved in LID, summarizing pre- and postsynaptic changes and describing the wide range of possibly involved transmitter systems. The second part of the manuscript focuses on the several pharmacological strategies applied to date for the treatment and prevention of LID. In order to point out the most recent and important findings, an extensive electronic search on PubMed and clinicaltrials.gov was conducted up to September 2014 using a combination of the following terms:

‘Parkinson’, ‘Parkinson’s disease’, ‘dyskinesia’, ‘L-dopa’, ‘levodopa’, ‘LID’, ‘levodopa-induced dyskinesia’, ‘dyskinesias’, ‘trial’. Only placebo-controlled randomized trials were included in the description of pharmacological management.

Fig. 1 Pathophysiology of levodopa-induced dyskinesia. Physiological activation, pathological alterations in LID, pathological alterations in LID, modulation, 11 increased activation, 22 decreased activation. A2A adenosine receptor, a2a and a2ab noradrenergic receptors, AMPA a-amino-3-hydroxy5-methyl-4-isoxazolepropionic acid, Cb1 cannabinoid receptor, D1

and D2 dopaminergic receptors, DA dopamine, DA-autoR dopamine autoreceptor, DAT dopamine transporter, GABA c-aminobutyric acid, Glu glutamate, H3 histamine receptor, LID levodopa-induced dyskinesia, mGluR metabotropic glutamate receptor, NA noradrenaline, nAchR nicotinic acetylcholine receptors, NMDA N-methyl-D-aspartate, OP opioid receptor

2 Pathophysiology Major insights into the pathophysiology of dyskinesia have been gained by two specific animal models of PD: the rodent model of the 6-hydroxydopamine (6-OHDA)lesioned rat, and the primate model of the 1-methyl-4phenyl-1,2,3,6-tetrahydropyridine (MPTP)-lesioned monkey. Both models are suitable for the induction and assessment of abnormal involuntary movements (AIMs), which are suggested to correspond to LID in humans [3–7]. Analyses of these two models, together with findings in

Pharmacological Strategies for the Management of Levodopa-Induced Dyskinesia

neurological patients, led to the overall impression that the development of LID is caused by a complex interaction of both pre- and postsynaptic changes, taking place not only in the dopaminergic system but also involving a variety of other neurotransmitters (Fig. 1). In combination, the basic pathophysiology can be seen in the combination of progressive nigrostriatal degeneration and the non-physiologic dopamine receptor stimulation during L-dopa therapy, both leading to a wide range of molecular and biochemical changes in the basal ganglia circuit (reviewed by Cenci and Konradi [8]). 2.1 Nigrostriatal Degeneration Disease severity has been suggested as one of the main risk factors for LID in PD [9]. Thus, it is reasonable to presume that progressive dopaminergic (DA) denervation plays a key role for the manifestation of LID [10, 11]. This presumption is confirmed by the finding that even long-term intake of L-dopa in general does not induce LID in patients without an impairment of the nigrostriatal DA system, for example in patients with restless legs syndrome or doparesponsive dystonia or in healthy individuals [12–14] (some exceptions do exist, please see below). Moreover, some studies have described that the first onset of dyskinesia tends to be on the body side more affected by parkinsonian symptoms, corresponding to the contralateral side of more advanced dopaminergic degeneration in the nigrostriatal system [15–17]. An association of LID development and disease duration has been observed, with a lower threshold for LID in the more advanced stages of PD [18]. This is supported in the MPTP-treated monkey model, in which more severe nigrostriatal DA lesions led to a dramatically lower threshold dose of L-dopa for LID [19, 20]. Regarding the rodent model, AIMs could only be observed in rats with more than 80 % nigrostriatal damage [21], and the implantation of ventral mesencephalic DA grafts reduced AIMs [22]. Disruption of the physiological dopamine release system provides an explanation for the essential role of nigrostriatal denervation for LID development. In the early stages of the disease, exogenously administered L-dopa is transformed into dopamine, stored and released in vesicles into the synaptic cleft by a sufficient number of DA terminals. The continuous release of dopamine is secured additionally by a system of auto-regulation, including dopamine-autoreceptors and the dopamine transporter (DAT) [23, 24]. Progressive loss of functional presynaptic DA neurons in the course of neurodegeneration entails a reduced storage capacity, and the release of dopamine is increasingly dependent on the pharmacokinetics of L-dopa (reviewed Cenci and Lundblad [25]). Subsequently, the administration of L-dopa leads to unphysiologically high extracellular

striatal dopamine levels, which could be demonstrated in dyskinetic rodent models after exogenous L-dopa administration [26, 27]. In positron emission tomography (PET) studies of PD patients, levels of synaptic DA after L-dopa administration correlated positively with more advanced disease stages and with the extent of LID [28, 29]. A recent functional magnetic resonance imaging (fMRI) study in PD patients with LID revealed a markedly increased activation of the putamen bilaterally after exposure to L-dopa, indicating increased metabolism exactly in the part of the brain most affected by DA degeneration [30]. Additionally, degeneration of presynaptic dopaminergic neurons entails the downregulation of DAT levels, which are essential for the homeostasis of synaptic dopamine [31]. The reduced reuptake of dopamine out of the synaptic cleft may first help to improve general PD symptoms; however, oscillating extracellular DA levels are the longterm consequence [32, 33]. Accordingly, Hong et al. [34] recently showed that decreased DAT activity in the putamen was associated with the development of LID in a prospective cohort of PD patients. Taken together, uncontrolled high extracellular dopamine levels in addition to the unphysiological pharmacokinetics of L-dopa lead to a pulsatile stimulation of striatal dopamine receptors, inducing abnormal postsynaptic responses and molecular changes. 2.2 Pharmacokinetics of L-Dopa By the time the degeneration of nigrostriatal neurons becomes more extensive, the release of dopamine into the synaptic cleft is increasingly dependent on the special pharmacokinetics of L-dopa (reviewed by Nutt [35]). The short half-life, variable gastrointestinal absorption, and usually intermittent administration of L-dopa eventually results in an unphysiological, pulsatile stimulation of dopaminergic receptors (reviewed in Olanow et al. [36– 38] and Contin and Martinelli [36–38]). As a consequence, the outflow of the basal ganglia changes, with different signal firing patterns between basal ganglia, thalamus, and cortex inducing abnormal motor responses [37]. It has been observed that another major risk factor for the development of LID is the daily dose of L-dopa [15, 39]. It can be argued that this finding is related to the above-mentioned advanced DA degeneration in patients and their need for higher doses of L-dopa [2], especially as dopamine agonists, which stimulate the postsynaptic receptors by passing the presynaptic neurons, induce less dyskinesia. However, studies in primate models showed an earlier development of LID in primates with L-dopa therapy than in primates with monotherapy of dopamine agonists, independently of the progression of dopaminergic denervation [40, 41]. In randomized, prospective

E. Schaeffer et al.

studies, PD patients receiving dopamine agonist monotherapy developed less LID than the patient group with early levodopa therapy [42, 43], indicating that L-dopa per se has an influence on LID development apart from DA degeneration. These findings hold especially true for the long-lasting agonists, with a markedly longer plasma halflife and therefore more physiological stimulation of striatal DA receptors [2, 42]. Moreover, some forms of continuous or longer-acting dopaminergic drug delivery have been shown to have considerable influence on the manifestation and severity of LID [44–47] (see also Sect. 3). In this respect, it is assumed that some sort of ‘priming’ mechanism, understood as a sensitization of the basal ganglia system, follow the repeated exposure to L-dopa. It is noteworthy that, once LID are established in a patient, they return frequently with increasing severity under the same dosage of L-dopa (reviewed in Brotchie [48, 49] and Nutt [48, 49]). Even after a ‘drug holiday’, patients may react again in a similar way to exposure to L-dopa (reviewed in Nadjar et al. [50]). As an exception to the observation of priming mechanisms, some primate models need to be considered, in which LID appeared after the very first administration of Ldopa [51, 52]. Here, it seems likely that striatal receptors try to compensate for the insufficient and non-physiological DA innervation by inducing a higher sensitivity to DA. This ‘supersensitivity’ is associated with a wide range of postsynaptic maladaptive plastic changes, which could be demonstrated in a variety of studies. Beyond that, there is ongoing debate that the lower incidence of LID under dopamine agonists is not so much due to their long-lasting pharmacokinetics, but rather dependent on their affinity to DA D2 instead of D1 receptors. As further described below, D1 receptors have been attributed an important role in the development of LID. Indeed, the long-lasting dopamine agonists are mostly D2 agonists (such as cabergoline and ropinirole), some dopamine agonists are even partly D1 antagonists (such as bromocriptine) [53–55], while levodopa is non-selective and therefore activates both D1 and D2 receptors. Furthermore, the antidyskinetic effect of clozapine might be due to its D1-antagonistic features [56]. However, this hypothesis alone does not seem to be sufficient. Apomorphine, similar to L-dopa, is a non-selective agonist at the D1 and D2 receptor, but induces far less LID than L-dopa [57]. In particular, the continuous application of apomorphine has been shown to induce less LID than intermittent application via injections in the MPTP primate [45]. Furthermore, the application of a selective D1-dopamine agonist (A-86929) induced less LID than L-dopa [58], contrary to the assumption that the activation of D1 receptors alone leads to LID.

2.3 Postsynaptic Changes Findings in patients with tyrosine hydroxylase deficiency and dopa-responsive dystonia argue for the additional involvement of postsynaptic changes in the pathophysiology of LID. Although nigrostriatal degeneration is not part of the pathophysiological process in these patients, treatment with L-dopa did induce LID in some [59, 60]. Additionally, the fact that dopamine agonists, which act independently of presynaptic terminals, may also induce LID [51] argues for an involvement of the postsynaptic system in the generation of LID. Medium spiny neurons (MSN) of the striatum have been identified as being involved in postsynaptic changes leading to LID [61–64] (reviewed by Iravani and Jenner [65]). Overall, 90 % of striatal neurons are MSN. They act as striatal output neurons, which receive their DA input via two different kinds of DA receptors: D1 and D2. Both receptors are part of the basal ganglia circuit and are stimulated by the DA input of the substantia nigra. As a result of the specific stimulation of D1 or D2 receptors, the GABAergic MSN activate different neuronal pathways, inducing different motor responses: (1) the stimulation of D2 receptors activates the indirect, striatopallidal pathway, which lessens locomotion by using GABA and enkephalin as neurotransmitters and co-transmitters; or (2) the stimulation of D1 receptors activates the direct, striatonigral pathway, which provides locomotion and uses GABA, substance P, and preproenkephalin p [66–68]. According to current studies, it seems that mainly the D1 receptors and therefore the direct, motor-activating pathway contribute to the development of LID [69]. The increased sensitivity of MSN neurons and particularly D1 receptors can be attributed to a multitude of molecular changes, following unphysiological postsynaptic stimulation and leading to maladaptive plasticity. In the rat and the primate model, variances of intracellular receptor trafficking lead to a higher density of postsynaptic D1 receptors at the plasma membrane of striatal MSN [61, 70]. Furthermore, the ‘supersensitivity’ of D1 receptors is induced by changes in signaling pathways, including modifications of second messenger systems and leading to alterations in gene and protein expression. In particular, rodent and primate models showed a decline of cyclic adenosine monosphosphate (cAMP), together with different activation and regulation of the extracellular signalingregulated protein kinases 1 and 2, mitogen-activated protein kinase, mitogen- and stress-activated kinase-1, arrestin, g-protein-coupled receptor kinases, and cAMPregulated phosphoprotein of 32 kDa. Abnormal regulation of transcription factors with different protein expression was demonstrated for, among others, c-Fos, FosB, prodynorphin, and mammalian target of rapamycin complex 1

Pharmacological Strategies for the Management of Levodopa-Induced Dyskinesia

[63, 64, 71–78]. Additionally, the transcriptional activity of striatal D1 receptors was found to be upregulated in dyskinetic rats, with a different messenger RNA (mRNA) expression profile [79]. 2.4 Involvement of Non-Dopaminergic Neurotransmitter Systems 2.4.1 Glutamatergic System A further important aspect of postsynaptic maladaptive changes seems to be the interplay between DA activation of the D1-mediated direct pathway and the corticostriatal glutamatergic input. As central switchpoint of the basal ganglia circuit, the striatum receives neuronal input of the cortex by the excitatory neurotransmitter glutamate. Glutamate excites postsynaptic neurons through ionotropic (iGluR) receptors, particularly N-methyl-D-aspartate (NMDA), a-amino-3hydroxy-5-methyl-4-isoxazolpropionic acid (AMPA), and metabotropic (mGluR) receptors. Thus, an increasing number of studies propose that another key mechanism for the development of LID may lie in an abnormal, supersensitive responsiveness of this glutamatergic corticostriatal input, interacting with DA innervation and thus leading to an overactivity of the direct striatonigral pathway [80– 82]. Glutamatergic and dopaminergic terminals converge at the same MSN synapses in the striatum. Here, DA D1 and glutamatergic NMDA receptors form heteromeric complexes [83]. Fiorentini et al. [84] postulated a downregulation of a special DA1/NMDA complex at corticostriatal synapses, involving both the glutamatergic and the DA system in the development of LID. Moreover, differences in NMDA-receptor trafficking were found to be modulated by D1-receptor activation [85]. Depending on the glutamatergic subtype, activation of corticostriatal synapses induces electrophysiological neuronal changes as long-term depression or long-term potentiation (LTP). Changes in these processes seem to be connected with ‘priming’ mechanisms in LID. Normally, after the stimulation of striatal neuronal inputs, LTP is followed by depotentation. A lack of depotentation after high-frequency stimulation was seen in dyskinetic rodent models and connected to an aberrant activation of D1 receptor-signaling pathways [86, 87]. However, changes in the glutamatergic system were also seen independently from DA innervations. Abnormal receptor trafficking with altered subcellular localization of striatal NMDA- and AMPA-receptor subunits could be demonstrated in the primate and rodent model of PD [88– 91]. Additionally, a plurality of molecular changes in glutamatergic subunits could be shown. Modified phosphorylation of some of these subunits enhanced the

sensitivity of striatal NMDA receptors and therefore altered glutamatergic neurotransmission in dyskinetic rodents [62, 92]. PET studies of PD patients showed higher activation of NMDA-receptors in the ‘on’ state of dyskinetic patients [93]. Higher striatal extracellular levels of glutamate could be demonstrated in rodent models with AIMs [94]. Finally, some studies showed higher binding to NMDA and mGluR receptors of the putamen in patients and monkeys with LID development [82, 95]. Further evidence for the participation of the glutamatergic systems is provided by the effectiveness of the NMDA-antagonist amantadine to alleviate LID, which has been proven in several patient studies [96–99]. Likewise, recent findings suggested a possible antidyskinetic effect by modulators of mGluR [100–102] (see also Sect. 3). 2.4.2 Serotonergic System An increasing number of studies indicate that the serotonergic system might play a key role in the development of LID [103, 104]. Serotonergic neurons contain similar enzymatic and transporter proteins as DA neurons, more precisely, the amino acid aromatic decarboxylase and vesicular monoamine transporter (VMAT)-2. They therefore have the capacity to decarboxylate exogenous administered L-dopa to dopamine and release it to the synaptic cleft [105–107]. The ‘false transmitter’ hypothesis postulates that, in the course of the progressive DA degeneration, serotonergic neurons take over parts of the DA function [108]. Immunohistochemical studies in rodents demonstrated DA reactions in serotonergic neurons after exogenous L-dopa administration [109]. However, since serotonergic neurons lack the mechanisms of autoregulation, dopamine is released in an uncontrolled and irregular manner, adding up to the pulsatile pharmacokinetics of L-dopa and the postulated supersensitivity of postsynaptic DA receptors (reviewed by Carta et al. [110, 111] and Carta and Bezard [110, 111]). Several dyskinesia studies of the rodent model underline this hypothesis. Exogenous L-dopa administered to 6-OHDA-lesioned rats with a completely destroyed nigrostriatal system was shown to still be efficient [112]. Clinical observation in patients in advanced disease stages are in line with the observation that, in spite of extensive presynaptic degeneration, L-dopa is still effective (although only for a short time interval). Thus, metabolization of Ldopa into dopamine under these circumstances needs to be provided by other mechanisms. Indeed, it could be demonstrated that the induction of toxic lesions in the serotonergic system of dyskinetic rats, more precisely in 5-hydroxytryptamin (5-HT) cells of the dorsal raphe nucleus, lowered extracellular dopamine levels after L-dopa administration and attenuated already manifest AIMs [113,

E. Schaeffer et al.

114]. On the other hand, the transplantation of embryonic serotonergic cells in the striatum exacerbated AIMs in the 6-OHDA-lesioned rat [115]. The initial manifestation of LID depended on the severity of DA denervation in these models, which indicates that the ratio of DA degeneration and serotonergic hyperactivity might be the determining factor for LID [22, 103]. As mentioned above, a significant correlation between higher extracellular dopamine levels in the striatum after Ldopa administration and the manifestation of LID has been described. Additionally, involvement of an over-activated serotonergic system could be demonstrated by parallel measurement of higher striatal levels of serotonin and its metabolite 5-hydroxyindoleacetic acid in the rodent model [116]. Moreover, 5-HT1A and 5-HT1B agonists have been seen to be able to reduce extracellular dopamine levels after exogenous administered L-dopa and to attenuate LID in parkinsonian rodent and primate models, as well as in patients [117–120] (see also Sect. 3). However, higher extracellular dopamine levels in correlation with serotonergic activity were seen not only in striatal neurons but also in neurons of the substantia nigra, hippocampus, and prefrontal cortex [121]. In this respect, interesting observations have also been made in patients after transplantation of ventral mesencephalic DA neurons. Those patients showed pronounced offstate dyskinesia that were also responsive to 5-HT1 agonists [122–124]. Post mortem immunohistochemical analyses of rodents, monkeys, and PD patients even indicated that L-dopa therapy may induce morphological changes of serotonergic axons. In the majority of studies, a serotonergic hyperinnervation of the striatum could be demonstrated, and evidence was seen for the induction of serotonergic axonsprouting by L-dopa [125–127]. Combined, these findings underline the assumption of an altered serotonergic function with unphysiological release of dopamine. 2.4.3 Noradrenergic System An important role in the pathophysiology of PD has been identified for the noradrenergic system. The loss of noradrenaline is a well described phenomenon in the development of parkinsonian symptoms (reviewed by Fornai et al. [128]). Several studies revealed an interaction between the DA and noradrenaline systems, leading to changes in motor function in PD patients [129, 130]. One of the most important aggregations of noradrenergic neurons can be found in the locus coeruleus [131], which directly stimulates nigrostriatal DA receptors in the midbrain and influences dopamine release in the basal ganglia [132, 133]. In this process, a2A and a2C receptors in particular have been found to occur in high concentrations in

the striatum [134, 135] and to modulate striatal GABA release [136, 137]. Alachkar et al. [138] showed changes of a2A and a2C mRNA expression in association with L-dopa treatment (see also Sect. 3). However, the role of the noradrenaline system in the development of LID is not yet fully understood, and contradictory results have been found, showing on one hand a reduction and on the other hand an exacerbation of AIMs in noradrenaline-lesioned rat models [139, 140]. 2.4.4 Other Neurotransmitter Systems Apart from the serotonergic, glutamatergic, and noradrenergic systems, a connection to the development of LID is being discussed for a variety of other neurotransmitter systems, which have led to pharmacological intervention with partly positive results (see also Sect. 3). Changes in the opioid system have been detected with different methods. Among others, PET studies have shown reduced striatal and thalamic opioid receptor binding in dyskinetic patients and MPTP monkeys [141, 142]. In the primate model, decreased opioid-derived neurotransmission with reduced opioid receptor binding could also be demonstrated by changes in mRNA levels [143]. It is presumed that stimulation of striatal cannabinoid receptors, particularly CB1, modulates GABAergic and glutamatergic neurotransmission within the basal ganglia circuit [144, 145]. Modifications in cannabinoid receptors therefore maybe connected to changes in striatal synaptic plasticity leading to the development of LID [146]. Furthermore, adenosine A2a receptors have been found to be represented in different areas of the basal ganglia, interacting closely with the DA system (especially D2 receptors) and modulating the striatopallidal pathway [147, 148]. An increase of A2A receptors could be associated with the development of LID in PD patients [149]. Moreover, the ablation of forebrain adenosine A2a receptors has been found to attenuate LID in the rodent model [150]. Another target for new treatment strategies is nicotine, which interacts with striatal pathways via nicotinic acetylcholine receptors (nAChRs). Several studies have shown an influence of nAChR activation on dopamine release [151, 152]. Furthermore, an interaction between striatal nAChR and adenosine A2a receptors was demonstrated recently [153]. Finally, histamine H3 receptors, which are distributed widely in the basal ganglia, have been found to influence glutamatergic and DA projections, as well as striatal GABAergic neurotransmission of the striatonigral and -pallidal pathway [154–156]. Gomez-Ramirez et al. [157] showed an antidyskinetic effect on chorea of histamine H3 agonists in the primate model.

Pharmacological Strategies for the Management of Levodopa-Induced Dyskinesia

2.5 Angiogenesis In the context of an increased D1 receptor stimulation, an induction of L-dopa levels by endothelial proliferation has been observed in the rodent model, which correlated positively with AIMs [158]. The additional impairment of blood–brain barrier integrity in the basal ganglia led to different pharmacokinetics of L-dopa in different areas of the brain and therefore may be an additional reason for higher extracellular dopamine levels in dyskinetic individuals [159]. Moreover, higher levels of vascular endothelial growth factor, together with higher density of striatal capillaries have been shown [160].

3.1.2 Controlled-Release and Rapid-Onset L-Dopa Formulations

3.1 L-Dopa Formulations

Multiple oral doses of standard L-dopa produce pulsatile plasma peaks, which are also presumed to be transferred in this pulsatile way to the striatum, thereby contributing to the development of late motor complications. L-dopa controlled-release formulations have been developed to smooth L-dopa plasma levels and prolong clinical effects due to more continuous dopamine receptor stimulation. Several studies compared these controlled-release formulations with standard L-dopa therapy in the amelioration of motor fluctuations without any conclusive result [165– 170] (see Table 1 for details). Furthermore, several studies showed that, after 5 years of follow-up, slow-release preparations did not show any efficacy in reducing the incidence of motor fluctuations and LID compared with standard L-dopa [171–173]. A number of strategies have been used to improve gastrointestinal absorption of L-dopa in order to minimize its pulsatile plasma peaks. MeL-dopa, a lipophilic L-dopa formulation that is 250 times more soluble than conventional L-dopa, has been evaluated for treatment of motor fluctuations in two different double-blind studies [174, 175]. No differences on total ‘on’ time and LID severity were detected compared with the standard formulation. L-Dopa ethylester, another highly soluble prodrug of L-dopa, was studied to overcome the impaired absorption of regular L-dopa. However, the only study comparing this formulation with standard L-dopa found no statistically significant difference in LID severity and only a small influence on fluctuation control [176].

3.1.1 Standard L-Dopa Formulations

3.1.3 Intraduodenal L-Dopa Formulations

The role of dopaminergic treatment on LID development has been indirectly addressed by several trials. In 2004, the ‘earlier versus later levodopa therapy in Parkinson’s disease study’, which primarily focused on the possible toxic effect of L-dopa on disease progression, obtained the most interesting results [164]. Naive PD subjects were randomized to placebo, 150, 300, or 600 mg L-dopa for 40 weeks, followed by a 2-week washout period. Without providing any solid result on the neuroprotective or toxic effect of L-dopa, a higher incidence of dyskinesia in the 600-mg group (16.5 %), compared with 3.3 % in the placebo and 150-mg/ day groups and 2.3 % in the 300-mg/day group (P \ 0.001), was found. Additionally, high-dose L-dopa was associated with dyskinesias and wearing-off anticipation [164]. This study supported the assumption that the L-dopa dose should be kept as low as possible, balancing between symptomatic effectiveness and motor complications.

To enable continuous levodopa administration and thus stable plasma levels, another strategy has been developed: direct intra-intestinal infusion using a suspension of micronized L-dopa in a methylcellulose gel (DuodopaÒ) and a portable infusion pump. Using this application, constant L-dopa delivery with similar plasma level profiles to intravenous infusions [177] could be demonstrated. A single double-blind, placebo-controlled, crossover study in ten patients with motor fluctuations showed a significantly improved functional ‘on’ time [178]. Several open-label and retrospective studies have also showed LID reduction and fluctuation improvement using intraduodenal duodopa compared with standard oral L-dopa–carbidopa [179–181]. In clinical studies, the most common reason for discontinuing therapy were problems with the gastrostomy or infusion device (such as infection, dislocation, skin irritation, and pain or leakage at the stoma). Thus, duodopa can

2.6 Interhemispheric Connections Recently, new studies in the MPTP-primate model showed some evidence that changes in interhemispheric striatal connections may be included into the pathophysiological concepts of LID [161]. Some primates with unilateral nigrostriatal degeneration of more than 90 % showed no signs of AIMs in the course of L-dopa treatment [162]. Moreover, in clinic, it has been observed that LID seem to only occur in patients with a bilateral affection of the disease [163]. Both observations have led to the hypothesis that, in the course of bilateral nigrostriatal degeneration, interhemispheric nigrostriatal connections get lost, entailing the manifestation of LID.

3 Pharmacological Management

L-dopa

standard

CO 24 weeks L-dopa vs. SR L-dopa (n = 170): increased ON times in SR arm, no differences in global evaluation CO 8 weeks L-dopa vs. SR-L-dopa (n = 103): no LID difference

Wolters et al. [170]

UK Madopar CR SG [306]

3 treated groups vs. PL and pramipexole 6 M (n = 606): improve parkinsonism, increase LID 28 weeks DF (n = 351): increase ON time without dyskinesia, not efficacious on LID already present

Poewe et al. [308]

LeWitt et al. [312]

Rotigotine

See sumanirole

Barone et al. [310]

24 weeks (n = 393): increase ON time without LID, decrease OFF time, decease Ldopa

16 weeks (n = 243): improve parkinsonism, increase LID

Pahwa et al. [311]

Phase III, 6 M (n = 347): decrease OFF time, partial LID reduction

Mizuno et al. [309]

See rotigotine

Poewe et al. [308]

NCT01154166

Phase III, 32 weeks (n = 354): improve parkinsonism, motivation/initiative, increase LID

CO blind PL-controlled (n = 10): improved ON-time 1.1 h, good safety

Moller et al. [307]

Kurth et al. [178]

4 weeks ? first morning and first post-lunch dose change (n = 62): turning on latency reduction (possible increase in ON time), no LID difference

CO 24 weeks L-dopa vs. SR L-dopa (n = 24): no parkinsonism difference, potential benefit on LID

Liebermann et al. [169]

12 ? 4 weeks L-dopa vs. melevodopa (n = 221): no parkinsonism/LID difference

CO 24 weeks L-dopa vs. SR L-dopa (n = 20): increased ON in SR arm; increased LID in open-label phase

Jankovic et al. [168]

Djaldetti et al. [176]

CO 24 weeks L-dopa vs. SR L-dopa (n = 202): no parkinsonism/LID difference

Hutton et al. [167]

Stocchi et al. [175]

CO 24 weeks L-dopa vs. SR L-dopa (n = 25): more pts with improvement in ON in SR L-dopa arm

Sage and Mark [165]

4 ? 8 weeks L-dopa vs. melevodopa afternoon dose (n = 74): no parkinsonism/LID difference

5 years standard L-dopa vs. SR L-dopa (n = 618): no parkinsonism/LID difference

Koller et al. [173]

Stocchi et al. [174]

5 years standard L-dopa vs. SR L-dopa (n = 134): no parkinsonism difference, no effect on LID incidence

40-week dose-response L-dopa study: 150, 300, 600 mg/day vs. PL (n = 361): parkinsonism increase in PL group, increased incidence of LID and fluctuations in L-dopa at high dose

Trial vs. PL: design and results

Dupont et al. [171]

Fahn [163]

Clinical trials

Ropinirole PR

Ropinirole

Pramipexole

Dopamine agonist

Duodenal infusion

Rapid-onset formulation (vs. standard L-dopa, no PL studies available)

Slow-release L-dopa (vs. standard L-dopa, no PL studies available)

L-Dopa

L-Dopa

Drug

Table 1 Placebo-controlled clinical trials for the treatment of levodopa-induced dyskinesia in Parkinson’s disease

Improvement PD: IE, I CMF: E, CU TL: IE, I

No effect/increased LID, parkinsonism

PD: E, CU CMF: E, CU TL: IE, I

No effect/increased LID, parkinsonism improvement

PD: E, CU CMF: E, CU TL: IE, I

No effect/increased LID, parkinsonism improvement

PD: E, CU CMF: E, CU TL: IE, I

Increased LID, parkinsonism improvement

PD: LE, PU CMF: IE, I TL: LE, PU (C)

Good safety SE: local infusion problems (skin irritation, dislocation, infections), improve ON-time no LID effect (seen in retrospective and open-label studies)

PD: IE, I CMF: IE, I TL: IE, I

Possible increased in ON time, no effect on LID

PD: IE, I CMF: IE, I TL: IE, I

SR L-dopa did not decrease LID incidence vs. standard L-dopa no differences on parkinsonism and LID control

High dose at risk of motor fluctuations, no conclusion on neuroprotective L-dopa role

Summary of safety, efficacy and SE

E. Schaeffer et al.

ACR325 (ordopidine)

NMDA antagonist

Tolcapone

Entacapone

COMT inhibitor

Safinamide

Selegiline

Rasagiline

MAO-B inhibitor

See perampanel

Rascol et al. [242]

Phase III, 3 M (n = 202): decrease L-dopa, decrease OFF time, increase LID

13 weeks (n = 341): increase ON time, increase LID

Mizuno et al. [319]

Rajput et al. [322]

See LARGO study, rasagiline

Rascol et al. [214]

Phase III, 6 M (n = 298): decrease L-dopa, mild increase LID

Phase IV, 13 weeks (n = 270): improvement ADL mild increase LID in treated and PL groups

Reichmann et al. [318]

Phase III, DF 3 M (n = 118): decrease OFF time, increase LID

Phase III, 6 M (172 PD with motor fluctuation ? 128 PD): increase mean ON time, Ldopa reduction, increase LID

Brooks and Sagar [317]

Waters et al. [321]

Phase III, DF 3 M (n = 162): L-dopa reduction in treated group, increase LID

Baas et al. [320]

Phase III, 24 weeks (n = 301): ameliorate ADL, mild LID increase

Borgohain et al. [313, 314]

Poewe et al. [315]

Phase III, 6 M (n = 669): increase ON time without dyskinesia Phase III, extension 2 years (n = 544): good safety, increase ON time without LID

NCT01187966

Fenelon et al. [316]

Phase II, 7 weeks DF with escalation (n = 26): study completed, not published Phase III, add-on therapy for MF (n = 549): study completed, not published

NCT01113320

Phase III, rapidly disintegrating tablet formulation (Zidys-selegiline): no parkinsonism/LID difference

Ondo et al. [201]

NCT00627640

Phase III, 12 weeks rapidly disintegrating tablet formulation (Zidys-selegiline) (n = 140): good safety, increase ON time without LID

Phase III, 18 weeks vs. PL vs. entacapone (n = 687): decrease OFF time, increase ON time without dyskinesia, no effect on LID already present

Rascol et al. LARGO Study [214]

Waters et al. [200]

Phase III, 26 weeks (n = 473): increase ON time without LID, decrease OFF time

Phase I, 4 weeks (n = 40): study completed, not published

PSG PRESTO study [196]

NCT01023282

Phase II, 3 M skin patch lisuride vs. L-dopa dose finding (n = 40) LID improvement Phase II–III SC minipump (n = 60): study completed, not published

NCT00089622

NCT00408915

Phase III, 7 ? 12 weeks (n = 294): decrease OFF time increase ON time without LID, high drop-out rate

Rascol et al. [194]

Lisuride

Phase II, 13 weeks vs. pramipexole (n = 30): study completed, not published

NCT00903838

Pardoprunox (SLV308)

3 treated group vs. ropinirole vs. PL 40 weeks (n = 948); improved parkinsonism, no LID difference

Trial vs. PL: design and results

Barone et al. [310]

Clinical trials

Sumanirole

Drug

Table 1 continued

PD:IE, I CMF: E, CU (A) TL: IE, I

No effect/increased LID parkinsonism improvement

PD:NE, NCU CMF: E, CU (A) TL: IE, I

No effect on/increased LID parkinsonism improvement

Good safety

Increased ON time without LID, less effective on already present LID, parkinsonism improvement

PD:NE, NCU CMF: IE, I TL: IE, I

Possible efficacy on LID with Zidys-selegiline formulation

PD:IE, I CMF: E, CU (A) TL: IE, I

Partial efficacy on LID, parkinsonism improvement

Study completed, not published

PD:IE, I CMF: IE, I TL: IE, I

Efficacy on LID, waiting for further results

Possible efficacy on LID parkinsonism improvement

No effect on LID, parkinsonism improvement

Summary of safety, efficacy and SE

Pharmacological Strategies for the Management of Levodopa-Induced Dyskinesia

mGLUr antagonist

NCT00607451

Neu-120

Phase II, 3 weeks (n = 15): no parkinsonism/LID difference

Bara-Jimenez [291]

Nutt et al. [231]

Braz et al. [226]

Riluzole

CP-101.606

Phase II short-term effect on apomorphine-induced dyskinesia(n = 16): mild parkinsonism reduction, no LID difference

Parkinson SG [323]

Remacemide

Phase I–II CO ascending dose administered as a single (n = 20): study completed, not published

Phase I–II dose augmentation study (n = 12): mild dyskinetic effect; dissociation and amnesia as very common side effect

Phase II, double-blind randomized DF study (n = 39): no LID difference

Phase II, CO 2 weeks; study under recruitment

NCT01767129

AVP-923

Phase II, CO study 3 weeks (n = 6): 30–50 % LID reduction vs. PL; confusion at high dose, complaints of decreased erection during DM use

Verhagen Metman et al. [225]

Phase IV, 13-week study to assess possible MEM effect on axial signs of PD (n = 25); effective in LID reduction, not effective on axial symptoms reduction

Moreau et al. [222]

DM

Phase II, CO (n = 12): no LID effects on single-dose L-dopa

Merello et al. [223]

Memantine

Phase III, 24-week study under recruitment Phase III, 16-week study under recruitment

NCT02136914

NCT02153645

Phase II–III DF with augmentation (n = 80): study completed, not published

Phase IV, 3 M wash out study IC: pts with stable AMA therapy, mean 3.4 years (n = 57); good long-term efficacy (develop of LID in wash out to PL)

Ory-Magne et al. [212]

NCT01397422 (Pahtwa 2013)

CO 4 weeks (n = 36): 50 % LID reduction vs. PL

3 weeks IC: patients with stable AMA therapy since 1 year (n = 32); no effect on LID or parkinsonism; good long-term efficacy in selected pts

Wolf et al. [99]

Phase IV, 8 weeks (n = 66): LID reduction vs. PL using different scales

3 weeks (n = 18): AMA decrease LID duration and disability vs. PL

da Silva-Junior et al. [209]

Sawada et al. [210]

12 M (n = 40): 45 % LID reduction in the first month the benefit of AMA lasted \8 months; common rebound effect after withdrawal

Thomas et al. [211]

Goetz et al. [302]

CO 2 weeks (n = 24): 26 % LID reduction vs. PL CO IV AMA infusion study (n = 9): 50 % acute LID reduction vs. PL

Snow et al. [97]

Del Dotto et al. [208]

Study completed, data NA

SE: depersonalization and amnesia

Possible efficacy on LID

Not effective on LID or parkinsonism

Not effective on LID or parkinsonism

Under recruitment

Effective on LID, limitation of use: CYP2D6-dependen kinetic

Possible effective on LID, good tolerability and safety

Under recruitment

One study completed, not published and one study under recruitment

PD:IE, I CMF: IE, I TL: E, CU (A)

Withdrawal effect: LID increase, delirium

SE: hallucinations, cognitive impairment or psychosis, livedo reticularis, pedal edema; common dizziness, and GI symptoms

1-year extension of previous study (n = 17); good long-term efficacy in selected pts CO 2 weeks (n = 11): 50 % LID reduction vs. PL

Metman et al. [98]

Luginger et al. [96]

Summary of safety, efficacy and SE Effective against LIDs, controversy concerning duration of antidyskinetic effect (long effect in selected pts)

Trial vs. PL: design and results CO 3 weeks (n = 18): 60 % LID reduction vs. PL improvement motor fluctuation

Verhagen Metman et al. [207]

Clinical trials

AMA XR

AMA XR (ADS-5102)

AMA

Drug

Table 1 continued

E. Schaeffer et al.

Phase II, 13 weeks DF (n = 197); LID reduction according to dose, significant at high dose vs. PL; possible dose-dependent illusions Phase II, 6 weeks associated with increase L-dopa (n = 23); study completed, data not available Phase II, 12-week fixed-dose (n = 78); study completed, data NA Phase II, 13 weeks AFQ056 modified-release (n = 154); study completed, data NA

Stocchi et al. [100]

NCT01092065

NCT01385592

NCT01491529

5HT agonist

Topiramate

12 weeks DF (n = 347): ameliorate parkinsonism, no LID difference

6 weeks (n = 13): poor tolerability, increase LID Phase II topiramate add therapy to AMA, under recruitment

Kobylecki et al. [248]

NCT01789047

6 ? 4 weeks washout CO (n = 16): no parkinsonism/LID difference

11 weeks DF (250–1,000 mg) (n = 32): no parkinsonism difference, partial LID improvement CO 1,000 mg/day (n = 38): no parkinsonism difference, increase ON time without dyskinesia

Levetiracetam

Wong et al. [250]

Murata et al. [247]

Wolz et al. [251]

Zonisamide

CO (n = 12): no parkinsonism/LID difference CO 4 weeks (n = 20): good tolerability, mild parkinsonism amelioration, no LID difference

Stathis et al. [249]

Price et al. [245]

Phase III, 3 arms vs. PL and vs. entacapone, 6 weeks L-dopa adjustment, 12 weeks maintenance (n = 480): good safety, no effect on motor symptoms or LID

Rascol et al. [242]

Van Blercom et al. [246]

Phase III, DF, 30 weeks vs. PL (n = 751); good safety, no parkinsonism/LID difference

Lees

Sodium valproate

Phase III DF, 20 weeks vs. PL (n = 743); good safety, no parkinsonism/LID difference

Lees et al. [241]

Phase I–II, 3 weeks (n = 22) vs. PL (n = 22): study completed, not published Phase II, DF 12 weeks vs. PL (n = 263); good safety, no parkinsonism/LID difference

NCT00036296

Eggert et al. [324]

Phase II, 3 weeks vs. PL (n = 20): study completed, not published Phase II DF, short-term effect with IV L-dopa (n = 40): study completed, not published

NCT00004576

NCT00108667

Phase II, 4 weeks placebo-controlled trial (n = 83); good safety, LID reduction, improve Parkinsonism, good safety profile (study completed and presented, not yet published)

Phase II, 16 days (severe LID patients) (n = 28): LID reduction vs. PL; one case of psychosis, possible increase dyskinesias at high dose

Berg et al. [101]

NCT01336088

Phase II, 16 days (moderate to severe LID) (n = 31): LID reduction vs. PL; possible increase dyskinesia at high dose

Trial vs. PL: design and results

Berg et al. [101]

Clinical trials

Gabapentin

Anticonvulsants

Perampanel

Talampanel

AMPA antagonist

Dipraglurant (AX48621)

Mavoglurant (AFQ056)

Drug

Table 1 continued

Not effective on LID poor tolerability

Possible efficacy on LID, no effect on parkinsonism

Not effective on LID

Not effective on LID

Not effective on LID

Not effective on LID, good safety

Studies completed, not published

Effective on LID and parkinsonism, good safety

SE: concerns regarding safety for dizziness, hallucinations, and illusions, data NA for the other studies

Effective on LID in the first 3 trials

Summary of safety, efficacy and SE

Pharmacological Strategies for the Management of Levodopa-Induced Dyskinesia

Phase III, DF (n = 381): no parkinsonism/LID difference

Goetz et al. [258]

Phase II, CO 10 weeks IV L-dopa DF (n = 40): study completed, not published

Katzenschlager et al. [183]

NCT00086294

Quetiapine

ACP-103 (pimvanserin)

CO 4 weeks (n = 19): good safety, no efficacy on LID reduction or motor fluctuation

Carroll et al. [280]

Cannabis

Preladenant (SCH420814)

Phase II, 12 weeks DF vs. PL (n = 253): efficacy in improving ‘off’ time, mild LID increase Phase II, 36 weeks open-label extension study (n = 108): reduction in ‘off’ state, LID increase

Hauser et al. [288]

Factor et al. [289]

Adenosine A2A receptor antagonist

CO (n = 7) single peak efficacy: effective on LID reduction, no data on long-term safety

CO (n = 14): L-dopa increase duration of efficacy no efficacy vs. PL on LID reduction

Sieradzan et al. [279]

Fox et al. [275]

CO 2.5 weeks (n = 10): 250–300 mg: small LID reduction on diary, no efficacy vs. PL in objective measures

Manson et al. [274]

Phase I–II, 5 weeks CO DF (n = 27): study completed, not published Phase I, 8 weeks vs. PL DF with escalation (n = 40): study completed, not published

NCT01149811

NCT01140841

CO 4 weeks (group 1 with motor fluctuation n = 10; group 2 with LID n = 8), 100 mg: no parkinsonism/LID difference

Phase II, dose escalating 4 weeks vs. PL 2 cohorts USA (n = 115) and India (n = 64): good safety, LID reduction only in US cohort

Lewitt et al. [271]

Rascol et al. [273]

Phase I, proof-of-concept study, increase dose vs. PL (n = 10): good safety increase ON time 20–30 %, LID reduction

Phase II, CO single-dose effect with dose finding after acute oral challenge of L-dopa (n = 18): possible HR and BP increase, nausea or flushing; small but significant reduction in LID

Rascol et al. [270]

Dimitrova et al. [326]

Phase II, 3 weeks CO vs. PL (n = 8): possible nausea and flushing, no effect on ‘on’ time and LID

Manson et al. [269]

Nabilone

Endocannabinoid agonist

Naloxone

Naltrexone

Opioid antagonist

Fipamezole

Idaxozan

a2 Adrenergic receptor antagonist

30 days CO 25 or 50 mg/day (n = 9): no parkinsonism/LID difference

Durif et al. [264]

Clozapine

10 weeks CO study (n = 50): effective on LID

Phase II (n = 27): study completed, not published

Phase III, 12 weeks (n = 596): no parkinsonism/LID difference

Mu¨ller and Russ (PADDI-II) [325]

NCT00623363

Phase III, 24 weeks (n = 609): no parkinsonism/LID difference

Trial vs. PL: design and results

Rascol et al. (PADDII) [42]

Clinical trials

5-HT antagonist/atypical antipsychotics

Piclozotan

Sarizotane

Drug

Table 1 continued

Increased LID, parkinsonism improvement

Controversial efficacy on LID

No effect on LID, increase L-dopa duration of efficacy

Good safety also at high dose, no effect on LID

Controversial efficacy on LID, good safety, data NA for the other studies

Controversial results for effectiveness and SE (BP and HR increase, nausea or flushing)

Study completed, not published

Not effective on LID

PD:IE, I CMF: IE, I TL: E, CU (C)

Effective on LID

Study completed, not published

Not effective on LID

Summary of safety, efficacy and SE

E. Schaeffer et al.

Phase II, 1 week CO response to IV L-dopa (n = 30): study completed, not published Phase II, DF 12 weeks (n = 420): good safety, improvement parkinsonism, no LID difference

Hauser et al. [334]

Phase III, 12 weeks DF under recruitment

NCT01968031

NCT00605553

Phase III, 12 weeks DF (n = 610): no parkinsonism/LID difference Phase III, 12 weeks DF (n = 373): good safety, reduction OFF time, increase LID

Mizuno and Kondo [332]

Mizuno and Kondo [332]

Pourcher et al. [333]

Phase II, 12 weeks (n = 196): good safety, reduction OFF time, mild LID increase Phase II, 12 weeks DF (n = 363): good safety, reduction OFF time, increase LID

LeWitt et al. [331]

Phase III, 12 weeks (n = 231): good safety reduction OFF time, no LID difference

Phase II, 12 weeks DF (n = 395): good safety, reduction OFF time, increase LID

Stacy et al. [329]

Hauser et al. [330]

Phase II, 12 weeks CO fixed dose (n = 64): study completed, not published Phase III, 16 weeks vs. PL and entacapone (n = 405): study completed, not published

NCT00199394

Phase II, 12 weeks DF (n = 591): good safety, reduction OFF time

LeWitt [328]

NCT00955526

Phase II, 6 weeks PC DF (n = 15): good safety, parkinsonism improvement, no LID difference

NCT01474421

Phase II, 4 weeks DF (n = 71): study completed, not published

Phase I–II, 14 weeks drug: NP002 (n = 65): study completed, not published Phase II, 28 weeks drug: transdermal nicotine (n = 40): study completed, not published

NCT00957918

NCT00873392

Waiting for definitive results

Waiting for definitive results low tolerability in previous studies

Not effective on LID, parkinsonism improvement

Good safety

No effect/increased LID parkinsonism improvement

Summary of safety, efficacy and SE

5-HT 5-hydroxytryptamine, (A) level of evidence A, ADL activities of daily living, AMA amantadine, AMPA a-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor, BP blood pressure, btw between, (C) level of evidence C, CMF control motor fluctuations, CO crossover, COMT catechol-O-methyltransferase, CYP cytochrome P450, DF dose finding or drugs at different dose, DM dextromethorphan, EFNS European Federation of Neurological Societies, GI gastrointestinal, HR heart rate, IC inclusion criteria, IV intravenous, LARGO lasting effect in adjunct therapy with rasagiline given once daily, LID levodopa-induced dyskinesia, MAO monoamine oxidase, MDS Movement Disorder Society, MDS-ES MDS-European section, MEM memantine, MF motor fluctuation, mGLUr metabotropic glutamate receptor, n number of participants (published or presented by http://www.clinicaltrials.gov), NA not available, NMDA N-methyl-D-aspartate, PC proof of concept study, PD preventing dyskinesia, PL placebo, PRESTO Parkinson’s rasagiline efficacy and safety in the treatment of ‘OFF’, PSG Parkinson study group, pts patients, RCT randomized controlled trial, SC subcutaneous, SE side effect, SG study group, SR sustained release, t titration period, TL treatment of LID, XR extended release

Clinical utility: (CU) [clinically useful]: for a given situation, evidence available is sufficient to conclude that the intervention provides clinical benefit. (PU) [possibly useful]: for a given situation, available evidence suggests but insufficient to conclude that the intervention provides clinical benefit. (I) [investigational]: available evidence is insufficient to support the use of the intervention in clinical practice; further study is warranted. (UU) [unlikely useful]: available evidence suggests that the intervention does not provide clinical benefit. (NU) [not useful]: for a given situation, available evidence is sufficient to say that the intervention provides no clinical benefit

Efficacy conclusion: (E) [efficacious]: evidence shows that the intervention has a positive effect on studied outcomes. Supported by data from at least one high-quality (score [75 %) RCT without conflicting level I data. (LE) [likely efficacious]: evidence suggests but is not sufficient to show that the intervention has a positive effect on studied outcomes. Supported by data from any level I trial without conflicting level I data. (UE) [unlikely efficacious]: evidence suggests that the intervention does not have a positive effect on studied outcomes. Supported by data from any level I trial without conflicting level I data. (NE) [non-efficacious]: evidence shows that the intervention does not have a positive side effect on studied outcomes. Supported by data from at least one high-quality (score[75 %) RCT without conflicting level I data. (IE) [insufficient evidence]: there is not enough evidence either for or against efficacy of the intervention in treatment of Parkinson’s disease. All the circumstances not covered by the previous statements

The status was verified on 5 September 2014. The MDS [188] reviewed the clinical evidence and utility in clinical practice of the different drugs to prevent dyskinesia (PD), control motor fluctuations, and treat LID (TL) in Parkinson’s disease pts. When available, the review conclusions have been included in the table. The level of evidence has also been added, according to the current EFNS/MDS-ES recommendation when available [206]

AQW051

Nicotine

Trial vs. PL: design and results

Bara-Jimenez et al. [327]

Clinical trials

Acetylcholine nicotinic receptor agonist

Tozadenant (SYN115)

Istradefylline

Drug

Table 1 continued

Pharmacological Strategies for the Management of Levodopa-Induced Dyskinesia

E. Schaeffer et al.

be regarded as a valid alternative for the treatment of advanced L-dopa-responsive PD when standard oral medication treatment fails. 3.2 Dopamine Agonists 3.2.1 Apomorphine Infusion Apomorphine is a nonergoline DA agonist acting on D2, D3, and D4 receptors, with an additional effect on adrenergic and serotonin 5-HT1A, 5-HT2A, 5-HT2B, and 5-HT2C receptors. Continuous subcutaneous apomorphine infusions are primarily used for the treatment of on–off motor fluctuations in PD patients. Its use leads to a reduction of oral dopamine intake and may, with less fluctuating plasma levels, diminish LID [182]. One prospective and one retrospective study confirmed significant LID reduction in patients with apomorphine infusion [183, 184]. Although placebo-controlled trials are lacking, it may thus be worth considering apomorphine in patients with severe motor complications in an adequate home monitoring situation. Cognitive or psychiatric disturbances (psychosis, hallucinations, confusion, sedation) have been reported as side effects. Moreover, dermatologic complications (skin nodules, panniculitis) are the most common reasons for treatment discontinuation. 3.2.2 Dopamine Agonists Delay L-Dopa-Induced Dyskinesias (LID) but May Exacerbate Dyskinesia

tone and prevent the appearance of motor fluctuations and LID, especially when a mild antiparkinsonian effect is required. Pardoprunox (SLV308), a partial D2 agonist, lead to anti-parkinsonian effects with less dyskinesia in animal models compared with L-dopa [190, 191] and was proposed for monotherapy or add-on therapy in PD patients [192, 193]. In advanced PD with motor fluctuations pardoprunox decreased ‘off’ time and increased ‘on’ time without LID [194]. However, when LID was already present, pardoprunox worsened the symptoms and was poorly tolerated at a higher dose ([25 mg). A smaller study evaluated the efficacy of pardoprunox on motor fluctuations compared with pramipexole (NCT00903838), but results are not yet available. An open-label study compared L-dopa therapy with lisuride, another partial dopamine agonist, which was given as an infusion. After 4 years, lisuride-treated patients improved their baseline dyskinesia scores (measured by AIMs) by 50 % and reduced ‘off’ time [195]. Two further placebo-controlled studies were completed, but data are yet to be published (see Table 1). Dopidines are a new pharmacological class of drugs acting at the same time as partial agonists and antagonists on DA D2 receptors. Based on unpublished preclinical data and phase I results, a clinical study on ACR325 (odopidine), focusing primarily on safety and tolerability and secondly on LIDs, has been initiated. The trial has ended, but no published data are available as yet (NCT01023282). 3.3 Monoamine Oxidase-B Inhibitors

Several placebo-controlled studies have shown that the long-term use of dopamine agonists such as ropinirole, pramipexole, and cabergoline decrease the amount of L-dopa needed and thus lower the risk for LID significantly [43, 185–187]. However, once a patient has established LID, none of the dopamine agonists showed efficacy in reducing dyskinesia despite a significant amelioration of parkinsonian symptoms (see Table 1 for recent studies) [188, 189]. Conversely, a large meta-analysis on placebocontrolled studies demonstrated that the use of dopamine agonists, especially pergolide and ropinirole, increases the threshold-time risk of dyskinesia in patients with motor fluctuations [189]. However, clear guidelines on the reduction of L-dopa in adjunct dopamine agonist therapy are still lacking. 3.2.3 Partial Dopamine Agonists Partial dopamine agonists are pharmacological agents able to occupy dopamine receptors completely (mainly D2 or D3), without producing the maximum pharmacological response, as full agonists do. This might stabilize the DA

Monoamine oxidase (MAO)-B inhibitors are commonly used as initial therapy in early PD and as add-on therapy to L-dopa. MAO-B inhibitors block brain enzymatic metabolism of dopamine, thus increasing its concentration and effectiveness. Rasagiline efficacy on motor fluctuations was tested in two blind controlled studies: against placebo (Parkinson’s rasagiline efficacy and safety in the treatment of ‘off’) and against entacapone and placebo (LARGO [lasting effect in adjunct therapy with rasagiline given once daily]). In both studies, rasagiline increased ‘on’ time and decreased ‘off’ time without a clear effect on LID [196]. Therefore, the Movement Disorder Society (MDS) suggests rasagiline for the treatment of motor complications, but not for dyskinesia treatment (see Table 1) [188]. Selegiline was also tested, with effects similar to rasagiline [197]. Its variable pharmacokinetic profile was ascribed to an extensive first-pass metabolism of the normal formulation, resulting in only 10 % bioavailability of the parent compound [198]. An orally disintegrated tablet (Zydis selegiline) was then developed and proposed as

Pharmacological Strategies for the Management of Levodopa-Induced Dyskinesia

adjunct therapy for PD with motor fluctuations [199]. Zydis selegiline showed an increase in ‘on’ time without LID in the placebo-controlled trial performed by Waters et al. [200], but these results were not replicated in a similarly designed trial [201]. Surprisingly, long-term selegiline treatment was associated with slower motor decline but higher risk of LID development [197], probably due to the increased availability of dopamine at the nigrostriatal synapse caused by the addition of the MAO-B inhibitor [202]. Moreover, it needs to be considered that MAO-B inhibitors are contraindicated in patients taking selective serotonin reuptake inhibitors (SSRIs) or serotonin norepinephrine reuptake inhibitors (SNRIs) and that relevant adverse events in later disease stages, such as confusion, hallucinations, and orthostatic hypotension have been reported. Safinamide, a novel MAO-B and glutamate-release inhibitor reduced LID and ameliorated parkinsonian symptoms in PD animal models [203]. The drug was tested as add-on therapy in early PD patients with promising results [204]. In PD with motor fluctuations, several placebo-controlled studies (see Table 1) evaluated the efficacy of safinamide in terms of LID reduction and control of parkinsonian symptoms. Safinamide showed good tolerability and safety and improved parkinsonism, with increasing ‘on’ time without LID after 2 years of follow-up [205]. 3.4 Catechol-O-Methyltransferase (COMT) Inhibitors The catechol-O-methyltransferase (COMT) inhibitors enhance bioavailability of L-dopa by inhibiting its metabolism. The currently available COMT inhibitors entacapone and tolcapone are added to L-dopa in patients with motor fluctuations. Two extensive meta-analyses [189, 202] and a review by the MDS [188] indicate that both drugs are efficacious in the treatment of motor fluctuations, but ineffective in reducing dyskinesia, as shown in several trials (see Table 1). On the contrary, both drugs, but especially tolcapone, increased the risk for developing or worsening of LID [189]. Therefore, European Federation of Neurological Societies (EFNS) guidelines for PD dyskinesia suggest discontinuation or reduction of COMT (which is the same for MAO-B), when LID develop [206]. However, for both classes of compounds, it needs to be considered that reduction of Ldopa metabolism inevitably leads to higher L-dopa concentrations with a LID-enhancing influence. Thus, clear schemes for L-dopa reduction are necessary for further evaluation of the effect of COMT (and MAO-B) inhibitors on LID.

3.5 Glutamatergic System 3.5.1 Amantadine: A Non-Selective NMDA Receptor Antagonist The NMDA glutamate receptor antagonist amantadine is, to date, the most commonly used drug for the treatment of dyskinesia. It is the sole compound considered as ‘clinically useful’ for the treatment of LID in a recent evidencebased review by the MDS [188]. Indeed, a significant number of clinical trials confirmed the efficacy of amantadine in reducing 30–50 % of LID in PD patients (see Table 1) [96, 97, 207–210], although some trials reported only minor efficacy compared with placebo [99, 209]. On the other hand, the duration of the antidyskinetic effect of amantadine is a topic of some controversy. Thomas et al. [211] reported that the benefit of amantadine in biphasic dyskinesia lasted\8 months, with a mean of\6 months. A much longer effect has been presented in the recent publication of the amantadine for dyskinesia trial, which could show, with class II evidence, that amantadine can retain its antidyskinetic properties over several years [212]. This is in line with one previous trial [98]. However, as the authors pointed out, these two trials evaluated only selected patients, who were treated with amantadine for several years. Both studies thus excluded patients discontinuing the treatment for efficacy loss or occurrence of side effects. The most important side effects reported in clinical studies using amantadine were hallucinations, livedo reticularis, and ankle edema. Given its anti-NMDA and partial anticholinergic profile, amantadine can also worsen psycho-cognitive functions and should be used carefully in patients with Parkinson dementia or psychiatric disturbances. As there is frequently an association between dyskinesia and cognitive impairment in the later stages of PD [213], amantadine still has a limited use in clinical practice to date. For example, in the large cohort of PD patients with motor fluctuations participating in the LARGO trial for rasagiline (n = 687), less than one-third (n = 211) received amantadine at baseline [214]. Furthermore, clinicians should remember the possible rebound effect (on motor and cognitive functions) during the amantadine withdrawal phase [215]. Given its proven efficacy, there is growing interest in understanding whether early use of amantadine may delay or reduce the incidence of LID. Indeed, some animal and indirect evidence suggests a protective role of amantadine on development of LID as well as a possible neuroprotective and antiparkinsonian effect [216, 217]. However, a retrospective study by Jahangirvand and Rajput [218] showed that amantadine treatment prior to use of L-dopa did not influence the incidence or the onset of LID. Thus,

E. Schaeffer et al.

prospective studies are needed to reach a definitive conclusion. ADS-5102 is an oral capsule containing an extendedrelease formulation of amantadine. The use of ADS-5102 once daily is expected to improve safety and tolerability of amantadine via the stabilization of its plasma concentrations during the day and overnight [219]. The final results of a multicenter, randomized, double-blind, placebo-controlled study (NCT01397422) evaluating the tolerability and efficacy of ADS-5102 have not yet been published. 3.5.2 Other Non-Selective NMDA Antagonists Memantine is a potent non-selective NMDA inhibitor. Its efficacy against LID has been proposed based on several observations when added to oral L-dopa therapy in PD patients [220, 221]. A recent placebo-controlled pilot study testing the possible effect of memantine on reduction of axial symptoms in PD patients with motor fluctuations showed a slight reduction of LID when compared with placebo [222]. However, the only placebo-controlled study, which specifically focused on the antidyskinetic effect of memantine, was a small crossover study in 12 patients with a single-dose L-dopa challenge, which could not show a potential benefit [223]. Compared with other anti-NMDA agents, memantine has a good tolerability and safety profile and has a potentially beneficial effect on cognition in PD patients with dementia [224]. Therefore, if a positive effect on LID is proven, the use of memantine may also be extended to those patients who cannot benefit from amantadine because of the presence of cognitive impairment or hallucinations. Thus, larger clinical trials, and especially a clinical trial evaluating the efficacy of memantine on LID in PD patients with different cognitive profiles, would be interesting. Dextromethorphan (DM), a widely used and well tolerated anti-tussive, is a relatively low-affinity, non-competitive NMDA receptor antagonist. In a single doubleblind study conducted by Verhagen Metman et al. [225], use of DM reduced LID about 30–50 % compared with placebo without affecting the L-dopa effect. DM is extensively metabolized by the hepatic cytochrome P450 (CYP)2D6, and its plasma levels are strictly dependent on CYP2D6 activation. Therefore, the effect of DM increases with the co-administration of a CYP2D6 inhibitor, like quinidine. A clinical trial to evaluate the efficacy, safety, and tolerability of AVP-923 capsules (containing 45 mg DM and 10 mg quinidine) versus placebo for the treatment of LID is currently recruiting participants (NCT01767129). Other non-selective NMDA antagonists have been evaluated for the treatment of LID: remacemide (a noncompetitive NMDA-receptor antagonist) and riluzole (a drug with several effects and a non-competitive NMDA

inhibitor) did not show any antidyskinetic effects in double-blind placebo-controlled trials [226]. 3.5.3 Selective NMDA Antagonists Recently, a number of studies in rodent and primate models of PD tested selective NMDA antagonists in the treatment of LID with promising results [227–229]. NMDA receptors contain four subunits, two NR1 subunits, for which there are seven isoforms and two NR2 subunits with four variants (A–D). NMDA receptors with different subunit compositions localize in the nervous system in regionally specific patterns. The highest concentration of NMDA receptors with NR2-B subunits is found in the motor cortical regions, hippocampus and the striatum [230]. Specific NMDA subunit antagonists may offer selective therapeutic benefit on LID without the cognitive and psychiatric adverse effects encountered with non-selective glutamate antagonists. The NMDA-subunit NR2B-specific inhibitor CP-101.606 (traxoprodil) has been tested in a randomized, double-blind, placebo-controlled clinical trial. Besides a mild antidyskinetic effect without improvement of parkinsonism, tolerability issues have been raised, since doserelated amnesia and depersonalization have been reported [231]. Neu-120 is a selective non-competitive NMDA-receptor modulator with MAO-B and glycogen synthase kinase-3b inhibition properties in vitro. After positive unpublished pre-clinical and early clinical phases, Neu-120 has been tested for LID in a double-blind clinical study (NCT00607451). Results have not yet been published. 3.5.4 Metabotropic Glutamate-Receptor Antagonist mGluRs are found abundantly in the basal ganglia. They regulate neuronal excitability and synaptic function and have been shown to be involved in the pathophysiology of LID. Animal dyskinesia models have shown a great variety of changes in all mGluRs, including the mGlu2/3 receptor [232]. However, most consistently, a distinct increase in the density of the postsynaptic mGluR5 was found [233]. Reduction of glutamatergic transmission by a wider range of selective mGluR5 antagonists (MPEP, MTEP, fenobam, and AFQ-056) was linked to a significant attenuation of LID in rodent and monkey models of PD [203, 234–236]. Antidyskinetic actions are thought to reflect blockade of abundant mGlu5 receptors on projection neurons of the direct striatal output pathway [236]. The antidyskinetic efficacy of AFQ056 in clinical practice has been tested in three different double-blind, randomized clinical trials in patients with moderate to severe LID. In the first two studies, AFQ056 successfully alleviated established LID without worsening other parkinsonian features [101]. A

Pharmacological Strategies for the Management of Levodopa-Induced Dyskinesia

larger trial tested AFQ056 in several doses versus placebo and confirmed that 200 mg AFQ056 daily had a significant antidyskinetic effect [100]. Mild side effects, such as fatigue, gastrointestinal disorders, and, most commonly, dizziness, were reported in all studies; the most common serious adverse event was worsening of dyskinesia after discontinuation of treatment. A small percentage of patients also developed delusions or transient psychiatric psychoses. Given these promising results, three more randomized trials and two open-label studies were conducted to expand the early observations and test the efficacy of a modified-release drug (see Table 1 for further details). All three placebo-controlled studies have been completed; however, the final results have not yet been published. Dipraglurant (AX48621), a negative allosteric modulator of mGluR5, has been recently tested in a randomized, double-blind, placebo-controlled clinical study. In this first experience, dipraglurant improved both dyskinesia and parkinsonian symptoms, reducing daily ‘off’ time. No safety or tolerability issues have been raised (NCT01336088) [237].

show any effect on LID reduction or parkinsonism improvement [245–248]. Levetiracetam, a promising drug within this class of agents, showed variable and very mild antidyskinetic effects in three different placebo-controlled trials, even with a wide range of doses [249–251].

3.5.5 AMPA Antagonists

Animal models have suggested that 5-HT1A-receptor agonists may interrupt the dopamine release of dorsal raphe neurons and reduce striatal glutamate levels [253]. In the clinical setting, different open-label trials with 5-HT1Areceptor agonists have shown a paradoxical worsening of parkinsonism instead of an amelioration of LID [254, 255]. Sarizotan, a full 5-HT1A-receptor agonist with high D3 and D4 receptor affinity, which initially showed promising antidyskinetic effects [254], failed to improve dyskinesia in two later double-blinded, placebo-controlled trials [256– 258]. Piclozotan, a new partial 5-HT1A-receptor agonist, significantly reduced hyperkinesias and improved motor complications in a rat model of advanced PD [259]. A small phase II trial testing the adverse effects and pharmacokinetic data of the drug has been recently completed (NCT00623363, see Table 1); however, data are not yet available. Specific stimulation of 5-HT1A receptors, or combined subthreshold doses of 5-HT1A and 5-HT1B-receptor agonists, has been suggested to treat dyskinesia without exacerbating parkinsonism. In the MPTP-lesioned macaque, this approach was more effective on LID than stimulation of each receptor alone [119]. Eltoprazine, a mixed 5-HT1B/1B-receptor agonist had promising results on LID in different animal models; thus, further clinical studies are warranted [260].

The glutamatergic AMPA receptor has been proposed to contribute to LID expression [232]. AMPA antagonists potentiate the antiparkinsonian effects of L-dopa and appear effective against LID in animal models [238, 239]. Several clinical studies (see Table 1) evaluated the effect of talampanel, an AMPA receptor antagonist in PD and LID, but no data have yet been published. Perampanel, a potent selective non-competitive AMPA antagonist, also reduced motor symptoms and LID in rats [240] and prompted examination of the antidyskinetic potential of this compound in patients. Disappointingly, four double-blinded, placebo-controlled trials, involving more than 1,500 subjects, failed to show any significant reduction in LID, besides the good safety and tolerability of this drug [241, 242]. Given the importance of calciumdependent signaling pathways in the pathogenesis of LID [243], it will be interesting to see whether or not selective calcium-permeable AMPA receptor antagonists like IEM1460, which are efficacious in animal models of LID [244], open up a more promising line of investigation. 3.5.6 Anticonvulsants Anticonvulsants have been considered as potential LID therapy for their desynchronizing properties on basal ganglia circuitry and for their secondary effects on neurotransmitters (particularly in dopamine- and NMDAdependent systems). However, valproate, gabapentin, zonisamide, and—recently—topiramate have failed to

3.6 Serotonergic System The serotonergic system modulates the two neurotransmitter systems with the greatest influence on the dyskinetic phenotype, the dopaminergic and glutamatergic systems. Following degeneration of the nigrostriatal system, serotonergic terminals of dorsal raphe neurons are capable of releasing dopamine, altering the normal autoregulatory mechanisms, and probably contributing to LID manifestation [252]. Both serotonergic 5-HT1A and 5-HT1B have been studied extensively in LID, identifying the anatomical sites of action and pharmacological mechanisms underlying a potential antidyskinetic effect. 3.6.1 5-HT1A and 5-HT1B Agonists

3.6.2 5-HT2A Antagonists Post mortem and pharmacological studies have provided indirect evidence of altered 5-HT22A-mediated

E. Schaeffer et al.

neurotransmission in the dyskinetic state. Antagonizing 5-HT2A receptors with the atypical antipsychotics clozapine and quetiapine reduced LID severity in rat and primate PD models [3, 261]. After some reports and open-label studies evaluating different 5-HT2A-receptor antagonists [262], several placebo-controlled trials focusing on LID control were performed. Quetiapine was not effective in LID reduction at a dose of 25 and 50 mg in a small crossover placebo-controlled trial [263]. On the other hand, in the double-blind placebo-controlled trial published by Durif et al. [264], clozapine increased the mean ‘on’ time without dyskinesia, with no changes in duration of ‘off’ time. Hence, the MDS evidence-based review indicates clozapine as ‘‘efficacious and possibly useful’’ in dyskinesia treatment [188]. Furthermore, it might be useful in patients with hallucinations. However, the use of clozapine in the treatment of LID is still limited in clinical practice because of its side effects, especially agranulocytosis. Several other specific 5-HT2A-receptor antagonists have been studied in animal models and pimvanserin (ACP-103) has also been tested in a small trial; however, results are not yet published (see Table 1). 3.7 a2-Adrenergic-Receptor Antagonists and Adrenergic Transmission a and b-adrenergic receptors appear to play an important role in the expression of a dyskinetic phenotype [265]. Most studies have focused on a2-adrenergic-receptors, abundantly expressed on GABAergic striatal neurons. While a2-receptor blockade appears to effectively alleviate LID without impairing L-dopa antiparkinsonian action, stimulation of a2 receptors also seems effective in reducing LID, although with deleterious effects on L-dopa antiparkinsonian benefit. The mechanisms underlying the antidyskinetic action of a2-receptor antagonists remain hypothetical: effect on striatal cell bodies was assumed, reducing excitability, or on terminals of the direct pathway in the output regions of the basal ganglia, reducing GABA release [266]. Although the a2-receptor antagonists idazoxan and fipamezole significantly reduced LID in animal models [267, 268], efficacy of these compounds is still controversial in the clinical setting (see Table 1). Two small clinical trials on idazoxan found, at the same dosage, different safety profiles and only partial efficacy, if any, on LID reduction [269, 270]. Fipamezole was tested in a large double-blind, randomized, placebo-controlled, dose-escalating study in patients from the USA and India and was found to be effective only in the American subpopulation [271]. Two other studies evaluating safety, tolerability, and the maximum tolerated dosage in L-dopa PD patients have been

conducted (NCT01140841, NCT01149811), but the findings have not yet been published. 3.8 Opioid Antagonists Opioid peptide transmission is enhanced in the striatum of animal models and PD patients with L-dopa-induced motor complications. Opioid receptor antagonists reduce LID in primate PD models [272]. Naltrexone, despite a good safety profile, was not associated with any significant change in motor function or in LID reduction in two different studies comparing low (100 mg) and high (250–300 mg) doses versus placebo [273, 274]. In 2004, the non-subtype-selective opioid receptor antagonist naloxone failed to reduce LID in a placebo-controlled study on 14 PD patients with motor fluctuations [275]. 3.9 Cannabinoid Antagonists Animal models have suggested that cannabinoid agonists may exert an antidyskinetic function by augmenting GABAergic transmission in the indirect pathway [276, 277]. This model is supported clinically by the finding that 14 % of PD patients report improvement of dyskinesia with cannabis use [278]. In a small randomized trial, nabilone was associated with a 22 % mean reduction of ‘on’-period LID compared with placebo [279]. However, only seven patients completed this study, and nabilone was administered only twice prior to an acute L-dopa challenge. In 2004, a 4-week dose-escalation cross-over placebo-controlled study evaluated the potential benefit of oral cannabis in 19 PD patients [280]. The drug administration was well tolerated, but no objective or subjective improvement on dyskinesias or parkinsonism was observed. 3.10 Adenosine A2A-Receptor Antagonists Adenosine is a purinergic neurotransmitter that exerts its effects via four types of receptors: A1, A2A, A2B, A3 [281]. In PD, attention was drawn to A2A receptors, localized in the caudate and accumbens nuclei, globus pallidus, and olfactory tubercles. A2A receptors on striatal neurons participate in the planning and execution of movements, probably by influencing D2- and D1-receptor activity, and are part of the indirect pathway [282, 283]. A2A-adenosine receptors can also modulate glutamatergic and GABAergic synaptic transmission by increasing the excitability of MSN in the striatum and by enhancing GABA inhibition in projection neurons of the globus pallidus [284]. Via regulation of specific neuropeptides, such as dynorphin and encephalin, the A2A-receptor antagonists modulate the effects of chronic L-dopa administration in synaptic plasticity and contribute to the expression of LID [285]. Thus,

Pharmacological Strategies for the Management of Levodopa-Induced Dyskinesia

regarding A2A, it may be supposed that receptor antagonists may modulate striatopallidal output balance and alleviate parkinsonian symptoms by preventing the excessive activity of striatopallidal neurons and restoring the D1/ D2 (direct/indirect) pathway equilibrium [284, 286]. Preladent, a potent and selective competitive A2A-receptor antagonist showed promising results in animal models of PD [287]. In two different phase II double-blind randomized studies, preladent showed efficacy on improvement of parkinsonism, but without any beneficial effect on LID. Conversely, in both studies, the treated group showed a mild increase in dyskinesia rates [288, 289]. Istradefylline, another adenosine A2A-receptor antagonist, was tested in numerous phase II–III trials and proved to be mildly effective in relieving wearing-off fluctuations in PD patients (see Table 1). However, the addition of the drug to L-dopa was associated with a slight increase of dyskinesia in ‘on’ time and dyskinesia as an adverse event was more commonly reported in the istradefylline groups. Tozadenant (SYN115), a selective A2a-receptor antagonist was recently tested in a phase II double-blind, randomized, placebo-controlled study. Interestingly, the treated patients reported a significant improvement in parkinsonism without any increase in LID [290]. In conclusion, A2A-receptor antagonists may be a promising target in patients with motor fluctuations. Despite the mild increase of LID observed in clinical trials, it may be interesting to again test the possible antidyskinetic effect, after lowering L-dopa once parkinsonian symptoms are ameliorated by the compound, as suggested in earlier studies [291]. 3.11 Nicotinic Acetylcholine-Receptor Agonists Nicotinic receptors co-localize with dopamine receptors and seem to be involved in the expression and modulation of LID [292]. There is growing evidence of improvement of LID with the use of different acetylcholine-receptor agonists in animal models [293–295]. In the clinical setting, nicotinicreceptor agonists have been tested as possible add-on therapy to improve parkinsonian symptoms. However, all studies showed low tolerability and no clear efficacy as an add-on therapy [296]. The antidyskinetic effect of nicotine was tested in two clinical trials, with oral capsule and transdermal patch. Another nicotin receptor partial agonist, namely AQW051, was tested in a Phase 2 double blind study. All studies are completed but no data have been published, yet.

4 Trial Limitations and Future Perspectives Despite extensive efforts in LID research, there remains a substantial difference between promising preclinical results

and the failure or inconsistent results of clinical trials. There are no easy answers to this issue. However, it is important to consider limitations in translation from animal studies into the clinic and study design as two major aspects. Most PD animal models are based mainly on monolateral denervation of the DA system [297, 298]. In contrast, patients with LID show a bilateral and complex pattern of neurodegeneration, involving also non-DA systems, as mentioned above. Moreover, age, adaptive processes, and L-dopa doses differ substantially between animals and PD patients. Furthermore, while a placebo effect is unlikely to play a role in experimental LID, it is particularly important in PD patients, especially in elderly subjects [299]. Moreover, to date, the role of genetics in LID modulation has not been clarified [300]. Whereas animal models can be based on homogenous study groups, it is well known that PD in humans is heterogeneous. This holds true not only for the pathophysiology of the development of neurodegeneration but also for the development of dyskinesia [301]. A compound effective in one person may thus not influence LID to the same extent in another. With regard to study design, it needs to be considered that assessment of dyskinesia remains a major challenge in clinical trials. To date the Unified Dyskinesia Rating Scale is considered the most sensitive tool to detect treatment effect, as it combines patient-based medical history and clinical evaluation of disability by a physician [302]. However, a single assessment during clinical evaluation offers insufficient insight into the severity and manifestation forms of dyskinesia in daily life. On the other hand, patients may be unaware of the severity of their LID, probably because of the complex interplay involving anosognosic and proprioceptive mechanisms [303]. Even if they realize their LID, most patients judge akinesia as worse than dyskinesia. Thus, amelioration of dyskinesia may not be weighted adequately by patients. Hence, both the physician’s assessment and the perception of the patient are subjective. Quantitative, objective techniques for dyskinesia assessment, for example as a wearable, ambulatory assessment, need to be further developed as a very promising approach and should be considered in the design of future trials [304, 305]. Moreover, compounds used as levodopa add-on therapy that do lead to an improvement of parkinsonism complicate the scenario in patients; however, often at the cost of worsening of dyskinesia (see Table 1). Indeed, patients with motor fluctuations and LID require antithetical multimodal therapy to balance these symptoms. A personalized approach with regard to patient stratification and assessment thus seems to be the future of LID therapy. In the meantime, research on promising compounds like safinamide and AFQ056 needs to be further

E. Schaeffer et al.

substantiated. Also, the extended-release formulation of amantadine is awaited with interest. Moreover, in the absence of clearly effective compounds, research on easily administered continuous (low-dose) L-dopa medication will need to be continued.

5 Conclusion Preclinical research has explored the wide spectrum of mechanisms involved in LID that may be effectively targeted in therapeutic strategies. However, many of the promising compounds developed in experimental models have failed to reproduce efficacy in clinical settings. At the present, according to evidence-based indications, amantadine and clozapine should be used in LID treatment. Nevertheless, data from as yet unpublished studies, and further understanding of the underlying pathophysiology leading to new translational approaches, may contribute to a more effective LID therapy in the future. Disclosure The authors Eva Schaeffer, Andrea Pilotto, and Daniela Berg report no conflicts of interest and no funding.

10.

11.

12.

13.

14.

15.

16.

17.

18.

References 19. 1. Warren Olanow C, Kieburtz K, Rascol O, Poewe W, Schapira AH, Emre M, et al. Factors predictive of the development of levodopa-induced dyskinesia and wearing-off in Parkinson’s disease. Mov Disord. 2013;28(8):1064–71. doi:10.1002/mds. 25364. 2. Ahlskog JE, Muenter MD. Frequency of levodopa-related dyskinesias and motor fluctuations as estimated from the cumulative literature. Mov Disord. 2001;16(3):448–58. 3. Lundblad M, Andersson M, Winkler C, Kirik D, Wierup N, Cenci MA. Pharmacological validation of behavioural measures of akinesia and dyskinesia in a rat model of Parkinson’s disease. Eur J Neurosci. 2002;15(1):120–32. 4. Lundblad M, Picconi B, Lindgren H, Cenci MA. A model of Ldopa-induced dyskinesia in 6-hydroxydopamine lesioned mice: relation to motor and cellular parameters of nigrostriatal function. Neurobiol Dis. 2004;16(1):110–23. doi:10.1016/j.nbd. 2004.01.007. 5. Cenci MA, Ohlin KE. Rodent models of treatment-induced motor complications in Parkinson’s disease. Parkinsonism Relat Disord. 2009;15(Suppl 4):S13–7. doi:10.1016/S1353-8020(09) 70828-4. 6. Jenner P. From the MPTP-treated primate to the treatment of motor complications in Parkinson’s disease. Parkinsonism Relat Disord. 2009;15(Suppl 4):S18–23. doi:10.1016/S1353-8020(09) 70829-6. 7. Boyce S, Rupniak NM, Steventon MJ, Iversen SD. Characterisation of dyskinesias induced by L-dopa in MPTP-treated squirrel monkeys. Psychopharmacology. 1990;102(1):21–7. 8. Cenci MA, Konradi C. Maladaptive striatal plasticity in L-dopainduced dyskinesia. Prog Brain Res. 2010;183:209–33. doi:10. 1016/S0079-6123(10)83011-0. 9. Hauser RA, McDermott MP, Messing S. Factors associated with the development of motor fluctuations and dyskinesias in

20.

21.

22.

23. 24.

25.

26.

27.

Parkinson disease. Arch Neurol. 2006;63(12):1756–60. doi:10. 1001/archneur.63.12.1756. Boyce S, Rupniak NM, Steventon MJ, Iversen SD. Nigrostriatal damage is required for induction of dyskinesias by L-dopa in squirrel monkeys. Clin Neuropharmacol. 1990;13(5):448–58. Horstink MW, Zijlmans JC, Pasman JW, Berger HJ, van’t Hof MA. Severity of Parkinson’s disease is a risk factor for peakdose dyskinesia. J Neurol Neurosurg Psychiatry. 1990;53(3): 224–6. Jankovic J, Rajput AH, McDermott MP, Perl DP. The evolution of diagnosis in early Parkinson disease. Parkinson Study Group. Arch Neurol. 2000;57(3):369–72. Becker PM, Jamieson AO, Brown WD. Dopaminergic agents in restless legs syndrome and periodic limb movements of sleep: response and complications of extended treatment in 49 cases. Sleep. 1993;16(8):713–6. Rajput AH, Fenton M, Birdi S, Macaulay R. Is levodopa toxic to human substantia nigra? Mov Disord. 1997;12(5):634–8. doi:10. 1002/mds.870120503. Grandas F, Galiano ML, Tabernero C. Risk factors for levodopa-induced dyskinesias in Parkinson’s disease. J Neurol. 1999;246(12):1127–33. Fabbrini G, Defazio G, Colosimo C, Suppa A, Bloise M, Berardelli A. Onset and spread of dyskinesias and motor symptoms in Parkinson’s disease. Mov Disord. 2009;24(14):2091–6. doi:10.1002/mds.22703. Mones RJ, Elizan TS, Siegel GJ. Analysis of L-dopa induced dyskinesias in 51 patients with Parkinsonism. J Neurol Neurosurg Psychiatry. 1971;34(6):668–73. Mouradian MM, Heuser IJ, Baronti F, Fabbrini G, Juncos JL, Chase TN. Pathogenesis of dyskinesias in Parkinson’s disease. Ann Neurol. 1989;25(5):523–6. doi:10.1002/ana.410250521. Di Monte DA, McCormack A, Petzinger G, Janson AM, Quik M, Langston WJ. Relationship among nigrostriatal denervation, parkinsonism, and dyskinesias in the MPTP primate model. Mov Disord. 2000;15(3):459–66. Schneider JS. Levodopa-induced dyskinesias in parkinsonian monkeys: relationship to extent of nigrostriatal damage. Pharmacol Biochem Behav. 1989;34(1):193–6. Winkler C, Kirik D, Bjorklund A, Cenci MA. L-dopa-induced dyskinesia in the intrastriatal 6-hydroxydopamine model of Parkinson’s disease: relation to motor and cellular parameters of nigrostriatal function. Neurobiol Dis. 2002;10(2):165–86. Carlsson T, Carta M, Munoz A, Mattsson B, Winkler C, Kirik D, et al. Impact of grafted serotonin and dopamine neurons on development of L-dopa-induced dyskinesias in parkinsonian rats is determined by the extent of dopamine neuron degeneration. Brain J Neurol. 2009;132(Pt 2):319–35. doi:10.1093/brain/ awn305. Cragg SJ, Rice ME. Dancing past the DAT at a DA synapse. Trends Neurosci. 2004;27(5):270–7. doi:10.1016/j.tins.2004.03.011. Miller DW, Abercrombie ED. Role of high-affinity dopamine uptake and impulse activity in the appearance of extracellular dopamine in striatum after administration of exogenous L-dopa: studies in intact and 6-hydroxydopamine-treated rats. J Neurochem. 1999;72(4):1516–22. Cenci MA, Lundblad M. Post-versus presynaptic plasticity in Ldopa-induced dyskinesia. J Neurochem. 2006;99(2):381–92. doi:10.1111/j.1471-4159.2006.04124.x. Abercrombie ED, Bonatz AE, Zigmond MJ. Effects of L-dopa on extracellular dopamine in striatum of normal and 6-hydroxydopamine-treated rats. Brain Res. 1990;525(1):36–44. Carta M, Lindgren HS, Lundblad M, Stancampiano R, Fadda F, Cenci MA. Role of striatal L-dopa in the production of dyskinesia in 6-hydroxydopamine lesioned rats. J Neurochem. 2006;96(6):1718–27. doi:10.1111/j.1471-4159.2006.03696.x.

Pharmacological Strategies for the Management of Levodopa-Induced Dyskinesia 28. Pavese N, Evans AH, Tai YF, Hotton G, Brooks DJ, Lees AJ, et al. Clinical correlates of levodopa-induced dopamine release in Parkinson disease: a PET study. Neurology. 2006;67(9): 1612–7. doi:10.1212/01.wnl.0000242888.30755.5d. 29. de la Fuente-Fernandez R, Sossi V, Huang Z, Furtado S, Lu JQ, Calne DB, et al. Levodopa-induced changes in synaptic dopamine levels increase with progression of Parkinson’s disease: implications for dyskinesias. Brain J Neurol. 2004;127(Pt 12):2747–54. doi:10.1093/brain/awh290. 30. Herz DM, Haagensen BN, Christensen MS, Madsen KH, Rowe JB, Lokkegaard A, et al. The acute brain response to levodopa heralds dyskinesias in Parkinson disease. Ann Neurol. 2014;75(6):829–36. doi:10.1002/ana.24138. 31. Raevskii KS, Gainetdinov RR, Budygin EA, Mannisto P, Wightman M. Dopaminergic transmission in the rat striatum in vivo in conditions of pharmacological modulation. Neurosci Behav Physiol. 2002;32(2):183–8. 32. Sossi V, Dinelle K, Topping GJ, Holden JE, Doudet D, Schulzer M, et al. Dopamine transporter relation to levodopa-derived synaptic dopamine in a rat model of Parkinson’s: an in vivo imaging study. J Neurochem. 2009;109(1):85–92. doi:10.1111/j. 1471-4159.2009.05904.x. 33. Troiano AR, de la Fuente-Fernandez R, Sossi V, Schulzer M, Mak E, Ruth TJ, et al. PET demonstrates reduced dopamine transporter expression in PD with dyskinesias. Neurology. 2009;72(14):1211–6. doi:10.1212/01.wnl.0000338631.73211. 56. 34. Hong JY, Oh JS, Lee I, Sunwoo MK, Ham JH, Lee JE, et al. Presynaptic dopamine depletion predicts levodopa-induced dyskinesia in de novo Parkinson disease. Neurology. 2014;82(18):1597–604. doi:10.1212/WNL.0000000000000385. 35. Nutt JG. Pharmacokinetics and pharmacodynamics of levodopa. Mov Disord. 2008;23(Suppl 3):S580–4. doi:10.1002/mds.22037. 36. Olanow CW, Obeso JA, Stocchi F. Continuous dopaminereceptor treatment of Parkinson’s disease: scientific rationale and clinical implications. Lancet Neurol. 2006;5(8):677–87. doi:10.1016/S1474-4422(06)70521-X. 37. Olanow CW, Obeso JA, Stocchi F. Drug insight: continuous dopaminergic stimulation in the treatment of Parkinson’s disease. Nat Clin Pract Neurol. 2006;2(7):382–92. doi:10.1038/ ncpneuro0222. 38. Contin M, Martinelli P. Pharmacokinetics of levodopa. J Neurol. 2010;257(Suppl 2):S253–61. doi:10.1007/s00415-010-5728-8. 39. Schrag A, Quinn N. Dyskinesias and motor fluctuations in Parkinson’s disease. A community-based study. Brain J Neurol. 2000;123(Pt 11):2297–305. 40. Jenner P. Factors influencing the onset and persistence of dyskinesia in MPTP-treated primates. Ann Neurol. 2000;47(4 Suppl 1):S90–9. (Discussion S9–104). 41. Smith LA, Tel BC, Jackson MJ, Hansard MJ, Braceras R, Bonhomme C, et al. Repeated administration of piribedil induces less dyskinesia than L-dopa in MPTP-treated common marmosets: a behavioural and biochemical investigation. Mov Disord. 2002;17(5):887–901. doi:10.1002/mds.10200. 42. Rascol O, Brooks DJ, Korczyn AD, De Deyn PP, Clarke CE, Lang AE, et al. Development of dyskinesias in a 5-year trial of ropinirole and L-dopa. Mov Disord. 2006;21(11):1844–50. doi:10.1002/mds.20988. 43. Holloway RG, Shoulson I, Fahn S, Kieburtz K, Lang A, Marek K, et al. Pramipexole vs levodopa as initial treatment for Parkinson disease: a 4-year randomized controlled trial. Arch Neurol. 2004;61(7):1044–53. doi:10.1001/archneur.61.7.1044. 44. Smith LA, Jackson MJ, Johnston L, Kuoppamaki M, Rose S, AlBarghouthy G, et al. Switching from levodopa to the long-acting dopamine D2/D3 agonist piribedil reduces the expression of dyskinesia while maintaining effective motor activity in MPTP-

45.

46.

47.

48.

49.

50.

51.

52.

53.

54.

55.

56.

57.

58.

59.

60.

61.

treated primates. Clin Neuropharmacol. 2006;29(3):112–25. doi:10.1097/01.WNF.0000220818.71231.DF. Bibbiani F, Costantini LC, Patel R, Chase TN. Continuous dopaminergic stimulation reduces risk of motor complications in parkinsonian primates. Exp Neurol. 2005;192(1):73–8. doi:10. 1016/j.expneurol.2004.11.013. Blanchet PJ, Calon F, Martel JC, Bedard PJ, Di Paolo T, Walters RR, et al. Continuous administration decreases and pulsatile administration increases behavioral sensitivity to a novel dopamine D2 agonist (U-91356A) in MPTP-exposed monkeys. J Pharmacol Exp Ther. 1995;272(2):854–9. Stockwell KA, Scheller DK, Smith LA, Rose S, Iravani MM, Jackson MJ, et al. Continuous rotigotine administration reduces dyskinesia resulting from pulsatile treatment with rotigotine or L-dopa in MPTP-treated common marmosets. Exp Neurol. 2010;221(1):79–85. doi:10.1016/j.expneurol.2009.10.004. Brotchie JM. Nondopaminergic mechanisms in levodopainduced dyskinesia. Mov Disord. 2005;20(8):919–31. doi:10. 1002/mds.20612. Nutt JG. Continuous dopaminergic stimulation: Is it the answer to the motor complications of levodopa? Mov Disord. 2007;22(1):1–9. doi:10.1002/mds.21060. Nadjar A, Gerfen CR, 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):1–9. doi:10.1016/j.pneurobio.2008.09.013. Maratos EC, Jackson MJ, Pearce RK, Jenner P. Antiparkinsonian activity and dyskinesia risk of ropinirole and L-dopa combination therapy in drug naive MPTP-lesioned common marmosets (Callithrix jacchus). Mov Disord. 2001;16(4):631–41. Hill MP, Brotchie JM, Crossman AR, Bezard E, Michel A, Grimee R, et al. Levetiracetam interferes with the L-dopa priming process in MPTP-lesioned drug-naive marmosets. Clin Neuropharmacol. 2004;27(4):171–7. Bedard PJ, Di Paolo T, Falardeau P, Boucher R. Chronic treatment with L-dopa, but not bromocriptine induces dyskinesia in MPTP-parkinsonian monkeys. Correlation with [3H] spiperone binding. Brain Res. 1986;379(2):294–9. Lera G, Vaamonde J, Muruzabal J, Obeso JA. Cabergoline: a long-acting dopamine agonist in Parkinson’s disease. Ann Neurol. 1990;28(4):593–4. doi:10.1002/ana.410280428. Montastruc JL, Rascol O, Senard JM, Rascol A. A randomised controlled study comparing bromocriptine to which levodopa was later added, with levodopa alone in previously untreated patients with Parkinson’s disease: a five year follow up. J Neurol Neurosurg Psychiatry. 1994;57(9):1034–8. Murray AM, Waddington JL. The interaction of clozapine with dopamine D1 versus dopamine D2 receptor-mediated function: behavioural indices. Eur J Pharmacol. 1990;186(1):79–86. Maratos EC, Jackson MJ, Pearce RK, Cannizzaro C, Jenner P. Both short- and long-acting D-1/D-2 dopamine agonists induce less dyskinesia than L-dopa in the MPTP-lesioned common marmoset (Callithrix jacchus). Exp Neurol. 2003;179(1):90–102. Grondin R, Bedard PJ, Britton DR, Shiosaki K. Potential therapeutic use of the selective dopamine D1 receptor agonist, A-86929: an acute study in parkinsonian levodopa-primed monkeys. Neurology. 1997;49(2):421–6. Pons R, Syrengelas D, Youroukos S, Orfanou I, Dinopoulos A, Cormand B, et al. Levodopa-induced dyskinesias in tyrosine hydroxylase deficiency. Mov Disord. 2013;28(8):1058–63. doi:10.1002/mds.25382. Hwang WJ, Calne DB, Tsui JK, de la Fuente-Fernandez R. The long-term response to levodopa in dopa-responsive dystonia. Parkinsonism Relat Disord. 2001;8(1):1–5. Berthet A, Porras G, Doudnikoff E, Stark H, Cador M, Bezard E, et al. Pharmacological analysis demonstrates dramatic alteration

E. Schaeffer et al.

62.

63.

64.

65.

66.

67.

68.

69.

70.

71.

72.

73.

74.

75.

76.

of D1 dopamine receptor neuronal distribution in the rat analog of L-dopa-induced dyskinesia. J Neurosci. 2009;29(15):4829–35. doi:10.1523/JNEUROSCI.5884-08.2009. Oh JD, Russell DS, Vaughan CL, Chase TN. Enhanced tyrosine phosphorylation of striatal NMDA receptor subunits: effect of dopaminergic denervation and L-dopa administration. Brain Res. 1998;813(1):150–9. Santini E, Sgambato-Faure V, Li Q, Savasta M, Dovero S, Fisone G, et al. Distinct changes in cAMP and extracellular signalregulated protein kinase signalling in L-dopa-induced dyskinesia. PloS One. 2010;5(8):e12322. doi:10.1371/journal.pone. 0012322. Santini E, Valjent E, Usiello A, Carta M, Borgkvist A, Girault JA, et al. Critical involvement of cAMP/DARPP-32 and extracellular signal-regulated protein kinase signaling in L-dopainduced dyskinesia. J Neurosci. 2007;27(26):6995–7005. doi:10. 1523/JNEUROSCI.0852-07.2007. Iravani MM, Jenner P. Mechanisms underlying the onset and expression of levodopa-induced dyskinesia and their pharmacological manipulation. J Neural Transm. 2011;118(12): 1661–90. doi:10.1007/s00702-011-0698-2. Gerfen CR, Young WS 3rd. Distribution of striatonigral and striatopallidal peptidergic neurons in both patch and matrix compartments: an in situ hybridization histochemistry and fluorescent retrograde tracing study. Brain Res. 1988;460(1): 161–7. Le Moine C, Normand E, Bloch B. Phenotypical characterization of the rat striatal neurons expressing the D1 dopamine receptor gene. Proc Natl Acad Sci USA. 1991;88(10):4205–9. Gertler TS, Chan CS, Surmeier DJ. Dichotomous anatomical properties of adult striatal medium spiny neurons. J Neurosci. 2008;28(43):10814–24. doi:10.1523/JNEUROSCI.2660-08. 2008. Darmopil S, Martin AB, De Diego IR, Ares S, Moratalla R. Genetic inactivation of dopamine D1 but not D2 receptors inhibits L-dopa-induced dyskinesia and histone activation. Biol Psychiatry. 2009;66(6):603–13. doi:10.1016/j.biopsych.2009.04. 025. Guigoni C, Doudnikoff E, Li Q, Bloch B, Bezard E. Altered D(1) dopamine receptor trafficking in parkinsonian and dyskinetic non-human primates. Neurobiol Dis. 2007;26(2):452–63. doi:10.1016/j.nbd.2007.02.001. Aubert I, Guigoni C, Hakansson K, Li Q, Dovero S, Barthe N, et al. Increased D1 dopamine receptor signaling in levodopainduced dyskinesia. Ann Neurol. 2005;57(1):17–26. doi:10. 1002/ana.20296. Santini E, Valjent E, Fisone G. mTORC1 signaling in Parkinson’s disease and L-dopa-induced dyskinesia: a sensitized matter. Cell Cycle. 2010;9(14):2713–8. Westin JE, Vercammen L, Strome EM, Konradi C, Cenci MA. Spatiotemporal pattern of striatal ERK1/2 phosphorylation in a rat model of L-dopa-induced dyskinesia and the role of dopamine D1 receptors. Biol Psychiatry. 2007;62(7):800–10. doi:10. 1016/j.biopsych.2006.11.032. Pavon N, Martin AB, Mendialdua A, Moratalla R. ERK phosphorylation and FosB expression are associated with L-dopainduced dyskinesia in hemiparkinsonian mice. Biol Psychiatry. 2006;59(1):64–74. doi:10.1016/j.biopsych.2005.05.044. Andersson M, Hilbertson A, Cenci MA. Striatal fosB expression is causally linked with L-dopa-induced abnormal involuntary movements and the associated upregulation of striatal prodynorphin mRNA in a rat model of Parkinson’s disease. Neurobiol Dis. 1999;6(6):461–74. doi:10.1006/nbdi.1999.0259. Gerfen CR, Miyachi S, Paletzki R, Brown P. D1 dopamine receptor supersensitivity in the dopamine-depleted striatum

77.

78.

79.

80.

81.

82.

83.

84.

85.

86.

87.

88.

89.

90.

results from a switch in the regulation of ERK1/2/MAP kinase. J Neurosci. 2002;22(12):5042–54. Gerfen CR. D1 dopamine receptor supersensitivity in the dopamine-depleted striatum animal model of Parkinson’s disease. Neuroscientist. 2003;9(6):455–62. doi:10.1177/1073858 403255839. Bezard E, Gross CE, Qin L, Gurevich VV, Benovic JL, Gurevich EV. L-dopa reverses the MPTP-induced elevation of the arrestin2 and GRK6 expression and enhanced ERK activation in monkey brain. Neurobiol Dis. 2005;18(2):323–35. doi:10.1016/ j.nbd.2004.10.005. Konradi C, Westin JE, Carta M, Eaton ME, Kuter K, Dekundy A, et al. Transcriptome analysis in a rat model of L-dopainduced dyskinesia. Neurobiol Dis. 2004;17(2):219–36. doi:10. 1016/j.nbd.2004.07.005. Chase TN, Oh JD. Striatal dopamine- and glutamate-mediated dysregulation in experimental parkinsonism. Trends Neurosci. 2000;23(10 Suppl):S86–91. Calabresi P, Giacomini P, Centonze D, Bernardi G. Levodopainduced dyskinesia: a pathological form of striatal synaptic plasticity? Ann Neurol. 2000;47(4 Suppl 1):S60–8. (Discussion S8–9). Calon F, Rajput AH, Hornykiewicz O, Bedard PJ, Di Paolo T. Levodopa-induced motor complications are associated with alterations of glutamate receptors in Parkinson’s disease. Neurobiol Dis. 2003;14(3):404–16. Fiorentini C, Busi C, Spano P, Missale C. Role of receptor heterodimers in the development of L-dopa-induced dyskinesias in the 6-hydroxydopamine rat model of Parkinson’s disease. Parkinsonism Relat Disord. 2008;14(Suppl 2):S159–64. doi:10. 1016/j.parkreldis.2008.04.022. Fiorentini C, Rizzetti MC, Busi C, Bontempi S, Collo G, Spano P, et al. Loss of synaptic D1 dopamine/N-methyl-D-aspartate glutamate receptor complexes in L-dopa-induced dyskinesia in the rat. Mol Pharmacol. 2006;69(3):805–12. doi:10.1124/mol. 105.016667. Hallett PJ, Spoelgen R, Hyman BT, Standaert DG, Dunah AW. Dopamine D1 activation potentiates striatal NMDA receptors by tyrosine phosphorylation-dependent subunit trafficking. J Neurosci. 2006;26(17):4690–700. doi:10.1523/JNEUROSCI.079206.2006. Picconi B, Paille V, Ghiglieri V, Bagetta V, Barone I, Lindgren HS, et al. L-dopa dosage is critically involved in dyskinesia via loss of synaptic depotentiation. Neurobiol Dis. 2008;29(2):327–35. doi:10.1016/j.nbd.2007.10.001. Picconi B, Centonze D, Hakansson K, Bernardi G, Greengard P, Fisone G, et al. Loss of bidirectional striatal synaptic plasticity in L-dopa-induced dyskinesia. Nat Neurosci. 2003;6(5):501–6. doi:10.1038/nn1040. Hallett PJ, Dunah AW, Ravenscroft P, Zhou S, Bezard E, Crossman AR, et al. 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(4):503–16. doi:10.1016/j.neuropharm. 2004.11.008. Silverdale MA, Kobylecki C, Hallett PJ, Li Q, Dunah AW, Ravenscroft P, et al. Synaptic recruitment of AMPA glutamate receptor subunits in levodopa-induced dyskinesia in the MPTPlesioned nonhuman primate. Synapse. 2010;64(2):177–80. doi:10.1002/syn.20739. Hurley MJ, Jackson MJ, Smith LA, Rose S, Jenner P. Immunoautoradiographic analysis of NMDA receptor subunits and associated postsynaptic density proteins in the brain of dyskinetic MPTP-treated common marmosets. Eur J Neurosci. 2005;21(12):3240–50. doi:10.1111/j.1460-9568.2005.04169.x.

Pharmacological Strategies for the Management of Levodopa-Induced Dyskinesia 91. Gardoni F, Picconi B, Ghiglieri V, Polli F, Bagetta V, Bernardi G, et al. A critical interaction between NR2B and MAGUK in Ldopa induced dyskinesia. J Neurosci. 2006;26(11):2914–22. doi:10.1523/JNEUROSCI.5326-05.2006. 92. Sgambato-Faure V, Cenci MA. 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(1):69–86. doi:10.1016/j. pneurobio.2011.10.005. 93. Ahmed I, Bose SK, Pavese N, Ramlackhansingh A, Turkheimer F, Hotton G, et al. Glutamate NMDA receptor dysregulation in Parkinson’s disease with dyskinesias. Brain J Neurol. 2011;134(Pt 4):979–86. doi:10.1093/brain/awr028. 94. Robelet S, Melon C, Guillet B, Salin P, Kerkerian-Le Goff L. Chronic L-dopa treatment increases extracellular glutamate levels and GLT1 expression in the basal ganglia in a rat model of Parkinson’s disease. Eur J Neurosci. 2004;20(5):1255–66. doi:10.1111/j.1460-9568.2004.03591.x. 95. Ouattara B, Gregoire L, Morissette M, Gasparini F, Vranesic I, Bilbe G, et al. Metabotropic glutamate receptor type 5 in levodopa-induced motor complications. Neurobiol Aging. 2011;32(7):1286–95. doi:10.1016/j.neurobiolaging.2009.07.014. 96. Luginger E, Wenning GK, Bosch S, Poewe W. Beneficial effects of amantadine on L-dopa-induced dyskinesias in Parkinson’s disease. Mov Disord. 2000;15(5):873–8. 97. Snow BJ, Macdonald L, McAuley D, Wallis W. The effect of amantadine on levodopa-induced dyskinesias in Parkinson’s disease: a double-blind, placebo-controlled study. Clin Neuropharmacol. 2000;23(2):82–5. 98. Metman LV, Del Dotto P, LePoole K, Konitsiotis S, Fang J, Chase TN. Amantadine for levodopa-induced dyskinesias: a 1-year follow-up study. Arch Neurol. 1999;56(11):1383–6. 99. Wolf E, Seppi K, Katzenschlager R, Hochschorner G, Ransmayr G, Schwingenschuh P, et al. Long-term antidyskinetic efficacy of amantadine in Parkinson’s disease. Mov Disord. 2010;25(10):1357–63. doi:10.1002/mds.23034. 100. Stocchi F, Rascol O, Destee A, Hattori N, Hauser RA, Lang AE, et al. AFQ056 in Parkinson patients with levodopa-induced dyskinesia: 13-week, randomized, dose-finding study. Mov Disord. 2013;28(13):1838–46. doi:10.1002/mds.25561. 101. Berg D, Godau J, Trenkwalder C, Eggert K, Csoti I, Storch A, et al. AFQ056 treatment of levodopa-induced dyskinesias: results of 2 randomized controlled trials. Mov Disord. 2011;26(7):1243–50. doi:10.1002/mds.23616. 102. Rylander D, Recchia A, Mela F, Dekundy A, Danysz W, Cenci MA. Pharmacological modulation of glutamate transmission in a rat model of L-dopa-induced dyskinesia: effects on motor behavior and striatal nuclear signaling. J Pharmacol Exp Ther. 2009;330(1):227–35. doi:10.1124/jpet.108.150425. 103. Carta M, Carlsson T, Kirik D, Bjorklund A. Dopamine released from 5-HT terminals is the cause of L-dopa-induced dyskinesia in parkinsonian rats. Brain J Neurol. 2007;130(Pt 7):1819–33. doi:10.1093/brain/awm082. 104. Carta M, Carlsson T, Munoz A, Kirik D, Bjorklund A. Serotonin-dopamine interaction in the induction and maintenance of Ldopa-induced dyskinesias. Prog Brain Res. 2008;172:465–78. doi:10.1016/S0079-6123(08)00922-9. 105. 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(3):195–8. 106. Arai R, Karasawa N, Geffard M, Nagatsu T, Nagatsu I. Immunohistochemical evidence that central serotonin neurons produce dopamine from exogenous L-dopa in the rat, with reference to the involvement of aromatic L-amino acid decarboxylase. Brain Res. 1994;667(2):295–9.

107. Lavoie B, Parent A. Immunohistochemical study of the serotoninergic innervation of the basal ganglia in the squirrel monkey. J Comp Neurol. 1990;299(1):1–16. doi:10.1002/cne.902990102. 108. Maeda T, Nagata K, Yoshida Y, Kannari K. Serotonergic hyperinnervation into the dopaminergic denervated striatum compensates for dopamine conversion from exogenously administered L-dopa. Brain Res. 2005;1046(1–2):230–3. doi:10. 1016/j.brainres.2005.04.019. 109. Yamada H, Aimi Y, Nagatsu I, Taki K, Kudo M, Arai R. Immunohistochemical detection of L-dopa-derived dopamine within serotonergic fibers in the striatum and the substantia nigra pars reticulata in Parkinsonian model rats. Neurosci Res. 2007;59(1):1–7. doi:10.1016/j.neures.2007.05.002. 110. Carta M, Carlsson T, Munoz A, Kirik D, Bjorklund A. Role of serotonin neurons in the induction of levodopa- and graftinduced dyskinesias in Parkinson’s disease. Mov Disord. 2010;25(Suppl 1):S174–9. doi:10.1002/mds.22792. 111. Carta M, Bezard E. Contribution of pre-synaptic mechanisms to L-dopa-induced dyskinesia. Neuroscience. 2011;198:245–51. doi:10.1016/j.neuroscience.2011.07.070. 112. Lopez A, Munoz A, Guerra MJ, Labandeira-Garcia JL. Mechanisms of the effects of exogenous levodopa on the dopaminedenervated striatum. Neuroscience. 2001;103(3):639–51. 113. 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(3):631–4. 114. Eskow KL, Dupre KB, Barnum CJ, Dickinson SO, Park JY, 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(7):610–20. doi:10.1002/syn.20630. 115. Carlsson T, Carta M, Winkler C, Bjorklund A, Kirik D. Serotonin neuron transplants exacerbate L-dopa-induced dyskinesias in a rat model of Parkinson’s disease. J Neurosci. 2007;27(30): 8011–22. doi:10.1523/JNEUROSCI.2079-07.2007. 116. Lindgren HS, Andersson DR, Lagerkvist S, Nissbrandt H, Cenci MA. L-dopa-induced dopamine efflux in the striatum and the substantia nigra in a rat model of Parkinson’s disease: temporal and quantitative relationship to the expression of dyskinesia. J Neurochem. 2010;112(6):1465–76. doi:10.1111/j.1471-4159. 2009.06556.x. 117. Kannari K, Yamato H, Shen H, Tomiyama M, Suda T, Matsunaga M. Activation of 5-HT(1A) but not 5-HT(1B) receptors attenuates an increase in extracellular dopamine derived from exogenously administered L-dopa in the striatum with nigrostriatal denervation. J Neurochem. 2001;76(5):1346–53. 118. Munoz A, Carlsson T, Tronci E, Kirik D, Bjorklund A, Carta M. Serotonin neuron-dependent and -independent reduction of dyskinesia by 5-HT1A and 5-HT1B receptor agonists in the rat Parkinson model. Exp Neurol. 2009;219(1):298–307. doi:10. 1016/j.expneurol.2009.05.033. 119. Munoz A, Li Q, Gardoni F, Marcello E, Qin C, Carlsson T, et al. Combined 5-HT1A and 5-HT1B receptor agonists for the treatment of L-dopa-induced dyskinesia. Brain J Neurol. 2008;131(Pt 12):3380–94. doi:10.1093/brain/awn235. 120. Nahimi A, Holtzermann M, Landau AM, Simonsen M, Jakobsen S, Alstrup AK, et al. Serotonergic modulation of receptor occupancy in rats treated with L-dopa after unilateral 6-OHDA lesioning. J Neurochem. 2012;120(5):806–17. doi:10.1111/j. 1471-4159.2011.07598.x. 121. Navailles S, Bioulac B, Gross C, De Deurwaerdere P. Serotonergic neurons mediate ectopic release of dopamine induced by L-dopa in a rat model of Parkinson’s disease. Neurobiol Dis. 2010;38(1):136–43. doi:10.1016/j.nbd.2010.01.012. 122. Politis M, Wu K, Loane C, Quinn NP, Brooks DJ, Rehncrona S, et al. Serotonergic neurons mediate dyskinesia side effects in

E. Schaeffer et al.

123.

124.

125.

126.

127.

128.

129.

130.

131.

132.

133.

134.

135.

136.

137.

138.

139.

Parkinson’s patients with neural transplants. Sci Transl Med. 2010;2(38):38ra46. doi:10.1126/scitranslmed.3000976. Politis M, Oertel WH, Wu K, Quinn NP, Pogarell O, Brooks DJ, et al. Graft-induced dyskinesias in Parkinson’s disease: High striatal serotonin/dopamine transporter ratio. Mov Disord. 2011;26(11):1997–2003. doi:10.1002/mds.23743. Hagell P, Piccini P, Bjorklund A, Brundin P, Rehncrona S, Widner H, et al. Dyskinesias following neural transplantation in Parkinson’s disease. Nat Neurosci. 2002;5(7):627–8. doi:10. 1038/nn863. Zeng BY, Iravani MM, Jackson MJ, Rose S, Parent A, Jenner P. Morphological changes in serotoninergic neurites in the striatum and globus pallidus in levodopa primed MPTP treated common marmosets with dyskinesia. Neurobiol Dis. 2010;40(3):599–607. doi:10.1016/j.nbd.2010.08.004. Rylander D, Parent M, O’Sullivan SS, Dovero S, Lees AJ, Bezard E, et al. Maladaptive plasticity of serotonin axon terminals in levodopa-induced dyskinesia. Ann Neurol. 2010;68(5):619–28. doi:10.1002/ana.22097. Guerra MJ, Liste I, Labandeira-Garcia JL. Effects of lesions of the nigrostriatal pathway and of nigral grafts on striatal serotonergic innervation in adult rats. Neuroreport. 1997;8(16):3485–8. Fornai F, di Poggio AB, Pellegrini A, Ruggieri S, Paparelli A. Noradrenaline in Parkinson’s disease: from disease progression to current therapeutics. Curr Med Chem. 2007;14(22):2330–4. Grenhoff J, Nisell M, Ferre S, Aston-Jones G, Svensson TH. Noradrenergic modulation of midbrain dopamine cell firing elicited by stimulation of the locus coeruleus in the rat. J Neural Transm Gen Sect. 1993;93(1):11–25. Bucheler MM, Hadamek K, Hein L. Two alpha(2)-adrenergic receptor subtypes, alpha(2A) and alpha(2C), inhibit transmitter release in the brain of gene-targeted mice. Neuroscience. 2002;109(4):819–26. Foote SL, Bloom FE, Aston-Jones G. Nucleus locus ceruleus: new evidence of anatomical and physiological specificity. Physiol Rev. 1983;63(3):844–914. Jones BE, Yang TZ. The efferent projections from the reticular formation and the locus coeruleus studied by anterograde and retrograde axonal transport in the rat. J Comp Neurol. 1985;242(1):56–92. doi:10.1002/cne.902420105. Gaspar P, Stepniewska I, Kaas JH. Topography and collateralization of the dopaminergic projections to motor and lateral prefrontal cortex in owl monkeys. J Comp Neurol. 1992;325(1):1–21. doi:10.1002/cne.903250102. Ordway GA, Jaconetta SM, Halaris AE. Characterization of subtypes of alpha-2 adrenoceptors in the human brain. J Pharmacol Exp Ther. 1993;264(2):967–76. Uhlen S, Lindblom J, Tiger G, Wikberg JE. Quantification of alpha2A and alpha2C adrenoceptors in the rat striatum and in different regions of the spinal cord. Acta Physiol Scand. 1997;160(4):407–12. doi:10.1046/j.1365-201X.1997.00175.x. Hill MP, Brotchie JM. The adrenergic receptor agonist, clonidine, potentiates the anti-parkinsonian action of the selective kappa-opioid receptor agonist, enadoline, in the monoaminedepleted rat. Br J Pharmacol. 1999;128(7):1577–85. doi:10. 1038/sj.bjp.0702943. Zhang W, Ordway GA. The alpha2C-adrenoceptor modulates GABA release in mouse striatum. Brain Res Mol Brain Res. 2003;112(1–2):24–32. Alachkar A, Brotchie JM, Jones OT. Changes in the mRNA levels of alpha2A and alpha2C adrenergic receptors in rat models of Parkinson’s disease and L-dopa-induced dyskinesia. J Mol Neurosci. 2012;46(1):145–52. doi:10.1007/s12031-0119539-x. Barnum CJ, Bhide N, Lindenbach D, Surrena MA, Goldenberg AA, Tignor S, et al. Effects of noradrenergic denervation on L-

140.

141.

142.

143.

144.

145.

146.

147.

148.

149.

150.

151.

152.

153.

154.

dopa-induced dyskinesia and its treatment by alpha- and betaadrenergic receptor antagonists in hemiparkinsonian rats. Pharmacol Biochem Behav. 2012;100(3):607–15. doi:10.1016/j.pbb. 2011.09.009. Fulceri F, Biagioni F, Ferrucci M, Lazzeri G, Bartalucci A, Galli V, et al. Abnormal involuntary movements (AIMs) following pulsatile dopaminergic stimulation: severe deterioration and morphological correlates following the loss of locus coeruleus neurons. Brain Res. 2007;1135(1):219–29. doi:10.1016/j.brainres. 2006.12.030. Piccini P, Weeks RA, Brooks DJ. Alterations in opioid receptor binding in Parkinson’s disease patients with levodopa-induced dyskinesias. Ann Neurol. 1997;42(5):720–6. doi:10.1002/ana. 410420508. Cohen RM, Carson RE, Aigner TG, Doudet DJ. Opiate receptor avidity is reduced in non-motor impaired MPTP-lesioned rhesus monkeys. Brain Res. 1998;806(2):292–6. Aubert I, Guigoni C, Li Q, Dovero S, Bioulac BH, Gross CE, et al. Enhanced preproenkephalin-B-derived opioid transmission in striatum and subthalamic nucleus converges upon globus pallidus internalis in L-3,4-dihydroxyphenylalanine-induced dyskinesia. Biol Psychiatry. 2007;61(7):836–44. doi:10.1016/j. biopsych.2006.06.038. Gerdeman G, Lovinger DM. CB1 cannabinoid receptor inhibits synaptic release of glutamate in rat dorsolateral striatum. J Neurophysiol. 2001;85(1):468–71. Mailleux P, Parmentier M, Vanderhaeghen JJ. Distribution of cannabinoid receptor messenger RNA in the human brain: an in situ hybridization histochemistry with oligonucleotides. Neurosci Lett. 1992;143(1–2):200–4. Wang Y, Zhang QJ, Wang HS, Wang T, Liu J. Genome-wide microarray analysis identifies a potential role for striatal retrograde endocannabinoid signaling in the pathogenesis of experimental Ldopa-induced dyskinesia. Synapse. 2014;. doi:10.1002/syn.21740. Ferre S, Popoli P, Gimenez-Llort L, Rimondini R, Muller CE, Stromberg I, et al. Adenosine/dopamine interaction: implications for the treatment of Parkinson’s disease. Parkinsonism Relat Disord. 2001;7(3):235–41. Kase H. New aspects of physiological and pathophysiological functions of adenosine A2A receptor in basal ganglia. Biosci Biotechnol Biochem. 2001;65(7):1447–57. Calon F, Dridi M, Hornykiewicz O, Bedard PJ, Rajput AH, Di Paolo T. Increased adenosine A2A receptors in the brain of Parkinson’s disease patients with dyskinesias. Brain J Neurol. 2004;127(Pt 5):1075–84. doi:10.1093/brain/awh128. Xiao D, Bastia E, Xu YH, Benn CL, Cha JH, Peterson TS, et al. Forebrain adenosine A2A receptors contribute to L-3,4-dihydroxyphenylalanine-induced dyskinesia in hemiparkinsonian mice. J Neurosci. 2006;26(52):13548–55. doi:10.1523/ JNEUROSCI.3554-06.2006. Koranda JL, Cone JJ, McGehee DS, Roitman MF, Beeler JA, Zhuang X. Nicotinic receptors regulate the dynamic range of dopamine release in vivo. J Neurophysiol. 2014;111(1):103–11. doi:10.1152/jn.00269.2013. Perez XA, O’Leary KT, Parameswaran N, McIntosh JM, Quik M. Prominent role of alpha3/alpha6beta2* nAChRs in regulating evoked dopamine release in primate putamen: effect of longterm nicotine treatment. Mol Pharmacol. 2009;75(4):938–46. doi:10.1124/mol.108.053801. Garcao P, Szabo EC, Wopereis S, Castro AA, Tome AR, Prediger RD, et al. Functional interaction between pre-synaptic alpha6beta2-containing nicotinic and adenosine A2A receptors in the control of dopamine release in the rat striatum. Br J Pharmacol. 2013;169(7):1600–11. doi:10.1111/bph.12234. Moreno E, Hoffmann H, Gonzalez-Sepulveda M, Navarro G, Casado V, Cortes A, et al. Dopamine D1-histamine H3 receptor

Pharmacological Strategies for the Management of Levodopa-Induced Dyskinesia

155.

156.

157.

158.

159.

160.

161.

162.

163. 164.

165.

166.

167.

168.

169.

170.

heteromers provide a selective link to MAPK signaling in GABAergic neurons of the direct striatal pathway. J Biol Chem. 2011;286(7):5846–54. doi:10.1074/jbc.M110.161489. Prast H, Tran MH, Fischer H, Kraus M, Lamberti C, Grass K, et al. Histaminergic neurons modulate acetylcholine release in the ventral striatum: role of H3 histamine receptors. Naunyn Schmiedebergs Arch Pharmacol. 1999;360(5):558–64. Molina-Hernandez A, Nunez A, Arias-Montano JA. Histamine H3-receptor activation inhibits dopamine synthesis in rat striatum. Neuroreport. 2000;11(1):163–6. Gomez-Ramirez J, Johnston TH, Visanji NP, Fox SH, Brotchie JM. Histamine H3 receptor agonists reduce L-dopa-induced chorea, but not dystonia, in the MPTP-lesioned nonhuman primate model of Parkinson’s disease. Mov Disord. 2006;21(6):839–46. doi:10.1002/mds.20828. Lindgren HS, Ohlin KE, Cenci MA. Differential involvement of D1 and D2 dopamine receptors in L-dopa-induced angiogenic activity in a rat model of Parkinson’s disease. Neuropsychopharmacology. 2009;34(12):2477–88. doi:10.1038/npp.2009.74. Westin JE, Lindgren HS, Gardi J, Nyengaard JR, Brundin P, Mohapel P, et al. Endothelial proliferation and increased bloodbrain barrier permeability in the basal ganglia in a rat model of 3,4-dihydroxyphenyl-L-alanine-induced dyskinesia. J Neurosci. 2006;26(37):9448–61. doi:10.1523/JNEUROSCI.0944-06.2006. Ohlin KE, Francardo V, Lindgren HS, Sillivan SE, O’Sullivan SS, Luksik AS, et al. Vascular endothelial growth factor is upregulated by L-dopa in the parkinsonian brain: implications for the development of dyskinesia. Brain J Neurol. 2011;134(Pt 8):2339–57. doi:10.1093/brain/awr165. Lieu CA, Subramanian T. The interhemispheric connections of the striatum: Implications for Parkinson’s disease and druginduced dyskinesias. Brain Res Bull. 2012;87(1):1–9. doi:10. 1016/j.brainresbull.2011.09.013. Lieu CA, Deogaonkar M, Bakay RA, Subramanian T. Dyskinesias do not develop after chronic intermittent levodopa therapy in clinically hemiparkinsonian rhesus monkeys. Parkinsonism Relat Disord. 2011;17(1):34–9. doi:10.1016/j. parkreldis.2010.10.010. Fahn S. A new look at levodopa based on the ELLDOPA study. J Neural Transm Suppl. 2006;70:419–26. Fahn S, Oakes D, Shoulson I, Kieburtz K, Rudolph A, Lang A, et al. Levodopa and the progression of Parkinson’s disease. N Engl J Med. 2004;351(24):2498–508. doi:10.1056/ NEJMoa033447. Sage JI, Mark MH. Comparison of controlled-release Sinemet (CR4) and standard Sinemet (25 mg/100 mg) in advanced Parkinson’s disease: a double-blind, crossover study. Clin Neuropharmacol. 1988;11(2):174–9. Ahlskog JE, Muenter MD, McManis PG, Bell GN, Bailey PA. Controlled-release Sinemet (CR-4): a double-blind crossover study in patients with fluctuating Parkinson’s disease. Mayo Clin Proc. 1988;63(9):876–86. Hutton JT, Morris JL, Bush DF, Smith ME, Liss CL, Reines S. Multicenter controlled study of Sinemet CR vs Sinemet (25/100) in advanced Parkinson’s disease. Neurology. 1989;39(11 Suppl 2):67–72. (Discussion 3). Jankovic J, Schwartz K, Vander Linden C. Comparison of Sinemet CR4 and standard Sinemet: double blind and long-term open trial in parkinsonian patients with fluctuations. Mov Disord. 1989;4(4):303–9. doi:10.1002/mds.870040403. Lieberman A, Gopinathan G, Miller E, Neophytides A, Baumann G, Chin L. Randomized double-blind cross-over study of Sinemet-controlled release (CR4 50/200) versus Sinemet 25/100 in Parkinson’s disease. Eur Neurol. 1990;30(2):75–8. Wolters EC, Horstink MW, Roos RA, Jansen EN. Clinical efficacy of Sinemet CR 50/200 versus Sinemet 25/100 in

171.

172.

173.

174.

175.

176.

177.

178.

179.

180.

181.

182.

183.

184.

patients with fluctuating Parkinson’s disease. An open, and a double-blind, double-dummy, multicenter treatment evaluation. The Dutch Sinemet CR Study Group. Clin Neurol Neurosurg. 1992;94(3):205–11. Dupont E, Andersen A, Boas J, Boisen E, Borgmann R, Helgetveit AC, et al. Sustained-release Madopar HBS compared with standard Madopar in the long-term treatment of de novo parkinsonian patients. Acta Neurol Scand. 1996;93(1):14–20. Block G, Liss C, Reines S, Irr J, Nibbelink D. Comparison of immediate-release and controlled release carbidopa/levodopa in Parkinson’s disease. A multicenter 5-year study. The CR First Study Group. Eur Neurol. 1997;37(1):23–7. Koller WC, Hutton JT, Tolosa E, Capilldeo R. Immediaterelease and controlled-release carbidopa/levodopa in PD: a 5-year randomized multicenter study. Carbidopa/Levodopa Study Group. Neurology. 1999;53(5):1012–9. Stocchi F, Fabbri L, Vecsei L, Krygowska-Wajs A, Monici Preti PA, Ruggieri SA. Clinical efficacy of a single afternoon dose of effervescent levodopa-carbidopa preparation (CHF 1512) in fluctuating Parkinson disease. Clin Neuropharmacol. 2007;30(1):18–24. doi:10.1097/01.WNF.0000236762.77913. C6. Stocchi F, Zappia M, Dall’Armi V, Kulisevsky J, Lamberti P, Obeso JA. Melevodopa/carbidopa effervescent formulation in the treatment of motor fluctuations in advanced Parkinson’s disease. Mov Disord. 2010;25(12):1881–7. doi:10.1002/mds. 23206. Djaldetti R, Inzelberg R, Giladi N, Korczyn AD, Peretz-Aharon Y, Rabey MJ, et al. Oral solution of levodopa ethylester for treatment of response fluctuations in patients with advanced Parkinson’s disease. Mov Disord. 2002;17(2):297–302. Metman LV, Hoff J, Mouradian MM, Chase TN. Fluctuations in plasma levodopa and motor responses with liquid and tablet levodopa/carbidopa. Mov Disord. 1994;9(4):463–5. doi:10. 1002/mds.870090416. Kurth MC, Tetrud JW, Tanner CM, Irwin I, Stebbins GT, Goetz CG, et al. Double-blind, placebo-controlled, crossover study of duodenal infusion of levodopa/carbidopa in Parkinson’s disease patients with ‘on-off’ fluctuations. Neurology. 1993;43(9):1698–703. Antonini A, Odin P. Pros and cons of apomorphine and L-dopa continuous infusion in advanced Parkinson’s disease. Parkinsonism Relat Disord. 2009;15(Suppl 4):S97–100. doi:10.1016/ S1353-8020(09)70844-2. Nilsson D, Nyholm D, Aquilonius SM. Duodenal levodopa infusion in Parkinson’s disease—long-term experience. Acta Neurol Scand. 2001;104(6):343–8. Eggert K, Schrader C, Hahn M, Stamelou M, Russmann A, Dengler R, et al. Continuous jejunal levodopa infusion in patients with advanced parkinson disease: practical aspects and outcome of motor and non-motor complications. Clin Neuropharmacol. 2008;31(3):151–66. doi:10.1097/wnf.0b013e31814b113e. Manson AJ, Turner K, Lees AJ. 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(6):1235–41. doi:10.1002/mds.10281. Katzenschlager R, Hughes A, Evans A, Manson AJ, Hoffman M, Swinn L, et al. Continuous subcutaneous apomorphine therapy improves dyskinesias in Parkinson’s disease: a prospective study using single-dose challenges. Mov Disord. 2005;20(2):151–7. doi:10.1002/mds.20276. Garcia Ruiz PJ, Sesar Ignacio A, Ares Pensado B, Castro Garcia A, Alonso Frech F, Alvarez Lopez M, et al. Efficacy of longterm continuous subcutaneous apomorphine infusion in advanced Parkinson’s disease with motor fluctuations: a multicenter study. Mov Disord. 2008;23(8):1130–6. doi:10.1002/mds. 22063.

E. Schaeffer et al. 185. Rascol O, Brooks DJ, Korczyn AD, De Deyn PP, Clarke CE, Lang AE. A five-year study of the incidence of dyskinesia in patients with early Parkinson’s disease who were treated with ropinirole or levodopa. 056 Study Group. N Engl J Med. 2000;342(20):1484–91. doi:10.1056/NEJM200005183422004. 186. Watts RL, Lyons KE, Pahwa R, Sethi K, Stern M, Hauser RA, et al. Onset of dyskinesia with adjunct ropinirole prolongedrelease or additional levodopa in early Parkinson’s disease. Mov Disord. 2010;25(7):858–66. doi:10.1002/mds.22890. 187. Bracco F, Battaglia A, Chouza C, Dupont E, Gershanik O, Marti Masso JF, et al. The long-acting dopamine receptor agonist cabergoline in early Parkinson’s disease: final results of a 5-year, double-blind, levodopa-controlled study. CNS Drugs. 2004;18(11):733–46. 188. Fox SH, Katzenschlager R, Lim SY, Ravina B, Seppi K, Coelho M, et al. The movement disorder society evidence-based medicine review update: treatments for the motor symptoms of Parkinson’s disease. Mov Disord. 2011;26(Suppl 3):S2–41. doi:10.1002/mds.23829. 189. Stowe R, Ives N, Clarke CE, Handley K, Furmston A, Deane K, et al. Meta-analysis of the comparative efficacy and safety of adjuvant treatment to levodopa in later Parkinson’s disease. Mov Disord. 2011;26(4):587–98. doi:10.1002/mds.23517. 190. Tayarani-Binazir K, Jackson MJ, Rose S, McCreary AC, Jenner P. The partial dopamine agonist pardoprunox (SLV308) administered in combination with L-dopa improves efficacy and decreases dyskinesia in MPTP treated common marmosets. Exp Neurol. 2010;226(2):320–7. doi:10.1016/j.expneurol.2010.09. 007. 191. Johnston LC, Jackson MJ, Rose S, McCreary AC, Jenner P. Pardoprunox reverses motor deficits but induces only mild dyskinesia in MPTP-treated common marmosets. Mov Disord. 2010;25(13):2059–66. doi:10.1002/mds.23249. 192. Bronzova J, Sampaio C, Hauser RA, Lang AE, Rascol O, Theeuwes A, et al. Double-blind study of pardoprunox, a new partial dopamine agonist, in early Parkinson’s disease. Mov Disord. 2010;25(6):738–46. doi:10.1002/mds.22948. 193. Sampaio C, Bronzova J, Hauser RA, Lang AE, Rascol O, van de Witte SV, et al. Pardoprunox in early Parkinson’s disease: results from 2 large, randomized double-blind trials. Mov Disord. 2011;26(8):1464–76. doi:10.1002/mds.23590. 194. Rascol O, Bronzova J, Hauser RA, Lang AE, Sampaio C, Theeuwes A, et al. Pardoprunox as adjunct therapy to levodopa in patients with Parkinson’s disease experiencing motor fluctuations: results of a double-blind, randomized, placebo-controlled, trial. Parkinsonism Relat Disord. 2012;18(4):370–6. doi:10.1016/j.parkreldis.2011.12.006. 195. Stocchi F, Ruggieri S, Vacca L, Olanow CW. Prospective randomized trial of lisuride infusion versus oral levodopa in patients with Parkinson’s disease. Brain J Neurol. 2002;125(Pt 9):2058–66. 196. Parkinson_Study_Group. A randomized placebo-controlled trial of rasagiline in levodopa-treated patients with Parkinson disease and motor fluctuations: the PRESTO study. Arch Neurol. 2005;62(2):241–8. doi:10.1001/archneur.62.2.241. 197. Shoulson I, Oakes D, Fahn S, Lang A, Langston JW, LeWitt P, 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(5):604–12. doi:10.1002/ana.10191. 198. Mahmood I. Clinical pharmacokinetics and pharmacodynamics of selegiline. An update. Clin Pharmacokinet. 1997;33(2):91–102. doi:10.2165/00003088-199733020-00002. 199. Lew MF. Selegiline orally disintegrating tablets for the treatment of Parkinson’s disease. Expert Rev Neurother. 2005;5(6):705–12. doi:10.1586/14737175.5.6.705.

200. Waters CH, Sethi KD, Hauser RA, Molho E, Bertoni JM. Zydis selegiline reduces off time in Parkinson’s disease patients with motor fluctuations: a 3-month, randomized, placebo-controlled study. Mov Disord. 2004;19(4):426–32. doi:10.1002/mds. 20036. 201. Ondo WG, Sethi KD, Kricorian G. Selegiline orally disintegrating tablets in patients with Parkinson disease and ‘‘wearing off’’ symptoms. Clin Neuropharmacol. 2007;30(5):295–300. doi:10.1097/WNF.0b013e3180616570. 202. Talati R, Reinhart K, Baker W, White CM, Coleman CI. Pharmacologic treatment of advanced Parkinson’s disease: a metaanalysis of COMT inhibitors and MAO-B inhibitors. Parkinsonism Relat Disord. 2009;15(7):500–5. doi:10.1016/j. parkreldis.2008.12.007. 203. Gregoire L, Morin N, Ouattara B, Gasparini F, Bilbe G, Johns D, et al. The acute antiparkinsonian and antidyskinetic effect of AFQ056, a novel metabotropic glutamate receptor type 5 antagonist, in L-dopa-treated parkinsonian monkeys. Parkinsonism Relat Disord. 2011;17(4):270–6. doi:10.1016/j. parkreldis.2011.01.008. 204. Schapira AH, Stocchi F, Borgohain R, Onofrj M, Bhatt M, Lorenzana P, et al. Long-term efficacy and safety of safinamide as add-on therapy in early Parkinson’s disease. Eur J Neurol. 2013;20(2):271–80. doi:10.1111/j.1468-1331.2012.03840.x. 205. Stocchi F, Borgohain R, Onofrj M, Schapira AH, Bhatt M, Lucini V, et al. A randomized, double-blind, placebo-controlled trial of safinamide as add-on therapy in early Parkinson’s disease patients. Mov Disord. 2012;27(1):106–12. doi:10.1002/mds. 23954. 206. Ferreira JJ, Katzenschlager R, Bloem BR, Bonuccelli U, Burn D, Deuschl G, et al. Summary of the recommendations of the EFNS/MDS-ES review on therapeutic management of Parkinson’s disease. Eur J Neurol. 2013;20(1):5–15. doi:10.1111/j. 1468-1331.2012.03866.x. 207. Verhagen Metman L, Del Dotto P, van den Munckhof P, Fang J, Mouradian MM, Chase TN. Amantadine as treatment for dyskinesias and motor fluctuations in Parkinson’s disease. Neurology. 1998;50(5):1323–6. 208. Del Dotto P, Pavese N, Gambaccini G, Bernardini S, Metman LV, Chase TN, et al. Intravenous amantadine improves levadopa-induced dyskinesias: an acute double-blind placebo-controlled study. Mov Disord. 2001;16(3):515–20. 209. da Silva-Junior FP, Braga-Neto P, Sueli Monte F, de Bruin VM. Amantadine reduces the duration of levodopa-induced dyskinesia: a randomized, double-blind, placebo-controlled study. Parkinsonism Relat Disord. 2005;11(7):449–52. doi:10.1016/j. parkreldis.2005.05.008. 210. Sawada H, Oeda T, Kuno S, Nomoto M, Yamamoto K, Yamamoto M, et al. Amantadine for dyskinesias in Parkinson’s disease: a randomized controlled trial. PLoS One. 2010;5(12):e15298. doi:10.1371/journal.pone.0015298. 211. Thomas A, Iacono D, Luciano AL, Armellino K, Di Iorio A, Onofrj M. Duration of amantadine benefit on dyskinesia of severe Parkinson’s disease. J Neurol Neurosurg Psychiatry. 2004;75(1):141–3. 212. Ory-Magne F, Corvol JC, Azulay JP, Bonnet AM, Brefel-Courbon C, Damier P, et al. Withdrawing amantadine in dyskinetic patients with Parkinson disease: the AMANDYSK trial. Neurology. 2014;82(4):300–7. doi:10.1212/WNL.0000000000000050. 213. Hely MA, Morris JG, Reid WG, Trafficante R. Sydney Multicenter Study of Parkinson’s disease: non-L-dopa-responsive problems dominate at 15 years. Mov Disord. 2005;20(2):190–9. doi:10.1002/mds.20324. 214. Rascol O, Brooks DJ, Melamed E, Oertel W, Poewe W, Stocchi F, et al. Rasagiline as an adjunct to levodopa in patients with Parkinson’s disease and motor fluctuations (LARGO, lasting

Pharmacological Strategies for the Management of Levodopa-Induced Dyskinesia

215.

216.

217.

218.

219.

220.

221.

222.

223.

224.

225.

226.

227.

228.

229.

effect in adjunct therapy with rasagiline given once daily, study): a randomised, double-blind, parallel-group trial. Lancet. 2005;365(9463):947–54. doi:10.1016/S0140-6736(05)71083-7. Factor SA, Molho ES, Brown DL. Acute delirium after withdrawal of amantadine in Parkinson’s disease. Neurology. 1998;50(5):1456–8. Ossola B, Schendzielorz N, Chen SH, Bird GS, Tuominen RK, Mannisto PT, et al. Amantadine protects dopamine neurons by a dual action: reducing activation of microglia and inducing expression of GDNF in astroglia [corrected]. Neuropharmacology. 2011;61(4):574–82. doi:10.1016/j.neuropharm.2011.04. 030. Inzelberg R, Bonuccelli U, Schechtman E, Miniowich A, Strugatsky R, Ceravolo R, et al. Association between amantadine and the onset of dementia in Parkinson’s disease. Mov Disord. 2006;21(9):1375–9. doi:10.1002/mds.20968. Jahangirvand A, Rajput A. Early use of amantadine to prevent or delay onset of levodopa-induced dyskinesia in Parkinson’s disease. Mov Disord. 2013;28(Suppl 1):S207. Pahwa RTC, Hauser RA. Randomized trial of extended release amantadine in Parkinson’s disease patients with L-dopa-induced dyskinesia (EASED study). Mov Disord. 2013;28(Suppl 1):S158. Hanagasi HA, Kaptanoglu G, Sahin HA, Emre M. The use of NMDA antagonist memantine in drug-resistant dyskinesias resulting from L-dopa. Mov Disord. 2000;15(5):1016–7. Varanese S, Howard J, Di Rocco A. NMDA antagonist memantine improves levodopa-induced dyskinesias and ‘‘on-off’’ phenomena in Parkinson’s disease. Mov Disord. 2010;25(4):508–10. doi:10.1002/mds.22917. Moreau C, Delval A, Tiffreau V, Defebvre L, Dujardin K, Duhamel A, et al. Memantine for axial signs in Parkinson’s disease: a randomised, double-blind, placebo-controlled pilot study. J Neurol Neurosurg Psychiatry. 2013;84(5):552–5. doi:10.1136/jnnp-2012-303182. Merello M, Nouzeilles MI, Cammarota A, Leiguarda R. Effect of memantine (NMDA antagonist) on Parkinson’s disease: a double-blind crossover randomized study. Clin Neuropharmacol. 1999;22(5):273–6. Emre M, Tsolaki M, Bonuccelli U, Destee A, Tolosa E, Kutzelnigg A, et al. Memantine for patients with Parkinson’s disease dementia or dementia with Lewy bodies: a randomised, double-blind, placebo-controlled trial. Lancet Neurol. 2010;9(10):969–77. doi:10.1016/S1474-4422(10)70194-0. Verhagen Metman L, Del Dotto P, Natte R, van den Munckhof P, Chase TN. Dextromethorphan improves levodopa-induced dyskinesias in Parkinson’s disease. Neurology. 1998;51(1):203–6. Braz CA, Borges V, Ferraz HB. Effect of riluzole on dyskinesia and duration of the on state in Parkinson disease patients: a double-blind, placebo-controlled pilot study. Clin Neuropharmacol. 2004;27(1):25–9. Wessell RH, Ahmed SM, Menniti FS, Dunbar GL, Chase TN, Oh JD. NR2B selective NMDA receptor antagonist CP-101,606 prevents levodopa-induced motor response alterations in hemiparkinsonian rats. Neuropharmacology. 2004;47(2):184–94. doi:10.1016/j.neuropharm.2004.03.011. Morissette M, Dridi M, Calon F, Hadj Tahar A, Meltzer LT, Bedard PJ, et al. Prevention of levodopa-induced dyskinesias by a selective NR1A/2B N-methyl-D-aspartate receptor antagonist in parkinsonian monkeys: implication of preproenkephalin. Mov Disord. 2006;21(1):9–17. doi:10.1002/mds.20654. Tamim MK, Samadi P, Morissette M, Gregoire L, Ouattara B, Levesque D, et al. Effect of non-dopaminergic drug treatment on Levodopa induced dyskinesias in MPTP monkeys: common implication of striatal neuropeptides. Neuropharmacology. 2010;58(1):286–96. doi:10.1016/j.neuropharm.2009.06.030.

230. Loftis JM, Janowsky A. The N-methyl-D-aspartate receptor subunit NR2B: localization, functional properties, regulation, and clinical implications. Pharmacol Ther. 2003;97(1):55–85. 231. Nutt JG, Gunzler SA, Kirchhoff T, Hogarth P, Weaver JL, Krams M, et al. Effects of a NR2B selective NMDA glutamate antagonist, CP-101,606, on dyskinesia and parkinsonism. Mov Disord. 2008;23(13):1860–6. doi:10.1002/mds.22169. 232. Duty S. Targeting glutamate receptors to tackle the pathogenesis, clinical symptoms and levodopa-induced dyskinesia associated with Parkinson’s disease. CNS Drugs. 2012;26(12):1017–32. doi:10.1007/s40263-012-0016-z. 233. Samadi P, Gregoire L, Morissette M, Calon F, Hadj Tahar A, Dridi M, et al. mGluR5 metabotropic glutamate receptors and dyskinesias in MPTP monkeys. Neurobiol Aging. 2008;29(7):1040–51. doi:10.1016/j.neurobiolaging.2007.02.005. 234. Mela F, Marti M, Dekundy A, Danysz W, Morari M, Cenci MA. Antagonism of metabotropic glutamate receptor type 5 attenuates L-dopa-induced dyskinesia and its molecular and neurochemical correlates in a rat model of Parkinson’s disease. J Neurochem. 2007;101(2):483–97. doi:10.1111/j.1471-4159. 2007.04456.x. 235. Morin N, Gregoire L, Gomez-Mancilla B, Gasparini F, Di Paolo T. Effect of the metabotropic glutamate receptor type 5 antagonists MPEP and MTEP in parkinsonian monkeys. Neuropharmacology. 2010;58(7):981–6. doi:10.1016/j.neuropharm.2009. 12.024. 236. Johnston TH, Fox SH, McIldowie MJ, Piggott MJ, Brotchie JM. Reduction of L-dopa-induced dyskinesia by the selective metabotropic glutamate receptor 5 antagonist 3-[(2-methyl-1,3thiazol-4-yl)ethynyl]pyridine in the 1-methyl-4-phenyl-1,2,3,6tetrahydropyridine-lesioned macaque model of Parkinson’s disease. J Pharmacol Exp Ther. 2010;333(3):865–73. doi:10. 1124/jpet.110.166629. 237. Tison F DF, Christophe J. Safety, tolerability and anti-dyskinetic efficacy of dipraglurant, a novel mGluR5 negative allosteric modulator (NAM) in Parkinson’s disease (PD) patients with Ldopa-induced dyskinesia. Poster presented at: 65th American Academy of Neurology Annual Meeting, San Diego; 2013. 238. Konitsiotis S, Blanchet PJ, Verhagen L, Lamers E, Chase TN. AMPA receptor blockade improves levodopa-induced dyskinesia in MPTP monkeys. Neurology. 2000;54(8):1589–95. 239. Klockgether T, Turski L, Honore T, Zhang ZM, Gash DM, Kurlan R, et al. The AMPA receptor antagonist NBQX has antiparkinsonian effects in monoamine-depleted rats and MPTPtreated monkeys. Ann Neurol. 1991;30(5):717–23. doi:10.1002/ ana.410300513. 240. Marin C, Jimenez A, Bonastre M, Vila M, Agid Y, Hirsch EC, et al. LY293558, an AMPA glutamate receptor antagonist, prevents and reverses levodopa-induced motor alterations in parkinsonian rats. Synapse. 2001;42(1):40–7. doi:10.1002/syn. 1097. 241. Lees A, Fahn S, Eggert KM, Jankovic J, Lang A, Micheli F, et al. Perampanel, an AMPA antagonist, found to have no benefit in reducing ‘‘off’’ time in Parkinson’s disease. Mov Disord. 2012;27(2):284–8. doi:10.1002/mds.23983. 242. Rascol O, Barone P, Behari M, Emre M, Giladi N, Olanow CW, et al. Perampanel in Parkinson disease fluctuations: a double-blind randomized trial with placebo and entacapone. Clin Neuropharmacol. 2012;35(1):15–20. doi:10.1097/WNF.0b013e318 241520b. 243. Calabresi P, Di Filippo M, Ghiglieri V, Tambasco N, Picconi B. Levodopa-induced dyskinesias in patients with Parkinson’s disease: filling the bench-to-bedside gap. Lancet Neurol. 2010;9(11):1106–17. doi:10.1016/S1474-4422(10)70218-0. 244. Kobylecki C, Cenci MA, Crossman AR, Ravenscroft P. Calcium-permeable AMPA receptors are involved in the induction

E. Schaeffer et al.

245.

246.

247.

248.

249.

250.

251.

252.

253.

254.

255.

256.

257.

258.

259.

and expression of L-dopa-induced dyskinesia in Parkinson’s disease. J Neurochem. 2010;114(2):499–511. doi:10.1111/j. 1471-4159.2010.06776.x. Price PA, Parkes JD, Marsden CD. Sodium valproate in the treatment of levodopa-induced dyskinesia. J Neurol Neurosurg Psychiatry. 1978;41(8):702–6. Van Blercom N, Lasa A, Verger K, Masramon X, Sastre VM, 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(3):124–8. Murata M, Hasegawa K, Kanazawa I. Zonisamide improves motor function in Parkinson disease: a randomized, double-blind study. Neurology. 2007;68(1):45–50. doi:10.1212/01.wnl. 0000250236.75053.16. Kobylecki C, Burn DJ, Kass-Iliyya L, Kellett MW, Crossman AR, Silverdale MA. Randomized clinical trial of topiramate for levodopa-induced dyskinesia in Parkinson’s disease. Parkinsonism Relat Disord. 2014;20(4):452–5. doi:10.1016/j. parkreldis.2014.01.016. Stathis P, Konitsiotis S, Tagaris G, Peterson D. Levetiracetam for the management of levodopa-induced dyskinesias in Parkinson’s disease. Mov Disord. 2011;26(2):264–70. doi:10.1002/ mds.23355. Wong KK, Alty JE, Goy AG, Raghav S, Reutens DC, Kempster PA. A randomized, double-blind, placebo-controlled trial of levetiracetam for dyskinesia in Parkinson’s disease. Mov Disord. 2011;26(8):1552–5. doi:10.1002/mds.23687. Wolz M, Lohle M, Strecker K, Schwanebeck U, Schneider C, Reichmann H, et al. Levetiracetam for levodopa-induced dyskinesia in Parkinson’s disease: a randomized, double-blind, placebocontrolled trial. J Neural Transm. 2010;117(11):1279–86. doi:10. 1007/s00702-010-0472-x. Nevalainen N, Af Bjerken S, Lundblad M, Gerhardt GA, Stromberg I. Dopamine release from serotonergic nerve fibers is reduced in L-dopa-induced dyskinesia. J Neurochem. 2011;118(1):12–23. doi:10.1111/j.1471-4159.2011.07292.x. Dupre KB, Ostock CY, Eskow Jaunarajs KL, Button T, Savage LM, Wolf W, et al. Local modulation of striatal glutamate efflux by serotonin 1A receptor stimulation in dyskinetic, hemiparkinsonian rats. Exp Neurol. 2011;229(2):288–99. doi:10.1016/j. expneurol.2011.02.012. Olanow CW, Damier P, Goetz CG, Mueller T, Nutt J, Rascol O, et al. Multicenter, open-label, trial of sarizotan in Parkinson disease patients with levodopa-induced dyskinesias (the SPLENDID Study). Clin Neuropharmacol. 2004;27(2):58–62. Kannari K, Kurahashi K, Tomiyama M, Maeda T, Arai A, Baba M, et al. Tandospirone citrate, a selective 5-HT1A agonist, alleviates L-dopa-induced dyskinesia in patients with Parkinson’s disease. No To Shinkei. 2002;54(2):133–7. Rascol ODP, Goetz CG. A large phase III study to evaluate the safety and efficacy of sarizotan in the treatment of L-dopainduced dyskinesia associated with Parkinson’s disease: the Paddy-1 study. Mov Disord. 2006;21(Suppl 15):S492–3. Mu¨ller TOC, Nutt J. The PADDY-2 study: the evaluation of sarizotan for treatment-associated dyskinesia in Parkinson’s disease patients. Mov Disord. 2006;21(Suppl 15):S591. Goetz CG, Damier P, Hicking C, Laska E, Muller T, Olanow CW, et al. Sarizotan as a treatment for dyskinesias in Parkinson’s disease: a double-blind placebo-controlled trial. Mov Disord. 2007;22(2):179–86. doi:10.1002/mds.21226. Tani Y, Ogata A, Koyama M, Inoue T. Effects of piclozotan (SUN N4057), a partial serotonin 1A receptor agonist, on motor complications induced by repeated administration of levodopa in parkinsonian rats. Eur J Pharmacol. 2010;649(1–3):218–23. doi:10.1016/j.ejphar.2010.09.013.

260. Bezard E, Tronci E, Pioli EY, Li Q, Porras G, Bjorklund A, et al. Study of the antidyskinetic effect of eltoprazine in animal models of levodopa-induced dyskinesia. Mov Disord. 2013;28(8):1088–96. doi:10.1002/mds.25366. 261. Visanji NP, Gomez-Ramirez J, Johnston TH, Pires D, Voon V, Brotchie JM, et al. Pharmacological characterization of psychosislike behavior in the MPTP-lesioned nonhuman primate model of Parkinson’s disease. Mov Disord. 2006;21(11):1879–91. doi:10. 1002/mds.21073. 262. Reddy S, Factor SA, Molho ES, Feustel PJ. The effect of quetiapine on psychosis and motor function in parkinsonian patients with and without dementia. Mov Disord. 2002;17(4):676–81. doi:10.1002/mds.10176. 263. Katzenschlager R, Manson AJ, Evans A, Watt H, Lees AJ. Low dose quetiapine for drug induced dyskinesias in Parkinson’s disease: a double blind cross over study. J Neurol Neurosurg Psychiatry. 2004;75(2):295–7. 264. Durif F, Debilly B, Galitzky M, Morand D, Viallet F, Borg M, et al. Clozapine improves dyskinesias in Parkinson disease: a double-blind, placebo-controlled study. Neurology. 2004;62(3): 381–8. 265. Huot P, Johnston TH, Koprich JB, Fox SH, Brotchie JM. The pharmacology of L-dopa-induced dyskinesia in Parkinson’s disease. Pharmacol Rev. 2013;65(1):171–222. doi:10.1124/pr. 111.005678. 266. Johnston TH, Fox SH, Piggott MJ, Savola JM, Brotchie JM. The alpha(2) adrenergic antagonist fipamezole improves quality of levodopa action in Parkinsonian primates. Mov Disord. 2010;25(13):2084–93. doi:10.1002/mds.23172. 267. Savola JM, Hill M, Engstrom M, Merivuori H, Wurster S, McGuire SG, et al. Fipamezole (JP-1730) is a potent alpha2 adrenergic receptor antagonist that reduces levodopa-induced dyskinesia in the MPTP-lesioned primate model of Parkinson’s disease. Mov Disord. 2003;18(8):872–83. doi:10.1002/mds.10464. 268. Henry B, Fox SH, Peggs D, Crossman AR, Brotchie JM. The alpha2-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. 1999;14(5):744–53. 269. Manson AJ, Iakovidou E, Lees AJ. Idazoxan is ineffective for levodopa-induced dyskinesias in Parkinson’s disease. Mov Disord. 2000;15(2):336–7. 270. Rascol O, Arnulf I, Peyro-Saint Paul H, Brefel-Courbon C, Vidailhet M, Thalamas C, et al. Idazoxan, an alpha-2 antagonist, and L-dopa-induced dyskinesias in patients with Parkinson’s disease. Mov Disord. 2001;16(4):708–13. 271. Lewitt PA, Hauser RA, Lu M, Nicholas AP, Weiner W, Coppard N, et al. Randomized clinical trial of fipamezole for dyskinesia in Parkinson disease (FJORD study). Neurology. 2012;79(2):163–9. doi:10.1212/WNL.0b013e31825f0451. 272. Henry B, Fox SH, Crossman AR, Brotchie JM. Mu- and deltaopioid receptor antagonists reduce levodopa-induced dyskinesia in the MPTP-lesioned primate model of Parkinson’s disease. Exp Neurol. 2001;171(1):139–46. doi:10.1006/exnr.2001.7727. 273. Rascol O, Fabre N, Blin O, Poulik J, Sabatini U, Senard JM, et al. Naltrexone, an opiate antagonist, fails to modify motor symptoms in patients with Parkinson’s disease. Mov Disord. 1994;9(4):437–40. doi:10.1002/mds.870090410. 274. Manson AJ, Katzenschlager R, Hobart J, Lees AJ. High dose naltrexone for dyskinesias induced by levodopa. J Neurol Neurosurg Psychiatry. 2001;70(4):554–6. 275. Fox S, Silverdale M, Kellett M, Davies R, Steiger M, Fletcher N, et al. Non-subtype-selective opioid receptor antagonism in treatment of levodopa-induced motor complications in Parkinson’s disease. Mov Disord. 2004;19(5):554–60. doi:10.1002/mds. 10693.

Pharmacological Strategies for the Management of Levodopa-Induced Dyskinesia 276. Di Marzo V, Hill MP, Bisogno T, Crossman AR, Brotchie JM. Enhanced levels of endogenous cannabinoids in the globus pallidus are associated with a reduction in movement in an animal model of Parkinson’s disease. FASEB J. 2000;14(10):1432–8. 277. Fox SH, 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. 2002;17(6):1180–7. doi:10.1002/mds.10289. 278. Venderova K, Ruzicka E, Vorisek V, Visnovsky P. Survey on cannabis use in Parkinson’s disease: subjective improvement of motor symptoms. Mov Disord. 2004;19(9):1102–6. doi:10.1002/ mds.20111. 279. Sieradzan KA, Fox SH, Hill M, Dick JP, Crossman AR, Brotchie JM. Cannabinoids reduce levodopa-induced dyskinesia in Parkinson’s disease: a pilot study. Neurology. 2001;57(11):2108–11. 280. Carroll CB, Bain PG, Teare L, Liu X, Joint C, Wroath C, et al. Cannabis for dyskinesia in Parkinson disease: a randomized double-blind crossover study. Neurology. 2004;63(7):1245–50. 281. Hasko G, Linden J, Cronstein B, Pacher P. Adenosine receptors: therapeutic aspects for inflammatory and immune diseases. Nat Rev Drug Discov. 2008;7(9):759–70. doi:10.1038/nrd2638. 282. Morelli M, Di Paolo T, Wardas J, Calon F, Xiao D, Schwarzschild MA. Role of adenosine A2A receptors in parkinsonian motor impairment and L-dopa-induced motor complications. Prog Neurobiol. 2007;83(5):293–309. doi:10.1016/j.pneurobio. 2007.07.001. 283. Schiffmann SN, Fisone G, Moresco R, Cunha RA, Ferre S. Adenosine A2A receptors and basal ganglia physiology. Prog Neurobiol. 2007;83(5):277–92. doi:10.1016/j.pneurobio.2007. 05.001. 284. Mori A, Shindou T. Modulation of GABAergic transmission in the striatopallidal system by adenosine A2A receptors: a potential mechanism for the antiparkinsonian effects of A2A antagonists. Neurology. 2003;61(11 Suppl 6):S44–8. 285. Carta AR, Pinna A, Cauli O, Morelli M. Differential regulation of GAD67, enkephalin and dynorphin mRNAs by chronicintermittent L-dopa and A2A receptor blockade plus L-dopa in dopamine-denervated rats. Synapse. 2002;44(3):166–74. doi:10. 1002/syn.10066. 286. Jenner P, Mori A, Hauser R, Morelli M, Fredholm BB, Chen JF. Adenosine, adenosine A 2A antagonists, and Parkinson’s disease. Parkinsonism Relat Disord. 2009;15(6):406–13. doi:10. 1016/j.parkreldis.2008.12.006. 287. Hodgson RA, Bedard PJ, Varty GB, Kazdoba TM, Di Paolo T, Grzelak ME, et al. Preladenant, a selective A(2A) receptor antagonist, is active in primate models of movement disorders. Exp Neurol. 2010;225(2):384–90. doi:10.1016/j.expneurol. 2010.07.011. 288. Hauser RA, Cantillon M, Pourcher E, Micheli F, Mok V, Onofrj M, et al. Preladenant in patients with Parkinson’s disease and motor fluctuations: a phase 2, double-blind, randomised trial. Lancet Neurol. 2011;10(3):221–9. doi:10.1016/S14744422(11)70012-6. 289. Factor SA, Wolski K, Togasaki DM, Huyck S, Cantillon M, Ho TW, et al. Long-term safety and efficacy of preladenant in subjects with fluctuating Parkinson’s disease. Mov Disord. 2013;28(6):817–20. doi:10.1002/mds.25395. 290. Hauser RA, Olanow C, Kieburtz K, Neale A, Resburg C, Maya U, Bandaket S. A phase 2, placebo-controlled, randomized, double-blind trial of tozadenant (SYN-115) in patients with Parkinson’s disease with wearing-off fluctuations on L-dopa. Mov Disord. 2013;28(Suppl 1):S158. 291. Bara-Jimenez W, Dimitrova TD, Sherzai A, Aksu M, Chase TN. Glutamate release inhibition ineffective in levodopa-induced motor complications. Mov Disord. 2006;21(9):1380–3. doi:10. 1002/mds.20976.

292. Zhou FM, Liang Y, Dani JA. Endogenous nicotinic cholinergic activity regulates dopamine release in the striatum. Nat Neurosci. 2001;4(12):1224–9. doi:10.1038/nn769. 293. Zhang D, Bordia T, McGregor M, McIntosh JM, Decker MW, Quik M. ABT-089 and ABT-894 reduce levodopa-induced dyskinesias in a monkey model of Parkinson’s disease. Mov Disord. 2014;29(4):508–17. doi:10.1002/mds.25817. 294. Xie CL, Pan JL, Zhang SF, Gan J, Liu ZG. Effect of nicotine on L-dopa-induced dyskinesia in animal models of Parkinson’s disease: a systematic review and meta-analysis. Neurol Sci. 2014;35(5):653–62. doi:10.1007/s10072-014-1652-5. 295. Bordia T, Campos C, McIntosh JM, Quik M. Nicotinic receptormediated reduction in L-dopa-induced dyskinesias may occur via desensitization. J Pharmacol Exp Ther. 2010;333(3):929–38. doi:10.1124/jpet.109.162396. 296. Parkinson_Study_Group. Randomized placebo-controlled study of the nicotinic agonist SIB-1508Y in Parkinson disease. Neurology. 2006;66(3):408–10. doi:10.1212/01.wnl.0000196466.99381.5c. 297. Mango D, Bonito-Oliva A, Ledonne A, Cappellacci L, Petrelli R, Nistico R, et al. Adenosine A1 receptor stimulation reduces D1 receptor-mediated GABAergic transmission from striatonigral terminals and attenuates L-dopa-induced dyskinesia in dopamine-denervated mice. Exp Neurol. 2014;. doi:10.1016/j. expneurol.2014.08.022. 298. Park HY, Kang YM, Kang Y, Park TS, Ryu YK, Hwang JH, et al. Inhibition of adenylyl cyclase type 5 prevents L-dopainduced dyskinesia in an animal model of Parkinson’s disease. J Neurosci. 2014;34(35):11744–53. doi:10.1523/JNEUROSCI. 0864-14.2014. 299. Goetz CG, Laska E, Hicking C, Damier P, Muller T, Nutt J, et al. Placebo influences on dyskinesia in Parkinson’s disease. Mov Disord. 2008;23(5):700–7. doi:10.1002/mds.21897. 300. Foltynie T, Cheeran B, Williams-Gray CH, Edwards MJ, Schneider SA, Weinberger D, et al. BDNF val66met influences time to onset of levodopa induced dyskinesia in Parkinson’s disease. J Neurol Neurosurg Psychiatry. 2009;80(2):141–4. doi:10.1136/jnnp.2008.154294. 301. Zappia M, Annesi G, Nicoletti G, Arabia G, Annesi F, Messina D, et al. Sex differences in clinical and genetic determinants of levodopa peak-dose dyskinesias in Parkinson disease: an exploratory study. Arch Neurol. 2005;62(4):601–5. doi:10.1001/ archneur.62.4.601. 302. Goetz CG, Stebbins GT, Chung KA, Hauser RA, Miyasaki JM, Nicholas AP, et al. Which dyskinesia scale best detects treatment response? Mov Disord. 2013;28(3):341–6. doi:10.1002/ mds.25321. 303. Pietracupa S, Fasano A, Fabbrini G, Sarchioto M, Bloise M, Latorre A, et al. Poor self-awareness of levodopa-induced dyskinesias in Parkinson’s disease: clinical features and mechanisms. Parkinsonism Relat Disord. 2013;19(11):1004–8. doi:10. 1016/j.parkreldis.2013.07.002. 304. Manson AJ, Brown P, O’Sullivan JD, Asselman P, Buckwell D, Lees AJ. An ambulatory dyskinesia monitor. J Neurol Neurosurg Psychiatry. 2000;68(2):196–201. 305. Maetzler W, Domingos J, Srulijes K, Ferreira JJ, Bloem BR. Quantitative wearable sensors for objective assessment of Parkinson’s disease. Mov Disord. 2013;28(12):1628–37. doi:10. 1002/mds.25628. 306. No authors listed. A comparison of Madopar CR and standard Madopar in the treatment of nocturnal and early-morning disability in Parkinson’s disease. The UK Madopar CR Study Group. Clin Neuropharmacol. 1989;12(6):498–505. 307. Moller JC, Oertel WH, Koster J, Pezzoli G, Provinciali L. Longterm efficacy and safety of pramipexole in advanced Parkinson’s disease: results from a European multicenter trial. Mov Disord. 2005;20(5):602–10. doi:10.1002/mds.20397.

E. Schaeffer et al. 308. Poewe WH, Rascol O, Quinn N, Tolosa E, Oertel WH, Martignoni E, et al. Efficacy of pramipexole and transdermal rotigotine in advanced Parkinson’s disease: a double-blind, doubledummy, randomised controlled trial. Lancet Neurol. 2007;6(6):513–20. doi:10.1016/S1474-4422(07)70108-4. 309. Mizuno Y, Abe T, Hasegawa K, Kuno S, Kondo T, Yamamoto M, et al. Ropinirole is effective on motor function when used as an adjunct to levodopa in Parkinson’s disease: STRONG study. Mov Disord. 2007;22(13):1860–5. doi:10.1002/mds.21313. 310. Barone P, Lamb J, Ellis A, Clarke Z. Sumanirole versus placebo or ropinirole for the adjunctive treatment of patients with advanced Parkinson’s disease. Mov Disord. 2007;22(4):483–9. doi:10.1002/mds.21191. 311. Pahwa R, Stacy MA, Factor SA, Lyons KE, Stocchi F, Hersh BP, et al. Ropinirole 24-hour prolonged release: randomized, controlled study in advanced Parkinson disease. Neurology. 2007;68(14):1108–15. doi:10.1212/01.wnl.0000258660.74391.c1. 312. LeWitt PA, Lyons KE, Pahwa R. Advanced Parkinson disease treated with rotigotine transdermal system: PREFER Study. Neurology. 2007;68(16):1262–7. doi:10.1212/01.wnl.0000259 516.61938.bb. 313. Borgohain R, Szasz J, Stanzione P, Meshram C, Bhatt MH, Chirilineau D, et al. Two-year, randomized, controlled study of safinamide as add-on to levodopa in mid to late Parkinson’s disease. Mov Disord. 2014;. doi:10.1002/mds.25961. 314. Borgohain R, Szasz J, Stanzione P, Meshram C, Bhatt M, Chirilineau D, et al. Randomized trial of safinamide add-on to levodopa in Parkinson’s disease with motor fluctuations. Mov Disord. 2014;29(2):229–37. doi:10.1002/mds.25751. 315. Poewe WH, Deuschl G, Gordin A, Kultalahti ER, Leinonen M. Efficacy and safety of entacapone in Parkinson’s disease patients with suboptimal levodopa response: a 6-month randomized placebo-controlled double-blind study in Germany and Austria (Celomen Study). Acta Neurol Scand. 2002;105(4):245–55. 316. Fenelon G, Gimenez-Roldan S, Montastruc JL, Bermejo F, Durif F, Bourdeix I, et al. Efficacy and tolerability of entacapone in patients with Parkinson’s disease treated with levodopa plus a dopamine agonist and experiencing wearing-off motor fluctuations. A randomized, double-blind, multicentre study. J Neural Transm. 2003;110(3):239–51. doi:10.1007/s00702-002-0799-z. 317. Brooks DJ, Sagar H. Entacapone is beneficial in both fluctuating and non-fluctuating patients with Parkinson’s disease: a randomised, placebo controlled, double blind, six month study. J Neurol Neurosurg Psychiatry. 2003;74(8):1071–9. 318. Reichmann H, Boas J, Macmahon D, Myllyla V, Hakala A, Reinikainen K. Efficacy of combining levodopa with entacapone on quality of life and activities of daily living in patients experiencing wearing-off type fluctuations. Acta Neurol Scand. 2005;111(1):21–8. doi:10.1111/j.1600-0404.2004.00363.x. 319. Mizuno Y, Kanazawa I, Kuno S, Yanagisawa N, Yamamoto M, Kondo T. Placebo-controlled, double-blind dose-finding study of entacapone in fluctuating parkinsonian patients. Mov Disord. 2007;22(1):75–80. doi:10.1002/mds.21218. 320. Baas H, Beiske AG, Ghika J, Jackson M, Oertel WH, Poewe W, et al. Catechol-O-methyltransferase inhibition with tolcapone reduces the ‘‘wearing off’’ phenomenon and levodopa

321.

322.

323.

324.

325.

326.

327.

328.

329.

330.

331.

332.

333.

334.

requirements in fluctuating parkinsonian patients. J Neurol Neurosurg Psychiatry. 1997;63(4):421–8. Waters CH, Kurth M, Bailey P, Shulman LM, LeWitt P, Dorflinger E, et al. Tolcapone in stable Parkinson’s disease: efficacy and safety of long-term treatment. The Tolcapone Stable Study Group. Neurology. 1997;49(3):665–71. Rajput AH, Martin W, Saint-Hilaire MH, Dorflinger E, Pedder S. Tolcapone improves motor function in parkinsonian patients with the ‘‘wearing-off’’ phenomenon: a double-blind, placebocontrolled, multicenter trial. Neurology. 1997;49(4):1066–71. Parkinson Study Group. Evaluation of dyskinesias in a pilot, randomized, placebo-controlled trial of remacemide in advanced Parkinson disease. Arch Neurol. 2001;58(10):1660–8. Eggert K, Squillacote D, Barone P, Dodel R, Katzenschlager R, Emre M, et al. Safety and efficacy of perampanel in advanced Parkinson’s disease: a randomized, placebo-controlled study. Mov Disord. 2010;25(7):896–905. doi:10.1002/mds.22974. Muller T, Russ H. Levodopa, motor fluctuations and dyskinesia in Parkinson’s disease. Expert Opin Pharmacother. 2006;7(13): 1715–30. doi:10.1517/14656566.7.13.1715. Dimitrova TDB-JW, Savola J-M. Alpha-2 adrenergic antagonist effects in Parkinson’s disease. Mov Disord. 2009;19(Suppl 9): S222. Bara-Jimenez W, Sherzai A, Dimitrova T, Favit A, Bibbiani F, Gillespie M, et al. Adenosine A(2A) receptor antagonist treatment of Parkinson’s disease. Neurology. 2003;61(3):293–6. LeWitt PA. ‘‘Off’’ time reduction from adjunctive use of istradefylline (KW-6002) in levodopa-treated patients with advanced Parkinson’s disease. Mov Disord. 2004;19(Suppl 9):S222. Stacy M, Silver D, Mendis T, Sutton J, Mori A, Chaikin P, et al. A 12-week, placebo-controlled study (6002-US-006) of istradefylline in Parkinson disease. Neurology. 2008;70(23):2233–40. doi:10.1212/01.wnl.0000313834.22171.17. Hauser RA, Shulman LM, Trugman JM, Roberts JW, Mori A, Ballerini R, et al. Study of istradefylline in patients with Parkinson’s disease on levodopa with motor fluctuations. Mov Disord. 2008;23(15):2177–85. doi:10.1002/mds.22095. LeWitt PA, Guttman M, Tetrud JW, Tuite PJ, Mori A, Chaikin P, et al. Adenosine A2A receptor antagonist istradefylline (KW6002) reduces ‘‘off’’ time in Parkinson’s disease: a double-blind, randomized, multicenter clinical trial (6002-US-005). Ann Neurol. 2008;63(3):295–302. doi:10.1002/ana.21315. Mizuno Y, Kondo T. Adenosine A2A receptor antagonist istradefylline reduces daily OFF time in Parkinson’s disease. Mov Disord. 2013;28(8):1138–41. doi:10.1002/mds.25418. Pourcher E, Fernandez HH, Stacy M, Mori A, Ballerini R, Chaikin P. Istradefylline for Parkinson’s disease patients experiencing motor fluctuations: results of the KW-6002-US-018 study. Parkinsonism Relat Disord. 2012;18(2):178–84. doi:10. 1016/j.parkreldis.2011.09.023. Hauser RA, Olanow CW, Kieburtz KD, Pourcher E, DocuAxelerad A, Lew M et al. 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(8):767–76. doi:10.1016/S1474-4422(14)70148-6.

Pharmacological strategies for the management of levodopa-induced dyskinesia in patients with Parkinson's disease.

L-Dopa-induced dyskinesias (LID) are the most common adverse effects of long-term dopaminergic therapy in Parkinson's disease (PD). However, the exact...
963KB Sizes 0 Downloads 7 Views