touchNEUROLOGY touchNEUROLOGY
Movement Disorders, Parkinson's Disease
Read Time: 11 mins

Dyskinesia – Advances in the Understanding of Pathophysiology and Possible Treatment Options

Copy Link
Published Online: Jun 4th 2011 European Neurological Review, 2010;5(2):34-40 DOI: http://doi.org/10.17925/ENR.2010.05.02.34
Authors: Hanna S Lindgren, M Angela Cenci, Emma L Lane
Quick Links:
Abstract
Article
Article Information
Abstract:
Overview

The degeneration of nigrostriatal dopaminergic neurons in Parkinson’s disease gives rise to tremor and slowness of movement, cardinal motor symptoms of the disease that can be alleviated by the dopamine precursor L-DOPA. Despite this, long-term L-DOPA treatment is hampered by the development of abnormal involuntary movements, i.e. dyskinesia, in the majority of patients. The pathophysiology of dyskinesia is complex and multifactorial, but excessive swings in extracellular dopamine causing aberrant plasticity in dopaminoceptive neurons are attributed a primary role. To date there are few effective treatment alternatives for patients with Parkinson’s disease experiencing dyskinesia, representing an unmet therapeutic need in the treatment strategy of the disease. This article reviews recent findings from both clinical and pre-clinical studies and their impact on the search for novel therapeutic approaches to levodopa-induced dyskinesia.

Keywords

L-3,4-dihydroxyphenylalanine (L-DOPA), dopamine replacement therapy, dyskinesia, animal models, plasticity, basal ganglia, dopamine, serotonin, noradrenalin, glutamate

Article:

Since its introduction in the 1960s, dopamine replacement therapy with its metabolic precursor L-3,4-dihydroxyphenylanaline (L-DOPA) has remained the most effective pharmacotherapy for Parkinson’s disease (PD).1 Particularly in early disease, it is efficiently converted into dopamine, replacing that which is lost through the degeneration of nigrostriatal dopaminergic neurons. In doing so, L-DOPA alleviates bradykinesia and rigidity, helping, for example, to restore gait and arm swing.2 However, its utility is limited by the development of motor complications that become increasingly evident as the disease progresses. These include:
• The development of ‘on-off’ fluctuations;
• periods in which the beneficial effects of L-DOPA are rapidly lost;
• a reduction in the effective duration of L-DOPA response; and
• the most troubling – choreic and dystonic abnormal involuntary movements, collectively termed L-DOPA-induced dyskinesia (LID).
The movements are present predominantly during the beneficial activity of L-DOPA,3 most commonly occurring as ‘peak dose’ dyskinesia but also occurring as dopamine levels decline as ‘end-of-dose’ LID. The reported incidence of LID varies between studies and patient groups, but can affect up to 40% of PD patients after only four to six years of L-DOPA therapy.4
Given that L-DOPA is still the most efficacious antiparkinsonian medication, providing therapy that circumvents LID remains a major unmet therapeutic need in the treatment strategy of PD. Currently, pharmacological interventions are limited to short-term treatment with the antiviral and weak NMDA (N-methyl-D-aspartic acid) antagonist amantadine or invasive neurosurgical intervention targeting the deep basal ganglia nuclei with deep brain stimulation (DBS). DBS stands alone as the only therapeutic option for severe LID and is both expensive and associated with acute and ongoing risks.5–7 The criteria for patient selection for this treatment is quite stringent and, critically, DBS is not offered to patients suffering from cognitive impairment, which affects as many as one-third of PD patients.8
Although the precise mechanisms underlying LID have remained elusive, three main risk factors for this condition have been conclusively identified in clinical studies. These are a young age of disease onset, disease severity (reflecting the extent of putaminal dopamine denervation) and high doses of L-DOPA.9,10 The latter two factors are readily reproduced in both non-human primates and rodents by the administration of neurotoxins, which ablate the nigrostriatal pathway, followed by daily treatment with L-DOPA at a sufficient dosage.11–16 The animals develop abnormal involuntary movements of hyperkinetic and/or dystonic phenotypes during the treatment period. These movements will gradually replace normal motor behaviour15 and overshadow the beneficial effect of L-DOPA.
Not all patients with PD develop dyskinesia and, similarly, a proportion of animals do not develop abnormal involuntary movements after chronic L-DOPA treatment. Thus, the animal models provide a valuable tool whereby molecular and neurochemical changes can be specifically correlated with dyskinesia rather than to a general effect of L-DOPA treatment. This article will provide a summary of the most important findings from these experimental models and their impact on the search for novel therapeutic options for LID.

Maladaptive Plasticity in L-DOPA-induced Dyskinesia
Pre-synaptic Plasticity

Generally, therapeutic doses of L-DOPA given to rodent or primate models of PD typically require a substantial dopaminergic depletion before LID can be established. Even so, some animals with extensive dopamine denervation remain non-dyskinetic after chronic L-DOPA treatment. Therefore, the extent of LID cannot be solely predicted by the degree or pattern of dopaminergic depletion.17 Dopaminergic depletion is a significant risk factor but additional, thus far undetermined factors, must also be key to the development of LID.
Large and rapid fluctuations in extracellular levels of dopamine have been regarded as the prime trigger of the motor complications associated with L-DOPA therapy.18 Accordingly, recent positron emission tomography studies in PD patients have demonstrated large increases in putaminal dopamine release after L-DOPA administration, which correlate positively with the severity of dyskinesia.19,20
In line with the clinical findings, high striatal levels of dopamine have been reported in L-DOPA-treated dyskinetic rats.21–23 The fluctuations were originally thought to be due to the loss of dopamine neurons. However, more recently it has been suggested that a reduced expression/dysfunction of the dopamine transporter in the remaining nigrostriatal terminals21,24 and a higher density of striatal serotonergic fibres25 may be critical contributors to changes in dopamine levels.
As the number of dopaminergic neurons decreases, the ability of the remaining terminals to convert L-DOPA into dopamine is reduced. Eventually, serotonergic neurons become the main site of L-DOPA decarboxylation in the brain. The lack of both dopamine autoreceptors and high-affinity reuptake capacity for dopamine, however, provides a source of unregulated dopamine efflux following exogenous L-DOPA administration.17
A causal link between the serotonin system and LID has recently been demonstrated in which LID was abolished by lesions of ascending 5-HT projections and greatly reduced by agonists of the serotonin autoreceptors 5-HT1A and 5-HT1B.26 These drugs were later found to blunt the surge of striatal extracellular dopamine following L-DOPA administration.22 Clinical trials with the 5-HT1A agonist sarizotan have demonstrated some improvements in dyskinesia, although the dopamine antagonist activity of this drug may have influenced the overall unimpressive clinical outcome.27,28 Pardoprunox (SLV308) is also being evaluated as a 5HT1A agonist with additional dopamine D2/3-receptor partial agonist activity. It has already shown some success as a de novo treatment in early PD.29
Dysregulated dopamine release may contribute substantially to the pulsatile levels of dopamine caused by intermittent L-DOPA administration.18 As these pulsatile events are closely correlated with LID, current pharmaceutical strategy has moved towards the concept of continuous dopaminergic stimulation (CDS). CDS aims to prolong the effects of L-DOPA, with the dual benefit of increases in ‘on-time’ and (hopefully) a reduction in dyskinesia. This approach is reportedly successful with the use of slow-release formulations of L-DOPA and adjuvant catechol-o-methyltransferase inhibitors improving ‘on’ time30 but two meta-analyses suggest that the consequence is significantly worse dyskinesia.31,32
The delivery of CDS moved forward significantly with the introduction of devices such as the duodenal L-DOPA pump (DuoDopa) and dopamine receptor-agonist patches. New gene therapy approaches should also assist in this endeavour, utilising viral vector-mediated delivery of coding sequences for essential enzymes in the synthesis of dopamine.33 The enzymes involved are tyrosine hydroxylase, the rate-limiting enzyme in dopamine synthesis, L-amino acid decarboxylase, which converts L-DOPA into dopamine, and guanosine 5’-triphosphate cyclohydrolase 1, which synthesises the co-factor biopterin necessary for tyrosine hydroxylase function.33 Gene therapy provides striatal neurons with the machinery to produce dopamine, obviating the need for exogenous L-DOPA delivery and providing a continuous supply of dopamine. This has been successfully demonstrated to protect against LID development in rodent and primate models of advanced PD34,35 and clinical trials are in progress.

Post-synaptic Plasticity

As dopaminergic innervation is lost, plastic changes occur at the post-synaptic membranes of striatal medium spinal neurons expressing D1 or D2 receptors. These are differentially altered again by the repeated pulsatile administration of L-DOPA.36 There are substantial data pointing towards a close correlation between the severity of LID and post-synaptic changes at the receptor level, downstream kinases, at gene regulation and transcription, largely focused around the D1 receptor subtype.11,12,37–40 Pre-clinical studies with dopamine D1 receptor agonists show both good antiparkinsonian and pro-dyskinetic effects, part of which may be attributed to their short half-lives, but also to the biochemical changes occurring at this receptor. D1 receptors are in fact G-protein-coupled receptors linked to G proteins, stimulating adenylate cyclase and activating gene transcription. With dopaminergic denervation and administration of L-DOPA, dopamine D1 receptors become more highly expressed and increasingly sensitive, the G-protein levels increasing, thereby altering activity through the whole signalling cascade.
Long-term dopamine D1 receptor sensitisation is controlled in part by G-protein receptor kinases (GRKs), which normally trigger receptor internalisation and halt the receptor response. GRK6, for example, is upregulated by MPTP administration to non-human primates and normalised by L-DOPA treatment.41 Reversing this normalisation significantly reduces LID, probably through the process of increasing internalisation of the D1 receptor.42
Further down the receptor cascade, components of the RAS/ERK signalling pathway become hyperactive in response to D1 supersensitivity and inhibitors of ERK, or an intermediary in this pathway, significantly reduce the severity of LID.43–45 The importance of these pathways is also that they are a point of convergence at which dopaminergic nigrostriatal and glutamatergic corticostriatal inputs to the medium spiny neurons are acting through co-localised dopaminergic and glutamatergic metabotropic receptors. Activation of the RAS/ERK pathway leads to phosphorylation of the GluR1 subunit on AMPA receptors located at the synapse promoting glutamatergic transmission.46
Further still down the cascade, the acetylation and phosphorylation states of histones H3 and H4 are altered (although reports are inconsistent as to the details).47–49 Nevertheless, this is indicative of chromatin rearrangements and thereby changes in transcriptional control. An example of these transcriptional changes is the persistent upregulation of the immediate early gene ΔFosB in the striatum of dyskinetic rodent and non-human primates.11,50,51 This is indicative of the long-term changes that cause the priming phenomena, meaning that dyskinesia can still be evoked after prolonged L-DOPA withdrawal. Experimental evidence suggests that it may be possible to ‘deprime’ the L-DOPA-treated striatum through suppression of these proteins.11,51
Dopamine D3 receptors that are co-expressed with D1 receptors on the direct pathway are increased in animal models of LID and may directly interact with them through intramembrane crosstalk.52 Nevertheless, in vivo reports are conflicting, suggesting either no effect or significant improvement on the effect of manipulating activity at the D3 receptor in LID, which may or may not be at the expense of the antiparkinsonian effect of L-DOPA.53–57
Dopamine D2 receptor agonists are effective at alleviating some of the symptoms of the motor disorder in early stages of the disease, partly because de novo therapies do not commonly induce significant levels of dyskinesia.58,59 Before concluding that D1 receptors are the LID culprits, however, it must be considered that D2 receptor agonists can produce motor sensitisation60 and are capable of inducing dyskinesia expression if L-DOPA priming has already taken place.61 Therefore, D2 receptors are not innocent bystanders concerning LID generation or expression.
Whether there are direct changes in D2 receptor expression in response to chronic L-DOPA administration is unclear,62 but there are indications that D2 receptor mechanisms are altered and as such could be potential therapeutic targets. D2 receptors are negatively coupled through G-protein Gi/o to adenylate cyclase, the activity of the G-protein being controlled by the speed at which guanosine triphosphate is converted back into guanosine diphosphate. Regulators of G-protein signalling (RGS) are GTPases, which mediate this conversion, effectively influencing the speed of inactivation of the active α G-protein subunit. RGS9 is upregulated following chronic L-DOPA administration63 and is thought to reflect an intrinsic compensation to the increase in D2 receptor activity. Despite this, further upregulation compromised the beneficial effects of L-DOPA.63

Microvascular Changes in L-DOPA-induced Dyskinesia

Maladaptive plasticity in LID is not restricted to neuronal activity but also includes the structural microenvironment of the basal ganglia. There is accumulating evidence of changes to the essential microvasculature in PD, accompanying both the progressive dopamine degeneration and the development of LID. Post-mortem studies of human PD brains have demonstrated pathological microvascular changes and altered levels of angiogenic cytokines in the basal ganglia.64–67
Recent experiments in rodents suggest that the microvascular changes may be attributable to dyskinesiogenic D1-receptor stimulation and activation of ERK1/2.43,68 These angiogenic vessels may represent areas with blood–brain barrier leakage and thus local foci of high L-DOPA concentration that may further exacerbate the fluctuations of dopamine. Interestingly, drugs with well-known effects on vascular physiology, such as α2-adrenergic receptor antagonists and nicotine, have well-documented antidyskinetic efficacy in animal models of PD.69–71 This suggests that possible effects on the microvasculature need to be taken into consideration when novel treatments for PD are evaluated.

Non-dopaminergic Modulators of Dyskinesia

As mentioned above, striatal function is not only modulated by nigrostriatal dopaminergic inputs. Indeed glutamatergic, corticostriatal and acetylcholine interneurons also have modulatory influences and post-synaptic dopamine receptors co-localise with other receptors (adenosine, glutamatergic). The activity of the striatum can also be modulated by monoamines apart from dopamine, such as noradrenaline and serotonin. It may also be possible to interfere with signalling in other nuclei of the basal ganglia circuitry, such as the globus pallidus, subthalamic nucleus and substantia nigra pars reticulata (cannabinoids and opioids). Using non-dopaminergic systems to regulate the activity of the basal ganglia is likely to be highly useful in the search for antidyskinetic agents that allow the maintenance of L-DOPA efficacy on motor function.

Glutamate

Both clinical and pre-clinical data point to the importance of the glutamate system in the pathophysiology of dyskinesia.72 Currently, the most commonly used drug for the treatment of dyskinesia is amantadine, a weak glutamatergic antagonist.73
Pre-clinical studies have demonstrated striatal changes in glutamate receptor expression, phosphorylation and synaptic organisation in LID.36,74 Picconi and co-workers recently showed defective corticostriatal synaptic plasticity in dyskinetic 6-OHDA-lesioned rats.75 It was demonstrated that the corticostriatal neurons recorded from rats with LID lost the capability to depotentiate the long-term potentiation previously induced by high-frequency stimulation.75 Lately, changes in the density of NMDA-receptor subunits due to altered trafficking in striatal neurons in LID versus non-dyskinetic rats have been demonstrated. Biochemical analysis of the post-synaptic density revealed increased expression of the NR2B subunit, whereas the levels of NR2A were decreased.76 The reduced expression of NR2B was discovered to be parallel to altered anchoring to the membrane-associated-guanylate kinase (MAGUK) protein family. Treatment of non-dyskinetic rats with a synthetic peptide that disrupted the binding between the NR2B subunit and the MAGUK protein resulted in the expression of abnormal involuntary movements.76
The fast glutamatergic transmission in the striatum from the cortex is also mediated by metabotrophic glutamate receptors (mGluR). Among the different classes of mGluR, the mGluR5 receptor has repeatedly been shown to have antidyskinetic properties.77–81 Antagonising the receptor blocks the maladaptive post-synaptic changes associated with LID, such as upregulation of ΔFosB,77 prodynorphin mRNA78,81 and pERK1/2.81
Abstract overactivity of striatal alpha-amino-3-hydroxy-5-methyl-4- isoxazolepropionic acid (AMPA) glutamate receptors has also been implicated in the expression of LID. Radioligand binding studies in primate models of LID have shown both upregulation82 and no change in striatal AMPA binding.83 A more complex subcellular alteration of striatal AMPA receptor localisation was recently demonstrated by Silverdale and co-workers. AMPA receptors were redistributed from the vesicular fraction and into the post-synaptic membrane during LID.84 Antagonising AMPA receptors attenuated the expression of LID in both MPTP-treated monkeys and 6-OHDA-lesioned rats, as well as blocking the associated molecular correlates of LID.85–87

Opioids

The striatal output neurons are GABAergic but also co-release the opioid peptides encephalin and dynorphin, which act at the mu, delta and kappa opioid receptors. The synthesis of these peptides, dynorphin in particular, is enhanced in dyskinetic states,12,88 while receptor binding and sensitivity is significantly altered.
Conflicting pre-clinical studies report that non-selective opioid receptor antagonists can improve or worsen LID,89,90 while a clinical trial illustrated their potential use as antidyskinetic agents.91 Mu- and delta-specific antagonists showed clear anti-dyskinesia activity in the MPTP-treated marmoset without compromising L-DOPA efficacy, while the kappa antagonist nor-binaltorphimine failed to show any effects.89 In the rodent model, infusion of the kappa receptor antagonist prevented behavioural sensitisation to apomorphine,92 while the kappa-receptor agonist TRK-820 reduced L-DOPA-mediated dyskinesia generation and LID expression in primed animals.93

Adenosine

The neural activity of the basal ganglia can also be modulated by adenosine, acting on adenosine A2A receptors that are abundantly expressed in several basal ganglia nuclei.94 The adenosine A2A receptors located on the dendritic spines of the striato-pallidal GABA neurons form functional complexes with dopamine D2 and metabotropic glutamate 5 (mGlu5) receptors95 and can modulate dopamine-mediated motor functions.
Antagonists of the A2A receptor have been put forward as potential targets for anti-parkinsonian drugs.94 Evidence from the use these antagnoists as antidyskinetic agents is less convincing, as are the clinical trial data obtained thus far. Pharmacological antagonism of the A2A receptor alleviated apomorphine-induced dyskinesia in MPTP-lesioned monkeys96 but the same pharmacological agent had no effect on LID in 6-OHDA-lesioned rats.97 Moreover, clinical trials with the A2A receptor-antagonist istradefylline actually produced a significant worsening of LID when given in conjunction with dopaminergic treatments.98 More recent studies suggest that the addition of an adenosine agonist may lower the dose of L-DOPA required to produce functional efficacy, which may as a consequence reduce LID.99,100

Noradrenaline

It has long been suggested that the effects of L-DOPA in PD are mediated by both the noradrenergic and dopamine systems.101–103 The noradrenergic system has also been implicated in LID.104 The α2-adrenoreceptor antagonists reduce LID in both the MPTP-primate model,70,105,106 and in 6-OHDA-lesioned rats.15,107,108 Unfortunately the antidyskinetic effect in patients with PD is somewhat disputable, with data showing both a reduction of109 and no effect on LID.110 Recently, Buck and co-workers showed that the attenuation of LID by the α2-adrenoceptor antagonist idazoxan in 6-OHDA-lesioned rats was associated with a reduction in striatal extracellular levels of dopamine,111 thus providing a possible mechanistic insight into the antidyskinetic effect of these noradrenergic drugs. New findings from the same investigators suggest that antagonists of the α1-adrenoceptor could also work as potential targets for new antidyskinetic drugs.107

Cannabinoids

The precise contribution of the cannabinoid (CB) system in LID has been difficult to define, since both agonists and antagonists of the CB1-receptor seem to alleviate LID.112 The attenuating effect on LID is well-documented in both animal models of PD113–116 and in patients,117 but mechanistic insights have been lacking. A recent study by Morgese and co-workers demonstrated that treatment with a CB1-receptor agonist alleviated LID in 6-OHDA-lesioned rats.114,115 The improvement in symptoms was inversely correlated with extracellular glutamate levels in the denervated striatum.115 Additionally, pharmacological elevation of endocannabinoid levels only attenuated LID when the transient receptor-potential vanilloid subtype (TRPV1) was blocked, suggesting that stimulation of the CB-receptors and TRPV1 have opposing effects on LID.114

The Translational Significance of Basic Experimental Data

The translation of antiparkinsonian and antidyskinetic agents into clinically successful drugs has been limited in recent years. As highlighted here, neither sarizotan nor istradefylline – both promising in pre-clinical studies – produced the anticipated results in clinical trials. It has been beyond the scope of this article to analyse why in detail, but key factors that may influence this translation are reviewed in detail elsewhere.118,119
While the animal models, in general, accord with each other and clinical post-mortem evaluations, there are issues with both clinical and pre-clinical investigative tools. A recent study of clinical dyskinesia rating scales raised concerns over their reliability and relevance and has recommended only two out of eight rating scales as having excellent clinometric properties in PD.120 There are also examples of conflicting findings using the same pharmacological approaches in pre-clinical studies. These discrepancies may be based on the precise animal model used, the strain, species or extent and selectivity of dopaminergic denervation.
The concept of lesion extent is similarly problematic in clinical trials. This is because patients may have variable degrees of degeneration and particular pharmacological approaches may only be effective at certain stages of the disease, thus only being appropriate for subgroups of patients.
Clinical trials need to take all of these issues into account to fully evaluate the true potential of new agents.

Conclusion

There are new pharmacological targets and strategies under investigation, at the receptor and downstream signalling cascade levels. Furthermore, is also looking towards how non-neural mechanisms influence the microenvironment of the striatum.
There are many proposed targets for treating dyskinesia, but where treatment strategies may require invasive procedures, such as neurosurgery, or risks, such as the use of viral vectors, the clinical benefit of alleviating PD symptoms through the same procedure should be a primary consideration.
Hopefully, advances in treatment will move the field into a new era of dyskinesia management, whether that be treating the dyskinesia or finding ways to avoid its development. ■

Article Information:
Disclosure

The authors have no conflicts of interest to declare.

Correspondence

Emma L Lane, Brain Repair Centre, Department of Biosciences, Cardiff University, Museum Avenue, Cardiff, CF10 3AX, Wales, UK. E: Laneel@cf.ac.uk

Received

2010-09-10T00:00:00

References

  1. Cotzias GC, Van Woert MH, Schiffer LM, Aromatic amino acids and modification of parkinsonism, N Engl J Med, 1967;276(7):374–9.
  2. Marsden CD, Parkinson’s disease, Lancet, 1990;335(8695):948–52.
  3. Nutt JG, Chung KA, Holford NH, Dyskinesia and the antiparkinsonian response always temporally coincide: a retrospective study, Neurology, 2010;74(15):1191–7.
  4. 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.
  5. Goetz CG, Poewe W, Rascol O, et al., Evidence-based medical review update: pharmacological and surgical treatments of Parkinson’s disease: 2001 to 2004, Mov Disord, 2005;20(5):523–39.
  6. Metman LV O’Leary ST, Role of surgery in the treatment of motor complications, Mov Disord, 2005;20(Suppl. 11): S45–56.
  7. Olanow CW, Double-blind, placebo-controlled trials for surgical interventions in Parkinson disease, Arch Neurol, 2005;62(9):1343–4.
  8. Obeso JA, Rodríguez MC, Gorospe A, et al., Surgical treatment of Parkinson’s disease, Baillieres Clin Neurol, 1997;6(1):125–45.
  9. Grandas F, Galiano ML, Tabernero C, Risk factors for levodopa-induced dyskinesias in Parkinson’s disease, J Neurol, 1999;246(12):1127–33.
  10. Schrag A, Quinn N, Dyskinesias and motor fluctuations in Parkinson’s disease. A community-based study, Brain, 2000;123(Pt 11):2297–305.
  11. 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.
  12. Cenci MA, Lee CS, Björklund A, L-DOPA-induced dyskinesia in the rat is associated with striatal overexpression of prodynorphin- and glutamic acid decarboxylase mRNA, Eur J Neurosci, 1998;10(8):2694–706.
  13. Clarke CE, Sambrook MA, Mitchell IJ, et al., Levodopa-induced dyskinesia and response fluctuations in primates rendered parkinsonian with 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), J Neurol Sci, 1987;78(3):273–80.
  14. Kuoppamäki M, Al-Barghouthy G, Jackson MJ, et al., L-dopa dose and the duration and severity of dyskinesia in primed MPTP-treated primates, J Neural Transm, 2007;114(9):1147–53.
  15. Lundblad M, Andersson M, Winkler C, et al., Pharmacological validation of behavioural measures of akinesia and dyskinesia in a rat model of Parkinson’s disease, Eur J Neurosci, 2002;15:120–32.
  16. Lindgren HS, Rylander D, Ohlin KE, et al., The “motor complication syndrome” in rats with 6-OHDA lesions treated chronically with L-DOPA: relation to dose and route of administration, Behav Brain Res, 2007;177(1): 150–9.
  17. Cenci MA, Lindgren HS, Advances in understanding L-DOPA-induced dyskinesia, Curr Opin Neurobiol, 2007;17(6):665–71.
  18. Chase TN, Levodopa therapy: consequences of the nonphysiologic replacement of dopamine, Neurology, 1998;50(5 Suppl. 5):S17–25.
  19. de la Fuente-Fernandez R, Sossi V, Huang Z, et al., Levodopa-induced changes in synaptic dopamine levels increase with progression of Parkinson’s disease: implications for dyskinesias, Brain, 2004;127(Pt 12): 2747–54.
  20. Pavese N, Evans AH, Tai YF, et al., Clinical correlates of levodopa-induced dopamine release in Parkinson disease: a PET study, Neurology, 2006;67(9):1612–7.
  21. Lee J, Zhu WM, Stanic D, et al., Sprouting of dopamine terminals and altered dopamine release and uptake in Parkinsonian dyskinaesia, Brain, 2008;131(Pt 6):1574–87.
  22. Lindgren HS, Andersson DR, Lagerkvist S, et al., 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.
  23. Meissner W, Ravenscroft P, Reese R, et al., Increased slow oscillatory activity in substantia nigra pars reticulata triggers abnormal involuntary movements in the 6-OHDA-lesioned rat in the presence of excessive extracelullar striatal dopamine, Neurobiol Dis, 2006;22(3): 586–98.
  24. Troiano AR, de la Fuente-Fernandez R, Sossi V, et al., PET demonstrates reduced dopamine transporter expression in PD with dyskinesias, Neurology, 2009;72(14):1211–6.
  25. Rylander D, Parent M, O’Sullivan S, et al., Maladaptive plasticity of serotonin axon terminals in L-DOPA-induced dyskinesia, Ann Neurol, 2010; [Epub ahead of print].
  26. Carta M, Carlsson T, Kirik D, et al., Dopamine released from 5-HT terminals is the cause of L-DOPA-induced dyskinesia in parkinsonian rats, Brain, 2007;130(Pt 7): 1819–33.
  27. Bara-Jimenez W, Bibbiani F, Morris MJ, et al., Effects of serotonin 5-HT1A agonist in advanced Parkinson’s disease, Mov Disord, 2005;20(8):932–6.
  28. Bartoszyk GD, Van Amsterdam C, Greiner HE, et al., Sarizotan, a serotonin 5-HT1A receptor agonist and dopamine receptor ligand. 1. Neurochemical profile, J Neural Transm, 2004;111(2):113–26.
  29. Bronzova J, Sampaio C, Hauser RA, et al., Double-blind study of pardoprunox, a new partial dopamine agonist, in early Parkinson’s disease, Mov Disord, 2010;25(6):730–8.
  30. Stocchi F, The therapeutic concept of continuous dopaminergic stimulation (CDS) in the treatment of Parkinson’s disease, Parkinsonism Relat Disord, 2009;15 (Suppl. 3):S68–71.
  31. Stowe R, Ives N, Clarke CE, et al., Evaluation of the efficacy and safety of adjuvant treatment to levodopa therapy in Parkinson’s disease patients with motor complications, Cochrane Database Syst Rev, 2010;7:CD007166.
  32. Talati R, Reinhart K, Baker W, et al., Pharmacologic treatment of advanced Parkinson’s disease: a meta-analysis of COMT inhibitors and MAO-B inhibitors, Parkinsonism Relat Disord, 2009;15(7):500–5.
  33. Carlsson T, Björklund T, Kirik D, Restoration of the striatal dopamine synthesis for Parkinson’s disease: viral vector-mediated enzyme replacement strategy, Curr Gene Ther, 2007;7(2):109–20.
  34. Björklund T, Carlsson T, Cederfjäll EA, et al., Optimized adeno-associated viral vector-mediated striatal DOPA delivery restores sensorimotor function and prevents dyskinesias in a model of advanced Parkinson’s disease, Brain, 2010;133(Pt 2):496–511.
  35. Jarraya B, Boulet S, Ralph GS, et al., Dopamine gene therapy for Parkinson’s disease in a nonhuman primate without associated dyskinesia, Sci Transl Med, 2009;1(2):2ra4.
  36. Cenci MA, Konradi C, Maladaptive striatal plasticity in L-DOPA-induced dyskinesia, Prog Brain Res, 2010;183: 209–33.
  37. Calon, F, Grondin R, Morissette M, et al., Molecular basis of levodopa-induced dyskinesias, Ann Neurol, 2000; 47(4 Suppl. 1):S70–8.
  38. Henry B, Duty S, Fox SH, et al., Increased striatal pre-proenkephalin B expression is associated with dyskinesia in Parkinson’s disease, Exp Neurol, 2003;183(2):458–68.
  39. Konradi C, Westin JE, Carta M, et al., Transcriptome analysis in a rat model of L-DOPA-induced dyskinesia, Neurobiol Dis, 2004;17(2):219–36.
  40. Westin JE, Vercammen L, Strome EM, et al., 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.
  41. Bezard E, Gross CE, Qin L, et al., 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.
  42. Ahmed MR, Berthet A, Bychkov E, et al., Lentiviral overexpression of GRK6 alleviates L-dopa-induced dyskinesia in experimental Parkinson’s disease, Sci Transl Med, 2010;2(28):28ra28.
  43. Lindgren HS, Ohlin KE, Cenci ME, 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.
  44. Santini E, Valjent E, Usiello A, et al., Critical involvement of cAMP/DARPP-32 and extracellular signal-regulated protein kinase signaling in L-DOPA-induced dyskinesia, J Neurosci, 2007;27(26):6995–7005.
  45. Schuster S, Nadjar A, Guo JT, et al., The 3-hydroxy-3- methylglutaryl-CoA reductase inhibitor lovastatin reduces severity of L-DOPA-induced abnormal involuntary movements in experimental Parkinson’s disease, J Neurosci, 2008;28(17):4311–6.
  46. Qin Y, Zhu Y, Baumgart JP, et al., State-dependent Ras signaling and AMPA receptor trafficking, Genes Dev, 2005;19(17):2000–15.
  47. Darmopil S, Martin AB, De Diego IR, et al., Genetic inactivation of dopamine D1 but not D2 receptors inhibits L-DOPA-induced dyskinesia and histone activation, Biol Psychiatry, 2009;66(6):603–13.
  48. Nicholas AP, Lubin FD, Hallett PJ, et al., Striatal histone modifications in models of levodopa-induced dyskinesia, J Neurochem, 2008;106(1):486–94.
  49. Santini E, Alcacer C, Cacciatore S, et al., L-DOPA activates ERK signaling and phosphorylates histone H3 in the striatonigral medium spiny neurons of
  50. Doucet JP, Nakabeppu Y, Bedard PJ, et al., Chronic alterations in dopaminergic neurotransmission produce a persistent elevation of deltaFosB-like protein(s) in both the rodent and primate striatum, Eur J Neurosci, 1996;8(2):365–81.
  51. Berton O, Guigoni C, Li Q, et al., Striatal overexpression of DeltaJunD resets L-DOPA-induced dyskinesia in a primate model of Parkinson disease, Biol Psychiatry, 2009;66(6): 554–61.
  52. Guigoni C, Aubert I, Li Q, et al., Pathogenesis of levodopa-induced dyskinesia: focus on D1 and D3 dopamine receptors, Parkinsonism Relat Disord, 2005;11(Suppl. 1):S25–9.
  53. Hsu A, Togasaki DM, Bezard E, et al., Effect of the D3 dopamine receptor partial agonist BP897 [N-[4-(4-(2- methoxyphenyl)piperazinyl)butyl]-2-naphthamide] on L-3,4-dihydroxyphenylalanine-induced dyskinesias and parkinsonism in squirrel monkeys, J Pharmacol Exp Ther, 2004;311(2):770–7.
  54. Kumar R, Riddle L, Griffin SA, et al., Evaluation of the D3 dopamine receptor selective antagonist PG01037 on Ldopa- dependent abnormal involuntary movements in rats, Neuropharmacology, 2009;56(6-7):944–55.
  55. Mela F, Millan MJ, Brocco M, et al., The selective D(3) receptor antagonist, S33084, improves parkinsonian-like motor dysfunction but does not affect L-DOPA-induced dyskinesia in 6-hydroxydopamine hemi-lesioned rats, Neuropharmacology, 2010;58(2):528–36.
  56. Silverdale MA, Nicholson SL, Ravenscroft P, et al., Selective blockade of D(3) dopamine receptors enhances the anti-parkinsonian properties of ropinirole and levodopa in the MPTP-lesioned primate, Exp Neurol, 2004;188(1):128–38.
  57. Visanji NP, Fox SH, Johnston T, et al., Dopamine D3 receptor stimulation underlies the development of LDOPA- induced dyskinesia in animal models of Parkinson’s disease, Neurobiol Dis, 2009;35(2):184–92.
  58. Lees AJ, Katzenschlager R, Head J, et al., Ten-year follow-up of three different initial treatments in de-novo PD: a randomized trial, Neurology, 2001;57(9):1687–94.
  59. Pearce RK, Banerji T, Jenner P, et al., De novo administration of ropinirole and bromocriptine induces less dyskinesia than L-dopa in the MPTP-treated marmoset, Mov Disord, 1998;13(2):234–41.
  60. Henry B, Crossman AR, Brotchie JM, Characterization of enhanced behavioral responses to L-DOPA following repeated administration in the 6-hydroxydopaminelesioned rat model of Parkinson’s disease, Exp Neurol, 1998;151(2):334–42.
  61. Gomez-Mancilla B, Bédard PJ, Effect of D1 and D2 agonists and antagonists on dyskinesia produced by L-dopa in 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridinetreated monkeys, J Pharmacol Exp Ther, 1991;259(1):409–13.
  62. Berthet A, Bezard E, Dopamine receptors and L-dopa-induced dyskinesia, Parkinsonism Relat Disord, 2009;15(Suppl. 4):S8–-12.
  63. Gold SJ, Hoang CV, Potts BW, et al., RGS9-2 negatively modulates L-3,4-dihydroxyphenylalanine-induced dyskinesia in experimental Parkinson’s disease, J Neurosci, 2007;27(52):14338–48.
  64. Farkas E, De Jong GI, de Vos RA, et al., Pathological features of cerebral cortical capillaries are doubled in Alzheimer’s disease and Parkinson’s disease, Acta Neuropathol, 2000;100(4):395–402.
  65. Faucheux BA, Bonnet AM, Agid Y, et al., Blood vessels change in the mesencephalon of patients with Parkinson’s disease, Lancet, 1999;353(9157):981–2.
  66. Wada K, Arai H, Takanashi M, et al., Expression levels of vascular endothelial growth factor and its receptors in Parkinson’s disease, Neuroreport, 2006;17(7):705–9.
  67. Yasuda T, Fukuda-Tani M, Nihira T, et al., Correlation between levels of pigment epithelium-derived factor and vascular endothelial growth factor in the striatum of patients with Parkinson’s disease, Exp Neurol, 2007;206(2):308–17.
  68. Westin JE, Lindgren HS, Gardi J, et al., Endothelial proliferation and increased blood-brain 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.
  69. Bordia T, Campos C, Huang L, et al., Continuous and intermittent nicotine treatment reduces L-3,4-dihydroxyphenylalanine (L-DOPA)-induced dyskinesias in a rat model of Parkinson’s disease, J Pharmacol Exp Ther, 2008;327(1):239–47.
  70. Fox SH, Henry B, Hill MP, et al., Neural mechanisms underlying peak-dose dyskinesia induced by levodopa and apomorphine are distinct: evidence from the effects of the alpha(2) adrenoceptor antagonist idazoxan, Mov Disord, 2001;16(4):642–50.
  71. Quik M, Cox H, Parameswaran N, et al., Nicotine reduces levodopa-induced dyskinesias in lesioned monkeys, Ann Neurol, 2007;62(6):588–96.
  72. Stocchi F, Tagliati M, Olanow CW, Treatment of levodopa-induced motor complications, Mov Disord, 2008;23(Suppl. 3):S599–612.
  73. Chase TN, Bibbiani F, Oh JD, Striatal glutamatergic mechanisms and extrapyramidal movement disorders, Neurotox Res, 2003;5(1-2):139–46.
  74. Chase TN, Oh JD, Konitsiotis S, Antiparkinsonian and antidyskinetic activity of drugs targeting central glutamatergic mechanisms, J Neurol, 2000;247 (Suppl. 2):II36–42.
  75. Picconi B, Centonze D, Håkansson K, et al., Loss of bidirectional striatal synaptic plasticity in L-DOPAinduced dyskinesia, Nat Neurosci, 2003;6(5):501–6.
  76. Gardoni F, Picconi B, Ghiglieri V, et al., A critical interaction between NR2B and MAGUK in L-DOPA induced dyskinesia, J Neurosci, 2006;26(11):2914–22.
  77. Levandis G, Bazzini E, Armentero MT, et al., Systemic administration of an mGluR5 antagonist, but not unilateral subthalamic lesion, counteracts l-DOPAinduced dyskinesias in a rodent model of Parkinson’s disease, Neurobiol Dis, 2008;29(1):161–8.
  78. Mela F, Marti M, Dekundy A, et al., 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.
  79. Morin N, Grégoire L, Gomez-Mancilla B, et al., Effect of the metabotropic glutamate receptor type 5 antagonists MPEP and MTEP in parkinsonian monkeys, Neuropharmacology, 2010;58(7):981–6.
  80. Rylander D, Iderberg H, Li Q, et al., A mGluR5 antagonist under clinical development improves L-DOPA-induced dyskinesia in parkinsonian rats and monkeys, Neurobiol Dis, 2010;39(3):352–61.
  81. Rylander D, Recchia A, Mela F, et al., 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):227–3
  82. Calon F, Morissette M, Ghribi O, et al., Alteration of glutamate receptors in the striatum of dyskinetic 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-treated monkeys following dopamine agonist treatment, Prog Neuropsychopharmacol Biol Psychiatry, 2002;26(1):127–38.
  83. Silverdale MA, Crossman AR, Brotchie JM, Striatal AMPA receptor binding is unaltered in the MPTP-lesioned macaque model of Parkinson’s disease and dyskinesia, Exp Neurol, 2002;174(1):21–8.
  84. Silverdale MA, Kobylecki C, Hallett PJ, et al., Synaptic recruitment of AMPA glutamate receptor subunits in levodopa-induced dyskinesia in the MPTP-lesioned nonhuman primate, Synapse, 2010;64(2):177–80.
  85. Bibbiani F, Oh JD, Kielaite A, et al., Combined blockade of AMPA and NMDA glutamate receptors reduces levodopa-induced motor complications in animal models of PD, Exp Neurol, 2005;196(2):422–9.
  86. Kobylecki C, Cenci MA, Crossman AR, et al., Calcium-permeable AMPA receptors are involved in the induction and expression of L-DOPA-induced dyskinesia in Parkinson’s disease, J Neurochem, 2010;114(2):499–511.
  87. Konitsiotis S, Blanchet PJ, Verhagen L, et al., AMPA receptor blockade improves levodopa-induced dyskinesia in MPTP monkeys, Neurology, 2000;54(8): 1589–95.
  88. Johansson PA, Andersson M, Andersson KE, et al., Alterations in cortical and basal ganglia levels of opioid receptor binding in a rat model of l-DOPA-induced dyskinesia, Neurobiol Dis, 2001;8(2):220–39.
  89. Henry B, FoxSH, Crossman AR, et al., Mu- and delta-opioid receptor antagonists reduce levodopainduced dyskinesia in the MPTP-lesioned primate model of Parkinson’s disease, Exp Neurol, 2001;171(1):139–46.
  90. Samadi P, Grégoire L, Bédard PJ, Opioid antagonists increase the dyskinetic response to dopaminergic agents in parkinsonian monkeys: interaction between dopamine and opioid systems, Neuropharmacology, 2003;45(7): 954–63.
  91. . Fox S, Silverdale M, Kellett M, 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.
  92. Newman DD, Rajakumar N, Flumerfelt BA, et al., A kappa opioid antagonist blocks sensitization in a rodent model of Parkinson’s disease, Neuroreport, 1997;8(3):669–72.
  93. Ikeda K, Yoshikawa S, Kurokawa T, et al., TRK-820, a selective kappa opioid receptor agonist, could effectively ameliorate L-DOPA-induced dyskinesia symptoms in a rat model of Parkinson’s disease, Eur J Pharmacol, 2009;620 (1-3):42–8.
  94. Morelli M, Di Paolo T, Wardas J, et al., Role of adenosine A2A receptors in parkinsonian motor impairment and l-DOPA-induced motor complications, Prog Neurobiol, 2007;83(5):293–309.
  95. Agnati LF, Ferré S, Lluis C, et al., Molecular mechanisms and therapeutical implications of intramembrane receptor/receptor interactions among heptahelical receptors with examples from the striatopallidal GABA
  96. Bibbiani F, Oh JD, Petzer JP, et al., A2A antagonist prevents dopamine agonist-induced motor complications in animal models of Parkinson’s disease, Exp Neurol, 2003;184(1):285–94.
  97. Lundblad M, Vaudano E, Cenci MA, Cellular and behavioural effects of the adenosine A2a receptor antagonist KW-6002 in a rat model of l-DOPA-induced dyskinesia, J Neurochem, 2003;84(6):1398–410.
  98. Hauser RA, Hubble JP, Truong DD, Randomized trial of the adenosine A(2A) receptor antagonist istradefylline in advanced PD, Neurology, 2003;61(3):297–303.
  99. Rose S, Jackson MJ, Smith LA, et al., The novel adenosine A2a receptor antagonist ST1535 potentiates the effects of a threshold dose of L-DOPA in MPTP treated common marmosets, Eur J Pharmacol, 2006;546 (1-3):82–7.
  100. Rose S, Ramsay Croft N, Jenner P, The novel adenosine A2a antagonist ST1535 potentiates the effects of a threshold dose of l-dopa in unilaterally 6-OHDA-lesioned rats, Brain Res, 2007;1133(1):110–4.
  101. Andén NE, Strömbom U, Svensson TH, Locomotor stimulation by L-dopa: relative importance of noradrenaline receptor activation, Psychopharmacology (Berl), 1977;54(3):243–8.
  102. Dolphin A, Jenner P, Marsden CD, Noradrenaline synthesis from L-DOPA in rodents and its relationship to motor activity, Pharmacol Biochem Behav, 1976;5(4): 431–9.
  103. Dolphin AC, Jenner P, Marsden CD, The relative importance of dopamine and noradrenaline receptor stimulation for the restoration of motor activity in reserpine or alpha-methyl-p-tyrosine pre-treated mice, Pharmacol Biochem Behav, 1976;4(6):661–70.
  104. Colosimo C, Craus A, Noradrenergic drugs for levodopa-induced dyskinesia, Clin Neuropharmacol, 2003;26(6):299–305.
  105. Grondin R, Hadj Tahar A, Doan VD, et al., Noradrenoceptor antagonism with idazoxan improves Ldopa- induced dyskinesias in MPTP monkeys, Naunyn Schmiedebergs Arch Pharmacol, 2000;361(2):181–6.
  106. Savola JM, Hill M, Engstrom M, 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.
  107. Buck K, Ferger B, The selective alpha1 adrenoceptor antagonist HEAT reduces L-DOPA-induced dyskinesia in a rat model of Parkinson’s disease, Synapse, 2010;64(2): 117–26.
  108. Dekundy A, Lundblad M, Danysz W, et al., Modulation of L-DOPA-induced abnormal involuntary movements by clinically tested compounds: further validation of the rat dyskinesia model, Behav Brain Res, 2007;179(1):76–89.
  109. Rascol O, Arnulf I, Peyro-Saint Paul H, 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.
  110. Manson AJ, Iakovidou E, Lees AJ, Idazoxan is ineffective for levodopa-induced dyskinesias in Parkinson’s disease, Mov Disord, 2000;15(2):336–7.
  111. Buck K, Voehringer P, Ferger B, The alpha(2) adrenoceptor antagonist idazoxan alleviates L-DOPAinduced dyskinesia by reduction of striatal dopamine levels: an in vivo microdialysis study in 6- hydroxydopamine-lesioned rats, J Neurochem, 2010;112(2):444–52.
  112. Brotchie JM, CB1 cannabinoid receptor signalling in Parkinson’s disease, Curr Opin Pharmacol, 2003;3(1):54–61.
  113. Fox SH, Henry B, Hill B, et al., 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.
  114. Morgese MG, Cassano T, Cuomo V, et al., Anti-dyskinetic effects of cannabinoids in a rat model of Parkinson’s disease: role of CB(1) and TRPV1 receptors, Exp Neurol, 2007;208(1):110–9.
  115. Morgese MG, Cassano T, Gaetani S, et al., Neurochemical changes in the striatum of dyskinetic rats after administration of the cannabinoid agonist WIN55,212-2, Neurochem Int, 2009;54(1):56–64.
  116. van der Stelt M, Fox SH, Hill M, et al., A role for endocannabinoids in the generation of parkinsonism and levodopa-induced dyskinesia in MPTP-lesioned non-human primate models of Parkinson’s disease, FASEB J, 2005;19(9):1140–2.
  117. Sieradzan KA, Fox SH, Hill M, et al., Cannabinoids reduce levodopa-induced dyskinesia in Parkinson’s disease: a pilot study, Neurology, 2001;57(11):2108–11.
  118. Fox SH, Lang AE, Brotchie JM, Translation of nondopaminergic treatments for levodopa-induced dyskinesia from MPTP-lesioned nonhuman primates to phase IIa clinical studies: keys to success and roads to failure, Mov Disord, 2006;21(10):1578–94.
  119. Lane E, Dunnett S, Animal models of Parkinson’s disease and L-dopa induced dyskinesia: how close are we to the clinic?, Psychopharmacology (Berl), 2008;199(3):303–12.
  120. Colosimo C, Martinez-Martin P, Fabbrini G, et al., Task force report on scales to assess dyskinesia in Parkinson’s disease: Critique and recommendations, Mov Disord, 2010;25(9):1131–42.

Further Resources

Share this Article
Related Content In Parkinson's Disease
  • Copied to clipboard!
    accredited arrow-down-editablearrow-downarrow_leftarrow-right-bluearrow-right-dark-bluearrow-right-greenarrow-right-greyarrow-right-orangearrow-right-whitearrow-right-bluearrow-up-orangeavatarcalendarchevron-down consultant-pathologist-nurseconsultant-pathologistcrosscrossdownloademailexclaimationfeedbackfiltergraph-arrowinterviewslinkmdt_iconmenumore_dots nurse-consultantpadlock patient-advocate-pathologistpatient-consultantpatientperson pharmacist-nurseplay_buttonplay-colour-tmcplay-colourAsset 1podcastprinter scenerysearch share single-doctor social_facebooksocial_googleplussocial_instagramsocial_linkedin_altsocial_linkedin_altsocial_pinterestlogo-twitter-glyph-32social_youtubeshape-star (1)tick-bluetick-orangetick-red tick-whiteticktimetranscriptup-arrowwebinar Sponsored Department Location NEW TMM Corporate Services Icons-07NEW TMM Corporate Services Icons-08NEW TMM Corporate Services Icons-09NEW TMM Corporate Services Icons-10NEW TMM Corporate Services Icons-11NEW TMM Corporate Services Icons-12Salary £ TMM-Corp-Site-Icons-01TMM-Corp-Site-Icons-02TMM-Corp-Site-Icons-03TMM-Corp-Site-Icons-04TMM-Corp-Site-Icons-05TMM-Corp-Site-Icons-06TMM-Corp-Site-Icons-07TMM-Corp-Site-Icons-08TMM-Corp-Site-Icons-09TMM-Corp-Site-Icons-10TMM-Corp-Site-Icons-11TMM-Corp-Site-Icons-12TMM-Corp-Site-Icons-13TMM-Corp-Site-Icons-14TMM-Corp-Site-Icons-15TMM-Corp-Site-Icons-16TMM-Corp-Site-Icons-17TMM-Corp-Site-Icons-18TMM-Corp-Site-Icons-19TMM-Corp-Site-Icons-20TMM-Corp-Site-Icons-21TMM-Corp-Site-Icons-22TMM-Corp-Site-Icons-23TMM-Corp-Site-Icons-24TMM-Corp-Site-Icons-25TMM-Corp-Site-Icons-26TMM-Corp-Site-Icons-27TMM-Corp-Site-Icons-28TMM-Corp-Site-Icons-29TMM-Corp-Site-Icons-30TMM-Corp-Site-Icons-31TMM-Corp-Site-Icons-32TMM-Corp-Site-Icons-33TMM-Corp-Site-Icons-34TMM-Corp-Site-Icons-35TMM-Corp-Site-Icons-36TMM-Corp-Site-Icons-37TMM-Corp-Site-Icons-38TMM-Corp-Site-Icons-39TMM-Corp-Site-Icons-40TMM-Corp-Site-Icons-41TMM-Corp-Site-Icons-42TMM-Corp-Site-Icons-43TMM-Corp-Site-Icons-44TMM-Corp-Site-Icons-45TMM-Corp-Site-Icons-46TMM-Corp-Site-Icons-47TMM-Corp-Site-Icons-48TMM-Corp-Site-Icons-49TMM-Corp-Site-Icons-50TMM-Corp-Site-Icons-51TMM-Corp-Site-Icons-52TMM-Corp-Site-Icons-53TMM-Corp-Site-Icons-54TMM-Corp-Site-Icons-55TMM-Corp-Site-Icons-56TMM-Corp-Site-Icons-57TMM-Corp-Site-Icons-58TMM-Corp-Site-Icons-59TMM-Corp-Site-Icons-60TMM-Corp-Site-Icons-61TMM-Corp-Site-Icons-62TMM-Corp-Site-Icons-63TMM-Corp-Site-Icons-64TMM-Corp-Site-Icons-65TMM-Corp-Site-Icons-66TMM-Corp-Site-Icons-67TMM-Corp-Site-Icons-68TMM-Corp-Site-Icons-69TMM-Corp-Site-Icons-70TMM-Corp-Site-Icons-71TMM-Corp-Site-Icons-72