Sample records for narcolepsy

  1. [NARCOLEPSY WITH CATAPLEXY: TYPE 1 NARCOLEPSY].

    PubMed

    Dauvilliers, Yves; Lopez, Régis

    2016-06-01

    Narcolepsy with cataplexy or narcolepsy type 1 in a rare, disabling sleep disorder, with a prevalence of 20 to 30 per 100,000. Its onset peaks in the second decade. The main features are excessive daytime sleepiness and cataplexy or sudden less of muscle tone triggered by emotional situations. Other less consistent symptoms include hypnagogic hallucinations, sleep paralysis, disturbed nighttime sleep, and weight gain. Narcolepsy with cataplexy remains a clinical diagnosis but nighttime and daytime polysomnography (multiple sleep latency tests) are useful to document mean sleep latency below 8 min and at least two sleep-onset REM periods. HLA typing shows an association with HLA DQB1*0602 in more than 92% of cases but was not included in the new diagnostic criteria. In contrast, a low hypocretin-1/orexin-A levels (values below 110 pg/mL) in the cerebrospinal fluid was highly specific for narcolepsy with cataplexy and was included in the recent diagnostic criteria for narcolepsy. The deficiency of the hypocretin system is well-established in human narcoleptics with a reduction of cerebrospinal fluid hypocretin levels in relation with an early loss of hypocretin neurons. The cause of human narcolepsy remains unknown, however an autoimmune process in most probable acting on a highly genetic background with environmental factors such as streptococcal infections, and H1N1 AS03-adjuvanted vaccine named Pandemrix.

  2. Narcolepsy: a review

    PubMed Central

    Akintomide, Gbolagade Sunmaila; Rickards, Hugh

    2011-01-01

    Narcolepsy is a lifelong sleep disorder characterized by a classic tetrad of excessive daytime sleepiness with irresistible sleep attacks, cataplexy (sudden bilateral loss of muscle tone), hypnagogic hallucination, and sleep paralysis. There are two distinct groups of patients, ie, those having narcolepsy with cataplexy and those having narcolepsy without cataplexy. Narcolepsy affects 0.05% of the population. It has a negative effect on the quality of life of its sufferers and can restrict them from certain careers and activities. There have been advances in the understanding of the pathogenesis of narcolepsy. It is thought that narcolepsy with cataplexy is secondary to loss of hypothalamic hypocretin neurons in those genetically predisposed to the disorder by possession of human leukocyte antigen DQB1*0602. The diagnostic criteria for narcolepsy are based on symptoms, laboratory sleep tests, and serum levels of hypocretin. There is no cure for narcolepsy, and the present mainstay of treatment is pharmacological treatment along with lifestyle changes. Some novel treatments are also being developed and tried. This article critically appraises the evidence for diagnosis and treatment of narcolepsy. PMID:21931493

  3. Misdiagnosis of narcolepsy.

    PubMed

    Dunne, Laura; Patel, Pallavi; Maschauer, Emily L; Morrison, Ian; Riha, Renata L

    2016-12-01

    Narcolepsy is a chronic primary sleep disorder, characterized by excessive daytime sleepiness and sleep dysfunction with or without cataplexy. Narcolepsy is uncommon, with a low prevalence rate which makes it difficult to diagnose definitively without a complex series of tests and a detailed history. The aim of this study was to review patients referred to a tertiary sleep centre who had been labelled with a diagnosis of narcolepsy prior to referral in order to assess if the diagnosis was accurate, and if not, to determine the cause of diagnostic misattribution. All patients seen at a sleep centre from 2007-2013 (n = 551) who underwent detailed objective testing including an MSLT PSG, as well as wearing an actigraphy watch and completing a sleep diary for 2 weeks, were assessed for a pre-referral and final diagnosis of narcolepsy. Of the 41 directly referred patients with a diagnostic label of narcolepsy, 19 (46 %) were subsequently confirmed to have narcolepsy on objective testing and assessment by a sleep physician using ICSD-2 criteria. The diagnosis of narcolepsy was incorrectly attributed to almost 50 % of patients labelled with a diagnosis of narcolepsy who were referred for further opinion by a variety of specialists and generalists. Accurate diagnosis of narcolepsy is critical for many reasons, such as the impact it has on quality of life, driving, employment, insurance and pregnancy in women as well as medication management.

  4. Medical comorbidity in narcolepsy: findings from the Burden of Narcolepsy Disease (BOND) study.

    PubMed

    Black, J; Reaven, N L; Funk, S E; McGaughey, K; Ohayon, M M; Guilleminault, C; Ruoff, C

    2017-05-01

    The objective of this study was to evaluate medical comorbidity patterns in patients with a narcolepsy diagnosis in the United States. This was a retrospective medical claims data analysis. Truven Health Analytics MarketScan® Research Databases were accessed to identify individuals ≥18 years of age with ≥1 diagnosis code for narcolepsy (International Classification of Diseases (ICD)-9, 347.0, 347.00, 347.01, 347.1, 347.10, or 347.11) continuously insured between 2006 and 2010, and controls without narcolepsy matched 5:1 on age, gender, region, and payer. Narcolepsy and control subjects were compared for frequency of comorbid conditions, identified by the appearance of >1 diagnosis code(s) mapped to a Clinical Classification System (CCS) level 1 category any time during the study period, and on specific subcategories, including recognized narcolepsy comorbidities of obstructive sleep apnea (OSA) and depression. The final study group included 9312 subjects with narcolepsy and 46,559 controls (each group: average age, 46.1 years; 59% female). As compared with controls, patients with narcolepsy showed a statistically significant excess prevalence in all the CCS multilevel categories, the only exceptions being conditions originating in the perinatal period and pregnancy/childbirth complications. The greatest excess prevalence in the narcolepsy cohort was seen for mental illness (31.1% excess prevalence; odds ratio (OR) 3.8, 95% confidence interval (CI) 3.6, 4.0), followed by diseases of the digestive system (21.4% excess prevalence; OR 2.7, 95% CI 2.5, 2.8) and nervous system/sense organs (excluding narcolepsy; 20.7% excess prevalence; OR 3.7, 95% CI 3.4, 3.9). In this claims analysis, a narcolepsy diagnosis was associated with a wide range of comorbid medical illness claims, at significantly higher rates than matched controls. Copyright © 2016. Published by Elsevier B.V.

  5. Disrupted Nighttime Sleep in Narcolepsy

    PubMed Central

    Roth, Thomas; Dauvilliers, Yves; Mignot, Emmanuel; Montplaisir, Jacques; Paul, Josh; Swick, Todd; Zee, Phyllis

    2013-01-01

    Study Objectives: Characterize disrupted nighttime sleep (DNS) in narcolepsy, an important symptom of narcolepsy. Methods: A panel of international narcolepsy experts was convened in 2011 to build a consensus characterization of DNS in patients with narcolepsy. A literature search of the Medline (1965 to date), Medline In-Process (latest weeks), Embase (1974 to date), Embase Alert (latest 8 weeks), and Biosis (1965 to date) databases was conducted using the following search terms: narcolepsy and disrupted nighttime sleep, disturbed nighttime sleep, fragmented sleep, consolidated sleep, sleep disruption, and narcolepsy questionnaire. The purpose of the literature search was to identify publications characterizing the nighttime sleep of patients with narcolepsy. The panel reviewed the literature. Nocturnal sleep can also be disturbed by REM sleep abnormalities such as vivid dreaming and REM sleep behavior disorder; however, these were not reviewed in the current paper, as we were evaluating for idiopathic sleep disturbances. Results: The literature reviewed provide a consistent characterization of nighttime sleep in patients with narcolepsy as fragmented, with reports of frequent, brief nightly awakenings with difficulties returning to sleep and associated reports of poor sleep quality. Polysomnographic studies consistently report frequent awakenings/arousals after sleep onset, more stage 1 (S1) sleep, and more frequent shifts to S1 sleep or wake from deeper stages of sleep. The consensus of the International Experts' Panel on Narcolepsy was that DNS can be distressing for patients with narcolepsy and that treatment of DNS warrants consideration. Conclusions: Clinicians involved in the management of patients with narcolepsy should investigate patients' quality of nighttime sleep, give weight and consideration to patient reports of nighttime sleep experience, and consider DNS a target for treatment. Citation: Roth T; Dauvilliers Y; Mignot E; Montplaisir J; Paul J

  6. Disrupted nighttime sleep in narcolepsy.

    PubMed

    Roth, Thomas; Dauvilliers, Yves; Mignot, Emmanuel; Montplaisir, Jacques; Paul, Josh; Swick, Todd; Zee, Phyllis

    2013-09-15

    Characterize disrupted nighttime sleep (DNS) in narcolepsy, an important symptom of narcolepsy. A panel of international narcolepsy experts was convened in 2011 to build a consensus characterization of DNS in patients with narcolepsy. A literature search of the Medline (1965 to date), Medline In-Process (latest weeks), Embase (1974 to date), Embase Alert (latest 8 weeks), and Biosis (1965 to date) databases was conducted using the following search terms: narcolepsy and disrupted nighttime sleep, disturbed nighttime sleep, fragmented sleep, consolidated sleep, sleep disruption, and narcolepsy questionnaire. The purpose of the literature search was to identify publications characterizing the nighttime sleep of patients with narcolepsy. The panel reviewed the literature. Nocturnal sleep can also be disturbed by REM sleep abnormalities such as vivid dreaming and REM sleep behavior disorder; however, these were not reviewed in the current paper, as we were evaluating for idiopathic sleep disturbances. The literature reviewed provide a consistent characterization of nighttime sleep in patients with narcolepsy as fragmented, with reports of frequent, brief nightly awakenings with difficulties returning to sleep and associated reports of poor sleep quality. Polysomnographic studies consistently report frequent awakenings/arousals after sleep onset, more stage 1 (S1) sleep, and more frequent shifts to S1 sleep or wake from deeper stages of sleep. The consensus of the International Experts' Panel on Narcolepsy was that DNS can be distressing for patients with narcolepsy and that treatment of DNS warrants consideration. Clinicians involved in the management of patients with narcolepsy should investigate patients' quality of nighttime sleep, give weight and consideration to patient reports of nighttime sleep experience, and consider DNS a target for treatment.

  7. Going to School with Narcolepsy--Perceptions of Families and Teachers of Children with Narcolepsy

    ERIC Educational Resources Information Center

    Karjalainen, Satu; Nyrhilä, Anna-Maria; Määttä, Kaarina; Uusiautti, Satu

    2014-01-01

    In 2009 and 2010, a large group of Finnish children and adolescents got narcolepsy after the vaccination campaign to prevent swine flu pandemic. A sample of children and adolescents who had gotten narcolepsy after 2009 participated in this study. The purpose of this research was to analyse how narcolepsy that developed from the swine flu…

  8. Autoimmunity in narcolepsy.

    PubMed

    Bonvalet, Melodie; Ollila, Hanna M; Ambati, Aditya; Mignot, Emmanuel

    2017-11-01

    Summarize the recent findings in narcolepsy focusing on the environmental and genetic risk factors in disease development. Both genetic and epidemiological evidence point towards an autoimmune mechanism in the destruction of orexin/hypocretin neurons. Recent studies suggest both humoral and cellular immune responses in the disease development. Narcolepsy is a severe sleep disorder, in which neurons producing orexin/hypocretin in the hypothalamus are destroyed. The core symptoms of narcolepsy are debilitating, extreme sleepiness, cataplexy, and abnormalities in the structure of sleep. Both genetic and epidemiological evidence point towards an autoimmune mechanism in the destruction of orexin/hypocretin neurons. Importantly, the highest environmental risk is seen with influenza-A infection and immunization. However, how the cells are destroyed is currently unknown. In this review we summarize the disease symptoms, and focus on the immunological findings in narcolepsy. We also discuss the environmental and genetic risk factors as well as propose a model for disease development.

  9. Clinical and Neurobiological Aspects of Narcolepsy

    PubMed Central

    Nishino, Seiji

    2007-01-01

    Narcolepsy is characterized by excessive daytime sleepiness (EDS), cataplexy and/or other dissociated manifestations of rapid eye movement (REM) sleep (hypnagogic hallucinations and sleep paralysis). Narcolepsy is currently treated with amphetamine-like central nervous system (CNS) stimulants (for EDS) and antidepressants (for cataplexy). Some other classes of compounds such as modafinil (a non-amphetamine wake-promoting compound for EDS) and gamma-hydroxybutyrate (GHB, a short-acting sedative for EDS/fragmented nighttime sleep and cataplexy) given at night are also employed. The major pathophysiology of human narcolepsy has been recently elucidated based on the discovery of narcolepsy genes in animals. Using forward (i.e., positional cloning in canine narcolepsy) and reverse (i.e., mouse gene knockout) genetics, the genes involved in the pathogenesis of narcolepsy (hypocretin/orexin ligand and its receptor) in animals have been identified. Hypocretins/orexins are novel hypothalamic neuropeptides also involved in various hypothalamic functions such as energy homeostasis and neuroendocrine functions. Mutations in hypocretin-related genes are rare in humans, but hypocretin-ligand deficiency is found in many narcolepsy-cataplexy cases. In this review, the clinical, pathophysiological and pharmacological aspects of narcolepsy are discussed. PMID:17470414

  10. Increased lucid dreaming frequency in narcolepsy.

    PubMed

    Rak, Michael; Beitinger, Pierre; Steiger, Axel; Schredl, Michael; Dresler, Martin

    2015-05-01

    Nightmares are a frequent symptom in narcolepsy. Lucid dreaming, i.e., the phenomenon of becoming aware of the dreaming state during dreaming, has been demonstrated to be of therapeutic value for recurrent nightmares. Data on lucid dreaming in narcolepsy patients, however, is sparse. The aim of this study was to evaluate the frequency of recalled dreams (DF), nightmares (NF), and lucid dreams (LDF) in narcolepsy patients compared to healthy controls. In addition, we explored if dream lucidity provides relief during nightmares in narcolepsy patients. We interviewed patients with narcolepsy and healthy controls. Telephone interview. 60 patients diagnosed with narcolepsy (23-82 years, 35 females) and 919 control subjects (14-93 years, 497 females). N/A. Logistic regression revealed significant (P < 0.001) differences in DF, NF, and LDF between narcolepsy patients and controls after controlling for age and gender, with effect sizes lying in the large range (Cohen's d > 0.8). The differences in NF and LDF between patients and controls stayed significant after controlling for DF. Comparison of 35 narcolepsy patients currently under medication with their former drug-free period revealed significant differences in DF and NF (z < 0.05, signed-rank test) but not LDF (z = 0.8). Irrespective of medication, 70% of narcolepsy patients with experience in lucid dreaming indicated that dream lucidity provides relief during nightmares. Narcolepsy patients experience a markedly higher lucid dreaming frequency compared to controls, and many patients report a positive impact of dream lucidity on the distress experienced from nightmares. © 2015 Associated Professional Sleep Societies, LLC.

  11. Increased Lucid Dreaming Frequency in Narcolepsy

    PubMed Central

    Rak, Michael; Beitinger, Pierre; Steiger, Axel; Schredl, Michael; Dresler, Martin

    2015-01-01

    Study Objective: Nightmares are a frequent symptom in narcolepsy. Lucid dreaming, i.e., the phenomenon of becoming aware of the dreaming state during dreaming, has been demonstrated to be of therapeutic value for recurrent nightmares. Data on lucid dreaming in narcolepsy patients, however, is sparse. The aim of this study was to evaluate the frequency of recalled dreams (DF), nightmares (NF), and lucid dreams (LDF) in narcolepsy patients compared to healthy controls. In addition, we explored if dream lucidity provides relief during nightmares in narcolepsy patients. Design: We interviewed patients with narcolepsy and healthy controls. Setting: Telephone interview. Patients: 60 patients diagnosed with narcolepsy (23–82 years, 35 females) and 919 control subjects (14–93 years, 497 females) Interventions: N/A. Measurements and Results: Logistic regression revealed significant (P < 0.001) differences in DF, NF, and LDF between narcolepsy patients and controls after controlling for age and gender, with effect sizes lying in the large range (Cohen's d > 0.8). The differences in NF and LDF between patients and controls stayed significant after controlling for DF. Comparison of 35 narcolepsy patients currently under medication with their former drug-free period revealed significant differences in DF and NF (z < 0.05, signed-rank test) but not LDF (z = 0.8). Irrespective of medication, 70% of narcolepsy patients with experience in lucid dreaming indicated that dream lucidity provides relief during nightmares. Conclusion: Narcolepsy patients experience a markedly higher lucid dreaming frequency compared to controls, and many patients report a positive impact of dream lucidity on the distress experienced from nightmares. Citation: Rak M, Beitinger P, Steiger A, Schredl M, Dresler M. Increased lucid dreaming frequency in narcolepsy. SLEEP 2015;38(5):787–792. PMID:25325481

  12. ApoE polymorphisms in narcolepsy

    PubMed Central

    Gencik, Martin; Dahmen, Norbert; Wieczorek, Stefan; Kasten, Meike; Gencikova, Alexandra; Epplen, Jorg T

    2001-01-01

    Background Narcolepsy is a common neuropsychiatric disorder characterized by increased daytime sleepiness, cataplexy and hypnagogic hallucinations. Deficiency of the hypocretin neurotransmitter system was shown to be involved in the pathogenesis of narcolepsy in animals and men. There are several hints that neurodegeneration of hypocretin producing neurons in the hypothalamus is the pathological correlate of narcolepsy. The ApoE4 allele is a major contributing factor to early-onset neuronal degeneration in Alzheimer disease and other neurodegenerative diseases as well. Methods To clarify whether the ApoE4 phenotype predisposes to narcolepsy or associates with an earlier disease onset, we have genotyped the ApoE gene in 103 patients with narcolepsy and 101 healthy controls. Results The frequency of the E4 allele of the ApoE gene was 11% in the patient and 15% in the control groups. Furthermore, the mean age of onset did not differ between the ApoE4+ and ApoE4- patient groups. Conclusion Our results exclude the ApoE4 allele as a major risk factor for narcolepsy. PMID:11560764

  13. Lucid Dreaming in Narcolepsy

    PubMed Central

    Dodet, Pauline; Chavez, Mario; Leu-Semenescu, Smaranda; Golmard, Jean-Louis; Arnulf, Isabelle

    2015-01-01

    Objective: To evaluate the frequency, determinants and sleep characteristics of lucid dreaming in narcolepsy Settings: University hospital sleep disorder unit Design: Case-control study Participants: Consecutive patients with narcolepsy and healthy controls Methods: Participants were interviewed regarding the frequency and determinants of lucid dreaming. Twelve narcolepsy patients and 5 controls who self-identified as frequent lucid dreamers underwent nighttime and daytime sleep monitoring after being given instructions regarding how to give an eye signal when lucid. Results: Compared to 53 healthy controls, the 53 narcolepsy patients reported more frequent dream recall, nightmares and recurrent dreams. Lucid dreaming was achieved by 77.4% of narcoleptic patients and 49.1% of controls (P < 0.05), with an average of 7.6 ± 11 vs. 0.3 ± 0.8 lucid dreams/month (P < 0.0001). The frequency of cataplexy, hallucinations, sleep paralysis, dyssomnia, HLA positivity, and the severity of sleepiness were similar in narcolepsy with and without lucid dreaming. Seven of 12 narcoleptic (and 0 non-narcoleptic) lucid dreamers achieved lucid REM sleep across a total of 33 naps, including 14 episodes with eye signal. The delta power in the electrode average, in delta, theta, and alpha powers in C4, and coherences between frontal electrodes were lower in lucid than non-lucid REM sleep in spectral EEG analysis. The duration of REM sleep was longer, the REM sleep onset latency tended to be shorter, and the percentage of atonia tended to be higher in lucid vs. non-lucid REM sleep; the arousal index and REM density and amplitude were unchanged. Conclusion: Narcoleptics have a high propensity for lucid dreaming without differing in REM sleep characteristics from people without narcolepsy. This suggests narcolepsy patients may provide useful information in future studies on the nature of lucid dreaming. Citation: Dodet P, Chavez M, Leu-Semenescu S, Golmard JL, Arnulf I. Lucid dreaming in

  14. Narcolepsy in African Americans

    PubMed Central

    Kawai, Makoto; O'Hara, Ruth; Einen, Mali; Lin, Ling; Mignot, Emmanuel

    2015-01-01

    Study Objectives: Although narcolepsy affects 0.02–0.05% of individuals in various ethnic groups, clinical presentation in different ethnicities has never been fully characterized. Our goal was to study phenotypic expression across ethnicities in the United States. Design/Setting: Cases of narcolepsy from 1992 to 2013 were identified from searches of the Stanford Center for Narcolepsy Research database. International Classification of Sleep Disorders, Third Edition diagnosis criteria for type 1 and type 2 narcolepsy were used for inclusion, but subjects were separated as with and without cataplexy for the purpose of data presentation. Information extracted included demographics, ethnicity and clinical data, HLA-DQB1*06:02, polysomnography (PSG), multiple sleep latency test (MSLT) data, and cerebrospinal fluid (CSF) hypocretin-1 level. Patients: 182 African-Americans, 839 Caucasians, 35 Asians, and 41 Latinos with narcolepsy. Results: Sex ratio, PSG, and MSLT findings did not differ across ethnicities. Epworth Sleepiness Scale (ESS) score was higher and age of onset of sleepiness earlier in African Americans compared with other ethnicities. HLA-DQB1*06:02 positivity was higher in African Americans (91.0%) versus others (76.6% in Caucasians, 80.0% in Asians, and 65.0% in Latinos). CSF hypocretin-1 level, obtained in 222 patients, was more frequently low (≤ 110 pg/ml) in African Americans (93.9%) versus Caucasians (61.5%), Asians (85.7%) and Latinos (75.0%). In subjects with low CSF hypocretin-1, African Americans (28.3%) were 4.5 fold more likely to be without cataplexy when compared with Caucasians (8.1%). Conclusions: Narcolepsy in African Americans is characterized by earlier symptom onset, higher Epworth Sleepiness Scale score, higher HLA-DQB1*06:02 positivity, and low cerebrospinal fluid hypocretin-1 level in the absence of cataplexy. In African Americans, more subjects without cataplexy have type 1 narcolepsy. Citation: Kawai M, O'Hara R, Einen M, Lin L

  15. Measurement of narcolepsy symptoms: The Narcolepsy Severity Scale.

    PubMed

    Dauvilliers, Yves; Beziat, Severine; Pesenti, Carole; Lopez, Regis; Barateau, Lucie; Carlander, Bertrand; Luca, Gianina; Tafti, Mehdi; Morin, Charles M; Billiard, Michel; Jaussent, Isabelle

    2017-04-04

    To validate the Narcolepsy Severity Scale (NSS), a brief clinical instrument to evaluate the severity and consequences of symptoms in patients with narcolepsy type 1 (NT1). A 15-item scale to assess the frequency and severity of excessive daytime sleepiness, cataplexy, hypnagogic hallucinations, sleep paralysis, and disrupted nighttime sleep was developed and validated by sleep experts with patients' feedback. Seventy untreated and 146 treated adult patients with NT1 were evaluated and completed the NSS in a single reference sleep center. The NSS psychometric properties, score changes with treatment, and convergent validity with other clinical parameters were assessed. The NSS showed good psychometric properties with significant item-total score correlations. The factor analysis indicated a 3-factor solution with good reliability, expressed by satisfactory Cronbach α values. The NSS total score temporal stability was good. Significant NSS score differences were observed between untreated and treated patients (dependent sample, 41 patients before and after sleep therapy; independent sample, 29 drug-free and 105 treated patients). Scores were lower in the treated populations (10-point difference between groups), without ceiling effect. Significant correlations were found among NSS total score and daytime sleepiness (Epworth Sleepiness Scale, Mean Sleep Latency Test), depressive symptoms, and health-related quality of life. The NSS can be considered a reliable and valid clinical tool for the quantification of narcolepsy symptoms to monitor and optimize narcolepsy management. © 2017 American Academy of Neurology.

  16. Multiple sleep latency test in narcolepsy type 1 and narcolepsy type 2: A 5-year follow-up study.

    PubMed

    Huang, Yu-Shu; Guilleminault, Christian; Lin, Cheng-Hui; Chen, Chia-Hsiang; Chin, Wei-Chih; Chen, Tzu-Shuang

    2018-05-29

    Excessively sleepy teenagers and young adults without sleep-disordered breathing are diagnosed with either narcolepsy type 1 or narcolepsy type 2, or hypersomnia, based on the presence/absence of cataplexy and the results of a multiple sleep latency test. However, there is controversy surrounding this nomenclature. We will try to find the differences between different diagnoses of hypersomnia from the results of the long-term follow-up evaluation of a sleep study. We diagnosed teenagers who had developed excessive daytime sleepiness based on the criteria of the International Classification of Sleep Disorders, 3rd edition. Each individual received the same clinical neurophysiologic testing every year for 5 years after the initial diagnosis of narcolepsy type 1 (n = 111) or type 2 (n = 46). The follow-up evaluation demonstrated that narcolepsy type 1 (narcolepsy-cataplexy) is a well-defined clinical entity, with very reproducible clinical neurophysiologic findings over time, whereas patients with narcolepsy type 2 presented clear clinical and test variability. By the fifth year of the follow-up evaluation, 17.6% of subjects did not meet the diagnostic criteria of narcolepsy type 2, and 23.9% didn't show any two sleep-onset rapid eye movement periods in multiple sleep latency during the 5-year follow-up. Therefore narcolepsy type 1 (narcolepsy-cataplexy) is a well-defined syndrome, with the presentation clearly related to the known consequences of destruction of hypocretin/orexin neurons. Narcolepsy type 2 covers patients with clinical and test variability over time, thus bringing into question the usage of the term "narcolepsy" to label these patients. © 2018 European Sleep Research Society.

  17. Narcolepsy in African Americans.

    PubMed

    Kawai, Makoto; O'Hara, Ruth; Einen, Mali; Lin, Ling; Mignot, Emmanuel

    2015-11-01

    Although narcolepsy affects 0.02-0.05% of individuals in various ethnic groups, clinical presentation in different ethnicities has never been fully characterized. Our goal was to study phenotypic expression across ethnicities in the United States. Cases of narcolepsy from 1992 to 2013 were identified from searches of the Stanford Center for Narcolepsy Research database. International Classification of Sleep Disorders, Third Edition diagnosis criteria for type 1 and type 2 narcolepsy were used for inclusion, but subjects were separated as with and without cataplexy for the purpose of data presentation. Information extracted included demographics, ethnicity and clinical data, HLA-DQB1*06:02, polysomnography (PSG), multiple sleep latency test (MSLT) data, and cerebrospinal fluid (CSF) hypocretin-1 level. 182 African-Americans, 839 Caucasians, 35 Asians, and 41 Latinos with narcolepsy. Sex ratio, PSG, and MSLT findings did not differ across ethnicities. Epworth Sleepiness Scale (ESS) score was higher and age of onset of sleepiness earlier in African Americans compared with other ethnicities. HLA-DQB1*06:02 positivity was higher in African Americans (91.0%) versus others (76.6% in Caucasians, 80.0% in Asians, and 65.0% in Latinos). CSF hypocretin-1 level, obtained in 222 patients, was more frequently low (≤ 110 pg/ml) in African Americans (93.9%) versus Caucasians (61.5%), Asians (85.7%) and Latinos (75.0%). In subjects with low CSF hypocretin-1, African Americans (28.3%) were 4.5 fold more likely to be without cataplexy when compared with Caucasians (8.1%). Narcolepsy in African Americans is characterized by earlier symptom onset, higher Epworth Sleepiness Scale score, higher HLA-DQB1*06:02 positivity, and low cerebrospinal fluid hypocretin-1 level in the absence of cataplexy. In African Americans, more subjects without cataplexy have type 1 narcolepsy. © 2015 Associated Professional Sleep Societies, LLC.

  18. Treatment Options for Narcolepsy.

    PubMed

    Barateau, Lucie; Lopez, Régis; Dauvilliers, Yves

    2016-05-01

    Narcolepsy type 1 and narcolepsy type 2 are central disorders of hypersomnolence. Narcolepsy type 1 is characterized by excessive daytime sleepiness and cataplexy and is associated with hypocretin-1 deficiency. On the other hand, in narcolepsy type 2, cerebrospinal fluid hypocretin-1 levels are normal and cataplexy absent. Despite major advances in our understanding of narcolepsy mechanisms, its current management is only symptomatic. Treatment options may vary from a single drug that targets several symptoms, or multiple medications that each treats a specific symptom. In recent years, narcolepsy treatment has changed with the widespread use of modafinil/armodafinil for daytime sleepiness, antidepressants (selective serotonin and dual serotonin and noradrenalin reuptake inhibitors) for cataplexy, and sodium oxybate for both symptoms. Other psychostimulants can also be used, such as methylphenidate, pitolisant and rarely amphetamines, as third-line therapy. Importantly, clinically relevant subjective and objective measures of daytime sleepiness are required to monitor the treatment efficacy and to provide guidance on whether the treatment goals are met. Associated symptoms and comorbid conditions, such as hypnagogic/hypnopompic hallucinations, sleep paralysis, disturbed nighttime sleep, unpleasant dreams, REM- and non REM-related parasomnias, depressive symptoms, overweight/obesity, and obstructive sleep apnea, should also be taken into account and managed, if required. In the near future, the efficacy of new wake-promoting drugs, anticataplectic agents, hypocretin replacement therapy and immunotherapy at the early stages of the disease should also be evaluated.

  19. Lucid dreaming in narcolepsy.

    PubMed

    Dodet, Pauline; Chavez, Mario; Leu-Semenescu, Smaranda; Golmard, Jean-Louis; Arnulf, Isabelle

    2015-03-01

    To evaluate the frequency, determinants and sleep characteristics of lucid dreaming in narcolepsy. University hospital sleep disorder unit. Case-control study. Consecutive patients with narcolepsy and healthy controls. Participants were interviewed regarding the frequency and determinants of lucid dreaming. Twelve narcolepsy patients and 5 controls who self-identified as frequent lucid dreamers underwent nighttime and daytime sleep monitoring after being given instructions regarding how to give an eye signal when lucid. Compared to 53 healthy controls, the 53 narcolepsy patients reported more frequent dream recall, nightmares and recurrent dreams. Lucid dreaming was achieved by 77.4% of narcoleptic patients and 49.1% of controls (P < 0.05), with an average of 7.6±11 vs. 0.3±0.8 lucid dreams/ month (P < 0.0001). The frequency of cataplexy, hallucinations, sleep paralysis, dyssomnia, HLA positivity, and the severity of sleepiness were similar in narcolepsy with and without lucid dreaming. Seven of 12 narcoleptic (and 0 non-narcoleptic) lucid dreamers achieved lucid REM sleep across a total of 33 naps, including 14 episodes with eye signal. The delta power in the electrode average, in delta, theta, and alpha powers in C4, and coherences between frontal electrodes were lower in lucid than non-lucid REM sleep in spectral EEG analysis. The duration of REM sleep was longer, the REM sleep onset latency tended to be shorter, and the percentage of atonia tended to be higher in lucid vs. non-lucid REM sleep; the arousal index and REM density and amplitude were unchanged. Narcolepsy is a novel, easy model for studying lucid dreaming. © 2015 Associated Professional Sleep Societies, LLC.

  20. The European Narcolepsy Network (EU-NN) database.

    PubMed

    Khatami, Ramin; Luca, Gianina; Baumann, Christian R; Bassetti, Claudio L; Bruni, Oliviero; Canellas, Francesca; Dauvilliers, Yves; Del Rio-Villegas, Rafael; Feketeova, Eva; Ferri, Raffaele; Geisler, Peter; Högl, Birgit; Jennum, Poul; Kornum, Birgitte R; Lecendreux, Michel; Martins-da-Silva, Antonio; Mathis, Johannes; Mayer, Geert; Paiva, Teresa; Partinen, Markku; Peraita-Adrados, Rosa; Plazzi, Guiseppe; Santamaria, Joan; Sonka, Karel; Riha, Renata; Tafti, Mehdi; Wierzbicka, Aleksandra; Young, Peter; Lammers, Gert Jan; Overeem, Sebastiaan

    2016-06-01

    Narcolepsy with cataplexy is a rare disease with an estimated prevalence of 0.02% in European populations. Narcolepsy shares many features of rare disorders, in particular the lack of awareness of the disease with serious consequences for healthcare supply. Similar to other rare diseases, only a few European countries have registered narcolepsy cases in databases of the International Classification of Diseases or in registries of the European health authorities. A promising approach to identify disease-specific adverse health effects and needs in healthcare delivery in the field of rare diseases is to establish a distributed expert network. A first and important step is to create a database that allows collection, storage and dissemination of data on narcolepsy in a comprehensive and systematic way. Here, the first prospective web-based European narcolepsy database hosted by the European Narcolepsy Network is introduced. The database structure, standardization of data acquisition and quality control procedures are described, and an overview provided of the first 1079 patients from 18 European specialized centres. Due to its standardization this continuously increasing data pool is most promising to provide a better insight into many unsolved aspects of narcolepsy and related disorders, including clear phenotype characterization of subtypes of narcolepsy, more precise epidemiological data and knowledge on the natural history of narcolepsy, expectations about treatment effects, identification of post-marketing medication side-effects, and will contribute to improve clinical trial designs and provide facilities to further develop phase III trials. © 2016 European Sleep Research Society.

  1. Incidence of Narcolepsy in Germany.

    PubMed

    Oberle, Doris; Drechsel-Bäuerle, Ursula; Schmidtmann, Irene; Mayer, Geert; Keller-Stanislawski, Brigitte

    2015-10-01

    Following the 2009 pandemic, reports of an association between an AS03 adjuvanted H1N1 pandemic influenza vaccine and narcolepsy were published. Besides determining background incidence rates for narcolepsy in Germany this study aimed at investigating whether there was a change in incidence rates of narcolepsy between the pre-pandemic, pandemic, and the post-pandemic period on the population level. Retrospective epidemiological study on the incidence of narcolepsy with additional capture-recapture analysis. German sleep centers. Eligible were patients with an initial diagnosis of narcolepsy (ICD10 Code G47.4) within the period from January 1, 2007 to December 31, 2011. None; observational study. A total of 342 sleep centers were invited to participate in the study. Adequate and suitable data were provided by 233 sleep centers (68.1%). A total of 1,198 patients with an initial diagnosis of narcolepsy within the observed period were included, of whom 106 (8.8%) were children and adolescents under the age of 18 years and 1,092 (91.2%) were adults. In children and adolescents, the age-standardized adjusted incidence rate significantly increased from 0.14/100,000 person-years in the pre-pandemic period to 0.50/100,000 person-years in the post-pandemic period (incidence density ratio, IDR 3.57; 95% CI 1.94-7.00). In adults, no significant change was detectable. This increase started in spring 2009. For the years 2007-2011, valid estimates for the incidence of narcolepsy in Germany were provided. In individuals under 18, the incidence rates continuously increased from spring 2009. © 2015 Associated Professional Sleep Societies, LLC.

  2. [Narcolepsy with cataplexy: an autoimmune disease?].

    PubMed

    Jacob, Louis; Dauvilliers, Yves

    2014-12-01

    Narcolepsy type 1 (also named narcolepsy-cataplexy or hypocretin deficiency syndrome) is a rare sleep disorder characterized by excessive daytime sleepiness and cataplexy, plus frequently hypnagogic hallucinations, sleep paralysis and nocturnal sleep disturbances. Narcolepsy type 1 is an immune system-associated disease linked with the destruction of 70.000-90.000 hypocretin neurons notably involved in wakefulness. Among narcoleptic patients, 98% are positive for HLA-DQB1*06:02, a HLA class II allele, against 20-25% in general population. Individuals carrying HLA-DQB1*06:02 have an extraordinary risk to develop narcolepsy (odd ratio: 251). Other genes involved in CD4+ T cells and immune system activation as T-cell receptor α are also associated with narcolepsy. The development of the disease is linked with environmental factors such as influenza and streptococcal infections. Narcolepsy type 1 incidence also increased in Europe following the use of Pandemrix, a 2009 H1N1 AS03-adjuvanted vaccine manufactured by GlaxoSmithKline. Interestingly, such increase was not observed with Arepanrix, another vaccine developed by GSK very similar to Pandemrix. © 2014 médecine/sciences – Inserm.

  3. Cases of pediatric narcolepsy after misdiagnoses.

    PubMed

    Kauta, Shilpa R; Marcus, Carole L

    2012-11-01

    Narcolepsy is characterized by recurrent brief attacks of irresistible sleepiness. Signs can begin during childhood. However, diagnoses are frequently delayed by 10-15 years because of unfamiliarity with pediatric narcolepsy and variable presentations of its associated features (cataplexy, hypnagogic/hypnopompic hallucinations, and sleep paralysis). Therefore, patients may remain untreated during their formative years. Three children with narcolepsy who were initially misdiagnosed are described. Each child's signs were initially related to depression, hypothyroidism, jaw dysfunction, or conversion disorder. However, after a multiple sleep latency test, the diagnosis of narcolepsy was established. All three patients were treated appropriately with stimulant medications, selective serotonin reuptake inhibitors, or sodium oxybate, and demonstrated positive responses. Although no definitive cure exists for narcolepsy, early recognition and appropriate symptomatic treatment with medications can allow affected children to improve quality of life and achieve normality, both academically and socially. Copyright © 2012 Elsevier Inc. All rights reserved.

  4. Update on therapy for narcolepsy.

    PubMed

    Thorpy, Michael J

    2015-05-01

    Narcolepsy is a severe, incurable, neurological disorder that is treated by pharmacological management of its symptoms. The main symptoms are excessive daytime sleepiness (EDS) and cataplexy, although addition symptoms that may require treatment include sleep paralysis, hypnagogic hallucinations, and disturbed nocturnal sleep. Sodium oxybate and modafinil/armodafinil are the first-line treatments for EDS, and sodium oxybate for cataplexy. Sodium oxybate treats all the symptoms of narcolepsy, whereas modafinil is effective for EDS only. Alternative medications for EDS include methylphenidate or amphetamines such as dextroamphetamine, lisdexamfetamine, methamphetamine, or combination amphetamine salts. Non-FDA approved medications for cataplexy include norepinephrine reuptake inhibitors such as venlafaxine or atomoxetine. Combination therapy can be more effective for sleepiness such as sodium oxybate and modafinil/armodafinil. Medication for narcolepsy is generally well tolerated and usually required life-long although does not eliminate all symptoms of narcolepsy.

  5. Narcolepsy and the hypocretins.

    PubMed

    Wurtman, Richard J

    2006-10-01

    Narcolepsy is a chronic neurologic disease characterized by excessive daytime sleepiness and one or more of three additional symptoms (cataplexy, or sudden loss of muscle tone; vivid hallucinations; and brief periods of total paralysis) related to the occurrence of rapid eye movement (REM) sleep at inappropriate times. The daytime sleepiness typically presents as a sudden overwhelming urge to sleep, followed by periods of sleep that last for seconds or minutes, or even longer. During daytime sleep episodes, patients may exhibit "automatic behavior," performing conventionalized functions (eg, taking notes), but not remembering having done so once they are awake. About 10% of narcoleptics are members of familial clusters; however, genetic factors alone are apparently insufficient to cause the disease, inasmuch as the most common genetic disorder, a mutation in chromosome 6 controlling the HLA antigen immune complex, although seen in 90% to 100% of patients, also occurs in as many as 50% of people without narcolepsy. A dog model of narcolepsy exhibits a mutation on chromosome 12 that disrupts the processing of the peptide neurotransmitter hypocretin. No such mutation characterizes human narcolepsy; however, cerebrospinal fluid (CSF) hypocretin levels are profoundly depressed in narcoleptic patients, and a specific reduction in hypocretin-containing neurons has been described. One hypothesis concerning the pathophysiology of narcolepsy proposes that the HLA subtype resulting from the mutation on chromosome 6 increases the susceptibility of hypocretin-containing brain neurons to immune attack. Because hypocretin may normally participate in the maintenance of wakefulness, the loss of neurons that release this peptide might allow REM sleep to occur at inappropriate times, ie, while the patient is awake, in contrast to its normal cyclic appearance after a period of slow-wave sleep. The cataplexy, hallucinations, and/or paralysis associated with REM episodes normally are

  6. Narcolepsy in children: a diagnostic and management approach.

    PubMed

    Babiker, Mohamed O E; Prasad, Manish

    2015-06-01

    To provide a diagnostic and management approach for narcolepsy in children. Narcolepsy is a chronic disabling disorder characterized by excessive daytime sleepiness, cataplexy, hypnogogic and/or hypnopompic hallucinations, and sleep paralysis. All four features are present in only half of the cases. Excessive daytime sleepiness is the essential feature of narcolepsy at any age and is usually the first symptom to manifest. A combination of excessive daytime sleepiness and definite cataplexy is considered pathognomonic of narcolepsy syndrome. New treatment options have become available over the past few years. Early diagnosis and management can significantly improve the quality of life of patients with narcolepsy with cataplexy. This review summarizes the pathophysiology, clinical features, and management options for children with narcolepsy. Copyright © 2015 Elsevier Inc. All rights reserved.

  7. HYPOCRETIN/OREXIN AND NARCOLEPSY NEW BASIC AND CLINICAL INSIGHTS

    PubMed Central

    NISHINO, Seiji; OKURO, Masashi; KOTORII, Nozomu; ANEGAWA, Emiko; ISHIMARU, Yuji; MATSUMURA, Mari; KANBAYASHI, Takashi

    2009-01-01

    Narcolepsy is a chronic sleep disorder, characterized by excessive daytime sleepiness (EDS), cataplexy, hypnagogic hallucinations, and sleep paralysis. Both sporadic (95%) and familial (5%) forms of narcolepsy exist in humans. The major pathophysiology of human narcolepsy has been recently discovered based on the discovery of narcolepsy genes in animals; the genes involved in the pathology of the hypocretin/orexin ligand and its receptor. Mutations in hypocretin-related genes are rare in humans, but hypocretin-ligand deficiency is found in a large majority of narcolepsy with cataplexy. Hypocretin ligand deficiency in human narcolepsy is likely due to the postnatal cell death of hypocretin neurons. Although tight association between human leukocyte antigen (HLA) association and human narcolepsy with cataplexy suggests an involvement of autoimmune mechanisms, this has not yet been proven. Hypocretin deficiency is also found in symptomatic cases of narcolepsy and EDS with various neurological conditions, including immune-mediated neurological disorders, such as Guillain-Barre syndrome, MA2-positive paraneoplastic syndrome and neuromyelitis optica (NMO) related disorder. These findings likely have significant clinical relevance and for understanding the mechanisms of hypocretin cell death and choice of treatment option. These series of discoveries in humans lead to the establishment of the new diagnostic test of narcolepsy (i.e. low cerebrospinal fluid [CSF] hypocretin-1 levels for narcolepsy with cataplexy and narcolepsy due to medical condition). Since a large majority of human narcolepsy patients are ligand deficient, hypocretin replacement therapy may be a promising new therapeutic option, and animal experiments using gene therapy and cell transplantations are in progress. PMID:19555382

  8. Misleading hallucinations in unrecognized narcolepsy.

    PubMed

    Szucs, A; Janszky, J; Holló, A; Migléczi, G; Halász, P

    2003-10-01

    To describe psychosis-like hallucinatory states in unrecognized narcolepsy. Two patients with hypnagogic/hypnapompic hallucinations are presented. Both patients had realistic and complex - multi-modal and scenic-daytime sexual hallucinations leading, in the first case, to a legal procedure because of false accusation, and in the second, to serious workplace conflicts. Both patients were convinced of the reality of their hallucinatory experiences but later both were able to recognize their hallucinatory character. Clinical data, a multiple sleep latency test, polysomnography, and HLA typing revealed that both patients suffered from narcolepsy. We suggest that in unrecognized narcolepsy with daytime hypnagogic/hypnapompic hallucinations the diagnostic procedure may mistakenly incline towards delusional psychoses. Daytime realistic hypnagogic/hypnapompic hallucinations may also have forensic consequences and mislead legal evaluation. Useful clinical features in differentiating narcolepsy from psychoses are: the presence of other narcoleptic symptoms, features of hallucinations, and response to adequate medication.

  9. Insufficient non-REM sleep intensity in narcolepsy-cataplexy.

    PubMed

    Khatami, Ramin; Landolt, Hans-Peter; Achermann, Peter; Rétey, Julia V; Werth, Esther; Mathis, Johannes; Bassetti, Claudio L

    2007-08-01

    To compare electroencephalogram (EEG) dynamics during nocturnal sleep in patients with narcolepsy-cataplexy and healthy controls. Fragmented nocturnal sleep is a prominent feature and contributes to excessive daytime sleepiness in narcolepsy-cataplexy. Only 3 studies have addressed changes in homeostatic sleep regulation as a possible mechanism underlying nocturnal sleep fragmentation in narcolepsy-cataplexy. Baseline sleep of 11 drug-naive patients with narcolepsy-cataplexy (19-37 years) and 11 matched controls (18-41 years) was polysomnographically recorded. The EEG was subjected to spectral analysis. None, baseline condition. All patients with narcolepsy-cataplexy but no control subjects showed a sleep-onset rapid eye movement (REM) episode. Non-REM (NREM)-REM sleep cycles were longer in patients with narcolepsy-cataplexy than in controls (P = 0.04). Mean slow-wave activity declined in both groups across the first 3 NREM sleep episodes (P<0.001). The rate of decline, however, appeared to be steeper in patients with narcolepsy-cataplexy (time constant: narcolepsy-cataplexy 51.1 +/- 23.8 minutes [mean +/- SEM], 95% confidence interval [CI]: 33.4-108.8 minutes) than in controls (169.4 +/- 81.5 minutes, 95% CI: 110.9-357.6 minutes) as concluded from nonoverlapping 95% confidence interval of the time constants. The steeper decline of SWA in narcolepsy-cataplexy compared to controls was related to an impaired build-up of slow-wave activity in the second cycle. Sleep in the second cycle was interrupted in patients with narcolepsy-cataplexy, when compared with controls, by an increased number (P = 0.01) and longer duration (P = 0.01) of short wake episodes. Insufficient NREM sleep intensity is associated with nonconsolidated nocturnal sleep in narcolepsy-cataplexy. The inability to consolidate sleep manifests itself when NREM sleep intensity has decayed below a certain level and is reflected in an altered time course of slow-wave activity across NREM sleep episodes.

  10. Increase of histaminergic tuberomammillary neurons in narcolepsy.

    PubMed

    Valko, Philipp O; Gavrilov, Yury V; Yamamoto, Mihoko; Reddy, Hasini; Haybaeck, Johannes; Mignot, Emmanuel; Baumann, Christian R; Scammell, Thomas E

    2013-12-01

    Narcolepsy is caused by loss of the hypothalamic neurons producing the orexin/hypocretin neuropeptides. One key target of the orexin system is the histaminergic neurons of the tuberomammillary nucleus (TMN), an essential wake-promoting system. As cerebrospinal fluid histamine levels may be low in patients with narcolepsy, we examined histaminergic neurons in patients with narcolepsy and in 2 mouse models of narcolepsy. We counted the number of hypothalamic neurons producing orexin, melanin-concentrating hormone, and histamine in 7 narcolepsy patients and 12 control subjects using stereological techniques. We identified histaminergic neurons using immunostaining for histidine decarboxylase. We also examined these systems in 6 wild-type mice, 6 orexin/ataxin-3 transgenic mice, and 5 orexin ligand knockout mice. Compared to controls, narcolepsy patients had 94% more histaminergic TMN neurons (233,572 ± 49,476 vs 120,455 ± 10,665, p < 0.001). This increase was higher in 5 narcolepsy patients with >90% orexin neuron loss than in 2 patients with ≤75% orexin neuron loss (252,279 ± 46,264 vs 186,804 ± 1,256, p = 0.03). Similarly, the number of histaminergic TMN neurons was increased 53% in orexin ligand knockout mice compared to wild-type mice, whereas orexin/ataxin-3 transgenic mice showed an intermediate 28% increase. This surprising increase in histaminergic neurons in narcolepsy may be a compensatory response to loss of excitatory drive from the orexin neurons and may contribute to some of the symptoms of narcolepsy such as preserved consciousness during cataplexy and fragmented nighttime sleep. In addition, this finding may have therapeutic implications, as medications that enhance histamine signaling are now under development. © 2013 American Neurological Association.

  11. Reward-Seeking Behavior in Human Narcolepsy

    PubMed Central

    Dimitrova, Alexandra; Fronczek, Rolf; Van der Ploeg, Janneke; Scammell, Thomas; Gautam, Shiva; Pascual-Leone, Alvaro; Lammers, Gert Jan

    2011-01-01

    Study Objectives: The hypocretin system enhances signaling in the mesolimbic pathways regulating reward processing and addiction. Because individuals with narcolepsy with cataplexy have low hypocretin levels, we hypothesized that they may be less prone to risk- and reward-seeking behaviors, including substance abuse. Design: Endpoints were performance on an array of psychometric tests (including the Eysenck Impulsiveness Scale, the Zuckerman Sensation Seeking Scale, the Gormally Binge Eating Scale, and the Beck Depression and Anxiety Inventory) and on the Balloon Analogue Risk Task (BART). Setting: Tertiary narcolepsy referral centers in Leiden (The Netherlands) and Boston (USA). Patients: Subjects with narcolepsy with cataplexy (n = 30), narcolepsy without cataplexy (n = 15), and controls (n = 32) matched for age, sex, and smoking behavior. Interventions: None. Measurements and Results: There was no difference in risk-taking behavior between narcolepsy with or without cataplexy and the control group, as measured using the BART and the array of questionnaires. However, subjects in the narcolepsy with cataplexy group had significantly higher scores on the Eysenck Impulsiveness Scale (p < 0.05), with 10.0% categorized as impulsive, compared to 6.7% of the narcolepsy without cataplexy group and none of the controls. Narcoleptics with cataplexy also scored significantly higher than controls on the Binge Eating Scale (p < 0.05), with moderate or severe binge eating in 23%. On the depression and anxiety scales, all narcolepsy patients, especially those with cataplexy, scored significantly higher than controls. Conclusions: We found that narcoleptics with or without cataplexy generally have normal risk-taking behavior, but narcoleptics with cataplexy were more impulsive and more prone to binge eating than patients without cataplexy and controls. Our findings shed new light on the relation between sleepiness and impulsiveness. Furthermore, rates of depression and anxiety

  12. Altered Brain Microstate Dynamics in Adolescents with Narcolepsy

    PubMed Central

    Drissi, Natasha M.; Szakács, Attila; Witt, Suzanne T.; Wretman, Anna; Ulander, Martin; Ståhlbrandt, Henriettae; Darin, Niklas; Hallböök, Tove; Landtblom, Anne-Marie; Engström, Maria

    2016-01-01

    Narcolepsy is a chronic sleep disorder caused by a loss of hypocretin-1 producing neurons in the hypothalamus. Previous neuroimaging studies have investigated brain function in narcolepsy during rest using positron emission tomography (PET) and single photon emission computed tomography (SPECT). In addition to hypothalamic and thalamic dysfunction they showed aberrant prefrontal perfusion and glucose metabolism in narcolepsy. Given these findings in brain structure and metabolism in narcolepsy, we anticipated that changes in functional magnetic resonance imaging (fMRI) resting state network (RSN) dynamics might also be apparent in patients with narcolepsy. The objective of this study was to investigate and describe brain microstate activity in adolescents with narcolepsy and correlate these to RSNs using simultaneous fMRI and electroencephalography (EEG). Sixteen adolescents (ages 13–20) with a confirmed diagnosis of narcolepsy were recruited and compared to age-matched healthy controls. Simultaneous EEG and fMRI data were collected during 10 min of wakeful rest. EEG data were analyzed for microstates, which are discrete epochs of stable global brain states obtained from topographical EEG analysis. Functional MRI data were analyzed for RSNs. Data showed that narcolepsy patients were less likely than controls to spend time in a microstate which we found to be related to the default mode network and may suggest a disruption of this network that is disease specific. We concluded that adolescents with narcolepsy have altered resting state brain dynamics. PMID:27536225

  13. Challenges in diagnosing narcolepsy without cataplexy: a consensus statement.

    PubMed

    Baumann, Christian R; Mignot, Emmanuel; Lammers, Gert Jan; Overeem, Sebastiaan; Arnulf, Isabelle; Rye, David; Dauvilliers, Yves; Honda, Makoto; Owens, Judith A; Plazzi, Giuseppe; Scammell, Thomas E

    2014-06-01

    Diagnosing narcolepsy without cataplexy is often a challenge as the symptoms are nonspecific, current diagnostic tests are limited, and there are no useful biomarkers. In this report, we review the clinical and physiological aspects of narcolepsy without cataplexy, the limitations of available diagnostic procedures, and the differential diagnoses, and we propose an approach for more accurate diagnosis of narcolepsy without cataplexy. A group of clinician-scientists experienced in narcolepsy reviewed the literature and convened to discuss current diagnostic tools, and to map out directions for research that should lead to a better understanding and more accurate diagnosis of narcolepsy without cataplexy. To aid in the identification of narcolepsy without cataplexy, we review key indicators of narcolepsy and present a diagnostic algorithm. A detailed clinical history is mainly helpful to rule out other possible causes of chronic sleepiness. The multiple sleep latency test remains the most important measure, and prior sleep deprivation, shift work, or circadian disorders should be excluded by actigraphy or sleep logs. A short REM sleep latency (≤ 15 minutes) on polysomnography can aid in the diagnosis of narcolepsy without cataplexy, although sensitivity is low. Finally, measurement of hypocretin levels can helpful, as levels are low to intermediate in 10% to 30% of narcolepsy without cataplexy patients.

  14. Degenerative pontine lesions in patients with familial narcolepsy.

    PubMed

    Stepień, Adam; Staszewski, Jacek; Domzał, Teofan M; Tomczykiewicz, Kazimierz; Skrobowska, Ewa; Durka-Kesy, Marta

    2010-01-01

    Narcolepsy is characterized by chronic excessive daytime sleepiness with episodic sleep attacks. There are several associated symptoms of narcolepsy: cataplexy (bilateral muscle weakness without loss of consciousness provoked by an emotional trigger, e.g. laughter), sleep paralysis and hypnagogic-hypnopompic hallucinations. Most cases are sporadic; familial narcolepsy contributes to only 1-5% of all cases. While most cases of narcolepsy are idiopathic and are not associated with clinical or radiographic evidence of brain pathology, symptomatic or secondary narcolepsy may occur occasionally in association with lesions caused by tumours, demyelination or strokes of the diencephalon, midbrain, and pons. There are some examples of non-specific brainstem lesions found in magnetic resonance imaging (MRI) in patients with idiopathic narcolepsy. The authors present eleven patients from a five-generation family with many members who suffer from episodic excessive daytime sleepiness. Narcolepsy was diagnosed in 9 patients. Sleepiness was frequently associated with cataplexy, hypnagogic-hypnopompic hallucinations and sleep paralysis. Improvement in their clinical state was observed during the treatment with modafinil. All probands had MRI of the brain, routine blood tests, EEG, polysomnography, examination of the level of hypocretin in cerebrospinal fluid and evaluation by means of Epworth and Stanford Sleepiness Scales. In 9 patients with narcolepsy, decreased thickness of the substantia nigra was found and in six of them degenerative lesions in the pontine substantia nigra were also noticed. The significance of these changes remains unclear. No data have been published until now concerning the presence of any brain lesions in patients with familial narcolepsy.

  15. Patient-Reported Measures of Narcolepsy: The Need for Better Assessment.

    PubMed

    Kallweit, Ulf; Schmidt, Markus; Bassetti, Claudio L

    2017-05-15

    Narcolepsy, a chronic disorder of the central nervous system, is clinically characterized by a symptom pentad that includes excessive daytime sleepiness, cataplexy, sleep paralysis, hypnopompic/hypnagogic hallucinations, and disrupted nighttime sleep. Ideally, screening and diagnosis instruments that assist physicians in evaluating a patient for type 1 or type 2 narcolepsy would be brief, easy for patients to understand and physicians to score, and would identify or rule out the need for electrophysiological testing. A search of the literature was conducted to review patient-reported measures used for the assessment of narcolepsy, mainly in clinical trials, with the goal of summarizing existing scales and identifying areas that may require additional screening questions and clinical practice scales. Of the seven scales reviewed, the Epworth Sleepiness Scale continues to be an important outcome measure to screen adults for excessive daytime sleepiness, which may be associated with narcolepsy. Several narcolepsy-specific scales have demonstrated utility, such as the Ullanlinna Narcolepsy Scale, Swiss Narcolepsy Scale, and Narcolepsy Symptom Assessment Questionnaire, but further validation is required. Although the narcolepsy-specific scales currently in use may identify type 1 narcolepsy, there are no validated questionnaires to identify type 2 narcolepsy. Thus, there remains a need for short, easily understood, and well-validated instruments that can be readily used in clinical practice to distinguish narcolepsy subtypes, as well as other hypersomnias, and for assessing symptoms of these conditions during treatment. © 2017 American Academy of Sleep Medicine

  16. Challenges in the development of therapeutics for narcolepsy

    PubMed Central

    Black, Sarah Wurts; Yamanaka, Akihiro; Kilduff, Thomas S.

    2016-01-01

    Narcolepsy is a neurological disorder that afflicts 1 in 2000 individuals and is characterized by excessive daytime sleepiness and cataplexy—a sudden loss of muscle tone triggered by positive emotions. Features of narcolepsy include dysregulation of arousal state boundaries as well as autonomic and metabolic disturbances. Disruption of neurotransmission through the hypocretin/orexin (Hcrt) system, usually by degeneration of the HCRT-producing neurons in the posterior hypothalamus, results in narcolepsy. The cause of Hcrt neurodegeneration is unknown but thought to be related to autoimmune processes. Current treatments for narcolepsy are symptomatic, including wake-promoting therapeutics that increase presynaptic dopamine release and anticataplectic agents that activate monoaminergic neurotransmission. Sodium oxybate is the only medication approved by the US Food and Drug Administration that alleviates both sleep/wake disturbances and cataplexy. Development of therapeutics for narcolepsy has been challenged by historical misunderstanding of the disease, its many disparate symptoms and, until recently, its unknown etiology. Animal models have been essential to elucidating the neuropathology underlying narcolepsy. These models have also aided understanding the neurobiology of the Hcrt system, mechanisms of cataplexy, and the pharmacology of narcolepsy medications. Transgenic rodent models will be critical in the development of novel therapeutics for the treatment of narcolepsy, particularly efforts directed to overcome challenges in the development of hypocretin replacement therapy. PMID:26721620

  17. Challenges in the development of therapeutics for narcolepsy.

    PubMed

    Black, Sarah Wurts; Yamanaka, Akihiro; Kilduff, Thomas S

    2017-05-01

    Narcolepsy is a neurological disorder that afflicts 1 in 2000 individuals and is characterized by excessive daytime sleepiness and cataplexy-a sudden loss of muscle tone triggered by positive emotions. Features of narcolepsy include dysregulation of arousal state boundaries as well as autonomic and metabolic disturbances. Disruption of neurotransmission through the hypocretin/orexin (Hcrt) system, usually by degeneration of the HCRT-producing neurons in the posterior hypothalamus, results in narcolepsy. The cause of Hcrt neurodegeneration is unknown but thought to be related to autoimmune processes. Current treatments for narcolepsy are symptomatic, including wake-promoting therapeutics that increase presynaptic dopamine release and anticataplectic agents that activate monoaminergic neurotransmission. Sodium oxybate is the only medication approved by the US Food and Drug Administration that alleviates both sleep/wake disturbances and cataplexy. Development of therapeutics for narcolepsy has been challenged by historical misunderstanding of the disease, its many disparate symptoms and, until recently, its unknown etiology. Animal models have been essential to elucidating the neuropathology underlying narcolepsy. These models have also aided understanding the neurobiology of the Hcrt system, mechanisms of cataplexy, and the pharmacology of narcolepsy medications. Transgenic rodent models will be critical in the development of novel therapeutics for the treatment of narcolepsy, particularly efforts directed to overcome challenges in the development of hypocretin replacement therapy. Copyright © 2015 Elsevier Ltd. All rights reserved.

  18. Pituitary Macrotumor Causing Narcolepsy-Cataplexy in a Dachshund.

    PubMed

    Schmid, S; Hodshon, A; Olin, S; Pfeiffer, I; Hecht, S

    2017-03-01

    Familial narcolepsy secondary to breed-specific mutations in the hypocretin receptor 2 gene and sporadic narcolepsy associated with hypocretin ligand deficiencies occur in dogs. In this report, a pituitary mass is described as a unique cause of narcolepsy-cataplexy in a dog. A 6-year-old male neutered Dachshund had presented for acute onset of feeding-induced cataplexy and was found to have a pituitary macrotumor on magnetic resonance imaging (MRI). Cerebral spinal fluid hypocretin-1 levels were normal, indicating that tumor effect on the ventral lateral nucleus of the hypothalamus was not the cause of the dog's narcolepsy-cataplexy. The dog was also negative for the hypocretin receptor 2 gene mutation associated with narcolepsy in Dachshunds, ruling out familial narcolepsy. The Dachshund underwent stereotactic radiotherapy (SRT), which resulted in reduction in the mass and coincident resolution of the cataplectic attacks. Nine months after SRT, the dog developed clinical hyperadrenocorticism, which was successfully managed with trilostane. These findings suggest that disruptions in downstream signaling of hypocretin secondary to an intracranial mass effect might result in narcolepsy-cataplexy in dogs and that brain MRI should be strongly considered in sporadic cases of narcolepsy-cataplexy. Copyright © 2017 The Authors. Journal of Veterinary Internal Medicine published by Wiley Periodicals, Inc. on behalf of the American College of Veterinary Internal Medicine.

  19. An overview of hypocretin based therapy in narcolepsy.

    PubMed

    Takenoshita, Shinichi; Sakai, Noriaki; Chiba, Yuhei; Matsumura, Mari; Yamaguchi, Mai; Nishino, Seiji

    2018-04-01

    Narcolepsy with cataplexy is most commonly caused by a loss of hypocretin/orexin peptide-producing neurons in the hypothalamus (i.e., Narcolepsy Type 1). Since hypocretin deficiency is assumed to be the main cause of narcoleptic symptoms, hypocretin replacement will be the most essential treatment for narcolepsy. Unfortunately, this option is still not available clinically. There are many potential approaches to replace hypocretin in the brain for narcolepsy such as intranasal administration of hypocretin peptides, developing small molecule hypocretin receptor agonists, hypocretin neuronal transplantation, transforming hypocretin stem cells into hypothalamic neurons, and hypocretin gene therapy. Together with these options, immunotherapy treatments to prevent hypocretin neuronal death should also be developed. Areas covered: In this review, we overview the pathophysiology of narcolepsy and the current and emerging treatments of narcolepsy especially focusing on hypocretin receptor based treatments. Expert opinion: Among hypocretin replacement strategies, developing non-peptide hypocretin receptor agonists is currently the most encouraging since systemic administration of a newly synthesized, selective hypocretin receptor 2 agonist (YNT-185) has been shown to ameliorate symptoms of narcolepsy in murine models. If this option is effective in humans, hypocretin cell transplants or gene therapy technology may become realistic in the future.

  20. Neuroimaging of Narcolepsy and Kleine-Levin Syndrome.

    PubMed

    Hong, Seung Bong

    2017-09-01

    Narcolepsy is a chronic neurologic disorder with the abnormal regulation of the sleep-wake cycle, resulting in excessive daytime sleepiness, disturbed nocturnal sleep, and manifestations related to rapid eye movement sleep, such as cataplexy, sleep paralysis, and hypnagogic hallucination. Over the past decade, numerous neuroimaging studies have been performed to characterize the pathophysiology and various clinical features of narcolepsy. This article reviews structural and functional brain imaging findings in narcolepsy and Kleine-Levin syndrome. Based on the current state of research, brain imaging is a useful tool to investigate and understand the neuroanatomic correlates and brain abnormalities of narcolepsy and other hypersomnia. Copyright © 2017 Elsevier Inc. All rights reserved.

  1. Patient-Reported Measures of Narcolepsy: The Need for Better Assessment

    PubMed Central

    Kallweit, Ulf; Schmidt, Markus; Bassetti, Claudio L.

    2017-01-01

    Study Objectives: Narcolepsy, a chronic disorder of the central nervous system, is clinically characterized by a symptom pentad that includes excessive daytime sleepiness, cataplexy, sleep paralysis, hypnopompic/hypnagogic hallucinations, and disrupted nighttime sleep. Ideally, screening and diagnosis instruments that assist physicians in evaluating a patient for type 1 or type 2 narcolepsy would be brief, easy for patients to understand and physicians to score, and would identify or rule out the need for electrophysiological testing. Methods: A search of the literature was conducted to review patient-reported measures used for the assessment of narcolepsy, mainly in clinical trials, with the goal of summarizing existing scales and identifying areas that may require additional screening questions and clinical practice scales. Results: Of the seven scales reviewed, the Epworth Sleepiness Scale continues to be an important outcome measure to screen adults for excessive daytime sleepiness, which may be associated with narcolepsy. Several narcolepsy-specific scales have demonstrated utility, such as the Ullanlinna Narcolepsy Scale, Swiss Narcolepsy Scale, and Narcolepsy Symptom Assessment Questionnaire, but further validation is required. Conclusions: Although the narcolepsy-specific scales currently in use may identify type 1 narcolepsy, there are no validated questionnaires to identify type 2 narcolepsy. Thus, there remains a need for short, easily understood, and well-validated instruments that can be readily used in clinical practice to distinguish narcolepsy subtypes, as well as other hypersomnias, and for assessing symptoms of these conditions during treatment. Citation: Kallweit U, Schmidt M, Bassetti CL. Patient-reported measures of narcolepsy: the need for better assessment. J Clin Sleep Med. 2017;13(5):737–744. PMID:28162143

  2. Challenges in Diagnosing Narcolepsy without Cataplexy: A Consensus Statement

    PubMed Central

    Baumann, Christian R.; Mignot, Emmanuel; Lammers, Gert Jan; Overeem, Sebastiaan; Arnulf, Isabelle; Rye, David; Dauvilliers, Yves; Honda, Makoto; Owens, Judith A.; Plazzi, Giuseppe; Scammell, Thomas E.

    2014-01-01

    Background: Diagnosing narcolepsy without cataplexy is often a challenge as the symptoms are nonspecific, current diagnostic tests are limited, and there are no useful biomarkers. In this report, we review the clinical and physiological aspects of narcolepsy without cataplexy, the limitations of available diagnostic procedures, and the differential diagnoses, and we propose an approach for more accurate diagnosis of narcolepsy without cataplexy. Methods: A group of clinician-scientists experienced in narcolepsy reviewed the literature and convened to discuss current diagnostic tools, and to map out directions for research that should lead to a better understanding and more accurate diagnosis of narcolepsy without cataplexy. Recommendations: To aid in the identification of narcolepsy without cataplexy, we review key indicators of narcolepsy and present a diagnostic algorithm. A detailed clinical history is mainly helpful to rule out other possible causes of chronic sleepiness. The multiple sleep latency test remains the most important measure, and prior sleep deprivation, shift work, or circadian disorders should be excluded by actigraphy or sleep logs. A short REM sleep latency (≤ 15 minutes) on polysomnography can aid in the diagnosis of narcolepsy without cataplexy, although sensitivity is low. Finally, measurement of hypocretin levels can helpful, as levels are low to intermediate in 10% to 30% of narcolepsy without cataplexy patients. Citation: Baumann CR, Mignot E, Lammers GJ, Overeem S, Arnulf I, Rye D, Dauvilliers Y, Honda M, Owens JA, Plazzi G, Scammell TE. Challenges in diagnosing narcolepsy without cataplexy: a consensus statement. SLEEP 2014;37(6):1035-1042. PMID:24882898

  3. Narcolepsy

    PubMed Central

    Mitler, Merrill M.; Hajdukovic, Roza; Erman, Milton; Koziol, James A.

    2008-01-01

    Summary Narcolepsy is a neurological condition with a prevalence of up to 1 per 1,000 that is characterized by irresistible bouts of sleep. Associated features include the pathological manifestations of rapid-eye-movement (REM) sleep: cataplexy, sleep paralysis, hypnagogic hallucinations, and abnormal sleep-onset REM periods and disturbed nocturnal sleep. The condition is strongly associated with the HLA-DR2 and DQw1 phenotype. The phenomenology of narcolepsy is discussed, and diagnostic procedures are reviewed. Treatment modalities involving central nervous system stimulants for somnolence and tricyclic drugs for REM-sleep abnormalities are discussed. Sleep laboratory studies on the treatment efficacy of methylphenidate, pemoline, dextroamphetamine, protriptyline, and viloxazine are presented. Data suggest that: (1) methylphenidate and dextro-amphetamine objectively improve somnolence; (2) pemoline, at doses up to 112.5 mg, is less effective in controlling somnolence but may improve certain aspects of performance; and (3) protriptyline and viloxazine are effective anticataplectic agents that produce little improvement in somnolence. PMID:1968069

  4. Pharmacological management of narcolepsy with and without cataplexy.

    PubMed

    Kallweit, Ulf; Bassetti, Claudio L

    2017-06-01

    Narcolepsy is an orphan neurological disease and presents with sleep-wake, motoric, neuropsychiatric and metabolic symptoms. Narcolepsy with cataplexy is most commonly caused by an immune-mediated process including genetic and environmental factors, resulting in the selective loss of hypocretin-producing neurons. Narcolepsy has a major impact on workableness and quality of life. Areas covered: This review provides an overview of the temporal available treatment options for narcolepsy (type 1 and 2) in adults, including authorization status by regulatory agencies. First- and second-line options are discussed as well as combination therapies. In addition, treatment options for frequent coexisting co-morbidities and different phenotypes of narcolepsy are presented. Finally, this review considers potential future management strategies. Non-pharmacological approaches are important in the management of narcolepsy but will not be covered in this review. Expert opinion: Concise evaluation of symptoms and type of narcolepsy, coexisting co-morbidities and patients´ distinct needs is mandatory in order to identify a suitable, individual pharmacological treatment. First-line options include Modafinil/Armodafinil (for excessive daytime sleepiness, EDS), Sodium Oxybate (for EDS and/with cataplexy), Pitolisant (for EDS and cataplexy) and Venlafaxine (for cataplexy (off-label) and co-morbid depression). New symptomatic and causal treatment most probably will be completed by hypocretin-replacement and immune-modifying strategies.

  5. IgG abnormality in narcolepsy and idiopathic hypersomnia.

    PubMed

    Tanaka, Susumu; Honda, Makoto

    2010-03-05

    A close association between narcolepsy and the Human Leukocyte Antigen (HLA)-DQB1*0602 allele suggests the involvement of the immune system, or possibly an autoimmune process. We investigated serum IgG levels in narcolepsy. We measured the serum total IgG levels in 159 Japanese narcolepsy-cataplexy patients positive for the HLA-DQB1*0602 allele, 28 idiopathic hypersomnia patients with long sleep time, and 123 healthy controls (the HLA-DQB1*0602 allele present in 45 subjects). The serum levels of each IgG subclass were subsequently measured. The distribution of serum IgG was significantly different among healthy controls negative for the HLA-DQB1*0602 allele (11.66+/-3.55 mg/ml), healthy controls positive for the HLA-DQB1*0602 allele (11.45+/-3.43), narcolepsy patients (9.67+/-3.38), and idiopathic hypersomnia patients (13.81+/-3.80). None of the following clinical variables, age, disease duration, Epworth Sleepiness Scale, smoking habit and BMI at the time of blood sampling, were associated with IgG levels in narcolepsy or idiopathic hypersomnia. Furthermore we found the decrease in IgG1 and IgG2 levels, stable expression of IgG3, and the increase in the proportion of IgG4 in narcolepsy patients with abnormally low IgG levels. The increase in the proportion of IgG4 levels was also found in narcolepsy patients with normal serum total IgG levels. Idiopathic hypersomnia patients showed a different pattern of IgG subclass distribution with high IgG3 and IgG4 level, low IgG2 level, and IgG1/IgG2 imbalance. Our study is the first to determine IgG abnormalities in narcolepsy and idiopathic hypersomnia by measuring the serum IgG levels in a large number of hypersomnia patients. The observed IgG abnormalities indicate humoral immune alterations in narcolepsy and idiopathic hypersomnia. Different IgG profiles suggest immunological differences between narcolepsy and idiopathic hypersomnia.

  6. Psychotic symptoms in narcolepsy: phenomenology and a comparison with schizophrenia.

    PubMed

    Fortuyn, Hal A Droogleever; Lappenschaar, G A; Nienhuis, Fokko J; Furer, Joop W; Hodiamont, Paul P; Rijnders, Cees A; Lammers, Gert Jan; Renier, Willy O; Buitelaar, Jan K; Overeem, Sebastiaan

    2009-01-01

    Patients with narcolepsy often experience pervasive hypnagogic hallucinations, sometimes even leading to confusion with schizophrenia. We aimed to provide a detailed qualitative description of hypnagogic hallucinations and other "psychotic" symptoms in patients with narcolepsy and contrast these with schizophrenia patients and healthy controls. We also compared the prevalence of formal psychotic disorders between narcolepsy patients and controls. We used SCAN 2.1 interviews to compare psychotic symptoms between 60 patients with narcolepsy, 102 with schizophrenia and 120 matched population controls. In addition, qualitative data was collected to enable a detailed description of hypnagogic hallucinations in narcolepsy. There were clear differences in the pattern of hallucinatory experiences in narcolepsy vs. schizophrenia patients. Narcoleptics reported multisensory "holistic" hallucinations rather than the predominantly verbal-auditory sensory mode of schizophrenia patients. Psychotic symptoms such as delusions were not more frequent in narcolepsy compared to population controls. In addition, the prevalence of formal psychotic disorders was not increased in patients with narcolepsy. Almost half of narcoleptics reported moderate interference with functioning due to hypnagogic hallucinations, mostly due to related anxiety. Hypnagogic hallucinations in narcolepsy can be differentiated on a phenomenological basis from hallucinations in schizophrenia which is useful in differential diagnostic dilemmas.

  7. Circadian Rest-Activity Rhythm in Pediatric Type 1 Narcolepsy

    PubMed Central

    Filardi, Marco; Pizza, Fabio; Bruni, Oliviero; Natale, Vincenzo; Plazzi, Giuseppe

    2016-01-01

    Study Objectives: Pediatric type 1 narcolepsy is often challenging to diagnose and remains largely undiagnosed. Excessive daytime sleepiness, disrupted nocturnal sleep, and a peculiar phenotype of cataplexy are the prominent features. The knowledge available about the regulation of circadian rhythms in affected children is scarce. This study compared circadian rest-activity rhythm and actigraphic estimated sleep measures of children with type 1 narcolepsy versus healthy controls. Methods: Twenty-two drug-naïve type 1 narcolepsy children and 21 age- and sex- matched controls were monitored for seven days during the school week by actigraphy. Circadian activity rhythms were analyzed through functional linear modeling; nocturnal and diurnal sleep measures were estimated from activity using a validated algorithm. Results: Children with type 1 narcolepsy presented an altered rest-activity rhythm characterized by enhanced motor activity throughout the night and blunted activity in the first afternoon. No difference was found between children with type 1 narcolepsy and controls in the timing of the circadian phase. Actigraphic sleep measures showed good discriminant capabilities in assessing type 1 narcolepsy nycthemeral disruption. Conclusions: Actigraphy reliably renders the nycthemeral disruption typical of narcolepsy type 1 in drug-naïve children with recent disease onset, indicating the sensibility of actigraphic assessment in the diagnostic work-up of childhood narcolepsy type 1. Citation: Filardi M, Pizza F, Bruni O, Natale V, Plazzi G. Circadian rest-activity rhythm in pediatric type 1 narcolepsy. SLEEP 2016;39(6):1241–1247. PMID:27091539

  8. High Rates of Psychiatric Comorbidity in Narcolepsy: Findings From the Burden of Narcolepsy Disease (BOND) Study of 9,312 Patients in the United States.

    PubMed

    Ruoff, Chad M; Reaven, Nancy L; Funk, Susan E; McGaughey, Karen J; Ohayon, Maurice M; Guilleminault, Christian; Black, Jed

    2017-02-01

    To evaluate psychiatric comorbidity patterns in patients with a narcolepsy diagnosis in the United States. Truven Health Analytics MarketScan Research Databases were accessed to identify individuals ≥ 18 years of age with ≥ 1 ICD-9 diagnosis code(s) for narcolepsy continuously insured between 2006 and 2010 and non-narcolepsy controls matched 5:1 (age, gender, region, payer). Extensive subanalyses were conducted to confirm the validity of narcolepsy definitions. Narcolepsy subjects and controls were compared for frequency of psychiatric comorbid conditions (based on ICD-9 codes/Clinical Classification Software [CCS] level 2 categories) and psychiatric medication use. The final population included 9,312 narcolepsy subjects and 46,559 controls (each group, mean age = 46.1 years; 59% female). All categories of mental illness were significantly more prevalent in patients with narcolepsy versus controls, with the highest excess prevalence noted for CCS 5.8 Mood disorders (37.9% vs 13.8%; odds ratio [OR] = 4.0; 95% CI, 3.8-4.2), CCS 5.8.2 Depressive disorders (35.8% vs 13.0%; OR = 3.9; 95% CI, 3.7-4.1), and CCS 5.2 Anxiety disorders (25.1% vs 11.9%; OR = 2.5; 95% CI, 2.4-2.7). Excess prevalence of anxiety and mood disorders (narcolepsy vs controls) was higher in younger age groups versus older age groups. Psychiatric medication usage was higher in the narcolepsy group versus controls in the following categories: selective serotonin reuptake inhibitors (36% vs 17%), anxiolytic benzodiazepines (34% vs 19%), hypnotics (29% vs 13%), serotonin-norepinephrine reuptake inhibitors (21% vs 6%), and tricyclic antidepressants (13% vs 4%) (all P values < .0001). Narcolepsy is associated with significant comorbid psychiatric illness burden and higher psychiatric medication usage compared with the non-narcolepsy population. © Copyright 2016 Physicians Postgraduate Press, Inc.

  9. Narcolepsy in Adolescence—A Missed Diagnosis: A Case Report

    PubMed Central

    Gupta, Anoop K.; Sahoo, Swapnajeet

    2017-01-01

    ABSTRACT: Narcolepsy is an uncommon sleep cycle disorder with a usual onset in adolescence, but it is often misdiagnosed and underdiagnosed. Rarely is the tetrad of excessive daytime sleepiness, cataplexy, hypnagogic hallucinations, and sleep paralysis seen in patients. The clinical characteristics of narcolepsy are often confused with many psychiatric and neurologic disorders. Lack of clinical awareness about narcolepsy leads to frequent prescriptions of antiepileptics and psychotropics, which can adversely affect the quality of life of children and adolescents. We report a case of an adolescent male who presented with all four cardinal symptoms of narcolepsy and had been misdiagnosed with epilepsy, psychosis, and depression. We discuss various issues regarding narcolepsy in children and adolescents. PMID:29616151

  10. Status Cataplecticus as Initial Presentation of Late Onset Narcolepsy

    PubMed Central

    Panda, Samhita

    2014-01-01

    Narcolepsy, one of the important causes of hypersomnia, is an under diagnosed sleep disorder. It has a bimodal age of onset around 15 and 35 years. It is characterized by the tetrad of excessive daytime sleepiness, cataplexy, hypnagogic/ hypnopompic hallucinations, and sleep paralysis. Cataplexy is by far the most predictive feature of narcolepsy. Status cataplecticus is the occurrence of cataplexy repeatedly for hours or days, a rare presentation of narcolepsy. This report describes an elderly gentleman with late onset narcolepsy in the sixth decade of life presenting with initial and chief symptom of status cataplecticus. Citation: Panda S. Status cataplecticus as initial presentation of late onset narcolepsy. J Clin Sleep Med 2014;10(2):207-209. PMID:24533005

  11. Narcolepsy in Adolescence-A Missed Diagnosis: A Case Report.

    PubMed

    Gupta, Anoop K; Sahoo, Swapnajeet; Grover, Sandeep

    2017-01-01

    Narcolepsy is an uncommon sleep cycle disorder with a usual onset in adolescence, but it is often misdiagnosed and underdiagnosed. Rarely is the tetrad of excessive daytime sleepiness, cataplexy, hypnagogic hallucinations, and sleep paralysis seen in patients. The clinical characteristics of narcolepsy are often confused with many psychiatric and neurologic disorders. Lack of clinical awareness about narcolepsy leads to frequent prescriptions of antiepileptics and psychotropics, which can adversely affect the quality of life of children and adolescents. We report a case of an adolescent male who presented with all four cardinal symptoms of narcolepsy and had been misdiagnosed with epilepsy, psychosis, and depression. We discuss various issues regarding narcolepsy in children and adolescents.

  12. Circadian Rest-Activity Rhythm in Pediatric Type 1 Narcolepsy.

    PubMed

    Filardi, Marco; Pizza, Fabio; Bruni, Oliviero; Natale, Vincenzo; Plazzi, Giuseppe

    2016-06-01

    Pediatric type 1 narcolepsy is often challenging to diagnose and remains largely undiagnosed. Excessive daytime sleepiness, disrupted nocturnal sleep, and a peculiar phenotype of cataplexy are the prominent features. The knowledge available about the regulation of circadian rhythms in affected children is scarce. This study compared circadian rest-activity rhythm and actigraphic estimated sleep measures of children with type 1 narcolepsy versus healthy controls. Twenty-two drug-naïve type 1 narcolepsy children and 21 age- and sex- matched controls were monitored for seven days during the school week by actigraphy. Circadian activity rhythms were analyzed through functional linear modeling; nocturnal and diurnal sleep measures were estimated from activity using a validated algorithm. Children with type 1 narcolepsy presented an altered rest-activity rhythm characterized by enhanced motor activity throughout the night and blunted activity in the first afternoon. No difference was found between children with type 1 narcolepsy and controls in the timing of the circadian phase. Actigraphic sleep measures showed good discriminant capabilities in assessing type 1 narcolepsy nycthemeral disruption. Actigraphy reliably renders the nycthemeral disruption typical of narcolepsy type 1 in drug-naïve children with recent disease onset, indicating the sensibility of actigraphic assessment in the diagnostic work-up of childhood narcolepsy type 1. © 2016 Associated Professional Sleep Societies, LLC.

  13. Narcolepsy and disorders of excessive somnolence.

    PubMed

    Dyken, Mark E; Yamada, Thoru

    2005-06-01

    Recent studies provide valid criteria that help differentiate idiopathic narcolepsy from other disorders of excessive daytime somnolence [3]. Research to date suggests that idiopathic narcolepsy might properly be considered a disorder of excessive sleepiness with dysfunctional REM-sleep mechanisms, clinically evidenced as cataplexy and electrophysiologically recognized as SOREMPs. Given these criteria, a diagnosis can generally be made using a combination of history, PSG, and MSLT. Traditionally, the medical treatment of idiopathic narcolepsy has centered on a two-drug regimen (stimulants for sleepiness and TCAs for cataplexy and auxiliary symptoms). Some newer medications are proving efficacious for sleepiness with minimal adverse effects, whereas others may provide a single-drug regimen that simultaneously addresses sleepiness and cataplexy [18]. New research has allowed some experts to hypothesize that idiopathic narcolepsy may be the result of a genetic predisposition to autoimmune disease [176]. It is possible that aberrant genetic coding of elements in the hypocretin/orexin systems allows a sensitivity to inducible and possibly virally mediated changes, which leave cells in the lateral hypothalamus susceptible to autoimmune attack [96]. As such, genetic screening of high-risk individuals might eventually rationalize the prophylactic use of immunosuppressants in some instances. In the future, for atypical cases(poorly responsive to therapy), genetic, CSF, and brain imaging studies, and possibly even neuronal transplantation may prove beneficial in the assessment and treatment of idiopathic narcolepsy.

  14. Autoantibody targets in vaccine-associated narcolepsy.

    PubMed

    Häggmark-Månberg, Anna; Zandian, Arash; Forsström, Björn; Khademi, Mohsen; Lima Bomfim, Izaura; Hellström, Cecilia; Arnheim-Dahlström, Lisen; Hallböök, Tove; Darin, Niklas; Lundberg, Ingrid E; Uhlén, Mathias; Partinen, Markku; Schwenk, Jochen M; Olsson, Tomas; Nilsson, Peter

    2016-09-01

    Narcolepsy is a chronic sleep disorder with a yet unknown cause, but the specific loss of hypocretin-producing neurons together with a strong human leukocyte antigen (HLA) association has led to the hypothesis that autoimmune mechanisms might be involved. Here, we describe an extensive effort to profile autoimmunity repertoires in serum with the aim to find disease-related autoantigens. Initially, 57 serum samples from vaccine-associated and sporadic narcolepsy patients and controls were screened for IgG reactivity towards 10 846 fragments of human proteins using planar microarrays. The discovered differential reactivities were verified on suspension bead arrays in the same sample collection followed by further investigation of 14 antigens in 176 independent samples, including 57 narcolepsy patients. Among these 14 antigens, methyltransferase-like 22 (METTL22) and 5'-nucleotidase cytosolic IA (NT5C1A) were recognized at a higher frequency in narcolepsy patients of both sample sets. Upon sequence analysis of the 14 proteins, polymerase family, member 3 (PARP3), acyl-CoA-binding domain containing 7 (ARID4B), glutaminase 2 (GLS2) and cyclin-dependent kinase-like 1 (CDKL1) were found to contain amino acid sequences with homology to proteins found in the H1N1 vaccine. These findings could become useful elements of further clinical assays that aim towards a better phenotypic understanding of narcolepsy and its triggers.

  15. Anti-Tribbles Homolog 2 Autoantibodies in Japanese Patients with Narcolepsy

    PubMed Central

    Toyoda, Hiromi; Tanaka, Susumu; Miyagawa, Taku; Honda, Yutaka; Tokunaga, Katsushi; Honda, Makoto

    2010-01-01

    Study Objectives: Narcolepsy is a sleep disorder characterized by excessive daytime sleepiness and cataplexy. The association with human leukocyte antigen (HLA)-DQB1*0602 and T-cell receptor alpha locus suggests that autoimmunity plays a role in narcolepsy. A recent study reported an increased prevalence of autoantibodies against Tribbles homolog 2 (TRIB2) in patients with narcolepsy. To replicate this finding, we examined anti-TRIB2 autoantibodies in Japanese patients with narcolepsy. Design: We examined anti-TRIB2 autoantibodies against a full-length [35S]-labeled TRIB2 antigen in Japanese patients with narcolepsy-cataplexy (n = 88), narcolepsy without cataplexy (n = 18), and idiopathic hypersomnia with long sleep time (n = 11). The results were compared to Japanese healthy controls (n = 87). Thirty-seven healthy control subjects were positive for HLA-DRB1*1501-DQB1*0602. We also examined autoantibodies against another Tribbles homolog, TRIB3, as an experimental control. Measurements and Results: Autoantibodies against TRIB2 were found in 26.1% of patients with narcolepsy-cataplexy, a significantly higher prevalence than the 2.3% in healthy controls. We found that anti-TRIB3 autoantibodies were rare in patients with narcolepsy and showed no association with anti-TRIB2 indices. No significant correlation was found between anti-TRIB2 positivity and clinical information. Conclusions: We confirmed the higher prevalence and specificity of anti-TRIB2 autoantibodies in Japanese patients with narcolepsy-cataplexy. This suggests a subgroup within narcolepsy-cataplexy might be affected by an anti-TRIB2 autoantibody-mediated autoimmune mechanism. Citation: Toyoda H; Tanaka S; Miyagawa T; Honda Y; Tokunaga K; Honda M. Anti-Tribbles homolog 2 autoantibodies in Japanese patients with narcolepsy. SLEEP 2010;33(7):875-878. PMID:20614847

  16. The Humanistic and Economic Burden of Narcolepsy

    PubMed Central

    Flores, Natalia M.; Villa, Kathleen F.; Black, Jed; Chervin, Ronald D.; Witt, Edward A.

    2016-01-01

    Study Objectives: To evaluate the burden of narcolepsy--with respect to psychiatric comorbidities, Health-Related Quality of Life (HRQoL), direct costs for healthcare resource utilization, and indirect costs for reported work loss–through comparison of patients to matched controls. Methods: This analysis was conducted on data from the 2011, 2012, and 2013 US National Health and Wellness Survey (NHWS; 2011 NHWS n = 75,000, 2012 NHWS n = 71,157, and 2013 NHWS n = 75,000). Patients who reported a narcolepsy diagnosis (n = 437) were matched 1:2 with controls (n = 874) on age, sex, race/ethnicity, marital status, education, household income, body mass index, smoking status, alcohol use, exercise, and physical comorbidity. Chi-square tests and one-way analyses of variance were used to assess whether the narcolepsy and control groups differed on psychiatric comorbidities, HRQoL, labor force participation, work productivity, and healthcare resource utilization. Results: Patients with narcolepsy, in comparison to matched controls, reported substantially (two to four times) greater psychiatric comorbidity, HRQoL impairment, prevalence of long-term disability, absenteeism, and presenteeism, and greater resource use in the past 6 mo as indicated by higher mean number of hospitalizations, emergency department visits, traditional healthcare professional visits, neurologist visits, and psychiatrist visits (each p < 0.05). Conclusions: These population-based data suggest that a narcolepsy diagnosis is associated with substantial adverse impact on mental health, HRQoL, and key economic burdens that include work impairment, resource use, and both direct and indirect costs. Although this study is cross-sectional, the results highlight the magnitude of the potential opportunity to improve mental health, lower costs, and augment work-related productivity through effective assessment and treatment of narcolepsy. Citation: Flores NM, Villa KF, Black J, Chervin RD, Witt EA. The

  17. EIF3G is associated with narcolepsy across ethnicities.

    PubMed

    Holm, Anja; Lin, Ling; Faraco, Juliette; Mostafavi, Sara; Battle, Alexis; Zhu, Xiaowei; Levinson, Douglas F; Han, Fang; Gammeltoft, Steen; Jennum, Poul; Mignot, Emmanuel; Kornum, Birgitte R

    2015-11-01

    Type 1 narcolepsy, an autoimmune disease affecting hypocretin (orexin) neurons, is strongly associated with HLA-DQB1*06:02. Among polymorphisms associated with the disease is single-nucleotide polymorphism rs2305795 (c.*638G>A) located within the P2RY11 gene. P2RY11 is in a region of synteny conserved in mammals and zebrafish containing PPAN, EIF3G and DNMT1 (DNA methyltransferase 1). As mutations in DNMT1 cause a rare dominant form of narcolepsy in association with deafness, cerebellar ataxia and dementia, we questioned whether the association with P2RY11 in sporadic narcolepsy could be secondary to linkage disequilibrium with DNMT1. Based on genome-wide association data from two cohorts of European and Chinese ancestry, we found that the narcolepsy association signal drops sharply between P2RY11/EIF3G and DNMT1, suggesting that the association with narcolepsy does not extend into the DNMT1 gene region. Interestingly, using transethnic mapping, we identified a novel single-nucleotide polymorphism rs3826784 (c.596-260A>G) in the EIF3G gene also associated with narcolepsy. The disease-associated allele increases EIF3G mRNA expression. EIF3G is located in the narcolepsy risk locus and EIF3G expression correlates with PPAN and P2RY11 expression. This suggests shared regulatory mechanisms that might be affected by the polymorphism and are of relevance to narcolepsy.

  18. Status cataplecticus as initial presentation of late onset narcolepsy.

    PubMed

    Panda, Samhita

    2014-02-15

    Narcolepsy, one of the important causes of hypersomnia, is an under diagnosed sleep disorder. It has a bimodal age of onset around 15 and 35 years. It is characterized by the tetrad of excessive daytime sleepiness, cataplexy, hypnagogic/ hypnopompic hallucinations, and sleep paralysis. Cataplexy is by far the most predictive feature of narcolepsy. Status cataplecticus is the occurrence of cataplexy repeatedly for hours or days, a rare presentation of narcolepsy. This report describes an elderly gentleman with late onset narcolepsy in the sixth decade of life presenting with initial and chief symptom of status cataplecticus.

  19. Complex Diagnostic and Treatment Issues in Psychotic Symptoms Associated with Narcolepsy

    PubMed Central

    Ivanenko, Anna

    2009-01-01

    Narcolepsy is an uncommon chronic, neurological disorder characterized by abnormal manifestations of rapid eye movement sleep and perturbations in the sleep-wake cycle. Accurate diagnosis of psychotic symptoms in a person with narcolepsy could be difficult due to side effects of stimulant treatment (e.g., hallucinations) as well as primary symptoms of narcolepsy (e.g., sleep paralysis and hypnagogic and/or hypnapompic hallucinations). Pertinent articles from peer-reviewed journals were identified to help understand the complex phenomenology of psychotic symptoms in patients with narcolepsy. In this ensuing review and discussion, we present an overview of narcolepsy and outline diagnostic and management approaches for psychotic symptoms in patients with narcolepsy. PMID:19724760

  20. Delusional Confusion of Dreaming and Reality in Narcolepsy

    PubMed Central

    Wamsley, Erin; Donjacour, Claire E.H.M.; Scammell, Thomas E.; Lammers, Gert Jan; Stickgold, Robert

    2014-01-01

    Study Objectives: We investigated a generally unappreciated feature of the sleep disorder narcolepsy, in which patients mistake the memory of a dream for a real experience and form sustained delusions about significant events. Design: We interviewed patients with narcolepsy and healthy controls to establish the prevalence of this complaint and identify its predictors. Setting: Academic medical centers in Boston, Massachusetts and Leiden, The Netherlands. Participants: Patients (n = 46) with a diagnosis of narcolepsy with cataplexy, and age-matched healthy healthy controls (n = 41). Interventions: N/A. Measurements and Results: “Dream delusions” were surprisingly common in narcolepsy and were often striking in their severity. As opposed to fleeting hypnagogic and hypnopompic hallucinations of the sleep/wake transition, dream delusions were false memories induced by the experience of a vivid dream, which led to false beliefs that could persist for days or weeks. Conclusions: The delusional confusion of dreamed events with reality is a prominent feature of narcolepsy, and suggests the possibility of source memory deficits in this disorder that have not yet been fully characterized. Citation: Wamsley E; Donjacour CE; Scammell TE; Lammers GJ; Stickgold R. Delusional confusion of dreaming and reality in narcolepsy. SLEEP 2014;37(2):419-422. PMID:24501437

  1. Narcolepsy: Let the Patient’s Voice Awaken Us!

    PubMed Central

    Flygare, Julie; Parthasarathy, Sairam

    2014-01-01

    This is a “patient-centered” review about narcolepsy that aims to awaken the reader to the narcolepsy condition and to the trials and tribulations of patients with sleep problems in general. Narcolepsy is a neurological disorder with a classic tetrad of symptoms consisting of excessive daytime sleepiness, cataplexy, sleep onset hallucinations, and sleep paralysis. The diagnosis of narcolepsy and other sleep disorders are often overlooked and could be attributed to other medical or even psychiatric conditions with years of missed diagnosis. Implementation of “two sleep-related questions” to the review of systems in the primary care physicians’ office visit may help address the issue of missed diagnosis and allow patients to seek prompt medical attention. Definitive diagnosis can be made by overnight sleep study followed by a nap test, “multiple sleep latency test” (MSLT). There is currently no cure for narcolepsy with the treatments addressing symptoms of excessive daytime sleepiness, cataplexy, and nighttime sleep disruption with stimulants (modafinil, methylphenidate, and amphetamines), anti-cataplexy medications (Serotonin-specific reuptake inhibitors and tricyclic antidepressants) and sedative-hypnotics including sodium oxybate. Narcolepsy, like other sleep disorders, can lead to marked reductions of health-related quality of life and affect patients’ social and work lives deleteriously. While traditional healthcare approaches are focused more on hard biomedical outcomes, a patient-centered approach with novel methods for better sleep assessment of patients, that can bypass the “impossibly crammed” physician office visit, would allow healthcare providers to better detect, diagnose and treat narcolepsy and other such sleep problems. PMID:24931392

  2. Hypocretin Receptor Expression in Canine and Murine Narcolepsy Models and in Hypocretin-Ligand Deficient Human Narcolepsy

    PubMed Central

    Mishima, Kazuo; Fujiki, Nobuhiro; Yoshida, Yasushi; Sakurai, Takeshi; Honda, Makoto; Mignot, Emmanuel; Nishino, Seiji

    2008-01-01

    Study Objective: To determine whether hypocretin receptor gene (hcrtR1 and hcrtR2) expression is affected after long-term hypocretin ligand loss in humans and animal models of narcolepsy. Design: Animal and human study. We measured hcrtR1 and hcrtR2 expression in the frontal cortex and pons using the RT-PCR method in murine models (8-week-old and 27-week-old orexin/ataxin-3 transgenic (TG) hypocretin cell ablated mice and wild-type mice from the same litter, 10 mice for each group), in canine models (8 genetically narcoleptic Dobermans with null mutations in the hcrtR2, 9 control Dobermans, 3 sporadic ligand-deficient narcoleptics, and 4 small breed controls), and in humans (5 narcolepsy-cataplexy patients with hypocretin deficiency (average age 77.0 years) and 5 control subjects (72.6 years). Measurement and Results: 27-week-old (but not 8-week-old) TG mice showed significant decreases in hcrtR1 expression, suggesting the influence of the long-term ligand loss on the receptor expression. Both sporadic narcoleptic dogs and human narcolepsy-cataplexy subjects showed a significant decrease in hcrtR1 expression, while declines in hcrtR2 expression were not significant in these cases. HcrtR2-mutated narcoleptic Dobermans (with normal ligand production) showed no alteration in hcrtR1 expression. Conclusions: Moderate declines in hcrtR expressions, possibly due to long-term postnatal loss of ligand production, were observed in hypocretin-ligand deficient narcoleptic subjects. These declines are not likely to be progressive and complete. The relative preservation of hcrtR2 expression also suggests that hypocretin based therapies are likely to be a viable therapeutic options in human narcolepsy-cataplexy. Citation: Mishima K; Fujiki N; Yoshida Y; Sakurai T; Honda M; Mignot E; Nishino S. Hypocretin receptor expression in canine and murine narcolepsy models and in hypocretin-ligand deficient human narcolepsy. SLEEP 2008;31(8):1119-1126. PMID:18714784

  3. Facial expression recognition and emotional regulation in narcolepsy with cataplexy.

    PubMed

    Bayard, Sophie; Croisier Langenier, Muriel; Dauvilliers, Yves

    2013-04-01

    Cataplexy is pathognomonic of narcolepsy with cataplexy, and defined by a transient loss of muscle tone triggered by strong emotions. Recent researches suggest abnormal amygdala function in narcolepsy with cataplexy. Emotion treatment and emotional regulation strategies are complex functions involving cortical and limbic structures, like the amygdala. As the amygdala has been shown to play a role in facial emotion recognition, we tested the hypothesis that patients with narcolepsy with cataplexy would have impaired recognition of facial emotional expressions compared with patients affected with central hypersomnia without cataplexy and healthy controls. We also aimed to determine whether cataplexy modulates emotional regulation strategies. Emotional intensity, arousal and valence ratings on Ekman faces displaying happiness, surprise, fear, anger, disgust, sadness and neutral expressions of 21 drug-free patients with narcolepsy with cataplexy were compared with 23 drug-free sex-, age- and intellectual level-matched adult patients with hypersomnia without cataplexy and 21 healthy controls. All participants underwent polysomnography recording and multiple sleep latency tests, and completed depression, anxiety and emotional regulation questionnaires. Performance of patients with narcolepsy with cataplexy did not differ from patients with hypersomnia without cataplexy or healthy controls on both intensity rating of each emotion on its prototypical label and mean ratings for valence and arousal. Moreover, patients with narcolepsy with cataplexy did not use different emotional regulation strategies. The level of depressive and anxious symptoms in narcolepsy with cataplexy did not differ from the other groups. Our results demonstrate that narcolepsy with cataplexy accurately perceives and discriminates facial emotions, and regulates emotions normally. The absence of alteration of perceived affective valence remains a major clinical interest in narcolepsy with cataplexy

  4. Investigational therapies for the treatment of narcolepsy.

    PubMed

    de Biase, Stefano; Nilo, Annacarmen; Gigli, Gian Luigi; Valente, Mariarosaria

    2017-08-01

    Narcolepsy is a chronic sleep disorder characterized by a pentad of excessive daytime sleepiness (EDS), cataplexy, sleep paralysis, hypnagogic/hypnopompic hallucinations, and disturbed nocturnal sleep. While non-pharmacological treatments are sometimes helpful, more than 90% of narcoleptic patients require a pharmacological treatment. Areas covered: The present review is based on an extensive Internet and PubMed search from 1994 to 2017. It is focused on drugs currently in development for the treatment of narcolepsy. Expert opinion: Currently there is no cure for narcolepsy, with treatment focusing on symptoms control. However, these symptomatic treatments are often unsatisfactory. The research is leading to a better understanding of narcolepsy and its symptoms. New classes of compounds with possible applications in the development of novel stimulant/anticataplectic medications are described. H3 receptor antagonists represent a new therapeutic option for EDS in narcolepsy. JZP-110, with its distinct mechanism of action, would be a new therapeutic option for the treatment of EDS in the coming years. In the future, hypocretin-based therapies and immune-based therapies, could modify the clinical course of the disease. However, more information would be necessary to completely understand the autoimmune process and also how this process can be altered for therapeutic benefits.

  5. Update on the treatment of narcolepsy: clinical efficacy of pitolisant

    PubMed Central

    Calik, Michael W

    2017-01-01

    Narcolepsy is a neurological disease that affects 1 in 2,000 individuals and is characterized by excessive daytime sleepiness (EDS). In 60–70% of individuals with narcolepsy, it is also characterized by cataplexy or a sudden loss of muscle tone that is triggered by positive or negative emotions. Narcolepsy decreases the quality of life of the afflicted individuals. Currently used drugs treat EDS alone (modafinil/armodafinil, methylphenidate, and amphetamine), cataplexy alone (“off-label” use of antidepressants), or both EDS and cataplexy (sodium oxybate). These drugs have abuse, tolerability, and adherence issues. A greater diversity of drug options is needed to treat narcolepsy. The small molecule drug, pitolisant, acts as an inverse agonist/antagonist at the H3 receptor, thus increasing histaminergic tone in the wake promoting system of the brain. Pitolisant has been studied in animal models of narcolepsy and used in clinical trials as a treatment for narcolepsy. A comprehensive search of online databases (eg, Medline, PubMed, EMBASE, the Cochrane Library Database, Ovid MEDLINE, Europe PubMed Central, EBSCOhost CINAHL, ProQuest Research Library, Google Scholar, and ClinicalTrials.gov) was performed. Nonrandomized and randomized studies were included. This review focuses on the outcomes of four clinical trials of pitolisant to treat narcolepsy. These four trials show that pitolisant is an effective drug to treat EDS and cataplexy in narcolepsy. PMID:28490912

  6. A Global View on Narcolepsy - A Review Study.

    PubMed

    Klimova, Blanka; Maresova, Petra; Novotny, Michal; Kuca, Kamil

    2018-02-14

    Narcolepsy is an incurable neurological disorder when the brain is not able to regulate a sleep and wakefulness cycle correctly. The affected person suddenly falls asleep during the day or he/she suffers from excessive day sleepiness. In addition, people may also suffer from cataplexy, hypnagogic hallucinations, sleep paralysis, and disturbed nighttime sleep. The purpose of this review study is to provide the latest information on both clinical and socioeconomic issues in the field of narcolepsy treatment and emphasize its benefits and limitations. The methodological approaches include a method of literature review of available sources exploring the issue of narcolepsy, both from a global and specific perspective point of view. On the basis of evaluation of these literature sources, the researched issue is examined. The main benefits (e.g., new drugs are being tested or non-invasive cognitive behavioral therapies are being applied) and limitations (e.g., late diagnosis of the disease or lifelong and costly treatment) of the treatment of narcolepsy are highlighted. The findings call for more research in the field of the development of novel drugs reflecting understanding of the neurological basis of narcolepsy and early diagnosis in order to eliminate the symptoms of narcolepsy and prevent the development of this disease. Copyright© Bentham Science Publishers; For any queries, please email at epub@benthamscience.org.

  7. Allergies and Disease Severity in Childhood Narcolepsy: Preliminary Findings.

    PubMed

    Aydinoz, Secil; Huang, Yu-Shu; Gozal, David; Inocente, Clara O; Franco, Patricia; Kheirandish-Gozal, Leila

    2015-12-01

    Narcolepsy frequently begins in childhood, and is characterized by excessive daytime sleepiness, with the presence of cataplexy reflecting a more severe phenotype. Narcolepsy may result from genetic predisposition involving deregulation of immune pathways, particularly involving T helper 2 cells (Th2). Increased activation of Th2 cells is usually manifested as allergic conditions such as rhinitis, atopic dermatitis, and asthma. We hypothesized that the presence of allergic conditions indicative of increased Th2 balance may dampen the severity of the phenotype in children with narcolepsy. A retrospective chart review of childhood narcolepsy patients was conducted at three major pediatric sleep centers. Patients were divided into those with narcolepsy without cataplexy (NC-) and narcolepsy with cataplexy (NC+). Demographics, polysomnographic and multiple sleep latency test data, and extraction of information on the presence of allergic diseases such allergic rhinitis, atopic dermatitis, and asthma was performed. There were 468 children identified, with 193 children in NC- group and 275 patients in the NC+ group. Overall, NC+ children were significantly younger, had higher body mass index, and had shorter mean sleep latencies and increased sleep onset rapid eye movement events. The frequency of allergic conditions, particularly asthma and allergic rhinitis, was markedly lower in NC+ (58/275) compared to NC- patients (94/193; P < 0.0001). Involvement of the immune system plays an important role in the pathophysiology of narcolepsy. Current findings further suggest that an increased shift toward T helper 2 cells, as indicated by the presence of allergic conditions, may modulate the severity of the phenotype in childhood narcolepsy, and reduce the prevalence of cataplexy in these patients. © 2015 Associated Professional Sleep Societies, LLC.

  8. The Medical and Economic Burden of Narcolepsy: Implications for Managed Care

    PubMed Central

    Thorpy, Michael J.; Hiller, George

    2017-01-01

    Background The neurologic disorder narcolepsy results from dysregulation of the sleep-wake cycle and is primarily characterized by chronic, severely excessive daytime sleepiness and cataplexy, an emotionally induced muscle weakness. The prevalence of narcolepsy is approximately 0.05%, and onset generally occurs during the first 2 decades of life. Narcolepsy is believed to be an autoimmune disorder with destruction of hypocretin-producing neurons in the lateral hypothalamus. Objectives To provide an enhanced understanding of narcolepsy and establish the need for early diagnosis and rapid initiation of effective treatment for patients with narcolepsy. Discussion Narcolepsy reduces daily functioning and is associated with a substantial medical and economic burden, with many patients being on full disability. The annual direct medical costs are approximately 2-fold higher in patients with narcolepsy than in matched controls without this condition ($11,702 vs $5261, respectively; P <.0001). Further contributing to the overall burden is a lack of recognition of the signs and symptoms of narcolepsy and an absence of easily measurable biomarkers, resulting in a diagnostic delay that often exceeds 10 years and may be associated with misdiagnosis and inappropriate resource utilization. Because narcolepsy generally has an onset in childhood or in adolescence, is often misdiagnosed, has no known cure, and requires lifelong treatment, it is an important disease from a managed care perspective. Clinical features, as well as objective testing, should be used to ensure the timely diagnosis and treatment of patients with narcolepsy. Conclusion Policies for the diagnosis and treatment of narcolepsy should be based on the current treatment guidelines, but they should also encourage shared decisions between clinicians and patients to allow for individualized diagnostic and treatment choices, as suggested in best practice recommendations. PMID:28975007

  9. Anesthetic Management of Narcolepsy Patients During Surgery: A Systematic Review.

    PubMed

    Hu, Sally; Singh, Mandeep; Wong, Jean; Auckley, Dennis; Hershner, Shelley; Kakkar, Rahul; Thorpy, Michael J; Chung, Frances

    2018-01-01

    Narcolepsy is a rare sleep disorder characterized by excessive daytime sleepiness, sleep paralysis, and/or hypnagogic/hypnopompic hallucinations, and in some cases cataplexy. The response to anesthetic medications and possible interactions in narcolepsy patients is unclear in the perioperative period. In this systematic review, we aim to evaluate the current evidence on the perioperative outcomes and anesthetic considerations in narcolepsy patients. Electronic literature search of Medline, Medline in-process, Embase, Cochrane Database of Systematic Reviews databases, international conference proceedings, and abstracts was conducted in November 2015 according to the Preferred Reporting Items for Systematic Reviews and Meta-Analysis Protocols guideline. A total of 3757 articles were screened using a 2-stage strategy (title-abstract followed by full text). We included case studies/series, cohort studies, and randomized controlled trials of narcolepsy patients undergoing surgical procedures under anesthesia or sedation. Preoperative narcolepsy symptoms and sleep study data, anesthetic technique, and perioperative complications were extracted. Screening of articles, data extraction, and compilation were conducted by 2 independent reviewers and any conflict was resolved by the senior author. A total of 19 studies including 16 case reports and 3 case series were included and evaluated. The majority of these patients received general anesthesia, whereas a small percentage of patients received regional anesthesia. Reported complications of narcolepsy patients undergoing surgeries were mainly related to autonomic dysregulation, or worsening of narcolepsy symptoms intra/postoperatively. Narcolepsy symptoms worsened only in those patient populations where the preoperative medications were either discontinued or reduced (mainly in obstetric patients). In narcolepsy patients, use of depth of anesthesia monitoring and total intravenous technique may have some advantage in terms

  10. A Consensus Definition of Cataplexy in Mouse Models of Narcolepsy

    PubMed Central

    Scammell, Thomas E.; Willie, Jon T.; Guilleminault, Christian; Siegel, Jerome M.

    2009-01-01

    People with narcolepsy often have episodes of cataplexy, brief periods of muscle weakness triggered by strong emotions. Many researchers are now studying mouse models of narcolepsy, but definitions of cataplexy-like behavior in mice differ across labs. To establish a common language, the International Working Group on Rodent Models of Narcolepsy reviewed the literature on cataplexy in people with narcolepsy and in dog and mouse models of narcolepsy and then developed a consensus definition of murine cataplexy. The group concluded that murine cataplexy is an abrupt episode of nuchal atonia lasting at least 10 seconds. In addition, theta activity dominates the EEG during the episode, and video recordings document immobility. To distinguish a cataplexy episode from REM sleep after a brief awakening, at least 40 seconds of wakefulness must precede the episode. Bouts of cataplexy fitting this definition are common in mice with disrupted orexin/hypocretin signaling, but these events almost never occur in wild type mice. It remains unclear whether murine cataplexy is triggered by strong emotions or whether mice remain conscious during the episodes as in people with narcolepsy. This working definition provides helpful insights into murine cataplexy and should allow objective and accurate comparisons of cataplexy in future studies using mouse models of narcolepsy. Citation: Scammell TE; Willie JT; Guilleminault C; Siegel JM. A consensus definition of cataplexy in mouse models of narcolepsy. SLEEP 2009;32(1):111-116. PMID:19189786

  11. Concomitant loss of dynorphin, NARP, and orexin in narcolepsy

    PubMed Central

    Crocker, Amanda; España, Rodrigo A.; Papadopoulou, Maria; Saper, Clifford B.; Faraco, Juliette; Sakurai, Takeshi; Honda, Makoto; Mignot, Emmanuel; Scammell, Thomas E.

    2008-01-01

    Background Narcolepsy with cataplexy is associated with a loss of orexin/hypocretin. It is speculated that an autoimmune process kills the orexin-producing neurons, but these cells may survive yet fail to produce orexin. Objective To examine whether other markers of the orexin neurons are lost in narcolepsy with cataplexy. Methods We used immunohistochemistry and in situ hybridization to examine the expression of orexin, neuronal activity-regulated pentraxin (NARP), and prodynorphin in hypothalami from five control and two narcoleptic individuals. Results In the control hypothalami, at least 80% of the orexin-producing neurons also contained prodynorphin mRNA and NARP. In the patients with narcolepsy, the number of cells producing these markers was reduced to about 5–10% of normal. Conclusions Narcolepsy with cataplexy is likely caused by a loss of the orexin-producing neurons. In addition, loss of dynorphin and NARP may contribute to the symptoms of narcolepsy. PMID:16247044

  12. Narcolepsy: current treatment options and future approaches

    PubMed Central

    Billiard, Michel

    2008-01-01

    The management of narcolepsy is presently at a turning point. Three main avenues are considered in this review: 1) Two tendencies characterize the conventional treatment of narcolepsy. Modafinil has replaced methylphenidate and amphetamine as the first-line treatment of excessive daytime sleepiness (EDS) and sleep attacks, based on randomized, double blind, placebo-controlled clinical trials of modafinil, but on no direct comparison of modafinil versus traditional stimulants. For cataplexy, sleep paralysis, and hypnagogic hallucinations, new antidepressants tend to replace tricyclic antidepressants and selective serotonin reuptake inhibitors (SSRIs) in spite of a lack of randomized, double blind, placebo-controlled clinical trials of these compounds; 2) The conventional treatment of narcolepsy is now challenged by sodium oxybate, the sodium salt of gammahydroxybutyrate, based on a series of randomized, double-blind, placebo-controlled clinical trials and a long-term open label study. This treatment has a fairly good efficacy and is active on all symptoms of narcolepsy. Careful titration up to an adequate level is essential both to obtain positive results and avoid adverse effects; 3) A series of new treatments are currently being tested, either in animal models or in humans, They include novel stimulant and anticataplectic drugs, endocrine therapy, and, more attractively, totally new approaches based on the present state of knowledge of the pathophysiology of narcolepsy with cataplexy, hypocretine-based therapies, and immunotherapy. PMID:18830438

  13. A consensus definition of cataplexy in mouse models of narcolepsy.

    PubMed

    Scammell, Thomas E; Willie, Jon T; Guilleminault, Christian; Siegel, Jerome M

    2009-01-01

    People with narcolepsy often have episodes of cataplexy, brief periods of muscle weakness triggered by strong emotions. Many researchers are now studying mouse models of narcolepsy, but definitions of cataplexy-like behavior in mice differ across labs. To establish a common language, the International Working Group on Rodent Models of Narcolepsy reviewed the literature on cataplexy in people with narcolepsy and in dog and mouse models of narcolepsy and then developed a consensus definition of murine cataplexy. The group concluded that murine cataplexy is an abrupt episode of nuchal atonia lasting at least 10 seconds. In addition, theta activity dominates the EEG during the episode, and video recordings document immobility. To distinguish a cataplexy episode from REM sleep after a brief awakening, at least 40 seconds of wakefulness must precede the episode. Bouts of cataplexy fitting this definition are common in mice with disrupted orexin/hypocretin signaling, but these events almost never occur in wild type mice. It remains unclear whether murine cataplexy is triggered by strong emotions or whether mice remain conscious during the episodes as in people with narcolepsy. This working definition provides helpful insights into murine cataplexy and should allow objective and accurate comparisons of cataplexy in future studies using mouse models of narcolepsy.

  14. The Familial Risk and HLA Susceptibility among Narcolepsy Patients in Hong Kong Chinese

    PubMed Central

    Chen, Lei; Fong, S.Y.Y.; Lam, Ching W.; Tang, Nelson L.S.; Ng, Margaret H. L.; Li, Albert M.; Ho, C.K.W.; Cheng, Suk-Hang; Lau, Kin-Mang; Wing, Yun Kwok

    2007-01-01

    Study Objectives: To explore the familial aggregation and HLA susceptibility of narcolepsy in Hong Kong Chinese by objective sleep measurements and HLA typing. Design: Case control design Participants: Twelve narcoleptic probands, 34 first-degree relatives, and 30 healthy controls. Interventions: N/A Measurements and Results: Each subject underwent a standardized nocturnal polysomnogram (PSG), followed by a daytime multiple sleep latency test (MSLT). HLA typing was performed for all subjects. One relative (2.9%) was diagnosed as suffering from narcolepsy with cataplexy. Nearly 30% of the relatives fulfilled the criteria of narcolepsy spectrum disorder (shortened mean sleep latency [MSL] and/or the presence of sleep onset REM periods [SOREMPs]). When using the population data for comparison, the relative risk of narcolepsy in first-degree relatives was 85.3. The odds ratio of narcolepsy spectrum disorder in first-degree relatives was 5.8 (95% CI: 1.2 – 29.3) when compared to healthy controls. There existed 6 multiplex families, in which all 10 relatives with narcolepsy spectrum disorders, including all 3 relatives with multiple SOREMPs, were positive for HLA DQB1*0602. Conclusions: Our study demonstrated a definitive familial aggregation of narcolepsy, narcolepsy spectrum disorders, and possibly cataplexy in Hong Kong Chinese. This familial aggregation supported an inherited basis for narcolepsy spectrum. The tight co-segregation of HLA DQB1*0602 and narcolepsy spectrum disorders might suggest that HLA typing, especially DQB1*0602, at least partly confer the familial risk of narcolepsy. In addition, our study suggested that the subjective questionnaire measurements including Ullanlinna Narcolepsy Scale and Epworth Sleepiness Scale were unable to detect the presence of narcolepsy spectrum disorders among the relatives. A stringent objective measurement-based design for family studies is suggested for future study. Further studies are indicated for the determination

  15. The Humanistic and Economic Burden of Narcolepsy.

    PubMed

    Flores, Natalia M; Villa, Kathleen F; Black, Jed; Chervin, Ronald D; Witt, Edward A

    2016-03-01

    To evaluate the burden of narcolepsy--with respect to psychiatric comorbidities, Health-Related Quality of Life (HRQoL), direct costs for healthcare resource utilization, and indirect costs for reported work loss-through comparison of patients to matched controls. This analysis was conducted on data from the 2011, 2012, and 2013 US National Health and Wellness Survey (NHWS; 2011 NHWS n = 75,000, 2012 NHWS n = 71,157, and 2013 NHWS n = 75,000). Patients who reported a narcolepsy diagnosis (n = 437) were matched 1:2 with controls (n = 874) on age, sex, race/ethnicity, marital status, education, household income, body mass index, smoking status, alcohol use, exercise, and physical comorbidity. Chi-square tests and one-way analyses of variance were used to assess whether the narcolepsy and control groups differed on psychiatric comorbidities, HRQoL, labor force participation, work productivity, and healthcare resource utilization. Patients with narcolepsy, in comparison to matched controls, reported substantially (two to four times) greater psychiatric comorbidity, HRQoL impairment, prevalence of long-term disability, absenteeism, and presenteeism, and greater resource use in the past 6 mo as indicated by higher mean number of hospitalizations, emergency department visits, traditional healthcare professional visits, neurologist visits, and psychiatrist visits (each p < 0.05). These population-based data suggest that a narcolepsy diagnosis is associated with substantial adverse impact on mental health, HRQoL, and key economic burdens that include work impairment, resource use, and both direct and indirect costs. Although this study is cross-sectional, the results highlight the magnitude of the potential opportunity to improve mental health, lower costs, and augment work-related productivity through effective assessment and treatment of narcolepsy. © 2016 American Academy of Sleep Medicine.

  16. Narcolepsy and Orexins: An Example of Progress in Sleep Research

    PubMed Central

    De la Herrán-Arita, Alberto K.; Guerra-Crespo, Magdalena; Drucker-Colín, René

    2011-01-01

    Narcolepsy is a chronic neurodegenerative disease caused by a deficiency of orexin-producing neurons in the lateral hypothalamus. It is clinically characterized by excessive daytime sleepiness and by intrusions into wakefulness of physiological aspects of rapid eye movement sleep such as cataplexy, sleep paralysis, and hypnagogic hallucinations. The major pathophysiology of narcolepsy has been recently described on the bases of the discovery of the neuropeptides named orexins (hypocretins) in 1998; considerable evidence, summarized below, demonstrates that narcolepsy is the result of alterations in the genes involved in the pathology of the orexin ligand or its receptor. Deficient orexin transmission is sufficient to produce narcolepsy, as we describe here, animal models with dysregulated orexin signaling exhibit a narcolepsy-like phenotype. Remarkably, these narcoleptic models have different alterations of the orexinergic circuit, this diversity provide us with the means for making comparison, and have a better understanding of orexin-cell physiology. It is of particular interest that the most remarkable findings regarding this sleep disorder were fortuitous and due to keen observations. Sleep is a highly intricate and regulated state, and narcolepsy is a disorder that still remains as one of the unsolved mysteries in science. Nevertheless, advances and development of technology in neuroscience will provide us with the necessary tools to unravel the narcolepsy puzzle in the near future. Through an evaluation of the scientific literature we traced an updated picture of narcolepsy and orexins in order to provide insight into the means by which neurobiological knowledge is constructed. PMID:21541306

  17. Parkinson's disease and narcolepsy-like symptoms.

    PubMed

    Ylikoski, Ari; Martikainen, Kirsti; Sarkanen, Tomi; Partinen, Markku

    2015-04-01

    Various sleep-related problems, for example, insomnia and symptoms of rapid eye movement behavior disorder (RBD), are common in patients with Parkinson's disease (PD). We studied the prevalence of symptoms of narcolepsy (NARC), hallucinations, and RBD and their association with other symptoms. Altogether, 1447 randomly selected patients with PD, aged 43-89 years, participated in a questionnaire study. A structured questionnaire with 207 items was based on the Basic Nordic Sleep Questionnaire. Questions on demographics, PD, RBD, and other issues were included. The response rate was 59.0%; of these patients, 73% had answered to all questions that were used in the analyses (N = 623). The occurrence of suspected narcolepsy (Ullanlinna Narcolepsy Scale ≥ 14 and Epworth Sleepiness Scale ≥ 11) was observed in 9.3% of the subjects (PD with NARC), RBD (REM Sleep Behavior Disorder Screening Questionnaire ≥ 6) in 39.2% of all patients with PD, and in 62.1% of those with PD and NARC. In patients with PD, hallucinations before going to bed in the evening occurred in 5.8%, hypnagogic hallucinations in 4.0%, hallucinations during night 8.3%, and hypnopompic hallucinations in 3.2%. Cataplexy symptoms occurred in 43.1% of subjects with PD and NARC. In a logistic regression analysis, PD with NARC was associated with RBD, all types of hallucinations, daytime sleepiness, fatigue, insomnia, and intense dreaming also when adjusted for age, sex, disease duration, and levodopa. Narcolepsy-like symptoms may be present in patients with PD. Symptoms of RBD were associated with symptoms of narcolepsy including symptoms of cataplexy. Copyright © 2014 Elsevier B.V. All rights reserved.

  18. Cerebrospinal Fluid Biomarkers of Neurodegeneration Are Decreased or Normal in Narcolepsy.

    PubMed

    Jørgen Jennum, Poul; Østergaard Pedersen, Lars; Czarna Bahl, Justyna Maria; Modvig, Signe; Fog, Karina; Holm, Anja; Rahbek Kornum, Birgitte; Gammeltoft, Steen

    2017-01-01

    To investigate whether cerebrospinal fluid (CSF) biomarkers of neurodegeneration are altered in narcolepsy in order to evaluate whether the hypocretin deficiency and abnormal sleep-wake pattern in narcolepsy leads to neurodegeneration. Twenty-one patients with central hypersomnia (10 type 1 narcolepsy, 5 type 2 narcolepsy, and 6 idiopathic hypersomnia cases), aged 33 years on average and with a disease duration of 2-29 years, and 12 healthy controls underwent CSF analyses of the levels of β-amyloid, total tau protein (T-tau), phosphorylated tau protein (P-tau181), α-synuclein, neurofilament light chain (NF-L), and chitinase 3-like protein-1 (CHI3L1). Levels of β-amyloid were lower in patients with type 1 narcolepsy (375.4 ± 143.5 pg/mL) and type 2 narcolepsy (455.9 ± 65.0 pg/mL) compared to controls (697.9 ± 167.3 pg/mL, p < .05). Furthermore, in patients with type 1 narcolepsy, levels of T-tau (79.0 ± 27.5 pg/mL) and P-tau181 (19.1 ± 4.3 pg/mL) were lower than in controls (162.2 ± 49.9 pg/mL and 33.8 ± 9.2 pg/mL, p < .05). Levels of α-synuclein, NF-L, and CHI3L1 in CSF from narcolepsy patients were similar to those of healthy individuals. Six CSF biomarkers of neurodegeneration were decreased or normal in narcolepsy indicating that taupathy, synucleinopathy, and immunopathy are not prevalent in narcolepsy patients with a disease duration of 2-29 years. Lower CSF levels of β-amyloid, T-tau protein, and P-tau181 in narcolepsy may indicate that hypocretin deficiency and an abnormal sleep-wake pattern alter the turnover of these proteins in the central nervous system. © Sleep Research Society 2016. Published by Oxford University Press on behalf of the Sleep Research Society. All rights reserved. For permissions, please e-mail journals.permissions@oup.com.

  19. Narcolepsy and cataplexy: a pediatric case report.

    PubMed

    Savaş, Tülin; Erol, Ilknur; Saygı, Semra; Habeşoğlu, Mehmet Ali

    2016-12-01

    Narcolepsy is characterized by excessive sleepiness, cataplexy, hypnagogic hallucinations, and sleep paralysis during the rapid eye movement period of sleep. Herein, we present a boy aged eight years who was diagnosed as having narcolepsy and cataplexy about thirteen months after his first presentation. He was admitted with symptoms of daytime sleepiness. In the follow-up, cataplexy in the form of head dropping attacks developed seven months after the first admission. The patient was investigated for different prediagnoses and was eventually diagnosed as having narcolepsy and cataplexy through polysomnography and multiple sleep latency tests thirteen months after the first presentation. He is being followed up and is under drug therapy; his symptoms have improved substantially.

  20. Central Disorders of Hypersomnolence: Focus on the Narcolepsies and Idiopathic Hypersomnia.

    PubMed

    Khan, Zeeshan; Trotti, Lynn Marie

    2015-07-01

    The central disorders of hypersomnolence are characterized by severe daytime sleepiness, which is present despite normal quality and timing of nocturnal sleep. Recent reclassification distinguishes three main subtypes: narcolepsy type 1, narcolepsy type 2, and idiopathic hypersomnia (IH), which are the focus of this review. Narcolepsy type 1 results from loss of hypothalamic hypocretin neurons, while the pathophysiology underlying narcolepsy type 2 and IH remains to be fully elucidated. Treatment of all three disorders focuses on the management of sleepiness, with additional treatment of cataplexy in those patients with narcolepsy type 1. Sleepiness can be treated with modafinil/armodafinil or sympathomimetic CNS stimulants, which have been shown to be beneficial in randomized controlled trials of narcolepsy and, quite recently, IH. In those patients with narcolepsy type 1, sodium oxybate is effective for the treatment of both sleepiness and cataplexy. Despite these treatments, there remains a subset of hypersomnolent patients with persistent sleepiness, in whom alternate therapies are needed. Emerging treatments for sleepiness include histamine H3 antagonists (eg, pitolisant) and possibly negative allosteric modulators of the gamma-aminobutyric acid-A receptor (eg, clarithromycin and flumazenil).

  1. Mechanistic insights into influenza vaccine-associated narcolepsy.

    PubMed

    Ahmed, S Sohail; Steinman, Lawrence

    2016-12-01

    We previously reported an increased frequency of antibodies to hypocretin (HCRT) receptor 2 in sera obtained from narcoleptic patients who received the European AS03-adjuvanted vaccine Pandemrix (GlaxoSmithKline Biologicals, s.a.) for the global influenza A H1N1 pandemic in 2009 [A(H1N1)pdm09]. These antibodies cross-reacted with a particular fragment of influenza nucleoprotein (NP) - one of the proteins naturally contained in the virus used to make seasonal influenza vaccine and pandemic influenza vaccines. The purpose of this commentary is to provide additional insights and interpretations of the findings and share additional data not presented in the original paper to help the reader appreciate the key messages of that publication. First, a brief background to narcolepsy and vaccine-induced narcolepsy will be provided. Then, additional insights and clarification will be provided on the following topics: 1) the critical difference identified in the adjuvanted A(H1N1)pdm09 vaccines, 2) the contributing factor likely for the discordant association of narcolepsy between the AS03-adjuvanted pandemic vaccines Pandemrix and Arepanrix (GlaxoSmithKline Biologicals, s.a.), 3) the significance of detecting HCRT receptor 2 (HCRTr2) antibodies in some Finnish control subjects, 4) the approach used for the detection of HCRTr2 antibodies in vaccine-associated narcolepsy, and 5) the plausibility of the proposed mechanism involving HCRTr2 modulation in vaccine-associated narcolepsy.

  2. Sleeping Beauty in the Classroom: What Do Teachers Know about Narcolepsy?

    ERIC Educational Resources Information Center

    Cosgrove, Maryellen S.

    2002-01-01

    Investigated teachers' awareness of narcolepsy and the accuracy of their knowledge. Found educators uninformed about narcolepsy and how to accommodate narcoleptic students. Recommended that teachers be aware of narcolepsy's symptoms, plan variety and movement into students' lessons, and allow students to redo assignments if a sleep attack…

  3. High-dimensional single-cell analysis reveals the immune signature of narcolepsy.

    PubMed

    Hartmann, Felix J; Bernard-Valnet, Raphaël; Quériault, Clémence; Mrdjen, Dunja; Weber, Lukas M; Galli, Edoardo; Krieg, Carsten; Robinson, Mark D; Nguyen, Xuan-Hung; Dauvilliers, Yves; Liblau, Roland S; Becher, Burkhard

    2016-11-14

    Narcolepsy type 1 is a devastating neurological sleep disorder resulting from the destruction of orexin-producing neurons in the central nervous system (CNS). Despite its striking association with the HLA-DQB1*06:02 allele, the autoimmune etiology of narcolepsy has remained largely hypothetical. Here, we compared peripheral mononucleated cells from narcolepsy patients with HLA-DQB1*06:02-matched healthy controls using high-dimensional mass cytometry in combination with algorithm-guided data analysis. Narcolepsy patients displayed multifaceted immune activation in CD4 + and CD8 + T cells dominated by elevated levels of B cell-supporting cytokines. Additionally, T cells from narcolepsy patients showed increased production of the proinflammatory cytokines IL-2 and TNF. Although it remains to be established whether these changes are primary to an autoimmune process in narcolepsy or secondary to orexin deficiency, these findings are indicative of inflammatory processes in the pathogenesis of this enigmatic disease. © 2016 Hartmann et al.

  4. High-dimensional single-cell analysis reveals the immune signature of narcolepsy

    PubMed Central

    Quériault, Clémence; Krieg, Carsten; Nguyen, Xuan-Hung

    2016-01-01

    Narcolepsy type 1 is a devastating neurological sleep disorder resulting from the destruction of orexin-producing neurons in the central nervous system (CNS). Despite its striking association with the HLA-DQB1*06:02 allele, the autoimmune etiology of narcolepsy has remained largely hypothetical. Here, we compared peripheral mononucleated cells from narcolepsy patients with HLA-DQB1*06:02-matched healthy controls using high-dimensional mass cytometry in combination with algorithm-guided data analysis. Narcolepsy patients displayed multifaceted immune activation in CD4+ and CD8+ T cells dominated by elevated levels of B cell–supporting cytokines. Additionally, T cells from narcolepsy patients showed increased production of the proinflammatory cytokines IL-2 and TNF. Although it remains to be established whether these changes are primary to an autoimmune process in narcolepsy or secondary to orexin deficiency, these findings are indicative of inflammatory processes in the pathogenesis of this enigmatic disease. PMID:27821550

  5. Health-related stigma as a determinant of functioning in young adults with narcolepsy.

    PubMed

    Kapella, Mary C; Berger, Barbara E; Vern, Boris A; Vispute, Sachin; Prasad, Bharati; Carley, David W

    2015-01-01

    Symptoms of narcolepsy tend to arise during adolescence or young adulthood, a formative time in human development during which people are usually completing their education and launching a career. Little is known about the impact of narcolepsy on the social aspects of health-related quality of life in young adults. The purpose of this study was to examine relationships between health-related stigma, mood (anxiety and depression) and daytime functioning in young adults with narcolepsy compared to those without narcolepsy. Young adults (age 18-35) with narcolepsy (N = 122) and without narcolepsy (N = 93) were mailed a packet that included questionnaires and a self-addressed postage paid envelope. The questionnaire included demographic information and a composite of instruments including the SF 36, Functional Outcomes of Sleep Questionnaire (FOSQ), Fife Stigma Scale (FSS), Epworth Sleepiness Scale (ESS) and Hospital Anxiety and Depression Scale (HADS). Variable associations were assessed using descriptive statistics, ANOVA, Mann-Whitney U Test, correlations, stepwise multiple regression and path analysis. Young adults with narcolepsy perceived significantly more stigma and lower mood and health-related quality of life than young adults without narcolepsy (p<0.01). Health-related stigma was directly and indirectly associated with lower functioning through depressed mood. Fifty-two percent of the variance in functioning was explained by the final model in the young adults with narcolepsy. Health-related stigma in young adults with narcolepsy is at a level consistent with other chronic medical illnesses. Health-related stigma may be an important determinant of functioning in young adults with narcolepsy. Future work is indicated toward further characterizing stigma and developing interventions that address various domains of stigma in people with narcolepsy.

  6. Health-Related Stigma as a Determinant of Functioning in Young Adults with Narcolepsy

    PubMed Central

    Kapella, Mary C.; Berger, Barbara E.; Vern, Boris A.; Vispute, Sachin; Prasad, Bharati; Carley, David W.

    2015-01-01

    Symptoms of narcolepsy tend to arise during adolescence or young adulthood, a formative time in human development during which people are usually completing their education and launching a career. Little is known about the impact of narcolepsy on the social aspects of health-related quality of life in young adults. The purpose of this study was to examine relationships between health-related stigma, mood (anxiety and depression) and daytime functioning in young adults with narcolepsy compared to those without narcolepsy. Young adults (age 18–35) with narcolepsy (N = 122) and without narcolepsy (N = 93) were mailed a packet that included questionnaires and a self-addressed postage paid envelope. The questionnaire included demographic information and a composite of instruments including the SF 36, Functional Outcomes of Sleep Questionnaire (FOSQ), Fife Stigma Scale (FSS), Epworth Sleepiness Scale (ESS) and Hospital Anxiety and Depression Scale (HADS). Variable associations were assessed using descriptive statistics, ANOVA, Mann-Whitney U Test, correlations, stepwise multiple regression and path analysis. Young adults with narcolepsy perceived significantly more stigma and lower mood and health-related quality of life than young adults without narcolepsy (p<0.01). Health-related stigma was directly and indirectly associated with lower functioning through depressed mood. Fifty-two percent of the variance in functioning was explained by the final model in the young adults with narcolepsy. Health-related stigma in young adults with narcolepsy is at a level consistent with other chronic medical illnesses. Health-related stigma may be an important determinant of functioning in young adults with narcolepsy. Future work is indicated toward further characterizing stigma and developing interventions that address various domains of stigma in people with narcolepsy. PMID:25898361

  7. Genetic association, seasonal infections and autoimmune basis of narcolepsy

    PubMed Central

    Singh, Abinav Kumar; Mahlios, Josh; Mignot, Emmanuel

    2014-01-01

    In recent years, a growing number of potential autoimmune disorders affecting neurons in the central nervous system have been identified, including narcolepsy. Narcolepsy is a lifelong sleep disorder characterized by excessive daytime sleepiness with irresistible sleep attacks, cataplexy (sudden bilateral loss of muscle tone), hypnagogic hallucinations, and abnormalities of Rapid Eye Movement sleep. Narcolepsy is generally a sporadic disorder and is caused by the loss of hypocretin (orexin)-producing neurons in the hypothalamus region of the brain. Studies have established that more than 90% of patients have a genetic association with HLA DQB1*06:02. Genome-wide association analysis shows a strong association between narcolepsy and polymorphisms in the TCRα locus and weaker associations within TNFSF4 (also called OX40L), Cathepsin H and the P2RY11-DNMT1 (purinergic receptor subtype P2Y11 to DNMT1, a DNA methytransferase) loci, suggesting an autoimmune basis. Mutations in DNMT1 have also been reported to cause narcolepsy in association with a complex neurological syndrome, suggesting the importance of DNA methylation in the pathology. More recently, narcolepsy was identified in association with seasonal streptococcus, H1N1 infections and following AS03-adjuvanted pH1N1 influenza vaccination in Northern Europe. Potential immunological pathways responsible for the loss of hypocretin producing neurons in these cases may be molecular mimicry or bystander activation. Specific autoantibodies or T cells cross-reactive with hypocretin neurons have not yet been identified, however, thus narcolepsy does not meet Witebsky’s criteria for an autoimmune disease. As the brain is not an easily accessible organ, mechanisms of disease initiation and progression remain a challenge to researchers. PMID:23497937

  8. Migraine and risk of narcolepsy in children: A nationwide longitudinal study.

    PubMed

    Yang, Chun-Pai; Hsieh, Meng-Lun; Chiang, Jen-Huai; Chang, Hsing-Yi; Hsieh, Vivian Chia-Rong

    2017-01-01

    The association between migraine and narcolepsy remains controversial. We aim to investigate whether migraine is associated with an increased risk of developing narcolepsy in children. In this longitudinal study, nationwide medical-claims data of pediatric patients (0-17y) with migraine are identified using the National Health Insurance Research Database (NHIRD) between 1997 and 2010 in Taiwan. Two cohorts are selected: migraine cases (n = 8,923) and propensity score-matched non-migraine controls (n = 35,692). Children with previous history of narcolepsy or headache before the index date are excluded. Cohorts are followed until the end of 2012, their withdrawal from the NHI program, or incidence of narcolepsy (ICD-9-CM: 347). Cox proportional hazards regression models are used to estimate hazard ratios (HRs) and 95% confidence intervals of developing narcolepsy in children with migraine compared to their non-migraine controls. A total of 13 incident cases with narcolepsy are observed during follow-up, with incidence rates of 0.1915 and 0.0278 per 1,000 person-years in migraine and non-migraine children, respectively. After a mean follow-up period of 4.68 and 5.04 years in the case and control cohort, respectively, the former exhibited a greater risk of developing narcolepsy compared to the latter (adjusted hazard ratio (aHR) = 5.30, 95% confidence interval (CI): 1.61, 17.4; p = 0.006). This finding persisted after controlling for potential confounders like baseline comorbidities and concurrent medication uptake, and in our analyses with migraine subtypes. Migraine is an independent risk factor for narcolepsy development in children. Further studies are needed to validate our findings and to explore the exact pathophysiological mechanisms linking migraine and narcolepsy.

  9. Evaluation of polygenic risks for narcolepsy and essential hypersomnia.

    PubMed

    Yamasaki, Maria; Miyagawa, Taku; Toyoda, Hiromi; Khor, Seik-Soon; Liu, Xiaoxi; Kuwabara, Hitoshi; Kano, Yukiko; Shimada, Takafumi; Sugiyama, Toshiro; Nishida, Hisami; Sugaya, Nagisa; Tochigi, Mamoru; Otowa, Takeshi; Okazaki, Yuji; Kaiya, Hisanobu; Kawamura, Yoshiya; Miyashita, Akinori; Kuwano, Ryozo; Kasai, Kiyoto; Tanii, Hisashi; Sasaki, Tsukasa; Honda, Yutaka; Honda, Makoto; Tokunaga, Katsushi

    2016-10-01

    In humans, narcolepsy is a sleep disorder that is characterized by sleepiness, cataplexy and rapid eye movement (REM) sleep abnormalities. Essential hypersomnia (EHS) is another type of sleep disorder that is characterized by excessive daytime sleepiness without cataplexy. A human leukocyte antigen (HLA) class II allele, HLA-DQB1*06:02, is a major genetic factor for narcolepsy. Almost all narcoleptic patients are carriers of this HLA allele, while 30-50% of EHS patients and 12% of all healthy individuals in Japan carry this allele. The pathogenesis of narcolepsy and EHS is thought to be partially shared. To evaluate the contribution of common single-nucleotide polymorphisms (SNPs) to narcolepsy onset and to assess the common genetic background of narcolepsy and EHS, we conducted a polygenic analysis that included 393 narcoleptic patients, 38 EHS patients with HLA-DQB1*06:02, 119 EHS patients without HLA-DQB1*06:02 and 1582 healthy individuals. We also included 376 individuals with panic disorder and 213 individuals with autism to confirm whether the results were biased. Polygenic risks in narcolepsy were estimated to explain 58.1% (P HLA-DQB1*06:02 =2.30 × 10 -48 , P whole genome without HLA-DQB1*06:02 =6.73 × 10 -2 ) including HLA-DQB1*06:02 effects and 1.3% (P whole genome without HLA-DQB1*06:02 =2.43 × 10 -2 ) excluding HLA-DQB1*06:02 effects. The results also indicated that small-effect SNPs contributed to the development of narcolepsy. Reported susceptibility SNPs for narcolepsy in the Japanese population, CPT1B (carnitine palmitoyltransferase 1B), TRA@ (T-cell receptor alpha) and P2RY11 (purinergic receptor P2Y, G-protein coupled, 11), were found to explain 0.8% of narcolepsy onset (P whole genome without HLA-DQB1*06:02 =9.74 × 10 -2 ). EHS patients with HLA-DQB1*06:02 were estimated to have higher shared genetic background to narcoleptic patients than EHS patients without HLA-DQB1*06:02 even when the effects of HLA-DQB1*06:02 were excluded (EHS with HLA

  10. Cognitive behavioral treatment for narcolepsy: can it complement pharmacotherapy?

    PubMed Central

    Marín Agudelo, Hernán Andrés; Jiménez Correa, Ulises; Carlos Sierra, Juan; Pandi-Perumal, Seithikurippu R.; Schenck, Carlos H.

    2014-01-01

    Sleep medicine in general and psychology in particular have recently developed cognitive behavioral treatment for narcolepsy (CBT-N). Despite a growing interest in this topic, most studies since 2007 have reviewed CBT applications for other sleep disorders. Currently, 6 reviews have been published on narcolepsy, with an expert consensus being reached that CBT represented an important adjunctive treatment for the disease. The current paper reviews the need for CBT applications for narcolepsy by generalizing the application of multicomponent treatments and performing studies that extrapolate the results obtained from multicenter studies. Nineteen studies were found in which the need-for-treatment guidelines identified the use of CBT for narcolepsy. Three additional studies were identified that evaluated the effectiveness of cognitive behavioral measures and multicomponent treatments for which treatment protocols have been proposed. PMID:26483898

  11. Narcolepsy in a three-year-old girl: A case report.

    PubMed

    Park, Eu Gene; Lee, Jiwon; Joo, Eun Yeon; Lee, Munhyang; Lee, Jeehun

    2016-01-01

    Narcolepsy is characterized by excessive daytime somnolence associated with sleep paralysis, hallucinations when falling asleep or awakening, and cataplexy. Early recognition of pediatric narcolepsy is essential for growth and development. We experienced a case of narcolepsy in a three-year-old girl. The patient underwent brain MRI and 24h video-electroencephalogram (EEG) monitoring. Polysomnography (PSG) with multiple sleep latency test (MSLT) and human leukocyte antigen (HLA) DQ typing was performed. The brain MRI was normal. 24h video-EEG monitoring revealed no abnormal slow or epileptiform discharge on interictal EEG, and no EEG change during tongue thrusting, dropping head with laughter, or flopping down, which was consistent with cataplexy associated with narcolepsy. A mean sleep latency of 2.5 min and four episodes of sleep-onset REM periods in five naps were observed in PSG with MSLT. She was positive in HLA-DQB1*0602. Based on these findings, she was diagnosed as narcoleptic with cataplexy. The history, combined with PSG and MSLT, was helpful in the diagnosis of narcolepsy. We report a case of early-onset narcolepsy presenting with excessive sleepiness and cataplexy. Copyright © 2015 The Japanese Society of Child Neurology. Published by Elsevier B.V. All rights reserved.

  12. Dual cases of type 1 narcolepsy with schizophrenia and other psychotic disorders.

    PubMed

    Canellas, Francesca; Lin, Ling; Julià, Maria Rosa; Clemente, Antonio; Vives-Bauza, Cristofol; Ollila, Hanna M; Hong, Seung Chul; Arboleya, Susana M; Einen, Mali A; Faraco, Juliette; Fernandez-Vina, Marcelo; Mignot, Emmanuel

    2014-09-15

    Cases of narcolepsy in association with psychotic features have been reported but never fully characterized. These patients present diagnostic and treatment challenges and may shed new light on immune associations in schizophrenia. Our case series was gathered at two narcolepsy specialty centers over a 9-year period. A questionnaire was created to improve diagnosis of schizophrenia or another psychotic disorder in patients with narcolepsy. Pathophysiological investigations included full HLA Class I and II typing, testing for known systemic and intracellular/synaptic neuronal antibodies, recently described neuronal surface antibodies, and immunocytochemistry on brain sections to detect new antigens. Ten cases were identified, one with schizoaffective disorder, one with delusional disorder, two with schizophreniform disorder, and 6 with schizophrenia. In all cases, narcolepsy manifested first in childhood or adolescence, followed by psychotic symptoms after a variable interval. These patients had auditory hallucinations, which was the most differentiating clinical feature in comparison to narcolepsy patients without psychosis. Narcolepsy therapy may have played a role in triggering psychotic symptoms but these did not reverse with changes in narcolepsy medications. Response to antipsychotic treatment was variable. Pathophysiological studies did not reveal any known autoantibodies or unusual brain immunostaining pattern. No strong HLA association outside of HLA DQB1*06:02 was found, although increased DRB3*03 and DPA1*02:01 was notable. Narcolepsy can occur in association with schizophrenia, with significant diagnostic and therapeutic challenges. Dual cases maybe under diagnosed, as onset is unusually early, often in childhood. Narcolepsy and psychosis may share an autoimmune pathology; thus, further investigations in larger samples are warranted. © 2014 American Academy of Sleep Medicine.

  13. Orexin neurons suppress narcolepsy via 2 distinct efferent pathways

    PubMed Central

    Hasegawa, Emi; Yanagisawa, Masashi; Sakurai, Takeshi; Mieda, Michihiro

    2014-01-01

    The loss of orexin neurons in humans is associated with the sleep disorder narcolepsy, which is characterized by excessive daytime sleepiness and cataplexy. Mice lacking orexin peptides, orexin neurons, or orexin receptors recapitulate human narcolepsy phenotypes, further highlighting a critical role for orexin signaling in the maintenance of wakefulness. Despite the known role of orexin neurons in narcolepsy, the precise neural mechanisms downstream of these neurons remain unknown. We found that targeted restoration of orexin receptor expression in the dorsal raphe (DR) and in the locus coeruleus (LC) of mice lacking orexin receptors inhibited cataplexy-like episodes and pathological fragmentation of wakefulness (i.e., sleepiness), respectively. The suppression of cataplexy-like episodes correlated with the number of serotonergic neurons restored with orexin receptor expression in the DR, while the consolidation of fragmented wakefulness correlated with the number of noradrenergic neurons restored in the LC. Furthermore, pharmacogenetic activation of these neurons using designer receptor exclusively activated by designer drug (DREADD) technology ameliorated narcolepsy in mice lacking orexin neurons. These results suggest that DR serotonergic and LC noradrenergic neurons play differential roles in orexin neuron–dependent regulation of sleep/wakefulness and highlight a pharmacogenetic approach for the amelioration of narcolepsy. PMID:24382351

  14. Psychosis in Patients with Narcolepsy as an Adverse Effect of Sodium Oxybate

    PubMed Central

    Sarkanen, Tomi; Niemelä, Valter; Landtblom, Anne-Marie; Partinen, Markku

    2014-01-01

    Aim: Hypnagogic and hypnopompic hallucinations are characteristic symptoms of narcolepsy, as are excessive daytime sleepiness, cataplexy, and sleep paralysis. Narcolepsy patients may also experience daytime hallucinations unrelated to sleep–wake transitions. The effect of medication on hallucinations is of interest since treatment of narcolepsy may provoke psychotic symptoms. We aim to analyze the relation between sodium oxybate (SXB) treatment and psychotic symptoms in narcolepsy patients. Furthermore, we analyze the characteristics of hallucinations to determine their nature as mainly psychotic or hypnagogic and raise a discussion about whether SXB causes psychosis or if psychosis occurs as an endogenous complication in narcolepsy. Method: We present altogether four patients with narcolepsy who experienced psychotic symptoms during treatment with SXB. In addition, we searched the literature for descriptions of hallucinations in narcolepsy and similarities and differences with psychotic symptoms in schizophrenia. Results: Three out of four patients had hallucinations typical for psychosis and one had symptoms that resembled aggravated hypnagogic hallucinations. Two patients also had delusional symptoms primarily associated with mental disorders. Tapering down SXB was tried and helped in two out of four cases. Adding antipsychotic treatment (risperidone) alleviated psychotic symptoms in two cases. Conclusion: Psychotic symptoms in narcolepsy may appear during SXB treatment. Hallucinations resemble those seen in schizophrenia; however, the insight that symptoms are delusional is usually preserved. In case of SXB-induced psychotic symptoms or hallucinations, reducing SXB dose or adding antipsychotic medication can be tried. PMID:25191304

  15. Psychosis in patients with narcolepsy as an adverse effect of sodium oxybate.

    PubMed

    Sarkanen, Tomi; Niemelä, Valter; Landtblom, Anne-Marie; Partinen, Markku

    2014-01-01

    Hypnagogic and hypnopompic hallucinations are characteristic symptoms of narcolepsy, as are excessive daytime sleepiness, cataplexy, and sleep paralysis. Narcolepsy patients may also experience daytime hallucinations unrelated to sleep-wake transitions. The effect of medication on hallucinations is of interest since treatment of narcolepsy may provoke psychotic symptoms. We aim to analyze the relation between sodium oxybate (SXB) treatment and psychotic symptoms in narcolepsy patients. Furthermore, we analyze the characteristics of hallucinations to determine their nature as mainly psychotic or hypnagogic and raise a discussion about whether SXB causes psychosis or if psychosis occurs as an endogenous complication in narcolepsy. We present altogether four patients with narcolepsy who experienced psychotic symptoms during treatment with SXB. In addition, we searched the literature for descriptions of hallucinations in narcolepsy and similarities and differences with psychotic symptoms in schizophrenia. Three out of four patients had hallucinations typical for psychosis and one had symptoms that resembled aggravated hypnagogic hallucinations. Two patients also had delusional symptoms primarily associated with mental disorders. Tapering down SXB was tried and helped in two out of four cases. Adding antipsychotic treatment (risperidone) alleviated psychotic symptoms in two cases. Psychotic symptoms in narcolepsy may appear during SXB treatment. Hallucinations resemble those seen in schizophrenia; however, the insight that symptoms are delusional is usually preserved. In case of SXB-induced psychotic symptoms or hallucinations, reducing SXB dose or adding antipsychotic medication can be tried.

  16. Frequencies and Associations of Narcolepsy-Related Symptoms: A Cross-Sectional Study

    PubMed Central

    Kim, Lenise Jihe; Coelho, Fernando Morgadinho; Hirotsu, Camila; Araujo, Paula; Bittencourt, Lia; Tufik, Sergio; Andersen, Monica Levy

    2015-01-01

    Objectives: Narcolepsy is a disabling disease with a delayed diagnosis. At least 3 years before the disorder identification, several comorbidities can be observed in patients with narcolepsy. The early recognition of narcolepsy symptoms may improve long-term prognosis of the patients. Thus, we aimed to investigate the prevalence of the symptoms associated with narcolepsy and its social and psychological association in a sample of Sao Paulo city inhabitants. Methods: We performed a cross-sectional evaluation with 1,008 individuals from the Sao Paulo Epidemiologic Sleep Study (EPISONO). Excessive daytime sleepiness (EDS) was assessed by the Epworth Sleepiness Scale. Volunteers were also asked about the occurrence of cataplectic-like, hypnagogic or hypnopompic hallucinations, and sleep paralysis symptoms. The participants underwent a full-night polysomnography and completed questionnaires about psychological, demographic, and quality of life parameters. Results: We observed a prevalence of 39.2% of EDS, 15.0% of cataplectic-like symptom, 9.2% of hypnagogic or hypnopompic hallucinations, and 14.9% of sleep paralysis in Sao Paulo city inhabitants. A frequency of 6.9% was observed when EDS and cataplectic-like symptoms were grouped. The other associations were EDS + hallucinations (4.7%) and EDS + sleep paralysis (7.5%). Symptomatic participants were predominantly women and younger compared with patients without any narcolepsy symptom (n = 451). Narcolepsy symptomatology was also associated with a poor quality of life and symptoms of depression, anxiety, and fatigue. Conclusions: Narcolepsy-related symptoms are associated with poor quality of life and worse psychological parameters. Citation: Kim LJ, Coelho FM, Hirotsu C, Araujo P, Bittencourt L, Tufik S, Andersen ML. Frequencies and associations of narcolepsy-related symptoms: a cross-sectional study. J Clin Sleep Med 2015;11(12):1377–1384. PMID:26235160

  17. Treatment of Narcolepsy and other Hypersomnias of Central Origin

    PubMed Central

    Wise, Merrill S.; Arand, Donna L.; Auger, R. Robert; Brooks, Stephen N.; Watson, Nathaniel F.

    2007-01-01

    Objective: The purpose of this paper is to summarize current knowledge about treatment of narcolepsy and other hypersomnias of central origin. Methods: The task force performed a systematic and comprehensive review of the relevant literature and graded the evidence using the Oxford grading system. This paper discusses the strengths and limitations of the available evidence regarding treatment of these conditions, and summarizes key information about safety of these medications. Our findings provide the foundation for development of evidence-based practice parameters on this topic by the Standards of Practice Committee of the American Academy of Sleep Medicine. Results: The majority of recent papers in this field provide information about use of modafinil or sodium oxybate for treatment of sleepiness associated with narcolepsy. Several large randomized, placebo-controlled studies indicate that modafinil and sodium oxybate are effective for treatment of hypersomnia due to narcolepsy. We identified no studies that report direct comparison of these newer medications versus traditional stimulants, or that indicate what proportion of patients treated initially with these medications require transition to traditional stimulants or to combination therapy to achieve adequate alertness. As with the traditional stimulants, modafinil and sodium oxybate provide, at best, only moderate improvement in alertness rather than full restoration of alertness in patients with narcolepsy. Several large randomized placebo-controlled studies demonstrate that sodium oxybate is effective for treatment of cataplexy associated with narcolepsy, and earlier studies provide limited data to support the effectiveness of fluoxetine and tricyclic antidepressants for treatment of cataplexy. Our findings indicate that very few reports provide information regarding treatment of special populations such as children, older adults, and pregnant or breastfeeding women. The available literature provides a

  18. Juvenile myoclonic epilepsy and narcolepsy: A series of three cases.

    PubMed

    Joshi, Puja Aggarwal; Poduri, Annapurna; Kothare, Sanjeev V

    2015-10-01

    This paper sets out to demonstrate the coexistence of juvenile myoclonic epilepsy (JME) and narcolepsy that raises the possibility of a shared genetic predisposition to both conditions. The electronic medical records (EMRs) were searched for narcolepsy and JME over 10years. We identified three young adult women diagnosed with JME in their teenage years, with myoclonic, generalized tonic-clonic, and absence seizure semiologies, along with psychiatric comorbidity, well managed on lamotrigine and/or levetiracetam. Our patients were also found to have disturbed sleep preceding the diagnosis of JME by many years, including excessive daytime sleepiness (EDS), fragmented nocturnal sleep, hypnagogic vivid hallucinations, and REM behavior disorder along with daytime cataplexy. They were ultimately diagnosed with coexisting narcolepsy, confirmed by sleep studies and multiple sleep latency testing, along with positive genetic testing for HLA-DQB1*0602 in all three patients. Stimulants, selective serotonin receptor inhibitors, and/or sodium oxybate were used to successfully treat their narcolepsy. The coexistence of JME and narcolepsy has not been well recognized and may be clinically relevant. In addition, it raises the possibility of a shared genetic predisposition to both conditions. Copyright © 2015 Elsevier Inc. All rights reserved.

  19. Listening to the Patient Voice in Narcolepsy: Diagnostic Delay, Disease Burden, and Treatment Efficacy.

    PubMed

    Maski, Kiran; Steinhart, Erin; Williams, David; Scammell, Thomas; Flygare, Julie; McCleary, Kimberly; Gow, Monica

    2017-03-15

    Describe common symptoms, comorbidities, functional limitations, and treatment responsiveness among patients with narcolepsy. Investigate the effect of pediatric onset of narcolepsy symptoms on time to diagnosis of narcolepsy and presence of comorbid depression. Cross-sectional survey of 1,699 people in the United States with self-reported diagnosis of narcolepsy. We utilized mixed-methods data analyses to report study findings. Most participants reported receiving a diagnosis of narcolepsy more than 1 y after symptom onset. We found that the strongest predictor of this delayed diagnosis was pediatric onset of symptoms (odds ratio = 2.4, p < 0.0005). Depression was the most common comorbidity but we detected no association with pediatric onset of narcolepsy symptoms. Overall, participants reported that fatigue and cognitive difficulties were their most burdensome symptoms in addition to sleepiness and cataplexy. The majority of participants reported residual daytime fatigue and/or sleepiness despite treatment. Most participants reported they could not perform at work or school as well as they would like because of narcolepsy symptoms. This study provides unique insight into the narcolepsy disease experience. The study quantifies the problem of diagnostic delay for narcolepsy patients in the United States and highlights that symptoms are more likely to be missed if they develop before 18 y of age. These results suggest that narcolepsy awareness efforts should be aimed at parents, pediatric health care providers, school professionals, and children/adolescents themselves. Disease burden is high because of problems with fatigue, cognition, and persistence of residual symptoms despite treatment. © 2017 American Academy of Sleep Medicine

  20. Listening to the Patient Voice in Narcolepsy: Diagnostic Delay, Disease Burden, and Treatment Efficacy

    PubMed Central

    Maski, Kiran; Steinhart, Erin; Williams, David; Scammell, Thomas; Flygare, Julie; McCleary, Kimberly; Gow, Monica

    2017-01-01

    Study Objectives: Describe common symptoms, comorbidities, functional limitations, and treatment responsiveness among patients with narcolepsy. Investigate the effect of pediatric onset of narcolepsy symptoms on time to diagnosis of narcolepsy and presence of comorbid depression. Methods: Cross-sectional survey of 1,699 people in the United States with self-reported diagnosis of narcolepsy. We utilized mixed-methods data analyses to report study findings. Results: Most participants reported receiving a diagnosis of narcolepsy more than 1 y after symptom onset. We found that the strongest predictor of this delayed diagnosis was pediatric onset of symptoms (odds ratio = 2.4, p < 0.0005). Depression was the most common comorbidity but we detected no association with pediatric onset of narcolepsy symptoms. Overall, participants reported that fatigue and cognitive difficulties were their most burdensome symptoms in addition to sleepiness and cataplexy. The majority of participants reported residual daytime fatigue and/or sleepiness despite treatment. Most participants reported they could not perform at work or school as well as they would like because of narcolepsy symptoms. Conclusions: This study provides unique insight into the narcolepsy disease experience. The study quantifies the problem of diagnostic delay for narcolepsy patients in the United States and highlights that symptoms are more likely to be missed if they develop before 18 y of age. These results suggest that narcolepsy awareness efforts should be aimed at parents, pediatric health care providers, school professionals, and children/adolescents themselves. Disease burden is high because of problems with fatigue, cognition, and persistence of residual symptoms despite treatment. Citation: Maski K, Steinhart E, Williams D, Scammell T, Flygare J, McCleary K, Gow M. Listening to the patient voice in narcolepsy: diagnostic delay, disease burden and treatment efficacy. J Clin Sleep Med. 2017;13(3):419–425

  1. Quality Measures for the Care of Patients with Narcolepsy

    PubMed Central

    Krahn, Lois E.; Hershner, Shelley; Loeding, Lauren D.; Maski, Kiran P.; Rifkin, Daniel I.; Selim, Bernardo; Watson, Nathaniel F.

    2015-01-01

    The American Academy of Sleep Medicine (AASM) commissioned a Workgroup to develop quality measures for the care of patients with narcolepsy. Following a comprehensive literature search, 306 publications were found addressing quality care or measures. Strength of association was graded between proposed process measures and desired outcomes. Following the AASM process for quality measure development, we identified three outcomes (including one outcome measure) and seven process measures. The first desired outcome was to reduce excessive daytime sleepiness by employing two process measures: quantifying sleepiness and initiating treatment. The second outcome was to improve the accuracy of diagnosis by employing the two process measures: completing both a comprehensive sleep history and an objective sleep assessment. The third outcome was to reduce adverse events through three steps: ensuring treatment follow-up, documenting medical comorbidities, and documenting safety measures counseling. All narcolepsy measures described in this report were developed by the Narcolepsy Quality Measures Work-group and approved by the AASM Quality Measures Task Force and the AASM Board of Directors. The AASM recommends the use of these measures as part of quality improvement programs that will enhance the ability to improve care for patients with narcolepsy. Citation: Krahn LE, Hershner S, Loeding LD, Maski KP, Rifkin DI, Selim B, Watson NF. Quality measures for the care of patients with narcolepsy. J Clin Sleep Med 2015;11(3):335–355. PMID:25700880

  2. Frequencies and Associations of Narcolepsy-Related Symptoms: A Cross-Sectional Study.

    PubMed

    Kim, Lenise Jihe; Coelho, Fernando Morgadinho; Hirotsu, Camila; Araujo, Paula; Bittencourt, Lia; Tufik, Sergio; Andersen, Monica Levy

    2015-12-15

    Narcolepsy is a disabling disease with a delayed diagnosis. At least 3 years before the disorder identification, several comorbidities can be observed in patients with narcolepsy. The early recognition of narcolepsy symptoms may improve long-term prognosis of the patients. Thus, we aimed to investigate the prevalence of the symptoms associated with narcolepsy and its social and psychological association in a sample of Sao Paulo city inhabitants. We performed a cross-sectional evaluation with 1,008 individuals from the Sao Paulo Epidemiologic Sleep Study (EPISONO). Excessive daytime sleepiness (EDS) was assessed by the Epworth Sleepiness Scale. Volunteers were also asked about the occurrence of cataplectic-like, hypnagogic or hypnopompic hallucinations, and sleep paralysis symptoms. The participants underwent a full-night polysomnography and completed questionnaires about psychological, demographic, and quality of life parameters. We observed a prevalence of 39.2% of EDS, 15.0% of cataplectic-like symptom, 9.2% of hypnagogic or hypnopompic hallucinations, and 14.9% of sleep paralysis in Sao Paulo city inhabitants. A frequency of 6.9% was observed when EDS and cataplectic-like symptoms were grouped. The other associations were EDS + hallucinations (4.7%) and EDS + sleep paralysis (7.5%). Symptomatic participants were predominantly women and younger compared with patients without any narcolepsy symptom (n = 451). Narcolepsy symptomatology was also associated with a poor quality of life and symptoms of depression, anxiety, and fatigue. Narcolepsy-related symptoms are associated with poor quality of life and worse psychological parameters. © 2015 American Academy of Sleep Medicine.

  3. Narcolepsy patients' blood-based miRNA expression profiling: miRNA expression differences with Pandemrix vaccination.

    PubMed

    Mosakhani, N; Sarhadi, V; Panula, P; Partinen, M; Knuutila, S

    2017-11-01

    Narcolepsy is a neurological sleep disorder characterized by excessive daytime sleepiness and nighttime sleep disturbance. Among children and adolescents vaccinated with Pandemrix vaccine in Finland and Sweden, the number of narcolepsy cases increased. Our aim was to identify miRNAs involved in narcolepsy and their association with Pandemrix vaccination. We performed global miRNA proofing by miRNA microarrays followed by RT-PCR verification on 20 narcolepsy patients (Pandemrix-associated and Pandemrix-non-associated) and 17 controls (vaccinated and non-vaccinated). Between all narcolepsy patients and controls, 11 miRNAs were differentially expressed; 17 miRNAs showed significantly differential expression between Pandemrix-non-associated narcolepsy patients and non-vaccinated healthy controls. MiR-188-5p and miR-4499 were over-expressed in narcolepsy patients vs healthy controls. Two miRNAs, miR-1470 and miR-4455, were under-expressed in Pandemrix-associated narcolepsy patients vs Pandemrix-non-associated narcolepsy patients. We identified miRNA expression patterns in narcolepsy patients that linked them to mRNA targets known to be involved in brain-related pathways or brain disorders. © 2017 John Wiley & Sons A/S. Published by John Wiley & Sons Ltd.

  4. Evidence for Metabolic Hypothalamo-Amygdala Dysfunction in Narcolepsy

    PubMed Central

    Poryazova, Rositsa; Schnepf, Betina; Werth, Esther; Khatami, Ramin; Dydak, Ulrike; Meier, Dieter; Boesiger, Peter; Bassetti, Claudio L.

    2009-01-01

    Study Objectives: Proton resonance spectroscopy (1H-MRS) allows noninvasive chemical tissue analysis in the living brain. As neuronal loss and gliosis have been described in narcolepsy, metabolites of primary interest are N-acetylaspartate (NAA), a marker of neuronal integrity and myo-Inositol (mI), a glial marker and second messenger involved in the regulation of intracellular calcium. One 1H-MRS study in narcolepsy found no metabolic changes in the pontomedullary junction. Another study showed a reduction in NAA/creatine-phosphocreatine (Cr) in the hypothalamus of narcolepsy patients with cataplexy. We aimed to test for metabolic changes in specific brain areas, “regions of interest,” thought to be involved in emotional processing, sleep regulation and pathophysiology of narcolepsy: hypothalamus, pontomesencephalic junction and both amygdalae. Design: We performed 1H-MRS using a 3T Philips Achieva whole body MR scanner. Single-voxel proton MR spectra were acquired and quantified with LCModel to determine metabolite concentration ratios. Setting: The participants in the study were recruited at the outpatient clinic for sleep medicine, Department of Neurology and magnetic resonance spectroscopy was performed at the MRI facility, University Hospital Zurich. Participants: 1H-MRS was performed in fourteen narcolepsy patients with cataplexy, CSF hypocretin deficiency (10/10) and HLA-DQB1*0602 positivity (14/14) and 14 age, gender and body mass index matched controls. Patients were treatment naïve or off therapy for at least 14 days before scanning. Measurements and Results: No differences were observed in the regions of interest for (total NAA)/Cr ratios. Myo-Inositol (mI)/Cr was significantly lower in the right amygdala of the patients, compared to controls (P < 0.042). Significant negative correlations only in the patients group were found between (total NAA)/Cr in hypothalamus and mI/Cr in the right amygdala (r = −0.89, P < 0.001), between mI/Cr in

  5. Comorbidity of narcolepsy and depressive disorders: a nationwide population-based study in Taiwan.

    PubMed

    Lee, Min-Jing; Lee, Sheng-Yu; Yuan, Shin-Sheng; Yang, Chun-Ju; Yang, Kang-Chung; Lee, Tung-Liang; Sun, Chi-Chin; Shyu, Yu-Chiau; Wang, Liang-Jen

    2017-11-01

    Narcolepsy is a chronic sleep disorder that is likely to have neuropsychiatric comorbidities. Depression is a serious mood disorder that affects individuals' daily activities and functions. The current study aimed to investigate the relationship between narcolepsy and depressive disorders. The study consisted of patients diagnosed with narcolepsy between January 2002, and December 2011 (n = 258), and age-matched and gender-matched controls (n = 2580) from Taiwan's National Health Insurance database. Both the patients and the controls were monitored through December 31, 2011, to identify the occurrence of a depressive disorder. A multivariate logistic regression model was used to assess the narcolepsy's potential influence on the comorbidity of a depressive disorder. During the study period, 32.7%, 24.8%, and 10.9% of the narcoleptic patients were comorbid with any depressive disorder, dysthymic disorder, and major depressive disorder, respectively. When compared to the control subjects, the patients with narcolepsy were at greater risks of having any depressive disorder (aOR 6.77; 95% CI 4.90-9.37), dysthymic disorder (aOR 6.62; 95% CI 4.61-9.57), and major depressive disorder (aOR 6.83; 95% CI 4.06-11.48). Of the narcoleptic patients that were comorbid with depression, >50% had been diagnosed with depression prior to being diagnosed with narcolepsy. This nationwide data study revealed that narcolepsy and depression commonly co-occurred. Since some symptoms of narcolepsy overlapped with those of depressive disorders, the findings serve as a reminder that clinicians must pay attention to the comorbidity of narcolepsy and depression. Copyright © 2017 Elsevier B.V. All rights reserved.

  6. Antibodies Against Hypocretin Receptor 2 Are Rare in Narcolepsy.

    PubMed

    Giannoccaro, Maria Pia; Waters, Patrick; Pizza, Fabio; Liguori, Rocco; Plazzi, Giuseppe; Vincent, Angela

    2017-02-01

    Recently, antibodies to the hypocretin receptor 2 (HCRTR2-Abs) were reported in a high proportion of narcolepsy patients who developed the disease following Pandemrix® vaccination. We tested a group of narcolepsy patients for the HCRTR2-Abs using a newly established cell-based assay. Sera from 50 narcolepsy type 1 (NT1) and 11 narcolepsy type 2 (NT2) patients, 22 patients with other sleep disorders, 15 healthy controls, and 93 disease controls were studied. Cerebrospinal fluid (CSFs) from three narcoleptic patients were subsequently included. Human embryonic kidney cells were transiently transfected with human HCRTR2, incubated with patients' sera for 1 hr at 1:20 dilution and then fixed. Binding of antibodies was detected by fluorescently labeled secondary antibodies to human immunoglobulin G (IgG) and the different IgG subclasses. A nonlinear visual scoring system was used from 0 to 4; samples scoring ≥1 were considered positive. Only 3 (5%) of 61 patients showed a score ≥1, one with IgG1- and two with IgG3-antibodies, but titers were low (1:40-1:100). CSFs from these patients were negative. The three positive patients included one NT1 case with associated psychotic features, one NT2 patient, and an NT1 patient with normal hypocretin CSF levels. Low levels of IgG1 or IgG3 antibodies against HCRTR2 were found in 3 of 61 patients with narcolepsy, although only 1 presented with full-blown NT1. HCRTR2-Abs are not common in narcolepsy unrelated to vaccination. © Sleep Research Society 2016. Published by Oxford University Press on behalf of the Sleep Research Society. All rights reserved. For permissions, please e-mail journals.permissions@oup.com.

  7. Complex movement disorders at disease onset in childhood narcolepsy with cataplexy

    PubMed Central

    Pizza, Fabio; Palaia, Vincenzo; Franceschini, Christian; Poli, Francesca; Moghadam, Keivan K.; Cortelli, Pietro; Nobili, Lino; Bruni, Oliviero; Dauvilliers, Yves; Lin, Ling; Edwards, Mark J.; Mignot, Emmanuel; Bhatia, Kailash P.

    2011-01-01

    Narcolepsy with cataplexy is characterized by daytime sleepiness, cataplexy (sudden loss of bilateral muscle tone triggered by emotions), sleep paralysis, hypnagogic hallucinations and disturbed nocturnal sleep. Narcolepsy with cataplexy is most often associated with human leucocyte antigen-DQB1*0602 and is caused by the loss of hypocretin-producing neurons in the hypothalamus of likely autoimmune aetiology. Noting that children with narcolepsy often display complex abnormal motor behaviours close to disease onset that do not meet the classical definition of cataplexy, we systematically analysed motor features in 39 children with narcolepsy with cataplexy in comparison with 25 age- and sex-matched healthy controls. We found that patients with narcolepsy with cataplexy displayed a complex array of ‘negative’ (hypotonia) and ‘active’ (ranging from perioral movements to dyskinetic–dystonic movements or stereotypies) motor disturbances. ‘Active’ and ‘negative’ motor scores correlated positively with the presence of hypotonic features at neurological examination and negatively with disease duration, whereas ‘negative’ motor scores also correlated negatively with age at disease onset. These observations suggest that paediatric narcolepsy with cataplexy often co-occurs with a complex movement disorder at disease onset, a phenomenon that may vanish later in the course of the disease. Further studies are warranted to assess clinical course and whether the associated movement disorder is also caused by hypocretin deficiency or by additional neurochemical abnormalities. PMID:21930661

  8. Does Narcolepsy Symptom Severity Vary According to HLA-DQB1*0602 Allele Status?

    PubMed Central

    Watson, Nathaniel F.; Ton, Thanh G.N.; Koepsell, Thomas D.; Gersuk, Vivian H.; Longstreth, W.T.

    2010-01-01

    Study Objectives: To investigate associations between HLA-DQB1*0602 allele status and measures of narcolepsy symptom severity. Design: Cross-sectional study of population-based narcolepsy patients. Setting: King County, Washington. Participants: All prevalent cases (n = 279) of physician-diagnosed narcolepsy ascertained from 2001-2005. Interventions: N/A Measurements: Narcolepsy diagnosis was based on cataplexy status, diagnostic sleep study results, and chart review. The number of HLA-DQB1 alleles was determined from buccal genomic DNA. Symptom severity instruments included the Epworth Sleepiness Scale (ESS), the Ullanlinna Narcolepsy Scale (UNS), age of symptom onset, subjective sleep latency and duration, and various clinical sleep parameters. We used linear regression adjusted for African American race and an extended chi-square test of trend to assess relationships across ordered groups defined by allele number (0, 1, or 2). Results: Narcolepsy patients were 63% female and 82% Caucasian, with a mean age of 47.6 years (SD = 17.1). One hundred forty-one (51%) patients had no DQB1*0602 alleles; 117 (42%) had one; and 21 (7%) had two. In the complete narcolepsy sample after adjustment for African American race, we observed a linear relationship between HLA-DQB1*0602 frequency and sleepiness as defined by the ESS (P < 0.01), narcolepsy severity as defined by UNS (P < 0.001), age of symptom onset (P < 0.05), and sleep latency (P < 0.001). In univariate analyses, HLA-DQB1*0602 frequency was also associated with napping (P < 0.05) and increased car and work accidents or near accidents (both P < 0.01). Habitual sleep duration was not associated with HLA status. These race-adjusted associations remained for the ESS (P < 0.05), UNS (P < 0.01), and sleep latency (P < 0.001) when restricting to narcolepsy with cataplexy. Conclusions: Narcolepsy symptom severity varies in a linear manner according to HLA-DQB1*0602 allele status. These findings support the notion that HLA

  9. Adenotonsillectomy outcomes in children with sleep apnea and narcolepsy.

    PubMed

    Biyani, Sneh; Cunningham, Tina D; Baldassari, Cristina M

    2017-09-01

    To identify improvements in daytime sleepiness following adenotonsillectomy in children with non-severe obstructive sleep apnea and narcolepsy. Case series with chart review over 15 years. Tertiary Children's Hospital. Children between 6 and 17 years of age with narcolepsy that underwent adenotonsillectomy for non-severe obstructive sleep apnea (OSA) were included. Narcolepsy was diagnosed based on clinical assessment and the Multiple Sleep Latency Test (MSLT) results. A standardized instrument, the pediatric Epworth Sleepiness Scale (ESS), was used to assess daytime sleepiness before and after adenotonsillectomy. Nine children with a mean age of 12.1 years were included. The majority of the subjects (78%, n = 7) were African American and six children (66.7%) were obese. Four children (44%) were treated with wake promoting agents during the study. The mean preoperative apnea hypopnea index on polysomnography was 4.89 (SD 1.86), while the mean sleep latency on MSLT was 6.32 min (SD 3.14). The mean preoperative ESS was 16.10 and the postoperative ESS was 10.80 (SD 3.96). There was significant improvement (p = 0.02) in the ESS following adenotonsillectomy with seven children (78%) reporting diminished daytime sleepiness. Children with non-severe OSA and narcolepsy experience significant improvement in daytime sleepiness following adenotonsillectomy. Future studies are needed to determine the incidence and clinical significance of non-severe OSA in children with narcolepsy. Copyright © 2017 Elsevier B.V. All rights reserved.

  10. Altered Sleep Stage Transitions of REM Sleep: A Novel and Stable Biomarker of Narcolepsy

    PubMed Central

    Liu, Yaping; Zhang, Jihui; Lam, Venny; Ho, Crover Kwok Wah; Zhou, Junying; Li, Shirley Xin; Lam, Siu Ping; Yu, Mandy Wai Man; Tang, Xiangdong; Wing, Yun-Kwok

    2015-01-01

    Objectives: To determine the diagnostic values, longitudinal stability, and HLA association of the sleep stage transitions in narcolepsy. Methods: To compare the baseline differences in the sleep stage transition to REM sleep among 35 patients with type 1 narcolepsy, 39 patients with type 2 narcolepsy, 26 unaffected relatives, and 159 non-narcoleptic sleep patient controls, followed by a reassessment at a mean duration of 37.4 months. Results: The highest prevalence of altered transition from stage non-N2/N3 to stage R in multiple sleep latency test (MSLT) and nocturnal polysomnography (NPSG) was found in patients with type 1 narcolepsy (92.0% and 57.1%), followed by patients with type 2 narcolepsy (69.4% and 12.8%), unaffected relatives (46.2% and 0%), and controls (39.3% and 1.3%). Individual sleep variables had varied sensitivity and specificity in diagnosing narcolepsy. By incorporating a combination of sleep variables, the decision tree analysis improved the sensitivity to 94.3% and 82.1% and enhanced specificity to 82.4% and 83% for the diagnosis of type 1 and type 2 narcolepsy, respectively. There was a significant association of DBQ1*0602 with the altered sleep stage transition (OR = 16.0, 95% CI: 1.7–149.8, p = 0.015). The persistence of the altered sleep stage transition in both MSLT and NPSG was high for both type 1 (90.5% and 64.7%) and type 2 narcolepsy (92.3% and 100%), respectively. Conclusions: Altered sleep stage transition is a significant and stable marker of narcolepsy, which suggests a vulnerable wake-sleep dysregulation trait in narcolepsy. Altered sleep stage transition has a significant diagnostic value in the differential diagnosis of hypersomnias, especially when combined with other diagnostic sleep variables in decision tree analysis. Citation: Liu Y, Zhang J, Lam V, Ho CK, Zhou J, Li SX, Lam SP, Yu MW, Tang X, Wing YK. Altered sleep stage transitions of REM sleep: a novel and stable biomarker of narcolepsy. J Clin Sleep Med 2015

  11. Altered Sleep Stage Transitions of REM Sleep: A Novel and Stable Biomarker of Narcolepsy.

    PubMed

    Liu, Yaping; Zhang, Jihui; Lam, Venny; Ho, Crover Kwok Wah; Zhou, Junying; Li, Shirley Xin; Lam, Siu Ping; Yu, Mandy Wai Man; Tang, Xiangdong; Wing, Yun-Kwok

    2015-08-15

    To determine the diagnostic values, longitudinal stability, and HLA association of the sleep stage transitions in narcolepsy. To compare the baseline differences in the sleep stage transition to REM sleep among 35 patients with type 1 narcolepsy, 39 patients with type 2 narcolepsy, 26 unaffected relatives, and 159 non-narcoleptic sleep patient controls, followed by a reassessment at a mean duration of 37.4 months. The highest prevalence of altered transition from stage non-N2/N3 to stage R in multiple sleep latency test (MSLT) and nocturnal polysomnography (NPSG) was found in patients with type 1 narcolepsy (92.0% and 57.1%), followed by patients with type 2 narcolepsy (69.4% and 12.8%), unaffected relatives (46.2% and 0%), and controls (39.3% and 1.3%). Individual sleep variables had varied sensitivity and specificity in diagnosing narcolepsy. By incorporating a combination of sleep variables, the decision tree analysis improved the sensitivity to 94.3% and 82.1% and enhanced specificity to 82.4% and 83% for the diagnosis of type 1 and type 2 narcolepsy, respectively. There was a significant association of DBQ1*0602 with the altered sleep stage transition (OR = 16.0, 95% CI: 1.7-149.8, p = 0.015). The persistence of the altered sleep stage transition in both MSLT and NPSG was high for both type 1 (90.5% and 64.7%) and type 2 narcolepsy (92.3% and 100%), respectively. Altered sleep stage transition is a significant and stable marker of narcolepsy, which suggests a vulnerable wake-sleep dysregulation trait in narcolepsy. Altered sleep stage transition has a significant diagnostic value in the differential diagnosis of hypersomnias, especially when combined with other diagnostic sleep variables in decision tree analysis. © 2015 American Academy of Sleep Medicine.

  12. CSF histamine contents in narcolepsy, idiopathic hypersomnia and obstructive sleep apnea syndrome.

    PubMed

    Kanbayashi, Takashi; Kodama, Tohru; Kondo, Hideaki; Satoh, Shinsuke; Inoue, Yuichi; Chiba, Shigeru; Shimizu, Tetsuo; Nishino, Seiji

    2009-02-01

    To (1) replicate our prior result of low cerebrospinal fluid (CSF) histamine levels in human narcolepsy in a different sample population and to (2) evaluate if histamine contents are altered in other types of hypersomnia with and without hypocretin deficiency. Cross sectional studies. Sixty-seven narcolepsy subjects, 26 idiopathic hypersomnia (IHS) subjects, 16 obstructive sleep apnea syndrome (OSAS) subjects, and 73 neurological controls were included. All patients were Japanese. Diagnoses were made according to ICSD-2. We found significant reductions in CSF histamine levels in hypocretin deficient narcolepsy with cataplexy (mean +/- SEM; 176.0 +/- 25.8 pg/mL), hypocretin non-deficient narcolepsy with cataplexy (97.8 +/- 38.4 pg/mL), hypocretin non-deficient narcolepsy without cataplexy (113.6 +/- 16.4 pg/mL), and idiopathic hypersomnia (161.0 +/- 29.3 pg/ mL); the levels in OSAS (259.3 +/- 46.6 pg/mL) did not statistically differ from those in the controls (333.8 +/- 22.0 pg/mL). Low CSF histamine levels were mostly observed in non-medicated patients; significant reductions in histamine levels were evident in non-medicated patients with hypocretin deficient narcolepsy with cataplexy (112.1 +/- 16.3 pg/ mL) and idiopathic hypersomnia (143.3 +/- 28.8 pg/mL), while the levels in the medicated patients were in the normal range. The study confirmed reduced CSF histamine levels in hypocretin-deficient narcolepsy with cataplexy. Similar degrees of reduction were also observed in hypocretin non-deficient narcolepsy and in idiopathic hypersomnia, while those in OSAS (non central nervous system hypersomnia) were not altered. The decrease in histamine in these subjects were more specifically observed in non-medicated subjects, suggesting CSF histamine is a biomarker reflecting the degree of hypersomnia of central origin.

  13. Body Mass Index-Independent Metabolic Alterations in Narcolepsy with Cataplexy

    PubMed Central

    Poli, Francesca; Plazzi, Giuseppe; Di Dalmazi, Guido; Ribichini, Danilo; Vicennati, Valentina; Pizza, Fabio; Mignot, Emmanuel; Montagna, Pasquale; Pasquali, Renato; Pagotto, Uberto

    2009-01-01

    Study Objectives: To contribute to the anthropometric and metabolic phenotyping of orexin-A–deficient narcoleptic patients, and to explore a possible risk of their developing a metabolic syndrome. Design: We performed a cross-sectional study comparing metabolic alterations in patients with narcolepsy with cataplexy (NC) and patients with idiopathic hypersomnia without long sleep time. Setting: University hospital. Patients: Fourteen patients with narcolepsy with cataplexy and 14 sex and age-matched patients with idiopathic hypersomnia without long sleep time. Interventions: N/A. Measurements and results: Metabolic parameters were evaluated by measuring body mass index (BMI), waist circumference (also with abdominal computed tomography), blood pressure, and daily calorie intake (3-day diary). Chronotypes were assessed through the morningness-eveningness questionnaire. Lumbar puncture for cerebrospinal fluid orexin-A determination and HLA typing were performed. Patients with narcolepsy with cataplexy (all HLA DQB1*0602 positive and with cerebrospinal fluid orexin-A levels < 110 pg/mL) had a higher BMI and BMI-independent metabolic alterations, namely waist circumference, high-density lipoprotein cholesterol, and glucose/insulin ratio (an insulin resistance index), with respect to patients with idiopathic hypersomnia without long sleep time (cerebrospinal fluid orexin-A levels > 300 pg/mL). Despite lower daily food intake, patients with narcolepsy with cataplexy displayed significant alterations in metabolic parameters resulting in a diagnosis of metabolic syndrome in more than half the cases. Conclusions: BMI-independent metabolic alterations and the relative hypophagia of patients with narcolepsy with cataplexy, as compared with patients with idiopathic hypersomnia without long sleep time, suggest that orexin-A influences the etiology of this phenotype. Moreover, considering that these dysmetabolic alterations are present from a young age, a careful metabolic

  14. An Interesting Case of Late Age at Onset of Narcolepsy with Cataplexy

    PubMed Central

    Krishnamurthy, Venkatesh B.; Nallamothu, Vijaya; Singareddy, Ravi

    2014-01-01

    The usual age at onset of narcolepsy with cataplexy is in the second or third decade. In cases with late onset narcolepsy with cataplexy, symptoms are usually mild with relatively less severe daytime sleepiness and less frequent cataplexy. Here we present a case of narcolepsy with cataplexy with onset of symptoms around sixty years of age. This case is unique, with severe daytime sleepiness both by subjective report as well as on objective Multiple Sleep Latency Test and having multiple cataplexy episodes in a day. Citation: Krishnamurthy VB; Nallamothu V; Singareddy R. An interesting case of late age at onset of narcolepsy with cataplexy. J Clin Sleep Med 2014;10(2):203-205. PMID:24533004

  15. Narcolepsy in Southern Chinese patients: clinical characteristics, HLA typing and seasonality of birth.

    PubMed

    Wing, Y K; Chen, L; Fong, S Y Y; Ng, M H L; Ho, C K W; Cheng, S H; Tang, N L S; Li, A M

    2008-11-01

    To report clinical characteristics, human leukocyte antigen (HLA) typing and seasonality of birth of a series of 54 Southern Chinese patients suffering from narcolepsy. All subjects underwent detailed medical and psychiatric interviews and a standardised nocturnal polysomnogram followed by a daytime Multiple Sleep Latency Test. Each subject also completed a set of sleep questionnaires. HLA typing was performed in 91% of subjects. A total of 78% and 22% of patients were diagnosed with suffering from cataplectic and non-cataplectic narcolepsy, respectively. The majority (n = 47, 87%) of patients were referred to our sleep clinic for excessive daytime sleepiness (EDS). The cataplectic narcolepsy differed from non-cataplectic narcolepsy by having more rapid eye movement (REM)-related clinical symptoms (more sleep paralysis and sleep-related hallucination) and sleep disturbances (shorter REM latency), as well as tighter association with HLA DQB1*0602. A bi-modal peak pattern was observed at 11 and 39 years old. A similar bi-modal pattern also occurred for EDS and cataplexy. Excess winter births were observed for this series of patients. 81% of patients with cataplectic narcolepsy were DQB1*0602-positive. There were no differences between early- and late-onset cases in the association with positive DQB1*0602 (71.4% vs 60%). Narcolepsy had prominent pernicious effects on various social, academic, family and mental aspects in our patients. In our Southern Chinese narcolepsy series, bi-modal peak pattern of age of onset, excess winter birth and tight association of HLA DQB1*0602 with cataplectic narcolepsy were found.

  16. Narcolepsy with cataplexy in a child with Charcot-Marie-Tooth disease. Case Report.

    PubMed

    Zheng, Feixia; Wang, Shuang

    2016-09-01

    We report an 8-year-old boy diagnosed with both CMT1 and narcolepsy, which were not reported simultaneously presenting in one person. The boy presented with a history of increased suddenly falling frequency and excessive daytime sleepiness for 3 months. CMT1 was diagnosed by electrophysiology and genetic testing. Narcolepsy had not been diagnosed until the frequently falling caused by sudden and transient episodes of legs weakness triggered by emotion was found. Multiple sleep latency test showed multiple sleep onset REM periods with reduced sleep latency. When CMT1 and narcolepsy were coexist in an individual, the latter might be overlooked. Cataplexy caused by narcolepsy might be disregard as distal muscle weakness of CMT1. The daytime sleepiness might also be ignored. Therefore, we recommend that patients with sleep disorders should be queried about the symptoms of narcolepsy.

  17. Identification of Differentially Expressed Genes in Blood Cells of Narcolepsy Patients

    PubMed Central

    Tanaka, Susumu; Honda, Yutaka; Honda, Makoto

    2007-01-01

    Study Objective: A close association between the human leukocyte antigen (HLA)-DRB1*1501/DQB1*0602 and abnormalities in some inflammatory cytokines have been demonstrated in narcolepsy. Specific alterations in the immune system have been suggested to occur in this disorder. We attempted to identify alterations in gene expression underlying the abnormalities in the blood cells of narcoleptic patients. Designs: Total RNA from 12 narcolepsy-cataplexy patients and from 12 age- and sex-matched healthy controls were pooled. The pooled samples were initially screened for candidate genes for narcolepsy by differential display analysis using annealing control primers (ACP). The second screening of the samples was carried out by semiquantitative PCR using gene-specific primers. Finally, the expression levels of the candidate genes were further confirmed by quantitative real-time PCR using a new set of samples (20 narcolepsy-cataplexy patients and 20 healthy controls). Results: The second screening revealed differential expression of 4 candidate genes. Among them, MX2 was confirmed as a significantly down-regulated gene in the white blood cells of narcoleptic patients by quantitative real-time PCR. Conclusion: We found the MX2 gene to be significantly less expressed in comparison with normal subjects in the white blood cells of narcoleptic patients. This gene is relevant to the immune system. Although differential display analysis using ACP technology has a limitation in that it does not help in determining the functional mechanism underlying sleep/wakefulness dysregulation, it is useful for identifying novel genetic factors related to narcolepsy, such as HLA molecules. Further studies are required to explore the functional relationship between the MX2 gene and narcolepsy pathophysiology. Citation: Tanaka S; Honda Y; Honda M. Identification of differentially expressed genes in blood cells of narcolepsy patients. SLEEP 2007;30(8):974-979. PMID:17702266

  18. Symptomatic narcolepsy, cataplexy and hypersomnia, and their implications in the hypothalamic hypocretin/orexin system.

    PubMed

    Nishino, Seiji; Kanbayashi, Takashi

    2005-08-01

    Human narcolepsy is a chronic sleep disorder affecting 1:2000 individuals. The disease is characterized by excessive daytime sleepiness, cataplexy and other abnormal manifestations of REM sleep, such as sleep paralysis and hypnagogic hallucinations. Recently, it was discovered that the pathophysiology of (idiopathic) narcolepsy-cataplexy is linked to hypocretin ligand deficiency in the brain and cerebrospinal fluid (CSF), as well as the positivity of the human leukocyte antigen (HLA) DR2/DQ6 (DQB1*0602). The symptoms of narcolepsy can also occur during the course of other neurological conditions (i.e. symptomatic narcolepsy). We define symptomatic narcolepsy as those cases that meet the International Sleep Disorders Narcolepsy Criteria, and which are also associated with a significant underlying neurological disorder that accounts for excessive daytime sleepiness (EDS) and temporal associations. To date, we have counted 116 symptomatic cases of narcolepsy reported in literature. As, several authors previously reported, inherited disorders (n=38), tumors (n=33), and head trauma (n=19) are the three most frequent causes for symptomatic narcolepsy. Of the 116 cases, 10 are associated with multiple sclerosis, one case of acute disseminated encephalomyelitis, and relatively rare cases were reported with vascular disorders (n=6), encephalitis (n=4) and degeneration (n=1), and hererodegenerative disorder (three cases in a family). EDS without cataplexy or any REM sleep abnormalities is also often associated with these neurological conditions, and defined as symptomatic cases of EDS. Although it is difficult to rule out the comorbidity of idiopathic narcolepsy in some cases, review of the literature reveals numerous unquestionable cases of symptomatic narcolepsy. These include cases with HLA negative and/or late onset, and cases in which the occurrences of the narcoleptic symptoms are parallel with the rise and fall of the causative disease. A review of these cases

  19. A population-based epidemiologic study of adult-onset narcolepsy incidence and associated risk factors, 2004-2013.

    PubMed

    Lee, Rachel U; Radin, Jennifer M

    2016-11-15

    An increase in narcolepsy incidence was noted after the novel pandemic influenza of 2009, leading to further interest in risk factors associated with this disease. However, there is limited data on the epidemiology of narcolepsy, particularly in the adult population. Therefore, we sought to examine narcolepsy incidence rates in the United States and describe associated characteristics. We performed a population based epidemiologic study of active duty military personnel. All outpatient clinics in the continental United States providing care for active duty military between 2004 through 2013 were included utilizing existing databases. Narcolepsy was defined in 3 ways: (1) 2 diagnoses of narcolepsy within 6months of each other, one made by a sleep expert; (2) 2 diagnoses by any provider followed by a narcolepsy prescription within 14days of last visit; and (3) procedure code for a sleep study followed by a narcolepsy diagnosis by a sleep expert within 6months. There were 1675 narcolepsy cases. Overall incidence of narcolepsy trended from 14.6 to 27.3 cases per 100,000 person-years, with an increase starting after 2005-2006 and peaking during the 2011-2012 influenza season. Higher frequencies were seen among females, non-Hispanic blacks, and members living in the south. Narcolepsy incidence rates among active duty military members are higher than previously described. The reason for the steady rise of incidence from 2005 to 2006 through 2011-2012 is unknown; however, these findings require further exploration. We detected risk factors associated with the development of narcolepsy which may aid in future study efforts. Published by Elsevier B.V.

  20. The diagnosis and treatment of pediatric narcolepsy.

    PubMed

    Nevsimalova, Sona

    2014-08-01

    Narcolepsy in children is a serious disorder marked by a chronic course and lifelong handicap in school performance and choice of employment, by free time activity limitation, and by behavior and personality changes, all of which constitute a major influence on the quality of life. Increased daytime sleepiness may be the only sign at the disease onset, with attacks of sleep becoming longer and lasting up to hours. Also present may be confusional arousals with features of sleep drunkenness. Paradoxically, preschool and young children may show inattentiveness, emotional lability, and hyperactive behavior. Cataplexy may develop after onset of sleepiness and affect mainly muscles of the face. Hypnagogic hallucinations and sleep paralysis are seldom present. Multiple Sleep Latency Test criteria are not available for children younger than 6 years. The haplotype (HLA-DQB1:0602) can be associated with the disorder; however, the best predictor of narcolepsy-cataplexy is hypocretin deficiency. The treatment generally used in adults is regarded as off-label in childhood, which is why the management of pediatric narcolepsy is difficult.

  1. Narcolepsy with Long Sleep Time: A Specific Entity?

    PubMed Central

    Vernet, Cyrille; Arnulf, Isabelle

    2009-01-01

    Background: The classical narcolepsy patient reports intense feelings of sleepiness (with/out cataplexy), normal or disrupted nighttime sleep, and takes short and restorative naps. However, with long-term monitoring, we identified some narcoleptics resembling patients with idiopathic hypersomnia. Objective: To isolate and describe a new subtype of narcolepsy with long sleep time). Setting: University Hospital Design: Controlled, prospective cohort Participants: Out of 160 narcoleptics newly diagnosed within the past 3 years, 29 (18%) had a long sleep time (more than 11 h/24 h). We compared narcoleptics with (n = 23) and without (n = 29) long sleep time to 25 hypersomniacs with long sleep time and 20 healthy subjects. Intervention: Patients and controls underwent face-to face interviews, questionnaires, human leukocyte antigen (HLA) genotype, an overnight polysomnography, multiple sleep latency tests, and 24-h ad libitum sleep monitoring. Results: Narcoleptics with long sleep time had a similar disease course and similar frequencies of cataplexy, sleep paralysis, hallucinations, multiple sleep onset in REM periods, short mean sleep latencies, and HLA DQB1*0602 positivity as narcoleptics with normal sleep time did. However, they had longer sleep time during 24 h, and higher sleep efficiency, lower Epworth Sleepiness Scale scores, and reported their naps were more often unrefreshing. Only 3/23 had core narcolepsy (HLA and cataplexy positive). Conclusions: The subgroup of narcoleptics with a long sleep time comprises 18% of narcoleptics. Their symptoms combine the disabilities of both narcolepsy (severe sleepiness) and idiopathic hypersomnia (long sleep time and unrefreshing naps). Thus, they may constitute a group with multiple arousal system dysfunctions. Citation: Vernet C; Arnulf I. Narcolepsy with long sleep time: a specific entity? SLEEP 2009;32(9):1229-1235. PMID:19750928

  2. Comparison of Polysomnography and Multiple Sleep Latency Test Findings in Subjects with Narcolepsy and İdiopathic Hypersomnia.

    PubMed

    Erdem, Murat; Bolu, Abdullah; Ünlü, A Gazi; Alper, Mustafa; Yetkin, Sinan

    2013-09-01

    Both narcolepsy and idiopathic hypersomnia are the main causes of excessive daytime sleepiness. In this study, we aimed to compare polysomnography (PSG) and multiple sleep latency test (MSLT) findings in narcolepsy and idiopathic hypersomnia patients. The files of patients with narcolepsy and hypersomnia who were admitted between 1995 and 2009 were reviewed. We evaluated data from 94 patients with narcolepsy with cataplexy, 49 with narcolepsy without cataplexy and 140 patients with idiopathic hypersomnia. Sleep latency and REM latency were longer in idiopathic hypersomnia group than in narcolepsy with and without cataplexy group. Mean sleep latency in MSLT was the shortest in narcolepsy with cataplexy group. There was no difference in sleep efficiency, percentage of sleep stage and number of awakenings in PSG between three groups. The findings of the study indicated that narcolepsy patients differ from idiopathic hypersomnia patients in terms of sleep latency and REM latency in PSG.

  3. CSF Histamine Contents in Narcolepsy, Idiopathic Hypersomnia and Obstructive Sleep Apnea Syndrome

    PubMed Central

    Kanbayashi, Takashi; Kodama, Tohru; Kondo, Hideaki; Satoh, Shinsuke; Inoue, Yuichi; Chiba, Shigeru; Shimizu, Tetsuo; Nishino, Seiji

    2009-01-01

    Study Objective: To (1) replicate our prior result of low cerebrospinal fluid (CSF) histamine levels in human narcolepsy in a different sample population and to (2) evaluate if histamine contents are altered in other types of hypersomnia with and without hypocretin deficiency. Design: Cross sectional studies. Setting and Patients: Sixty-seven narcolepsy subjects, 26 idiopathic hypersomnia (IHS) subjects, 16 obstructive sleep apnea syndrome (OSAS) subjects, and 73 neurological controls were included. All patients were Japanese. Diagnoses were made according to ICSD-2. Results: We found significant reductions in CSF histamine levels in hypocretin deficient narcolepsy with cataplexy (mean ± SEM; 176.0 ± 25.8 pg/mL), hypocretin non-deficient narcolepsy with cataplexy (97.8 ± 38.4 pg/mL), hypocretin non-deficient narcolepsy without cataplexy (113.6 ± 16.4 pg/mL), and idiopathic hypersomnia (161.0 ± 29.3 pg/mL); the levels in OSAS (259.3 ± 46.6 pg/mL) did not statistically differ from those in the controls (333.8 ± 22.0 pg/mL). Low CSF histamine levels were mostly observed in non-medicated patients; significant reductions in histamine levels were evident in non-medicated patients with hypocretin deficient narcolepsy with cataplexy (112.1 ± 16.3 pg/mL) and idiopathic hypersomnia (143.3 ± 28.8 pg/mL), while the levels in the medicated patients were in the normal range. Conclusion: The study confirmed reduced CSF histamine levels in hypocretin-deficient narcolepsy with cataplexy. Similar degrees of reduction were also observed in hypocretin non-deficient narcolepsy and in idiopathic hypersomnia, while those in OSAS (non central nervous system hypersomnia) were not altered. The decrease in histamine in these subjects were more specifically observed in non-medicated subjects, suggesting CSF histamine is a biomarker reflecting the degree of hypersomnia of central origin. Citation: Kanbayashi T; Kodama T; Kondo H; Satoh S; Inoue Y; Chiba S; Shimizu T; Nishino S. CSF

  4. Narcolepsy induced by chronic heavy alcohol consumption: a case report

    PubMed Central

    Wang, Xinyuan

    2012-01-01

    Summary Narcolepsy is a chronic neurological disorder, characterized by uncontrollable excessive daytime sleepiness, cataplectic episodes, sleep paralysis, hypnagogic hallucinations, and night time sleep disruption. The paper reviewed the related literature and reported a case of long-term drinking induced narcolepsy which was significantly improved after treatment with paroxetine and dexzopiclone. PMID:25328357

  5. The neurobiology, diagnosis, and treatment of narcolepsy.

    PubMed

    Scammell, Thomas E

    2003-02-01

    Narcolepsy is a common cause of chronic sleepiness distinguished by intrusions into wakefulness of physiological aspects of rapid eye movement sleep such as cataplexy and hallucinations. Recent advances provide compelling evidence that narcolepsy may be a neurodegenerative or autoimmune disorder resulting in a loss of hypothalamic neurons containing the neuropeptide orexin (also known as hypocretin). Because orexin promotes wakefulness and inhibits rapid eye movement sleep, its absence may permit inappropriate transitions between wakefulness and sleep. These discoveries have considerably improved our understanding of the neurobiology of sleep and should foster the development of rational treatments for a variety of sleep disorders.

  6. Positive effects of massage therapy on a patient with narcolepsy.

    PubMed

    Hill, Robyn; Baskwill, Amanda

    2013-01-01

    The purpose of this case report was to investigate the effects of massage therapy on the sleep patterns of a woman with narcolepsy. The 23-year-old woman's primary symptoms included excessive daytime sleepiness and periodic leg movements (PLM), which were associated with her diagnoses of both narcolepsy and cataplexy. Five 45-minute massage therapy treatments were administered over a five-week period. The patient's sleep patterns were recorded each week before the treatment. A final measurement was recorded in the sixth week. The sleep patterns were monitored using the Leeds Sleep Evaluation Questionnaire, which included ten visual analogue scales. The results of this case report included an improvement in getting to sleep by 148%, an improvement in quality of sleep by 1100%, an improvement in awake following sleep by 121%, and an improvement in behaviour following wakening by 28% using the Leeds Sleep Evaluation Questionnaire. This case report suggests that massage therapy had a positive effect on this patient with narcolepsy. Further research is needed to investigate the effects of massage therapy on narcolepsy and sleep patterns.

  7. Gray matter atrophy in narcolepsy: An activation likelihood estimation meta-analysis.

    PubMed

    Weng, Hsu-Huei; Chen, Chih-Feng; Tsai, Yuan-Hsiung; Wu, Chih-Ying; Lee, Meng; Lin, Yu-Ching; Yang, Cheng-Ta; Tsai, Ying-Huang; Yang, Chun-Yuh

    2015-12-01

    The authors reviewed the literature on the use of voxel-based morphometry (VBM) in narcolepsy magnetic resonance imaging (MRI) studies via the use of a meta-analysis of neuroimaging to identify concordant and specific structural deficits in patients with narcolepsy as compared with healthy subjects. We used PubMed to retrieve articles published between January 2000 and March 2014. The authors included all VBM research on narcolepsy and compared the findings of the studies by using gray matter volume (GMV) or gray matter concentration (GMC) to index differences in gray matter. Stereotactic data were extracted from 8 VBM studies of 149 narcoleptic patients and 162 control subjects. We applied activation likelihood estimation (ALE) technique and found significant regional gray matter reduction in the bilateral hypothalamus, thalamus, globus pallidus, extending to nucleus accumbens (NAcc) and anterior cingulate cortex (ACC), left mid orbital and rectal gyri (BAs 10 and 11), right inferior frontal gyrus (BA 47), and the right superior temporal gyrus (BA 41) in patients with narcolepsy. The significant gray matter deficits in narcoleptic patients occurred in the bilateral hypothalamus and frontotemporal regions, which may be related to the emotional processing abnormalities and orexin/hypocretin pathway common among populations of patients with narcolepsy. Copyright © 2015. Published by Elsevier Ltd.

  8. Effective treatment of narcolepsy-like symptoms with high-frequency repetitive transcranial magnetic stimulation

    PubMed Central

    Lai, Jian-bo; Han, Mao-mao; Xu, Yi; Hu, Shao-hua

    2017-01-01

    Abstract Rationale: Narcolepsy is a rare sleep disorder with disrupted sleep-architecture. Clinical management of narcolepsy lies dominantly on symptom-driven pharmacotherapy. The treatment role of repetitive transcranial magnetic stimulation (rTMS) for narcolepsy remains unexplored. Patient concerns: In this paper, we present a case of a 14-year-old young girl with excessive daytime sleepiness (EDS), cataplexy and hypnagogic hallucinations. Diagnoses: After excluding other possible medical conditions, this patient was primarily diagnosed with narcolepsy. Interventions: The patient received 25 sessions of high-frequency rTMS over the left dorsolateral prefrontal cortex (DLPFC). Outcomes: The symptoms of EDS and cataplexy significantly improved after rTMS treatment. Meanwhile, her score in the Epworth sleep scale (ESS) also remarkably decreased. Lessons: This case indicates that rTMS may be selected as a safe and effective alternative strategy for treating narcolepsy-like symptoms. Well-designed researches are warranted in future investigations on this topic. PMID:29145290

  9. Clinical and practical considerations in the pharmacologic management of narcolepsy.

    PubMed

    Thorpy, Michael J; Dauvilliers, Yves

    2015-01-01

    Despite published treatment recommendations and the availability of approved and off-label pharmacologic therapies for narcolepsy, the clinical management of this incurable, chronic neurologic disorder remains challenging. While treatment is generally symptomatically driven, decisions regarding which drug(s) to use need to take into account a variety of factors that may affect adherence, efficacy, and tolerability. Type 1 narcolepsy (predominantly excessive daytime sleepiness with cataplexy) or type 2 narcolepsy (excessive daytime sleepiness without cataplexy) may drive treatment decisions, with consideration given either to a single drug that targets multiple symptoms or to multiple drugs that each treat a specific symptom. Other drug-related characteristics that affect drug choice are dosing regimens, tolerability, and potential drug-drug interactions. Additionally, the patient should be an active participant in treatment decisions, and the main symptomatic complaints, treatment goals, psychosocial setting, and use of lifestyle substances (ie, alcohol, nicotine, caffeine, and cannabis) need to be discussed with respect to treatment decisions. Although there is a lack of narcolepsy-specific instruments for monitoring therapeutic effects, clinically relevant subjective and objective measures of daytime sleepiness (eg, Epworth Sleepiness Scale and Maintenance of Wakefulness Test) can be used to provide guidance on whether treatment goals are being met. These considerations are discussed with the objective of providing clinically relevant recommendations for making treatment decisions that can enhance the effective management of patients with narcolepsy. Copyright © 2014 The Authors. Published by Elsevier B.V. All rights reserved.

  10. Complex Movement Disorders at Disease Onset in Childhood Narcolepsy with Cataplexy

    ERIC Educational Resources Information Center

    Plazzi, Giuseppe; Pizza, Fabio; Palaia, Vincenzo; Franceschini, Christian; Poli, Francesca; Moghadam, Keivan K.; Cortelli, Pietro; Nobili, Lino; Bruni, Oliviero; Dauvilliers, Yves; Lin, Ling; Edwards, Mark J.; Mignot, Emmanuel; Bhatia, Kailash P.

    2011-01-01

    Narcolepsy with cataplexy is characterized by daytime sleepiness, cataplexy (sudden loss of bilateral muscle tone triggered by emotions), sleep paralysis, hypnagogic hallucinations and disturbed nocturnal sleep. Narcolepsy with cataplexy is most often associated with human leucocyte antigen-DQB1*0602 and is caused by the loss of…

  11. Rare missense mutations in P2RY11 in narcolepsy with cataplexy.

    PubMed

    Degn, Matilda; Dauvilliers, Yves; Dreisig, Karin; Lopez, Régis; Pfister, Corinne; Pradervand, Sylvain; Rahbek Kornum, Birgitte; Tafti, Mehdi

    2017-06-01

    The sleep disorder narcolepsy with cataplexy is characterized by a highly specific loss of hypocretin (orexin) neurons, leading to the hypothesis that the condition is caused by an immune or autoimmune mechanism. All genetic variants associated with narcolepsy are immune-related. Among these are single nucleotide polymorphisms in the P2RY11-EIF3G locus. It is unknown how these genetic variants affect narcolepsy pathogenesis and whether the effect is directly related to P2Y11 signalling or EIF3G function. Exome sequencing in 18 families with at least two affected narcolepsy with cataplexy subjects revealed non-synonymous mutations in the second exon of P2RY11 in two families, and P2RY11 re-sequencing in 250 non-familial cases and 135 healthy control subjects revealed further six different non-synonymous mutations in the second exon of P2RY11 in seven patients. No mutations were found in healthy controls. Six of the eight narcolepsy-associated P2Y11 mutations resulted in significant functional deficits in P2Y11 signalling through both Ca2+ and cAMP signalling pathways. In conclusion, our data show that decreased P2Y11 signalling plays an important role in the development of narcolepsy with cataplexy. © The Author (2017). Published by Oxford University Press on behalf of the Guarantors of Brain. All rights reserved. For Permissions, please email: journals.permissions@oup.com.

  12. Impact of cytokine in type 1 narcolepsy: Role of pandemic H1N1 vaccination ?

    PubMed

    Lecendreux, Michel; Libri, Valentina; Jaussent, Isabelle; Mottez, Estelle; Lopez, Régis; Lavault, Sophie; Regnault, Armelle; Arnulf, Isabelle; Dauvilliers, Yves

    2015-06-01

    Recent advances in the identification of susceptibility genes and environmental exposures (pandemic influenza 2009 vaccination) provide strong support that narcolepsy type 1 is an immune-mediated disease. Considering the limited knowledge regarding the immune mechanisms involved in narcolepsy whether related to flu vaccination or not and the recent progresses in cytokine measurement technology, we assessed 30 cytokines, chemokines and growth factors using the Luminex technology in either peripheral (serum) or central (CSF) compartments in a large population of 90 children and adult patients with narcolepsy type 1 in comparison to 58 non-hypocretin deficient hypersomniacs and 41 healthy controls. Furthermore, we compared their levels in patients with narcolepsy whether exposed to pandemic flu vaccine or not, and analyzed the effect of age, duration of disease and symptom severity. Comparison for sera biomarkers between narcolepsy (n = 84, 54 males, median age: 15.5 years old) and healthy controls (n = 41, 13 males, median age: 20 years old) revealed an increased stimulation of the immune system with high release of several pro- and anti-inflammatory serum cytokines and growth factors with interferon-γ, CCL11, epidermal growth factor, and interleukin-2 receptor being independently associated with narcolepsy. Increased levels of interferon-γ, CCL11, and interleukin-12 were found when close to narcolepsy onset. After several adjustments, only one CSF biomarker differed between narcolepsy (n = 44, 26 males, median age: 15 years old) and non-hypocretin deficient hypersomnias (n = 57, 24 males, median age: 36 years old) with higher CCL 3 levels found in narcolepsy. Comparison for sera biomarkers between patients with narcolepsy who developed the disease post-pandemic flu vaccination (n = 36) to those without vaccination (n = 48) revealed an increased stimulation of the immune system with high release of three cytokines, regulated upon activation normal T-cell expressed

  13. Beta-amyloid and phosphorylated tau metabolism changes in narcolepsy over time.

    PubMed

    Liguori, Claudio; Placidi, Fabio; Izzi, Francesca; Nuccetelli, Marzia; Bernardini, Sergio; Sarpa, Maria Giovanna; Cum, Fabrizio; Marciani, Maria Grazia; Mercuri, Nicola Biagio; Romigi, Andrea

    2016-03-01

    The aim od this study is to test whether metabolism of beta-amyloid and tau proteins changes in narcolepsy along with the disease course. We analyzed a population of narcoleptic drug-naïve patients compared to a sample of healthy controls. Patients and controls underwent lumbar puncture for assessment of cerebrospinal fluid (CSF) beta-amyloid1-42 (Aβ42), total tau (t-tau), and phosphorylated tau (p-tau) levels. Moreover, based on the median disease duration of the whole narcolepsy group, the patients were divided into two subgroups: patients with a short disease duration (SdN, <5 years) and patients with a long disease duration (LdN, >5 years). We found significantly lower CSF Aβ42 levels in the whole narcolepsy group with respect to controls. Taking into account the patient subgroups, we documented reduced CSF Aβ42 levels in SdN compared to both LdN and controls. Even LdN patients showed lower CSF Aβ42 levels with respect to controls. Moreover, we documented higher CSF p-tau levels in LdN patients compared to both SdN and controls. Finally, a significant positive correlation between CSF Aβ42 levels and disease duration was evident. We hypothesize that beta-amyloid metabolism and cascade may be impaired in narcolepsy not only at the onset but also along with the disease course, although they show a compensatory profile over time. Concurrently, also CSF biomarkers indicative of neural structure (p-tau) appear to be altered in narcolepsy patients with a long disease duration. However, the mechanism underlying beta-amyloid and tau metabolism impairment in narcolepsy remains still unclear and deserves to be better elucidated.

  14. A polymorphism in CCR1/CCR3 is associated with narcolepsy.

    PubMed

    Toyoda, Hiromi; Miyagawa, Taku; Koike, Asako; Kanbayashi, Takashi; Imanishi, Aya; Sagawa, Yohei; Kotorii, Nozomu; Kotorii, Tatayu; Hashizume, Yuji; Ogi, Kimihiro; Hiejima, Hiroshi; Kamei, Yuichi; Hida, Akiko; Miyamoto, Masayuki; Imai, Makoto; Fujimura, Yota; Tamura, Yoshiyuki; Ikegami, Azusa; Wada, Yamato; Moriya, Shunpei; Furuya, Hirokazu; Takeuchi, Masaki; Kirino, Yohei; Meguro, Akira; Remmers, Elaine F; Kawamura, Yoshiya; Otowa, Takeshi; Miyashita, Akinori; Kashiwase, Koichi; Khor, Seik-Soon; Yamasaki, Maria; Kuwano, Ryozo; Sasaki, Tsukasa; Ishigooka, Jun; Kuroda, Kenji; Kume, Kazuhiko; Chiba, Shigeru; Yamada, Naoto; Okawa, Masako; Hirata, Koichi; Mizuki, Nobuhisa; Uchimura, Naohisa; Shimizu, Tetsuo; Inoue, Yuichi; Honda, Yutaka; Mishima, Kazuo; Honda, Makoto; Tokunaga, Katsushi

    2015-10-01

    Etiology of narcolepsy-cataplexy involves multiple genetic and environmental factors. While the human leukocyte antigen (HLA)-DRB1*15:01-DQB1*06:02 haplotype is strongly associated with narcolepsy, it is not sufficient for disease development. To identify additional, non-HLA susceptibility genes, we conducted a genome-wide association study (GWAS) using Japanese samples. An initial sample set comprising 409 cases and 1562 controls was used for the GWAS of 525,196 single nucleotide polymorphisms (SNPs) located outside the HLA region. An independent sample set comprising 240 cases and 869 controls was then genotyped at 37 SNPs identified in the GWAS. We found that narcolepsy was associated with a SNP in the promoter region of chemokine (C-C motif) receptor 1 (CCR1) (rs3181077, P=1.6×10(-5), odds ratio [OR]=1.86). This rs3181077 association was replicated with the independent sample set (P=0.032, OR=1.36). We measured mRNA levels of candidate genes in peripheral blood samples of 38 cases and 37 controls. CCR1 and CCR3 mRNA levels were significantly lower in patients than in healthy controls, and CCR1 mRNA levels were associated with rs3181077 genotypes. In vitro chemotaxis assays were also performed to measure monocyte migration. We observed that monocytes from carriers of the rs3181077 risk allele had lower migration indices with a CCR1 ligand. CCR1 and CCR3 are newly discovered susceptibility genes for narcolepsy. These results highlight the potential role of CCR genes in narcolepsy and support the hypothesis that patients with narcolepsy have impaired immune function. Copyright © 2015 Elsevier Inc. All rights reserved.

  15. [Workplace accomodations for two workers with narcolepsy].

    PubMed

    Vico Garcerán, Belén; Monzó Salas, Monserrat; Cuenca Esteve, Francisco; Luis Domingo, José

    2013-01-01

    We describe the case of two workers evaluated in our occupational health unit. The first worker was a kitchen aide; the second was a primary care physician. Both had been diagnosed with narcolepsy and had obvious disability.We assessed occupational hazards related to their jobs, analysed their tasks, and performed medical examinations. Afterwards, we offered recommendations to the patients, consisting of avoidance of situations involving a risk of work accidents and improving their sleep habits. Narcolepsy is a rare disorder, but it has important social and occupational consequences. A better understanding of the disease and some work accommodations can help improve the quality of life of affected workers.

  16. Epigenome-wide association study of DNA methylation in narcolepsy: an integrated genetic and epigenetic approach.

    PubMed

    Shimada, Mihoko; Miyagawa, Taku; Toyoda, Hiromi; Tokunaga, Katsushi; Honda, Makoto

    2018-04-01

    Narcolepsy with cataplexy, which is a hypersomnia characterized by excessive daytime sleepiness and cataplexy, is a multifactorial disease caused by both genetic and environmental factors. Several genetic factors including HLA-DQB1*06:02 have been identified; however, the disease etiology is still unclear. Epigenetic modifications, such as DNA methylation, have been suggested to play an important role in the pathogenesis of complex diseases. Here, we examined DNA methylation profiles of blood samples from narcolepsy and healthy control individuals and performed an epigenome-wide association study (EWAS) to investigate methylation loci associated with narcolepsy. Moreover, data from the EWAS and a previously performed narcolepsy genome-wide association study were integrated to search for methylation loci with causal links to the disease. We found that (1) genes annotated to the top-ranked differentially methylated positions (DMPs) in narcolepsy were associated with pathways of hormone secretion and monocarboxylic acid metabolism. (2) Top-ranked narcolepsy-associated DMPs were significantly more abundant in non-CpG island regions and more than 95 per cent of such sites were hypomethylated in narcolepsy patients. (3) The integrative analysis identified the CCR3 region where both a single methylation site and multiple single-nucleotide polymorphisms were found to be associated with the disease as a candidate region responsible for narcolepsy. The findings of this study suggest the importance of future replication studies, using methylation technologies with wider genome coverage and/or larger number of samples, to confirm and expand on these results.

  17. Disrupted Sleep in Narcolepsy: Exploring the Integrity of Galanin Neurons in the Ventrolateral Preoptic Area.

    PubMed

    Gavrilov, Yury V; Ellison, Brian A; Yamamoto, Mihoko; Reddy, Hasini; Haybaeck, Johannes; Mignot, Emmanuel; Baumann, Christian R; Scammell, Thomas E; Valko, Philipp O

    2016-05-01

    To examine the integrity of sleep-promoting neurons of the ventrolateral preoptic nucleus (VLPO) in postmortem brains of narcolepsy type 1 patients. Postmortem examination of five narcolepsy and eight control brains. VLPO galanin neuron count did not differ between narcolepsy patients (11,151 ± 3,656) and controls (13,526 ± 9,544). A normal number of galanin-immunoreactive VLPO neurons in narcolepsy type 1 brains at autopsy suggests that VLPO cell loss is an unlikely explanation for the sleep fragmentation that often accompanies the disease. © 2016 Associated Professional Sleep Societies, LLC.

  18. Positive Effects of Massage Therapy on a Patient with Narcolepsy

    PubMed Central

    Hill, Robyn; Baskwill, Amanda

    2013-01-01

    Purpose The purpose of this case report was to investigate the effects of massage therapy on the sleep patterns of a woman with narcolepsy. Participant The 23-year-old woman’s primary symptoms included excessive daytime sleepiness and periodic leg movements (PLM), which were associated with her diagnoses of both narcolepsy and cataplexy. Intervention Five 45-minute massage therapy treatments were administered over a five-week period. The patient’s sleep patterns were recorded each week before the treatment. A final measurement was recorded in the sixth week. The sleep patterns were monitored using the Leeds Sleep Evaluation Questionnaire, which included ten visual analogue scales. Results The results of this case report included an improvement in getting to sleep by 148%, an improvement in quality of sleep by 1100%, an improvement in awake following sleep by 121%, and an improvement in behaviour following wakening by 28% using the Leeds Sleep Evaluation Questionnaire. Conclusion This case report suggests that massage therapy had a positive effect on this patient with narcolepsy. Further research is needed to investigate the effects of massage therapy on narcolepsy and sleep patterns. PMID:23730398

  19. CSF Hypocretin-1 Levels and Clinical Profiles in Narcolepsy and Idiopathic CNS Hypersomnia in Norway

    PubMed Central

    Heier, Mona Skard; Evsiukova, Tatiana; Vilming, Steinar; Gjerstad, Michaela D.; Schrader, Harald; Gautvik, Kaare

    2007-01-01

    Objective: To evaluate the relationship between CSF hypocretin-1 levels and clinical profiles in narcolepsy and CNS hypersomnia in Norwegian patients. Method: CSF hypocretin-1 was measured by a sensitive radioimmunoassay in 47 patients with narcolepsy with cataplexy, 7 with narcolepsy without cataplexy, 10 with idiopathic CNS hypersomnia, and a control group. Results: Low hypocretin-1 values were found in 72% of the HLA DQB1*0602 positive patients with narcolepsy and cataplexy. Patients with low CSF hypocretin-1 levels reported more extensive muscular involvement during cataplectic attacks than patients with normal levels. Hypnagogic hallucinations and sleep paralysis occurred more frequently in patients with cataplexy than in the other patient groups, but with no correlation to hypocretin-1 levels. Conclusion: About three quarters of the HLA DQB1*0602 positive patients with narcolepsy and cataplexy had low CSF hypocretin-1 values, and appear to form a distinct clinical entity. Narcolepsy without cataplexy could not be distinguished from idiopathic CNS hypersomnia by clinical symptoms or biochemical findings. Citation: Heier MS; Evsiukova T; Vilming S; Gjerstad MD; Schrader H; Gautvik K. CSF hypocretin-1 levels and clinical profiles in narcolepsy and idiopathic CNS hypersomnia in norway. SLEEP 2007;30(8):969-973. PMID:17702265

  20. Disrupted Sleep in Narcolepsy: Exploring the Integrity of Galanin Neurons in the Ventrolateral Preoptic Area

    PubMed Central

    Gavrilov, Yury V.; Ellison, Brian A.; Yamamoto, Mihoko; Reddy, Hasini; Haybaeck, Johannes; Mignot, Emmanuel; Baumann, Christian R.; Scammell, Thomas E.; Valko, Philipp O.

    2016-01-01

    Study Objectives: To examine the integrity of sleep-promoting neurons of the ventrolateral preoptic nucleus (VLPO) in postmortem brains of narcolepsy type 1 patients. Methods: Postmortem examination of five narcolepsy and eight control brains. Results: VLPO galanin neuron count did not differ between narcolepsy patients (11,151 ± 3,656) and controls (13,526 ± 9,544). Conclusions: A normal number of galanin-immunoreactive VLPO neurons in narcolepsy type 1 brains at autopsy suggests that VLPO cell loss is an unlikely explanation for the sleep fragmentation that often accompanies the disease. Citation: Gavrilov YV, Ellison BA, Yamamoto M, Reddy H, Haybaeck J, Mignot E, Baumann CR, Scammell TE, Valko PO. Disrupted sleep in narcolepsy: exploring the integrity of galanin neurons in the ventrolateral preoptic area. SLEEP 2016;39(5):1059–1062. PMID:26951397

  1. Tolerance and Efficacy of Sodium Oxybate in Childhood Narcolepsy with Cataplexy: A Retrospective Study

    PubMed Central

    Lecendreux, Michel; Poli, Francesca; Oudiette, Delphine; Benazzouz, Fatima; Donjacour, Claire E.H.M; Franceschini, Christian; Finotti, Elena; Pizza, Fabio; Bruni, Oliviero; Plazzi, Giuseppe

    2012-01-01

    Narcolepsy with cataplexy is a sleep disorder characterized by excessive daytime sleepiness, irresistible sleep episodes, and sudden loss of muscle tone (cataplexy) mostly triggered by emotions. Narcolepsy with cataplexy is a disabling lifelong disorder frequently arising during childhood. Pediatric narcolepsy often results in severe learning and social impairment. Improving awareness about this condition increases early diagnosis and may allow patients to rapidly access adequate treatments, including pharmacotherapy and/or non-medication-based approaches. Even though children currently undergo pharmacotherapy, data about safety and efficacy in the pediatric population are scarce. Lacking international guidelines as well as drugs registered for childhood narcolepsy with cataplexy, physicians have no other alternative but to prescribe in an off-label manner medications identical to those recommended for adults. We retrospectively evaluated 27 children ranging from 6 to 16 years old, suffering from narcolepsy with cataplexy, who had been treated with off-label sodium oxybate and had been followed in a clinical setting. Throughout a semi-structured interview, we documented the good efficacy and tolerability of sodium oxybate in the majority of the patients. This study constitutes a preliminary step towards a further randomized controlled trial in childhood narcolepsy with cataplexy. Citation: Lecendreux M; Poli F; Oudiette D; Benazzouz F; Donjacour CEHM; Franceschini C; Finotti E; Pizza F; Bruni O; Plazzi G. Tolerance and efficacy of sodium oxybate in childhood narcolepsy with cataplexy: a retrospective study. SLEEP 2012;35(5):709-711. PMID:22547897

  2. Levothyroxine Improves Subjective Sleepiness in a Euthyroid Patient with Narcolepsy without Cataplexy

    PubMed Central

    Sobol, Danielle L.; Spector, Andrew R.

    2014-01-01

    Objective: We discuss the use of levothyroxine for excessive daytime sleepiness (EDS) and prolonged nocturnal sleep time in a euthyroid patient with narcolepsy. Methods: After failure of first-line narcolepsy treatments, a 48-year-old female began levothyroxine (25 mcg/day). After 12 weeks of treatment, the patient was evaluated for improvement in total sleep time and subjective daytime sleepiness assessed by Epworth Sleepiness Scale (ESS). Results: At baseline, ESS score was 16 and total sleep time averaged 16 h/day. After 12 weeks, ESS was 13 and reported total sleep time was 13 h/day. Conclusions: Levothyroxine improved EDS and total sleep time in a euthyroid patient with narcolepsy without cataplexy after 12 weeks without side effects. Citation: Sobol DL, Spector AR. Levothyroxine improves subjective sleepiness in a euthyroid patient with narcolepsy without cataplexy. J Clin Sleep Med 2014;10(11):1231-1232. PMID:25325591

  3. Differences in Brain Morphological Findings between Narcolepsy with and without Cataplexy

    PubMed Central

    Nakamura, Masaki; Nishida, Shingo; Hayashida, Kenichi; Ueki, Yoichiro; Dauvilliers, Yves; Inoue, Yuichi

    2013-01-01

    Objective Maps of fractional anisotropy (FA) and apparent diffusion coefficient (ADC) obtained by diffusion tensor imaging (DTI) can detect microscopic axonal changes by estimating the diffusivity of water molecules using magnetic resonance imaging (MRI). We applied an MRI voxel-based statistical approach to FA and ADC maps to evaluate microstructural abnormalities in the brain in narcolepsy and to investigate differences between patients having narcolepsy with and without cataplexy. Methods Twelve patients with drug-naive narcolepsy with cataplexy (NA/CA), 12 with drug-naive narcolepsy without cataplexy (NA w/o CA) and 12 age-matched healthy normal controls (NC) were enrolled. FA and ADC maps for these 3 groups were statistically compared by using voxel-based one-way ANOVA. In addition, we investigated the correlation between FA and ADC values and clinical variables in the patient groups. Results Compared to the NC group, the NA/CA group showed higher ADC values in the left inferior frontal gyrus and left amygdala, and a lower ADC value in the left postcentral gyrus. The ADC value in the right inferior frontal gyrus and FA value in the right precuneus were higher for NA/CA group than for the NA w/o CA group. However, no significant differences were observed in FA and ADC values between the NA w/o CA and NC groups in any of the areas investigated. In addition, no correlation was found between the clinical variables and ADC and FA values of any brain areas in these patient groups. Conclusions Several microstructural changes were noted in the inferior frontal gyrus and amygdala in the NA/CA but not in the NA w/o CA group. These findings suggest that these 2 narcolepsy conditions have different pathological mechanisms: narcolepsy without cataplexy form appears to be a potentially broader condition without any significant brain imaging differences from normal controls. PMID:24312261

  4. Narcolepsy

    PubMed Central

    Murray, T. J.; Foley, Anita

    1974-01-01

    Narcolepsy is a disorder of sleep control characterized by a tetrad of symptoms: sleep attacks, cataplexy, sleep paralysis and hypnagogic hallucinations. A diagnosis is made from a careful history. The incidence is estimated as high as 0.3% of the population. Unfortunately patients go for many years before the diagnosis is made and often have experienced disruption of their employment, social and family life, and may have experienced a number of car accidents because of falling asleep at the wheel. An unknown number of narcoleptics kill themselves on the highways before the diagnosis is ever made. Sleep attacks can usually be controlled by methylphenidate, and if the other symptoms persist they can often be effectively managed by imipramine. PMID:4809449

  5. Sleep spindle density in narcolepsy.

    PubMed

    Christensen, Julie Anja Engelhard; Nikolic, Miki; Hvidtfelt, Mathias; Kornum, Birgitte Rahbek; Jennum, Poul

    2017-06-01

    Patients with narcolepsy type 1 (NT1) show alterations in sleep stage transitions, rapid-eye-movement (REM) and non-REM sleep due to the loss of hypocretinergic signaling. However, the sleep microstructure has not yet been evaluated in these patients. We aimed to evaluate whether the sleep spindle (SS) density is altered in patients with NT1 compared to controls and patients with narcolepsy type 2 (NT2). All-night polysomnographic recordings from 28 NT1 patients, 19 NT2 patients, 20 controls (C) with narcolepsy-like symptoms, but with normal cerebrospinal fluid hypocretin levels and multiple sleep latency tests, and 18 healthy controls (HC) were included. Unspecified, slow, and fast SS were automatically detected, and SS densities were defined as number per minute and were computed across sleep stages and sleep cycles. The between-cycle trends of SS densities in N2 and NREM sleep were evaluated within and between groups. Between-group comparisons in sleep stages revealed no significant differences in any type of SS. Within-group analyses of the SS trends revealed significant decreasing trends for NT1, HC, and C between first and last sleep cycle. Between-group analyses of SS trends between first and last sleep cycle revealed that NT2 differ from NT1 patients in the unspecified SS density in NREM sleep, and from HC in the slow SS density in N2 sleep. SS activity is preserved in NT1, suggesting that the ascending neurons to thalamic activation of SS are not significantly affected by the hypocretinergic system. NT2 patients show an abnormal pattern of SS distribution. Copyright © 2017 Elsevier B.V. All rights reserved.

  6. CD8 T cell-mediated killing of orexinergic neurons induces a narcolepsy-like phenotype in mice.

    PubMed

    Bernard-Valnet, Raphaël; Yshii, Lidia; Quériault, Clémence; Nguyen, Xuan-Hung; Arthaud, Sébastien; Rodrigues, Magda; Canivet, Astrid; Morel, Anne-Laure; Matthys, Arthur; Bauer, Jan; Pignolet, Béatrice; Dauvilliers, Yves; Peyron, Christelle; Liblau, Roland S

    2016-09-27

    Narcolepsy with cataplexy is a rare and severe sleep disorder caused by the destruction of orexinergic neurons in the lateral hypothalamus. The genetic and environmental factors associated with narcolepsy, together with serologic data, collectively point to an autoimmune origin. The current animal models of narcolepsy, based on either disruption of the orexinergic neurotransmission or neurons, do not allow study of the potential autoimmune etiology. Here, we sought to generate a mouse model that allows deciphering of the immune mechanisms leading to orexin(+) neuron loss and narcolepsy development. We generated mice expressing the hemagglutinin (HA) as a "neo-self-antigen" specifically in hypothalamic orexin(+) neurons (called Orex-HA), which were transferred with effector neo-self-antigen-specific T cells to assess whether an autoimmune process could be at play in narcolepsy. Given the tight association of narcolepsy with the human leukocyte antigen (HLA) HLA-DQB1*06:02 allele, we first tested the pathogenic contribution of CD4 Th1 cells. Although these T cells readily infiltrated the hypothalamus and triggered local inflammation, they did not elicit the loss of orexin(+) neurons or clinical manifestations of narcolepsy. In contrast, the transfer of cytotoxic CD8 T cells (CTLs) led to both T-cell infiltration and specific destruction of orexin(+) neurons. This phenotype was further aggravated upon repeated injections of CTLs. In situ, CTLs interacted directly with MHC class I-expressing orexin(+) neurons, resulting in cytolytic granule polarization toward neurons. Finally, drastic neuronal loss caused manifestations mimicking human narcolepsy, such as cataplexy and sleep attacks. This work demonstrates the potential role of CTLs as final effectors of the immunopathological process in narcolepsy.

  7. CD8 T cell-mediated killing of orexinergic neurons induces a narcolepsy-like phenotype in mice

    PubMed Central

    Bernard-Valnet, Raphaël; Yshii, Lidia; Quériault, Clémence; Nguyen, Xuan-Hung; Arthaud, Sébastien; Rodrigues, Magda; Canivet, Astrid; Morel, Anne-Laure; Matthys, Arthur; Bauer, Jan; Pignolet, Béatrice; Dauvilliers, Yves; Peyron, Christelle; Liblau, Roland S.

    2016-01-01

    Narcolepsy with cataplexy is a rare and severe sleep disorder caused by the destruction of orexinergic neurons in the lateral hypothalamus. The genetic and environmental factors associated with narcolepsy, together with serologic data, collectively point to an autoimmune origin. The current animal models of narcolepsy, based on either disruption of the orexinergic neurotransmission or neurons, do not allow study of the potential autoimmune etiology. Here, we sought to generate a mouse model that allows deciphering of the immune mechanisms leading to orexin+ neuron loss and narcolepsy development. We generated mice expressing the hemagglutinin (HA) as a “neo-self-antigen” specifically in hypothalamic orexin+ neurons (called Orex-HA), which were transferred with effector neo-self-antigen–specific T cells to assess whether an autoimmune process could be at play in narcolepsy. Given the tight association of narcolepsy with the human leukocyte antigen (HLA) HLA-DQB1*06:02 allele, we first tested the pathogenic contribution of CD4 Th1 cells. Although these T cells readily infiltrated the hypothalamus and triggered local inflammation, they did not elicit the loss of orexin+ neurons or clinical manifestations of narcolepsy. In contrast, the transfer of cytotoxic CD8 T cells (CTLs) led to both T-cell infiltration and specific destruction of orexin+ neurons. This phenotype was further aggravated upon repeated injections of CTLs. In situ, CTLs interacted directly with MHC class I-expressing orexin+ neurons, resulting in cytolytic granule polarization toward neurons. Finally, drastic neuronal loss caused manifestations mimicking human narcolepsy, such as cataplexy and sleep attacks. This work demonstrates the potential role of CTLs as final effectors of the immunopathological process in narcolepsy. PMID:27621438

  8. Cataplexy in anxious patients: is subclinical narcolepsy underrecognized in anxiety disorders?

    PubMed

    Flosnik, Dawn L; Cortese, Bernadette M; Uhde, Thomas W

    2009-06-01

    Excessive daytime sleepiness, hypnagogic-hypnopompic hallucinations, sleep paralysis, and cataplexy are symptoms associated with narcolepsy. Recent findings indicate that anxiety disorders also are associated with excessive daytime sleepiness, hypnagogic-hypnopompic hallucinations, and sleep paralysis. These observations suggest a possible relationship between anxiety disorders and narcolepsy. Cataplexy is considered the most specific symptom of narcolepsy, but its association with anxiety disorders is unknown. This preliminary investigation examined the prevalence and types of cataplexy in patients with primary anxiety disorders. Sex- and age-matched patients with anxiety disorders (N = 33) and healthy volunteers (N = 33) were assessed on standardized and validated measures of subjective sleep quality (Pittsburgh Sleep Quality Index) and subclinical narcoleptic events in the form of cataplexy (Stanford Center for Narcolepsy Revised Sleep Inventory). Patients were recruited from October 2006 to January 2007 from 2 programs of the Penn State Behavioral Health Clinic. Anxiety disorder patients as a group reported poorer sleep quality and endorsed a larger number of different types of situations (e.g., surprise, embarrassment) associated with cataplectic events. Among anxious patients, 33.3% (11 of 33) endorsed events specific for classic cataplexy, as opposed to 9.1% (3 of 33) of healthy volunteers (chi(2) = 5.80, p = .016). Our preliminary findings suggest that anxiety disorders are associated with increased rates of cataplexy. Future research is indicated to elucidate the relationship between anxiety and narcolepsy, with a particular focus on panic and generalized anxiety disorders. Copyright 2009 Physicians Postgraduate Press, Inc.

  9. Association Study of HLA-DQB1*0602 Allele in Iranian Patients with Narcolepsy.

    PubMed

    Geremew, Demeke; Rahimi-Golkhandan, Ania; Sadeghniiat-Haghighi, Khosro; Shakiba, Yadollah; Khajeh-Mehrizi, Ahmad; Ansaripour, Bita; Izad, Maryam

    2017-10-01

    Narcolepsy is a rare, disabling disorder characterized by excessive daytime sleepiness, cataplexy, hypnagogic hallucinations and sleep paralysis. Several studies demonstrated its association with HLA-DQB1*0602 in various ethnic groups. Our study aimed to determine the prevalence of HLA-DQB1*0602 allele in Iranian patients with narcolepsy and assess its predictive parameters for diagnosing narcolepsy. In addition, car accidents and job problems were assessed among narcoleptic patients. We studied 44 narcoleptic patients, 30 patients with other types of excessive daytime sleepiness (EDS)  and 50 healthy age and sex matched individuals in this case-control study. Patients and controls filled out a questionnaire including items about car accidents due to sleepiness and job problems. International classification of sleep disorders-2 criteria was used as the gold standard for diagnosis of narcolepsy. The DNAs isolated from whole blood samples were collected from the patients and controls to assess the presence of HLA-DQB1*0602. The results showed that HLA DQB1*0602 was present in 4 (8%) individual of controls and 20 (45.5%) patients with higher prevalence in patients with cataplexy (78.9%) than patients without cataplexy (p<0.001). The sensitivities of the DQB1*0602 for diagnosing narcolepsy with cataplexy and narcolepsy without cataplexy were 78.9 and 20; specificities were 88 and 72.4, respectively. 18.2% of patients had car accidents due to sleepiness and 68.2% suffered from job problems. Our study shows that evaluation of DQB1*0602 in patients suspected to narcolepsy could be helpful especially in complex cases with atypical cataplexy and indistinguishable multiple sleep latency test MSLT results. Moreover, high rates of car accidents and job problems are found among narcoleptic patients.

  10. Diagnosis of narcolepsy and idiopathic hypersomnia. An update based on the International classification of sleep disorders, 2nd edition.

    PubMed

    Billiard, Michel

    2007-10-01

    Defining the precise nosological limits of narcolepsy and idiopathic hypersomnia is an ongoing process dating back to the first description of the two conditions. The most recent step forward has been done within the preparation of the second edition of the "International classification of sleep disorders" published in June 2005. Appointed by Dr Emmanuel Mignot, the Task Force on "Hypersomnias of central origin, not due to a circadian rhythm sleep disorder, sleep related breathing disorder, or other causes of disturbed nocturnal sleep" thoroughly revisited the nosology of narcolepsy and of idiopathic hypersomnia. Narcolepsy is now distinguished into three different entities, narcolepsy with cataplexy, narcolepsy without cataplexy and narcolepsy due to medical condition, and idiopathic hypersomnia into two entities, idiopathic hypersomnia with long sleep time and idiopathic hypersomnia without long sleep time. Nevertheless there are still a number of pending issues. What are the limits of narcolepsy without cataplexy? Is there a continuum in the pathophysiology of narcolepsy with and without cataplexy? Should sporadic and familial forms of narcolepsy with cataplexy appear as subgroups in the classification? Are idiopathic hypersomnia with long sleep time and idiopathic hypersomnia without long sleep time, two forms of the same condition or two different conditions? Is there a pathophysiological relationship between narcolepsy without cataplexy and idiopathic hypersomnia without long sleep time?

  11. Parental Fitness Questioned on the Grounds of Narcolepsy: Presentation of Two Cases

    PubMed Central

    Barbero, Laura; Govi, Annamaria; Pizza, Fabio; Plazzi, Giuseppe; Ingravallo, Francesca

    2017-01-01

    We report two cases of fathers whose parental fitness was questioned during divorce and custody litigation because of narcolepsy type 2 and type 1, respectively. These cases highlighted both the existence of a narcolepsy-related stigma and the need to involve sleep experts in custody assessments when concerns about the parental fitness are related to a sleep disorder, expanding the field of interest of the growing “sleep forensics.” Citation: Barbero L, Govi A, Pizza F, Plazzi G, Ingravallo F. Parental fitness questioned on the grounds of narcolepsy: presentation of two cases. J Clin Sleep Med. 2017;13(8):1017–1018. PMID:28728625

  12. Absence of anti-hypocretin receptor 2 autoantibodies in post pandemrix narcolepsy cases.

    PubMed

    Luo, Guo; Lin, Ling; Jacob, Louis; Bonvalet, Mélodie; Ambati, Aditya; Plazzi, Giuseppe; Pizza, Fabio; Leib, Ryan; Adams, Christopher M; Partinen, Markku; Mignot, Emmanuel Jean-Marie

    2017-01-01

    A recent publication suggested molecular mimicry of a nucleoprotein (NP) sequence from A/Puerto Rico/8/1934 (PR8) strain, the backbone used in the construction of the reassortant strain X-179A that was used in Pandemrix® vaccine, and reported on anti-hypocretin (HCRT) receptor 2 (anti-HCRTR2) autoantibodies in narcolepsy, mostly in post Pandemrix® narcolepsy cases (17 of 20 sera). In this study, we re-examined this hypothesis through mass spectrometry (MS) characterization of Pandemrix®, and two other pandemic H1N1 (pH1N1)-2009 vaccines, Arepanrix® and Focetria®, and analyzed anti-HCRTR2 autoantibodies in narcolepsy patients and controls using three independent strategies. MS characterization of Pandemrix® (2 batches), Arepanrix® (4 batches) and Focetria® (1 batch) was conducted with mapping of NP 116I or 116M spectrogram. Two sets of narcolepsy cases and controls were used: 40 post Pandemrix® narcolepsy (PP-N) cases and 18 age-matched post Pandemrix® controls (PP-C), and 48 recent (≤6 months) early onset narcolepsy (EO-N) cases and 70 age-matched other controls (O-C). Anti-HCRTR2 autoantibodies were detected using three strategies: (1) Human embryonic kidney (HEK) 293T cells with transient expression of HCRTR2 were stained with human sera and then analyzed by flow cytometer; (2) In vitro translation of [35S]-radiolabelled HCRTR2 was incubated with human sera and immune complexes of autoantibody and [35S]-radiolabelled HCRTR2 were quantified using a radioligand-binding assay; (3) Optical density (OD) at 450 nm (OD450) of human serum immunoglobulin G (IgG) binding to HCRTR2 stably expressed in Chinese hamster ovary (CHO)-K1 cell line was measured using an in-cell enzyme-linked immunosorbent assay (ELISA). NP 116M mutations were predominantly present in all batches of Pandemrix®, Arepanrix® and Focetria®. The wild-type NP109-123 (ILYDKEEIRRIWRQA), a mimic to HCRTR234-45 (YDDEEFLRYLWR), was not found to bind to DQ0602. Three or four subjects were found

  13. Absence of anti-hypocretin receptor 2 autoantibodies in post pandemrix narcolepsy cases

    PubMed Central

    Lin, Ling; Jacob, Louis; Bonvalet, Mélodie; Ambati, Aditya; Plazzi, Giuseppe; Pizza, Fabio; Leib, Ryan; Adams, Christopher M.; Partinen, Markku; Mignot, Emmanuel Jean-Marie

    2017-01-01

    Background A recent publication suggested molecular mimicry of a nucleoprotein (NP) sequence from A/Puerto Rico/8/1934 (PR8) strain, the backbone used in the construction of the reassortant strain X-179A that was used in Pandemrix® vaccine, and reported on anti-hypocretin (HCRT) receptor 2 (anti-HCRTR2) autoantibodies in narcolepsy, mostly in post Pandemrix® narcolepsy cases (17 of 20 sera). In this study, we re-examined this hypothesis through mass spectrometry (MS) characterization of Pandemrix®, and two other pandemic H1N1 (pH1N1)-2009 vaccines, Arepanrix® and Focetria®, and analyzed anti-HCRTR2 autoantibodies in narcolepsy patients and controls using three independent strategies. Methods MS characterization of Pandemrix® (2 batches), Arepanrix® (4 batches) and Focetria® (1 batch) was conducted with mapping of NP 116I or 116M spectrogram. Two sets of narcolepsy cases and controls were used: 40 post Pandemrix® narcolepsy (PP-N) cases and 18 age-matched post Pandemrix® controls (PP-C), and 48 recent (≤6 months) early onset narcolepsy (EO-N) cases and 70 age-matched other controls (O-C). Anti-HCRTR2 autoantibodies were detected using three strategies: (1) Human embryonic kidney (HEK) 293T cells with transient expression of HCRTR2 were stained with human sera and then analyzed by flow cytometer; (2) In vitro translation of [35S]-radiolabelled HCRTR2 was incubated with human sera and immune complexes of autoantibody and [35S]-radiolabelled HCRTR2 were quantified using a radioligand-binding assay; (3) Optical density (OD) at 450 nm (OD450) of human serum immunoglobulin G (IgG) binding to HCRTR2 stably expressed in Chinese hamster ovary (CHO)-K1 cell line was measured using an in-cell enzyme-linked immunosorbent assay (ELISA). Results NP 116M mutations were predominantly present in all batches of Pandemrix®, Arepanrix® and Focetria®. The wild-type NP109-123 (ILYDKEEIRRIWRQA), a mimic to HCRTR234-45 (YDDEEFLRYLWR), was not found to bind to DQ0602. Three

  14. Attention-Deficit/Hyperactivity Disorder (ADHD) Symptoms in Pediatric Narcolepsy: A Cross-Sectional Study

    PubMed Central

    Lecendreux, Michel; Lavault, Sophie; Lopez, Régis; Inocente, Clara Odilia; Konofal, Eric; Cortese, Samuele; Franco, Patricia; Arnulf, Isabelle; Dauvilliers, Yves

    2015-01-01

    Study Objectives: To evaluate the frequency, severity, and associations of symptoms of attention-deficit/hyperactivity disorder (ADHD) in children with narcolepsy with and without cataplexy. Design: Cross-sectional survey. Setting: Four French national reference centers for narcolepsy. Patients: One hundred eight consecutively referred children aged younger than 18 y with narcolepsy, with (NwC, n = 86) or without cataplexy (NwoC, n = 22), and 67 healthy controls. Interventions: The participants, their families, and sleep specialists completed a structured interview and questionnaires about sleep, daytime sleepiness, fatigue, and ADHD symptoms (ADHD-rating scale based upon Diagnostic and Statistical Manual of Mental Disorders, Fourth Edition, Text Revision [DSM-IV-TR] symptoms), and use of psychostimulants for the treatment of narcolepsy (administered in 68.2%). Polysomnographic measures were collected. Measurements and Results: Clinically significant levels of ADHD symptoms were found in 4.8% of controls compared with 35.3% in patients with NwoC (P < 0.001) and 19.7% in patients with NwC (P < 0.01). Total ADHD scores were 6.4 (95% confidence interval [CI]: 4.5, 9.0) in controls compared with 14.2 (95% CI: 10.6, 18.9; P < 0.001), in patients with NwoC and 12.2 (95% CI: 9.8, 15.3; P < 0.01) in patients with NwC; subscores of inattention and hyperactivity/impulsivity were also significantly higher in both narcolepsy groups compared with controls. No difference was found between the NwC and NwoC groups for any ADHD measure. ADHD symptom severity was associated with increased levels of sleepiness, fatigue, and insomnia. Compared with the 34 untreated patients, the 73 patients treated with psychostimulants (modafinil in 91%) showed a trend toward lower narcolepsy symptoms but not lower ADHD symptoms. Conclusions: Pediatric patients with narcolepsy have high levels of treatment-resistant attention-deficit/hyperactivity disorder (ADHD) symptoms. The optimal treatment for

  15. miRNA Profiles in Plasma from Patients with Sleep Disorders Reveal Dysregulation of miRNAs in Narcolepsy and Other Central Hypersomnias

    PubMed Central

    Holm, Anja; Bang-Berthelsen, Claus Heiner; Knudsen, Stine; Kornum, Birgitte R.; Modvig, Signe; Jennum, Poul; Gammeltoft, Steen

    2014-01-01

    Study Objectives: MicroRNAs (miRNAs) have been implicated in the pathogenesis of human diseases including neurological disorders. The aim is to address the involvement of miRNAs in the pathophysiology of central hypersomnias including autoimmune narcolepsy with cataplexy and hypocretin deficiency (type 1 narcolepsy), narcolepsy without cataplexy (type 2 narcolepsy), and idiopathic hypersomnia. Design: We conducted high-throughput analysis of miRNA in plasma from three groups of patients—with type 1 narcolepsy, type 2 narcolepsy, and idiopathic hypersomnia, respectively—in comparison with healthy controls using quantitative real-time polymerase chain reaction (qPCR) panels. Setting: University hospital based sleep clinic and research laboratories. Patients: Twelve patients with type 1 narcolepsy, 12 patients with type 2 narcolepsy, 12 patients with idiopathic hypersomnia, and 12 healthy controls. Measurements and Results: By analyzing miRNA in plasma with qPCR we identified 50, 24, and 6 miRNAs that were different in patients with type 1 narcolepsy, type 2 narcolepsy, and idiopathic hypersomnia, respectively, compared with healthy controls. Twenty miRNA candidates who fulfilled the criteria of at least two-fold difference and p-value < 0.05 were selected to validate the miRNA changes in an independent cohort of patients. Four miRNAs differed significantly between type 1 narcolepsy patients and healthy controls. Levels of miR-30c, let-7f, and miR-26a were higher, whereas the level of miR-130a was lower in type 1 narcolepsy than healthy controls. The miRNA differences were not specific for type 1 narcolepsy, since the levels of the four miRNAs were also altered in patients with type 2 narcolepsy and idiopathic hypersomnia compared with healthy controls. Conclusion: The levels of four miRNAs differed in plasma from patients with type 1 narcolepsy, type 2 narcolepsy and idiopathic hypersomnia suggesting that alterations of miRNAs may be involved in the

  16. Is there a relationship between narcolepsy, multiple sclerosis and HLA-DQB1*06:02?

    PubMed

    Lorenzoni, Paulo José; Werneck, Lineu Cesar; Crippa, Ana Christina de Souza; Zanatta, Alessandra; Kay, Cláudia S Kamoi; Silvado, Carlos Eduardo S; Scola, Rosana Herminia

    2017-06-01

    We studied multiple sclerosis (MS) patients with the HLA-DQB1*06:02 allele and compared them with MS patients who did not carry the HLA-DQB1*06:02 allele. We analyzed clinical and neurophysiological criteria for narcolepsy in six MS patients with HLA-DQB1*06:02, compared with 12 MS patients who were HLA-DQB1*06:02 non-carriers. Only two patients with HLA-DQB1*06:02 allele scored higher than 10 on the Epworth Sleepiness Scale. Polysomnography recording parameters and the multiple sleep latency test showed an absence of narcolepsy in the study group. Our study suggested no significant correlation between narcolepsy, MS and HLA-DQB1*06:02. The HLA-DQB1*06:02 allele alone was not sufficient to cause MS patients to develop narcolepsy.

  17. Effective treatment of narcolepsy-like symptoms with high-frequency repetitive transcranial magnetic stimulation: A case report.

    PubMed

    Lai, Jian-Bo; Han, Mao-Mao; Xu, Yi; Hu, Shao-Hua

    2017-11-01

    Narcolepsy is a rare sleep disorder with disrupted sleep-architecture. Clinical management of narcolepsy lies dominantly on symptom-driven pharmacotherapy. The treatment role of repetitive transcranial magnetic stimulation (rTMS) for narcolepsy remains unexplored. In this paper, we present a case of a 14-year-old young girl with excessive daytime sleepiness (EDS), cataplexy and hypnagogic hallucinations. After excluding other possible medical conditions, this patient was primarily diagnosed with narcolepsy. The patient received 25 sessions of high-frequency rTMS over the left dorsolateral prefrontal cortex (DLPFC). The symptoms of EDS and cataplexy significantly improved after rTMS treatment. Meanwhile, her score in the Epworth sleep scale (ESS) also remarkably decreased. This case indicates that rTMS may be selected as a safe and effective alternative strategy for treating narcolepsy-like symptoms. Well-designed researches are warranted in future investigations on this topic.

  18. Increased GABA Levels in Medial Prefrontal Cortex of Young Adults with Narcolepsy

    PubMed Central

    Kim, Seog Ju; Lyoo, In Kyoon; Lee, Yujin S.; Sung, Young Hoon; Kim, Hengjun J.; Kim, Jihyun H.; Kim, Kye Hyun; Jeong, Do-Un

    2008-01-01

    Study Objectives: To explore absolute concentrations of brain metabolites including gamma amino-butyric acid (GABA) in the medial prefrontal cortex and basal ganglia of young adults with narcolepsy. Design: Proton magnetic resonance (MR) spectroscopy centered on the medial prefrontal cortex and the basal ganglia was acquired. The absolute concentrations of brain metabolites including GABA and glutamate were assessed and compared between narcoleptic patients and healthy comparison subjects. Setting: Sleep and Chronobiology Center at Seoul National University Hospital; A high strength 3.0 Tesla MR scanner in the Department of Radiology at Seoul National University Hospital. Patients or Participants: Seventeen young adults with a sole diagnosis of HLA DQB1 0602 positive narcolepsy with cataplexy (25.1 ± 4.6 years old) and 17 healthy comparison subjects (26.8 ± 4.8 years old). Interventions: N/A. Measurements and Results: Relative to comparison subjects, narcoleptic patients had higher GABA concentration in the medial prefrontal cortex (t = 4.10, P <0.001). Narcoleptic patients with nocturnal sleep disturbance had higher GABA concentration in the medial prefrontal cortex than those without nocturnal sleep disturbance (t = 2.45, P= 0.03), but had lower GABA concentration than comparison subjects (t = 2.30, P = 0.03). Conclusions: The current study reports that young adults with narcolepsy had a higher GABA concentration in the medial prefrontal cortex, which was more prominent in patients without nocturnal sleep disturbance. Our findings suggest that the medial prefrontal GABA level may be increased in narcolepsy, and the increased medial prefrontal GABA might be a compensatory mechanism to reduce nocturnal sleep disturbances in narcolepsy. Citation: Kim SJ; Lyoo IK; Lee YS; Sung YH; Kim HJ; Kim JH; Kim KH; Jeong DU. Increased GABA levels in medial prefrontal cortex of young adults with narcolepsy. SLEEP 2008;31(3):342-347. PMID:18363310

  19. Aberrant Food Choices after Satiation in Human Orexin-Deficient Narcolepsy Type 1

    PubMed Central

    van Holst, Ruth Janke; van der Cruijsen, Lisa; van Mierlo, Petra; Lammers, Gert Jan; Cools, Roshan; Overeem, Sebastiaan; Aarts, Esther

    2016-01-01

    Study Objectives: Besides influencing vigilance, orexin neurotransmission serves a variety of functions, including reward, motivation, and appetite regulation. As obesity is an important symptom in orexin-deficient narcolepsy, we explored the effects of satiety on food-related choices and spontaneous snack intake in patients with narcolepsy type 1 (n = 24) compared with healthy matched controls (n = 19). In additional analyses, we also included patients with idiopathic hypersomnia (n = 14) to assess sleepiness-related influences. Methods: Participants were first trained on a choice task to earn salty and sweet snacks. Next, one of the snack outcomes was devalued by having participants consume it until satiation (i.e., sensory-specific satiety). We then measured the selective reduction in choices for the devalued snack outcome. Finally, we assessed the number of calories that participants consumed spontaneously from ad libitum available snacks afterwards. Results: After satiety, all participants reported reduced hunger and less wanting for the devalued snack. However, while controls and idiopathic hypersomnia patients chose the devalued snack less often in the choice task, patients with narcolepsy still chose the devalued snack as often as before satiety. Subsequently, narcolepsy patients spontaneously consumed almost 4 times more calories during ad libitum snack intake. Conclusions: We show that the manipulation of food-specific satiety has reduced effects on food choices and caloric intake in narcolepsy type 1 patients. These mechanisms may contribute to their obesity, and suggest an important functional role for orexin in human eating behavior. Clinical Trials Registration: Study registered at Netherlands Trial Register. URL: www.trialregister.nl. Trial ID: NTR4508. Citation: van Holst RJ, van der Cruijsen L, van Mierlo P, Lammers GJ, Cools R, Overeem S, Aarts E. Aberrant food choices after satiation in human orexindeficient narcolepsy type 1. SLEEP 2016

  20. Narcolepsy: regional cerebral blood flow during sleep and wakefulness

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Sakai, F.; Meyer, J.S.; Karacan, I.

    Serial measurements of regional cerebral blood flow were made by the 135Xe inhalation method during the early stages of sleep and wakefulness in eight normal volunteers and 12 patients with narcolepsy. Electroencephalogram, electro-oculogram, and submental electromyogram were recorded simultaneously. In normals, mean hemispheric gray matter blood flow (Fg) during stages I and II sleep was significantly less than waking values. Maximum regional blood flow decreases during sleep occurred in the brainstem-cerebellar, right inferior temporal, and bilateral frontal regions. In patients with narcolepsy, mean hemispheric Fg while awake was 80.5 +- 13 ml per 100 gm brain per minute. During REMmore » sleep, mean hemispheric Fg increased concurrently with large increases in brainstem-cerebellar region flow. During stages I and II sleep without REM, there were significant increases in mean hemispheric Fg and brainstem-cerebellar Fg, just the opposite of changes in normals. In narcolepsy, there appears to be a reversal of normal cerebral deactivation patterns, particularly involving the brainstem, during stages I and II sleep.« less

  1. New developments in the management of narcolepsy.

    PubMed

    Abad, Vivien C; Guilleminault, Christian

    2017-01-01

    Narcolepsy is a life-long, underrecognized sleep disorder that affects 0.02%-0.18% of the US and Western European populations. Genetic predisposition is suspected because of narcolepsy's strong association with HLA DQB1*06-02, and genome-wide association studies have identified polymorphisms in T-cell receptor loci. Narcolepsy pathophysiology is linked to loss of signaling by hypocretin-producing neurons; an autoimmune etiology possibly triggered by some environmental agent may precipitate hypocretin neuronal loss. Current treatment modalities alleviate the main symptoms of excessive daytime somnolence (EDS) and cataplexy and, to a lesser extent, reduce nocturnal sleep disruption, hypnagogic hallucinations, and sleep paralysis. Sodium oxybate (SXB), a sodium salt of γ hydroxybutyric acid, is a first-line agent for cataplexy and EDS and may help sleep disruption, hypnagogic hallucinations, and sleep paralysis. Various antidepressant medications including norepinephrine serotonin reuptake inhibitors, selective serotonin reuptake inhibitors, and tricyclic antidepressants are second-line agents for treating cataplexy. In addition to SXB, modafinil and armodafinil are first-line agents to treat EDS. Second-line agents for EDS are stimulants such as methylphenidate and extended-release amphetamines. Emerging therapies include non-hypocretin-based therapy, hypocretin-based treatments, and immunotherapy to prevent hypocretin neuronal death. Non-hypocretin-based novel treatments for narcolepsy include pitolisant (BF2.649, tiprolisant); JZP-110 (ADX-N05) for EDS in adults; JZP 13-005 for children; JZP-386, a deuterated sodium oxybate oral suspension; FT 218 an extended-release formulation of SXB; and JNJ-17216498, a new formulation of modafinil. Clinical trials are investigating efficacy and safety of SXB, modafinil, and armodafinil in children. γ-amino butyric acid (GABA) modulation with GABA A receptor agonists clarithromycin and flumazenil may help daytime somnolence

  2. Transient Impact of Rituximab in H1N1 Vaccination-associated Narcolepsy With Severe Psychiatric Symptoms.

    PubMed

    Sarkanen, Tomi; Alén, Reija; Partinen, Markku

    2016-09-01

    Narcolepsy type 1 is an organic sleep disorder caused by the destruction of hypocretin producing neurons in hypothalamus. In addition to daytime sleepiness, the spectrum and severity of symptoms are very variable. Psychiatric comorbidity and phenomena resembling psychotic symptoms are also common. Current treatment options for narcolepsy are symptomatic but there are few case reports of positive effect of immunotherapy. We report a very severely affected young boy treated with rituximab (RXB). A 12-year-old boy developed narcolepsy after Pandemrix H1N1 vaccination in 2010. He started to express severe psychiatric symptoms shortly after the onset. Cataplexy and sleepiness were devastatingly disabling. Conventional treatments did not have any effect on symptoms so we decided to try RXB, chimeric human monoclonal antibody against CD20 expressed in B lymphocytes. After the first treatment his condition ameliorated dramatically. Unfortunately, the effect lasted only for 2 months. Following attempts did not show any effect. Effect of RXB on narcolepsy has not been reported before. Remarkable but short-lasting effect of RXB in narcolepsy is intriguing as it could imply that there is still ongoing B cell-mediated autoimmune response possible contributing to symptoms in narcolepsy.

  3. Attention-Deficit/Hyperactivity Disorder (ADHD) Symptoms in Pediatric Narcolepsy: A Cross-Sectional Study.

    PubMed

    Lecendreux, Michel; Lavault, Sophie; Lopez, Régis; Inocente, Clara Odilia; Konofal, Eric; Cortese, Samuele; Franco, Patricia; Arnulf, Isabelle; Dauvilliers, Yves

    2015-08-01

    To evaluate the frequency, severity, and associations of symptoms of attention-deficit/hyperactivity disorder (ADHD) in children with narcolepsy with and without cataplexy. Cross-sectional survey. Four French national reference centers for narcolepsy. One hundred eight consecutively referred children aged younger than 18 y with narcolepsy, with (NwC, n = 86) or without cataplexy (NwoC, n = 22), and 67 healthy controls. The participants, their families, and sleep specialists completed a structured interview and questionnaires about sleep, daytime sleepiness, fatigue, and ADHD symptoms (ADHD-rating scale based upon Diagnostic and Statistical Manual of Mental Disorders, Fourth Edition, Text Revision [DSM-IV-TR] symptoms), and use of psychostimulants for the treatment of narcolepsy (administered in 68.2%). Polysomnographic measures were collected. Clinically significant levels of ADHD symptoms were found in 4.8% of controls compared with 35.3% in patients with NwoC (P < 0.001) and 19.7% in patients with NwC (P < 0.01). Total ADHD scores were 6.4 (95% confidence interval [CI]: 4.5, 9.0) in controls compared with 14.2 (95% CI: 10.6, 18.9; P < 0.001), in patients with NwoC and 12.2 (95% CI: 9.8, 15.3; P < 0.01) in patients with NwC; subscores of inattention and hyperactivity/impulsivity were also significantly higher in both narcolepsy groups compared with controls. No difference was found between the NwC and NwoC groups for any ADHD measure. ADHD symptom severity was associated with increased levels of sleepiness, fatigue, and insomnia. Compared with the 34 untreated patients, the 73 patients treated with psychostimulants (modafinil in 91%) showed a trend toward lower narcolepsy symptoms but not lower ADHD symptoms. Pediatric patients with narcolepsy have high levels of treatment-resistant attention-deficit/hyperactivity disorder (ADHD) symptoms. The optimal treatment for ADHD symptoms in these patients warrants further evaluation in longitudinal intervention studies.

  4. Clinical and polysomnographic course of childhood narcolepsy with cataplexy.

    PubMed

    Pizza, Fabio; Franceschini, Christian; Peltola, Hanna; Vandi, Stefano; Finotti, Elena; Ingravallo, Francesca; Nobili, Lino; Bruni, Oliviero; Lin, Ling; Edwards, Mark J; Partinen, Markku; Dauvilliers, Yves; Mignot, Emmanuel; Bhatia, Kailash P; Plazzi, Giuseppe

    2013-12-01

    Our aim was to investigate the natural evolution of cataplexy and polysomnographic features in untreated children with narcolepsy with cataplexy. To this end, clinical, polysomnographic, and cataplexy-video assessments were performed at diagnosis (mean age of 10 ± 3 and disease duration of 1 ± 1 years) and after a median follow-up of 3 years from symptom onset (mean age of 12 ± 4 years) in 21 children with narcolepsy with cataplexy and hypocretin 1 deficiency (tested in 19 subjects). Video assessment was also performed in two control groups matched for age and sex at first evaluation and follow-up and was blindly scored for presence of hypotonic (negative) and active movements. Patients' data at diagnosis and at follow-up were contrasted, compared with controls, and related with age and disease duration. At diagnosis children with narcolepsy with cataplexy showed an increase of sleep time during the 24 h; at follow-up sleep time and nocturnal sleep latency shortened, in the absence of other polysomnographic or clinical (including body mass index) changes. Hypotonic phenomena and selected facial movements decreased over time and, tested against disease duration and age, appeared as age-dependent. At onset, childhood narcolepsy with cataplexy is characterized by an abrupt increase of total sleep over the 24 h, generalized hypotonia and motor overactivity. With time, the picture of cataplexy evolves into classic presentation (i.e., brief muscle weakness episodes triggered by emotions), whereas total sleep time across the 24 h decreases, returning to more age-appropriate levels.

  5. Clinical and polysomnographic course of childhood narcolepsy with cataplexy

    PubMed Central

    Pizza, Fabio; Franceschini, Christian; Peltola, Hanna; Vandi, Stefano; Finotti, Elena; Ingravallo, Francesca; Nobili, Lino; Bruni, Oliviero; Lin, Ling; Edwards, Mark J.; Partinen, Markku; Dauvilliers, Yves; Mignot, Emmanuel; Bhatia, Kailash P.

    2013-01-01

    Our aim was to investigate the natural evolution of cataplexy and polysomnographic features in untreated children with narcolepsy with cataplexy. To this end, clinical, polysomnographic, and cataplexy-video assessments were performed at diagnosis (mean age of 10 ± 3 and disease duration of 1 ± 1 years) and after a median follow-up of 3 years from symptom onset (mean age of 12 ± 4 years) in 21 children with narcolepsy with cataplexy and hypocretin 1 deficiency (tested in 19 subjects). Video assessment was also performed in two control groups matched for age and sex at first evaluation and follow-up and was blindly scored for presence of hypotonic (negative) and active movements. Patients’ data at diagnosis and at follow-up were contrasted, compared with controls, and related with age and disease duration. At diagnosis children with narcolepsy with cataplexy showed an increase of sleep time during the 24 h; at follow-up sleep time and nocturnal sleep latency shortened, in the absence of other polysomnographic or clinical (including body mass index) changes. Hypotonic phenomena and selected facial movements decreased over time and, tested against disease duration and age, appeared as age-dependent. At onset, childhood narcolepsy with cataplexy is characterized by an abrupt increase of total sleep over the 24 h, generalized hypotonia and motor overactivity. With time, the picture of cataplexy evolves into classic presentation (i.e. brief muscle weakness episodes triggered by emotions), whereas total sleep time across the 24 h decreases, returning to more age-appropriate levels. PMID:24142146

  6. Morbidity of childhood onset narcolepsy: a controlled national study.

    PubMed

    Jennum, Poul; Pickering, Line; Thorstensen, Eva Wiberg; Ibsen, Rikke; Kjellberg, Jakob

    2017-01-01

    Narcolepsy is associated with significant morbidities. We evaluated the morbidities and mortality in a national group of child and adolescent patients after a first diagnosis of narcolepsy. Identified from the Danish National Patient Registry (NPR), 243 patients (128 boys) aged 0-19 years diagnosed with narcolepsy between 1998 and 2012 with follow-up until 2014 were compared with 970 controls who were randomly chosen from the Danish Civil Registration System Statistics and matched by age, gender and geography. Comorbidities were calculated three years before and after diagnoses. In addition to the more frequent health contacts due to neurological diseases, patients showed elevated odds ratios before and after diagnosis of endocrine and metabolic conditions (4.4 (95% CI, 1.9-10.4); 3.8 (1.7-8.4)), nervous disorders (16.6 (8.0-34.4); 198 (49.0-804)), psychiatric illnesses (4.5 (2.3-9.1)/5.8 (2.8-12.1)), pulmonary diseases, and other diseases (3.1 (2.0-4.9); 3.1 (2.0-4.9)). Congenital abnormalities (2.5 (1.1-5.5)), respiratory (2.9 (1.5-5-5)) and eye (5.7 (2.2-15.0)) diseases were more common before diagnosis. Injuries were also more common after diagnosis (1.5 (1.0-2.1)). Narcoleptic children presented significantly more diagnoses of multiple comorbidities than controls before and after diagnosis. Before and after a diagnosis of narcolepsy in children, morbidity is more frequent in several domains, including metabolic, psychiatric, neurological and other diseases. Copyright © 2016 Elsevier B.V. All rights reserved.

  7. Benefit and risk of modafinil in idiopathic hypersomnia vs. narcolepsy with cataplexy.

    PubMed

    Lavault, Sophie; Dauvilliers, Yves; Drouot, Xavier; Leu-Semenescu, Smaranda; Golmard, Jean-Louis; Lecendreux, Michel; Franco, Patricia; Arnulf, Isabelle

    2011-06-01

    The benefit/risk ratio of modafinil was recently re-evaluated by the European Medicines Agency and was shown to be negative for idiopathic hypersomnia (IH) because of insufficient data. To evaluate the benefit/risk ratio of modafinil in idiopathic hypersomnia (with and without long sleep time) vs. narcolepsy/cataplexy. The benefit (Epworth sleepiness score, ESS; visual analog scale, patient and clinician opinions) and risks (habituation, adverse effects) of modafinil were studied in a consecutive clinical cohort of 104 IH patients (59 with long sleep time) and 126 patients with narcolepsy/cataplexy. Modafinil was the first line treatment in 96-99% of patients. It produced a similar ESS change in IH patients and in narcolepsy patients (-2.6±5.1 vs. -3±5.1) and a similar benefit as estimated by the patients (6.9±2.7 vs. 6.5±2.5 on a visual analog scale) and clinicians. The ESS change was lower in IH patients with long sleep time than in those without. Sudden loss of efficacy and habituation were rare in both groups. Patients with IH reported similar but more frequent adverse effects with modafinil than narcolepsy patients: nervousness (14%), palpitations (13%), and headache (11%). Modafinil has an excellent benefit/risk ratio in idiopathic hypersomnia, with or without long sleep time, similar to its effect on narcolepsy/cataplexy. Copyright © 2011 Elsevier B.V. All rights reserved.

  8. Narcolepsy susceptibility gene CCR3 modulates sleep-wake patterns in mice.

    PubMed

    Toyoda, Hiromi; Honda, Yoshiko; Tanaka, Susumu; Miyagawa, Taku; Honda, Makoto; Honda, Kazuki; Tokunaga, Katsushi; Kodama, Tohru

    2017-01-01

    Narcolepsy is caused by the loss of hypocretin (Hcrt) neurons and is associated with multiple genetic and environmental factors. Although abnormalities in immunity are suggested to be involved in the etiology of narcolepsy, no decisive mechanism has been established. We previously reported chemokine (C-C motif) receptor 3 (CCR3) as a novel susceptibility gene for narcolepsy. To understand the role of CCR3 in the development of narcolepsy, we investigated sleep-wake patterns of Ccr3 knockout (KO) mice. Ccr3 KO mice exhibited fragmented sleep patterns in the light phase, whereas the overall sleep structure in the dark phase did not differ between Ccr3 KO mice and wild-type (WT) littermates. Intraperitoneal injection of lipopolysaccharide (LPS) promoted wakefulness and suppressed both REM and NREM sleep in the light phase in both Ccr3 KO and WT mice. Conversely, LPS suppressed wakefulness and promoted NREM sleep in the dark phase in both genotypes. After LPS administration, the proportion of time spent in wakefulness was higher, and the proportion of time spent in NREM sleep was lower in Ccr3 KO compared to WT mice only in the light phase. LPS-induced changes in sleep patterns were larger in Ccr3 KO compared to WT mice. Furthermore, we quantified the number of Hcrt neurons and found that Ccr3 KO mice had fewer Hcrt neurons in the lateral hypothalamus compared to WT mice. We found abnormalities in sleep patterns in the resting phase and in the number of Hcrt neurons in Ccr3 KO mice. These observations suggest a role for CCR3 in sleep-wake regulation in narcolepsy patients.

  9. Subjective deficits of attention, cognition and depression in patients with narcolepsy.

    PubMed

    Zamarian, Laura; Högl, Birgit; Delazer, Margarete; Hingerl, Katharina; Gabelia, David; Mitterling, Thomas; Brandauer, Elisabeth; Frauscher, Birgit

    2015-01-01

    Patients with narcolepsy often complain about attention deficits in everyday situations. In comparison with these subjective complaints, deficits in objective testing are subtler. The present study assessed the relationships between subjective complaints, objectively measured cognitive performance, disease-related variables, and mood. A total of 51 patients with narcolepsy and 35 healthy controls responded to questionnaires regarding subjectively perceived attention deficits, sleepiness, anxiety and depression. Moreover, they performed an extensive neuropsychological assessment tapping into attention, executive functions, and memory. Patients rated their level of attention in everyday situations to be relatively poor. In an objective assessment of cognitive functioning, they showed only slight attention and executive function deficits. The subjective ratings of attention deficits significantly correlated with ratings of momentary sleepiness, anxiety, and depression, but not with objectively measured cognitive performance. Momentary sleepiness and depression predicted almost 39% of the variance in the ratings of subjectively perceived attention deficits. The present study showed that sleepiness and depression, more than objective cognitive deficits, might play a role in the subjectively perceived attention deficits of patients with narcolepsy. The results suggested that when counselling and treating patients with narcolepsy, clinicians should pay attention to potential depression because subjective cognitive complaints may not relate to objective cognitive impairments. Copyright © 2014 Elsevier B.V. All rights reserved.

  10. ImmunoChip Study Implicates Antigen Presentation to T Cells in Narcolepsy

    PubMed Central

    Kornum, Birgitte Rahbek; Kenny, Eimear E.; Trynka, Gosia; Einen, Mali; Rico, Tom J.; Lichtner, Peter; Dauvilliers, Yves; Arnulf, Isabelle; Lecendreux, Michel; Javidi, Sirous; Geisler, Peter; Mayer, Geert; Pizza, Fabio; Poli, Francesca; Plazzi, Giuseppe; Overeem, Sebastiaan; Lammers, Gert Jan; Kemlink, David; Sonka, Karel; Nevsimalova, Sona; Rouleau, Guy; Desautels, Alex; Montplaisir, Jacques; Frauscher, Birgit; Ehrmann, Laura; Högl, Birgit; Jennum, Poul; Bourgin, Patrice; Peraita-Adrados, Rosa; Iranzo, Alex; Bassetti, Claudio; Chen, Wei-Min; Concannon, Patrick; Thompson, Susan D.; Damotte, Vincent; Fontaine, Bertrand; Breban, Maxime; Gieger, Christian; Klopp, Norman; Deloukas, Panos; Wijmenga, Cisca; Hallmayer, Joachim; Onengut-Gumuscu, Suna; Rich, Stephen S.; Winkelmann, Juliane; Mignot, Emmanuel

    2013-01-01

    Recent advances in the identification of susceptibility genes and environmental exposures provide broad support for a post-infectious autoimmune basis for narcolepsy/hypocretin (orexin) deficiency. We genotyped loci associated with other autoimmune and inflammatory diseases in 1,886 individuals with hypocretin-deficient narcolepsy and 10,421 controls, all of European ancestry, using a custom genotyping array (ImmunoChip). Three loci located outside the Human Leukocyte Antigen (HLA) region on chromosome 6 were significantly associated with disease risk. In addition to a strong signal in the T cell receptor alpha (TRA@), variants in two additional narcolepsy loci, Cathepsin H (CTSH) and Tumor necrosis factor (ligand) superfamily member 4 (TNFSF4, also called OX40L), attained genome-wide significance. These findings underline the importance of antigen presentation by HLA Class II to T cells in the pathophysiology of this autoimmune disease. PMID:23459209

  11. Intravenous Immunoglobulin Therapy in Pediatric Narcolepsy: A Nonrandomized, Open-Label, Controlled, Longitudinal Observational Study

    PubMed Central

    Lecendreux, Michel; Berthier, Johanna; Corny, Jennifer; Bourdon, Olivier; Dossier, Claire; Delclaux, Christophe

    2017-01-01

    Study Objectives: Previous case reports of intravenous immunoglobulins (IVIg) in pediatric narcolepsy have shown contradictory results. Methods: This was a nonrandomized, open-label, controlled, longitudinal observational study of IVIg use in pediatric narcolepsy with retrospective data collection from medical files obtained from a single pediatric national reference center for the treatment of narcolepsy in France. Of 56 consecutively referred patients with narcolepsy, 24 received IVIg (3 infusions administered at 1-mo intervals) in addition to standard care (psychostimulants and/or anticataplectic agents), and 32 continued on standard care alone (controls). Results: For two patients in each group, medical files were unavailable. Of the 22 IVIg patients, all had cerebrospinal fluid (CSF) hypocretin ≤ 110 pg/mL and were HLA-DQB1*06:02 positive. Of the 30 control patients, 29 were HLA-DQB1*06:02 positive and of those with available CSF measurements, all 12 had hypocretin ≤ 110 pg/mL. Compared with control patients, IVIg patients had shorter disease duration, shorter latency to sleep onset, and more had received H1N1 vaccination. Mean (standard deviation) follow-up length was 2.4 (1.1) y in the IVIg group and 3.9 (1.7) y in controls. In multivariate-adjusted linear mixed-effects analyses of change from baseline in Ullanlinna Narcolepsy Scale (UNS) scores, high baseline UNS, but not IVIg treatment, was associated with a reduction in narcolepsy symptoms. On time-to-event analysis, among patients with high baseline UNS scores, control patients achieved a UNS score < 14 (indicating remission) less rapidly than IVIg patients (adjusted hazard ratio 0.18; 95% confidence interval: 95% confidence interval: 0.03, 0.95; p = 0.043). Shorter or longer disease duration did not influence treatment response in any analysis. Conclusions: Overall, narcolepsy symptoms were not significantly reduced by IVIg. However, in patients with high baseline symptoms, a subset of IVIg

  12. Mechanism of action of narcolepsy medications.

    PubMed

    Gowda, Chandan R; Lundt, Leslie P

    2014-12-01

    The medications used to treat narcolepsy are targeted toward alleviating symptoms such as excessive sleepiness and cataplexy. The cause of this neurological sleep disorder is still not completely clear, though a destruction of hypocretin/orexin neurons has been implicated. The destruction of these neurons is linked to inactivity of neurotransmitters including histamine, norepinephrine, acetylcholine, and serotonin, causing a disturbance in the sleep/wake cycles of narcoleptic patients. Stimulants and MAOIs have traditionally been used to counteract excessive daytime sleepiness and sleep attacks by inhibiting the breakdown of catecholamines. Newer drugs, called wake-promoting agents, have recently become first-line agents due to their better side-effect profile, efficacy, and lesser potential for abuse. These agents similarly inhibit reuptake of dopamine, but have a novel mechanism of action, as they have been found to increase neuronal activity in the tuberomamillary nucleus and in orexin neurons. Sodium oxybate, a sodium salt of gamma-hydroxybutyrate (GHB), is another class that is used to treat many symptoms of narcolepsy, and is the only U.S. Food and Drug Administration (FDA)-approved medication for cataplexy. It has a different mechanism of action than either stimulants or wake-promoting agents, as it binds to its own unique receptor. Antidepressants, like selective serotonin re-uptake inhibitors (SSRIs) and tricyclic antidepressants (TCAs), have also been used, as similar to stimulants, they inhibit reuptake of specific catecholamines. In this article, we seek to review the mechanisms behind these classes of drugs in relation to the proposed pathophysiology of narcolepsy. Appropriate clinical strategies will be discussed, including specific combinations of medications that have been shown to be effective.

  13. Aberrant Food Choices after Satiation in Human Orexin-Deficient Narcolepsy Type 1.

    PubMed

    van Holst, Ruth Janke; van der Cruijsen, Lisa; van Mierlo, Petra; Lammers, Gert Jan; Cools, Roshan; Overeem, Sebastiaan; Aarts, Esther

    2016-11-01

    Besides influencing vigilance, orexin neurotransmission serves a variety of functions, including reward, motivation, and appetite regulation. As obesity is an important symptom in orexin-deficient narcolepsy, we explored the effects of satiety on food-related choices and spontaneous snack intake in patients with narcolepsy type 1 (n = 24) compared with healthy matched controls (n = 19). In additional analyses, we also included patients with idiopathic hypersomnia (n = 14) to assess sleepiness-related influences. Participants were first trained on a choice task to earn salty and sweet snacks. Next, one of the snack outcomes was devalued by having participants consume it until satiation (i.e., sensory-specific satiety). We then measured the selective reduction in choices for the devalued snack outcome. Finally, we assessed the number of calories that participants consumed spontaneously from ad libitum available snacks afterwards. After satiety, all participants reported reduced hunger and less wanting for the devalued snack. However, while controls and idiopathic hypersomnia patients chose the devalued snack less often in the choice task, patients with narcolepsy still chose the devalued snack as often as before satiety. Subsequently, narcolepsy patients spontaneously consumed almost 4 times more calories during ad libitum snack intake. We show that the manipulation of food-specific satiety has reduced effects on food choices and caloric intake in narcolepsy type 1 patients. These mechanisms may contribute to their obesity, and suggest an important functional role for orexin in human eating behavior. Study registered at Netherlands Trial Register. URL: www.trialregister.nl. Trial ID: NTR4508. © 2016 Associated Professional Sleep Societies, LLC.

  14. Body weight and basal metabolic rate in childhood narcolepsy: a longitudinal study.

    PubMed

    Wang, Zongwen; Wu, Huijuan; Stone, William S; Zhuang, Jianhua; Qiu, Linli; Xu, Xing; Wang, Yan; Zhao, Zhengqing; Han, Fang; Zhao, Zhongxin

    2016-09-01

    The aims of this study were to document the trajectory of weight gain and body mass index (BMI) in children with type 1 narcolepsy, and to analyze basal metabolic rate (BMR). A total of 65 Chinese children with type 1 narcolepsy with a disease duration ≤12 months were included. In addition, 79 healthy age-matched students were enrolled as controls. Height and body weight were measured every six months for up to 36 months to calculate BMI growth. BMR was measured using COSMED K4b2 indirect calorimetry in 34 patients and 30 healthy controls at six months. At the end of 36 months, the BMR was compared among 18 patients and 16 healthy controls. The children with type 1 narcolepsy showed higher BMIs at follow-up assessments. At the end of the study, 38.46% of the patients were obese and an additional 26.15% were overweight. The patients' BMI growth at six, 12, 18, 24 and 30 months of follow-up was significantly higher, but not at month 36. The patients' basal energy expenditure was significantly lower than that of the controls at six months but not at 36 months. BMI increased rapidly in children with type 1 narcolepsy after disease onset, but BMI growth decreased gradually with prolonged disease. Decreased BMR is an important cause underlying rapid weight gain. The gradual restoration of BMI growth and BMR in narcolepsy emphasizes the importance of compensatory metabolic mechanisms in this disease. Copyright © 2016 Elsevier B.V. All rights reserved.

  15. Two cases of childhood narcolepsy mimicking epileptic seizures in video-EEG/EMG.

    PubMed

    Yanagishita, Tomoe; Ito, Susumu; Ohtani, Yui; Eto, Kaoru; Kanbayashi, Takashi; Oguni, Hirokazu; Nagata, Satoru

    2018-06-06

    Narcolepsy is characterized by excessive sleepiness, hypnagogic hallucinations, and sleep paralysis, and can occur with or without cataplexy. Here, we report two children with narcolepsy presenting with cataplexy mimicking epileptic seizures as determined by long-term video-electroencephalography (EEG) and electromyography (EMG) monitoring. Case 1 was a 15-year-old girl presenting with recurrent episodes of "convulsions" and loss of consciousness, who was referred to our hospital with a diagnosis of epilepsy showing "convulsions" and "complex partial seizures". The long-term video-polygraph showed a clonic attack lasting for 15 s, which corresponded to 1-2 Hz with interruption of mentalis EMG discharges lasting for 70-300 ms without any EEG changes. Narcolepsy was suspected due to the attack induced by hearty laughs and the presence of sleep attacks, and confirmed by low orexin levels in cerebrospinal fluid (CSF). Case 2 was an 11-year-old girl presenting with recurrent episodes of myoclonic attacks simultaneously with dropping objects immediately after hearty laughs, in addition to sleep attacks, hypnagogic hallucinations, and sleep paralysis. The long-term video-polygraph showed a subtle attack, characterized by dropping chopsticks from her hand, which corresponded to an interruption of ongoing deltoid EMG discharges lasting 140 ms without any EEG changes. A diagnosis of narcolepsy was confirmed by the low orexin levels in CSF. These cases demonstrate that children with narcolepsy may have attacks of cataplexy that resemble clonic or myoclonic seizures. Copyright © 2018 The Japanese Society of Child Neurology. Published by Elsevier B.V. All rights reserved.

  16. Narcolepsy and Predictors of Positive MSLTs in the Wisconsin Sleep Cohort

    PubMed Central

    Goldbart, Aviv; Peppard, Paul; Finn, Laurel; Ruoff, Chad M.; Barnet, Jodi; Young, Terry; Mignot, Emmanuel

    2014-01-01

    Study Objectives: To study whether positive multiple sleep latency tests (MSLTs, mean sleep latency [MSL] ≤ 8 minutes, ≥ 2 sleep onset REM sleep periods [SOREMPs]) and/or nocturnal SOREMP (REM sleep latency ≤ 15 minutes during nocturnal polysomonography [NPSG]) are stable traits and can reflect incipient narcolepsy. Design and Setting: Cross-sectional and longitudinal investigation of the Wisconsin Sleep Cohort Study. Participants: Adults (44% females, 30-81 years) underwent NPSG (n = 4,866 in 1,518 subjects), and clinical MSLT (n = 1,135), with 823 having a repeat NPSG-MSLT at 4-year intervals, totaling 1725 NPSG with MSLT studies. Data were analyzed using linear mixed-effects models, and the stability of positive MSLTs was explored using κ statistics. Measurements and Results: Prevalence of a nocturnal SOREMP on a NPSG, of ≥ 2 SOREMPs on the MSLT, of MSL ≤ 8 minutes on the MSLT, and of a positive MSLT (MSL ≤ 8 minutes plus ≥ 2 SOREMPs) were 0.35%, 7.0%, 22%, and 3.4%, respectively. Correlates of a positive MSLT were shift work (OR = 7.8, P = 0.0001) and short sleep (OR = 1.51/h, P = 0.04). Test-retest for these parameters was poor, with κ < 0.2 (n.s.) after excluding shift workers and short sleepers. Excluding shift-work, short sleep, and subjects with negative MSLTs, we found one undiagnosed subject with possible cataplexy (≥ 1/month) and a NPSG SOREMPs; one subject previously diagnosed with narcolepsy without cataplexy with 2 NPSG SOREMPs and a positive MSLT, and two subjects with 2 independently positive MSLTs (66% human leukocyte antigen [HLA] positive). The proportions for narcolepsy with and without cataplexy were 0.07% (95% CI: 0.02-0.37%) and 0.20% (95% CI: 0.07-0.58%), respectively. Conclusions: The diagnostic value of multiple sleep latency tests is strongly altered by shift work and to a lesser extent by chronic sleep deprivation. The prevalence of narcolepsy without cataplexy may be 3-fold higher than that of narcolepsy

  17. Comparison of driving simulator performance and neuropsychological testing in narcolepsy.

    PubMed

    Kotterba, Sylvia; Mueller, Nicole; Leidag, Markus; Widdig, Walter; Rasche, Kurt; Malin, Jean-Pierre; Schultze-Werninghaus, Gerhard; Orth, Maritta

    2004-09-01

    Daytime sleepiness and cataplexy can increase automobile accident rates in narcolepsy. Several countries have produced guidelines for issuing a driving license. The aim of the study was to compare driving simulator performance and neuropsychological test results in narcolepsy in order to evaluate their predictive value regarding driving ability. Thirteen patients with narcolepsy (age: 41.5+/-12.9 years) and 10 healthy control patients (age: 55.1+/-7.8 years) were investigated. By computer-assisted neuropsychological testing, vigilance, alertness and divided attention were assessed. In a driving simulator patients and controls had to drive on a highway for 60 min (mean speed of 100 km/h). Different weather and daytime conditions and obstacles were presented. Epworth Sleepiness Scale-Scores were significantly raised (narcolepsy patients: 16.7+/-5.1, controls: 6.6+/-3.6, P < or = 0.001). The accident rate of the control patients increased (3.2+/-1.8 versus 1.3+/-1.5, P < or = 0.01). Significant differences in concentration lapses (e.g. tracking errors and deviation from speed limit) could not be revealed (9.8+/-3.5 versus 7.1+/-3.2, pns). Follow-up investigation in five patients after an optimising therapy could demonstrate the decrease in accidents due to concentration lapses (P < or = 0.05). Neuropsychological testing (expressed as percentage compared to a standardised control population) revealed deficits in alertness (32.3+/-28.6). Mean percentage scores of divided attention (56.9+/-25.4) and vigilance (58.7+/-26.8) were in a normal range. There was, however, a high inter-individual difference. There was no correlation between driving performance and neuropsychological test results or ESS Score. Neuropsychological test results did not significantly change in the follow-up. The difficulties encountered by the narcolepsy patient in remaining alert may account for sleep-related motor vehicle accidents. Driving simulator investigations are closely related to real

  18. Novel method for evaluation of eye movements in patients with narcolepsy.

    PubMed

    Christensen, Julie A E; Kempfner, Lykke; Leonthin, Helle L; Hvidtfelt, Mathias; Nikolic, Miki; Kornum, Birgitte Rahbek; Jennum, Poul

    2017-05-01

    Narcolepsy causes abnormalities in the control of wake-sleep, non-rapid-eye-movement (non-REM) sleep and REM sleep, which includes specific eye movements (EMs). In this study, we aim to evaluate EM characteristics in narcolepsy as compared to controls using an automated detector. We developed a data-driven method to detect EMs during sleep based on two EOG signals recorded as part of a polysomnography (PSG). The method was optimized using the manually scored hypnograms from 36 control subjects. The detector was applied on a clinical sample with subjects suspected for central hypersomnias. Based on PSG, multiple sleep latency test and cerebrospinal fluid hypocretin-1 measures, they were divided into clinical controls (N = 20), narcolepsy type 2 (NT2, N = 19), and narcolepsy type 1 (NT1, N = 28). We investigated the distribution of EMs across sleep stages and cycles. NT1 patients had significantly less EMs during wake, N1, and N2 sleep and more EMs during REM sleep compared to clinical controls, and significantly less EMs during wake and N1 sleep compared to NT2 patients. Furthermore, NT1 patients showed less EMs during NREM sleep in the first sleep cycle and more EMs during NREM sleep in the second sleep cycle compared to clinical controls and NT2 patients. NT1 patients show an altered distribution of EMs across sleep stages and cycles compared to NT2 patients and clinical controls, suggesting that EMs are directly or indirectly controlled by the hypocretinergic system. A data-driven EM detector may contribute to the evaluation of narcolepsy and other disorders involving the control of EMs. Copyright © 2016 Elsevier B.V. All rights reserved.

  19. Incidence of narcolepsy before and after MF59-adjuvanted influenza A(H1N1)pdm09 vaccination in South Korean soldiers.

    PubMed

    Kim, Woo Jung; Lee, Sang Don; Lee, Eun; Namkoong, Kee; Choe, Kang-Won; Song, Joon Young; Cheong, Hee Jin; Jeong, Hye Won; Heo, Jung Yeon

    2015-09-11

    Previous reports mostly from Europe suggested an association between an occurrence of narcolepsy and an influenza A(H1N1)pdm09 vaccine adjuvanted with AS03 (Pandemrix(®)). During the 2009 H1N1 pandemic vaccination campaign, the Korean military performed a vaccination campaign with one type of influenza vaccine containing MF59-adjuvants. This study was conducted to investigate the background incidence rate of narcolepsy in South Korean soldiers and the association of the MF59-adjuvanted vaccine with the occurrence of narcolepsy in a young adult group. To assess the incidence of narcolepsy, we retrospectively reviewed medical records of suspicious cases of narcolepsy in 2007-2013 in the whole 20 military hospitals of the Korean military. The screened cases were classified according to the Brighton Collaboration case definition of narcolepsy. After obtaining the number of confirmed cases of narcolepsy per 3 months in 2007-2013, we compared the crude incidence rate of narcolepsy before and after the vaccination campaign. We included 218 narcolepsy suspicious cases in the initial review, which were screened by the diagnostic code on the computerized disease registry in 2007-2013. Forty-one cases were finally diagnosed with narcolepsy in 2007-2013 (male sex, 95%; median age, 21 years). The average background incidence rate of narcolepsy in Korean soldiers was 0.91 cases per 100,000 persons per year. During the 9 months before vaccination implementation (April to December 2009), 6 narcolepsy cases occurred, whereas during the next 9 months (January to September 2010) including the 3-month vaccination campaign, 5 cases occurred. The incidence of narcolepsy in South Korean soldiers was not increased after the pandemic vaccination campaign using the MF59-adjuvanted vaccine. Our results suggest that the MF59-adjuvanted H1N1 vaccine did not contribute to the occurrence of narcolepsy in this young adult group. Copyright © 2015 Elsevier Ltd. All rights reserved.

  20. Cardiovascular fitness in narcolepsy is inversely related to sleepiness and the number of cataplexy episodes.

    PubMed

    Matoulek, Martin; Tuka, Vladimír; Fialová, Magdalena; Nevšímalová, Soňa; Šonka, Karel

    2017-06-01

    Cardiopulmonary fitness depends on daily energy expenditure or the amount of daily exercise. Patients with narcolepsy spent more time being sleepy or asleep than controls; thus we may speculate that they have a lower quantity and quality of physical activity. The aim of the present study was thus to test the hypothesis that exercise tolerance in narcolepsy negatively depends on sleepiness. The cross-sectional study included 32 patients with narcolepsy with cataplexy, 10 patients with narcolepsy without cataplexy, and 36 age- and gender-matched control subjects, in whom a symptom-limited exercise stress test with expired gas analysis was performed. A linear regression analysis with multivariate models was used with stepwise variable selection. In narcolepsy patients, maximal oxygen uptake (VO 2peak ) was 30.1 ± 7.5 mL/kg/min, which was lower than 36.0 ± 7.8 mL/kg/min, p = 0.001, in controls and corresponded to 86.4% ± 20.0% of the population norm (VO 2peak %) and to a standard deviation (VO 2peak SD) of -1.08 ± 1.63 mL/kg/min of the population norm. VO 2peak depended primarily on gender (p = 0.007) and on sleepiness (p = 0.046). VO 2peak % depended on sleepiness (p = 0.028) and on age (p = 0.039). VO 2peak SD depended on the number of cataplexy episodes per month (p = 0.015) and on age (p = 0.030). Cardiopulmonary fitness in narcolepsy and in narcolepsy without cataplexy is inversely related to the degree of sleepiness and cataplexy episode frequency. Copyright © 2017 Elsevier B.V. All rights reserved.

  1. Intravenous Immunoglobulin Therapy in Pediatric Narcolepsy: A Nonrandomized, Open-Label, Controlled, Longitudinal Observational Study.

    PubMed

    Lecendreux, Michel; Berthier, Johanna; Corny, Jennifer; Bourdon, Olivier; Dossier, Claire; Delclaux, Christophe

    2017-03-15

    Previous case reports of intravenous immunoglobulins (IVIg) in pediatric narcolepsy have shown contradictory results. This was a nonrandomized, open-label, controlled, longitudinal observational study of IVIg use in pediatric narcolepsy with retrospective data collection from medical files obtained from a single pediatric national reference center for the treatment of narcolepsy in France. Of 56 consecutively referred patients with narcolepsy, 24 received IVIg (3 infusions administered at 1-mo intervals) in addition to standard care (psychostimulants and/or anticataplectic agents), and 32 continued on standard care alone (controls). For two patients in each group, medical files were unavailable. Of the 22 IVIg patients, all had cerebrospinal fluid (CSF) hypocretin ≤ 110 pg/mL and were HLA-DQB1*06:02 positive. Of the 30 control patients, 29 were HLA-DQB1*06:02 positive and of those with available CSF measurements, all 12 had hypocretin ≤ 110 pg/mL. Compared with control patients, IVIg patients had shorter disease duration, shorter latency to sleep onset, and more had received H1N1 vaccination. Mean (standard deviation) follow-up length was 2.4 (1.1) y in the IVIg group and 3.9 (1.7) y in controls. In multivariate-adjusted linear mixed-effects analyses of change from baseline in Ullanlinna Narcolepsy Scale (UNS) scores, high baseline UNS, but not IVIg treatment, was associated with a reduction in narcolepsy symptoms. On time-to-event analysis, among patients with high baseline UNS scores, control patients achieved a UNS score < 14 (indicating remission) less rapidly than IVIg patients (adjusted hazard ratio 0.18; 95% confidence interval: 95% confidence interval: 0.03, 0.95; p = 0.043). Shorter or longer disease duration did not influence treatment response in any analysis. Overall, narcolepsy symptoms were not significantly reduced by IVIg. However, in patients with high baseline symptoms, a subset of IVIg-treated patients achieved remission more rapidly than

  2. Nocturnal Sleep Dynamics Identify Narcolepsy Type 1

    PubMed Central

    Pizza, Fabio; Vandi, Stefano; Iloti, Martina; Franceschini, Christian; Liguori, Rocco; Mignot, Emmanuel; Plazzi, Giuseppe

    2015-01-01

    Study Objectives: To evaluate the reliability of nocturnal sleep dynamics in the differential diagnosis of central disorders of hypersomnolence. Design: Cross-sectional. Setting: Sleep laboratory. Patients: One hundred seventy-five patients with hypocretin-deficient narcolepsy type 1 (NT1, n = 79), narcolepsy type 2 (NT2, n = 22), idiopathic hypersomnia (IH, n = 22), and “subjective” hypersomnolence (sHS, n = 52). Interventions: None. Methods: Polysomnographic (PSG) work-up included 48 h of continuous PSG recording. From nocturnal PSG conventional sleep macrostructure, occurrence of sleep onset rapid eye movement period (SOREMP), sleep stages distribution, and sleep stage transitions were calculated. Patient groups were compared, and receiver operating characteristic (ROC) curve analysis was used to test the diagnostic utility of nocturnal PSG data to identify NT1. Results: Sleep macrostructure was substantially stable in the 2 nights of each diagnostic group. NT1 and NT2 patients had lower latency to rapid eye movement (REM) sleep, and NT1 patients showed the highest number of awakenings, sleep stage transitions, and more time spent in N1 sleep, as well as most SOREMPs at daytime PSG and at multiple sleep latency test (MSLT) than all other groups. ROC curve analysis showed that nocturnal SOREMP (area under the curve of 0.724 ± 0.041, P < 0.0001), percent of total sleep time spent in N1 (0.896 ± 0.023, P < 0.0001), and the wakefulness-sleep transition index (0.796 ± 0.034, P < 0.0001) had a good sensitivity and specificity profile to identify NT1 sleep, especially when used in combination (0.903 ± 0.023, P < 0.0001), similarly to SOREMP number at continuous daytime PSG (0.899 ± 0.026, P < 0.0001) and at MSLT (0.956 ± 0.015, P < 0.0001). Conclusions: Sleep macrostructure (i.e. SOREMP, N1 timing) including stage transitions reliably identifies hypocretin-deficient narcolepsy type 1 among central disorders of hypersomnolence. Citation: Pizza F, Vandi S

  3. Genome-wide analysis of CNV (copy number variation) and their associations with narcolepsy in a Japanese population.

    PubMed

    Yamasaki, Maria; Miyagawa, Taku; Toyoda, Hiromi; Khor, Seik-Soon; Koike, Asako; Nitta, Aino; Akiyama, Kumi; Sasaki, Tsukasa; Honda, Yutaka; Honda, Makoto; Tokunaga, Katsushi

    2014-05-01

    In humans, narcolepsy with cataplexy (narcolepsy) is a sleep disorder that is characterized by sleepiness, cataplexy and rapid eye movement (REM) sleep abnormalities. Narcolepsy is caused by a reduction in the number of neurons that produce hypocretin (orexin) neuropeptide. Both genetic and environmental factors contribute to the development of narcolepsy.Rare and large copy number variations (CNVs) reportedly play a role in the etiology of a number of neuropsychiatric disorders. Narcolepsy is considered a neurological disorder; therefore, we sought to investigate any possible association between rare and large CNVs and human narcolepsy. We used DNA microarray data and a CNV detection software application, PennCNV-Affy, to detect CNVs in 426 Japanese narcoleptic patients and 562 healthy individuals. Overall, we found a significant enrichment of rare and large CNVs (frequency ≤1%, size ≥100 kb) in the patients (case-control ratio of CNV count=1.54, P=5.00 × 10(-4)). Next, we extended a region-based association analysis by including CNVs with its size ≥30 kb. Rare and large CNVs in PARK2 region showed a significant association with narcolepsy. Four patients were assessed to carry duplications of the gene region, whereas no controls carried the duplication, which was further confirmed by quantitative PCR assay. This duplication was also found in 2 essential hypersomnia (EHS) patients out of 171 patients. Furthermore, a pathway analysis revealed enrichments of gene disruptions by rare and large CNVs in immune response, acetyltransferase activity, cell cycle regulation and regulation of cell development. This study constitutes the first report on the risk association between multiple rare and large CNVs and the pathogenesis of narcolepsy. In the future, replication studies are needed to confirm the associations.

  4. Nocturnal Dynamics of Sleep-Wake Transitions in Patients With Narcolepsy.

    PubMed

    Zhang, Xiaozhe; Kantelhardt, Jan W; Dong, Xiao Song; Krefting, Dagmar; Li, Jing; Yan, Han; Pillmann, Frank; Fietze, Ingo; Penzel, Thomas; Zhao, Long; Han, Fang

    2017-02-01

    We investigate how characteristics of sleep-wake dynamics in humans are modified by narcolepsy, a clinical condition that is supposed to destabilize sleep-wake regulation. Subjects with and without cataplexy are considered separately. Differences in sleep scoring habits as a possible confounder have been examined. Four groups of subjects are considered: narcolepsy patients from China with (n = 88) and without (n = 15) cataplexy, healthy controls from China (n = 110) and from Europe (n = 187, 2 nights each). After sleep-stage scoring and calculation of sleep characteristic parameters, the distributions of wake-episode durations and sleep-episode durations are determined for each group and fitted by power laws (exponent α) and by exponentials (decay time τ). We find that wake duration distributions are consistent with power laws for healthy subjects (China: α = 0.88, Europe: α = 1.02). Wake durations in all groups of narcolepsy patients, however, follow the exponential law (τ = 6.2-8.1 min). All sleep duration distributions are best fitted by exponentials on long time scales (τ = 34-82 min). We conclude that narcolepsy mainly alters the control of wake-episode durations but not sleep-episode durations, irrespective of cataplexy. Observed distributions of shortest wake and sleep durations suggest that differences in scoring habits regarding the scoring of short-term sleep stages may notably influence the fitting parameters but do not affect the main conclusion. © Sleep Research Society 2016. Published by Oxford University Press on behalf of the Sleep Research Society. All rights reserved. For permissions, please e-mail journals.permissions@oup.com.

  5. Narcolepsy with and without cataplexy, idiopathic hypersomnia with and without long sleep time: a cluster analysis.

    PubMed

    Šonka, Karel; Šusta, Marek; Billiard, Michel

    2015-02-01

    The successive editions of the International Classification of Sleep Disorders (ICSD) reflect the evolution of the concepts of various sleep disorders. This is particularly the case for central disorders of hypersomnolence, with continuous changes in terminology and divisions of narcolepsy, idiopathic hypersomnia, and recurrent hypersomnia. According to the ICSD 2nd Edition (ICSD-2), narcolepsy with cataplexy (NwithC), narcolepsy without cataplexy (Nw/oC), idiopathic hypersomnia with long sleep time (IHwithLST), and idiopathic hypersomnia without long sleep time (IHw/oLST) are four, well-defined hypersomnias of central origin. However, in the absence of biological markers, doubts have been raised as to the relevance of a division of idiopathic hypersomnia into two forms, and it is not yet clear whether Nw/oC and IHw/oLST are two distinct entities. With this in mind, it was decided to empirically review the ICSD-2 classification by using a hierarchical cluster analysis to see whether this division has some relevance, even though the terms "with long sleep time" and "without long sleep time" are inappropriate. The cluster analysis differentiated three main clusters: Cluster 1, "combined monosymptomatic hypersomnia/narcolepsy type 2" (people initially diagnosed with IHw/oLST and Nw/oC); Cluster 2 "polysymptomatic hypersomnia" (people initially diagnosed with IHwithLST); and Cluster 3, narcolepsy type 1 (people initially diagnosed with NwithC). Cluster analysis confirmed that narcolepsy type 1 and polysymptomatic hypersomnia are independent sleep disorders. People who were initially diagnosed with Nw/oC and IHw/oLST formed a single cluster, referred to as "combined monosymptomatic hypersomnia/narcolepsy type 2." Copyright © 2014 Elsevier B.V. All rights reserved.

  6. White matter alterations in narcolepsy patients with cataplexy: tract-based spatial statistics.

    PubMed

    Park, Yun K; Kwon, Oh-Hun; Joo, Eun Yeon; Kim, Jae-Hun; Lee, Jong M; Kim, Sung T; Hong, Seung B

    2016-04-01

    Functional imaging studies and voxel-based morphometry analysis of brain magnetic resonance imaging showed abnormalities in the hypothalamus-thalamus-orbitofrontal pathway, demonstrating altered hypocretin pathway in narcolepsy. Those distinct morphometric changes account for problems in wake-sleep control, attention and memory. It also raised the necessity to evaluate white matter changes. To investigate brain white matter alterations in drug-naïve narcolepsy patients with cataplexy and to explore relationships between white matter changes and patient clinical characteristics, drug-naïve narcolepsy patients with cataplexy (n = 22) and healthy age- and gender-matched controls (n = 26) were studied. Fractional anisotropy and mean diffusivity images were obtained from whole-brain diffusion tensor imaging, and tract-based spatial statistics were used to localize white matter abnormalities. Compared with controls, patients showed significant decreases in fractional anisotropy of white matter of the bilateral anterior cingulate, fronto-orbital area, frontal lobe, anterior limb of the internal capsule and corpus callosum, as well as the left anterior and medial thalamus. Patients and controls showed no differences in mean diffusivity. Among patients, mean diffusivity values of white matter in the bilateral superior frontal gyri, bilateral fronto-orbital gyri and right superior parietal gyrus were positively correlated with depressive mood. This tract-based spatial statistics study demonstrated that drug-naïve patients with narcolepsy had reduced fractional anisotropy of white matter in multiple brain areas and significant relationship between increased mean diffusivity of white matter in frontal/cingulate and depression. It suggests the widespread disruption of white matter integrity and prevalent brain degeneration of frontal lobes according to a depressive symptom in narcolepsy. © 2015 European Sleep Research Society.

  7. Executive Control of Attention in Narcolepsy

    PubMed Central

    Bayard, Sophie; Croisier Langenier, Muriel; Cochen De Cock, Valérie; Scholz, Sabine; Dauvilliers, Yves

    2012-01-01

    Background Narcolepsy with cataplexy (NC) is a disabling sleep disorder characterized by early loss of hypocretin neurons that project to areas involved in the attention network. We characterized the executive control of attention in drug-free patients with NC to determine whether the executive deficits observed in patients with NC are specific to the disease itself or whether they reflect performance changes due to the severity of excessive daytime sleepiness. Methodology Twenty-two patients with NC compared to 22 patients with narcolepsy without cataplexy (NwC) matched for age, gender, intellectual level, objective daytime sleepiness and number of sleep onset REM periods (SOREMPs) were studied. Thirty-two matched healthy controls were included. All participants underwent a standardized interview, completed questionnaires, and neuropsychological tests. All patients underwent a polysomnography followed by multiple sleep latency tests (MSLT), with neuropsychological evaluation performed the same day between MSLT sessions. Principal Findings Irrespective of diagnosis, patients reported higher self-reported attentional complaints associated with the intensity of depressive symptoms. Patients with NC performed slower and more variably on simple reaction time tasks than patients with NwC, who did not differ from controls. Patients with NC and NwC generally performed slower, reacted more variably, and made more errors than controls on executive functioning tests. Individual profile analyses showed a clear heterogeneity of the severity of executive deficit. This severity was related to objective sleepiness, higher number of SOREMPs on the MSLT, and lower intelligence quotient. The nature and severity of the executive deficits were unrelated to NC and NwC diagnosis. Conclusions We demonstrated that drug-free patients with NC and NwC complained of attention deficit, with altered executive control of attention being explained by the severity of objective sleepiness and

  8. [Narcolepsy in childhood and adolescence: symptoms, diagnosis, and therapy. A case report].

    PubMed

    Gehrmann, Jochen; Siegler, Dominik; Ignacy, Evelin; Reimer, Inga

    2017-03-01

    Narcolepsy is a rare, multifactorial disease of the hypothalamus characterized by its leading symptoms of excessive daytime sleepiness and cataplexy. Sleep-EEG and a HLA-DR-genotype serve to secure the diagnosis. We report here on a 14-year-old girl suffering from anxieties, depression, school refusal, social withdrawal as well as very frequent attacks of sleep during the day and cataplexy. Currently, there is no approved drug for children and adolescents suffering from narcolepsy. Our patient benefited significantly and quickly from an off-label treatment with methylphenidate in combination with psychoeducation, cognitive behavioral therapy, and family therapy. Narcolepsy is a very rare but probably underestimated differential diagnosis applied to unclear daytime sleepiness, anxieties, or depression in childhood and adolescence. Both the key symptoms and the comorbid symptoms improve significantly under treatment with stimulants, albeit at a higher dosage.

  9. An autoantibody in narcolepsy disrupts colonic migrating motor complexes.

    PubMed

    Jackson, Michael W; Reed, Joanne H; Smith, Anthony J F; Gordon, Tom P

    2008-12-03

    Despite strong circumstantial evidence for the autoimmune hypothesis of narcolepsy, conventional immunological methods have failed to detect an autoantibody. This study investigated the real-time effects of narcoleptic immunoglobulins on a spontaneous colonic migrating motor complex (CMMC) preparation. IgG from patients with narcolepsy with cataplexy or healthy controls was added directly to isolated mouse colons undergoing CMMC activity to test for autoantibodies that disrupt colonic motility. The effect of immunoglobulins prepared for clinical intravenous treatment (IVIg) on autoantibody-mediated colonic disruption was also assessed. Narcoleptic IgGs markedly reduced the frequency of CMMCs or irreversibly abolished them. Abrogation of CMMCs was followed by an increase in the resting tension of the colon preparation and appearance of atropine-sensitive phasic smooth muscle contractions. IVIg partially neutralized the inhibitory effect of narcoleptic IgG on the CMMCs. The dramatic effect of narcoleptic IgG on CMMC generation is consistent with an autoantibody-mediated disruption of enteric neural pathways. The ex vivo whole-organ approach allows real-time examination of the physiological effects of the narcoleptic autoantibody and offers a new avenue for exploring the autoimmune basis of narcolepsy. The neutralizing effect of IVIg on the autoantibody provides a rationale for the reported clinical improvement in cataplexy when IVIg are given at disease onset.

  10. Narcolepsy Type 1 and Idiopathic Generalized Epilepsy: Diagnostic and Therapeutic Challenges in Dual Cases

    PubMed Central

    Baiardi, Simone; Vandi, Stefano; Pizza, Fabio; Alvisi, Lara; Toscani, Lucia; Zambrelli, Elena; Tinuper, Paolo; Mayer, Geert; Plazzi, Giuseppe

    2015-01-01

    Study Objectives: The aim of this study is to describe the possible co-occurrence of narcolepsy type 1 and generalized epilepsy, focusing on diagnostic challenge and safety of dual treatments. Methods and Results: Four patients with comorbidity for narcolepsy type 1 and idiopathic generalized epilepsy are reported: in three cases the onset of epilepsy preceded narcolepsy type 1 appearance, whereas in one case epileptic spells onset was subsequent. Patients presented with absences, myoclonic and tonic-clonic seizure type: in the patient with tonic-clonic seizures the dual pathology was easily recognized, in the other cases the first diagnosis caused the comorbid disease to be overlooked, independent of the time-course sequence. All four patients underwent neurological examination, video-electroencephalogram during which ictal and interictal epileptic discharges were recorded, and sleep polysomnographic studies. Repeated sleep onset rapid eye movement periods (SOREMPs) were documented with the multiple sleep latency test (MLST) in all the four cases. All patients had unremarkable brain magnetic resonance imaging studies and cerebrospinal hypocretin-1 was assessed in two patients, revealing undetectable levels. The association of antiepileptic drugs and substances currently used to treat narcolepsy type 1, including sodium oxybate, was effective in improving seizures, sleep disturbance, and cataplexy. Conclusions: Narcolepsy type 1 may occur in association with idiopathic generalized epilepsy, leading to remarkable diagnostic and therapeutic challenges. Electrophysiological studies as well as a comprehensive somnologic interview can help confirm the diagnosis in patients with ambiguous neurological history. Sodium oxybate in combination with antiepileptic drugs is safe and effective in treating cataplexy and excessive daytime sleepiness. Citation: Baiardi S, Vandi S, Pizza F, Alvisi L, Toscani L, Zambrelli E, Tinuper P, Mayer G, Plazzi G. Narcolepsy type 1 and

  11. Incidence of narcolepsy after H1N1 influenza and vaccinations: Systematic review and meta-analysis.

    PubMed

    Sarkanen, Tomi O; Alakuijala, Anniina P E; Dauvilliers, Yves A; Partinen, Markku M

    2018-04-01

    An increased incidence of narcolepsy was seen in many countries after the pandemic H1N1 influenza vaccination campaign in 2009-2010. The H1N1 vaccine - narcolepsy connection is based on observational studies that are prone to various biases, e.g., confounding by H1N1 infection, and ascertainment, recall and selection biases. A direct pathogenic link has, however, remained elusive. We conducted a systematic review and meta-analysis to analyze the magnitude of H1N1 vaccination related risk and to examine if there was any association with H1N1 infection itself. We searched all articles from PubMed, Web of Science and Scopus, and other relevant sources reporting the incidence and risk of post-vaccine narcolepsy. In our paper, we show that the risk appears to be limited to only one vaccine (Pandemrix ® ). During the first year after vaccination, the relative risk of narcolepsy was increased 5 to 14-fold in children and adolescents and 2 to 7-fold in adults. The vaccine attributable risk in children and adolescents was around 1 per 18,400 vaccine doses. Studies from Finland and Sweden also appear to demonstrate an extended risk of narcolepsy into the second year following vaccination, but such conclusions should be interpreted with a word of caution due to possible biases. Benefits of immunization outweigh the risk of vaccination-associated narcolepsy, which remains a rare disease. Copyright © 2017 Elsevier Ltd. All rights reserved.

  12. Case report of narcolepsy in a six-year-old child initially misdiagnosed as atypical epilepsy.

    PubMed

    Zhou, Jinquan; Zhang, Xi; Dong, Zaiwen

    2014-08-01

    This report describes a case of first-onset narcolepsy in a six-year-old female that was misdiagnosed as atypical epilepsy and other diagnoses at eight different hospitals over a period of 10 months before the correct diagnosis was made. The diagnosis of narcolepsy is more difficult in children because very few of them experience all four cardinal symptoms of narcolepsy - paroxysmal sleep, cataplexy, hypnagogic hallucination, and sleep paralysis - and they often have a more prolonged onset and diverse symptoms. To decrease the time lag between initial presentation and accurate diagnosis, we recommend that in all cases in which children report excessive sleep of unknown etiology - regardless of the associated symptoms - that sleep monitoring and sleep latency tests be conducted to rule out the possibility of narcolepsy. The case highlights the wide variety of presentations of uncommon psychiatric conditions, particularly in children, and the need for clinicians to be aware of the atypical presentations of these conditions when collecting medical histories.

  13. Increased risk of narcolepsy in children and adults after pandemic H1N1 vaccination in France.

    PubMed

    Dauvilliers, Yves; Arnulf, Isabelle; Lecendreux, Michel; Monaca Charley, Christelle; Franco, Patricia; Drouot, Xavier; d'Ortho, Marie-Pia; Launois, Sandrine; Lignot, Séverine; Bourgin, Patrice; Nogues, Béatrice; Rey, Marc; Bayard, Sophie; Scholz, Sabine; Lavault, Sophie; Tubert-Bitter, Pascale; Saussier, Cristel; Pariente, Antoine

    2013-08-01

    An increased incidence of narcolepsy in children was detected in Scandinavian countries where pandemic H1N1 influenza ASO3-adjuvanted vaccine was used. A campaign of vaccination against pandemic H1N1 influenza was implemented in France using both ASO3-adjuvanted and non-adjuvanted vaccines. As part of a study considering all-type narcolepsy, we investigated the association between H1N1 vaccination and narcolepsy with cataplexy in children and adults compared with matched controls; and compared the phenotype of narcolepsy with cataplexy according to exposure to the H1N1 vaccination. Patients with narcolepsy-cataplexy were included from 14 expert centres in France. Date of diagnosis constituted the index date. Validation of cases was performed by independent experts using the Brighton collaboration criteria. Up to four controls were individually matched to cases according to age, gender and geographic location. A structured telephone interview was performed to collect information on medical history, past infections and vaccinations. Eighty-five cases with narcolepsy-cataplexy were included; 23 being further excluded regarding eligibility criteria. Of the 62 eligible cases, 59 (64% males, 57.6% children) could be matched with 135 control subjects. H1N1 vaccination was associated with narcolepsy-cataplexy with an odds ratio of 6.5 (2.1-19.9) in subjects aged<18 years, and 4.7 (1.6-13.9) in those aged 18 and over. Sensitivity analyses considering date of referral for diagnosis or the date of onset of symptoms as the index date gave similar results, as did analyses focusing only on exposure to ASO3-adjuvanted vaccine. Slight differences were found when comparing cases with narcolepsy-cataplexy exposed to H1N1 vaccination (n=32; mostly AS03-adjuvanted vaccine, n=28) to non-exposed cases (n=30), including shorter delay of diagnosis and a higher number of sleep onset rapid eye movement periods for exposed cases. No difference was found regarding history of infections. In

  14. Case report of narcolepsy in a six-year-old child initially misdiagnosed as atypical epilepsy

    PubMed Central

    ZHOU, Jinquan; ZHANG, Xi; DONG, Zaiwen

    2014-01-01

    Summary This report describes a case of first-onset narcolepsy in a six-year-old female that was misdiagnosed as atypical epilepsy and other diagnoses at eight different hospitals over a period of 10 months before the correct diagnosis was made. The diagnosis of narcolepsy is more difficult in children because very few of them experience all four cardinal symptoms of narcolepsy – paroxysmal sleep, cataplexy, hypnagogic hallucination, and sleep paralysis – and they often have a more prolonged onset and diverse symptoms. To decrease the time lag between initial presentation and accurate diagnosis, we recommend that in all cases in which children report excessive sleep of unknown etiology – regardless of the associated symptoms – that sleep monitoring and sleep latency tests be conducted to rule out the possibility of narcolepsy. The case highlights the wide variety of presentations of uncommon psychiatric conditions, particularly in children, and the need for clinicians to be aware of the atypical presentations of these conditions when collecting medical histories. PMID:25317010

  15. Narcolepsy Type 1 and Idiopathic Generalized Epilepsy: Diagnostic and Therapeutic Challenges in Dual Cases.

    PubMed

    Baiardi, Simone; Vandi, Stefano; Pizza, Fabio; Alvisi, Lara; Toscani, Lucia; Zambrelli, Elena; Tinuper, Paolo; Mayer, Geert; Plazzi, Giuseppe

    2015-11-15

    The aim of this study is to describe the possible co-occurrence of narcolepsy type 1 and generalized epilepsy, focusing on diagnostic challenge and safety of dual treatments. Four patients with comorbidity for narcolepsy type 1 and idiopathic generalized epilepsy are reported: in three cases the onset of epilepsy preceded narcolepsy type 1 appearance, whereas in one case epileptic spells onset was subsequent. Patients presented with absences, myoclonic and tonic-clonic seizure type: in the patient with tonic-clonic seizures the dual pathology was easily recognized, in the other cases the first diagnosis caused the comorbid disease to be overlooked, independent of the time-course sequence. All four patients underwent neurological examination, video-electroencephalogram during which ictal and interictal epileptic discharges were recorded, and sleep polysomnographic studies. Repeated sleep onset rapid eye movement periods (SOREMPs) were documented with the multiple sleep latency test (MLST) in all the four cases. All patients had unremarkable brain magnetic resonance imaging studies and cerebrospinal hypocretin-1 was assessed in two patients, revealing undetectable levels. The association of antiepileptic drugs and substances currently used to treat narcolepsy type 1, including sodium oxybate, was effective in improving seizures, sleep disturbance, and cataplexy. Narcolepsy type 1 may occur in association with idiopathic generalized epilepsy, leading to remarkable diagnostic and therapeutic challenges. Electrophysiological studies as well as a comprehensive somnologic interview can help confirm the diagnosis in patients with ambiguous neurological history. Sodium oxybate in combination with antiepileptic drugs is safe and effective in treating cataplexy and excessive daytime sleepiness. © 2015 American Academy of Sleep Medicine.

  16. Predictors of Hypocretin (Orexin) Deficiency in Narcolepsy Without Cataplexy

    PubMed Central

    Andlauer, Olivier; Moore, Hyatt; Hong, Seung-Chul; Dauvilliers, Yves; Kanbayashi, Takashi; Nishino, Seiji; Han, Fang; Silber, Michael H.; Rico, Tom; Einen, Mali; Kornum, Birgitte R.; Jennum, Poul; Knudsen, Stine; Nevsimalova, Sona; Poli, Francesca; Plazzi, Giuseppe; Mignot, Emmanuel

    2012-01-01

    Study Objectives: To compare clinical, electrophysiologic, and biologic data in narcolepsy without cataplexy with low (≤ 110 pg/ml), intermediate (110–200 pg/ml), and normal (> 200 pg/ml) concentrations of cerebrospinal fluid (CSF) hypocretin-1. Setting: University-based sleep clinics and laboratories. Patients: Narcolepsy without cataplexy (n = 171) and control patients (n = 170), all with available CSF hypocretin-1. Design and interventions: Retrospective comparison and receiver operating characteristics curve analysis. Patients were also recontacted to evaluate if they developed cataplexy by survival curve analysis. Measurements and Results: The optimal cutoff of CSF hypocretin-1 for narcolepsy without cataplexy diagnosis was 200 pg/ml rather than 110 pg/ml (sensitivity 33%, specificity 99%). Forty-one patients (24%), all HLA DQB1*06:02 positive, had low concentrations (≤ 110 pg/ml) of CSF hypocretin-1. Patients with low concentrations of hypocretin-1 only differed subjectively from other groups by a higher Epworth Sleepiness Scale score and more frequent sleep paralysis. Compared with patients with normal hypocretin-1 concentration (n = 117, 68%), those with low hypocretin-1 concentration had higher HLA DQB1*06:02 frequencies, were more frequently non-Caucasians (notably African Americans), with lower age of onset, and longer duration of illness. They also had more frequently short rapid-eye movement (REM) sleep latency (≤ 15 min) during polysomnography (64% versus 23%), and shorter sleep latencies (2.7 ± 0.3 versus 4.4 ± 0.2 min) and more sleep-onset REM periods (3.6 ± 0.1 versus 2.9 ± 0.1 min) during the Multiple Sleep Latency Test (MSLT). Patients with intermediate concentrations of CSF hypocretin-1 (n = 13, 8%) had intermediate HLA DQB1*06:02 and polysomnography results, suggesting heterogeneity. Of the 127 patients we were able to recontact, survival analysis showed that almost half (48%) with low concentration of CSF hypocretin-1 had developed

  17. Neuronal Antibodies in Children with or without Narcolepsy following H1N1-AS03 Vaccination

    PubMed Central

    Thebault, Simon; Waters, Patrick; Snape, Matthew D.; Cottrell, Dominic; Darin, Niklas; Hallböök, Tove; Huutoniemi, Anne; Partinen, Markku; Pollard, Andrew J.; Vincent, Angela

    2015-01-01

    Type 1 narcolepsy is caused by deficiency of hypothalamic orexin/hypocretin. An autoimmune basis is suspected, but no specific antibodies, either causative or as biomarkers, have been identified. However, the AS03 adjuvanted split virion H1N1 (H1N1-AS03) vaccine, created to protect against the 2009 Pandemic, has been implicated as a trigger of narcolepsy particularly in children. Sera and CSFs from 13 H1N1-AS03-vaccinated patients (12 children, 1 young adult) with type 1 narcolepsy were tested for autoantibodies to known neuronal antigens including the N-methyl-D-aspartate receptor (NMDAR) and contactin-associated protein 2 (CASPR2), both associated with encephalopathies that include disordered sleep, to rodent brain tissue including the lateral hypothalamus, and to live hippocampal neurons in culture. When sufficient sample was available, CSF levels of melanin-concentrating hormone (MCH) were measured. Sera from 44 H1N1-ASO3-vaccinated children without narcolepsy were also examined. None of these patients’ CSFs or sera was positive for NMDAR or CASPR2 antibodies or binding to neurons; 4/13 sera bound to orexin-neurons in rat brain tissue, but also to other neurons. MCH levels were a marginally raised (n = 8; p = 0.054) in orexin-deficient narcolepsy patients compared with orexin-normal children (n = 6). In the 44 H1N1-AS03-vaccinated healthy children, there was no rise in total IgG levels or in CASPR2 or NMDAR antibodies three weeks following vaccination. In conclusion, there were no narcolepsy-specific autoantibodies identified in type 1 narcolepsy sera or CSFs, and no evidence for a general increase in immune reactivity following H1N1-AS03 vaccination in the healthy children. Antibodies to other neuronal specific membrane targets, with their potential for directing use of immunotherapies, are still an important goal for future research. PMID:26090827

  18. Neuronal Antibodies in Children with or without Narcolepsy following H1N1-AS03 Vaccination.

    PubMed

    Thebault, Simon; Waters, Patrick; Snape, Matthew D; Cottrell, Dominic; Darin, Niklas; Hallböök, Tove; Huutoniemi, Anne; Partinen, Markku; Pollard, Andrew J; Vincent, Angela

    2015-01-01

    Type 1 narcolepsy is caused by deficiency of hypothalamic orexin/hypocretin. An autoimmune basis is suspected, but no specific antibodies, either causative or as biomarkers, have been identified. However, the AS03 adjuvanted split virion H1N1 (H1N1-AS03) vaccine, created to protect against the 2009 Pandemic, has been implicated as a trigger of narcolepsy particularly in children. Sera and CSFs from 13 H1N1-AS03-vaccinated patients (12 children, 1 young adult) with type 1 narcolepsy were tested for autoantibodies to known neuronal antigens including the N-methyl-D-aspartate receptor (NMDAR) and contactin-associated protein 2 (CASPR2), both associated with encephalopathies that include disordered sleep, to rodent brain tissue including the lateral hypothalamus, and to live hippocampal neurons in culture. When sufficient sample was available, CSF levels of melanin-concentrating hormone (MCH) were measured. Sera from 44 H1N1-ASO3-vaccinated children without narcolepsy were also examined. None of these patients' CSFs or sera was positive for NMDAR or CASPR2 antibodies or binding to neurons; 4/13 sera bound to orexin-neurons in rat brain tissue, but also to other neurons. MCH levels were a marginally raised (n = 8; p = 0.054) in orexin-deficient narcolepsy patients compared with orexin-normal children (n = 6). In the 44 H1N1-AS03-vaccinated healthy children, there was no rise in total IgG levels or in CASPR2 or NMDAR antibodies three weeks following vaccination. In conclusion, there were no narcolepsy-specific autoantibodies identified in type 1 narcolepsy sera or CSFs, and no evidence for a general increase in immune reactivity following H1N1-AS03 vaccination in the healthy children. Antibodies to other neuronal specific membrane targets, with their potential for directing use of immunotherapies, are still an important goal for future research.

  19. Health-Related Quality of Life Among Drug-Naïve Patients with Narcolepsy with Cataplexy, Narcolepsy Without Cataplexy, and Idiopathic Hypersomnia Without Long Sleep Time

    PubMed Central

    Ozaki, Akiko; Inoue, Yuichi; Nakajima, Toru; Hayashida, Kenichi; Honda, Makoto; Komada, Yoko; Takahashi, Kiyohisa

    2008-01-01

    Objective: To evaluate the health-related quality life (HRQOL) of drugnaïve patients with narcolepsy with cataplexy (NA with CA), narcolepsy without cataplexy (NA without CA) and idiopathic hypersomnia without long sleep time (IHS without LST), and to explore the factors influencing the HRQOL. Factors associated with the occurrence of automobile accidents are also discussed. Methods: A total of 137 consecutive drug naïve patients who met the criteria of the 2nd edition of the International Classification of Sleep Disorders (NA with CA, n = 28; NA without CA, n = 27; IHS without LST, n = 82) were enrolled. The patients were asked to fill out questionnaires, including the SF-36, Epworth Sleepiness Scale (ESS), sociodemographic variables, and items regarding driving habits and the experiences related to automobile accidents. Results: All 3 diagnostic groups had significantly lower scores in most SF-36 domains compared with Japanese normative data. Significant differences among the 3 diagnostic groups were not observed. Specific factors in SF-36 domains were not found with multiple linear regression analyses, while disease duration was positively correlated with mental health among all subjects. Among the patients reporting driving habits, ESS score (≥16) was positively associated with the experience of automobile accidents. Conclusions: Our results indicated that HRQOL decreases in drugnaïve patients with hypersomnia, but neither disease category nor severity of the disorder appears as an associated factor. Increased severity of hypersomnia, however, was thought to play an important role in the occurrence of automobile accidents. Citation: Ozaki A; Inoue Y; Nakajima T; Hayashida K; Honda M; Komada Y; Takahashi K. Health-related quality of life among drug-naïve patients with narcolepsy with cataplexy, narcolepsy without cataplexy, and idiopathic hypersomnia without long sleep time. J Clin Sleep Med 2008;4(6):572-578. PMID:19110887

  20. Narcolepsy and syndromes of primary excessive daytime somnolence.

    PubMed

    Black, Jed E; Brooks, Stephen N; Nishino, Seiji

    2004-09-01

    Excessive daytime sleepiness (EDS) or somnolence is common in our patients and in society in general. The most common cause of EDS is "voluntary" sleep restriction. Other common causes include sleep-fragmenting disorders such as the obstructive sleep apnea syndrome. Somewhat less familiar to the clinician are EDS conditions arising from central nervous system dysfunction. Of these so-called primary disorders of somnolence, narcolepsy is the most well known and extensively studied, yet often misunderstood and misdiagnosed. Idiopathic hypersomnia, the recurrent hypersomnias, and EDS associated with nervous system disorders also must be well-understood to provide appropriate evaluation and management of the patient with EDS. This review summarizes the distinguishing features of these clinical syndromes of primary EDS. A brief overview of the pharmacological management of primary EDS is included. Finally, in view of the tremendous advances that have occurred in the past few years in our understanding of the pathophysiology of canine and human narcolepsy, we also highlight these discoveries.

  1. Cataplectic facies: clinical marker in the diagnosis of childhood narcolepsy-report of two cases.

    PubMed

    Prasad, Manish; Setty, Gururaj; Ponnusamy, Athi; Hussain, Nahin; Desurkar, Archana

    2014-05-01

    Narcolepsy is a chronic disease and is commonly diagnosed in adulthood. However, more than half of the patients have onset of symptoms in childhood and/or adolescence. The full spectrum of clinical manifestations, namely excessive daytime sleepiness, cataplexy, hypnagogic hallucinations, and sleep paralysis, is usually not present at disease onset, delaying diagnosis during childhood. Mean delay in diagnosis since symptom onset is known to be several years. Initial manifestations can sometimes be as subtle as only partial drooping of eyelids leading to confusion with a myasthenic condition. We present two children who presented with "cataplectic facies," an unusual facial feature only recently described in children with narcolepsy with cataplexy. The diagnosis of narcolepsy was confirmed by multiple sleep latency test along with human leukocyte antigen typing and cerebrospinal fluid hypocretin assay. The diagnosis of narcolepsy with cataplexy at onset can be challenging in young children. With more awareness of subtle signs such as cataplectic facies, earlier diagnosis is possible. To date, only 11 children between 6 and 18 years of age presenting with typical cataplectic facies have been reported in the literature. We present two patients, one of whom is the youngest individual (4 years old) yet described with the typical cataplectic facies. Copyright © 2014 Elsevier Inc. All rights reserved.

  2. Antigenic Differences between AS03 Adjuvanted Influenza A (H1N1) Pandemic Vaccines: Implications for Pandemrix-Associated Narcolepsy Risk

    PubMed Central

    Vaarala, Outi; Vuorela, Arja; Partinen, Markku; Baumann, Marc; Freitag, Tobias L.; Meri, Seppo; Saavalainen, Päivi; Jauhiainen, Matti; Soliymani, Rabah; Kirjavainen, Turkka; Olsen, Päivi; Saarenpää-Heikkilä, Outi; Rouvinen, Juha; Roivainen, Merja; Nohynek, Hanna; Jokinen, Jukka; Julkunen, Ilkka; Kilpi, Terhi

    2014-01-01

    Background Narcolepsy results from immune-mediated destruction of hypocretin secreting neurons in hypothalamus, however the triggers and disease mechanisms are poorly understood. Vaccine-attributable risk of narcolepsy reported so far with the AS03 adjuvanted H1N1 vaccination Pandemrix has been manifold compared to the AS03 adjuvanted Arepanrix, which contained differently produced H1N1 viral antigen preparation. Hence, antigenic differences and antibody response to these vaccines were investigated. Methods and Findings Increased circulating IgG-antibody levels to Pandemrix H1N1 antigen were found in 47 children with Pandemrix-associated narcolepsy when compared to 57 healthy children vaccinated with Pandemrix. H1N1 antigen of Arepanrix inhibited poorly these antibodies indicating antigenic difference between Arepanrix and Pandemrix. High-resolution gel electrophoresis quantitation and mass spectrometry identification analyses revealed higher amounts of structurally altered viral nucleoprotein (NP) in Pandemrix. Increased antibody levels to hemagglutinin (HA) and NP, particularly to detergent treated NP, was seen in narcolepsy. Higher levels of antibodies to NP were found in children with DQB1*06∶02 risk allele and in DQB1*06∶02 transgenic mice immunized with Pandemrix when compared to controls. Conclusions This work identified 1) higher amounts of structurally altered viral NP in Pandemrix than in Arepanrix, 2) detergent-induced antigenic changes of viral NP, that are recognized by antibodies from children with narcolepsy, and 3) increased antibody response to NP in association of DQB1*06∶02 risk allele of narcolepsy. These findings provide a link between Pandemrix and narcolepsy. Although detailed mechanisms of Pandemrix in narcolepsy remain elusive, our results move the focus from adjuvant(s) onto the H1N1 viral proteins. PMID:25501681

  3. Sodium Oxybate treatment in pediatric type 1 narcolepsy.

    PubMed

    Moresco, Monica; Pizza, Fabio; Antelmi, Elena; Plazzi, Giuseppe

    2018-03-05

    Narcolepsy type 1 (NT1) is a rare chronic neurologic disorder characterized by excessive daytime sleepiness, cataplexy, sleep paralysis, hallucinations and disrupted nocturnal sleep, usually with onset during childhood/adolescence. Pediatric NT1 is associated with limitations on children's activities and achievements, especially poor performance at school, difficulty with peers due to disease symptoms and comorbidities including depression, obesity, and precocious puberty. NT1 disease is caused by the selective loss of hypocretin-producing neurons in the lateral hypothalamus, most probably related to an autoimmune pathophysiology. Indeed a strong genetic predisposition including the HLA system and other associations in genes involved in immune responses has been found together with the triggering role of environmental agents such as H1N1 influenza infections and vaccinations. Sodium Oxybate (SO) is a sodium salt of γ-hydroxybutyric (GHB) acid that is synthetized by neurons in the brain and functions as neurotransmitter. GHB is a central nervous system depressant and produces dose-dependent sedation. SO is a first line medication for cataplexy and excessive daytime sleepiness in adults with NT1, but can be helpful also for sleep disruption, hypnagogic hallucination and sleep paralysis in these patients. Although in the majority of patients narcolepsy develops before 15 years of age, there are no approved treatments for pediatric NT1. However, SO has been widely used off-label to treat narcolepsy symptoms in children and adolescents with NT1 in non-controlled studies, showing a similar safety profile and therapeutic response to adult patients. Current therapy is based only on empirical data shared among expert sleep disorders clinicians. Copyright© Bentham Science Publishers; For any queries, please email at epub@benthamscience.org.

  4. Changes in Medical Services and Drug Utilization and Associated Costs After Narcolepsy Diagnosis in the United States

    PubMed Central

    Villa, Kathleen F.; Reaven, Nancy L.; Funk, Susan E.; McGaughey, Karen; Black, Jed

    2018-01-01

    Background Healthcare utilization and the cost implications associated with undiagnosed and/or misdiagnosed narcolepsy have not been evaluated, and there is scant literature characterizing the newly diagnosed population with narcolepsy with respect to treatment patterns and resource utilization. Objective To analyze the changes in medication use, healthcare utilization, and the associated costs after a new diagnosis of narcolepsy. Methods In this retrospective cohort study, we used data from the Truven Health Analytics MarketScan Research Databases, between January 2006 and March 2013, to identify patients who had a probable new diagnosis of narcolepsy—defined as a de novo medical claim for a multiple sleep latency test—which was preceded by ≥6 months of continuous insurance and was followed by a de novo diagnosis of narcolepsy. The utilization and cost of medical services and the percentage of patients filling prescriptions for narcolepsy-related medications were evaluated in 3 consecutive 1-year periods from the date of a positive multiple sleep latency test result (ie, index date), and each year's findings were compared with the annualized results from the 6-month preindex period. Results A total of 3757 patients who met the definition of a new diagnosis of narcolepsy were identified. The total medical service utilization decreased each year from a preindex average of 28.2 visits per patient per year (PPPY) to 26.9 visits (P <.05), 23.1 visits (P <.0001), and 22.5 visits (P <.0001) PPPY in years 1, 2, and 3 postdiagnosis, respectively. In each outpatient service category, the medical services utilization decreased from preindex to year 3 postdiagnosis, including hospital outpatient and physician visits (P <.0001), and other outpatient and emergency department visits (P <.05). The percentage of patients receiving narcolepsy-related medications increased from 54.0% preindex to 77.4%, 70.0%, and 66.9% for years 1, 2, and 3 postindex (all P <.0001 vs preindex

  5. Clinical course of H1N1-vaccine-related narcolepsy.

    PubMed

    Sarkanen, Tomi; Alakuijala, Anniina; Partinen, Markku

    2016-03-01

    To follow and analyze the clinical course and quality of life of Pandemrix H1N1-vaccine-related narcolepsy (pNT1). Twenty-six drug-naïve confirmed pNT1 subjects completed Epworth Sleepiness Scale (ESS), Ullanlinna Narcolepsy Scale (UNS), Swiss Narcolepsy Scale (SNS), Rimon's Brief Depression scale (RDS), and WHO-5 Well-being index questionnaires near the disease onset and in a follow-up a minimum of two years later. The number of cataplexies and body mass index (BMI) were recorded. The effects of hypocretin-1 levels and sleep recording results were analyzed. The findings at the follow-up visit were compared with 25 non-vaccine-related type 1 narcolepsy (NT1) subjects. In pNT1, RDS score decreased significantly (mean 10.2, SD 4.7 vs mean 6.7, SD 4.5, p = 0.003). Median of BMI increased from 20.8 kg m(-2) to 23.4 kg m(-2), p <0.001. There were no significant differences in other sleep scores. However, deviation and range in questionnaire scores at the follow-up were wide. Subjects with very low or undetectable hypocretin-1 levels had worse scores in UNS (mean 26.4, SD 6.95 vs mean 19.1, SD 3.83, p = 0.006) and ESS (mean 17.9, SD = 4.29 vs mean 14.1, SD = 3.70, p = 0.047) than those with hypocretin-1 levels of 20-110 pg/mL. Most disabling symptoms were excessive daytime sleepiness and disturbed sleep. There were no significant differences between the scores in pNT1 and NT1. Clinical course of pNT1 is heterogeneous but the evolution of pNT1 seems similar to NT1. Lower hypocretin levels in pNT1 are associated with a more severe phenotype. Copyright © 2015 Elsevier B.V. All rights reserved.

  6. Delayed Diagnosis, Range of Severity, and Multiple Sleep Comorbidities: A Clinical and Polysomnographic Analysis of 100 Patients of the Innsbruck Narcolepsy Cohort

    PubMed Central

    Frauscher, Birgit; Ehrmann, Laura; Mitterling, Thomas; Gabelia, David; Gschliesser, Viola; Brandauer, Elisabeth; Poewe, Werner; Högl, Birgit

    2013-01-01

    Study Objectives: Narcolepsy is reported to affect 26-56/100,000 in the general population. We aimed to describe clinical and polysomnographic features of a large narcolepsy cohort in order to comprehensively characterize the narcoleptic spectrum. Methods: We performed a chart- and polysomnographybased review of all narcolepsy patients of the Innsbruck narcolepsy cohort. Results: A total of 100 consecutive narcolepsy patients (87 with cataplexy [NC], 13 without cataplexy [N]) were included in the analysis. All subjects had either excessive daytime sleepiness or cataplexy as their initial presenting clinical feature. Age at symptom onset was 20 (6-69) years. Diagnostic delay was 6.5 (0-39) years. The complete narcolepsy tetrad was present in 36/100 patients; 28/100 patients had three cardinal symptoms; 29/100 had two; and 7/100 had only excessive daytime sleepiness. Severity varied broadly with respect to excessive daytime sleepiness (median Epworth Sleepiness Scale score: 18, range 10-24), cataplexy (8-point Likert scale: median 4.5, range 1-8), hypnagogic hallucinations (median 4.5, range 1-7), and sleep paralysis (median 3, range 1-7). Sleep comorbidity was highly prevalent and ranged from sleeprelated movement disorders (n = 55/100), parasomnias (n = 34/100), and sleeprelated breathing disorders (n = 24/100), to insomnia (n = 28/100). REM sleep without atonia or a periodic limb movement in sleep index > 5/h were present in most patients (90/100 and 75/100). A high percentage of narcoleptic patients in the present study had high frequency leg movements (35%) and excessive fragmentary myoclonus (22%). Of the narcolepsy patients with clinical features of REM sleep behavior disorder (RBD), 76.5% had EMG evidence for RBD on the multiple sleep latency test (MSLT), based on a standard cutoff of a minimum of 18% of 3-sec miniepochs. Conclusion: This study is one of the largest monocentric polysomnographic studies to date of patients with narcolepsy and confirms the

  7. Associations Between Neuropsychological, Neurobehavioral and Emotional Functioning and Either Narcolepsy or Idiopathic Hypersomnia in Children and Adolescents.

    PubMed

    Ludwig, Beris; Smith, Simon; Heussler, Helen

    2018-04-15

    Narcolepsy and idiopathic hypersomnia are chronic neurological sleep disorders characterized by hypersomnolence or excessive daytime sleepiness. This review aims to systematically examine the scientific literature on the associations between narcolepsy and idiopathic hypersomnia and their effect on intellectual functioning, academic achievement, behavior, and emotion. Published studies that examined those associations in children and adolescents were included. Studies in which children or adolescents received a clinical diagnosis, and in which the associated function was measured with at least one objective instrument were included. Twenty studies published between 1968 and 2017 were eligible for inclusion in this review. There does not appear to be a clear association between intellectual functioning and narcolepsy or idiopathic hypersomnia; however, limited research is an obstacle to obtaining generalizability. The variability in results from studies investigating associations between academic achievement and these two hypersomnolence disorders suggests that further research using standardized and validated assessment instruments is required to determine if there is an association. Behavior and emotion appear to be significantly affected by narcolepsy. Only two studies included populations of children and adolescents with idiopathic hypersomnia. Further research using larger populations of children and adolescents with narcolepsy or idiopathic hypersomnia while utilizing standardized and validated instruments is required, because the effect of these conditions of hypersomnolence varies and is significant for each individual. © 2018 American Academy of Sleep Medicine.

  8. DQB1 Locus Alone Explains Most of the Risk and Protection in Narcolepsy with Cataplexy in Europe

    PubMed Central

    Tafti, Mehdi; Hor, Hyun; Dauvilliers, Yves; Lammers, Gert J.; Overeem, Sebastiaan; Mayer, Geert; Javidi, Sirous; Iranzo, Alex; Santamaria, Joan; Peraita-Adrados, Rosa; Vicario, José L.; Arnulf, Isabelle; Plazzi, Giuseppe; Bayard, Sophie; Poli, Francesca; Pizza, Fabio; Geisler, Peter; Wierzbicka, Aleksandra; Bassetti, Claudio L.; Mathis, Johannes; Lecendreux, Michel; Donjacour, Claire E.H.M.; van der Heide, Astrid; Heinzer, Raphaël; Haba-Rubio, José; Feketeova, Eva; Högl, Birgit; Frauscher, Birgit; Benetó, Antonio; Khatami, Ramin; Cañellas, Francesca; Pfister, Corinne; Scholz, Sabine; Billiard, Michel; Baumann, Christian R.; Ercilla, Guadalupe; Verduijn, Willem; Claas, Frans H.J.; Dubois, Valérie; Nowak, Jacek; Eberhard, Hans-Peter; Pradervand, Sylvain; Hor, Charlotte N.; Testi, Manuela; Tiercy, Jean-Marie; Kutalik, Zoltán

    2014-01-01

    Study Objective: Prior research has identified five common genetic variants associated with narcolepsy with cataplexy in Caucasian patients. To replicate and/or extend these findings, we have tested HLA-DQB1, the previously identified 5 variants, and 10 other potential variants in a large European sample of narcolepsy with cataplexy subjects. Design: Retrospective case-control study. Setting: A recent study showed that over 76% of significant genome-wide association variants lie within DNase I hypersensitive sites (DHSs). From our previous GWAS, we identified 30 single nucleotide polymorphisms (SNPs) with P < 10-4 mapping to DHSs. Ten SNPs tagging these sites, HLADQB1, and all previously reported SNPs significantly associated with narcolepsy were tested for replication. Patients and Participants: For GWAS, 1,261 narcolepsy patients and 1,422 HLA-DQB1*06:02-matched controls were included. For HLA study, 1,218 patients and 3,541 controls were included. Measurements and Results: None of the top variants within DHSs were replicated. Out of the five previously reported SNPs, only rs2858884 within the HLA region (P < 2x10-9) and rs1154155 within the TRA locus (P < 2x10-8) replicated. DQB1 typing confirmed that DQB1*06:02 confers an extraordinary risk (odds ratio 251). Four protective alleles (DQB1*06:03, odds ratio 0.17, DQB1*05:01, odds ratio 0.56, DQB1*06:09 odds ratio 0.21, DQB1*02 odds ratio 0.76) were also identified. Conclusion: An overwhelming portion of genetic risk for narcolepsy with cataplexy is found at DQB1 locus. Since DQB1*06:02 positive subjects are at 251-fold increase in risk for narcolepsy, and all recent cases of narcolepsy after H1N1 vaccination are positive for this allele, DQB1 genotyping may be relevant to public health policy. Citation: Tafti M; Hor H; Dauvilliers Y; Lammers GJ; Overeem S; Mayer G; Javidi S; Iranzo A; Santamaria J; Peraita-Adrados R; Vicario JL; Arnulf I; Plazzi G; Bayard S; Poli F; Pizza F; Geisler P; Wierzbicka A; Bassetti CL

  9. The Prevalence and Characteristics of Primary Headache and Dream-Enacting Behaviour in Japanese Patients with Narcolepsy or Idiopathic Hypersomnia: A Multi-Centre Cross-Sectional Study.

    PubMed

    Suzuki, Keisuke; Miyamoto, Masayuki; Miyamoto, Tomoyuki; Inoue, Yuichi; Matsui, Kentaro; Nishida, Shingo; Hayashida, Kenichi; Usui, Akira; Ueki, Yoichiro; Nakamura, Masaki; Murata, Momoyo; Numao, Ayaka; Watanabe, Yuji; Suzuki, Shiho; Hirata, Koichi

    2015-01-01

    Because the prevalence and characteristics of primary headache have yet to be thoroughly studied in patients with hypersomnia disorders, including narcolepsy and idiopathic hypersomnia, we examined these parameters in the Japanese population. In a multicentre cross-sectional survey, among 576 consecutive outpatients with sleep disorders, 68 narcolepsy patients and 35 idiopathic hypersomnia patients were included. Additionally, 61 healthy control subjects participated. Semi-structured headache questionnaires were administered to all participants. The patients with narcolepsy (52.9%) and idiopathic hypersomnia (77.1%) more frequently experienced headache than the healthy controls (24.6%; p<0.0001). The prevalence rates were 23.5%, 41.2% and 4.9% for migraine (p<0.0001) and 16.2%, 23.5% and 14.8% (p = 0.58) for tension-type headache among the narcolepsy patients, the idiopathic hypersomnia patients and the control subjects, respectively. Those who experienced migraine more frequently experienced excessive daytime sleepiness, defined as an Epworth Sleepiness Scale score of ≥10, than those who did not experience headache among the patients with narcolepsy (93.8% vs. 65.6%, p = 0.040) and idiopathic hypersomnia (86.7% vs. 37.5%, p = 0.026). Dream-enacting behaviour (DEB), as evaluated by the rapid eye movement sleep disorders questionnaire, was more frequently observed in the narcolepsy patients than in the idiopathic hypersomnia patients and the control subjects. An increased DEB frequency was observed in the narcolepsy patients with migraines compared to those without headache. Migraines were frequently observed in patients with narcolepsy and idiopathic hypersomnia. DEB is a characteristic of narcolepsy patients. Further studies are required to assess the factors that contribute to migraines in narcolepsy and idiopathic hypersomnia patients.

  10. HLA DQB1*06:02 Negative Narcolepsy with Hypocretin/Orexin Deficiency

    PubMed Central

    Han, Fang; Lin, Ling; Schormair, Barbara; Pizza, Fabio; Plazzi, Giuseppe; Ollila, Hanna M.; Nevsimalova, Sona; Jennum, Poul; Knudsen, Stine; Winkelmann, Juliane; Coquillard, Cristin; Babrzadeh, Farbod; Strom, Tim M.; Wang, Chunlin; Mindrinos, Michael; Vina, Marcelo Fernandez; Mignot, Emmanuel

    2014-01-01

    Study Objectives: To identify rare allelic variants and HLA alleles in narcolepsy patients with hypocretin (orexin, HCRT) deficiency but lacking DQB1*06:02. Settings: China (Peking University People's Hospital), Czech Republic (Charles University), Denmark (Golstrup Hospital), Italy (University of Bologna), Korea (Catholic University), and USA (Stanford University). Design: CSF hypocretin-1, DQB1*06:02, clinical and polysomnographic data were collected in narcolepsy patients (552 with and 144 without cataplexy) from 6 sites. Numbers of cases with and without DQB1*06:02 and low CSF hypocretin-1 were compiled. HLA class I (A, B, C), class II (DRBs, DQA1, DQB1, DPA1, and DPB1), and whole exome sequencing were conducted in 9 DQB1*06:02 negative cases with low CSF hypocretin-1. Sanger sequencing of selected exons in DNMT1, HCRT, and MOG was performed to exclude mutations in known narcolepsy-associated genes. Measurements and Results: Classic narcolepsy markers DQB1*06:02 and low CSF hypocretin-1 were found in 87.4% of cases with cataplexy, and in 20.0% without cataplexy. Nine cases (all with cataplexy) were DQB1*06:02 negative with low CSF hypocretin-1, constituting 1.7% [0.8%-3.4%] of all cases with cataplexy and 1.8% [0.8%-3.4%] of cases with low CSF hypocretin independent of cataplexy across sites. Five HLA negative subjects had severe cataplexy, often occurring without clear triggers. Subjects had diverse ethnic backgrounds and HLA alleles at all loci, suggesting no single secondary HLA association. The rare subtype DPB1*0901, and homologous DPB1*10:01 subtype, were present in 5 subjects, suggesting a secondary association with HLA-DP. Preprohypocretin sequencing revealed no mutations beyond one previously reported in a very early onset case. No new MOG or DNMT1 mutations were found, nor were suspicious or private variants in novel genes identified through exome sequencing. Conclusions: Hypocretin, MOG, or DNMT1 mutations are exceptional findings in DQB1

  11. Nocturnal Sleep Dynamics Identify Narcolepsy Type 1.

    PubMed

    Pizza, Fabio; Vandi, Stefano; Iloti, Martina; Franceschini, Christian; Liguori, Rocco; Mignot, Emmanuel; Plazzi, Giuseppe

    2015-08-01

    To evaluate the reliability of nocturnal sleep dynamics in the differential diagnosis of central disorders of hypersomnolence. Cross-sectional. Sleep laboratory. One hundred seventy-five patients with hypocretin-deficient narcolepsy type 1 (NT1, n = 79), narcolepsy type 2 (NT2, n = 22), idiopathic hypersomnia (IH, n = 22), and "subjective" hypersomnolence (sHS, n = 52). None. Polysomnographic (PSG) work-up included 48 h of continuous PSG recording. From nocturnal PSG conventional sleep macrostructure, occurrence of sleep onset rapid eye movement period (SOREMP), sleep stages distribution, and sleep stage transitions were calculated. Patient groups were compared, and receiver operating characteristic (ROC) curve analysis was used to test the diagnostic utility of nocturnal PSG data to identify NT1. Sleep macrostructure was substantially stable in the 2 nights of each diagnostic group. NT1 and NT2 patients had lower latency to rapid eye movement (REM) sleep, and NT1 patients showed the highest number of awakenings, sleep stage transitions, and more time spent in N1 sleep, as well as most SOREMPs at daytime PSG and at multiple sleep latency test (MSLT) than all other groups. ROC curve analysis showed that nocturnal SOREMP (area under the curve of 0.724 ± 0.041, P < 0.0001), percent of total sleep time spent in N1 (0.896 ± 0.023, P < 0.0001), and the wakefulness-sleep transition index (0.796 ± 0.034, P < 0.0001) had a good sensitivity and specificity profile to identify NT1 sleep, especially when used in combination (0.903 ± 0.023, P < 0.0001), similarly to SOREMP number at continuous daytime PSG (0.899 ± 0.026, P < 0.0001) and at MSLT (0.956 ± 0.015, P < 0.0001). Sleep macrostructure (i.e. SOREMP, N1 timing) including stage transitions reliably identifies hypocretin-deficient narcolepsy type 1 among central disorders of hypersomnolence. © 2015 Associated Professional Sleep Societies, LLC.

  12. New developments in the management of narcolepsy

    PubMed Central

    Abad, Vivien C; Guilleminault, Christian

    2017-01-01

    Narcolepsy is a life-long, underrecognized sleep disorder that affects 0.02%–0.18% of the US and Western European populations. Genetic predisposition is suspected because of narcolepsy’s strong association with HLA DQB1*06-02, and genome-wide association studies have identified polymorphisms in T-cell receptor loci. Narcolepsy pathophysiology is linked to loss of signaling by hypocretin-producing neurons; an autoimmune etiology possibly triggered by some environmental agent may precipitate hypocretin neuronal loss. Current treatment modalities alleviate the main symptoms of excessive daytime somnolence (EDS) and cataplexy and, to a lesser extent, reduce nocturnal sleep disruption, hypnagogic hallucinations, and sleep paralysis. Sodium oxybate (SXB), a sodium salt of γ hydroxybutyric acid, is a first-line agent for cataplexy and EDS and may help sleep disruption, hypnagogic hallucinations, and sleep paralysis. Various antidepressant medications including norepinephrine serotonin reuptake inhibitors, selective serotonin reuptake inhibitors, and tricyclic antidepressants are second-line agents for treating cataplexy. In addition to SXB, modafinil and armodafinil are first-line agents to treat EDS. Second-line agents for EDS are stimulants such as methylphenidate and extended-release amphetamines. Emerging therapies include non-hypocretin-based therapy, hypocretin-based treatments, and immunotherapy to prevent hypocretin neuronal death. Non-hypocretin-based novel treatments for narcolepsy include pitolisant (BF2.649, tiprolisant); JZP-110 (ADX-N05) for EDS in adults; JZP 13-005 for children; JZP-386, a deuterated sodium oxybate oral suspension; FT 218 an extended-release formulation of SXB; and JNJ-17216498, a new formulation of modafinil. Clinical trials are investigating efficacy and safety of SXB, modafinil, and armodafinil in children. γ-amino butyric acid (GABA) modulation with GABAA receptor agonists clarithromycin and flumazenil may help daytime somnolence

  13. Hypocretin-1 Levels Associate with Fragmented Sleep in Patients with Narcolepsy Type 1.

    PubMed

    Alakuijala, Anniina; Sarkanen, Tomi; Partinen, Markku

    2016-05-01

    We aimed to analyze nocturnal sleep characteristics of patients with narcolepsy type 1 (narcolepsy with cataplexy) measured by actigraphy in respect to cerebrospinal fluid hypocretin-1 levels of the same patients. Actigraphy recording of 1-2 w and hypocretin-1 concentration analysis were done to thirty-six unmedicated patients, aged 7 to 63 y, 50% female. Twenty-six of them had hypocretin-1 levels under 30 pg/mL and the rest had levels of 31-79 pg/mL. According to actigraphy, patients with very low hypocretin levels had statistically significantly longer sleep latency (P = 0.033) and more fragmented sleep, indicated by both the number of immobile phases of 1 min (P = 0.020) and movement + fragmentation index (P = 0.049). There were no statistically significant differences in the actual sleep time or circadian rhythm parameters measured by actigraphy. Actigraphy gives additional information about the stabilization of sleep in patients with narcolepsy type 1. Very low hypocretin levels associate with more wake intruding into sleep. © 2016 Associated Professional Sleep Societies, LLC.

  14. Increased Incidence and Clinical Picture of Childhood Narcolepsy following the 2009 H1N1 Pandemic Vaccination Campaign in Finland

    PubMed Central

    Partinen, Markku; Saarenpää-Heikkilä, Outi; Ilveskoski, Ismo; Hublin, Christer; Linna, Miika; Olsén, Päivi; Nokelainen, Pekka; Alén, Reija; Wallden, Tiina; Espo, Merimaaria; Rusanen, Harri; Olme, Jan; Sätilä, Heli; Arikka, Harri; Kaipainen, Pekka; Julkunen, Ilkka; Kirjavainen, Turkka

    2012-01-01

    Background Narcolepsy is a rare neurological sleep disorder especially in children who are younger than 10 years. In the beginning of 2010, an exceptionally large number of Finnish children suffered from an abrupt onset of excessive daytime sleepiness (EDS) and cataplexy. Therefore, we carried out a systematic analysis of the incidence of narcolepsy in Finland between the years 2002–2010. Methods All Finnish hospitals and sleep clinics were contacted to find out the incidence of narcolepsy in 2010. The national hospital discharge register from 2002 to 2009 was used as a reference. Findings Altogether 335 cases (all ages) of narcolepsy were diagnosed in Finland during 2002–2009 giving an annual incidence of 0.79 per 100 000 inhabitants (95% confidence interval 0.62–0.96). The average annual incidence among subjects under 17 years of age was 0.31 (0.12–0.51) per 100 000 inhabitants. In 2010, 54 children under age 17 were diagnosed with narcolepsy (5.3/100 000; 17-fold increase). Among adults ≥20 years of age the incidence rate in 2010 was 0.87/100 000, which equals that in 2002–2009. Thirty-four of the 54 children were HLA-typed, and they were all positive for narcolepsy risk allele DQB1*0602/DRB1*15. 50/54 children had received Pandemrix vaccination 0 to 242 days (median 42) before onset. All 50 had EDS with abnormal multiple sleep latency test (sleep latency <8 min and ≥2 sleep onset REM periods). The symptoms started abruptly. Forty-seven (94%) had cataplexy, which started at the same time or soon after the onset of EDS. Psychiatric symptoms were common. Otherwise the clinical picture was similar to that described in childhood narcolepsy. Interpretation A sudden increase in the incidence of abrupt childhood narcolepsy was observed in Finland in 2010. We consider it likely that Pandemrix vaccination contributed, perhaps together with other environmental factors, to this increase in genetically susceptible children. PMID:22470463

  15. The Prevalence and Characteristics of Primary Headache and Dream-Enacting Behaviour in Japanese Patients with Narcolepsy or Idiopathic Hypersomnia: A Multi-Centre Cross-Sectional Study

    PubMed Central

    Suzuki, Keisuke; Miyamoto, Masayuki; Miyamoto, Tomoyuki; Inoue, Yuichi; Matsui, Kentaro; Nishida, Shingo; Hayashida, Kenichi; Usui, Akira; Ueki, Yoichiro; Nakamura, Masaki; Murata, Momoyo; Numao, Ayaka; Watanabe, Yuji; Suzuki, Shiho; Hirata, Koichi

    2015-01-01

    Background Because the prevalence and characteristics of primary headache have yet to be thoroughly studied in patients with hypersomnia disorders, including narcolepsy and idiopathic hypersomnia, we examined these parameters in the Japanese population. Methods In a multicentre cross-sectional survey, among 576 consecutive outpatients with sleep disorders, 68 narcolepsy patients and 35 idiopathic hypersomnia patients were included. Additionally, 61 healthy control subjects participated. Semi-structured headache questionnaires were administered to all participants. Results The patients with narcolepsy (52.9%) and idiopathic hypersomnia (77.1%) more frequently experienced headache than the healthy controls (24.6%; p<0.0001). The prevalence rates were 23.5%, 41.2% and 4.9% for migraine (p<0.0001) and 16.2%, 23.5% and 14.8% (p = 0.58) for tension-type headache among the narcolepsy patients, the idiopathic hypersomnia patients and the control subjects, respectively. Those who experienced migraine more frequently experienced excessive daytime sleepiness, defined as an Epworth Sleepiness Scale score of ≥10, than those who did not experience headache among the patients with narcolepsy (93.8% vs. 65.6%, p = 0.040) and idiopathic hypersomnia (86.7% vs. 37.5%, p = 0.026). Dream-enacting behaviour (DEB), as evaluated by the rapid eye movement sleep disorders questionnaire, was more frequently observed in the narcolepsy patients than in the idiopathic hypersomnia patients and the control subjects. An increased DEB frequency was observed in the narcolepsy patients with migraines compared to those without headache. Conclusions Migraines were frequently observed in patients with narcolepsy and idiopathic hypersomnia. DEB is a characteristic of narcolepsy patients. Further studies are required to assess the factors that contribute to migraines in narcolepsy and idiopathic hypersomnia patients. PMID:26418536

  16. The Safety of Adjuvanted Vaccines Revisited: Vaccine-Induced Narcolepsy.

    PubMed

    Ahmed, S Sohail; Montomoli, Emanuele; Pasini, Franco Laghi; Steinman, Lawrence

    2016-01-01

    Despite the very high benefit-to-risk ratio of vaccines, the fear of negative side effects has discouraged many people from getting vaccinated, resulting in the reemergence of previously controlled diseases such as measles, pertussis and diphtheria. This fear has been amplified more recently by multiple epidemiologic studies that confirmed the link of an AS03-adjuvanted pandemic influenza vaccine (Pandemrix, GlaxoSmithKline Biologicals, Germany) used in Europe during the 2009 H1N1 influenza pandemic [A(H1N1) pdm09] with the development of narcolepsy, a chronic sleep disorder, in children and adolescents. However, public misperceptions of what adjuvants are and why they are used in vaccines has created in some individuals a closed "black box" attitude towards all vaccines. The focus of this review article is to revisit this "black box" using the example of narcolepsy associated with the European AS03-adjuvanted pandemic influenza vaccine.

  17. CD4+ T-Cell Reactivity to Orexin/Hypocretin in Patients With Narcolepsy Type 1.

    PubMed

    Ramberger, Melanie; Högl, Birgit; Stefani, Ambra; Mitterling, Thomas; Reindl, Markus; Lutterotti, Andreas

    2017-03-01

    Narcolepsy type 1 is accompanied by a selective loss of orexin/hypocretin (hcrt) neurons in the lateral hypothalamus caused by yet unknown mechanisms. Epidemiologic and genetic associations strongly suggest an immune-mediated pathogenesis of the disease. We compared specific T-cell reactivity to orexin/hcrt peptides in peripheral blood mononuclear cells of narcolepsy type 1 patients to healthy controls by a carboxyfluorescein succinimidyl ester proliferation assay. Orexin/hcrt-specific T-cell reactivity was also determined by cytokine (interferon gamma and granulocyte-macrophage colony-stimulating factor) analysis. Individuals were considered as responders if the cell division index of CD3+CD4+ T cells and both stimulation indices of cytokine secretion exceeded the cutoff 3. Additionally, T-cell reactivity to orexin/hcrt had to be confirmed by showing reactivity to single peptides present in different peptide pools. Using these criteria, 3/15 patients (20%) and 0/13 controls (0%) showed orexin/hcrt-specific CD4+ T-cell proliferation (p = .2262). The heterogeneous reactivity pattern did not allow the identification of a preferential target epitope. A significant role of orexin/hcrt-specific T cells in narcolepsy type 1 patients could not be confirmed in this study. Further studies are needed to assess the exact role of CD4+ T cells and possible target antigens in narcolepsy type 1 patients. © Sleep Research Society 2016. Published by Oxford University Press [on behalf of the Sleep Research Society].

  18. Medical exposures in youth and the frequency of narcolepsy with cataplexy: a population-based case-control study in genetically predisposed people

    PubMed Central

    Koepsell, Thomas D.; Longstreth, William T.; Ton, Thanh G. N.

    2012-01-01

    SUMMARY Epidemiological observations suggest that exposures in youth may trigger narcolepsy in genetically predisposed individuals. In this population-based case–control study, we sought to identify all prevalent cases of narcolepsy with cataplexy aged 18–50 years as of 1 July 2001, in King County, Washington. The 45 eligible cases who were DQB1*0602-positive were compared with 95 controls with this allele, identified through random-digit dialing and buccal smears. Cases and controls were interviewed in person about physician-diagnosed infectious and non-infectious illnesses, immunizations, head trauma and parasomnias or psychiatric problems during youth. Narcolepsy with cataplexy was more frequent in African-Americans and in poorer households. Adjusting for these factors, the condition was 5.4-fold more common [95% confidence interval (CI) = 1.5–19.1] among people reporting a physician-diagnosed strep throat before the age of 21 years. No other significant associations with childhood diseases, immunizations or head trauma were found. However, prevalence was increased 16.3-fold (95% CI = 6.1–44.1) in subjects who reported having had ‘night terrors’. Strep throat may be related to narcolepsy with cataplexy in genetically susceptible individuals. The association with night terrors could simply reflect early symptoms of narcolepsy, or they could be a prodromal sign of disturbed sleep physiology. keywords epidemiology, head injuries, immunization, narcolepsy, night terrors, streptococcal infections PMID:19732319

  19. Paraneoplastic limbic encephalitis and possible narcolepsy in a patient with testicular cancer: case study.

    PubMed Central

    Landolfi, Joseph C.; Nadkarni, Mangala

    2003-01-01

    We describe a patient who presented with a clinical syndrome of limbic encephalitis, narcolepsy, and cataplexy. The anti-Ma2 antibody was positive. Although there was no mass on imaging, orchiectomy was performed in this patient, and testicular carcinoma was found. This is the first known case of limbic encephalitis and anti-Ma2 antibody to be associated with cataplexy and possible narcolepsy. Neurological symptoms precede the diagnosis of cancer in 50% of patients with paraneoplastic syndromes, and clinicians are therefore strongly advised to evaluate patients with neurological symptoms for this condition. PMID:12816728

  20. Usefulness of a Nocturnal SOREMP for Diagnosing Narcolepsy with Cataplexy in a Pediatric Population

    PubMed Central

    Reiter, Joel; Katz, Eliot; Scammell, Thomas E.; Maski, Kiran

    2015-01-01

    Study Objectives: We investigated the diagnostic accuracy of a nocturnal sleep onset rapid eye movement sleep period (nSOREMP) for the identification of narcolepsy with cataplexy (N+C) among children and adolescents referred to the sleep laboratory for an overnight polysomnography (PSG) and multiple sleep latency test (MSLT). Design: Retrospective chart review of sleep clinic notes and PSG and MSLT reports. Setting: Boston Children's Hospital sleep laboratory and outpatient clinics. Patients: All patients 6–18 y old, referred for consecutive PSG and MSLT for the evaluation of central hypersomnias, between January 2005 and January 2014. Measurements and Results: We analyzed the records of 148 patients and established diagnostic categories using the International Classification of Sleep Disorders, 2nd Edition. Patient diagnoses included narcolepsy with cataplexy (28.4%), narcolepsy without cataplexy (8.1%), other hypersomnia conditions (9.5%), delayed sleep phase syndrome (12.2%), behaviorally induced insufficient sleep syndrome (4.1%), other sleep disorders (obstructive sleep apnea, periodic limb movements of sleep; 6.8%), isolated cataplexy (2%), and various diagnoses (29.1%). There were 54.8% of the N+C patients who had an nSOREMP, but only 2.4% of all other patients had an nSOREMP. The specificity of an nSOREMP for detection of N+C was high at 97.3% (95% confidence interval [CI]: 92.2–99.4%), but the sensitivity was moderate at 54.8% (95% CI: 38.7–70.2%). Overall, the positive predictive value of an nSOREMP for the diagnosis of N+C was 88.5% (95% CI: 69.8–97.4%). Conclusions: In children, the presence of an nocturnal sleep onset rapid eye movement sleep period is highly suggestive of narcolepsy with cataplexy and provides further evidence of rapid eye movement sleep dysregulation in this condition. Citation: Reiter J, Katz E, Scammell TE, Maski K. Usefulness of a nocturnal SOREMP for diagnosing narcolepsy with cataplexy in a pediatric population. SLEEP

  1. The ICSD-3 and DSM-5 guidelines for diagnosing narcolepsy: clinical relevance and practicality.

    PubMed

    Ruoff, Chad; Rye, David

    2016-07-20

    Narcolepsy is a chronic neurological disease manifesting as difficulty with maintaining continuous wake and sleep. Clinical presentation varies but requires excessive daytime sleepiness (EDS) occurring alone or together with features of rapid-eye movement (REM) sleep dissociation (e.g., cataplexy, hypnagogic/hypnopompic hallucinations, sleep paralysis), and disrupted nighttime sleep. Narcolepsy with cataplexy is associated with reductions of cerebrospinal fluid (CSF) hypocretin due to destruction of hypocretin peptide-producing neurons in the hypothalamus in individuals with a specific genetic predisposition. Updated diagnostic criteria include the Diagnostic and Statistical Manual of Mental Disorders Fifth Edition (DSM-5) and International Classification of Sleep Disorders Third Edition (ICSD-3). DSM-5 criteria require EDS in association with any one of the following: (1) cataplexy; (2) CSF hypocretin deficiency; (3) REM sleep latency ≤15 minutes on nocturnal polysomnography (PSG); or (4) mean sleep latency ≤8 minutes on multiple sleep latency testing (MSLT) with ≥2 sleep-onset REM-sleep periods (SOREMPs). ICSD-3 relies more upon objective data in addition to EDS, somewhat complicating the diagnostic criteria: 1) cataplexy and either positive MSLT/PSG findings or CSF hypocretin deficiency; (2) MSLT criteria similar to DSM-5 except that a SOREMP on PSG may count as one of the SOREMPs required on MSLT; and (3) distinct division of narcolepsy into type 1, which requires the presence of cataplexy or documented CSF hypocretin deficiency, and type 2, where cataplexy is absent, and CSF hypocretin levels are either normal or undocumented. We discuss limitations of these criteria such as variability in clinical presentation of cataplexy, particularly when cataplexy may be ambiguous, as well as by age; multiple and/or invasive CSF diagnostic test requirements; and lack of normative diagnostic test data (e.g., MSLT) in certain populations. While ICSD-3 criteria

  2. Use of PCR with Sequence-specific Primers for High-Resolution Human Leukocyte Antigen Typing of Patients with Narcolepsy

    PubMed Central

    Woo, Hye In; Joo, Eun Yeon; Lee, Kyung Wha

    2012-01-01

    Background Narcolepsy is a neurologic disorder characterized by excessive daytime sleepiness, symptoms of abnormal rapid eye movement (REM) sleep, and a strong association with HLA-DRB1*1501, -DQA1*0102, and -DQB1*0602. Here, we investigated the clinico-physical characteristics of Korean patients with narcolepsy, their HLA types, and the clinical utility of high-resolution PCR with sequence-specific primers (PCR-SSP) as a simple typing method for identifying DRB1*15/16, DQA1, and DQB1 alleles. Methods The study population consisted of 67 consecutively enrolled patients having unexplained daytime sleepiness and diagnosed narcolepsy based on clinical and neurological findings. Clinical data and the results of the multiple sleep latency test and polysomnography were reviewed, and HLA typing was performed using both high-resolution PCR-SSP and sequence-based typing (SBT). Results The 44 narcolepsy patients with cataplexy displayed significantly higher frequencies of DRB1*1501 (Pc= 0.003), DQA1*0102 (Pc=0.001), and DQB1*0602 (Pc=0.014) than the patients without cataplexy. Among patients carrying DRB1*1501-DQB1*0602 or DQA1*0102, the frequencies of a mean REM sleep latency of less than 20 min in nocturnal polysomnography and clinical findings, including sleep paralysis and hypnagogic hallucination were significantly higher. SBT and PCR-SSP showed 100% concordance for high-resolution typing of DRB1*15/16 alleles and DQA1 and DQB1 loci. Conclusions The clinical characteristics and somnographic findings of narcolepsy patients were associated with specific HLA alleles, including DRB1*1501, DQA1*0102, and DQB1*0602. Application of high-resolution PCR-SSP, a reliable and simple method, for both allele- and locus-specific HLA typing of DRB1*15/16, DQA1, and DQB1 would be useful for characterizing clinical status among subjects with narcolepsy. PMID:22259780

  3. Effects of Oral L-Carnitine Administration in Narcolepsy Patients: A Randomized, Double-Blind, Cross-Over and Placebo-Controlled Trial

    PubMed Central

    Miyagawa, Taku; Kawamura, Hiromi; Obuchi, Mariko; Ikesaki, Asuka; Ozaki, Akiko; Tokunaga, Katsushi; Inoue, Yuichi; Honda, Makoto

    2013-01-01

    Narcolepsy is a sleep disorder characterized by excessive daytime sleepiness, cataplexy, and rapid eye movement (REM) sleep abnormalities. A genome-wide association study (GWAS) identified a novel narcolepsy-related single nucleotide polymorphism (SNP), which is located adjacent to the carnitine palmitoyltransferase 1B (CPT1B) gene encoding an enzyme involved in β-oxidation of long-chain fatty acids. The mRNA expression levels of CPT1B were associated with this SNP. In addition, we recently reported that acylcarnitine levels were abnormally low in narcolepsy patients. To assess the efficacy of oral l-carnitine for the treatment of narcolepsy, we performed a clinical trial administering l-carnitine (510 mg/day) to patients with the disease. The study design was a randomized, double-blind, cross-over and placebo-controlled trial. Thirty narcolepsy patients were enrolled in our study. Two patients were withdrawn and 28 patients were included in the statistical analysis (15 males and 13 females, all with HLA-DQB1*06:02). l-carnitine treatment significantly improved the total time for dozing off during the daytime, calculated from the sleep logs, compared with that of placebo-treated periods. l-carnitine efficiently increased serum acylcarnitine levels, and reduced serum triglycerides concentration. Differences in the Japanese version of the Epworth Sleepiness Scale (ESS) and the Medical Outcomes Study 36-Item Short-Form Health Survey (SF-36) vitality and mental health subscales did not reach statistical significance between l-carnitine and placebo. This study suggests that oral l-carnitine can be effective in reducing excessive daytime sleepiness in narcolepsy patients. Trial Registration University hospital Medical Information Network (UMIN) UMIN000003760 PMID:23349733

  4. Potentiation of a functional autoantibody in narcolepsy by a cholinesterase inhibitor.

    PubMed

    Jackson, Michael W; Spencer, Nicolas J; Reed, Joanne H; Smith, Anthony J F; Gordon, Tom P

    2009-12-01

    We have recently reported the presence of an immunoglobulin G (IgG) autoantibody (Ab) in patients with narcolepsy with cataplexy that abolishes spontaneous colonic migrating motor complexes (CMMCs) and increases smooth muscle tension and atropine-sensitive phasic contractions in a physiological assay of an isolated colon. In this study, we used the cholinesterase inhibitor, neostigmine, to explore the mechanism of the narcoleptic IgG-mediated disruption of enteric motor function in four patients with narcolepsy with cataplexy and to identify a pharmacological mimic of the Ab. Neostigmine potentiated the narcoleptic IgG-mediated increase in smooth muscle resting tension and phasic smooth muscle contractions by an atropine-sensitive mechanism but exerted no effect on resting tension in the presence of control IgG. Decreased frequency of CMMCs mediated by IgG with anti-M3R activity was reversed by neostigmine. Therefore, a challenge with a cholinesterase inhibitor improves the specificity of the CMMC assay for narcoleptic IgG. Tetrodotoxin (TTX), a neuronal sodium channel blocker, also abolished CMMCs and increased resting tone, and a similar potentiation was observed with neostigmine; thus, TTX is a mimic of the functional effects of the narcoleptic IgG in this bioassay. These findings provide a link to pharmacological studies of canine narcolepsy and are consistent with a functional blockade of both excitatory and inhibitory motor neurons by the narcoleptic Ab, similar to the TTX mimic, presumably by binding to an autoantigenic target expressed in both populations of neurons.

  5. Environmental factors in the development of narcolepsy with cataplexy. A case-control study.

    PubMed

    Peraita-Adrados, R; del Rio-Villegas, R; Vela-Bueno, A

    2015-06-16

    Epidemiological studies suggest the importance of environmental factors in the etiology of narcolepsy-cataplexy in genetically predisposed subjects. To assess the role of environmental factors in the development of narcolepsy-cataplexy, using a case-control design with control subjects being matched for ethnicity and age. All patients were recruited through two outpatient clinics at the community of Madrid, ant the diagnosis of narcolepsy fulfilled the criteria of the International Classification on Sleep Disorders-2005. A questionnaire, including 54 environmental psychological stressor life events and 42 infectious diseases items, was administered to 54 patients. We specifically assessed the stressful factors and infectious diseases that occurred in the year preceding the onset of the first symptom of narcolepsy (excessive daytime sleepiness and/or cataplexy). The same questionnaire was administered to 84 control subjects recruited from non-related family members of the same community. Fifty four patients (55.6% males) answered the questionnaire, The mean age at onset of the first symptom was 21.6 ± 9.3 years, and the mean age at diagnosis was 36.5 ± 12.4 years. The main finding in narcoleptic patients as compared to control subjects was major changes in the 'number of arguments with partner, family, or friends' (odds ratio: 5.2; 95% confidence interval: 1.8-14.5). This can be interpreted as having a protective function and it suggests that psychological mechanisms are present since the beginning of the disease. As for the infectious factors, chickenpox was the most frequently reported. No significant differences were found in terms of total numbers of stress-related and infectious factors between cases and controls. Prospective studies regarding the interaction between environmental and genetic factors are warranted.

  6. Diagnostic delay in narcolepsy type 1: combining the patients' and the doctors' perspectives.

    PubMed

    Taddei, Raquel N; Werth, Esther; Poryazova, Rositsa; Baumann, Christian R; Valko, Philipp O

    2016-12-01

    Narcolepsy type 1 is a neurological disorder characterized by a unique syndrome, including the pathognomonic symptom of cataplexy. The diagnosis can be confirmed by objective measures, such as typical findings in the multiple sleep latency test, reduced or undetectable levels of orexin (hypocretin) in the cerebrospinal fluid, and linkage to a specific HLA haplotype. Nevertheless, the mean time that elapses from symptom onset to the correct diagnosis ranges between 10 and 20 years, and the causes and correlates of this delay are poorly understood. Diagnostic delay was assessed on 52 well-defined patients with narcolepsy type 1, evaluating clinical, electrophysiological and neurochemical parameters and the results of a 41-item questionnaire developed to obtain the patients' perspective on various aspects of the diagnostic process. The mean time gap between disease onset and first medical consultation was 3.2 ± 5.1 years; the mean diagnostic delay was 8.9 ± 11.0 years. Prior to correct diagnosis, patients received a wide variety of misdiagnoses. The self-ratings of the patients revealed that the undiagnosed symptoms caused high levels of anxiety and unjustified criticism by family, friends and employers. Multiple regression analysis identified higher cerebrospinal fluid orexin levels (β = 0.311, P = 0.01), and a longer interval between the onset of excessive daytime sleepiness and cataplexy (β = 0.368, P = 0.002) as independent associates of longer diagnostic delay. The diagnostic delay decreased over the last decades (β = -0.672, P < 0.001). In conclusion, delayed diagnosis of narcolepsy type 1 is very common, associated with many adverse consequences, and requires educational efforts to improve awareness on narcolepsy among healthcare providers and the general population. © 2016 European Sleep Research Society.

  7. Narcolepsy (with and without cataplexy) and commercial motor vehicle driver safety.

    DOT National Transportation Integrated Search

    2009-10-01

    The purpose of this evidence report is to address several questions posed by FMCSA regarding the topic of narcolepsy and commercial motor vehicle (CMV) driver safety. In the early scope development work conducted by the Agency and the Medical Review ...

  8. Clarifying the association of genes within the major histocompatibility complex with narcolepsy

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Acton, R.T.; Watson, B.; Rivers, C.

    1994-09-01

    HLA-DR2 and DQwl has been reported to be strongly associated with narcolepsy. The particular phenotype and strength of these associations varies between races. For example DQB*0601 has been reported associated with some African American (AA) narcoleptics while some Caucasian American (CA) narcoleptics do not possess DR2 or DQw1. We have sought to clarify the relationship of MHC genes with narcolepsy in the local CA and AA population. There was no significant difference in the frequency of DR phenotypes in CA or AA narcoleptics compared to race, age, sex and geographic region-matched controls. DR2 was increased in CA cataplexy positive (Cat+)more » narcoleptics compared to controls (p=0.028, odds ratio (OR)=2.4) and to Cat- narcoleptics (p=<0.001, OR=8.8). DR11 was increased in AA Cat+ narcoleptics compared to controls (p=0.004, OR=11.2) and to Cat- narcoleptics (p=0.002). DQB1*0601 was not significantly associated with narcolepsy in our AA population. We have assessed the frequency of the TNFa (13 alleles, 1.1Mb telomeric to DQ{alpha}), D6S105 (13 alleles, 1kb telomeric of HLA-A), and GLP-1R (19 alleles, 18.5 Mb centromeric of DQ{alpha}), dinucleotide repeats in narcoleptics compared to controls. The TNFa allele 117 was increased in CA Cat+ vs. controls (p=0.003). The GLP-1R allele 144 was increased in CA Cat- vs. controls (p=0.02). In AA narcoleptics, the TNFa allele 109 was significantly increased (p=0.04) along with the D6S105 allele 130 (p=0.02) compared to controls. The D6S105 allele 130 was increased in AA Cat- vs. controls (p=0.03). The GLP-1R allele 154 was significantly decreased in AA Cat+ vs. Cat- (p=0.04). These data suggest that DR and/or DQ genes are not responsible for narcolepsy and that cataplexy is associated with different regions around the MHC in various racial groups.« less

  9. Retrospective multicenter matched case-control study on the risk factors for narcolepsy with special focus on vaccinations (including pandemic influenza vaccination) and infections in Germany.

    PubMed

    Oberle, Doris; Pavel, Jutta; Mayer, Geert; Geisler, Peter; Keller-Stanislawski, Brigitte

    2017-06-01

    Studies associate pandemic influenza vaccination with narcolepsy. In Germany, a retrospective, multicenter, matched case-control study was performed to identify risk factors for narcolepsy, particularly regarding vaccinations (seasonal and pandemic influenza vaccination) and infections (seasonal and pandemic influenza) and to quantify the detected risks. Patients with excessive daytime sleepiness who had been referred to a sleep center between April 2009 and December 2012 for multiple sleep latency test (MSLT) were eligible. Case report forms were validated according to the criteria for narcolepsy defined by the Brighton Collaboration (BC). Confirmed cases of narcolepsy (BC level of diagnostic certainty 1-4a) were matched with population-based controls by year of birth, gender, and place of residence. A second control group was established including patients in whom narcolepsy was definitely excluded (test-negative controls). A total of 103 validated cases of narcolepsy were matched with 264 population-based controls. The second control group included 29 test-negative controls. A significantly increased odd ratio (OR) to develop narcolepsy (crude OR [cOR] = 3.9, 95% confidence interval [CI] = 1.8-8.5; adjusted OR [aOR] = 4.5, 95% CI = 2.0-9.9) was detected in individuals immunized with pandemic influenza A/H1N1/v vaccine prior to symptoms onset as compared to nonvaccinated individuals. Using test-negative controls, in individuals immunized with pandemic influenza A/H1N1/v vaccine prior to symptoms onset, a nonsignificantly increased OR of narcolepsy was detected when compared to nonvaccinated individuals (whole study population, BC levels 1-4a: cOR = 1.9, 95% CI = 0.5-6.9; aOR = 1.8, 95% CI = 0.3-10.1). The findings of this study support an increased risk for narcolepsy after immunization with pandemic influenza A/H1N1/v vaccine. Copyright © 2017 Elsevier B.V. All rights reserved.

  10. [Clinical effect of atomoxetine hydrochloride in 66 children with narcolepsy].

    PubMed

    Zhang, Shen; Ding, Changhong; Wu, Husheng; Fang, Fang; Wang, Xiaohui; Ren, Xiaotun

    2015-10-01

    To observe the efficacy and safety of atomoxetine hydrochloride in children with narcolepsy. Totally 66 patients with narcolepsy who were conformed international classification of sleep disturbances (ICSD-2) diagnostic criteria treated with atomoxetine hydrochloride seen from November 2010 to December 2014 were enrolled into this study, 42 of them were male and 24 female, mean age of onset was 7.5 years (3.75-13.00 years), mean duration before diagnosis was 1.75 years (0.25-5.00 years). Complete blood count, liver and kidney function, multiple sleep latency test (MSLT), polysomnography (PGS), neuroimaging and electroencephalography (EEG) were performed for each patient. For some of the children HLA-DR2 gene and serum markers of infection were tested. The 66 cases were followed up from 2 to 49 months (average 18 months) to observe the clinical efficacy and adverse reactions. In 62 cases excessive daytime sleepiness was improved, in 11 cases (16.7%) it was controlled (16.7%), in 29 cases (43.9%) the treatment was obviously effective and in 22 (33.3%) it was effective; cataplexy occurred in 54 cases, in 18 (33.3%) it was controlled, in 19 (35.2%) the treatment was obviously effective and in 10 (18.5%) effective; night sleep disorders existed in 55 cases, in 47 cases it was improved, in 14 (25.5%) it was controlled, in 20 (36.4%) the treatment was obviously effective and in 13 (23.6%) effective; hypnagogic or hypnopompic hallucination was present in 13 cases, in only 4 these symptoms were controlled. Sleep paralysis existed in 4 cases, it was controlled in only 1 case. In 18 cases attention and learning efficiency improved.Anorexia occurred in 18 cases, mood disorder in 5 cases, depression in 2 cases, nocturia, muscle tremors, involuntary tongue movement each occurred in 1 case. P-R interval prolongation and atrial premature contraction were found in 1 case. Atomoxetine hydrochloride showed good effects in patients with narcolepsy on excessive daytime sleepiness

  11. Reducing the Clinical and Socioeconomic Burden of Narcolepsy by Earlier Diagnosis and Effective Treatment.

    PubMed

    Thorpy, Michael; Morse, Anne Marie

    2017-03-01

    The burden of narcolepsy is likely the result of 2 main aspects: the clinical difficulties and disability incurred as a direct effect of the disorder and the socioeconomic burden. The clinical burden includes the symptoms, diagnosis, comorbidities, treatment, and even mortality that can be associated with narcolepsy. Lifelong therapy is necessary for these patients. Effective treatment results in long-term benefits from both patient and societal perspectives by improving clinical outcomes, potentially enabling improved education and increased employment and work productivity, and quality of life. Thus, reducing the time to appropriate management results in improved outcomes in these patients. Copyright © 2017 Elsevier Inc. All rights reserved.

  12. Improved vigilance after sodium oxybate treatment in narcolepsy: a comparison between in-field and in-laboratory measurements.

    PubMed

    van Schie, Mojca K M; Werth, Esther; Lammers, Gert Jan; Overeem, Sebastiaan; Baumann, Christian R; Fronczek, Rolf

    2016-08-01

    This two-centre observational study of vigilance measurements assessed the feasibility of vigilance measurements on multiple days using the Sustained Attention to Response Task and the Psychomotor Vigilance Test with portable task equipment, and subsequently assessed the effect of sodium oxybate treatment on vigilance in patients with narcolepsy. Twenty-six patients with narcolepsy and 15 healthy controls were included. The study comprised two in-laboratory days for the Maintenance of Wakefulness Test and the Oxford Sleep Resistance test, followed by 7-day portable vigilance battery measurements. This procedure was repeated for patients with narcolepsy after at least 3 months of stable treatment with sodium oxybate. Patients with narcolepsy had a higher Sustained Attention to Response Task error count, lower Psychomotor Vigilance Test reciprocal reaction time, higher Oxford Sleep Resistance test omission error count adjusted for test duration (Oxford Sleep Resistance testOMIS / MIN ), and lower Oxford Sleep Resistance test and Maintenance of Wakefulness Test sleep latency compared with controls (all P < 0.01). Treatment with sodium oxybate was associated with a longer Maintenance of Wakefulness Test sleep latency (P < 0.01), lower Oxford Sleep Resistance testOMIS / MIN (P = 0.01) and a lower Sustained Attention to Response Task error count (P = 0.01) in patients with narcolepsy, but not with absolute changes in Oxford Sleep Resistance test sleep latency or Psychomotor Vigilance Test reciprocal reaction time. It was concluded that portable measurements of sustained attention as well as in-laboratory Oxford Sleep Resistance test and Maintenance of Wakefulness Test measurements revealed worse performance for narcoleptic patients compared with controls, and that sodium oxybate was associated with an improvement of sustained attention and a better resistance to sleep. © 2016 European Sleep Research Society.

  13. Sleep-stage sequencing of sleep-onset REM periods in MSLT predicts treatment response in patients with narcolepsy.

    PubMed

    Drakatos, Panagis; Patel, Kishankumar; Thakrar, Chiraag; Williams, Adrian J; Kent, Brian D; Leschziner, Guy D

    2016-04-01

    Current treatment recommendations for narcolepsy suggest that modafinil should be used as a first-line treatment ahead of conventional stimulants or sodium oxybate. In this study, performed in a tertiary sleep disorders centre, treatment responses were examined following these recommendations, and the ability of sleep-stage sequencing of sleep-onset rapid eye movement periods in the multiple sleep latency test to predict treatment response. Over a 3.5-year period, 255 patients were retrospectively identified in the authors' database as patients diagnosed with narcolepsy, type 1 (with cataplexy) or type 2 (without) using clinical and polysomnographic criteria. Eligible patients were examined in detail, sleep study data were abstracted and sleep-stage sequencing of sleep-onset rapid eye movement periods were analysed. Response to treatment was graded utilizing an internally developed scale. Seventy-five patients were included (39% males). Forty (53%) were diagnosed with type 1 narcolepsy with a mean follow-up of 2.37 ± 1.35 years. Ninety-seven percent of the patients were initially started on modafinil, and overall 59% reported complete response on the last follow-up. Twenty-nine patients (39%) had the sequence of sleep stage 1 or wake to rapid eye movement in all of their sleep-onset rapid eye movement periods, with most of these diagnosed as narcolepsy type 1 (72%). The presence of this specific sleep-stage sequence in all sleep-onset rapid eye movement periods was associated with worse treatment response (P = 0.0023). Sleep-stage sequence analysis of sleep-onset rapid eye movement periods in the multiple sleep latency test may aid the prediction of treatment response in narcoleptics and provide a useful prognostic tool in clinical practice, above and beyond their classification as narcolepsy type 1 or 2. © 2015 European Sleep Research Society.

  14. Abnormal activity in reward brain circuits in human narcolepsy with cataplexy.

    PubMed

    Ponz, Aurélie; Khatami, Ramin; Poryazova, Rositsa; Werth, Esther; Boesiger, Peter; Bassetti, Claudio L; Schwartz, Sophie

    2010-02-01

    Hypothalamic hypocretins (or orexins) regulate energy metabolism and arousal maintenance. Recent animal research suggests that hypocretins may also influence reward-related behaviors. In humans, the loss of hypocretin-containing neurons results in a major sleep-wake disorder called narcolepsy-cataplexy, which is associated with emotional disturbances. Here, we aim to test whether narcoleptic patients show an abnormal pattern of brain activity during reward processing. We used functional magnetic resonance imaging in 12 unmedicated patients with narcolepsy-cataplexy to measure the neural responses to expectancy and experience of monetary gains and losses. We statistically compared the patients' data with those obtained in a group of 12 healthy matched controls. Our results reveal that activity in the dopaminergic ventral midbrain (ventral tegmental area) was not modulated in narcolepsy-cataplexy patients during high reward expectancy (unlike controls), and that ventral striatum activity was reduced during winning. By contrast, the patients showed abnormal activity increases in the amygdala and in dorsal striatum for positive outcomes. In addition, we found that activity in the nucleus accumbens and the ventral-medial prefrontal cortex correlated with disease duration, suggesting that an alternate neural circuit could be privileged over the years to control affective responses to emotional challenges and compensate for the lack of influence from ventral midbrain regions. Our study offers a detailed picture of the distributed brain network involved during distinct stages of reward processing and shows for the first time, to our knowledge, how this network is affected in hypocretin-deficient narcoleptic patients.

  15. Practice Parameters for the Treatment of Narcolepsy and other Hypersomnias of Central Origin An American Academy of Sleep Medicine Report

    PubMed Central

    Morgenthaler, Timothy I.; Kapur, Vishesh K.; Brown, Terry; Swick, Todd J.; Alessi, Cathy; Aurora, R. Nisha; Boehlecke, Brian; Chesson, Andrew L.; Friedman, Leah; Maganti, Rama; Owens, Judith; Pancer, Jeffrey; Zak, Rochelle

    2007-01-01

    These practice parameters pertain to the treatment of hypersomnias of central origin. They serve as both an update of previous practice parameters for the therapy of narcolepsy and as the first practice parameters to address treatment of other hypersomnias of central origin. They are based on evidence analyzed in the accompanying review paper. The specific disorders addressed by these parameters are narcolepsy (with cataplexy, without cataplexy, due to medical condition and unspecified), idiopathic hypersomnia (with long sleep time and without long sleep time), recurrent hypersomnia and hypersomnia due to medical condition. Successful treatment of hypersomnia of central origin requires an accurate diagnosis, individual tailoring of therapy to produce the fullest possible return of normal function, and regular follow-up to monitor response to treatment. Modafinil, sodium oxybate, amphetamine, methamphetamine, dextroamphetamine, methylphenidate, and selegiline are effective treatments for excessive sleepiness associated with narcolepsy, while tricyclic antidepressants and fluoxetine are effective treatments for cataplexy, sleep paralysis, and hypnagogic hallucinations; but the quality of published clinical evidence supporting them varies. Scheduled naps can be beneficial to combat sleepiness in narcolepsy patients. Based on available evidence, modafinil is an effective therapy for sleepiness due to idiopathic hypersomnia, Parkinson's disease, myotonic dystrophy, and multiple sclerosis. Based on evidence and/or long history of use in the therapy of narcolepsy committee consensus was that modafinil, amphetamine, methamphetamine, dextroamphetamine, and methylphenidate are reasonable options for the therapy of hypersomnias of central origin. Citation: Morgenthaler TI; Kapur VK; Brown T; Swick TJ; Alessi C; Aurora RN; Boehlecke B; Chesson AL; Friedman L; Maganti R; Owens J; Pancer J; Zak R; Standards of Practice Committee of the AASM. Practice parameters for the treatment

  16. Comorbidity of Narcolepsy Type 1 With Autoimmune Diseases and Other Immunopathological Disorders: A Case-Control Study

    PubMed Central

    Martinez-Orozco, Francisco Javier; Vicario, Jose Luis; De Andres, Clara; Fernandez-Arquero, Miguel; Peraita-Adrados, Rosa

    2016-01-01

    Background Several evidences suggest that autoimmune diseases (ADs) tend to co-occur in an individual and within the same family. Narcolepsy type 1 (NT1) is a chronic sleep disorder caused by a selective loss of hypocretin-producing neurons due to a mechanism of neural destruction that indicates an autoimmune pathogenesis, although no evidence is available. We report on the comorbidity of ADs and other immunopathological diseases (including allergy diseases) in narcolepsy. Methods We studied 158 Caucasian NT1 patients (60.7% male; mean age 49.4 ± 19.7 years), in whom the diagnosis was confirmed by polysomnography followed by a multiple sleep latency test, or by hypocretin-1 levels measurements. Results Thirty out of 158 patients (18.99%; 53.3% female; 29 sporadic and one familial cases) had one or more immunopathological diseases associated. A control group of 151 subjects were matched by gender and age with the narcolepsy patients. Results demonstrated that there was a higher frequency of ADs in our series of narcolepsy patients compared to the sample of general population (odds ratio: 3.17; 95% confidence interval: 1.01 - 10.07; P = 0.040). A temporal relationship with the age at onset of the diseases was found. Conclusions Cataplexy was significantly more severe in NT1 patients with immunopathological diseases, and immunopathological diseases are a risk factor for severe forms of cataplexy in our series (odds ratio: 23.6; 95% confidence interval: 5.5 - 100.1). PMID:27298657

  17. Sleep-wake patterns, non-rapid eye movement, and rapid eye movement sleep cycles in teenage narcolepsy.

    PubMed

    Xu, Xing; Wu, Huijuan; Zhuang, Jianhua; Chen, Kun; Huang, Bei; Zhao, Zhengqing; Zhao, Zhongxin

    2017-05-01

    To further characterize sleep disorders associated with narcolepsy, we assessed the sleep-wake patterns, rapid eye movement (REM), and non-REM (NREM) sleep cycles in Chinese teenagers with narcolepsy. A total of 14 Chinese type 1 narcoleptic patients (13.4 ± 2.6 years of age) and 14 healthy age- and sex-matched control subjects (13.6 ± 1.8 years of age) were recruited. Ambulatory 24-h polysomnography was recorded for two days, with test subjects adapting to the instruments on day one and the study data collection performed on day two. Compared with the controls, the narcoleptic patients showed a 1.5-fold increase in total sleep time over 24 h, characterized by enhanced slow-wave sleep and REM sleep. Frequent sleep-wake transitions were identified in nocturnal sleep with all sleep stages switching to wakefulness, with more awakenings and time spent in wakefulness after sleep onset. Despite eight cases of narcolepsy with sleep onset REM periods at night, the mean duration of NREM-REM sleep cycle episode and the ratio of REM/NREM sleep between patients and controls were not significantly different. Our study identified hypersomnia in teenage narcolepsy despite excessive daytime sleepiness. Sleep fragmentation extended to all sleep stages, indicating impaired sleep-wake cycles and instability of sleep stages. The limited effects on NREM-REM sleep cycles suggest the relative conservation of ultradian regulation of sleep. Copyright © 2016 Elsevier B.V. All rights reserved.

  18. 78 FR 43209 - Narcolepsy Public Meeting on Patient-Focused Drug Development

    Federal Register 2010, 2011, 2012, 2013, 2014

    2013-07-19

    ... DEPARTMENT OF HEALTH AND HUMAN SERVICES Food and Drug Administration [Docket No. FDA-2013-N-0815] Narcolepsy Public Meeting on Patient-Focused Drug Development AGENCY: Food and Drug Administration, HHS. ACTION: Notice of public meeting; request for comments. SUMMARY: The Food and Drug Administration (FDA...

  19. Continuous intrathecal orexin delivery inhibits cataplexy in a murine model of narcolepsy.

    PubMed

    Kaushik, Mahesh K; Aritake, Kosuke; Imanishi, Aya; Kanbayashi, Takashi; Ichikawa, Tadashi; Shimizu, Tetsuo; Urade, Yoshihiro; Yanagisawa, Masashi

    2018-06-05

    Narcolepsy-cataplexy is a chronic neurological disorder caused by loss of orexin (hypocretin)-producing neurons, associated with excessive daytime sleepiness, sleep attacks, cataplexy, sleep paralysis, hypnagogic hallucinations, and fragmentation of nighttime sleep. Currently, human narcolepsy is treated by providing symptomatic therapies, which can be associated with an array of side effects. Although peripherally administered orexin does not efficiently penetrate the blood-brain barrier, centrally delivered orexin can effectively alleviate narcoleptic symptoms in animal models. Chronic intrathecal drug infusion through an implantable pump is a clinically available strategy to treat a number of neurological diseases. Here we demonstrate that the narcoleptic symptoms of orexin knockout mice can be reversed by lumbar-level intrathecal orexin delivery. Orexin was delivered via a chronically implanted intrathecal catheter at the upper lumbar level. The computed tomographic scan confirmed that intrathecally administered contrast agent rapidly moved from the spinal cord to the brain. Intrathecally delivered orexin was detected in the brain by radioimmunoassay at levels comparable to endogenous orexin levels. Cataplexy and sleep-onset REM sleep were significantly decreased in orexin knockout mice during and long after slow infusion of orexin (1 nmol/1 µL/h). Sleep/wake states remained unchanged both quantitatively as well as qualitatively. Intrathecal orexin failed to induce any changes in double orexin receptor-1 and -2 knockout mice. This study supports the concept of intrathecal orexin delivery as a potential therapy for narcolepsy-cataplexy to improve the well-being of patients.

  20. Narcolepsy type 1 and hypersomnia associated with a psychiatric disorder show different slow wave activity dynamics.

    PubMed

    Walacik-Ufnal, Ewa; Piotrowska, Anna Justyna; Wołyńczyk-Gmaj, Dorota; Januszko, Piotr; Gmaj, Bartłomiej; Ufnal, Marcin; Kabat, Marek; Wojnar, Marcin

    2017-01-01

    The aim of the study was to compare electrophysiological parameters of night sleep in narcolepsy type 1 and hypersomnia associated with a psychiatric disorder. Fortyfour patients: 15 with narcolepsy type 1, 14 with hypersomnia associated with a psychiatric disorder and 15 age- and sex-matched controls participated in the study. The study subjects filled in the Athens Insomnia Scale (AIS) and the Beck Depression Inventory (BDI). The severity of daytime sleepiness was quantified subjectively using the Epworth Sleepiness Scale (ESS) and the Stanford Sleepiness Scale (SSS), and objectively using the Multiple Sleep Latency Test (MSLT). All subjects underwent polysomnography (PSG) on the two consecutive nights. The data from the second night was analysed. The slow wave activity (SWA, 1-4 Hz) was calculated for the three consecutive sleep cycles, and topographic delta power maps were plotted. In contrast to narcoleptics, psychiatric hypersomniacs had undisturbed nocturnal sleep, high sleep efficiency, normal non-rapid eye movement (NREM) and rapid eye movement (REM) sleep proportions, normal REM latency and sleep latencies on MSLT and PSG. The subjective and objective sleepiness was significantly higher in narcolepsy group than in psychiatric hypersomnia group. In all the study groups SWA was the most prominent in frontal areas, while the greatest between-group differences were found in the central areas. There were significant differences between the groups in SWA in the second NREM episode. The highest SWA was observed in the hypersomnia group, while the lowest in the narcolepsy group. Psychiatric hypersomniacs and controls did not differ in the SWA exponential decline over consecutive NREM episodes, whereas narcoleptics exhibited a steeper dissipation of sleep pressure from the first to the second NREM episode. In conclusion, narcolepsy type1 and hypersomnia associated with psychiatric disorder differ in the SWA dynamics. Narcoleptics presented with the altered

  1. Clinical utility of the Chinese version of the Pediatric Daytime Sleepiness Scale in children with obstructive sleep apnea syndrome and narcolepsy.

    PubMed

    Yang, Chien-Ming; Huang, Yu-Shu; Song, Yu-Chen

    2010-04-01

    The present study examined the psychometric properties of the Chinese version of the Pediatric Daytime Sleepiness Scale (PDSS) and the utility of the PDSS as a screening tool for pathological daytime sleepiness in teenagers with obstructive sleep apnea (OSA) and narcolepsy. The PDSS was first administered to 238 middle and high school students to assess the reliability of the scale, and then administered to 28 teenagers with OSA, 31 teenagers with narcolepsy, and 34 normal controls to evaluate its clinical utility. Test-retest reliability and internal consistency were acceptable. The PDSS scores were significantly higher in narcoleptic subjects than in subjects with OSA, and higher in OSA syndrome (OSAS) subjects than normal controls. Furthermore, the scores decreased in narcoleptic subjects after medical treatment. Both reliability and validity were proven to be good. As a screening tool for narcolepsy, receiver operator characteristic (ROC) curve analysis showed that the PDSS, with a cut-off score of 16/17, had good sensitivity (87.1%) and fair specificity (74.3%) for identifying individuals with narcolepsy. When used for screening OSA, however, the differentiating power was not as good. The PDSS is a reliable and valid tool for the measurement of sleepiness in clinical youth populations. When used as a screening tool, it is useful for sleep disorders involving more severe pathological sleepiness, as in narcolepsy.

  2. Narcolepsy with Cataplexy Mimicry: The Strange Case of Two Sisters

    PubMed Central

    Pizza, Fabio; Vandi, Stefano; Poli, Francesca; Moghadam, Keivan Kaveh; Franceschini, Christian; Bellucci, Claudia; Cipolli, Carlo; Ingravallo, Francesca; Natalini, Giuliana; Mignot, Emmanuel; Plazzi, Giuseppe

    2013-01-01

    We report on two sisters, 17 and 12 years of age, with clinical features suggesting narcolepsy with cataplexy (NC): daytime sleepiness, spontaneous and emotionally triggered sudden falls to the ground, and overweight/obesity. MSLT showed borderline sleep latency, with 1 and 0 sleep onset REM periods. HLA typing disclosed the DQB1*0602 allele. Video-polygraphy of the spells ruled out NC diagnosis by demonstrating their easy elicitation by suggestion, with wake EEG, electromyo-graphic persistence of muscle tone, and stable presence of tendon reflexes (i.e., pseudo-cataplexy), together with normal cerebrospinal hypocretin-1 levels. Our cases emphasize the need of a clear depiction of cataplexy pattern at the different ages, the usefulness of examining ictal neurophysiology, and collecting all available disease markers in ambiguous cases. Citation: Pizza F; Vandi S; Poli F; Moghadam KK; Fran-ceschini C; Bellucci C; Cipolli C; Ingravallo F; Natalini G; Mignot E; Plazzi G. Narcolepsy with cataplexy mimicry: the strange case of two sisters. J Clin Sleep Med 2013;9(6):611-612. PMID:23772196

  3. Modafinil : A Review of its Pharmacology and Clinical Efficacy in the Management of Narcolepsy.

    PubMed

    McClellan, K J; Spencer, C M

    1998-04-01

    Modafinil promotes wakefulness through an as yet unknown mechanism of action. It significantly increases daytime sleep latency and reduces excessive daytime sleepiness (EDS) compared with placebo in patients with narcolepsy. However, the drug does not suppress cataplexy. Although direct comparative data are lacking, modafinil offers advantages over amphetamines and methylphenidate in patients with narcolepsy because of its lack of rebound phenomena after treatment withdrawal and its low abuse potential. Clinical trials have shown modafinil to be well tolerated in patients with narcolepsy. Except for headache, which was reported with a significantly greater frequency in modafinil recipients, the tolerability profile of modafinil 200 to 400 mg/day was similar to that of placebo in patients treated for 9 weeks. Preliminary data suggest that the tolerability of modafinil is maintained long term (40 weeks). Thus, modafinil is effective in the treatment of EDS in patients with narcolepsy, although it is not effective against cataplexy. Preliminary findings indicate that, unlike other psychostimulants, the drug is unlikely to be abused and is not associated with withdrawal phenomena. Therefore, modafinil is likely to be an effective therapeutic option for the treatment of EDS in patients with narcolepsy. The mechanism of action of modafinil has not been clearly established. However, it may indirectly increase wakefulness, at least in part, through inhibition of cortical γ-aminobutyric acid (GABA) release via serotonergic mechanisms. Modafinil induces wakefulness and increases locomotor activity in a variety of animal species without causing stereotyped behaviour. In rhesus monkeys, the effects of oral modafinil were not associated with changes in blood pressure or heart rate. In contrast to dexamphetamine 20mg, single night-time doses of modafinil 100 or 200mg had no significant effects on objective sleep variables or sleep structure in young or elderly healthy

  4. Narcolepsy: Autoimmunity, Effector T Cell Activation Due to Infection, or T Cell Independent, Major Histocompatibility Complex Class II Induced Neuronal Loss?

    ERIC Educational Resources Information Center

    Fontana, Adriano; Gast, Heidemarie; Reith, Walter; Recher, Mike; Birchler, Thomas; Bassetti, Claudio L.

    2010-01-01

    Human narcolepsy with cataplexy is a neurological disorder, which develops due to a deficiency in hypocretin producing neurons in the hypothalamus. There is a strong association with human leucocyte antigens HLA-DR2 and HLA-DQB1*0602. The disease typically starts in adolescence. Recent developments in narcolepsy research support the hypothesis of…

  5. Efficiency of a Combination of Pharmacological Treatment and Nondrug Interventions in Childhood Narcolepsy.

    PubMed

    Kacar Bayram, Ayşe; Per, Hüseyin; Ismailoğullari, Sevda; Canpolat, Mehmet; Gumus, Hakan; Aksu, Murat

    2016-12-01

    Objective  Narcolepsy is a chronic sleep disorder characterized by excessive daytime sleepiness, cataplexy, hypnagogic and/or hypnopompic hallucinations, and sleep paralysis. It is one of the most important causes of excessive daytime sleepiness in the pediatric population. The aim of this study is to present the clinical and laboratory findings, and treatment results of pediatric patients with narcolepsy. Materials and Methods  We studied five unrelated consecutive children with narcolepsy, focusing on clinical and laboratory features, the therapy and outcome over the 33-month follow-up period. Results  The study subjects included two boys and three girls. The mean age at diagnosis was 11.8 ± 3.3 years (range: 8-16 years). Three patients had cataplexy. There were no hypnagogic hallucinations and/or sleep paralysis in any patients. All patients were educated about sleep hygiene, appropriate nutrition, and regular exercise. Three patients were treated with modafinil, while two patients received methylphenidate. Sodium oxybate was added to existing treatment in patients with cataplexy. Cataplexy attacks did not respond well to the treatment in one patient; therefore intravenous immunoglobulin therapy was given. Conclusions  Early diagnosis is important to help narcoleptic patients in improving their quality of life. A combination of pharmacological treatment and nondrug interventions can greatly improve children's clinical symptoms. Georg Thieme Verlag KG Stuttgart · New York.

  6. Patterns of spontaneous reports on narcolepsy following administration of pandemic influenza vaccine; a case series of individual case safety reports in Eudravigilance.

    PubMed

    Gadroen, Kartini; Straus, Sabine M J M; Pacurariu, Alexandra; Weibel, Daniel; Kurz, Xavier; Sturkenboom, Miriam C J M

    2016-09-22

    This study aims to describe the frequency and quality of spontaneous narcolepsy case reports following administration of pandemic influenza vaccine as captured in the Eudravigilance database. We conducted a retrospective descriptive study of spontaneous Individual Case Safety Reports (ICSRs), reporting narcolepsy following administration of pandemic influenza vaccine as received by Eudravigilance until July 2014. De-duplication was carried out by Eudravigilance. Frequency of reporting is described as number of ICSRs received per month over time. The quality of the ICSRs was evaluated by completeness of information and diagnostic certainty using the Automated Brighton Collaboration case definition tool (ABC-tool) for narcolepsy. After de-duplication, a total of 1333 ICSRs of narcolepsy and/or cataplexy following pandemic influenza vaccine were identified, originating from 18 countries worldwide. Most of the ICSRs (61.9%) originated from the signaling countries, Sweden and Finland. Although de-duplication of case reports was carried out, it is suspected that many duplicates exist, in particular from Sweden. The majority of the ICSRs (95.3%), reported exposure to Pandemrix®. Only few reports were received for Arepanrix® (1.6%) or Focetria® (0.5%), and Celvapan® (0.1%). Of those ICSRs reporting age, 73.1% concerned persons below age of 20years. When using the ABC-tool, all ICSRs were classified as having insufficient information to meet the Brighton Collaboration case definition of narcolepsy. An increase in reporting of narcolepsy appeared in Eudravigilance only after awareness was raised by the national authorities. Most narcolepsy reports were received from countries where the signal initially occurred, and were related to Pandemrix® in children/adolescents. Basic information about the patient and the exposure was present in most of the ICSRs. The ICSRs captured by Eudravigilance however, do not collect enough information to assess the diagnostic certainty

  7. The spectrum of REM sleep-related episodes in children with type 1 narcolepsy.

    PubMed

    Antelmi, Elena; Pizza, Fabio; Vandi, Stefano; Neccia, Giulia; Ferri, Raffaele; Bruni, Oliviero; Filardi, Marco; Cantalupo, Gaetano; Liguori, Rocco; Plazzi, Giuseppe

    2017-06-01

    Type 1 narcolepsy is a central hypersomnia due to the loss of hypocretin-producing neurons and characterized by cataplexy, excessive daytime sleepiness, sleep paralysis, hypnagogic hallucinations and disturbed nocturnal sleep. In children, close to the disease onset, type 1 narcolepsy has peculiar clinical features with severe cataplexy and a complex admixture of movement disorders occurring while awake. Motor dyscontrol during sleep has never been systematically investigated. Suspecting that abnormal motor control might affect also sleep, we systematically analysed motor events recorded by means of video polysomnography in 40 children with type 1 narcolepsy (20 females; mean age 11.8 ± 2.6 years) and compared these data with those recorded in 22 age- and sex-matched healthy controls. Motor events were classified as elementary movements, if brief and non-purposeful and complex behaviours, if simulating purposeful behaviours. Complex behaviours occurring during REM sleep were further classified as 'classically-defined' and 'pantomime-like' REM sleep behaviour disorder episodes, based on their duration and on their pattern (i.e. brief and vivid-energetic in the first case, longer and with subcontinuous gesturing mimicking daily life activity in the second case). Elementary movements emerging either from non-REM or REM sleep were present in both groups, even if those emerging from REM sleep were more numerous in the group of patients. Conversely, complex behaviours could be detected only in children with type 1 narcolepsy and were observed in 13 patients, with six having 'classically-defined' REM sleep behaviour disorder episodes and seven having 'pantomime-like' REM sleep behaviour disorder episodes. Complex behaviours during REM sleep tended to recur in a stereotyped fashion for several times during the night, up to be almost continuous. Patients displaying a more severe motor dyscontrol during REM sleep had also more severe motor disorder during daytime (i

  8. Treatment with venlafaxine in six cases of children with narcolepsy and with cataplexy and hypnagogic hallucinations.

    PubMed

    Møller, Lene Ruge; Østergaard, John R

    2009-04-01

    Narcolepsy with cataplexy is a chronic neuropsychiatric disorder associated with inappropriate control of rapid eye movement (REM) sleep. The main symptoms are excessive daytime sleepiness, cataplexy, hypnagogic hallucinations, and disturbed nocturnal sleep. Cataplexy is marked by episodes of muscular weakness and may cause the patient to collapse to the ground. So far, pharmacotherapy of cataplexy and hypnagogic hallucinations has been predominantly based on tricyclic antidepressants. Recently, new drugs that block the reuptake of norepineprine and serotonin (e.g., venlafaxine) have been suggested as first-line treatment. These drugs have become our choice in treating children with cataplexy and nightmares as a symptom in narcolepsy. We describe clinical case reports of venlafaxine treatment in 6 children aged 7-12 years old when diagnosed with narcolepsy-cataplexy. In 2 cases with up to 50 daily cataplectic attacks, an initial effect of 37.5 mg of venlafaxine was initially observed. However, during the first year, the dose had to be increased to 112.5 mg daily to avoid cataplexy. A third patient with partial cataplexy was treated with 75 mg of venlafaxine daily. In 2 cases, hypnagogic hallucinations, described by the patients as nightmares, were the most troubling symptom and were successfully treated with only 37.5 mg of venlafaxine daily. Side effects included an increase of disturbed nocturnal sleep when venlafaxine was taken after 2:00 p.m. No major aggressive or suicidal thoughts and no raised blood pressure were recorded. Venlafaxine has proven to be an effective treatment of cataplexy and hypnagogic hallucinations in 6 children with narcolepsy. No severe side effects were observed.

  9. English Translations Of The First Clinical Reports On Narcolepsy And Cataplexy By Westphal And Gélineau In The Late 19th Century, With Commentary

    PubMed Central

    Schenck, Carlos H.; Bassetti, Claudio L.; Arnulf, Isabelle; Mignot, Emmanuel

    2007-01-01

    Study Objectives: To publish the first English translations, with commentary, of the original reports describing narcolepsy and cataplexy by Westphal in German (1877) and by Gélineau in French (1880). Methods: A professional translation service translated the 2 reports from either German or French to English, with each translation then being slightly edited by one of the authors. All authors then provided commentary. Results: Both Westphal and Gélineau correctly identified and described the new clinical entities of cataplexy and narcolepsy, with recurrent, self-limited sleep attacks and/or cataplectic attacks affecting 2 otherwise healthy people. Narcolepsy was named by Gélineau (and cataplexy was named by Henneberg in 1916). The evidence in both cases is sufficiently convincing to conclude that they were likely each HLA-DQB1*0602 positive and hypocretin deficient. Conclusions: The original descriptions of narcolepsy and cataplexy are now available in English, allowing for extensive clinical and historical commentary. Citations: Schenck CH; Bassetti CL; Arnulf I et al. English translations of the first clinical reports on narcolepsy and cataplexy by Westphal and Gélineau in the late 19th century, with commentary. J Clin Sleep Med 2007;3(3):301–311 PMID:17561602

  10. The Nightly Use of Sodium Oxybate Is Associated with a Reduction in Nocturnal Sleep Disruption: A Double-Blind, Placebo-Controlled Study in Patients with Narcolepsy

    PubMed Central

    Black, Jed; Pardi, Daniel; Hornfeldt, Carl S.; Inhaber, Neil

    2010-01-01

    Objective: To further explore the effects of sodium oxybate (SXB) administration on nocturnal sleep in narcolepsy patients during a double-blind, placebo-controlled, parallel group study conducted with 228 adult patients with narcolepsy/cataplexy in the United States, Canada, and Europe. Method: Patients were withdrawn from antidepressants and sedative/hypnotics, and then randomized to receive 4.5, 6, or 9 g SXB or placebo nightly for 8 weeks. Patients receiving 6 and 9 g/night doses were titrated to their final dose in weekly 1.5 g increments, while patients receiving placebo were randomized to undergo a similar mock dose titration. The use of stimulant therapy continued unchanged. Changes in sleep architecture were measured using centrally scored nocturnal polysomnograms. Daily diaries were used to record changes in narcolepsy symptoms and adverse events. Results: Following 8 weeks of SXB treatment, study patients demonstrated significant dose-related increases in the duration of stage 3 and 4 sleep, reaching a median increase of 52.5 minutes in patients receiving 9 g nightly. Compared to placebo-treated patients, delta power was significantly increased in all dose groups. Stage 1 sleep and the frequency of nocturnal awakenings were each significantly decreased at the 6 and 9 g/night doses. The changes in nocturnal sleep coincided with significant decreases in the severity and frequency of narcolepsy symptoms. Conclusions: The nightly administration of SXB to narcolepsy patients significantly impacts measures of slow wave sleep, wake after sleep onset, awakenings, total sleep time, and stage 1 sleep in a dose-related manner. The frequency and severity of narcolepsy symptoms decreased with treatment. Citation: Black J; Pardi D; Hornfeldt CS; Inhaber N. The nightly use of sodium oxybate is associated with a reduction in nocturnal sleep disruption: a double-blind, placebo-controlled study in patients with narcolepsy. J Clin Sleep Med 2010;6(6):596-602. PMID:21206549

  11. Where are we in our understanding of the association between narcolepsy and one of the 2009 adjuvanted influenza A (H1N1) vaccines?

    PubMed

    Johansen, K; Brasseur, D; MacDonald, N; Nohynek, H; Vandeputte, J; Wood, D; Neels, P

    2016-07-01

    Evaluating new rare serious vaccine safety signals is difficult and complex work. To further assess the observed increase in narcolepsy cases seen in Europe with the 2009 pandemic H1N1 influenza vaccine, the International Alliance for Biological Standardization (IABS) invited a wide range of experts to a one day meeting in Geneva in October 2015 to present data and to discuss the implications. The presentations covered the following topics: clinical picture of childhood narcolepsy following the 2009 H1N1 pandemic vaccination campaigns; epidemiological studies conducted to assess the risk of narcolepsy, other neurological and immune-related diseases following 2009 pandemic H1N1 influenza vaccine; potential biases influencing the different epidemiological study designs; potential genetic contribution to the development of narcolepsy; potential biological mechanisms for development of narcolepsy in this setting including the role of the virus itself, antigenic differences between the vaccines and differences in AS03-adjuvanted vaccines. The presentations were followed by fulsome roundtable discussions. Members from affected families also attended and made informal comments to round out the day's deliberations. This meeting emphasized the value added in bringing together in a neutral setting a wide range of experts and vaccine producers to discuss such a complex new serious adverse event following immunization. Copyright © 2016.

  12. Normal Morning Melanin-Concentrating Hormone Levels and No Association with Rapid Eye Movement or Non-Rapid Eye Movement Sleep Parameters in Narcolepsy Type 1 and Type 2

    PubMed Central

    Schrölkamp, Maren; Jennum, Poul J.; Gammeltoft, Steen; Holm, Anja; Kornum, Birgitte R.; Knudsen, Stine

    2017-01-01

    Study Objectives: Other than hypocretin-1 (HCRT-1) deficiency in narcolepsy type 1 (NT1), the neurochemical imbalance of NT1 and narcolepsy type 2 (NT2) with normal HCRT-1 levels is largely unknown. The neuropeptide melanin-concentrating hormone (MCH) is mainly secreted during sleep and is involved in rapid eye movement (REM) and non-rapid eye movement (NREM) sleep regulation. Hypocretin neurons reciprocally interact with MCH neurons. We hypothesized that altered MCH secretion contributes to the symptoms and sleep abnormalities of narcolepsy and that this is reflected in morning cerebrospinal fluid (CSF) MCH levels, in contrast to previously reported normal evening/afternoon levels. Methods: Lumbar CSF and plasma were collected from 07:00 to 10:00 from 57 patients with narcolepsy (subtypes: 47 NT1; 10 NT2) diagnosed according to International Classification of Sleep Disorders, Third Edition (ICSD-3) and 20 healthy controls. HCRT-1 and MCH levels were quantified by radioimmunoassay and correlated with clinical symptoms, polysomnography (PSG), and Multiple Sleep Latency Test (MSLT) parameters. Results: CSF and plasma MCH levels were not significantly different between narcolepsy patients regardless of ICSD-3 subtype, HCRT-1 levels, or compared to controls. CSF MCH and HCRT-1 levels were not significantly correlated. Multivariate regression models of CSF MCH levels, age, sex, and body mass index predicting clinical, PSG, and MSLT parameters did not reveal any significant associations to CSF MCH levels. Conclusions: Our study shows that MCH levels in CSF collected in the morning are normal in narcolepsy and not associated with the clinical symptoms, REM sleep abnormalities, nor number of muscle movements during REM or NREM sleep of the patients. We conclude that morning lumbar CSF MCH measurement is not an informative diagnostic marker for narcolepsy. Citation: Schrölkamp M, Jennum PJ, Gammeltoft S, Holm A, Kornum BR, Knudsen S. Normal morning melanin

  13. Risk of Narcolepsy after AS03 Adjuvanted Pandemic A/H1N1 2009 Influenza Vaccine in Adults: A Case-Coverage Study in England.

    PubMed

    Stowe, Julia; Andrews, Nicholas; Kosky, Christopher; Dennis, Gary; Eriksson, Sofia; Hall, Andrew; Leschziner, Guy; Reading, Paul; Shneerson, John M; Donegan, Katherine; Miller, Elizabeth

    2016-05-01

    An increased risk of narcolepsy has been observed in children following ASO3-adjuvanted pandemic A/H1N1 2009 (Pandemrix) vaccine. We investigated whether this risk extends to adults in England. Six adult sleep centers in England were visited between November 2012 and February 2014 and vaccination/clinical histories obtained from general practitioners. Suspected narcolepsy cases aged older than 17 y were selected. The risk of narcolepsy following Pandemrix was calculated using cases diagnosed by the time of the center visits and those with a diagnosis by November 30, 2011 after which there was increased awareness of the risk in children. The odds of vaccination in cases and in matched population data were compared using a case-coverage design. Of 1,446 possible cases identified, most had onset before 2009 or were clearly not narcolepsy. Of the 60 remaining cases, 20 were excluded after expert review, leaving 40 cases with narcolepsy; 5 had received Pandemrix between 3 and 18 mo before onset. All the vaccinated cases had cataplexy, two received a diagnosis by November 2011 and two were aged 40 y or older. The odds ratio for vaccination in cases compared to the population was 4.24 (95% confidence interval 1.45-12.38) using all cases and 9.06 (1.90-43.17) using cases with a diagnosis by November 2011, giving an attributable risk of 0.59 cases per 100,000 doses. We found a significantly increased risk of narcolepsy in adults following Pandemrix vaccination in England. The risk was lower than that seen in children using a similar study design. © 2016 Associated Professional Sleep Societies, LLC.

  14. The relationship between orexin levels and blood pressure changes in patients with narcolepsy.

    PubMed

    Sieminski, Mariusz; Chwojnicki, Kamil; Sarkanen, Tomi; Partinen, Markku

    2017-01-01

    Narcolepsy type 1 (NT1) is caused by a deficiency or absence of the neurotransmitter orexin. NT1 is also associated with a reduced nocturnal "dipping" of blood pressure (BP). The study objective was to analyze whether nocturnal BP values differed in patients depleted of orexin, versus those in whom production was preserved. We performed a retrospective analysis of the polysomnographic recordings, orexin levels, and BP values of patients with NT1. Data was collected from a total of 21 patients, divided into two groups as follows: those with a complete depletion of orexin (n = 11) (Group1), and those with a remaining, limited presence of orexin (n = 10) (Group 2). The groups did not differ in terms of the clinical features of NT1 or sleep characteristics, with an exception of increased number of cataplexy episodes and increased percentage of sleep stage 2 in the Group 1. Daytime and nocturnal BP did not differ between the groups. Most patients, regardless of group, had a non-dipping blood pressure pattern, and no difference in dipping prevalence was observed between groups. The amplitude of the daytime to nighttime change in BP did not differ between the groups. Non-dipping BP patterns are frequent among patients with narcolepsy type 1, but we saw no evidence that they depended on whether orexin levels were above or below the assay detection threshold. Therefore, our results do not support the hypothesis that in patients with narcolepsy type 1 residual orexin levels play a role in the control of nocturnal BP dipping.

  15. Narcolepsy-cataplexy and loss of sphincter control.

    PubMed Central

    Vgontzas, A. N.; Sollenberger, S. E.; Kales, A.; Bixler, E. O.; Vela-Bueno, A.

    1996-01-01

    We describe the case of a 34-year-old man who presented intermittent faecal incontinence as a manifestation of cataplexy. The patient's sleep history was positive for the full narcoleptic tetrad (sleep attacks, cataplexy, sleep paralysis and hypnagogic hallucinations) while extensive neuropsychiatric work up was negative for any neurologic or psychiatric illness. Repeat polysomnograms (including a polysomnogram with a full seizure montage) were positive for pathologic sleepiness, but there was no evidence of a seizure disorder. The course of the patient's symptomatology and the favourable response of his symptoms to stimulants and imipramine support the theory that his intermittent loss of sphincter control is part of his narcolepsy-cataplexy. PMID:8796217

  16. No Evidence for Disease History as a Risk Factor for Narcolepsy after A(H1N1)pdm09 Vaccination.

    PubMed

    Lamb, Favelle; Ploner, Alexander; Fink, Katharina; Maeurer, Markus; Bergman, Peter; Piehl, Fredrik; Weibel, Daniel; Sparén, Pär; Dahlström, Lisen Arnheim

    2016-01-01

    To investigate disease history before A(H1N1)pdm09 vaccination as a risk factor for narcolepsy. Case-control study in Sweden. Cases included persons referred for a Multiple Sleep Latency Test between 2009 and 2010, identified through diagnostic sleep centres and confirmed through independent review of medical charts. Controls, selected from the total population register, were matched to cases on age, gender, MSLT-referral date and county of residence. Disease history (prescriptions and diagnoses) and vaccination history was collected through telephone interviews and population-based healthcare registers. Conditional logistic regression was used to investigate disease history before A(H1N1)pdm09 vaccination as a risk-factor for narcolepsy. In total, 72 narcolepsy cases and 251 controls were included (range 3-69 years mean19-years). Risk of narcolepsy was increased in individuals with a disease history of nervous system disorders (OR range = 3.6-8.8) and mental and behavioural disorders (OR = 3.8, 95% CI 1.6-8.8) before referral. In a second analysis of vaccinated individuals only, nearly all initial associations were no longer statistically significant and effect sizes were smaller (OR range = 1.3-2.6). A significant effect for antibiotics (OR = 0.4, 95% CI 0.2-0.8) and a marginally significant effect for nervous system disorders was observed. In a third case-only analysis, comparing cases referred before vaccination to those referred after; prescriptions for nervous system disorders (OR = 26.0 95% CI 4.0-170.2) and ADHD (OR = 35.3 95% CI 3.4-369.9) were statistically significant during the vaccination period, suggesting initial associations were due to confounding by indication. The findings of this study do not support disease history before A(H1N1)pdm09 vaccination as a risk factor for narcolepsy.

  17. Identifying clinically important difference on the Epworth Sleepiness Scale: results from a narcolepsy clinical trial of JZP-110.

    PubMed

    Scrima, Lawrence; Emsellem, Helene A; Becker, Philip M; Ruoff, Chad; Lankford, Alan; Bream, Gary; Khayrallah, Moise; Lu, Yuan; Black, Jed

    2017-10-01

    While scores ≤10 on the Epworth Sleepiness Scale (ESS) are within the normal range, the reduction in elevated ESS score that is clinically meaningful in patients with narcolepsy has not been established. This post hoc analysis of a clinical trial of patients with narcolepsy evaluated correlations between Patient Global Impression of Change (PGI-C) and ESS. Data of adult patients with narcolepsy from a double-blind, 12-week placebo-controlled study of JZP-110, a wake-promoting agent, were used in this analysis. Descriptive statistics and receiver operating characteristic (ROC) analysis compared PGI-C (anchor measure) to percent change from baseline in ESS to establish the responder criterion from patients taking either placebo or JZP-110 (treatments). At week 12, patients (n = 10) who reported being "very much improved" on the PGI-C had a mean 76.7% reduction in ESS score, and patients (n = 33) who reported being "much improved" on the PGI-C had a mean 49.1% reduction in ESS score. ROC analysis showed that patients who improved were almost exclusively from JZP-110 treatment group, with an area-under-the-curve of 0.9, and revealed that a 25% reduction in ESS (sensitivity, 81.4%; specificity, 80.9%) may be an appropriate threshold for defining a meaningful patient response to JZP-110 and placebo. A ≥25% reduction in patients' subjective ESS score may be useful as a threshold to identify patients with narcolepsy who respond to JZP-110 treatment. Copyright © 2017 The Authors. Published by Elsevier B.V. All rights reserved.

  18. Normal Morning Melanin-Concentrating Hormone Levels and No Association with Rapid Eye Movement or Non-Rapid Eye Movement Sleep Parameters in Narcolepsy Type 1 and Type 2.

    PubMed

    Schrölkamp, Maren; Jennum, Poul J; Gammeltoft, Steen; Holm, Anja; Kornum, Birgitte R; Knudsen, Stine

    2017-02-15

    Other than hypocretin-1 (HCRT-1) deficiency in narcolepsy type 1 (NT1), the neurochemical imbalance of NT1 and narcolepsy type 2 (NT2) with normal HCRT-1 levels is largely unknown. The neuropeptide melanin-concentrating hormone (MCH) is mainly secreted during sleep and is involved in rapid eye movement (REM) and non-rapid eye movement (NREM) sleep regulation. Hypocretin neurons reciprocally interact with MCH neurons. We hypothesized that altered MCH secretion contributes to the symptoms and sleep abnormalities of narcolepsy and that this is reflected in morning cerebrospinal fluid (CSF) MCH levels, in contrast to previously reported normal evening/afternoon levels. Lumbar CSF and plasma were collected from 07:00 to 10:00 from 57 patients with narcolepsy (subtypes: 47 NT1; 10 NT2) diagnosed according to International Classification of Sleep Disorders, Third Edition (ICSD-3) and 20 healthy controls. HCRT-1 and MCH levels were quantified by radioimmunoassay and correlated with clinical symptoms, polysomnography (PSG), and Multiple Sleep Latency Test (MSLT) parameters. CSF and plasma MCH levels were not significantly different between narcolepsy patients regardless of ICSD-3 subtype, HCRT-1 levels, or compared to controls. CSF MCH and HCRT-1 levels were not significantly correlated. Multivariate regression models of CSF MCH levels, age, sex, and body mass index predicting clinical, PSG, and MSLT parameters did not reveal any significant associations to CSF MCH levels. Our study shows that MCH levels in CSF collected in the morning are normal in narcolepsy and not associated with the clinical symptoms, REM sleep abnormalities, nor number of muscle movements during REM or NREM sleep of the patients. We conclude that morning lumbar CSF MCH measurement is not an informative diagnostic marker for narcolepsy. © 2017 American Academy of Sleep Medicine

  19. Prevalence of the HLA-DQB1*0602 allele in narcolepsy and idiopathic hypersomnia patients seen at a sleep disorders outpatient unit in São Paulo.

    PubMed

    Coelho, Fernando Morgadinho Santos; Pradella-Hallinan, Márcia; Predazzoli Neto, Mario; Bittencourt, Lia Rita Azeredo; Tufik, Sérgio

    2009-03-01

    Narcolepsy (with and without cataplexy) and idiopathic hypersomnia, are disorders with common features but with different HLA-DQB1*0602 allele prevalence. The present study describes the prevalence of HLA-DQB1*0602 allele in narcoleptics with and without cataplexy and in patients with idiopathic hypersomnia. Subjects comprised 68 patients who were diagnosed for narcolepsy or idiopathic hypersomnia and 23 healthy controls according to the International Classification of Sleep Disorders-2. Subjects comprised 43 patients with narcolepsy and cataplexy, 11 patients with narcolepsy but without cataplexy, 14 patients with idiopathic hypersomnia and 23 healthy controls. Genotyping of HLA-DQB1*0602 allele was performed for all subjects. The prevalence of the HLA-DQB1*0602 allele was increased in idiopathic hypersomnia and in narcoleptic patients with and without cataplexy when compared to healthy subjects (p = 0.04; p = 0.03 and p < 0.0001, respectively). This finding is in accordance with those of previous studies. The gold standard exam of narcolepsy with cataplexy is Hypocretin-1 dosage, but in patients without cataplexy and idiopathic hypersomnia, there are no specific diagnostic lab findings. The presence of the HLA-DQB1* 0602 allele may be important for the differential diagnosis of situations that resemble those sleep disorders such as secondary changes in sleep structure due to drugs' consumption.

  20. Psychiatric comorbidity and cognitive profile in children with narcolepsy with or without association to the H1N1 influenza vaccination.

    PubMed

    Szakács, Attila; Hallböök, Tove; Tideman, Pontus; Darin, Niklas; Wentz, Elisabet

    2015-04-01

    To evaluate psychiatric comorbidity and the cognitive profile in children and adolescents with narcolepsy in western Sweden and the relationship of these problems to H1N1 vaccination. Thirty-eight patients were included in the study. We performed a population-based, cross-sectional study to investigate psychiatric comorbidity using a test battery of semistructured interviews generating Diagnostic and Statistical Manual of Mental Disorders, 4th Edition diagnoses, including the Development and Well-Being Assessment and the attention deficit hyperactivity disorder rating scale. The Autism Spectrum Screening Questionnaire and the Positive and Negative Syndrome Scale were used to screen for autistic traits and psychotic symptoms, respectively. The cognitive assessments were made by a clinical psychologist using the Wechsler Preschool and Primary Scale of Intelligence, Third Edition, the Wechsler Intelligence Scale for Children, Fourth Edition, or the Wechsler Adult Intelligence Scale, Fourth Edition. In the post-H1N1 vaccination (PHV) narcolepsy group (n = 31), 43% of patients had psychiatric comorbidity, 29% had attention deficit hyperactivity disorder (ADHD) inattentive type, 20% had major depression, 10% had general anxiety disorder, 7% had oppositional defiant disorder (ODD), 3% had pervasive developmental disorder not otherwise specified (i.e., atypical autism), and 3% had eating disorder not otherwise specified (anorectic type). In the non-post-H1N1 vaccination (nPHV) narcolepsy group, one of seven patients had ADHD, inattentive type and ODD. The most frequent psychiatric symptom was temper tantrums, which occurred in 94% of the patients in the PHV group and 71% of the patients in the nPHV narcolepsy group. The cognitive assessment profile was similar in both groups and showed normal results for mean full-scale IQ and perceptual speed but decreased verbal comprehension and working memory. Patients with psychiatric comorbidity had a significantly lower full

  1. Quality of life and trust among young people with narcolepsy and their families, after the Pandemrix® vaccination: protocol for a case-control study.

    PubMed

    Blomberg, Karin; Carlsson, Agneta Anderzén; Hagberg, Lars; Jonsson, Östen; Leissner, Lena; Eriksson, Mats H

    2017-08-23

    The extensive vaccination programme against swine flu resulted in an increased incidence of narcolepsy among children and adolescents. There is a need to explore if these young persons' experiences have affected their trust in healthcare, their willingness to participate in future prevention programmes, and their contacts with the healthcare system. The overall aim is to identify factors important for the life-situation of children and adolescents with narcolepsy and their families, and factors that correlate with trust in healthcare. Data will be collected via questionnaires from all available children with narcolepsy following the vaccination and their families, as well as a control group of children with diabetes and their families. Longitudinal descriptive interviews will also be conducted with a selection of 20-25 children and their families. Techniques from media research will be used for Internet-based data collection and analysis of information relating to narcolepsy from social media. This project will use the situation of young persons with narcolepsy after the swine flu vaccination as a case to build a model that can be used in situations where trust in healthcare is essential. This model will be based on findings from the included studies on how trust is influenced by support, quality of life, burden of disease, impact on family, and use of social media. The model developed in this project will be beneficial in future situations where trust in healthcare is essential, such as new pandemic outbreaks but also for "everyday" adherence to health advice.

  2. Assessing narcolepsy with cataplexy in children and adolescents: development of a cataplexy diary and the ESS-CHAD

    PubMed Central

    Wang, Y Grace; Benmedjahed, Khadra; Lambert, Jérémy; Evans, Christopher J; Hwang, Steve; Black, Jed; Johns, Murray W

    2017-01-01

    Objective The aim of this study was to qualitatively evaluate concepts for incorporation into a daily diary to capture cataplexy frequency and to assess the content validity of the Epworth Sleepiness Scale for Children and Adolescents (ESS-CHAD) in pediatric patients with narcolepsy. Patients and methods Face-to-face concept elicitation and cognitive interviews were conducted with children (7–9 years; n=13) and adolescents (10–17 years; n=16) who have narcolepsy with cataplexy, and their parents/caregivers. Results Similarities and differences were noted between narcolepsy concepts described by children and their parents/caregivers, suggesting some different but complementary perspectives; parents may not recognize cataplexy symptoms/triggers as well as children, but parents have greater recognition of the circumstances of falling asleep. Cataplexy diary modifications included changes in definitions and examples of cataplexy, using child-friendly terminology, adding a quantitative question to determine daily frequency, and standardizing the questionnaire for evening administration with self-completion by the child. Modifications were made to ESS-CHAD for child-friendly wording and to ensure that items reflect activities (eating, watching TV/video) and environments (school, bus/car transport) in which children are likely to participate. Two ESS-CHAD versions were proposed: one with a 1-month recall period, for general use, and the other with a recall period of “since your last study visit,” for research, which could be shorter or longer than 1 month (as short as 1 week). Conclusion The cataplexy diary and ESS-CHAD were modified for the assessment of children and adolescents. Further psychometric validation is recommended. These measures are being used in a Phase III, placebo-controlled clinical trial of sodium oxybate in children and adolescents with narcolepsy. PMID:28860883

  3. Increased β-haemolytic group A streptococcal M6 serotype and streptodornase B-specific cellular immune responses in Swedish narcolepsy cases.

    PubMed

    Ambati, A; Poiret, T; Svahn, B-M; Valentini, D; Khademi, M; Kockum, I; Lima, I; Arnheim-Dahlström, L; Lamb, F; Fink, K; Meng, Q; Kumar, A; Rane, L; Olsson, T; Maeurer, M

    2015-09-01

    Type 1 narcolepsy is a neurological disorder characterized by excessive daytime sleepiness and cataplexy associated with the HLA allele DQB1*06:02. Genetic predisposition along with external triggering factors may drive autoimmune responses, ultimately leading to the selective loss of hypocretin-positive neurons. The aim of this study was to investigate potential aetiological factors in Swedish cases of postvaccination (Pandemrix) narcolepsy defined by interferon-gamma (IFNγ) production from immune cells in response to molecularly defined targets. Cellular reactivity defined by IFNγ production was examined in blood from 38 (HLA-DQB1*06:02(+) ) Pandemrix-vaccinated narcolepsy cases and 76 (23 HLA-DQB1*06:02(+) and 53 HLA-DQB1*06:02(-) ) control subjects, matched for age, sex and exposure, using a variety of different antigens: β-haemolytic group A streptococcal (GAS) antigens (M5, M6 and streptodornase B), influenza (the pandemic A/H1N1/California/7/09 NYMC X-179A and A/H1N1/California/7/09 NYMC X-181 vaccine antigens, previous Flu-A and -B vaccine targets, A/H1N1/Brisbane/59/2007, A/H1N1/Solomon Islands/3/2006, A/H3N2/Uruguay/716/2007, A/H3N2/Wisconsin/67/2005, A/H5N1/Vietnam/1203/2004 and B/Malaysia/2506/2004), noninfluenza viral targets (CMVpp65, EBNA-1 and EBNA-3) and auto-antigens (hypocretin peptide, Tribbles homolog 2 peptide cocktail and extract from rat hypothalamus tissue). IFN-γ production was significantly increased in whole blood from narcolepsy cases in response to streptococcus serotype M6 (P = 0.0065) and streptodornase B protein (P = 0.0050). T-cell recognition of M6 and streptodornase B was confirmed at the single-cell level by intracellular cytokine (IL-2, IFNγ, tumour necrosis factor-alpha and IL-17) production after stimulation with synthetic M6 or streptodornase B peptides. Significantly, higher (P = 0.02) titres of serum antistreptolysin O were observed in narcolepsy cases, compared to vaccinated controls. β-haemolytic GAS may be

  4. Characterization of REM sleep without atonia in patients with narcolepsy and idiopathic hypersomnia using AASM scoring manual criteria.

    PubMed

    DelRosso, Lourdes M; Chesson, Andrew L; Hoque, Romy

    2013-07-15

    The AASM Manual for the Scoring of Sleep and Associated Events (Manual) has provided standardized definitions for tonic and phasic REM sleep without atonia (RSWA). This study used Manual criteria to characterize REM sleep in patients with narcolepsy and idiopathic hypersomnia (IH). A retrospective review of PSG data from ICSD-2 defined patients with narcolepsy or IH, performed by two board certified sleep medicine physicians. Data compiled included REM sleep epochs and the presence in REM sleep of epochs scored as sustained muscle activity (tonic), and excessive transient muscle activity (phasic) as defined by Manual criteria. PSG data from 8 narcolepsy patients (mean age: 27.5 years; age range: 11-55) showed mean ± standard deviation values for: total REM sleep epochs 205 ± 46.1; RSWA/ phasic epochs 56.1 ± 25.4; and RSWA/tonic epochs 15.0 ± 10.7. PSG data from 8 IH patients (mean age: 33.1 years; age range: 20-57) showed mean ± standard deviation values of total REM sleep epochs 163.8 ± 67.9; RSWA/phasic epochs 6.2 ± 3.5; and RSWA/tonic epochs 0.2 ± 0.4. Comparison revealed intergroup differences in phasic REM sleep (p < 0.01) and tonic REM sleep (p < 0.01) were significantly increased in narcoleptics compared to IH. Our retrospective analysis showed that RSWA phasic activity and RSWA tonic activity are significantly increased in patients meeting ICSD-2 criteria for narcolepsy compared to patients meeting ICSD-2 criteria for IH. This robust difference, with further validation, could be useful as electrophysiological criteria differentiating the two disorders and understanding the physiological differences.

  5. Diagnostic value of sleep stage dissociation as visualized on a 2-dimensional sleep state space in human narcolepsy.

    PubMed

    Olsen, Anders Vinther; Stephansen, Jens; Leary, Eileen; Peppard, Paul E; Sheungshul, Hong; Jennum, Poul Jørgen; Sorensen, Helge; Mignot, Emmanuel

    2017-04-15

    Type 1 narcolepsy (NT1) is characterized by symptoms believed to represent Rapid Eye Movement (REM) sleep stage dissociations, occurrences where features of wake and REM sleep are intermingled, resulting in a mixed state. We hypothesized that sleep stage dissociations can be objectively detected through the analysis of nocturnal Polysomnography (PSG) data, and that those affecting REM sleep can be used as a diagnostic feature for narcolepsy. A Linear Discriminant Analysis (LDA) model using 38 features extracted from EOG, EMG and EEG was used in control subjects to select features differentiating wake, stage N1, N2, N3 and REM sleep. Sleep stage differentiation was next represented in a 2D projection. Features characteristic of sleep stage differences were estimated from the residual sleep stage probability in the 2D space. Using this model we evaluated PSG data from NT1 and non-narcoleptic subjects. An LDA classifier was used to determine the best separation plane. This method replicates the specificity/sensitivity from the training set to the validation set better than many other methods. Eight prominent features could differentiate narcolepsy and controls in the validation dataset. Using a composite measure and a specificity cut off 95% in the training dataset, sensitivity was 43%. Specificity/sensitivity was 94%/38% in the validation set. Using hypersomnia subjects, specificity/sensitivity was 84%/15%. Analyzing treated narcoleptics the specificity/sensitivity was 94%/10%. Sleep stage dissociation can be used for the diagnosis of narcolepsy. However the use of some medications and presence of undiagnosed hypersomnolence patients impacts the result. Copyright © 2017 Elsevier B.V. All rights reserved.

  6. Safety and efficacy of long-term use of sodium oxybate for narcolepsy with cataplexy in routine clinical practice.

    PubMed

    Drakatos, Panagis; Lykouras, Dimosthenis; D'Ancona, Grainne; Higgins, Sean; Gildeh, Nadia; Macavei, Raluca; Rosenzweig, Ivana; Steier, Joerg; Williams, Adrian J; Muza, Rexford; Kent, Brian D; Leschziner, Guy

    2017-07-01

    Sodium oxybate is licensed in Europe for the treatment of narcolepsy with cataplexy in adults. The aim of this study was to assess the efficacy and safety of sodium oxybate in clinical practice in patients with narcolepsy and cataplexy refractory to other treatments. This was a retrospective single centre study including patients with severe narcolepsy with cataplexy refractory to other treatments, who were initiated on sodium oxybate between 2009 and 2015. Patients were allowed to be on other stimulants or/and anti-cataplectic agents. Epworth sleepiness scale (ESS) and weekly cataplexy events were recorded. Side effects (SEs) were recorded at every follow-up visit. 90 patients were prescribed sodium oxybate, with a total of 3116 patient-months of drug exposure. ESS and weekly cataplexy events were significantly reduced by sodium oxybate for all patients (ΔESS = 4.3 ± 4.4 and Δcataplexy = 21.8 ± 18.5 events/week; p < 0.0001, respectively). The required maintenance dose could not be predicted based upon gender, body mass index, or clinical factors. 60% of patients were able to reduce or come off other medications. Half of the patients experienced at least one SE, and 26.6% had to stop treatment due to limiting SEs. Nausea, mood swings and enuresis were the most commonly reported SEs. SEs that led to drug discontinuation, particularly psychosis, were associated with increasing age and were observed early after the initiation of the drug. Sodium oxybate provides a good clinical efficacy and acceptable safety profile in routine clinical practice for the treatment of patients suffering from narcolepsy with cataplexy. A quarter of patients experience SEs requiring withdrawal of the drug with older patients being more vulnerable to the more serious SEs. Copyright © 2017 The Authors. Published by Elsevier B.V. All rights reserved.

  7. [18F]Fludeoxyglucose-Positron Emission Tomography Evidence for Cerebral Hypermetabolism in the Awake State in Narcolepsy and Idiopathic Hypersomnia.

    PubMed

    Dauvilliers, Yves; Evangelista, Elisa; de Verbizier, Delphine; Barateau, Lucie; Peigneux, Philippe

    2017-01-01

    Changes in structural and functional central nervous system have been reported in narcolepsy, with large discrepancies between studies. No study has investigated yet spontaneous brain activity at wake in idiopathic hypersomnia (IH). We compared relative changes in regional brain metabolism in two central hypersomnia conditions with different clinical features, namely narcolepsy type 1 (NT1) and IH, and in healthy controls. Sixteen patients [12 males, median age 30 years (17-78)] with NT1, nine patients [2 males, median age 27 years (20-60)] with IH and 19 healthy controls [16 males, median age 36 years (17-78)] were included. 18 F-fludeoxyglucose positron emission tomography (PET) was performed in all drug-free subjects under similar conditions and instructions to stay in a wake resting state. We found increased metabolism in the anterior and middle cingulate and the insula in the two pathological conditions as compared to healthy controls. The reverse contrast failed to evidence hypometabolism in patients vs. controls. Comparisons between patient groups were non-significant. At sub-statistical threshold, we found higher right superior occipital gyrus glucose metabolism in narcolepsy and higher middle orbital cortex and supplementary motor area metabolism in IH, findings that require further confirmation. There is significant hypermetabolism in narcolepsy and IH in the wake resting state in a set of brain regions constitutive of the salience cortical network that may reflect a compensatory neurocircuitry activity secondary to sleepiness. Metabolic differences between the two disorders within the executive-control network may be a signature of abnormally functioning neural system leading to persistent drowsiness typical of IH.

  8. Core Body and Skin Temperature in Type 1 Narcolepsy in Daily Life; Effects of Sodium Oxybate and Prediction of Sleep Attacks

    PubMed Central

    van der Heide, Astrid; Werth, Esther; Donjacour, Claire E.H.M.; Reijntjes, Robert H.A.M.; Lammers, Gert Jan; Van Someren, Eus J.W.; Baumann, Christian R.; Fronczek, Rolf

    2016-01-01

    Study Objectives: Previous laboratory studies in narcolepsy patients showed altered core body and skin temperatures, which are hypothesised to be related to a disturbed sleep wake regulation. In this ambulatory study we assessed temperature profiles in normal daily life, and whether sleep attacks are heralded by changes in skin temperature. Furthermore, the effects of three months of treatment with sodium oxybate (SXB) were investigated. Methods: Twenty-five narcolepsy patients and 15 healthy controls were included. Core body, proximal and distal skin temperatures, and sleep-wake state were measured simultaneously for 24 hours in ambulatory patients. This procedure was repeated in 16 narcolepsy patients after at least 3 months of stable treatment with SXB. Results: Increases in distal skin temperature and distal-to-proximal temperature gradient (DPG) strongly predicted daytime sleep attacks (P < 0.001). As compared to controls, patients had a higher proximal and distal skin temperature in the morning, and a lower distal skin temperature during the night (all P < 0.05). Furthermore, they had a higher core body temperature during the first part of the night (P < 0.05), which SXB decreased (F = 4.99, df = 1, P = 0.03) to a level similar to controls. SXB did not affect skin temperature. Conclusions: This ambulatory study demonstrates that daytime sleep attacks were preceded by clear changes in distal skin temperature and DPG. Furthermore, changes in core body and skin temperature in narcolepsy, previously only studied in laboratory settings, were partially confirmed. Treatment with SXB resulted in a normalisation of the core body temperature profile. Future studies should explore whether predictive temperature changes can be used to signal or even prevent sleep attacks. Citation: van der Heide A, Werth E, Donjacour CE, Reijntjes RH, Lammers GJ, Van Someren EJ, Baumann CR, Fronczek R. Core body and skin temperature in type 1 narcolepsy in daily life; effects of sodium

  9. Decreased sleep stage transition pattern complexity in narcolepsy type 1.

    PubMed

    Ferri, Raffaele; Pizza, Fabio; Vandi, Stefano; Iloti, Martina; Plazzi, Giuseppe

    2016-08-01

    To analyze the complexity of the nocturnal sleep stage sequence in central disorders of hypersomnolence (CDH), with the hypothesis that narcolepsy type 1 (NT1) might exhibit distinctive sleep stage sequence organization and complexity. Seventy-nine NT1 patients, 22 narcolepsy type 2 (NT2), 22 idiopathic hypersomnia (IH), and 52 patients with subjective hypersomnolence (sHS) were recruited and their nocturnal sleep was polysomnographically recorded and scored. Group between-stage transition probability matrices were obtained and compared. Patients with NT1 differed significantly from all the other patient groups, the latter, in turn, were not different between each other. The individual probability of the R-to-N2 transition was found to be the parameter showing the difference of highest significance between the groups (lowest in NT1) and classified patients with or without NT1 with an accuracy of 78.9% (sensitivity 78.5% and specificity 79.2%), by applying a cut-off value of 0.15. The main result of this study is that the structure of the sleep stage transition pattern of hypocretin-deficient NT1 patients is significantly different from that of other forms of CDH and sHS, with normal hypocretin levels. The lower probability of R-to-N2 transition occurrence in NT1 appears to be a reliable polysomnographic feature with potential application at the individual level, for supportive diagnostic purposes. Copyright © 2016 International Federation of Clinical Neurophysiology. Published by Elsevier Ireland Ltd. All rights reserved.

  10. Health-related quality of life among drug-naïve patients with narcolepsy with cataplexy, narcolepsy without cataplexy, and idiopathic hypersomnia without long sleep time.

    PubMed

    Ozaki, Akiko; Inoue, Yuichi; Nakajima, Toru; Hayashida, Kenichi; Honda, Makoto; Komada, Yoko; Takahashi, Kiyohisa

    2008-12-15

    To evaluate the health-related quality life (HRQOL) of drug-naïve patients with narcolepsy with cataplexy (NAwith CA), narcolepsy without cataplexy (NA without CA) and idiopathic hypersomnia without long sleep time (IHS without LST), and to explore the factors influencing the HRQOL. Factors associated with the occurrence of automobile accidents are also discussed. A total of 137 consecutive drug naïve patients who met the criteria of the 2nd edition of the International Classification of Sleep Disorders (NA with CA, n = 28; NA without CA, n = 27; IHS without LST, n = 82) were enrolled. The patients were asked to fill out questionnaires, including the SF-36, Epworth Sleepiness Scale (ESS), sociodemographic variables, and items regarding driving habits and the experiences related to automobile accidents. All 3 diagnostic groups had significantly lower scores in most SF-36 domains compared with Japanese normative data. Significant differences among the 3 diagnostic groups were not observed. Specific factors in SF-36 domains were not found with multiple linear regression analyses, while disease duration was positively correlated with mental health among all subjects. Among the patients reporting driving habits, ESS score (> or =16) was positively associated with the experience of automobile accidents. Our results indicated that HRQOL decreases in drug-naïve patients with hypersomnia, but neither disease category nor severity of the disorder appears as an associated factor. Increased severity of hypersomnia, however, was thought to play an important role in the occurrence of automobile accidents.

  11. A coordinated cross-disciplinary research initiative to address an increased incidence of narcolepsy following the 2009-2010 Pandemrix vaccination programme in Sweden.

    PubMed

    Feltelius, N; Persson, I; Ahlqvist-Rastad, J; Andersson, M; Arnheim-Dahlström, L; Bergman, P; Granath, F; Adori, C; Hökfelt, T; Kühlmann-Berenzon, S; Liljeström, P; Maeurer, M; Olsson, T; Örtqvist, Å; Partinen, M; Salmonson, T; Zethelius, B

    2015-10-01

    In response to the 2009-2010 influenza A(H1N1)pdm09 pandemic, a mass vaccination programme with the AS03-adjuvanted influenza A(H1N1) vaccine Pandemrix was initiated in Sweden. Unexpectedly, there were a number of narcolepsy cases amongst vaccinated children and adolescents reported. In this review, we summarize the results of a joint cross-disciplinary national research effort to investigate the adverse reaction signal from the spontaneous reporting system and to better understand possible causative mechanisms. A three- to fourfold increased risk of narcolepsy in vaccinated children and adolescents was verified by epidemiological studies. Of importance, no risk increase was observed for the other neurological and autoimmune diseases studied. Genetic studies confirmed the association with the allele HLA-DQB1*06:02, which is known to be related to sporadic narcolepsy. Furthermore, a number of studies using cellular and molecular experimental models investigated possible links between influenza vaccination and narcolepsy. Serum analysis, using a peptide microarray platform, showed that individuals who received Pandemrix exhibited a different epitope reactivity pattern to neuraminidase and haemagglutinin, as compared to individuals who were infected with H1N1. Patients with narcolepsy were also found to have increased levels of interferon-gamma production in response to streptococcus-associated antigens. The chain of patient-related events and the study results emerging over time were subjected to intense nationwide media attention. The importance of transparent communication and collaboration with patient representatives to maintain public trust in vaccination programmes is also discussed in the review. Organizational challenges due to this unexpected event delayed the initiation of some of the research projects, still the main objectives of this joint, cross-disciplinary research effort were reached, and important insights were acquired for future, similar

  12. A patient with coexisting narcolepsy and morbid jealousy showing favourable response to fluoxetine.

    PubMed Central

    Wing, Y. K.; Lee, S.; Chiu, H. F.; Ho, C. K.; Chen, C. N.

    1994-01-01

    A 37 year old Chinese man suffered from coexisting narcolepsy and morbid jealousy which were precipitated by head injury 5 years previously. Fluoxetine 20 mg/day reduced his narcoleptic symptoms and morbid jealousy but not his sleepiness. On defaulting treatment, the patient's symptoms and marital problem recurred. A common central serotonin disturbance might be involved in mediating the sleep disorder and associated psychopathology. PMID:8140016

  13. Core Body and Skin Temperature in Type 1 Narcolepsy in Daily Life; Effects of Sodium Oxybate and Prediction of Sleep Attacks.

    PubMed

    van der Heide, Astrid; Werth, Esther; Donjacour, Claire E H M; Reijntjes, Robert H A M; Lammers, Gert Jan; Van Someren, Eus J W; Baumann, Christian R; Fronczek, Rolf

    2016-11-01

    Previous laboratory studies in narcolepsy patients showed altered core body and skin temperatures, which are hypothesised to be related to a disturbed sleep wake regulation. In this ambulatory study we assessed temperature profiles in normal daily life, and whether sleep attacks are heralded by changes in skin temperature. Furthermore, the effects of three months of treatment with sodium oxybate (SXB) were investigated. Twenty-five narcolepsy patients and 15 healthy controls were included. Core body, proximal and distal skin temperatures, and sleep-wake state were measured simultaneously for 24 hours in ambulatory patients. This procedure was repeated in 16 narcolepsy patients after at least 3 months of stable treatment with SXB. Increases in distal skin temperature and distal-to-proximal temperature gradient (DPG) strongly predicted daytime sleep attacks (P < 0.001). As compared to controls, patients had a higher proximal and distal skin temperature in the morning, and a lower distal skin temperature during the night (all P < 0.05). Furthermore, they had a higher core body temperature during the first part of the night (P < 0.05), which SXB decreased (F = 4.99, df = 1, P = 0.03) to a level similar to controls. SXB did not affect skin temperature. This ambulatory study demonstrates that daytime sleep attacks were preceded by clear changes in distal skin temperature and DPG. Furthermore, changes in core body and skin temperature in narcolepsy, previously only studied in laboratory settings, were partially confirmed. Treatment with SXB resulted in a normalisation of the core body temperature profile. Future studies should explore whether predictive temperature changes can be used to signal or even prevent sleep attacks. © 2016 Associated Professional Sleep Societies, LLC.

  14. The Psychosocial Problems of Children with Narcolepsy and Those with Excessive Daytime Sleepiness of Uncertain Origin

    ERIC Educational Resources Information Center

    Stores, Gregory; Montgomery, Paul; Wiggs, Luci

    2007-01-01

    Background: Narcolepsy is a predominantly rapid eye movement sleep disorder with onset usually in the second decade but often in earlier childhood. Classically it is characterized by combinations of excessive sleepiness especially sleep attacks, cataplexy, hypnagogic hallucinations, and sleep paralysis. The psychosocial effects of this lifelong…

  15. Disorders of Excessive Daytime Sleepiness Including Narcolepsy and Idiopathic Hypersomnia.

    PubMed

    Berkowski, Joseph Andrew; Shelgikar, Anita Valanju

    2016-09-01

    Central disorders of hypersomnolence are rare conditions with a poorly understood pathophysiology, making the identification and management challenging for sleep clinicians. Clinical history is essential for ruling out secondary causes of hypersomnolence and distinguishing among diagnoses. Current diagnostic criteria rely heavily on the polysomnogram and multiple sleep latency test. The current focus of treatment of hypersomnolence is on drugs that promote alertness. Additionally, in the case of narcolepsy type 1, medication management addresses control of cataplexy, the hallmark symptom of this disorder. Elucidation of pathophysiology of these disorders in the future will be essential to better categorization and management. Published by Elsevier Inc.

  16. Comparing Treatment Effect Measurements in Narcolepsy: The Sustained Attention to Response Task, Epworth Sleepiness Scale and Maintenance of Wakefulness Test.

    PubMed

    van der Heide, Astrid; van Schie, Mojca K M; Lammers, Gert Jan; Dauvilliers, Yves; Arnulf, Isabelle; Mayer, Geert; Bassetti, Claudio L; Ding, Claire-Li; Lehert, Philippe; van Dijk, J Gert

    2015-07-01

    To validate the Sustained Attention to Response Task (SART) as a treatment effect measure in narcolepsy, and to compare the SART with the Maintenance of Wakefulness Test (MWT) and the Epworth Sleepiness Scale (ESS). Validation of treatment effect measurements within a randomized controlled trial (RCT). Ninety-five patients with narcolepsy with or without cataplexy. The RCT comprised a double-blind, parallel-group, multicenter trial comparing the effects of 8-w treatments with pitolisant (BF2.649), modafinil, or placebo (NCT01067222). MWT, ESS, and SART were administered at baseline and after an 8-w treatment period. The severity of excessive daytime sleepiness and cataplexy was also assessed using the Clinical Global Impression scale (CGI-C). The SART, MWT, and ESS all had good reliability, obtained for the SART and MWT using two to three sessions in 1 day. The ability to distinguish responders from nonresponders, classified using the CGI-C score, was high for all measures, with a high performance for the SART (r = 0.61) and the ESS (r = 0.54). The Sustained Attention to Response Task is a valid and easy-to-administer measure to assess treatment effects in narcolepsy, enhanced by combining it with the Epworth Sleepiness Scale. © 2015 Associated Professional Sleep Societies, LLC.

  17. Test-Retest Reliability of the Multiple Sleep Latency Test in Narcolepsy without Cataplexy and Idiopathic Hypersomnia

    PubMed Central

    Trotti, Lynn Marie; Staab, Beth A.; Rye, David B.

    2013-01-01

    Study Objectives: Differentiation of narcolepsy without cataplexy from idiopathic hypersomnia relies entirely upon the multiple sleep latency test (MSLT). However, the test-retest reliability for these central nervous system hypersomnias has never been determined. Methods: Patients with narcolepsy without cataplexy, idiopathic hypersomnia, and physiologic hypersomnia who underwent two diagnostic multiple sleep latency tests were identified retrospectively. Correlations between the mean sleep latencies on the two studies were evaluated, and we probed for demographic and clinical features associated with reproducibility versus change in diagnosis. Results: Thirty-six patients (58% women, mean age 34 years) were included. Inter -test interval was 4.2 ± 3.8 years (range 2.5 months to 16.9 years). Mean sleep latencies on the first and second tests were 5.5 (± 3.7 SD) and 7.3 (± 3.9) minutes, respectively, with no significant correlation (r = 0.17, p = 0.31). A change in diagnosis occurred in 53% of patients, and was accounted for by a difference in the mean sleep latency (N = 15, 42%) or the number of sleep onset REM periods (N = 11, 31%). The only feature predictive of a diagnosis change was a history of hypnagogic or hypnopompic hallucinations. Conclusions: The multiple sleep latency test demonstrates poor test-retest reliability in a clinical population of patients with central nervous system hypersomnia evaluated in a tertiary referral center. Alternative diagnostic tools are needed. Citation: Trotti LM; Staab BA; Rye DB. Test- retest reliability of the multiple sleep latency test in narcolepsy without cataplexy and idiopathic hypersomnia. J Clin Sleep Med 2013;9(8):789-795. PMID:23946709

  18. Low-resolution brain electromagnetic tomography (LORETA) identifies brain regions linked to psychometric performance under modafinil in narcolepsy.

    PubMed

    Saletu, Michael; Anderer, Peter; Semlitsch, Heribert V; Saletu-Zyhlarz, Gerda Maria; Mandl, Magdalena; Zeitlhofer, Josef; Saletu, Bernd

    2007-01-15

    Low-resolution brain electromagnetic tomography (LORETA) showed a functional deterioration of the fronto-temporo-parietal network of the right hemispheric vigilance system in narcolepsy and a therapeutic effect of modafinil. The aim of this study was to determine the effects of modafinil on cognitive and thymopsychic variables in patients with narcolepsy and investigate whether neurophysiological vigilance changes correlate with cognitive and subjective vigilance alterations at the behavioral level. In a double-blind, placebo-controlled crossover design, EEG-LORETA and psychometric data were obtained during midmorning hours in 15 narcoleptics before and after 3 weeks of placebo or 400 mg modafinil. Cognitive investigations included the Pauli Test and complex reaction time. Thymopsychic/psychophysiological evaluation comprised drive, mood, affectivity, wakefulness, depression, anxiety, the Symptom Checklist 90 and critical flicker frequency. The Multiple Sleep Latency Test (MSLT) and the Epworth Sleepiness Scale (ESS) were performed too. Cognitive performance (Pauli Test) was significantly better after modafinil than after placebo. Concerning reaction time and thymopsychic variables, no significant differences were observed. Correlation analyses revealed that a decrease in prefrontal delta, theta and alpha-1 power correlated with an improvement in cognitive performance. Moreover, drowsiness was positively correlated with theta power in parietal and medial prefrontal regions and beta-1 and beta-2 power in occipital regions. A less significant correlation was observed between midmorning EEG LORETA and the MSLT; between EEG LORETA and the ESS, the correlation was even weaker. In conclusion, modafinil did not influence thymopsychic variables in narcolepsy, but it significantly improved cognitive performance, which may be related to medial prefrontal activity processes identified by LORETA.

  19. Tolerability and efficacy of armodafinil in naïve patients with excessive sleepiness associated with obstructive sleep apnea, shift work disorder, or narcolepsy: a 12-month, open-label, flexible-dose study with an extension period.

    PubMed

    Schwartz, Jonathan R L; Khan, Arifulla; McCall, W Vaughn; Weintraub, James; Tiller, Jane

    2010-10-15

    This 12-month, open-label, flexible-dose study with an extension period evaluated the tolerability and efficacy of armodafinil in patients with excessive sleepiness associated with treated obstructive sleep apnea (OSA), shift work disorder (SWD), or narcolepsy. Armodafinil-naïve, adult patients with excessive sleepiness associated with treated OSA (n = 170), SWD (n = 108), or narcolepsy (n = 50) received armodafinil (100-250 mg) once daily (treated OSA or narcolepsy) or before night shifts (SWD). Patients with OSA were regular users of continuous positive airway pressure (CPAP) therapy. Efficacy measures included the Clinical Global Impression of Improvement (CGI-I) and the Epworth Sleepiness Scale (ESS). Across the diagnosis groups, the most commonly occurring adverse event was headache (14%-24%). Forty-three patients (13%) and 13 patients (4%) were withdrawn because of adverse events and insufficient efficacy, respectively. Armodafinil did not adversely affect CPAP therapy. At the final visit, 80% (95% CI: 74.1, 86.7) of patients with treated OSA and 84% (72.7, 94.8) of patients with narcolepsy were rated on the CGI-I as at least minimally improved with regard to overall clinical condition; 98% (95.2, 100.0) of patients with SWD were rated as improved with regard to sleepiness during night shifts, including the commute to and from work. Armodafinil improved ESS total scores in patients with treated OSA (mean [SD] [95% CI] change from baseline, -7.3 [5.6] [-8.39, -6.30]) and patients with narcolepsy (-4.7 [6.0] [-7.41, -1.93]). Armodafinil administered for 12 months or more was generally well tolerated and improved wakefulness in patients with excessive sleepiness associated with treated OSA, SWD, or narcolepsy. Armodafinil improved the overall clinical condition of patients with treated OSA or narcolepsy.

  20. Growing Up with Type 1 Narcolepsy: Its Anthropometric and Endocrine Features

    PubMed Central

    Ponziani, Virginia; Gennari, Monia; Pizza, Fabio; Balsamo, Antonio; Bernardi, Filippo; Plazzi, Giuseppe

    2016-01-01

    Study Objectives: To evaluate the effect of type 1 narcolepsy (NT1) on anthropometric and endocrine features in childhood/adolescence, focusing on patterns and correlates of weight, pubertal development, and growth in treated and untreated patients. Methods: We collected anthropometric (height, weight, body mass index (BMI) z-scores), pubertal, metabolic, and endocrine data from 72 NT1 patients at diagnosis and all available premorbid anthropometric parameters of patients from their pediatric files (n = 30). New measurements at 1-y reassessment in patients undergoing different treatments were compared with baseline data. Results: We detected a high prevalence of overweight (29.2%), obesity (25%), metabolic syndrome (18.8%), and precocious puberty (16.1%), but no signs of linear growth alterations at diagnosis. According to anthropometric records, weight gain started soon after NT1 onset. At 1-y follow-up reassessment, sodium oxybate treatment was associated with a significant BMI z-score reduction (−1.29 ± 0.30, p < 0.0005) after adjusting for baseline age, sex, sleepiness, and BMI. Conclusions: NT1 onset in children/adolescents is associated with rapid weight gain up to overweight/obesity and precocious puberty without affecting growth. In our study, sodium oxybate treatment resulted in a significant weight reduction in NT1 overweight/obese patients at 1-y follow-up. Citation: Ponziani V, Gennari M, Pizza F, Balsamo A, Bernardi F, Plazzi G. Growing up with type 1 narcolepsy: its anthropometric and endocrine features. J Clin Sleep Med 2016;12(12):1649–1657. PMID:27707443

  1. Narcolepsy with cataplexy and hyperthyroidism sudden appeared after H1N1 vaccination

    PubMed Central

    Leiva, Silvia; Madrazo, Jimena; Podesta, Claudio

    2018-01-01

    Narcolepsy type 1 (NT1) is a chronic sleep disorder, characterized by excessive daytime sleepiness, cataplexy and fragmented nocturnal sleep. It is caused by a hypocretin deficiency due to a significant reduction of the neurons producing it. In the last years, it has been postulated that an autoimmune mechanism would be responsible for the destruction of these neurons in those genetically predisposed patients. The increased incidence of narcolepsy after the pandemic H1N1 influenza vaccination campaign in 2009-2010 is known. We present below the case of an adult patient who, 10 days after receiving H1N1 vaccination, suffers a traffic accident after falling asleep. Subsequent studies revealed hyperthyroidism due to Graves disease. In spite of the treatment, the patient persisted with daily and disabling daytime sleepiness, sleep attacks and episodes of generalized muscle atony with preservation of consciousness. A nocturnal polysomnography and multiple sleep latency test (MSLT) were performed with a diagnosis of NT1. The particularity of this case is the presentation of 2 autoimmune diseases triggered by an H1N1 vaccine without adjuvant, so far there is only evidence of NT1 associated with vaccines with adjuvant and viral infection. The association of both entities has made us reflect on the autoimmune mechanism, reinforcing the theory of its role in the onset of the disease. PMID:29796199

  2. Identification of the variations in the CPT1B and CHKB genes along with the HLA-DQB1*06:02 allele in Turkish narcolepsy patients and healthy persons.

    PubMed

    Cingoz, Sultan; Agilkaya, Sinem; Oztura, Ibrahim; Eroglu, Secil; Karadeniz, Derya; Evlice, Ahmet; Altungoz, Oguz; Yilmaz, Hikmet; Baklan, Baris

    2014-04-01

    The HLA-DQB1*06:02 allele across all ethnic groups and the rs5770917 variation between CPT1B and CHKB genes in Japanese and Koreans are common genetic susceptibility factors for narcolepsy. This comprehensive genetic study sought to assess variations in CHKB and CPT1B susceptibility genes and HLA-DQB1*06:02 allele status in Turkish patients with narcolepsy and healthy persons. CHKB/CPT1B genes were sequenced in patients with narcolepsy (n=37) and healthy persons (n=100) to detect variations. The HLA-DQB1*06:02 allele status was determined by sequence specific polymerase chain reaction. The HLA-DQB1*06:02 allele was significantly more frequent in narcoleptic patients than in healthy persons (p=2×10(-7)) and in patients with narcolepsy and cataplexy than in those without (p=0.018). The mean of the multiple sleep latency test, sleep-onset rapid eye movement periods, and frequency of sleep paralysis significantly differed in the HLA-DQB1*06:02-positive patients. rs5770917, rs5770911, rs2269381, and rs2269382 were detected together as a haplotype in three patients and 11 healthy persons. In addition to this haplotype, the indel variation (rs144647670) was detected in the 5' upstream region of the human CHKB gene in the patients and healthy persons carrying four variants together. This study identified a novel haplotype consisting of the indel variation, which had not been detected in previous studies in Japanese and Korean populations, and observed four single-nucleotide polymorphisms in CHKB/CPT1B. The study confirmed the association of the HLA-DQB1*06:02 allele with narcolepsy and cataplexy susceptibility. The findings suggest that the presence of HLA-DQB1*06:02 may be a predictor of cataplexy in narcoleptic patients and could therefore be used as an additional diagnostic marker alongside hypocretin.

  3. Environmental toxins and risk of narcolepsy among people with HLA DQB1*0602

    PubMed Central

    Ton, Thanh G.N.; Longstreth, WT; Koepsell, Thomas D.

    2010-01-01

    One etiologic model for narcolepsy suggests that some environmental toxin selectively and irreversibly destroys hypocretin-producing cells in individuals with human leukocyte antigen (HLA) DQB1*0602. Between 2001-2005, the authors conducted a population-based case-control study in King County, Washington to examine narcolepsy risk in relation to toxins found in jobs, hobbies and other non-vocational activities. Sixty-seven cases and 95 controls were enrolled; all were between ages 18-50 and positive for HLA DQB1*0602. All were administered in-person interviews about jobs, hobbies or other non-vocational activities before age 21. All analyses were adjusted for African American race and income. Risk increased significantly for jobs involving heavy metals (odds ratio [OR]=4.7; 95% confidence interval [CI]: 1.5, 14.5) and for highest levels of exposure to woodwork (OR: 3.0; 95% CI: 1.0, 8.9), fertilizer (OR=3.1; 95% CI: 1.1, 9.1), and bug or weed killer (OR=4.5; 95% CI: 1.5, 13.4). Associations were of borderline significance for activities involving ceramics, pesticides, and painting projects. Significant dose-response relationships were evident for jobs involving metals (p<0.03), paints (p<0.03), and bug or weed killer (p<0.02). Additional studies are needed to replicate these findings and continue the search for specific toxins that could damage hypocretin neurons in genetically susceptible people. PMID:20519130

  4. Narcolepsy with cataplexy mimicry: the strange case of two sisters.

    PubMed

    Pizza, Fabio; Vandi, Stefano; Poli, Francesca; Moghadam, Keivan Kaveh; Franceschini, Christian; Bellucci, Claudia; Cipolli, Carlo; Ingravallo, Francesca; Natalini, Giuliana; Mignot, Emmanuel; Plazzi, Giuseppe

    2013-06-15

    We report on two sisters, 17 and 12 years of age, with clinical features suggesting narcolepsy with cataplexy (NC): daytime sleepiness, spontaneous and emotionally triggered sudden falls to the ground, and overweight/obesity. MSLT showed borderline sleep latency, with 1 and 0 sleep onset REM periods. HLA typing disclosed the DQB1*0602 allele. Video-polygraphy of the spells ruled out NC diagnosis by demonstrating their easy elicitation by suggestion, with wake EEG, electromyographic persistence of muscle tone, and stable presence of tendon reflexes (i.e., pseudo-cataplexy), together with normal cerebrospinal hypocretin-1 levels. Our cases emphasize the need of a clear depiction of cataplexy pattern at the different ages, the usefulness of examining ictal neurophysiology, and collecting all available disease markers in ambiguous cases.

  5. Psychosocial Profile and Quality of Life in Children With Type 1 Narcolepsy: A Case-Control Study

    PubMed Central

    Rocca, Francesca Letizia; Finotti, Elena; Pizza, Fabio; Ingravallo, Francesca; Gatta, Michela; Bruni, Oliviero; Plazzi, Giuseppe

    2016-01-01

    Study Objectives: To investigate behavioral aspects and quality of life in children and adolescents with type 1 narcolepsy (NT1). Methods: We performed a case-control study comparing 29 patients with NT1 versus sex- and age-matched patients with idiopathic epilepsy (n = 39) and healthy controls (n = 39). Behavior and quality of life were evaluated by self-administered questionnaires (Child Behavior Checklist, Pediatric Quality of Life Inventory). Patient groups were contrasted and scale results were correlated with clinical and polysomnographic parameters, and cerebrospinal fluid hypocretin-1 levels. Results: Young patients with NT1 showed increased internalizing problems associated with aggressive behavior. Emotional profile in patients with NT1 positively correlated with age at onset, diagnostic delay, and subjective sleepiness, whereas treatment and disease duration were associated with fewer behavioral problems (attention problems, aggressive behavior, and attention deficit/hyperactivity disorder). Psychosocial health domains of pediatric NT1 were worse than in healthy controls, whereas the physical health domains were comparable. Conclusions: Young NT1 patients show a discrete pattern of altered behavioral, thought, and mood profile in comparison with healthy controls and with idiopathic epilepsy patients thus suggesting a direct link with sleepiness. Further studies investigating behavior in patients with idiopathic hypersomnia or type 2 narcolepsy are needed to disentangle the role of REM sleep dysfunction and hypocretin deficiency in psychiatric disorders. Symptoms of withdrawal, depression, somatic complaints, thought problems, and aggressiveness were common, NT1 children perceived lower school competencies than healthy children, and their parents also reported worse psychosocial health. Our data suggest that early effective treatment and disease self-awareness should be promoted in NT1 children for their positive effect on behavior and psychosocial health

  6. Autonomic Response to Periodic Leg Movements during Sleep in Narcolepsy-Cataplexy

    PubMed Central

    Dauvilliers, Yves; Pennestri, Marie-Hélène; Whittom, Shirley; Lanfranchi, Paola A.; Montplaisir, Jacques Y.

    2011-01-01

    Study Objectives: To test the hypothesis of autonomic nervous system dysfunction in patients with narcolepsy-cataplexy (NC) by assessing the physiologic activations associated with periodic limb movements during sleep (PLMS). Design: Sleep and heart rate (HR) were recorded during 1 night of polysomnography. Setting: Data were collected at the Sleep Disorders Center, Sacre-Coeur Hospital, Montreal, Canada. Participants: Data from 14 patients with NC (6 men, 8 women, mean age: 52.5 ± 11.9 years) were compared with data from 14 healthy control subjects matched for age and sex. Interventions: NA. Measurements and Results: Analyses included sleep stages, PLMS, microarousals, RR intervals converted into beats per minute on segments lasting 25 heartbeats (10 RR intervals before PLMS and 15 after), and cardiac-activation amplitudes. A Group-by-Heartbeat interaction was noted for PLMS without microarousals; the patients had a tachycardia of lower amplitude and a delayed and lower-amplitude bradycardia, compared with normal control subjects. Similar significant HR modifications were observed for PLMS with microarousals between patients with NC and control subjects. Patients with NC had a reduced magnitude of cardiac activation associated with PLMS with and without microarousals, as compared with control subjects. A negative correlation was noted between cardiac-activation amplitude and age in patients with NC, but no correlation with PLMS index was found in either patients with NC or control subjects. Conclusion: A significant reduction in the amplitude of PLMS-related HR responses in both tachycardia and bradycardia was found in patients with NC. These findings favor the physiologic relevance of the action of hypocretin on autonomic function that may be of clinical significance, i.e., increasing the risk of cardiovascular diseases. Citation: Dauvilliers Y; Pennestri MH; Whittom S; Lanfranchi PA; Montplaisir JY. Autonomic response to periodic leg movements during sleep in

  7. Antibody Affinity Against 2009 A/H1N1 Influenza and Pandemrix Vaccine Nucleoproteins Differs Between Childhood Narcolepsy Patients and Controls.

    PubMed

    Lind, Alexander; Freyhult, Eva; Ramelius, Anita; Olsson, Tomas; Arnheim-Dahlström, Lisen; Lamb, Favelle; Khademi, Mohsen; Ambati, Aditya; Maeurer, Markus; Lima Bomfim, Izaura; Fink, Katharina; Fex, Malin; Törn, Carina; Elding Larsson, Helena; Lernmark, Åke

    2017-10-01

    Increased narcolepsy incidence was observed in Sweden following the 2009 influenza vaccination with Pandemrix ® . A substitution of the 2009 nucleoprotein for the 1934 variant has been implicated in narcolepsy development. The aims were to determine (a) antibody levels toward wild-type A/H1N1-2009[A/California/04/2009(H1N1)] (NP-CA2009) and Pandemrix-[A/Puerto Rico/8/1934(H1N1)] (NP-PR1934) nucleoproteins in 43 patients and 64 age-matched controls; (b) antibody affinity in reciprocal competitive assays in 11 childhood narcolepsy patients compared with 21 age-matched controls; and (c) antibody levels toward wild-type A/H1N1-2009[A/California/04/2009(H1N1)] (H1N1 NS1), not a component of the Pandemrix vaccine. In vitro transcribed and translated 35 S-methionine-labeled H1N1 influenza A virus proteins were used in radiobinding reciprocal competition assays to estimate antibody levels and affinity (Kd). Childhood patients had higher NP-CA2009 (p = 0.0339) and NP-PR1934 (p = 0.0246) antibody levels compared with age-matched controls. These childhood controls had lower NP-CA2009 (p = 0.0221) and NP-PR1934 (p = 0.00619) antibodies compared with controls 13 years or older. In contrast, in patients 13 years or older, the levels of NP-PR1934 (p = 0.279) and NP-CA2009 (p = 0.0644) antibodies did not differ from the older controls. Childhood antibody affinity (Kd) against NP-CA2009 was comparable between controls (68 ng/mL) and patients (74 ng/mL; p = 0.21) with NP-CA2009 and NP-PR1934 displacement (controls: 165 ng/mL; patients: 199 ng/mL; p = 0.48). In contrast, antibody affinity against NP-PR1934 was higher in controls with either NP-PR1934 (controls: 9 ng/mL; patients: 20 ng/mL; p = 0.0031) or NP-CA2009 (controls: 14 ng/mL; patients: 23 ng/mL; p = 0.0048). A/H1N1-NS1 antibodies were detected in 0/43 of the narcolepsy patients compared with 3/64 (4.7%) controls (p = 0.272). Similarly, none (0/11) of the childhood

  8. Facing emotions in narcolepsy with cataplexy: haemodynamic and behavioural responses during emotional stimulation.

    PubMed

    de Zambotti, Massimiliano; Pizza, Fabio; Covassin, Naima; Vandi, Stefano; Cellini, Nicola; Stegagno, Luciano; Plazzi, Giuseppe

    2014-08-01

    Narcolepsy with cataplexy is a complex sleep disorder that affects the modulation of emotions: cataplexy, the key symptom of narcolepsy, is indeed strongly linked with emotions that usually trigger the episodes. Our study aimed to investigate haemodynamic and behavioural responses during emotional stimulation in narco-cataplexy. Twelve adult drug-naive narcoleptic patients (five males; age: 33.3 ± 9.4 years) and 12 healthy controls (five males; age: 30.9 ± 9.5 years) were exposed to emotional stimuli (pleasant, unpleasant and neutral pictures). Heart rate, arterial blood pressure and mean cerebral blood flow velocity of the middle cerebral arteries were continuously recorded using photoplethysmography and Doppler ultrasound. Ratings of valence and arousal and coping strategies were scored by the Self-Assessment Manikin and by questionnaires, respectively. Narcoleptic patients' haemodynamic responses to pictures overlapped with the data obtained from controls: decrease of heart rate and increase of mean cerebral blood flow velocity regardless of pictures' content, increase of systolic blood pressure during the pleasant condition, and relative reduction of heart rate during pleasant and unpleasant conditions. However, when compared with controls, narcoleptic patients reported lower arousal scores during the pleasant and neutral stimulation, and lower valence scores during the pleasant condition, respectively, and also a lower score at the 'focus on and venting of emotions' dimensions of coping. Our results suggested that adult narcoleptic patients, compared with healthy controls, inhibited their emotion-expressive behaviour to emotional stimulation, and that may be related to the development of adaptive cognitive strategies to face emotions avoiding cataplexy. © 2014 European Sleep Research Society.

  9. Absence of NMDA receptor antibodies in the rare association between Type 1 Narcolepsy and Psychosis

    PubMed Central

    Dauvilliers, Y.; Gaig, C.; Barateau, L.; Graus, F.; Iranzo, A.; Lopez, R.; Santamaria, J.

    2016-01-01

    Frequency and mechanisms underlying the association between narcolepsy type 1 (NT1) and psychosis remain unclear with potential role for a common immune pathway. We estimated the frequency of psychosis and its characteristics in NT1 at two European sleep centers (France, n = 381; Spain, n = 161) and measured IgG autoantibodies that recognize the GluN1 subunit of the NMDAR in 9 patients with NT1 with psychosis, and 25 NT1 patients without psychosis. Ten NT1 patients (6 in France, 4 in Spain) were diagnosed with comorbid psychosis, a frequency of 1.8%. One patient reported psychotic symptoms few months before narcolepsy onset, two patients few months after onset, and one patient one year after onset but after modafinil introduction. The six remaining patients reported long delays between NT1 and psychosis onset. Half the patients, mostly male adults, reported onset or worsening of psychotic symptoms after medication. We found no IgG antibodies to NR1/NR2B heteromers of the NMDARs in patients with NT1 with or without psychosis. To conclude, psychosis is rare in NT1, with limited evidence for a key impact of stimulants, and no association with anti-NMDAR antibodies. However, dramatic NT1 and schizophrenia exists especially in early onset NT1, which may lead to inappropriate diagnosis and management. PMID:27143278

  10. Selected sleep disorders: restless legs syndrome and periodic limb movement disorder, sleep apnea syndrome, and narcolepsy.

    PubMed

    Erman, Milton K

    2006-12-01

    Sleep disorders, including restless legs syndrome and periodic limb movement disorder, sleep apnea syndrome, and narcolepsy, are prevalent medical conditions, likely to be seen by practicing psychiatrists. Awareness of these conditions and their presentations, pathophysiology, and treatment allows psychiatrists to treat these conditions where appropriate, to minimize complications and health consequences associated with delayed diagnosis, and to reduce the burden of disease that these conditions may place on patients already experiencing primary psychiatric disorders.

  11. Almorexant promotes sleep and exacerbates cataplexy in a murine model of narcolepsy.

    PubMed

    Black, Sarah Wurts; Morairty, Stephen R; Fisher, Simon P; Chen, Tsui-Ming; Warrier, Deepti R; Kilduff, Thomas S

    2013-03-01

    Humans with narcolepsy and orexin/ataxin-3 transgenic (TG) mice exhibit extensive, but incomplete, degeneration of hypo-cretin (Hcrt) neurons. Partial Hcrt cell loss also occurs in Parkinson disease and other neurologic conditions. Whether Hcrt antagonists such as almorexant (ALM) can exert an effect on the Hcrt that remains after Hcrt neurodegeneration has not yet been determined. The current study was designed to evaluate the hypnotic and cataplexy-inducing efficacy of a Hcrt antagonist in an animal model with low Hcrt tone and compare the ALM efficacy profile in the disease model to that produced in wild-type (WT) control animals. Counterbalanced crossover study. Home cage. Nine TG mice and 10 WT mice. ALM (30, 100, 300 mg/kg), vehicle and positive control injections, dark/active phase onset. During the 12-h dark period after dosing, ALM exacerbated cataplexy in TG mice and increased nonrapid eye movement sleep with heightened sleep/wake fragmentation in both genotypes. ALM showed greater hypnotic potency in WT mice than in TG mice. The 100 mg/kg dose conferred maximal promotion of cataplexy in TG mice and maximal promotion of REM sleep in WT mice. In TG mice, ALM (30 mg/ kg) paradoxically induced a transient increase in active wakefulness. Core body temperature (Tb) decreased after acute Hcrt receptor blockade, but the reduction in Tb that normally accompanies the wake-to-sleep transition was blunted in TG mice. These complex dose- and genotype-dependent interactions underscore the importance of effector mechanisms downstream from Hcrt receptors that regulate arousal state. Cataplexy promotion by ALM warrants cautious use of Hcrt antagonists in patient populations with Hcrt neurodegeneration, but may also facilitate the discovery of anticataplectic medications. Black SW; Morairty SR; Fisher SP; Chen TM; Warrier DR; Kilduff TS. Almorexant promotes sleep and exacerbates cataplexy in a murine model of narcolepsy. SLEEP 2013;36(3):325-336.

  12. Pseudotumor cerebri syndrome in a patient with narcolepsy type 1.

    PubMed

    Rossor, Thomas; Lim, Ming; VanDenEshof, Kirandeep; Gringras, Paul

    2018-01-01

    Type 1 narcolepsy (NT1) is a chronic primary disorder of hypersomnolence characterized by excessive daytime sleepiness, cataplexy, sleep paralysis, hypnagogic hallucinations and disrupted nocturnal sleep. NT1 is linked to hypothalamic hypocretin deficiency, strongly associated with Human Leukocyte Antigen (HLA) marker DQB1*06:02 and of probable autoimmune origin. NT1 is usually associated with increased rates of overweight and obesity, and sometimes with increases in overnight blood pressure and increased rates of hypoventilation with raised CO 2 levels overnight. Many of these are predisposing factors for pseudotumor cerebri syndrome (PTCS). We present a case of a young girl with both NT1 and PTCS that responded well to treatment with acetazolamide after early identification, with improvement of headache and resolution of hypoventilation. Copyright © 2017 European Paediatric Neurology Society. Published by Elsevier Ltd. All rights reserved.

  13. Risks of neurological and immune-related diseases, including narcolepsy, after vaccination with Pandemrix: a population- and registry-based cohort study with over 2 years of follow-up.

    PubMed

    Persson, I; Granath, F; Askling, J; Ludvigsson, J F; Olsson, T; Feltelius, N

    2014-02-01

    To investigate the association between vaccination with Pandemrix and risk of selected neurological and immune-related diseases including narcolepsy. Population-based prospective cohort study using data from regional vaccination registries and national health registries. Seven healthcare regions in Sweden comprising 61% of the Swedish population. Study population of 3,347,467 vaccinated and 2,497,572 nonvaccinated individuals (vaccination coverage ≈ 60%) followed between 2009 and 2011 for 6.9 million person-years after exposure and 6.0 million person-years without exposure. First recorded diagnosis of neurological and immune-related diseases. Relative risks [hazard ratios (HRs) with 95% confidence intervals (CIs)] assessed using Cox regression, adjusted for covariates. For all selected neurological and immune-related outcomes under study, other than allergic vaccine reactions (for which we verified an expected increase in risk) and narcolepsy, HRs were close to 1.0 and always below 1.3. We observed a three-fold increased risk of a diagnosis of narcolepsy (HR: 2.92, 95% CI: 1.78-4.79; that is, four additional cases per 100,000 person-years) in individuals ≤ 20 years of age at vaccination and a two-fold increase (HR: 2.18, 95% CI: 1.00-4.75) amongst young adults between 21 and 30 years of age. The excess risk declined successively with increasing age at vaccination; no increase in risk was seen after 40 years of age. For a large number of selected neurological and immune-related diseases, we could neither confirm any causal association with Pandemrix nor refute entirely a small excess risk. We confirmed an increased risk for a diagnosis of narcolepsy in individuals ≤ 20 years of age and observed a trend towards an increased risk also amongst young adults between 21 and 30 years. © 2013 The Association for the Publication of the Journal of Internal Medicine.

  14. Test-retest reliability of the multiple sleep latency test in narcolepsy without cataplexy and idiopathic hypersomnia.

    PubMed

    Trotti, Lynn Marie; Staab, Beth A; Rye, David B

    2013-08-15

    Differentiation of narcolepsy without cataplexy from idiopathic hypersomnia relies entirely upon the multiple sleep latency test (MSLT). However, the test-retest reliability for these central nervous system hypersomnias has never been determined. Patients with narcolepsy without cataplexy, idiopathic hypersomnia, and physiologic hypersomnia who underwent two diagnostic multiple sleep latency tests were identified retrospectively. Correlations between the mean sleep latencies on the two studies were evaluated, and we probed for demographic and clinical features associated with reproducibility versus change in diagnosis. Thirty-six patients (58% women, mean age 34 years) were included. Inter -test interval was 4.2 ± 3.8 years (range 2.5 months to 16.9 years). Mean sleep latencies on the first and second tests were 5.5 (± 3.7 SD) and 7.3 (± 3.9) minutes, respectively, with no significant correlation (r = 0.17, p = 0.31). A change in diagnosis occurred in 53% of patients, and was accounted for by a difference in the mean sleep latency (N = 15, 42%) or the number of sleep onset REM periods (N = 11, 31%). The only feature predictive of a diagnosis change was a history of hypnagogic or hypnopompic hallucinations. The multiple sleep latency test demonstrates poor test-retest reliability in a clinical population of patients with central nervous system hypersomnia evaluated in a tertiary referral center. Alternative diagnostic tools are needed.

  15. Effect of Oral JZP-110 (ADX-N05) on Wakefulness and Sleepiness in Adults with Narcolepsy: A Phase 2b Study

    PubMed Central

    Ruoff, Chad; Swick, Todd J.; Doekel, Robert; Emsellem, Helene A.; Feldman, Neil T.; Rosenberg, Russell; Bream, Gary; Khayrallah, Moise A.; Lu, Yuan; Black, Jed

    2016-01-01

    Study Objectives: To evaluate the efficacy and safety of oral JZP-110, a second-generation wake-promoting agent with dopaminergic and noradrenergic activity, for treatment of impaired wakefulness and excessive sleepiness in adults with narcolepsy. Methods: This was a phase 2b, randomized, double-blind, placebo-controlled, parallel-group trial conducted at 28 centers in the United States. Patients were adults with narcolepsy who had baseline scores ≥ 10 on the Epworth Sleepiness Scale (ESS) and baseline sleep latency ≤ 10 min on the Maintenance of Wakefulness Test (MWT). Patients received a daily placebo (n = 49) or JZP-110 (n = 44) 150 mg/day weeks 1–4 and 300 mg/day weeks 5–12. Primary efficacy endpoints were change from baseline in average MWT sleep latency, and the Clinical Global Impression-Change (CGI-C); secondary endpoints were change from baseline in ESS score and Patient Global Impression-Change. Results: Improvements were significantly greater with JZP-110 versus placebo on mean MWT sleep latency (4 w, 9.5 versus 1.4 min, P < 0.0001; 12 w, 12.8 versus 2.1 min, P < 0.0001), percentage of patients with CGI-C improvement (4 w, 80% versus 51%, P = 0.0066; 12 w, 86% versus 38%, P < 0.0001), and mean change in ESS (4 w, −5.6 versus −2.4, P = 0.0038; 12 w, −8.5 versus −2.5, P < 0.0001). Three JZP-110-treated patients (6.8%) discontinued due to adverse events (AEs). The most common AEs with JZP-110 versus placebo were insomnia (23% versus 8%), headache (16% versus 10%), nausea (14% versus 6%), diarrhea (11% versus 6%), decreased appetite (14% versus 0%), and anxiety (11% versus 0%). Conclusions: At doses of 150–300 mg/day, JZP-110 was well tolerated and significantly improved the ability to stay awake and subjective symptoms of excessive sleepiness in adults with narcolepsy. Clinical Trials Registration: Clinicaltrials.gov identifier NCT01681121. Citation: Ruoff C, Swick TJ, Doekel R, Emsellem HA, Feldman NT, Rosenberg R, Bream G, Khayrallah MA

  16. Healthcare Costs Among Patients with Excessive Sleepiness Associated with Obstructive Sleep Apnea, Shift Work Disorder, or Narcolepsy

    PubMed Central

    Carlton, Rashad; Lunacsek, Orsolya; Regan, Timothy; Carroll, Cathryn A.

    2014-01-01

    Background Excessive daytime sleepiness affects nearly 20% of the general population and is associated with many medical conditions, including shift work disorder (SWD), obstructive sleep apnea (OSA), and narcolepsy. Excessive sleepiness imposes a significant clinical, quality-of-life, safety, and economic burden on society. Objective To compare healthcare costs for patients receiving initial therapy with armodafinil or with modafinil for the treatment of excessive sleepiness associated with OSA, SWD, or narcolepsy. Methods A retrospective cohort analysis of medical and pharmacy claims was conducted using the IMS LifeLink Health Plan Claims Database. Patients aged ≥18 years who had a pharmacy claim for armodafinil or for modafinil between June 1, 2009, and February 28, 2012, and had 6 months of continuous eligibility before the index prescription date, as well as International Classification of Diseases, Ninth Revision diagnosis for either OSA (327.23), SWD (327.36), or narcolepsy (347.0x) were included in the study. Patients were placed into 1 of 2 treatment cohorts based on their index prescription and followed for 1 month minimum and 34 months maximum. The annualized all-cause costs were calculated by multiplying the average per-month medical and pharmacy costs for each patient by 12 months. The daily average consumption (DACON) for armodafinil or for modafinil was calculated by dividing the total units dispensed of either drug by the prescription days supply. Results A total of 5693 patients receiving armodafinil and 9212 patients receiving modafinil were included in this study. A lower DACON was observed for armodafinil (1.04) compared with modafinil (1.47). The postindex mean medical costs were significantly lower for the armodafinil cohort compared with the modafinil cohort after adjusting for baseline differences ($11,363 vs $13,775, respectively; P = .005). The mean monthly drug-specific pharmacy costs were lower for the armodafinil cohort compared with

  17. Absence of autoreactive CD4+ T-cells targeting HLA-DQA1*01:02/DQB1*06:02 restricted hypocretin/orexin epitopes in narcolepsy type 1 when detected by EliSpot.

    PubMed

    Kornum, Birgitte Rahbek; Burgdorf, Kristoffer Sølvsten; Holm, Anja; Ullum, Henrik; Jennum, Poul; Knudsen, Stine

    2017-08-15

    Narcolepsy type 1, a neurological sleep disorder strongly associated with Human Leukocyte Antigen (HLA-)DQB1*06:02, is caused by the loss of hypothalamic neurons producing the wake-promoting neuropeptide hypocretin (hcrt, also known as orexin). This loss is believed to be caused by an autoimmune reaction. To test whether hcrt itself could be a possible target in the autoimmune attack, CD4 + T-cell reactivity towards six different 15-mer peptides from prepro-hypocretin with high predicted affinity to the DQA1*01:02/DQB1*06:02 MHC class II dimer was tested using EliSpot in a cohort of 22 narcolepsy patients with low CSF hcrt levels, and 23 DQB1*06:02 positive healthy controls. Our ELISpot assay had a detection limit of 1:10,000 cells. We present data showing that autoreactive CD4 + T-cells targeting epitopes from the hcrt precursor in the context of MHC-DQA1*01:02/DQB1*06:02 are either not present or present in a frequency is <1:10,000 among peripheral CD4 + T-cells from narcolepsy type 1 patients. Copyright © 2017 Elsevier B.V. All rights reserved.

  18. Smoking, Alcohol, Drug Use, Abuse and Dependence in Narcolepsy and Idiopathic Hypersomnia: A Case-Control Study

    PubMed Central

    Barateau, Lucie; Jaussent, Isabelle; Lopez, Régis; Boutrel, Benjamin; Leu-Semenescu, Smaranda; Arnulf, Isabelle; Dauvilliers, Yves

    2016-01-01

    Study Objectives: Basic experiments support the impact of hypocretin on hyperarousal and motivated state required for increasing drug craving. Our aim was to assess the frequencies of smoking, alcohol and drug use, abuse and dependence in narcolepsy type 1 (NT1, hypocretin-deficient), narcolepsy type 2 (NT2), idiopathic hypersomnia (IH) (non-hypocretin-deficient conditions), in comparison to controls. We hypothesized that NT1 patients would be less vulnerable to drug abuse and addiction compared to other hypersomniac patients and controls from general population. Methods: We performed a cross-sectional study in French reference centres for rare hypersomnia diseases and included 450 adult patients (median age 35 years; 41.3% men) with NT1 (n = 243), NT2 (n = 116), IH (n = 91), and 710 adult controls. All participants were evaluated for alcohol consumption, smoking habits, and substance (alcohol and illicit drug) abuse and dependence diagnosis during the past year using the Mini International Neuropsychiatric Interview. Results: An increased proportion of both tobacco and heavy tobacco smokers was found in NT1 compared to controls and other hypersomniacs, despite adjustments for potential confounders. We reported an increased regular and frequent alcohol drinking habit in NT1 versus controls but not compared to other hypersomniacs in adjusted models. In contrast, heavy drinkers were significantly reduced in NT1 versus controls but not compared to other hypersomniacs. The proportion of patients with excessive drug use (codeine, cocaine, and cannabis), substance dependence, or abuse was low in all subgroups, without significant differences between either hypersomnia disorder categories or compared with controls. Conclusions: We first described a low frequency of illicit drug use, dependence, or abuse in patients with central hypersomnia, whether Hcrt-deficient or not, and whether drug-free or medicated, in the same range as in controls. Conversely, heavy drinkers were

  19. Narcolepsy Treated with Racemic Amphetamine during Pregnancy and Breastfeeding.

    PubMed

    Öhman, Inger; Wikner, Birgitta Norstedt; Beck, Olof; Sarman, Ihsan

    2015-08-01

    This case report describes a woman with narcolepsy treated with racemic amphetamine (rac-amphetamine) during pregnancy and breastfeeding with follow-up on the infant's development up to 10 months of age. The pregnancy outcome and the pharmacokinetics of rac-amphetamine were studied during breastfeeding. The pregnancy and the delivery were uneventful. Concentrations of rac-amphetamine were determined in the plasma of the mother and infant, and in the breast milk with a liquid chromatography-mass spectrometry method. Samples were obtained at 2, 5, and 9 weeks postpartum. The transfer of rac-amphetamine to the breast milk was extensive (mean milk/maternal plasma concentration ratio approximately 3). The breastfed infant had a low plasma concentration of rac-amphetamine (about 9% of the maternal plasma level) and the calculated relative infant dose was low (2%). No adverse effects were observed in the breastfed infant. The infant's somatic and psychomotor development up to 10 months of age was normal. Further studies of amphetamine prescribed for medical reasons during pregnancy and lactation are needed. © The Author(s) 2015.

  20. High Prevalence of Precocious Puberty and Obesity in Childhood Narcolepsy with Cataplexy

    PubMed Central

    Poli, Francesca; Pizza, Fabio; Mignot, Emmanuel; Ferri, Raffaele; Pagotto, Uberto; Taheri, Shahrad; Finotti, Elena; Bernardi, Filippo; Pirazzoli, Piero; Cicognani, Alessandro; Balsamo, Antonio; Nobili, Lino; Bruni, Oliviero; Plazzi, Giuseppe

    2013-01-01

    Study Objectives: We analyzed the potential predictive factors for precocious puberty, observed in some cases of childhood narcolepsy with cataplexy (NC) and for obesity, a much more common feature of NC, through a systematic assessment of pubertal staging, body mass index (BMI), and metabolic/endocrine biochemical analyses. Design: Cross-sectional on consecutive recruitment. Setting: Hospital sleep center and pediatric unit. Patients: Forty-three children and adolescents with NC versus 52 age-matched obese children as controls. Interventions: N/A. Measurements and Results: Patients underwent clinical interview, polysomnographic recordings, cerebrospinal fluid hypocretin-1 measurement, and human leukocyte antigen typing. Height, weight, arterial blood pressure, and Tanner pubertal stage were evaluated. Plasma lipid and glucose profiles were analyzed. When an altered pubertal development was clinically suspected, plasma concentrations of hypothalamic-pituitary-gonadal axis hormones were determined. Children with NC showed a high prevalence of overweight/obesity (74%) and a higher occurrence of precocious puberty (17%) than obese controls (1.9%). Isolated signs of accelerated pubertal development (thelarche, pubic hair, advanced bone age) were also present (41%). Precocious puberty was significantly predicted by a younger age at first NC symptom onset but not by overweight/obesity or other factors. In addition, overweight/obesity was predicted by younger age at diagnosis; additional predictors were found for overweight/obesity (short disease duration, younger age at weight gain and lower high-density lipoprotein cholesterol), which did not include precocious puberty. NC symptoms, pubertal signs appearance, and body weight gain developed in close temporal sequence. Conclusions: NC occurring during prepubertal age is frequently accompanied by precocious puberty and overweight/obesity, suggesting an extended hypothalamic dysfunction. The severity of these comorbidities

  1. The inappropriate occurrence of rapid eye movement sleep in narcolepsy is not due to a defect in homeostatic regulation of rapid eye movement sleep.

    PubMed

    Roman, Alexis; Meftah, Soraya; Arthaud, Sébastien; Luppi, Pierre-Hervé; Peyron, Christelle

    2018-06-01

    Narcolepsy type 1 is a disabling disorder with four primary symptoms: excessive-daytime-sleepiness, cataplexy, hypnagogic hallucinations, and sleep paralysis. The later three symptoms together with a short rapid eye movement (REM) sleep latency have suggested impairment in REM sleep homeostatic regulation with an enhanced propensity for (i.e. tendency to enter) REM sleep. To test this hypothesis, we challenged REM sleep homeostatic regulation in a recognized model of narcolepsy, the orexin knock-out (Orex-KO) mice and their wild-type (WT) littermates. We first performed 48 hr of REM sleep deprivation using the classic small-platforms-over-water method. We found that narcoleptic mice are similarly REM sleep deprived to WT mice. Although they had shorter sleep latency, Orex-KO mice recovered similarly to WT during the following 10 hr of recovery. Interestingly, Orex-KO mice also had cataplexy episodes immediately after REM sleep deprivation, anticipating REM sleep rebound, at a time of day when cataplexy does not occur in baseline condition. We then evaluated REM sleep propensity using our new automated method of deprivation that performs a specific and efficient REM sleep deprivation. We showed that REM sleep propensity is similar during light phase in Orex-KO and WT mice. However, during the dark phase, REM sleep propensity was not suppressed in Orex-KO mice when hypocretin/orexin neuropeptides are normally released. Altogether our data suggest that in addition to the well-known wake-promoting role of hypocretin/orexin, these neuropeptides would also suppress REM sleep. Therefore, hypocretin/orexin deficiency would facilitate the occurrence of REM sleep at any time of day in an opportunistic manner as seen in human narcolepsy.

  2. Correlation of Changes in Patient-Reported Quality of Life With Physician-Rated Global Impression of Change in Patients With Narcolepsy Participating in a Clinical Trial of Sodium Oxybate: A Post Hoc Analysis.

    PubMed

    Bogan, Richard K; Black, Jed; Swick, Todd; Mamelak, Mortimer; Kovacevic-Ristanovic, Ruzica; Villa, Kathleen F; Mori, Fannie; Montplaisir, Jacques

    2017-12-01

    Narcolepsy patients report lower health-related quality of life (HRQoL) than the general population, as measured by the Short Form-36 Health Survey (SF-36). This analysis evaluated whether changes in SF-36 correlated with physician-rated Clinical Global Impression of Change (CGI-C). Data were from 209 of 228 narcolepsy patients participating in an 8-week clinical trial of sodium oxybate. Changes from baseline for SF-36 subscales (Physical Functioning, Role Physical, Bodily Pain, General Health, Vitality, Social Functioning, Role Emotional, and Mental Health) and the summary scores were evaluated for correlation with CGI-C overall and by treatment group. Correlations were calculated using the Pearson product-moment correlation coefficient (r). Correlations described an inverse relationship in scores, but a direct relationship in improvement; lower CGI-C scores (i.e., better) were associated with higher SF-36 subscale scores (i.e., improved HRQoL). Moderate and significant correlations were observed for Vitality (r = -0.464; P < 0.0001) and Role Physical (r = -0.310; P < 0.0001) subscales, but weak correlations were observed with other subscales including summary scores. Correlations were stronger at higher sodium oxybate doses for most SF-36 subscales. Some aspects of HRQoL, measured by the SF-36, may be associated with narcolepsy. In particular, Vitality (indicative of energy and tiredness) and Role Physical (impact of physical function on daily roles) moderately correlated with overall change in status observed by clinicians. However, lack of strong correlations between SF-36 and CGI-C indicates differences in patient and clinician perspectives of disease, and suggest a need for broader assessment of the impact of narcolepsy and its treatment on patients. Jazz Pharmaceuticals.

  3. Narcolepsy in pediatric age – Experience of a tertiary pediatric hospital

    PubMed Central

    Dias Costa, Filipa; Barreto, Maria Inês; Clemente, Vanda; Vasconcelos, Mónica; Estêvão, Maria Helena; Madureira, Núria

    2014-01-01

    Narcolepsy, a chronic disorder of the sleep–wake cycle of multifactorial etiology, is characterized by excessive daytime sleepiness, often associated with cataplexy, hypnagogic/hypnopompic hallucinations and sleep paralysis. Both early clinical suspicion and therapeutic approach are essential for promotion of cognitive development and social integration of these children. The authors present a descriptive retrospective study of a series of eight children in whom symptoms first started between 6.8 and 10.5 years of age. Diagnostic delay ranged from 4 months to 2 years. One child had H1N1 flu vaccination eight months before the clinical onset. The first multiple sleep latency test was positive in 6 of 8 cases. All cases were treated with methylphenidate, and venlafaxine was associated in 4 of them. In one case the initial therapy was exclusively behavioral. In all cases, symptomatic improvement, better school performance and social integration were achieved after therapeutic adjustment. PMID:26483902

  4. Event-related-potential low-resolution brain electromagnetic tomography (ERP-LORETA) suggests decreased energetic resources for cognitive processing in narcolepsy.

    PubMed

    Saletu, Michael; Anderer, Peter; Saletu-Zyhlarz, Gerda Maria; Mandl, Magdalena; Zeitlhofer, Josef; Saletu, Bernd

    2008-08-01

    Event-related potentials (ERPs) are sensitive measures of both perceptual and cognitive processes. The aim of the present study was to identify brain regions involved in the processes of cognitive dysfunction in narcolepsy by means of ERP tomography. In 17 drug-free patients with narcolepsy and 17 controls, ERPs were recorded (auditory odd-ball paradigm). Latencies, amplitudes and LORETA sources were determined for standard (N1 and P2) and target (N2 and P300) ERP components. Psychometry included measures of mental performance, affect and critical flicker fusion frequency (CFF). In the ERPs patients demonstrated delayed cognitive N2 and P300 components and reduced amplitudes in midline regions, while N1 and P2 components did not differ from controls. LORETA suggested reduced P300 sources bilaterally in the precuneus, the anterior and posterior cingulate gyri, the ventrolateral prefrontal cortex and the parahippocampal gyrus. In psychometry, patients demonstrated deteriorated mood, increased trait anxiety, decreased CFF and a trend toward reduced general verbal memory and psychomotor activity. Narcoleptic patients showed prolonged information processing, as indexed by N2 and P300 latencies and decreased energetic resources for cognitive processing. Electrophysiological aberrations in brain areas related to the 'executive attention network' and the 'limbic system' may contribute to a deterioration in mental performance and mood at the behavioral level.

  5. Narcolepsy with Cataplexy Associated with Nocturnal Compulsive Behaviors: A Case-Control Study

    PubMed Central

    Palaia, Vincenzo; Poli, Francesca; Pizza, Fabio; Antelmi, Elena; Franceschini, Christian; Moghadam, Keivan Kaveh; Provini, Frederica; Pagotto, Uberto; Montagna, Pasquale; Schenck, Carlos H.; Mignot, Emmanuel; Plazzi, Giuseppe

    2011-01-01

    Study Objectives: To assess the prevalence of sleep related-eating disorder (SRED) and nocturnal smoking (NS) in patients with narcolepsy with cataplexy (NC). Design: Case-control study. Setting: University hospital. Patients or Participants: 65 consecutive adult NC patients (33 men; mean age 43.9 ± 19.2 years) and 65 age-, sex-, and geographical origin-matched controls. Interventions: Validated questionnaires were used to investigate SRED, NS, restless legs syndrome (RLS), and psychopathological traits (using Eating Disorder Inventory-2 [EDI-2]; Maudsley Obsessive-Compulsive Inventory [MOCI]; and Beck Depression Inventory [BDI]). Measurements and Results: NC patients showed a higher prevalence of SRED (32% vs 3%, P = 0.00001), NS (21% vs 0%, P = 0.00006), and RLS (18% vs 5%, P = 0.013) than controls. Moreover, NC patients presented more frequently with an eating-related pathological profile on the EDI-2 (80% vs 46%, P = 0.00006) and had a higher prevalence of depressed mood on the BDI (41% vs 18%, P = 0.004). In comparison to patients without SRED, NC patients with SRED were more frequently women (71% vs 39%, P = 0.013), had higher “bulimic” (29% vs 2%, P = 0.004) and “social insecurity” (48% vs 18%, P = 0.013) traits on the EDI-2, had higher obsessive-compulsiveness on the MOCI (29% vs 4%, P = 0.009), and were more depressed on the BDI (67% vs 29%, P = 0.005). NC patients with NS showed more frequent pathological profiles on the EDI-2 (100% vs 75%, P = 0.035), including the “bulimic” (29% vs 6%, P = 0.015), “perfectionism” (43% vs 14%, P = 0.016), and “social insecurity” (50% vs 22, P = 0.035) profiles. Conclusion: Our study shows a strong association of the compulsive nocturnal behaviors SRED and NS with adult NC. Citation: Palaia V; Poli F; Pizza F; Antelmi E; Franceschini C; Moghadam KK; Provini F; Pagotto U; Montagna P; Schenck CH; Mignot E; Plazzi G. Narcolepsy with cataplexy associated with nocturnal compulsive behaviors: a case

  6. Time-on-task decrements in "steer clear" performance of patients with sleep apnea and narcolepsy

    NASA Technical Reports Server (NTRS)

    Findley, L. J.; Suratt, P. M.; Dinges, D. F.

    1999-01-01

    Loss of attention with time-on-task reflects the increasing instability of the waking state during performance in experimentally induced sleepiness. To determine whether patients with disorders of excessive sleepiness also displayed time-on-task decrements indicative of wake state instability, visual sustained attention performance on "Steer Clear," a computerized simple RT driving simulation task, was compared among 31 patients with untreated sleep apnea, 16 patients with narcolepsy, and 14 healthy control subjects. Vigilance decrement functions were generated by analyzing the number of collisions in each of six four-minute periods of Steer Clear task performance in a mixed-model analysis of variance and linear regression equations. As expected, patients had more Steer Clear collisions than control subjects (p=0.006). However, the inter-subject variability in errors among the narcoleptic patients was four-fold that of the apnea patients, and 100-fold that of the controls volunteers; the variance in errors among untreated apnea patients was 27-times that of controls. The results of transformed collision data revealed main effects for group (p=0.006), time-on-task (p=0.001), and a significant interaction (p=0.022). Control subjects showed no clear evidence of increasing collision errors with time-on-task (adjusted R2=0.22), while apnea patients showed a trend toward vigilance decrement (adjusted R2=0.42, p=0.097), and narcolepsy patients evidenced a robust linear vigilance decrement (adjusted R2=0.87, p=0.004). The association of disorders of excessive somnolence with escalating time-on-task decrements makes it imperative that when assessment of neurobehavioral performance is conducted in patients, it involves task durations and analyses that will evaluate the underlying vulnerability of potentially sleepy patients to decrements over time in tasks that require sustained attention and timely responses, both of which are key components in safe driving performance.

  7. Executive functions, information sampling, and decision making in narcolepsy with cataplexy.

    PubMed

    Delazer, Margarete; Högl, Birgit; Zamarian, Laura; Wenter, Johanna; Gschliesser, Viola; Ehrmann, Laura; Brandauer, Elisabeth; Cevikkol, Zehra; Frauscher, Birgit

    2011-07-01

    Narcolepsy with cataplexy (NC) affects neurotransmitter systems regulating emotions and cognitive functions. This study aimed to assess executive functions, information sampling, reward processing, and decision making in NC. Twenty-one NC patients and 58 healthy participants performed an extensive neuropsychological test battery. NC patients scored as controls in executive function tasks assessing set shifting, reversal learning, working memory, and planning. Group differences appeared in a task measuring information sampling and reward sensitivity. NC patients gathered less information, tolerated a higher level of uncertainty, and were less influenced by reward contingencies than controls. NC patients also showed reduced learning in decision making and had significantly lower scores than controls in the fifth block of the IOWA gambling task. No correlations were found with measures of sleepiness. NC patients may achieve high performance in several neuropsychological domains, including executive functions. Specific differences between NC patients and controls highlight the importance of the hypocretin system in reward processing and decision making and are in line with previous neuroimaging and neurophysiological studies. PsycINFO Database Record (c) 2011 APA, all rights reserved.

  8. Mazindol in narcolepsy and idiopathic and symptomatic hypersomnia refractory to stimulants: a long-term chart review.

    PubMed

    Nittur, Nandini; Konofal, Eric; Dauvilliers, Yves; Franco, Patricia; Leu-Semenescu, Smaranda; Cock, Valérie Cochen De; Inocente, Clara O; Bayard, Sophie; Scholtz, Sabine; Lecendreux, Michel; Arnulf, Isabelle

    2013-01-01

    Mazindol is a tricyclic, anorectic, non-amphetamine stimulant used in narcolepsy and obesity since 1970. This study aimed to evaluate the long-term benefit/risk ratio in drug-resistant hypersomniacs and cataplexy sufferers. By retrospective analysis of the patients' files in the hospitals of Paris-Salpêtrière (n=91), Montpellier (n=40) and Lyon (n=8), the benefit (Epworth Sleepiness Score (ESS), cataplexy frequency, authorization renewal) and tolerance (side-effects, vital signs, electrocardiogram and cardiac echography) of mazindol were assessed. The 139 patients (45% men) aged 36±15years (range: 9-74) suffered narcolepsy (n=94, 66% with cataplexy), idiopathic (n=37) and symptomatic hypersomnia (n=8) refractory to modafinil, methylphenidate and sodium oxybate. Under mazindol (3.4±1.3mg/day, 1-6mg) for an average of 30months, the ESS decreased from 17.7±3.5 to 12.8±5.1, with an average fall of -4.6±4.7 (p<0.0001) and the frequency of cataplexy fell from 4.6±3.1 to 2±2.8 episodes per week. The cataplexy was eliminated in 14.5% of patients, improved in 27.5%, and unchanged in 29% (missing data in 29%). The treatment was maintained long term in 83 (60%) patients, and stopped because of a lack of efficacy (22%) and/or secondary effects (9%). There was no pulmonary hypertension in the 45 patients who underwent a cardiac echography. The most common adverse effects were dry mouth (13%), palpitations (10%, including one with ventricular hyperexcitability), anorexia (6%), nervousness (6%) and headaches (6%). Mazindol has a long-term, favorable benefit/risk ratio in 60% of drug-resistant hypersomniacs, including a clear benefit on cataplexy. Copyright © 2012 Elsevier B.V. All rights reserved.

  9. Psychosocial intervention for children with narcolepsy: Parents' expectations and perceived support.

    PubMed

    Kippola-Pääkkönen, Anu; Härkäpää, Kristiina; Valkonen, Jukka; Tuulio-Henriksson, Annamari; Autti-Rämö, Ilona

    2016-04-18

    The study focuses on the parents of children who were affected by narcolepsy after a pandemic influenza and vaccination campaign in Finland. The main aim of the study was to clarify parents' expectations and perceived support from the intervention and to assess their need for additional support. The data were gathered using questionnaires. Fifty-eight parents answered the baseline questionnaire and 40 parents the final questionnaire. Parents' expectations of and perceived support from the intervention mainly related to peer support. The intervention offered an arena for sharing information and experiences and provided encouragement for coping in everyday life. Many expectations were not met, especially those concerning information about needed services, financial benefits and availability of local support. The results highlight that for persons with rare disorders and their families, an inpatient psychosocial intervention can offer an important arena to receive both informal and professionally led peer support. Comprehensive psychosocial and other support services are also needed in the community. Listening to parents' perspectives on the intervention and perceived support can help to establish multiform family-centred support for families with children affected by a rare chronic disabling condition. © The Author(s) 2016.

  10. HLA-DPB1 and HLA Class I Confer Risk of and Protection from Narcolepsy

    PubMed Central

    Ollila, Hanna M.; Ravel, Jean-Marie; Han, Fang; Faraco, Juliette; Lin, Ling; Zheng, Xiuwen; Plazzi, Giuseppe; Dauvilliers, Yves; Pizza, Fabio; Hong, Seung-Chul; Jennum, Poul; Knudsen, Stine; Kornum, Birgitte R.; Dong, Xiao Song; Yan, Han; Hong, Heeseung; Coquillard, Cristin; Mahlios, Joshua; Jolanki, Otto; Einen, Mali; Lavault, Sophie; Högl, Birgit; Frauscher, Birgit; Crowe, Catherine; Partinen, Markku; Huang, Yu Shu; Bourgin, Patrice; Vaarala, Outi; Désautels, Alex; Montplaisir, Jacques; Mack, Steven J.; Mindrinos, Michael; Fernandez-Vina, Marcelo; Mignot, Emmanuel

    2015-01-01

    Type 1 narcolepsy, a disorder caused by a lack of hypocretin (orexin), is so strongly associated with human leukocyte antigen (HLA) class II HLA-DQA1∗01:02-DQB1∗06:02 (DQ0602) that very few non-DQ0602 cases have been reported. A known triggering factor for narcolepsy is pandemic 2009 influenza H1N1, suggesting autoimmunity triggered by upper-airway infections. Additional effects of other HLA-DQ alleles have been reported consistently across multiple ethnic groups. Using over 3,000 case and 10,000 control individuals of European and Chinese background, we examined the effects of other HLA loci. After careful matching of HLA-DR and HLA-DQ in case and control individuals, we found strong protective effects of HLA-DPA1∗01:03-DPB1∗04:02 (DP0402; odds ratio [OR] = 0.51 [0.38–0.67], p = 1.01 × 10−6) and HLA-DPA1∗01:03-DPB1∗04:01 (DP0401; OR = 0.61 [0.47–0.80], p = 2.07 × 10−4) and predisposing effects of HLA-DPB1∗05:01 in Asians (OR = 1.76 [1.34–2.31], p = 4.71 × 10−05). Similar effects were found by conditional analysis controlling for HLA-DR and HLA-DQ with DP0402 (OR = 0.45 [0.38–0.55] p = 8.99 × 10−17) and DP0501 (OR = 1.38 [1.18–1.61], p = 7.11 × 10−5). HLA-class-II-independent associations with HLA-A∗11:01 (OR = 1.32 [1.13–1.54], p = 4.92 × 10−4), HLA-B∗35:03 (OR = 1.96 [1.41–2.70], p = 5.14 × 10−5), and HLA-B∗51:01 (OR = 1.49 [1.25–1.78], p = 1.09 × 10−5) were also seen across ethnic groups in the HLA class I region. These effects might reflect modulation of autoimmunity or indirect effects of HLA class I and HLA-DP alleles on response to viral infections such as that of influenza. PMID:25574827

  11. Smoking, Alcohol, Drug Use, Abuse and Dependence in Narcolepsy and Idiopathic Hypersomnia: A Case-Control Study.

    PubMed

    Barateau, Lucie; Jaussent, Isabelle; Lopez, Régis; Boutrel, Benjamin; Leu-Semenescu, Smaranda; Arnulf, Isabelle; Dauvilliers, Yves

    2016-03-01

    Basic experiments support the impact of hypocretin on hyperarousal and motivated state required for increasing drug craving. Our aim was to assess the frequencies of smoking, alcohol and drug use, abuse and dependence in narcolepsy type 1 (NT1, hypocretin-deficient), narcolepsy type 2 (NT2), idiopathic hypersomnia (IH) (non-hypocretin-deficient conditions), in comparison to controls. We hypothesized that NT1 patients would be less vulnerable to drug abuse and addiction compared to other hypersomniac patients and controls from general population. We performed a cross-sectional study in French reference centres for rare hypersomnia diseases and included 450 adult patients (median age 35 years; 41.3% men) with NT1 (n = 243), NT2 (n = 116), IH (n = 91), and 710 adult controls. All participants were evaluated for alcohol consumption, smoking habits, and substance (alcohol and illicit drug) abuse and dependence diagnosis during the past year using the Mini International Neuropsychiatric Interview. An increased proportion of both tobacco and heavy tobacco smokers was found in NT1 compared to controls and other hypersomniacs, despite adjustments for potential confounders. We reported an increased regular and frequent alcohol drinking habit in NT1 versus controls but not compared to other hypersomniacs in adjusted models. In contrast, heavy drinkers were significantly reduced in NT1 versus controls but not compared to other hypersomniacs. The proportion of patients with excessive drug use (codeine, cocaine, and cannabis), substance dependence, or abuse was low in all subgroups, without significant differences between either hypersomnia disorder categories or compared with controls. We first described a low frequency of illicit drug use, dependence, or abuse in patients with central hypersomnia, whether Hcrt-deficient or not, and whether drug-free or medicated, in the same range as in controls. Conversely, heavy drinkers were rare in NT1 compared to controls but not to other

  12. Comparison of clinical characteristics among narcolepsy with and without cataplexy and idiopathic hypersomnia without long sleep time, focusing on HLA-DRB1( *)1501/DQB1( *)0602 finding.

    PubMed

    Sasai, Taeko; Inoue, Yuichi; Komada, Yoko; Sugiura, Tatsuki; Matsushima, Eisuke

    2009-10-01

    Clinical characteristics of narcolepsy without cataplexy (NA w/o CA) and its relation to positivity of HLA-DRB1( *)1501/DQB1( *)0602 remain unclarified. We investigated clinical features of NA w/o CA, particularly addressing HLA-DRB1( *)1501/DQB1( *)0602. Comparisons of the Epworth Sleepiness Scale (ESS), multiple sleep latency test (MSLT) variables, rapid eye movement (REM)-related symptoms, and treatment response to psychostimulant medication were made for four patient groups (narcolepsy with cataplexy; NA-CA, NA w/o CA HLA-positive, NA w/o CA HLA-negative, and idiopathic hypersomnia without long sleep time; IHS w/o LST). Mean sleep latency was significantly shorter and the rate of reduction of ESS after medication was lower in both NA-CA and NA w/o CA HLA-positive groups than those in the IHS w/o LST group. Among the three narcoleptic groups, the NA w/o CA HLA-negative group showed the lowest REM latency and the highest reduction rate of ESS after treatment. Neither these subjective and objective sleepiness measures nor the treatment response measure was significantly different between this group and the IHS w/o LST group. In NA w/o CA, HLA-positivity might affect hypersomnia severity and REM propensity. The NA w/o CA HLA-negative group and the IHS w/o LST group exhibit equivalent hypersomnia severity.

  13. Time to Response with Sodium Oxybate for the Treatment of Excessive Daytime Sleepiness and Cataplexy in Patients with Narcolepsy

    PubMed Central

    Bogan, Richard K.; Roth, Thomas; Schwartz, Jonathan; Miloslavsky, Maja

    2015-01-01

    Study Objectives: This post hoc analysis evaluated the time to response that can be expected with sodium oxybate (SXB) for treatment of excessive daytime sleepiness (EDS) and cataplexy in patients with narcolepsy. Methods: Data were from a 4-week, double-blind, randomized, placebo-controlled trial (GHB-2; N = 136) of oral SXB 3 g, 6 g, and 9 g nightly, and its 12-month open-label extension (GHB-3). Two response definitions were utilized: ≥ 20% improvement in Epworth Sleepiness Scale (ESS) score (EDS responders), and ≥ 50% reduction in weekly cataplexy attacks (cataplexy responders). These thresholds were previously determined to be clinically relevant based on analysis of the relationship of Clinical Global Impression of Change with ESS and number of cataplexy attacks. Kaplan-Meier curves and median times to first response, based on above criteria, and to maximum response were estimated. Results: Among 86 patients randomized to SXB in GHB-2 and continued into GHB-3, 77.6% and 90.7% were EDS and cataplexy responders, respectively. The median (95% CI) times to first response were 37 (31–50) days for EDS and 25 (17–29) days for cataplexy, and median times to maximum response were 106 (85–164) days for EDS and 213 (94–279) days for cataplexy. GHB-3 results among 31 patients initially randomized to placebo were consistent with those treated with SXB throughout, but with longer times to maximum response. Conclusions: Response onset, assessed as clinically meaningful improvements in EDS and cataplexy, was observed in most patients within 2 months; a longer period is needed to achieve maximum response. Clinicians should recognize that time to initial and maximum response may take weeks to months. Citation: Bogan RK, Roth T, Schwartz J, Miloslavsky M. Time to response with sodium oxybate for the treatment of excessive daytime sleepiness and cataplexy in patients with narcolepsy. J Clin Sleep Med 2015;11(4):427–432. PMID:25580605

  14. Schizophrenia-like symptoms in narcolepsy type 1: shared and distinctive clinical characteristics.

    PubMed

    Plazzi, Giuseppe; Fabbri, Chiara; Pizza, Fabio; Serretti, Alessandro

    2015-01-01

    The occurrence of psychotic symptoms in narcolepsy type 1 (NT1) has been reported as responsible for delayed diagnosis due to the misdiagnosis of schizophrenia. This study aimed to identify shared and distinctive clinical characteristics between NT1 and schizophrenia, with the focus on psychotic symptoms. A total of 28 NT1 and 21 schizophrenia patients were included. Hallucination characteristics and PANSS (Positive and Negative Syndrome Scale), HRSD (Hamilton Rating Scale for Depression), DES (Dissociative Experiences Scale), and STAI (State-Trait Anxiety Inventory) scores were collected. Symptom overlap was investigated by χ(2), Fisher's or t tests and multiple logistic regression models. Hallucinations and illusions frequently occurred in both diseases. Unimodal hallucinations were more common in schizophrenia (p = 6.30e-07) and multimodal hallucinations in NT1, but no clear difference was identified in their sensory modality. Hypnagogic/hypnopompic hallucinations were typical of NT1 (p = 5.22e-07), and 25% of NT1 patients exhibited some degree of insight deficit. Hypnagogic/hypnopompic hallucinations, unimodal hallucinations and PANSS score were the most distinctive clinical characteristics. Clinical overlap was found in the dissociative and anxiety domains, while higher depressive scores were observed in schizophrenia. The overlap between NT1 and schizophrenia should be further investigated under a clinical and pathogenetic point of view to improve diagnostic and therapeutic approaches. © 2015 S. Karger AG, Basel.

  15. Narcolepsy, REM sleep behavior disorder, and supranuclear gaze palsy associated with Ma1 and Ma2 antibodies and tonsillar carcinoma.

    PubMed

    Adams, Chris; McKeon, Andrew; Silber, Michael H; Kumar, Rajeev

    2011-04-01

    To describe a patient with diencephalic and mesencephalic presentation of a Ma1 and Ma2 antibody-associated paraneoplastic neurological disorder. Case report. The Colorado Neurological Institute Movement Disorders Center in Englewood, Colorado, and the Mayo Clinic in Rochester, Minnesota. A 55-year-old man with a paraneoplastic neurological disorder characterized by rapid eye movement sleep behavior disorder, narcolepsy, and a progressive supranuclear palsy-like syndrome in the setting of tonsillar carcinoma. Immunotherapy for paraneoplastic neurological disorder, surgery and radiotherapy for cancer, and symptomatic treatment for parkinsonism and sleep disorders. Polysomnography, multiple sleep latency test, and neurological examination. The cancer was detected at a limited stage and treatable. After oncological therapy and immunotherapy, symptoms stabilized. Treatment with modafinil improved daytime somnolence. Rapid onset and progression of multifocal deficits may be a clue to paraneoplastic etiology. Early treatment of a limited stage cancer (with or without immunotherapy) may possibly slow progression of neurological symptoms. Symptomatic treatment may be beneficial.

  16. Dopaminergic regulation of sleep and cataplexy in a murine model of narcolepsy.

    PubMed

    Burgess, Christian R; Tse, Gavin; Gillis, Lauren; Peever, John H

    2010-10-01

    To determine if the dopaminergic system modulates cataplexy, sleep attacks and sleep-wake behavior in narcoleptic mice. Hypocretin/orexin knockout (i.e., narcoleptic) and wild-type mice were administered amphetamine and specific dopamine receptor modulators to determine their effects on sleep, cataplexy and sleep attacks. Hypocretin knockout (n = 17) and wild-type mice (n = 21). Cataplexy, sleep attacks and sleep-wake behavior were identified using electroencephalogram, electromyogram and videography. These behaviors were monitored for 4 hours after an i.p. injection of saline, amphetamine and specific dopamine receptor modulators (D1- and D2-like receptor modulators). Amphetamine (2 mg/kg), which increases brain dopamine levels, decreased sleep attacks and cataplexy by 61% and 67%, suggesting that dopamine transmission modulates such behaviors. Dopamine receptor modulation also had powerful effects on sleep attacks and cataplexy. Activation (SKF 38393; 20 mg/kg) and blockade (SCH 23390; 1 mg/kg) of D1-like receptors decreased and increased sleep attacks by 77% and 88%, without affecting cataplexy. Pharmacological activation of D2-like receptors (quinpirole; 0.5 mg/kg) increased cataplectic attacks by 172% and blockade of these receptors (eticlopride; 1 mg/kg) potently suppressed them by 97%. Manipulation of D2-like receptors did not affect sleep attacks. We show that the dopaminergic system plays a role in regulating both cataplexy and sleep attacks in narcoleptic mice. We found that cataplexy is modulated by a D2-like receptor mechanism, whereas dopamine modulates sleep attacks by a D1-like receptor mechanism. These results support a role for the dopamine system in regulating sleep attacks and cataplexy in a murine model of narcolepsy.

  17. Narcolepsy Type 1 Is Associated with a Systemic Increase and Activation of Regulatory T Cells and with a Systemic Activation of Global T Cells

    PubMed Central

    Pitoiset, Fabien; Regnault, Armelle; Tran, Tu Anh; Liblau, Roland; Klatzmann, David; Rosenzwajg, Michelle

    2017-01-01

    Narcolepsy is a rare neurologic disorder characterized by excessive daytime sleepiness, cataplexy and disturbed nocturnal sleep patterns. Narcolepsy type 1 (NT1) has been shown to result from a selective loss of hypothalamic hypocretin-secreting neurons with patients typically showing low CSF-hypocretin levels (<110 pg/ml). This specific loss of hypocretin and the strong association with the HLA-DQB1*06:02 allele led to the hypothesis that NT1 could be an immune-mediated pathology. Moreover, susceptibility to NT1 has recently been associated with several pathogens, particularly with influenza A H1N1 virus either through infection or vaccination. The goal of this study was to compare peripheral blood immune cell populations in recent onset pediatric NT1 subjects (post or non-post 2009-influenza A H1N1 vaccination) to healthy donors. We demonstrated an increased number of central memory CD4+ T cells (CD62L+ CD45RA-) associated to an activated phenotype (increase in CD69 and CD25 expression) in NT1 patients. Percentage and absolute count of regulatory T cells (Tregs) in NT1 patients were increased associated with an activated phenotype (increase in GITR and LAP expression), and of activated memory phenotype. Cytokine production by CD4+ and CD8+ T cells after activation was not modified in NT1 patients. In H1N1 vaccinated NT1 patients, absolute counts of CD3+, CD8+ T cells, and B cells were increased compared to non-vaccinated NT1 patients. These results support a global T cell activation in NT1 patients and thus support a T cell-mediated autoimmune origin of NT1, but do not demonstrate the pathological role of H1N1 prophylactic vaccination. They should prompt further studies of T cells, particularly of Tregs (such as suppression and proliferation antigen specific assays, and also T-cell receptor sequencing), in NT1. PMID:28107375

  18. Narcolepsy Type 1 Is Associated with a Systemic Increase and Activation of Regulatory T Cells and with a Systemic Activation of Global T Cells.

    PubMed

    Lecendreux, Michel; Churlaud, Guillaume; Pitoiset, Fabien; Regnault, Armelle; Tran, Tu Anh; Liblau, Roland; Klatzmann, David; Rosenzwajg, Michelle

    2017-01-01

    Narcolepsy is a rare neurologic disorder characterized by excessive daytime sleepiness, cataplexy and disturbed nocturnal sleep patterns. Narcolepsy type 1 (NT1) has been shown to result from a selective loss of hypothalamic hypocretin-secreting neurons with patients typically showing low CSF-hypocretin levels (<110 pg/ml). This specific loss of hypocretin and the strong association with the HLA-DQB1*06:02 allele led to the hypothesis that NT1 could be an immune-mediated pathology. Moreover, susceptibility to NT1 has recently been associated with several pathogens, particularly with influenza A H1N1 virus either through infection or vaccination. The goal of this study was to compare peripheral blood immune cell populations in recent onset pediatric NT1 subjects (post or non-post 2009-influenza A H1N1 vaccination) to healthy donors. We demonstrated an increased number of central memory CD4+ T cells (CD62L+ CD45RA-) associated to an activated phenotype (increase in CD69 and CD25 expression) in NT1 patients. Percentage and absolute count of regulatory T cells (Tregs) in NT1 patients were increased associated with an activated phenotype (increase in GITR and LAP expression), and of activated memory phenotype. Cytokine production by CD4+ and CD8+ T cells after activation was not modified in NT1 patients. In H1N1 vaccinated NT1 patients, absolute counts of CD3+, CD8+ T cells, and B cells were increased compared to non-vaccinated NT1 patients. These results support a global T cell activation in NT1 patients and thus support a T cell-mediated autoimmune origin of NT1, but do not demonstrate the pathological role of H1N1 prophylactic vaccination. They should prompt further studies of T cells, particularly of Tregs (such as suppression and proliferation antigen specific assays, and also T-cell receptor sequencing), in NT1.

  19. Childhood narcolepsy with cataplexy: comparison between post-H1N1 vaccination and sporadic cases.

    PubMed

    Pizza, Fabio; Peltola, Hanna; Sarkanen, Tomi; Moghadam, Keivan K; Plazzi, Giuseppe; Partinen, Markku

    2014-02-01

    We aimed to compare post-Pandemrix vaccination (postvaccine) childhood narcolepsy with cataplexy (NC) vs. sporadic pre-H1N1 pandemic (pre-H1N1) cases. Clinical, anthropometric, polysomnographic, and cerebrospinal hypocretin 1 (hcrt-1) measurements were collected together with the video recordings of cataplexy in 27 Finnish patients with NC onset after H1N1 Pandemrix vaccination (mean age, 12±4 years; 52% boys) and 42 Italian NC patients with NC onset before the H1N1 pandemic (mean age, 11±3 years; 48% boys). All subjects carried the HLA-DQB1*0602 allele. Postvaccine subjects were older at NC onset (12±3 vs. 9±3 years; P=.008) and displayed a shorter mean sleep latency in multiple sleep latency tests (MSLT) (2.3±2.2 vs. 3.7±2.9 min; P=.026) compared to pre-H1N1 cases. Anthropometric, clinical (core NC symptoms), hcrt-1 deficiency, and polysomnographic data did not differ among groups, but higher disrupted nocturnal sleep was observed in postvaccine subjects. Comparison of cataplexy features at video assessment showed an overlapping picture with the exception for hyperkinetic movements which appeared to be more evident in pre-H1N1 subjects. The clinical picture of childhood NC was similar in postvaccine and pre-H1N1 children. Copyright © 2013 Elsevier B.V. All rights reserved.

  20. A neural network method for detection of obstructive sleep apnea and narcolepsy based on pupil size and EEG.

    PubMed

    Liu, D; Pang, Z; Lloyd, S R

    2008-02-01

    Electroencephalogram (EEG) is able to indicate states of mental activity ranging from concentrated cognitive efforts to sleepiness. Such mental activity can be reflected by EEG energy. In particular, intrusion of EEG theta wave activity into the beta activity of active wakefulness has been interpreted as ensuing sleepiness. Pupil behavior can also provide information regarding alertness. This paper develops an innovative signal classification method that is capable of differentiating subjects with sleep disorders which cause excessive daytime sleepiness (EDS) from normal control subjects who do not have a sleep disorder based on EEG and pupil size. Subjects with sleep disorders include persons with untreated obstructive sleep apnea (OSA) and narcolepsy. The Yoss pupil staging rule is used to scale levels of wakefulness and at the same time theta energy ratios are calculated from the same 2-s sliding windows by Fourier or wavelet transforms. Then, an artificial neural network (NN) of modified adaptive resonance theory (ART2) is utilized to identify the two groups within a combined group of subjects including those with OSA and healthy controls. This grouping from the NN is then compared with the actual diagnostic classification of subjects as OSA or controls and is found to be 91% accurate in differentiating between the two groups. The same algorithm results in 90% correct differentiation between narcoleptic and control subjects.

  1. Central Disorders of Hypersomnolence

    PubMed Central

    Khan, Zeeshan

    2015-01-01

    The central disorders of hypersomnolence are characterized by severe daytime sleepiness, which is present despite normal quality and timing of nocturnal sleep. Recent reclassification distinguishes three main subtypes: narcolepsy type 1, narcolepsy type 2, and idiopathic hypersomnia (IH), which are the focus of this review. Narcolepsy type 1 results from loss of hypothalamic hypocretin neurons, while the pathophysiology underlying narcolepsy type 2 and IH remains to be fully elucidated. Treatment of all three disorders focuses on the management of sleepiness, with additional treatment of cataplexy in those patients with narcolepsy type 1. Sleepiness can be treated with modafinil/armodafinil or sympathomimetic CNS stimulants, which have been shown to be beneficial in randomized controlled trials of narcolepsy and, quite recently, IH. In those patients with narcolepsy type 1, sodium oxybate is effective for the treatment of both sleepiness and cataplexy. Despite these treatments, there remains a subset of hypersomnolent patients with persistent sleepiness, in whom alternate therapies are needed. Emerging treatments for sleepiness include histamine H3 antagonists (eg, pitolisant) and possibly negative allosteric modulators of the gamma-aminobutyric acid-A receptor (eg, clarithromycin and flumazenil). PMID:26149554

  2. Multiple sleep latency measures in narcolepsy and behaviourally induced insufficient sleep syndrome.

    PubMed

    Marti, Isabelle; Valko, Philipp O; Khatami, Ramin; Bassetti, Claudio L; Baumann, Christian R

    2009-12-01

    Short mean latencies to the first epoch of non-rapid eye movement sleep stage 1 (NREM1) and the presence of >or= 2 sleep onset REM (SOREM) periods on multiple sleep latency test (MSLT) occur in both narcolepsy-cataplexy (NC) and behaviourally induced insufficient sleep syndrome (BIISS). It is not known whether specific MSLT findings help differentiate the two disorders. We analyzed MSLT data including sleep latencies to and between different sleep stages of 60 age-, gender- and body mass index (BMI)-matched subjects (hypocretin-deficient NC, actigraphy-confirmed BIISS, healthy controls: each 20). Mean latency (in minutes) to NREM1 sleep was significantly shorter in NC (1.8+/-1.5) than in BIISS (4.7+/-2.1, p<0.001) and controls (11.4+/-3.3, p<0.001). Mean latency to NREM2 sleep was similar in NC (8.6+/-4.7) and BIISS (8.1+/-2.7, p=0.64); latency to either NREM2 or rapid eye movement (REM) sleep (i.e., the sum of the sleep latency to NREM1 and the duration of the first NREM1 sleep sequence), however, was shorter in NC (4.4+/-2.9) than in BIISS (7.9+/-3.5, p<0.001). Referring to all naps with SOREM periods, the sequence NREM1-REM-NREM2 was more common (71%) in NC than in BIISS (15%, p<0.001), reflecting the shorter latency from NREM1 to NREM2 in BIISS (3.7+/-2.5) than in NC (6.1+/-5.9, p<0.001). Our findings show that both sleepiness (as measured by NREM1 sleep latency) and REM sleep propensity are higher in NC than in BIISS. Furthermore, our finding of frequent REM sleep prior to NREM2 sleep in NC is in line with the recent assumption of an insufficient NREM sleep intensity in NC. Together with detailed clinical interviews, sleep logs, actigraphy, and nocturnal polysomnography, mean sleep latencies to NREM1

  3. IGFBP3 Colocalizes with and Regulates Hypocretin (Orexin)

    PubMed Central

    Honda, Makoto; Eriksson, Krister S.; Zhang, Shengwen; Tanaka, Susumu; Lin, Ling; Salehi, Ahmad; Hesla, Per Egil; Maehlen, Jan; Gaus, Stephanie E.; Yanagisawa, Masashi; Sakurai, Takeshi; Taheri, Shahrad; Tsuchiya, Kuniaki; Honda, Yutaka; Mignot, Emmanuel

    2009-01-01

    Background The sleep disorder narcolepsy is caused by a vast reduction in neurons producing the hypocretin (orexin) neuropeptides. Based on the tight association with HLA, narcolepsy is believed to result from an autoimmune attack, but the cause of hypocretin cell loss is still unknown. We performed gene expression profiling in the hypothalamus to identify novel genes dysregulated in narcolepsy, as these may be the target of autoimmune attack or modulate hypocretin gene expression. Methodology/Principal Findings We used microarrays to compare the transcriptome in the posterior hypothalamus of (1) narcoleptic versus control postmortem human brains and (2) transgenic mice lacking hypocretin neurons versus wild type mice. Hypocretin was the most downregulated gene in human narcolepsy brains. Among many additional candidates, only one, insulin-like growth factor binding protein 3 (IGFBP3), was downregulated in both human and mouse models and co-expressed in hypocretin neurons. Functional analysis indicated decreased hypocretin messenger RNA and peptide content, and increased sleep in transgenic mice overexpressing human IGFBP3, an effect possibly mediated through decreased hypocretin promotor activity in the presence of excessive IGFBP3. Although we found no IGFBP3 autoantibodies nor a genetic association with IGFBP3 polymorphisms in human narcolepsy, we found that an IGFBP3 polymorphism known to increase serum IGFBP3 levels was associated with lower CSF hypocretin-1 in normal individuals. Conclusions/Significance Comparison of the transcriptome in narcolepsy and narcolepsy model mouse brains revealed a novel dysregulated gene which colocalized in hypocretin cells. Functional analysis indicated that the identified IGFBP3 is a new regulator of hypocretin cell physiology that may be involved not only in the pathophysiology of narcolepsy, but also in the regulation of sleep in normal individuals, most notably during adolescence. Further studies are required to address the

  4. Sleep-disordered breathing and periodic limb movements in narcolepsy with cataplexy: a systematic analysis of 35 consecutive patients.

    PubMed

    Pizza, Fabio; Tartarotti, Sonja; Poryazova, Rositsa; Baumann, Christian R; Bassetti, Claudio L

    2013-01-01

    Disturbed sleep is a core feature of narcolepsy with cataplexy (NC). Few studies have independently assessed sleep-disordered breathing (SDB) and periodic limb movements (PLMs) in non-homogeneous series of patients with and without cataplexy. We systematically assessed both SDB and PLMs in well-defined NC patients. We analyzed the clinical and polysomnographic features of 35 consecutive NC patients (mean age 40 ± 16 years, 51% males, 23/23 hypocretin-deficient) to assess the prevalence of SDB (apnea-hypopnea index >5) and PLMs (periodic leg movements in sleep (PLMI) >15) together with their impact on nocturnal sleep and daytime sleepiness using the multiple sleep latency test. 11 (31%) and 14 (40%) patients had SDB and PLMs, respectively. SDB was associated with older age (49 ± 16 vs. 35 ± 13 years, p = 0.02), higher BMI (30 ± 5 vs. 27 ± 6, p = 0.05), and a trend towards higher PLMI (25 ± 20 vs. 12 ± 23, p = 0.052), whereas PLMs with older age (50 ± 16 vs. 33 ± 11 years, p = 0.002) and reduced and fragmented sleep (e.g. sleep efficiency of 82 ± 12% vs. 91 ± 6%, p = 0.015; sleep time of 353 ± 66 vs. 395 ± 28, p = 0.010). SDB and PLMs were also mutually associated (p = 0.007), but not correlated to daytime sleepiness. SDB and PLMs are highly prevalent and associated in NC. Nevertheless, SDB and PLMs are rarely severe, suggesting an overall limited effect on clinical manifestations. Copyright © 2013 S. Karger AG, Basel.

  5. Interactions of the histamine and hypocretin systems in CNS disorders.

    PubMed

    Shan, Ling; Dauvilliers, Yves; Siegel, Jerome M

    2015-07-01

    Histamine and hypocretin neurons are localized to the hypothalamus, a brain area critical to autonomic function and sleep. Narcolepsy type 1, also known as narcolepsy with cataplexy, is a neurological disorder characterized by excessive daytime sleepiness, impaired night-time sleep, cataplexy, sleep paralysis and short latency to rapid eye movement (REM) sleep after sleep onset. In narcolepsy, 90% of hypocretin neurons are lost; in addition, two groups reported in 2014 that the number of histamine neurons is increased by 64% or more in human patients with narcolepsy, suggesting involvement of histamine in the aetiology of this disorder. Here, we review the role of the histamine and hypocretin systems in sleep-wake modulation. Furthermore, we summarize the neuropathological changes to these two systems in narcolepsy and discuss the possibility that narcolepsy-associated histamine abnormalities could mediate or result from the same processes that cause the hypocretin cell loss. We also review the changes in the hypocretin and histamine systems, and the associated sleep disruptions, in Parkinson disease, Alzheimer disease, Huntington disease and Tourette syndrome. Finally, we discuss novel therapeutic approaches for manipulation of the histamine system.

  6. [Hypersomnia: a diagnostic problem].

    PubMed

    Kranenburg, Guido; Teunissen, Laurien L

    2014-01-01

    Hypersomnia is a frequently occurring problem. When taking a medical history it is important to distinguish between fatigue and sleepiness. We present a 14-year-old girl with narcolepsy and a 59-year-old man with idiopathic hypersomnia. Features that are typical of narcolepsy are cataplexy and weight gain. Features that are typical of both narcolepsy and idiopathic hypersomnia are daytime naps, insomnia, sleep paralysis and hypnagogic hallucinations. Additional testing in patients with hypersomnia should include a polysomnography in order to exclude other sleeping disorders, and a mean sleep latency test. Practice shows that both patients with narcolepsy and those with idiopathic hypersomnia benefit from treatment with stimulating drugs such as modafinil.

  7. EEG-tomographic studies with LORETA on vigilance differences between narcolepsy patients and controls and subsequent double-blind, placebo-controlled studies with modafinil.

    PubMed

    Saletu, M; Anderer, P; Saletu-Zyhlarz, G M; Mandl, M; Arnold, O; Zeitlhofer, J; Saletu, B

    2004-11-01

    The aim of the present study was to identify brain regions associated with vigilance in untreated and modafinil-treated narcoleptic patients by means of low-resolution brain electromagnetic tomography (LORETA). 16 drug-free narcoleptics and 16 normal controls were included in the baseline investigation. Subsequently patients participated in a double-blind, placebo-controlled crossover study receiving a three-week fixed titration of modafinil (200, 300, 400 mg) and placebo. Measurements comprised LORETA, the Multiple Sleep Latency Test (MSLT) and the Epworth Sleepiness Scale (ESS) obtained before and after three weeks' therapy. Statistical overall analysis by means of the omnibus significance test demonstrated significant inter-group differences in the resting (R-EEG), but not in the vigilance-controlled recordings (V-EEG). Subsequent univariate analysis revealed a decrease in alpha-2 and beta 1-3 power in prefrontal, temporal and parietal cortices, with the right hemisphere slightly more involved in this vigilance decrement. Modafinil 400 mg/d as compared with placebo induced changes opposite to the aforementioned baseline differences (key-lock principle) with a preponderance in the left hemisphere. This increase in vigilance resulted in an improvement in the MSLT and the ESS. LORETA provided evidence of a functional deterioration of the fronto-temporo-parietal network of the right-hemispheric vigilance system in narcolepsy and a therapeutic effect of modafinil on the left hemisphere, which is less affected by the disease.

  8. Diagnosis and management of central hypersomnias

    PubMed Central

    Susta, Marek

    2012-01-01

    Central hypersomnias are diseases manifested in excessive daytime sleepiness (EDS) not caused by disturbed nocturnal sleep or misaligned circadian rhythms. Central hypersomnias includes narcolepsy with and without cataplexy, recurrent hypersomnia, idiopathic hypersomnia, with and without long sleep time, behaviorally induced insufficient sleep syndrome, hypersomnia and narcolepsy due to medical conditions, and finally hypersomnia induced by substance intake. The Epworth Sleepiness Scale is a subjective tool mostly used for EDS assessment, while the Multiple Sleep Latency Test serves as an objective diagnostic method for narcolepsy and idiopathic hypersomnias. As for symptomatic therapy of EDS, the central nervous system stimulants modafinil and methylphenidate seem to work well in most cases and in narcolepsy and Parkinson’s disease; sodium oxybate also has notable therapeutic value. PMID:22973425

  9. Orexin gene therapy restores the timing and maintenance of wakefulness in narcoleptic mice.

    PubMed

    Kantor, Sandor; Mochizuki, Takatoshi; Lops, Stefan N; Ko, Brian; Clain, Elizabeth; Clark, Erika; Yamamoto, Mihoko; Scammell, Thomas E

    2013-08-01

    Narcolepsy is caused by selective loss of the orexin/hypocretin-producing neurons of the hypothalamus. For patients with narcolepsy, chronic sleepiness is often the most disabling symptom, but current therapies rarely normalize alertness and do not address the underlying orexin deficiency. We hypothesized that the sleepiness of narcolepsy would substantially improve if orexin signaling were restored in specific brain regions at appropriate times of day. We used gene therapy to restore orexin signaling in a mouse model of narcolepsy. In these Atx mice, expression of a toxic protein (ataxin-3) selectively kills the orexin neurons. To induce ectopic expression of the orexin neuropeptides, we microinjected an adeno-associated viral vector coding for prepro-orexin plus a red fluorescence protein (AAV-orexin) into the mediobasal hypothalamus of Atx and wild-type mice. Control mice received an AAV coding only for red fluorescence protein. Two weeks later, we recorded sleep/wake behavior, locomotor activity, and body temperature and examined the patterns of orexin expression. Atx mice rescued with AAV-orexin produced long bouts of wakefulness and had a normal diurnal pattern of arousal, with the longest bouts of wake and the highest amounts of locomotor activity in the first hours of the night. In addition, AAV-orexin improved the timing of rapid eye movement sleep and the consolidation of nonrapid eye movement sleep in Atx mice. These substantial improvements in sleepiness and other symptoms of narcolepsy demonstrate the effectiveness of orexin gene therapy in a mouse model of narcolepsy. Additional work is needed to optimize this approach, but in time, AAV-orexin could become a useful therapeutic option for patients with narcolepsy.

  10. Benefits and risk of sodium oxybate in idiopathic hypersomnia versus narcolepsy type 1: a chart review.

    PubMed

    Leu-Semenescu, Smaranda; Louis, Pauline; Arnulf, Isabelle

    2016-01-01

    Few stimulants have been evaluated for the treatment of idiopathic hypersomnia (IH). Sodium oxybate (indicated in narcolepsy type 1, NT1) has not been tested in IH patients. The aim of this study is to retrospectively evaluate the benefit/risk ratio of sodium oxybate in IH versus NT1 using a chart review. We reviewed the files of 46 patients with IH (35.7 ± 12.6 years old, 78% women) and 47 patients with NT1 (44.1 ± 18 years old, 47% women) and evaluated the benefits of sodium oxybate using the Epworth sleepiness scale (ESS) and a four-point scale assessing the global benefit, sleep inertia, sleepiness, sleep duration, and sleep onset latency. The spontaneously reported side effects were collected. Sodium oxybate was prescribed at a lower dose in IH than in NT1 (4.3 ± 2.2 vs. 6.6 ± 2.8 g/night, p <0.0001) patients after having tried more (3.2 ± 1.4 vs. 2.2 ± 1, p <0.0001) stimulants, but it produced a similar ESS change (-3.5 ± 4.5 vs. -3.2 ± 4.2 points) in the IH and NT1 groups. Severe morning inertia was improved in 24/34 (71%) patients with IH. During the follow-up period (15.8 months in IH vs. 35 months in NT1 groups), 53% IH and 68% NT1 patients dropped out. The side effects were as frequent in the IH group as in the NT1 group (67% vs. 52%), but nausea (40% vs. 13%) and dizziness (34.3% vs. 4.3%) were more frequent in the IH group. The benefit/risk ratio of sodium oxybate in IH- was similar to NT1-associated sleepiness, with additional benefits on severe morning inertia, despite using smaller doses in more refractory patients. Copyright © 2015 Elsevier B.V. All rights reserved.

  11. Hypocretin-1 CSF levels in anti-Ma2 associated encephalitis.

    PubMed

    Overeem, S; Dalmau, J; Bataller, L; Nishino, S; Mignot, E; Verschuuren, J; Lammers, G J

    2004-01-13

    Idiopathic narcolepsy is associated with deficient hypocretin transmission. Narcoleptic symptoms have recently been described in paraneoplastic encephalitis with anti-Ma2 antibodies. The authors measured CSF hypocretin-1 levels in six patients with anti-Ma2 encephalitis, and screened for anti-Ma antibodies in patients with idiopathic narcolepsy. Anti-Ma autoantibodies were not detected in patients with idiopathic narcolepsy. Four patients with anti-Ma2 encephalitis had excessive daytime sleepiness; hypocretin-1 was not detectable in their cerebrospinal fluid, suggesting an immune-mediated hypocretin dysfunction.

  12. Hypocretin-1 CSF levels in anti-Ma2 associated encephalitis

    PubMed Central

    Overeem, S.; Dalmau, J.; Bataller, L.; Nishino, S.; Mignot, E.; Verschuuren, J.; Lammers, G.J.

    2008-01-01

    Idiopathic narcolepsy is associated with deficient hypocretin transmission. Narcoleptic symptoms have recently been described in paraneoplastic encephalitis with anti-Ma2 antibodies. The authors measured CSF hypocretin-1 levels in six patients with anti-Ma2 encephalitis, and screened for anti-Ma antibodies in patients with ideopathic narcolepsy. Anti-Ma autoantibodies were not detected in patients with idiopathic narcolepsy. Four patients with anti-Ma2 encephalitis had excessive daytime sleepiness; hypocretin-1 was not detectable in their cerebrospinal fluid, suggesting an immune-mediated hypocretin dysfunction. PMID:14718718

  13. Polysomnographic and neurometabolic features may mark preclinical autosomal dominant cerebellar ataxia, deafness, and narcolepsy due to a mutation in the DNA (cytosine-5-)-methyltransferase gene, DNMT1.

    PubMed

    Moghadam, Keivan Kaveh; Pizza, Fabio; Tonon, Caterina; Lodi, Raffaele; Carelli, Valerio; Poli, Francesca; Franceschini, Christian; Barboni, Piero; Seri, Marco; Ferrari, Simona; La Morgia, Chiara; Testa, Claudia; Cornelio, Ferdinando; Liguori, Rocco; Winkelmann, Juliane; Lin, Ling; Mignot, Emmanuel; Plazzi, Giuseppe

    2014-05-01

    We aimed to report the clinical picture of two asymptomatic daughters of a patient with autosomal dominant cerebellar ataxia, deafness, and narcolepsy (ADCA-DN) due to a mutation in the DNA (cytosine-5-)-methyltransferase gene, DNMT1. Clinical assessment based on history, neurologic examination, sleep recordings, neurophysiologic neuroimaging, and genetic tests was performed. History and neurologic examination in both subjects were unremarkable. Genetic analysis disclosed in both the paternally-inherited heterozygous point mutation in the DNMT1 gene. Sleep recordings found sleep-onset rapid eye movement periods (SOREMPs) and proton magnetic resonance spectroscopy (MRS) revealed increased cerebellar myoinositol (mI) in both subjects. Auditory and ophthalmologic investigations as well as structural brain magnetic resonance imaging (MRI) scans revealed no abnormalities. The two asymptomatic carriers of the heterozygous DNMT1 mutation for ADCA-DN, a late-onset neurodegenerative disease, presented with SOREMPs associated with an increase of mI in the brain, a marker of glial cell activity and density characteristic of early stages of neurodegenerative diseases. Therefore, SOREMPs may precede the clinical picture of ADCA-DN as an early polysomnographic marker of central nervous system involvement detected by MRS. Copyright © 2014 Elsevier B.V. All rights reserved.

  14. MX2 Gene Expression Tends to be Downregulated in Subjects with HLA-DQB1*0602

    PubMed Central

    Tanaka, Susumu; Honda, Yutaka; Honda, Makoto

    2008-01-01

    Objective: There is a close association between narcolepsy and the human leukocyte antigen (HLA)-DQB1*0602. The detailed influence and function of this specific HLA allele with regard to narcolepsy have not yet been elucidated. Our previous report identified the myxovirus resistance 2 (MX2) gene as a narcolepsy-specific dysregulated gene; however, the report had a limitation—the control groups were not HLA matched. In this study, we examined the possibility of an association between MX2 expression and HLA haplotypes. Designs: The expression levels of the MX2 gene in 3 groups (24 narcolepsy with cataplexy patients; 24 age-, sex-, and HLA-DQB1 genotype-matched controls; and 24 age- and sex-matched controls without the HLA-DQB1*0602 allele) were measured by quantitative real-time RT-PCR. Results: The expression level of the MX2 gene tended to be downregulated in subjects carrying HLA-DQB1*0602, compared with that of the control subjects without this allele. There was no difference in the MX2 expression level between the narcolepsy subjects and the HLA-DQB1 genotype-matched control subjects. Conclusion: Our previous finding—the narcolepsy-specific reduction of MX2 gene expression—was not replicated in this follow-up study. The expression level of the MX2 gene in white blood cells was found to be lower in subjects with the HLA-DQB1*0602 than in subjects without this allele, suggesting that there exists a relationship between the HLA-DQB1*0602 allele and MX2 gene expression. This might be a possible explanation for the strong HLA association observed in narcolepsy. Citation: Tanaka S; Honda Y; Honda M. MX2 gene expression tends to be downregulated in subjects with HLA-DQB1*0602. SLEEP 2008;31(5):749-751. PMID:18517045

  15. PubMed

    Monaca, C; Franco, P; Philip, P; Dauvilliers, Y

    In the new international classification of sleep disorders (ICSD-3), narcolepsy is differentiated into two distinct pathologies: type 1 narcolepsy (NT1) and type 2 narcolepsy (NT2). NT1 is characterised by periods of an irrepressible need to sleep, cataplexy (a sudden loss of muscle tone triggered by emotion) and in some cases the presence of symptoms such as hypnagogic hallucinations, sleep paralysis and disturbed night-time sleep. Its physiopathology is based on the loss of hypocretin neurons in the hypothalamus, seemingly connected to an auto-immune process. By definition, cataplexy is absent and the hypocretin levels in the CSF are normal in NT2. Confirming the diagnosis requires polysomnography and multiple sleep latency tests. The choice of further investigations is based on the presence or absence of typical cataplexy. Further investigations include HLA typing, lumbar puncture to measure the hypocretin level in the CSF, or even brain imagery in the case of narcolepsy suspected to be secondary to an underlying pathology. In this consensus we propose recommendations for the work-up to be carried out during diagnosis and follow-up for patients suffering from narcolepsy. Copyright © 2016 Elsevier Masson SAS. All rights reserved.

  16. Flow cytometry analysis of T-cell subsets in cerebrospinal fluid of narcolepsy type 1 patients with long-lasting disease.

    PubMed

    Moresco, Monica; Lecciso, Mariangela; Ocadlikova, Darina; Filardi, Marco; Melzi, Silvia; Kornum, Birgitte Rahbek; Antelmi, Elena; Pizza, Fabio; Mignot, Emmanuel; Curti, Antonio; Plazzi, Giuseppe

    2018-04-01

    Type 1 narcolepsy (NT1) is a central hypersomnia linked to the destruction of hypocretin-producing neurons. A great body of genetic and epidemiological data points to likely autoimmune disease aetiology. Recent reports have characterized peripheral blood T-cell subsets in NT1, whereas data regarding the cerebrospinal fluid (CSF) immune cell composition are lacking. The current study aimed to characterize the T-cell and natural killer (NK) cell subsets in NT1 patients with long disease course. Immune cell subsets from CSF and peripheral blood mononuclear cell (PBMC) samples were analysed by flow cytometry in two age-balanced and sex-balanced groups of 14 NT1 patients versus 14 healthy controls. The frequency of CSF cell groups was compared with PBMCs. Non-parametric tests were used for statistical analyses. The NT1 patients did not show significant differences of CSF immune cell subsets compared to controls, despite a trend towards higher CD4 + terminally differentiated effector memory T cells. T cells preferentially displayed a memory phenotype in the CSF compared to PBMCs. Furthermore, a reduced frequency of CD4 + terminally differentiated effector memory T cells and an increased frequency of NK CD56 bright cells was observed in PBMCs from patients compared to controls. Finally, the ratio between CSF and peripheral CD4 + terminally differentiated effector memory T cells was two-fold increased in NT1 patients versus controls. Significant differences in PBMCs and in CSF/PBMC ratios of immune cell profile were found in NT1 patients compared to healthy controls. These differences might have arisen from the different HLA status, or be primary or secondary to hypocretin deficiency. Further functional studies in patients close to disease onset are required to understand NT1 pathophysiology. Copyright © 2017 Elsevier B.V. All rights reserved.

  17. Hypnagogic and hypnopompic hallucinations: pathological phenomena?

    PubMed

    Ohayon, M M; Priest, R G; Caulet, M; Guilleminault, C

    1996-10-01

    Hypnagogic and hypnopompic hallucinations are common in narcolepsy. However, the prevalence of these phenomena in the general population is uncertain. A representative community sample of 4972 people in the UK, aged 15-100, was interviewed by telephone (79.6% of those contacted). Interviews were performed by lay interviewers using a computerised system that guided the interviewer through the interview process. Thirty-seven per cent of the sample reported experiencing hypnagogic hallucinations and 12.5% reported hypnopompic hallucinations. Both types of hallucinations were significantly more common among subjects with symptoms of insomnia, excessive daytime sleepiness or mental disorders. According to this study, the prevalence of narcolepsy in the UK is 0.04%. Hypnagogic and hypnopompic hallucinations were much more common than expected, with a prevalence that far exceeds that which can be explained by the association with narcolepsy. Hypnopompic hallucinations may be a better indicator of narcolepsy than hypnagogic hallucinations in subjects reporting excessive daytime sleepiness.

  18. Orexin Gene Therapy Restores the Timing and Maintenance of Wakefulness in Narcoleptic Mice

    PubMed Central

    Kantor, Sandor; Mochizuki, Takatoshi; Lops, Stefan N.; Ko, Brian; Clain, Elizabeth; Clark, Erika; Yamamoto, Mihoko; Scammell, Thomas E.

    2013-01-01

    Study Objectives: Narcolepsy is caused by selective loss of the orexin/hypocretin-producing neurons of the hypothalamus. For patients with narcolepsy, chronic sleepiness is often the most disabling symptom, but current therapies rarely normalize alertness and do not address the underlying orexin deficiency. We hypothesized that the sleepiness of narcolepsy would substantially improve if orexin signaling were restored in specific brain regions at appropriate times of day. Design: We used gene therapy to restore orexin signaling in a mouse model of narcolepsy. In these Atx mice, expression of a toxic protein (ataxin-3) selectively kills the orexin neurons. Interventions: To induce ectopic expression of the orexin neuropeptides, we microinjected an adeno-associated viral vector coding for prepro-orexin plus a red fluorescence protein (AAV-orexin) into the mediobasal hypothalamus of Atx and wild-type mice. Control mice received an AAV coding only for red fluorescence protein. Two weeks later, we recorded sleep/wake behavior, locomotor activity, and body temperature and examined the patterns of orexin expression. Measurements and Results: Atx mice rescued with AAV-orexin produced long bouts of wakefulness and had a normal diurnal pattern of arousal, with the longest bouts of wake and the highest amounts of locomotor activity in the first hours of the night. In addition, AAV-orexin improved the timing of rapid eye movement sleep and the consolidation of nonrapid eye movement sleep in Atx mice. Conclusions: These substantial improvements in sleepiness and other symptoms of narcolepsy demonstrate the effectiveness of orexin gene therapy in a mouse model of narcolepsy. Additional work is needed to optimize this approach, but in time, AAV-orexin could become a useful therapeutic option for patients with narcolepsy. Citation: Kantor S; Mochizuki T; Lops SN; Ko B; Clain E; Clark E; Yamamoto M; Scammell TE. Orexin gene therapy restores the timing and maintenance of wakefulness

  19. What Does One Sleep-Onset REM Period—During Either Nocturnal Polysomnography or Multiple Sleep Latency Test—Mean in Differential Diagnosis of Central Hypersomnias?

    PubMed

    Bozluolcay, Melda; Nalbantoglu, Mecbure; Benbir Senel, Gulcin; Karadeniz, Derya

    2015-08-01

    The differentiation of narcolepsy without cataplexy from idiopathic hypersomnia is based on the number of sleep-onset rapid eye movement periods (SOREMPs) observed by multiple sleep latency test (MSLT) and nocturnal polysomnography. The main aim of this study was to investigate the utility of SOREMP in differential diagnosis of central hypersomnias. The authors retrospectively evaluated consecutive 101 patients with a normal polysomnography other than the presence of SOREMP and/or REM without atonia and a latency of ≤8 minutes in MSLT. The authors classified patients as follows: 52 patients had at least 2 SOREMPs (narcolepsy group), 23 had no SOREMPs (idiopathic hypersomnia group), and 26 patients had only 1 SOREMP (intermediate group). In polysomnographic recordings, both mean sleep latency and REM latency were significantly shorter in the narcolepsy (P = 0.012, P < 0.001, respectively) and intermediate groups (P = 0.005 and P = 0.035, respectively) compared with the idiopathic hypersomnia group. In MSLT recordings, sleep latency was 2.7 ± 2.2 minutes in the narcolepsy group, 3.6 ± 1.4 minutes in the intermediate group, and 5.2 ± 2.7 minutes in the idiopathic hypersomnia group (P < 0.001). The mean REM latency and sleep stages SOREMPs arised from were similar between the narcolepsy and intermediate groups. To date, SOREMPs in MSLT and polysomnography remain the sole electrodiagnostic feature that discriminates narcolepsy without cataplexy from idiopathic hypersomnia. Different parameters or combined criteria are being increasingly investigated to increase the sensitivity and specificity of MSLT. The findings showed an altered instability of REM sleep not only in patients with 2 or more SOREMPs in MSLT but also in patients with one SOREMP.

  20. Narcolepsy

    MedlinePlus

    ... when you fall asleep. This is called a sleep attack. These periods can last from a few seconds to a few minutes. They may happen after eating, while talking to someone, or during other situations. Most often, ...

  1. Presentations of primary hypersomnia in Chinese children.

    PubMed

    Han, Fang; Lin, Ling; Li, Jing; Aran, Adi; Dong, Song X; An, Pei; Zhao, Long; Li, Ming; Li, Qian Y; Yan, Han; Wang, Jie S; Gao, Hui Y; Li, Mei; Gao, Zhan C; Strohl, Kingman P; Mignot, Emmanuel

    2011-05-01

    To retrospectively describe childhood presentations of primary hypersomnia with an emphasis on narcolepsy-cataplexy in a Chinese population. A total of 417 children (< 18 years old) successively presenting with complaints of hypersomnia without anatomic cause or sleep apnea risk were evaluated using the Stanford Sleep Inventory, human leukocyte antigen (HLA) DQB1*0602 typing, and MSLT recordings. CSF hypocretin-1 was measured in 47 cases to document hypocretin deficiency. A subgroup ("narcolepsy/hypocretin deficiency") with likely hypocretin deficiency (low hypocretin-1 or HLA positive with clear-cut cataplexy) was further examined for presentations prior to, around, or after puberty. Narcolepsy with (n = 361) or without (n = 17) cataplexy presented at an earlier age and with increased male predominance when compared to idiopathic hypersomnia (n = 39, P < 0.01). Nearly 70% of those with narcolepsy/hypocretin deficiency (n = 271) had disease onset before age 10 y, and 15% had onset before age 6, an unusually young age distribution. Onset was prior to puberty in 78% of cases. Clinical features were similar in presentations across puberty groups except for sleep paralysis, which increased in frequency with age/puberty. Mean sleep latency (MSL) decreased and the number of sleep onset REM periods (SOREMPs) increased with age/puberty, but MSLT diagnosis criteria (MSL ≤ 8 min, ≥ 2 SOREMPs) were similarly positive across groups. Familial clustering was present in only 1.7% of probands. In children presenting with a complaint of primary hypersomnia to a sleep clinic in China, 86% (361/417) meet criteria for narcolepsy with cataplexy. Puberty did not affect positivity on the MSLT as a diagnostic feature. Sleep paralysis was the only symptom that increased with increasing age. In addition, narcolepsy with cataplexy in our clinic population appeared to begin at a younger age than usually reported in other studies.

  2. Cataplexy as a side effect of modafinil in a patient without narcolepsy☆

    PubMed Central

    Lopes, Eduardo; Pereira, Danielle; da Silva Behrens, Nilce Sanny Costa; de Almeida Fonseca, Hassana; Calvancanti, Paola Oliveira; de Araújo Lima, Taís Figueiredo; Pradella-Hallinan, Marcia; Castro, Juliana; Tufik, Sergio; Coelho, Fernando Morgadinho Santos

    2014-01-01

    Narcolepsy is a disease in which there is diurnal excessive sleepiness with sleep attacks and a prevalence in the general population of 1/4000 individuals. Classically, it is characterized by cataplexy, sleep paralysis, hypnagogic hallucinations and fragmented sleep. The use of modafinil in the treatment of narcolepsy is the first option of treatment for diurnal excessive sleepiness. Although considered a safe drug for use in patients with narcolepsy, being utilized for more than 20 years, modafinil possesses a series of side effects, some of them still not fully researched or described. Side effects such as headache, nausea, anxiety, insomnia, lumbago, diarrhea, dyspepsia, rhinitis and vertigo are the most frequent. However, the clinical follow-up of patients under treatment with modafinil must be intensive and the side effects ought to be noted and evaluated. The under-response to treatment or the unexpected side effects must always be directed to differential diagnostics. The objective of this article is to describe an unexpected side effect of the use of modafinil in a patient with incorrect diagnosis of narcolepsy. PMID:26483900

  3. Proton spectroscopy in the narcoleptic syndrome. Is there evidence of a brainstem lesion?

    PubMed

    Ellis, C M; Simmons, A; Lemmens, G; Williams, S C; Parkes, J D

    1998-02-01

    There is controversy regarding the relationship of structural or biochemical brainstem lesions to "idiopathic" narcolepsy. Most cases of the narcoleptic syndrome are considered to be idiopathic because no structural lesion is detectable, although some cases of secondary narcolepsy are known to be associated with no structural brainstem lesions. Using proton spectroscopy, we determined levels of ventral pontine metabolite pools in 12 normal subjects and 12 subjects with idiopathic narcolepsy. REM sleep is generated in ventral pontine areas. Proton spectroscopy was used to study levels of N-acetyl aspartate (NAA) as a marker of cell mass, creatine and phosphocreatine (Cr + PCr), and choline (Cho). The intensity of the peaks, as determined by the area under the peak (AUP), was measured. The AUP correlates with the quantity of chemical present. In this study, the ratios of NAA to Cr + PCr were similar in normal subjects and in narcoleptic subjects with idiopathic narcolepsy. No differences in measured metabolic ratio were observed in subjects who slept during the scan procedure compared with those who remained awake. Subjects with "symptomatic" narcolepsy accompanied by an obvious structural brain lesion were not studied. Proton spectroscopy of the brain initiates a new kind of neurochemistry, allowing the noninvasive study of metabolic pools in the living human brain without the use of any kind of tracer or radioactive molecule. In this study, there was no evidence of cell loss in the ventral pontine areas of subjects with the narcoleptic syndrome.

  4. Key sleep neurologic disorders

    PubMed Central

    St. Louis, Erik K.

    2014-01-01

    Summary Sleep disorders are frequent comorbidities in neurologic patients. This review focuses on clinical aspects and prognosis of 3 neurologic sleep disorders: narcolepsy, restless legs syndrome/Willis-Ekbom disease (RLS/WED), and REM sleep behavior disorder (RBD). Narcolepsy causes pervasive, enduring excessive daytime sleepiness, adversely affecting patients' daily functioning. RLS/WED is characterized by an uncomfortable urge to move the legs before sleep, often evolving toward augmentation and resulting in daylong bothersome symptoms. RBD causes potentially injurious dream enactment behaviors that often signify future evolution of overt synucleinopathy neurodegeneration in as many as 81% of patients. Timely recognition, referral for polysomnography, and longitudinal follow-up of narcolepsy, RLS/WED, and RBD patients are imperatives for neurologists in providing quality comprehensive patient care. PMID:24605270

  5. An Overview of Sleep Deprivation and The Ameliorative Effects of Modafinil

    DTIC Science & Technology

    2002-11-01

    H., and Jouvet, M., Successful treatment of idiopathic hypersomnia and narcolepsy with modafinil. Progress in Neuro-Psychopharmacology and Biological...Laffont, F., Cathala, H.P., and Kohler, F. Effect of modafinil on narcolepsy and idiopathic hypersomnia . in the 5th European Congress of Sleep Research...amphetamine and modafinil on the sleep/wake cycle during experimental hypersomnia induced by sleep deprivation in the cat. Journal of Sleep Research, 2000. 9

  6. Presentations of Primary Hypersomnia in Chinese Children

    PubMed Central

    Han, Fang; Lin, Ling; Li, Jing; Aran, Adi; Dong, Song X.; An, Pei; Zhao, Long; Li, Ming; Li, Qian Y.; Yan, Han; Wang, Jie S.; Gao, Hui Y.; Li, Mei; Gao, Zhan C.; Strohl, Kingman P.; Mignot, Emmanuel

    2011-01-01

    Objective: To retrospectively describe childhood presentations of primary hypersomnia with an emphasis on narcolepsy-cataplexy in a Chinese population. Methods: A total of 417 children (< 18 years old) successively presenting with complaints of hypersomnia without anatomic cause or sleep apnea risk were evaluated using the Stanford Sleep Inventory, human leukocyte antigen (HLA) DQB1*0602 typing, and MSLT recordings. CSF hypocretin-1 was measured in 47 cases to document hypocretin deficiency. A subgroup (“narcolepsy/hypocretin deficiency”) with likely hypocretin deficiency (low hypocretin-1 or HLA positive with clear-cut cataplexy) was further examined for presentations prior to, around, or after puberty. Results: Narcolepsy with (n = 361) or without (n = 17) cataplexy presented at an earlier age and with increased male predominance when compared to idiopathic hypersomnia (n = 39, P < 0.01). Nearly 70% of those with narcolepsy/hypocretin deficiency (n = 271) had disease onset before age 10 y, and 15% had onset before age 6, an unusually young age distribution. Onset was prior to puberty in 78% of cases. Clinical features were similar in presentations across puberty groups except for sleep paralysis, which increased in frequency with age/puberty. Mean sleep latency (MSL) decreased and the number of sleep onset REM periods (SOREMPs) increased with age/puberty, but MSLT diagnosis criteria (MSL ≤ 8 min, ≥ 2 SOREMPs) were similarly positive across groups. Familial clustering was present in only 1.7% of probands. Conclusion: In children presenting with a complaint of primary hypersomnia to a sleep clinic in China, 86% (361/417) meet criteria for narcolepsy with cataplexy. Puberty did not affect positivity on the MSLT as a diagnostic feature. Sleep paralysis was the only symptom that increased with increasing age. In addition, narcolepsy with cataplexy in our clinic population appeared to begin at a younger age than usually reported in other studies. Citation: Han

  7. [Symptomatic hypersomnia due to orexin deficiency in hypothalamic lesions].

    PubMed

    Kanbayashi, Takashi; Arii, Junko; Kubota, Hiroaki; Yano, Tamami; Kashiwagi, Mitsuru; Yoshikawa, Sousuke; Tohyama, Jun; Sawaishi, Yukio

    2006-09-01

    Narcolepsy is characterized by excessive daytime sleepiness (EDS), cataplexy and other abnormal manifestations of REM sleep. Recently, it was discovered that the pathophysiology of idiopathic narcolepsy-cataplexy is linked to orexin ligand deficiency in the brain and cerebrospinal fluid. Orexin neurons localize in the posterior hypothalamic area, which was previously described as "waking center" by von Economo in 1920s. Hypersomnia due to orexin ligand deficiency can also occur during the course of other neurological conditions, such as hypothalamic tumor, encephalopathy and demyelinating disorder (i.e. symptomatic hypersomnia). We experienced 8 pediatric cases with symptomatic hypersomnia. These cases were diagnosed as brain tumor (n = 2), head trauma (n = 1), encephalopathy (n = 1), demyelinating disorder (n = 3) and infarction (n = 1). Six pediatric cases with orexin measurements from the literatures were additionally included and total 14 cases were studied. Although it is difficult to rule out the comorbidity of idiopathic narcolepsy in some cases, a review of the case histories reveals numerous unquestionable cases of symptomatic hypersomnia. In these cases, the occurrences of the hypersomnia run parallel with the rise and fall of the causative diseases. Most of symptomatic hypersomnia cases show both extended nocturnal sleep time and EDS consisting of prolonged sleep episodes of NREM sleep. The features of nocturnal sleep and EDS in symptomatic hypersomnia are more similar to idiopathic hypersomnia than to narcolepsy.

  8. Role of the medial prefrontal cortex in cataplexy.

    PubMed

    Oishi, Yo; Williams, Rhiannan H; Agostinelli, Lindsay; Arrigoni, Elda; Fuller, Patrick M; Mochizuki, Takatoshi; Saper, Clifford B; Scammell, Thomas E

    2013-06-05

    Narcolepsy is characterized by chronic sleepiness and cataplexy, episodes of profound muscle weakness that are often triggered by strong, positive emotions. Narcolepsy with cataplexy is caused by a loss of orexin (also known as hypocretin) signaling, but almost nothing is known about the neural mechanisms through which positive emotions trigger cataplexy. Using orexin knock-out mice as a model of narcolepsy, we found that palatable foods, especially chocolate, markedly increased cataplexy and activated neurons in the medial prefrontal cortex (mPFC). Reversible suppression of mPFC activity using an engineered chloride channel substantially reduced cataplexy induced by chocolate but did not affect spontaneous cataplexy. In addition, neurons in the mPFC innervated parts of the amygdala and lateral hypothalamus that contain neurons active during cataplexy and that innervate brainstem regions known to regulate motor tone. These observations indicate that the mPFC is a critical site through which positive emotions trigger cataplexy.

  9. Polymorphism Located between CPT1B and CHKB, and HLA-DRB1*1501-DQB1*0602 Haplotype Confer Susceptibility to CNS Hypersomnias (Essential Hypersomnia)

    PubMed Central

    Miyagawa, Taku; Honda, Makoto; Kawashima, Minae; Shimada, Mihoko; Tanaka, Susumu; Honda, Yutaka; Tokunaga, Katsushi

    2009-01-01

    Background SNP rs5770917 located between CPT1B and CHKB, and HLA-DRB1*1501-DQB1*0602 haplotype were previously identified as susceptibility loci for narcolepsy with cataplexy. This study was conducted in order to investigate whether these genetic markers are associated with Japanese CNS hypersomnias (essential hypersomnia: EHS) other than narcolepsy with cataplexy. Principal Findings EHS was significantly associated with SNP rs5770917 (Pallele = 3.6×10−3; OR = 1.56; 95% c.i.: 1.12–2.15) and HLA-DRB1*1501-DQB1*0602 haplotype (P positivity = 9.2×10−11; OR = 3.97; 95% c.i.: 2.55–6.19). No interaction between the two markers (SNP rs5770917 and HLA-DRB1*1501-DQB1*0602 haplotype) was observed in EHS. Conclusion CPT1B, CHKB and HLA are candidates for susceptibility to CNS hypersomnias (EHS), as well as narcolepsy with cataplexy. PMID:19404393

  10. Role of the medial prefrontal cortex in cataplexy

    PubMed Central

    Oishi, Yo; Williams, Rhiannan H.; Agostinelli, Lindsay; Arrigoni, Elda; Fuller, Patrick M.; Mochizuki, Takatoshi; Saper, Clifford B.; Scammell, Thomas E.

    2013-01-01

    Narcolepsy is characterized by chronic sleepiness and cataplexy - episodes of profound muscle weakness that are often triggered by strong, positive emotions. Narcolepsy with cataplexy is caused by a loss of orexin (also known as hypocretin) signaling, but almost nothing is known about the neural mechanisms through which positive emotions trigger cataplexy. Using orexin knockout mice as a model of narcolepsy, we found that palatable foods, especially chocolate, markedly increased cataplexy and activated neurons in the medial prefrontal cortex (mPFC). Reversible suppression of mPFC activity using an engineered chloride channel substantially reduced cataplexy induced by chocolate but did not affect spontaneous cataplexy. In addition, neurons in the mPFC innervated parts of the amygdala and lateral hypothalamus that contain neurons active during cataplexy, and that innervate brainstem regions known to regulate motor tone. These observations indicate that the mPFC is a critical site through which positive emotions trigger cataplexy. PMID:23739971

  11. Prevalence and Clinical Correlates of a Short Onset REM Period (SOREMP) during Routine PSG

    PubMed Central

    Cairns, Alyssa; Bogan, Richard

    2015-01-01

    Study Objectives: The objectives of this study were to quantify the (1) sensitivity and specificity of nocturnal PSG SOREMP (REM latency ≤ 15 min) for narcolepsy in those being evaluated for hypersomnolence and (2) prevalence and predictors of SOREMP during baseline PSG for patients being evaluated for various sleep disorders. Design: This was a retrospective analysis of a large repository of de-identified PSG and MSLT test results from 2007 to 2013. Setting and Patients: Patient records were retrieved from a repository of studies completed at a variety of sleep laboratories across the USA. Included in the analyses were 79,651 general sleep clinic patients (without an MSLT; 48% male; 72% Caucasian) and an additional 3,059 patients (31.3% male; 72% Caucasian) being evaluated for hypersomnolence (with a consecutive MSLT). Interventions: NA. Measurements and Results: For patients being evaluated for hypersomnolence, the prevalence of PSG SOREMP increased in a dose-response fashion with the number of REM onsets that occurred on a consecutive MSLT (0.5% for no MSLT SOREMPs to > 33.0% for those with 5 MSLT SOREMPs). Overall, having a PSG SOREMP was highly specific (99.5%; 95% CI: 99.1–99.7%) but not sensitive (6.7%; 95% CI: 4.7–9.2%) for narcolepsy. The prevalence of PSG SOREMP for patients in the general sleep clinic sample (i.e., not being evaluated by a consecutive MSLT) was 0.8% and was much higher in those that work night/swing shift. In adjusted models, African American race contributed to the most variance in PSG SOREMP. Conclusions: A short onset rapid eye movement (REM) latency occurs rarely in general sleep clinic samples (< 1.0%), but is highly specific for the diagnosis of narcolepsy. Although rare, the prevalence of the phenomenon is much higher than the estimated prevalence of narcolepsy and may provide a critical opportunity for practitioners to identify narcolepsy in sleep clinic patients. These data also suggest that the utility of polysomnography

  12. Sodium oxybate for cataplexy.

    PubMed

    Lemon, Michael D; Strain, Joe D; Farver, Debra K

    2006-03-01

    To review the pharmacology, pharmacokinetics, clinical efficacy, adverse effects, drug interactions, precautions, dosing recommendations, and patient counseling of sodium oxybate for the treatment of cataplexy in patients with narcolepsy. OVID and PubMed databases were searched (1966-January 2006) using the key words sodium oxybate, gamma-hydroxybutyrate, narcolepsy, and cataplexy. Only English-language articles were selected. All information on sodium oxybate related to narcolepsy and cataplexy was considered. Study selection included human trials evaluating safety and efficacy of sodium oxybate for the treatment of cataplexy. Sodium oxybate is approved by the Food and Drug Administration for the treatment of excessive daytime sleepiness and cataplexy in patients with narcolepsy. In placebo-controlled trials, sodium oxybate demonstrated efficacy in reducing the number of cataplexy attacks. The dosing regimen includes a split dose given at bedtime and 2.5-4 hours later due to its short elimination half-life. The drug is generally well tolerated, with headache, nausea, dizziness, pain, and somnolence being the most common adverse events. Sodium oxybate is safe and effective for the treatment of cataplexy. Potential disadvantages include a multiple dosing regimen, abuse potential, cost, and a closed distribution system. Potential advantages demonstrated in clinical trials include significant decreases in the number of weekly cataplexy attacks, improvement in daytime sleepiness, and improvement in the Clinical Global Impression of Change score and nighttime awakenings. Overall, sodium oxybate provides a new option for the treatment of cataplexy.

  13. Hypocretin-1 deficiency in a girl with ROHHAD syndrome.

    PubMed

    Dhondt, Karlien; Verloo, Patrick; Verhelst, Hélène; Van Coster, Rudy; Overeem, Sebastiaan

    2013-09-01

    Rapid-onset obesity with hypothalamic dysfunction, hypoventilation, and autonomic dysregulation (ROHHAD) is a rare and complex pediatric syndrome, essentially caused by dysfunction of 3 vital systems regulating endocrine, respiratory, and autonomic nervous system functioning. The clinical spectrum of ROHHAD is broad, but sleep/wake disorders have received relatively little attention so far, although the central hypothalamic dysfunction would make the occurrence of sleep symptoms likely. In this case report, we expand the phenotype of ROHHAD with a number of striking sleep symptoms that together can be classified as a secondary form of narcolepsy. We present a 7-year-old girl with ROHHAD who displayed the classic features of narcolepsy with cataplexy: excessive daytime sleepiness with daytime naps, visual hallucinations, and partial cataplexy reflected in intermittent loss of facial muscle tone. Nocturnal polysomnography revealed sleep fragmentation and a sleep-onset REM period characteristic for narcolepsy. The diagnosis was confirmed by showing an absence of hypocretin-1 in the cerebrospinal fluid. We discuss potential pathophysiological implications as well as symptomatic treatment options.

  14. Genetic ablation of hypocretin neurons alters behavioral state transitions in zebrafish.

    PubMed

    Elbaz, Idan; Yelin-Bekerman, Laura; Nicenboim, Julian; Vatine, Gad; Appelbaum, Lior

    2012-09-12

    Sleep is an essential biological need of all animals studied to date. The sleep disorder narcolepsy is characterized by excessive daytime sleepiness, fragmentation of nighttime sleep, and cataplexy. Narcolepsy is caused by selective degeneration of hypothalamic hypocretin/orexin (HCRT) neurons. In mammals, HCRT neurons primarily regulate the sleep/wake cycle, feeding, reward-seeking, and addiction. The role of HCRT neurons in zebrafish is implicated in both sleep and wake regulation. We established a transgenic zebrafish model enabling inducible ablation of HCRT neurons and used these animals to understand the function of HCRT neurons and narcolepsy. Loss of HCRT neurons increased the expression of the HCRT receptor (hcrtr). Behavioral assays revealed that HCRT neuron-ablated larvae had normal locomotor activity, but demonstrated an increase in sleep time during the day and an increased number of sleep/wake transitions during both day and night. Mild sleep disturbance reduced sleep and increased c-fos expression in HCRT neuron-ablated larvae. Furthermore, ablation of HCRT neurons altered the behavioral response to external stimuli. Exposure to light during the night decreased locomotor activity of wild-type siblings, but induced an opposite response in HCRT neuron-ablated larvae. Sound stimulus during the day reduced the locomotor activity of wild-type sibling larvae, while HCRT neuron-ablated larvae demonstrated a hyposensitive response. This study establishes zebrafish as a model for narcolepsy, and indicating a role of HCRT neurons in regulation of sleep/wake transitions during both day and night. Our results further suggest a key role of HCRT neurons in mediating behavioral state transitions in response to external stimuli.

  15. Narcolepsy: Fact Sheet

    MedlinePlus

    ... go limp or unable to move (cataplexy), vivid dream-like images or hallucinations, and total paralysis just ... REM) sleep after about 60 to 90 minutes. Dreams occur during REM sleep, and the brain keeps ...

  16. What Is Narcolepsy?

    MedlinePlus

    Skip to main content U.S. Department of Health & Human Services Health Topics Health Topics A-Z Clinical Trials Publications and Resources Health Education and Awareness The Science Science Home Blood Disorders and ...

  17. Daytime continuous polysomnography predicts MSLT results in hypersomnias of central origin.

    PubMed

    Pizza, Fabio; Moghadam, Keivan K; Vandi, Stefano; Detto, Stefania; Poli, Francesca; Mignot, Emmanuel; Ferri, Raffaele; Plazzi, Giuseppe

    2013-02-01

    In the diagnostic work-up of hypersomnias of central origin, the complaint of excessive daytime sleepiness should be objectively confirmed by MSLT findings. Indeed, the features and diagnostic utility of spontaneous daytime sleep at 24 h continuous polysomnography (PSG) have never been investigated. We compared daytime PSG features to MSLT data in 98 consecutive patients presenting with excessive daytime sleepiness and with a final diagnosis of narcolepsy with cataplexy/hypocretin deficiency (n = 39), narcolepsy without cataplexy (n = 7), idiopathic hypersomnia without long sleep time (n = 19), and 'hypersomnia' with normal sleep latency at MSLT (n = 33). Daytime sleep time was significantly higher in narcolepsy-cataplexy but similar in the other groups. Receiver operating characteristics (ROC) curves showed that the number of naps during daytime PSG predicted a mean sleep latency ≤8 min at MSLT with an area under the curve of 0.67 ± 0.05 (P = 0.005). The number of daytime sleep-onset REM periods (SOREMPs) in spontaneous naps strikingly predicted the scheduled occurrence of two or more SOREMPs at MSLT, with an area under the ROC curve of 0.93 ± 0.03 (P < 10(-12) ). One spontaneous SOREMP during daytime had a sensitivity of 96% with specificity of 74%, whereas two SOREMPs had a sensitivity of 75%, with a specificity of 95% for a pathological REM sleep propensity at MSLT. The features of spontaneous daytime sleep well correlated with MSLT findings. Notably, the occurrence of multiple spontaneous SOREMPs during daytime clearly identified patients with narcolepsy, as well as during the MSLT. © 2012 European Sleep Research Society.

  18. Idiopathic hypersomnia: a study of 77 cases.

    PubMed

    Anderson, Kirstie N; Pilsworth, Samantha; Sharples, Linda D; Smith, Ian E; Shneerson, John M

    2007-10-01

    To review the clinical and polysomnographic characteristics of idiopathic hypersomnia as well as the long-term response to treatment. The Respiratory Support and Sleep Centre at Papworth Hospital, Cambridge, UK. A large database of more than 6000 patients with sleep disorders was reviewed. A retrospective study of the clinical and polysomnographic characteristics of 77 patients with idiopathic hypersomnia was performed. Comparison with a similar group of patients with narcolepsy was performed. The response to drug treatment was assessed in 61 patients over a mean follow-up of 3.8 years. Idiopathic hypersomnia was 60% as prevalent as narcolepsy. Comparison with a similar group of patients with narcolepsy showed that those with idiopathic hypersomnia were more likely to have prolonged unrefreshing daytime naps, a positive family history, increased slow-wave sleep, and a longer sleep latency on the Multiple Sleep Latency Test. The results of the Multiple Sleep Latency Test were not helpful in predicting disease severity or treatment response. The clinical features were heterogeneous and of variable severity. The majority of patients with idiopathic hypersomnia had symptoms that remained stable over many years, but 11% had spontaneous remission, which was never seen in narcolepsy. Two thirds of patients with idiopathic hypersomnolence had a sustained improvement in daytime somnolence with medication, although a third needed high doses or combinations of drugs. Idiopathic hypersomnolence has characteristic clinical and polysomnographic features but the prolonged latency on the Multiple Sleep Latency Test raises doubt about the validity of this test within the current diagnostic criteria. The disease often responds well to treatment and a substantial minority of patients appear to spontaneously improve.

  19. First rapid eye movement sleep periods and sleep-onset rapid eye movement periods in sleep-stage sequencing of hypersomnias.

    PubMed

    Drakatos, Panagis; Kosky, Christopher A; Higgins, Sean E; Muza, Rexford T; Williams, Adrian J; Leschziner, Guy D

    2013-09-01

    Discrimination between narcolepsy, idiopathic hypersomnia, and behavior-induced inadequate sleep syndrome (BIISS) is based on clinical features and on specific nocturnal polysomnography (NPSG) and multiple sleep latency test (MSLT) results. However, previous studies have cast doubt on the specificity and sensitivity of these diagnostic tools. Eleven variables of the NPSG were analyzed in 101 patients who were retrospectively diagnosed with narcolepsy with cataplexy (N+C) (n=24), narcolepsy without cataplexy (N-C) (n=38), idiopathic hypersomnia with long sleep period (IHL) (n=21), and BIISS (n=18). Fifteen out of 24 N+C and 8 out of 38 N-C entered the first rapid eye movement (REM) sleep period (FREMP) from sleep stage 1 (N1) or wake (W), though this sleep-stage sequence did not arise in the other patient groups. FREMP stage sequence was a function of REM sleep latency (REML) for both N+C and N-C groups. FREMP stage sequence was not associated with mean sleep latency (MSL) in N+C but was associated in N-C, which implies heterogeneity within the N-C group. REML also was a useful discriminator. Depending on the cutoff period, REML had a sensitivity and specificity of up to 85.5% and 97.4%, respectively. The FREMP stage sequence may be a useful tool in the diagnosis of narcolepsy, particularly in conjunction with sleep-stage sequence analysis of sleep-onset REM periods (SOREMPs) in the MSLT; it also may provide a helpful intermediate phenotype in the clarification of heterogeneity in the N-C diagnostic group. However, larger prospective studies are necessary to confirm these findings. Copyright © 2013 Elsevier B.V. All rights reserved.

  20. Prader Willi syndrome and obstructive sleep apnea: co-occurrence in the pediatric population.

    PubMed

    Sedky, Karim; Bennett, David S; Pumariega, Andres

    2014-04-15

    A high prevalence of obstructive sleep apnea (OSA) occurs in children with Prader-Willi syndrome (PWS). Yet, due in part to the relatively small samples previously used, the prevalence of OSA has varied greatly across studies. It is also unclear if factors such as age, gender, body mass index (BMI), or type of genetic imprinting are associated with increased risk for OSA among children with PWS. To evaluate the (a) prevalence of OSA, as well as narcolepsy, in pediatric populations diagnosed with PWS; (b) effects of age, gender, body mass index, and genetic imprinting on OSA severity; and (c) efficacy of adenotonsillectomy (AT) for decreasing OSA severity in this population. All studies assessing OSA among children with PWS through August 2013 were identified using the PubMed/Medline, Psych Info, Cochrane library, and Google Scholar data bases. Fourteen studies of children diagnosed with PWS and who were assessed for OSA using polysomnography (PSG) met inclusion criteria (n = 224 children). The prevalence of OSA across studies was 79.91% (n = 179/224). Among youths with OSA, 53.07% had mild OSA, 22.35% moderate OSA, and 24.58% severe OSA. Narcolepsy was found to occur in 35.71% of children with PWS. Adenotonsillectomy was associated with improvement in OSA for most children with PWS. However, residual OSA was present in the majority of cases post-surgery. This study confirms the high prevalence of OSA and narcolepsy among children with PWS. Screening for OSA and narcolepsy among children with PWS is recommended. In addition, while adenotonsillectomy was effective in reducing OSA for some children, alternative treatments may need to be considered, given the only moderate response rate.

  1. From radioimmunoassay to mass spectrometry: a new method to quantify orexin-A (hypocretin-1) in cerebrospinal fluid

    PubMed Central

    Hirtz, Christophe; Vialaret, Jérôme; Gabelle, Audrey; Nowak, Nora; Dauvilliers, Yves; Lehmann, Sylvain

    2016-01-01

    I125 radioimmunoassay (RIA) is currently the standard technique for quantifying cerebrospinal fluid (CSF) orexin-A/hypocretin-1, a biomarker used to diagnose narcolepsy type 1. However, orexin-A RIA is liable to undergo cross-reactions with matrix constituents generating interference, high variability between batches, low precision and accuracy, and requires special radioactivity precautions. Here we developed the first quantitative mass spectrometry assay of orexin-A based on a multiple reaction monitoring (MRM) approach. This method was tested in keeping with the Clinical and Laboratory Standards Institute (CLSI) guidelines and its clinical relevance was confirmed by comparing patients with narcolepsy type 1 versus patients with other neurological conditions. The results obtained using MRM and RIA methods were highly correlated, and Bland–Altman analysis established their interchangeability. However, the MRM values had a wider distribution and were 2.5 time lower than the RIA findings. In conclusion, this method of assay provides a useful alternative to RIA to quantify orexin-A, and may well replace it not only in narcolepsy type 1, but also in the increasing number of pathologies in which the quantification of this analyte is relevant. PMID:27165941

  2. Normal cerebrospinal fluid histamine and tele-methylhistamine levels in hypersomnia conditions.

    PubMed

    Dauvilliers, Yves; Delallée, Nathalie; Jaussent, Isabelle; Scholz, Sabine; Bayard, Sophie; Croyal, Mickael; Schwartz, Jean-Charles; Robert, Philippe

    2012-10-01

    To determine the activity of cerebral histaminergic system evaluated by CSF levels of histamine (HA) and tele-methylhistamine (t-MHA), its major metabolite, and their relationships with hypocretin-1 levels in a large population of patients with hypersomnia and neurological conditions. sensitive liquid chromatographic-electrospray/tandem mass spectrometric assay was developed for the simultaneous quantification of CSF HA and t-MHA. ata were collected and CSF hypocretin-1 levels were measured using radioimmunoassay at the Sleep Disorders Center, Montpellier, France. CSF HA and t-MHA were measured in Bioprojet-Biotech, France One hundred fourteen unrelated patients with a suspicion of central hypersomnia underwent one night of polysomnography followed by the multiple sleep latency test. Sleep disorders were diagnosed clinically and using sleep studies: narcolepsy-cataplexy NC (n = 56), narcolepsy without cataplexy NwC (n = 27), idiopathic hypersomnia IH (n = 11), secondary narcolepsy (n = 3), and unspecified hypersomnia Uns EDS (n = 17). Fifty neurological patients without daytime sleepiness were included as controls. No between-hypersomnia group differences were found for CSF HA levels (median 708.62 pM extreme range [55.92-3335.50] in NC; 781.34 [174.08-4391.50] in NwC; 489.42 [177.45-906.70] in IH, and 1155.40 [134.80-2736.59] in Uns EDS) or for t-MHA levels. No association was found between CSF HA, t-MHA, or HA + t-MHA, sleepiness, treatment intake, and frequency of cataplexy. A slight negative correlation was found between age and HA levels. Further adjustment for the age revealed no significant HA levels difference between hypersomnia patients and controls. CSF histamine and tele-methylhistamine did not significantly differ between patients with narcolepsy-cataplexy and other etiologies of non-hypocretin-1 deficient central hypersomnias; these measurements, therefore, are not useful in assessing the etiology or severity of centrally mediated hypersomnia.

  3. Idiopathic Hypersomnia: A Study of 77 Cases

    PubMed Central

    Anderson, Kirstie N.; Pilsworth, Samantha; Sharples, Linda D.; Smith, Ian E.; Shneerson, John M.

    2007-01-01

    Study Objectives: To review the clinical and polysomnographic characteristics of idiopathic hypersomnia as well as the long-term response to treatment. Setting: The Respiratory Support and Sleep Centre at Papworth Hospital, Cambridge, UK. Patients and Design: A large database of more than 6000 patients with sleep disorders was reviewed. A retrospective study of the clinical and polysomnographic characteristics of 77 patients with idiopathic hypersomnia was performed. Comparison with a similar group of patients with narcolepsy was performed. The response to drug treatment was assessed in 61 patients over a mean follow-up of 3.8 years. Measurements and Results: Idiopathic hypersomnia was 60% as prevalent as narcolepsy. Comparison with a similar group of patients with narcolepsy showed that those with idiopathic hypersomnia were more likely to have prolonged unrefreshing daytime naps, a positive family history, increased slow-wave sleep, and a longer sleep latency on the Multiple Sleep Latency Test. The results of the Multiple Sleep Latency Test were not helpful in predicting disease severity or treatment response. The clinical features were heterogeneous and of variable severity. The majority of patients with idiopathic hypersomnia had symptoms that remained stable over many years, but 11% had spontaneous remission, which was never seen in narcolepsy. Two thirds of patients with idiopathic hypersomnolence had a sustained improvement in daytime somnolence with medication, although a third needed high doses or combinations of drugs. Conclusions: Idiopathic hypersomnolence has characteristic clinical and polysomnographic features but the prolonged latency on the Multiple Sleep Latency Test raises doubt about the validity of this test within the current diagnostic criteria. The disease often responds well to treatment and a substantial minority of patients appear to spontaneously improve. Citation: Anderson KN; Pilsworth S; Sharples LD; Smith IE; Shneerson JM. Idiopathic

  4. 77 FR 14769 - Meeting of the Uniform Formulary Beneficiary Advisory Panel

    Federal Register 2010, 2011, 2012, 2013, 2014

    2012-03-13

    ... Citizen Comments. 4. Scheduled Therapeutic Class Reviews (Comments will follow each agenda item). a. Attention Deficit Hyperactivity Disorder-Narcolepsy Agents. b. Anti-Platelet Hemorhelogic Agents. c...

  5. Modafinil in the treatment of excessive sleepiness

    PubMed Central

    Schwartz, Jonathan RL

    2008-01-01

    The wake-promoting agent modafinil is approved for the treatment of excessive sleepiness associated with obstructive sleep apnea (OSA), shift work disorder (SWD), and narcolepsy. In OSA, modafinil is recommended for use as an adjunct to standard therapies that treat the underlying airway obstruction. This article reviews the literature on modafinil (pharmacology, pharmacokinetics, efficacy, tolerability, and abuse potential), with emphasis on use of modafinil in the treatment of excessive sleepiness in patients with OSA, SWD, and narcolepsy. In large-scale, double-blind, placebo-controlled studies, modafinil improved objectively determined sleep latency, improved overall clinical condition related to severity of sleepiness, and reduced patient-reported sleepiness. Improvements in wakefulness were accompanied by improvements in behavioral alertness, functional status, and health-related quality of life. In patients with SWD, diary data showed modafinil reduced the maximum level of sleepiness during night shift work, level of sleepiness during the commute home, and incidence of accidents or near-accidents during the commute home when compared with placebo. Modafinil was well tolerated, without adversely affecting cardiovascular parameters or scheduled sleep. These findings and those of extension studies which reported improvements were maintained suggest modafinil has a beneficial effect on daily life and well-being in patients with excessive sleepiness associated with OSA, SWD, or narcolepsy. PMID:19920895

  6. Genetics Home Reference: narcolepsy

    MedlinePlus

    ... result from a loss of particular brain cells (neurons) in a part of the brain called the hypothalamus. These cells normally produce chemicals called hypocretins (also known as orexins), which have many important functions in the body. In particular, hypocretins regulate the ...

  7. Increased anxiety but normal fear and safety learning in orexin-deficient mice.

    PubMed

    Khalil, Radwa; Fendt, Markus

    2017-03-01

    The loss of orexin neurons in humans leads to the disease narcolepsy, characterized by daytime sleepiness and cataplexy. Recent data suggest that orexin is also involved in emotional processing. The goal of the present study was to evaluate fear and safety learning as well as unconditioned fear (anxiety) in orexin-deficient animals. Orexin-deficient mice are an established animal model used to investigate the neuropathology and potential treatments for narcolepsy. Here, we present novel data showing that orexin-deficient mice express increased anxiety in the open field, light-dark box test and carnivore odor-induced avoidance, but are normal in fear and safety learning. These findings suggest an important role of orexin in brain areas involved in anxiety. Copyright © 2016 Elsevier B.V. All rights reserved.

  8. Sleep Disorders

    MedlinePlus

    ... Narcolepsy - daytime "sleep attacks" Nightmares, night terrors, sleepwalking, sleep talking, head banging, wetting the bed and grinding your teeth are kinds of sleep problems called parasomnias. There are treatments for most ...

  9. [Physiopathology of idiopathic hypersomnia. Current studies and new orientations].

    PubMed

    Billiard, M; Rondouin, G; Espa, F; Dauvilliers, Y; Besset, A

    2001-11-01

    In 1976 Bedrich Roth coined the term "idiopathic hypersomnia" and described two forms of the disease, one monosymptomatic, manifested only by excessive daytime sleepiness, and one polysymptomatic, characterized by excessive daytime sleepiness, nocturnal sleep of abnormally long duration and signs of "sleep drunkenness" on awakening. In comparison with that of narcolepsy, the pathophysiology of idiopathic hypersomnia remains poorly known. There are two main reasons for that: the absence of clinical and polysomnographic criteria pathognomonic or at least characteristic of the condition, as the cataplexies and the sleep onset REM periods of narcolepsy, and also the absence of a natural animal model comparable with the canine model of narcolepsy. The first investigations have stressed the frequent familial pattern of idiopathic hypersomnia. Later on biochemical assays have been performed in the CSF with results in favour of a dysfunction of noradrenergic systems. In the light of the two process model of sleep regulation in which sleep propensity is determined by a homeostatic process S and a circadian process C and of the later three-process model of regulation in which sleepiness/alertness are simulated by the combined action of a homeostatic process, a circadian process and sleep inertia, we suggest that idiopathic hypersomnia is not a pathological entity in itself, but rather the consequence of chronic sleep deprivation in very long sleepers.

  10. Orexin Gene Transfer into Zona Incerta Neurons Suppresses Muscle Paralysis in Narcoleptic Mice

    PubMed Central

    Liu, Meng; Konadhode, RodaRani; Begum, Suraiya; Pelluru, Dheeraj; Gerashchenko, Dmitry; Sakurai, Takeshi; Yanagisawa, Masashi; van den Pol, Anthony N.

    2011-01-01

    Cataplexy, a sudden unexpected muscle paralysis, is a debilitating symptom of the neurodegenerative sleep disorder, narcolepsy. During these attacks, the person is paralyzed, but fully conscious and aware of their surroundings. To identify potential neurons that might serve as surrogate orexin neurons to suppress such attacks, the gene for orexin (hypocretin), a peptide lost in most human narcoleptics, was delivered into the brains of the orexin-ataxin-3 transgenic mouse model of human narcolepsy. Three weeks after the recombinant adenoassociated virus (rAAV)-mediated orexin gene transfer, sleep–wake behavior was assessed. rAAV-orexin gene delivery into neurons of the zona incerta (ZI), or the lateral hypothalamus (LH) blocked cataplexy. Orexin gene transfer into the striatum or in the melanin-concentrating hormone neurons in the ZI or LH had no such effect, indicating site specificity. In transgenic mice lacking orexin neurons but given rAAV-orexin, detectable levels of orexin-A were evident in the CSF, indicating release of the peptide from the surrogate neurons. Retrograde tracer studies showed that the amygdala innervates the ZI consistent with evidence that strong emotions trigger cataplexy. In turn, the ZI projects to the locus ceruleus, indicating that the ZI is part of a circuit that stabilizes motor tone. Our results indicate that these neurons might also be recruited to block the muscle paralysis in narcolepsy. PMID:21508228

  11. Future Perspectives in Sleep Medicine.

    PubMed

    Huon, Leh-Kiong Anne; Guilleminault, Christian

    2017-01-01

    "Sleep Medicine" is now a specialty in its own right. Currently, there is increasing recognition of the very negative impact sleep disorders have on learning, education, safety, and quality of life. Technological advances will help us to break down diagnoses (e.g., narcolepsy has now been subdivided into types 1 and 2, depending upon the hypocretin levels in the spinal fluid) and to discover relationships to other bodily systems (e.g., type 1 narcolepsy potentially being an autoimmune disorder). The modern lifestyle of many, as characterized by a shortening of sleep periods, shift work, jet lag, and the need to be constantly available, means that advances in sleep medicine may result in a major understanding of more balanced "work-rest lifestyle" modifications. © 2017 S. Karger AG, Basel.

  12. Your Wake-Up Call

    MedlinePlus

    ... sleep comprise a broad range of problems, including sleep apnea, narcolepsy, insomnia, parasomnia, jet-lag syndrome, and disturbed ... health risks." Photo courtesy of NIH/NHLBI/NCSDR Sleep Apnea and Children Snoring is common in children and ...

  13. Get the Facts: Prescription Drug Abuse on College Campuses

    MedlinePlus

    ... a stimulant for non- medical reasons in the past year compared to those who aren’t in college ... disorder (ADHD) and narcolepsy) non-medically in the past year were also binge drinkers and more than half ...

  14. Methamphetamine/Dextroamphetamine and Pregnancy

    MedlinePlus

    ... d-amphetamine is a prescription medication used for attention deficit hyperactivity disorder (ADHD) and narcolepsy (extreme daytime sleepiness). ... of dexamphetamine into breast milk during treatment for attention deficit hyperactivity disorder. Br J Clin Pharmacol. 63(3): ...

  15. Daytime REM Sleep in Parkinson’s Disease

    PubMed Central

    Bliwise, Donald L.; Trotti, Lynn Marie; Juncos, Jorge J.; Factor, Stewart A.; Freeman, Alan; Rye, David B.

    2012-01-01

    Background Previous studies have demonstrated both clinical and neurochemical similarities between Parkinson’s disease (PD) and narcolepsy. The intrusion of REM sleep into the daytime remains a cardinal feature of narcolepsy, but the importance of these intrusions in PD remains unclear. In this study we examined REM sleep during daytime Maintenance of Wakefulness Testing (MWT) in PD patients. Methods Patients spent 2 consecutive nights and days in the sleep laboratory. During the daytime, we employed a modified MWT procedure in which each daytime nap opportunity (4 per day) was extended to 40 minutes, regardless of whether the patient was able to sleep or how much the patient slept. We examined each nap opportunity for the presence of REM sleep and time to fall asleep. Results Eleven of 63 PD patients studied showed 2 or more REM episodes and 10 showed 1 REM episode on their daytime MWTs. Nocturnal sleep characteristics and sleep disorders were unrelated to the presence of daytime REM sleep, however, patients with daytime REM were significantly sleepier during the daytime than those patients without REM. Demographic and clinical variables, including Unified Parkinson’s Disease Rating Scale motor scores and levodopa dose equivalents, were unrelated to the presence of REM sleep. Conclusions A sizeable proportion of PD patients demonstrated REM sleep and daytime sleep tendency during daytime nap testing. These data confirm similarities in REM intrusions between narcolepsy and PD, perhaps suggesting parallel neurodegenerative conditions of hypocretin deficiency. PMID:22939103

  16. Use of recurrence plots in the analysis of pupil diameter dynamics in narcoleptics

    NASA Astrophysics Data System (ADS)

    Keegan, Andrew P.; Zbilut, J. P.; Merritt, S. L.; Mercer, P. J.

    1993-11-01

    Recurrence plots were used to evaluate pupil dynamics of subjects with narcolepsy. Preliminary data indicate that this nonlinear method of analyses may be more useful in revealing underlying deterministic differences than traditional methods like FFT and counting statistics.

  17. Treatment of Narcolepsy with Methamphetamine

    PubMed Central

    Miller, Merrill M.; Hajdukovic, Roza; Erman, Milton K.

    2008-01-01

    Summary Eight pairs of subjects (each consisting of a narcoleptic and a control matched on the basis of age, sex, educational background and job) were evaluated under the following double-blind, randomized treatment conditions: baseline, placebo, low dose and high dose methamphetamine. Subjects were drug-free for 2 weeks prior to beginning the protocol. Methamphetamine was the only drug taken during the protocol and was given in a single morning dose of 0, 20 or 40–60 mg to narcoleptics and 0, 5 or 10 mg to controls. The protocol was 28 days long, with each of the four treatment conditions lasting 4 days followed by 3 days of washout. Nighttime polysomnography and daytime testing were done during the last 24 hours of each treatment condition. Daytime sleep tendency was assessed with the multiple sleep latency test (MSLT). Daytime performance was assessed with performance tests including a simple, computer-based driving task. Narcoleptics’ mean MSLT sleep latency increased from 4.3 minutes on placebo to 9.3 minutes on high dose, compared with an increase from 10.4 to 17.1 minutes for controls. Narcoleptics’ error rate on the driving task decreased from 2.53% on placebo to 0.33% on high dose, compared with a decrease from 0.22% to 0.16% for controls. The effects of methamphetamine on nocturnal sleep were generally dose-dependent and affected sleep continuity and rapid eye movement (REM) sleep. Elimination half life was estimated to be between 15.9 and 22.0 hours. Mild side effects emerged in a dose-dependent fashion and most often involved the central nervous system and gastrointestinal tract. We concluded that methamphetamine caused a dose-dependent decrease in daytime sleep tendency and improvement in performance in both narcoleptics and controls. Methamphetamine at doses of 40–60 mg allowed narcoleptics to function at levels comparable to those of unmedicated controls. PMID:8341891

  18. Normal Cerebrospinal Fluid Histamine and tele-Methylhistamine Levels in Hypersomnia Conditions

    PubMed Central

    Dauvilliers, Yves; Delallée, Nathalie; Jaussent, Isabelle; Scholz, Sabine; Bayard, Sophie; Croyal, Mickael; Schwartz, Jean-Charles; Robert, Philippe

    2012-01-01

    Study Objectives: To determine the activity of cerebral histaminergic system evaluated by CSF levels of histamine (HA) and tele-methylhistamine (t-MHA), its major metabolite, and their relationships with hypocretin-1 levels in a large population of patients with hypersomnia and neurological conditions. Design: sensitive liquid chromatographic-electrospray/tandem mass spectrometric assay was developed for the simultaneous quantification of CSF HA and t-MHA. Setting: ata were collected and CSF hypocretin-1 levels were measured using radioimmunoassay at the Sleep Disorders Center, Montpellier, France. CSF HA and t-MHA were measured in Bioprojet-Biotech, France Participants: One hundred fourteen unrelated patients with a suspicion of central hypersomnia underwent one night of polysomnography followed by the multiple sleep latency test. Sleep disorders were diagnosed clinically and using sleep studies: narcolepsy-cataplexy NC (n = 56), narcolepsy without cataplexy NwC (n = 27), idiopathic hypersomnia IH (n = 11), secondary narcolepsy (n = 3), and unspecified hypersomnia Uns EDS (n = 17). Fifty neurological patients without daytime sleepiness were included as controls. Measurements and Results: No between-hypersomnia group differences were found for CSF HA levels (median 708.62 pM extreme range [55.92-3335.50] in NC; 781.34 [174.08-4391.50] in NwC; 489.42 [177.45-906.70] in IH, and 1155.40 [134.80-2736.59] in Uns EDS) or for t-MHA levels. No association was found between CSF HA, t-MHA, or HA + t-MHA, sleepiness, treatment intake, and frequency of cataplexy. A slight negative correlation was found between age and HA levels. Further adjustment for the age revealed no significant HA levels difference between hypersomnia patients and controls. Conclusion: CSF histamine and tele-methylhistamine did not significantly differ between patients with narcolepsy-cataplexy and other etiologies of non-hypocretin-1 deficient central hypersomnias; these measurements, therefore, are not

  19. 77 FR 58849 - Prescription Drug User Fee Act Patient-Focused Drug Development; Public Meeting and Request for...

    Federal Register 2010, 2011, 2012, 2013, 2014

    2012-09-24

    ... the impact of the disease on patients, the spectrum of severity for those who have the disease, the... glomerular diseases. Narcolepsy. Huntington's Disease. Depression. Autism. Peripheral neuropathy.... Cancer and depression. Clotting disorders (e.g., hemophilia A (factor VIII deficiency) and von Willebrand...

  20. [Performance and personality of patients with hypersomnia].

    PubMed

    Mayer, G; Leonhardt, E

    1996-01-01

    5 groups of patients with hypersomnia (narcolepsy, posttraumatic, psychophysiologic, idiopathic hypersomnia and circadian sleep-wake disorders) were tested with a battery of psychometric tests (FPI, MMPI, BVND, BIV, Benton, d2, WIP), visual vigilance test, polysomnography and MSLT in order to investigate the context between personality and performance. MSLT showed a range from clear pathologic to borderline sleep latencies among all groups, only patients with posttraumatic hypersomnia and narcolepsy displayed sleep onset REM. Correct results of vigilance tests correlated negatively with performance-motivation and orientation in patients with narcolepsy and posttraumatic hypersomnia, whereas there was positive correlation for patients with idiopathic hypersomnia. In patients with psychophysiologic hypersomnia performance orientation and false reactions correlate negatively. Patients with posttraumatic hypersomnia have better results on d2. Benton and vigilance tests than all other groups. Results of personality diagnosis are similar to those of healthy subjects, while patients with psychophysiologic hypersomnia are more sensible than all other groups with high social fears and the highest disposition among all groups toward somatic complaints. Patients with idiopathic hypersomnia show strong introversion and inhibition. Patients with circadian sleep-wake disorders display the most striking personality disorders, which are most probably sequelae of their strong disease-dependent impairment. The degree of personality disorder seems to be strongly dependent on the duration of the hypersomnias. The assessment of the whole set of tests can only be recommended for patients with psychophysiologic hypersomnia and circadian sleep-wake disorders, a few tests suffice to describe the other groups.

  1. Abnormal sleep/wake dynamics in orexin knockout mice.

    PubMed

    Diniz Behn, Cecilia G; Klerman, Elizabeth B; Mochizuki, Takatoshi; Lin, Shih-Chieh; Scammell, Thomas E

    2010-03-01

    Narcolepsy with cataplexy is caused by a loss of orexin (hypocretin) signaling, but the physiologic mechanisms that result in poor maintenance of wakefulness and fragmented sleep remain unknown. Conventional scoring of sleep cannot reveal much about the process of transitioning between states or the variations within states. We developed an EEG spectral analysis technique to determine whether the state instability in a mouse model of narcolepsy reflects abnormal sleep or wake states, faster movements between states, or abnormal transitions between states. We analyzed sleep recordings in orexin knockout (OXKO) mice and wild type (WT) littermates using a state space analysis technique. This non-categorical approach allows quantitative and unbiased examination of sleep/wake states and state transitions. OXKO mice spent less time in deep, delta-rich NREM sleep and in active, theta-rich wake and instead spent more time near the transition zones between states. In addition, while in the midst of what should be stable wake, OXKO mice initiated rapid changes into NREM sleep with high velocities normally seen only in transition regions. Consequently, state transitions were much more frequent and rapid even though the EEG progressions during state transitions were normal. State space analysis enables visualization of the boundaries between sleep and wake and shows that narcoleptic mice have less distinct and more labile states of sleep and wakefulness. These observations provide new perspectives on the abnormal state dynamics resulting from disrupted orexin signaling and highlight the usefulness of state space analysis in understanding narcolepsy and other sleep disorders.

  2. Psychosocial Characteristics of Children with Central Disorders of Hypersomnolence Versus Matched Healthy Children.

    PubMed

    Avis, Kristin T; Shen, Jiabin; Weaver, Patrick; Schwebel, David C

    2015-11-15

    Hypersomnia of central origin from narcolepsy or idiopathic hypersomnia (IHS) is characterized by pathological levels of excessive daytime sleepiness (EDS). Central hypersomnia has historically been underdiagnosed and poorly understood, especially with respect to its impact on daytime functioning and quality of life in children. Describe the psychosocial adjustment of children treated for narcolepsy or IHS on school performance, quality of life, and physical/extracurricular activities. Using a matched case control design, we compared child self- and parent-reported data from thirty-three 8- to 16-year-olds with an established diagnosis of narcolepsy or IHS, according to ICSD-2 criteria, to that of 33 healthy children matched by age, race/ethnicity, gender, and household income. Assessments evaluated academic performance, quality of life and wellness, sleepiness, and participation in extracurricular activities. Compared to healthy controls, children with central hypersomnia had poorer daytime functioning in multiple domains. Children with hypersomnia missed more days of school and had lower grades than healthy controls. Children with hypersomnia had poorer quality of life by both parent and child report. Children with hypersomnia were significantly sleepier, had higher BMI, and were more likely to report a history of recent injury. Finally, children with hypersomnia engaged in fewer after-school activities than healthy controls. A range of significant psychosocial consequences are reported in children with hypersomnia even after a diagnosis has been made and treatments initiated. Health care professionals should be mindful of the psychosocial problems that may present in children with hypersomnia over the course of treatment. © 2015 American Academy of Sleep Medicine.

  3. Examining the Frequency of Stimulant Misuse among Patients with Primary Disorders of Hypersomnolence: A Retrospective Cohort Study.

    PubMed

    Mantyh, William G; Auger, R Robert; Morgenthaler, Timothy I; Silber, Michael H; Moore, Wendy R

    2016-05-15

    Narcolepsy and idiopathic hypersomnia are commonly treated by sleep specialists and encountered by other medical providers. Although pharmacotherapy with modafinil and traditional stimulants is considered the mainstay of treatment, physicians are often uncomfortable with their prescription because of concerns regarding misuse. The goal of this study was to assess the frequency of stimulant misuse in this population. A retrospective cohort study was performed evaluating patients 18 years and older diagnosed with narcolepsy with and without cataplexy and idiopathic hypersomnia with and without long sleep between 2003-2008. Patients were included if they obtained stimulant prescriptions from and had at least one follow-up visit subsequent to initial diagnosis at our center. Stimulant misuse was defined by multiple prescription sources or early refill requests, which are systematically entered into the record by nursing staff. A total of 105 patients met inclusion criteria for the study; 45 (42%) were male. Mean age at multiple sleep latency test was 42 (± 16). Twelve (11%) patients had a history of illicit substance misuse, and one (1%) patient demonstrated previous stimulant misuse. Fifty-seven (54%) patients carried psychiatric diagnoses, 88% of whom reported depression. Median duration of monitored stimulant therapy was 26 months (range 1-250). None of the 105 patients was found to have evidence of stimulant misuse. This study suggests that the frequency of stimulant misuse in patients with narcolepsy and idiopathic hypersomnia is extremely low. Concerns regarding drug misuse should not leverage decisions to provide long-term therapy. © 2016 American Academy of Sleep Medicine.

  4. French consensus. Management of patients with hypersomnia: Which strategy?

    PubMed

    Lopez, R; Arnulf, I; Drouot, X; Lecendreux, M; Dauvilliers, Y

    Central hypersomnias principally involves type 1 narcolepsy (NT1), type 2 narcolepsy (NT2) and idiopathic hypersomnia (IH). Despite great progress made in understanding the physiopathology of NT1 with low cerebrospinal fluid hypocretin-1 levels, current treatment remains symptomatic. The same applies to NT2 and IH, for which the physiopathology is still largely unknown. Controlling excessive daytime sleepiness (EDS), cataplexy, hypnagogic hallucinations, sleep paralysis and disturbed night-time sleep are key therapeutic targets in NT1. For IH and NT2, reducing EDS is the main objective. Based on European and American directives for the treatment of narcolepsy, we propose French recommendations for managing central hypersomnias as well as strategies in the case of drug-resistance. Stimulating treatments target EDS, and Modafinil is the first-line treatment. Other stimulants such as methylphenidate, pitolisant, and exceptionally dextro-amphetamine can be prescribed. Selective serotonin and noradrenaline reuptake inhibitor antidepressants are effective for the management of cataplexy in NT1. Sodium oxybate is an effective treatment for several symptoms, including EDS, cataplexy and disturbed night-time sleep. Treatment of central hypersomnia must also take into consideration frequent cardiovascular, metabolic and psychiatric comorbidities, particularly in NT1. New therapies are currently under study with the development of new stimulants and anti-cataplectics. The next few years will see innovative emerging therapies, based on a physiopathological approach, aiming to restore hypocretinergic transmission or to interrupt the autoimmune processes causing the loss of hypocretin neurons. Copyright © 2016 Elsevier Masson SAS. All rights reserved.

  5. Attention Deficit Hyperactivity Disorder: The Differential Diagnosis.

    ERIC Educational Resources Information Center

    Weinberg, Warren A.; Emslie, Graham J.

    This paper presents information on the diagnostic criteria and management of disorders that may be wrongly identified as Attention Deficit Hyperactivity Disorder (ADHD) or may coexist with ADHD thus complicating identification and treatment. The disorders discussed are: depression, mania, primary disorder of vigilance, narcolepsy, developmental…

  6. 78 FR 21613 - Prescription Drug User Fee Act Patient-Focused Drug Development; Announcement of Disease Areas...

    Federal Register 2010, 2011, 2012, 2013, 2014

    2013-04-11

    ... syndrome; narcolepsy; neurological manifestations of inborn errors of metabolism; Parkinson's disease and... available therapies. Patients who live with a disease have a direct stake in the outcome of FDA's decisions... of applications for new drugs in certain disease areas. For FDA's review divisions, this kind of...

  7. Text Mining of Journal Articles for Sleep Disorder Terminologies.

    PubMed

    Lam, Calvin; Lai, Fu-Chih; Wang, Chia-Hui; Lai, Mei-Hsin; Hsu, Nanly; Chung, Min-Huey

    2016-01-01

    Research on publication trends in journal articles on sleep disorders (SDs) and the associated methodologies by using text mining has been limited. The present study involved text mining for terms to determine the publication trends in sleep-related journal articles published during 2000-2013 and to identify associations between SD and methodology terms as well as conducting statistical analyses of the text mining findings. SD and methodology terms were extracted from 3,720 sleep-related journal articles in the PubMed database by using MetaMap. The extracted data set was analyzed using hierarchical cluster analyses and adjusted logistic regression models to investigate publication trends and associations between SD and methodology terms. MetaMap had a text mining precision, recall, and false positive rate of 0.70, 0.77, and 11.51%, respectively. The most common SD term was breathing-related sleep disorder, whereas narcolepsy was the least common. Cluster analyses showed similar methodology clusters for each SD term, except narcolepsy. The logistic regression models showed an increasing prevalence of insomnia, parasomnia, and other sleep disorders but a decreasing prevalence of breathing-related sleep disorder during 2000-2013. Different SD terms were positively associated with different methodology terms regarding research design terms, measure terms, and analysis terms. Insomnia-, parasomnia-, and other sleep disorder-related articles showed an increasing publication trend, whereas those related to breathing-related sleep disorder showed a decreasing trend. Furthermore, experimental studies more commonly focused on hypersomnia and other SDs and less commonly on insomnia, breathing-related sleep disorder, narcolepsy, and parasomnia. Thus, text mining may facilitate the exploration of the publication trends in SDs and the associated methodologies.

  8. Specificity of Direct Transition from Wake to REM Sleep in Orexin/ataxin-3 Transgenic Narcoleptic Mice

    PubMed Central

    Fujiki, Nobuhiro; Cheng, Timothy; Yoshino, Fuyumi; Nishino, Seiji

    2009-01-01

    To create operational criteria for polygraphic assessments of direct transitions from wake to REM sleep (DREM), as a murine analog of human cataplexy, we have analyzed DREM episodes in congenic lines of orexin/ataxin-3 transgenic [TG] mice and wild-type littermates. The sleep stage of each 10-second epoch was visually scored using our standard criteria. Specificity of DREM for narcoleptic TG mice and sensitivity to detect DREM was evaluated using different DREM criteria. We found that DREM transitions by 10-second epoch scoring are not specific for narcoleptic TG mice and also occur in WT mice during light period. These wake-to-REM transitions in WT mice (also seen in TG mice during light period) were characteristically different from DREM transitions in TG mice during dark period; they tended to occur as brief bouts of wakefulness interrupting extended episodes of REM sleep, suggesting that these transitions do not represent abnormal manifestations of REM sleep. We therefore defined the DREM transitions by requiring a minimum number of preceding wake epochs. Requiring no fewer than four consecutive epochs of wakefulness produced the best combination of specificity (95.9%) and sensitivity (66.0%). By definition, DREM in dark-period is 100% specific to narcolepsy and was 95.9% specific overall. In addition, we found that desipramine, a trycyclic anticataplectic, potently reduces DREM, while two wake-promoting compounds have moderate (d-amphetamine) and no (modafinil) effect on DREM; the effects mirror the anticataplectic effects of these compounds reported in canine and human narcolepsy. Our definition of DREM in murine narcolepsy may provide good electrophysiological measure for cataplexy-equivalent episodes. PMID:19416673

  9. Treatment of Sleep Disorders after Traumatic Brain Injury

    PubMed Central

    Castriotta, Richard J.; Atanasov, Strahil; Wilde, Mark C.; Masel, Brent E.; Lai, Jenny M.; Kuna, Samuel T.

    2009-01-01

    Study Objectives: Determine whether treatment of sleep disorders identified in brain injured adults would result in resolution of those sleep disorders and improvement of symptoms and daytime function. Methods: Prospective evaluation of unselected traumatic brain injury patients with nocturnal polysomnography (NPSG), multiple sleep latency test (MSLT), Epworth Sleepiness Scale (ESS), and neuropsychological testing including Psychomotor Vigilance Test (PVT), Profile of Mood States (POMS), and Functional Outcome of Sleep Questionnaire (FOSQ) before and after treatment with continuous positive airway pressure (CPAP) for obstructive sleep apnea (OSA), modafinil (200 mg) for narcolepsy and posttraumatic hypersomnia (PTH), or pramipexole (0.375 mg) for periodic limb movements in sleep (PLMS). Setting: Three academic medical centers. Participants: Fifty-seven (57) adults ≥ 3 months post traumatic brain injury (TBI). Measurements And Results: Abnormal sleep studies were found in 22 subjects (39%), of whom 13 (23%) had OSA, 2 (3%) had PTH, 3 (5%) had narcolepsy, 4 (7%) had PLMS, and 12 had objective excessive daytime sleepiness with MSLT score < 10 minutes. Apneas, hypopneas, and snoring were eliminated by CPAP in OSA subjects, but there was no significant change in MSLT scores. Periodic limb movements were eliminated with pramipexole. One of 3 narcolepsy subjects and 1 of 2 PTH subjects had resolution of hypersomnia with modafinil. There was no significant change in FOSQ, POMS, or PVT results after treatment. Conclusions: Treatment of sleep disorders after TBI may result in polysomnographic resolution without change in sleepiness or neuropsychological function. Citation: Castriotta RJ; Atanasov S; Wilde MC; Masel BE; Lai JM; Kuna ST. Treatment of sleep disorders after traumatic brain injury. J Clin Sleep Med 2009;5(2):137-144. PMID:19968047

  10. Comorbidity between central disorders of hypersomnolence and immune-based disorders.

    PubMed

    Barateau, Lucie; Lopez, Régis; Arnulf, Isabelle; Lecendreux, Michel; Franco, Patricia; Drouot, Xavier; Leu-Semenescu, Smaranda; Jaussent, Isabelle; Dauvilliers, Yves

    2017-01-03

    To assess and compare the frequencies of personal and family history of autoimmune diseases (AID), autoinflammatory disorders (ID), and allergies in a population of patients, adults and children, with narcolepsy type 1 (NT1), narcolepsy type 2 (NT2), and idiopathic hypersomnia (IH), 3 central hypersomnia disorders, and healthy controls. Personal and family history of AID, ID, and allergies were assessed by questionnaire and medical interview in a large cohort of 450 consecutive adult patients (206 NT1, 106 NT2, 138 IH) and 95 pediatric patients (80 NT1) diagnosed according to the third International Classification of Sleep Disorders criteria in national reference centers for narcolepsy in France and 751 controls (700 adults, 51 children) from the general population. Ten adults with NT1 (4.9%) had a comorbid AID vs 3.4% of adult controls, without between-group differences in adjusted models. AID frequency did not differ between children with NT1 and controls. Conversely, compared with controls, AID frequency was higher in adults with NT2 (p = 0.002), whereas ID (p = 0.0002) and allergy (p = 0.003) frequencies were higher in adults with IH. A positive family history of AID was found in the NT1 group and of ID in the IH group. NT1 is not associated with increased risk of comorbid immune disorders, in favor of a potentially unique pathophysiology. Conversely, compared with controls, the frequency of autoimmune diseases was higher in adults with NT2, whereas allergies and autoinflammatory disorders were more common in adults with IH, suggesting an immune dysregulation mechanism in these conditions. © 2016 American Academy of Neurology.

  11. Methylphenidate Enhances Extinction of Contextual Fear

    ERIC Educational Resources Information Center

    Abraham, Antony D.; Cunningham, Christopher L.; Lattal, K. Matthew

    2012-01-01

    Methylphenidate (MPH, Ritalin) is a norepinephrine and dopamine transporter blocker that is widely used in humans for treatment of attention deficit disorder and narcolepsy. Although there is some evidence that targeted microinjections of MPH may enhance fear acquisition, little is known about the effect of MPH on fear extinction. Here, we show…

  12. The Effects of Modafinil on Convergent and Divergent Thinking of Creativity: A Randomized Controlled Trial

    ERIC Educational Resources Information Center

    Mohamed, Ahmed Dahir

    2016-01-01

    Modafinil is a drug licensed for the treatment of narcolepsy and sleep apnea. Recently, modafinil has been reported to be used as a pharmacological cognitive enhancer by healthy individuals with no psychiatric disorders. This paper reports on a study that investigated the effects of modafinil on divergent and convergent thinking tasks of…

  13. Characteristics of REM Sleep Behavior Disorder in Childhood

    PubMed Central

    Lloyd, Robin; Tippmann-Peikert, Maja; Slocumb, Nancy; Kotagal, Suresh

    2012-01-01

    Study Objective: To describe our experience regarding the clinical and polysomnographic features of REM sleep behavior disorder (RBD) in childhood. Methods: This was a retrospective chart review of children and adolescents with RBD and REM sleep without atonia. Demographics, and clinical and polysomnographic information were tabulated. Our findings were compared with those in the existing literature. Results: The 15 subjects identified (13 RBD and 2 having REM sleep without atonia) had a mean age at diagnosis of 9.5 years (range 3-17 years); 11/15 (73%) were male. Nightmares were reported in 13/15 and excessive daytime sleepiness in 6/15. Two children had caused bodily harm to bedmate siblings. Comorbidities, which were multiple in some subjects, included anxiety (8/15), attention deficit disorder (10/15), nonspecific developmental delay (6/15), Smith-Magenis syndrome (1/15), pervasive developmental disorder (1/15), narcolepsy (1/15), idiopathic hypersomnia (1/15), and Moebius Syndrome (1/15). Abnormal MRI scans were seen in 5/8 evaluated subjects. Treatments consisted of clonazepam (10/15), melatonin (2/15), and discontinuation of a tricyclic agent (1/15), with a favorable response in 11 of 13. Two of 15 patients with REM sleep without atonia did not require pharmacotherapy. Conclusions: RBD in children may be associated with neurodevelopmental disabilities, narcolepsy, or medication use. It seems to be modestly responsive to benzodiazepines or melatonin. The etiology is distinct from that of common childhood arousal parasomnias and RBD in adults; congenital and neurodevelopmental disorders, medication effect, and narcolepsy coexisted in some, but none had an extrapyramidal neurodegenerative disorder. Citation: Lloyd R; Tippmann-Peikert M; Slocumb N; Kotagal S. Characteristics of REM sleep behavior disorder in childhood. J Clin Sleep Med 2012;8(2):127-131. PMID:22505856

  14. Abnormal Sleep/Wake Dynamics in Orexin Knockout Mice

    PubMed Central

    Diniz Behn, Cecilia G.; Klerman, Elizabeth B.; Mochizuki, Takatoshi; Lin, Shih-Chieh; Scammell, Thomas E.

    2010-01-01

    Study Objectives: Narcolepsy with cataplexy is caused by a loss of orexin (hypocretin) signaling, but the physiologic mechanisms that result in poor maintenance of wakefulness and fragmented sleep remain unknown. Conventional scoring of sleep cannot reveal much about the process of transitioning between states or the variations within states. We developed an EEG spectral analysis technique to determine whether the state instability in a mouse model of narcolepsy reflects abnormal sleep or wake states, faster movements between states, or abnormal transitions between states. Design: We analyzed sleep recordings in orexin knockout (OXKO) mice and wild type (WT) littermates using a state space analysis technique. This non-categorical approach allows quantitative and unbiased examination of sleep/wake states and state transitions. Measurements and Results: OXKO mice spent less time in deep, delta-rich NREM sleep and in active, theta-rich wake and instead spent more time near the transition zones between states. In addition, while in the midst of what should be stable wake, OXKO mice initiated rapid changes into NREM sleep with high velocities normally seen only in transition regions. Consequently, state transitions were much more frequent and rapid even though the EEG progressions during state transitions were normal. Conclusions: State space analysis enables visualization of the boundaries between sleep and wake and shows that narcoleptic mice have less distinct and more labile states of sleep and wakefulness. These observations provide new perspectives on the abnormal state dynamics resulting from disrupted orexin signaling and highlight the usefulness of state space analysis in understanding narcolepsy and other sleep disorders. Citation: Diniz Behn CG; Klerman EB; Mochizuki T; Lin S; Scammell TE. Abnormal sleep/wake dynamics in orexin knockout mice. SLEEP 2010;33(3):297-306. PMID:20337187

  15. Feeding-elicited cataplexy in orexin knockout mice

    PubMed Central

    Clark, Erika L.; Baumann, Christian R.; Cano, Georgina; Scammell, Thomas E.; Mochizuki, Takatoshi

    2009-01-01

    Mice lacking orexin/hypocretin signaling have sudden episodes of atonia and paralysis during active wakefulness. These events strongly resemble cataplexy, episodes of sudden muscle weakness triggered by strong positive emotions in people with narcolepsy, but it remains unknown whether murine cataplexy is triggered by positive emotions. To determine whether positive emotions elicit murine cataplexy, we placed orexin knockout (KO) mice on a scheduled feeding protocol with regular or highly palatable food. Baseline sleep/wake behavior was recorded with ad lib regular chow. Mice were then placed on a scheduled feeding protocol in which they received 60% of their normal amount of chow 3 hr after dark onset for the next 10 days. Wild-type and KO mice rapidly entrained to scheduled feeding with regular chow, with more wake and locomotor activity prior to the feeding time. On day 10 of scheduled feeding, orexin KO mice had slightly more cataplexy during the food-anticipation period and more cataplexy in the second half of the dark period, when they may have been foraging for residual food. To test whether more palatable food increases cataplexy, mice were then switched to scheduled feeding with an isocaloric amount of Froot Loops, a food often used as a reward in behavioral studies. With this highly palatable food, orexin KO mice had much more cataplexy during the food-anticipation period and throughout the dark period. The increase in cataplexy with scheduled feeding, especially with highly palatable food, suggests that positive emotions may trigger cataplexy in mice, just as in people with narcolepsy. Establishing this connection helps validate orexin KO mice as an excellent model of human narcolepsy and provides an opportunity to better understand the mechanisms that trigger cataplexy. PMID:19362119

  16. Psychosocial Characteristics of Children with Central Disorders of Hypersomnolence Versus Matched Healthy Children

    PubMed Central

    Avis, Kristin T.; Shen, Jiabin; Weaver, Patrick; Schwebel, David C.

    2015-01-01

    Background: Hypersomnia of central origin from narcolepsy or idiopathic hypersomnia (IHS) is characterized by pathological levels of excessive daytime sleepiness (EDS). Central hypersomnia has historically been underdiagnosed and poorly understood, especially with respect to its impact on daytime functioning and quality of life in children. Objective: Describe the psychosocial adjustment of children treated for narcolepsy or IHS on school performance, quality of life, and physical/extracurricular activities. Methods: Using a matched case control design, we compared child self- and parent-reported data from thirty-three 8- to 16-year-olds with an established diagnosis of narcolepsy or IHS, according to ICSD-2 criteria, to that of 33 healthy children matched by age, race/ethnicity, gender, and household income. Assessments evaluated academic performance, quality of life and wellness, sleepiness, and participation in extracurricular activities. Results: Compared to healthy controls, children with central hypersomnia had poorer daytime functioning in multiple domains. Children with hypersomnia missed more days of school and had lower grades than healthy controls. Children with hypersomnia had poorer quality of life by both parent and child report. Children with hypersomnia were significantly sleepier, had higher BMI, and were more likely to report a history of recent injury. Finally, children with hypersomnia engaged in fewer after-school activities than healthy controls. Conclusions: A range of significant psychosocial consequences are reported in children with hypersomnia even after a diagnosis has been made and treatments initiated. Health care professionals should be mindful of the psychosocial problems that may present in children with hypersomnia over the course of treatment. Citation: Avis KT, Shen J, Weaver P, Schwebel DC. Psychosocial characteristics of children with central disorders of hypersomnolence versus matched healthy children. J Clin Sleep Med 2015

  17. Examining the Frequency of Stimulant Misuse among Patients with Primary Disorders of Hypersomnolence: A Retrospective Cohort Study

    PubMed Central

    Mantyh, William G.; Auger, R. Robert; Morgenthaler, Timothy I.; Silber, Michael H.; Moore, Wendy R.

    2016-01-01

    Study Objectives: Narcolepsy and idiopathic hypersomnia are commonly treated by sleep specialists and encountered by other medical providers. Although pharmacotherapy with modafinil and traditional stimulants is considered the mainstay of treatment, physicians are often uncomfortable with their prescription because of concerns regarding misuse. The goal of this study was to assess the frequency of stimulant misuse in this population. Methods: A retrospective cohort study was performed evaluating patients 18 years and older diagnosed with narcolepsy with and without cataplexy and idiopathic hypersomnia with and without long sleep between 2003–2008. Patients were included if they obtained stimulant prescriptions from and had at least one follow-up visit subsequent to initial diagnosis at our center. Stimulant misuse was defined by multiple prescription sources or early refill requests, which are systematically entered into the record by nursing staff. Results: A total of 105 patients met inclusion criteria for the study; 45 (42%) were male. Mean age at multiple sleep latency test was 42 (± 16). Twelve (11%) patients had a history of illicit substance misuse, and one (1%) patient demonstrated previous stimulant misuse. Fifty-seven (54%) patients carried psychiatric diagnoses, 88% of whom reported depression. Median duration of monitored stimulant therapy was 26 months (range 1–250). None of the 105 patients was found to have evidence of stimulant misuse. Conclusion: This study suggests that the frequency of stimulant misuse in patients with narcolepsy and idiopathic hypersomnia is extremely low. Concerns regarding drug misuse should not leverage decisions to provide long-term therapy. Citation: Mantyh WG, Auger RR, Morgenthaler TI, Silber MH, Moore WR. Examining the frequency of stimulant misuse among patients with primary disorders of hypersomnolence: a retrospective cohort study. J Clin Sleep Med 2016;12(5):659–662. PMID:26943713

  18. A standardized test to document cataplexy.

    PubMed

    Vandi, Stefano; Pizza, Fabio; Antelmi, Elena; Neccia, Giulia; Iloti, Martina; Mazzoni, Alice; Avoni, Patrizia; Plazzi, Giuseppe

    2017-10-09

    Cataplexy is the pathognomonic symptom of narcolepsy type 1 (NT1). Since it is considered difficult to be directly observed or documented by clinicians, its diagnosis relies mainly on history taking. Our study aimed at testing the feasibility of a standardized video recording procedure under emotional stimulation to document cataplexy in the diagnostic work-up of suspected hypersomnia of central origin. Two-hundred-eight consecutive patients underwent the diagnostic work-up and reached the final diagnosis of NT1 (n = 133), idiopathic hypersomnia or narcolepsy type 2 (IH/NT2 group, n = 33), or subjective excessive daytime sleepiness (sEDS group, n = 42). All subjects underwent a standardized video recording procedure while watching funny movies selected according to individual preferences, and a technician blind to clinical features reviewed the recordings to identify hypotonic phenomena that were finally confirmed by patients. The video recording under emotional stimulation captured hypotonic phenomena in 72.2%, 9.1% and 4.8% of NT1, IH/NT2, and sEDS subjects (p < 0.0001), respectively. When tested against CSF hypocretin deficiency, the documentation of a hypotonic episode at the test showed an area under the ROC curve of 0.823 ± 0.033 (p < 0.0001). NT1 patients under anticataplectic medications showed less frequently hypotonic episodes than untreated ones (48.0% vs 77.8%, p = 0.003). A standardized video recording procedure under emotional stimulation can help in the characterization of suspected hypersomnia of central origin. Further multi-center studies are warranted to extend the present findings and integrate a shared procedure for the laboratory work-up of narcolepsy. Copyright © 2017 Elsevier B.V. All rights reserved.

  19. Driving Safety and Fitness to Drive in Sleep Disorders.

    PubMed

    Tippin, Jon; Dyken, Mark Eric

    2017-08-01

    Driving an automobile while sleepy increases the risk of crash-related injury and death. Neurologists see patients with sleepiness due to obstructive sleep apnea, narcolepsy, and a wide variety of neurologic disorders. When addressing fitness to drive, the physician must weigh patient and societal health risks and regional legal mandates. The Driver Fitness Medical Guidelines published by the National Highway Traffic Safety Administration (NHTSA) and the American Association of Motor Vehicle Administrators (AAMVA) provide assistance to clinicians. Drivers with obstructive sleep apnea may continue to drive if they have no excessive daytime sleepiness and their apnea-hypopnea index is less than 20 per hour. Those with excessive daytime sleepiness or an apnea-hypopnea index of 20 per hour or more may not drive until their condition is effectively treated. Drivers with sleep disorders amenable to pharmaceutical treatment (eg, narcolepsy) may resume driving as long as the therapy has eliminated excessive daytime sleepiness. Following these guidelines, documenting compliance to recommended therapy, and using the Epworth Sleepiness Scale to assess subjective sleepiness can be helpful in determining patients' fitness to drive.

  20. Cognition-Enhancing Drugs and Their Appropriateness for Aviation and Ground Troops: A Meta-Analysis

    DTIC Science & Technology

    2010-12-01

    individuals is not approved by the Food and Drug Administration (FDA). Current indications include narcolepsy, obstructive sleep apnea/hypopnea syndrome, and...to caffeine and perceived effects of caffeine in moderate and high regular caffeine consumers . Psychopharmacology. 190: 469-477...J. 2004. The effect of caffeinated tube food on cognitive performance during fatigue/circadian desynchronosis. Brooks City-Base, TX

  1. Serotonin neurons in the dorsal raphe mediate the anticataplectic action of orexin neurons by reducing amygdala activity.

    PubMed

    Hasegawa, Emi; Maejima, Takashi; Yoshida, Takayuki; Masseck, Olivia A; Herlitze, Stefan; Yoshioka, Mitsuhiro; Sakurai, Takeshi; Mieda, Michihiro

    2017-04-25

    Narcolepsy is a sleep disorder caused by the loss of orexin (hypocretin)-producing neurons and marked by excessive daytime sleepiness and a sudden weakening of muscle tone, or cataplexy, often triggered by strong emotions. In a mouse model for narcolepsy, we previously demonstrated that serotonin neurons of the dorsal raphe nucleus (DRN) mediate the suppression of cataplexy-like episodes (CLEs) by orexin neurons. Using an optogenetic tool, in this paper we show that the acute activation of DRN serotonin neuron terminals in the amygdala, but not in nuclei involved in regulating rapid eye-movement sleep and atonia, suppressed CLEs. Not only did stimulating serotonin nerve terminals reduce amygdala activity, but the chemogenetic inhibition of the amygdala using designer receptors exclusively activated by designer drugs also drastically decreased CLEs, whereas chemogenetic activation increased them. Moreover, the optogenetic inhibition of serotonin nerve terminals in the amygdala blocked the anticataplectic effects of orexin signaling in DRN serotonin neurons. Taken together, the results suggest that DRN serotonin neurons, as a downstream target of orexin neurons, inhibit cataplexy by reducing the activity of amygdala as a center for emotional processing.

  2. Orexin research: patent news from 2016.

    PubMed

    Boss, Christoph; Roch, Catherine

    2017-10-01

    The orexin system consists of two G-protein-coupled receptors, orexin 1 and orexin 2 and two endogenous ligands, orexin A and orexin B . It is evolutionarily highly conserved. It is involved in the promotion of wakefulness as well as in anxiety and addictive disorders. In addition, its activation via the Ox1 receptor triggers apoptosis in several cancer cell lines. Dual orexin receptor antagonists are successfully used to treat primary insomnia. The major open questions are now related to the clinical validation of Ox1 selective antagonists. A strong rationale exists for orexin agonism in the treatment of narcolepsy with cataplexy. Areas covered: The patent applications from Thomson Reuters Integrity Database added in 2016 are summarized and discussed together with the most important findings published in the scientific literature. Expert opinion: The large number of patents shows the continuing interest in the orexin receptors as targets. The structural scope covered is narrow. Questions about novelty and inventiveness are evident. The additional information published on X-ray structures on both orexin receptors opens new ways of optimizing antagonists. It might also influence the efforts in the identification of orexin receptor agonists. Being potential treatments for narcolepsy with cataplexy.

  3. REM Sleep at its Core – Circuits, Neurotransmitters, and Pathophysiology

    PubMed Central

    Fraigne, Jimmy J.; Torontali, Zoltan A.; Snow, Matthew B.; Peever, John H.

    2015-01-01

    Rapid eye movement (REM) sleep is generated and maintained by the interaction of a variety of neurotransmitter systems in the brainstem, forebrain, and hypothalamus. Within these circuits lies a core region that is active during REM sleep, known as the subcoeruleus nucleus (SubC) or sublaterodorsal nucleus. It is hypothesized that glutamatergic SubC neurons regulate REM sleep and its defining features such as muscle paralysis and cortical activation. REM sleep paralysis is initiated when glutamatergic SubC cells activate neurons in the ventral medial medulla, which causes release of GABA and glycine onto skeletal motoneurons. REM sleep timing is controlled by activity of GABAergic neurons in the ventrolateral periaqueductal gray and dorsal paragigantocellular reticular nucleus as well as melanin-concentrating hormone neurons in the hypothalamus and cholinergic cells in the laterodorsal and pedunculo-pontine tegmentum in the brainstem. Determining how these circuits interact with the SubC is important because breakdown in their communication is hypothesized to underlie narcolepsy/cataplexy and REM sleep behavior disorder (RBD). This review synthesizes our current understanding of mechanisms generating healthy REM sleep and how dysfunction of these circuits contributes to common REM sleep disorders such as cataplexy/narcolepsy and RBD. PMID:26074874

  4. Serotonin neurons in the dorsal raphe mediate the anticataplectic action of orexin neurons by reducing amygdala activity

    PubMed Central

    Hasegawa, Emi; Maejima, Takashi; Yoshida, Takayuki; Herlitze, Stefan; Yoshioka, Mitsuhiro; Sakurai, Takeshi; Mieda, Michihiro

    2017-01-01

    Narcolepsy is a sleep disorder caused by the loss of orexin (hypocretin)-producing neurons and marked by excessive daytime sleepiness and a sudden weakening of muscle tone, or cataplexy, often triggered by strong emotions. In a mouse model for narcolepsy, we previously demonstrated that serotonin neurons of the dorsal raphe nucleus (DRN) mediate the suppression of cataplexy-like episodes (CLEs) by orexin neurons. Using an optogenetic tool, in this paper we show that the acute activation of DRN serotonin neuron terminals in the amygdala, but not in nuclei involved in regulating rapid eye-movement sleep and atonia, suppressed CLEs. Not only did stimulating serotonin nerve terminals reduce amygdala activity, but the chemogenetic inhibition of the amygdala using designer receptors exclusively activated by designer drugs also drastically decreased CLEs, whereas chemogenetic activation increased them. Moreover, the optogenetic inhibition of serotonin nerve terminals in the amygdala blocked the anticataplectic effects of orexin signaling in DRN serotonin neurons. Taken together, the results suggest that DRN serotonin neurons, as a downstream target of orexin neurons, inhibit cataplexy by reducing the activity of amygdala as a center for emotional processing. PMID:28396432

  5. The Use of Modafinil in Operational Settings: Individual Difference Implications

    DTIC Science & Technology

    2000-03-01

    them "skylark". On another treatment of narcolepsia and idiopathic hypersomnia . hand, some people are eveningness type, they have the The wakening...showed that modafinil acts as idiopathic hypersomnia and narcolepsy with modafinil. an agonist of ct1-adrenergic post-synaptic receptors (4). Prog...of During sustained operations (SUSOPS) it is impossible narcolepsia and idiopatic hypersomnia . It is synthetized for the soldier to sleep, sometimes

  6. Idiopathic hypersomnia.

    PubMed

    Billiard, Michel; Sonka, Karel

    2016-10-01

    Idiopathic hypersomnia continues to evolve from the concept of "sleep drunkenness" introduced by Bedrich Roth in Prague in 1956 and the description of idiopathic hypersomnia with two forms, polysymptomatic and monosymptomatic, by the same Bedrich Roth in 1976. The diagnostic criteria of idiopathic hypersomnia have varied with the successive revisions of the International classifications of sleep disorders, including the recent 3rd edition. No epidemiological studies have been conducted so far. Disease onset occurs most often during adolescence or young adulthood. A familial background is often present but rigorous studies are still lacking. The key manifestation is hypersomnolence. It is often accompanied by sleep of long duration and debilitating sleep inertia. Polysomnography (PSG) followed by a multiple sleep latency test (MSLT) is mandatory, as well as a 24 h PSG or a 2-wk actigraphy in association with a sleep log to ensure a total 24-h sleep time longer than or equal to 66O minutes, when the mean sleep latency on the MSLT is longer than 8 min. Yet, MSLT is neither sensitive nor specific and the polysomnographic diagnostic criteria require continuous readjustment and biologic markers are still lacking. Idiopathic hypersomnia is most often a chronic condition though spontaneous remission may occur. The condition is disabling, sometimes even more so than narcolepsy type 1 or 2. Based on neurochemical, genetic and immunological analyses as well as on exploration of the homeostatic and circadian processes of sleep, various pathophysiological hypotheses have been proposed. Differential diagnosis involves a number of diseases and it is not yet clear whether idiopathic hypersomnia and narcolepsy type 2 are not the same condition. Until now, the treatment of idiopathic hypersomnia has mirrored that of the sleepiness of narcolepsy type 1 or 2. The first randomized, double-blind, placebo-controlled trials of modafinil have just been published, as well as a double

  7. Prevalence and Phenotype of Sleep Disorders in 60 Adults With Prader-Willi Syndrome.

    PubMed

    Ghergan, Adelina; Coupaye, Muriel; Leu-Semenescu, Smaranda; Attali, Valérie; Oppert, Jean-Michel; Arnulf, Isabelle; Poitou, Christine; Redolfi, Stefania

    2017-12-01

    Excessive sleepiness is a common symptom in Prader-Willi syndrome (PWS), and it negatively impacts the quality of life. Obstructive sleep apnea and narcolepsy phenotypes have been reported in PWS. We characterized sleep disorders in a large cohort of adults with PWS. All consecutive patients with genetically confirmed PWS unselected for sleep-related symptoms, underwent a clinical interview, polysomnography, and multiple sleep latency tests (MSLT, n = 60), followed by long-term (24 hours) polysomnography (n = 22/60). Among 60 adults evaluated (57% female, aged 25 ± 10 years, body mass index: 39 ± 12 kg/m2), 67% reported excessive sleepiness. According to the sleep study results, 43% had a previously unrecognized hypersomnia disorder, 15% had an isolated sleep breathing disorder, 12% had combined hypersomnia disorder and untreated breathing sleep disorder, and only 30% had normal sleep. Isolated hypersomnia disorder included narcolepsy in 35% (type 1, n = 1, and type 2, n = 8), hypersomnia in 12% (total sleep time >11 hours, n = 2, and MSLT <8 minutes, n = 1), and borderline phenotype in 53% (≥2 sleep onset in REM periods and MSLT >8 minutes, n = 10, and 8 minutes < MSLT < 10 minutes, n = 4). Sleep breathing disorders, isolated and combined, included obstructive sleep apnea (n = 14, already treated in seven), sleep hypoxemia (n = 1) and previously undiagnosed hypoventilation (n = 5). Modafinil was taken by 16 patients (well tolerated in 10), resulting in improved sleepiness over a mean 5-year follow-up period. Sleepiness affects more than half of adult patients with PWS, with a variety of hypersomnia disorder (narcolepsy, hypersomnia, and borderline phenotypes) and breathing sleep disorders. Earlier diagnosis and management of sleep disorders may improve sleepiness, cognition, and behavior in these patients. © Sleep Research Society 2017. Published by Oxford University Press [on behalf of the Sleep Research Society]. All rights reserved. For permissions, please

  8. Neurosarcoidosis associated with hypersomnolence treated with corticosteroids and brain irradiation

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Rubinstein, I.; Gray, T.A.; Moldofsky, H.

    1988-07-01

    Narcoleptic features developed in a young man with CNS sarcoidosis. This was associated with a structural lesion in the hypothalamus as demonstrated on CT scans of the head. The diagnosis of narcolepsy was established by compatible clinical history and the Multiple Sleep Latency Test. Treatment with high-dose corticosteroids was ineffective, but when the low-dose, whole-brain irradiation was added, complete resolution of the narcoleptic features ensued.

  9. Virtual-reality-based system for controlled study of cataplexy

    NASA Astrophysics Data System (ADS)

    Augustine, Kurt E.; Cameron, Bruce M.; Camp, Jon J.; Krahn, Lois E.; Robb, Richard A.

    2002-05-01

    Cataplexy is a sudden loss of voluntary muscle control experienced by narcolepsy patients. It is usually triggered by strong, spontaneous emotions and is more common in times of stress. The Sleep Disorders Unit and the Biomedical Imaging Resource at Mayo Clinic are developing interactive display technology for reliably inducing cataplexy during clinical monitoring. The project is referred to as the Cataplexy/Narcolepsy Activation Program, or CatNAP. We have developed an automobile driving simulation that introduces humorous, surprising, and stress-inducing events and objects as the patient attempts to navigate a vehicle through a virtual town. The patient wears a head-mounted display and controls the vehicle via a driving simulator steering wheel and pedal cluster. As the patient attempts to drive through the town, various objects, sounds or conditions occur that distract, startle, frustrate or amuse. These responses may trigger a cataplectic episode, which can then be clinically evaluated. We believe CatNAP is a novel and innovative example of the effective application of virtual reality technology to study an important clinical problem that has resisted previous approaches. An evaluation phase with volunteer patients previously diagnosed with cataplexy has been completed. The prototype system is being prepared for a full clinical study.

  10. REM sleep enhancement and behavioral cataplexy following orexin (hypocretin)-II receptor antisense perfusion in the pontine reticular formation.

    PubMed

    Thakkar, M M; Ramesh, V; Cape, E G; Winston, S; Strecker, R E; McCarley, R W

    1999-01-01

    Orexin (hypocretin)-containing neurons of the hypothalamus project to brainstem sites that are involved in the neural control of REM sleep, including the locus coeruleus, the dorsal raphe nucleus, the cholinergic zone of the mesopontine tegmentum, and the pontine reticular formation (PRF). Orexin knockout mice exhibit narcolepsy/cataplexy, and a mutant and defective gene for the orexin type II receptor is present in dogs with an inherited form of narcolepsy/cataplexy. However, the physiological systems mediating these effects have not been described. We reasoned that, since the effector neurons for the majority of REM sleep signs, including muscle atonia, were located in the PRF, this region was likely implicated in the production of these orexin-related abnormalities. To test this possibility, we used microdialysis perfusion of orexin type II receptor antisense in the PRF of rats. Ten to 24 hours after antisense perfusion, REM sleep increased two- to three-fold during both the light period (quiescent phase) and the dark period (active phase), and infrared video showed episodes of behavioral cataplexy. Moreover, preliminary data indicated no REM-related effects following perfusion with nonsense DNA, or when perfusion sites were outside the PRF. More work is needed to provide precise localization of the most effective site of orexin-induced inhibition of REM sleep phenomena.

  11. The utility of a 5(th) nap in multiple sleep latency test.

    PubMed

    Muza, Rexford; Lykouras, Dimosthenis; Rees, Kate

    2016-02-01

    This is the first study that aimed to look specifically at the utility of the 5(th) nap in the multiple sleep latency test (MSLT), a test used to assist in the diagnosis of narcolepsy. Data was retrospectively collected from the Sleep Disorders Centre of a Tertiary Hospital on patients that had a 5(th) nap during their MSLT from the 08(th) November 2011 to 12(th) November 2014. Fifty-three patients had a 5(th) nap performed out of 378 MSLT studies. In 16% of cases a diagnosis of narcolepsy was given directly due to the inclusion of the 5(th) nap on the MSLT. Here a 5(th) nap allowed diagnostic criteria of mean sleep latency <8 minutes and >2 SOREMPS to be met. In 53% of cases the mean sleep latency increased due to 5(th) nap inclusion; the mean sleep latency of the first four naps was 5.6 vs. 6.7 after inclusion of the 5(th) nap. The 5(th) nap is not often performed within the MSLT studies. Our study shows that only a few patients may benefit from a 5(th) nap opportunity which also led to increase of the mean sleep latency at the expense of extra time, cost, labour and increased patient anxiety.

  12. Orexin receptor 2 expression in the posterior hypothalamus rescues sleepiness in narcoleptic mice.

    PubMed

    Mochizuki, Takatoshi; Arrigoni, Elda; Marcus, Jacob N; Clark, Erika L; Yamamoto, Mihoko; Honer, Michael; Borroni, Edilio; Lowell, Bradford B; Elmquist, Joel K; Scammell, Thomas E

    2011-03-15

    Narcolepsy is caused by a loss of orexin/hypocretin signaling, resulting in chronic sleepiness, fragmented non-rapid eye movement sleep, and cataplexy. To identify the neuronal circuits underlying narcolepsy, we produced a mouse model in which a loxP-flanked gene cassette disrupts production of the orexin receptor type 2 (OX2R; also known as HCRTR2), but normal OX2R expression can be restored by Cre recombinase. Mice lacking OX2R signaling had poor maintenance of wakefulness indicative of sleepiness and fragmented sleep and lacked any electrophysiological response to orexin-A in the wake-promoting neurons of the tuberomammillary nucleus. These defects were completely recovered by crossing them with mice that express Cre in the female germline, thus globally deleting the transcription-disrupter cassette. Then, by using an adeno-associated viral vector coding for Cre recombinase, we found that focal restoration of OX2R in neurons of the tuberomammillary nucleus and adjacent parts of the posterior hypothalamus completely rescued the sleepiness of these mice, but their fragmented sleep was unimproved. These observations demonstrate that the tuberomammillary region plays an essential role in the wake-promoting effects of orexins, but orexins must stabilize sleep through other targets.

  13. Persistence of deep-tendon reflexes during partial cataplexy.

    PubMed

    Barateau, Lucie; Pizza, Fabio; Lopez, Régis; Antelmi, Elena; Plazzi, Giuseppe; Dauvilliers, Yves

    2018-05-01

    Deep-tendon reflexes are abolished during generalized cataplexy, but whether this is the case in partial cataplexy currently remains unknown. Partial cataplexy may mimic other neurologic/psychiatric phenomena, and knowledge of the reflexes status may provide information for differential diagnosis. We assessed whether deep-tendon reflexes are persistent during partial cataplexy. Five drug-free patients with typical diagnoses of narcolepsy and clear-cut partial cataplexy were diagnosed in Reference Narcolepsy Centers in France and Italy. Biceps and patellar reflexes were elicited by physicians in charge and video-documented during cataplexy. Reflexes were assessed several times for each patient in different conditions and for various localizations of cataplexy. The absence of tendon reflexes and complete loss of muscle tone during generalized cataplexy was confirmed, but the persistence of those reflexes during several partial cataplectic attacks at different ages, gender, localization of cataplexy (upper limbs, face) and reflexes (biceps, patellar) in drug-naive or withdrawal conditions was documented. The persistence of tendon reflexes during several partial cataplexy episodes contrasts with their absence during generalized cataplexy. This discovery has clinical implications: the persistence of tendon reflexes does not rule out cataplexy diagnosis for partial attacks, whereas their transient abolishment or persistence during generalized attacks indicates cataplexy or pseudocataplexy, respectively. Copyright © 2018. Published by Elsevier B.V.

  14. Which diagnostic findings in disorders with excessive daytime sleepiness are really helpful? A retrospective study.

    PubMed

    Kretzschmar, Ute; Werth, Esther; Sturzenegger, Christian; Khatami, Ramin; Bassetti, Claudio L; Baumann, Christian R

    2016-06-01

    Due to extensive clinical and electrophysiological overlaps, the correct diagnosis of disorders with excessive daytime sleepiness is often challenging. The aim of this study was to provide diagnostic measures that help discriminating such disorders, and to identify parameters, which don't. In this single-center study, we retrospectively identified consecutive treatment-naïve patients who suffered from excessive daytime sleepiness, and analyzed clinical and electrophysiological measures in those patients in whom a doubtless final diagnosis could be made. Of 588 patients, 287 reported subjective excessive daytime sleepiness. Obstructive sleep apnea is the only disorder that could be identified by polysomnography alone. The diagnosis of insufficient sleep syndrome relies on actigraphy as patients underestimate their sleep need and the disorder shares several clinical and electrophysiological properties with both narcolepsy type 1 and idiopathic hypersomnia. Sleep stage sequencing on MSLT appears helpful to discriminate between insufficient sleep syndrome and narcolepsy. Sleep inertia is a strong indicator for idiopathic hypersomnia. There are no distinctive electrophysiological findings for the diagnosis of restless legs syndrome. Altogether, EDS disorders are common in neurological sleep laboratories, but usually cannot be diagnosed based on PSG and MSLT findings alone. The diagnostic value of actigraphy recordings can hardly be overestimated. © 2016 European Sleep Research Society.

  15. Primary hypersomnias of central origin.

    PubMed

    Frenette, Eric; Kushida, Clete A

    2009-09-01

    Hypersomnia is a frequently encountered symptom in clinical practice. The cardinal manifestation is inappropriate daytime sleepiness, common to all types of hypersomnias. Hypersomnias of central origin are a rare cause of excessive daytime sleepiness, much rarer than the hypersomnia related to other pathologies, such as sleep-disordered breathing. Narcolepsy, with or without cataplexy, remains the most well studied of the primary hypersomnias. Although recognized more than a century ago, it was not until the end of the 20th century that major breakthroughs led to a better understanding of the disease, with hope of more specific therapies. The authors review the major aspects of this disorder, including the newer treatment modalities. Idiopathic hypersomnia is also part of the primary hypersomnias. Although difficult to diagnose, certain peculiarities stand out to help us differentiate it from the more commonly seen narcolepsy. The recurrent hypersomnias, particularly the Kleine-Levin syndrome, will be discussed. This rare disorder has been studied more closely in the last few years with abundant epidemiologic data assembled through literature and worldwide case reviews. Understanding the primary central hypersomnias warrants a thorough look from the original description, as well as a peek at the future, while more efficacious diagnostic and therapeutic interventions are currently being developed. Thieme Medical Publishers.

  16. Ictal SPECT in patients with rapid eye movement sleep behaviour disorder.

    PubMed

    Mayer, Geert; Bitterlich, Marion; Kuwert, Torsten; Ritt, Philipp; Stefan, Hermann

    2015-05-01

    Rapid eye movement sleep behaviour disorder is a rapid eye movement parasomnia clinically characterized by acting out dreams due to disinhibition of muscle tone in rapid eye movement sleep. Up to 80-90% of the patients with rapid eye movement sleep behaviour disorder develop neurodegenerative disorders within 10-15 years after symptom onset. The disorder is reported in 45-60% of all narcoleptic patients. Whether rapid eye movement sleep behaviour disorder is also a predictor for neurodegeneration in narcolepsy is not known. Although the pathophysiology causing the disinhibition of muscle tone in rapid eye movement sleep behaviour disorder has been studied extensively in animals, little is known about the mechanisms in humans. Most of the human data are from imaging or post-mortem studies. Recent studies show altered functional connectivity between substantia nigra and striatum in patients with rapid eye movement sleep behaviour disorder. We were interested to study which regions are activated in rapid eye movement sleep behaviour disorder during actual episodes by performing ictal single photon emission tomography. We studied one patient with idiopathic rapid eye movement sleep behaviour disorder, one with Parkinson's disease and rapid eye movement sleep behaviour disorder, and two patients with narcolepsy and rapid eye movement sleep behaviour disorder. All patients underwent extended video polysomnography. The tracer was injected after at least 10 s of consecutive rapid eye movement sleep and 10 s of disinhibited muscle tone accompanied by movements registered by an experienced sleep technician. Ictal single photon emission tomography displayed the same activation in the bilateral premotor areas, the interhemispheric cleft, the periaqueductal area, the dorsal and ventral pons and the anterior lobe of the cerebellum in all patients. Our study shows that in patients with Parkinson's disease and rapid eye movement sleep behaviour disorder-in contrast to wakefulness

  17. From wakefulness to excessive sleepiness: what we know and still need to know.

    PubMed

    Ohayon, Maurice Moyses

    2008-04-01

    The epidemiological study of hypersomnia symptoms is still in its infancy; most epidemiological surveys on this topic were published in the last decade. More than two dozen representative community studies can be found. These studies assessed two aspects of hypersomnia: excessive quantity of sleep and sleep propensity during wakefulness excessive daytime sleepiness. The prevalence of excessive quantity of sleep when referring to the subjective evaluation of sleep duration is around 4% of the population. Excessive daytime sleepiness has been mostly investigated in terms of frequency or severity; duration of the symptom has rarely been investigated. Excessive daytime sleepiness occurring at least 3 days per week has been reported in between 4% and 20.6% of the population, while severe excessive daytime sleepiness was reported at 5%. In most studies, men and women are equally affected. In the International Classification of Sleep Disorders, hypersomnia symptoms are the essential feature of three disorders: insufficient sleep syndrome, hypersomnia (idiopathic, recurrent or posttraumatic) and narcolepsy. Insufficient sleep syndrome and hypersomnia diagnoses are poorly documented. The co-occurrence of insufficient sleep and excessive daytime sleepiness has been explored in some studies and prevalence has been found in around 8% of the general population. However, these subjects often have other conditions such as insomnia, depression or sleep apnea. Therefore, the prevalence of insufficient sleep syndrome is more likely to be between 1% and 4% of the population. Idiopathic hypersomnia would be rare in the general population with prevalence, around 0.3%. Narcolepsy has been more extensively studied, with a prevalence around 0.045% in the general population. Genetic epidemiological studies of narcolepsy have shown that between 1.5% and 20.8% of narcoleptic individuals have at least one family member with the disease. The large variation is mostly due to the method used to

  18. From wakefulness to excessive sleepiness: what we know and still need to know

    PubMed Central

    Ohayon, Maurice M.

    2008-01-01

    The epidemiological study of hypersomnia symptoms is still in its infancy; most epidemiological surveys on this topic were published in the last decade. More than two dozen representative community studies can be found. These studies assessed two aspects of hypersomnia: excessive quantity of sleep and sleep propensity during wakefulness (excessive daytime sleepiness). The prevalence of excessive quantity of sleep when referring to the subjective evaluation of sleep duration is around 4% of the population. Excessive daytime sleepiness (EDS) has been mostly investigated in terms of frequency or severity; duration of the symptom has rarely been investigated. EDS occurring at least 3 days per week has been reported in between 4% and 20.6% of the population, while severe EDS was reported at 5%. In most studies men and women are equally affected. In the International Classification of Sleep Disorders, hypersomnia symptoms are the essential feature of 3 disorders: insufficient sleep syndrome, hypersomnia (idiopathic, recurrent or posttraumatic) and narcolepsy. Insufficient sleep syndrome and hypersomnia diagnoses are poorly documented. The co-occurrence of insufficient sleep and EDS has been explored in some studies and prevalence has been found in around 8% of the general population. However, these subjects often have other conditions such as insomnia, depression or sleep apnea. Therefore, the prevalence of insufficient sleep syndrome is more likely to be between 1% and 4% of the population. Idiopathic hypersomnia would be rare in the general population with prevalence, around 0.3%. Narcolepsy has been more extensively studied, with a prevalence around 0.045% in the general population. Genetic epidemiological studies of narcolepsy have shown that between 1.5% and 20.8% of narcoleptic individuals have at least one family member with the disease. The large variation is mostly due to the method used to collect the information on the family members; systematic investigation

  19. The distinguishing motor features of cataplexy: a study from video-recorded attacks.

    PubMed

    Pizza, Fabio; Antelmi, Elena; Vandi, Stefano; Meletti, Stefano; Erro, Roberto; Baumann, Christian R; Bhatia, Kailash P; Dauvilliers, Yves; Edwards, Mark J; Iranzo, Alex; Overeem, Sebastiaan; Tinazzi, Michele; Liguori, Rocco; Plazzi, Giuseppe

    2018-05-01

    To describe the motor pattern of cataplexy and to determine its phenomenological differences from pseudocataplexy in the differential diagnosis of episodic falls. We selected 30 video-recorded cataplexy and 21 pseudocataplexy attacks in 17 and 10 patients evaluated for suspected narcolepsy and with final diagnosis of narcolepsy type 1 and conversion disorder, respectively, together with self-reported attacks features, and asked expert neurologists to blindly evaluate the motor features of the attacks. Video documented and self-reported attack features of cataplexy and pseudocataplexy were contrasted. Video-recorded cataplexy can be positively differentiated from pseudocataplexy by the occurrence of facial hypotonia (ptosis, mouth opening, tongue protrusion) intermingled by jerks and grimaces abruptly interrupting laughter behavior (i.e. smile, facial expression) and postural control (head drops, trunk fall) under clear emotional trigger. Facial involvement is present in both partial and generalized cataplexy. Conversely, generalized pseudocataplexy is associated with persistence of deep tendon reflexes during the attack. Self-reported features confirmed the important role of positive emotions (laughter, telling a joke) in triggering the attacks, as well as the more frequent occurrence of partial body involvement in cataplexy compared with pseudocataplexy. Cataplexy is characterized by abrupt facial involvement during laughter behavior. Video recording of suspected cataplexy attacks allows the identification of positive clinical signs useful for diagnosis and, possibly in the future, for severity assessment.

  20. Behavioral Sleep Medicine Services for Hypersomnia Disorders: A Survey Study.

    PubMed

    Neikrug, Ariel B; Crawford, Megan R; Ong, Jason C

    2017-01-01

    Patients with hypersomnia disorders (HD) suffer from debilitating symptoms that result in reduced functioning, depression, anxiety, and overall worse quality of life. Little is known about the need and desire of this population to utilize behavioral sleep medicine (BSM) interventions that focus on psychosocial functioning and quality of life, and there have been limited attempts to develop such interventions. The purpose of this survey study was to gather patient-centered data on engagement in pharmacological and nonpharmacological interventions, the psychosocial impact of HD symptoms on quality of life and mental health, and potential interest in BSM services, such as cognitive behavioral therapy, mindfulness or yoga, and support groups. We obtained responses from 371 individuals with HD (65.2% narcolepsy and 34.8% idiopathic hypersomnia) to an Internet-based survey. Overall, HD patients reported engagement in pharmacological and nonpharmacological interventions, with narcolepsy patients reporting more perceived effectiveness than those with idiopathic hypersomnia. In addition, HD patients reported a strong negative impact on psychosocial functioning, with elevations in depression and anxiety symptoms along with significant impact on functioning and quality of life. The majority (71.7-85.5%) voiced at least some interest in BSM services. These data suggest that there is substantial interest and need for BSM services that focus on assessment and treatment of psychosocial functioning related to HD.

  1. A new DRB1*1202 allele (DRB1*12022) found in association with DQA1*0102 and DQB1*0602 in two Black narcoleptic subjects

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Behar, E.; Grumet, F.C.; Lin, X.

    1995-01-01

    DQB1*0602 is a better genetic marker than DR2 for narcolepsy susceptibility across all ethnic groups; for instance, only 75% of African American narcoleptics are DR2+ compared with 96% DQB1*0602+. We studied DRB1 genes of DR2- but DQB1*0602+ African American patients with cataplexy and observed two with an unusual DR12, DQA1*0102, DQB1*0602 haplotype; a new allelic variant of DRB1*1202 has been designated DRB*12022. 8 refs.

  2. Emerging therapeutics in sleep.

    PubMed

    Saper, Clifford B; Scammell, Thomas E

    2013-09-01

    The development of new therapeutics for sleep disorders is increasingly dependent upon understanding the basic brain circuitry that underlies sleep-wake regulation, and how it may be pharmacologically manipulated. In this review, we consider the pathophysiological basis of major sleep disorders that often are seen by neurologists, including excessive daytime sleepiness, insomnia, narcolepsy, rapid eye movement sleep behavior disorder, and restless legs syndrome, as well as circadian disorders, and we review the current and potential future therapeutic approaches. Copyright © 2013 American Neurological Association.

  3. The Xyrem risk management program.

    PubMed

    Fuller, David E; Hornfeldt, Carl S; Kelloway, Judy S; Stahl, Pamela J; Anderson, Todd F

    2004-01-01

    Sodium oxybate, also known as gamma-hydroxybutyric acid (GHB), was discovered in 1960 and has been described both as a therapeutic agent with high medical value and, more recently, a substance of abuse. The naturally occurring form of this drug is found in various body tissues but has been studied most extensively in the CNS where its possible function as a neurotransmitter continues to be studied. Sodium oxybate has been approved in different countries for such varied uses as general anaesthesia, the treatment of alcohol withdrawal and addiction, and, most recently, cataplexy associated with narcolepsy. During the 1980s, easy access to GHB-containing products led to various unapproved uses, including weight loss, bodybuilding and the treatment of sleeplessness, sometimes with serious long-term effects. The availability of these unapproved and unregulated forms of the drug led to GHB and its analogues being popularised as substances of abuse and subsequent notoriety as agents used in drug-facilitated sexual assault, or 'date rape', eventually leading to the prohibition of GHB sales in the US. Legal efforts to control the sale and distribution of GHB and its analogues nearly prevented the clinical development of sodium oxybate for narcolepsy in the US. However, following extensive discussions with a variety of interested parties, a satisfactory solution was devised, including legislative action and the development of the Xyrem Risk Management Program. Amendments to the US Controlled Substances Act made GHB a schedule I drug, but also contained provisions that allow US FDA-approved products to be placed under schedule III. This unique, bifurcated schedule for sodium oxybate/GHB allowed the clinical development of sodium oxybate to proceed and, in July 2002, it was approved by the FDA as an orphan drug for the treatment of cataplexy in patients with narcolepsy as Xyrem(sodium oxybate) oral solution. To promote the safe use of sodium oxybate, as well as alleviate

  4. Sleep disorders, obesity, and aging: the role of orexin

    PubMed Central

    Nixon, Joshua P.; Mavanji, Vijayakumar; Butterick, Tammy A.; Billington, Charles J.; Kotz, Catherine M.; Teske, Jennifer A.

    2015-01-01

    The hypothalamic neuropeptides orexin A and B (hypocretin 1 and 2) are important homeostatic mediators of central control of energy metabolism and maintenance of sleep/wake states. Dysregulation or loss of orexin signaling has been linked to narcolepsy, obesity, and age-related disorders. In this review, we present an overview of our current understanding of orexin function, focusing on sleep disorders, energy balance, and aging, in both rodents and humans. We first discuss animal models used in studies of obesity and sleep, including loss of function using transgenic or viral-mediated approaches, gain of function models using exogenous delivery of orexin receptor agonist, and naturally-occurring models in which orexin responsiveness varies by individual. We next explore rodent models of orexin in aging, presenting evidence that orexin loss contributes to age-related changes in sleep and energy balance. In the next section, we focus on clinical importance of orexin in human obesity, sleep, and aging. We include discussion of orexin loss in narcolepsy and potential importance of orexin in insomnia, correlations between animal and human studies of age-related decline, and evidence for orexin involvement in age-related changes in cognitive performance. Finally, we present a summary of recent studies of orexin in neurodegenerative disease. We conclude that orexin acts as an integrative homeostatic signal influencing numerous brain regions, and that this pivotal role results in potential dysregulation of multiple physiological processes when orexin signaling is disrupted or lost. PMID:25462194

  5. Sleep disorders, obesity, and aging: the role of orexin.

    PubMed

    Nixon, Joshua P; Mavanji, Vijayakumar; Butterick, Tammy A; Billington, Charles J; Kotz, Catherine M; Teske, Jennifer A

    2015-03-01

    The hypothalamic neuropeptides orexin A and B (hypocretin 1 and 2) are important homeostatic mediators of central control of energy metabolism and maintenance of sleep/wake states. Dysregulation or loss of orexin signaling has been linked to narcolepsy, obesity, and age-related disorders. In this review, we present an overview of our current understanding of orexin function, focusing on sleep disorders, energy balance, and aging, in both rodents and humans. We first discuss animal models used in studies of obesity and sleep, including loss of function using transgenic or viral-mediated approaches, gain of function models using exogenous delivery of orexin receptor agonist, and naturally-occurring models in which orexin responsiveness varies by individual. We next explore rodent models of orexin in aging, presenting evidence that orexin loss contributes to age-related changes in sleep and energy balance. In the next section, we focus on clinical importance of orexin in human obesity, sleep, and aging. We include discussion of orexin loss in narcolepsy and potential importance of orexin in insomnia, correlations between animal and human studies of age-related decline, and evidence for orexin involvement in age-related changes in cognitive performance. Finally, we present a summary of recent studies of orexin in neurodegenerative disease. We conclude that orexin acts as an integrative homeostatic signal influencing numerous brain regions, and that this pivotal role results in potential dysregulation of multiple physiological processes when orexin signaling is disrupted or lost. Published by Elsevier B.V.

  6. Amygdala Lesions Reduce Cataplexy in Orexin KO mice

    PubMed Central

    Burgess, C.R.; Oishi, Y.; Mochizuki, T.; Peever, J.H.; Scammell, T.E.

    2013-01-01

    Narcolepsy is characterized by excessive sleepiness and cataplexy, sudden episodes of muscle weakness during waking that are thought to be an intrusion of REM sleep muscle atonia into wakefulness. One of the most striking aspects of cataplexy is that it is often triggered by strong, generally positive emotions, but little is known about the neural pathways through which positive emotions trigger muscle atonia. We hypothesized that the amygdala is functionally important for cataplexy because the amygdala has a role in processing emotional stimuli and it contains neurons that are active during cataplexy. Using anterograde and retrograde tracing in mice, we found that GABAergic neurons in the central nucleus of the amygdala heavily innervate neurons that maintain waking muscle tone such as those in the ventrolateral periaqueductal grey, lateral pontine tegmentum, locus coeruleus, and dorsal raphe. We then found that bilateral, excitotoxic lesions of the amygdala markedly reduced cataplexy in orexin knockout mice, a model of narcolepsy. These lesions did not alter basic sleep/wake behavior, but substantially reduced the triggering of cataplexy. Lesions also reduced the cataplexy events triggered by conditions associated with high arousal and positive emotions (i.e., wheel running and chocolate). These observations demonstrate that the amygdala is a functionally important part of the circuitry underlying cataplexy and suggest that increased amygdala activity in response to emotional stimuli could directly trigger cataplexy by inhibiting brainstem regions that suppress muscle atonia. PMID:23739970

  7. Activity of dorsal raphe cells across the sleep–waking cycle and during cataplexy in narcoleptic dogs

    PubMed Central

    Wu, M-F; John, J; Boehmer, L N; Yau, D; Nguyen, G B; Siegel, J M

    2004-01-01

    Cataplexy, a symptom associated with narcolepsy, represents a unique dissociation of behavioural states. During cataplectic attacks, awareness of the environment is maintained, as in waking, but muscle tone is lost, as in REM sleep. We have previously reported that, in the narcoleptic dog, noradrenergic cells of the locus coeruleus cease discharge during cataplexy. In the current study, we report on the activity of serotonergic cells of the dorsal raphe nucleus. The discharge patterns of serotonergic dorsal raphe cells across sleep–waking states did not differ from those of dorsal raphe and locus coeruleus cells recorded in normal rats, cats and monkeys, with tonic discharge in waking, reduced activity in non-REM sleep and cessation of activity in REM sleep. However, in contrast with locus coeruleus cells, dorsal raphe REM sleep-off neurones did not cease discharge during cataplexy. Instead, discharge continued at a level significantly higher than that seen in REM sleep and comparable to that seen in non-REM sleep. We also identified several cells in the dorsal raphe whose pattern of activity was the opposite of that of the presumed serotonergic cells. These cells were maximally active in REM sleep and minimally active in waking and increased activity during cataplexy. The difference between noradrenergic and serotonergic cell discharge profiles in cataplexy suggests different roles for these cell groups in the normal regulation of environmental awareness and muscle tone and in the pathophysiology of narcolepsy. PMID:14678502

  8. Amygdala lesions reduce cataplexy in orexin knock-out mice.

    PubMed

    Burgess, Christian R; Oishi, Yo; Mochizuki, Takatoshi; Peever, John H; Scammell, Thomas E

    2013-06-05

    Narcolepsy is characterized by excessive sleepiness and cataplexy, sudden episodes of muscle weakness during waking that are thought to be an intrusion of rapid eye movement sleep muscle atonia into wakefulness. One of the most striking aspects of cataplexy is that it is often triggered by strong, generally positive emotions, but little is known about the neural pathways through which positive emotions trigger muscle atonia. We hypothesized that the amygdala is functionally important for cataplexy because the amygdala has a role in processing emotional stimuli and it contains neurons that are active during cataplexy. Using anterograde and retrograde tracing in mice, we found that GABAergic neurons in the central nucleus of the amygdala heavily innervate neurons that maintain waking muscle tone such as those in the ventrolateral periaqueductal gray, lateral pontine tegmentum, locus ceruleus, and dorsal raphe. We then found that bilateral, excitotoxic lesions of the amygdala markedly reduced cataplexy in orexin knock-out mice, a model of narcolepsy. These lesions did not alter basic sleep-wake behavior but substantially reduced the triggering of cataplexy. Lesions also reduced the cataplexy events triggered by conditions associated with high arousal and positive emotions (i.e., wheel running and chocolate). These observations demonstrate that the amygdala is a functionally important part of the circuitry underlying cataplexy and suggest that increased amygdala activity in response to emotional stimuli could directly trigger cataplexy by inhibiting brainstem regions that suppress muscle atonia.

  9. Running promotes wakefulness and increases cataplexy in orexin knockout mice.

    PubMed

    España, Rodrigo A; McCormack, Sarah L; Mochizuki, Takatoshi; Scammell, Thomas E

    2007-11-01

    People with narcolepsy and mice lacking orexin/hypocretin have disrupted sleep/wake behavior and reduced physical activity. Our objective was to identify physiologic mechanisms through which orexin deficiency reduces locomotor activity. We examined spontaneous wheel running activity and its relationship to sleep/wake behavior in wild type (WT) and orexin knockout (KO) mice. Additionally, given that physical activity promotes alertness, we also studied whether orexin deficiency reduces the wake-promoting effects of exercise. Orexin KO mice ran 42% less than WT mice. Their ability to run appeared normal as they initiated running as often as WT mice and ran at normal speeds. However, their running bouts were considerably shorter, and they often had cataplexy or quick transitions into sleep after running. Wheel running increased the total amount of wakefulness in WT and orexin KO mice similarly, however, KO mice continued to have moderately fragmented sleep/wake behavior. Wheel running also doubled the amount of cataplexy by increasing the probability of transitioning into cataplexy. Orexin KO mice run significantly less than normal, likely due to sleepiness, imminent cataplexy, or a reduced motivation to run. Orexin is not required for the wake-promoting effects of wheel running given that both WT and KO mice had similar increases in wakefulness with running wheels. In addition, the clear increase in cataplexy with wheel running suggests the possibility that positive emotions or reward can trigger murine cataplexy, similar to that seen in people and dogs with narcolepsy.

  10. Elevated body temperature during sleep in orexin knockout mice

    PubMed Central

    Mochizuki, Takatoshi; Klerman, Elizabeth B.; Sakurai, Takeshi; Scammell, Thomas E.

    2008-01-01

    Core body temperature (Tb) is influenced by many physiological factors, including behavioral state, locomotor activity, and biological rhythms. To determine the relative roles of these factors, we examined Tb in orexin knockout (KO) mice, which have a narcolepsy-like phenotype with severe sleep-wake fragmentation. Because orexin is thought to promote heat production during wakefulness, we hypothesized that orexin KO mice would have lower Tb while awake. Surprisingly, the Tb of orexin KO mice was 0.4°C higher than wild-type (WT) littermates during the dark period. Orexin KO mice had normal diurnal variations in Tb, but the ultradian rhythms of Tb, locomotor activity, and wakefulness were markedly reduced. During sustained wakefulness, Tb was the same in both groups. During the first 15 min of spontaneous sleep, the Tb of WT mice decreased by 1.0°C, but Tb in orexin KO mice decreased only 0.4°C. Even during intense recovery sleep after 8 hr of sleep deprivation, the Tb of orexin KO mice remained 0.7°C higher than in WT mice. This blunted fall in Tb during sleep may be due to inadequate activation of heat loss mechanisms or sustained activity in heat-generating systems. These observations reveal an unexpected role for orexin in thermoregulation. In addition, because heat loss is an essential aspect of sleep, the blunted fall in Tb of orexin KO mice may provide an explanation for the fragmented sleep of narcolepsy. PMID:16556901

  11. Different fates of excessive daytime sleepiness: survival analysis for remission.

    PubMed

    Kim, T; Lee, J H; Lee, C S; Yoon, I Y

    2016-07-01

    Excessive daytime sleepiness (EDS) is a symptom frequently presented in sleep clinics. Only a paucity of data has addressed clinical courses of sleep disorders with EDS. Therefore, we sought to compare clinical outcomes of patients presenting EDS. A retrospective observational study was performed in the setting of sleep laboratory and outpatient department in a university hospital. One hundred and eight patients who presented EDS underwent polysomnography and multiple sleep latency test. Each patient was diagnosed as one of the following four categories: (1) narcolepsy with cataplexy (N + C; n = 29); (2) narcolepsy without cataplexy (N - C; n = 22); (3) idiopathic hypersomnia (IH; n = 24); and (4) subjective hypersomnolence (SH; n = 33) with mean sleep latency >8 min. Remission of EDS and treatment response were determined based on clinical evaluation. Kaplan-Meier survival analysis was performed. Remission rates were significantly different (P < 0.001, overall log-rank test) among four groups except those between N - C and IH (P = 0.489). While N + C showed no remission, predicted remission rates of N - C and IH group were 44.6% at 5 years and 32.5% at 5.5 years after diagnosis. The predicted remission rate of SH group was 71.7% at 3 years after diagnosis. The similarity of clinical courses between N - C and IH suggests that N - C may be more related to IH compared to N + C. Considering different clinical courses among EDS patients, thorough evaluation of EDS should be warranted before starting treatment. © 2015 John Wiley & Sons A/S. Published by John Wiley & Sons Ltd.

  12. Modafinil improves real driving performance in patients with hypersomnia: a randomized double-blind placebo-controlled crossover clinical trial.

    PubMed

    Philip, Pierre; Chaufton, Cyril; Taillard, Jacques; Capelli, Aurore; Coste, Olivier; Léger, Damien; Moore, Nicholas; Sagaspe, Patricia

    2014-03-01

    Patients with excessive daytime sleepiness (EDS) are at high risk for driving accidents, and physicians are concerned by the effect of alerting drugs on driving skills of sleepy patients. No study has up to now investigated the effect of modafinil (a reference drug to treat EDS in patients with hypersomnia) on on-road driving performance of patients suffering from central hypersomnia. The objective is to evaluate in patients with central hypersomnia the effect of a wake-promoting drug on real driving performance and to assess the relationship between objective sleepiness and driving performance. Randomized, crossover, double-blind placebo-controlled trial conducted among 13 patients with narcolepsy and 14 patients with idiopathic hypersomnia. Patients were randomly assigned to receive modafinil (400 mg) or placebo for 5 days prior to the driving test. Each condition was separated by at least 3 weeks of washout. Mean number of Inappropriate Line Crossings, Standard Deviation of Lateral Position of the vehicle and mean sleep latency in the Maintenance of Wakefulness Test were assessed. Modafinil reduced the mean number of Inappropriate Line Crossings and Standard Deviation of Lateral Position of the vehicle compared to placebo (F(1,25) = 4.88, P < 0.05 and F(1,25) = 3.87, P = 0.06 tendency). Mean sleep latency at the Maintenance of Wakefulness Test significantly correlated with the mean number of Inappropriate Line Crossings (r = -0.41, P < 0.001). Modafinil improves driving performance in patients with narcolepsy and idiopathic hypersomnia. The Maintenance of Wakefulness Test is a suitable clinical tool to assess fitness to drive in this population.

  13. Risks of high-dose stimulants in the treatment of disorders of excessive somnolence: a case-control study.

    PubMed

    Auger, R Robert; Goodman, Scott H; Silber, Michael H; Krahn, Lois E; Pankratz, V Shane; Slocumb, Nancy L

    2005-06-01

    To ascertain complications associated with high-dose stimulant therapy in patients with narcolepsy or idiopathic hypersomnia. Case-control, retrospective chart review. Sleep center in an academic hospital. 116 patients with narcolepsy or idiopathic hypersomnia were individually matched by sex, diagnosis, age of onset, and duration of follow-up from both onset and diagnosis. Members of the high-dose group (n = 58) had received at least 1 stimulant at a dosage > or = 120% of the maximum recommended by the American Academy of Sleep Medicine Standards of Practice Committee. The standard-dose control group (n = 58) had received stimulants at a dosage < or = 100% of the American Academy of Sleep Medicine guidelines. N/A. The prevalence of psychosis (odds ratio = 12.0 [1.6-92.0]), alcohol or polysubstance misuse (odds ratio = 4.3 [1.2-15.2]), and psychiatric hospitalization (odds ratio = 3.2 [1.1-10.0]) was significantly increased in the high-dose group. More high-dose patients also experienced tachyarrhythmias (odds ratio = 3.3 [0.92-12.1] and anorexia or weight loss (odds ratio = 11.0 [1.4-85.2]). The frequency of physician-diagnosed depression, drug-seeking and suicide-related behaviors, hypertension, and cardiovascular disease did not differ significantly between the groups. This study demonstrated a significantly higher occurrence of psychosis, substance misuse, and psychiatric hospitalizations in patients using high-dose stimulants compared to those using standard doses. Tachyarrhythmias and anorexia or weight loss were also more common in this group as compared with controls. Clinicians should be very cautious in prescribing dosages that exceed maximum guidelines.

  14. Recommended treatment strategies for patients with excessive daytime sleepiness.

    PubMed

    Rosenberg, Russell P

    2015-10-01

    Excessive daytime sleepiness (EDS) is a common and bothersome phenomenon. It can be associated with insufficient sleep syndrome, narcolepsy, idiopathic hypersomnia, obstructive sleep apnea, shift work disorder, Kleine-Levin syndrome, or Parkinson's disease. Once the underlying cause of the excessive sleepiness is determined, clinicians must select the most appropriate behavioral and pharmacologic interventions to reduce daytime sleepiness, alleviate other symptoms, improve functioning, and ensure the safety of patients and those around them. Patient history, adverse effects, and efficacy in specific conditions should be considered in pharmacologic treatment options for patients with EDS. © Copyright 2015 Physicians Postgraduate Press, Inc.

  15. Sleep State Switching

    PubMed Central

    Saper, Clifford B.; Fuller, Patrick M.; Pedersen, Nigel P.; Lu, Jun; Scammell, Thomas E.

    2010-01-01

    We take for granted the ability to fall asleep or to snap out of sleep into wakefulness, but these changes in behavioral state require specific switching mechanisms in the brain that allow well-defined state transitions. In this review, we examine the basic circuitry underlying the regulation of sleep and wakefulness, and discuss a theoretical framework wherein the interactions between reciprocal neuronal circuits enable relatively rapid and complete state transitions. We also review how homeostatic, circadian, and allostatic drives help regulate sleep state switching, and discuss how breakdown of the switching mechanism may contribute to sleep disorders such as narcolepsy. PMID:21172606

  16. Loss of hypocretin (orexin) neurons with traumatic brain injury.

    PubMed

    Baumann, Christian R; Bassetti, Claudio L; Valko, Philipp O; Haybaeck, Johannes; Keller, Morten; Clark, Erika; Stocker, Reto; Tolnay, Markus; Scammell, Thomas E

    2009-10-01

    Chronic, daytime sleepiness is a major, disabling symptom for many patients with traumatic brain injury (TBI), but thus far, its etiology is not well understood. Extensive loss of the hypothalamic neurons that produce the wake-promoting neuropeptide hypocretin (orexin) causes the severe sleepiness of narcolepsy, and partial loss of these cells may contribute to the sleepiness of Parkinson disease and other disorders. We have found that the number of hypocretin neurons is significantly reduced in patients with severe TBI. This observation highlights the often overlooked hypothalamic injury in TBI and provides new insights into the causes of chronic sleepiness in patients with TBI.

  17. Hypersomnia in children: interface with psychiatric disorders.

    PubMed

    Kotagal, Suresh

    2009-10-01

    Patients being evaluated in child psychiatry clinics for behavior and mood disturbances frequently exhibit daytime sleepiness. Conversely, patients being evaluated for hypersomnia by sleep specialists may have depressed mood or hyperactive and aggressive behavior. The etiology of daytime sleepiness in children and adolescents is diverse and includes inadequate sleep hygiene, obstructive sleep apnea, delayed sleep phase syndrome, idiopathic hypersomnia, periodic hypersomnia, narcolepsy, and mood disorders per se. Treatment of a sleep disorder can have a favorable impact on alertness and quality of life. A high index of suspicion for sleep problems should be maintained in children and adolescents with psychiatric disorders.

  18. Defects of mutant DNMT1 are linked to a spectrum of neurological disorders

    PubMed Central

    Baets, Jonathan; Duan, Xiaohui; Wu, Yanhong; Smith, Gordon; Seeley, William W.; Mademan, Inès; McGrath, Nicole M.; Beadell, Noah C.; Khoury, Julie; Botuyan, Maria-Victoria; Mer, Georges; Worrell, Gregory A.; Hojo, Kaori; DeLeon, Jessica; Laura, Matilde; Liu, Yo-Tsen; Senderek, Jan; Weis, Joachim; Van den Bergh, Peter; Merrill, Shana L.; Reilly, Mary M.; Houlden, Henry; Grossman, Murray; Scherer, Steven S.; De Jonghe, Peter; Dyck, Peter J.

    2015-01-01

    We report a broader than previously appreciated clinical spectrum for hereditary sensory and autonomic neuropathy type 1E (HSAN1E) and a potential pathogenic mechanism for DNA methyltransferase (DNMT1) mutations. The clinical presentations and genetic characteristics of nine newly identified HSAN1E kinships (45 affected subjects) were investigated. Five novel mutations of DNMT1 were discovered; p.C353F, p.T481P, p.P491L, p.Y524D and p.I531N, all within the target-sequence domain, and two mutations (p.T481P, p.P491L) arising de novo. Recently, HSAN1E has been suggested as an allelic disorder of autosomal dominant cerebellar ataxia, deafness and narcolepsy. Our results indicate that all the mutations causal for HSAN1E are located in the middle part or N-terminus end of the TS domain, whereas all the mutations causal for autosomal dominant cerebellar ataxia, deafness and narcolepsy are located in the C-terminus end of the TS domain. The impact of the seven causal mutations in this cohort was studied by cellular localization experiments. The binding efficiency of the mutant DNMT proteins at the replication foci and heterochromatin were evaluated. Phenotypic characterizations included electromyography, brain magnetic resonance and nuclear imaging, electroencephalography, sural nerve biopsies, sleep evaluation and neuropsychometric testing. The average survival of HSAN1E was 53.6 years. [standard deviation = 7.7, range 43–75 years], and mean onset age was 37.7 years. (standard deviation = 8.6, range 18–51 years). Expanded phenotypes include myoclonic seizures, auditory or visual hallucinations, and renal failure. Hypersomnia, rapid eye movement sleep disorder and/or narcolepsy were identified in 11 subjects. Global brain atrophy was found in 12 of 14 who had brain MRI. EEGs showed low frequency (delta waves) frontal-predominant abnormality in five of six patients. Marked variability in cognitive deficits was observed, but the majority of patients (89%) developed

  19. Genetics Home Reference: autosomal dominant cerebellar ataxia, deafness, and narcolepsy

    MedlinePlus

    ... This Page Baets J, Duan X, Wu Y, Smith G, Seeley WW, Mademan I, McGrath NM, Beadell ... Central Kernohan KD, Cigana Schenkel L, Huang L, Smith A, Pare G, Ainsworth P; Care4Rare Canada Consortium, ...

  20. Narcolepsy and its treatment with stimulants. ASDA standards of practice.

    PubMed

    Mitler, M M; Aldrich, M S; Koob, G F; Zarcone, V P

    1994-06-01

    This review is part of the standards of practice recommendations. It has been commended and reviewed by the Board of the ASDA. It reflects recommendations of the Board for the practice of sleep medicine in North America. The subcommittee is responsible for the presented write-up.

  1. Effectiveness and side-effect profile of stimulant therapy as monotherapy and in combination in the central hypersomnias in clinical practice.

    PubMed

    Thakrar, Chiraag; Patel, Kishankumar; D'ancona, Grainne; Kent, Brian D; Nesbitt, Alexander; Selsick, Hugh; Steier, Joerg; Rosenzweig, Ivana; Williams, Adrian J; Leschziner, Guy D; Drakatos, Panagis

    2017-10-19

    Effectiveness and side-effect profile data on pharmacotherapy for daytime sleepiness in central hypersomnias are based largely upon randomized controlled trials. Evidence regarding the use of combination therapy is scant. The aim of this study was to examine the effectiveness and occurrence of drug-related side effects of these drugs in routine clinical practice. Adult patients diagnosed with a central hypersomnia during a 54-month period at a tertiary sleep disorders centre were identified retrospectively. Side effects were recorded at every follow-up visit. A total of 126 patients, with 3275 patient-months of drug exposure, were categorized into narcolepsy type 1 (n = 70), narcolepsy type 2 (n = 47) and idiopathic hypersomnia (n = 9). Modafinil was the most common drug used as a first-line treatment (93%) and in combination therapy (70%). Thirty-nine per cent of the patients demonstrated a complete, 25% partial and 36% a poor response to treatment. Combination treatment improved daytime sleepiness in 55% of the patients with residual symptoms despite monotherapy. Sixty per cent of patients reported side effects, and 30% reported treatment-limiting side effects. Drugs had similar side-effect incidence (P = 0.363) and their side-effect profile met those reported in the literature. Twenty-seven per cent of the patients received combination treatment and had fewer side effects compared to monotherapy (29.4% versus 60%, respectively, P = 0.001). Monotherapy appears to achieve satisfactory symptom control in most patients with central hypersomnia, but significant side effects are common. Combination therapy appears to be a useful and safe option in patients with refractory symptoms. © 2017 European Sleep Research Society.

  2. Running Promotes Wakefulness and Increases Cataplexy in Orexin Knockout Mice

    PubMed Central

    España, Rodrigo A.; McCormack, Sarah L.; Mochizuki, Takatoshi; Scammell, Thomas E.

    2007-01-01

    Study Objective: People with narcolepsy and mice lacking orexin/hypocretin have disrupted sleep/wake behavior and reduced physical activity. Our objective was to identify physiologic mechanisms through which orexin deficiency reduces locomotor activity. Design: We examined spontaneous wheel running activity and its relationship to sleep/wake behavior in wild type (WT) and orexin knockout (KO) mice. Additionally, given that physical activity promotes alertness, we also studied whether orexin deficiency reduces the wake-promoting effects of exercise. Measurements and Results: Orexin KO mice ran 42% less than WT mice. Their ability to run appeared normal as they initiated running as often as WT mice and ran at normal speeds. However, their running bouts were considerably shorter, and they often had cataplexy or quick transitions into sleep after running. Wheel running increased the total amount of wakefulness in WT and orexin KO mice similarly, however, KO mice continued to have moderately fragmented sleep/wake behavior. Wheel running also doubled the amount of cataplexy by increasing the probability of transitioning into cataplexy. Conclusions: Orexin KO mice run significantly less than normal, likely due to sleepiness, imminent cataplexy, or a reduced motivation to run. Orexin is not required for the wake-promoting effects of wheel running given that both WT and KO mice had similar increases in wakefulness with running wheels. In addition, the clear increase in cataplexy with wheel running suggests the possibility that positive emotions or reward can trigger murine cataplexy, similar to that seen in people and dogs with narcolepsy. Citation: España RA; McCormack SL; Mochizuki T; Scammell TE. Running promotes wakefulness and increases cataplexy in orexin knockout mice. SLEEP 2007;30(11):1417-1425. PMID:18041476

  3. Disinhibition of perifornical hypothalamic neurones activates noradrenergic neurones and blocks pontine carbachol-induced REM sleep-like episodes in rats

    PubMed Central

    Lu, Jackie W; Fenik, Victor B; Branconi, Jennifer L; Mann, Graziella L; Rukhadze, Irma; Kubin, Leszek

    2007-01-01

    Studies in behaving animals suggest that neurones located in the perifornical (PF) region of the posterior hypothalamus promote wakefulness and suppress sleep. Among such cells are those that synthesize the excitatory peptides, orexins (ORX). Lack of ORX, or their receptors, is associated with narcolepsy/cataplexy, a disorder characterized by an increased pressure for rapid eye movement (REM) sleep. We used anaesthetized rats in which pontine microinjections of a cholinergic agonist, carbachol, can repeatedly elicit REM sleep-like episodes to test whether activation of PF cells induced by antagonism of endogenous, GABAA receptor-mediated, inhibition suppresses the ability of the brainstem to generate REM sleep-like state. Microinjections of the GABAA receptor antagonist, bicuculline (20 nl, 1 mm), into the PF region elicited cortical and hippocampal activation, increased the respiratory rate and hypoglossal nerve activity, induced c-fos expression in ORX and other PF neurones, and increased c-fos expression in pontine A7 and other noradrenergic neurones. The ability of pontine carbachol to elicit any cortical, hippocampal or brainstem component of the REM sleep-like response was abolished during the period of bicuculline-induced activation. The activating and REM sleep-suppressing effect of PF bicuculline was not attenuated by systemic administration of the ORX type 1 receptor antagonist, SB334867. Thus, activation of PF neurones that are endogenously inhibited by GABAA receptors is sufficient to turn off the brainstem REM sleep-generating network; the effect is, at least in part, due to activation of pontine noradrenergic neurones, but is not mediated by ORX type 1 receptors. A malfunction of the pathway that originates in GABAA receptor-expressing PF neurones may cause narcolepsy/cataplexy. PMID:17495048

  4. Sustained attention to response task (SART) shows impaired vigilance in a spectrum of disorders of excessive daytime sleepiness.

    PubMed

    Van Schie, Mojca K M; Thijs, Roland D; Fronczek, Rolf; Middelkoop, Huub A M; Lammers, Gert Jan; Van Dijk, J Gert

    2012-08-01

    The sustained attention to response task comprises withholding key presses to one in nine of 225 target stimuli; it proved to be a sensitive measure of vigilance in a small group of narcoleptics. We studied sustained attention to response task results in 96 patients from a tertiary narcolepsy referral centre. Diagnoses according to ICSD-2 criteria were narcolepsy with (n=42) and without cataplexy (n=5), idiopathic hypersomnia without long sleep time (n=37), and obstructive sleep apnoea syndrome (n=12). The sustained attention to response task was administered prior to each of five multiple sleep latency test sessions. Analysis concerned error rates, mean reaction time, reaction time variability and post-error slowing, as well as the correlation of sustained attention to response task results with mean latency of the multiple sleep latency test and possible time of day influences. Median sustained attention to response task error scores ranged from 8.4 to 11.1, and mean reaction times from 332 to 366ms. Sustained attention to response task error score and mean reaction time did not differ significantly between patient groups. Sustained attention to response task error score did not correlate with multiple sleep latency test sleep latency. Reaction time was more variable as the error score was higher. Sustained attention to response task error score was highest for the first session. We conclude that a high sustained attention to response task error rate reflects vigilance impairment in excessive daytime sleepiness irrespective of its cause. The sustained attention to response task and the multiple sleep latency test reflect different aspects of sleep/wakefulness and are complementary. © 2011 European Sleep Research Society.

  5. Recent advances in the treatment and management of excessive daytime sleepiness.

    PubMed

    Black, Jed; Duntley, Stephen P; Bogan, Richard K; O'Malley, Mary B

    2007-02-01

    Excessive daytime sleepiness (EDS) is a prevalent complaint among patients in psychiatric care. Patients with conditions of EDS have often been misdiagnosed with depression due to their complaints of lack of energy, poor concentration, memory disturbance, and a reduced interest in life. Impaired alertness associated with EDS can be detrimental to a person's quality of life by causing decreased work performance, self-consciousness, low self esteem, and social isolation. Excessive sleepiness is also associated with various health problems, comorbid medical and psychiatric conditions, and fatal accidents occurring after the driver has fallen asleep at the wheel. Contributing factors leading to EDS range from insufficient sleep hours to central nervous system-mediated debilitating hypersomnolence. Circadian rhythm disorders, sleep disorders such as obstructive sleep apnea and narcolepsy, and medications that cause sleepiness may also contribute to symptoms of EDS. Recognition of the symptoms of sleep deprivation is essential, as many such patients do not have a clear awareness of their own sleepiness. Treatment options, depending upon the condition, include light therapy or appropriate airway management techniques such as nasal continuous positive airway pressure (CPAP). Occasionally, wakefulness-promoting medications are necessary, particularly in patients with narcolepsy. In this expert roundtable supplement, Stephen P. Duntley, MD, reviews the definition and prevalence of EDS and discusses the contributing factors and consequences of daytime sleepiness. Next, Richard K. Bogan, MD, FCCP, gives an overview of the differential diagnosis of EDS and the assessment tools available for identifying sleepiness in symptomatic patients. Finally, Mary B. O'Malley, MD, PhD, reviews treatment of EDS, including counseling on sleep hygiene and duration of sleep, mechanical treatments, bright-light therapy, and wake-promoting medications.

  6. Migraine, arousal and sleep deprivation: comment on: "sleep quality, arousal and pain thresholds in migraineurs: a blinded controlled polysomnographic study".

    PubMed

    Vollono, Catello; Testani, Elisa; Losurdo, Anna; Mazza, Salvatore; Della Marca, Giacomo

    2013-06-10

    We discuss the hypothesis proposed by Engstrom and coworkers that Migraineurs have a relative sleep deprivation, which lowers the pain threshold and predispose to attacks. Previous data indicate that Migraineurs have a reduction of Cyclic Alternating Pattern (CAP), an essential mechanism of NREM sleep regulation which allows to dump the effect of incoming disruptive stimuli, and to protect sleep. The modifications of CAP observed in Migraineurs are similar to those observed in patients with impaired arousal (narcolepsy) and after sleep deprivation. The impairment of this mechanism makes Migraineurs more vulnerable to stimuli triggering attacks during sleep, and represents part of a more general vulnerability to incoming stimuli.

  7. Migraine, arousal and sleep deprivation: comment on: “sleep quality, arousal and pain thresholds in migraineurs: a blinded controlled polysomnographic study”

    PubMed Central

    2013-01-01

    We discuss the hypothesis proposed by Engstrom and coworkers that Migraineurs have a relative sleep deprivation, which lowers the pain threshold and predispose to attacks. Previous data indicate that Migraineurs have a reduction of Cyclic Alternating Pattern (CAP), an essential mechanism of NREM sleep regulation which allows to dump the effect of incoming disruptive stimuli, and to protect sleep. The modifications of CAP observed in Migraineurs are similar to those observed in patients with impaired arousal (narcolepsy) and after sleep deprivation. The impairment of this mechanism makes Migraineurs more vulnerable to stimuli triggering attacks during sleep, and represents part of a more general vulnerability to incoming stimuli. PMID:23758606

  8. [Hypernatremia caused by treatment with GHB obtained via a doctor's prescription].

    PubMed

    Rood, I M; Seijger, C G W; van Waarde, J A; de Maat, M M R; Verhave, J C; Blans, M J

    In the last few years, gamma hydroxybutyric acid (GHB) has been used increasingly as a party drug; this has led to a marked increase in the number of requests for professional help with the treatment of GHB addiction. Pharmaceutical GHB (sodium oxybate, the sodium-salt of GHB), registered for cataplexia in narcolepsy patients, is used off-label to treat the withdrawal symptoms associated with GHB addiction. Pharmaceutical GHB has a high sodium load. In this report we present the cases of two patients who developed symptomatic hypernatremia following treatment with pharmaceutical GHB and who thereafter needed intensive care for the severe withdrawal symptoms that they experienced.

  9. Sleep state switching.

    PubMed

    Saper, Clifford B; Fuller, Patrick M; Pedersen, Nigel P; Lu, Jun; Scammell, Thomas E

    2010-12-22

    We take for granted the ability to fall asleep or to snap out of sleep into wakefulness, but these changes in behavioral state require specific switching mechanisms in the brain that allow well-defined state transitions. In this review, we examine the basic circuitry underlying the regulation of sleep and wakefulness and discuss a theoretical framework wherein the interactions between reciprocal neuronal circuits enable relatively rapid and complete state transitions. We also review how homeostatic, circadian, and allostatic drives help regulate sleep state switching and discuss how breakdown of the switching mechanism may contribute to sleep disorders such as narcolepsy. Copyright © 2010 Elsevier Inc. All rights reserved.

  10. Loss of hypocretin (orexin) neurons with traumatic brain injury

    PubMed Central

    Baumann, Christian R.; Bassetti, Claudio L.; Valko, Philipp O.; Haybaeck, Johannes; Keller, Morten; Clark, Erika; Stocker, Reto; Tolnay, Markus; Scammell, Thomas E.

    2009-01-01

    Chronic, daytime sleepiness is a major, disabling symptom for many patients with traumatic brain injury (TBI), but thus far, its etiology is not well understood. Extensive loss of the hypothalamic neurons that produce the wake-promoting neuropeptide hypocretin (orexin) causes the severe sleepiness of narcolepsy, and partial loss of these cells may contribute to the sleepiness of Parkinson’s disease and other disorders. We have found that the number of hypocretin neurons is significantly reduced in patients with severe TBI. This observation highlights the often overlooked hypothalamic injury in TBI and provides new insights into the causes of chronic sleepiness in patients with TBI. PMID:19847903

  11. Hypersomnia: Evaluation, Treatment, and Social and Economic Aspects.

    PubMed

    Saini, Prabhjyot; Rye, David B

    2017-03-01

    Most central disorders of hypersomnolence are conditions with poorly understood pathophysiologies, making their identification, treatment, and management challenging for sleep clinicians. The most challenging to diagnose and treat is idiopathic hypersomnia. There are no FDA-approved treatments, and off-label usage of narcolepsy treatments seldom provide benefit. Patients are largely left on their own to alleviate the compound effects of this disorder on their quality of life. This review covers the major points regarding clinical features and diagnosis of idiopathic hypersomnia, reviews current evidence supporting the available treatment options, and discusses the psychosocial impact and effects of idiopathic hypersomnia. Copyright © 2016 Elsevier Inc. All rights reserved.

  12. Clinical assessment of excessive daytime sleepiness in the diagnosis of sleep disorders.

    PubMed

    Rosenberg, Russell P

    2015-12-01

    Daytime sleepiness is common, but, in some individuals, it can be excessive and lead to distress and impairment. For many of these individuals, excessive daytime sleepiness is simply caused by poor sleep habits or self-imposed sleep times that are not sufficient to maintain alertness throughout the day. For others, daytime sleepiness may be related to a more serious disorder or condition such as narcolepsy, idiopathic hypersomnia, or obstructive sleep apnea. Clinicians must be familiar with the disorders associated with excessive daytime sleepiness and the assessment methods used to diagnose these disorders in order to identify patients who need treatment. © Copyright 2015 Physicians Postgraduate Press, Inc.

  13. Pharmacological treatment of sleep disorders and its relationship with neuroplasticity.

    PubMed

    Abad, Vivien C; Guilleminault, Christian

    2015-01-01

    Sleep and wakefulness are regulated by complex brain circuits located in the brain stem, thalamus, subthalamus, hypothalamus, basal forebrain, and cerebral cortex. Wakefulness and NREM and REM sleep are modulated by the interactions between neurotransmitters that promote arousal and neurotransmitters that promote sleep. Various lines of evidence suggest that sleep disorders may negatively affect neuronal plasticity and cognitive function. Pharmacological treatments may alleviate these effects but may also have adverse side effects by themselves. This chapter discusses the relationship between sleep disorders, pharmacological treatments, and brain plasticity, including the treatment of insomnia, hypersomnias such as narcolepsy, restless legs syndrome (RLS), obstructive sleep apnea (OSA), and parasomnias.

  14. Defects of mutant DNMT1 are linked to a spectrum of neurological disorders.

    PubMed

    Baets, Jonathan; Duan, Xiaohui; Wu, Yanhong; Smith, Gordon; Seeley, William W; Mademan, Inès; McGrath, Nicole M; Beadell, Noah C; Khoury, Julie; Botuyan, Maria-Victoria; Mer, Georges; Worrell, Gregory A; Hojo, Kaori; DeLeon, Jessica; Laura, Matilde; Liu, Yo-Tsen; Senderek, Jan; Weis, Joachim; Van den Bergh, Peter; Merrill, Shana L; Reilly, Mary M; Houlden, Henry; Grossman, Murray; Scherer, Steven S; De Jonghe, Peter; Dyck, Peter J; Klein, Christopher J

    2015-04-01

    We report a broader than previously appreciated clinical spectrum for hereditary sensory and autonomic neuropathy type 1E (HSAN1E) and a potential pathogenic mechanism for DNA methyltransferase (DNMT1) mutations. The clinical presentations and genetic characteristics of nine newly identified HSAN1E kinships (45 affected subjects) were investigated. Five novel mutations of DNMT1 were discovered; p.C353F, p.T481P, p.P491L, p.Y524D and p.I531N, all within the target-sequence domain, and two mutations (p.T481P, p.P491L) arising de novo. Recently, HSAN1E has been suggested as an allelic disorder of autosomal dominant cerebellar ataxia, deafness and narcolepsy. Our results indicate that all the mutations causal for HSAN1E are located in the middle part or N-terminus end of the TS domain, whereas all the mutations causal for autosomal dominant cerebellar ataxia, deafness and narcolepsy are located in the C-terminus end of the TS domain. The impact of the seven causal mutations in this cohort was studied by cellular localization experiments. The binding efficiency of the mutant DNMT proteins at the replication foci and heterochromatin were evaluated. Phenotypic characterizations included electromyography, brain magnetic resonance and nuclear imaging, electroencephalography, sural nerve biopsies, sleep evaluation and neuropsychometric testing. The average survival of HSAN1E was 53.6 years. [standard deviation = 7.7, range 43-75 years], and mean onset age was 37.7 years. (standard deviation = 8.6, range 18-51 years). Expanded phenotypes include myoclonic seizures, auditory or visual hallucinations, and renal failure. Hypersomnia, rapid eye movement sleep disorder and/or narcolepsy were identified in 11 subjects. Global brain atrophy was found in 12 of 14 who had brain MRI. EEGs showed low frequency (delta waves) frontal-predominant abnormality in five of six patients. Marked variability in cognitive deficits was observed, but the majority of patients (89%) developed

  15. Different sleep onset criteria at the multiple sleep latency test (MSLT): an additional marker to differentiate central nervous system (CNS) hypersomnias.

    PubMed

    Pizza, Fabio; Vandi, Stefano; Detto, Stefania; Poli, Francesca; Franceschini, Christian; Montagna, Pasquale; Plazzi, Giuseppe

    2011-03-01

    Excessive daytime sleepiness (EDS) has different correlates in non-rapid eye movement (NREM) [idiopathic hypersomnia (IH) without long sleep time] and REM sleep [narcolepsy without cataplexy (NwoC) and narcolepsy with cataplexy (NC)]-related hypersomnias of central origin. We analysed sleep onset characteristics at the multiple sleep latency test (MSLT) applying simultaneously two sleep onset criteria in 44 NC, seven NwoC and 16 IH consecutive patients referred for subjective EDS complaint. Sleep latency (SL) at MSLT was assessed both as the time elapsed to the occurrence of a single epoch of sleep Stage 1 NREM (SL) and of unequivocal sleep [three sleep Stage 1 NREM epochs or any other sleep stage epoch, sustained SL (SusSL)]. Idiopathic hypersomnia patients showed significantly (P<0.0001) longer SusSL than SL (7.7±2.5 versus 5.6±1.3 min, respectively) compared to NwoC (5.8±2.5 versus 5.3±2.2 min) and NC patients (4.1±3 versus 3.9±3 min). A mean difference threshold between SusSL and SL ≥27 s reached a diagnostic value to discriminate IH versus NC and NwoC sufferers (sensitivity 88%; specificity 82%). Moreover, NC patients showed better subjective sleepiness perception than NwoC and IH cases in the comparison between naps with or without sleep occurrence. Simultaneous application of the two widely used sleep onset criteria differentiates IH further from NC and NwoC patients: IH fluctuate through a wake-Stage 1 NREM sleep state before the onset of sustained sleep, while NC and NwoC shift abruptly into a sustained sleep. The combination of SusSL and SL determination at MSLT should be tested as an additional objective differential criterion for EDS disorders. © 2010 European Sleep Research Society.

  16. Car Crashes and Central Disorders of Hypersomnolence: A French Study.

    PubMed

    Pizza, Fabio; Jaussent, Isabelle; Lopez, Regis; Pesenti, Carole; Plazzi, Giuseppe; Drouot, Xavier; Leu-Semenescu, Smaranda; Beziat, Severine; Arnulf, Isabelle; Dauvilliers, Yves

    2015-01-01

    Drowsiness compromises driving ability by reducing alertness and attentiveness, and delayed reaction times. Sleep-related car crashes account for a considerable proportion of accident at the wheel. Narcolepsy type 1 (NT1), narcolepsy type 2 (NT2) and idiopathic hypersomnia (IH) are rare central disorders of hypersomnolence, the most severe causes of sleepiness thus being potential dangerous conditions for both personal and public safety with increasing scientific, social, and political attention. Our main objective was to assess the frequency of recent car crashes in a large cohort of patients affected with well-defined central disorders of hypersomnolence versus subjects from the general population. We performed a cross-sectional study in French reference centres for rare hypersomnia diseases and included 527 patients and 781 healthy subjects. All participants included needed to have a driving license, information available on potential accident events during the last 5 years, and on potential confounders; thus analyses were performed on 282 cases (71 IH, 82 NT2, 129 NT1) and 470 healthy subjects. Patients reported more frequently than healthy subjects the occurrence of recent car crashes (in the previous five years), a risk that was confirmed in both treated and untreated subjects at study inclusion (Untreated, OR = 2.21 95%CI = [1.30-3.76], Treated OR = 2.04 95%CI = [1.26-3.30]), as well as in all disease categories, and was modulated by subjective sleepiness level (Epworth scale and naps). Conversely, the risk of car accidents of patients treated for at least 5 years was not different to healthy subjects (OR = 1.23 95%CI = [0.56-2.69]). Main risk factors were analogous in patients and healthy subjects. Patients affected with central disorders of hypersomnolence had increased risk of recent car crashes compared to subjects from the general population, a finding potentially reversed by long-term treatment.

  17. Sleep disorders and medical conditions in women. Proceedings of the Women & Sleep Workshop, National Sleep Foundation, Washington, DC, March 5-6, 2007.

    PubMed

    Phillips, Barbara A; Collop, Nancy A; Drake, Christopher; Consens, Flavia; Vgontzas, Alexandros N; Weaver, Terri E

    2008-09-01

    Sleep disorders affect women differently than they affect men and may have different manifestations and prevalences. With regard to obstructive sleep apnea (OSA), variations in symptoms may cause misdiagnoses and delay of appropriate treatment. The prevalence of OSA appears to increase markedly after the time of menopause. Although OSA as defined by the numbers of apneas/hypopneas may be less severe in women, its consequences are similar and perhaps worse. Therapeutic issues related to gender should be factored into the management of OSA. The prevalence of insomnia is significantly greater in women than in men throughout most of the life span. The ratio of insomnia in women to men is approximately 1.4:1.0, but the difference is minimal before puberty and increases steadily with age. Although much of the higher prevalence of insomnia in women may be attributable to the hormonal or psychological changes associated with major life transitions, some of the gender differences may result from the higher prevalence of depression and pain in women. Insomnia's negative impact on quality of life is important to address in women, given the high relative prevalence of insomnia as well as the comorbid disorders in this population. Gender differences in etiology and symptom manifestation in narcolepsy remain understudied in humans. There is little available scientific information to evaluate the clinical significance and specific consequences of the diagnosis of narcolepsy in women. Restless legs syndrome (RLS) is characterized by an urge to move the legs or other limbs during periods of rest or inactivity and may affect as much as 10% of the population. This condition is more likely to afflict women than men, and its risk is increased by pregnancy. Although RLS is associated with impaired quality of life, highly effective treatment is available.

  18. Progressive Loss of the Orexin Neurons Reveals Dual Effects on Wakefulness

    PubMed Central

    Branch, Abigail F.; Navidi, William; Tabuchi, Sawako; Terao, Akira; Yamanaka, Akihiro; Scammell, Thomas E.; Diniz Behn, Cecilia

    2016-01-01

    Study Objectives: Narcolepsy is caused by loss of the orexin (also known as hypocretin) neurons. In addition to the orexin peptides, these neurons release additional neurotransmitters, which may produce complex effects on sleep/wake behavior. Currently, it remains unknown whether the orexin neurons promote the initiation as well as the maintenance of wakefulness, and whether the orexin neurons influence initiation or maintenance of sleep. To determine the effects of the orexin neurons on the dynamics of sleep/wake behavior, we analyzed sleep/wake architecture in a novel mouse model of acute orexin neuron loss. Methods: We used survival analysis and other statistical methods to analyze sleep/wake architecture in orexin-tTA ; TetO diphtheria toxin A mice at different stages of orexin neuron degeneration. Results: Progressive loss of the orexin neurons dramatically reduced survival of long wake bouts, but it also improved survival of brief wake bouts. In addition, with loss of the orexin neurons, mice were more likely to wake during the first 30 sec of nonrapid eye movement sleep and then less likely to return to sleep during the first 60 sec of wakefulness. Conclusions: These findings help explain the sleepiness and fragmented sleep that are characteristic of narcolepsy. Orexin neuron loss impairs survival of long wake bouts resulting in poor maintenance of wakefulness, but this neuronal loss also fragments sleep by increasing the risk of awakening at the beginning of sleep and then reducing the likelihood of quickly returning to sleep. Citation: Branch AF, Navidi W, Tabuchi S, Terao A, Yamanaka A, Scammell TE, Diniz Behn C. Progressive loss of the orexin neurons reveals dual effects on wakefulness. SLEEP 2016;39(2):369–377. PMID:26446125

  19. A Physiologically Based Model of Orexinergic Stabilization of Sleep and Wake

    PubMed Central

    Fulcher, Ben D.; Phillips, Andrew J. K.; Postnova, Svetlana; Robinson, Peter A.

    2014-01-01

    The orexinergic neurons of the lateral hypothalamus (Orx) are essential for regulating sleep-wake dynamics, and their loss causes narcolepsy, a disorder characterized by severe instability of sleep and wake states. However, the mechanisms through which Orx stabilize sleep and wake are not well understood. In this work, an explanation of the stabilizing effects of Orx is presented using a quantitative model of important physiological connections between Orx and the sleep-wake switch. In addition to Orx and the sleep-wake switch, which is composed of mutually inhibitory wake-active monoaminergic neurons in brainstem and hypothalamus (MA) and the sleep-active ventrolateral preoptic neurons of the hypothalamus (VLPO), the model also includes the circadian and homeostatic sleep drives. It is shown that Orx stabilizes prolonged waking episodes via its excitatory input to MA and by relaying a circadian input to MA, thus sustaining MA firing activity during the circadian day. During sleep, both Orx and MA are inhibited by the VLPO, and the subsequent reduction in Orx input to the MA indirectly stabilizes sustained sleep episodes. Simulating a loss of Orx, the model produces dynamics resembling narcolepsy, including frequent transitions between states, reduced waking arousal levels, and a normal daily amount of total sleep. The model predicts a change in sleep timing with differences in orexin levels, with higher orexin levels delaying the normal sleep episode, suggesting that individual differences in Orx signaling may contribute to chronotype. Dynamics resembling sleep inertia also emerge from the model as a gradual sleep-to-wake transition on a timescale that varies with that of Orx dynamics. The quantitative, physiologically based model developed in this work thus provides a new explanation of how Orx stabilizes prolonged episodes of sleep and wake, and makes a range of experimentally testable predictions, including a role for Orx in chronotype and sleep inertia. PMID

  20. Absence of γ-aminobutyric acid-a receptor potentiation in central hypersomnolence disorders.

    PubMed

    Dauvilliers, Yves; Evangelista, Elisa; Lopez, Regis; Barateau, Lucie; Jaussent, Isabelle; Cens, Thierry; Rousset, Matthieu; Charnet, Pierre

    2016-08-01

    The pathophysiology of idiopathic hypersomnia (IH) remains unclear. Recently, cerebrospinal fluid (CSF)-induced enhancement of γ-aminobutyric acid (GABA)-A receptor activity was found in patients with IH compared to controls. Fifteen unrelated patients (2 males and 13 females) affected with typical IH, 12 patients (9 males and 3 females) with narcolepsy type 1, and 15 controls (9 males and 6 females) with unspecified hypersomnolence (n = 7) and miscellaneous neurological conditions (n = 8) were included. A lumbar puncture was performed in all participants to measure CSF hypocretin-1 and GABA-A response. We used a voltage-clamp assay on Xenopus oocytes injected with the RNAs that encode the α1 β2 γ2 or the α2 β2 γ2 subunits of the human GABA-A receptor. A sequence of 6 different applications (GABA, GABA/CSF, and CSF alone) with 2 to 4 oocytes per CSF sample was performed in a whole-cell voltage-clamp assay. Representative current traces from oocytes expressing human α1 β2 γ2 or α2 β2 γ2 GABA-A receptors were recorded in response to 6 successive puffs of GABA diluted in the survival medium (SM), showing stable and reliable response. GABA puffs diluted in SM/CSF solution or SM/CSF solution alone showed no significant differences in the CSF of IH, narcolepsy, or control groups. No associations were found between GABA responses, demographic features, disease duration, or disease severity in the whole population or within groups. Using the Xenopus oocyte assay, we found an absence of GABA-A receptor potentiation with CSF from patients with central hypersomnolence disorders, with no significant differences between hypocretin-deficient and non-hypocretin-deficient patients compared to controls. Ann Neurol 2016;80:259-268. © 2016 American Neurological Association.

  1. The hypocretins/orexins: integrators of multiple physiological functions

    PubMed Central

    Li, Jingcheng; Hu, Zhian; Lecea, Luis

    2014-01-01

    The hypocretins (Hcrts), also known as orexins, are two peptides derived from a single precursor produced in the posterior lateral hypothalamus. Over the past decade, the orexin system has been associated with numerous physiological functions, including sleep/arousal, energy homeostasis, endocrine, visceral functions and pathological states, such as narcolepsy and drug abuse. Here, we review the discovery of Hcrt/orexins and their receptors and propose a hypothesis as to how the orexin system orchestrates these multifaceted physiological functions. Linked ArticlesThis article is part of a themed section on Orexin Receptors. To view the other articles in this section visit http://dx.doi.org/10.1111/bph.2014.171.issue-2 PMID:24102345

  2. Do enteric neurons make hypocretin? ☆

    PubMed Central

    Baumann, Christian R.; Clark, Erika L.; Pedersen, Nigel P.; Hecht, Jonathan L.; Scammell, Thomas E.

    2008-01-01

    Hypocretins (orexins) are wake-promoting neuropeptides produced by hypothalamic neurons. These hypocretin-producing cells are lost in people with narcolepsy, possibly due to an autoimmune attack. Prior studies described hypocretin neurons in the enteric nervous system, and these cells could be an additional target of an autoimmune process. We sought to determine whether enteric hypocretin neurons are lost in narcoleptic subjects. Even though we tried several methods (including whole mounts, sectioned tissue, pre-treatment of mice with colchicine, and the use of various primary antisera), we could not identify hypocretin-producing cells in enteric nervous tissue collected from mice or normal human subjects. These results raise doubts about whether enteric neurons produce hypocretin. PMID:18191238

  3. Do enteric neurons make hypocretin?

    PubMed

    Baumann, Christian R; Clark, Erika L; Pedersen, Nigel P; Hecht, Jonathan L; Scammell, Thomas E

    2008-04-10

    Hypocretins (orexins) are wake-promoting neuropeptides produced by hypothalamic neurons. These hypocretin-producing cells are lost in people with narcolepsy, possibly due to an autoimmune attack. Prior studies described hypocretin neurons in the enteric nervous system, and these cells could be an additional target of an autoimmune process. We sought to determine whether enteric hypocretin neurons are lost in narcoleptic subjects. Even though we tried several methods (including whole mounts, sectioned tissue, pre-treatment of mice with colchicine, and the use of various primary antisera), we could not identify hypocretin-producing cells in enteric nervous tissue collected from mice or normal human subjects. These results raise doubts about whether enteric neurons produce hypocretin.

  4. PubMed

    Pelizza, Paolo

    2017-08-28

    For a long time, European legislation has not considered the Obstructive Sleep Apnea Syndrome (OSAS) a relevant factor for road safety. In Italy, in 2010 the Scientific Society of Legal Medicine COMLAS published Guidelines for the assessment of fitness to drive. An entire chapter, written in collaboration with the Commission "Sleepiness, Safety and Transportation" of the Italian Association of Sleep Medicine (AIMS), is dedicated to the criteria for assessing fitness to drive of people with OSAS or narcolepsy. Only in 2016, following the adoption of European Commission Directive 2014/85/EU, Italian law recognized OSAS as a condition to be considered in the assessment of fitness to drive. The author discusses the ascertainment criteria and the medico-legal implications regarding information to patients and reporting duty.

  5. Modafinil Improves Real Driving Performance in Patients with Hypersomnia: A Randomized Double-Blind Placebo-Controlled Crossover Clinical Trial

    PubMed Central

    Philip, Pierre; Chaufton, Cyril; Taillard, Jacques; Capelli, Aurore; Coste, Olivier; Léger, Damien; Moore, Nicholas; Sagaspe, Patricia

    2014-01-01

    Study Objective: Patients with excessive daytime sleepiness (EDS) are at high risk for driving accidents, and physicians are concerned by the effect of alerting drugs on driving skills of sleepy patients. No study has up to now investigated the effect of modafinil (a reference drug to treat EDS in patients with hypersomnia) on on-road driving performance of patients suffering from central hypersomnia. The objective is to evaluate in patients with central hypersomnia the effect of a wake-promoting drug on real driving performance and to assess the relationship between objective sleepiness and driving performance. Design and Participants: Randomized, crossover, double-blind placebo-controlled trial conducted among 13 patients with narcolepsy and 14 patients with idiopathic hypersomnia. Patients were randomly assigned to receive modafinil (400 mg) or placebo for 5 days prior to the driving test. Each condition was separated by at least 3 weeks of washout. Measurements: Mean number of Inappropriate Line Crossings, Standard Deviation of Lateral Position of the vehicle and mean sleep latency in the Maintenance of Wakefulness Test were assessed. Results: Modafinil reduced the mean number of Inappropriate Line Crossings and Standard Deviation of Lateral Position of the vehicle compared to placebo (F(1,25) = 4.88, P < 0.05 and F(1,25) = 3.87, P = 0.06 tendency). Mean sleep latency at the Maintenance of Wakefulness Test significantly correlated with the mean number of Inappropriate Line Crossings (r = -0.41, P < 0.001). Conclusions: Modafinil improves driving performance in patients with narcolepsy and idiopathic hypersomnia. The Maintenance of Wakefulness Test is a suitable clinical tool to assess fitness to drive in this population. Citation: Philip P; Chaufton C; Taillard J; Capelli A; Coste O; Léger D; Moore N; Sagaspe P. Modafinil improves real driving performance in patients with hypersomnia: a randomized double-blind placebo-controlled crossover clinical trial. SLEEP

  6. Car Crashes and Central Disorders of Hypersomnolence: A French Study

    PubMed Central

    Lopez, Regis; Pesenti, Carole; Plazzi, Giuseppe; Drouot, Xavier; Leu-Semenescu, Smaranda; Beziat, Severine; Arnulf, Isabelle; Dauvilliers, Yves

    2015-01-01

    Background Drowsiness compromises driving ability by reducing alertness and attentiveness, and delayed reaction times. Sleep-related car crashes account for a considerable proportion of accident at the wheel. Narcolepsy type 1 (NT1), narcolepsy type 2 (NT2) and idiopathic hypersomnia (IH) are rare central disorders of hypersomnolence, the most severe causes of sleepiness thus being potential dangerous conditions for both personal and public safety with increasing scientific, social, and political attention. Our main objective was to assess the frequency of recent car crashes in a large cohort of patients affected with well-defined central disorders of hypersomnolence versus subjects from the general population. Methods We performed a cross-sectional study in French reference centres for rare hypersomnia diseases and included 527 patients and 781 healthy subjects. All participants included needed to have a driving license, information available on potential accident events during the last 5 years, and on potential confounders; thus analyses were performed on 282 cases (71 IH, 82 NT2, 129 NT1) and 470 healthy subjects. Results Patients reported more frequently than healthy subjects the occurrence of recent car crashes (in the previous five years), a risk that was confirmed in both treated and untreated subjects at study inclusion (Untreated, OR = 2.21 95%CI = [1.30–3.76], Treated OR = 2.04 95%CI = [1.26–3.30]), as well as in all disease categories, and was modulated by subjective sleepiness level (Epworth scale and naps). Conversely, the risk of car accidents of patients treated for at least 5 years was not different to healthy subjects (OR = 1.23 95%CI = [0.56–2.69]). Main risk factors were analogous in patients and healthy subjects. Conclusion Patients affected with central disorders of hypersomnolence had increased risk of recent car crashes compared to subjects from the general population, a finding potentially reversed by long-term treatment. PMID:26052938

  7. Orexin Receptors: Pharmacology and Therapeutic Opportunities

    PubMed Central

    Scammell, Thomas E.; Winrow, Christopher J.

    2011-01-01

    Orexin-A and -B (also known as hypocretin-1 and -2) are neuropeptides produced in the lateral hypothalamus that promote many aspects of arousal through the OX1 and OX2 receptors. In fact, they are necessary for normal wakefulness, as loss of the orexin-producing neurons causes narcolepsy in humans and rodents. This has generated considerable interest in developing small-molecule orexin receptor antagonists as a novel therapy for the treatment of insomnia. Orexin antagonists, especially those that block OX2 or both OX1 and OX2 receptors, clearly promote sleep in animals, and clinical results are encouraging: Several compounds are in Phase III trials. As the orexin system mainly promotes arousal, these new compounds will likely improve insomnia without incurring many of the side effects encountered with current medications. PMID:21034217

  8. Orexin receptors: pharmacology and therapeutic opportunities.

    PubMed

    Scammell, Thomas E; Winrow, Christopher J

    2011-01-01

    Orexin-A and -B (also known as hypocretin-1 and -2) are neuropeptides produced in the lateral hypothalamus that promote many aspects of arousal through the OX1 and OX2 receptors. In fact, they are necessary for normal wakefulness, as loss of the orexin-producing neurons causes narcolepsy in humans and rodents. This has generated considerable interest in developing small-molecule orexin receptor antagonists as a novel therapy for the treatment of insomnia. Orexin antagonists, especially those that block OX2 or both OX1 and OX2 receptors, clearly promote sleep in animals, and clinical results are encouraging: Several compounds are in Phase III trials. As the orexin system mainly promotes arousal, these new compounds will likely improve insomnia without incurring many of the side effects encountered with current medications.

  9. New Advances in the Treatment of Neurological Diseases Using High Dose Intravenous Immunoglobulins

    PubMed Central

    2008-01-01

    Since the incidental discovery in 1981 that intravenous immunoglobulins (IVIg) are immunomodulatory, they have been investigated in a large number of putative autoimmune diseases. This has led to licensing for idiopathic thrombocytopenic purpura, Kawasaki disease, and in neurological disorders for Guillain-Barré syndrome (GBS). Although not licensed, randomized controlled trials have also shown IVIg efficacy in other neuroimmunological diseases such as multifocal motor neuropathy (MMN), chronic inflammatory demyelinating neuropathy (CIDP), myasthenia gravis, dermatomyositis, and stiff-person syndrome. However, other indications are currently being explored including Alzheimer's disease, postpolio syndrome, and narcolepsy. There are even reports from experimental studies in stroke. The results of recently published clinical trials in both the classical neuroimmunological disorders as well as for new indications are reported and their role in clinical practice is discussed. PMID:21180569

  10. Sustained attention assessment of narcoleptic patients: two case reports.

    PubMed

    Moraes, Mirleny; Wilson, Barbara A; Rossini, Sueli; Osternack-Pinto, Kátia; Reimão, Rubens

    2008-01-01

    Narcolepsy is a sleep disorder characterized by uncontrollable REM sleep attacks which alter the patients wake state and can lead to difficulties in attention aspects, such as maintaining attention when performing activities or tasks. This study aimed to evaluate sustained attention performance of two narcoleptic patients on the d2 Test, Epworth Sleepiness Scale (ESS), Pittsburgh Sleep Quality Index (PSQI) and Hamilton Rating Scale for Depression (HAM-D). Results showed that the maintenance of attention was associated with a slowing of the target symbols processing function in visual scanning with accuracy in task performance. A high degree of excessive sleepiness was observed, along with mild and moderate degrees of depressive signs and symptoms. One subject also presented with a nocturnal sleep disorder which could represent an important factor affecting attentional and affective capacity.

  11. Orexin in sleep, addiction and more: is the perfect insomnia drug at hand?

    PubMed

    Hoyer, Daniel; Jacobson, Laura H

    2013-12-01

    Orexins A and B (hypocretins 1 and 2) and their two receptors (OX1R and OX2R) were discovered in 1998 by two different groups. Orexin A and B are derived from the differential processing of a common precursor, the prepro-orexin peptide. The neuropeptides are expressed in a few thousand cells located in the lateral hypothalamus (LH), but their projections and receptor distribution are widespread throughout the brain. Remarkably, prepro peptide and double (OX1R/OX2R) receptor knock out (KO) mice reproduce a sleep phenotype known in humans and dogs as narcolepsy/cataplexy. In humans, this disease is characterized by the absence of orexin producing cells in the LH, and severely depleted levels of orexin the cerebrospinal fluid. Null mutation of the individual OX1R or OX2R in mice substantially ameliorates the narcolepsy/cataplexy phenotype compared to the OX1R/OX2R KO, and highlights specific roles of the individual receptors in sleep architecture, the OX1R KO demonstrating an a attenuated sleep phenotype relative to the OX2R KO. It has therefore been suggested that orexin is a master regulator of the sleep-wake cycle, with high activity of the LH orexin cells during wake and almost none during sleep. Less than 10years later, the first orexin antagonist, almorexant, a dual orexin receptor antagonist (DORA), was reported to be effective in inducing sleep in volunteers and insomnia patients. Although development was stopped for almorexant and for Glaxo's DORA SB-649868, no less than 4 orexin receptor antagonists have reached phase II for insomnia, including Filorexant (MK-6096) and Suvorexant (MK-4305) from Merck. Suvorexant has since progressed to Phase III and dossier submission to the FDA. These four compounds are reported as DORAs, however, they equilibrate very slowly at one and/or the other orexin receptor, and thus at equilibrium may show more or less selectivity for OX1R or OX2R. The appropriate balance of antagonism of the two receptors for sleep is a point of

  12. Cerebrospinal fluid hypocretin (orexin) levels are elevated by play but are not raised by exercise and its associated heart rate, blood pressure, respiration or body temperature changes.

    PubMed

    Wu, M-F; Nienhuis, R; Maidment, N; Lam, H A; Siegel, J M

    2011-11-07

    Hypocretin (Hcrt) has been implicated in the control of motor activity and in respiration and cardiovascular changes. Loss of Hcrt in narcolepsy is linked to sleepiness and to cataplexy, a sudden loss of muscle tone which is triggered by sudden strong emotions. In the current study we have compared the effects of treadmill running, to yard play on cerebrospinal fluid (CSF) Hcrt level in normal dogs. We find that treadmill locomotion, at a wide range of speeds, does not increase Hcrt level beyond baseline, whereas yard play produces a substantial increase in Hcrt, even though both activities produce comparable increases in heart rate, respiration and body temperature. We conclude that motor and cardiovascular changes are not sufficient to elevate CSF levels of Hcrt and we hypothesize that the emotional aspects of yard play account for the observed increase in Hcrt.

  13. Neurologic Complications of Psychomotor Stimulant Abuse.

    PubMed

    Sanchez-Ramos, Juan

    2015-01-01

    Psychomotor stimulants are drugs that act on the central nervous system (CNS) to increase alertness, elevate mood, and produce a sense of well-being. These drugs also decrease appetite and the need for sleep. Stimulants can enhance stamina and improve performance in tasks that have been impaired by fatigue or boredom. Approved therapeutic applications of stimulants include attention deficit hyperactivity disorder (ADHD), narcolepsy, and obesity. These agents also possess potent reinforcing properties that can result in excessive self-administration and abuse. Chronic use is associated with adverse effects including psychosis, seizures, and cerebrovascular accidents, though these complications usually occur in individuals with preexisting risk factors. This chapter reviews the adverse neurologic consequences of chronic psychomotor stimulant use and abuse, with a focus on two prototypical stimulants methamphetamine and cocaine. © 2015 Elsevier Inc. All rights reserved.

  14. Ritalin revisited: does it really help in neurological injury?

    PubMed

    Kajs-Wyllie, Marylyn

    2002-12-01

    Methylphenidate (Ritalin) is a commonly used central nervous stimulant. It has been used in various neurological conditions, including attention deficit disorder, depression, and narcolepsy. Methylphenidate has been advocated in patients with traumatic brain injury and stroke for a variety of cognitive, attention, and behavioral problems. It also has been shown to speed recovery from poststroke depression so that patients can participate more fully in rehabilitation programs. Research suggests that it also may have a role in augmenting activity of injured neuronal tissue in the comatose patient, thus facilitating a return to consciousness. The neuroscience nurse plays an important role in monitoring response to Ritalin, including identifying its side effects. A review of the limited studies on the use of Ritalin, its mechanisms of action, dosing, and weaning provide a current understanding of this adjunctive agent's role in treatment for the neurological population.

  15. Genetics of Sleep and Sleep disorders

    PubMed Central

    Sehgal, Amita; Mignot, Emmanuel

    2011-01-01

    Sleep remains one of the least understood phenomena in biology – even its role in synaptic plasticity remains debatable. Since sleep was recognized to be regulated genetically, intense research has launched on two fronts: the development of model organisms for deciphering the molecular mechanisms of sleep and attempts to identify genetic underpinnings of human sleep disorders. In this Review, we describe how unbiased, high-throughput screens in model organisms are uncovering sleep regulatory mechanisms and how pathways, such as the circadian clock network and specific neurotransmitter signals, have conserved effects on sleep from Drosophila to humans. At the same time, genome-wide association (GWA) studies have uncovered ~14 loci increasing susceptibility to sleep disorders, such as narcolepsy and restless leg syndrome. To conclude, we discuss how these different strategies will be critical to unambiguously defining the function of sleep. PMID:21784243

  16. Modafinil: the unique properties of a new stimulant.

    PubMed

    Lyons, T J; French, J

    1991-05-01

    Modafinil, a novel stimulant which has several remarkable features that distinguish it from other stimulants, has been developed by Lafon, a French pharmaceutical company. Unlike the amphetamines, for example, modafinil is reported to have minimal peripheral side effects at therapeutic doses. It also appears to have a low abuse potential, does not interfere with normal sleep, and does not seem to produce tolerance. It improves vigilance especially in sleep-deprived subjects. It has been used clinically for up to 3 years in the treatment of narcolepsy and idiopathic hypersomnia. It could be an ideal replacement for amphetamine in short-term operations in which fatigue might threaten the successful completion of a mission. We recommend that military laboratories experienced in studying sustained performance include modafinil or perhaps a more selective alpha 1 receptor agonist in their investigations.

  17. Diagnosis and treatment of sleep disorders: a brief review for clinicians

    PubMed Central

    Abad, Vivien C.; Guilleminault, Christian

    2003-01-01

    Sleep disorders encompass a wide spectrum of diseases with significant individual health consequences and high economic costs to society. To facilitate the diagnosis and treatment of sleep disorders, this review provides a framework using the International Classification of Sleep Disorders, Primary and secondary insomnia are differentiated, and pharmacological and nonpharmacological treatments are discussed. Common circadian rhythm disorders are described in conjunction with interventions, including chronotherapy and light therapy. The diagnosis and treatment of restless legs syndrome/periodic limb movement disorder is addressed. Attention is focused on obstructive sleep apnea and upper airway resistance syndrome, and their treatment. The constellation of symptoms and findings in narcolepsy are reviewed together with diagnostic testing and therapy, Parasomnias, including sleep terrors, somnambulism, and rapid eye movement (REM) behavior sleep disorders are described, together with associated laboratory testing results and treatment. PMID:22033666

  18. Idiopathic hypersomnia: a report of three adolescent-onset cases in a two-generation family.

    PubMed

    Janácková, Sona; Motte, Jacques; Bakchine, Serge; Sforza, Emilia

    2011-04-01

    Idiopathic hypersomnia is an uncommon sleep disorder characterized by prolonged sleep time and excessive daytime sleepiness without cataplexy. This study concerned a case of familial occurrence. The proband expressed an idiopathic hypersomnia with long sleep time at the age of 12 years. Clinical interview and ad libitum polysomnographic study did not reveal any symptoms of narcolepsy or other sleep disorders. Family history revealed that a 20-year-old sister had experienced symptoms of hypersomnia from the age of 16 and their mother had been diagnosed with idiopathic hypersomnia previously. The diagnosis of idiopathic hypersomnia with long sleep time was confirmed in the sister by clinical interview and ad libitum polysomnography. Human leukocyte antigen (HLA) did not reveal the DQB1-0602 phenotype in the proband and relatives. This report confirms the hypothesis of a genetic predisposition in idiopathic hypersomnia.

  19. Cataplexy leading to the diagnosis of Niemann-Pick disease type C.

    PubMed

    Smit, Liesbeth S; Lammers, Gert Jan; Catsman-Berrevoets, Coriene E

    2006-07-01

    Cataplexy in childhood is a rare and often misdiagnosed symptom. It is described as a brief episode of bilateral loss of muscle tone with intact consciousness, triggered by a variety of strong emotions and in particular with unexpected laughter. This report presents a 9-year old male with progressive cerebellar and pyramidal symptoms and a cognitive decline since the age of 4. His recently developed "drop attacks" on laughter were recognized as cataplexy and led to the diagnosis of Niemann-Pick type C disease. With biochemical studies this diagnosis, a lysosomal storage disease, was confirmed. With cataplexy narcolepsy, Niemann-Pick type C disease, Norrie disease, Prader-Willi syndrome, and Coffin-Lowry syndrome are associated disorders. Recognition of cataplexy in children with concomitant neurologic symptoms may lead to an early and straight diagnosis of one of these disorders.

  20. Reclassification of clinical sleep disorders using traditional models of syndromic, neuroanatomic, pathophysiological and etiological diagnosis.

    PubMed

    Spitzer, A Robert

    2014-09-01

    Existing classifications of central nervous system sleep disorders do not often provide tools to diagnose the majority of patients complaining of sleep-related symptoms, nor always guide effective treatment. I present a novel classification system that completely separates clinical syndromes from anatomical localization, pathophysiology, and etiology. The clinical syndrome I present can describe the majority of patients, but can be fractionated into individual subgroups for further study. By then separating the anatomy and physiology from the symptoms, an avenue of research becomes available to study the different possible structures that regulate sleep, that may be damaged and cause syndromes of sleep dysfunction. Some of these may produce symptoms that overlap with narcolepsy and some may be distinct. Because the clinical syndrome should be distinguished from anatomy or physiology, I have proposed the term narcoleptiform syndrome for the clinical syndrome. The model also clearly separates etiology from anatomy in a classical neurological manner. This allows etiology, localization and symptoms to be studied separately. It is likely that different etiologies may produce damage in areas that produce similar syndromes. For example, in this model, different causes of damage to the orexin nucleus would result in the same clinical syndrome. This reinforces the concept of studying anatomy, symptoms and etiology separately. By studying the relationship of syndromes or symptoms to anatomic localization and pathophysiology, it should be possible to test novel approaches to treatment based on different underlying structure or function. For example, patients with lesions in the ventrolateral preoptic nucleus or the thalamic intralaminar nuclei may both present with insomnia symptoms but need different treatment; or they might present with symptoms overlapping narcolepsy (a narcoleptiform syndrome) yet need different treatment. In some cases, a single treatment may cross over

  1. The hypocretin system and psychiatric disorders.

    PubMed

    Pizza, Fabio; Magnani, Michele; Indrio, Camilla; Plazzi, Giuseppe

    2014-02-01

    The hypocretin system is constituted by a small group of hypothalamic neurons with widespread connections within the entire central nervous system producing two neuropeptides involved in several key physiological functions such as the regulation of sleep and wakefulness, motor control, autonomic functions, metabolism, feeding behavior, and reward. Narcolepsy with cataplexy is a neurological disorder regarded as a disease model for the selective hypocretin system damage, and also shares several psychopatological traits and comorbidities with psychiatric disorders. We reviewed the available literature on the involvement of the hypocretin system in psychiatric nosography. Different evidences such as cerebrospinal hypocretin-1 levels, genetic polymorphisms of the neuropeptides or their receptors, response to treatments, clinical, experimental and functional data directly or indirectly linked the hypocretin system to schizophrenia, mood, anxiety and eating disorders, as well as to addiction. Future genetic and pharmacological studies will disentangle the hypocretin system role in the field of psychiatry.

  2. [Non-convulsive paroxysmal disorders in exogenous-organic diseases of the brain].

    PubMed

    Piven', B N; Koveva, O P

    1999-01-01

    Examination of 273 patients with exogenous-organic diseases of the brain revealed nonconvulsive paroxysmal disorders of traumatic, toxic, infectious, radioactive and combined origin in 112 cases (41.0%). Such disorders were characterised by pronounced polymorphism and presented with viscero-vegetative (36.6%), affective (27.7%), psychosensory (19.6%), sensory (15.2%), ideatoric (11.6%) paroxysms as well as with twilight states of consciousness (16.1%), absence seizures (10.7%), narcolepsy (2.7%), catalepcy (1.8%) and the states of "déjà vu" and "jamais vu" (5.4%). In most of the patients such paroxysms were found 5 or more years after exogenous influences, i.e. when the severity of the organic brain damage increased. A resemblance of nonconvulsive paroxysms was observed in the patients with different etiology of the disease. The disorders were seldom detected in routine medical practice which may cause in adequate therapy.

  3. Dyssomnias, parasomnias, and sleep disorders associated with medical and psychiatric diseases.

    PubMed

    Barthlen, G M; Stacy, C

    1994-03-01

    Sleep disorders can be intrinsic, as are insomnia or narcolepsy, or can be accounted for by external factors, such as noise, altitude, drug or alcohol abuse, or shift work. The arousal disorders, common in children, are usually benign and disappear by puberty. Sleep-wake transition disorders such as sleep starts are benign as well, and may occur at any age. The parasomnias comprise different entities such as nightmares, REM-sleep behavior disorder, sleep enuresis, and bruxism. Diagnosis and treatment often require a multidisciplinary approach. Virtually every psychiatric, neurologic, or medical disease, when of sufficient severity, leaves its specific fingerprint on sleep; some disorders, such as peptic ulcer disease, gastroesophageal reflux, or epilepsy, tend to be exacerbated during sleep. Fortunately, most sleep disorders are amenable to therapy, which can include counseling, sleep hygiene, withholding of an offending agent, behavioral therapy, light therapy, or cautious drug therapy.

  4. Unraveling the Neurobiology of Sleep and Sleep Disorders Using Drosophila.

    PubMed

    Chakravarti, L; Moscato, E H; Kayser, M S

    2017-01-01

    Sleep disorders in humans are increasingly appreciated to be not only widespread but also detrimental to multiple facets of physical and mental health. Recent work has begun to shed light on the mechanistic basis of sleep disorders like insomnia, restless legs syndrome, narcolepsy, and a host of others, but a more detailed genetic and molecular understanding of how sleep goes awry is lacking. Over the past 15 years, studies in Drosophila have yielded new insights into basic questions regarding sleep function and regulation. More recently, powerful genetic approaches in the fly have been applied toward studying primary human sleep disorders and other disease states associated with dysregulated sleep. In this review, we discuss the contribution of Drosophila to the landscape of sleep biology, examining not only fundamental advances in sleep neurobiology but also how flies have begun to inform pathological sleep states in humans. © 2017 Elsevier Inc. All rights reserved.

  5. Metabolic consequences of sleep and circadian disorders.

    PubMed

    Depner, Christopher M; Stothard, Ellen R; Wright, Kenneth P

    2014-07-01

    Sleep and circadian rhythms modulate or control daily physiological patterns with importance for normal metabolic health. Sleep deficiencies associated with insufficient sleep schedules, insomnia with short-sleep duration, sleep apnea, narcolepsy, circadian misalignment, shift work, night eating syndrome, and sleep-related eating disorder may all contribute to metabolic dysregulation. Sleep deficiencies and circadian disruption associated with metabolic dysregulation may contribute to weight gain, obesity, and type 2 diabetes potentially by altering timing and amount of food intake, disrupting energy balance, inflammation, impairing glucose tolerance, and insulin sensitivity. Given the rapidly increasing prevalence of metabolic diseases, it is important to recognize the role of sleep and circadian disruption in the development, progression, and morbidity of metabolic disease. Some findings indicate sleep treatments and countermeasures improve metabolic health, but future clinical research investigating prevention and treatment of chronic metabolic disorders through treatment of sleep and circadian disruption is needed.

  6. Metabolic consequences of sleep and circadian disorders

    PubMed Central

    Depner, Christopher M.; Stothard, Ellen R.; Wright, Kenneth P.

    2014-01-01

    Sleep and circadian rhythms modulate or control daily physiological patterns with importance for normal metabolic health. Sleep deficiencies associated with insufficient sleep schedules, insomnia with short-sleep duration, sleep apnea, narcolepsy, circadian misalignment, shift work, night eating syndrome and sleep-related eating disorder may all contribute to metabolic dysregulation. Sleep deficiencies and circadian disruption associated with metabolic dysregulation may contribute to weight gain, obesity, and type 2 diabetes potentially by altering timing and amount of food intake, disrupting energy balance, inflammation, impairing glucose tolerance and insulin sensitivity. Given the rapidly increasing prevalence of metabolic diseases, it is important to recognize the role of sleep and circadian disruption in the development, progression, and morbidity of metabolic disease. Some findings indicate sleep treatments and countermeasures improve metabolic health, but future clinical research investigating prevention and treatment of chronic metabolic disorders through treatment of sleep and circadian disruption is needed. PMID:24816752

  7. Sleep disorders in pregnancy

    PubMed Central

    Bourjeily, Ghada

    2009-01-01

    Sleep complaints are a common occurrence in pregnancy that are in part due to pregnancy-associated anatomic and physiological changes but may also be due to pathological causes. In the non-pregnant population, sleep deprivation has been associated with physical and cognitive issues; poor sleep may even be associated with adverse maternal outcomes. Maternal obesity, one of the most prevalent risk factors in obstetric practices, together with physiologic changes of pregnancy predispose to the development of sleep disordered breathing. Symptoms of sleep disordered breathing have also been associated with poor maternal outcomes. Management options of restless legs syndrome and narcolepsy pose a challenge in pregnancy; benefits of therapy need to be weighed against the potential harm to the fetus. This article briefly reviews the normal changes in pregnancy affecting sleep, gives an overview of certain sleep disorders occurring in pregnancy, and suggests management options specific for this population. PMID:27582822

  8. Hypersomnolence, Hypersomnia, and Mood Disorders.

    PubMed

    Barateau, Lucie; Lopez, Régis; Franchi, Jean Arthur Micoulaud; Dauvilliers, Yves

    2017-02-01

    Relationships between symptoms of hypersomnolence, psychiatric disorders, and hypersomnia disorders (i.e., narcolepsy and idiopathic hypersomnia) are complex and multidirectional. Hypersomnolence is a common complaint across mood disorders; however, patients suffering from mood disorders and hypersomnolence rarely have objective daytime sleepiness, as assessed by the current gold standard test, the Multiple Sleep Latency Test. An iatrogenic origin of symptoms of hypersomnolence, and sleep apnea syndrome must be considered in a population of psychiatric patients, often overweight and treated with sedative drugs. On the other hand, psychiatric comorbidities, especially depression symptoms, are often reported in patients with hypersomnia disorders, and an endogenous origin cannot be ruled out. A great challenge for sleep specialists and psychiatrists is to differentiate psychiatric hypersomnolence and a central hypersomnia disorder with comorbid psychiatric symptoms. The current diagnostic tools seem to be limited in that condition, and further research in that field is warranted.

  9. [Excessive daytime sleepiness].

    PubMed

    Bittencourt, Lia Rita Azeredo; Silva, Rogério Santos; Santos, Ruth Ferreira; Pires, Maria Laura Nogueira; Mello, Marco Túlio de

    2005-05-01

    Sleepiness is a physiological function, and can be defined as increased propension to fall asleep. However, excessive sleepiness (ES) or hypersomnia refer to an abnormal increase in the probability to fall asleep, to take involuntary naps, or to have sleep atacks, when sleep is not desired. The main causes of excessive sleepiness is chronic sleep deprivation, sleep apnea syndrome, narcolepsy, movement disorders during sleep, circadian sleep disorders, use of drugs and medications, or idiopathic hypersomnia. Social, familial, work, and cognitive impairment are among the consequences of hypersomnia. Moreover, it has also been reported increased risk of accidents. The treatment of excessive sleepiness includes treating the primary cause, whenever identified. Sleep hygiene for sleep deprivation, positive pressure (CPAP) for sleep apnea, dopaminergic agents and exercises for sleep-related movement disorders, phototherapy and/or melatonin for circadian disorders, and use of stimulants are the treatment modalities of first choice.

  10. Precipitating pharmakologies and capital entrapments: narcolepsy and the strange cases of Provigil and Xyrem.

    PubMed

    Wolf-Meyer, Matthew

    2009-01-01

    Since the mid-1990s, Americans have been made more aware of chronic sleep deprivation and sleep disorders exacerbated by dominant temporal regimes of work, school, and family life, primarily through increased medical and media attention. Concomitantly, Americans have turned to medical treatments and pharmaceutical cocktails to achieve normalcy rather than attending to the social and cultural causes of sleep sickness. This turn toward pharmaceuticalization is aided in part by the proliferation of medical disorders and the pharmaceuticals marketed to treat them (e.g., "excessive daytime sleepiness" requires treatment once reserved for narcoleptics). These cocktails have explicit and implicit components: the former consist of pharmaceuticals, the latter of capital dependencies, including ties to medical insurance companies, stable employment, and familial networks. In this article, I examine the proliferation of pharmaceutical cocktails through the concept of the pharmakon-something simultaneously remedy and cause-to illuminate the causes and effects of such pharmaceutical regimens in contemporary American society, specifically those relating to sleepiness. Specific cases of this struggle between chemical dependence and normalcy are offered from my ethnographic work with patients who suffer from sleep disorders.

  11. Absence of mutations in HCRT, HCRTR1 and HCRTR2 in patients with ROHHAD.

    PubMed

    Barclay, Sarah F; Rand, Casey M; Gray, Paul A; Gibson, William T; Wilson, Richard J A; Berry-Kravis, Elizabeth M; Ize-Ludlow, Diego; Bech-Hansen, N Torben; Weese-Mayer, Debra E

    2016-01-15

    Rapid-onset obesity with hypothalamic dysfunction, hypoventilation, and autonomic dysregulation (ROHHAD) is a rare pediatric disease of unknown cause. Here, in response to a recent case report describing a ROHHAD patient who suffered from secondary narcolepsy confirmed by an absence of hypocretin-1 in the cerebrospinal fluid, we consider whether the ROHHAD phenotype is owing to one or more mutations in genes specific to hypocretin protein signalling. DNA samples from 16 ROHHAD patients were analyzed using a combination of next-generation and Sanger sequencing to identify exonic sequence variations in three genes: HCRT, HCRTR1, and HCRTR2. No rare or novel mutations were identified in the exons of HCRT, HCRTR1, or HCRTR2 genes in a set of 16 ROHHAD patients. ROHHAD is highly unlikely to be caused by mutations in the exons of the genes for hypocretin and its two receptors. Copyright © 2015 Elsevier B.V. All rights reserved.

  12. Unusual Sleep Experiences, Dissociation, and Schizotypy: Evidence for a Common Domain

    PubMed Central

    Koffel, Erin; Watson, David

    2009-01-01

    This paper reviews studies that have examined associations between unusual sleep experiences (including nightmares, vivid dreaming, narcolepsy symptoms, and complex nighttime behaviors) and dissociation and schizotypy. Using correlational studies and structural analyses, evidence is provided that unusual sleep experiences, dissociation, and schizotypy belong to a common domain. It is demonstrated that unusual sleep experiences show specificity to dissociation and schizotypy compared to other daytime symptoms (e.g., anxiety, depression, substance use) and other sleep disturbances (e.g., insomnia, lassitude/fatigue). The paper also outlines the methodological limitations of the existing evidence and makes suggestions for future research. Finally, three models for the overlap of daytime and nighttime symptoms are reviewed, including biological abnormalities, trauma, and personality traits. Although further research is needed, it is suggested that daytime and nighttime symptoms result from problems with sleep-wake state boundaries, which may be precipitated by stress or trauma. In addition, association between daytime and nighttime symptoms can be attributed to the higher order personality trait of Oddity. PMID:19581031

  13. Melanin-concentrating hormone neurons discharge in a reciprocal manner to orexin neurons across the sleep–wake cycle

    PubMed Central

    Hassani, Oum Kaltoum; Lee, Maan Gee; Jones, Barbara E.

    2009-01-01

    Neurons containing melanin-concentrating hormone (MCH) are codistributed with neurons containing orexin (Orx or hypocretin) in the lateral hypothalamus, a peptide and region known to be critical for maintaining wakefulness. Evidence from knockout and c-Fos studies suggests, however, that the MCH neurons might play a different role than Orx neurons in regulating activity and sleep–wake states. To examine this possibility, neurons were recorded across natural sleep–wake states in head-fixed rats and labeled by using the juxtacellular technique for subsequent immunohistochemical identification. Neurons identified as MCH+ did not fire during wake (W); they fired selectively during sleep, occasionally during slow wave sleep (SWS) and maximally during paradoxical sleep (PS). As W-Off/Sleep-On, the MCH neurons discharged in a reciprocal manner to the W-On/Sleep-Off Orx neurons and could accordingly play a complementary role to Orx neurons in sleep–wake state regulation and contribute to the pathophysiology of certain sleep disorders, such as narcolepsy with cataplexy. PMID:19188611

  14. New Technologies for Integrating Genomic, Environmental and Trait Data

    PubMed Central

    Church, George M.

    2011-01-01

    Rare diseases, which (by definition) occur at a frequency less than 1/2000 per allele – are individually rare, yet common collectively (10% affected and 50% carrier rates). There are 1800 genes which have tests considered highly predictive and actionable. Human genes with known variants causing insomnia, narcolepsy, and circadian variation include Prion Protein Fatal Familial Insomnia (PRNP), hypocretin (HCRT), DQ beta 1 (DQB1), and period circadian protein homolog (PER2). We have developed human genome sequencing technology that lowered costs a million-fold over the past 6 yr. This has increasingly enabled the use of the causative alleles above, which are far more valuable than merely correlated or common variants. To expand this further we have established community resources for open access collection, integration and interpretation of diverse personal genomic, environmental and trait data evidence.personalgenomes.org) Citation: Church GM. New technologies for integrating genomic, environmental and trait data. J Clin Sleep Med 2011;7(5):Supplement S43–S44. PMID:22003332

  15. The Biology of REM Sleep

    PubMed Central

    Peever, John; Fuller, Patrick M.

    2018-01-01

    Considerable advances in our understanding of the mechanisms and functions of rapid-eye-movement (REM) sleep have occurred over the past decade. Much of this progress can be attributed to the development of new neuroscience tools that have enabled high-precision interrogation of brain circuitry linked with REM sleep control, in turn revealing how REM sleep mechanisms themselves impact processes such as sensorimotor function. This review is intended to update the general scientific community about the recent mechanistic, functional and conceptual developments in our current understanding of REM sleep biology and pathobiology. Specifically, this review outlines the historical origins of the discovery of REM sleep, the diversity of REM sleep expression across and within species, the potential functions of REM sleep (e.g., memory consolidation), the neural circuits that control REM sleep, and how dysfunction of REM sleep mechanisms underlie debilitating sleep disorders such as REM sleep behaviour disorder and narcolepsy. PMID:26766231

  16. The hypocretins (orexins) mediate the “phasic” components of REM sleep: A new hypothesis

    PubMed Central

    Torterolo, Pablo; Chase, Michael H.

    2014-01-01

    In 1998, a group of phenotypically distinct neurons were discovered in the postero-lateral hypothalamus which contained the neuropeptides hypocretin 1 and hypocretin 2 (also called orexin A and orexin B), which are excitatory neuromodulators. Hypocretinergic neurons project throughout the central nervous system and have been involved in the generation and maintenance of wakefulness. The sleep disorder narcolepsy, characterized by hypersomnia and cataplexy, is produced by degeneration of these neurons. The hypocretinergic neurons are active during wakefulness in conjunction with the presence of motor activity that occurs during survival-related behaviors. These neurons decrease their firing rate during non-REM sleep; however there is still controversy upon the activity and role of these neurons during REM sleep. Hence, in the present report we conducted a critical review of the literature of the hypocretinergic system during REM sleep, and hypothesize a possible role of this system in the generation of REM sleep. PMID:26483897

  17. Improving amphetamine therapeutic selectivity: N,N-dimethyl-MTA has dopaminergic effects and does not produce aortic contraction.

    PubMed

    Sotomayor-Zárate, Ramón; Jara, Pablo; Araos, Patricio; Vinet, Raúl; Quiroz, Gabriel; Renard, Georgina M; Espinosa, Pedro; Hurtado-Guzmán, Claudio; Moya, Pablo R; Iturriaga-Vásquez, Patricio; Gysling, Katia; Reyes-Parada, Miguel

    2014-05-01

    Amphetamine derivatives have therapeutic potential in diseases such as attention deficit hyperactivity disorder, narcolepsy and obesity. However, their prolonged use has been associated with cardiovascular toxicity and addiction. In recent years, we have studied the pharmacological effects of amphetamine derivatives such as methylthioamphetamine (MTA) and N,N-dimethyl-thioamphetamine, with the aim of improving their therapeutic selectivity. In this work, we show that similarly to MTA, N,N-dimethyl-thioamphetamine has effects on the dopamine system, producing a significant increase in extracellular levels of dopamine (as measured by in vivo brain microdialysis) and locomotor activity, which is a behavioural measure of dopaminergic activation. However, unlike MTA, N,N-dimethyl- thioamphetamine does not produce aortic contraction in vitro. Our results show that N,N-dimethyl-thioamphetamine is a drug that retains the dopaminergic effects of amphetamine derivatives but exhibits a lower potential for producing cardiovascular side effects. © 2013 Nordic Association for the Publication of BCPT (former Nordic Pharmacological Society).

  18. Hypersomnia.

    PubMed

    Dauvilliers, Yves; Buguet, Alain

    2005-01-01

    Hypersomnia, a complaint of excessive daytime sleep or sleepiness, affects 4% to 6% of the population, with an impact on the everyday life of the patient Methodological tools to explore sleep and wakefulness (interview, questionnaires, sleep diary, polysomnography, Multiple Sleep Latency Test, Maintenance of Wakefulness Test) and psychomotor tests (for example, psychomotor vigilance task and Oxford Sleep Resistance or Osler Test) help distinguish between the causes of hypersomnia. In this article, the causes of hypersomnia are detailed following the conventional classification of hypersomnic syndromes: narcolepsy, idiopathic hypersomnia, recurrent hypersomnia, insufficient sleep syndrome, medication- and toxin-dependent sleepiness, hypersomnia associated with psychiatric disorders, hypersomnia associated with neurological disorders, posttraumatic hypersomnia, infection (with a special emphasis on the differences between bacterial and viral diseases compared with parasitic diseases, such as sleeping sickness) and hypersomnia, hypersomnia associated with metabolic or endocrine diseases, breathing-related sleep disorders and sleep apnea syndromes, and periodic limb movements in sleep.

  19. Differential diagnosis in hypersomnia.

    PubMed

    Dauvilliers, Yves

    2006-03-01

    Hypersomnia includes a group of disorders in which the primary complaint is excessive daytime sleepiness. Chronic hypersomnia is characterized by at least 3 months of excessive sleepiness prior to diagnosis and may affect 4% to 6% of the population. The severity of daytime sleepiness needs to be quantified by subjective scales (at least the Epworth sleepiness scale) and objective tests such as the multiple sleep latency test. Chronic hypersomnia does not correspond to an individual clinical entity but includes numerous different etiologies of hypersomnia as recently reported in the revised International Classification of Sleep Disorders. This review details most of those disorders, including narcolepsy with and without cataplexy, idiopathic hypersomnia with and without long sleep time, recurrent hypersomnia, behaviorally induced insufficient sleep syndrome, hypersomnia due to medical condition, hypersomnia due to drug or substance, hypersomnia not due to a substance or known physiologic condition, and also sleep-related disordered breathing and periodic leg movement disorders.

  20. Hypersomnia

    PubMed Central

    Dauvilliers, Yves; Buguet, Alain

    2005-01-01

    Hypersomnia, a complaint of excessive daytime sleep or sleepiness, affects 4% to 6% of the population, with an impact on the everyday life of the patient Methodological tools to explore sleep and wakefulness (interview, questionnaires, sleep diary, polysomnography Multiple Sleep Latency Test, Maintenance of Wakefulness Test) and psy-chomotor tests (for example, psychomotor vigilance task and Oxford Sleep Resistance or Osier Test) help distinguish between the causes of hypersomnia. In this article, the causes of hypersomnia are detailed following the conventional classification of hypersomnic syndromes: narcolepsy, idiopathic hypersomnia, recurrent hypersomnia, insufficient sleep syndrome, medication- and toxin-dependent sleepiness, hypersomnia associated with psychiatric disorders, hypersomnia associated with neurological disorders, posttraumatic hypersomnia, infection (with a special emphasis on the differences between bacterial and viral diseases compared with parasitic diseases, such as sleeping sickness) and hypersomnia, hypersomnia associated with metabolic or endocrine diseases, breathing-related sleep disorders and sleep apnea syndromes, and periodic limb movements in sleep. PMID:16416710