Parkinson s disease (PD) is a neurodegenerative. The Nonmotor Symptoms of Parkinson s Disease: Update on Diagnosis and Treatment

Clinical Review The Nonmotor Symptoms of Parkinson’s Disease: Update on Diagnosis and Treatment Amber D. Van Laar, MD, and Samay Jain, MD, MS ABSTRAC...
2 downloads 3 Views 1MB Size
Clinical Review

The Nonmotor Symptoms of Parkinson’s Disease: Update on Diagnosis and Treatment Amber D. Van Laar, MD, and Samay Jain, MD, MS ABSTRACT • Objective: To review the prevalence, diagnosis, and treatment of the nonmotor symptoms (NMS) associated with Parkinson’s disease (PD). • Methods: Narrative review of the literature. • Results: The NMS of PD are becoming increasingly recognized as having a critical role in the impact of this neurodegenerative movement disorder. This has led to significant investigative efforts to identify new or better NMS therapies. The preponderance of PD patients will be diagnosed with 1 or multiple NMS during the course of their disease, with many of these symptoms occurring months or even years prior to receiving the PD diagnosis. Despite the high prevalence and impact on disease burden, NMS often go undetected due to a lack of reporting by patients or insufficient interrogation by physicians. Further complicating NMS management is that only a few therapies have the level of evidence needed to support their use in the treatment of NMS. • Conclusion: The practitioner needs to be aware of NMS and conduct thorough patient questioning in order to recognize, diagnose, and address NMS in PD patients.

P

arkinson’s disease (PD) is a neurodegenerative movement disorder with an estimated prevalence of 1% to 2% among the population over the age of 65 years [1]. Recognition and clinical diagnosis of PD is primarily made based on the cardinal motor features, including rigidity, tremor, bradykinesia, and postural instability. The motor symptoms are neuropathologically associated with accumulation of alpha-synuclein with Lewy body formation and neurodegeneration of the nigrostriatal dopamine system. Postmortem evaluation of the brains of PD patients has revealed more widespread degeneration in nondopaminergic systems, including

www.jcomjournal.com

several brainstem nuclei (raphe nucleus, locus ceruleus, dorsal vagal nucleus), limbic and neocortical structures, as well as the peripheral autonomic system [2,3]. The nonmotor symptoms (NMS) of PD are the clinical manifestations of this extensive degeneration, which suggests that NMS are intrinsic and fundamental features of PD. NMS are exceedingly common, and up to 90% of PD patients will experience nonmotor features, including depression, anxiety, sleep disturbances, cognitive impairment, and dysautonomia [4,5] (Table). NMS have a greater impact on quality of life as compared to the motor symptoms [6,7], but are frequently underrecognized [8]. Evidence suggests that unless there is systematic and specific interrogation by practioners, NMS will elude recognition [9–11]. Recognizing NMS as part of PD is complicated by the fact that these symptoms are common in the general population and not specific for PD [12,13]. NMS can occur at any stage of the disease and may predate diagnosis [12], although as PD progresses the NMS become more prevalent, with a greater impact on health care costs and institutionalization rates than motor features [14,15]. NEUROPSYCHIATRIC SYMPTOMS

Depression Epidemiology and Diagnosis Depression is one of the most common neuropsychiatric manifestations observed in PD patients, with prevalence reports between 4% and 72%, though likely to be closer to 30% to 45% [16–20]. The severity of depression in the PD population has been shown to be greater than in patients with matched chronic disabilities [21,22] and also greater than in the general population over the age

From the Department of Neurology, Movement Disorders Division, University of Pittsburgh Medical Center, Pittsburgh, PA. Vol. 21, No. 2 February 2014 JCOM 71

Nonmotor Symptoms of Parkinson’s Disease Table. Nonmotor Features of Parkinson’s Disease Feature

Pharmacologic Treatment

Nonpharmacologic Treatment

Pathophysiology

Prevalence

Symptomatology

Depression

Degeneration of locus ceruleus and raphe nucleus [3]

30%–45%

Somatic complaints, appetite/sleep changes, fatigue, etc.

Pramipexole [32,33], SSRIs [27], TCAs [30]

Anxiety

Dysfunction of NA, LC, and limbic structures [39]

40%

Generalized anxiety, panic, phobias

SSRIs [42], adjust DA therapy [39]



Hallucinations

Hypersensitization of DA receptors; dysfunction of visual association cortex; lesion of PPN, LC, RN, and superior parietal [176]

20%–40%

Visual hallucinations (animal passing in peripheral vision), tactile hallucination

↓DA-agonists, clozapine [54,55], quetiapine [53,54]



ICD/DDS/ punding

DA agonist use; ↑activity in VTA, NA, mesolimbic DA system [75]

15%–20%

Compulsive behaviors; overuse of levodopa; obsessive handling of objects

↓DA agonists; amantadine for gambling [79], SSRI or atypical antipsychotic for punding [83]



Dementia

Dysfunction ventromedial prefrontal, limbic system, amygdala [177]

20%–40%

Cognitive impairment

Rivastigmine [69], memantine [70]

Apathy

Altered cingulate and dopaminergic connections [178]

30%–50%

Amotivation, indifference, lack of spontaneous activity

Occasional improvement with DA therapy [90]





Neuropsychiatric Cognitive behavioral therapy [37], physical activity? [38], ECT [35]

Physical activity [72], cognitive training [73]

Sleep Disorders REM behavior disorder

Degeneration w/ LBs in 25%–50% LC, PPN, reticulospinal tracts, medial medulla [3]

Dream enactment, vocalizations

Clonazepam [97], melatonin [99]

Restless leg syndrome/periodic limb movements in sleep

Unclear; iron deficiency in SN [179,180]

< 10%–50%

Urge to move legs, discomfort

Pramipexole [110], ropinirole [112], gabapentin enacarbil [115,116], opioids [107]

Insomnia

Dysregulation of raphe nucleus, LC, and PPN [181]

37%–88%

Difficulty with sleep initiation and maintenance

Eszopiclone [119], Good sleep hygiene zolpidem [120], benzodiazepines [120], doxepin [121]

Excessive daytime sleepiness/ fatigue

DA agonist use; neurodegeneration of LC, PPN, RAS [182]

15%–50%

Daytime somnolence, sleep attacks; excessive tiredness

Modafinil [126,127]

Obstructive sleep apnea

Unclear, mechanical obstruction

0%–50%

Snoring, daytime fatigue

of 65 years [23]. The onset of depression can occur at any stage of the disease, even predating the diagnosis. Additionally, depression has more than twice the impact on health status than motor symptoms [24]. Though the mechanisms are not fully understood, it is suspected that psychosocial as well as neuropathological 72 JCOM February 2014 Vol. 21, No. 2



Good sleep hygiene, exercise, leg massage [107]

Stimulating environment, bright light exposure, mild exercise [117] CPAP application

changes contribute to the pathogenesis of depression in PD. In a study comparing 104 PD patients and 61 patients with equivalent disability scores, functional disability was found to be responsible for only 9% of the variation of depression scores [22]. The increased prevalence of depression in PD patients can in part be explained www.jcomjournal.com

Clinical Review Table. Nonmotor Features of Parkinson’s Disease, continued Feature

Pharmacologic Treatment

Nonpharmacologic Treatment

Pathophysiology

Prevalence

Symptomatology

Orthostatic hypotension

Degeneration of autonomic centers: ventrolateral medulla, nucleus tractus solitarius, descending para/sympathetics [3,183]

30%–60%

Dizziness, drowsiness, palpitations, nausea

Fludricortisone [141], midodrine[143], pyridostigmine [142], droxidopa [144]

Compression stockings, sleep w/ head elevated, ↑salt and water intake, frequent small meals, ↓carbohydrate [140,141]

Constipation/ gastroparesis

Loss of enteric DA cells, degeneration dorsal vagal nuclei [155]

60%

↓# of BMs; early satiety

Macrogol [149], lubiprostone [150]

Dietary modification, mild exercise, ↑fluid intake

Urinary disturbance

Detrusor hyperactivity: disinhibition of micturition centers in pons [15]

50%

Urinary urgency, incontinence

Anticholinergics (oxybutynin, solifenacin, tolterodine) [164], botulinum injections [165]



Erectile dysfunction

Hypothalamic dysfunction and dysregulation of DA-oxytocin pathway [166]

60%

Inability to maintain an erection

Phosphodiesterase inhibitors (sildenafil, vardenafil, tadarafil) [166,168], apomorphine [169]



Sialorrhea

Impaired handling of oral secretions [157]

75%

Drooling, impaired speaking/eating

Botulinum injections [158,160], glycopyrrolate [161]

Gum chewing [162]

NSAIDs, adjust DA therapy or add CR formulation [173,174]

Physical therapy, exercise [173], DBS [174,175]

Dysautonomia

Sensory Pain

BG dysfunction in modulation of sensory input; altered 5-HT pathways [184]

30%–85%

Musculoskeletal/ shoulder pain, dystonic pain, vague discomfort

Hyposmia

Degeneration of olfactory structures and amygdala [3,185]

40%–100% [20]

Inability to distinguish odors





Visual disturbances

Retinal DA dysfunction, degeneration of visual cortex [186]

30%–40% [186]

Impaired contrast sensitivity, blurred vision, diplopia





BG = basal ganglia; DA = dopamine; DDS = dopamine dysregulation syndrome; ICD = impulse control disorders; LB = Lewy bodies; LC = locus ceruleus; NA = nucleus acumbens; PPN = pedunculopontine nucleus; RAS = reticular activating system; SN = substantia nigra; VTA = ventral tegmental area.

by the neuropathological changes seen in post-mortem studies. Two neurotransmitters that are fundamental in the pathogenesis of depression are serotonin, from the raphe nuclei, and norepinepherine, from the locus ceruleus [20]. Both of these brainstem structures demonstrate alpha-synucleinopathy-associated degeneration and these changes can precede the development of motor dysfunction [3]. Diagnosing depression in PD is complicated by the fact that there is overlap between other PD symptoms and clinical features of depression (ie, amotivation, bradykinesia, fatigue, and sleep disturbances). However, many www.jcomjournal.com

depressed PD patients are less likely to report feelings of guilt or failure and tend to have higher rates of anxiety [9,20,25]. Typically, PD patients are more likely to be diagnosed with minor depression or dysthymia rather than a major depressive disorder [19,20]. Formal testing through systematic questionnaires are diagnostically useful in the clinic, and serial testing can reveal changes over time to guide more effective treatment. Validated tools to evaluate depression in PD include the Beck Depression Inventory, Hamilton Depression Rating Scale, Montgomery-Asberg Depression Rating Scale, Geriatric DRS, and Hospital Anxiety and Depression scale [20].

Vol. 21, No. 2 February 2014 JCOM 73

Nonmotor Symptoms of Parkinson’s Disease Treatment Options Treatment of depression in PD demonstrates generally poorer responses to typical antidepressants and side effects that may worsen other PD symptoms. Selective serotonin reuptake inhibitors (SSRIs) have been widely used as there are generally few drug-drug interactions and minimal effect on motor symptoms; however, several studies have demonstrated little benefit on depression in PD [26]. In a randomized, double-blind, placebo-controlled trial of the antidepressants paroxetine and venlafaxine, both were found to be effective and well tolerated [27]. Tricyclic antidepressants (TCAs) have also demonstrated efficacy. In randomized controlled trials comparing TCAs to SSRIs, a greater benefit on depression symptoms has been found with TCAs [28–30]. The use of TCAs, however, is limited by anticholinergic side effects that occasionally worsen orthostatic hypotension or cognitive impairment [15,31]. Dopamine agonists have also been studied in depressed PD patients. In a randomized, double-blind, placebocontrolled trial [32] and a prospective observational study [33], pramipexole demonstrated significant improvements in depression symptoms. Ropinirole also demonstrated significant symptomatic improvement [34]. These studies suggest that while SSRIs are commonly used, evidence is accumulating to support the role of TCAs, SNRIs, and dopamine agonists in the treatment of depression in PD. Other therapies have also been tried in pharmacologicresistant patients. Electroconvulsive therapy has been reported to improve both depression and motor symptoms [35,36]; however, this is a treatment reserved for patients with severe and drug-refractory depression. A randomized controlled trial investigating cognitive behavioral therapy has also demonstrated improvement of depression scores [37]. The role of physical activity as treatment for depression in PD patients is unclear. As described in a recent review by Loprinzi et al [38], the literature is contradictory, with one group experiencing reduced depression but with no signficant effect in several other studies.

Anxiety Epidemiology and Diagnosis The prevalence of anxiety in PD patients is about 40% [39], which is 2 times greater than in the general population [9]. Anxiety may worsen PD symptoms, especially tremor and cognition. Risk factors for anxiety include the female gender, greater motor fluctuations, prior history of anxiety, and younger age of PD onset [40]. As with 74 JCOM February 2014 Vol. 21, No. 2

depression, some patients also report worsening of anxious symptoms during “off” states [41]. Screening tools that have been validated to help practitioners identify anxiety in PD include the Hospital Anxiety and Depression Scale, Beck Anxiety Inventory, Zung Self-rating Anxiety Scale, Spielberger State Trait Anxiety Inventory, and Hamilton Anxiety Rating Scale [15].

Treatment Options The treatment of diagnosed anxiety in PD is primarily with benzodiazepines, which are particularly beneficial in patients whose tremors are exacerbated by anxiety or stress. The use of benzodiazepines has not been evaluated by a randomized controlled trial and use should be limited given the potential risks of sedation, cognitive effects, and psychomotor agitation. Other case studies have found benefit with serotonergic medications like fluoxetine or citalopram (especially with concomitant depression) or with optimization of levodopa therapy [42,43].

Hallucinations, Delusions, and Psychosis Epidemiology The prevalence of visual hallucinations in PD patients is about 20% to 40% [44,45]. Risk factors for psychotic symptoms include cognitive impairment, advanced age, prolonged duration of disease, depression, severe dysautonomia, and sleep disorders [46–48]. Early recognition of hallucinations is critical because of a strong correlation between the manifestation of psychosis and the need for nursing home placement or hospitalization. With early and effective treatment there is a decreased need for placement and a reduction on caregiver burden [44,49]. Treatment Options Hallucinations can occur in delirium and it is important to first rule out an underlying infection or an offending medication, especially if there is a sudden onset or worsening of symptoms. Psychotic symptoms have been reported in drug-naive patients, though they are often iatrogenically induced with dopaminergic agents. All antiparkinsonian medications are capable of inducing or exacerbating hallucinations [9,50]. Additionally, psychotic symptoms tend to improve when dopaminergic agonists are reduced or eliminated. However, there is no clear relationship between the dose of dopaminergic agents and manifestation of hallucinations [48,51,52]. If hallucinations persist or there are motor complications www.jcomjournal.com

Clinical Review that arise from reduction of dopaminergic agents, initiation of clozapine has been demonstrated to be efficacious in a rater-blinded prospective study and in a retrospective analysis [53–55]; however, regular monitoring for neutropenia is required. Quetiapine has demonstrated similar benefit without significant effects on motor symptoms in a randomized, rater-blinded study and in an evidence-based review [56,57]. It is also important to review or eliminate other medications that may contribute to hallucinations.

Cognitive Impairment Epidemiology The prevalence of dementia in the PD population is 20% to 40% [58], though almost 80% of PD patients ultimately develop cognitive decline [59]. Overall, a PD patient is 6 times more likely to develop dementia than someone in the general population [60]. There may be parallel progression of cognitive impairment and motor symptoms, but there is no correlation with overall duration of disease [60,61]. Risk factors linked with the presence of dementia include older age at onset of PD, presence of hallucinations, and male gender [62,63]. Cognitive dysfunction can be detected early in PD through neuropsychological testing; however, impairment of cognition is often insidious and may not be appreciated until symptoms become severe. Several screening tools have been used to evaluate for cognitive impairment in PD including the Mini-Mental State Exam (MMSE), Montreal Cognitive Assessment (MoCA), Mini-Mental Parkinson, Scales for Outcomes of Parkinson’s disease–Cognition, and others. Accumulating evidence, however, is suggestive of the superiority of the MoCA in the detection of cognitive deficits associated with PD [64]. Dementia is a substantial burden for the caregiver and is a significant contributor to mortality in PD patients [65]. Cognitive impairment often presents with other behavioral symptoms, which further hastens placement outside the home and increases cost of caring for PD patients [49,66]. Cognitive impairment in Parkinson’s disease is typically associated with degeneration of primarily subcortical structures. PD patients with mild cognitive impairment were found to have deficits most significantly in memory, executive function, memory, and language abilities [67]. A recent study by Mak et al evaluated grey matter volumes by structural MRI in PD patients with www.jcomjournal.com

evidence of mild cognitive impairment by MMSE and MoCA as compared with findings in cognitively intact patients. This demonstrated decreased brain volumes in areas that correlate with affected cognitive domains including the left insula, left superior frontal and left middle temporal areas [68].

Treatment Options Prior to initiation of therapy, it is important to evaluate the patient for depression and to rule out pseudodementia. Bradyphrenia, or slowness of thought, should also be considered, as this symptom may also lead to an incorrect dementia diagnosis. Lastly, a thorough review of medications should be performed and offending agents including anticholinergics, TCAs, dopamine agonists, and amantadine should be discontinued as these can worsen cognition. Rivastigmine has demonstrated modest improvement in cognitive performance in PD patients with dementia in a large multicenter, placebo-controlled study [69]. Other cholinesterase inhibitors (ie, donepezil or galantamine) are not recommended at this time due to limited studies or contradictory results in the literature [31,54]. Caution is advised with use of cholinesterase inhibitors as they may worsen tremor or autonomic dysfunction; also, use is limited by nausea or other gastrointestinal symptoms. Memantine, an NMDA receptor antagonist, has also been investigated in randomized, double-blind, placebocontrolled trials and demonstrated modest improvement of cognition and is generally well tolerated [70,71]. Nonpharmacologic therapy includes physical exercise, which has demonstrated improvement in memory tasks and processing speed [72]. Cognitive training has been less rigorously studied; however, a recent single-blinded controlled study demonstrated significant improvement of learning and memory in PD patients who completed computer-based cognitive training [73].

Compulsive Disorders Impulse Control Disorders Impulse control disorders (ICDs) are inappropriate behaviors resulting from a failure to resist an impulse, which leads to pleasure-seeking activities at the expense of relationships and ability to function socially. In PD, ICDs are expressed as pathologic gambling, hypersexuality, binge eating, compulsive shopping, and excessive spending [9,66]. The prevalence of all ICDs in PD is 15% to 20% and a patient may be diagnosed with multiple

Vol. 21, No. 2 February 2014 JCOM 75

Nonmotor Symptoms of Parkinson’s Disease ICDs [74]. Dopamine agonist use has been implicated in the development of ICDs and this risk is further increased with the addition of levodopa [75,76]. Clinical features associated with ICDs include young age of onset, male gender, family history of addiction, depression or anxiety, and disinhibition or impulsive traits [77,78]. Traditionally, treatment consists of reduction or elimination of dopamine agonists, though adjustment of levodopa therapy may also be necessary. Amantadine as an adjunct therapy has been shown in a randomized, double-blind crossover study to reduce impulsivity in a few patients with pathologic gambling [79].

Dopamine Dysregulation Syndrome Dopamine dysregulation syndrome (DDS) is characterized by compulsive use of dopaminergic medications beyond what is needed to treat parkinsonian symptoms, and is associated with social impairment. Patients describe addictive symptoms like craving or intense desire to obtain more dopaminergic medication [9,74]. Like ICDs, treatment of DDS consists of modification to dopaminergic medications, though patients with DDS may also require psychiatric evaluation and treatment. Punding Punding is another compulsive disorder that is defined as an intense fascination with objects and is associated with repetitive handling, manipulation, sorting, or arrangement of the items [80]. Occurrence of punding has been associated with higher total daily levels of levodopa, although one study has also implicated dopamine agonists [15,81]. As with the other compulsive disorders, punding also tends to respond well to reduction or discontinuation of levodopa. Studies have demonstrated modest benefit with SSRIs or atypical antipsychotics in long-term follow-up [82,83], though one study reported worsening of punding with quetiapine [84].

Apathy Epidemiology and Treatment Apathy is often characterized by a loss of motivation or inability to initiate goal-directed behavior, which results in dependence on others for activities of daily living and increases caregiver burden [85]. Patients demonstrate indifference, lack of interest, or inability to express or describe emotion. The apathetic patient may lack spontaneous and voluntary activity, and their affect display is often flattened [86]. 76 JCOM February 2014 Vol. 21, No. 2

With a prevalence of 30% to 50% [87], apathy is as common as depression in PD patients [66,88]. Risk factors associated with apathy include advanced age, severity of depression, severity of motor dysfunction, and dementia [89]. Apathy is frequently mistaken for depression given the significant overlap in symptoms; however, the patient with pure apathy will deny sadness or depressed feelings. It is also important to distinguish apathy from motor impairment or cognitive dysfunction that could explain the behavioral changes. No medications have reliably been shown to improve apathy, though it may be improved with initiation of dopaminergic therapy, especially early in the course [86,90]. SLEEP DISORDERS The original report of PD by James Parkinson describes sleep disturbances and daytime somnolence [91], which suggests that sleep disorders may be an intrinsic feature of the neurodegenerative process of PD itself.

REM Behavioral Disorder Epidemiology and Diagnosis Rapid eye movement behavioral disorder (RBD) is a parasomnia characterized by vocalizations and motor activity during dreaming due to loss of normal atonia associated with rapid eye movement (REM) sleep. Patients enact their dreams, which may lead to violent behaviors that can injure the patient or their bed partner. RBD is seen in 25% to 50% of PD patients [92,93], with variability depending on diagnostic technique and patient selection. Polysomnography is the most important diagnostic tool and demonstrates increased chin tone and limb movements during REM sleep in RBD [94,95]. Diagnosis can also be made clinically with patient and bed partner reports, though sensitivity is only approximately 30% [15]. Interestingly, many studies are now investigating the relationship between presence of RBD and later onset of neurodegenerative disorders. Multiple studies have shown that 40% to 65% of patients diagnosed with idiopathic RBD later develop an alpha-synucleinopathy, which includes PD, dementia with Lewy bodies, or multiple system atrophy within 10 years [92,95]. Prior studies report that as many as 90% of patients with idiopathic RBD develop neurodegenerative synucleinopathy when followed over 14 years [96]. Idiopathic RBD is currently being investigated as a potential clinical marker of presymptomatic PD in a multicenter observational study. If RBD is an early marker for neurodegenerative disease, it www.jcomjournal.com

Clinical Review may be used to identify patients for neuroprotective trials as treatments are developed.

Treatment Options Low-dose clonazepam (0.25–1 mg) is the mainstay of therapy, especially for patients that injure themselves or bed partners [97]; however, the use of benzodiazepines is historical and there remain no randomized controlled double-blind studies to evaluate the efficacy of clonazepam. Use of clonazepam may be limited by daytime sedation, confusion, or psychomotor agitation [31,97,98]. Melatonin (doses between 3–12 mg at bedtime) has also demonstrated benefit in RBD in a double-blind, placebo-controlled trial and in a small case series, with fewer side effects and no addiction potential as compared to clonazepam [99,100]. Case reports also support the use of several other effective medications, including cholinesterase inhibitors (rivastigmine and donepezil) and dopaminergic agents (pramipexole and levodopa) [15,20].

Restless Leg Syndrome and Periodic Limb Movements in Sleep Epidemiology Restless leg syndrome (RLS) and periodic limb movements in sleep (PLMS) cause disruptions of sleep and have an important impact on quality of sleep in PD patients. RLS is described as a strong urge to move the legs, accompanied by an uncomfortable sensation that is exacerbated at rest and relieved by movement. RLS is more frequently diagnosed in patients with PD, though prevalence reports vary widely [15]. Secondary causes for RLS should be investigated including iron deficiency, uremia and polyneuropathy. Several case reports demonstrate onset or worsening of RLS with use of antidepressants [101, 102] or antipsychotics like risperidone, aripiprazole, and quetiapine [103,104]. PLMS occurs in approximately 80% to 90% of patients with RLS, though may be present independently, and when seen on polysomnography is supportive of RLS [105]. PLMS is characterized by repetitive dorsiflexion of the foot, extension of the great toe, and may be accompanied by flexion of the knee and hip. The prevalence of PLMS in PD is approximately 60% and correlates with severity of PD motor features [106]. Treatment Options Treatment of RLS should be initiated with nonpharmacologic therapies including good sleep hygiene, exercise, www.jcomjournal.com

leg massage, and heat or ice packs [105,107]. Dopamine (DA) agonists are the primary treatment for RLS; however, even modest adjustments in levodopa can be helpful. One drawback to levodopa therapy is augmentation (a worsening or reappearance of symptoms) when serum levels fall due to the short half-life of levodopa [107,108]. DA agonists are less likely to cause augmentation. Both pramipexole and ropinirole have been extensively investigated in controlled, randomized, double-blind studies with benefits in 70% to 90% of patients with RLS and PLMS; however, there is a risk of developing compulsive behaviors [109–112]. Another option for PD patients is rotigotine, which has demonstrated improvement of RLS symptoms in a randomized, double-blind, placebocontrolled trial and has the added benefit that it may also help with motor symptoms [113,114]. More recently, gabapentin enacarbil has demonstrated improvement of moderate to severe RLS and was well tolerated in multiple randomized, double-blind, placebocontrolled trials [107,115,116]. Lastly, opioids (tramadol, oxycodone, codeine) have been shown to be effective, especially in the treatment of RLS that is refractory to other treatments [105,107].

Insomnia Epidemiology The most common sleep disorder in PD is insomnia, with a prevalence between 37% to 88% [14,117]. Insomnia is associated with difficulty in initiation or maintenance of sleep. Disruption of sleep typically leads to daytime somnolence and patient reports of a strong impact on motor disability and overall quality of life. There are several contributors to insomnia in PD patients including nocturia, depression, RLS, dystonia, and akinesia/rigidity/ difficulty turning in bed [118]. Treatment Options The use of carbidopa/levodopa controlled-release formulations at bedtime is associated with improved sleep duration and nocturnal akinesia, although it does not demonstrate a significant improvement in overall sleep ratings [54]. Hypnotics like eszopiclone and zolpidem have also demonstrated improved quality of sleep in limited controlled trials and a meta-analysis, but use is limited by sedation, dizziness, and falls [54,119]. Benzodiazepines improve sleep latency, but there is a risk of cognitive impairment, tolerance, and falls [117,120]. Melatonin at 3 to 5 mg and 50 mg doses have been investigated in

Vol. 21, No. 2 February 2014 JCOM 77

Nonmotor Symptoms of Parkinson’s Disease 2 randomized, double-blind, placebo-controlled trials; however, there was a modest benefit and it was concluded that there is insufficient evidence to support the use of melatonin [54]. Nevertheless, melatonin is well tolerated and may be tried with minimal risk [54]. More recently, a randomized controlled trial using doxepin has demonstrated improvement of insomnia scores and was generally well tolerated [121].

Excessive Daytime Sleepiness and Abrupt Sleep Onset EDS and Fatigue: Epidemiology and Treatment A common complaint by PD patients is excessive daytime sleepiness (EDS), which can be verified with multiple sleep latency testing. EDS frequency varies in the literature, but is seen in approximately 15% to 50% of PD patients [4,122]. The etiology is usually multifactorial, with insomnia, dysautonomia, and depression as contributing factors [117]. A longer duration of symptoms, greater total load of levodopa, cognitive decline, and male gender are all risk factors for EDS [122,123]. It has been proposed that EDS is an intrinsic feature of PD; however, there is also an association with the use of antiparkinsonian medications. A randomized controlled trial demonstrated that use of the dopamine agonist pramipexole was associated with greater somnolence as compared to levodopa therapy (35% vs. 13%); however, this difference was only seen during the initial escalation phase [124]. Additionally, the combined use of dopamine agonists and levodopa has shown an even greater risk of EDS [125]. The evidence for the use of stimulants for EDS is lacking. The few studies conducted with modafinil have not demonstrated a robust improvement of EDS [126–128]. Other stimulants like methylphenidate have been studied with improvement of Epworth Sleepiness Score, though no randomized control trials have been undertaken [129]. It is important to distinguish EDS, a propensity for daytime sleep, from fatigue or excessive tiredness associated with mental or physical exertion [117]. Fatigue is often multifactorial and may be related to insomnia, sleep apnea, sedating effects of medications, frequent awakenings from nocturia, and degeneration of brain areas regulating sleep/wake cycles related to the underlying disease process [20, 117]. It is also important to consider depression and dementia in the differential, as these disorders may be erroneously be diagnosed as fatigue. Treatment of fatigue should include regular mild exer78 JCOM February 2014 Vol. 21, No. 2

cise, maintenance of a stimulating environment, removal of sedating medications, and management of intrinsic sleep disorders if present [117]. The use of stimulants for fatigue is controversial. A small randomized controlled trial (n = 48) using modafinil demonstrated improvement on the global clinical impression scale for fatigue but no significant change on the Fatigue Severity Scale; this study was limited by the power and points to the need for a larger study [130].

Sleep Attacks: Epidemiology and Treatment Abrupt sleep onset, or “sleep attacks,” occurs when transition from wake to sleep is unavoidable and may occur without warning. Sleep attacks are threefold more likely to occur in patients using DA agonists, with an associated dose-related increase in risk [131]. Adjustment or elimination of DA agonists often improves sleep attacks, though it is important to address concurrent EDS if present. Nonpharmacologic treatments to consider include mild exercise, early morning bright light exposure, and a stimulating environment [117].

Sleep-Disordered Breathing/Obstructive Sleep Apnea Epidemiology and Treatment Sleep-disordered breathing (SDB) consists of either a deficit in the drive to breathe as in central sleep apnea, or may be due to an blockage of the airway as seen in obstructive sleep apnea (OSA). Apnea leads to oxygen desaturations that consequently trigger awakenings throughout the night, which in turn is experienced by the patient as daytime somnolence [117]. The prevalence of SDB and OSA is variable in the literature, ranging from no increased risk in PD patients [132,133] to 50% prevalence in PD patients [134,135]. Discussions with bed partners, history of snoring, and clinical reports of EDS or daytime fatigue are important indicators of SDB. Polysomnography confirms the diagnosis and can direct treatment, which frequently includes application of CPAP devices during sleep. AUTONOMIC DYSFUNCTION

Orthostatic Hypotension Epidemiology and Diagnosis Orthostatic hypotension (OH) is defined as a 20-mm Hg fall in systolic blood pressure or 10-mm Hg drop in diastolic blood pressure within 3 minutes of a change in position. The prevalence of OH in PD patients is 30% to www.jcomjournal.com

Clinical Review 60% [136,137]. Symptoms of OH can occur early in the disease and may precede diagnosis of PD [137]. Patients experience OH as dizziness, drowsiness, palpitations, nausea, or loss of consciousness. Additionally, falls and supine hypertension that accompany OH are associated with increased risk of morbidity and mortality in PD patients [138]. Several medications used in the treatment of PD can exacerbate OH, including levodopa, DA agonists, MAO-B inhibitors, and TCAs [139].

Treatment Options First-line therapies for OH include nonpharmacologic methods such as compression stockings, sleeping with head elevated to 30 degrees, increased water and salt intake, more frequent small meals, and slowly changing position [140]. Additionally, it is important to discuss the removal or reduction of all antihypertensives with the patient’s PCP. Fludrocortisone (a mineralacorticoid) and domperidone (a peripheral dopamine antagonist not currently approved for use in the United States) modestly improved OH in a 2-phase, randomized, controlled, double-blind, crossover trial [141]. Pyridostigmine has also demonstrated improvement of standing blood pressure and OH symptoms in a double-blind, randomized cross-over study and has the additional benefit of not worsening supine hypertension [142]. Other effective treatments include midodrine, per a randomized, double-blind multicenter study [143], as well as droxidopa in a double-blind, crossover, placebocontrolled study [144]. Currently there is insufficient evidence to support the preferential use of any specific agent in the treatment of OH in PD.

Gastrointestinal Dysmotility Constipation: Epidemiology and Treatment Constipation is reported by nearly 60% of PD patients [145]. Constipation can precede the development of motor symptoms of PD, and the prevalence of GI disturbances increases with age and longer duration of disease. Nearly one third of patients will have been diagnosed with a GI disturbance within the year prior to PD diagnosis [146], which is associated with an increased risk for the development PD [147]. People with constipation (defined as < 1 bowel movement per day) but without a PD diagnosis had more nigral Lewy body degeneration postmortem [148] compared with people without constipation. Treatments for constipation include dietary modification, increased fluid intake, and mild exercise. Macrogol www.jcomjournal.com

significantly improved constipation in PD patients and was very well tolerated in a randomized placebo-controlled study [149]. Lubiprostone, a GI active prostaglandin, is also effective in the short-term treatment of constipation in a placebo-controlled trial [150].

Gastroparesis: Epidemiology and Treatment Gastroparesis, like constipation, is related to enteric dopaminergic cell loss and degeneration of the dorsal motor nucleus of the vagus [151]. Patients experience gastroparesis as early satiety, full sensation, and nausea. Decreased gastric motility leads to retention of food as well as medications, which can slow absorption and delay onset of action for many medications including levodopa. Domperidone has both prokinetic and antiemetic properties, which have been beneficial in the treatment of gastroparesis [152], but its use is not currently approved in the United States. Dysphagia: Epidemiology and Treatment Dysphagia is associated with more advanced stages of PD as well as a significant increase in morbidity. Swallow exercises have demonstrated improvement of dysphagia [153]. The impact of levodopa therapy on dysphagia in the literature is controversial. Videofluoroscopic examination is the most common method for evaluation of swallowing disorders and provides important information for speech-language pathologists regarding recommendations for dietary modifications [154]. Adjustment of medication regimens to avoid an oral route is also helpful. This includes Parcopa, orally disintegrating carbidopa/levodopa tablets, and transdermal approaches like the rotigotine patch. For some patients, enteral nutrition is needed and placement of nasogastric tubes or percutaneous endoscopic gastrostomy tubes are an option.

Sialorrhea (Drooling) Epidemiology Difficulty handling oral secretions due to impaired or infrequent swallowing results in sialorrhea in up to 75% of PD patients [155], which is a significant embarrassment for most patients [156]. PD patients with drooling have difficulty speaking, eating, and engaging in social interactions, which significantly impacts perceived quality of life [157]. Treatment Options Botulinum toxin (A and B) injections into the submandibular or parotid glands have demonstrated efficacy in

Vol. 21, No. 2 February 2014 JCOM 79

Nonmotor Symptoms of Parkinson’s Disease multiple double-blind, randomized, placebo-controlled studies for the treatment of sialorrhea in PD patients; however, injections are associated with greater invasiveness and cost [158–160]. Glycopyrrolate, an anticholinergic drug, was also efficacious in the treatment of sialorrhea in the short term in a double-blind, randomized, placebo-controlled study [161]. Alternatively, gum chewing increases swallow frequency, improves drooling, and also shows a benefit with dysphagia [162].

Genitourinary Disturbances Bladdery dysfunction: Epidemiology and Treatment Bladder dysfunction in PD is often secondary to hyperactivity of the detrusor muscle leading to urinary urgency, increased urinary frequency, and nocturia. Less commonly, hypoactive detrusor muscle causes difficulty with initiation of urination, delayed bladder emptying, and recurrent infections. Urinary disturbances may occur before the onset of motor symptoms or early on in the disease course [12]. Disease severity is associated with greater urinary disturbances, and more than 50% of advanced PD patients report severe bladder symptoms [163]. Anticholinergic medications such as oxybutynin, solifenacin, and tolterodine are commonly used in the treatment of detrusor hyperactivity and demonstrate significant improvement in detrusor pressure in a recent systemic review and meta-analysis [164]. PD patients on these agents should be closely monitored for side effects including cognitive impairment, somnolence, hallucinations, confusion, and blurred vision. Other treatments include botulinum toxin injections into the detrusor muscle, which has demonstrated safety and efficacy in a recent systematic review [165]. Erectile dysfunction: Epidemiology and Treatment Erectile dysfunction (ED) is reported by more than 60% of male PD patients [145] and is thought to be related to hypothalamic dysfunction and modification of the dopamine-oxytocin pathway [166]. Effects of PD medications, cognitive impairment, fatigue, apathy, and low testosterone contribute to loss of libido and ED [20,167]. Phosphodiesterase inhibitors such as sildenafil, vardenafil, and tadalafil are possibly useful in the treatment of ED in PD patients, though randomized trials have been limited [166,168]. Apomorphine sublingually is another medication that has demonstrated improvement of ED 80 JCOM February 2014 Vol. 21, No. 2

in a double-blind, crossover study and can be considered for patients with contraindications to phosphodiesterase inhibitors [169]. SENSORY SYMPTOMS

Pain Epidemiology Sensory disturbances in PD include diminished ability to identify odors, visual abnormalities (blurred vision, abnormal color perception, double vision), and pain. Pain is the most disabling sensory disturbance, though frequently underreported. Nearly two thirds of PD patients report pain, [170], though only half of patients receive any treatment [171]. Pain may also be a presenting symptom that precedes the clinical diagnosis of PD [172,173]. Treatment Options There are several types of pain described by PD patients, the most common of which is musculoskeletal, typically involving the shoulder. Other types include dystonic, radicular, and central pain [174]. First-line treatment of musculoskeletal complaints includes nonsteroidal antiinflammatory drugs (NSAIDs) and physiotherapy. Modification of levodopa regimen (including altering timing and frequency or adding controlled release formulations) can often provide relief for dystonic pain, and also for central pain for some patients [173, 174]. Deep brain stimulation, with subthalamic nucleus or globus pallidus targets, has demonstrated improvement with dystonic, central, and musculoskeletal pain in a small clinical study [175].

CONCLUSION NMS are an intrinsic part of PD, may predate diagnosis, and substantially affect the majority of patients with PD. For many of these patients, NMS have a greater impact on quality of life and health care costs than the cardinal motor symptoms that define the disease. Many of these symptoms are not recognized by practioners and often are not volunteered by PD patients, making it important for practitioners to routinely and directly inquire about NMS. Treatment of NMS in PD is challenging, and only a few therapies have the level of evidence needed to support their use in the treatment of these problems. Nevertheless, proper recognition and addressing of these symptoms afford the clinician an opportunity to make a positive and potentially significant impact on the PD patient’s quality of life. www.jcomjournal.com

Clinical Review Corresponding author: Samay Jain, MD, MS, Dept of Neurology, 811 Kaufmann Bldg, Pittsburgh, PA 15213, [email protected]. Financial disclosures: None. REFERENCES 1. Alves G, Forsaa EB, Pedersen KF, et al. Epidemiology of Parkinson's disease. J Neurol 2008;255 Suppl 5:18–32. 2. Stern MB, Lang A, Poewe W. Toward a redefinition of Parkinson's disease. Mov Disord 2012;27:54–60. 3. Braak H, Del Tredici K, Rub U, et al. Staging of brain pathology related to sporadic Parkinson's disease. Neurobiol Aging 2003;24:197–211. 4. Tandberg E, Larsen JP, Karlsen K. A community-based study of sleep disorders in patients with Parkinson's disease. Mov Disord 1998;13:895–9. 5. Shulman LM, Taback RL, Bean J, Weiner WJ. Comorbidity of the nonmotor symptoms of Parkinson's disease. Mov Disord 2001;16:507–10. 6. Park A, Stacy M. Non-motor symptoms in Parkinson's disease. J Neurol 2009;256 Suppl 3:293–8. 7. Chaudhuri KR, Schapira AH. Non-motor symptoms of Parkinson's disease: dopaminergic pathophysiology and treatment. Lancet Neurol 2009;8:464–74. 8. Shulman LM, Taback RL, Rabinstein AA, Weiner WJ. Nonrecognition of depression and other non-motor symptoms in Parkinson's disease. Parkinsonism Relat Disord 2002;8:193–7. 9. Bonnet AM, Jutras MF, Czernecki V, et al. Nonmotor symptoms in Parkinson's disease in 2012: relevant clinical aspects. Parkinsons Dis 2012:2012:198316. 10. Chaudhuri KR, Healy DG, Schapira AH. Non-motor symptoms of Parkinson's disease: diagnosis and management. Lancet Neurol 2006;5:235–45. 11. Hussl AK, Seppi K, Poewe W. Nonmotor symptoms in Parkinson's disease. Expert Rev Neurother 2013;13:581–3. 12. O'Sullivan SS, Williams DR, Gallagher DA, et al. Nonmotor symptoms as presenting complaints in Parkinson's disease: a clinicopathological study. Mov Disord 2008;23:101–6. 13. Lang AE. A critical appraisal of the premotor symptoms of Parkinson's disease: potential usefulness in early diagnosis and design of neuroprotective trials. Mov Disord 2011;26:775–83. 14. Barone P, Antonini A, Colosimo C, et al. The PRIAMO study: A multicenter assessment of nonmotor symptoms and their impact on quality of life in Parkinson's disease. Mov Disord 2009;24:1641–9. 15. Bernal-Pacheco O, Limotai N, Go CL, Fernandez HH. Nonmotor manifestations in Parkinson disease. Neurologist 2012;18:1–16. 16. Lemke MR, Fuchs G, Gemende I, et al. Depression and Parkinson's disease. J Neurol 2004;251 Suppl 6: VI/24–7. 17. Ravina, B, Camicioli R, Como PG, et al, The impact of depressive symptoms in early Parkinson disease. Neurology 2007;69:342–7. 18. Jasinska-Myga B, Putzke JD Wider C, et al. Depression in Parkinson's disease. Can J Neurol Sci 2010;37:61–6. www.jcomjournal.com

19. Slaughter JR, Slaughter KA, Nichols D, et al. Prevalence, clinical manifestations, etiology, and treatment of depression in Parkinson's disease. J Neuropsychiatry Clin Neurosci 2001; 13:187–96. 20. Simuni T, Sethi K. Nonmotor manifestations of Parkinson's disease. Ann Neurol 2008;64 Suppl 2:S65–80. 21. Ehmann TS, Beninger RJ, Gawel MJ, Riopelle RJ. Depressive symptoms in Parkinson's disease: a comparison with disabled control subjects. J Geriatr Psychiatry Neurol 1990;3:3–9. 22. Menza MA, Mark MH. Parkinson's disease and depression: the relationship to disability and personality. J Neuropsychiatry Clin Neurosci 1994;6:165–9. 23. CDC. Current depression among adults–United States, 2006 and 2008. Morb Mort Weekly Rep 2010;59:1229–35. 24. Hinnell C, Hurt CS, Landau S, et al. Nonmotor versus motor symptoms: how much do they matter to health status in Parkinson's disease? Mov Disord 2012;27:236–41. 25. Cummings JL. Depression and Parkinson's disease: a review. Am J Psychiatry 1992;149:443–54. 26. Weintraub D, Morales KH, Moberg PJ, et al. Antidepressant studies in Parkinson's disease: a review and meta-analysis. Mov Disord 2005;20:1161–9. 27. Richard IH, McDermott MP, Kurlan R, et al. A randomized, double-blind, placebo-controlled trial of antidepressants in Parkinson disease. Neurology 2012;78:1229–36. 28. Menza M, Dobkin RD, Marin H, et al. A controlled trial of antidepressants in patients with Parkinson disease and depression. Neurology 2009;72:886–92. 29. Okun MS, Fernandez HH. Will tricyclic antidepressants make a comeback for depressed Parkinson disease patients? Neurology 2009;72:868–9. 30. Antonini A, Tesei S, Zecchinelli A, et al. Randomized study of sertraline and low-dose amitriptyline in patients with Parkinson's disease and depression: effect on quality of life. Mov Disord 2006;21:1119–22. 31. Pedrosa DJ, Timmermann L. Review: management of Parkinson's disease. Neuropsychiatr Dis Treat 2013;9: 321–40. 32. Barone P, Poewe W, Albrecht S, et al. Pramipexole for the treatment of depressive symptoms in patients with Parkinson's disease: a randomised, double-blind, placebo-controlled trial. Lancet Neurol 2010;9:573–80. 33. Lemke MR, Brecht HM, Koester J, et al. Anhedonia, depression, and motor functioning in Parkinson's disease during treatment with pramipexole. J Neuropsychiatry Clin Neurosci 2005;17:214–20. 34. Rektorova I, Balaz M, Svatova J, et al. Effects of ropinirole on nonmotor symptoms of Parkinson disease: a prospective multicenter study. Clin Neuropharmacol 2008;31:261–6. 35. Van der Wurff FB, Stek ML, Hoogendijk WL, Beekman AT. Electroconvulsive therapy for the depressed elderly. Cochrane Database Syst Rev 2003(2):CD003593. 36. Okun MS, Watts RL. Depression associated with Parkinson's disease: clinical features and treatment. Neurology 2002;58(4 Suppl 1):S63–70. 37. Dobkin RD, Menza M, Allen LA, et al. Cognitive-behavioral therapy for depression in Parkinson's disease: a randomized, controlled trial. Am J Psychiatry 2013;168:1066–74.

Vol. 21, No. 2 February 2014 JCOM 81

Nonmotor Symptoms of Parkinson’s Disease 38. Loprinzi PD, Herod SM, Cardinal BJ, Noakes TD. Physical activity and the brain: A review of this dynamic, bi-directional relationship. Brain Res 2013;1539:95–104. 39. Richard IH. Anxiety disorders in Parkinson's disease. Adv Neurol 2005;96:42–55. 40. Leentjens AF, Dujardin K, Marsh L, et al. Symptomatology and markers of anxiety disorders in Parkinson's disease: a cross-sectional study. Mov Disord 2011;26:484–92. 41. Witjas T, Kaphan E, Azulay JP, et al. Nonmotor fluctuations in Parkinson's disease: frequent and disabling. Neurology 2002;59:408–13. 42. Olanow CW, Stern MB, Sethi K. The scientific and clinical basis for the treatment of Parkinson disease Neurology 2009;72(21 Suppl 4):S1–136. 43. Richard IH, Schiffer RB, Kurlan R. Anxiety and Parkinson's disease. J Neuropsychiatry Clin Neurosci 1996;8:383–92. 44. Fenelon G, Alves G. Epidemiology of psychosis in Parkinson's disease. J Neurol Sci 2010;289:12–7. 45. Papapetropoulos S, Katzen H, Schrag A, et al. A questionnaire-based (UM-PDHQ) study of hallucinations in Parkinson's disease. BMC Neurol 2008;8:21. 46. Fenelon G, Mahieux F, Huon R, Ziegler M. Hallucinations in Parkinson's disease: prevalence, phenomenology and risk factors. Brain 2000;123 (Pt 4):733–45. 47. Papapetropoulos S, Mash DC. Psychotic symptoms in Parkinson's disease. From description to etiology. J Neurol 2005;252:753–64. 48. Aarsland D, Larsen JP, Cummins JL, Laake K. Prevalence and clinical correlates of psychotic symptoms in Parkinson disease: a community-based study. Arch Neurol 1999;56:595–601. 49. Aarsland D, Larsen JP, Tandberg E, Laake K. Predictors of nursing home placement in Parkinson's disease: a populationbased, prospective study. J Am Geriatr Soc 2000;48:938–42. 50. Lohle M, Storch A, Reichmann H, Beyond tremor and rigidity: non-motor features of Parkinson's disease. J Neural Transm 2009;116:1483–92. 51. Fenelon G. Psychosis in Parkinson's disease: phenomenology, frequency, risk factors, and current understanding of pathophysiologic mechanisms. CNS Spectr 2008;13(3 Suppl 4): 18–25. 52. Merims D, Shabtai H, Korczyn AD, et al. Antiparkinsonian medication is not a risk factor for the development of hallucinations in Parkinson's disease. J Neural Transm 2004; 111:1447–53. 53. Merims D, Balas M, Peretz C, et al. Rater-blinded, prospective comparison: quetiapine versus clozapine for Parkinson's disease psychosis. Clin Neuropharmacol 2006;29:331–7. 54. Seppi K, Weintraub D, Coelho M, et al. The Movement Disorder Society Evidence-Based Medicine Review Update: Treatments for the non-motor symptoms of Parkinson's disease. Mov Disord 2011;26 Suppl 3:S42–80. 55. Fernandez HH, Donnelly EM, Friedman JH. Long-term outcome of clozapine use for psychosis in parkinsonian patients. Mov Disord 2004;19:831–3. 56. Morgante L, Epifanio A, Spina E, et al. Quetiapine and clozapine in parkinsonian patients with dopaminergic psychosis. Clin Neuropharmacol 2004;27:153–6.

82 JCOM February 2014 Vol. 21, No. 2

57. Miyasaki JM, Shannon K, Voon V, et al. Practice parameter: evaluation and treatment of depression, psychosis, and dementia in Parkinson disease (an evidence-based review): report of the Quality Standards Subcommittee of the American Academy of Neurology. Neurology 2006;66:996–1002. 58. Riedel O, Klotsche J, Spottke A, et al. Cognitive impairment in 873 patients with idiopathic Parkinson's disease. Results from the German Study on Epidemiology of Parkinson's Disease with Dementia (GEPAD). J Neurol 2008;255:255–64. 59. Aarsland D, Andersen K, Larsen JP, et al. Prevalence and characteristics of dementia in Parkinson disease: an 8-year prospective study. Arch Neurol 2003;60:387–92. 60. Aarsland D, Andersen K, Larsen JP, et al. Risk of dementia in Parkinson's disease: a community-based, prospective study. Neurology 2001;56:730–6. 61. Riggeal BD, Crucian GP, Seignourel P, et al. Cognitive decline tracks motor progression and not disease duration in Parkinson patients. Neuropsychiatr Dis Treat 2007;3:955–8. 62. Hughes TA, Ross HF, Musa S, et al. A 10-year study of the incidence of and factors predicting dementia in Parkinson's disease. Neurology 2000;54:1596–602. 63. Hobson P, Meara J. Risk and incidence of dementia in a cohort of older subjects with Parkinson's disease in the United Kingdom. Mov Disord 2004;19:1043–9. 64. Hoops S, Nazem S, Siderowf AD, et al. Validity of the MoCA and MMSE in the detection of MCI and dementia in Parkinson disease. Neurology 2009;73:1738–45. 65. Louis ED, Marder K, Cote L, et al. Mortality from Parkinson disease. Arch Neurol 1997;54:260–4. 66. Fernandez HH. Nonmotor complications of Parkinson disease. Cleve Clin J Med 2012;79 Suppl 2:S14–8. 67. Sollinger AB, Goldstein FC, Lah JJ, et al. Mild cognitive impairment in Parkinson's disease: subtypes and motor characteristics. Parkinsonism Relat Disord 2010;16:177–80. 68. Mak E, Zhou J, Tan LC, et al. Cognitive deficits in mild Parkinson's disease are associated with distinct areas of grey matter atrophy. J Neurol Neurosurg Psychiatry 2013 Oct 16. 69. Emre M, Aarsland D, Albanese A, et al. Rivastigmine for dementia associated with Parkinson's disease. N Engl J Med 2004;351:2509–18. 70. Aarsland D, Ballard C, Walker Z, et al. Memantine in patients with Parkinson's disease dementia or dementia with Lewy bodies: a double-blind, placebo-controlled, multicentre trial. Lancet Neurol 2009;8:613–8. 71. Leroi I, Overshott R, Byrne EJ, et al. Randomized controlled trial of memantine in dementia associated with Parkinson's disease. Mov Disord 2009;24:1217–21. 72. Speelman AD, van de Warrenburg BP, van Nimwegen M, et al. How might physical activity benefit patients with Parkinson disease? Nat Rev Neurol 2013;7:528–34. 73. Naismith SL, Mowszowski L, Diamond K, Lewis SJ. Improving memory in Parkinson's disease: a healthy brain ageing cognitive training program. Mov Disord 2013;28:1097–103. 74. Weintraub D, Koester J, Potenza MN, et al. Impulse control disorders in Parkinson disease: a cross-sectional study of 3090 patients. Arch Neurol 2010;67:589–95. 75. Voon V, Reynolds B, Brezing C, et al. Impulsive choice and www.jcomjournal.com

Clinical Review response in dopamine agonist-related impulse control behaviors. Psychopharmacology (Berl) 2010;207:645–59. 76. Weintraub D, Siderowf AD, Potenza MN, et al. Association of dopamine agonist use with impulse control disorders in Parkinson disease. Arch Neurol 2006; 63:969–73. 77. Pontone G, Williams JR, Bassett SS, Marsh L. Clinical features associated with impulse control disorders in Parkinson disease. Neurology 2006;67:1258–61. 78. Voon V, Mehta AR, Hallett M. Impulse control disorders in Parkinson's disease: recent advances. Curr Opin Neurol 2011;24:324–30. 79. Thomas A, Bonanni L, Gambi F, et al. Pathological gambling in Parkinson disease is reduced by amantadine. Ann Neurol 2010;68:400–4. 80. Miyasaki JM, Al Hassan K, Lang AE, Voon V. Punding prevalence in Parkinson's disease. Mov Disord 2007;22:1179–81. 81. Nguyen FN, Chang YL, Okun MS, et al. Prevalence and characteristics of punding and repetitive behaviors among Parkinson patients in North-Central Florida. Int J Geriatr Psychiatry 2010;25:540–1. 82. Sohtaoglu M, Demiray DY, Kenangil G, et al. Long term follow-up of Parkinson's disease patients with impulse control disorders. Parkinsonism Relat Disord 2010;16:334–7. 83. Antonini A, Cilia R. Behavioural adverse effects of dopaminergic treatments in Parkinson's disease: incidence, neurobiological basis, management and prevention. Drug Saf 2009; 32:475–88. 84. Miwa H, Morita S, Nakanishi I, Kondo T, Stereotyped behaviors or punding after quetiapine administration in Parkinson's disease. Parkinsonism Relat Disord 2004;10:177–80. 85. Skorvanek M, Rosenberger J, Gdovinova Z, et al. Apathy in elderly nondemented patients with parkinson's disease: clinical determinants and relationship to quality of life. J Geriatr Psychiatry Neurol 2013;26:237–43. 86. Marin RS, Fogel BS, Hawkins J, et al. Apathy: a treatable syndrome. J Neuropsychiatry Clin Neurosci 1995;7:23–30. 87. Pluck GC, Brown RG. Apathy in Parkinson's disease. J Neurol Neurosurg Psychiatry 2002;73:636–42. 88. Kirsch-Darrow L, Fernandez HH, Marsiske M, et al. Dissociating apathy and depression in Parkinson disease. Neurology 2006;67:33–8. 89. Pedersen KF, Alves G, Aarsland D, Larsen JP. Occurrence and risk factors for apathy in Parkinson disease: a 4-year prospective longitudinal study. J Neurol Neurosurg Psychiatry 2009;80:1279–82. 90. Czernecki V, Pillon B, Houeto JL, et al. Motivation, reward, and Parkinson's disease: influence of dopatherapy. Neuropsychologia 2002;40:2257–67. 91. Parkinson J. An essay on the shaking palsy. Sherwood, Neely, and Jones; 1817. 92. Schenck CH, Mahowald WM. REM sleep behavior disorder: clinical, developmental, and neuroscience perspectives 16 years after its formal identification in SLEEP. Sleep 2002;25:120–38. 93. Sixel-Doring F, Trautmann E, Mollenhauer B, Trenkwalder C. Associated factors for REM sleep behavior disorder in Parkinson disease. Neurology 2011;77:1048–54. www.jcomjournal.com

94. Eisensehr I, v Lindeiner H, Jager M, Noachtar S. REM sleep behavior disorder in sleep-disordered patients with versus without Parkinson's disease: is there a need for polysomnography? J Neurol Sci 2001;186:7–11. 95. Postuma RB, Gagnon JF, Montplaisir JY. REM sleep behavior disorder and prodromal neurodegeneration - where are we headed? Tremor Other Hyperkinet Mov (N Y) 2013;3. 96. Iranzo A, Tolosa E, Gelpi E, et al. Neurodegenerative disease status and post-mortem pathology in idiopathic rapid-eyemovement sleep behaviour disorder: an observational cohort study. Lancet Neurol 2013;12: 443–53. 97. Schenck CH, Mahowald MW. Rapid eye movement sleep parasomnias. Neurol Clin 2005;23:1107–26. 98. Anderson KN, Shneerson JM. Drug treatment of REM sleep behavior disorder: the use of drug therapies other than clonazepam. J Clin Sleep Med 2009;5:235–9. 99. Boeve BF, Silber MH, Ferman TJ. Melatonin for treatment of REM sleep behavior disorder in neurologic disorders: results in 14 patients. Sleep Med 2003;4:281–4. 100. Kunz D, Mahlberg R. A two-part, double-blind, placebocontrolled trial of exogenous melatonin in REM sleep behaviour disorder. J Sleep Res 2010;19:591–6. 101. Perroud N, Lazignac C, Baleydier B, et al. Restless legs syndrome induced by citalopram: a psychiatric emergency? Gen Hosp Psychiatry 2007;29:72–4. 102. Buskova J, Vorlova T, Pisko J, Sonka K, Severe sleeprelated movement disorder induced by sertraline. Sleep Med 2012;13:769–70. 103. Rittmannsberger H, Werl R. Restless legs syndrome induced by quetiapine: report of seven cases and review of the literature. Int J Neuropsychopharmacol 2013;16:1427–31. 104. Perez-Lloret S, Rey MV, Bondon-Guitton E, et al. Drugs associated with restless legs syndrome: a case/noncase study in the French pharmacovigilance database. J Clin Psychopharmacol 2012;32:824–7. 105. Aurora RN, Kristo DA, Bista SR, et al. The treatment of restless legs syndrome and periodic limb movement disorder in adults-an update for 2012: practice parameters with an evidence-based systematic review and meta-analyses: an American Academy of Sleep Medicine Clinical Practice Guideline. Sleep 2012;35:1039–62. 106. Covassin N, Neikrug AB, Liu L, et al. Clinical correlates of periodic limb movements in sleep in Parkinson's disease. J Neurol Sci 2012;316:131–6. 107. Rios Romenets S, Postuma RB. Treatment of restless legs syndrome. Curr Treat Options Neurol 2013;15:396-409. 108. Hogl B, Paulus W, Clarenbach P, Trenkwalder C, Restless legs syndrome: diagnostic assessment and the advantages and risks of dopaminergic treatment. J Neurol 2006;253 Suppl 4: IV22-8. 109. Garcia-Borreguero D, Kohnen R, Silber MH, et al. The long-term treatment of restless legs syndrome/Willis-Ekbom disease: evidence-based guidelines and clinical consensus best practice guidance: a report from the International Restless Legs Syndrome Study Group. Sleep Med 2013;14:675–84. 110. Montagna P, Hornyak M, Ulfberg J, et al. Randomized trial of pramipexole for patients with restless legs syndrome

Vol. 21, No. 2 February 2014 JCOM 83

Nonmotor Symptoms of Parkinson’s Disease (RLS) and RLS-related impairment of mood. Sleep Med 2011;12:34–40. 111. Partinen M, Hirvonen K, Jama L, et al. Efficacy and safety of pramipexole in idiopathic restless legs syndrome: a polysomnographic dose-finding study--the PRELUDE study. Sleep Med 2006;7:407–17. 112. Trenkwalder C Garcia-Borreguero D, Montagna P, et al. Ropinirole in the treatment of restless legs syndrome: results from the TREAT RLS 1 study, a 12 week, randomised, placebo controlled study in 10 European countries. J Neurol Neurosurg Psychiatry 2004;75:92–7. 113. Hening WA, Allen RP, Ondo WG, et al. Rotigotine improves restless legs syndrome: a 6-month randomized, double-blind, placebo-controlled trial in the United States. Mov Disord 2010;25:1675–83. 114. Oertel WH, Benes H, Garcia-Borreguero D, et al. Rotigotine transdermal patch in moderate to severe idiopathic restless legs syndrome: a randomized, placebo-controlled polysomnographic study. Sleep Med 2010;11:848–56. 115. Lee DO, Ziman RB, Perkins AT, et al. A randomized, doubleblind, placebo-controlled study to assess the efficacy and tolerability of gabapentin enacarbil in subjects with restless legs syndrome. J Clin Sleep Med 2011;7:282–92. 116. Kushida CA, Walters AS, Becker P, et al. A randomized, double-blind, placebo-controlled, crossover study of XP13512/ GSK1838262 in the treatment of patients with primary restless legs syndrome. Sleep 2009;32:159–68. 117. Menza M, Dobkin RD, Marin H, Bienfait K. Sleep disturbances in Parkinson's disease. Mov Disord 2010;25 Suppl 1:S117–22. 118. Louter M, van Sloun RJ, Pevernagie DA, et al. Subjectively impaired bed mobility in Parkinson disease affects sleep efficiency. Sleep Med 2013;14:668–74. 119. Menza M, Dobkin RD, Marin H, et al. Treatment of insomnia in Parkinson's disease: a controlled trial of eszopiclone and placebo. Mov Disord 2010;25:1708–14. 120. Nowell PD, Mazumdar S, Buysse DJ, et al. Benzodiazepines and zolpidem for chronic insomnia: a meta-analysis of treatment efficacy. JAMA 1997;278:2170–7. 121. Rios Romenets S, Creti L, Fichten C, et al. Doxepin and cognitive behavioural therapy for insomnia in patients with Parkinson's disease – a randomized study. Parkinsonism Relat Disord 2013;19:670–5. 122. Ondo WG, Dat Vuong K, Khan H, et al. Daytime sleepiness and other sleep disorders in Parkinson's disease. Neurology 2001;57:1392–6. 123. Razmy A, Lang AE, Shapiro CM, Predictors of impaired daytime sleep and wakefulness in patients with Parkinson disease treated with older (ergot) vs newer (nonergot) dopamine agonists. Arch Neurol 2004;61:97–102. 124. Holloway RG, Shoulson I, Fahn S, et al. Pramipexole vs levodopa as initial treatment for Parkinson disease: a 4-year randomized controlled trial. Arch Neurol 2004;61:1044–53. 125. Paus S, Brecht HM, Koster J, et al. Sleep attacks, daytime sleepiness, and dopamine agonists in Parkinson's disease. Mov Disord 2003;18:659–67. 126. Hogl B, Saletu M, Brandauer E, et al. Modafinil for the treat-

84 JCOM February 2014 Vol. 21, No. 2

ment of daytime sleepiness in Parkinson's disease: a doubleblind, randomized, crossover, placebo-controlled polygraphic trial. Sleep 2002;25:905–9. 127. Ondo WG, Fayle R, Atassi F, Jankovic J. Modafinil for daytime somnolence in Parkinson's disease: double blind, placebo controlled parallel trial. J Neurol Neurosurg Psychiatry 2005;76:1636-–9. 128. Adler CH, Caviness JN, Hentz JG, et al. Randomized trial of modafinil for treating subjective daytime sleepiness in patients with Parkinson's disease. Mov Disord 2003;18:287–93. 129. Devos D, Krystkowiak P, Clement F, et al. Improvement of gait by chronic, high doses of methylphenidate in patients with advanced Parkinson's disease. J Neurol Neurosurg Psychiatry 2007;78:470–5. 130. Tyne HL, Taylor J, Baker GA, Steiger MJ. Modafinil for Parkinson's disease fatigue. J Neurol 2010;257:452–6. 131. Avorn J, Schneeweiss S, Sudarsky LR, et al. Sudden uncontrollable somnolence and medication use in Parkinson disease. Arch Neurol 2005;62:1242–8. 132. Trotti LM, Bliwise DL. No increased risk of obstructive sleep apnea in Parkinson's disease. Mov Disord 2010;25:2246–9. 133. da Silva-Junior FP, do Prado GF, Barbosa ER, et al. Sleep disordered breathing in Parkinson's disease: A critical appraisal. Sleep Med Rev 2013 Jul 22. 134. Noradina AT, Karim NA, Hamidon BB, et al. Sleep-disordered breathing in patients with Parkinson's disease. Singapore Med J 2010;51:60–4. 135. Oerlemans WG, de Weerd AW. The prevalence of sleep disorders in patients with Parkinson's disease. A self-reported, community-based survey. Sleep Med 2002;3:147–9. 136. Low PA. Prevalence of orthostatic hypotension. Clin Auton Res 2008;18 Suppl 1:8–13. 137. Goldstein DS. Orthostatic hypotension as an early finding in Parkinson's disease. Clin Auton Res 2006;16:46–54. 138. Sharabi Y, Goldstein DS. Mechanisms of orthostatic hypotension and supine hypertension in Parkinson disease. J Neurol Sci 2011;310:123–8. 139. Sanchez-Ferro A, Benito-Leon J, Gomez-Esteban JC. The management of orthostatic hypotension in Parkinson's disease. Front Neurol 2013;4:64. 140. Lahrmann H, Cortelli P, Hilz M, et al. EFNS guidelines on the diagnosis and management of orthostatic hypotension. Eur J Neurol 2006;13:930–6. 141. Schoffer KL, Henderson RD, O'Maley K, O'Sullivan JD. Nonpharmacological treatment, fludrocortisone, and domperidone for orthostatic hypotension in Parkinson's disease. Mov Disord 2007;22:1543–9. 142. Singer W, Sandroni P, Opfer-Gehrking TL, et al. Pyridostigmine treatment trial in neurogenic orthostatic hypotension. Arch Neurol 2006;63:513–8. 143. Low PA, Gilden JL, Freeman R, et al. Efficacy of midodrine vs placebo in neurogenic orthostatic hypotension. A randomized, double-blind multicenter study. Midodrine Study Group. JAMA 1997;277:1046–51. 144. Kaufmann H. L-dihydroxyphenylserine (Droxidopa): a new therapy for neurogenic orthostatic hypotension: the US experience. Clin Auton Res 2008;18 Suppl 1:19–24. www.jcomjournal.com

Clinical Review 145. Magerkurth C, Schnitzer R, Braune S. Symptoms of autonomic failure in Parkinson's disease: prevalence and impact on daily life. Clin Auton Res 2005;15:76–82. 146. Makaroff L, Gunn A, Gervasoni C, Richy F. Gastrointestinal Disorders in Parkinson's Disease: Prevalence and Health Outcomes in a US Claims Database. J Parkinsons Dis 2011;1: 65–74. 147. Abbott RD, Petrovitch H, White LR, et al. Frequency of bowel movements and the future risk of Parkinson's disease. Neurology 2001;57:456–62. 148. Abbott RD, Ross GW, Petrovitch H, et al. Bowel movement frequency in late-life and incidental Lewy bodies. Mov Disord 2007;22:1581–6. 149. Zangaglia R, Martignoni E, Glorioso M, et al. Macrogol for the treatment of constipation in Parkinson's disease. A randomized placebo-controlled study. Mov Disord 2007;22:1239–44. 150. Ondo WG, Kenney C, Sullivan K, et al. Placebo-controlled trial of lubiprostone for constipation associated with Parkinson disease. Neurology 2012;78:1650–4. 151. Cersosimo MG, Benarroch EE. Neural control of the gastrointestinal tract: implications for Parkinson disease. Mov Disord 2008;23:1065–75. 152. Reddymasu SC, Soykan I, McCallum RW. Domperidone: review of pharmacology and clinical applications in gastroenterology. Am J Gastroenterol 2007;102:2036–45. 153. Argolo N, Sampaio M, Pinho P, et al. Do swallowing exercises improve swallowing dynamic and quality of life in Parkinson's disease? NeuroRehabilitation 2013;32:949–55. 154. Troche MS, Sapienza CM, Rosenbek JC. Effects of bolus consistency on timing and safety of swallow in patients with Parkinson's disease. Dysphagia 2008;23:26–32. 155. Edwards LL, Quigley EM, Pfeiffer RF. Gastrointestinal dysfunction in Parkinson's disease: frequency and pathophysiology. Neurology 1992;42:726–32. 156. Kalf JG, Smit AM, Bloem BR, et al. Impact of drooling in Parkinson's disease. J Neurol 2007;254:1227–32. 157. Leibner J, Ramjit A, Sedig L, et al. The impact of and the factors associated with drooling in Parkinson's disease. Parkinsonism Relat Disord 2010;16:475–7. 158. Mancini F, Zangaglia R, Cristina S, et al. Double-blind, placebo-controlled study to evaluate the efficacy and safety of botulinum toxin type A in the treatment of drooling in parkinsonism. Mov Disord 2003;18:685–8. 159. Lagalla G, Millevolte M, Capecci M, et al. Long-lasting benefits of botulinum toxin type B in Parkinson's disease-related drooling. J Neurol 2009;256:563–7. 160. Lagalla G, Millevolte N, Capecci M, et al. Botulinum toxin type A for drooling in Parkinson's disease: a doubleblind, randomized, placebo-controlled study. Mov Disord 2006;21:704–7. 161. Arbouw ME, Movig KL, Koopmann M, et al. Glycopyrrolate for sialorrhea in Parkinson disease: a randomized, doubleblind, crossover trial. Neurology 2010; 74:1203–7. 162. South AR, Somers, Jog MS. Gum chewing improves swallow frequency and latency in Parkinson patients: a preliminary study. Neurology 2010;74:1198–202. www.jcomjournal.com

163. Winge K, Nielsen KK. Bladder dysfunction in advanced Parkinson's disease. Neurourol Urodyn 2012;31:1279–83. 164. Madhuvrata P, Singh PM, Hasafa Z, Abdel-Fattah M. Anticholinergic drugs for adult neurogenic detrusor overactivity: a systematic review and meta-analysis. Eur Urol 2012;62: 816–30. 165. Soljanik I. Efficacy and safety of botulinum toxin a intradetrusor injections in adults with neurogenic detrusor overactivity/ neurogenic overactive bladder: a systematic review. Drugs 2013;73:1055–66. 166. Sakakibara R, Uchiyama T, Yamanishi T, Kishi M. Genitourinary dysfunction in Parkinson's disease. Mov Disord 2010; 25:2–12. 167. Kummer A, Cardoso F, Teixeira AL. Loss of libido in Parkinson's disease. J Sex Med 2009;6:1024–31. 168. Lombardi G, Nelli F, Celso M, et al. Treating erectile dysfunction and central neurological diseases with oral phosphodiesterase type 5 inhibitors. Review of the literature. J Sex Med 2012 9(4): 970–85. 169. Dula E, Bukofzer S, Perdok R, George M. Double-blind, crossover comparison of 3 mg apomorphine SL with placebo and with 4 mg apomorphine SL in male erectile dysfunction. Eur Urol 2001;39(5): 558–3 170. Negre-Pages L, Regragui W, Bouhassira D, et al. Chronic pain in Parkinson's disease: the cross-sectional French DoPaMiP survey. Mov Disord 2008;23:1361–9. 171. Beiske AG, Loge JH, Ronningen A, Svensson E. Pain in Parkinson's disease: Prevalence and characteristics. Pain 2009. 141:173–7. 172. Lin CH, Wu RM, H.Y. Chang HY, et al. Preceding pain symptoms and Parkinson's disease: a nationwide populationbased cohort study. Eur J Neurol 2013;20:1398–404. 173. Ha AD, Jankovic J. Pain in Parkinson's disease. Mov Disord 2012; 27:485–91. 174. Ford B. Pain in Parkinson's disease. Mov Disord 2010;25 Suppl 1:S98–103. 175. Loher TJ, Burgunder JM, Weber S, et al. Effect of chronic pallidal deep brain stimulation on off period dystonia and sensory symptoms in advanced Parkinson's disease. J Neurol Neurosurg Psychiatry 2002;73:395–9. 176. Ramirez-Ruiz B, Marti MJ, Tolosa E, et al. Longitudinal evaluation of cerebral morphological changes in Parkinson's disease with and without dementia. J Neurol 2005;252: 1345–52. 177. Jellinger KA. Formation and development of Lewy pathology: a critical update. J Neurol 2009;256 Suppl 3:270–9. 178. Levy R, Dubois B. Apathy and the functional anatomy of the prefrontal cortex-basal ganglia circuits. Cereb Cortex 2006;16:916–28. 179. Connor JR, Boyer PJ, Menzies SL, et al. Neuropathological examination suggests impaired brain iron acquisition in restless legs syndrome. Neurology 2003;61: 304–9. 180. Earley CJ, Barker P, Horská A, Allen RP. MRI-determined regional brain iron concentrations in early- and late-onset restless legs syndrome. Sleep Med 2006;7: 458–61. 181. Saper CB, Chou TC, Scammell TE. The sleep switch: hypothalamic control of sleep and wakefulness. Trends Neurosci

Vol. 21, No. 2 February 2014 JCOM 85

Nonmotor Symptoms of Parkinson’s Disease 2001;24:726–31. 182. Boeve BF, Silber MH, Saper CB, et al. Pathophysiology of REM sleep behaviour disorder and relevance to neurodegenerative disease. Brain 2007;130(Pt 11): 2770–88. 183. Benarroch EE, Schmeichel AM, Paris J. Involvement of the ventrolateral medulla in parkinsonism with autonomic failure. Neurology 2000;54:963–8.

184. Chudler EH, Dong WK. The role of the basal ganglia in nociception and pain. Pain 1995;60:3–38. 185. Wolters E. Non-motor extranigral signs and symptoms in Parkinson's disease. Parkinsonism Relat Disord 2009;15 Suppl 3:S6–12. 186. Davidsdottir S, Cronin-Golomb A, Lee A. Visual and spatial symptoms in Parkinson's disease. Vision Res 2005;45:1285–96.

Copyright 2014 by Turner White Communications Inc., Wayne, PA. All rights reserved.

86 JCOM February 2014 Vol. 21, No. 2

www.jcomjournal.com

Suggest Documents