Sample records for quality daytime sleepiness

  1. Sleep Quality Assessment and Daytime Sleepiness of Liver Transplantation Candidates.

    PubMed

    Marques, D M; Teixeira, H R S; Lopes, A R F; Martins-Pedersoli, T A; Ziviani, L C; Mente, Ê D; Castro-E-Silva, O; Galvão, C M; Mendes, K S

    2016-09-01

    The goal of this study was to evaluate the sleep quality and daytime sleepiness of patients eligible for liver transplants. A cross-sectional prospective study was conducted on liver transplant candidates from a transplant center in the interior of São Paulo State. The Pittsburgh Sleep Quality Index and Epworth Sleepiness Scale questionnaires were applied to obtain demographic and clinical characteristics and to assess sleep quality and daytime sleepiness. The mean (±SD) score on the Epworth Sleepiness Scale of the 45 liver transplantation candidates was 7.00 ± 2.83 points, with 28.89% having scores >10 points, indicating excessive daytime sleepiness. The mean score on the Pittsburgh Sleep Quality Index was 6.64 ± 4.95 points, with 60% of the subjects showing impaired sleep quality, with scores >5 points. The average sleep duration was 07:16 h. Regarding sleep quality self-classification, 31.11% reported poor or very poor quality. It is noteworthy that 73.33% of patients had to go to the bathroom, 53.33% woke up in the middle of the night, and 40.00% reported pain related to sleeping difficulties. Comparison of subjects with good and poor sleep quality revealed a significant difference in time to sleep (P = .0002), sleep hours (P = .0003), and sleep quality self-classification (P = .000072). Liver transplant candidates have a compromised quality of sleep and excessive daytime sleepiness. In clinical practice, we recommend the evaluation and implementation of interventions aimed at improving the sleep and wakefulness cycle, contributing to a better quality of life. Copyright © 2016 Elsevier Inc. All rights reserved.

  2. Associations of Subjective Sleep Quality and Daytime Sleepiness With Cognitive Impairment in Adults and Elders With Heart Failure.

    PubMed

    Byun, Eeeseung; Kim, Jinyoung; Riegel, Barbara

    2017-01-01

    This study examined the association of subjective nighttime sleep quality and daytime sleepiness with cognitive impairment in 105 adults (< 60 years old) and 167 elders (≥ 60 years old) with heart failure. Nighttime sleep quality and daytime sleepiness were measured by the Pittsburgh Sleep Quality Index and the Epworth Sleepiness Scale. Cognitive impairment was assessed using a neuropsychological battery measuring attention, memory, and processing speed. Multivariate logistic regression was used. In adults, daytime sleepiness was associated with cognitive impairment, whereas poor nighttime sleep quality was associated with cognitive impairment in elders. Age may play an important role in how sleep impacts cognition in persons with heart failure. Improving nighttime sleep quality and daytime sleepiness in this population may improve cognition.

  3. [SLEEP QUALITY, EXCESSIVE DAYTIME SLEEPINESS AND INSOMNIA IN CHILEAN PARALYMPIC ATHLETES].

    PubMed

    Durán Agüero, Samuel; Arroyo Jofre, Patricio; Varas Standen, Camila; Herrera-Valenzuela, Tomas; Moya Cantillana, Cristobal; Pereira Robledo, Rodolfo; Valdés-Badilla, Pablo

    2015-12-01

    the sleep takes part in diverse biological and physiological functions, associating his restriction, with minor performance in the sport, nevertheless the quantity and quality of sleep is not known in paralympic athletes. to determine the sleep quality, insomnia and excessive daytime sleepiness in Chilean paralympic athletes. descriptive transverse Study, the sample included 33 paralympic athletes (24.2% women), those who were practicing swimming, tennis of table, football 5, powerlifting and tennis chair. The studied variables measured up across two surveys of dream: the Questionnaire of Insomnia and the Pittsburgh Sleep Quality Index. the paralympic athletes sleep were 6.9 } 1.4 hours, 27.7% presents daytime sleepiness, 69.6 % insomnia (Survey of insomnia =7), whereas 78.7 % exhibits a bad sleep quality. The age showed a positive correlation with latency to the sleep (r=0.417 *), the insomnia with latency to the sleep (r=0.462 **), the Pittsburg score was correlated negatively by the sleep duration (r =-0.323) and latency to the sleep is correlated positively by the Pittsburgh score (r=0.603 **). the chilean paralympic athletes, present a low sleep quality, insomnia and excessive daytime sleepiness, situation that might influence negatively the sports performance. Copyright AULA MEDICA EDICIONES 2014. Published by AULA MEDICA. All rights reserved.

  4. Depressive symptoms are associated with daytime sleepiness and subjective sleep quality in dementia with Lewy bodies.

    PubMed

    Elder, Greg J; Colloby, Sean J; Lett, Debra J; O'Brien, John T; Anderson, Kirstie N; Burn, David J; McKeith, Ian G; Taylor, John-Paul

    2016-07-01

    Sleep problems and depression are common symptoms in dementia with Lewy bodies (DLB), where patients typically experience subjectively poor sleep quality, fatigue and excessive daytime sleepiness. However, whilst sleep disturbances have been linked to depression, this relationship has not received much attention in DLB. The present cross-sectional study addresses this by examining whether depressive symptoms are specifically associated with subjective sleep quality and daytime sleepiness in DLB, and by examining other contributory factors. DLB patients (n = 32) completed the Pittsburgh Sleep Quality Index (PSQI), Epworth Sleepiness Scale (ESS) and the 15-item Geriatric Depression Scale (GDS-15). Motor and cognitive functioning was also assessed. Pearson correlations were used to assess the relationship between GDS-15, ESS and PSQI scores. GDS-15 scores were positively associated with both ESS (r = 0.51, p < 0.01) and PSQI (r = 0.59, p < 0.001) scores. Subjective poor sleep and daytime sleepiness were associated with depressive symptoms in DLB. Given the cross-sectional nature of the present study, the directionality of this relationship cannot be determined, although this association did not appear to be mediated by sleep quality or daytime sleepiness. Nevertheless, these findings have clinical relevance; daytime sleepiness or poor sleep quality might indicate depression in DLB, and subsequent work should examine whether the treatment of depression can reduce excessive daytime sleepiness and improve sleep quality in DLB patients. Alternatively, more rigorous screening for sleep problems in DLB might assist the treatment of depression. © 2015 The Authors. International Journal of Geriatric Psychiatry published by John Wiley & Sons, Ltd. © 2015 The Authors. International Journal of Geriatric Psychiatry published by John Wiley & Sons, Ltd.

  5. Determinants of daytime sleepiness in first-year nursing students: a questionnaire survey.

    PubMed

    Huang, Ching-Feng; Yang, Li-Yu; Wu, Li-Min; Liu, Yi; Chen, Hsing-Mei

    2014-06-01

    Daytime sleepiness may affect student learning achievement. Research studies have found that daytime sleepiness is common in university students; however, information regarding the determinants of daytime sleepiness in this population is still lacking. The purpose of this study was to investigate the determinants of daytime sleepiness in first-year nursing students. In particular, we looked for the relationship between perceived symptoms, nocturnal sleep quality, and daytime sleepiness. A cross-sectional and correlational design was employed. Participants were recruited from two nursing programs at an institute of technology located in southern Taiwan. Ninety-three nursing students completed the questionnaires one month after enrollment into their program. Approximately 35% of the participants experienced excessive daytime sleepiness at the beginning of the semester. Six variables (joining a student club, perceived symptoms, daytime dysfunction, sleep disturbances, sleep latency, and subjective sleep quality) were significantly correlated with daytime sleepiness. Among them, daytime dysfunction and perceived symptoms were two major determinants of daytime sleepiness, both accounting for 37.2% of the variance. Daytime sleepiness in students should not be ignored. It is necessary to help first-year students identify and mitigate physical and psychological symptoms early on, as well as improve daytime functioning, to maintain their daytime performance and promote learning achievement. Copyright © 2013 Elsevier Ltd. All rights reserved.

  6. The effects of individual biological rhythm differences on sleep quality, daytime sleepiness, and dissociative experiences.

    PubMed

    Selvi, Yavuz; Kandeger, Ali; Boysan, Murat; Akbaba, Nursel; Sayin, Ayca A; Tekinarslan, Emine; Koc, Basak O; Uygur, Omer F; Sar, Vedat

    2017-10-01

    Individuals who differ markedly by sleep chronotype, i.e., morning-type or evening-type also differ on a number of psychological, behavioral, and biological variables. Among several other psychological functions, dissociation may also lead to disruption and alteration of consciousness, which may facilitate dream-like experiences. Our study was aimed at an inquiry into the effects of individual biological rhythm differences on sleep quality and daytime sleepiness in conjunction with dissociative experiences. Participants were 372 undergraduate college students, completed a package of psychological instruments, including the Morningness-Eveningness Questionnaire, Dissociative Experiences Scale, Insomnia Severity Index, and Epworth Sleepiness Scale. Using logistic regression models, direct relations of pathological dissociation with sleepiness, sleep quality and circadian preferences were investigated. Poor sleep quality and sleepiness significantly contributed to the variance of dissociative symptomatology. Although there was no substantial linear association between circadian preferences and pathological dissociation, having evening-type preferences of sleep was indirectly associated with higher dissociation mediated by poor sleep quality. Poor sleep quality and daytime sleepiness seems to be significant antecedents of pathological dissociation. Sleep chronotype preferences underlie this relational pattern that chronobiological characteristics seem to influence indirectly on dissociative tendency via sleep quality. Copyright © 2017 Elsevier B.V. All rights reserved.

  7. Association of sociodemographic, lifestyle, and health factors with sleep quality and daytime sleepiness in women: findings from the 2007 National Sleep Foundation "Sleep in America Poll".

    PubMed

    Baker, Fiona C; Wolfson, Amy R; Lee, Kathryn A

    2009-06-01

    To investigate factors associated with poor sleep quality and daytime sleepiness in women living in the United States. Data are presented from the National Sleep Foundation's 2007 Sleep in America Poll that included 959 women (18-64 years of age) surveyed by telephone about their sleep quality, daytime sleepiness, and sociodemographic, health, and lifestyle factors. Poor sleep quality was reported by 27% and daytime sleepiness was reported by 21% of respondents. Logistic multivariate regression analyses revealed that poor sleep quality and daytime sleepiness were both independently associated with poor health, having a sleep disorder, and psychological distress. Also, multivariate analyses showed that women who consumed more caffeinated beverages and those who had more than one job were more likely to report poor sleep quality but not daytime sleepiness. Daytime sleepiness, on the other hand, was independently associated with being black/African American, younger, disabled, having less education, and daytime napping. Poor sleep quality and daytime sleepiness are common in American women and are associated with health-related, as well as sociodemographic, factors. Addressing sleep-related complaints in women is important to improve their daytime functioning and quality of life.

  8. Prevalence and correlates of poor sleep quality and daytime sleepiness in Belgian truck drivers.

    PubMed

    Braeckman, Lutgart; Verpraet, Rini; Van Risseghem, Marleen; Pevernagie, Dirk; De Bacquer, Dirk

    2011-03-01

    Sleepiness and sleep complaints are common among professional drivers. Sleepiness is a considerable problem not only because it affects the drivers' well-being, but also because of the consequences for performance and safety. Assessment of the (self-reported) prevalence and research into the risk factors are thus an important health issue and are also indispensable to prevent productivity loss and work-related accidents and injuries. Therefore, the aim of this study was to describe sleeping, driving, and health characteristics of Belgian truck drivers and to determine occupational and individual factors associated with poor sleep quality and daytime sleepiness. Cross-sectional data were collected using a self-administered questionnaire that included the Pittsburgh Sleep Quality Index (PSQI), Epworth Sleepiness Scale (ESS), and Berlin Questionnaire (BQ). The mean (SD) age of the 476 studied truck drivers was 42.7 (10.2) yrs and the mean (SD) body mass index was 27.3 (5.1) kg/m(2). Approximately 47% declared that they drove >50 h/wk and found their work schedule unrealistic. The mean (SD) PSQI score was 4.45 (2.7); poor quality of sleep (PSQI >5) was found in 27.2%. The mean (SD) ESS score was 6.79 (4.17); 18% had a score >10. The BQ indicated that 21.5% had a higher risk on obstructive sleep apnea. In multiple logistic regression analysis, low educational level (odds ratio [OR] 1.86), current smoking (OR 1.75), unrealistic work schedule (OR 1.75), and risk for obstructive sleep apnea (OR 2.97) were found to be independent correlates of daytime sleepiness. Poor sleep quality was significantly associated with poor self-perceived health (OR 1.95), unrealistic work schedule (OR 2.85), low job satisfaction (OR 1.91), and less driving experience (OR 1.73). These results show that poor sleep quality and daytime sleepiness were prevalent in Belgian truck drivers. Taking into account that several significant correlates with respect to these sleep problems were identified

  9. Nocturnal sleep, daytime sleepiness, and quality of life in stable patients on hemodialysis

    PubMed Central

    Parker, Kathy P; Kutner, Nancy G; Bliwise, Donald L; Bailey, James L; Rye, David B

    2003-01-01

    Background Although considerable progress has been made in the treatment of chronic kidney disease, compromised quality of life continues to be a significant problem for patients receiving hemodialysis (HD). However, in spite of the high prevalence of sleep complaints and disorders in this population, the relationship between these problems and quality of life remains to be well characterized. Thus, we studied a sample of stable HD patients to explore relationships between quality of life and both subjective and objective measures of nocturnal sleep and daytime sleepiness Methods The sample included forty-six HD patients, 24 men and 22 women, with a mean age of 51.6 (10.8) years. Subjects underwent one night of polysomnography followed the next morning by a Multiple Sleep Latency Test (MSLT), an objective measure of daytime sleepiness. Subjects also completed: 1) a brief nocturnal sleep questionnaire; 2) the Epworth Sleepiness Scale; and, 3) the Quality of Life Index (QLI, Dialysis Version) which provides an overall QLI score and four subscale scores for Health & Functioning (H&F), Social & Economic (S&E), Psychological & Spiritual (P&S), and Family (F). (The range of scores is 0 to 30 with higher scores indicating better quality of life.) Results The mean (standard deviation; SD) of the overall QLI was 22.8 (4.0). The mean (SD) of the four subscales were as follows: H&F – 21.1 (4.7); S&E – 22.0 (4.8); P&S – 24.5 (4.4); and, F – 26.8 (3.5). H&F (rs = -0.326, p = 0.013) and F (rs = -0.248, p = 0.048) subscale scores were negatively correlated with periodic limb movement index but not other polysomnographic measures. The H&F subscale score were positively correlated with nocturnal sleep latency (rs = 0.248, p = 0.048) while the H&F (rs = 0.278, p = 0.030) and total QLI (rs = 0.263, p = 0.038) scores were positively associated with MSLT scores. Both of these latter findings indicate that higher life quality is associated with lower sleepiness levels. ESS

  10. Randomized controlled trial of exercise interventions to improve sleep quality and daytime sleepiness in individuals with multiple sclerosis: A pilot study.

    PubMed

    Siengsukon, Catherine F; Aldughmi, Mayis; Kahya, Melike; Bruce, Jared; Lynch, Sharon; Ness Norouzinia, Abigail; Glusman, Morgan; Billinger, Sandra

    2016-01-01

    Nearly 70% of individuals with multiple sclerosis (MS) experience sleep disturbances. Increasing physical activity in people with MS has been shown to produce a moderate improvement in sleep quality, and exercise has been shown to improve sleep quality in non-neurologically impaired adults. The purpose of this pilot randomized controlled trial study was to examine the effect of two exercise interventions on sleep quality and daytime sleepiness in individuals with MS. Twenty-eight individuals with relapsing-remitting or secondary progressive MS were randomized into one of two 12-week exercise interventions: a supervised, moderate-intensity aerobic exercise (AE) program or an unsupervised, low-intensity walking and stretching (WS) program. Only individuals who were ≥ 70% compliant with the programs were included in analysis ( n  = 12 AE; n  = 10 WS). Both groups demonstrated a moderate improvement in sleep quality, although only the improvement by the WS group was statistically significant. Only the AE group demonstrated a significant improvement in daytime sleepiness. Change in sleep quality and daytime sleepiness was not correlated with disease severity or with change in cardiovascular fitness, depression, or fatigue. The mechanisms for improvement in sleep quality and daytime sleepiness need further investigation, but may be due to introduction of zeitgebers to improve circadian rhythm.

  11. Circadian melatonin rhythm and excessive daytime sleepiness in Parkinson disease.

    PubMed

    Videnovic, Aleksandar; Noble, Charleston; Reid, Kathryn J; Peng, Jie; Turek, Fred W; Marconi, Angelica; Rademaker, Alfred W; Simuni, Tanya; Zadikoff, Cindy; Zee, Phyllis C

    2014-04-01

    Diurnal fluctuations of motor and nonmotor symptoms and a high prevalence of sleep-wake disturbances in Parkinson disease (PD) suggest a role of the circadian system in the modulation of these symptoms. However, surprisingly little is known regarding circadian function in PD and whether circadian dysfunction is involved in the development of sleep-wake disturbances in PD. To determine the relationship between the timing and amplitude of the 24-hour melatonin rhythm, a marker of endogenous circadian rhythmicity, with self-reported sleep quality, the severity of daytime sleepiness, and disease metrics. A cross-sectional study from January 1, 2009, through December 31, 2012, of 20 patients with PD receiving stable dopaminergic therapy and 15 age-matched control participants. Both groups underwent blood sampling for the measurement of serum melatonin levels at 30-minute intervals for 24 hours under modified constant routine conditions at the Parkinson's Disease and Movement Disorders Center of Northwestern University. Twenty-four hour monitoring of serum melatonin secretion. Clinical and demographic data, self-reported measures of sleep quality (Pittsburgh Sleep Quality Index) and daytime sleepiness (Epworth Sleepiness Scale), and circadian markers of the melatonin rhythm, including the amplitude, area under the curve (AUC), and phase of the 24-hour rhythm. Patients with PD had blunted circadian rhythms of melatonin secretion compared with controls; the amplitude of the melatonin rhythm and the 24-hour AUC for circulating melatonin levels were significantly lower in PD patients (P < .001). Markers of the circadian phase were not significantly different between the 2 groups. Compared with PD patients without excessive daytime sleepiness, patients with excessive daytime sleepiness (Epworth Sleepiness Scale score ≥10) had a significantly lower amplitude of the melatonin rhythm and 24-hour melatonin AUC (P = .001). Disease duration, Unified Parkinson's Disease

  12. Caffeine: sleep and daytime sleepiness.

    PubMed

    Roehrs, Timothy; Roth, Thomas

    2008-04-01

    Caffeine is one of the most widely consumed psychoactive substances and it has profound effects on sleep and wake function. Laboratory studies have documented its sleep-disruptive effects. It clearly enhances alertness and performance in studies with explicit sleep deprivation, restriction, or circadian sleep schedule reversals. But, under conditions of habitual sleep the evidence indicates that caffeine, rather then enhancing performance, is merely restoring performance degraded by sleepiness. The sleepiness and degraded function may be due to basal sleep insufficiency, circadian sleep schedule reversals, rebound sleepiness, and/or a withdrawal syndrome after the acute, over-night, caffeine discontinuation typical of most studies. Studies have shown that caffeine dependence develops at relatively low daily doses and after short periods of regular daily use. Large sample and population-based studies indicate that regular daily dietary caffeine intake is associated with disturbed sleep and associated daytime sleepiness. Further, children and adolescents, while reporting lower daily, weight-corrected caffeine intake, similarly experience sleep disturbance and daytime sleepiness associated with their caffeine use. The risks to sleep and alertness of regular caffeine use are greatly underestimated by both the general population and physicians.

  13. Associations among daytime sleepiness, depression and suicidal ideation in Korean adolescents.

    PubMed

    Yang, Boksun; Choe, Kwisoon; Park, Youngrye; Kang, Youngmi

    2017-06-09

    The aim of this study was to examine the effects of daytime sleepiness on depression and suicidal ideation in adolescent high-school students. A survey of 538 high school students aged 16-17 years attending two academic schools was conducted. The Epworth Sleepiness Scale (ESS), the Beck Depression Inventory and the Scale for Suicide Ideation were used to assess subjects' daytime sleepiness, depression and suicidal ideation. The mean score for daytime sleepiness was 8.52, which indicates a sleep deficit. Significant positive correlations were found between daytime sleepiness and depression, between daytime sleepiness and suicidal ideation and between depression and suicidal ideation. Gender and depression were significant predictors of suicidal ideation, accounting for 48% of the variance in this measure. Depression acts as a mediator of the relationship between daytime sleepiness and suicidal ideation. High school students in Korea generally have insufficient sleep time and feel sleepy during the day; insufficient sleep during adolescence may be associated with depression and suicidal ideation.

  14. The comparison of nasal surgery and CPAP on daytime sleepiness in patients with OSAS.

    PubMed

    Tagaya, M; Otake, H; Suzuki, K; Yasuma, F; Yamamoto, H; Noda, A; Nishimura, Y; Sone, M; Nakashima, T; Nakata, S

    2017-09-01

    Residual sleepiness after continuous positive airway pressure (CPAP) is a critical problem in some patients with obstructive sleep apnea syndrome (OSAS). However, nasal surgery is likely to reduce daytime sleepiness and feelings of unrefreshed sleep. The aim of this study is to clarify the effects of nasal surgery and CPAP on daytime sleepiness. This is a retrospective and matched-case control study. The participants were consecutive 40 patients with OSAS who underwent nasal surgery (Surgery group) and 40 matched patients who were treated with CPAP (CPAP group). In the Surgery group, although the nasal surgery did not decrease either apnea or hypopnea, it improved oxygenation, the quality of sleep. In the CPAP Group, the CPAP treatment reduced apnea and hypopnea, and improved oxygenation, quality of sleep. The degree of relief from daytime sleepiness was different between the two groups. The improvement of Epworth Sleepiness Scale was more significant in the Surgery Group than those in the CPAP Group (Surgery from 11.0 to 5.1, CPAP from 10.0 to 6.2). These findings suggest that the results of the nasal surgery is more satisfactory for some patients with OSAS than CPAP on daytime sleepiness.

  15. Symptomatic endometriosis of the posterior cul-de-sac is associated with impaired sleep quality, excessive daytime sleepiness and insomnia: a case-control study.

    PubMed

    Leone Roberti Maggiore, Umberto; Bizzarri, Nicolò; Scala, Carolina; Tafi, Emanuela; Siesto, Gabriele; Alessandri, Franco; Ferrero, Simone

    2017-02-01

    To assess the impact of endometriosis of the posterior cul-de-sac on quality of sleep, average daytime sleepiness and insomnia. This age-matched case-control study was conducted in a tertiary referral centre for the diagnosis and treatment of endometriosis between May 2012 and December 2013. It included 145 women with endometriosis of the posterior cul-de-sac (cases; group E) and 145 patients referred to our Institution because of routine gynaecologic consultations (controls; group C). This study investigated whether sleep is impaired in patients with endometriosis of the posterior cul-de-sac. Sleep quality, daytime sleepiness and insomnia were assessed using the following self-administered questionnaires: the Pittsburgh Sleep Quality Index, the Epworth sleepiness scale and the Insomnia Severity Index, respectively. The primary objective of the study was to evaluate sleep quality in the two study groups. Secondary outcomes of the study were to assess average daytime sleepiness and insomnia in the two study groups. The prevalence of poor sleep quality was significantly higher in group E (64.8%) than in group C (15.1%; p<0.001). The prevalence of excessive daytime sleepiness was significantly higher in group E (23.4%) than in group C (12.9%; p=0.033). Patients of group E experienced subthreshold insomnia (29.0%) and moderate clinical insomnia (16.6%) significantly more frequently than patients in group C (24.4% and 5.0%; p=0.002). A substantial proportion of women with endometriosis of the posterior cul-de-sac experiences poor sleep quality, excessive daytime sleepiness and insomnia. Copyright © 2015 Elsevier Ireland Ltd. All rights reserved.

  16. Daytime Sleepiness, Poor Sleep Quality, Eveningness Chronotype, and Common Mental Disorders among Chilean College Students

    ERIC Educational Resources Information Center

    Concepcion, Tessa; Barbosa, Clarita; Vélez, Juan Carlos; Pepper, Micah; Andrade, Asterio; Gelaye, Bizu; Yanez, David; Williams, Michelle A.

    2014-01-01

    Objectives: To evaluate whether daytime sleepiness, poor sleep quality, and morningness and eveningness preferences are associated with common mental disorders (CMDs) among college students. Methods: A total of 963 college students completed self-administered questionnaires that collected information about sociodemographic characteristics, sleep…

  17. Is the Epworth Sleepiness Scale a useful tool for screening excessive daytime sleepiness in commercial drivers?

    PubMed

    Baiardi, Simone; La Morgia, Chiara; Sciamanna, Lucia; Gerosa, Alberto; Cirignotta, Fabio; Mondini, Susanna

    2018-01-01

    The significant social and economic impact of excessive daytime sleepiness makes sleep evaluation a primary medical need in commercial drivers. However, the best screening tool is still matter of debate. In our cohort of 221 commercial drivers, only ten (4.5%) had Epworth Sleepiness Scale scores indicative of excessive daytime sleepiness. These findings and the lack of concordance in estimating excessive daytime sleepiness among commercial drivers in previous studies using the same psychometric measure indicate that the Epworth Sleepiness Scale is not a reliable tool. This may be due to the low internal consistency of the scale in non-clinical samples and the possible intentional underscoring of sleepiness due to a perceived threat of driver's license suspension. Moreover, the reliability of the Epworth Sleepiness Scale results may be strongly influenced by the administration setting. The clinical application of inexpensive less time-consuming new tools like performance tests should be considered for the objective evaluation of excessive daytime sleepiness in commercial drivers. Copyright © 2017 Elsevier Ltd. All rights reserved.

  18. Daytime sleepiness, sleep habits and occupational accidents among hospital nurses.

    PubMed

    Suzuki, Kenshu; Ohida, Takashi; Kaneita, Yoshitaka; Yokoyama, Eise; Uchiyama, Makoto

    2005-11-01

    This paper reports a study to determine the prevalence of excessive daytime sleepiness and sleep habits among hospital nurses and to analyse associations between excessive daytime sleepiness and different types of medical error. It has been reported that sleep disorders, and the tiredness and sleepiness brought about by sleep disorders may be associated with occupational accidents. However, to our knowledge, there has so far been no report on associations between sleep disorders, excessive daytime sleepiness in particular, and occupational accidents among hospital nurses. The study was a cross-sectional study targeting 4407 nurses working in eight large general hospitals in Japan. An anonymous self-administered questionnaire was used to investigate their sleep patterns and experience of occupational accidents. The data were collected in 2003. The prevalence of excessive daytime sleepiness among hospital nurses in the present study was 26.0%. A statistically significant relationship was observed between having or not having occupational accidents during the past 12 months and excessive daytime sleepiness. Multiple logistic regression analyses on factors leading to occupational accidents during the past 12 months showed statistically significant associations between (1) drug administration errors and (2) shift work and age, between (1) incorrect operation of medical equipment and (2) excessive daytime sleepiness and age, and between needlestick injuries and age. Excessive daytime sleepiness is an important occupational health issue in hospital nurses. It is possible that occupational policies and health promotion measures, such as a provision of sleep hygiene advice and social support at worksites, would be effective in preventing occupational accidents among hospital nurses.

  19. Parents of children referred to a sleep laboratory for disordered breathing reported anxiety, daytime sleepiness and poor sleep quality.

    PubMed

    Cadart, Marion; De Sanctis, Livio; Khirani, Sonia; Amaddeo, Alessandro; Ouss, Lisa; Fauroux, Brigitte

    2018-07-01

    We evaluated the impact that having a child with sleep-disordered breathing had on their parents, including their own sleep quality. Questionnaires were completed by 96 parents of 86 children referred for a sleep study or control of continuous positive airway pressure (CPAP) or noninvasive ventilation (NIV) at the sleep laboratory of the Necker Hospital, Paris, France, between October 2015 and January 2016. The questionnaires evaluated anxiety and depression, family functioning, the parents' quality of life, daytime sleepiness and sleep quality. The children had a mean age of seven ±five years and most of the responses (79%) came from their mothers. These showed that 26% of parents showed moderate-to-severe anxiety, 8% moderate-to-severe depression, 6% complex family cohesion, 59% moderate-to-severe daytime sleepiness and 54% poor sleep quality. Anxiety was higher in mothers than in fathers (p < 0.001). The questionnaire scores did not differ according to the child's age, the results of the sleep studies or the CPAP or NIV treatment. The symptoms seem to be more commonly related to the child's underlying disease than their sleep-disordered breathing. The parents of children referred to a sleep laboratory reported frequent anxiety, daytime sleepiness and poor sleep quality. ©2018 Foundation Acta Paediatrica. Published by John Wiley & Sons Ltd.

  20. Wake-up stroke: Clinical characteristics, sedentary lifestyle, and daytime sleepiness.

    PubMed

    Diniz, Deborath Lucia de Oliveira; Barreto, Pedro Rodrigues; Bruin, Pedro Felipe Carvalhedo de; Bruin, Veralice Meireles Sales de

    2016-10-01

    Wake-up stroke (WUS) is defined when the exact time of the beginning of the symptoms cannot be determined, for the deficits are perceived upon awakening. Sleep alterations are important risk factors for stroke and cardiovascular diseases. This study evaluates the characteristics of patients with and without WUS, the presence of daytime sleepiness, and associated risk factors. Patients with ischemic stroke were investigated about the presence of WUS. Clinical and demographic characteristics were evaluated. Stroke severity was studied by the National Institutes of Health Stroke Scale (NIHSS) and the Modified Rankin Scale (MRS), and daytime sleepiness severity was studied by the Epworth Sleepiness Scale (ESS). Seventy patients (57.1% men) aged from 32 to 80 years (58.5±13.3) were studied. WUS was observed in 24.3%. Arterial hypertension (67.1%), type 2 diabetes (27.1%), and hyperlipidemia (22.8%) were frequent. Type 2 diabetes and sedentary lifestyle were more common in patients with WUS (p<0.05). Overall, mild, moderate or very few symptoms of stroke (NIHSS<5) were predominant (62.3%). Among all cases, 20% had excessive daytime sleepiness (ESS>10). No differences were found between patients with and without WUS as regards stroke severity or excessive daytime sleepiness. Patients with excessive daytime sleepiness were younger and had more sedentary lifestyle (p<0.05). Individuals with previous history of heavy drinking had more daytime sleepiness (p=0.03). Wake-up stroke occurs in approximately 25% of stroke cases. In this study, patients with WUS had more diabetes and sedentary lifestyle. Daytime sleepiness is frequent and is associated with sedentary lifestyle and heavy drinking.

  1. Daytime Sleepiness in Obesity: Mechanisms Beyond Obstructive Sleep Apnea—A Review

    PubMed Central

    Panossian, Lori A.; Veasey, Sigrid C.

    2012-01-01

    Increasing numbers of overweight children and adults are presenting to sleep medicine clinics for evaluation and treatment of sleepiness. Sleepiness negatively affects quality of life, mental health, productivity, and safety. Thus, it is essential to comprehensively address all potential causes of sleepiness. While many obese individuals presenting with hypersomnolence will be diagnosed with obstructive sleep apnea and their sleepiness will improve with effective therapy for sleep apnea, a significant proportion of patients will continue to have hypersomnolence. Clinical studies demonstrate that obesity without sleep apnea is also associated with a higher prevalence of hypersomnolence and that bariatric surgery can markedly improve hypersomnolence before resolution of obstructive sleep apnea. High fat diet in both humans and animals is associated with hypersomnolence. This review critically examines the relationships between sleepiness, feeding, obesity, and sleep apnea and then discusses the hormonal, metabolic, and inflammatory mechanisms potentially contributing to hypersomnolence in obesity, independent of sleep apnea and other established causes of excessive daytime sleepiness. Citation: Panossian LA; Veasey SC. Daytime sleepiness in obesity: mechanisms beyond obstructive sleep apnea—a review. SLEEP 2012;35(5):605-615. PMID:22547886

  2. Smartphone addiction risk and daytime sleepiness in Korean adolescents.

    PubMed

    Chung, Jee Eun; Choi, Soo An; Kim, Ki Tai; Yee, Jeong; Kim, Joo Hee; Seong, Jin Won; Seong, Jong Mi; Kim, Ju Young; Lee, Kyung Eun; Gwak, Hye Sun

    2018-04-06

    Smartphone overuse can cause not only mobility problems in the wrists, fingers and neck but also interference with sleep habits. However, research on smartphone addiction and sleep disturbances is scarce. Therefore, we aimed to investigate daytime sleepiness in association with smartphone addiction risk in Korean adolescents. A cross-sectional survey method was used in this study. The Pediatric Daytime Sleepiness Scale was used to assess daytime sleepiness, and the Korean Smartphone Addiction Proneness Scale index was used to evaluate the degree of risk for smartphone addiction. The analyses were performed in 1796 adolescents using smartphones, including 820 boys and 976 girls. The at-risk smartphone users made up 15.1% of boys and 23.9% of girls. Our multivariate analyses demonstrated that students who were female, consumed alcohol, had lower academic performance, did not feel refreshed in the morning and initiated sleep after 12 am were at a significantly higher risk of smartphone addiction. The at-risk smartphone user group was independently associated with the upper quartile Pediatric Daytime Sleepiness Scale score in students with the following factors: Female gender, alcohol consumption, poor self-perceived health level, initiating sleep after 12 am, longer time taken to fall asleep and duration of night sleep less than 6 h. The quality of sleep in adolescence affects growth, emotional stability and learning skills. Therefore, the management of smartphone addiction seems to be essential for proper sleeping habits. There is a critical need to develop a means of preventing smartphone addiction on a social level. © 2018 Paediatrics and Child Health Division (The Royal Australasian College of Physicians).

  3. Daytime sleepiness, exercise, and physical function in older adults.

    PubMed

    Chasens, Eileen R; Sereika, Susan M; Weaver, Terri E; Umlauf, Mary Grace

    2007-03-01

    The purpose of this study was to describe the association between sleepiness, exercise, and physical function in older adults, testing the hypothesis that sleepiness predicts decreased exercise and impaired physical function in this population. We performed a secondary analysis of data from the National Sleep Foundation's Sleep in America Poll, comparing frequency of exercise and ability to perform functional tasks between sleepy and non-sleepy subjects. Trained interviewers administered a scripted telephone survey. Participants (n = 1506) were community-dwelling older Americans (55-84 years) randomly chosen from geographically representative households with listed telephone numbers. Sleepiness 'so severe that it interferes with daytime activity' was dichotomized as 'daily/frequently' or 'never/rare'. Exercise frequency was scored 1-4 ('less than once a week' to 'more than five times a week'). Responses to five questions (walk 0.5 mile, climb stairs, push/pull heavy object, stoop/crouch/or kneel, write, handle small objects), rated 1-5 ('no difficulty' to 'unable to do'), were summed; a mean score of > or = 2.5 was considered impaired physical function. Daytime sleepiness predicted low exercise frequency while controlling for age and body mass index (BMI) (OR = 1.40, 95% CI 1.031-1.897, P = 0.031). Frequent daytime sleepiness predicted impaired physical function (OR = 2.76, 95%CI = 0.237-0.553, P = 0.001) after controlling for age, BMI, income and number of co-morbid conditions. The conclusion was that daytime sleepiness in older adults is associated with physical functional impairments and decreased exercise frequency. The findings suggest that sleepiness in older adults is not benign but has implications for continued physical decline and warrants attention.

  4. Objective daytime sleepiness in patients with somnambulism or sleep terrors.

    PubMed

    Lopez, Régis; Jaussent, Isabelle; Dauvilliers, Yves

    2014-11-25

    To objectively measure daytime sleepiness and to assess for clinical and polysomnographic determinants of mean sleep latency in adult patients with somnambulism (sleepwalking [SW]) or sleep terrors (ST) compared with controls. Thirty drug-free adult patients with primary SW or ST, and age-, sex-, and body mass index-matched healthy controls underwent a standardized clinical interview, completed questionnaires including the Epworth Sleepiness Scale, and underwent one night of video polysomnography followed by the Multiple Sleep Latency Test (MSLT). Excessive daytime sleepiness defined as Epworth Sleepiness Scale score >10 was reported in 66.7% of patients and 6.7% of controls. The temporal pattern of sleep latencies in individual MSLT trials differed between patients and controls, with progressive increased sleep latency in patients across the trials in contrast to a "U curve" for controls. We did not find between-group differences regarding the mean sleep latency on the 5 MSLT trials, but did observe reduced sleep latencies in patients for the first 2 trials. Despite increased slow-wave sleep disruptions found in patients (i.e, more micro-arousals and hypersynchronous high-voltage delta waves arousals), we did not find polysomnographic characteristic differences when comparing sleepy patients for either subjective or objective daytime sleepiness on the MSLT compared with alert patients. Excessive daytime sleepiness is a common complaint in subjects with SW or ST and shorter sleep latencies in the early morning hours. Despite an increased slow-wave sleep fragmentation found in these patients, we did not identify any association with the level of daytime sleepiness. © 2014 American Academy of Neurology.

  5. Home exercise improves the quality of sleep and daytime sleepiness of elderlies: a randomized controlled trial.

    PubMed

    Brandão, Glauber Sá; Gomes, Glaucia Sá Brandão Freitas; Brandão, Glaudson Sá; Callou Sampaio, Antônia A; Donner, Claudio F; Oliveira, Luis V F; Camelier, Aquiles Assunção

    2018-01-01

    Aging causes physiological changes which affect the quality of sleep. Supervised physical exercise is an important therapeutic resource to improve the sleep of the elderlies, however there is a low adherence to those type of programs, so it is necessary to implement an exercise program which is feasible and effective. The study aimed to test the hypothesis that a semi-supervised home exercise program, improves sleep quality and daytime sleepiness of elderlies of the community who present poor sleep quality. This was a randomized controlled trial study, conducted from May to September 2017, in Northeastern Brazil, with elderlies of the community aging 60 years old or older, sedentary, with lower scores or equal to 5 at the Pittsburgh Sleep Quality Index (PSQI) and without cognitive decline. From one hundred ninety-one potential participants twenty-eight refused to participate, therefore, one hundred thirty-one (mean age 68 ± 7 years), and 88% female, were randomly assigned to an intervention group - IG (home exercise and sleep hygiene, n  = 65) and a control group - CG (sleep hygiene only, n  = 66). Sleep assessment tools were used: PSQI, Epworth sleepiness scale (ESS) and clinical questionnaire of Berlin. The level of physical activity has been assessed by means of International Physical Activity Questionnaire adapted for the elderly (IPAQ) and Mini-Mental State Examination for cognitive decline. All participants were assessed before and after the 12-week intervention period and, also, the assessors were blind. The IG showed significant improvement in quality of sleep with a mean reduction of 4.9 ± 2.7 points in the overall PSQI ( p  < 0.01) and in all its 7 components of evaluation ( p  < 0.05), and improvement of secondary endpoint, daytime sleepiness, a decline of 2.8 ± 2.2 points in the ESS (p < 0.01). Our results suggest that semi-supervised home exercise is effective in improving the quality of sleep and self

  6. 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.

  7. 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.

  8. Sleep and daytime sleepiness of patients with left ventricular assist devices: a longitudinal pilot study.

    PubMed

    Casida, Jesus M; Davis, Jean E; Brewer, Robert J; Smith, Cheryl; Yarandi, Hossein

    2011-06-01

    No empirical longitudinal data on sleep and daytime sleepiness patterns in patients with an implantable left ventricular assist device (LVAD) exist. (1) To describe the sleep patterns (sleep onset latency, sleep efficiency, sleep fragmentation index, total sleep time, and wake after sleep onset), sleep quality, and daytime sleepiness variables and (2) to determine the change in the pattern of these variables before and up to 6 months after LVAD implantation. A longitudinal descriptive repeated-measures design was used. Patients wore wrist actigraphs (AW64 Actiwatch), which objectively measured sleep, for 3 consecutive days and nights before LVAD implant and at the first and second week and first, third, and sixth month after implantation. During these periods, patients also completed questionnaires on sleep quality and daytime sleepiness. Patients-Twelve of 15 patients completed the 6-month data. Data were analyzed by using descriptive statistics and repeated-measures analysis of variance. We found long sleep onset latencies and low sleep efficiency across time periods. High sleep fragmentation index was noted at baseline and 1 week after LVAD. Short total sleep times, long wake-after-sleep-onset durations, and poor sleep quality were evident at baseline and persisted up to 6 months after LVAD implantation. Low alertness level, a manifestation of sleepiness, was common during late morning to early evening hours. However, only sleep efficiency and wake after sleep onset showed significant changes in pattern (P < .05). Sleep disturbance and daytime sleepiness may be prevalent before and up to 6 months after LVAD implantation, warranting further investigation.

  9. Excessive daytime sleepiness and metabolic syndrome in men with obstructive sleep apnea: a large cross-sectional study.

    PubMed

    Fu, Yiqun; Xu, Huajun; Xia, Yunyan; Qian, Yingjun; Li, Xinyi; Zou, Jianyin; Wang, Yuyu; Meng, Lili; Tang, Xulan; Zhu, Huaming; Zhou, Huiqun; Su, Kaiming; Yu, Dongzhen; Yi, Hongliang; Guan, Jian; Yin, Shankai

    2017-10-03

    Excessive daytime sleepiness is a common symptom in obstructive sleep apnea (OSA). Previous studies have showed that excessive daytime sleepiness is associated with some individual components of metabolic syndrome. We performed a large cross-sectional study to explore the relationship between excessive daytime sleepiness and metabolic syndrome in male OSA patients. A total of 2241 suspected male OSA patients were consecutively recruited from 2007 to 2013. Subjective daytime sleepiness was assessed using the Epworth sleepiness scale. Anthropometric, metabolic, and polysomnographic parameters were measured. Metabolic score was used to evaluate the severity of metabolic syndrome. Among the male OSA patients, most metabolic parameters varied by excessive daytime sleepiness. In the severe group, male OSA patients with excessive daytime sleepiness were more obese, with higher blood pressure, more severe insulin resistance and dyslipidemia than non-sleepy patients. Patients with metabolic syndrome also had a higher prevalence of excessive daytime sleepiness and scored higher on the Epworth sleepiness scale. Excessive daytime sleepiness was independently associated with an increased risk of metabolic syndrome (odds ratio =1.242, 95% confidence interval: 1.019-1.512). No substantial interaction was observed between excessive daytime sleepiness and OSA/ obesity. Excessive daytime sleepiness was related to metabolic disorders and independently associated with an increased risk of metabolic syndrome in men with OSA. Excessive daytime sleepiness should be taken into consideration for OSA patients, as it may be a simple and useful clinical indicator for evaluating the risk of metabolic syndrome.

  10. The pediatric daytime sleepiness scale (PDSS): sleep habits and school outcomes in middle-school children.

    PubMed

    Drake, Christopher; Nickel, Chelsea; Burduvali, Eleni; Roth, Thomas; Jefferson, Catherine; Pietro, Badia

    2003-06-15

    To develop a measure of daytime sleepiness suitable for middle-school children and examine the relationship between daytime sleepiness and school-related outcomes. Self-report questionnaire. Four hundred fifty, 11- to 15-year-old students, from grades 6, 7, and 8 of a public middle school in Dayton, Ohio. A pediatric daytime sleepiness questionnaire was developed using factor analysis of questions regarding sleep-related behaviors. Results of the sleepiness questionnaire were then compared across other variables, including daily sleep patterns, school achievement, mood, and extracurricular activities. Factor analysis on the 13 questions related to daytime sleepiness yielded 1 primary factor ("pediatric daytime sleepiness"; 32% of variance). Only items with factor loadings above .4 were included in the final sleepiness scale. Internal consistency (Chronbach's alpha) for the final 8-item scale was .80. Separate one-way analyses of variance and trend analyses were performed comparing pediatric daytime sleepiness scores at the 5 different levels of total sleep time and academic achievement. Participants who reported low school achievement, high rates of absenteeism, low school enjoyment, low total sleep time, and frequent illness reported significantly higher levels of daytime sleepiness compared to children with better school-related outcomes. The self-report scale developed in the present work is suitable for middle-school-age children and may be useful in future research given its ease of administration and robust psychometric properties. Daytime sleepiness is related to reduced educational achievement and other negative school-related outcomes.

  11. Morningness/eveningness chronotype, poor sleep quality, and daytime sleepiness in relation to common mental disorders among Peruvian college students.

    PubMed

    Rose, Deborah; Gelaye, Bizu; Sanchez, Sixto; Castañeda, Benjamín; Sanchez, Elena; Yanez, N David; Williams, Michelle A

    2015-01-01

    The study was designed to investigate the association between sleep disturbances and common mental disorders (CMDs) among Peruvian college students. A total of 2538 undergraduate students completed a self-administered questionnaire to gather information about sleep characteristics, sociodemographic, and lifestyle data. Evening chronotype, sleep quality, and daytime sleepiness were assessed using the Horne and Ostberg morningness-eveningness questionnaire, Pittsburgh Sleep Quality Index, and Epworth Sleepiness Scale, respectivelty. Presence of CMDs was evaluated using the General Health Questionnaire. Logistic regression procedures were used to examine the associations of sleep disturbances with CMDs while accounting for possible confounding factors. Overall, 32.9% of the participants had prevalent CMDs (39.3% among females and 24.4% among males). In multivariable-adjusted logistic models, those with evening chronotype (odds ratios (OR) = 1.43; 95% CI 1.00-2.05), poor sleep quality (OR = 4.50; 95% CI 3.69-5.49), and excessive daytime sleepiness (OR = 1.68; 95% CI 1.41-2.01) were at a relative increased odds of CMDs compared with those without sleep disturbances. In conclusion, we found strong associations between sleep disturbances and CMDs among Peruvian college students. Early education and preventative interventions designed to improve sleep habits may effectively alter the possibility of developing CMDs among young adults.

  12. Acute modafinil exposure reduces daytime sleepiness in abstinent methamphetamine-dependent volunteers

    PubMed Central

    Mahoney, James J.; Jackson, Brian J.; Kalechstein, Ari D.; De La Garza, Richard; Chang, Lee C.; Newton, Thomas F.

    2012-01-01

    The purpose of this study was to evaluate the effects of acute, oral modafinil (200 mg) exposure on daytime sleepiness in methamphetamine (Meth)-dependent individuals. Eighteen Meth-dependent subjects were enrolled in a 7-d inpatient study and were administered placebo or modafinil on day 6 and the counter-condition on day 7 (randomized) of the protocol. Subjects completed several subjective daily assessments (such as the Epworth Sleepiness Scale, Pittsburgh Sleep Quality Index, Beck Depression Inventory and visual analogue scale) throughout the protocol as well as objective assessments on days 5–7, when the Multiple Sleep Latency Test was performed. The results of the current study suggest that short-term abstinence from Meth is associated with increased daytime sleepiness and that a single dose of 200 mg modafinil reduces daytime somnolence in this population. In addition, a positive correlation was found between subjective reporting of the likelihood of taking a nap and craving and desire for Meth, as well as the likelihood of using Meth and whether Meth would make the participant feel better. The results of this study should be considered when investigating candidate medications for Meth-dependence, especially in those individuals who attribute their Meth use to overcoming deficits resulting from sleep abnormalities. PMID:22214752

  13. Acute modafinil exposure reduces daytime sleepiness in abstinent methamphetamine-dependent volunteers.

    PubMed

    Mahoney, James J; Jackson, Brian J; Kalechstein, Ari D; De La Garza, Richard; Chang, Lee C; Newton, Thomas F

    2012-10-01

    The purpose of this study was to evaluate the effects of acute, oral modafinil (200 mg) exposure on daytime sleepiness in methamphetamine (Meth)-dependent individuals. Eighteen Meth-dependent subjects were enrolled in a 7-d inpatient study and were administered placebo or modafinil on day 6 and the counter-condition on day 7 (randomized) of the protocol. Subjects completed several subjective daily assessments (such as the Epworth Sleepiness Scale, Pittsburgh Sleep Quality Index, Beck Depression Inventory and visual analogue scale) throughout the protocol as well as objective assessments on days 5-7, when the Multiple Sleep Latency Test was performed. The results of the current study suggest that short-term abstinence from Meth is associated with increased daytime sleepiness and that a single dose of 200 mg modafinil reduces daytime somnolence in this population. In addition, a positive correlation was found between subjective reporting of the likelihood of taking a nap and craving and desire for Meth, as well as the likelihood of using Meth and whether Meth would make the participant feel better. The results of this study should be considered when investigating candidate medications for Meth-dependence, especially in those individuals who attribute their Meth use to overcoming deficits resulting from sleep abnormalities.

  14. Daytime sleepiness and insomnia as correlates of depression.

    PubMed

    Fava, Maurizio

    2004-01-01

    Insomnia and daytime sleepiness are often associated with depression. The possible relationships between sleep difficulties and depression are numerous. Insomnia and other sleep disturbances can be precursors to the onset of major depressive disorder, so they may act as risk factors for or predictors of depression. The symptomatology of depression also prominently includes insomnia, and sleep disturbances may be residual symptoms after response to antidepressant treatment. Insomnia and the resultant daytime sleepiness may be short-term or long-term side effects of antidepressant treatment as well. Whether insomnia is a precursor, symptom, residual symptom, or side effect of depression or its treatment, clinicians must give serious attention to and attempt to resolve sleep disturbances because of the risk of depression onset, worsening of depressive symptoms, and relapse of depression after response to antidepressant treatment. Remission of depression cannot be fully achieved until the associated insomnia and daytime sleepiness are resolved. This article describes the relationships between insomnia and depression and discusses the effects of various antidepressants on sleep. Finally, several different treatment options, including antidepressant monotherapy and augmentation of antidepressants with other medications, are explored.

  15. [Excessive daytime sleepiness--as possible suicidal equivalent in people with depressive disorder].

    PubMed

    Koić, Elvira; Strkalj-Ivezić, Sladana; Venus, Miroslav; John, Nada

    2011-01-01

    This paper presents the case of a patient suffering from atypical depression, with excessive daytime sleepiness, which was a suicidal equivalent. Patients who suffer from depression and sleep disorders can have an increased risk of suicidality, and to them should pay increased attention. Field of suicide as a means of resolving conflicts and unconscious mental escape from the reality of the obstacles, including sleep and daytime sleepiness, which may represent a form of "temporary suicide". The authors recommend that hypersomnia, daytime sleepiness in depressed patients and other sleep disorders should be treated as a potential risk factor for suicidal behavior.

  16. [Health-related consequences of obstructive sleep apnea: daytime sleepiness, accident risk and legal aspects].

    PubMed

    Orth, M; Kotterba, S

    2012-04-01

    Daytime sleepiness for any reason leads to impairment of daytime performance and an increased accident rate. The consequences are an increase of illness- and accident-related costs for the health system. Obstructive sleep apnea (OSA) is one of the major reasons for increased daytime sleepiness, especially in professional drivers. The accident frequency in OSA can be significantly reduced by adequate continuous positive airway pressure (CPAP) therapy. Up till now there are no uniform legal regulations about the handling of OSAS patients or patients with daytime sleepiness due to other diseases as far as driving ability is concerned.

  17. [Excessive Daytime Sleepiness, Poor Quality Sleep, and Low Academic Performance in Medical Students].

    PubMed

    Machado-Duque, Manuel Enrique; Echeverri Chabur, Jorge Enrique; Machado-Alba, Jorge Enrique

    2015-01-01

    Quality of sleep and excessive daytime sleepiness (EDS) affect cognitive ability and performance of medical students. This study attempts to determine the prevalence of EDS, sleep quality, and assess their association with poor academic performance in this population. A descriptive, observational study was conducted on a random sample of 217 medical students from the Universidad Tecnológica de Pereira, who completed the Pittsburgh Sleep Quality Index (PSQI) questionnaire and the Epworth sleepiness scale. Sociodemographic, clinic and academic variables were also measured. Multivariate analyses for poor academic performance were performed. The included students had a mean age of 21.7±3.3 years, of whom 59.4% were men. Almost half (49.8%) had EDS criteria, and 79.3% were poor sleepers (PSQI ≥ 5), while 43.3% had poor academic performance during the last semester. The bivariate analysis showed that having used tobacco or alcohol until intoxicated, fairly bad subjective sleep quality, sleep efficiency < 65%, and being a poor sleeper were associated with increased risk of low performance. Sleep efficiency < 65% was statistically associated with poor academic performance (P=.024; OR = 4.23; 95% CI, 1.12-15.42) in the multivariate analysis. A poor sleep quality determined by low efficiency was related to poor academic achievement at the end of semester in medical students. Copyright © 2015 Asociación Colombiana de Psiquiatría. Publicado por Elsevier España. All rights reserved.

  18. Daytime Sleepiness Increases With Age in Early Adolescence: A Sleep Restriction Dose-Response Study.

    PubMed

    Campbell, Ian G; Burright, Christopher S; Kraus, Amanda M; Grimm, Kevin J; Feinberg, Irwin

    2017-05-01

    Daytime sleepiness increases across adolescence. This increase is commonly attributed to insufficient sleep durations resulting from increasingly limited time in bed. We tested the effects of 3 sleep schedules on daytime sleepiness and whether these effects changed with age in early adolescence. In 77 children ranging in age from 9.9 to 14 years, objective (multiple sleep latency test [MSLT]) and subjective (Karolinska sleepiness scale [KSS]) sleepiness was measured following 4 consecutive nights of either 7, 8.5, or 10 hours in bed. All participants completed all 3 sleep schedules. The order in which they completed the schedules was not randomized but was accounted for in all statistical analyses. Time in bed restriction decreased sleep duration and increased objective and subjective daytime sleepiness. Although the sleep durations did not change with age, the likelihood of falling asleep during the MSLT increased with age. Nevertheless, sleep restriction produced a greater increase in MSLT-measured sleepiness in younger participants. Subjective sleepiness measured with the KSS increased with shorter sleep duration, but this effect did not change with age. Increasing objective daytime sleepiness in early adolescence cannot simply be attributed to reduced sleep due to restricted sleep schedules. We propose that some of the increased daytime sleepiness of adolescents is a consequence of adolescent brain reorganization driven by synaptic pruning which decreases the intensity of waking brain activity. © Sleep Research Society 2017. 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. 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.

  20. The prevalence of excessive daytime sleepiness among academic physicians and its impact on the quality of life and occupational performance.

    PubMed

    Ozder, Aclan; Eker, Hasan Huseyin

    2015-01-01

    Sleep disorders can affect health and occupational performance of physicians as well as outcomes in patients. The purpose of this study was to assess the prevalence of excessive daytime sleepiness (EDS) measured by the Epworth Sleepiness Scale (ESS) among academic physicians at a tertiary academic medical center in an urban area in the northwest region of Turkey, and to establish a relationship between the self-perceived sleepiness and the quality of life using the EuroQol-5 dimensions (EQ-5D). A questionnaire prepared by the researchers after scanning the literature on the subject was e-mailed to the academic physicians of a tertiary academic medical center in Istanbul. The ESS and the EQ-5D were also included in the survey. The e-mail database of the institution directory was used to compile a list of active academic physicians who practiced clinical medicine. Paired and independent t tests were used for the data analysis at a significance level of p < 0.05. Three hundred and ninety six academic physicians were e-mailed and a total of 252 subjects replied resulting in a 63.6% response rate. There were 84 (33.3%) female and 168 (66.7%) male academic physicians participating in the study. One hundred and eight out of 252 (42.8%) academic physicians were taking night calls (p < 0.001). Ninety study subjects (35.7%) felt they had enough sleep and 84 (33.3%) reported napping daily (p < 0.001). In our sample, 28.6% (N = 72) of the physicians felt sleepy during the day (ESS score > 10) (p < 0.001). In the case of the EQ-5D index and visual analogue scale of the EQ-5D questionnaire (EQ-5D VAS), the status of sleepiness of academic physicians was associated with a poorer quality of life (p < 0.001). More than a 1/4 of the academic physicians suffered from sleepiness. There was an association between the poor quality of life and daytime sleepiness. There was also a positive relationship between habitual napping and being sleepy during the day. This work is available in

  1. Association between Screen Viewing Duration and Sleep Duration, Sleep Quality, and Excessive Daytime Sleepiness among Adolescents in Hong Kong

    PubMed Central

    Mak, Yim Wah; Wu, Cynthia Sau Ting; Hui, Donna Wing Shun; Lam, Siu Ping; Tse, Hei Yin; Yu, Wing Yan; Wong, Ho Ting

    2014-01-01

    Screen viewing is considered to have adverse impacts on the sleep of adolescents. Although there has been a considerable amount of research on the association between screen viewing and sleep, most studies have focused on specific types of screen viewing devices such as televisions and computers. The present study investigated the duration with which currently prevalent screen viewing devices (including televisions, personal computers, mobile phones, and portable video devices) are viewed in relation to sleep duration, sleep quality, and daytime sleepiness among Hong Kong adolescents (N = 762). Television and computer viewing remain prevalent, but were not correlated with sleep variables. Mobile phone viewing was correlated with all sleep variables, while portable video device viewing was shown to be correlated only with daytime sleepiness. The results demonstrated a trend of increase in the prevalence and types of screen viewing and their effects on the sleep patterns of adolescents. PMID:25353062

  2. Association between screen viewing duration and sleep duration, sleep quality, and excessive daytime sleepiness among adolescents in Hong Kong.

    PubMed

    Mak, Yim Wah; Wu, Cynthia Sau Ting; Hui, Donna Wing Shun; Lam, Siu Ping; Tse, Hei Yin; Yu, Wing Yan; Wong, Ho Ting

    2014-10-28

    Screen viewing is considered to have adverse impacts on the sleep of adolescents. Although there has been a considerable amount of research on the association between screen viewing and sleep, most studies have focused on specific types of screen viewing devices such as televisions and computers. The present study investigated the duration with which currently prevalent screen viewing devices (including televisions, personal computers, mobile phones, and portable video devices) are viewed in relation to sleep duration, sleep quality, and daytime sleepiness among Hong Kong adolescents (N = 762). Television and computer viewing remain prevalent, but were not correlated with sleep variables. Mobile phone viewing was correlated with all sleep variables, while portable video device viewing was shown to be correlated only with daytime sleepiness. The results demonstrated a trend of increase in the prevalence and types of screen viewing and their effects on the sleep patterns of adolescents.

  3. Pharmacological interventions for daytime sleepiness and sleep disorders in Parkinson's disease: Systematic review and meta-analysis.

    PubMed

    Rodrigues, Tiago Martins; Castro Caldas, Ana; Ferreira, Joaquim J

    2016-06-01

    Daytime sleepiness and sleep disorders are frequently reported in Parkinson's disease (PD). However, their impact on quality of life has been underestimated and few clinical trials have been performed. We aimed to assess the efficacy and safety of pharmacological interventions for daytime sleepiness and sleep disorders in PD. Systematic review of randomized controlled trials comparing any pharmacological intervention with no intervention or placebo for the treatment of daytime sleepiness and sleep problems in PD patients. Ten studies (n = 338 patients) were included. Four trials addressed interventions for excessive daytime sleepiness. Meta-analysis of the three trials evaluating modafinil showed a significant reduction in sleepiness, as assessed by the Epworth Sleepiness Scale (ESS) (- 2.24 points, 95% CI - 3.90 to - 0.57, p < 0.05). In one study, treatment with caffeine was associated with a non-significant improvement of 1.71 points in ESS (95% CI, - 3.57 to 0.13). The six remaining trials assessed interventions for insomnia and REM sleep Behaviour Disorder (RBD). Single study results suggest that doxepin and YXQN granules might be efficacious, while pergolide may be deleterious for insomnia and that rivastigmine may be used to treat RBD in PD patients. However, there is insufficient evidence to support or refute the efficacy of any of these interventions. No relevant side effects were reported. Whilst providing recommendations, this systematic review depicts the lack of a body of evidence regarding the treatment of sleep disorders in PD patients; hence, further studies are warranted. Copyright © 2016 Elsevier Ltd. All rights reserved.

  4. Sleep Disordered Breathing And Daytime Sleepiness Are Associated With Poor Academic Performance In Teenagers. A Study Using The Pediatric Daytime Sleepiness Scale (PDSS)

    PubMed Central

    Perez-Chada, Daniel; Perez-Lloret, Santiago; Videla, Alejandro J.; Cardinali, Daniel; Bergna, Miguel A.; Fernández-Acquier, Mariano; Larrateguy, Luis; Zabert, Gustavo E.; Drake, Christopher

    2007-01-01

    Study Objectives: Inadequate sleep and sleep disordered breathing (SDB) can impair learning skills. Questionnaires used to evaluate sleepiness in adults are usually inadequate for adolescents. We conducted a study to evaluate the performance of a Spanish version of the Pediatric Daytime Sleepiness Scale (PDSS) and to assess the impact of sleepiness and SDB on academic performance. Design: A cross-sectional survey of students from 7 schools in 4 cities of Argentina. Measurements: A questionnaire with a Spanish version of the PDSS was used. Questions on the occurrence of snoring and witnessed apneas were answered by the parents. Mathematics and language grades were used as indicators of academic performance. Participants: The sample included 2,884 students (50% males; age: 13.3 ± 1.5 years) Results: Response rate was 85%; 678 cases were excluded due to missing data. Half the students slept <9 h per night on weekdays. The mean PDSS value was 15.74 ± 5.93. Parental reporting of snoring occurred in 511 subjects (23%); snoring was occasional in 14% and frequent in 9%. Apneas were witnessed in 237 cases (11%), being frequent in 4% and occasional in 7%. Frequent snorers had higher mean PDSS scores than occasional or nonsnorers (18 ± 5, 15.7 ± 6 and 15.5 ± 6, respectively; P < 0.001). Reported snoring or apneas and the PDSS were significant univariate predictors of failure and remained significant in multivariate logistic regression analysis after adjusting for age, sex, body mass index, specific school attended, and sleep habits. Conclusions: Insufficient hours of sleep were prevalent in this population. The Spanish version of the PDSS was a reliable tool in middle–school-aged children. Reports of snoring or witnessed apneas and daytime sleepiness as measured by PDSS were independent predictors of poor academic performance. Citation: Perez-Chada D; Perez-Lloret S; Videla AJ; Cardinali D; Bergna MA; Fernández-Acquier M; Larrateguy L; Zabert GE; Drake C. Sleep

  5. Does Excessive Daytime Sleepiness Affect Children's Pedestrian Safety?

    PubMed Central

    Avis, Kristin T.; Gamble, Karen L.; Schwebel, David C.

    2014-01-01

    Study Objectives: Many cognitive factors contribute to unintentional pedestrian injury, including reaction time, impulsivity, risk-taking, attention, and decision-making. These same factors are negatively influenced by excessive daytime sleepiness (EDS), which may place children with EDS at greater risk for pedestrian injury. Design, Participants, and Methods: Using a case-control design, 33 children age 8 to 16 y with EDS from an established diagnosis of narcolepsy or idiopathic hypersomnia (IHS) engaged in a virtual reality pedestrian environment while unmedicated. Thirty-three healthy children matched by age, race, sex, and household income served as controls. Results: Children with EDS were riskier pedestrians than healthy children. They were twice as likely to be struck by a virtual vehicle in the virtual pedestrian environment than healthy controls. Attentional skills of looking at oncoming traffic were not impaired among children with EDS, but decision-making for when to cross the street safely was significantly impaired. Conclusions: Results suggest excessive daytime sleepiness (EDS) from the clinical sleep disorders known as the hypersomnias of central origin may have significant consequences on children's daytime functioning in a critical domain of personal safety, pedestrian skills. Cognitive processes involved in safe pedestrian crossings may be impaired in children with EDS. In the pedestrian simulation, children with EDS appeared to show a pattern consistent with inattentional blindness, in that they “looked but did not process” information in their pedestrian environment. Results highlight the need for heightened awareness of potentially irreversible consequences of untreated sleep disorders and identify a possible target for pediatric injury prevention. Citation: Avis KT; Gamble KL; Schwebel DC. Does excessive daytime sleepiness affect children's pedestrian safety? SLEEP 2014;37(2):283-287. PMID:24497656

  6. Obstructive Sleep Apnea With Objective Daytime Sleepiness Is Associated With Hypertension.

    PubMed

    Ren, Rong; Li, Yun; Zhang, Jihui; Zhou, Junying; Sun, Yuanfeng; Tan, Lu; Li, Taomei; Wing, Yun-Kwok; Tang, Xiangdong

    2016-11-01

    Subjective daytime sleepiness is considered a significant risk factor of hypertension in patients with obstructive sleep apnea (OSA). In this study, our goal was to examine the joint effect on hypertension of OSA and objective daytime sleepiness measured by the multiple sleep latency test (MSLT). A total of 1338 Chinese patients with OSA and 484 primary snorers were included in the study. All subjects underwent 1 night polysomnography followed by MSLT. The MSLT values were classified into 3 categories: >8 minutes, 5 to 8 minutes, and <5 minutes. Hypertension was defined based either on direct blood pressure measures or on diagnosis by a physician. After controlling for confounders, OSA combined with MSLT of 5 to 8 minutes increased the odds of hypertension by 95% (odds ratio, 1.95; 95% confidence interval, 1.10-3.46), whereas OSA combined with MSLT <5 minutes further increased the odds of hypertension by 111% (odds ratio, 2.11; 95% confidence interval, 1.22-3.31) compared with primary snorers with MSLT >8 minutes. In stratified analyses, the association of hypertension with MSLT in OSA patients was seen among both sexes, younger ages, both obese and nonobese patients, and patients with and without subjective excessive daytime sleepiness. We conclude that objective daytime sleepiness is associated with hypertension in patients with OSA. © 2016 American Heart Association, Inc.

  7. Sleep hygiene and its association with daytime sleepiness, depressive symptoms, and quality of life in patients with mild obstructive sleep apnea.

    PubMed

    Lee, Sang-Ahm; Paek, Joon-Hyun; Han, Su-Hyun

    2015-12-15

    To investigate the direct and indirect associations of sleep hygiene with daytime sleepiness, depressive symptoms, and quality of life (QoL), in newly diagnosed, untreated patients with mild obstructive sleep apnea (OSA). Data were collected from adults with mild OSA. The Sleep Hygiene Index (SHI), Sleep Problems Index-1 (SPI-1) of the Medical Outcomes Study-Sleep Scale, Epworth Sleepiness Scale (ESS), Beck Depression Inventory (BDI), and Medical Outcomes Study Short-Form Health survey (SF-36) were used to evaluate patients. To determine the indirect and direct associations between SHI and disease outcomes, the Sobel test and multiple linear regression analyses were used, respectively. When we evaluated the direct associations, we excluded 3 items of the original SHI which were more reflective of general health rather than sleep-specific habits and environments. In total, 260 patients with mild OSA participated in this study. The average age, AHI, and SHI scores were 49.1 years (SD 12.5), 9.3/h (SD 2.9), and 24.7 (SD 6.0), respectively. Here, ≥ 10% of participants indicated poor sleep hygiene behaviors on 7 of 13 items. Young age and men were associated with higher SHI scores (both p<0.01). The 13-item SHI scores were indirectly related to ESS, BDI, and SF-36 scores via SPI-1 (all p<0.05). The 10-item SHI scores were related to ESS (p=0.049) and SF-36 (p=0.001), but not to BDI, independently of SPI-1 or other confounding factors in mild OSA patients. Age, sex, AHI, and body mass index were not related to ESS, BDI, or total SF-36 scores. Sleep hygiene is indirectly related to daytime sleepiness, depressive symptoms, QoL via sleep quality and also related to daytime sleepiness and QoL independent of sleep quality in mild OSA patients. Copyright © 2015 Elsevier B.V. All rights reserved.

  8. Gender differences in sleep habits and quality and daytime sleepiness in elementary and high school teachers.

    PubMed

    de Souza, Jane Carla; Oliveira, Maria Luiza Cruz de; de Sousa, Ivanise Cortez; Azevedo, Carolina V M de

    2018-04-01

    The extensive workload of teachers inside and outside the classroom may contribute to sleep problems. Such problems may occur more frequently in women due to the combination of professional demands, domestic tasks, and their relatively greater sleep needs compared to men. The objective of this cross-sectional study was to evaluate the influence of gender on sleep habits and quality, and daytime sleepiness in a sample of 243 teachers (77 men and 166 women) using questionnaires. Linear regression models were used to examine the effect of gender on sleep measures; the unadjusted model considered only gender and the adjusted model considered chronotype and work characteristics as potential confounders. Bedtimes of women were significantly earlier than men during the week, but not on weekends, in the unadjusted and adjusted models. Time in bed was longer for women throughout the week and weekend in the unadjusted model. However, in the adjusted model, this statistical significance disappeared, and longer time in bed during the week was associated with teaching in one shift and for both levels of education. In addition, the female gender was associated with higher sleepiness scores compared to males in both models, and worse sleep quality in the adjusted model. Also, sleep quality was worse in subjects working in three shifts and in both types of schools (public and private). The tendency to eveningness was associated with later bedtimes and wake up times during both week days and weekends, higher irregularity of bedtimes and wake up times, and higher sleepiness scores in the adjusted model. Therefore, we suggest that female teachers do not fulfill their sleep needs and show higher levels of diurnal sleepiness and poor sleep quality that can be modulated by chronotype and some work characteristics. More studies are needed to evaluate the role of double workload on this pattern.

  9. Psychometric Properties of Turkish Version of Pediatric Daytime Sleepiness Scale (PDSS-T).

    PubMed

    Bektas, Murat; Bektas, Ilknur; Ayar, Dijle; Selekoglu, Yasemin; Ayar, Ugur; Kudubes, Aslı Akdeniz; Altan, Sema Sal; Armstrong, Merry

    2016-03-01

    The aim of the research was to evaluate the psychometric properties of the Pediatric Daytime Sleepiness Scale-Turkish Version (PDSS-T). The researchers chose a study sample of 522 grade 5-11 students. Data were collected using a demographic data collection form and the PDSS-T. Cronbach α for the scale was .79 and Kaiser-Meyer-Olkin coefficient was .78. Item-total correlations for the scale varied between .53 and .73 (p < .001). The indices of model fit were determined to be the root mean square error of approximation at .07, the goodness of fit index at .97, and the comparative fit index at .97. The study's results showed that PDSS-T is a valid and reliable instrument for detecting Turkish-speaking children's and adolescents' daytime sleepiness. PDSS-T is convenient for professionals to prevent and manage daytime sleepiness. Copyright © 2016. Published by Elsevier B.V.

  10. School Maladjustment and External Locus of Control Predict the Daytime Sleepiness of College Students With ADHD.

    PubMed

    Langberg, Joshua M; Dvorsky, Melissa R; Becker, Stephen P; Molitor, Stephen J

    2016-09-01

    The primary aim of this study was to evaluate whether school maladjustment longitudinally predicts the daytime sleepiness of college students with ADHD above and beyond symptoms of ADHD and to determine whether internalizing dimensions mediate the relationship between maladjustment and sleepiness. A prospective longitudinal study of 59 college students comprehensively diagnosed with ADHD who completed ratings at the beginning, middle, and end of the school year. School maladjustment at the beginning of the year significantly predicted daytime sleepiness at the end of the year above and beyond symptoms of ADHD. Locus of control mediated the relationship between maladjustment and daytime sleepiness. The significant school maladjustment difficulties that students with ADHD experience following the transition to college may lead to the development of problems with daytime sleepiness, particularly for those students with high external locus of control. This pattern is likely reciprocal, whereby sleep problems in turn result in greater school impairment, reinforcing the idea that life events are outside of one's control. © The Author(s) 2014.

  11. Objective measurement of daytime napping, cognitive dysfunction and subjective sleepiness in Parkinson's disease.

    PubMed

    Bolitho, Samuel J; Naismith, Sharon L; Salahuddin, Pierre; Terpening, Zoe; Grunstein, Ron R; Lewis, Simon J G

    2013-01-01

    Sleep-wake disturbances and concomitant cognitive dysfunction in Parkinson's disease (PD) contribute significantly to morbidity in patients and their carers. Subjectively reported daytime sleep disturbance is observed in over half of all patients with PD and has been linked to executive cognitive dysfunction. The current study used daytime actigraphy, a novel objective measure of napping and related this to neuropsychological performance in a sample of PD patients and healthy, age and gender-matched controls. Furthermore this study aimed to identify patients with PD who may benefit from pharmacologic and behavioural intervention to improve these symptoms. Eighty-five PD patients and 21 healthy, age-matched controls completed 14 days of wrist actigraphy within two weeks of neuropsychological testing. Objective napping measures were derived from actigraphy using a standardised protocol and subjective daytime sleepiness was recorded by the previously validated Epworth Sleepiness Scale. Patients with PD had a 225% increase in the mean nap time per day (minutes) as recorded by actigraphy compared to age matched controls (39.2 ± 35.2 vs. 11.5 ± 11.0 minutes respectively, p < 0.001). Significantly, differences in napping duration between patients, as recorded by actigraphy were not distinguished by their ratings on the subjective measurement of excessive daytime sleepiness. Finally, those patients with excessive daytime napping showed greater cognitive deficits in the domains of attention, semantic verbal fluency and processing speed. This study confirms increased levels of napping in PD, a finding that is concordant with subjective reports. However, subjective self-report measures of excessive daytime sleepiness do not robustly identify excessive napping in PD. Fronto-subcortical cognitive dysfunction was observed in those patients who napped excessively. Furthermore, this study suggests that daytime actigraphy, a non-invasive and inexpensive objective measure of

  12. Psychosocial work characteristics predicting daytime sleepiness in day and shift workers.

    PubMed

    Takahashi, Masaya; Nakata, Akinori; Haratani, Takashi; Otsuka, Yasumasa; Kaida, Kosuke; Fukasawa, Kenji

    2006-01-01

    Characteristics of work organization other than working time arrangements may contribute importantly to daytime sleepiness. The present study was designed to identify the psychosocial factors at work that predict daytime sleepiness in a sample of day and shift workers. Participants working at a pulp and chemical factory completed an annual questionnaire regarding psychosocial factors at work using the U.S. National Institute for Occupational Safety and Health Generic Job Stress Questionnaire (i.e., quantitative workload, variance in workload, job control, support from supervisor, coworkers, or family/friends, job satisfaction, and depressive symptoms), as well as daytime sleepiness (through the Epworth Sleepiness Scale [ESS]) and sleep disturbances for three years starting in 2002 (response rates, 94.6-99.0%). The present analysis included 55 day workers (11 women) and 57 shift workers (all men) who participated in all three years of the study, worked under the same work schedule throughout the study period, and had no missing data on any of the daytime sleep items. A repeated-measures analysis of covariance (ANCOVA) was used to test the effects of work schedule (day vs. shift work) and psychosocial factors at work in 2002 on the ESS scores in subsequent years, with sleep duration, insomnia symptoms, chronic diseases, and sleepiness levels at baseline as covariates. Given significant and near-significant interactions of work schedules with psychosocial factor or study year, the ANCOVA, with the factors of psychosocial work characteristics and study year, was performed by type of work schedule. The results indicated a significant main effect of psychosocial work characteristics (p = 0.010, partial eng2 = 0.14) and an almost significant main effect of study year (p = 0.067, partial eng2 = 0.06) and interaction between psychosocial work characteristics and study year (p = 0.085, partial eng2 = 0.06) for variance in workload among the day work group. The day workers

  13. Short Daytime Naps Briefly Attenuate Objectively Measured Sleepiness Under Chronic Sleep Restriction.

    PubMed

    Saletin, Jared M; Hilditch, Cassie J; Dement, William C; Carskadon, Mary A

    2017-09-01

    Napping is a useful countermeasure to the negative effects of acute sleep loss on alertness. The efficacy of naps to recover from chronic sleep loss is less well understood. Following 2 baseline nights (10 hours' time-in-bed), participants were restricted to 7 nights of 5-hour sleep opportunity. Ten adults participated in the No-Nap condition, and a further 9 were assigned to a Nap condition with a daily 45-minute nap opportunity at 1300 h. Sleepiness was assessed using the multiple sleep latency test and a visual analogue scale at 2-hour intervals. Both objective and subjective indexes of sleepiness were normalized within subject as a difference from those at baseline prior to sleep restriction. Mixed-effects models examined how the daytime nap opportunity altered sleepiness across the day and across the protocol. Short daytime naps attenuated sleepiness due to chronic sleep restriction for up to 6-8 hours after the nap. Benefits of the nap did not extend late into evening. Subjective sleepiness demonstrated a similar short-lived benefit that emerged later in the day when objective sleepiness already returned to pre-nap levels. Neither measure showed a benefit of the nap the following morning after the subsequent restriction night. These data indicate a short daytime nap may attenuate sleepiness in chronic sleep restriction, yet subjective and objective benefits emerge at different time scales. Because neither measure showed a benefit the next day, the current study underscores the need for careful consideration before naps are used as routine countermeasures to chronic sleep loss. © Sleep Research Society 2017. Published by Oxford University Press on behalf of the Sleep Research Society. All rights reserved. For permissions, please e-mail journals.permissions@oup.com.

  14. Excessive daytime sleepiness and subsequent development of Parkinson disease.

    PubMed

    Abbott, R D; Ross, G W; White, L R; Tanner, C M; Masaki, K H; Nelson, J S; Curb, J D; Petrovitch, H

    2005-11-08

    To determine if excessive daytime sleepiness (EDS) can predate future Parkinson disease (PD). EDS was assessed in 3,078 men aged 71 to 93 years in the Honolulu-Asia Aging Study from 1991 to 1993. All were free of prevalent PD and dementia. Follow-up for incident PD was based on three repeat neurologic assessments from 1994 to 2001. During the course of follow-up, 43 men developed PD (19.9/10,000 person-years). After age adjustment, there was more than a threefold excess in the risk of PD in men with EDS vs men without EDS (55.3 vs 17.0/10,000 person-years; odds ratio [OR] = 3.3; 95% CI = 1.4 to 7.0; p = 0.004). Additional adjustment for insomnia, cognitive function, depressed mood, midlife cigarette smoking and coffee drinking, and other factors failed to alter the association between EDS and PD (OR = 2.8; 95% CI = 1.1 to 6.4; p = 0.014). Other sleep related features such as insomnia, daytime napping, early morning grogginess, and frequent nocturnal awakening showed little relation with the risk of PD. Excessive daytime sleepiness may be associated with an increased risk of developing Parkinson disease.

  15. Impact of social media usage on daytime sleepiness: A study in a sample of tertiary students in Singapore.

    PubMed

    Nasirudeen, A M A; Lee Chin Adeline, Lau; Wat Neo Josephine, Koh; Lay Seng, Lim; Wenjie, Li

    2017-01-01

    Many tertiary students access social networking sites on a daily basis. With the increased usage of smartphones, accessing social networking sites while commuting, in schools, waiting for friends, television commercial breaks has become prevalent among tertiary students. What started as a lifestyle choice has now become a daily necessity. Such behavior among tertiary students raises an important question for educators: how does social media usage affect tertiary students' sleep patterns and daytime sleepiness, their attention difficulties, especially in school? Thus, the aim of this study was to examine the relationships between tertiary students' self-reports of social media usage and daytime sleepiness. The design was a cross-sectional, quantitative research study. We used a survey that contained questions concerning demographic data, daytime sleepiness, total sleep time and social media usage and a version of the Cleveland Adolescent Sleepiness Questionnaire, modified for use in tertiary students, were used for data collection ( n  = 969). The most preferred tool for accessing social networking sites was smartphones and WhatsApp was the most accessed site. Results indicated that nocturnal technology use has a weak, negative impact on tertiary students' quantity of sleep that may lead to daytime sleepiness. Local Singapore students spent significantly more time on social networking sites at night compared to foreign students. As a result, local students experienced more daytime sleepiness compared to foreign students. Prolonged social media usage, especially in bed, has a negative impact on tertiary students' daytime sleepiness. Since the technology is such an integral part of most tertiary students' lives, it is important to understand the impact it has on their sleep and daytime sleepiness.

  16. Sleep habits, insomnia, and daytime sleepiness in a large and healthy community-based sample of New Zealanders.

    PubMed

    Wilsmore, Bradley R; Grunstein, Ronald R; Fransen, Marlene; Woodward, Mark; Norton, Robyn; Ameratunga, Shanthi

    2013-06-15

    To determine the relationship between sleep complaints, primary insomnia, excessive daytime sleepiness, and lifestyle factors in a large community-based sample. Cross-sectional study. Blood donor sites in New Zealand. 22,389 individuals aged 16-84 years volunteering to donate blood. N/A. A comprehensive self-administered questionnaire including personal demographics and validated questions assessing sleep disorders (snoring, apnea), sleep complaints (sleep quantity, sleep dissatisfaction), insomnia symptoms, excessive daytime sleepiness, mood, and lifestyle factors such as work patterns, smoking, alcohol, and illicit substance use. Additionally, direct measurements of height and weight were obtained. One in three participants report < 7-8 h sleep, 5 or more nights per week, and 60% would like more sleep. Almost half the participants (45%) report suffering the symptoms of insomnia at least once per week, with one in 5 meeting more stringent criteria for primary insomnia. Excessive daytime sleepiness (evident in 9% of this large, predominantly healthy sample) was associated with insomnia (odds ratio [OR] 1.75, 95% confidence interval [CI] 1.50 to 2.05), depression (OR 2.01, CI 1.74 to 2.32), and sleep disordered breathing (OR 1.92, CI 1.59 to 2.32). Long work hours, alcohol dependence, and rotating work shifts also increase the risk of daytime sleepiness. Even in this relatively young, healthy, non-clinical sample, sleep complaints and primary insomnia with subsequent excess daytime sleepiness were common. There were clear associations between many personal and lifestyle factors-such as depression, long work hours, alcohol dependence, and rotating shift work-and sleep problems or excessive daytime sleepiness.

  17. Amount of Sleep, Daytime Sleepiness, Hazardous Driving, and Quality of Life of Second Year Medical Students.

    PubMed

    Johnson, Kay M; Simon, Nancy; Wicks, Mark; Barr, Karen; O'Connor, Kim; Schaad, Doug

    2017-10-01

    The authors describe the sleep habits of second year medical students and look for associations between reported sleep duration and depression, burnout, overall quality of life, self-reported academic success, and falling asleep while driving. The authors conducted a cross-sectional descriptive study of two consecutive cohorts of second year medical students at a large public university in the USA. Participants completed an anonymous survey about their sleep habits, daytime sleepiness (Epworth sleepiness scale), burnout (Maslach burnout inventory), depression (PRIME MD), and perceived stress (perceived stress scale). Categorical and continuous variables were compared using chi square tests and t tests, respectively. Sixty-eight percent of the students responded. Many (34.3%) reported fewer than 7 h of sleep on typical weeknights, including 6.5% who typically sleep less than 6 h. Twenty-five students (8.4%) reported nodding off while driving during the current academic year. Low typical weeknight sleep (fewer than 6 h vs 6-6.9 h vs 7 or more hours) was associated with (1) higher Epworth sleepiness scale scores, (2) nodding off while driving, (3) symptoms of burnout or depression, (4) decreased satisfaction with quality of life, and (5) lower perceived academic success (all p values ≤0.01). Students reporting under 6 h of sleep were four times more likely to nod off while driving than those reporting 7 h or more. Educational, behavioral, and curricular interventions should be explored to help pre-clinical medical students obtain at least 7 h of sleep most on weeknights.

  18. Does excessive daytime sleepiness affect children's pedestrian safety?

    PubMed

    Avis, Kristin T; Gamble, Karen L; Schwebel, David C

    2014-02-01

    Many cognitive factors contribute to unintentional pedestrian injury, including reaction time, impulsivity, risk-taking, attention, and decision-making. These same factors are negatively influenced by excessive daytime sleepiness (EDS), which may place children with EDS at greater risk for pedestrian injury. Using a case-control design, 33 children age 8 to 16 y with EDS from an established diagnosis of narcolepsy or idiopathic hypersomnia (IHS) engaged in a virtual reality pedestrian environment while unmedicated. Thirty-three healthy children matched by age, race, sex, and household income served as controls. Children with EDS were riskier pedestrians than healthy children. They were twice as likely to be struck by a virtual vehicle in the virtual pedestrian environment than healthy controls. Attentional skills of looking at oncoming traffic were not impaired among children with EDS, but decision-making for when to cross the street safely was significantly impaired. Results suggest excessive daytime sleepiness (EDS) from the clinical sleep disorders known as the hypersomnias of central origin may have significant consequences on children's daytime functioning in a critical domain of personal safety, pedestrian skills. Cognitive processes involved in safe pedestrian crossings may be impaired in children with EDS. In the pedestrian simulation, children with EDS appeared to show a pattern consistent with inattentional blindness, in that they "looked but did not process" information in their pedestrian environment. Results highlight the need for heightened awareness of potentially irreversible consequences of untreated sleep disorders and identify a possible target for pediatric injury prevention.

  19. Fatigue as a cause, not a consequence of depression and daytime sleepiness: a cross-lagged analysis.

    PubMed

    Schönberger, Michael; Herrberg, Marlene; Ponsford, Jennie

    2014-01-01

    To examine the temporal relation between fatigue, depression, and daytime sleepiness after traumatic brain injury. Fatigue is a frequent and disabling consequence of traumatic brain injury (TBI). However, it is unclear whether fatigue is a primary consequence of the structural brain injury or a secondary consequence of injury-related sequelae such as depression and daytime sleepiness. Eighty-eight adults with complicated mild-severe TBI (69% male). Fatigue Severity Scale; depression subscale of the Hospital Anxiety and Depression Scale; Epworth Sleepiness scale at baseline and 6-month follow-up. A cross-lagged path analysis computed within a structural equation modeling framework revealed that fatigue was predictive of depression (β = .20, P < .05) and sleepiness (β = .25, P < .05). However, depression and sleepiness did not predict fatigue (P > .05). The results support the view of fatigue after TBI as "primary fatigue"-that is, a consequence of the structural brain injury rather than a secondary consequence of depression or daytime sleepiness. A rehabilitation approach that assists individuals with brain injury in learning to cope with their neuropsychological and physical limitations in everyday life might attenuate their experience with fatigue.

  20. Sleep Habits, Insomnia, and Daytime Sleepiness in a Large and Healthy Community-Based Sample of New Zealanders

    PubMed Central

    Wilsmore, Bradley R.; Grunstein, Ronald R.; Fransen, Marlene; Woodward, Mark; Norton, Robyn; Ameratunga, Shanthi

    2013-01-01

    Study Objectives: To determine the relationship between sleep complaints, primary insomnia, excessive daytime sleepiness, and lifestyle factors in a large community-based sample. Design: Cross-sectional study. Setting: Blood donor sites in New Zealand. Patients or Participants: 22,389 individuals aged 16-84 years volunteering to donate blood. Interventions: N/A. Measurements: A comprehensive self-administered questionnaire including personal demographics and validated questions assessing sleep disorders (snoring, apnea), sleep complaints (sleep quantity, sleep dissatisfaction), insomnia symptoms, excessive daytime sleepiness, mood, and lifestyle factors such as work patterns, smoking, alcohol, and illicit substance use. Additionally, direct measurements of height and weight were obtained. Results: One in three participants report < 7-8 h sleep, 5 or more nights per week, and 60% would like more sleep. Almost half the participants (45%) report suffering the symptoms of insomnia at least once per week, with one in 5 meeting more stringent criteria for primary insomnia. Excessive daytime sleepiness (evident in 9% of this large, predominantly healthy sample) was associated with insomnia (odds ratio [OR] 1.75, 95% confidence interval [CI] 1.50 to 2.05), depression (OR 2.01, CI 1.74 to 2.32), and sleep disordered breathing (OR 1.92, CI 1.59 to 2.32). Long work hours, alcohol dependence, and rotating work shifts also increase the risk of daytime sleepiness. Conclusions: Even in this relatively young, healthy, non-clinical sample, sleep complaints and primary insomnia with subsequent excess daytime sleepiness were common. There were clear associations between many personal and lifestyle factors—such as depression, long work hours, alcohol dependence, and rotating shift work—and sleep problems or excessive daytime sleepiness. Citation: Wilsmore BR; Grunstein RR; Fransen M; Woodward M; Norton R; Ameratunga S. Sleep habits, insomnia, and daytime sleepiness in a large and

  1. Objective Measurement of Daytime Napping, Cognitive Dysfunction and Subjective Sleepiness in Parkinson’s Disease

    PubMed Central

    Bolitho, Samuel J.; Naismith, Sharon L.; Salahuddin, Pierre; Terpening, Zoe; Grunstein, Ron R.; Lewis, Simon J. G.

    2013-01-01

    Introduction Sleep-wake disturbances and concomitant cognitive dysfunction in Parkinson’s disease (PD) contribute significantly to morbidity in patients and their carers. Subjectively reported daytime sleep disturbance is observed in over half of all patients with PD and has been linked to executive cognitive dysfunction. The current study used daytime actigraphy, a novel objective measure of napping and related this to neuropsychological performance in a sample of PD patients and healthy, age and gender-matched controls. Furthermore this study aimed to identify patients with PD who may benefit from pharmacologic and behavioural intervention to improve these symptoms. Methods Eighty-five PD patients and 21 healthy, age-matched controls completed 14 days of wrist actigraphy within two weeks of neuropsychological testing. Objective napping measures were derived from actigraphy using a standardised protocol and subjective daytime sleepiness was recorded by the previously validated Epworth Sleepiness Scale. Results Patients with PD had a 225% increase in the mean nap time per day (minutes) as recorded by actigraphy compared to age matched controls (39.2 ± 35.2 vs. 11.5 ± 11.0 minutes respectively, p < 0.001). Significantly, differences in napping duration between patients, as recorded by actigraphy were not distinguished by their ratings on the subjective measurement of excessive daytime sleepiness. Finally, those patients with excessive daytime napping showed greater cognitive deficits in the domains of attention, semantic verbal fluency and processing speed. Conclusion This study confirms increased levels of napping in PD, a finding that is concordant with subjective reports. However, subjective self-report measures of excessive daytime sleepiness do not robustly identify excessive napping in PD. Fronto-subcortical cognitive dysfunction was observed in those patients who napped excessively. Furthermore, this study suggests that daytime actigraphy, a non

  2. Determinants of Excessive Daytime Sleepiness and Fatigue in Adults with Heart Failure

    PubMed Central

    Riegel, Barbara; Ratcliffe, Sarah J.; Sayers, Steven L.; Potashnik, Sheryl; Buck, Harleah; Jurkovitz, Claudine; Fontana, Sarah; Weaver, Terri E.; Weintraub, William S.; Goldberg, Lee R.

    2012-01-01

    Little is known about excessive daytime sleepiness (EDS) in heart failure (HF). The aim of this cross-sectional descriptive study was to describe the prevalence of EDS and factors associated with it in HF. A secondary purpose was to explore the correlates of fatigue. We enrolled a consecutive sample of 280 adults with a confirmed diagnosis of chronic HF from three outpatient settings in the northeastern US. Patients with major depressive illness were excluded. Clinical, sociodemographic, behavioral, and perceptual factors were explored as possible correlates of EDS. Using an Epworth Sleepiness Scale score >10, the prevalence of EDS was 23.6%. Significant determinants of EDS were worse sleep quality (p=0.048), worse functional class (p=0.004), not taking a diuretic (p=0.005), and lack of physical activity (p=0.04). Only sleep quality was associated with fatigue (p<0.001). Sleep disordered breathing was not significantly associated with EDS or with fatigue. These factors may be amenable to intervention. PMID:21878581

  3. Objective and Subjective Socioeconomic Gradients Exist for Sleep Quality, Sleep Latency, Sleep Duration, Weekend Oversleep, and Daytime Sleepiness in Adults

    PubMed Central

    Jarrin, Denise Christina; McGrath, Jennifer J.; Silverstein, Janice E.; Drake, Christopher

    2017-01-01

    Socioeconomic gradients exist for multiple health outcomes. Lower objective socioeconomic position (SEP), whether measured by income, education, or occupation, is associated with inadequate sleep. Less is known about whether one’s perceived ranking of their social status, or subjective SEP, affects sleep. This study examined whether a subjective socioeconomic gradient exists for sleep while controlling for objective SEP. Participants (N = 177; age, M = 45.3 years, SD = 6.3 years) completed the Pittsburgh Sleep Quality Index, Epworth Sleepiness Scale, MacArthur Ladder, and other self-report measures to assess sleep and objective SEP. Subjective SEP trumped objective SEP as a better predictor of sleep duration, daytime sleepiness, and weekend oversleep. These findings highlight the need to expand our framework to better understand the mechanisms underlying socioeconomic gradients and sleep. PMID:23136841

  4. Excessive daytime sleepiness and its pattern among Indian college students.

    PubMed

    Kaur, Gurjeet; Singh, Amarjeet

    2017-01-01

    The objective of this study was to determine the prevalence and pattern of excessive daytime sleepiness (EDTS) in Indian college students. This was a cross-sectional study that was conducted among 1215 undergraduate students, using the Epworth Sleepiness Scale (ESS) and a sociodemographic survey. A high proportion (45%) of EDTS was observed, and the problem was significantly greater in participants from professional courses. A probability of association of EDTS with coffee/tea consumption, alcohol consumption and smoking was also observed in the study. Copyright © 2016 Elsevier B.V. All rights reserved.

  5. Severe excessive daytime sleepiness induced by hydroxyurea.

    PubMed

    Revol, Bruno; Joyeux-Faure, Marie; Albahary, Marie-Victoire; Gressin, Remy; Mallaret, Michel; Pepin, Jean-Louis; Launois, Sandrine H

    2017-06-01

    Excessive daytime sleepiness (EDS) has been reported with many drugs, either as an extension of a hypnotic effect (e.g. central nervous system depressants) or as an idiosyncratic response of the patient. Here, we report unexpected and severe subjective and objective EDS induced by hydroxyurea therapy, with a favorable outcome after withdrawal. Clinical history, sleep log, polysomnography, and multiple sleep latency tests confirming the absence of other EDS causes are presented. © 2016 Société Française de Pharmacologie et de Thérapeutique.

  6. Clinical implications of daytime sleepiness for the academic performance of middle school-aged adolescents with attention deficit hyperactivity disorder.

    PubMed

    Langberg, Joshua M; Dvorsky, Melissa R; Marshall, Stephen; Evans, Steven W

    2013-10-01

    This study investigated the relative impact of total time slept per night and daytime sleepiness on the academic functioning of 100 middle school-aged youth (mean age = 11.9) with attention deficit hyperactivity disorder (ADHD). The primary goal of the study was to determine if total time slept per night and/or daytime sleepiness, as measured by youth self-report on the Pediatric Daytime Sleepiness Scale (PDSS), predicted academic functioning above and beyond symptoms of ADHD and relevant covariates, such as intelligence, achievement scores and parent education level. Self-reported daytime sleepiness but not self-reported total time slept per night was associated significantly with all academic outcomes. When examined in a hierarchical regression model, self-reported daytime sleepiness significantly predicted parent-rated homework problems and academic impairment and teacher-rated academic competence above and beyond symptoms of ADHD and relevant covariates, but did not predict grade point average or teacher-rated academic impairment. The implications of these findings for understanding more clearly the association between ADHD and sleep and the functional implications of this relationship are discussed. © 2013 European Sleep Research Society.

  7. Symptoms of Daytime Sleepiness and Sleep Apnea in Liver Cirrhosis Patients.

    PubMed

    Enezi, Abdullah; Al-Jahdali, Fares; Ahmed, Anwar; Shirbini, Nahid; Harbi, Abdullah; Salim, Baharoon; Ali, Yosra; Abdulrahman, Aljumah; Khan, Mohd; Khaleid, Abdullah; Hamdan, Al-Jahdali

    2017-01-01

    Background/propose. Sleep disturbance and excessive daytime sleepiness (EDS) have been reported in patients with hepatic cirrhosis with no hepatic encephalopathy (HE). The objective of this study was to evaluate daytime sleepiness and risk of obstructive sleep apnea (OSA) among liver cirrhosis patients. A cross-sectional study was conducted at King Abdulaziz Medical City (KAMC)-Riyadh over a period of six months, using a structured questionnaire that investigated: 1) Sleep patterns and daytime sleepiness using the Epworth Sleeping Scale (ESS), and 2) The risk for sleep apnea using the Berlin Questionnaire (BQ). We enrolled patients with a confirmed diagnosis of liver cirrhosis who were being followed at the hepatology and pre-liver transplant clinics. We enrolled 200 patients with liver cirrhosis, 57.5% of whom were male. The mean age was 60 (± SD 12.2). The reported prevalence of EDS, OSA, and both EDS and OSA were 29.5%, 42.9%, and 13.6%, respectively. The prevalence of EDS was higher in patients with Hepatitis-C and patients with DM, who experienced short sleep duration. We did not find any association between the severity of liver disease and EDS or OSA as measured by Child-Pugh scores (CPS). The risk of OSA and EDS is high among liver cirrhosis patients. Those patients with cirrhosis secondary to Hepatitis C are at higher risk of EDS and OSA. Both EDS and OSA affect patients designated as CPS Class A more frequently than patients designated as CPS Class B.

  8. Frequent nocturnal awakening in children: prevalence, risk factors, and associations with subjective sleep perception and daytime sleepiness.

    PubMed

    Li, Liwen; Ren, Jiwei; Shi, Lei; Jin, Xinming; Yan, Chonghuai; Jiang, Fan; Shen, Xiaoming; Li, Shenghui

    2014-07-30

    Nocturnal awakening is the most frequent insomnia complaint in the general population. In contrast to a growing knowledge based on adults, little is known about its prevalence, correlated factors, and associations with subjective sleep perception and daytime sleepiness in children. This study was designed to assess the prevalence and the correlate factors of frequent nocturnal awakening (FNA) among Chinese school-aged children. Furthermore, the associations of FNA with subjective sleep perception and daytime sleepiness were examined. A random sample of 20,505 children aged 5.00 to 11.92 years old (boys: 49.5% vs. girls: 50.5%) participated in a cross-sectional survey, which was conducted in eight cities of China. Parent-administered questionnaires were used to collect information on children's sleep behaviors, sleep perception, and potential influential factors of FNA from six domains. Univariate and multivariate logistic regression models were performed. The prevalence of FNA was 9.8% (10.0% for boys vs. 8.9% for girls) in our sampled children. The prominent FNA-related factors inclued biological health problems, such as overweight/obesity (OR = 1.70), chronic pain during night (OR = 2.47), and chronic respiratory condition (OR = 1.23), poor psychosocial condition, such as poor mental and emotional functioning (OR = 1.34), poor sleep hygiene, such as frequently doing exciting activities before bedtime (OR = 1.24) and bedtime resistance (OR = 1.42), and parents' history of insomnia (OR = 1.31). FNA was associated with subjective poor sleep quality (OR = 1.24), subjective insufficient sleep (OR = 1.21), and daytime sleepiness (OR = 1.35). FNA was associated with poor sleep and daytime sleepiness. Compared to sleep environment and family susceptibility, chronic health problems, poor psychosocial condition, and poor sleep hygiene had greater impact on FNA, indicating childhood FNA could be partly prevented by health promotion, by psychological intervention, and by

  9. Sleep disordered breathing and daytime sleepiness are associated with poor academic performance in teenagers. A study using the Pediatric Daytime Sleepiness Scale (PDSS).

    PubMed

    Perez-Chada, Daniel; Perez-Lloret, Santiago; Videla, Alejandro J; Cardinali, Daniel; Bergna, Miguel A; Fernández-Acquier, Mariano; Larrateguy, Luis; Zabert, Gustavo E; Drake, Christopher

    2007-12-01

    Inadequate sleep and sleep disordered breathing (SDB) can impair learning skills. Questionnaires used to evaluate sleepiness in adults are usually inadequate for adolescents. We conducted a study to evaluate the performance of a Spanish version of the Pediatric Daytime Sleepiness Scale (PDSS) and to assess the impact of sleepiness and SDB on academic performance. A cross-sectional survey of students from 7 schools in 4 cities of Argentina. A questionnaire with a Spanish version of the PDSS was used. Questions on the occurrence of snoring and witnessed apneas were answered by the parents. Mathematics and language grades were used as indicators of academic performance. The sample included 2,884 students (50% males; age: 13.3 +/- 1.5 years) Response rate was 85%; 678 cases were excluded due to missing data. Half the students slept <9 h per night on weekdays. The mean PDSS value was 15.74 +/- 5.93. Parental reporting of snoring occurred in 511 subjects (23%); snoring was occasional in 14% and frequent in 9%. Apneas were witnessed in 237 cases (11%), being frequent in 4% and occasional in 7%. Frequent snorers had higher mean PDSS scores than occasional or nonsnorers (18 +/- 5, 15.7 +/- 6 and 15.5 +/- 6, respectively; P < 0.001). Reported snoring or apneas and the PDSS were significant univariate predictors of failure and remained significant in multivariate logistic regression analysis after adjusting for age, sex, body mass index, specific school attended, and sleep habits. Insufficient hours of sleep were prevalent in this population. The Spanish version of the PDSS was a reliable tool in middle-school-aged children. Reports of snoring or witnessed apneas and daytime sleepiness as measured by PDSS were independent predictors of poor academic performance.

  10. Impaired Patient-Reported Outcomes Predict Poor School Functioning and Daytime Sleepiness: The PROMIS Pediatric Asthma Study.

    PubMed

    Jones, Conor M; DeWalt, Darren A; Huang, I-Chan

    Poor asthma control in children is related to impaired patient-reported outcomes (PROs; eg, fatigue, depressive symptoms, anxiety), but less well studied is the effect of PROs on children's school performance and sleep outcomes. In this study we investigated whether the consistency status of PROs over time affected school functioning and daytime sleepiness in children with asthma. Of the 238 children with asthma enrolled in the Patient-Reported Outcomes Measurement Information System (PROMIS) Pediatric Asthma Study, 169 children who provided survey data for all 4 time points were used in the analysis. The child's PROs, school functioning, and daytime sleepiness were measured 4 times within a 15-month period. PRO domains included asthma impact, pain interference, fatigue, depressive symptoms, anxiety, and mobility. Each child was classified as having poor/fair versus good PROs per meaningful cut points. The consistency status of each domain was classified as consistently poor/fair if poor/fair status was present for at least 3 time points; otherwise, the status was classified as consistently good. Seemingly unrelated regression was performed to test if consistently poor/fair PROs predicted impaired school functioning and daytime sleepiness at the fourth time point. Consistently poor/fair in all PRO domains was significantly associated with impaired school functioning and excessive daytime sleepiness (Ps < .01) after controlling for the influence of the child's age, sex, and race/ethnicity. Children with asthma with consistently poor/fair PROs are at risk of poor school functioning and daytime sleepiness. Developing child-friendly PRO assessment systems to track PROs can inform potential problems in the school setting. Copyright © 2017 Academic Pediatric Association. Published by Elsevier Inc. All rights reserved.

  11. Excessive Daytime sleepiness in idiopathic restless legs syndrome: characteristics and evolution under dopaminergic treatment.

    PubMed

    Kallweit, Ulf; Siccoli, Massimiliano M; Poryazova, Rositsa; Werth, Esther; Bassetti, Claudio L

    2009-01-01

    Whereas insomnia is frequent in restless legs syndrome (RLS), little is known about daytime sleepiness. We studied a series of 27 consecutive patients with idiopathic RLS in order to identify the characteristics and evolution of excessive daytime sleepiness (EDS) under dopaminergic treatment. Patients were assessed by clinical examination, questionnaires and video-polysomnography (PSG). Sleepy patients, as defined by Epworth Sleepiness Scale (ESS) >10, were also assessed by the multiple sleep latency test (MSLT). We excluded RLS patients with other sleep-wake disorders, in particular chronic sleep deprivation. Mean age was 56 years, the mean International RLS Study Group Rating Scale score was 24 at baseline. Ten (37%) of the 27 patients reported EDS. RLS patients with sleepiness had a higher amount of total sleep time (p = 0.029) on PSG and a mean sleep latency of 6.4 min on MSLT. No other differences regarding clinical or polysomnographic parameters were found. RLS severity improved in all patients under dopaminergic treatment (p = 0.001); this was also the case for the ESS score in sleepy patients (p = 0.007). In our series of RLS patients, EDS was common, characterized by longer sleep (PSG) and reduced sleep latencies on MSLT. Under dopaminergic treatment, both RLS severity and ESS improved. Copyright 2009 S. Karger AG, Basel.

  12. [Nutritional status, healthy habits, quality of life and daytime sleepiness in nightlife workers of Córdoba].

    PubMed

    Moreno Linares, Vicente; Diéguez Cantueso, Inmaculada; Lara Carmona, Juan José; Molina Recio, Guillermo

    2015-04-01

    This study is aimed to analyze the factors that affect body composition, nutritional status, dietary habits, substance abuse (alcohol and smoking), physical activity, sleepiness disorders and self-rated health status in people working in nightlife in the city of Cordoba. Representative sample of 144 subjects (88 men and 56 women) with a mean age of 26.88 (± 4.7) years was studied. Individuals were analized for their body composition. Besides, a personal interview was used to administrate validated questionnaires to get other important data related to the aim of the study. The male group showed higher body mass index (p<0.05), showing overweight in more than half of the sample, and higher levels of body fat in 42% of subjects. Adherence to the Mediterranean diet was low and up to 48.6% presents a risk alcohol consumption, being higher in the male group. 40% of the subjects suffer from disorders of daytime sleepiness, however they spend a big amount of time in physical activities. The sample shows a high obesity and overweight prevalence and a low adherence to the mediterranean diet. Although they are not sedentary, the sample has unhealthy habits such as drinking alcohol and smoking and at the same time they suffer from sleepiness daytime disorders. In spite of they seems to have a high self-awareness about their own health status, 1 from every 5 individuals recognize that they could improve it. Copyright AULA MEDICA EDICIONES 2014. Published by AULA MEDICA. All rights reserved.

  13. Excessive daytime sleepiness in patients on intrathecal analgesia for chronic pain.

    PubMed

    Demartini, Laura; Fanfulla, Francesco; Bonezzi, Cesare; Armiento, Luciana; Buonocore, Michelangelo

    2015-01-01

    Intrathecal (IT) drug administration is an advanced technique in pain treatment algorithm for patients poorly responsive to systemic pharmacological treatment or less invasive techniques. The aim is to improve analgesia lowering side effects; despite this premise, many side effects of long-term IT therapy have been described, mainly related to opioid administration. We observed, in some of the patients regularly followed for pump refills in our Pain Unit, the appearance of excessive daytime sleepiness (EDS) interfering with daily life and work activity; this study aims to investigate the incidence of EDS in patients on IT analgesia with opioid or non-opioid drugs and its possible relationship with respiratory problems during sleep. 21 patients on IT therapy for chronic pain answered the Epworth Sleepiness Scale (ESS). The incidence of EDS in patients receiving IT opioids was compared to a control group not receiving opioids. In 10 patients, who performed polysomnography (PSG) and maintenance of wakefulness test (MWT) for sleep complaints, we studied the relationship between PSG data and ESS scores and we verified the concordance of ESS and MWT results. 38% of the patients reported EDS, according to ESS data; all the patients with EDS were receiving an IT opioid. Even if some patients presented sleep apneas, we failed to correlate this data with daytime sleepiness. Subjective sleepiness is confirmed by the results of MWT. Our data demonstrate that EDS is a frequent and important side effect of IT analgesia and it seems related to opioids administration.

  14. Sleep-Wake Cycle, Daytime Sleepiness, and Attention Components in Children Attending Preschool in the Morning and Afternoon Shifts

    ERIC Educational Resources Information Center

    Belísio, Aline S.; Kolodiuk, Fernanda F.; Louzada, Fernando M.; Valdez, Pablo; Azevedo, Carolina V. M.

    2017-01-01

    Children tend to sleep and wake up early and to exhibit daytime sleep episodes. To evaluate the impact of school start times on sleepiness and attention in preschool children, this study compared the temporal patterns of sleep, daytime sleepiness, and the components of attention between children aged 4-6 years that study in the morning (n = 66)…

  15. Spatial clusters of daytime sleepiness and association with nighttime noise levels in a Swiss general population (GeoHypnoLaus).

    PubMed

    Joost, Stéphane; Haba-Rubio, José; Himsl, Rebecca; Vollenweider, Peter; Preisig, Martin; Waeber, Gérard; Marques-Vidal, Pedro; Heinzer, Raphaël; Guessous, Idris

    2018-05-31

    Daytime sleepiness is highly prevalent in the general adult population and has been linked to an increased risk of workplace and vehicle accidents, lower professional performance and poorer health. Despite the established relationship between noise and daytime sleepiness, little research has explored the individual-level spatial distribution of noise-related sleep disturbances. We assessed the spatial dependence of daytime sleepiness and tested whether clusters of individuals exhibiting higher daytime sleepiness were characterized by higher nocturnal noise levels than other clusters. Population-based cross-sectional study, in the city of Lausanne, Switzerland. Sleepiness was measured using the Epworth Sleepiness Scale (ESS) for 3697 georeferenced individuals from the CoLaus|PsyCoLaus cohort (period = 2009-2012). We used the sonBASE georeferenced database produced by the Swiss Federal Office for the Environment to characterize nighttime road traffic noise exposure throughout the city. We used the GeoDa software program to calculate the Getis-Ord G i * statistics for unadjusted and adjusted ESS in order to detect spatial clusters of high and low ESS values. Modeled nighttime noise exposure from road and rail traffic was compared across ESS clusters. Daytime sleepiness was not randomly distributed and showed a significant spatial dependence. The median nighttime traffic noise exposure was significantly different across the three ESS Getis cluster classes (p < 0.001). The mean nighttime noise exposure in the high ESS cluster class was 47.6, dB(A) 5.2 dB(A) higher than in low clusters (p < 0.001) and 2.1 dB(A) higher than in the neutral class (p < 0.001). These associations were independent of major potential confounders including body mass index and neighborhood income level. Clusters of higher daytime sleepiness in adults are associated with higher median nighttime noise levels. The identification of these clusters can guide tailored public health

  16. Prevalence of poor sleep quality, sleepiness and obstructive sleep apnoea risk factors in athletes.

    PubMed

    Swinbourne, Richard; Gill, Nicholas; Vaile, Joanna; Smart, Daniel

    2016-10-01

    Despite the perceived importance of sleep for athletes, little is known regarding athlete sleep quality, their prevalence of daytime sleepiness or risk factors for obstructive sleep apnoea (OSA) such as snoring and witnessed apnoeic episodes. The purpose of the present study was to characterise normative sleep quality among highly trained team sport athletes. 175 elite or highly trained rugby sevens, rugby union and cricket athletes completed the Pittsburgh Sleep Quality Index (PSQI), Epworth Sleepiness Score (ESS) and Quality of Life questionnaires and an OSA risk factor screen. On average, athletes reported 7.9 ± 1.3 h of sleep per night. The average PSQI score was 5.9 ± 2.6, and 50% of athletes were found to be poor sleepers (PSQI > 5). Daytime sleepiness was prevalent throughout the population (average global score of 8.5) and clinically significant (ESS score of ≥10) in 28% of athletes. OSA may be an important clinical consideration within athletic populations, as a considerable number of athletes (38%) defined themselves as snorers and 8% reported having a witnessed apnoeic episode. The relationship between self-rated sleep quality and actual PSQI score was strong (Pearson correlation of 0.4 ± 0.1, 90% confidence limits). These findings suggest that this cohort of team sport athletes suffer a preponderance of poor sleep quality, with associated high levels of daytime sleepiness. Athletes should receive education about how to improve sleep wake schedules, extend total sleep time and improve sleep quality.

  17. Rotigotine Transdermal Patch Does Not Make Parkinson Disease Patients Sleepy During Daytime.

    PubMed

    Ohta, Kouichi; Osada, Takashi

    2015-01-01

    To assess quantitatively the influence of rotigotine transdermal patch on daytime sleepiness, the most common adverse event by non-ergot dopamine agonists (DAs), in Parkinson disease (PD) patients. An open-label study enrolled PD patients with unsatisfactory control of motor symptoms. Treatment with rotigotine transdermal patch was titrated to optimal dose (4-8 mg/24 hours) over 2 to 4 weeks. Primary outcome was Epworth Sleepiness Scale (ESS) for daytime sleepiness. Secondary outcomes included Hoehn&Yahr stage, time spent with dyskinesia, Clinical Global Impression of Improvement (CGI-I) of motor symptoms, adverse events, and compliance. The subjects were 31 PD patients (age 72 ± 8, Hoehn &Yahr stage 2.7 ± 0.9, mean ± SD). The ESS did not increase after rotigotine treatment (7.2 ± 4.9 before treatment, 6.2 ± 4.0 with 4 mg/24 hour, and 8.1 ± 6.4 with 8 mg/24 hour). The CGI-I score improved after treatment; responder rate reached 88.9% with 8 mg/24 hours. No patients showed worsening in other secondary outcomes. In 13 patients treated with equivalent doses of rotigotine switched from other DAs (pramipexole, ropinirole, and cabergoline), ESS did not increase after treatment (10.0 ± 4.6 before and 8.6 ± 4.5 after treatment) and decreased without worsening of CGI-I in 54% patients. Other secondary outcomes did not worsen after treatment. Twenty four-hour transdermal delivery of rotigotine at doses up to 8 mg/24 hours does not worsen the daytime sleepiness in PD patients and often improves it when switched from other non-ergot DAs. This is achieved together with satisfactory improvement in motor symptoms, demonstrating that this new modality of non-ergot DA is well tolerated and beneficial in PD patients.

  18. Trends in insomnia and excessive daytime sleepiness among U.S. adults from 2002 to 2012.

    PubMed

    Ford, Earl S; Cunningham, Timothy J; Giles, Wayne H; Croft, Janet B

    2015-03-01

    Insomnia is a prevalent disorder in the United States and elsewhere. It has been associated with a range of somatic and psychiatric conditions, and adversely affects quality of life, productivity at work, and school performance. The objective of this study was to examine the trend in self-reported insomnia and excessive daytime sleepiness among US adults. We used data of participants aged ≥18 years from the National Health Interview Survey for the years 2002 (30,970 participants), 2007 (23,344 participants), and 2012 (34,509 participants). The unadjusted prevalence of insomnia or trouble sleeping increased from 17.5% (representing 37.5 million adults) in 2002 to 19.2% (representing 46.2 million adults) in 2012 (relative increase: +8.0%) (P trend <0.001). The age-adjusted prevalence increased from 17.4% to 18.8%. Significant increases were present among participants aged 18-24, 25-34, 55-64, and 65-74 years, men, women, whites, Hispanics, participants with diabetes, and participants with joint pain. Large relative increases occurred among participants aged 18-24 years (+30.9%) and participants with diabetes (+27.0%). The age-adjusted percentage of participants who reported regularly having excessive daytime sleepiness increased from 9.8% to 12.7% (P trend <0.001). Significant increases were present in most demographic groups. The largest relative increase was among participants aged 25-34 years (+49%). Increases were also found among participants with hypertension, chronic obstructive pulmonary disease, asthma, and joint pain. Given the deleterious effects of insomnia on health and performance, the increasing prevalence of insomnia and excessive daytime sleepiness among US adults is a potentially troubling development. Published by Elsevier B.V.

  19. Excessive daytime sleepiness in Parkinson disease: a SPECT study.

    PubMed

    Matsui, Hideaki; Nishinaka, Kazuto; Oda, Masaya; Hara, Narihiro; Komatsu, Kenichi; Kubori, Tamotsu; Udaka, Fukashi

    2006-07-01

    The underlying pathologic mechanism of excessive daytime sleepiness (EDS) in Parkinson disease and the relative contributions of brain function to this process are poorly understood. We compared brain perfusion images between patients with Parkinson disease and EDS and those without EDS using n-isopropyl-p-1231 iodoamphetamine single photon emission computed tomography. Clinical study. Sumitomo Hospital. Thirteen patients with Parkinson disease with EDS (EDS group) and 27 patients with Parkinson disease without EDS (no-EDS group) were studied. Whether or not each case had EDS was determined according to the response to the Epworth Sleepiness Scale: patients with an Epworth Sleepiness Scale score > or = 10 were included in the EDS group, and patients with an Epworth Sleepiness Scale score < or = 9 were included in the no-EDS group. There were significant hypoperfusions in the left parietal and temporal association cortex in the EDS group. In the multivariable logistic regression model, attention and decreased regional cerebral blood flow of the left parietal association cortex and right caudate and increased regional cerebral blood flow of the right thalamus were the independent and significant factors. The cortical hypofunction relative to hyperfunction of the brain stem may relate to EDS in Parkinson disease. This is the first imaging study about EDS in Parkinson disease, and further studies are required.

  20. [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.

  1. Excessive Daytime Sleepiness as an Indicator of Depression in Hispanic Americans

    PubMed Central

    Malcarne, Vanessa L.; Wachsman, Solenne I.; Sadler, Georgia Robins

    2016-01-01

    Introduction Excessive daytime sleepiness (EDS) has been shown to be associated with depression; however, this relationship has not been confirmed among Hispanic Americans. Method This study examined the link between EDS and depression among Hispanic Americans (N = 411) and explored the potential moderating roles of age, gender, income, education, health status, and acculturation. The Epworth Sleepiness Scale and Patient Health Questionnaire–9 measured EDS and depression, respectively. Results Hierarchical linear regression demonstrated that EDS was significantly related to depression. Receiver operating characteristic curve analysis suggested that the Epworth Sleepiness Scale discriminated with adequate sensitivity and specificity between participants with moderately severe depression and those with less severe symptoms. No sociodemographic variables moderated the EDS–depression relationship. Conclusion These findings suggest that depression should be considered when Hispanic Americans present with EDS. PMID:27465932

  2. Excessive Daytime Sleepiness as an Indicator of Depression in Hispanic Americans.

    PubMed

    Nuyen, Brian A; Fox, Rina S; Malcarne, Vanessa L; Wachsman, Solenne I; Sadler, Georgia Robins

    2016-09-01

    Excessive daytime sleepiness (EDS) has been shown to be associated with depression; however, this relationship has not been confirmed among Hispanic Americans. This study examined the link between EDS and depression among Hispanic Americans (N = 411) and explored the potential moderating roles of age, gender, income, education, health status, and acculturation. The Epworth Sleepiness Scale and Patient Health Questionnaire-9 measured EDS and depression, respectively. Hierarchical linear regression demonstrated that EDS was significantly related to depression. Receiver operating characteristic curve analysis suggested that the Epworth Sleepiness Scale discriminated with adequate sensitivity and specificity between participants with moderately severe depression and those with less severe symptoms. No sociodemographic variables moderated the EDS-depression relationship. These findings suggest that depression should be considered when Hispanic Americans present with EDS. © The Author(s) 2016.

  3. Association of daytime sleepiness with obstructive sleep apnoea and comorbidities varies by sleepiness definition in a population cohort of men.

    PubMed

    Adams, Robert J; Appleton, Sarah L; Vakulin, Andrew; Lang, Carol; Martin, Sean A; Taylor, Anne W; McEvoy, R Doug; Antic, Nick A; Catcheside, Peter G; Wittert, Gary A

    2016-10-01

    To determine correlates of excessive daytime sleepiness (EDS) identified with the Epworth Sleepiness Scale (ESS) and a more broad definition, while accounting for obstructive sleep apnoea (OSA) in community dwelling men. Participants of the Men Androgens Inflammation Lifestyle Environment and Stress (MAILES) Study (n = 837, ≥ 40 years) without a prior OSA diagnosis, underwent in-home full unattended polysomnography (PSG, Embletta X100), completed the ESS, STOP questionnaire and Pittsburgh Sleep Quality Index in 2010-2011. In 2007-2010, questionnaires and biomedical assessment (in South Australian public hospital-based clinics) identified medical conditions. An alternate EDS definition (EDSAlt ) consisted of ≥ 2 of 3 problems (feeling sleepy sitting quietly; feeling tired/fatigued/sleepy; trouble staying awake). EDSAlt (30.4%, n = 253), but not ESS ≥ 11 (EDSESS , 12.6%, n = 104), increased significantly across OSA severity and body mass index categories. In adjusted analyses, EDSESS was significantly associated with depression: odds ratio (OR), 95%CI: 2.2 (1.3-3.8) and nocturia: 2.0 (1.3-3.2). EDSAlt was associated with depression, financial stress, relationship, work-life balance problems and associations with nocturia and diabetes were borderline. After excluding men with EDSESS , EDSAlt was associated with oxygen desaturation index (3%) ≥ 16 and the highest arousal index quartile but not with comorbidities. Sleepiness not necessarily leading to dozing, but not ESS ≥ 11, was related to sleep disordered breathing. Clinicians should be alert to (1) differing perspectives of sleepiness for investigation and treatment of OSA, and (2) the presence of depression and nocturia in men presenting with significant Epworth sleepiness regardless of the presence of OSA. © 2016 Asian Pacific Society of Respirology.

  4. Associations between physical activity, sedentary time, sleep duration and daytime sleepiness in US adults.

    PubMed

    McClain, James J; Lewin, Daniel S; Laposky, Aaron D; Kahle, Lisa; Berrigan, David

    2014-09-01

    To examine the associations between objectively measured physical activity (PA) or sedentary behavior and self-reported sleep duration or daytime sleepiness in a nationally representative sample of healthy US adults (N=2128). We report analyses of four aspects of sedentary behavior and PA derived from accelerometry data (minutes of sedentary time, activity counts/minute, Minutes of Moderate and Vigorous PA [MVPA], and MVPA in 10-minute bouts) versus self-report of sleep duration and frequency of daytime sleepiness from the 2005-2006 National Health and Nutrition Examination Survey. Age and sex dependence of associations between PA and sleep were observed. Aspects of PA were significantly lower in adults reporting more frequent daytime sleepiness in younger (20-39) and older (≥ 60) age groups, but not in middle-aged (40-59), respondents. In younger respondents, PA increased with sleep duration, but in middle aged and older respondents PA was either unrelated to sleep duration or lower in those reporting ≥ 8 h of sleep. Objectively measured sedentary time showed limited evidence of associations with sleep duration. Further research delineating the relationships between sleep and PA is important because both activities have been implicated in diverse health outcomes as well as in the etiology of obesity. Published by Elsevier Inc.

  5. Association of Mild Obstructive Sleep Apnea With Cognitive Performance, Excessive Daytime Sleepiness, and Quality of Life in the General Population: The Korean Genome and Epidemiology Study (KoGES).

    PubMed

    Kim, Hyun; Thomas, Robert J; Yun, Chang-Ho; Au, Rhoda; Lee, Seung Ku; Lee, Sunghee; Shin, Chol

    2017-05-01

    Research points to impaired cognitive performance in sleep clinic patients with obstructive sleep apnea (OSA). However, inconsistent findings from various epidemiologic studies make this relationship less generalizable. The current study investigated the association between OSA and functional outcome measures, such as cognition, daytime sleepiness, and quality of life, in a Korean general population sample. A total of 1492 participants from the Korean Genome and Epidemiology Study (KoGES) were included in the analyses. The presence of OSA measured by overnight polysomnography (PSG) was defined by apnea-hypopnea index (AHI) >5. Cognitive performance was determined with scores from a comprehensive neuropsychological battery. Excessive daytime sleepiness and quality of life were additionally measured through subjective reports. After adjusting for various demographic and medical characteristics, OSA was independently associated with lower performance in the Digit Symbol Test (52.73 ± 17.08 vs. 58.72 ± 18.03, OSA vs. not, p = .02). Hypoxia measures were not related to cognitive performance. OSA was associated with higher odds of displaying excessive daytime sleepiness (odds ratio = 1.72, 95% CI: 1.05-2.80), but there was no significant relationship between OSA and quality of life. Cognition was unexpectedly unaffected overall. However, OSA was associated with impairment in a multidomain test that taps skills generally associated with frontal lobe function. The results suggest that research on protective and adaptive brain mechanisms to OSA stress can provide unique insights into the brain-sleep interface. As the study runs longitudinally, it will enable future studies on the impact of OSA on cognitive decline. © Sleep Research Society 2017. Published by Oxford University Press on behalf of the Sleep Research Society. All rights reserved. For permissions, please e-mail journals.permissions@oup.com.

  6. Associations of Caffeinated Beverage Consumption and Screen Time with Excessive Daytime Sleepiness in Korean High School Students.

    PubMed

    Jun, Nuri; Lee, Aeri; Baik, Inkyung

    2017-01-01

    The present study investigated caffeinated beverage consumption and screen time in the association with excessive daytime sleepiness (EDS) and sleep duration. We conducted a cross-sectional study including 249 Korean male high school students. These participants responded to a questionnaire inquiring the information on lifestyle factors, consumption of caffeinated beverages, time spent for screen media, and sleep duration as well as to the Epworth Sleepiness Scale (ESS) questionnaire. EDS was defined as ESS scores of 9 or greater. Students with EDS consumed greater amount of chocolate/cocoa drinks and spent longer time for a TV and a mobile phone than those without EDS (p < 0.05). In addition, students with short sleep (≤ 6 hours) consumed greater amount of coffee than others whereas students with long sleep (> 8 hours) consumed greater amount of chocolate/cocoa drinks than others (p < 0.05). Screen time did not differ according to the categories of sleep duration. Although these findings do not support causal relationships, they suggest that screen time is associated with EDS, but not with sleep duration, and that consumption of certain types of caffeinated beverages is associated with EDS and sleep duration. Adolescents may need to reduce screen time and caffeine consumption to improve sleep quality and avoid daytime sleepiness.

  7. Sleep quality evaluation, chronotype, sleepiness and anxiety of Paralympic Brazilian athletes: Beijing 2008 Paralympic Games.

    PubMed

    Silva, Andressa; Queiroz, Sandra Souza; Winckler, Ciro; Vital, Roberto; Sousa, Ronnie Andrade; Fagundes, Vander; Tufik, Sergio; de Mello, Marco Túlio

    2012-02-01

    The objective of this study was to evaluate the sleep quality, sleepiness, chronotype and the anxiety level of Brazilian Paralympics athletes before the 2008 Beijing Paralympic Games. Cross-sectional study. Setting Exercise and Psychobiology Studies Center (CEPE) and Universidade Federal de São Paulo, an urban city in Brazil. A total of 27 Paralympics athletes of both genders (16 men and 11 women) with an average age of 28±6 years who practised athletics (track and field events) were evaluated. Sleep quality was evaluated using the Pittsburgh Scale and the Epworth Sleepiness Scale to evaluate sleepiness. Chronotype was determined by the Horne and Östberg questionnaire and anxiety through the State-Trait Anxiety Inventory. The evaluations were performed in Brazil 10 days before the competition. The study's results demonstrate that 83.3% of the athletes that presented excessive daytime sleepiness also had poor sleep quality. The authors noted that 71.4% were classified into the morning type and 72% of the athletes who presented a medium anxiety level also presented poor sleep quality. Athletes with poor sleep quality showed significantly lower sleep efficiency (p=0.0119) and greater sleep latency (p=0.0068) than athletes with good sleep quality. Athletes who presented excessive daytime sleepiness presented lower sleep efficiency compared to non-sleepy athletes (p=0.0241). The authors conclude that the majority of athletes presented poor sleep quality before the competition. This information should be taken into consideration whenever possible when scheduling rest, training and competition times.

  8. Dissociating Effects of Global SWS Disruption and Healthy Aging on Waking Performance and Daytime Sleepiness

    PubMed Central

    Groeger, John A.; Stanley, Neil; Deacon, Stephen; Dijk, Derk-Jan

    2014-01-01

    Study Objective: To contrast the effects of slow wave sleep (SWS) disruption and age on daytime functioning. Design: Daytime functioning was contrasted in three age cohorts, across two parallel 4-night randomized groups (baseline, two nights of SWS disruption or control, recovery sleep). Setting: Sleep research laboratory. Participants: 44 healthy young (20-30 y), 35 middle-aged (40-55 y), and 31 older (66-83 y) men and women. Interventions: Acoustic stimulation contingent on appearance of slow waves. Measurements and Results: Cognitive performance was assessed before sleep latency tests at five daily time-points. SWS disruption resulted in less positive affect, slower or impaired information processing and sustained attention, less precise motor control, and erroneous implementation, rather than inhibition, of well-practiced actions. These performance impairments had far smaller effect sizes than the increase in daytime sleepiness and differed from baseline to the same extent for each age group. At baseline, younger participants performed better than older participants across many cognitive domains, with largest effects on executive function, response time, sustained attention, and motor control. At baseline, the young were sleepier than other age groups. Conclusions: SWS has been considered a potential mediator of age-related decline in performance, although the effects of SWS disruption on daytime functioning have not been quantified across different cognitive domains nor directly compared to age-related changes in performance. The data imply that two nights of SWS disruption primarily leads to an increase in sleepiness with minor effects on other aspects of daytime functioning, which are different from the substantial effects of age. Citation: Groeger JA, Stanley N, Deacon S, Dijk DJ. Dissociating effects of global sws disruption and healthy aging on waking performance and daytime sleepiness. SLEEP 2014;37(6):1127-1142. PMID:24882908

  9. 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.

  10. Daytime exposure to bright light, as compared to dim light, decreases sleepiness and improves psychomotor vigilance performance.

    PubMed

    Phipps-Nelson, Jo; Redman, Jennifer R; Dijk, Derk-Jan; Rajaratnam, Shantha M W

    2003-09-01

    This study examined the effects of bright light exposure, as compared to dim light, on daytime subjective sleepiness, incidences of slow eye movements (SEMs), and psychomotor vigilance task (PVT) performance following 2 nights of sleep restriction. The study had a mixed factorial design with 2 independent variables: light condition (bright light, 1,000 lux; dim light, < 5 lux) and time of day. The dependent variables were subjective sleepiness, PVT performance, incidences of SEMs, and salivary melatonin levels. Sleep research laboratory at Monash University. Sixteen healthy adults (10 women and 6 men) aged 18 to 35 years (mean age 25 years, 3 months). Following 2 nights of sleep restriction (5 hours each night), participants were exposed to modified constant routine conditions. Eight participants were exposed to bright light from noon until 5:00 pm. Outside the bright light exposure period (9:00 am to noon, 5:00 pm to 9:00 pm) light levels were maintained at less than 5 lux. A second group of 8 participants served as controls for the bright light exposure and were exposed to dim light throughout the entire protocol. Bright light exposure reduced subjective sleepiness, decreased SEMs, and improved PVT performance compared to dim light. Bright lights had no effect on salivary melatonin. A significant positive correlation between PVT reaction times and subjective sleepiness was observed for both groups. Changes in SEMs did not correlate significantly with either subjective sleepiness or PVT performance. Daytime bright light exposure can reduce the impact of sleep loss on sleepiness levels and performance, as compared to dim light. These effects appear to be mediated by mechanisms that are separate from melatonin suppression. The results may assist in the development of treatments for daytime sleepiness.

  11. Effects of a selective educational system on fatigue, sleep problems, daytime sleepiness, and depression among senior high school adolescents in Taiwan.

    PubMed

    Chen, Tien-Yu; Chou, Yu-Ching; Tzeng, Nian-Sheng; Chang, Hsin-An; Kuo, Shin-Chang; Pan, Pei-Yin; Yeh, Yi-Wei; Yeh, Chin-Bin; Mao, Wei-Chung

    2015-01-01

    The aim of the study reported here was to clarify the effects of academic pressure on fatigue, sleep problems, daytime sleepiness, and depression among senior high school adolescents in Taiwan. This cross-sectional study enrolled 757 senior high school adolescents who were classified into four groups: Grade 1 (n=261), Grade 2 (n=228), Grade 3T (n=199; Grade 3 students who had another college entrance test to take), and Grade 3S (n=69; Grade 3 students who had succeeded in their college application). Fatigue, sleep quality, daytime sleepiness, and depression were assessed using the Chinese version of the Multidimensional Fatigue Symptom Inventory - Short Form, Pittsburgh Sleep Quality Index-Taiwan Form, the Chinese version of the Epworth Sleepiness Scale, and the Chinese version of the Beck Depression Inventory(®)-II (BDI-II), respectively. Physical, emotional, and mental fatigue scores were all higher in higher-grade groups. The Grade 3T (test) students had the worst fatigue severity, and the Grade 3S (success) students had the least fatigue severity. More than half of the students (60.9%) went to bed after 12 am, and they had on average 6.0 hours of sleep per night. More than 30% of the students in Grade 2 (37.3%) and Grades 3T/S (30.2%/30.4%) possibly had daily sleepiness problems. The students in Grade 3T had the worst BDI-II score (13.27±9.24), and the Grade 3S students had a much lower BDI-II score (7.91±6.13). Relatively high proportions of fatigue, sleep problems, daytime sleepiness, and depression among senior high school adolescents were found in our study. The severities of fatigue, sleep problems, and depression were significantly diminished in the group under less academic stress (Grade 3S). Our findings may increase the understanding of the mental health of senior high school students under academic pressure in Taiwan. Further large sample size and population-based study should be done for better understanding about this topic.

  12. Excessive Daytime Sleepiness is Associated with Longer Culprit Lesion and Adverse Outcomes in Patients with Coronary Artery Disease

    PubMed Central

    Lee, Chi-Hang; Ng, Wai-Yee; Hau, William; Ho, Hee-Hwa; Tai, Bee-Choo; Chan, Mark Y.; Richards, A. Mark; Tan, Huay-Cheem

    2013-01-01

    Study Objectives: We assessed whether excessive daytime sleepiness was associated with coronary plaque phenotype and subsequent adverse cardiovascular events. Methods: Prospective cohort study. Intravascular ultrasound (IVUS) examination of the culprit coronary stenosis was performed. The Epworth Sleepiness Scale (ESS) questionnaire was administered, and the patients were divided into 2 groups—(1) sleepier and (2) less sleepy—based on the ESS score. Adverse cardiovascular outcomes were defined as cardiac death, myocardial infarction, stroke, unplanned revascularization, or heart failure admission. Results: One hundred seventeen patients undergoing urgent or non-urgent coronary angiography were recruited. Compared with the less sleepy group (ESS ≤ 10, n = 87), the sleepier group (ESS > 10, n = 30) had higher serum levels of total cholesterol and of low-density-lipoprotein cholesterols (p < 0.05 for both). The IVUS examinations indicated coronary stenoses were longer in the sleepier group than in the less sleepy group (p = 0.011). The cumulative incidence of adverse cardiovascular events at 16-month follow-up was higher in the sleepier than the less sleepy group (12.5% versus 6.9%, p = 0.03). Cox regression analysis adjusting for age and smoking showed increased hazard of adverse cardiovascular events in sleepier group as compared to less sleepy group (HR = 3.44, 95% CI 1.01-11.72). Conclusion: In patients presenting with coronary artery disease, excessive daytime sleepiness based on ESS > 10 was associated with longer culprit lesions and future adverse cardiovascular events. Citation: Lee CH; Ng WY; Hau W; Ho HH; Tai BC; Chan MY; Richards AM; Tan HC. Excessive daytime sleepiness is associated with longer culprit lesion and adverse outcomes in patients with coronary artery disease. J Clin Sleep Med 2013;9(12):1267-1272. PMID:24340288

  13. Excessive daytime sleepiness in adult patients with ADHD as measured by the Maintenance of Wakefulness Test, an electrophysiologic measure.

    PubMed

    Bioulac, Stéphanie; Chaufton, Cyril; Taillard, Jacques; Claret, Astrid; Sagaspe, Patricia; Fabrigoule, Colette; Bouvard, Manuel P; Philip, Pierre

    2015-07-01

    To quantify the objective level of sleepiness in adult attention-deficit/hyperactivity disorder (ADHD) patients and to determine the relationship between excessive daytime sleepiness and simulated driving performance. Forty adult ADHD patients (DSM-IV criteria) and 19 matched healthy control subjects were included between June 30, 2010, and June 19, 2013. All participants completed the Epworth Sleepiness Scale and the Manchester Driving Behavior Questionnaire. After nocturnal polysomnography, they performed 2 neuropsychological tests, a 4 × 40-minute Maintenance of Wakefulness Test, and a 1-hour driving session. The primary outcome measure was the mean sleep latency on the Maintenance of Wakefulness Test. ADHD patients were divided into 3 groups defined by their Maintenance of Wakefulness Test scores. Participants (patients and control subjects) were allocated as follows: sleepy ADHD (0-19 min), intermediate ADHD (20-33 min), alert ADHD (34-40 min), and control group (34-40 min). The driving performance outcome was the mean standard deviation of lateral position of the vehicle during the simulated session. The group mean (SD) Epworth Sleepiness Scale score was higher in ADHD patients (12.1 [4.4]) than in controls (6.0 [2.7]) (P < .001). On the basis of the Maintenance of Wakefulness Test scores, 14 patients (35%) were in the sleepy group, 20 (50%) were in the intermediate group, and only 6 (15%) were in the alert group. Sleepy ADHD patients exhibited significantly deteriorated driving performance compared to the other 3 groups (P < .01). Our study shows that a significant proportion of adult ADHD patients exhibit an objective excessive daytime sleepiness, which, in addition, has an impact on simulated driving performance. Excessive daytime sleepiness, therefore, may be a key element needed to better evaluate these ADHD patients. ClinicalTrials.gov identifier: NCT01160874. © Copyright 2015 Physicians Postgraduate Press, Inc.

  14. Cross-Sectional Study of Obstructive Sleep Apnea Syndrome in Japanese Public Transportation Drivers: Its Prevalence and Association With Pathological Objective Daytime Sleepiness.

    PubMed

    Sasai-Sakuma, Taeko; Kikuchi, Katsunori; Inoue, Yuichi

    2016-05-01

    This study investigates obstructive sleep apnea syndrome (OSAS) prevalence among Japanese occupational drivers and factors associated with a pathological level of objective daytime sleepiness. Portable monitoring device (PMD) screening was applied to 2389 Japanese male public transportation traffic drivers. Nocturnal polysomnography (n-PSG) and multiple sleep latency tests (MSLT) were administered to subjects with apnea-hypopnea index (AHI) at least 15 on PMD. In all, 235 subjects were diagnosed as having OSAS (9.8%). AHI on n-PSG at least 40 and Epworth Sleepiness Scale score at least 11 were extracted as factors associated with mean sleep latency on MSLT less than 5 minutes. Prevalence of OSAS in male Japanese public transportation traffic drivers was 9.8% or greater. Individuals aware of excessive daytime sleepiness and with severe OSAS were inferred as exhibiting a pathological level of objective daytime sleepiness.

  15. Effects of an irregular bedtime schedule on sleep quality, daytime sleepiness, and fatigue among university students in Taiwan.

    PubMed

    Kang, Jiunn-Horng; Chen, Shih-Ching

    2009-07-19

    An irregular bedtime schedule is a prevalent problem in young adults, and could be a factor detrimentally affecting sleep quality. The goal of the present study was to explore the association between an irregular bedtime schedule and sleep quality, daytime sleepiness, and fatigue among undergraduate students in Taiwan. A total of 160 students underwent a semi-structured interview and completed a survey comprising 4 parts: Pittsburgh Sleep Quality Index (PSQI), Epworth Sleepiness Scale (ESS), Fatigue Severity Scale (FSS), and a rating of irregular bedtime frequency. Participants were grouped into 3 groups in terms of irregular bedtime frequency: low, intermediate, or high according to their 2-week sleep log. To screen for psychological disorders or distress that may have affected responses on the sleep assessment measures, the Chinese health questionnaire-12 (CHQ-12) was also administered. We found an increase in bedtime schedule irregularity to be significantly associated with a decrease in average sleep time per day (Spearman r = -0.22, p = 0.05). Multivariate regression analysis revealed that irregular bedtime frequency and average sleep time per day were correlated with PSQI scores, but not with ESS or FSS scores. A significant positive correlation between irregular bedtime frequency and PSQI scores was evident in the intermediate (partial r = 0.18, p = 0.02) and high (partial r = 0.15, p = 0.05) frequency groups as compared to low frequency group. The results of our study suggest a high prevalence of both an irregular bedtime schedule and insufficient sleep among university students in Taiwan. Students with an irregular bedtime schedule may experience poor sleep quality. We suggest further research that explores the mechanisms involved in an irregular bedtime schedule and the effectiveness of interventions for improving this condition.

  16. Sleep Disordered Breathing Symptoms and Daytime Sleepiness are Associated with Emotional Problems and Poor School Performance in Children

    PubMed Central

    Liu, Jianghong; Liu, Xianchen; Ji, Xiaopeng; Wang, Yingjie; Zhou, Guoping; Chen, Xinyin

    2016-01-01

    This study examined the prevalence of sleep disordered breathing (SDB) symptoms and their associations with daytime sleepiness, emotional problems, and school performance in Chinese children. Participants included 3,979 children (10.99 ± 0.99 years old) from four elementary schools in Jintan City, Jiangsu Province, China. Children completed a self-administered questionnaire on sleep behavior and emotional problems, while parents completed the Child Sleep Habit Questionnaire (CSHQ). SDB symptoms included 3 items: loud snoring, stopped breathing, and snorting/ gasping during sleep. Teachers rated the children's school performance. The prevalence rates of parent- and self-reported SDB symptoms were 17.2% and 10.1% for “sometimes” and 8.9% and 5.6% for “usually”. SDB symptoms, more prevalent in boys than in girls, increased the risks for depression, loneliness, and poor school performance. Daytime sleepiness mediated the relationship between SDB symptoms and depression, loneliness, and poor school performance. This study suggests the importance of early screening and intervention of SDB and daytime sleepiness in child behavioral and cognitive development. PMID:27289327

  17. A functional polymorphism in the promoter region of MAOA gene is associated with daytime sleepiness in healthy subjects.

    PubMed

    Ojeda, Diego A; Niño, Carmen L; López-León, Sandra; Camargo, Andrés; Adan, Ana; Forero, Diego A

    2014-02-15

    Excessive daytime sleepiness (EDS) is one of the main causes of car and industrial accidents and it is associated with increased morbidity and alterations in quality of life. Prevalence of EDS in the general population around the world ranges from 6.2 to 32.4%, with a heritability of 38-40%. However, few studies have explored the role of candidate genes in EDS. Monoamine oxidase A (MAOA) gene has an important role in the regulation of neurotransmitter levels and a large number of human behaviors. We hypothesized that a functional VNTR in the promoter region of the MAOA gene might be associated with daytime sleepiness in healthy individuals. The Epworth sleepiness scale (ESS) was applied to 210 Colombian healthy subjects (university students), which were genotyped for MAOA-uVNTR. MAOA-uVNTR showed a significant association with ESS scores (p = 0.01): 3/3 genotype carriers had the lowest scores. These results were supported by differences in MAOA-uVNTR frequencies between diurnal somnolence categories (p = 0.03). Our finding provides evidence for the first time that MAOA-uVNTR has a significant association with EDS in healthy subjects. Finally, these data suggest that functional variations in MAOA gene could have a role in other phenotypes of neuropsychiatric relevance. Copyright © 2013 Elsevier B.V. All rights reserved.

  18. Normative data on the diurnal pattern of the Karolinska Sleepiness Scale ratings and its relation to age, sex, work, stress, sleep quality and sickness absence/illness in a large sample of daytime workers.

    PubMed

    Åkerstedt, Torbjorn; Hallvig, David; Kecklund, Göran

    2017-10-01

    Self-rated sleepiness responds to sleep loss, time of day and work schedules. There is, however, a lack of a normative reference showing the diurnal pattern during a normal working day, compared with a day off, as well as differences depending on stress, sleep quality, sex, age and being sick listed. The present study sought to provide such data for the Karolinska Sleepiness Scale. Participants were 431 individuals working in medium-sized public service units. Sleepiness (Karolinska Sleepiness Scale, scale 1-9) was rated at six times a day for a working week and 2 days off (>90.000 ratings). The results show a clear circadian pattern, with high values during the morning (4.5 at 07:00 hours) and evening (6.0 at 22:00 hours), and with low values (3-4) during the 10:00-16:00 hours span. Women had significantly higher (0.5 units) Karolinska Sleepiness Scale values than men, as did younger individuals (0.3 units), those with stress (1.3 units above the low-stress group) and those with poor sleep quality (1.0 units above those with qood sleep quality). Days off showed reduced sleepiness (0.7 units), while being sick listed was associated with an increased sleepiness (0.8 units). Multiple regression analysis of mean sleepiness during the working week yielded mean daytime stress, mean sleep quality, age, and sex as predictors (not sleep duration). Improved sleep quality accounted for the reduced sleepiness during days off, but reduced stress was a second factor. Similar results were obtained in a longitudinal mixed-model regression analysis across the 7 days of the week. The percentage of ratings at Karolinska Sleepiness Scale risk levels (8 + 9) was 6.6%, but most of these were obtained at 22:00 hours. It was concluded that sleepiness ratings are strongly associated with time of day, sleep quality, stress, work day/day off, being ill, age, and sex. © 2017 European Sleep Research Society.

  19. Self-reported exposure to traffic pollution in relation to daytime sleepiness and habitual snoring: a questionnaire study in seven North-European cities.

    PubMed

    Gislason, Thorarinn; Bertelsen, Randi J; Real, Francisco Gomez; Sigsgaard, Torben; Franklin, Karl A; Lindberg, Eva; Janson, Christer; Arnardottir, Erna Sif; Hellgren, Johan; Benediktsdottir, Bryndis; Forsberg, Bertil; Johannessen, Ane

    2016-08-01

    Little is known about associations between traffic exposure and sleep disturbances. We examined if self-reported exposure to traffic is associated with habitual snoring and daytime sleepiness in a general population. In the RHINE III study, 12184 adults answered questions on sleep disturbances and traffic exposure. We analysed bedrooms near roads with traffic, bedrooms with traffic noise, and travelling regularly along busy roads as proxies for traffic exposures, using logistic regression. Adjustment factors were study centre, gender, age, smoking habits, educational level, body mass index, physical activity, obstructive sleep apnoea, and sleep duration. One in ten lived near a busy road, 6% slept in a bedroom with traffic noise, and 11% travelled regularly along busy roads. Habitual snoring affected 25% and daytime sleepiness 21%. More men reported snoring and more women reported daytime sleepiness. Having a bedroom with traffic noise was associated with snoring (adjusted OR 1.29, [95% CI 1.12, 1.48]). For daytime sleepiness, on the other hand, bedroom with traffic noise and high exposure to traffic pollution have significant risk factors (adjusted ORs 1.46 [1.11, 1.92] and 1.65 [1.11, 2.45]). Results were consistent across study centres. Daytime sleepiness is associated with traffic pollution and traffic noise, while habitual snoring is only associated with traffic noise. Self-reported traffic exposure should be taken into account when diagnosing and planning treatment for patients with sleep disturbances, because reducing noise and pollution exposure in the bedroom may have a beneficial effect. Copyright © 2016 Elsevier B.V. All rights reserved.

  20. The association between use of electronic media in bed before going to sleep and insomnia symptoms, daytime sleepiness, morningness, and chronotype.

    PubMed

    Fossum, Ingrid Nesdal; Nordnes, Linn Tinnesand; Storemark, Sunniva Straume; Bjorvatn, Bjørn; Pallesen, Ståle

    2014-09-03

    This study investigated whether the use of a television, computer, gaming console, tablet, mobile phone, or audio player in bed before going to sleep was associated with insomnia, daytime sleepiness, morningness, or chronotype. 532 students aged 18-39 were recruited from lectures or via e-mail. Respondents reported the frequency and average duration of their in-bed media use, as well as insomnia symptoms, daytime sleepiness, morningness-eveningness preference and bedtime/rise time on days off. Mean time of media use per night was 46.6 minutes. The results showed that computer usage for playing/surfing/reading was positively associated with insomnia, and negatively associated with morningness. Mobile phone usage for playing/surfing/texting was positively associated with insomnia and chronotype, and negatively associated with morningness. None of the other media devices were related to either of these variables, and no type of media use was related to daytime sleepiness.

  1. Does physical exercise reduce excessive daytime sleepiness by improving inflammatory profiles in obstructive sleep apnea patients?

    PubMed

    Alves, Eduardo da Silva; Ackel-D'Elia, Carolina; Luz, Gabriela Pontes; Cunha, Thays Crosara Abrahão; Carneiro, Gláucia; Tufik, Sergio; Bittencourt, Lia Rita Azeredo; de Mello, Marco Tulio

    2013-05-01

    Obstructive sleep apnea syndrome (OSAS) is associated with a variety of long-term consequences such as high rates of morbidity and mortality, due to excessive diurnal somnolence as well as cardiovascular and metabolic diseases. Obesity, recurrent episodes of upper airway obstruction, progressive hypoxemia, and sleep fragmentation during sleep cause neural, cardiovascular, and metabolic changes. These changes include activation of peripheral sympathetic nervous system and the hypothalamic-pituitary-adrenal axis, insulin sensitivity, and inflammatory cytokines alterations, which predispose an individual to vascular damage. Previous studies proposed that OSAS modulated the expression and secretion of inflammatory cytokines from fat and other tissues. Independent of obesity, patients with OSAS exhibited elevated levels of C-reactive protein, tumor necrosis factor-α and interleukin-6, which are associated with sleepiness, fatigue, and the development of a variety of metabolic and cardiovascular diseases. OSAS and obesity are strongly associated with each other and share many common pathways that induce chronic inflammation. Previous studies suggested that the protective effect of exercise may be partially attributed to the anti-inflammatory effect of regular exercise, and this effect was observed in obese patients. Although some studies assessed the effects of physical exercise on objective and subjective sleep parameters, the quality of life, and mood in patients with OSAS, no study has evaluated the effects of this treatment on inflammatory profiles. In this review, we cited some studies that directed our opinion to believe that since OSAS causes increased inflammation and has excessive daytime sleepiness as a symptom and being that physical exercise improves inflammatory profiles and possibly OSAS symptoms, it must be that physical exercise improves excessive daytime sleepiness due to its improvement in inflammatory profiles.

  2. Effects of armodafinil and cognitive behavior therapy for insomnia on sleep continuity and daytime sleepiness in cancer survivors.

    PubMed

    Garland, Sheila N; Roscoe, Joseph A; Heckler, Charles E; Barilla, Holly; Gehrman, Philip; Findley, James C; Peoples, Anita R; Morrow, Gary R; Kamen, Charles; Perlis, Michael L

    2016-04-01

    This study involves the analysis of a secondary outcome of a trial examining whether cognitive behavior therapy for insomnia (CBT-I), a wake-promoting medication (armodafinil), or both results in greater improvement in prospectively assessed sleep continuity and daytime sleepiness than a placebo-alone group among a heterogeneous group of cancer survivors. Whether or not armodafinil alone, and/or when combined with CBT-I, affected adherence with CBT-I was evaluated. This study is a randomized, placebo-controlled, clinical trial. This study was conducted at two northeastern academic medical centers. Eighty-eight cancer survivors with chronic insomnia were recruited between October 2008 and November 2012. Participants were assigned to one of four conditions: 1) CBT-I and placebo (CBT-I+P); 2) CBT-I and armodafinil (CBT-I + A); 2) armodafinil alone (ARM); or 4) placebo alone (PLA). CBT-I was delivered in seven weekly individual therapy sessions (three in person, four via telephone). The armodafinil dosage was 50 mg BID. Sleep continuity was measured with daily sleep diaries assessing sleep latency (SL), wake after sleep onset (WASO), and total sleep time (TST). The Epworth Sleepiness Scale (ESS) measured daytime sleepiness. Compared to the PLA group, the CBT-I+P and CBT-I+A groups reported a significant reduction in SL with effect sizes of 0.67 and 0.58, respectively. A significant reduction was observed in WASO in the CBT-I+A group with an effect size of 0.64. An increasing trend of TST was observed in the CBT-I+P, CBT-I+A, and PLA groups, but not in the ARM group. No statistically significant reductions in daytime sleepiness (ESS) were observed for any of the groups. CBT-I alone and in combination with armodafinil caused significant improvement in sleep continuity. The addition of armodafinil did not appear to improve daytime sleepiness or enhance adherence to CBT-I. Copyright © 2015 Elsevier B.V. All rights reserved.

  3. Effects of Armodafinil and Cognitive Behavior Therapy for Insomnia on Sleep Continuity and Daytime Sleepiness in Cancer Survivors

    PubMed Central

    Garland, Sheila N.; Roscoe, Joseph A.; Heckler, Charles E.; Barilla, Holly; Gehrman, Philip; Findley, James C.; Peoples, Anita R.; Morrow, Gary R.; Kamen, Charles; Perlis, Michael L.

    2016-01-01

    Study Objectives This study involves the analysis of a secondary outcome of a trial examining whether cognitive behavior therapy for insomnia (CBT-I), a wake-promoting medication (armodafinil), or both results in greater improvement in prospectively assessed sleep continuity and daytime sleepiness than a placebo-alone group among a heterogeneous group of cancer survivors. Whether or not armodafinil alone, and/or when combined with CBT-I, affected adherence with CBT-I was evaluated. Design This study is a randomized, placebo-controlled, clinical trial. Setting This study was conducted at two northeastern academic medical centers. Participants Eighty-eight cancer survivors with chronic insomnia were recruited between October 2008 and November 2012. Participants were assigned to one of four conditions: 1) CBT-I and placebo (CBTI+P); 2) CBT-I and armodafinil (CBT-I+A); 2) armodafinil alone (ARM); or 4) placebo alone (PLA). Interventions CBT-I was delivered in seven weekly individual therapy sessions (three in person, four via telephone). The armodafinil dosage was 50 mg BID. Measurements and Results Sleep continuity was measured with daily sleep diaries assessing sleep latency (SL), wake after sleep onset (WASO), and total sleep time (TST). The Epworth Sleepiness Scale (ESS) measured daytime sleepiness. Compared to the PLA group, the CBT+P and CBT+A groups reported a significant reduction in SL with effect sizes of 0.67 and 0.58, respectively. A significant reduction was observed in WASO in the CBT+A group with an effect size of 0.64. An increasing trend of TST was observed in the CBT+P, CBT+A, and PLA groups, but not in the ARM group. No statistically significant reductions in daytime sleepiness (ESS) were observed for any of the groups. Conclusion CBT-I alone and in combination with armodafinil caused significant improvement in sleep continuity. The addition of armodafinil did not appear to improve daytime sleepiness or enhance adherence to CBT-I. PMID:27318221

  4. Daytime Sleepiness and Driving Performance in Patients with Obstructive Sleep Apnea: Comparison of the MSLT, the MWT, and a Simulated Driving Task

    PubMed Central

    Pizza, Fabio; Contardi, Sara; Mondini, Susanna; Trentin, Lino; Cirignotta, Fabio

    2009-01-01

    Study Objectives: To test the reliability of a driving-simulation test for the objective measurement of daytime alertness compared with the Multiple Sleep Latency Test (MSLT) and with the Maintenance of Wakefulness Test (MWT), and to test the ability to drive safely, in comparison with on-road history, in the clinical setting of untreated severe obstructive sleep apnea. Design: N/A. Setting: Sleep laboratory. Patients or Participants: Twenty-four patients with severe obstructive sleep apnea and reported daytime sleepiness varying in severity (as measured by the Epworth Sleepiness Scale). Interventions: N/A. Measurements and Results: Patients underwent MSLT and MWT coupled with 4 sessions of driving-simulation test on 2 different days randomly distributed 1 week apart. Simulated-driving performance (in terms of lane-position variability and crash occurrence) was correlated with sleep latency on the MSLT and more significantly on the MWT, showing a predictive validity toward the detection of sleepy versus alert patients with obstructive sleep apnea. In addition, patients reporting excessive daytime sleepiness or a history of car crashes showed poorer performances on the driving simulator. Conclusions: A simulated driving test is a suitable tool for objective measurement of daytime alertness in patients with obstructive sleep apnea. Further studies are needed to clarify the association between simulated-driving performance and on-road crash risk of patients with sleep disordered breathing. Citation: Pizza F; Contardi S; Mondini S; Trentin L; Cirignotta F. Daytime sleepiness and driving performance in patients with obstructive sleep apnea: comparison of the MSLT, the MWT, and a simulated driving task. SLEEP 2009;32(3):382-391. PMID:19294958

  5. Effect of CPAP Therapy in Improving Daytime Sleepiness in Indian Patients with Moderate and Severe OSA.

    PubMed

    Battan, Gulshan; Kumar, Sanjeev; Panwar, Ajay; Atam, Virendra; Kumar, Pradeep; Gangwar, Anil; Roy, Ujjawal

    2016-11-01

    Obstructive Sleep Apnoea (OSA) is a highly prevalent disease and a major public health issue in India. Excessive daytime sleepiness is an almost ubiquitous symptom of OSA. Epworth Sleepiness Scale (ESS) score is a validated objective score to measure the degree of daytime sleepiness. Continuous Positive Airway Pressure (CPAP) therapy has been established as the gold standard treatment modality for OSA patients. A few Indian studies have reported the effectiveness of CPAP therapy in improving ESS scores after 1 st month of CPAP use. To observe both, short-term (one month) and long-term (three month) effects of CPAP therapy on ESS scores in moderate to severe OSA patients. The patients complaining of excessive day-time sleepiness, snoring and choking episodes during sleep, consecutively presenting to medicine OPD over a period of 2 years, were subjected to Polysomnography (PSG). Seventy-three patients with apnoea-hypopnea index (AHI) ≥15 were categorised as having moderate to severe forms of OSA (moderate OSA with AHI=15-30 and severe OSA with AHI >30), and were scheduled for an initial trial of CPAP therapy. Forty-seven patients reported good tolerance to CPAP therapy after a trial period of 2 weeks and comprised the final study group. ESS scores in these patients were recorded at the baseline, and after 1 st and 3 rd month of CPAP therapy, and statistically analysed for significance. Mean ESS score at the baseline among moderate and severe OSA patients were 13.67±2.29 and 16.56 ±1.87, respectively. ESS score in both these subgroups improved significantly to 11.63±3.79, p=0.022, CI (0.3293-4.0106)} and 14.13 ±3.74, p < 0.001, CI (1.2991-4.5408), respectively after one month of CPAP therapy. Likewise, mean ESS scores among moderate and severe OSA patients improved significantly to 9.84 ±2.97, p = 0.022, CI (0.3293-4.0106) and 12.29 ±3.97, p <0.001, CI (2.9414-6.1385), respectively after three months of CPAP therapy. The result of the present study shows that

  6. Personal sleep debt and daytime sleepiness mediate the relationship between sleep and mental health outcomes in young adults.

    PubMed

    Dickinson, David L; Wolkow, Alexander P; Rajaratnam, Shantha M W; Drummond, Sean P A

    2018-05-22

    Sleep duration and chronotype (i.e., morningness-eveningness) are associated with increased depression and anxiety risk, but differences in individual sleep need and lifestyle may mean these sleep parameters do not present the same risk across all individuals. This study explored the mediating role of sleep debt and daytime sleepiness in the relationship between sleep and mental health symptoms in young adults, a particularly vulnerable population. Young adult university students (n = 2,218) and young adults from the general population in the United States (n = 992) provided estimates of actual and optimal sleep duration, and completed validated measures of sleepiness, chronotype, and depression and anxiety risk. Mediation models examining sleepiness and sleep debt (i.e., difference between optimal and actual sleep) as parallel mediators were tested. Sleepiness and sleep debt mediated the relationship between short sleep and depression and anxiety risk in the university sample, while sleepiness mediated these relationships in the general population sample. Sleepiness and sleep debt also mediated the impact of evening-type preferences on depression and anxiety risk in university students, but no mediation of this effect was found in young adults from the general population. This study reports potential mediating mechanisms related to the increased mental health risk conferred by short sleep and evening chronotype. These results have implications for how primary care physicians assess psychopathology risk, arguing for a focus on the assessment of daytime sleepiness and sleep debt in university populations, while for young adults in the general population, these factors may be less important. © 2018 Wiley Periodicals, Inc.

  7. Associations of Bullying, Victimization, and Daytime Sleepiness With Academic Problems in Adolescents Attending an Alternative High School.

    PubMed

    Rubens, Sonia L; Miller, Molly A; Zeringue, Megan M; Laird, Robert D

    2018-01-22

    Adolescents attending alternative high schools often present with high rates of academic and behavior problems. They are also at increased risk of poor health behaviors and engaging in physical violence compared with students in traditional high school settings. To address the needs of students in these educational settings, examining factors that influence academic problems in this population is essential. Research has established that both bullying/victimization and sleep problems increase adolescents' risk for academic problems. Little is known about how these 2 factors together may exacerbate risk for academic problems among students attending an alternative high school. The current study investigated the interaction between teacher-reported bullying, victimization and daytime sleepiness on academic concerns (attention and learning problems) among a sample of 172 students (56% female; age M = 18.07 years, SD = 1.42) attending an alternative high school in a large, Southeastern U.S. city. Findings from path models indicated that daytime sleepiness, bullying, and victimization were uniquely associated with attention and learning problems. Further, significant interactions indicated that the association between victimization/bullying and attention/learning problems weakened as levels of daytime sleepiness increased. Results suggest the importance of assessing and addressing multiple contextual risk factors in adolescents attending alternative high schools to provide comprehensive intervention for students in these settings. (PsycINFO Database Record (c) 2018 APA, all rights reserved).

  8. Lack of regular exercise, depression, and degree of apnea are predictors of excessive daytime sleepiness in patients with sleep apnea: sex differences.

    PubMed

    Basta, Maria; Lin, Hung-Mo; Pejovic, Slobodanka; Sarrigiannidis, Alexios; Bixler, Edward; Vgontzas, Alexandros N

    2008-02-15

    Apnea, depression, and metabolic abnormalities are independent predictors of excessive daytime sleepiness (EDS) in patients with sleep apnea. Exercise is beneficial for apnea, depression, and metabolic abnormalities; however, its association with EDS is not known. To evaluate the contribution of lack of regular exercise, depression, and apnea severity on daytime sleepiness in patients with sleep apnea. One thousand one hundred six consecutive patients (741 men and 365 women) referred to the sleep disorders clinic for symptoms consistent with sleep apnea. Daytime sleepiness was assessed with the Epworth Sleepiness Scale and activity was evaluated with a quantifiable Physical Activity Questionnaire. Compared with women, men had a higher apnea hypopnea index (AHI) (40.4 +/- 1.2 vs 31.0 +/- 1.8), lower body mass index (BMI) (35.3 +/- 0.3 kg/m2 vs 39.6 +/- 0.5 kg/m2), and higher rate of regular exercise (39.1% vs 28.8%) ( p < 0.05). Linear regression analysis of the total sample after adjusting for age, BMI, sex, central nervous system medication, and diabetes showed that logAHI, depression, and lack of regular exercise were significant predictors of sleepiness. Predictors of mild or moderate sleepiness for both sexes were depression and logAHI, whereas predictors of severe sleepiness for men were lack of regular exercise, depression, and minimum SaO2 and, for women, logAHI. In obese apneic patients, lack of regular exercise (only in men), depression, and degree of apnea are significant predictors of EDS. This association is modified by sex and degree of sleepiness. Assessment and management of depression and physical exercise should be part of a thorough evaluation of patients with sleep apnea.

  9. The prevalence of neck-shoulder pain, back pain and psychological symptoms in association with daytime sleepiness - a prospective follow-up study of school children aged 10 to 15.

    PubMed

    Gustafsson, Marja-Liisa; Laaksonen, Camilla; Aromaa, Minna; Löyttyniemi, Eliisa; Salanterä, Sanna

    2018-04-17

    Chronic and recurrent pain is prevalent in adolescents and generally girls report more pain symptoms than boys. Also, pain symptoms and sleep problems often co-occur. Pain symptoms have negative effects on school achievement, emotional well-being, sleep, and overall health and well-being. For effective intervention and prevention there is a need for defining factors associated with pain symptoms and daytime sleepiness. The aim of this longitudinal study was to investigate the prevalence and association between neck-shoulder pain, back pain, psychological symptoms and daytime sleepiness in 10-, 12- and 15-year-old children. This study is the first that followed up the same cohort of children from the age of 10 to 15. A cohort study design with three measurement points was used. Participants (n=568) were recruited from an elementary school cohort in a city of 1,75,000 inhabitants in South-Western Finland. Symptoms and daytime sleepiness were measured with self-administered questionnaires. Regression models were used to analyze the associations. Frequent neck-shoulder pain and back pain, and psychological symptoms, as well as daytime sleepiness, are already common at the age of 10 and increase strongly between the ages 12 and 15. Overall a greater proportion of girls suffered from pain symptoms and daytime sleepiness compared to boys. Daytime sleepiness in all ages associated positively with the frequency of neck-shoulder pain and back pain. The more that daytime sleepiness existed, the more neck-shoulder pain and back pain occurred. Daytime sleepiness at the age of 10 predicted neck-shoulder pain at the age of 15, and back pain at the age of 10 indicated that there would also be back pain at the age of 15. In addition, positive associations between psychological symptoms and neck-shoulder pain, as well as back pain, were observed. Subjects with psychological problems suffered neck-shoulder pain and back pain more frequently. This study is the first study that has

  10. The association of dopamine agonists with daytime sleepiness, sleep problems and quality of life in patients with Parkinson's disease--a prospective study.

    PubMed

    Happe, S; Berger, K

    2001-12-01

    Reports that dopamine agonists (DA) precipitate sudden daytime sleep episodes in Parkinson's disease (PD) patients have received widespread attention. It remains unclear if non-ergoline and ergoline DAs have differential sedating effects or if sedation rather represents a class effect of DAs. The aim of this study was the evaluation of sleep disturbances and the quality of life (QoL) in PD patients with different dopaminergic treatment strategies. This analysis is part of the FAQT-study, a prospective German cohort study evaluating determinants of QoL in PD patients. A subgroup of 111 PD patients was evaluated twice, at baseline and after one year of follow-up, using standardised and validated questionnaires (Unified Parkinson's disease rating scale (UPDRS), Hoehn and Yahr classification, Center for Epidemiologic Studies Depression Scale (CESD), Short Form-36 (SF-36), Parkinson Disease Questionnaire (PDQ-39)). The impact of treatment strategies on sleep problems, daytime sleepiness, bad dreams and hallucinations, depression and QoL in PD patients was analysed separately for ergoline DAs, non-ergoline DAs and the patient group taking no DA. At baseline, sleep problems were reported by about one third of the patients with and without DA medication. Excessive daytime sleepiness (EDS) was higher in the two DA groups (ergoline 11.9%, non-ergoline 9.1%) than among patients not taking DAs (4.5 %). At follow-up, sleep problems in general had decreased among patients taking DAs continuously and among those newly taking DAs, while the sleep problems increased in patients discontinuing DAs. However, EDS had increased to 25% in patients newly taking DAs, and decreased to 15.9% in those taking them continuously. QoL scores at follow-up were slightly increased in the patient groups newly taking and discontinuing DAs (the latter except in physical functioning) while those on continuing DA-medication remained unchanged. No differential effects of ergoline or non-ergoline DAs on

  11. Timed Light Therapy for Sleep and Daytime Sleepiness Associated With Parkinson Disease: A Randomized Clinical Trial.

    PubMed

    Videnovic, Aleksandar; Klerman, Elizabeth B; Wang, Wei; Marconi, Angelica; Kuhta, Teresa; Zee, Phyllis C

    2017-04-01

    Impaired sleep and alertness are some of the most common nonmotor manifestations of Parkinson disease (PD) and currently have only limited treatment options. Light therapy (LT), a widely available treatment modality in sleep medicine, has not been systematically studied in the PD population. To determine the safety and efficacy of LT on excessive daytime sleepiness (EDS) associated with PD. This randomized, placebo-controlled, clinical intervention study was set in PD centers at Northwestern University and Rush University. Participants were 31 patients with PD receiving stable dopaminergic therapy with coexistent EDS, as assessed by an Epworth Sleepiness Scale score of 12 or greater, and without cognitive impairment or primary sleep disorder. Participants were randomized 1:1 to receive bright LT or dim-red LT (controlled condition) twice daily in 1-hour intervals for 14 days. This trial was conducted between March 1, 2007, and October 31, 2012. Data analysis of the intention-to-treat population was conducted from November 1, 2012, through April 30, 2016. The primary outcome measure was the change in the Epworth Sleepiness Scale score comparing the bright LT with the dim-red LT. Secondary outcome measures included the Pittsburgh Sleep Quality Index score, the Parkinson's Disease Sleep Scale score, the visual analog scale score for daytime sleepiness, and sleep log-derived and actigraphy-derived metrics. Among the 31 patients (13 males and 18 females; mean [SD] disease duration, 5.9 [3.6] years), bright LT resulted in significant improvements in EDS, as assessed by the Epworth Sleepiness Scale score (mean [SD], 15.81 [3.10] at baseline vs 11.19 [3.31] after the intervention). Both bright LT and dim-red LT were associated with improvements in sleep quality as captured by mean (SD) scores on the Pittsburg Sleep Quality Index (7.88 [4.11] at baseline vs 6.25 [4.27] after bright LT, and 8.87 [2.83] at baseline vs 7.33 [3.52] after dim-red LT) and the Parkinson's Disease

  12. Effects of an alternating work shift on air traffic controllers and the relationship with excessive daytime sleepiness and stress.

    PubMed

    Freitas, Ângela M; Portuguez, Mirna Wetters; Russomano, Thaís; Freitas, Marcos de; Silvello, Silvio Luis da Silva; Costa, Jaderson Costa da

    2017-10-01

    To evaluate symptoms of stress and excessive daytime sleepiness (EDS) in air traffic control (ATC) officers in Brazil. Fifty-two ATC officers participated, based at three air traffic control units, identified as A, B and C. Stress symptoms were assessed using the Lipp Inventory of Stress Symptoms for Adults, and EDS by the Epworth Sleepiness Scale. The sample mean age was 37 years, 76.9% of whom were male. Excessive daytime sleepiness was identified in 25% of the ATC officers, with 84.6% of these based at air traffic control unit A, which has greater air traffic flow, operating a 24-hour alternating work shift schedule. A total of 16% of the ATC officers had stress symptoms, and of these, 62% showed a predominance of physical symptoms. The high percentage of ATC officers with EDS identified in group A may be related to chronodisruption due to night work and alternating shifts.

  13. Myotonic dystrophy type 1, daytime sleepiness and REM sleep dysregulation.

    PubMed

    Dauvilliers, Yves A; Laberge, Luc

    2012-12-01

    Myotonic dystrophy type 1 (DM1), or Steinert's disease, is the most common adult-onset form of muscular dystrophy. DM1 also constitutes the neuromuscular condition with the most significant sleep disorders including excessive daytime sleepiness (EDS), central and obstructive sleep apneas, restless legs syndrome (RLS), periodic leg movements in wake (PLMW) and periodic leg movements in sleep (PLMS) as well as nocturnal and diurnal rapid eye movement (REM) sleep dysregulation. EDS is the most frequent non-muscular complaint in DM1, being present in about 70-80% of patients. Different phenotypes of sleep-related problems may mimic several sleep disorders, including idiopathic hypersomnia, narcolepsy without cataplexy, sleep apnea syndrome, and periodic leg movement disorder. Subjective and objective daytime sleepiness may be associated with the degree of muscular impairment. However, available evidence suggests that DM1-related EDS is primarily caused by a central dysfunction of sleep regulation rather than by sleep fragmentation, sleep-related respiratory events or periodic leg movements. EDS also tends to persist despite successful treatment of sleep-disordered breathing in DM1 patients. As EDS clearly impacts on physical and social functioning of DM1 patients, studies are needed to identify the best appropriate tools to identify hypersomnia, and clarify the indications for polysomnography (PSG) and multiple sleep latency test (MSLT) in DM1. In addition, further structured trials of assisted nocturnal ventilation and randomized trials of central nervous system (CNS) stimulant drugs in large samples of DM1 patients are required to optimally treat patients affected by this progressive, incurable condition. Copyright © 2012 Elsevier Ltd. All rights reserved.

  14. Natural History of Excessive Daytime Sleepiness: Role of Obesity, Weight Loss, Depression, and Sleep Propensity

    PubMed Central

    Fernandez-Mendoza, Julio; Vgontzas, Alexandros N.; Kritikou, Ilia; Calhoun, Susan L.; Liao, Duanping; Bixler, Edward O.

    2015-01-01

    Study Objectives: Excessive daytime sleepiness (EDS) is highly prevalent in the general population and is associated with occupational and public safety hazards. However, no study has examined the clinical and polysomnographic (PSG) predictors of the natural history of EDS. Design: Representative longitudinal study. Setting: Sleep laboratory. Participants: From a random, general population sample of 1,741 individuals of the Penn State Adult Cohort, 1,395 were followed up after 7.5 years. Measurements and Results: Full medical evaluation and 1-night PSG at baseline and standardized telephone interview at follow-up. The incidence of EDS was 8.2%, while its persistence and remission were 38% and 62%, respectively. Obesity and weight gain were associated with the incidence and persistence of EDS, while weight loss was associated with its remission. Significant interactions between depression and PSG parameters on incident EDS showed that, in depressed individuals, incident EDS was associated with sleep disturbances, while in non-depressed individuals, incident EDS was associated with increased physiologic sleep propensity. Diabetes, allergy/asthma, anemia, and sleep complaints also predicted the natural history of EDS. Conclusions: Obesity, a disorder of epidemic proportions, is a major risk factor for the incidence and chronicity of excessive daytime sleepiness (EDS), while weight loss is associated with its remission. Interestingly, objective sleep disturbances predict incident EDS in depressed individuals, whereas physiologic sleep propensity predicts incident EDS in those without depression. Weight management and treatment of depression and sleep disorders should be part of public health policies. Citation: Fernandez-Mendoza J, Vgontzas AN, Kritikou I, Calhoun SL, Liao D, Bixler EO. Natural history of excessive daytime sleepiness: role of obesity, weight loss, depression, and sleep propensity. SLEEP 2015;38(3):351–360. PMID:25581913

  15. Morningness-eveningness and daytime functioning in university students: the mediating role of sleep characteristics.

    PubMed

    Bakotic, Marija; Radosevic-Vidacek, Biserka; Koscec Bjelajac, Adrijana

    2017-04-01

    The aim of this study was to explore the mediating role of sleep characteristics in the relationship between morningness-eveningness and three different aspects of daytime functioning: daytime sleepiness, depressive mood and substance use in university students. A multiple mediator model was proposed with sleep debt, poor sleep quality and bedtime delay at weekends as parallel mediators in these relationships. We analysed the data of 1052 university students aged 18-25 years who completed a modified version of the School Sleep Habits Survey, which included questions on sleep and the Composite Scale of Morningness, Sleepiness Scale, Depressive Mood Scale and Substance Use Scale. Students with more pronounced eveningness reported greater daytime sleepiness, greater depressive mood and more frequent substance use, as well as greater sleep debt, poorer sleep quality and greater bedtime delay at weekends. Mediation analyses indicated that morningness-eveningness affected daytime sleepiness and substance use both directly and indirectly through all proposed sleep-related mediators. However, the effect of morningness-eveningness on depressive mood was entirely indirect and was accounted for more by poor sleep than by sleep debt or bedtime irregularity. In conclusion, there are multiple possible mechanisms through which morningness-eveningness affects daytime functioning in university students, and sleep characteristics are a significant mechanism. Sleep debt, poor sleep quality and bedtime irregularity can, to a significant extent, explain the feeling of daytime sleepiness and greater substance use in students with eveningness preferences. However, more depressed mood in the evening-orientated students is primarily a consequence of their poor sleep quality. © 2016 European Sleep Research Society.

  16. Continuous positive airway pressure improves sleep and daytime sleepiness in patients with Parkinson disease and sleep apnea.

    PubMed

    Neikrug, Ariel B; Liu, Lianqi; Avanzino, Julie A; Maglione, Jeanne E; Natarajan, Loki; Bradley, Lenette; Maugeri, Alex; Corey-Bloom, Jody; Palmer, Barton W; Loredo, Jose S; Ancoli-Israel, Sonia

    2014-01-01

    Obstructive sleep apnea (OSA), common in Parkinson disease (PD), contributes to sleep disturbances and daytime sleepiness. We assessed the effect of continuous positive airway pressure (CPAP) on OSA, sleep, and daytime sleepiness in patients with PD. This was a randomized placebo-controlled, crossover design. Patients with PD and OSA were randomized into 6 w of therapeutic treatment or 3 w of placebo followed by 3 w of therapeutic treatment. Patients were evaluated by polysomnography (PSG) and multiple sleep latency test (MSLT) pretreatment (baseline), after 3 w, and after 6 w of CPAP treatment. Analyses included mixed models, paired analysis, and within-group analyses comparing 3 w to 6 w of treatment. Sleep laboratory. Thirty-eight patients with PD (mean age = 67.2 ± 9.2 y; 12 females). Continuous positive airway pressure. PSG OUTCOME MEASURES: sleep efficiency, %sleep stages (N1, N2, N3, R), arousal index, apnea-hypopnea index (AHI), and % time oxygen saturation < 90% (%time SaO2 < 90%). MSLT outcome measures: mean sleep-onset latency (MSL). There were significant group-by-time interactions for AHI (P < 0.001), % time SaO2 < 90% (P = 0.02), %N2 (P = 0.015) and %N3 (P = 0.014). Subjects receiving therapeutic CPAP showed significant decrease in AHI, %time SaO2 < 90%, %N2, and significant increase in %N3 indicating effectiveness of CPAP in the treatment of OSA, improvement in nighttime oxygenation, and in deepening sleep. The paired sample analyses revealed that 3 w of therapeutic treatment resulted in significant decreases in arousal index (t = 3.4, P = 0.002). All improvements after 3 w were maintained at 6 w. Finally, 3 w of therapeutic CPAP also resulted in overall decreases in daytime sleepiness (P = 0.011). Therapeutic continuous positive airway pressure versus placebo was effective in reducing apnea events, improving oxygen saturation, and deepening sleep in patients with Parkinson disease and obstructive sleep apnea. Additionally, arousal index was reduced

  17. How is good and poor sleep in older adults and college students related to daytime sleepiness, fatigue, and ability to concentrate?

    PubMed

    Alapin, I; Fichten, C S; Libman, E; Creti, L; Bailes, S; Wright, J

    2000-11-01

    We compared good sleepers with minimally and highly distressed poor sleepers on three measures of daytime functioning: self-reported fatigue, sleepiness, and cognitive inefficiency. In two samples (194 older adults, 136 college students), we tested the hypotheses that (1) poor sleepers experience more problems with daytime functioning than good sleepers, (2) highly distressed poor sleepers report greater impairment in functioning during the day than either good sleepers or minimally distressed poor sleepers, (3) daytime symptoms are more closely related to psychological adjustment and to psychologically laden sleep variables than to quantitative sleep parameters, and (4) daytime symptoms are more closely related to longer nocturnal wake times than to shorter sleep times. Results in both samples indicated that poor sleepers reported more daytime difficulties than good sleepers. While low- and high-distress poor sleepers did not differ on sleep parameters, highly distressed poor sleepers reported consistently more difficulty in functioning during the day and experienced greater tension and depression than minimally distressed poor sleepers. Severity of all three daytime problems was generally significantly and positively related to poor psychological adjustment, psychologically laden sleep variables, and, with the exception of sleepiness, to quantitative sleep parameters. Results are used to discuss discrepancies between experiential and quantitative measures of daytime functioning.

  18. Poor actigraphic and self-reported sleep patterns predict delinquency and daytime impairment among at-risk adolescents.

    PubMed

    Stone, Kristen C; Cuellar, Crystal R; Miller-Loncar, Cynthia L; LaGasse, Linda L; Lester, Barry M

    2015-09-01

    To evaluate associations between actigraphic sleep patterns, subjective sleep quality, and daytime functioning (ie, sleepiness, symptoms of depression, and delinquency and other conduct problems) in at-risk adolescents. Prospective, observational cohort study. Providence, RI, predominantly home and school and 2 visits to the Brown Center for the Study of Children at Risk. A diverse group of low-income 13-year-olds (n = 49) with and without prenatal drug exposure. None. Actigraphy, sleep diaries, and sleep and health questionnaires. Above and beyond the effects of prenatal drug exposure and postnatal adversity, actigraphic daytime sleep was a significant predictor of daytime sleepiness and delinquency. Subjective sleep quality was a significant predictor of daytime sleepiness, delinquency, and depressive symptoms. Later bed times predicted increased delinquency. There was a unique effect of actigraphic daytime sleep duration, subjective nighttime sleep quality, and bedtime on daytime functioning (ie, sleepiness, symptoms of depression, and delinquency and other conduct problems) of at-risk adolescents. In these vulnerable youth, these problematic sleep patterns may contribute to feeling and behaving poorly. Intervention studies with at-risk teens should be conducted to further explore the role of these sleep parameters on daytime functioning. Copyright © 2015 National Sleep Foundation. Published by Elsevier Inc. All rights reserved.

  19. [Relationship between serum substance P levels and excessive daytime sleepiness in patients with obstructive sleep apnea-hypopnea syndrome].

    PubMed

    Zhuang, Xi-bin; Huang, Hong-bo; Chen, Wei-wen; Huang, Hong; Guo, Wei-feng

    2012-08-01

    To investigate the relationship between serum Substance P levels and excessive daytime sleepiness (EDS) in patients with obstructive sleep apnea-hypopnea syndrome (OSAHS). A total of 120 adult habitual snorers treated by respiratory physicians in First Hospital of Quanzhou Affiliated to Fujian Medical University were selected for this study. The patients were grouped as simple snorers and OSAHS by the results of polysomnography test. Thirty patients were in the simple snorer group, among whom 24 were male and 6 were female. Their average age was (48 ± 15) years and average AHI was (2.8 ± 1.6) events/hour. Ninety patients were in the OSAHS group, among whom 78 were male and 12 were female. Their average age was (49 ± 12) years and average AHI was (37.1 ± 23.7) events/hour. EDS was assessed using the Epworth sleepiness scale (ESS). Substance P levels were analyzed with a radioimmunoassay. There was no significant difference in gender, age, and body mass index between the 2 groups. The ESS score for patients with OSAHS was (13 ± 5), higher than that for patients in the simple snorer group (F = 10.299, P < 0.05). With increasing severity of OSAHS, the score increased. The serum Substance P level for OSAHS group was (132 ± 27) ng/L, which was lower than that in the control (F = 3.048, P = 0.031), and decrease in Substance P level was most significant in patients with severe OSAHS. Pearson correlation analysis showed that Substance P levels in OSAHS patients were negatively correlated with ESS scores (r = -0.238, P < 0.05). Substance P levels were lower in OSAHS patients with higher degree of daytime sleepiness. Daytime sleepiness and Substance P level were interrelated in patients with OSAHS.

  20. Excessive daytime sleepiness does not correlate with sympathetic nervous system activation and arterial stiffening in patients with mild-to-moderate obstructive sleep apnoea: A proof-of-principle study.

    PubMed

    Bisogni, Valeria; Pengo, Martino F; Drakatos, Panagis; Maiolino, Giuseppe; Kent, Brian; Rossitto, Giacomo; Steier, Joerg; Rossi, Gian Paolo

    2017-06-01

    Increased arterial stiffness and sympathetic nervous system activity, independent markers of cardiovascular risk, are common in patients with severe obstructive sleep apnoea, who have excessive daytime sleepiness. Among patients with mild-to-moderate obstructive sleep apnoea, however, it remains unknown whether arterial stiffness and/or increased sympathetic nervous system activity correlate with excessive daytime sleepiness. We measured heart rate variability, as an index of autonomic nervous system activity, and arterial stiffness index, as a marker of vascular damage and cardiovascular risk, in 56 men aged 18 to 75years, with mild-to-moderate obstructive sleep apnoea, and matched into two groups, "sleepy" (Epworth Sleepiness Scale≥10) and "non-sleepy" (Epworth Sleepiness Scale<10). We found no association of excessive daytime sleepiness with sympathetic nervous system activation (very low frequency power 18,947±11,207ms 2 vs 15,893±8,272ms 2 , p=0.28; low frequency (LH) power 17,753±8,441ms 2 vs 15,414±5,666ms 2 , p=0.26; high frequency (HF) power 7,527±1,979ms 2 vs 8,257±3,416ms 2 , p=0.36; LF/HF ratio 3.04±1.37 vs 2.55±1.01, p=0.15) and mean arterial stiffness index (6.97±0.83 vs 7.26±0.66, p=0.18) in mild-to-moderate obstructive sleep apnoea patients. Symptoms of excessive daytime sleepiness are not associated with sympathetic nervous system activation and arterial stiffness in male subjects with mild-to-moderate obstructive sleep apnoea. Copyright © 2017 Elsevier B.V. All rights reserved.

  1. Comparative levels of excessive daytime sleepiness in common medical disorders.

    PubMed

    Stroe, Alice F; Roth, Thomas; Jefferson, Catherine; Hudgel, David W; Roehrs, Timothy; Moss, Kenneth; Drake, Christopher L

    2010-10-01

    Sleep restriction and sleep disorders are common causes of excessive daytime sleepiness (EDS). Medical disorders (MD) can also cause EDS, but previous studies have used non-standardized measures, selected samples, or have examined EDS in singular disorders. This study describes the relative degree of EDS associated with medical disorders to provide comparative data across a range of common medical conditions in a large unselected community-based sample. Responses of 2612 individuals (aged 18-65) were assessed after excluding those with suspected sleep disordered breathing, narcolepsy, and shift workers. Participants across a range of medical disorders were evaluated using the Epworth Sleepiness Scale (ESS) and patient reports of nocturnal sleep. Sixty-seven percent of the sample reported a MD. The prevalence of EDS (ESS>or=10) was 31.4% in individuals with MD and increased as a function of a number of MD (0 MD=29.4%, 1 MD=28.4%, 2 MD=31.0%, 3 MD=35.3%, 4 MD=38.4%). Disorders which were independent predictors of EDS were ulcers OR=2.21 (95% CI=1.35-3.61), migraines OR=1.36 (95% CI=1.08-1.72), and depression OR=1.46 (95% CI=1.16-1.83) after controlling for other conditions, age, gender, time in bed, caffeine, smoking and alcohol use. Participants with ulcers had the highest prevalence of sleepiness, 50.0%, as well as the highest level of problems falling asleep (40.8%) and awakenings during the night (62.5%). Individuals with ulcers, migraines, and depression have independent and clinically significant levels of EDS relative to other common MD.

  2. The impact of daytime sleepiness on the school performance of college students with attention deficit hyperactivity disorder (ADHD): a prospective longitudinal study.

    PubMed

    Langberg, Joshua M; Dvorsky, Melissa R; Becker, Stephen P; Molitor, Stephen J

    2014-06-01

    This prospective longitudinal study evaluated the impact of daytime sleepiness on the school performance of 62 college students diagnosed comprehensively with attention deficit hyperactivity disorder. The primary goal of the study was to determine if self-reported daytime sleepiness rated at the beginning of the academic year could predict academic and overall functioning at the end of the academic year while also considering potentially important covariates, including symptoms of inattention, hyperactivity and impulsivity, medication status and whether or not students lived at home or on-campus. Self-reported daytime sleepiness predicted longitudinally school maladjustment, overall functional impairment and the number of D and F grades (i.e. poor and failing) students received in courses above and beyond both self- and parent-report of symptoms, but did not predict overall grade point average. Living at home served as a protective factor and was associated with less school maladjustment and overall impairment. Gender was the only significant predictor in the overall grade point average model, with female gender associated with higher overall grades. The implications of these findings for monitoring and treatment of sleep disturbances in college students with attention deficit hyperactivity disorder are discussed. © 2013 European Sleep Research Society.

  3. A survey study of the association between mobile phone use and daytime sleepiness in California high school students.

    PubMed

    Nathan, Nila; Zeitzer, Jamie

    2013-09-12

    Mobile phone use is near ubiquitous in teenagers. Paralleling the rise in mobile phone use is an equally rapid decline in the amount of time teenagers are spending asleep at night. Prior research indicates that there might be a relationship between daytime sleepiness and nocturnal mobile phone use in teenagers in a variety of countries. As such, the aim of this study was to see if there was an association between mobile phone use, especially at night, and sleepiness in a group of U.S. teenagers. A questionnaire containing an Epworth Sleepiness Scale (ESS) modified for use in teens and questions about qualitative and quantitative use of the mobile phone was completed by students attending Mountain View High School in Mountain View, California (n = 211). Multivariate regression analysis indicated that ESS score was significantly associated with being female, feeling a need to be accessible by mobile phone all of the time, and a past attempt to reduce mobile phone use. The number of daily texts or phone calls was not directly associated with ESS. Those individuals who felt they needed to be accessible and those who had attempted to reduce mobile phone use were also ones who stayed up later to use the mobile phone and were awakened more often at night by the mobile phone. The relationship between daytime sleepiness and mobile phone use was not directly related to the volume of texting but may be related to the temporal pattern of mobile phone use.

  4. Lack of association between MAOA-uVNTR variants and excessive daytime sleepiness.

    PubMed

    Ozen, Filiz; Yegin, Zeynep; Yavlal, Figen; Saglam, Zuhal Aydan; Koc, Haydar; Berber, Ismet

    2017-05-01

    Sleep disorders are highly prevalent in the population and have dramatic health, social, and economic impacts. However, their treatments may remain symptomatic due to ignorance of molecular factors which may provide fundamental insights into the neurological bases of sleep. Excessive daytime sleepiness (EDS) is a common complaint encountered in neurological practice with significant effects both on individuals and on society. We aimed to investigate the role of monoamine oxidase A (MAOA) as a candidate gene in EDS. Epworth sleepiness scale (ESS) was applied to 221 subjects who were also genotyped for MAOA upstream variable number of tandem repeats (MAOA-uVNTR). Patient group displayed higher ESS values (mean 12.67) when compared with the control group (mean 6.38). However, MAOA-uVNTR genotypes did not show a significant association with ESS scores neither on women nor on men. Finally, these data suggest further replications in different populations. Moreover, the investigation of some other genes together with MAOA and/or some possible regulatory molecular mechanisms may offer a more comprehensive approach in the role of genetic factors contributing to EDS.

  5. The Association between Daytime Napping and Cognitive Functioning in Chronic Fatigue Syndrome

    PubMed Central

    Gotts, Zoe M.; Ellis, Jason G.; Deary, Vincent; Barclay, Nicola; Newton, Julia L.

    2015-01-01

    Objectives The precise relationship between sleep and physical and mental functioning in chronic fatigue syndrome (CFS) has not been examined directly, nor has the impact of daytime napping. This study aimed to examine self-reported sleep in patients with CFS and explore whether sleep quality and daytime napping, specific patient characteristics (gender, illness length) and levels of anxiety and depression, predicted daytime fatigue severity, levels of daytime sleepiness and cognitive functioning, all key dimensions of the illness experience. Methods 118 adults meeting the 1994 CDC case criteria for CFS completed a standardised sleep diary over 14 days. Momentary functional assessments of fatigue, sleepiness, cognition and mood were completed by patients as part of usual care. Levels of daytime functioning and disability were quantified using symptom assessment tools, measuring fatigue (Chalder Fatigue Scale), sleepiness (Epworth Sleepiness Scale), cognitive functioning (Trail Making Test, Cognitive Failures Questionnaire), and mood (Hospital Anxiety and Depression Scale). Results Hierarchical Regressions demonstrated that a shorter time since diagnosis, higher depression and longer wake time after sleep onset predicted 23.4% of the variance in fatigue severity (p <.001). Being male, higher depression and more afternoon naps predicted 25.6% of the variance in objective cognitive dysfunction (p <.001). Higher anxiety and depression and morning napping predicted 32.2% of the variance in subjective cognitive dysfunction (p <.001). When patients were classified into groups of mild and moderate sleepiness, those with longer daytime naps, those who mainly napped in the afternoon, and those with higher levels of anxiety, were more likely to be in the moderately sleepy group. Conclusions Napping, particularly in the afternoon is associated with poorer cognitive functioning and more daytime sleepiness in CFS. These findings have clinical implications for symptom management

  6. The association between daytime napping and cognitive functioning in chronic fatigue syndrome.

    PubMed

    Gotts, Zoe M; Ellis, Jason G; Deary, Vincent; Barclay, Nicola; Newton, Julia L

    2015-01-01

    The precise relationship between sleep and physical and mental functioning in chronic fatigue syndrome (CFS) has not been examined directly, nor has the impact of daytime napping. This study aimed to examine self-reported sleep in patients with CFS and explore whether sleep quality and daytime napping, specific patient characteristics (gender, illness length) and levels of anxiety and depression, predicted daytime fatigue severity, levels of daytime sleepiness and cognitive functioning, all key dimensions of the illness experience. 118 adults meeting the 1994 CDC case criteria for CFS completed a standardised sleep diary over 14 days. Momentary functional assessments of fatigue, sleepiness, cognition and mood were completed by patients as part of usual care. Levels of daytime functioning and disability were quantified using symptom assessment tools, measuring fatigue (Chalder Fatigue Scale), sleepiness (Epworth Sleepiness Scale), cognitive functioning (Trail Making Test, Cognitive Failures Questionnaire), and mood (Hospital Anxiety and Depression Scale). Hierarchical Regressions demonstrated that a shorter time since diagnosis, higher depression and longer wake time after sleep onset predicted 23.4% of the variance in fatigue severity (p <.001). Being male, higher depression and more afternoon naps predicted 25.6% of the variance in objective cognitive dysfunction (p <.001). Higher anxiety and depression and morning napping predicted 32.2% of the variance in subjective cognitive dysfunction (p <.001). When patients were classified into groups of mild and moderate sleepiness, those with longer daytime naps, those who mainly napped in the afternoon, and those with higher levels of anxiety, were more likely to be in the moderately sleepy group. Napping, particularly in the afternoon is associated with poorer cognitive functioning and more daytime sleepiness in CFS. These findings have clinical implications for symptom management strategies.

  7. Efficacy and Safety of Adjunctive Modafinil Treatment on Residual Excessive Daytime Sleepiness among Nasal Continuous Positive Airway Pressure-Treated Japanese Patients with Obstructive Sleep Apnea Syndrome: A Double-Blind Placebo-Controlled Study

    PubMed Central

    Inoue, Yuichi; Takasaki, Yuji; Yamashiro, Yoshihiro

    2013-01-01

    Study Objectives: This double-blind study evaluated the efficacy and safety of modafinil for treating excessive daytime sleepiness in Japanese patients with obstructive sleep apnea syndrome (OSAS). Methods: Patients with residual excessive sleepiness (Epworth Sleepiness Scale [ESS] ≥ 11) on optimal nasal continuous positive airway pressure (nCPAP) therapy (apnea-hypopnea index ≤ 10) were randomized to either 200 mg modafinil (n = 52) or placebo (n = 62) once daily for 4 weeks. Outcomes included baseline-week 4 changes in ESS total score, sleep latency on maintenance of wakefulness test (SL-MWT), nocturnal polysomnography, Pittsburgh Sleep Quality Index (PSQI), and safety. Results: All 114 randomized patients completed the study. Mean change in ESS total score (-6.6 vs -2.4, p < 0.001) and SL-MWT (+2.8 vs -0.4 minutes, p = 0.009) were significantly greater with modafinil than with placebo. ESS total score decreased from > 11 to < 11 at the final assessment in 69.2% of modafinil-treated patients and 30.6% of placebo-treated patients (p < 0.001). Corresponding rates at week 1 were 57.7% and 33.9% (p = 0.014). Changes in nocturnal polysomnography, PSQI, and apnea-hypopnea index from baseline to the final assessment were similar in both groups. Adverse drug reactions occurred in 36.5% and 22.6% of patients in the modafinil and placebo groups, respectively (p = 0.146). Conclusions: Once-daily modafinil was effective and well tolerated for managing residual daytime sleepiness in Japanese OSAS patients with residual excessive daytime sleepiness on optimal nCPAP therapy. Clinical Trial Registration: JapicCTI-No.090777 Citation: Inoue Y; Takasaki Y; Yamashiro Y. Efficacy and safety of adjunctive modafinil treatment on residual excessive daytime sleepiness among nasal continuous positive airway pressure-treated Japanese patients with obstructive sleep apnea syndrome: a double-blind placebo-controlled study. J Clin Sleep Med 2013;9(8):751-757. PMID:23946704

  8. Efficacy and safety of adjunctive modafinil treatment on residual excessive daytime sleepiness among nasal continuous positive airway pressure-treated japanese patients with obstructive sleep apnea syndrome: a double-blind placebo-controlled study.

    PubMed

    Inoue, Yuichi; Takasaki, Yuji; Yamashiro, Yoshihiro

    2013-08-15

    This double-blind study evaluated the efficacy and safety of modafinil for treating excessive daytime sleepiness in Japanese patients with obstructive sleep apnea syndrome (OSAS). Patients with residual excessive sleepiness (Epworth Sleepiness Scale [ESS] ≥ 11) on optimal nasal continuous positive airway pressure (nCPAP) therapy (apnea-hypopnea index ≤ 10) were randomized to either 200 mg modafinil (n = 52) or placebo (n = 62) once daily for 4 weeks. Outcomes included baseline-week 4 changes in ESS total score, sleep latency on maintenance of wakefulness test (SL-MWT), nocturnal polysomnography, Pittsburgh Sleep Quality Index (PSQI), and safety. All 114 randomized patients completed the study. Mean change in ESS total score (-6.6 vs -2.4, p < 0.001) and SL-MWT (+2.8 vs -0.4 minutes, p = 0.009) were significantly greater with modafinil than with placebo. ESS total score decreased from > 11 to < 11 at the final assessment in 69.2% of modafinil-treated patients and 30.6% of placebo-treated patients (p < 0.001). Corresponding rates at week 1 were 57.7% and 33.9% (p = 0.014). Changes in nocturnal polysomnography, PSQI, and apnea-hypopnea index from baseline to the final assessment were similar in both groups. Adverse drug reactions occurred in 36.5% and 22.6% of patients in the modafinil and placebo groups, respectively (p = 0.146). Once-daily modafinil was effective and well tolerated for managing residual daytime sleepiness in Japanese OSAS patients with residual excessive daytime sleepiness on optimal nCPAP therapy.

  9. Longitudinal assessment of excessive daytime sleepiness in early Parkinson's disease.

    PubMed

    Amara, Amy W; Chahine, Lama M; Caspell-Garcia, Chelsea; Long, Jeffrey D; Coffey, Christopher; Högl, Birgit; Videnovic, Aleksandar; Iranzo, Alex; Mayer, Geert; Foldvary-Schaefer, Nancy; Postuma, Ron; Oertel, Wolfgang; Lasch, Shirley; Marek, Ken; Simuni, Tanya

    2017-08-01

    Excessive daytime sleepiness (EDS) is common and disabling in Parkinson's disease (PD). Predictors of EDS are unclear, and data on biological correlates of EDS in PD are limited. We investigated clinical, imaging and biological variables associated with longitudinal changes in sleepiness in early PD. The Parkinson's Progression Markers Initiative is a prospective cohort study evaluating progression markers in participants with PD who are unmedicated at baseline (n=423) and healthy controls (HC; n=196). EDS was measured with the Epworth Sleepiness Scale (ESS). Clinical, biological and imaging variables were assessed for associations with EDS for up to 3 years. A machine learning approach (random survival forests) was used to investigate baseline predictors of incident EDS. ESS increased in PD from baseline to year 3 (mean±SD 5.8±3.5 to 7.55±4.6, p<0.0001), with no change in HC. Longitudinally, EDS in PD was associated with non-tremor dominant phenotype, autonomic dysfunction, depression, anxiety and probable behaviour disorder, but not cognitive dysfunction or motor severity. Dopaminergic therapy was associated with EDS at years 2 and 3, as dose increased. EDS was also associated with presynaptic dopaminergic dysfunction, whereas biofluid markers at year 1 showed no significant associations with EDS. A predictive index for EDS was generated, which included seven baseline characteristics, including non-motor symptoms and cerebrospinal fluid phosphorylated-tau/total-tau ratio. In early PD, EDS increases significantly over time and is associated with several clinical variables. The influence of dopaminergic therapy on EDS is dose dependent. Further longitudinal analyses will better characterise associations with imaging and biomarkers. © Article author(s) (or their employer(s) unless otherwise stated in the text of the article) 2017. All rights reserved. No commercial use is permitted unless otherwise expressly granted.

  10. Prevalence, Patterns, and Predictors of Sleep Problems and Daytime Sleepiness in Young Adolescents with Attention-Deficit/Hyperactivity Disorder

    ERIC Educational Resources Information Center

    Langberg, Joshua M.; Molitor, Stephen J.; Oddo, Lauren E.; Eadeh, Hana-May; Dvorsky, Melissa R.; Becker, Stephen P.

    2017-01-01

    Objective: The primary objective of this study was to evaluate the prevalence of multiple types of sleep problems in young adolescents with ADHD. Method: 262 adolescents comprehensively diagnosed with ADHD and their caregivers completed well-validated measures of sleep problems and daytime sleepiness. Participants also completed measures related…

  11. A survey study of the association between mobile phone use and daytime sleepiness in California high school students

    PubMed Central

    2013-01-01

    Background Mobile phone use is near ubiquitous in teenagers. Paralleling the rise in mobile phone use is an equally rapid decline in the amount of time teenagers are spending asleep at night. Prior research indicates that there might be a relationship between daytime sleepiness and nocturnal mobile phone use in teenagers in a variety of countries. As such, the aim of this study was to see if there was an association between mobile phone use, especially at night, and sleepiness in a group of U.S. teenagers. Methods A questionnaire containing an Epworth Sleepiness Scale (ESS) modified for use in teens and questions about qualitative and quantitative use of the mobile phone was completed by students attending Mountain View High School in Mountain View, California (n = 211). Results Multivariate regression analysis indicated that ESS score was significantly associated with being female, feeling a need to be accessible by mobile phone all of the time, and a past attempt to reduce mobile phone use. The number of daily texts or phone calls was not directly associated with ESS. Those individuals who felt they needed to be accessible and those who had attempted to reduce mobile phone use were also ones who stayed up later to use the mobile phone and were awakened more often at night by the mobile phone. Conclusions The relationship between daytime sleepiness and mobile phone use was not directly related to the volume of texting but may be related to the temporal pattern of mobile phone use. PMID:24028604

  12. Insomnia Subtypes and Their Relationship to Excessive Daytime Sleepiness in Brazilian Community-Dwelling Older Adults

    PubMed Central

    Hara, Cláudia; Stewart, Robert; Lima-Costa, Maria Fernanda; Rocha, Fábio Lopes; Fuzikawa, Cíntia; Uchoa, Elizabeth; Firmo, Josélia O.A.; Castro-Costa, Érico

    2011-01-01

    Study Objectives: To investigate the association between different types of insomnia as exposures and excessive daytime sleepiness (EDS) as a binary outcome in older Brazilian residents. Design: The baseline examination of the Bambuí Health and Ageing Study (BHAS), which is an ongoing population-based prospective cohort study of older adults. Setting: Bambuí (15,000 inhabitants), a city in the State of Minas Gerais, Southeast Brazil Participants: All residents aged ≥ 60 years were eligible to take part in the BHAS baseline. Of 1742 residents identified who were ≥ 60 years, 1606 (92.2%) were interviewed and received comprehensive examinations of health status. Interventions: None Measurements and Results: EDS was defined as the presence of sleepiness ≥ 3 times per week in the last month, causing any interference in usual activities. All insomnia subtypes were significantly associated with EDS in unadjusted analyses, and these associations were only modestly altered after adjusting incrementally for the other covariates. In a final model, the 3 insomnia subtypes were entered into a fully adjusted model simultaneously to investigate mutual independence, giving prevalence ratios of 1.63 (95% CI 1.14-2.31) for initial insomnia, 2.13 (95% CI 1.48-3.07) for middle insomnia, and 1.36 (95% CI 0.94-1.96) for terminal insomnia. The population attributable fractions for initial, middle, and terminal insomnia on prevalence of EDS were 17.6%, 32.9%, and 9.7%, respectively. Conclusions: Middle insomnia emerged as the insomnia subtype most strongly associated with EDS. Further research is required to clarify causal pathways underlying this cross-sectional association. Citation: Hara C; Stewart R; Lima-Costa MF; Rocha FL; Fuzikawa C; Uchoa E; Firmo JOA; Castro-Costa E. Insomnia subtypes and their relationship to excessive daytime sleepiness in Brazilian community-dwelling older adults. SLEEP 2011;34(8):1111-1117. PMID:21804673

  13. Adolescent Sleepiness: Causes and Consequences.

    PubMed

    Hansen, Shana L; Capener, Dale; Daly, Christopher

    2017-09-01

    Insufficient sleep duration and poor sleep quality are common among adolescents. The multidimensional causes of insufficient sleep duration and poor sleep quality include biological, health-related, environmental, and lifestyle factors. The most common direct consequence of insufficient and/or poor sleep quality is excessive daytime sleepiness, which may contribute to poor academic performance, behavioral health problems, substance use, and drowsy driving. Evaluation of sleepiness includes a detailed sleep history and sleep diary, with polysomnography only required for the assessment of specific sleep disorders. Management involves encouraging healthy sleep practices such as having consistent bed and wake times, limiting caffeine and electronics at night before bed, and eliminating napping, in addition to treating any existing sleep or medical disorders. [Pediatr Ann. 2017;46(9):e340-e344.]. Copyright 2017, SLACK Incorporated.

  14. Obesity and sleepiness in women with fibromyalgia.

    PubMed

    de Araújo, Tânia Aparecida; Mota, Maria Carliana; Crispim, Cibele Aparecida

    2015-02-01

    Fibromyalgia (FM) is associated with a number of comorbidities, including chronic widespread pain, fatigue and non-restorative sleep. Evidence has shown that FM is closely associated with overweight and obesity. The objective of the present study was to investigate the relationship between obesity and sleepiness in women with FM. A total of 100 adult female patients with a prior medical diagnosis of FM participated in the study. Body mass, height and waist circumference were measured, and body mass index (BMI) was calculated. The diet quality was evaluated by the Healthy Eating Index. Subjective analyses of daytime sleepiness [Epworth Sleepiness Scale (ESS)] and sleep quality (Pittsburgh Sleep Quality) were performed. An obesity rate of 41 % was found in all women (56.1 % were sleepy and 43.9 % were not, p = 0.04). Obese women showed a greater level of sleepiness when compared with non-obese (10.2 and 7.0, respectively, p = 0.004). Sleepy women showed a greater weight gain after the diagnosis of FM when compared with non-sleepy women (11.7 and 6.4 kg, respectively, p = 0.04). A positive and significant correlation between BMI and sleepiness (r = 0.35, p = 0.02) was also found. In multivariate logistic regression, moderate or severe sleepiness (ESS >12) was associated with obesity (odds ratio 3.44, 95 % CI 1.31-9.01, p = 0.04). These results demonstrate an important association between sleepiness and FM, suggesting that the occurrence of obesity may be involved with sleepiness in these patients.

  15. 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.

  16. Slow Wave Sleep Enhancement with Gaboxadol Reduces Daytime Sleepiness During Sleep Restriction

    PubMed Central

    Walsh, James K.; Snyder, Ellen; Hall, Janine; Randazzo, Angela C.; Griffin, Kara; Groeger, John; Eisenstein, Rhody; Feren, Stephen D.; Dickey, Pam; Schweitzer, Paula K.

    2008-01-01

    physiological sleep tendency and introspective sleepiness and fatigue which typically result from sleep restriction. Citation: Walsh JK; Snyder E; Hall J; Randazzo AC; Griffin K; Groeger J; Eisenstein R; Feren SD; Dickey P; Schweitzer PK. Slow Wave Sleep Enhancement with Gaboxadol Reduces Daytime Sleepiness During Sleep Restriction. SLEEP 2008;31(5):659–672. PMID:18517036

  17. Double jeopardy: the influence of excessive daytime sleepiness and impaired cognition on health-related quality of life in adults with heart failure

    PubMed Central

    Riegel, Barbara; Ratcliffe, Sarah J.; Weintraub, William S.; Sayers, Steven L.; Goldberg, Lee R.; Potashnik, Sheryl; Weaver, Terri E.; Pressler, Susan J.

    2012-01-01

    Aims To determine how excessive daytime sleepiness (EDS) and impaired cognition contribute to health-related quality of life (HRQL) in heart failure (HF). Methods and results Adults with chronic HF were enrolled into a prospective cohort study. Data were obtained from 280 subjects enrolled from three sites in the northeastern USA; 242 completed the 6-month study. At baseline, cohorts with and without EDS were identified using the Epworth Sleepiness Scale. Each EDS group was further subdivided into those with and without impaired cognition using a battery of five neuropsychological tests. Two disease-specific measures, the Kansas City Cardiomyopathy Questionnaire (KCCQ) and the Functional Outcomes of Sleep Questionnaire (FOSQ), were used to measure HRQL. General linear modelling of square-transformed variables was used to test the hypothesis that cohort membership was a significant predictor of HRQL. At 6 months the remaining sample was 62.5 [standard deviation (SD) 12] years old, mostly male (63%), white (65%), and functionally compromised [72% New York Heart Association (NYHA) class III/IV]. The cohort with both EDS and impaired cognition had the lowest KCCQ overall summary score (60.5 ± 22.5) compared with the cohort without EDS or impaired cognition (74.6 ± 17.4, P ≤ 0.001). A similar effect was seen on the FOSQ (16.0 ± 2.8 vs. 18.5 ± 2.2, P < 0.001). Conclusion Impaired cognition alone did not explain poor HRQL, but the addition of EDS poses a significant risk for poor HRQL. Interventions designed to influence EDS may improve HRQL in this population. PMID:22510422

  18. Construct validity and factor structure of the pittsburgh sleep quality index and epworth sleepiness scale in a multi-national study of African, South East Asian and South American college students.

    PubMed

    Gelaye, Bizu; Lohsoonthorn, Vitool; Lertmeharit, Somrat; Pensuksan, Wipawan C; Sanchez, Sixto E; Lemma, Seblewengel; Berhane, Yemane; Zhu, Xiaotong; Vélez, Juan Carlos; Barbosa, Clarita; Anderade, Asterio; Tadesse, Mahlet G; Williams, Michelle A

    2014-01-01

    The Pittsburgh Sleep Quality Index (PSQI) and the Epworth Sleepiness Scale (ESS) are questionnaires used to assess sleep quality and excessive daytime sleepiness in clinical and population-based studies. The present study aimed to evaluate the construct validity and factor structure of the PSQI and ESS questionnaires among young adults in four countries (Chile, Ethiopia, Peru and Thailand). A cross-sectional study was conducted among 8,481 undergraduate students. Students were invited to complete a self-administered questionnaire that collected information about lifestyle, demographic, and sleep characteristics. In each country, the construct validity and factorial structures of PSQI and ESS questionnaires were tested through exploratory and confirmatory factor analyses (EFA and CFA). The largest component-total correlation coefficient for sleep quality as assessed using PSQI was noted in Chile (r = 0.71) while the smallest component-total correlation coefficient was noted for sleep medication use in Peru (r = 0.28). The largest component-total correlation coefficient for excessive daytime sleepiness as assessed using ESS was found for item 1 (sitting/reading) in Chile (r = 0.65) while the lowest item-total correlation was observed for item 6 (sitting and talking to someone) in Thailand (r = 0.35). Using both EFA and CFA a two-factor model was found for PSQI questionnaire in Chile, Ethiopia and Thailand while a three-factor model was found for Peru. For the ESS questionnaire, we noted two factors for all four countries. Overall, we documented cross-cultural comparability of sleep quality and excessive daytime sleepiness measures using the PSQI and ESS questionnaires among Asian, South American and African young adults. Although both the PSQI and ESS were originally developed as single-factor questionnaires, the results of our EFA and CFA revealed the multi- dimensionality of the scales suggesting limited usefulness of the global PSQI and ESS scores

  19. Associations of impaired sleep quality, insomnia, and sleepiness with epilepsy: A questionnaire-based case-control study.

    PubMed

    Im, Hee-Jin; Park, Seong-Ho; Baek, Shin-Hye; Chu, Min Kyung; Yang, Kwang Ik; Kim, Won-Joo; Yun, Chang-Ho

    2016-04-01

    The purpose of this study was to document the frequency of sleep problems including poor sleep quality, excessive daytime sleepiness, and insomnia in subjects with epilepsy compared with healthy controls and to determine the factors associated with these sleep disturbances. We recruited 180 patients with epilepsy (age: 43.2 ± 15.6 years, men: 50.0%) and 2836 healthy subjects (age: 44.5 ± 15.0 years, men: 49.8%). Sleep and the anxiety/mood profiles were measured using the Epworth Sleepiness Scale, Pittsburgh Sleep Quality Index, Insomnia Severity Index, Goldberg Anxiety Scale, and Patient Health Questionnaire-9 depression scale. Associations of sleep problems with epilepsy and other factors were tested by multiple logistic regression analysis, adjusted for age, gender, body mass index, alcohol intake, smoking, perceived sleep insufficiency, and habitual snoring. Sleep disturbances were more common in the group with epilepsy than in the controls (53.3% vs. 25.5%; p<0.001). Poor sleep quality, excessive daytime sleepiness, and insomnia were significantly associated with epilepsy (odds ratio [95% confidence interval]: 3.52 [2.45-5.05], 2.10 [1.41-3.12], 5.91 [3.43-10.16], respectively). Depressive mood, anxiety, and perceived sleep insufficiency contributed to the presence of sleep disturbances. In the group with epilepsy, seizure remission for the past year related to a lower frequency of insomnia, whereas age, sex, type of epilepsy, and number of antiepileptic drugs were not correlated with sleep problems. Epilepsy was significantly associated with the higher frequency of sleep disturbances, which supports the importance of screening sleep problems in patients with epilepsy and providing available intervention. Copyright © 2016 Elsevier Inc. All rights reserved.

  20. Gender differences in excessive daytime sleepiness among Japanese workers.

    PubMed

    Doi, Yuriko; Minowa, Masumi

    2003-02-01

    Excessive daytime sleepiness (EDS) is serious concern in the workplace with respect to errors, accidents, absenteeism, reduced productivity and impaired personal or professional life. Previous community studies found a female preponderance of EDS, however, there is little research on EDS and gender in occupational settings. We examined the gender differences in prevalence and risk factors of EDS among employees working at a telecommunications company in the Tokyo metropolitan area. Our outcome measure of EDS was the Epworth Sleepiness Scale (ESS). A self-administered questionnaire on health and sleep including ESS was distributed to 5,571 workers between December 1999 and January 2000, and 5,072 responses were returned (91.0%). A total of 4,722 full-time, non-manual and non-shift employees aged 20-59 were used for analysis (3,909 men and 813 women). Chi-squared tests and multiple logistic regression analyses were applied for examining the gender differences in the prevalence and risk factors of EDS. The prevalence rates of EDS were 13.3% for women and 7.2% for men (P<0.001). We identified that deprived nocturnal sleep, an irregular sleep-wake schedule and depression were the risk factors of EDS for both genders, and being married worked as a protective factor against EDS for men alone. It is obvious that a ban on overtime work and a provision of mental health hygiene are the general strategies for reducing EDS at worksites. In the case of women, we suggest the formation of effective strategies for improving women's status at home and in the workplace must also be a solution for the prevention of EDS (e.g. promoting gender equality in the division of labor at home and strengthening family care policies for working women).

  1. Myotonic dystrophytype 1 - report of non-24-h sleep-wake disorder with excessive daytime sleepiness.

    PubMed

    Filho, Lucio Huebra Pimentel; Gomes, Ana Carolina Dias; Gonçalves, Bruno; Tufik, Sergio; Coelho, Fernando Morgadinho

    2018-05-15

    Myotonic dystrophy (MD) is a neuromuscular disease with myotonia, progressive weakness, and involvement of CNS, heart, and gastrointestinal system. Excessive daytime sleepiness (EDS) in myotonic dystrophy type 1 (MD1) is related to sleep breathing diseases, restless leg syndrome, periodic limb movements during sleep and narcoleptic-like phenotype. However, authors highlight a central dysfunction of sleep regulation. We describe a 26-year-old, female, MD1 patient with EDS. Sleep diary/actigraphy evidenced two different circadian periods with values of 1442 and 1522 min. Agomelatine, 50 mg at night, was prescribed with improvement of the circadian rhythm and complaints of sleepiness. The identification of unanticipated causes of EDS, such as circadian rhythm disorders permits an appropriated treatment. As we know, it is the first relate of non-24-h sleep-wake disorder in patient with MD1. Sleep diary and actigraphy could be good options to investigate sleep-wake cycle disorder in patients with MD and EDS.

  2. Effects of recovery sleep after one work week of mild sleep restriction on interleukin-6 and cortisol secretion and daytime sleepiness and performance.

    PubMed

    Pejovic, Slobodanka; Basta, Maria; Vgontzas, Alexandros N; Kritikou, Ilia; Shaffer, Michele L; Tsaoussoglou, Marina; Stiffler, David; Stefanakis, Zacharias; Bixler, Edward O; Chrousos, George P

    2013-10-01

    One workweek of mild sleep restriction adversely impacts sleepiness, performance, and proinflammatory cytokines. Many individuals try to overcome these adverse effects by extending their sleep on weekends. To assess whether extended recovery sleep reverses the effects of mild sleep restriction on sleepiness/alertness, inflammation, and stress hormones, 30 healthy young men and women (mean age ± SD, 24.7 ± 3.5 yr; mean body mass index ± SD, 23.6 ± 2.4 kg/m(2)) participated in a sleep laboratory experiment of 13 nights [4 baseline nights (8 h/night), followed by 6 sleep restriction nights (6 h/night) and 3 recovery nights (10 h/night)]. Twenty-four-hour profiles of circulating IL-6 and cortisol, objective and subjective daytime sleepiness (Multiple Sleep Latency Test and Stanford Sleepiness Scale), and performance (Psychomotor Vigilance Task) were assessed on days 4 (baseline), 10 (after 1 wk of sleep restriction), and 13 (after 2 nights of recovery sleep). Serial 24-h IL-6 plasma levels increased significantly during sleep restriction and returned to baseline after recovery sleep. Serial 24-h cortisol levels during restriction did not change compared with baseline, but after recovery they were significantly lower. Subjective and objective sleepiness increased significantly after restriction and returned to baseline after recovery. In contrast, performance deteriorated significantly after restriction and did not improve after recovery. Extended recovery sleep over the weekend reverses the impact of one work week of mild sleep restriction on daytime sleepiness, fatigue, and IL-6 levels, reduces cortisol levels, but does not correct performance deficits. The long-term effects of a repeated sleep restriction/sleep recovery weekly cycle in humans remain unknown.

  3. Effects of recovery sleep after one work week of mild sleep restriction on interleukin-6 and cortisol secretion and daytime sleepiness and performance

    PubMed Central

    Pejovic, Slobodanka; Basta, Maria; Kritikou, Ilia; Shaffer, Michele L.; Tsaoussoglou, Marina; Stiffler, David; Stefanakis, Zacharias; Bixler, Edward O.; Chrousos, George P.

    2013-01-01

    One workweek of mild sleep restriction adversely impacts sleepiness, performance, and proinflammatory cytokines. Many individuals try to overcome these adverse effects by extending their sleep on weekends. To assess whether extended recovery sleep reverses the effects of mild sleep restriction on sleepiness/alertness, inflammation, and stress hormones, 30 healthy young men and women (mean age ± SD, 24.7 ± 3.5 yr; mean body mass index ± SD, 23.6 ± 2.4 kg/m2) participated in a sleep laboratory experiment of 13 nights [4 baseline nights (8 h/night), followed by 6 sleep restriction nights (6 h/night) and 3 recovery nights (10 h/night)]. Twenty-four-hour profiles of circulating IL-6 and cortisol, objective and subjective daytime sleepiness (Multiple Sleep Latency Test and Stanford Sleepiness Scale), and performance (Psychomotor Vigilance Task) were assessed on days 4 (baseline), 10 (after 1 wk of sleep restriction), and 13 (after 2 nights of recovery sleep). Serial 24-h IL-6 plasma levels increased significantly during sleep restriction and returned to baseline after recovery sleep. Serial 24-h cortisol levels during restriction did not change compared with baseline, but after recovery they were significantly lower. Subjective and objective sleepiness increased significantly after restriction and returned to baseline after recovery. In contrast, performance deteriorated significantly after restriction and did not improve after recovery. Extended recovery sleep over the weekend reverses the impact of one work week of mild sleep restriction on daytime sleepiness, fatigue, and IL-6 levels, reduces cortisol levels, but does not correct performance deficits. The long-term effects of a repeated sleep restriction/sleep recovery weekly cycle in humans remain unknown. PMID:23941878

  4. The sleepy adolescent: causes and consequences of sleepiness in teens.

    PubMed

    Moore, Melisa; Meltzer, Lisa J

    2008-06-01

    The majority of adolescents do not obtain the recommended amount of sleep, resulting in significant daytime sleepiness. For most adolescents, insufficient sleep results from the interaction between intrinsic factors such as puberty and extrinsic factors such as school start times. Insufficient sleep and sleepiness impact all areas of adolescent functioning, including academic, psychological and behavioural, which underscores the importance of evaluating sleepy adolescents. While polysomnography is required for the diagnosis of certain sleep disorders, causes of sleepiness are generally best identified with a detailed sleep history and daily sleep diary. The management of sleep problems in adolescents involves treating any underlying sleep disorders, increasing total sleep time and improving other environmental factors that impact sleep. Recognition and management of insufficient sleep and sleepiness is important for the health and functioning of adolescents.

  5. Sleepiness and activity in heart failure patients with reduced ejection fraction and central sleep-disordered breathing.

    PubMed

    Atalla, Angela; Carlisle, Thomas W; Simonds, Anita K; Cowie, Martin R; Morrell, Mary J

    2017-06-01

    Patients with heart failure (HF) and sleep disordered breathing (SDB) are typically not sleepy, unlike patients without heart failure. Previous work in HF patients with obstructive SDB suggested that sleepiness was associated with a reduction in daytime activity. The consequences of predominately central SDB on sleepiness in HF are less well understood. The aim of this study was to test the hypothesis that subjective sleepiness is associated with reduced daytime activity in HF patients with central SDB, compared to those without SDB. The Epworth Sleepiness Scale (ESS), nocturnal polysomnography, and 14 days of wrist watch actigraphy were used to assess subjective daytime sleepiness, nocturnal sleep and breathing, and 24-h activity levels, respectively. A total of 54 patients with HF were studied, nine had obstructive SDB and were removed from further analysis. Of the patients, 23 had HF with predominantly central SDB (HF-CSA; apnea-hypopnea index (AHI) median 20.6 (IQR 12.9-40.2)/h), and 22 had noSDB (HF-noSDB; AHI 3.7 (2.5-5.9)/h). The median patient age was 68 years (range 59-73 years). There were no significant differences either in ESS score (HF-CSA; 8 [4-10] vs. HF-noSDB; 8 (6-12); p = 0.49) or in duration of daytime activity (HF-CSA 14.5 (14.1-15.2) and HF-noSDB 15.1 (14.4-15.3) hours; p = 0.10) between the groups. HF patients with predominately central SDB are not subjectively sleepy compared to those without SDB, despite reduced sleep quality. We speculate that the lack of sleepiness (based on ESS score) may be due to increased sympathetic nerve activity, although further studies are needed due to the small number (n = 5) of sleepy HF-CSA patients. Daytime activity was not different between HF-noSDB and HF-CSA patients. Copyright © 2017. Published by Elsevier B.V.

  6. [Dietary factors associated with daytime somnolence in healthy elderly of Chile].

    PubMed

    Durán Agüero, Samuel; Sánchez Reyes, Hugo; Díaz Narváez, Víctor; Araya Pérez, Mónica

    2015-01-01

    To determine the prevalence of mild and excessive somnolence and the associated factors with the presence of daytime sleepiness in the elderly. A total of 1780 independent individuals 60 years and older of both sexes (70.9±7.9 years old; females 62%), were included, of which 1704 of them completed all the information. All of them were assessed using an Epworth sleepiness scale (ESE), an Pittsburgh sleep quality index, plus information of cigarettes smoking, dinner time, and an anthropometric evaluation. An ESE score>10 was considered drowsiness and scores>15 excessive or severe drowsiness. Among the population under 80 years, 5.3% showed ESE score>15 and 26.2% an ESE score>10. For over 80 years, the prevalence of sleepiness was 6.3% for an ESE score>15 and 32.5% for an ESE score>10. In the adjusted model, the factors associated with increased risk of sleepiness (ESE>10) were age older than 80 years (OR=1.58; 95% CI=1.14 to 2.19) and dinner after 21 hours (OR=1.3; 95% CI=1.01 to 1.68). By contrast, only age older than 80 years was independently associated with severe sleepiness (OR=1.81; 95% CI=1.01 to 3.29). Meals after 21 hours and age above 80 years are associated with increased likelihood of daytime sleepiness. Instead, only older than 80 years is associated with severe daytime sleepiness. Copyright © 2015 SEGG. Published by Elsevier Espana. All rights reserved.

  7. Scatter Plot Analysis of Excessive Daytime Sleepiness and Severe Disruptive Behavior in Adults with Prader-Willi Syndrome: A Pilot Study

    ERIC Educational Resources Information Center

    Maas, Anneke P. H. M.; Didden, Robert; Bouts, Lex; Smits, Marcel G.; Curfs, Leopold M. G.

    2009-01-01

    Individuals with Prader-Willi syndrome (PWS) are at risk for excessive daytime sleepiness (EDS) and disruptive behavior. This pilot study explores temporal characteristics of EDS and severe disruptive behavior across time of day and day of week in seven individuals with PWS (aged between 33 and 49 years) of whom five were matched to controls.…

  8. Sleepiness in sleepwalking and sleep terrors: a higher sleep pressure?

    PubMed

    Carrillo-Solano, Marisol; Leu-Semenescu, Smaranda; Golmard, Jean-Louis; Groos, Elisabeth; Arnulf, Isabelle

    2016-10-01

    To identify the determinants of excessive daytime sleepiness in adults with sleepwalking or sleep terrors (SW/ST). We collected the charts of all consecutive adult patients admitted from 2012 to 2014 for SW/ST. They had completed the Paris Arousal Disorders Severity Scale and the Epworth Sleepiness Scale, and had undergone one (n = 34) or two consecutive (n = 124) nocturnal videopolysomnographies. The demographic, clinical, and sleep determinants of excessive daytime sleepiness (defined as an Epworth Sleepiness Scale score of greater than 10) were analyzed. Almost half (46.8%) of the 158 adult patients with SW/ST reported excessive daytime sleepiness. They had shorter sleep onset latencies (in night 1 and night 2), shorter REM sleep latencies, longer total sleep time, and higher REM sleep percentages in night 2, but no greater clinical severity of the parasomnia than patients without sleepiness. The level of sleepiness correlated with the same measures (sleep onset latency on both nights, REM sleep onset latency, and total sleep time in night 2), plus the latency to N3. In the regression model, higher sleepiness was determined by shorter sleep onset latency on night 1, lower number of awakenings in N3 on night 1, and higher total sleep time on night 2. Daytime sleepiness in patients with SW/ST is not the consequence of disturbed sleep but is associated with a specific polygraphic phenotype (rapid sleep onset, long sleep time, lower numbers of awakenings on N3) that is suggestive of a higher sleep pressure that may contribute to incomplete arousal from N3. Copyright © 2015 Elsevier B.V. All rights reserved.

  9. Comprehensive assessment of the impact of life habits on sleep disturbance, chronotype, and daytime sleepiness among high-school students.

    PubMed

    Shimura, Akiyoshi; Hideo, Sakai; Takaesu, Yoshikazu; Nomura, Ryota; Komada, Yoko; Inoue, Takeshi

    2018-04-01

    Sleep affects adolescents in various ways. However, the effects of multiple factors on sleep hygiene remain unclear. A comprehensive assessment of the effects of life habits on sleep in high-school students was conducted. A cross-sectional survey of 344 high school students (age range 15-17; 171 boys, 173 girls) in Tokyo, Japan was conducted in 2015. Complete responses were provided by 294 students. Demographic variables, Pittsburgh Sleep Quality Index (PSQI), diurnal type scale, Pediatric Daytime Sleepiness Scale (PDSS), and life habits such as dinnertime, viewing electronic displays, caffeine intake, sunlight in the morning, and the brightness of the room in the night were asked. The mean scores were PSQI: 5.9 (±2.3), PDSS: 19.0 (±5.8), and the diurnal type scale: 16.7 (±3.4). Using an electronic display in bed (OR = 3.01; (95%CI) 1.24-7.30), caffeine intake at night always (OR = 2.22; 1.01-4.90), and waking up before dawn (OR = 3.25; 1.34-7.88) were significantly associated with sleep disturbance. Irregular timing of the evening meal (OR = 2.06; 1.10-3.84) and display viewing within 2 h before bedtime (OR = 2.50; 1.01-6.18) or in bed (OR = 3.60; 1.41-9.21) were significantly associated with excessive daytime sleepiness. Using an electronic display within 2 h before bedtime (OR = 2.64; 1.10-6.38) or in bed (OR = 3.50; 1.40-8.76) and a living room which is bright at night (OR = 1.89; 1.06-3.36) were significantly associated with eveningness. Each type of sleep-related problem had its own associated life habit factors. Copyright © 2017 The Author(s). Published by Elsevier B.V. All rights reserved.

  10. Sleep Disturbance, Daytime Symptoms, and Functional Performance in Patients With Stable Heart Failure: A Mediation Analysis.

    PubMed

    Jeon, Sangchoon; Redeker, Nancy S

    2016-01-01

    Sleep disturbance is common among patients with heart failure (HF) who also experience symptom burden and poor functional performance. We evaluated the extent to which sleep-related, daytime symptoms (fatigue, excessive daytime sleepiness, and depressive symptoms) mediate the relationship between sleep disturbance and functional performance among patients with stable HF. We recruited patients with stable HF for this secondary analysis of data from a cross-sectional, observational study. Participants completed unattended ambulatory polysomnography from which the Respiratory Disturbance Index was calculated, along with a Six-Minute Walk Test, questionnaires to elicit sleep disturbance (Pittsburgh Sleep Quality Index, Insomnia Symptoms from the Sleep Habits Questionnaire), daytime symptoms (Center for Epidemiologic Studies Depression Scale, Global Fatigue Index, Epworth Sleepiness Scale), and self-reported functional performance (Medical Outcomes Study SF36 V2 Physical Function Scale). We used structural equation modeling with latent variables for the key analysis. Follow-up, exploratory regression analysis with bootstrapped samples was used to examine the extent to which individual daytime symptoms mediated effects of sleep disturbance on functional performance after controlling for clinical and demographic covariates. The sample included 173 New York Heart Association Class I-IV HF patients (n = 60/34.7% women; M = 60.7, SD = 16.07 years of age). Daytime symptoms mediated the relationship between sleep disturbance and functional performance. Fatigue and depression mediated the relationship between insomnia symptoms and self-reported functional performance, whereas fatigue and sleepiness mediated the relationship between sleep quality and functional performance. Sleepiness mediated the relationship between the respiratory index and self-reported functional performance only in people who did not report insomnia. Daytime symptoms explain the relationships between sleep

  11. Impact of shift work on sleep and daytime performance among health care professionals.

    PubMed

    Alshahrani, Sultan M; Baqays, Abdulsalam A; Alenazi, Abdelelah A; AlAngari, Abdulaziz M; AlHadi, Ahmad N

    2017-08-01

    To evaluate sleep quality and daytime sleepiness in health care professionals who are performing shift work. Methods: This cross-sectional study was conducted on 510 health care professionals at Prince Sultan Military Medical City and King Khalid University Hospital, King Saud University, Riyadh, Kingdom of Saudi Arabia between December 2015 and April 2016. Data were collected using the Pittsburgh Sleep Quality Index (PSQI) and the Epworth Sleepiness Scale (ESS). Participants were divided into 2 groups: shift workers and non-shift workers. Results: We compared both groups regarding the effect of shift work on the total score of PSQI and ESS. We found that the PSQI global score (p less than 0.001) and the total ESS score (p=0.003) were significantly higher in shift work health care professionals.  Conclusion: Shift work among health care professionals is associated with poor sleep quality but not excessive daytime sleepiness. Health care professionals performing shift work have PSQI and ESS scores slightly higher than non-shift work health professionals.

  12. [Circadian rhythm : Influence on Epworth Sleepiness Scale score].

    PubMed

    Herzog, M; Bedorf, A; Rohrmeier, C; Kühnel, T; Herzog, B; Bremert, T; Plontke, S; Plößl, S

    2017-02-01

    The Epworth Sleepiness Scale (ESS) is frequently used to determine daytime sleepiness in patients with sleep-disordered breathing. It is still unclear whether different levels of alertness induced by the circadian rhythm influence ESS score. The aim of this study is to investigate the influence of circadian rhythm-dependent alertness on ESS performance. In a monocentric prospective noninterventional observation study, 97 patients with suspected sleep-disordered breathing were investigated with respect to daytime sleepiness in temporal relationship to polysomnographic examination and treatment. The Karolinska Sleepiness Scale (KSS) and the Stanford Sleepiness Scale (SSS) served as references for the detection of present sleepiness at three different measurement times (morning, noon, evening), prior to and following a diagnostic polysomnography night as well as after a continuous positive airway pressure (CPAP) titration night (9 measurements in total). The KSS, SSS, and ESS were performed at these times in a randomized order. The KSS and SSS scores revealed a circadian rhythm-dependent curve with increased sleepiness at noon and in the evening. Following a diagnostic polysomnography night, the scores were increased compared to the measurements prior to the night. After the CPAP titration night, sleepiness in the morning was reduced. KSS and SSS reflect the changes in alertness induced by the circadian rhythm. The ESS score war neither altered by the intra-daily nor by the inter-daily changes in the level of alertness. According to the present data, the ESS serves as a reliable instrument to detect the level of daytime sleepiness independently of the circadian rhythm-dependent level of alertness.

  13. Effects of a sleep education program with self-help treatment on sleeping patterns and daytime sleepiness in Japanese adolescents: A cluster randomized trial.

    PubMed

    Tamura, Norihisa; Tanaka, Hideki

    2016-01-01

    Subjective insufficient sleep and delayed sleep-wake patterns have been reported as the primary causes for daytime sleepiness, a reasonably significant and prevalent problem for adolescents worldwide. Systematic reviews have indicated that the success of sleep education programs has thus far been inconsistent, due to the lack of a tailored approach that allows for evaluation of individual differences in behavior patterns. One way to resolve this problem is to assess the individual sleep behaviors of adolescents by using a checklist containing the recommended behaviors for promoting sleep health. Such self-help education programs have already been implemented for elementary school children, school nurses and the elderly. The present study aimed to verify the effects of a sleep education program with supplementary self-help treatment, based on a checklist of sleep-promoting behaviors, in addition to evaluation of changes in sleeping patterns, sleep-promoting behaviors and daytime sleepiness in adolescents. A cluster randomized controlled trial involving 5 Japanese junior high schools was conducted, and 243 students (sleep education: n = 122; waiting list: n = 121; 50.6% female; 7th grade) were included in the final analysis. The sleep education group was provided with information on proper sleep health and sleep-promoting behaviors. The students in this group were asked to practice one sleep-promoting behavior as a goal for 2 weeks and to monitor their practice using sleep diaries. Both pre- and post-treatment questionnaires were administered to students in order to assess knowledge of sleep-promoting behaviors, sleeping patterns and daytime functioning. Students in the sleep education group showed significant improvement in their knowledge of sleep health (F1,121 = 648.05, p < 0.001) and in their sleep-promoting behaviors (F1,121 = 55.66, p < 0.001). Bedtime on both school nights (F1,121 = 50.86, p < 0.001) and weekends (F1,121 = 15.03, p < 0.001), sleep

  14. Sleepiness, driving, and motor vehicle accidents: A questionnaire-based survey.

    PubMed

    Zwahlen, Daniel; Jackowski, Christian; Pfäffli, Matthias

    2016-11-01

    In Switzerland, the prevalence of an excessive daytime sleepiness (EDS) in drivers undergoing a driving capacity assessment is currently not known. In this study, private and professional drivers were evaluated by means of a paper-based questionnaire, including Epworth Sleepiness Scale, Berlin Questionnaire, and additional questions to sleepiness-related accidents, near-miss accidents, health issues, and demographic data. Of the 435 distributed questionnaires, 128 completed were returned. The response rate was 29%. The mean age of the investigated drivers was 42.5 years (20-85 years). According to the Epworth Sleepiness Scale, 9% of the participants are likely to suffer from excessive daytime sleepiness. An equal percentage has a high risk for obstructive sleep apnea syndrome based on the Berlin Questionnaire. 16% admitted an involuntary nodding off while driving a motor vehicle. This subset of the participants scored statistically significant higher on the Epworth Sleepiness Scale (p = 0.036). 8% of the participants already suffered an accident because of being sleepy while driving. An equal number experienced a sleepiness-related near-miss accident on the road. The study shows that a medical workup of excessive daytime sleepiness is highly recommended in each driver undergoing a driving capacity assessment. Routine application of easily available and time-saving assessment tools such as the Epworth Sleepiness Scale questionnaire could prevent accidents in a simple way. The applicability of the Berlin Questionnaire to screen suspected fatal sleepiness-related motor vehicle accidents is discussed. Copyright © 2016 Elsevier Ltd and Faculty of Forensic and Legal Medicine. All rights reserved.

  15. Excessive Daytime Sleepiness Predicts Neurodegeneration in Idiopathic REM Sleep Behavior Disorder.

    PubMed

    Zhou, Junying; Zhang, Jihui; Lam, Siu Ping; Chan, Joey Wy; Mok, Vincent; Chan, Anne; Li, Shirley Xin; Liu, Yaping; Tang, Xiangdong; Yung, Wing Ho; Wing, Yun Kwok

    2017-05-01

    To determine the association of excessive daytime sleepiness (EDS) with the conversion of neurodegenerative diseases in patients with idiopathic REM sleep behavior disorder (iRBD). A total of 179 patients with iRBD (79.1% males, mean age = 66.3 ± 9.8 years) were consecutively recruited. Forty-five patients with Epworth Sleepiness Scale score ≥14 were defined as having EDS. Demographic, clinical, and polysomnographic data were compared between iRBD patients with and without EDS. The risk of developing neurodegenerative diseases was examined using Cox proportional hazards model. After a mean follow-up of 5.8 years (SD = 4.3 years), 50 (27.9%) patients developed neurodegenerative diseases. There was a significantly higher proportion of conversion in patients with EDS compared to those without EDS (42.2 % vs. 23.1%, p = .01). EDS significantly predicted an increased risk of developing neurodegenerative diseases (adjusted hazard ratios [HR] = 2.56, 95% confidence interval [CI] 1.37 to 4.77) after adjusting for age, sex, body mass index, current depression, obstructive sleep apnea, and periodic limb movements during sleep. Further analyses demonstrated that EDS predicted the conversion of Parkinson's disease (PD) (adjusted HR = 3.55, 95% CI 1.59 to 7.89) but not dementia (adjusted HR = 1.48, 95% CI 0.44 to 4.97). EDS is associated with an increased risk of developing neurodegenerative diseases, especially PD, in patients with iRBD. Our findings suggest that EDS is a potential clinical biomarker of α-synucleinopathies in iRBD. © Sleep Research Society 2017. Published by Oxford University Press on behalf of the Sleep Research Society. All rights reserved. For permissions, please e-mail journals.permissions@oup.com.

  16. Validation of the Epworth Sleepiness Scale for Children and Adolescents using Rasch analysis.

    PubMed

    Janssen, Kitty C; Phillipson, Sivanes; O'Connor, Justen; Johns, Murray W

    2017-05-01

    A validated measure of daytime sleepiness for adolescents is needed to better explore emerging relationships between sleepiness and the mental and physical health of adolescents. The Epworth Sleepiness Scale (ESS) is a widely used scale for daytime sleepiness in adults but contains references to alcohol and driving. The Epworth Sleepiness Scale for Children and Adolescents (ESS-CHAD) has been proposed as the official modified version of the ESS for children and adolescents. This study describes the psychometric analysis of the ESS-CHAD as a measure of daytime sleepiness for adolescents. The ESS-CHAD was completed by 297 adolescents, 12-18 years old, from two independent schools in Victoria, Australia. Exploratory factor analysis and Rasch analysis was conducted to determine the validity of the scale. Exploratory factor analysis and Rasch analysis indicated that ESS-CHAD has internal validity and a unidimensional structure with good model fit. Rasch analysis of four subgroups based on gender and year-level were consistent with the overall results. The results were consistent with published ESS results, which strongly indicates that the changes to the scale do not affect the scale's capacity to measure daytime sleepiness. It is concluded that the ESS-CHAD is a reliable and internally valid measure of daytime sleepiness in adolescents 12-18 years old. Further studies are needed to establish the internal validity of the ESS-CHAD for children under 12 years, and to establish external validity and accurate cut-off points for children and adolescents. Copyright © 2017 Elsevier B.V. All rights reserved.

  17. Effects of school start time on students' sleep duration, daytime sleepiness, and attendance: a meta-analysis.

    PubMed

    Bowers, Jennifer M; Moyer, Anne

    2017-12-01

    Research conducted over the past three decades finds that many children and adolescents do not meet recommended sleep guidelines. Lack of sleep is a predictor of a number of consequences, including issues at school such as sleepiness and tardiness. Considering the severity of this public health issue, it is essential to understand more about the factors that may compromise children's and adolescents' sleep. This meta-analysis examined the effects of school start time (SST) on sleep duration of students by aggregating the results of five longitudinal studies and 15 cross-sectional comparison group studies. Results indicated that later starting school times are associated with longer sleep durations. Additionally, later start times were associated with less daytime sleepiness (7 studies) and tardiness to school (3 studies). However, methodological considerations, such as a need for more longitudinal primary research, lead to a cautious interpretation. Overall, this systematic analysis of SST studies suggests that delaying SST is associated with benefits for students' sleep and, thus, their general well-being. Copyright © 2017 National Sleep Foundation. Published by Elsevier Inc. All rights reserved.

  18. Poor sleep quality and insufficient sleep of a collegiate student-athlete population.

    PubMed

    Mah, Cheri D; Kezirian, Eric J; Marcello, Brandon M; Dement, William C

    2018-06-01

    Poor and inadequate sleep negatively impact cognitive and physical functioning and may also affect sports performance. The study aim is to examine sleep quality, sleep duration, and daytime sleepiness in collegiate student-athletes across a wide range of sports. Questionnaire. University setting. 628 athletes across 29 varsity teams at Stanford University. Athletes completed a questionnaire inquiring about sleep quality via a modified Pittsburgh Sleep Quality Index (PSQI), sleep duration, and daytime sleepiness via Epworth Sleepiness Scale. Sleep quality on campus and while traveling for competition was rated on a 10-point scale. Collegiate athletes were classified as poor sleepers (PSQI 5.38 ± 2.45), and 42.4% of athletes experience poor sleep quality (reporting PSQI global scores >5). Athletes reported lower sleep quality on campus than when traveling for competition (7.1 vs 7.6, P< .001). Inadequate sleep was demonstrated by 39.1% of athletes that regularly obtain <7 hours of sleep on weekdays. Fifty-one percent of athletes reported high levels of daytime sleepiness with Epworth scores ≥10. Teen student-athletes in the first and second year of college reported the highest mean levels of daytime sleepiness. Greater total sleep time was associated with daytime functioning including lower frequency of difficulty waking up for practice or class (P< .001) and lower frequency of trouble staying awake during daily activities (P< .001). Collegiate athletes frequently experience poor sleep quality, regularly obtain insufficient sleep, and commonly exhibit daytime sleepiness. Copyright © 2018 National Sleep Foundation. All rights reserved.

  19. Causes and consequences of sleepiness among college students

    PubMed Central

    Hershner, Shelley D; Chervin, Ronald D

    2014-01-01

    Daytime sleepiness, sleep deprivation, and irregular sleep schedules are highly prevalent among college students, as 50% report daytime sleepiness and 70% attain insufficient sleep. The consequences of sleep deprivation and daytime sleepiness are especially problematic to college students and can result in lower grade point averages, increased risk of academic failure, compromised learning, impaired mood, and increased risk of motor vehicle accidents. This article reviews the current prevalence of sleepiness and sleep deprivation among college students, contributing factors for sleep deprivation, and the role of sleep in learning and memory. The impact of sleep and sleep disorders on academics, grade point average, driving, and mood will be examined. Most importantly, effective and viable interventions to decrease sleepiness and sleep deprivation through sleep education classes, online programs, encouragement of naps, and adjustment of class time will be reviewed. This paper highlights that addressing sleep issues, which are not often considered as a risk factor for depression and academic failure, should be encouraged. Promotion of university and college policies and class schedules that encourage healthy and adequate sleep could have a significant impact on the sleep, learning, and health of college students. Future research to investigate effective and feasible interventions, which disseminate both sleep knowledge and encouragement of healthy sleep habits to college students in a time and cost effective manner, is a priority. PMID:25018659

  20. Causes and consequences of sleepiness among college students.

    PubMed

    Hershner, Shelley D; Chervin, Ronald D

    2014-01-01

    Daytime sleepiness, sleep deprivation, and irregular sleep schedules are highly prevalent among college students, as 50% report daytime sleepiness and 70% attain insufficient sleep. The consequences of sleep deprivation and daytime sleepiness are especially problematic to college students and can result in lower grade point averages, increased risk of academic failure, compromised learning, impaired mood, and increased risk of motor vehicle accidents. This article reviews the current prevalence of sleepiness and sleep deprivation among college students, contributing factors for sleep deprivation, and the role of sleep in learning and memory. The impact of sleep and sleep disorders on academics, grade point average, driving, and mood will be examined. Most importantly, effective and viable interventions to decrease sleepiness and sleep deprivation through sleep education classes, online programs, encouragement of naps, and adjustment of class time will be reviewed. This paper highlights that addressing sleep issues, which are not often considered as a risk factor for depression and academic failure, should be encouraged. Promotion of university and college policies and class schedules that encourage healthy and adequate sleep could have a significant impact on the sleep, learning, and health of college students. Future research to investigate effective and feasible interventions, which disseminate both sleep knowledge and encouragement of healthy sleep habits to college students in a time and cost effective manner, is a priority.

  1. Pharmacological interventions for sleepiness and sleep disturbances caused by shift work.

    PubMed

    Liira, Juha; Verbeek, Jos H; Costa, Giovanni; Driscoll, Tim R; Sallinen, Mikael; Isotalo, Leena K; Ruotsalainen, Jani H

    2014-08-12

    Shift work results in sleep-wake disturbances, which cause sleepiness during night shifts and reduce sleep length and quality in daytime sleep after the night shift. In its serious form it is also called shift work sleep disorder. Various pharmacological products are used to ameliorate symptoms of sleepiness or poor sleep length and quality. To evaluate the effects of pharmacological interventions to reduce sleepiness or to improve alertness at work and decrease sleep disturbances whilst off work, or both, in workers undertaking shift work in their present job and to assess their cost-effectiveness. We searched CENTRAL, MEDLINE, EMBASE, PubMed and PsycINFO up to 20 September 2013 and ClinicalTrials.gov up to July 2013. We also screened reference lists of included trials and relevant reviews. We included all eligible randomised controlled trials (RCTs), including cross-over RCTs, of pharmacological products among workers who were engaged in shift work (including night shifts) in their present jobs and who may or may not have had sleep problems. Primary outcomes were sleep length and sleep quality while off work, alertness and sleepiness, or fatigue at work. Two authors independently selected studies, extracted data and assessed risk of bias in included trials. We performed meta-analyses where appropriate. We included 15 randomised placebo-controlled trials with 718 participants. Nine trials evaluated the effect of melatonin and two the effect of hypnotics for improving sleep problems. One trial assessed the effect of modafinil, two of armodafinil and one examined caffeine plus naps to decrease sleepiness or to increase alertness.Melatonin (1 to 10 mg) after the night shift may increase sleep length during daytime sleep (mean difference (MD) 24 minutes, 95% confidence interval (CI) 9.8 to 38.9; seven trials, 263 participants, low quality evidence) and night-time sleep (MD 17 minutes, 95% CI 3.71 to 30.22; three trials, 234 participants, low quality evidence) compared

  2. The effect of nocturia on sleep quality and daytime function in patients with lower urinary tract symptoms: a cross-sectional study.

    PubMed

    Shao, I-Hung; Wu, Chia-Chen; Hsu, Hueih-Shing; Chang, Shyh-Chyi; Wang, Hsu-Hsiang; Chuang, Heng-Chang; Tam, Yuan-Yun

    2016-01-01

    Nocturia has been proven to have a negative impact on the quality of life and sleep quality in general elderly population. However, there are limited studies on the quantitative effect of nocturia on sleep quality and daytime dysfunction, specifically in patients with lower urinary tract symptoms. During March 1, 2015 to December 31, 2015, a total of 728 patients who visited our urology department due to voiding dysfunction and experienced nocturia at least once per night were enrolled. Three questionnaires were administered to them after obtaining their written consents. Pittsburgh Sleep Quality Index (PSQI) questionnaire, Epworth Sleepiness Scale (ESS) questionnaire, and International Prostate Symptom Score (IPSS) questionnaire were applied to evaluate their sleep quality, daytime dysfunction, and voiding problems, respectively. Statistical analysis of the impact of nocturia on sleep quality and daytime dysfunction was performed. The mean age of patients was 61 years, with a male-to-female ratio of 2.7. The mean nocturia number was 3.03. The IPSS, PSQI, and ESS scores were 17.56, 8.35, and 8.22, respectively. The nocturia number increased with age and was significantly correlated to ESS score (daytime dysfunction) and PSQI total score (sleep quality) in overall group. Among subgroups divided by age and sex, there was a significant correlation between nocturia number and daytime dysfunction in male patients or patients younger than 65 years. In patients with lower urinary tract symptoms, nocturia number increased with age and was significantly correlated with poor sleep quality. Nocturia plays an important role in patients younger than 65 years in daytime dysfunction.

  3. Sleep-Disordered Breathing and Excessive Daytime Sleepiness in Patients With Atrial Fibrillation

    PubMed Central

    Albuquerque, Felipe N.; Calvin, Andrew D.; Sert Kuniyoshi, Fatima H.; Konecny, Tomas; Lopez-Jimenez, Francisco; Pressman, Gregg S.; Kara, Thomas; Friedman, Paul; Ammash, Naser; Somers, Virend K.

    2012-01-01

    Background: An important consequence of sleep-disordered breathing (SDB) is excessive daytime sleepiness (EDS). EDS often predicts a favorable response to treatment of SDB, although in the setting of cardiovascular disease, particularly heart failure, SDB and EDS do not reliably correlate. Atrial fibrillation (AF) is another highly prevalent condition strongly associated with SDB. We sought to assess the relationship between EDS and SDB in patients with AF. Methods: We conducted a prospective study of 151 patients referred for direct current cardioversion for AF who also underwent sleep evaluation and nocturnal polysomnography. The Epworth Sleepiness Scale (ESS) was administered prior to polysomnography and considered positive if the score was ≥ 11. The apnea-hypopnea index (AHI) was tested for correlation with the ESS, with a cutoff of ≥ 5 events/h for the diagnosis of SDB. Results: Among the study participants, mean age was 69.1 ± 11.7 years, mean BMI was 34.1 ± 8.4 kg/m2, and 76% were men. The prevalence of SDB in this population was 81.4%, and 35% had EDS. The association between ESS score and AHI was low (R2 = 0.014, P = .64). The sensitivity and specificity of the ESS for the detection of SDB in patients with AF were 32.2% and 54.5%, respectively. Conclusions: Despite a high prevalence of SDB in this population with AF, most patients do not report EDS. Furthermore, EDS does not appear to correlate with severity of SDB or to accurately predict the presence of SDB. Further research is needed to determine whether EDS affects the natural history of AF or modifies the response to SDB treatment. PMID:21903736

  4. Internet overuse and excessive daytime sleepiness in adolescents.

    PubMed

    Choi, Kwisook; Son, Hyunsook; Park, Myunghee; Han, Jinkyu; Kim, Kitai; Lee, Byungkoo; Gwak, Hyesun

    2009-08-01

    The purpose of this study was to examine the association of Internet overuse with excessive daytime sleepiness (EDS). A total of 2336 high school students in South Korea (boys, 57.5%; girls, 42.5%) completed the structured questionnaire. The severity of Internet addiction was evaluated using Young's Internet addiction test. The proportions of boys who were classified as Internet addicts and possible Internet addicts were 2.5% and 53.7%, respectively. For girls, the corresponding proportions were 1.9% and 38.9%, respectively. The prevalence of EDS was 11.2% (boys, 11.2%; girls, 11.1%). When Internet addicts were compared with non-addicts, they consisted of more boys, drank alcohol more, and considered their own health condition as poor. But smoking was not related with Internet addiction. The prevalence rate of EDS for Internet addicts was 37.7%, whereas that for possible Internet addicts and non-addicts was 13.9% and 7.4%, respectively. The prevalence of insomnia, witnessed snoring, apnea, teeth grinding, and nightmares was highest in Internet addicts, middle in possible addicts, and lowest in non-addicts. With adjustment for duration of Internet use, duration of sleep time, age, gender, smoking, taking painkillers due to headache, insomnia symptoms, witnessed apnea, and nightmares, the odds of EDS were 5.2-fold greater (95% confidence interval [CI]: 2.7-10.2) in Internet addicts and 1.9-fold greater (95%CI: 1.4-2.6) in possible Internet addicts compared to non-addicts. Internet addiction is strongly associated with EDS in adolescents. Clinicians should consider examining Internet addiction in adolescent cases of EDS.

  5. Learning, Attention/Hyperactivity, and Conduct Problems as Sequelae of Excessive Daytime Sleepiness in a General Population Study of Young Children

    PubMed Central

    Calhoun, Susan L.; Fernandez-Mendoza, Julio; Vgontzas, Alexandros N.; Mayes, Susan D.; Tsaoussoglou, Marina; Rodriguez-Muñoz, Alfredo; Bixler, Edward O.

    2012-01-01

    Study Objectives: Although excessive daytime sleepiness (EDS) is a common problem in children, with estimates of 15%; few studies have investigated the sequelae of EDS in young children. We investigated the association of EDS with objective neurocognitive measures and parent reported learning, attention/hyperactivity, and conduct problems in a large general population sample of children. Design: Cross-sectional. Setting: Population based. Participants: 508 children from The Penn State Child Cohort. Interventions: N/A. Measurements and Results: Children underwent a 9-h polysomnogram, comprehensive neurocognitive testing, and parent rating scales. Children were divided into 2 groups: those with and without parent-reported EDS. Structural equation modeling was used to examine whether processing speed and working memory performance would mediate the relationship between EDS and learning, attention/hyperactivity, and conduct problems. Logistic regression models suggest that parent-reported learning, attention/hyperactivity, and conduct problems, as well as objective measurement of processing speed and working memory are significant sequelae of EDS, even when controlling for AHI and objective markers of sleep. Path analysis demonstrates that processing speed and working memory performance are strong mediators of the association of EDS with learning and attention/hyperactivity problems, while to a slightly lesser degree are mediators from EDS to conduct problems. Conclusions: This study suggests that in a large general population sample of young children, parent-reported EDS is associated with neurobehavioral (learning, attention/hyperactivity, conduct) problems and poorer performance in processing speed and working memory. Impairment due to EDS in daytime cognitive and behavioral functioning can have a significant impact on children's development. Citation: Calhoun SL; Fernandez-Mendoza J; Vgontzas AN; Mayes SD; Tsaoussoglou M; Rodriguez-Muñoz A; Bixler EO. Learning

  6. Sleep and sleepiness in environmental intolerances: a population-based study.

    PubMed

    Nordin, Maria; Nordin, Steven

    2016-08-01

    About one fourth of the general population report environmental intolerance (EI) to odorous/pungent chemicals, certain buildings, electromagnetic fields (EMFs), and/or sounds. EI sufferers show various clinical features, of which sleep disturbance is one. Sleep disturbance is common also in the general population, but it is not known whether the disturbance is more prominent in EI sufferers than in individuals who do not experience EI. Therefore, EI was compared on various sleep aspects with referents without EI. A population-based sample of 3406 individuals, aged 18-79 years, was recruited from Northern Sweden. Sleep quality, non-restorative sleep, daytime sleepiness, obstructive breathing, and nocturnal insomnia were assessed with the Karolinska Sleep Questionnaire. Single questions assessed time slept, amount of hours of needed sleep, and extent of enough time slept. All four EI groups, compared to the referents, reported significantly poorer sleep quality, more non-restorative sleep, more daytime sleepiness, more obstructive breathing and higher prevalence of nocturnal insomnia than the referents. Nocturnal insomnia was an important factor for EI groups attributing their most prevalent symptoms to chemicals and sounds, irrespective of distress and certain syndromes. None of the EI groups differed significantly from the referents on time slept, but reported needing more sleep time (the EMF-intolerance group showing only a tendency), and all four groups reported to perceive enough sleep to a significantly lesser extent. Sleep disturbance and daytime sleepiness are more common in individuals reporting EI compared to normal referents. Moreover, nocturnal insomnia is an important symptom in its own right in various types of EI. This evokes the question of whether or not sleep therapy may attenuate the severity of the EI. Copyright © 2016. Published by Elsevier B.V.

  7. Exercise training improves selected aspects of daytime functioning in adults with obstructive sleep apnea.

    PubMed

    Kline, Christopher E; Ewing, Gary B; Burch, James B; Blair, Steven N; Durstine, J Larry; Davis, J Mark; Youngstedt, Shawn D

    2012-08-15

    To explore the utility of exercise training for improving daytime functioning in adults with obstructive sleep apnea (OSA). Forty-three sedentary and overweight/obese adults aged 18-55 years with at least moderate-severity untreated OSA (apnea-hypopnea index ≥ 15) were randomized to 12 weeks of moderate-intensity aerobic and resistance exercise training (n = 27) or low-intensity stretching control treatment (n = 16). As part of a trial investigating the efficacy of exercise training on OSA severity, daytime functioning was assessed before and following the intervention. Sleepiness, functional impairment due to sleepiness, depressive symptoms, mood, and quality of life (QOL) were evaluated with validated questionnaires, and cognitive function was assessed with a neurobehavioral performance battery. OSA severity was measured with one night of laboratory polysomnography before and following the intervention. Compared with stretching control, exercise training resulted in significant improvements in depressive symptoms, fatigue and vigor, and aspects of QOL (p < 0.05). Sleepiness and functional impairment due to sleepiness also were improved following exercise versus control to a similar degree in terms of effect sizes (d > 0.5), though these changes were not statistically significant. No neurobehavioral performance improvements were found. Reduced fatigue following exercise training was mediated by a reduction in OSA severity, but changes in OSA severity did not significantly mediate improvement in any other measure of daytime functioning. These data provide preliminary evidence that exercise training may be helpful for improving aspects of daytime functioning of adults with OSA. Larger trials are needed to further verify the observed improvements.

  8. 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

  9. Insomnia and Sleepiness in Parkinson Disease: Associations with Symptoms and Comorbidities

    PubMed Central

    Chung, Seockhoon; Bohnen, Nicolaas I.; Albin, Roger L.; Frey, Kirk A.; Müller, Martijn L. T. M.; Chervin, Ronald D.

    2013-01-01

    Study Objectives: Insomnia and daytime sleepiness are common complaints in Parkinson disease (PD), but the main causes remain unclear. We examined the potential impact of both motor and non-motor symptoms of PD on sleep problems. Methods: Patients with PD (n = 128) were assessed using the Insomnia Severity Index, Epworth Sleepiness Scale, Unified Parkinson Disease Rating Scale, Beck Depression Inventory, Fatigue Severity Scale, Survey of Autonomic Symptoms, and the 39-item Parkinson Disease Questionnaire. A subset of subjects (n = 38, 30%) also completed nocturnal polysomnography and a multiple sleep latency test (MSLT). Results: Multivariate stepwise logistic regression models revealed that subjective insomnia was independently associated with depressed mood (odds ratio [OR] = 1.79; 95% confidence interval (CI) [1.01-3.19]), autonomic symptoms (1.77 [1.08-2.90]), fatigue (1.19 [1.02-1.38]), and age (0.61 [0.39-0.96]). Subjective daytime sleepiness was associated with dosage of dopaminergic medication (1.74 [1.08-2.80]) and fatigue (1.14 [1.02-1.28]). On polysomnography, longer sleep latency correlated with autonomic symptoms (rho = 0.40, p = 0.01) and part I (non-motor symptoms) of the Unified PD Rating Scale (rho = 0.38, p = 0.02). Decreased sleep efficiency correlated with autonomic symptoms (rho = -0.42, p < 0.0001). However, no significant difference emerged on polysomnography and MSLTs between patients with or without insomnia or daytime sleepiness. Higher rates of apneic events did predict shorter sleep latencies on the MSLTs. Conclusions: Non-motor symptoms appear to be associated with subjective insomnia, whereas fatigue and dopaminergic medication are associated with subjective daytime sleepiness. Objective sleep laboratory data provided little insight into complaints of insomnia and sleepiness, though obstructive sleep apnea predicted worsened sleepiness when measured objectively. Citation: Chung S; Bohnen NI; Albin RL; Frey KA; Müller MLTM; Chervin RD

  10. Association of Current Work and Sleep Situations with Excessive Daytime Sleepiness and Medical Incidents among Japanese Physicians

    PubMed Central

    Kaneita, Yoshitaka; Ohida, Takashi

    2011-01-01

    Objective: The aim of the present study was to clarify the current work and sleep situations of physicians in Japan and to clarify the association between these situations and excessive daytime sleepiness as well as medical incidents. Methods: A self-administered questionnaire survey was conducted among the members of the Japan Medical Association in 2008. The randomly selected subjects comprised 3,000 male physicians and 1,500 female physicians. Results: Valid responses were obtained from 3,486 physicians (2,298 men and 1,188 women). Mean sleep duration was 6 h 36 min for men and 6 h 8 min for women. The prevalence of lack of rest due to sleep deprivation was 30.4% among men and 36.6% among women; the prevalence of insomnia was 21.0% and 18.1%, respectively; and the prevalence of EDS was 3.5%. The adjusted odds ratio for EDS was high for physicians who reported short sleep duration, lack of rest due to sleep deprivation, and a high frequency of on-call/overnight work. Physicians who had experienced a medical incident within the previous one month accounted for 19.0% of participants. The adjusted odds ratio for medical incidents was high for those subjected to long working hours, high frequency of on-call/overnight works, lack of rest due to sleep deprivation, and insomnia. Conclusion: In order to facilitate optimal health management for physicians as well as securing medical safety, it is important to fully consider the work and sleep situations of physicians. Citation: Kaneita Y; Ohida T. Association of current work and sleep situations with excessive daytime sleepiness and medical incidents among Japanese physicians. J Clin Sleep Med 2011;7(5):512-522. PMID:22003348

  11. Exercise Training Improves Selected Aspects of Daytime Functioning in Adults with Obstructive Sleep Apnea

    PubMed Central

    Kline, Christopher E.; Ewing, Gary B.; Burch, James B.; Blair, Steven N.; Durstine, J. Larry; Davis, J. Mark; Youngstedt, Shawn D.

    2012-01-01

    Study Objectives: To explore the utility of exercise training for improving daytime functioning in adults with obstructive sleep apnea (OSA). Methods: Forty-three sedentary and overweight/obese adults aged 18-55 years with at least moderate-severity untreated OSA (apnea-hypopnea index ≥ 15) were randomized to 12 weeks of moderate-intensity aerobic and resistance exercise training (n = 27) or low-intensity stretching control treatment (n = 16). As part of a trial investigating the efficacy of exercise training on OSA severity, daytime functioning was assessed before and following the intervention. Sleepiness, functional impairment due to sleepiness, depressive symptoms, mood, and quality of life (QOL) were evaluated with validated questionnaires, and cognitive function was assessed with a neurobehavioral performance battery. OSA severity was measured with one night of laboratory polysomnography before and following the intervention. Results: Compared with stretching control, exercise training resulted in significant improvements in depressive symptoms, fatigue and vigor, and aspects of QOL (p < 0.05). Sleepiness and functional impairment due to sleepiness also were improved following exercise versus control to a similar degree in terms of effect sizes (d > 0.5), though these changes were not statistically significant. No neurobehavioral performance improvements were found. Reduced fatigue following exercise training was mediated by a reduction in OSA severity, but changes in OSA severity did not significantly mediate improvement in any other measure of daytime functioning. Conclusions: These data provide preliminary evidence that exercise training may be helpful for improving aspects of daytime functioning of adults with OSA. Larger trials are needed to further verify the observed improvements. Trial Registration: Clinicaltrials.gov identification number NCT00956423. Citation: Kline CE; Ewing GB; Burch JB; Blair SN; Durstine JL; Davis JM; Youngstedt SD. Exercise

  12. Habitual Sleep Duration, Unmet Sleep Need, and Excessive Daytime Sleepiness in Korean Adults.

    PubMed

    Hwangbo, Young; Kim, Won Joo; Chu, Min Kyung; Yun, Chang Ho; Yang, Kwang Ik

    2016-04-01

    Sleep need differs between individuals, and so the same duration of sleep will lead to sleep insufficiency in some individuals but not others. The aim of this study was to determine the separate and combined associations of both sleep duration and unmet sleep need with excessive daytime sleepiness (EDS) in Korean adults. The participants comprised 2,769 Korean adults aged 19 years or older. They completed questionnaires about their sleep habits over the previous month. The question regarding sleep need was "How much sleep do you need to be at your best during the day?" Unmet sleep need was calculated as sleep need minus habitual sleep duration. Participants with a score of >10 on the Epworth Sleepiness Scale were considered to have EDS. The overall prevalence of EDS was 11.9%. Approximately one-third of the participants (31.9%) reported not getting at least 7 hours of sleep. An unmet sleep need of >0 hours was present in 30.2% of the participants. An adjusted multivariate logistic regression analysis revealed a significant excess risk of EDS in the groups with unmet sleep needs of ≥2 hours [odds ratio (OR), 1.80; 95% confidence interval (CI), 1.27-2.54] and 0.01-2 hours (OR, 1.42; 95% CI, 1.02-1.98). However, habitual sleep duration was not significantly related to EDS. EDS was found to be associated with unmet sleep need but not with habitual sleep duration when both factors were examined together. We suggest that individual unmet sleep need is more important than habitual sleep duration in terms of the relation to EDS.

  13. Sleep quality of German soldiers before, during and after deployment in Afghanistan-a prospective study.

    PubMed

    Danker-Hopfe, Heidi; Sauter, Cornelia; Kowalski, Jens T; Kropp, Stefan; Ströhle, Andreas; Wesemann, Ulrich; Zimmermann, Peter L

    2017-06-01

    In this prospective study, subjective sleep quality and excessive daytime sleepiness prior to, during and after deployment of German soldiers in Afghanistan were examined. Sleep quality (Pittsburgh Sleep Quality Index; PSQI) and daytime sleepiness (Epworth Sleepiness Scale; ESS) were assessed in 118 soldiers of the German army, who were deployed in Afghanistan for 6 months (deployment group: DG) and in 146 soldiers of a non-deployed control group (CG) at baseline. Results of the longitudinal analysis are reported, based on assessments conducted prior to, during the deployment and afterwards in the DG, and in the CG in parallel. Sleep quality and daytime sleepiness in the DG were already impaired during the predeployment training phase and remained at that level during the deployment phase, which clearly indicates the need for more attention on sleep in young soldiers, already at this early stage. The percentage of impaired sleepers decreased significantly after deployment. Programmes to teach techniques to improve sleep and reduce stress should be implemented prior to deployment to reduce sleep difficulties and excessive daytime sleepiness and subsequent psychiatric disorders. © 2017 European Sleep Research Society.

  14. The relationship between driving simulation performance and obstructive sleep apnoea risk, daytime sleepiness, obesity and road traffic accident history of commercial drivers in Turkey.

    PubMed

    Demirdöğen Çetinoğlu, Ezgi; Görek Dilektaşlı, Aslı; Demir, Nefise Ateş; Özkaya, Güven; Acet, Nilüfer Aylin; Durmuş, Eda; Ursavaş, Ahmet; Karadağ, Mehmet; Ege, Ercüment

    2015-09-01

    Driving performance is known to be very sensitive to cognitive-psychomotor impairment. The aim of the study was to determine the relationship between obesity, risk of obstructive sleep apnoea (OSA), daytime sleepiness, history of road traffic accident (RTA) and performance on a driving simulator, among commercial drivers. We examined commercial vehicle drivers admitted to Psycho-Technical Assessment System (PTAS), which is a computer-aided system that includes a driving simulator test and tests assessing psychomotor-cognitive skills required for driving. Risk of OSA and daytime sleepiness were assessed by the Berlin Questionnaire and the Epworth Sleepiness Scale (ESS), respectively. A total of 282 commercial vehicle drivers were consecutively enrolled. The age range was 29-76 years. Thirty drivers were at high risk of OSA. Median ESS of the group was 2 (0-20). Forty-seven percent of the subjects at high risk of OSA failed in early reaction time test, while 28% of the drivers with low risk of OSA failed (p = 0.03). The obese drivers failed the peripheral vision test when compared with non-obese drivers (p = 0.02). ESS was higher for drivers with a history of RTA when compared to those without RTA (p = 0.02). Cognitive-psychomotor functions can be impaired in obese and high risk of OSA patients. In our opinion, requiring obese and/or high risk of OSA drivers to take PTAS tests that assess driving skills and psychomotor-cognitive functions crucial to those skills would significantly improve road traffic safety, which is of considerable importance to public health.

  15. Pharmacological interventions for sleepiness and sleep disturbances caused by shift work.

    PubMed

    Liira, Juha; Verbeek, Jos H; Costa, Giovanni; Driscoll, Tim R; Sallinen, Mikael; Isotalo, Leena K; Ruotsalainen, Jani H

    2015-02-01

    Shift work results in sleep-wake disturbances, which cause sleepiness during night shifts and reduce sleep length and quality in daytime sleep after the night shift. In its serious form it is also called shift work sleep disorder. Various pharmacological products are used to ameliorate symptoms of sleepiness or poor sleep length and quality. To evaluate the effects of pharmacological interventions to reduce sleepiness or to improve alertness at work and decrease sleep disturbances whilst of work, or both, in workers undertaking shift work. We searched CENTRAL, MEDLINE, EMBASE, PubMed and PsycINFO up to 20 September 2013 and ClinicalTrials.gov up to July 2013. We also screened reference lists of included trials and relevant reviews. We included all eligible randomised controlled trials (RCTs), including cross-over RCTs, of pharmacological products among workers who were engaged in shift work (including night shifts) in their present jobs and who may or may not have had sleep problems. Primary outcomes were sleep length and sleep quality while of work, alertness and sleepiness, or fatigue at work. Two authors independently selected studies, extracted data and assessed risk of bias in included trials. We performed meta-analyses where appropriate. We included 15 randomised placebo-controlled trials with 718 participants. Nine trials evaluated the effect of melatonin and two the effect of hypnotics for improving sleep problems. One trial assessed the effect of modafinil, two of armodafinil and one examined caffeine plus naps to decrease sleepiness or to increase alertness.

  16. Drug treatment of patients with insomnia and excessive daytime sleepiness: pharmacokinetic considerations.

    PubMed

    Nishino, S; Mignot, E

    1999-10-01

    Insomnia and excessive daytime sleepiness (EDS) are frequently observed conditions in the general public. A national survey in the USA in 1979 indicated that 35% of American adults experience insomnia in the course of a year. The prevalence of EDS varies depending on the survey (0.3 to 13.3%), but a recent study stated that 2.4% of individuals reported that they continually fell asleep at work. These problems are often long term and negatively affect the individuals' quality of life. People with these sleep problems often have difficulties maintaining high levels of productivity at work or pursuing their daily activities; individuals with insomnia lack the feeling of being rested or refreshed in the morning and EDS is unavoidable in most cases. Behavioural therapy has been shown to be effective for many people affected with insomnia and EDS. However, pharmacological treatments using hypnosedatives and central nervous system (CNS) stimulants are usually necessary, and effective, for those with more severe cases. These compounds have thus been widely prescribed in clinical practice (e.g., 2.6% of all adults surveyed used medically prescribed hypnosedatives and 4.5% used over-the-counter drugs to promote sleep). The onset and duration of action of these hypnosedatives and CNS stimulant drugs are important factors to be considered when prescribing these compounds. These factors primarily depend on physicochemical properties (lipid solubility and protein binding), as well as the pharmacokinetic profile (absorption, distribution, elimination and clearance) of the compounds. Significant differences in profile exist amongst hypnosedatives and CNS stimulants, and these differences may account for the observed variations in clinical action and adverse effects during and after treatment. In this review, we will introduce recently obtained knowledge of the pharmacokinetics of hypnosedatives and CNS stimulants and their applications for patients affected with insomnia and EDS.

  17. Perceived sleepiness in Canadian anesthesia residents: a national survey.

    PubMed

    Hanlon, John G; Hayter, Megan A; Bould, M Dylan; Joo, Hwan S; Naik, Viren N

    2009-01-01

    To compare the self-perceived sleepiness of Canadian anesthesia residents providing modified on-call duties (12-16 h) vs. traditional on-call duties (24 h). A 25-item online survey was distributed to all Canadian anesthesia residents who, at that time, were on anesthesia rotations. The survey assessed resident demographics, perceived work patterns, and sleepiness, as well as their opinions on resident work hour reform. Self-perceived sleepiness was quantified using the validated Epworth sleepiness scale (ESS). Three hundred eight of 400 (77%) eligible Canadian anesthesia residents completed the survey. Forty-three percent of residents who worked traditional on-call (duration 24.1 +/- 0.5 h) shifts and 48% of residents who worked modified on-call (duration 15.5 +/- 1.8 h) shifts met ESS criteria for excessive daytime sleepiness. Overall mean ESS scores did not differ significantly between the traditional (9.1 +/- 4.9) and the modified call groups (9.5 +/- 4.8). Residents with an on-call frequency of >or=1:4 days or those who slept sleepy (P = 0.045 and P = 0.008, respectively). Six percent of residents admitted to taking "something other than caffeine" to stay awake on call. Many anesthesia residents do exhibit excessive daytime sleepiness, with a similar incidence for those working within either modified or traditional call systems. Our study suggests that sleepiness may be reduced by scheduling on-call duties no more frequently than one in every five nights and by ensuring that residents sleep more than 2 h while on call.

  18. Mutual information measures applied to EEG signals for sleepiness characterization.

    PubMed

    Melia, Umberto; Guaita, Marc; Vallverdú, Montserrat; Embid, Cristina; Vilaseca, Isabel; Salamero, Manel; Santamaria, Joan

    2015-03-01

    Excessive daytime sleepiness (EDS) is one of the main symptoms of several sleep related disorders with a great impact on the patient lives. While many studies have been carried out in order to assess daytime sleepiness, the automatic EDS detection still remains an open problem. In this work, a novel approach to this issue based on non-linear dynamical analysis of EEG signal was proposed. Multichannel EEG signals were recorded during five maintenance of wakefulness (MWT) and multiple sleep latency (MSLT) tests alternated throughout the day from patients suffering from sleep disordered breathing. A group of 20 patients with excessive daytime sleepiness (EDS) was compared with a group of 20 patients without daytime sleepiness (WDS), by analyzing 60-s EEG windows in waking state. Measures obtained from cross-mutual information function (CMIF) and auto-mutual-information function (AMIF) were calculated in the EEG. These functions permitted a quantification of the complexity properties of the EEG signal and the non-linear couplings between different zones of the scalp. Statistical differences between EDS and WDS groups were found in β band during MSLT events (p-value < 0.0001). WDS group presented more complexity than EDS in the occipital zone, while a stronger nonlinear coupling between occipital and frontal zones was detected in EDS patients than in WDS. The AMIF and CMIF measures yielded sensitivity and specificity above 80% and AUC of ROC above 0.85 in classifying EDS and WDS patients. Copyright © 2015 IPEM. Published by Elsevier Ltd. All rights reserved.

  19. Relationship between serum substance P levels and daytime sleepiness in obstructive sleep apnea syndrome.

    PubMed

    Ursavas, Ahmet; Karadag, Mehmet; Ilcol, Yesim Ozarda; Burgazlioglu, Basak; Ercan, Ilker; Gozu, R Oktay

    2007-05-01

    We hypothesized that intermittent hypoxia might influence serum substance P levels, and that this effect might in turn contribute in excessive daytime sleepiness (EDS) in patients with obstructive sleep apnea syndrome (OSAS). Fifty-five patients with newly diagnosed OSAS and 15 age-matched nonapneic control subjects were enrolled in this study. Full polysomnography was performed in all patients. Single blood samples were drawn between 8:00 am and 9:00 am after the sleep study. Substance P levels were analyzed with a competitive enzyme immunoassay (substance P EIA kit; Cayman Chemical; Ann Arbor, MI). There were no significant differences in age, gender, body mass index, smoking habit, and snoring between the two groups. Serum substance P levels in the OSAS group were significantly lower than that in the control group (p < 0.0001). Serum substance P levels were positively correlated with rapid eye movement sleep (r = 0.330, p = 0.049) and slow-wave sleep (r = 0.324, p = 0.049) phases. Serum substance P levels were negatively correlated with Epworth sleepiness scale score (r = - 0.253, p = 0.048), number of total apneas during the night (r = - 0.247, p = 0.036), number of respiratory events during the night (r = - 0.266, p = 0.024), apnea-hypopnea index (r = - 0.287, p = 0.015), respiratory arousal index (r = - 0.267, p = 0.026), time spent in apnea and hypopnea (r = - 0.307, p = 0.01), average oxygen desaturation (r = - 0.265, p = 0.026), and oxygen desaturation index (r = - 0.254, p = 0.031). We concluded that EDS seen in some of the OSAS patients might be associated with various pathophysiologic mechanisms including substance P levels.

  20. 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.

  1. 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

  2. "To sleep, perchance to tweet": in-bed electronic social media use and its associations with insomnia, daytime sleepiness, mood, and sleep duration in adults.

    PubMed

    Bhat, Sushanth; Pinto-Zipp, Genevieve; Upadhyay, Hinesh; Polos, Peter G

    2018-04-01

    The use of mobile device-based electronic social media (ESM) in bed is rapidly becoming commonplace, with potentially adverse impacts on sleep and daytime functioning. The purpose of this study was to determine the extent to which in-bed ESM use is associated with insomnia, daytime sleepiness, mood, and sleep duration in adults. This was a cross-sectional observational study conducted among 855 hospital employees and university students (mean age, 43.6years; 85% female) via an online questionnaire. Nearly 70% of participants indulged in in-bed ESM use, with nearly 15% spending an hour or more a night doing so. The degree of in-bed ESM use did not vary by gender, but higher levels of in-bed ESM use were seen in younger and middle-aged than elderly participants. Compared with participants with no in-bed ESM use and controlling for age, gender, and ethnicity, participants with high in-bed ESM use were more likely to have insomnia, anxiety, and short sleep duration on weeknights, but not depression or daytime sleepiness; low in-bed ESM use only increased the likelihood of short sleep duration on weeknights. In-bed ESM use by a bed partner did not have an adverse association with sleep or mood. In-bed ESM use is associated with sleep and mood dysfunction in adults. These findings are of relevance to clinicians, therapists, and the public at large, as they suggest that limitation of in-bed ESM use is a potential interventional strategy in the overall management of sleep hygiene and mental health. Copyright © 2017 National Sleep Foundation. Published by Elsevier Inc. All rights reserved.

  3. Adolescent Sleep and the Impact of Technology Use Before Sleep on Daytime Function.

    PubMed

    Johansson, Ann E E; Petrisko, Maria A; Chasens, Eileen R

    2016-01-01

    Technology has become pervasive in our culture, particularly among adolescents. The purpose of this study is to examine associations between use of technology before sleep and daytime function in adolescents. This study is a secondary analysis of respondents aged 13 to 21 years (N = 259) from the 2011 National Sleep Foundation's Sleep in America Poll. The survey included questions on demographics, sleep habits, and use of technology in the hour before bedtime. Daytime sleepiness was assessed with the Epworth Sleepiness Scale (ESS). Student's t-tests, Mann-Whitney U, and Fischer's exact tests were performed to detect differences in demographics, sleep duration, and technology use in the total sample, and between respondents with "adequate" compared to "inadequate" sleep. Correlations were calculated between technology frequency and daytime function. Adolescents had mean sleep duration of 7.3 ± 1.3 h. Almost all respondents (97%) used some form of technology before sleep. Increased technology use and the frequency of being awoken in the night by a cell phone were significantly associated with waking too early, waking unrefreshed, and daytime sleepiness (p < 0.05). Adolescents who reported "inadequate" sleep had shorter sleep duration, greater frequency of technology use before bedtime, feeling unrefreshed on waking, and greater daytime sleepiness than those reporting "adequate" sleep (all p-values < 0.05). Technology use before sleep by adolescents had negative consequences on nighttime sleep and on daytime function. Healthcare professionals who interact with adolescents should encourage technology to be curtailed before bedtime and for adolescents to value obtaining adequate sleep. Copyright © 2016 Elsevier Inc. All rights reserved.

  4. Non-REM sleep EEG power distribution in fatigue and sleepiness.

    PubMed

    Neu, Daniel; Mairesse, Olivier; Verbanck, Paul; Linkowski, Paul; Le Bon, Olivier

    2014-04-01

    The aim of this study is to contribute to the sleep-related differentiation between daytime fatigue and sleepiness. 135 subjects presenting with sleep apnea-hypopnea syndrome (SAHS, n=58) or chronic fatigue syndrome (CFS, n=52) with respective sleepiness or fatigue complaints and a control group (n=25) underwent polysomnography and psychometric assessments for fatigue, sleepiness, affective symptoms and perceived sleep quality. Sleep EEG spectral analysis for ultra slow, delta, theta, alpha, sigma and beta power bands was performed on frontal, central and occipital derivations. Patient groups presented with impaired subjective sleep quality and higher affective symptom intensity. CFS patients presented with highest fatigue and SAHS patients with highest sleepiness levels. All groups showed similar total sleep time. Subject groups mainly differed in sleep efficiency, wake after sleep onset, duration of light sleep (N1, N2) and slow wave sleep, as well as in sleep fragmentation and respiratory disturbance. Relative non-REM sleep power spectra distributions suggest a pattern of power exchange in higher frequency bands at the expense of central ultra slow power in CFS patients during all non-REM stages. In SAHS patients, however, we found an opposite pattern at occipital sites during N1 and N2. Slow wave activity presents as a crossroad of fatigue and sleepiness with, however, different spectral power band distributions during non-REM sleep. The homeostatic function of sleep might be compromised in CFS patients and could explain why, in contrast to sleepiness, fatigue does not resolve with sleep in these patients. The present findings thus contribute to the differentiation of both phenomena. Copyright © 2014 Elsevier Inc. All rights reserved.

  5. Neurobehavioral performance impairment in insomnia: relationships with self-reported sleep and daytime functioning.

    PubMed

    Shekleton, Julia A; Flynn-Evans, Erin E; Miller, Belinda; Epstein, Lawrence J; Kirsch, Douglas; Brogna, Lauren A; Burke, Liza M; Bremer, Erin; Murray, Jade M; Gehrman, Philip; Lockley, Steven W; Rajaratnam, Shantha M W

    2014-01-01

    Despite the high prevalence of insomnia, daytime consequences of the disorder are poorly characterized. This study aimed to identify neurobehavioral impairments associated with insomnia, and to investigate relationships between these impairments and subjective ratings of sleep and daytime dysfunction. Cross-sectional, multicenter study. Three sleep laboratories in the USA and Australia. Seventy-six individuals who met the Research Diagnostic Criteria (RDC) for Primary Insomnia, Psychophysiological Insomnia, Paradoxical Insomnia, and/or Idiopathic Childhood Insomnia (44F, 35.8 ± 12.0 years [mean ± SD]) and 20 healthy controls (14F, 34.8 ± 12.1 years). N/A. Participants completed a 7-day sleep-wake diary, questionnaires assessing daytime dysfunction, and a neurobehavioral test battery every 60-180 minutes during an afternoon/evening sleep laboratory visit. Included were tasks assessing sustained and switching attention, working memory, subjective sleepiness, and effort. Switching attention and working memory were significantly worse in insomnia patients than controls, while no differences were found for simple or complex sustained attention tasks. Poorer sustained attention in the control, but not the insomnia group, was significantly associated with increased subjective sleepiness. In insomnia patients, poorer sustained attention performance was associated with reduced health-related quality of life and increased insomnia severity. We found that insomnia patients exhibit deficits in higher level neurobehavioral functioning, but not in basic attention. The findings indicate that neurobehavioral deficits in insomnia are due to neurobiological alterations, rather than sleepiness resulting from chronic sleep deficiency.

  6. Adolescent Sleep and the Impact of Technology Use Before Sleep on Daytime Function

    PubMed Central

    Johansson, Ann E. E.; Petrisko, Maria A.; Chasens, Eileen R.

    2016-01-01

    Purpose Technology has become pervasive in our culture, particularly among adolescents. The purpose of this study is to examine associations between use of technology before sleep and daytime function in adolescents. Design and Methods This study is a secondary analysis of respondents aged 13 to 21 years (N= 259) from the 2011 National Sleep Foundation’s Sleep in America Poll. The survey included questions on demographics, sleep habits, and use of technology in the hour before bedtime. Daytime sleepiness was assessed with the Epworth Sleepiness Scale (ESS). Student’s t-tests, Mann-Whitney U, and Fischer’s exact tests were performed to detect differences in demographics, sleep duration, and technology use in the total sample, and between respondents with “adequate” compared to “inadequate” sleep. Correlations were calculated between technology frequency and daytime function. Results Adolescents had mean sleep duration of 7.3±1.3 hours. Almost all respondents (97%) used some form of technology before sleep. Increased technology use and the frequency of being awoken in the night by a cell phone were significantly associated with waking too early, waking unrefreshed, and daytime sleepiness (p<0.05). Adolescents who reported “inadequate” sleep had shorter sleep duration, greater frequency of technology use before bedtime, feeling unrefreshed on waking, and greater daytime sleepiness than those reporting “adequate” sleep (all p-values<0.05). Conclusion Technology use before sleep by adolescents had negative consequences on nighttime sleep and on daytime function. Practice Implications Healthcare professionals who interact with adolescents should encourage technology to be curtailed before bedtime and for adolescents to value obtaining adequate sleep. PMID:27184356

  7. Mood and neural correlates of excessive daytime sleepiness in Parkinson's disease.

    PubMed

    Wen, M-C; Chan, L L; Tan, L C S; Tan, E K

    2017-08-01

    For patients with Parkinson's disease (PD), excessive daytime sleepiness (PD-EDS) is a debilitating non-motor symptom and may be affected by mood symptoms, especially depression and anxiety. Few neuroimaging works have attempted to identify the neural features of PD-EDS, but various findings were reported. The purpose of this study was to systematically review the literature on mood and neuroimaging correlates of PD-EDS. A MEDLINE, PubMed, EMBASE, and PsycInfo search for peer-reviewed original research articles on depression, anxiety, and neuroimaging in PD-EDS identified 26 studies on depression, nine on anxiety, and eight on neuroimaging. Half of the studies reported greater depression in PD-EDS-positive patients compared with PD-EDS-negative patients. There was a significantly positive correlation between depression and PD-EDS. Limited studies on anxiety in PD-EDS suggested a weak correlation between anxiety and EDS. For depression and anxiety, the effect sizes were medium when EDS was subjectively measured, but became small when EDS was objective measured. Current neuroimaging studies generally suggested diminished neural structural and functional features (eg, brain volume, white matter integrity as indicated by fractional anisotropy, and cerebral metabolism) in patients with PD-EDS. Future studies should apply objective and subjective measures of mood symptoms and EDS and improve the neuroimaging methodology via using multimodal techniques and whole-brain analysis to provide new clues on the mood and neural correlates of PD-EDS. © 2016 John Wiley & Sons A/S. Published by John Wiley & Sons Ltd.

  8. Study of a patient population investigated for excessive daytime sleepiness (EDS).

    PubMed

    Laffont, F; Mallet, A; Mayer, G; Meunier, S; Minz, M; N'Doye, S; Quilfen-Buzare, M A

    2002-12-01

    This study included all patients referred to the out-patient department of our sleep disorders centre from 1993 to 1999 on account of excessive daytime sleepiness (EDS). As a first step, patients in whom a diagnosis was established following appropriate polysomnography were excluded: this included sleep apnea syndrome, increased upper airway resistance syndrome, narcolepsy, periodic movements during sleep or other parasomnia, and epilepsy. Patients regularly taking psychotropic substances or with psychiatric disorders were also excluded. Finally, 128 patients remained in whom no clear diagnosis had been established for EDS, 70 women and 58 men, their ages ranging from 16 to 77 years. They underwent a 48-h recording (night 1-MSLT-night 2-continuous day). The aim of the study was to establish, define and characterise different groups of undiagnosed EDS patients using clinical, electrophysiological and immunological data with the help of hierarchical cluster analysis. Eight groups were characterised: group 1: mild hypersomnia type 1 (n = 11); group 2: hypersomnia frequently associated with HLA type DR2-DQw1 (n = 11); group 3: mild hypersomnia type 2 (n = 28); group 4: morning recovery from disrupted sleep (n = 19); group 5: young "long sleepers with difficulty at waking up" (n = 17); group 6: idiopathic hypersomnia (n = 15); group 7: poor or short sleepers since childhood (n = 8); group 8: older poor sleepers with a late onset of symptoms (n = 19). Characteristic features of these different groups provided consistent and objective arguments leading to a more precise diagnosis for these patients, and helped the initiation of appropriate management and treatment.

  9. Perceived Stress and Coffee and Energy Drink Consumption Predict Poor Sleep Quality in Podiatric Medical Students A Cross-sectional Study.

    PubMed

    Sawah, Mohomad Al; Ruffin, Naeemah; Rimawi, Mohammad; Concerto, Carmen; Aguglia, Eugenio; Chusid, Eileen; Infortuna, Carmenrita; Battaglia, Fortunato

    2015-09-01

    A cross-sectional survey administered to first- and second-year podiatric medical students aimed to investigate the effect of coffee intake, energy drink consumption, and perceived stress on sleep quality in medical students during their preclinical studies. Ninety-eight of 183 students contacted (53.6%) completed a questionnaire comprising standard instruments measuring sleep quality (Pittsburgh Sleep Quality Index), daytime sleepiness (Epworth Sleepiness scale), and perceived stress (ten-item Perceived Stress Scale). Furthermore, we investigated coffee and energy drink consumption. Logistic regression was conducted to identify factors associated with poor sleep quality and the relation between sleep quality and academic performance (grade point average). High prevalences of poor sleep quality, excessive daytime sleepiness, and perceived stress were reported. In addition, higher odds of developing poor sleep quality were associated with coffee and energy drink intake, perceived stress, and excessive daytime sleepiness. The total Pittsburgh Sleep Quality Index score was inversely correlated with grade point average. First- and second-year podiatric medical students have poor sleep quality. Further research is needed to identify effective strategies to reduce stress and decrease coffee and energy drink intake to minimize their negative effect on sleep quality and academic performance in podiatric medical students.

  10. Individual vulnerability to insomnia, excessive sleepiness and shift work disorder amongst healthcare shift workers. A systematic review.

    PubMed

    Booker, Lauren A; Magee, Michelle; Rajaratnam, Shantha M W; Sletten, Tracey L; Howard, Mark E

    2018-03-27

    Shift workers often experience reduced sleep quality, duration and/or excessive sleepiness due to the imposed conflict between work and their circadian system. About 20-30% of shift workers experience prominent insomnia symptoms and excessive daytime sleepiness consistent with the circadian rhythm sleep disorder known as shift work disorder. Individual factors may influence this vulnerability to shift work disorder or sleep-related impairment associated with shift work. This paper was registered with Prospero and was conducted using recommended standards for systematic reviews and meta-analyses. Published literature that measured sleep-related impairment associated with shift work including reduced sleep quality and duration and increased daytime sleepiness amongst healthcare shift workers and explored characteristics associated with individual variability were reviewed. Fifty-eight studies were included. Older age, morning-type, circadian flexibility, being married or having children, increased caffeine intake, higher scores on neuroticism and lower on hardiness were related to a higher risk of sleep-related impairment in response to shift work, whereas physical activity was a protective factor. The review highlights the diverse range of measurement tools used to evaluate the impact of shift work on sleep. Use of standardised and validated tools would enable cross-study comparisons. Longitudinal studies are required to establish causal relationships between individual factors and the development of shift work disorder. Copyright © 2018 Elsevier Ltd. All rights reserved.

  11. Sleep quality and daytime function in adults with cystic fibrosis and severe lung disease.

    PubMed

    Dancey, D R; Tullis, E D; Heslegrave, R; Thornley, K; Hanly, P J

    2002-03-01

    It was hypothesized that adult cystic fibrosis (CF) patients with severe lung disease have impaired daytime function related to nocturnal hypoxaemia and sleep disruption. Nineteen CF patients (forced expiratory volume in one second 28+/-7% predicted) and 10 healthy subjects completed sleep diaries, overnight polysomnography (PSG), and assessment of daytime sleepiness and neurocognitive function. CF patients tended to report more awakenings (0.7+/-0.5 versus 0.3+/-0.2 x h(-1), p=0.08), and PSG revealed reduced sleep efficiency (71+/-25 versus 93+/-4%, p=0.004) and a higher frequency of awakenings (4.2+/-2.7 versus 2.4+/-1.4 x h(-1), p=0.06). Mean arterial oxygen saturation during sleep was lower in CF patients (84.4+/-6.8 versus 94.3+/-1.5%, p<0.0001) and was associated with reduced sleep efficiency (regression coefficient (r)=0.57, p=0.014). CF patients had short sleep latency on the multiple sleep latency test (6.7+/-3 min). The CF group reported lower levels of activation and happiness and greater levels of fatigue (p<0.01), which correlated with indices of sleep loss, such as sleep efficiency (r=0.47, p=10.05). Objective neurocognitive performance was also impaired in CF patients, reflected by lower throughput for simple addition/subtraction, serial reaction and colour-word conflict. The authors concluded that adult cystic fibrosis patients with severe lung disease have impaired neurocognitive function and daytime sleepiness, which is partly related to chronic sleep loss and nocturnal hypoxaemia.

  12. Neurobehavioral Performance Impairment in Insomnia: Relationships with Self-Reported Sleep and Daytime Functioning

    PubMed Central

    Shekleton, Julia A.; Flynn-Evans, Erin E.; Miller, Belinda; Epstein, Lawrence J.; Kirsch, Douglas; Brogna, Lauren A.; Burke, Liza M.; Bremer, Erin; Murray, Jade M.; Gehrman, Philip; Lockley, Steven W.; Rajaratnam, Shantha M. W.

    2014-01-01

    Study Objectives: Despite the high prevalence of insomnia, daytime consequences of the disorder are poorly characterized. This study aimed to identify neurobehavioral impairments associated with insomnia, and to investigate relationships between these impairments and subjective ratings of sleep and daytime dysfunction. Design: Cross-sectional, multicenter study. Setting: Three sleep laboratories in the USA and Australia. Patients: Seventy-six individuals who met the Research Diagnostic Criteria (RDC) for Primary Insomnia, Psychophysiological Insomnia, Paradoxical Insomnia, and/or Idiopathic Childhood Insomnia (44F, 35.8 ± 12.0 years [mean ± SD]) and 20 healthy controls (14F, 34.8 ± 12.1 years). Interventions: N/A. Measurements and Results: Participants completed a 7-day sleep-wake diary, questionnaires assessing daytime dysfunction, and a neurobehavioral test battery every 60-180 minutes during an afternoon/evening sleep laboratory visit. Included were tasks assessing sustained and switching attention, working memory, subjective sleepiness, and effort. Switching attention and working memory were significantly worse in insomnia patients than controls, while no differences were found for simple or complex sustained attention tasks. Poorer sustained attention in the control, but not the insomnia group, was significantly associated with increased subjective sleepiness. In insomnia patients, poorer sustained attention performance was associated with reduced health-related quality of life and increased insomnia severity. Conclusions: We found that insomnia patients exhibit deficits in higher level neurobehavioral functioning, but not in basic attention. The findings indicate that neurobehavioral deficits in insomnia are due to neurobiological alterations, rather than sleepiness resulting from chronic sleep deficiency. Citation: Shekleton JA; Flynn-Evans EE; Miller B; Epstein LJ; Kirsch D; Brogna LA; Burke LM; Cremer E; Murray JM; Gehrman P; Lockley SW; Rajaratnam SMW

  13. The effect of daylight versus darkness on driver sleepiness: a driving simulator study.

    PubMed

    Ahlström, Christer; Anund, Anna; Fors, Carina; Åkerstedt, Torbjörn

    2018-06-01

    Driver sleepiness studies are often carried out with alert drivers during daytime and sleep-deprived drivers during night-time. This design results in a mixture of different factors (e.g. circadian effects, homeostatic effects, light conditions) that may confound the results. The aim of this study was to investigate the effect of light conditions on driver sleepiness. Thirty young male drivers (23.6 ± 1.7 years old) participated in a driving simulator experiment where they drove on a rural road. A 2 × 2 design was used with the conditions daylight versus darkness, and daytime (full sleep) versus night-time (sleep deprived). The results show that light condition had an independent effect on the sleepiness variables. The subjective sleepiness measured by Karolinska Sleepiness Scale was higher, lateral position more left-oriented, speed lower, electroencephalogram alpha and theta higher, and blink durations were longer during darkness. The number of line crossings did not change significantly with light condition. The day/night condition had profound effects on most sleepiness indicators while controlling for light condition. The number of line crossings was higher during night driving, Karolinska Sleepiness Scale was higher, blink durations were longer and speed was lower. There were no significant interactions, indicating that light conditions have an additive effect on sleepiness. In conclusion, Karolinska Sleepiness Scale and blink durations increase primarily with sleep deprivation, but also as an effect of darkness. Line crossings are mainly driven by the need for sleep and the reduced alertness at the circadian nadir. Lane position is, however, more determined by light conditions than by sleepiness. © 2017 European Sleep Research Society.

  14. Effects of a Single Night of Postpartum Sleep on Childless Women’s Daytime Functioning

    PubMed Central

    McBean, Amanda L.; Kinsey, Steven G.; Montgomery-Downs, Hawley E.

    2017-01-01

    Study Objectives The maternal postpartum period is characterized by sleep fragmentation, which is associated with daytime impairment, mental health disturbances, and changes in melatonin patterns. In addition to sleep fragmentation, women undergo a complex set of physiological and environmental changes upon entering the postpartum period, confounding our understanding of effects of postpartum sleep disturbance. The primary study aim was to understand the basic impact of a single night of postpartum-like sleep fragmentation on sleep architecture, nocturnal melatonin levels, mood, daytime sleepiness, and neurobehavioral performance. Measurements and Results For one week prior to entry into the laboratory, eleven healthy nulliparous women kept a stable sleep-wake schedule (verified via actigraphy). Participants contributed three consecutive nights of laboratory overnight polysomnography: (1) a habituation/sleep disorder screening night; (2) a baseline night; and (3) a sleep fragmentation night, when participants were awakened three times for ~30 min each. Self-reported sleep quality and mood (Profile of Mood States Survey) both decreased significantly after sleep fragmentation compared to baseline measurements. Unexpectedly, daytime sleepiness (Multiple Sleep Latency Test) decreased significantly after sleep fragmentation. Experimental fragmentation had no significant effect on time spent in nocturnal sleep stages, urinary 6-sulphatoxymelatonin concentration, or psychomotor vigilance test performance. Participants continued to provide actigraphy data, and daily PVTs and self-reported sleep quality assessments at home for one week following sleep fragmentation; these assessments did not differ from baseline values. Conclusions While there were no changes in measured physiological components of a single night of postpartum-like experimental sleep fragmentation, there were decreases in self-reported measures of mood and sleep quality. Future research should examine the

  15. Subjective sleepiness in heart failure patients with sleep-related breathing disorder.

    PubMed

    Wang, Han-Qiao; Chen, Gang; Li, Jing; Hao, Shu-Min; Gu, Xin-Shun; Pang, Jiang-Na; Fu, Xiang-Hua

    2009-06-20

    Previous studies show that sleep-related breathing disorder (SRBD) is common in patients with heart failure (HF) and is associated with increased mortality. This study aimed to determine whether there was significant difference of subjective daytime sleepiness between HF patients with and without SRBD. We enrolled, prospectively, 195 consecutive HF patients with left ventricular ejection fractions (LVEF) < or = 45% and all subjects underwent polysomnography to measure the sleep structure between 2005 and 2008. Patients were then assigned to those with SRBD including obstructive and central sleep apnea (apnea-hypopnea index (AHI) > or = 5/hour of sleep) and those without SRBD (AHI < 5/hour) according to the sleep study. The subjective sleepiness was assessed with Epworth sleepiness scale (ESS). Among 195 HF patients, the prevalence of obstructive sleep apnea (OSA) was 53% and of central sleep apnea (CSA) was 27%. There was no significant difference of ESS scores between patients without SRBD (NSA) and with SRBD (NSA vs OSA: 6.7 +/- 0.6 vs 7.6 +/- 0.4, P = 0.105 and NSA vs CSA: 6.7 +/- 0.6 vs 7.4 +/- 0.5, P = 0.235, respectively), indicating that SRBD patients had no more subjective daytime sleepiness. Compared with NSA, patients with SRBD had increased arousal index (ArI) (NSA vs OSA: 14.1 +/- 1.4 vs 26.3 +/- 1.5, P < 0.001 and NSA vs CSA: 14.1 +/- 1.4 vs 31.3 +/- 3.5, P < 0.001, respectively), more awake number after sleep onset (NSA vs OSA: 19.2 +/- 1.5 vs 26.2 +/- 1.4, P = 0.01 and NSA vs CSA: 19.2 +/- 1.5 vs 36.9 +/- 4.4, P < 0.001, respectively), and reduced proportion of slow-wave sleep (SWS) (NSA vs OSA: 13.8 +/- 1.7 vs 9.3 +/- 0.7, P = 0.024 and NSA vs CSA: 13.8 +/- 1.7 vs 8.9 +/- 0.9, P = 0.024, respectively). OSA and CSA remain common in patients with HF on optimal contemporary therapy. Patients with both HF and SRBD have no significant subjective daytime sleepiness compared with patients without SRBD, despite of significantly increased awake number

  16. 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

  17. Daytime Napping, Nighttime Sleeping, and Parkinson Disease

    PubMed Central

    Gao, Jianjun; Huang, Xuemei; Park, Yikyung; Hollenbeck, Albert; Blair, Aaron; Schatzkin, Arthur; Chen, Honglei

    2011-01-01

    Preliminary evidence suggests that daytime sleepiness may predate clinical diagnosis of Parkinson disease. The authors examined daytime napping and nighttime sleeping durations, reported in 1996–1997 by 220,934 US NIH-AARP Diet and Health Study participants, in relation to Parkinson disease diagnoses at 3 clinical stages: established (cases diagnosed before 1995, n = 267), recent (1995–1999, n = 396), and prediagnostic (2000 and after, n = 770). Odds ratios and 95% confidence intervals were derived from multivariate logistic regression models. Longer daytime napping was associated with higher odds of Parkinson disease at all 3 clinical stages: the odds ratios comparing long nappers (>1 hour/day) with nonnappers were 3.9 (95% confidence interval: 2.8, 5.6) for established cases, 2.2 (95% confidence interval: 1.7, 3.0) for recent cases, and 1.5 (95% confidence interval: 1.2, 1.9) for prediagnostic cases. Further control for health status or nighttime sleeping duration attenuated the association for established cases but made little difference for recent or prediagnostic cases. In the nighttime sleeping analysis, a clear U-shaped association with Parkinson disease was observed for established cases; however, this association was attenuated markedly for recent cases and disappeared for prediagnostic cases. This study supports the notion that daytime sleepiness, but not nighttime sleeping duration, is one of the early nonmotor symptoms of Parkinson disease. PMID:21402730

  18. A spiritual sleepiness scale: the Friday prayer.

    PubMed

    Abouda, Maher; Turki, Senda; Hachicha, Amani; Yangui, Ferdaous; Triki, Miriam; Charfi, Med Ridha

    2016-03-01

    Excessive daytime sleepiness (EDS) affects 5% to 20% of the population and is involved in a large number of traffic accidents. EDS is a major symptom in sleep disorders, especially obstructive sleep apnea syndrome (OSA). The daytime sleepiness is evaluated subjectively using scales and questionnaires based on perception. This study is aimed to build a new questionnaire more suited to our lifestyle and then to compare it to the Epworth sleepiness scales (ESS). We administered to 91 adult's patients (76 men and 15 women) consulting for sleep disturbance the ESS and a single subjective question tendency to drowsiness during the Friday prayer. Patients were listed in four groups according to their response to the question «During the past month, have you ever doze or fall asleep during the sermon of the Friday prayer? » By G1 never dozes, G2 low chance of falling asleep, G3 average chance of dozing, G4 high chance of falling asleep. Only 63 patients (58 men and 5 women) responded to both questionnaires. Group 1 included 14 patients with a ESS of 5.5 +/- 1.8, Group 2 included 18 patients with a ESS of 7.3 +/- 1.9, Group 3 included 18 patients with a ESS of 11.05 +/- 2 and Group 4 included 13 patients with a ESS of 14.69 +/- 2.3. The Rho correlation coefficient was high (0.86) and shows a strong correlation between the Results of the two questionnaires. the answer to the question " During the past month, have you ever doze or fall asleep during the sermon of the Friday prayer? » seems to be an appropriate Sleepiness Scale among Muslim patients.

  19. [Sleep disorders in Parkinson's disease: insomnia and sleep fragmentation, daytime hypersomnia, alterations to the circadian rhythm and sleep apnea syndrome].

    PubMed

    Mondragón-Rezola, E; Arratíbel-Echarren, I; Ruiz-Martínez, J; Martí-Massó, J F

    2010-02-08

    Sleep disorders in Parkinson's disease are present in 60-98% of patients and reduce their quality of life. To review the pathophysiology, diagnostic approach and management of the different sleep disorders. We describe the pathophysiology associated with neurodegeneration, due to symptoms (motor and nonmotor) and drug therapies. This article reviews insomnia, excessive daytime sleepiness, circadian sleep disorders and sleep apnea. Subjective or objective sleepiness assessment should routinely be performed by physicians looking after Parkinson's disease patients. Management is difficult and should be targeted to the specific sleep disorder and its likely cause.

  20. Night sleep electroencephalogram power spectral analysis in excessive daytime sleepiness disorders.

    PubMed

    Reimão, R

    1991-06-01

    A group of 53 patients (40 males, 13 females) with mean age of 49 years, ranging from 30 to 70 years, was evaluated in the following excessive daytime sleepiness (EDS) disorders: obstructive sleep apnea syndrome (B4a), periodic movements in sleep (B5a), affective disorder (B2a), functional psychiatric non affective disorder (B2b). We considered all adult patients referred to the Center sequentially with no other distinctions but these three criteria: (a) EDS was the main complaint; (b) right handed; (c) not using psychotropic drugs for two weeks prior to the all-night polysomnography. EEG (C3/A1, C4/A2) samples from 2 to 10 minutes of each stage of the first REM cycle were chosen. The data was recorded simultaneously in magnetic tape and then fed into a computer for power spectral analysis. The percentage of power (PP) in each band calculated in relation to the total EEG power was determined of subsequent sections of 20.4 s for the following frequency bands: delta, theta, alpha and beta. The PP in all EDS patients sample had a tendency to decrease progressively from the slowest to the fastest frequency bands, in every sleep stage. PP distribution in the delta range increased progressively from stage 1 to stage 4; stage REM levels were close to stage 2 levels. In an EDS patients interhemispheric coherence was high in every band and sleep stage. B4a patients sample PP had a tendency to decrease progressively from the slowest to the fastest frequency bands, in every sleep stage; PP distribution in the delta range increased progressively from stage 1 to stage 4; stage REM levels were between stage 1 and stage 2 levels.(ABSTRACT TRUNCATED AT 250 WORDS)

  1. Perceived sleepiness of non-shift working men in two different types of work organization.

    PubMed

    Fukasawa, Kenji; Aikawa, Hiroyuki; Okazaki, Isao; Haratani, Takashi; Takahashi, Masaya; Nakata, Akinori; Otsuka, Yasumasa; Kaida, Kosuke; Hanada, Takanobu

    2006-07-01

    Increased sleepiness at work is increasingly being focused on as a safety and health issue. However, research on workers' sleepiness is very limited in scope and the characteristics of work organization, including the impact of job stress, have not been fully addressed. A questionnaire survey was conducted to investigate the prevalence of daytime sleepiness and its associated factors among non-shift working men at two manufacturing businesses: Company A, having a rapid rate of development and growth, with 564 workers (19-61 yr old, mean age: 32.7, response rate: 81.4%); and Company B, long established, possessing a huge production facility, with 1,654 workers (20-63 yr old, mean age: 37.1, response rate: 78.2%). The prevalence of daytime sleepiness was 11.3% in company A and 16.8% in company B. Multivariate logistic regression analysis revealed that, in company A, perceived sleepiness was associated with long sleep duration on non-working days and high cognitive demands and, in company B, with insufficient daily sleep, single, and depression. Psychosomatic exhaustion resulting from jobs requiring high adaptivity due to rapid frequency of operational change as in company A may have the potential to become an important factor in perceived sleepiness. However, in a comparatively stable work organization, as in company B, increased sleepiness may be mainly linked to factors outside work. It is suggested that not only lifestyle and sleep habits, but also the characteristics and dynamics of a work organization should be a focus of attention when planning measures to prevent sleepiness at work.

  2. The Relation Between Use of Mobile Electronic Devices and Bedtime Resistance, Sleep Duration, and Daytime Sleepiness Among Preschoolers.

    PubMed

    Nathanson, Amy I; Beyens, Ine

    2018-01-01

    This study investigated the relation between preschoolers' mobile electronic device (MED) use and sleep disturbances. A national sample of 402 predominantly college-educated and Caucasian mothers of 3-5-year-olds completed a survey assessing their preschoolers' MED use, bedtime resistance, sleep duration, and daytime sleepiness. Heavier evening and daily tablet use (and to some extent, smartphone use) were related to sleep disturbances. Other forms of MED use were not consistently related to sleep disturbances. In addition, playing games on MEDs at bedtime was related to compromised sleep duration, although other forms of MED use at bedtime were not related to sleep outcomes. Although the relations between MED use and sleep disturbances were small in size, they were larger than the relations between sleep and other predictors in the models. Continued work should investigate how MED exposure is related to children's cognitive, psychological, emotional, and physiological development, particularly given the popularity and widespread use of these devices.

  3. 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.

  4. The effect of optokinetic stimulation on daytime sleepiness

    NASA Technical Reports Server (NTRS)

    Leslie, K. R.; Stickgold, R.; Dizio, P.; Lackner, J. R.; Hobson, J. A.

    1997-01-01

    This study examined the effect of optokinetic stimulation on objective sleepiness, as measured by the Multiple Sleep Latency Test (MSLT). The Nightcap, a portable sleep monitor, was used in a novel way to perform MSLTs, as well as record sleep in the home. Subjects wore the Nightcap for seven consecutive nights. On days 3 and 5 of the protocol, subjects came into the lab for an MSLT. On the experimental day, subjects underwent 10 minutes optokinetic stimulation (OKS), resulting in moderate motion sickness prior to each MSLT trial. Although subjects in the OKS condition reported significantly more drowsiness than controls, this did not result in significantly reduced sleep latencies.

  5. Improvement in daytime sleepiness with clarithromycin in patients with GABA-related hypersomnia: Clinical experience.

    PubMed

    Trotti, Lynn Marie; Saini, Prabhjyot; Freeman, Amanda A; Bliwise, Donald L; García, Paul S; Jenkins, Andrew; Rye, David B

    2014-07-01

    The macrolide antibiotic clarithromycin can enhance central nervous system excitability, possibly by antagonism of GABA-A receptors. Enhancement of GABA signaling has recently been demonstrated in a significant proportion of patients with central nervous system hypersomnias, so we sought to determine whether clarithromycin might provide symptomatic benefit in these patients. We performed a retrospective review of all patients treated with clarithromycin for hypersomnia, in whom cerebrospinal fluid enhanced GABA-A receptor activity in vitro in excess of controls, excluding those with hypocretin deficiency or definite cataplexy. Subjective reports of benefit and objective measures of psychomotor vigilance were collected to assess clarithromycin's effects. Clinical and demographic characteristics were compared in responders and non-responders. In total, 53 patients (38 women, mean age 35.2 (SD 12.8 years)) were prescribed clarithromycin. Of these, 34 (64%) reported improvement in daytime sleepiness, while 10 (19%) did not tolerate its side effects, and nine (17%) found it tolerable but without symptomatic benefit. In those who reported subjective benefit, objective corroboration of improved vigilance was evident on the psychomotor vigilance task. Twenty patients (38%) elected to continue clarithromycin therapy. Clarithromycin responders were significantly younger than non-responders. Clarithromycin may be useful in the treatment of hypersomnia associated with enhancement of GABA-A receptor function. Further evaluation of this novel therapy is needed. © The Author(s) 2013.

  6. Objective and quantitative analysis of daytime sleepiness in physicians after night duties.

    PubMed

    Wilhelm, Barbara J; Widmann, Anja; Durst, Wilhelm; Heine, Christian; Otto, Gerhard

    2009-06-01

    Work place studies often have the disadvantage of lacking objective data less prone to subject bias. The aim of this study was to contribute objective data to the discussion about safety aspects of night shifts in physicians. For this purpose we applied the Pupillographic Sleepiness Test (PST). The PST allows recording and analyses of pupillary sleepiness-related oscillations in darkness for 11 min in the sitting subject. The parameter of evaluation is the Pupillary Unrest Index (PUI; mm/min). For statistical analysis the natural logarithm of this parameter is used (lnPUI). Thirty-four physicians were examined by the PST and subjective scales during the first half of the day. Data taken during a day work period (D) were compared to those taken directly after night duty (N) by a Wilcoxon signed rank test. Night duty caused a mean sleep reduction of 3 h (Difference N-D: median 3 h, minimum 0 h, maximum 7 h, p < 0.001). Time since the last sleep period was about equal in both conditions (Difference N-D: median -0.25 h, min. -4 h, max. 20 h, p = 0.2). The lnPUI was larger after night duty (Difference N-D: median 0.19, min. -0.71, max. 1.29, p = 0.03). The increase of physiologically measured sleepiness correlated significantly with changes in subjective measures (PUI/SSS, Spearman Rho 0.41, p = 0.02; PUI/VAS, Spearman Rho 0.38, p = 0.02). Despite a mean sleep duration of 4 h, considerable sleepiness in physicians after nights on duty was found, implying lower safety levels for both patients (if physicians remaining on duty) and physicians while commuting home.

  7. Excessive sleepiness in adolescents and young adults: causes, consequences, and treatment strategies.

    PubMed

    Millman, Richard P

    2005-06-01

    Adolescents and young adults are often excessively sleepy. This excessive sleepiness can have a profound negative effect on school performance, cognitive function, and mood and has been associated with other serious consequences such as increased incidence of automobile crashes. In this article we review available scientific knowledge about normal sleep changes in adolescents (13-22 years of age), the factors associated with chronic insufficient sleep, the effect of insufficient sleep on a variety of systems and functions, and the primary sleep disorders or organic dysfunctions that, if untreated, can cause excessive daytime sleepiness in this population.

  8. Pupil miosis within 5 minutes in darkness is a valid and sensitive quantitative measure of alertness: application in daytime sleepiness associated with sleep apnea.

    PubMed

    Bitsios, Panos; Schiza, Sophia E; Giakoumaki, Stella G; Savidou, Kyriaki; Alegakis, Athanasios K; Siafakas, Nikolaos

    2006-11-01

    The regulation of arousal and pupillary functions may be intimately linked via activity in the nucleus locus coeruleus. In this preliminary study, we tested the validity of the gradual pupillary miosis during 5 minutes in darkness, as a quantitative physiologic index of the arousal state of the brain. Cross-sectional assessment of 2 groups with between-group comparison and correlational analyses within the patient group. Eleven unmedicated male patients recently diagnosed with obstructive sleep apnea (OSA) with no comorbid conditions who had undergone polysomnography to assess OSA severity and sleep variables, and 11 sex- and age-matched healthy controls. Sampling of the resting pupil diameter (RPD) over 5 minutes in darkness in the morning and in the afternoon hours, using an infrared video pupillometer. The RPD was smaller, indicating a lower level of arousal, in the patient group compared with controls in both the morning and the afternoon; the RPD showed a significant circadian reduction in the afternoon only in the patient group. Within the patient group, the RPD correlated negatively with Epworth Sleepiness Scale scores and Arousal Index and positively with the lowest oxygen saturation during the night. Controlling for the effect of body mass index, the relationship between RPD and subjective sleepiness was lost, whereas the relationship with most of the objective indexes of OSA severity was improved. The 5-minute pupillary miosis in darkness holds promise as a simple, fast-to-administer, valid, and sensitive test for the objective assessment of excessive daytime sleepiness.

  9. Association of Calf Muscle Pump Stimulation With Sleep Quality in Adults.

    PubMed

    Baniak, Lynn M; Pierce, Carolyn S; McLeod, Kenneth J; Chasens, Eileen R

    2016-12-01

    Prevention of lower extremity fluid pooling (LEFP) is associated with improved sleep quality. Physical activity and compression stockings are non-invasive methods used to manage LEFP, but both are associated with low adherence. Calf muscle pump (CMP) stimulation is an alternative and more convenient approach. Convenience sampling was used to recruit 11 participants between ages 45 and 65 with poor sleep quality. A within-person single-group pre-test-post-test design was used to evaluate changes in sleep quality, daytime sleepiness, and functional outcomes sensitive to impaired sleep as measured by the Pittsburgh Sleep Quality Index (PSQI), Functional Outcomes of Sleep Questionnaire, and Epworth Sleepiness Scale after 4 weeks of CMP stimulation. Statistical analysis included effect size (ES) calculations. After daily use of CMP stimulation, participants demonstrated improvement in overall sleep quality (ES = -.97) and a large reduction in daily disturbance from poor sleep (ES = -1.25). Moderate improvements were observed in daytime sleepiness (ES = -.53) and functional outcomes sensitive to sleepiness (ES = .49). Although causality could not be determined with this study design, these results support further research to determine whether CMP stimulation can improve sleep quality. © 2016 Wiley Periodicals, Inc. © 2016 Wiley Periodicals, Inc.

  10. Excessive daytime sleepiness and falls among older men and women: cross-sectional examination of a population-based sample.

    PubMed

    Hayley, Amie C; Williams, Lana J; Kennedy, Gerard A; Holloway, Kara L; Berk, Michael; Brennan-Olsen, Sharon L; Pasco, Julie A

    2015-07-05

    Excessive daytime sleepiness (EDS) has been associated with an increased risk for falls among clinical samples of older adults. However, there is little detailed information among population-representative samples. The current study aimed to assess the relationship between EDS and falls among a cohort of population-based older adults. This study assessed 367 women aged 60-93 years (median 72, interquartile range 65-79) and 451 men aged 60-92 years (median 73, interquartile range 66-80) who participated in the Geelong Osteoporosis Study between the years 2001 and 2008. Falls during the prior year were documented via self-report, and for men, falls risk score was obtained using an Elderly Fall Screening Test (EFST). Sleepiness was assessed using the Epworth Sleepiness Scale (ESS), and scores of  ≥ 10 indicated EDS. Differences among those with and without EDS in regard to falls were tested using logistic regression models. Among women, 50 (13.6%) individuals reported EDS. Women with EDS were more likely to report a fall, and were more likely to report the fall occurring outside. EDS was similarly associated with an increased risk of a fall following adjustment for use of a walking aid, cases of nocturia and antidepressant medication use (adjusted OR = 2.54, 95% CI 1.24-5.21). Multivariate modelling revealed antidepressant use (current) as an effect modifier (p < .001 for the interaction term). After stratifying the data by antidepressant medication use, the association between EDS and falls was sustained following adjustment for nocturia among antidepressant non-users (adjusted OR = 2.63, 95% CI 1.31-5.30). Among men, 72 (16.0%) individuals reported EDS. No differences were detected for men with and without EDS in regard to reported falls, and a trend towards significance was noted between EDS and a high falls risk as assessed by the EFST (p = 0.06), however, age explained this relationship (age adjusted OR = 2.20, 95% CI 1.03-1.10). For

  11. The effects of armodafinil on objective sleepiness and performance in a shift work disorder sample unselected for objective sleepiness.

    PubMed

    Howard, Ryan; Roth, Thomas; Drake, Christopher L

    2014-06-01

    Armodafinil is a medication used to treat excessive sleepiness in individuals with shift work disorder (SWD). In the present study, we investigate whether armodafinil can normalize nocturnal sleepiness in a group of typical SWD patients. Participants were 12 night workers (aged 33.8 ± 8.57 years, 7 female subjects) with excessive sleepiness (≥10 on the Epworth Sleepiness Scale; mean, 14.8 ± 3.16), meeting the International Classification of Sleep Disorders, Second Edition criteria for SWD, with no other sleep or medical disorders verified by polysomnogram. The multiple sleep latency test (MSLT) was not used as an entry criteria. Armodafinil was administered at 10:30 pm in a randomized, double-blind, placebo-controlled, crossover design with experimental nights separated by 1 week. Primary end point was the MSLT, with naps at 1:30, 3:30, 5:30, and 7:30 am. Other study measures included a sleepiness-alertness visual analog scale administered before each nap, and 2 computer-based performance tests evaluating attention and memory. Subjects with SWD had a mean MSLT of 5.3 ± 3.25 minutes, indicating a mean level of pathological sleepiness. Armodafinil significantly improved MSLT score to 11.1 ± 4.79 minutes (P = 0.006). Subjective levels of alertness on the visual analog scale also improved (P = 0.008). For performance, reaction time to central (P = 0.006) and peripheral (P = 0.003) stimuli and free recall memory (P = 0.05) were also improved. Armodafinil 150 mg administered at the beginning of a night shift normalizes nocturnal sleepiness in individuals with SWD unselected for objective sleepiness. Subjective measures of sleepiness and cognitive performance are also improved. This suggests that armodafinil can improve levels of nocturnal alertness to within normal daytime levels in the majority of patients with SWD.

  12. Mind wandering, sleep quality, affect and chronotype: an exploratory study.

    PubMed

    Carciofo, Richard; Du, Feng; Song, Nan; Zhang, Kan

    2014-01-01

    Poor sleep quality impairs cognition, including executive functions and concentration, but there has been little direct research on the relationships between sleep quality and mind wandering or daydreaming. Evening chronotype is associated with poor sleep quality, more mind wandering and more daydreaming; negative affect is also a mutual correlate. This exploratory study investigated how mind wandering and daydreaming are related to different aspects of sleep quality, and whether sleep quality influences the relationships between mind wandering/daydreaming and negative affect, and mind wandering/daydreaming and chronotype. Three surveys (Ns = 213; 190; 270) were completed with Chinese adults aged 18-50, including measures of sleep quality, daytime sleepiness, mind wandering, daydreaming, chronotype and affect (positive and negative). Higher frequencies of mind wandering and daydreaming were associated with poorer sleep quality, in particular with poor subjective sleep quality and increased sleep latency, night-time disturbance, daytime dysfunction and daytime sleepiness. Poor sleep quality was found to partially mediate the relationships between daydreaming and negative affect, and mind wandering and negative affect. Additionally, low positive affect and poor sleep quality, in conjunction, fully mediated the relationships between chronotype and mind wandering, and chronotype and daydreaming. The relationships between mind wandering/daydreaming and positive affect were also moderated by chronotype, being weaker in those with a morning preference. Finally, while daytime sleepiness was positively correlated with daydream frequency, it was negatively correlated with a measure of problem-solving daydreams, indicating that more refined distinctions between different forms of daydreaming or mind wandering are warranted. Overall, the evidence is suggestive of a bi-directional relationship between poor sleep quality and mind wandering/daydreaming, which may be important in

  13. Mind Wandering, Sleep Quality, Affect and Chronotype: An Exploratory Study

    PubMed Central

    Carciofo, Richard; Du, Feng; Song, Nan; Zhang, Kan

    2014-01-01

    Poor sleep quality impairs cognition, including executive functions and concentration, but there has been little direct research on the relationships between sleep quality and mind wandering or daydreaming. Evening chronotype is associated with poor sleep quality, more mind wandering and more daydreaming; negative affect is also a mutual correlate. This exploratory study investigated how mind wandering and daydreaming are related to different aspects of sleep quality, and whether sleep quality influences the relationships between mind wandering/daydreaming and negative affect, and mind wandering/daydreaming and chronotype. Three surveys (Ns = 213; 190; 270) were completed with Chinese adults aged 18–50, including measures of sleep quality, daytime sleepiness, mind wandering, daydreaming, chronotype and affect (positive and negative). Higher frequencies of mind wandering and daydreaming were associated with poorer sleep quality, in particular with poor subjective sleep quality and increased sleep latency, night-time disturbance, daytime dysfunction and daytime sleepiness. Poor sleep quality was found to partially mediate the relationships between daydreaming and negative affect, and mind wandering and negative affect. Additionally, low positive affect and poor sleep quality, in conjunction, fully mediated the relationships between chronotype and mind wandering, and chronotype and daydreaming. The relationships between mind wandering/daydreaming and positive affect were also moderated by chronotype, being weaker in those with a morning preference. Finally, while daytime sleepiness was positively correlated with daydream frequency, it was negatively correlated with a measure of problem-solving daydreams, indicating that more refined distinctions between different forms of daydreaming or mind wandering are warranted. Overall, the evidence is suggestive of a bi-directional relationship between poor sleep quality and mind wandering/daydreaming, which may be

  14. The impact of obstructive sleep apnea and daytime sleepiness on work limitation.

    PubMed

    Mulgrew, A T; Ryan, C F; Fleetham, J A; Cheema, R; Fox, N; Koehoorn, M; Fitzgerald, J M; Marra, C; Ayas, N T

    2007-12-01

    Many patients with obstructive sleep apnea (OSA) participate in the work force. However, the impact of OSA and sleepiness on work performance is unclear. To address this issue, we administered the Epworth Sleepiness Scale (ESS), the Work Limitations Questionnaire (WLQ), and an occupational survey to patients undergoing full-night polysomnography for the investigation of sleep-disordered breathing. Of 498 patients enrolled in the study, 428 (86.0%) completed the questionnaires. Their mean age+/-standard deviation (SD) was 49+/-12 years, mean body mass index (BMI) was 31+/-7 kg/m(2) mean apnea hypopnea index (AHI) was 21+/-22 events/h, and mean ESS score was 10+/-5. Subjects worked a mean of 39+/-18 h per week. The first 100 patients to complete the survey were followed up at two years. In the group as a whole, there was no significant relationship between severity of OSA and the four dimensions of work limitation. However, in blue-collar workers, significant differences were detected between patients with mild OSA (AHI 5-15/h) and those with severe OSA (AHI>30/h) with respect to time management (limited 23.1% of the time vs. 43.8%, p=0.05) and mental/personnel interactions (17.9% vs. 33.0%, p=0.05). In contrast, there were strong associations between subjective sleepiness (as assessed by the ESS) and three of the four scales of work limitation. That is, patients with an ESS of 5 had much less work limitation compared to those with an ESS 18 in terms of time management (19.7% vs. 38.6 %, p<0.001), mental-interpersonal relationships (15.5% vs. 36.0%, p<0.001) and work output (16.8% vs. 36.0%; p<0.001). Of the group followed up, 49 returned surveys and 33 who were using continuous positive airway pressure (CPAP) showed significant improvements between the initial and second follow-up in time management (26% vs. 9%, p=0.0005), mental-interpersonal relationships (16% vs. 11.0%, p=0.014) and work output (18% vs. 10%; p<0.009). We have demonstrated a clear relationship

  15. Restless leg syndrome, sleep quality and fatigue in multiple sclerosis patients.

    PubMed

    Moreira, N C V; Damasceno, R S; Medeiros, C A M; Bruin, P F C de; Teixeira, C A C; Horta, W G; Bruin, V M S de

    2008-10-01

    We have tested the hypothesis that restless leg syndrome (RLS) is related to quality of sleep, fatigue and clinical disability in multiple sclerosis (MS). The diagnosis of RLS used the four minimum criteria defined by the International Restless Legs Syndrome Study Group. Fatigue was assessed by the Fatigue Severity Scale (FSS >27), quality of sleep by the Pittsburgh Sleep Quality Index (PSQI >6), excessive daytime sleepiness by the Epworth Sleepiness Scale (ESS >10) and clinical disability by the Expanded Disability Status Scale (EDSS). Forty-four patients (32 women) aged 14 to 64 years (43 +/- 14) with disease from 0.4 to 23 years (6.7 +/- 5.9) were evaluated. Thirty-five were classified as relapsing-remitting, 5 as primary progressive and 4 as secondary progressive. EDSS varied from 0 to 8.0 (3.6 +/- 2.0). RLS was detected in 12 cases (27%). Patients with RLS presented greater disability (P = 0.01), poorer sleep (P = 0.02) and greater levels of fatigue (P = 0.03). Impaired sleep was present in 23 (52%) and excessive daytime sleepiness in 3 cases (6.8%). Fatigue was present in 32 subjects (73%) and was associated with clinical disability (P = 0.000) and sleep quality (P = 0.002). Age, gender, disease duration, MS pattern, excessive daytime sleepiness and the presence of upper motor neuron signs were not associated with the presence of RLS. Fatigue was best explained by clinical disability and poor sleep quality. Awareness of RLS among health care professionals may contribute to improvement in MS management.

  16. Relationships Among Daytime Napping and Fatigue, Sleep Quality, and Quality of Life in Cancer Patients.

    PubMed

    Sun, Jia-Ling; Lin, Chia-Chin

    2016-01-01

    The relationships among napping and sleep quality, fatigue, and quality of life (QOL) in cancer patients are not clearly understood. The aim of the study was to determine whether daytime napping is associated with nighttime sleep, fatigue, and QOL in cancer patients. In total, 187 cancer patients were recruited. Daytime napping, nighttime self-reported sleep, fatigue, and QOL were assessed using a questionnaire. Objective sleep parameters were collected using a wrist actigraph. According to waking-after-sleep-onset measurements, patients who napped during the day experienced poorer nighttime sleep than did patients who did not (t = -2.44, P = .02). Daytime napping duration was significantly negatively correlated with QOL. Patients who napped after 4 PM had poorer sleep quality (t = -1.93, P = .05) and a poorer Short-Form Health Survey mental component score (t = 2.06, P = .04) than did patients who did not. Fatigue, daytime napping duration, and sleep quality were significant predictors of the mental component score and physical component score, accounting for 45.7% and 39.3% of the variance, respectively. Daytime napping duration was negatively associated with QOL. Napping should be avoided after 4 PM. Daytime napping affects the QOL of cancer patients. Future research can determine the role of napping in the sleep hygiene of cancer patients.

  17. Work and excessive sleepiness among Brazilian evening high school students: effects on days off.

    PubMed

    Teixeira, Liliane; Lowden, Arne; Moreno, Claudia Roberta; Turte, Samantha; Nagai, Roberta; Latorre, Maria Do Rosário; Valente, Daniel; Fischer, Frida Marina

    2010-01-01

    Previous studies have revealed that students who work and study build up sleep deficits during the workweek, which can trigger a sleep rebound during days off. The objective of this study was to investigate the impact of working/non-working on sleepiness during days off among high school students. The study population, aged 14-21 years, attended evening classes in São Paulo, Brazil. For the study, the students completed questionnaires on living conditions, health, and work; wore actigraphs; and completed the Karolinska Sleepiness Scale (KSS). To predict sleepiness, a logistic regression analysis was performed. Excessive sleepiness was observed on the first day off among working students. Results suggest that working is a significant predictor for sleepiness and that two shifts of daily systematic activities, study and work, might lead to excessive daytime sleepiness on the first day off. Further, this observed excessive sleepiness may reflect the sleep debt accumulated during the workweek.

  18. The impact of Sleep Time-Related Information and Communication Technology (STRICT) on sleep patterns and daytime functioning in American adolescents.

    PubMed

    Polos, Peter G; Bhat, Sushanth; Gupta, Divya; O'Malley, Richard J; DeBari, Vincent A; Upadhyay, Hinesh; Chaudhry, Saqib; Nimma, Anitha; Pinto-Zipp, Genevieve; Chokroverty, Sudhansu

    2015-10-01

    This cross-sectional study explored the extent and impact of mobile device-based Sleep Time-Related Information and Communication Technology (STRICT) use among American adolescents (N = 3139, 49.3% female, mean age = 13.3 years). Nearly 62% used STRICT after bedtime, 56.7% texted/tweeted/messaged in bed, and 20.8% awoke to texts. STRICT use was associated with insomnia, daytime sleepiness, eveningness, academic underperformance, later bedtimes and shorter sleep duration. Moderation analysis demonstrated that the association between STRICT use and insomnia increased with age, the association between STRICT use and daytime sleepiness decreased with age, and the association between STRICT use and shorter sleep duration decreased with age and was stronger in girls. Insomnia and daytime sleepiness partially mediated the relationship between STRICT use and academic underperformance. Our results illustrate the adverse interactions between adolescent STRICT use and sleep, with deleterious effects on daytime functioning. These worrisome findings suggest that placing reasonable limitations on adolescent STRICT use may be appropriate. Copyright © 2015 The Foundation for Professionals in Services for Adolescents. Published by Elsevier Ltd. All rights reserved.

  19. Examining courses of sleep quality and sleepiness in full 2 weeks on/2 weeks off offshore day shift rotations.

    PubMed

    Riethmeister, V; Bültmann, U; De Boer, M R; Gordijn, M; Brouwer, S

    2018-05-16

    To better understand sleep quality and sleepiness problems offshore, we examined courses of sleep quality and sleepiness in full 2-weeks on/2-weeks off offshore day shift rotations by comparing pre-offshore (1 week), offshore (2 weeks) and post-offshore (1 week) work periods. A longitudinal observational study was conducted among N=42 offshore workers. Sleep quality was measured subjectively with two daily questions and objectively with actigraphy, measuring: time in bed (TIB), total sleep time (TST), sleep latency (SL) and sleep efficiency percentage (SE%). Sleepiness was measured twice a day (morning and evening) with the Karolinska Sleepiness Scale. Changes in sleep and sleepiness parameters during the pre/post and offshore work periods were investigated using (generalized) linear mixed models. In the pre-offshore work period, courses of SE% significantly decreased (p=.038). During offshore work periods, the courses of evening sleepiness scores significantly increased (p<.001) and significantly decreased during post-offshore work periods (p=.004). During offshore work periods, TIB (p<.001) and TST (p<.001) were significantly shorter, SE% was significantly higher (p=.002), perceived sleep quality was significantly lower (p<.001) and level of rest after wake was significantly worse (p<.001) than during the pre- and post-offshore work periods. Morning sleepiness was significantly higher during offshore work periods (p=.015) and evening sleepiness was significantly higher in the post-offshore work period (p=.005) compared to the other periods. No significant changes in SL were observed. Courses of sleep quality and sleepiness parameters significantly changed during full 2-weeks on/2-weeks off offshore day shift rotation periods. These changes should be considered in offshore fatigue risk management programmes.

  20. Ready for takeoff? A critical review of armodafinil and modafinil for the treatment of sleepiness associated with jet lag

    PubMed Central

    McCarty, David E

    2010-01-01

    Jet lag syndrome (JLS) is a clinical syndrome of disrupted nocturnal sleep and daytime neurocognitive impairment which occurs in the context of rapid transmeridian travel. Many strategies for treatment of JLS exist, and include hypnotics to enhance nocturnal sleep, chronotherapeutic approaches (eg, light therapy, melatonin, or gradual schedule shifting), and alerting agents to counter daytime sleepiness. Safety concerns have prompted renewed interest in managing JLS-associated excessive daytime sleepiness (JLSAEDS). Off-label use of the newer alerting agents modafinil and armodafinil is increasing for this indication, often at the specific request of patients. In order to better evaluate the potential risks and benefits of these medications for the management of JLSAEDS, clinicians must be aware of what is known – and still not known. In this article, the pharmacology and pharmacokinetics of modafinil and armodafinil are reviewed, along with evidence for their efficacy in treating sleepiness associated with narcolepsy, obstructive sleep apnea and shift work sleep disorder. Clinical trial data for use of alerting agents in the management of JLSAEDS are limited to one three-day trial involving armodafinil, dosed in the morning to treat JLSAEDS in the setting of eastbound transmeridian travel. This study showed improvement in objective measures of daytime sleepiness at doses of 50 and 150 mg per day. However, global impression of clinical severity of symptom scores only improved on day 1 for those patients receiving 150 mg, and were otherwise not superior to placebo. Consideration for the use of modafinil or armodafinil for the treatment of sleepiness associated with JLS involves careful integration of patient-reported goals, a review of medical contraindications, and an awareness of rare adverse events. More research is needed in order to identify those who are most likely to benefit from this intervention and better define the risk-benefit ratio for this

  1. Sleep and Daytime Functioning: A Short-Term Longitudinal Study of Three Preschool-Age Comparison Groups

    ERIC Educational Resources Information Center

    Anders, Thomas; Iosif, Ana-Maria; Schwichtenberg, A. J.; Tang, Karen; Goodlin-Jones, Beth

    2012-01-01

    This study examined sleep, sleepiness, and daytime performance in 68 children with autism, 57 children with intellectual disability (ID), and 69 typically developing preschool children. Children in the autism and ID groups had poorer daytime performance and behaviors than the typically developing children. Children in the ID group also were…

  2. High-risk of obstructive sleep apnea and excessive daytime sleepiness among commercial intra-city drivers in Lagos metropolis

    PubMed Central

    Ozoh, Obianuju B.; Okubadejo, Njideka U.; Akanbi, Maxwell O.; Dania, Michelle G.

    2013-01-01

    Background: The burden of obstructive sleep apnea among commercial drivers in Nigeria is not known. Aim: To assess the prevalence of high risk of obstructive sleep apnea (OSA) and excessive daytime sleepiness (EDS) among intra-city commercial drivers. Setting and Design: A descriptive cross-sectional study in three major motor parks in Lagos metropolis. Materials and Methods: Demographic, anthropometric and historical data was obtained. The risk of OSA and EDS was assessed using the STOP BANG questionnaire and the Epworth Sleepiness Scale, respectively. Statistical Analysis: The relationship between the OSA risk, EDS risk and past road traffic accident (RTA) was explored using the Pearson's chi square. Independent determinants of OSA risk, EDS risk and past RTA, respectively, were assessed by multiple logistic regression models. Result: Five hundred male commercial drivers (mean age (years) ±SD = 42.36 ± 11.17 and mean BMI (kg/m2) ±SD = 25.68 ± 3.79) were recruited. OSA risk was high in 244 (48.8%) drivers and 72 (14.4%) had EDS. There was a positive relationship between OSA risk and the risk of EDS (Pearson's X2 = 28.2, P < 0.001). Sixty-one (12.2%) drivers had a past history of RTA but there was no significant relationship between a past RTA and either OSA risk (X2 = 2.05, P = 0.15) or EDS risk (X2 = 2.7, P = 0.1), respectively. Abdominal adiposity, regular alcohol use and EDS were independent determinants of OSA risk while the use of cannabis and OSA risk were independent determinants of EDS. No independent risk factor for past RTA was identified. Conclusion: A significant proportion of commercial drivers in Lagos metropolis are at high risk of OSA and EDS. PMID:24249946

  3. The role of sleepiness on arterial stiffness improvement after CPAP therapy in males with obstructive sleep apnea: a prospective cohort study.

    PubMed

    Mineiro, Maria Alexandra; Silva, Pedro Marques da; Alves, Marta; Papoila, Ana Luísa; Marques Gomes, Maria João; Cardoso, João

    2017-12-08

    Obstructive sleep apnea (OSA) is associated with increased cardiovascular risk. This study aim to assess differences in changes in arterial stiffness of two groups of patients, defined as having daytime sleepiness or not, after continuous positive airway pressure (CPAP) treatment. A selected cohort of consecutive male patients, under 65 years old, with moderate to severe OSA and without great number of comorbidities was studied. The diagnosis was confirmed by home respiratory poligraphy. Sleepiness was considered with an Epworth Sleepiness Scale (ESS) > 10. An ambulatory blood pressure (BP) monitoring and carotid-femoral pulse wave velocity (cf-PWV) measurements were performed, before and after four months under CPAP. Compliant patients, sleepy and non-sleepy, were compared using linear mixed effects regression models. A further stratified analysis was performed with non-sleepy patients. Thirty-four patients were recruited, with mean age 55.2 (7.9) years, 38.2% were sleepy, 79.4% with hypertension, 61.8% with metabolic syndrome and 82.4% with dyslipidaemia. In univariable analysis, cf-PWV was strongly related to systolic BP parameters and age, but also to antihypertensive drugs (p = 0.030), metabolic syndrome (p = 0.025) and daytime sleepiness (p = 0.004). Sleepy patients had a more severe OSA, with AHI 44.8 (19.0) vs 29.7 (15.7) events/h (p = 0.018), but sleep study parameters were not associated with cf-PWV values. On multivariable regression, a significant interaction between time (CPAP) and sleepiness (p = 0.033) was found. There was a weak evidence of a cf-PWV reduction after CPAP treatment (p = 0.086), but the effects of treatment differed significantly between groups, with no changes in non-sleepy patients, while in sleepy patients a significant decrease was observed (p = 0.012). Evaluating non-sleepy patients group under CPAP therapy, results showed that both higher pulse pressure (p = 0.001) and lower LDL-cholesterol levels (p

  4. Relationship between circadian rhythm amplitude and stability with sleep quality and sleepiness among shift nurses and health care workers.

    PubMed

    Jafari Roodbandi, Akram; Choobineh, Alireza; Daneshvar, Somayeh

    2015-01-01

    Sleep is affected by the circadian cycle and its features. Amplitude and stability of circadian rhythm are important parameters of the circadian cycle. This study aims to examine the relationship between amplitude and stability of circadian rhythm with sleep quality and sleepiness. In this cross-sectional research, 315 shift nurses and health care workers from educational hospitals of Kerman University of Medical Sciences (KUMS), Iran, were selected using a random sampling method. The Pittsburgh Sleep Quality Index (PSQI), Epworth Sleepiness Scale (ESS) and Circadian Type Inventory (CTI) were used to collect the required data. In this study, 83.2% suffered from poor sleep and one-half had moderate and excessive sleepiness. The results showed that flexibility in circadian rhythm stability, job stress and sleepiness are among the factors affecting quality sleep in shift workers. Those whose circadian rhythm amplitude was languid suffered more from sleepiness and those whose circadian stability was flexible had a better sleep. Variables including circadian rhythm stability (flexible/rigid) and amplitude (languid/vigorous) can act as predictive indices in order to employ people in a shift work system so that sleepiness and a drop in quality of sleep are prevented.

  5. [CLINICAL INVESTIGATION OF AN EXCESSIVE SLEEPINESS COMPLAINT].

    PubMed

    Evangelista, Elisa; Barateau, Lucie; Dauvilliers, Yves

    2016-06-01

    Excessive sleepiness is a common problem, defined by a complaint of excessive daytime sleepiness almost daily with an inability to stay awake and alert dosing periods at sleep, with episodes of irresistible sleep need or drowsiness or non-intentional sleep, or by a night's sleep time overly extended often associated with sleep inertia. This sleepiness is variable in terms of phenotype and severity to be specified by the out-patient clinic. It is considered to be chronic beyond three months and often responsible for significant functional impairment of school and professional performance, of the accidents and cardiovascular risk. We need to decipher the causes of excessive sleepiness: sleep deprivation, toxic and iatrogenic, psychiatric disorders (including depression), non-psychiatric medical problems (obesity, neurological pathologies...), sleep disorders (as for example the sleep apnea syndrome), and finally the central hypersomnias namely narcolepsy type 1 and 2, idiopathic hypersomnia, and Kleine-Levin syndrome. If careful questioning often towards one of these etiologies, need most of the time a paraclinical balance with a sleep recording to confirm the diagnosis. Patients affected with potential central hypersomnia must be referred to the Sleep Study Centers that have the skills and the appropriate means to achieve this balance sheet.

  6. Lower plasma choline levels are associated with sleepiness symptoms.

    PubMed

    Pak, Victoria M; Dai, Feng; Keenan, Brendan T; Gooneratne, Nalaka S; Pack, Allan I

    2018-04-01

    Sleepiness and cardiovascular disease share common molecular pathways; thus, metabolic risk factors for sleepiness may also predict cardiovascular disease risk. Daytime sleepiness predicts mortality and cardiovascular disease, although the mechanism is unidentified. This study explored the associations between subjective sleepiness and metabolite concentrations in human blood plasma within the oxidative and inflammatory pathways, in order to identify mechanisms that may contribute to sleepiness and cardiovascular disease risk. An exploratory case-control sample of 36 subjects, categorized based on the Epworth Sleepiness Scale (ESS) questionnaire as sleepy (ESS ≥ 10) or non-sleepy (ESS < 10), was recruited among subjects undergoing an overnight sleep study for suspected sleep apnea at the University of Pennsylvania Sleep Center. The average age was 42.4 ± 10.5 years, the mean body mass index (BMI) was 40.0 ± 9.36 kg/m 2 , median Apnea Hypopnea Index (AHI) was 8.2 (IQR: 2.5-26.5), and 52% were male. Fasting morning blood plasma samples were collected after an overnight sleep study. Biomarkers were explored in subjects with sleepiness versus those without using the multiple linear regression adjusting for age, BMI, smoking, Apnea Hypopnea Index (sleep apnea severity), study cohort, and hypertension. The level of choline is significantly lower (P = 0.003) in sleepy subjects (N = 18; mean plasma choline concentration of 8.19 ± 2.62 μmol/L) compared with non-sleepy subjects (N = 18; mean plasma choline concentration of 9.14 ± 2.25 μmol/L). Other markers with suggestive differences (P < 0.1) include isovalerylcarnitine, Alpha-Amino apidipic acid, Spingosine 1 Phosphate, Aspartic Acid, Propionylcarnitine, and Ceramides (fatty acids; C14-C16 and C-18). This pilot study is the first to show that lower levels of plasma choline metabolites are associated with sleepiness. Further exploration of choline and other noted metabolites and their

  7. Subjective symptoms in idiopathic hypersomnia: beyond excessive sleepiness.

    PubMed

    Vernet, Cyrille; Leu-Semenescu, Smaranda; Buzare, Marie-Annick; Arnulf, Isabelle

    2010-12-01

    Patients with idiopathic hypersomnia never feel fully alert despite a normal or long sleep night. The spectrum of the symptoms is insufficiently studied. We interviewed 62 consecutive patients with idiopathic hypersomnia (with a mean sleep latency lower than 8 min or a sleep time longer than 11 h) and 50 healthy controls using a questionnaire on sleep, awakening, sleepiness, alertness and cognitive, psychological and functional problems during daily life conditions. Patients slept 3 h more on weekends, holidays and in the sleep unit than on working days. In the morning, the patients needed somebody to wake them, or to be stressed, while routine, light, alarm clocks and motivation were inefficient. Three-quarters of the patients did not feel refreshed after short naps. During the daytime, their alertness was modulated by the same external conditions as controls, but they felt more sedated in darkness, in a quiet environment, when listening to music or conversation. Being hyperactive helped them more than controls to resist sleepiness. They were more frequently evening-type and more alert in the evening than in the morning. The patients were able to focus only for 1 h (versus 4 h in the controls). They complained of attention and memory deficit. Half of them had problems regulating their body temperature and were near-sighted. Mental fatigability, dependence on other people for awakening them, and a reduced benefit from usually alerting conditions (except being hyperactive or stressed) seem to be more specific of the daily problems of patients with idiopathic hypersomnia than daytime sleepiness. © 2010 European Sleep Research Society.

  8. Final report on the portable computerized assessments of sleepy drivers in operational environments.

    DOT National Transportation Integrated Search

    2011-06-01

    Excessive daytime sleepiness underpins a large number of the reported motor vehicle crashes. Fair and accurate field : measures are needed to identify at-risk drivers who have been identified as potentially driving in a sleep deprived state on : the ...

  9. Attention Deficit Hyperactivity Disorder Symptoms, Sleepiness and Accidental Risk in 36140 Regularly Registered Highway Drivers

    PubMed Central

    Philip, Pierre; Micoulaud-Franchi, Jean-Arthur; Lagarde, Emmanuel; Taillard, Jacques; Canel, Annick; Sagaspe, Patricia; Bioulac, Stéphanie

    2015-01-01

    Background Attention Deficit Hyperactivity Disorder (ADHD) is a frequent neurodevelopmental disorder that increases accidental risk. Recent studies show that some patients with ADHD can also suffer from excessive daytime sleepiness but there are no data assessing the role of sleepiness in road safety in patients with ADHD. We conducted an epidemiological study to explore sleep complaints, inattention and driving risks among automobile drivers. Methods and Findings From August to September 2014, 491186 regular highway users were invited to participate in an Internet survey on driving habits. 36140 drivers answered a questionnaire exploring driving risks, sleep complaints, sleepiness at the wheel, ADHD symptoms (Adult ADHD Self-Report Scale) and distraction at the wheel. 1.7% of all drivers reported inattention-related driving accidents and 0.3% sleep-related driving accidents in the previous year. 1543 drivers (4.3%) reported ADHD symptoms and were more likely to report accidents than drivers without ADHD symptoms (adjusted OR = 1.24, [1.03–1.51], p < .021). 14.2% of drivers with ADHD symptoms reported severe excessive daytime sleepiness (Epworth Sleepiness Scale >15) versus 3.2% of drivers without ADHD symptoms and 20.5% reported severe sleepiness at the wheel versus 7.3%. Drivers with ADHD symptoms reported significantly more sleep-related (adjusted OR = 1.4, [1.21–1.60], p < .0001) and inattention-related (adjusted OR = 1.9, [1.71–2.14], p<0001) near misses than drivers without ADHD symptoms. The fraction of near-misses attributable to severe sleepiness at the wheel was 4.24% for drivers without ADHD symptoms versus 10,35% for drivers with ADHD symptoms. Conclusion Our study shows that drivers with ADHD symptoms have more accidents and a higher level of sleepiness at the wheel than drivers without ADHD symptoms. Drivers with ADHD symptoms report more sleep-related and inattention-related near misses, thus confirming the clinical importance of exploring both

  10. Attention Deficit Hyperactivity Disorder Symptoms, Sleepiness and Accidental Risk in 36140 Regularly Registered Highway Drivers.

    PubMed

    Philip, Pierre; Micoulaud-Franchi, Jean-Arthur; Lagarde, Emmanuel; Taillard, Jacques; Canel, Annick; Sagaspe, Patricia; Bioulac, Stéphanie

    2015-01-01

    Attention Deficit Hyperactivity Disorder (ADHD) is a frequent neurodevelopmental disorder that increases accidental risk. Recent studies show that some patients with ADHD can also suffer from excessive daytime sleepiness but there are no data assessing the role of sleepiness in road safety in patients with ADHD. We conducted an epidemiological study to explore sleep complaints, inattention and driving risks among automobile drivers. From August to September 2014, 491186 regular highway users were invited to participate in an Internet survey on driving habits. 36140 drivers answered a questionnaire exploring driving risks, sleep complaints, sleepiness at the wheel, ADHD symptoms (Adult ADHD Self-Report Scale) and distraction at the wheel. 1.7% of all drivers reported inattention-related driving accidents and 0.3% sleep-related driving accidents in the previous year. 1543 drivers (4.3%) reported ADHD symptoms and were more likely to report accidents than drivers without ADHD symptoms (adjusted OR = 1.24, [1.03-1.51], p < .021). 14.2% of drivers with ADHD symptoms reported severe excessive daytime sleepiness (Epworth Sleepiness Scale >15) versus 3.2% of drivers without ADHD symptoms and 20.5% reported severe sleepiness at the wheel versus 7.3%. Drivers with ADHD symptoms reported significantly more sleep-related (adjusted OR = 1.4, [1.21-1.60], p < .0001) and inattention-related (adjusted OR = 1.9, [1.71-2.14], p<0001) near misses than drivers without ADHD symptoms. The fraction of near-misses attributable to severe sleepiness at the wheel was 4.24% for drivers without ADHD symptoms versus 10,35% for drivers with ADHD symptoms. Our study shows that drivers with ADHD symptoms have more accidents and a higher level of sleepiness at the wheel than drivers without ADHD symptoms. Drivers with ADHD symptoms report more sleep-related and inattention-related near misses, thus confirming the clinical importance of exploring both attentional deficits and sleepiness at the wheel

  11. Natural history of excessive daytime sleepiness: role of obesity, weight loss, depression, and sleep propensity.

    PubMed

    Fernandez-Mendoza, Julio; Vgontzas, Alexandros N; Kritikou, Ilia; Calhoun, Susan L; Liao, Duanping; Bixler, Edward O

    2015-03-01

    Excessive daytime sleepiness (EDS) is highly prevalent in the general population and is associated with occupational and public safety hazards. However, no study has examined the clinical and polysomnographic (PSG) predictors of the natural history of EDS. Representative longitudinal study. Sleep laboratory. From a random, general population sample of 1,741 individuals of the Penn State Adult Cohort, 1,395 were followed up after 7.5 years. Full medical evaluation and 1-night PSG at baseline and standardized telephone interview at follow-up. The incidence of EDS was 8.2%, while its persistence and remission were 38% and 62%, respectively. Obesity and weight gain were associated with the incidence and persistence of EDS, while weight loss was associated with its remission. Significant interactions between depression and PSG parameters on incident EDS showed that, in depressed individuals, incident EDS was associated with sleep disturbances, while in non-depressed individuals, incident EDS was associated with increased physiologic sleep propensity. Diabetes, allergy/ asthma, anemia, and sleep complaints also predicted the natural history of EDS. Obesity, a disorder of epidemic proportions, is a major risk factor for the incidence and chronicity of EDS, while weight loss is associated with its remission. Interestingly, objective sleep disturbances predict incident EDS in depressed individuals, whereas physiologic sleep propensity predicts incident EDS in those without depression. Weight management and treatment of depression and sleep disorders should be part of our public health policies. © 2015 Associated Professional Sleep Societies, LLC.

  12. Insomnia, Sleepiness, and Depression in Adolescents Living in Residential Care Facilities

    ERIC Educational Resources Information Center

    Moreau, Vincent; Belanger, Lynda; Begin, Gilles; Morin, Charles M.

    2009-01-01

    The main objective of this study was to document sleep patterns and disturbances reported by youths temporarily living in residential care facilities. A secondary objective was to examine the relationships between sleep disturbances and mood and daytime sleepiness. A self-reported questionnaire on sleep patterns and habits assessing duration,…

  13. Validation of the Urdu version of the Epworth Sleepiness Scale.

    PubMed

    Surani, Asif Anwar; Ramar, Kannan; Surani, Arif Anwar; Khaliqdina, Jehangir Shehryar; Subramanian, Shyam; Surani, Salim

    2012-09-01

    To translate and validate the Epworth Sleepiness Scale (ESS) for use in Urdu-speaking population. The original Epworth Sleepiness Scale was translated into the Urdu version (ESS-Ur) in three phases - translation and back-translation; committee-based translation; and testing in bilingual individuals. The final was subsequently tested on 89 healthy bilingual subjects between February and April, 2010, to assess the validity of the translation compared to the original version. The subjects were students and employees of Dow University of Health Sciences, Karachi. Both English and Urdu versions of the Epworth Sleepiness Scale were administered to 59 (67%) women and 30 (33%) men. The mean composite Epworth score was 7.53 in English language and 7.7 in the Urdu version (p=0.76). The translated version was found to be highly correlated with the original scale (rho=0.938; p<.01). The study validated the scale's Urdu version as an effective tool for measuring daytime sleepiness in Urdu-speaking population. Future studies assessing the validity of such patients with sleep disorders need to be undertaken.

  14. Sleep, Glucose, and Daytime Functioning in Youth with Type 1 Diabetes

    PubMed Central

    Perfect, Michelle M.; Patel, Priti G.; Scott, Roxanne E.; Wheeler, Mark D.; Patel, Chetanbabu; Griffin, Kurt; Sorensen, Seth T.; Goodwin, James L.; Quan, Stuart F.

    2012-01-01

    Study Hypotheses: 1) Youth with evidence of SDB (total apnea-hypopnea index [Total-AHI] ≥ 1.5) would have significantly worse glucose control than those without SDB; 2) Elevated self-reported sleepiness in youth with T1DM would be related to compromised psychosocial functioning; and 3) Youth with T1DM would have significantly less slow wave sleep (SWS) than controls. Design: The study utilized home-based polysomnography, actigraphy, and questionnaires to assess sleep, and continuous glucose monitors and hemoglobin A1C (HbA1C) values to assess glucose control in youth with T1DM. We compared sleep of youth with T1DM to sleep of a matched control sample. Setting: Diabetic participants were recruited in a pediatric endocrinology clinic. Participants: Participants were youth (10 through 16 years) with T1DM. Controls, matched for sex, age, and BMI percentile, were from the Tucson Children's Assessment of Sleep Apnea study. Results: Participants with a Total-AHI ≥ 1.5 had higher glucose levels. Sleepiness and/or poor sleep habits correlated with reduced quality of life, depressed mood, lower grades, and lower state standardized reading scores. Diabetic youth spent more time (%) in stage N2 and less time in stage N3. Findings related to sleep architecture included associations between reduced SWS and higher HbA1C, worse quality of life, and sleepiness. More time (%) spent in stage N2 related to higher glucose levels/hyperglycemia, behavioral difficulties, reduced quality of life, lower grades, depression, sleep-wake behavior problems, poor sleep quality, sleepiness, and lower state standardized math scores. Conclusions: Sleep should be routinely assessed as part of diabetes management in youth with T1DM. Citation: Perfect MM; Patel PG; Scott RE; Wheeler MD; Patel C; Griffin K; Sorensen ST; Goodwin JL; Quan SF. Sleep, glucose, and daytime functioning in youth with type 1 diabetes. SLEEP 2012;35(1):81-88. PMID:22215921

  15. Daytime Symptoms in Primary Insomnia: A Prospective Analysis Using Ecological Momentary Assessment

    PubMed Central

    Buysse, Daniel J.; Thompson, Wesley; Scott, John; Franzen, Peter L.; Germain, Anne; Hall, Martica L.; Moul, Douglas E.; Nofzinger, Eric A.; Kupfer, David J.

    2007-01-01

    Objectives To prospectively characterize and compare daytime symptoms in primary insomnia (PI) and good sleeper control (GSC) subjects using ecological momentary assessment; to examine relationships between daytime symptom factors, retrospective psychological and sleep reports, and concurrent sleep diary reports. Methods Subjects included 47 PI and 18 GSC. Retrospective self-reports of daytime and sleep symptoms were collected. Daytime symptoms and sleep diary information were then collected for one week on hand-held computers. The Daytime Insomnia Symptom Scale (DISS) consisted of 19 visual analog scales completed four times per day. Factors for the DISS were derived using functional principal components analysis. Nonparametric tests were used to contrast DISS, retrospective symptom ratings, and sleep diary results in PI and GSC subjects, and to examine relationships among them. Results Four principal components were identified for the DISS: Alert Cognition, Negative Mood, Positive Mood, and Sleepiness/Fatigue. PI scored significantly worse than GSC on all four factors (p < .0003 for each). Among PI subjects DISS scales and retrospective psychological symptoms were related to each other in plausible ways. DISS factors were also related to self-report measures of sleep, whereas retrospective psychological symptom measures were not. Conclusions Daytime symptom factors of alertness, positive and negative mood, and sleepiness/fatigue, collected with ecological momentary assessment, showed impairment in PI versus GSC. DISS factors showed stronger relationships to retrospective sleep symptoms and concurrent sleep diary reports than retrospective psychological symptoms. The diurnal pattern of symptoms may inform studies of the pathophysiology and treatment outcome of insomnia. PMID:17368098

  16. Job demands and resting and napping opportunities for nurses during night shifts: impact on sleepiness and self-evaluated quality of healthcare.

    PubMed

    Barthe, Béatrice; Tirilly, Ghislaine; Gentil, Catherine; Toupin, Cathy

    2016-01-01

    The aim of this field study is to describe night shift resting and napping strategies and to examine their beneficial effects on sleepiness and quality of work. The study was carried out with 16 nurses working in an intensive care unit. Data collected during 20 night shifts were related to job demands (systematic observations), to the duration and timing of rests and naps taken by nurses (systematic observations, sleep diaries), to sleepiness (Karolinska Sleepiness Scale), and to quality of work scores (visual analog scale). The results showed that the number of rests and naps depended on the job demands. Resting and napping lowered the levels of sleepiness at the end of the shift. There was no direct relationship between sleepiness and the quality of work score. Discussions about the choice of indicators for the quality of work are necessary. Suggestions for implementing regulations for prescribed napping during night shifts are presented.

  17. Armodafinil for treatment of excessive sleepiness associated with shift work disorder: a randomized controlled study.

    PubMed

    Czeisler, Charles A; Walsh, James K; Wesnes, Keith A; Arora, Sanjay; Roth, Thomas

    2009-11-01

    To assess the effect of armodafinil, 150 mg, on the physiologic propensity for sleep and cognitive performance during usual night shift hours in patients with excessive sleepiness associated with chronic (> or =3 months) shift work disorder (SWD) of moderate or greater severity. This 12-week, randomized controlled study was conducted at 42 sleep research facilities in North America from April 2 through December 23, 2004, and enrolled 254 permanent or rotating night shift workers with SWD. Entry criteria included excessive sleepiness during usual night shifts for 3 months or longer (corroborated by mean sleep latency of < or =6 minutes on a Multiple Sleep Latency Test), insomnia (sleep efficiency < or =87.5% during daytime sleep), and SWD that was judged clinically to be of moderate or greater severity. Patients received armodafinil, 150 mg, or placebo 30 to 60 minutes before each night shift. Physiologic sleep propensity during night shift hours, clinical impression of severity, patient-reported sleepiness, and cognitive function were assessed during laboratory night shifts at weeks 4, 8, and 12. Armodafinil significantly improved mean (SD) sleep latency from 2.3 (1.6) minutes at baseline to 5.3 (5.0) minutes at final visit, compared with a change from 2.4 (1.6) minutes to 2.8 (2.9) minutes in the placebo group (P<.001). Clinical condition ratings improved in more patients receiving armodafinil (79%) vs placebo (59%) (P=.001). As reported by patients' diaries, armodafinil significantly reduced sleepiness during laboratory nights (P<.001), night shifts at work (P<.001), and the commute home (P=.003). Armodafinil improved performance on standardized memory (P<.001) and attention (power, P=.001; continuity, P<.001) tests compared with placebo. Armodafinil was well tolerated and did not affect daytime sleep, as measured by polysomnography. In patients with excessive sleepiness associated with chronic SWD of moderate or greater severity, armodafinil significantly

  18. The relation between burnout and sleep disorders in medical students.

    PubMed

    Pagnin, Daniel; de Queiroz, Valéria; Carvalho, Yeska Talita Maia Santos; Dutra, Augusto Sergio Soares; Amaral, Monique Bastos; Queiroz, Thiago Thomasin

    2014-08-01

    The aim of this study is to assess the mutual relationships between burnout and sleep disorders in students in the preclinical phase of medical school. This study collected data on 127 medical students who filled in the Maslach Burnout Inventory-Student Survey, Pittsburgh Sleep Quality Index, Epworth Sleepiness Scale, Beck Depression Inventory, and Beck Anxiety Inventory. Hierarchical logistic regressions tested the reciprocal influence between sleep disorders and burnout, controlling for depression and anxiety. Regular occurrence of emotional exhaustion, poor sleep quality, and excessive daytime sleepiness affected 60, 65, and 63% of medical students, respectively. Emotional exhaustion and daytime sleepiness influenced each other. Daytime sleep dysfunctions affected unidirectionally the occurrence of cynicism and academic efficacy. The odds of emotional exhaustion (odds ratio (OR)=1.21, 95% confidence interval (CI)=1.08 to 1.35) and cynicism (OR=2.47, 95% CI=1.25 to 4.90) increased when daytime sleepiness increased. Reciprocally, the odds of excessive daytime sleepiness (OR=2.13, 95% CI=1.22 to 3.73) increased when emotional exhaustion worsened. Finally, the odds of academic efficacy decreased (OR=0.86, 95% CI=0.75 to 0.98) when daytime sleepiness increased. Burnout and sleep disorders have relevant bidirectional effects in medical students in the early phase of medical school. Emotional exhaustion and daytime sleepiness showed an important mutual influence. Daytime sleepiness linked unidirectionally with cynicism and academic efficacy.

  19. Judgment of Daytime Sleepiness in Self-Reported Short, Long and Midrange Sleepers

    ERIC Educational Resources Information Center

    Mairesse, Olivier; Neu, Daniel; Migeotte, Pierre-Francois; Pattyn, Nathalie; Hofmans, Joeri; Theuns, Peter; Cluydts, Raymond; De Valck, Elke

    2012-01-01

    Sleep-wake behavior, as well as sleepiness, is regulated by the joint action of an exponentially increasing drive for sleep--sleep homeostasis--and by variations in sleep propensity due to a biological circadian oscillator. However, large inter-individual differences remain. Short and long sleepers have been known to differ in the amount of…

  20. Sleep Problems, Sleepiness and Daytime Behavior in Preschool-Age Children

    ERIC Educational Resources Information Center

    Goodlin-Jones, Beth; Tang, Karen; Liu, Jingyi; Anders, Thomas F.

    2009-01-01

    Background: Sleep problems are a common complaint of parents of preschool children. Children with neurodevelopmental disorders have even more disrupted sleep than typically developing children. Although disrupted nighttime sleep has been reported to affect daytime behavior, the pathway from sleep disruption to sleep problems, to impairments in…

  1. Fatigue and excessive daytime sleepiness in idiopathic Parkinson's disease differently correlate with motor symptoms, depression and dopaminergic treatment.

    PubMed

    Valko, P O; Waldvogel, D; Weller, M; Bassetti, C L; Held, U; Baumann, C R

    2010-12-01

    A comprehensive study of both fatigue and excessive daytime sleepiness (EDS) in association with Parkinson's disease (PD)-related symptoms and treatment has not been performed yet. To assess the frequency and severity of fatigue and EDS in patients with idiopathic PD and to study their relation to motor and non-motor symptoms and dopaminergic treatment. We prospectively assessed Fatigue Severity Scale (FSS) scores, Epworth Sleepiness Scale (ESS) scores, Beck Depression Inventory (BDI) scores, severity (Unified PD Rating Scale, UPDRS, part III; Hoehn & Yahr staging) and duration of the disease, and the current dopaminergic treatment in 88 consecutive patients with idiopathic PD. Fatigue was found in 52 (59%), EDS in 42 (48%), and both complaints in 31 (35%) patients. Fatigued patients had higher UPDRS III scores (23.5 ± 11.1 vs. 18.6 ± 7.6, P = 0.03), higher Hoehn & Yahr staging (2.4 ± 0.9 vs. 2.1 ± 0.7, P = 0.03), and higher BDI scores (13.4 ± 7.1 vs. 9.1 ± 5.8, P = 0.004) than non-fatigued patients. In contrast, UPDRS III, Hoehn & Yahr, and BDI scores did not differ between patients with or without EDS. However, the type of dopaminergic treatment (levodopa monotherapy versus combination of levodopa/dopamine agonists) was associated with significant differences in ESS (8.5 ± 5.2 vs. 10.8 ± 4.3, P = 0.04), but not FSS scores (4.1 ± 1.5 vs. 4.3 ± 1.5, P = 0.55). Disease duration correlated with ESS scores (r = 0.32, P = 0.003), but not with FSS scores (r = -0.02, P = 0.82). In PD, there is a significant overlap of fatigue and EDS, but the two symptoms are differently correlated with the severity of motor symptoms, disease duration, depression, and dopaminergic treatment. © 2010 The Author(s). European Journal of Neurology © 2010 EFNS.

  2. Sleep and sleepiness among working and non-working high school evening students.

    PubMed

    Teixeira, Liliane Reis; Lowden, Arne; Turte, Samantha Lemos; Nagai, Roberta; Moreno, Claudia Roberta de Castro; Latorre, Maria do Rosário Dias de Oliveira; Fischer, Frida Marina

    2007-01-01

    The aim of this study was to evaluate patterns of sleepiness, comparing working and non-working students. The study was conducted on high school students attending evening classes (19:00-22:30 h) at a public school in São Paulo, Brazil. The study group consisted of working (n=51) and non-working (n=41) students, aged 14-21 yrs. The students answered a questionnaire about working and living conditions and reported health symptoms and diseases. For seven consecutive days, actigraphy measurements were recorded, and the students also filled in a sleep diary. Sleepiness ratings were given six times per day, including upon waking and at bedtime, using the Karolinska Sleepiness Scale. Statistical analyses included three-way ANOVA and t-test. The mean sleep duration during weekdays was shorter among workers (7.2 h) than non-workers (8.8 h) (t=4.34; p<.01). The mean duration of night awakenings was longer among workers on Tuesdays and Wednesdays (28.2 min) and shorter on Mondays (24.2 min) (t=2.57; p=.03). Among workers, mean napping duration was longer on Mondays and Tuesdays (89.9 min) (t=2.27; p=.03) but shorter on Fridays and Sundays (31.4 min) (t=3.13; p=.03). Sleep efficiency was lower on Fridays among non-workers. Working students were moderately sleepier than non-workers during the week and also during class on specific days: Mondays (13:00-15:00 h), Wednesdays (19:00-22:00 h), and Fridays (22:00-00:59 h). The study found that daytime sleepiness of workers is moderately higher in the evening. This might be due to a work effect, reducing the available time for sleep and shortening the sleep duration. Sleepiness and shorter sleep duration can have a negative impact on the quality of life and school development of high school students.

  3. CPAP and behavioral therapies in patients with obstructive sleep apnea: effects on daytime sleepiness, mood, and cognitive function.

    PubMed

    Sánchez, Ana Isabel; Martínez, Pilar; Miró, Elena; Bardwell, Wayne A; Buela-Casal, Gualberto

    2009-06-01

    Obstructive sleep apnea (OSA) is a disorder characterized by repeated episodes of complete (apneas) or partial (hypopneas) cessations of breathing while sleeping. While continuous positive airway pressure (CPAP) treatment is commonly chosen to treat OSA, various conservative behavioral therapies are also used, particularly in patients unable to tolerate or benefit from CPAP or who have mild OSA. The principal purpose of these behavioral measures is to reduce risk factors which may underlie or exacerbate the disorder (e.g., weight reduction, smoking cessation, reduction/elimination of alcohol consumption, change in sleeping posture and sleep hygiene). Numerous studies have been conducted to evaluate the efficacy and/or effectiveness of CPAP in treating a wide range of OSA symptomatology. The present study consists of an exhaustive bibliographic search in Medline, PsycINFO, and Cochrane Review (1994-2007) databases and selection of works which have evaluated the efficacy and/or effectiveness of CPAP vis-a-vis daytime sleepiness, depression and cognitive functioning in OSA patients. The selected studies include randomized clinical trials in which CPAP was compared with more conservative measures, sham CPAP and oral placebos. The most important studies which evaluate the efficacy of behavioral treatments for OSA are also reviewed and the most remarkable results are presented. Various conclusions derived from the studies are discussed.

  4. Sleepiness, On-Task Behavior and Attention in Children with Epilepsy Who Visited a School for Special Education: A Comparative Study

    ERIC Educational Resources Information Center

    Didden, Robert; de Moor, Jan M. H.; Korzilius, Hubert

    2009-01-01

    Children with epilepsy are at risk for problems in daytime functioning. We assessed daytime sleepiness, on-task behavior and attention in 17 children (aged between 7 and 11 years) with epilepsy who visited a school for special education and compared these to 17 children from a control group who visited a regular school. Within the group of…

  5. Sleep-wake and melatonin pattern in craniopharyngioma patients.

    PubMed

    Pickering, Line; Jennum, Poul; Gammeltoft, Steen; Poulsgaard, Lars; Feldt-Rasmussen, Ulla; Klose, Marianne

    2014-06-01

    To assess the influence of craniopharyngioma or consequent surgery on melatonin secretion, and the association with fatigue, sleepiness, sleep pattern and sleep quality. Cross-sectional study. A total of 15 craniopharyngioma patients were individually matched to healthy controls. In this study, 24-h salivary melatonin and cortisol were measured. Sleep-wake patterns were characterised by actigraphy and sleep diaries recorded for 2 weeks. Sleepiness, fatigue, sleep quality and general health were assessed by Multidimensional Fatigue Inventory, Pittsburgh Sleep Quality Index, Epworth Sleepiness Scale and Short-Form 36. Patients had increased mental fatigue, daytime dysfunction, sleep latency and lower general health (all, P≤0.05), and they tended to have increased daytime sleepiness, general fatigue and impaired sleep quality compared with controls. The degree of hypothalamic injury was associated with an increased BMI and lower mental health (P=0.01). High BMI was associated with increased daytime sleepiness, daytime dysfunction, mental fatigue and lower mental health (all, P≤0.01). Low midnight melatonin was associated with reduced sleep time and efficiency (P≤0.03) and a tendency for increased sleepiness, impaired sleep quality and physical health. Midnight melatonin remained independently related to sleep time after adjustment for cortisol. Three different patterns of melatonin profiles were observed; normal (n=6), absent midnight peak (n=6) and phase-shifted peak (n=2). Only patients with absent midnight peak had impaired sleep quality, increased daytime sleepiness and general and mental fatigue. Craniopharyngioma patients present with changes in circadian pattern and daytime symptoms, which may be due to the influence of the craniopharyngioma or its treatment on the hypothalamic circadian and sleep regulatory nuclei. © 2014 European Society of Endocrinology.

  6. Can standardized sleep questionnaires be used to identify excessive daytime sleeping in older post-acute rehabilitation patients?

    PubMed

    Skibitsky, Megan; Edelen, Maria Orlando; Martin, Jennifer L; Harker, Judith; Alessi, Cathy; Saliba, Debra

    2012-02-01

    Excessive daytime sleeping is associated with poorer functional outcomes in rehabilitation populations and may be improved with targeted interventions. The purpose of this study was to test simple methods of screening for excessive daytime sleeping among older adults admitted for postacute rehabilitation. Secondary analysis of data from 2 clinical samples. Two postacute rehabilitation (PAR) units in southern California. Two hundred twenty-six patients older than 65 years with Mini-Mental State Examination (MMSE) score higher than 11 undergoing rehabilitation. The primary outcome was excessive daytime sleeping, defined as greater than 15% (1.8 hours) of daytime hours (8 am to 8 pm) sleeping as measured by actigraphy. Participants spent, on average, 16.2% (SD 12.5%) of daytime hours sleeping as measured by actigraphy. Thirty-nine percent of participants had excessive daytime sleeping. The Pittsburgh Sleep Quality Index (PSQI) was significantly associated with actigraphically measured daytime sleeping (P = .0038), but the Epworth Sleepiness Scale (ESS) was not (P = .49). Neither the ESS nor the PSQI achieved sufficient sensitivity and specificity to be used as a screening tool for excessive daytime sleeping. Two additional models using items from these questionnaires were not significantly associated with the outcome. In an older PAR population, self-report items from existing sleep questionnaires do not identify excessive daytime sleeping. Therefore we recommend objective measures for the evaluation of excessive daytime sleeping as well as further research to identify new self-report items that may be more applicable in PAR populations. Copyright © 2012 American Medical Directors Association, Inc. Published by Elsevier Inc. All rights reserved.

  7. Drivers' misjudgement of vigilance state during prolonged monotonous daytime driving.

    PubMed

    Schmidt, Eike A; Schrauf, Michael; Simon, Michael; Fritzsche, Martin; Buchner, Axel; Kincses, Wilhelm E

    2009-09-01

    To investigate the effects of monotonous daytime driving on vigilance state and particularly the ability to judge this state, a real road driving study was conducted. To objectively assess vigilance state, performance (auditory reaction time) and physiological measures (EEG: alpha spindle rate, P3 amplitude; ECG: heart rate) were recorded continuously. Drivers judged sleepiness, attention to the driving task and monotony retrospectively every 20 min. Results showed that prolonged daytime driving under monotonous conditions leads to a continuous reduction in vigilance. Towards the end of the drive, drivers reported a subjectively improved vigilance state, which was contrary to the continued decrease in vigilance as indicated by all performance and physiological measures. These findings indicate a lack of self-assessment abilities after approximately 3h of continuous monotonous daytime driving.

  8. Chronotype, Light Exposure, Sleep, and Daytime Functioning in High School Students Attending Morning or Afternoon School Shifts: An Actigraphic Study.

    PubMed

    Martin, Jeanne Sophie; Gaudreault, Michael M; Perron, Michel; Laberge, Luc

    2016-04-01

    Adolescent maturation is associated with delays of the endogenous circadian phase. Consequently, early school schedules may lead to a mismatch between internal and external time, which can be detrimental to adolescent sleep and health. In parallel, chronotype is known to play a role in adolescent health; evening chronotype adolescents are at higher risk for sleep problems and lower academic achievement. In the summer of 2008, Kénogami High School (Saguenay, Canada) was destroyed by fire. Kénogami students were subsequently relocated to Arvida High School (situated 5.3 km away) for the 2008-2009 academic year. A dual school schedule was implemented, with Arvida students attending a morning schedule (0740-1305 h) and Kénogami students an afternoon schedule (1325-1845 h). This study aimed to investigate the effects of such school schedules and chronotype on sleep, light exposure, and daytime functioning. Twenty-four morning and 33 afternoon schedule students wore an actigraph during 7 days to measure sleep and light exposure. Academic achievement was obtained from school. Subjects completed validated questionnaires on daytime sleepiness, psychological distress, social rhythms, school satisfaction, alcohol, and chronotype. Overall, afternoon schedule students had longer sleep duration, lower sleepiness, and lower light exposure than morning schedule students. Evening chronotypes (E-types) reported higher levels of sleepiness than morning chronotypes (M-types) in both morning and afternoon schedules. Furthermore, M-types attending the morning schedule reported higher sleepiness than M-types attending the afternoon schedule. No difference was found between morning and afternoon schedule students with regard to academic achievement, psychological distress, social rhythms, school satisfaction, and alcohol consumption. However, in both schedules, M-type had more regular social rhythms and lower alcohol consumption. In summary, this study emphasizes that an early school

  9. Definitions of sleeplessness in children with attention-deficit hyperactivity disorder (ADHD): implications for mothers' mental state, daytime sleepiness and sleep-related cognitions.

    PubMed

    Montgomery, P; Wiggs, L

    2015-01-01

    Sleep disturbances are common in children with attention-deficit hyperactivity disorder (ADHD). Sleeplessness is frequently reported although results are inconsistent perhaps because different definitions for it are applied. This study looked at maternal functioning and child objective sleep patterns in relation to different definitions of sleeplessness in children with ADHD. The study included 45 children (aged 3-14 years) with ADHD and their mothers. Sleeplessness was defined according to: (i) yes/no report of whether mothers thought their children had a problem with sleeplessness (Maternal definition MD) and (ii) mothers' responses to a quantitative standardized questionnaire (Quantitative definition QD) designed to detect the frequency and duration of parent-reported problems with settling, night waking and early waking. Objective sleep patterns were also assessed by means of actigraphy. Maternal mental health, daytime sleepiness and cognitions related to child sleep were assessed by questionnaire. Both definitions appeared to tap similar although slightly different constructs. There were no group differences in objective sleep patterns. Maternal mental health was found to be significantly worse in the mothers who considered their child to be sleepless (MD) (P < 0.025), but not in those mothers whose child was found to be sleepless according to the standardized criteria (QD). Maternal sleepiness did not differ between groups. For both categories of sleepless children (MD and QD), the mothers had significantly more doubts about their competency as a parent (P < 0.01 and P < 0.025, respectively) compared to mothers of children without sleeplessness. Two different maternal assessments of child sleeplessness in children with ADHD may assess subtly different constructs, but both may provide useful information about potential problems across the family. © 2014 John Wiley & Sons Ltd.

  10. Self-reported body silhouette trajectories across the lifespan and excessive daytime sleepiness in adulthood: a retrospective analysis. The Paris Prospective Study III

    PubMed Central

    Tafflet, Muriel; Charles, Marie-Aline; Thomas, Frédérique; Boutouyrie, Pierre; Guibout, Catherine; Haba-Rubio, José; Périer, Marie Cécile; Pannier, Bruno; Marques-Vidal, Pedro; Jouven, Xavier; Empana, Jean-Philippe

    2018-01-01

    Objectives Excessive daytime sleepiness (EDS) is a common sleep complaint in the population and is increasingly recognised as deleterious for health. Simple and sensitive tools allowing identifying individuals at greater risk of EDS would be of public health importance. Hence, we determined trajectories of body silhouette from early childhood to adulthood and evaluated their association with EDS in adulthood. Design A retrospective analysis in a prospective community-based study. Participants 6820 men and women self-reported their silhouette at ages 8, 15, 25, 35 and 45 using the body silhouettes proposed by Stunkard et al. EDS was defined by an Epworth Sleepiness Scale score ≥11. Main outcome measure Presence of EDS in adulthood. Results The study population comprised 6820 participants (mean age 59.8 years, 61.1% men). Five distinct body silhouettes trajectories over the lifespan were identified: 31.9% ‘lean stable’, 11.1% ‘lean increase’, 16.1% ‘lean-marked increase’, 32.5% ‘moderate stable’ and 8.4% ‘heavy stable’. Subjects with a ‘heavy-stable’ trajectory (OR 1.24, 95% CI 0.94 to 1.62) and those with a ‘lean-marked increase’ trajectory (OR 1.46, 95% CI 1.18 to 1.81) were more likely to have EDS when compared with the ‘lean-stable’ group after adjusting for confounding. Further adjustment for birth weight strengthened the magnitude of the ORs. Conclusion Increasing body silhouette and to a lesser extent constantly high body silhouette trajectory from childhood to adulthood are associated with increased likelihood of EDS, independently of major confounding variables. Trial registration number NCT00741728; Pre-results. PMID:29593025

  11. Age-Related Reduction in Daytime Sleep Propensity and Nocturnal Slow Wave Sleep

    PubMed Central

    Dijk, Derk-Jan; Groeger, John A.; Stanley, Neil; Deacon, Stephen

    2010-01-01

    Objective: To investigate whether age-related and experimental reductions in SWS and sleep continuity are associated with increased daytime sleep propensity. Methods: Assessment of daytime sleep propensity under baseline conditions and following experimental disruption of SWS. Healthy young (20-30 y, n = 44), middle-aged (40-55 y, n = 35) and older (66-83 y, n = 31) men and women, completed a 2-way parallel group study. After an 8-h baseline sleep episode, subjects were randomized to 2 nights with selective SWS disruption by acoustic stimuli, or without disruption, followed by 1 recovery night. Objective and subjective sleep propensity were assessed using the Multiple Sleep Latency Test (MSLT) and the Karolinska Sleepiness Scale (KSS). Findings: During baseline sleep, SWS decreased (P < 0.001) and the number of awakenings increased (P < 0.001) across the 3 age groups. During the baseline day, MSLT values increased across the three age groups (P < 0.0001) with mean values of 8.7min (SD: 4.5), 11.7 (5.1) and 14.2 (4.1) in the young, middle-aged, and older adults, respectively. KSS values were 3.7 (1.0), 3.2 (0.9), and 3.4 (0.6) (age-group: P = 0.031). Two nights of SWS disruption led to a reduction in MSLT and increase in KSS in all 3 age groups (SWS disruption vs. control: P < 0.05 in all cases). Conclusions: Healthy aging is associated with a reduction in daytime sleep propensity, sleep continuity, and SWS. In contrast, experimental disruption of SWS leads to an increase in daytime sleep propensity. The age-related decline in SWS and reduction in daytime sleep propensity may reflect a lessening in homeostatic sleep requirement. Healthy older adults without sleep disorders can expect to be less sleepy during the daytime than young adults. Citation: Dijk DJ; Groeger JA; Stanley N; Deacon S. Age-related reduction in daytime sleep propensity and nocturnal slow wave sleep. SLEEP 2010;33(2):211-223. PMID:20175405

  12. Association of current work and sleep situations with excessive daytime sleepiness and medical incidents among Japanese physicians.

    PubMed

    Kaneita, Yoshitaka; Ohida, Takashi

    2011-10-15

    The aim of the present study was to clarify the current work and sleep situations of physicians in Japan and to clarify the association between these situations and excessive daytime sleepiness as well as medical incidents. A self-administered questionnaire survey was conducted among the members of the Japan Medical Association in 2008. The randomly selected subjects comprised 3,000 male physicians and 1,500 female physicians. Valid responses were obtained from 3,486 physicians (2,298 men and 1,188 women). Mean sleep duration was 6 h 36 min for men and 6 h 8 min for women. The prevalence of lack of rest due to sleep deprivation was 30.4% among men and 36.6% among women; the prevalence of insomnia was 21.0% and 18.1%, respectively; and the prevalence of EDS was 3.5%. The adjusted odds ratio for EDS was high for physicians who reported short sleep duration, lack of rest due to sleep deprivation, and a high frequency of on-call/overnight work. Physicians who had experienced a medical incident within the previous one month accounted for 19.0% of participants. The adjusted odds ratio for medical incidents was high for those subjected to long working hours, high frequency of on-call/overnight works, lack of rest due to sleep deprivation, and insomnia. In order to facilitate optimal health management for physicians as well as securing medical safety, it is important to fully consider the work and sleep situations of physicians.

  13. A multi-step pathway connecting short sleep duration to daytime somnolence, reduced attention, and poor academic performance: an exploratory cross-sectional study in teenagers.

    PubMed

    Perez-Lloret, Santiago; Videla, Alejandro J; Richaudeau, Alba; Vigo, Daniel; Rossi, Malco; Cardinali, Daniel P; Perez-Chada, Daniel

    2013-05-15

    A multi-step causality pathway connecting short sleep duration to daytime somnolence and sleepiness leading to reduced attention and poor academic performance as the final result can be envisaged. However this hypothesis has never been explored. To explore consecutive correlations between sleep duration, daytime somnolence, attention levels, and academic performance in a sample of school-aged teenagers. We carried out a survey assessing sleep duration and daytime somnolence using the Pediatric Daytime Sleepiness Scale (PDSS). Sleep duration variables included week-days' total sleep time, usual bedtimes, and absolute weekday to-weekend sleep time difference. Attention was assessed by d2 test and by the coding subtest from the WISC-IV scale. Academic performance was obtained from literature and math grades. Structural equation modeling was used to assess the independent relationships between these variables, while controlling for confounding effects of other variables, in one single model. Standardized regression weights (SWR) for relationships between these variables are reported. Study sample included 1,194 teenagers (mean age: 15 years; range: 13-17 y). Sleep duration was inversely associated with daytime somnolence (SWR = -0.36, p < 0.01) while sleepiness was negatively associated with attention (SWR = -0.13, p < 0.01). Attention scores correlated positively with academic results (SWR = 0.18, p < 0.01). Daytime somnolence correlated negatively with academic achievements (SWR = -0.16, p < 0.01). The model offered an acceptable fit according to usual measures (RMSEA = 0.0548, CFI = 0.874, NFI = 0.838). A Sobel test confirmed that short sleep duration influenced attention through daytime somnolence (p < 0.02), which in turn influenced academic achievements through reduced attention (p < 0.002). Poor academic achievements correlated with reduced attention, which in turn was related to daytime somnolence. Somnolence correlated with short sleep duration.

  14. The relationship between sleep quality, depression, and anxiety in patients with epilepsy and suicidal ideation.

    PubMed

    Wigg, Cristina Maria Duarte; Filgueiras, Alberto; Gomes, Marleide da Mota

    2014-05-01

    The relationships among suicidal ideation, sleep, depression, anxiety, and effects on epilepsy require more research. The aim of this study was to estimate the prevalence of suicidal ideation in outpatients with epilepsy, and relate this to sleep quality, daytime sleepiness, depression, and anxiety. Ninety-eight non-selected patients were evaluated. The subjects were classified as "suicidal ideators" or "non-ideators", based on their response to item 9 of the Beck Depression Inventory. The prevalence of suicidal ideation was 13.3% (χ2=50.46, p<0.001). The differences between cases with or without suicidal ideation were statistically significant in relation to sleep quality (p=0.005) and symptoms of depression (p=0.001) and anxiety (p=0.002). Our results revealed that depression and anxiety were associated with sleep quality, daytime sleepiness, and suicidal ideation and that depression and sleep disturbance were good predictors of suicide in subjects with epilepsy.

  15. Sleepiness and health in midlife women: results of the National Sleep Foundation's 2007 Sleep in America poll.

    PubMed

    Chasens, Eileen R; Twerski, Sarah R; Yang, Kyeongra; Umlauf, Mary Grace

    2010-01-01

    The 2007 Sleep in America poll, a random-sample telephone survey, provided data for this study of sleep in community-dwelling women aged 40 to 60 years. The majority of the respondents were post- or perimenopausal, overweight, married or living with someone, and reported good health. A subsample (20%) reported sleepiness that consistently interfered with daily life; the sleepy subsample reported more symptoms of insomnia, restless legs syndrome, obstructive sleep apnea, depression and anxiety, as well as more problems with health-promoting behaviors, drowsy driving, job performance, household duties, and personal relationships. Hierarchical regression showed that sleepiness along with depressive symptoms, medical comorbidities, obesity, and lower education were associated with poor self-rated health, whereas menopause status (pre-, peri- or post-) was not. These results suggest that sleep disruptions and daytime sleepiness negatively affect the daily life of midlife women.

  16. Ecological momentary assessment of fatigue, sleepiness, and exhaustion in ESKD.

    PubMed

    Abdel-Kader, Khaled; Jhamb, Manisha; Mandich, Lee Anne; Yabes, Jonathan; Keene, Robert M; Beach, Scott; Buysse, Daniel J; Unruh, Mark L

    2014-02-06

    Many patients on maintenance dialysis experience significant sleepiness and fatigue. However, the influence of the hemodialysis (HD) day and circadian rhythms on patients' symptoms have not been well characterized. We sought to use ecological momentary assessment to evaluate day-to-day and diurnal variability of fatigue, sleepiness, exhaustion and related symptoms in thrice-weekly maintenance HD patients. Subjects used a modified cellular phone to access an interactive voice response system that administered the Daytime Insomnia Symptom Scale (DISS). The DISS assessed subjective vitality, mood, and alertness through 19 questions using 7- point Likert scales. Subjects completed the DISS 4 times daily for 7 consecutive days. Factor analysis was conducted and a mean composite score of fatigue-sleepiness-exhaustion was created. Linear mixed regression models (LMM) were used to examine the association of time of day, dialysis day and fatigue, sleepiness, and exhaustion composite scores. The 55 participants completed 1,252 of 1,540 (81%) possible assessments over the 7 day period. Multiple symptoms related to mood (e.g., feeling sad, feeling tense), cognition (e.g., difficulty concentrating), and fatigue (e.g., exhaustion, feeling sleepy) demonstrated significant daily and diurnal variation, with higher overall symptom scores noted on hemodialysis days and later in the day. In factor analysis, 4 factors explained the majority of the observed variance for DISS symptoms. Fatigue, sleepiness, and exhaustion loaded onto the same factor and were highly intercorrelated. In LMM, mean composite fatigue-sleepiness-exhaustion scores were associated with dialysis day (coefficient and 95% confidence interval [CI] 0.21 [0.02 - 0.39]) and time of day (coefficient and 95% CI 0.33 [0.25 - 0.41]. Observed associations were minimally affected by adjustment for demographics and common confounders. Maintenance HD patients experience fatigue-sleepiness-exhaustion symptoms that demonstrate

  17. Trajectories of sleep disturbance and daytime sleepiness in women before and after surgery for breast cancer.

    PubMed

    Van Onselen, Christina; Paul, Steven M; Lee, Kathryn; Dunn, Laura; Aouizerat, Bradley E; West, Claudia; Dodd, Marylin; Cooper, Bruce; Miaskowski, Christine

    2013-02-01

    Sleep disturbance is a problem for oncology patients. To evaluate how sleep disturbance and daytime sleepiness (DS) changed from before to six months following surgery and whether certain characteristics predicted initial levels and/or the trajectories of these parameters. Patients (n=396) were enrolled prior to surgery and completed monthly assessments for six months following surgery. The General Sleep Disturbance Scale was used to assess sleep disturbance and DS. Using hierarchical linear modeling, demographic, clinical, symptom, and psychosocial adjustment characteristics were evaluated as predictors of initial levels and trajectories of sleep disturbance and DS. All seven General Sleep Disturbance Scale scores were above the cutoff for clinically meaningful levels of sleep disturbance. Lower performance status; higher comorbidity, attentional fatigue, and physical fatigue; and more severe hot flashes predicted higher preoperative levels of sleep disturbance. Higher levels of education predicted higher sleep disturbance scores over time. Higher levels of depressive symptoms predicted higher preoperative levels of sleep disturbance, which declined over time. Lower performance status; higher body mass index; higher fear of future diagnostic tests; not having had sentinel lymph node biopsy; having had an axillary lymph node dissection; and higher depression, physical fatigue, and attentional fatigue predicted higher DS prior to surgery. Higher levels of education, not working for pay, and not having undergone neo-adjuvant chemotherapy predicted higher DS scores over time. Sleep disturbance is a persistent problem for patients with breast cancer. The effects of interventions that can address modifiable risk factors need to be evaluated. Copyright © 2013 U.S. Cancer Pain Relief Committee. Published by Elsevier Inc. All rights reserved.

  18. Sleep disturbance and depressive affect in patients treated with haemodialysis.

    PubMed

    Maung, Stephanie; Sara, Ammar El; Cohen, Danielle; Chapman, Cherylle; Saggi, Subodh; Cukor, Daniel

    2017-03-01

    Sleep disorders and depression are prevalent conditions in patients with end-stage kidney disease. These co-morbidities have significant overlap and compounded morbidity and mortality burden. This overlap presents challenges to optimal clinical assessment and treatment. The goal of this study was to assess the prevalence of sleep disturbance in patients on maintenance haemodialysis, and to assess the impact of depressive affect. This was a single-site, single group, cross-sectional study of 69 English-speaking patients undergoing maintenance haemodialysis. Self-reported assessments included those of sleep quality (Pittsburgh Sleep Quality Index), depression (Beck Depression Inventory), daytime sleepiness (Epworth's Sleepiness Scale), a dialysis-specific sleep questionnaire, and standard laboratory values. No objective sleep information was collected. All participants were well dialysed, and represented all four daily shifts. Fifty-eight per cent reported clinically significant sleep difficulty, with elevated yet sub-threshold daytime sleepiness. Mean depressive affect was also elevated, yet sub-diagnostic and was positively correlated with increased age. Participants scoring above the diagnostic threshold for depression had significantly more disturbed sleep quality, more daytime sleepiness and had more problems sleeping due to restless leg syndrome than people with minimal depressive affect. Poor sleep quality is prevalent in patients on maintenance haemodialysis, and is associated with increased daytime sleepiness. Depression further compounds this relationship, and is significantly associated with increased daytime sleepiness and restless leg syndrome. © 2016 European Dialysis and Transplant Nurses Association/European Renal Care Association.

  19. Continuing to drive while sleepy: the influence of sleepiness countermeasures, motivation for driving sleepy, and risk perception.

    PubMed

    Watling, Christopher N; Armstrong, Kerry A; Obst, Patricia L; Smith, Simon S

    2014-12-01

    Driver sleepiness is a major contributor to road crashes. The current study sought to examine the association between perceptions of effectiveness of six sleepiness countermeasures and their relationship with self-reports of continuing to drive while sleepy among 309 drivers after controlling for the influence of age, sex, motivation for driving sleepy, and risk perception of sleepy driving. The results demonstrate that the variables of age, sex, motivation, and risk perception were significantly associated with self-reports of continuing to drive while sleepy and only one countermeasure was associated with self-reports of continuing to drive while sleepy. Further, it was found that age differences in self-reports of continuing to drive while sleepy was mediated by participants' motivation and risk perception. These findings highlight modifiable factors that could be focused on with interventions that seek to modify drivers' attitudes and behaviours of driving while sleepy. Copyright © 2014 Elsevier Ltd. All rights reserved.

  20. [Daytime consequences of insomnia complaints in the French general population].

    PubMed

    Ohayon, M M; Lemoine, P

    2004-01-01

    Insomnia is a frequent symptom in the general population; numerous studies have proven this. In the past years, classifications have gradually given more emphasis to daytime repercussions of insomnia and to their consequences on social and cognitive functioning. They are now integrated in the definition of insomnia and are used to quantify its severity. If the daytime consequences of insomnia are well known at the clinical level, there are few epidemiological data on this matter. The aim of this study was to assess the daytime repercussions of insomnia complaints in the general population of France. A representative sample (n=5,622) aged 15 or older was surveyed by telephone with the help of the sleep-EVAL expert system, a computer program specially designed to evaluate sleep disorders and to manage epidemiological investigations. Interviews have been completed for 80.8% of the solicited subjects (n=5,622). The variables considered comprised insomnia and its daytime repercussions on cognitive functioning, affective tone, daytime sleepiness and diurnal fatigue. Insomnia was found in 18.6% of the sample. The prevalence was higher in women (22.4%) than in men (14.5%, p<0.001) with a relative risk of 1.7 (95% confidence interval 1.5 to 2) and was twice more frequent for subjects 65 years of age or older compared to subjects younger than 45 years. Approximately 30% of subjects reporting insomnia had difficulties initiating sleep. Nearly 75% of insomnia complainers reported having a disrupted sleep or waking up too early in the morning and about 40% said they had a non-restorative sleep. Repercussions on daytime functioning were reported by most insomnia subjects (67%). Repercussions on cognitive functioning changed according age, number of insomnia symptoms and the use of a psychotropic medication. A decreased efficiency was more likely to be reported by subjects between 15 and 44 years of age (OR: 2.9), those using a psychotropic (OR: 1.5), those reporting at least

  1. Examining signs of driver sleepiness, usage of sleepiness countermeasures and the associations with sleepy driving behaviours and individual factors.

    PubMed

    Watling, Christopher N; Armstrong, Kerry A; Radun, Igor

    2015-12-01

    The impairing effect from sleepiness is a major contributor to road crashes. The ability of a sleepy driver to perceive their level of sleepiness is an important consideration for road safety as well as the type of sleepiness countermeasure used by drivers as some sleepiness countermeasures are more effective than others. The aims of the current study were to determine the extent that the signs of driver sleepiness were associated with sleepy driving behaviours, as well as determining which individual factors (demographic, work, driving, and sleep-related factors) were associated with using a roadside or in-vehicle sleepiness countermeasure. A sample of 1518 Australian drivers from the Australian State of New South Wales and the neighbouring Australian Capital Territory took part in the study. The participants' experiences with the signs of sleepiness were reasonably extensive. A number of the early signs of sleepiness (e.g., yawning, frequent eye blinks) were related with continuing to drive while sleepy, with the more advanced signs of sleepiness (e.g., difficulty keeping eyes open, dreamlike state of consciousness) associated with having a sleep-related close call. The individual factors associated with using a roadside sleepiness countermeasure included age (being older), education (tertiary level), difficulties getting to sleep, not continuing to drive while sleepy, and having experienced many signs of sleepiness. The results suggest that these participants have a reasonable awareness and experience with the signs of driver sleepiness. Factors related to previous experiences with sleepiness were associated with implementing a roadside countermeasure. Nonetheless, the high proportions of drivers performing sleepy driving behaviours suggest that concerted efforts are needed with road safety campaigns regarding the dangers of driving while sleepy. Copyright © 2015 Elsevier Ltd. All rights reserved.

  2. The Influence of Sleep Disorders on Voice Quality.

    PubMed

    Rocha, Bruna Rainho; Behlau, Mara

    2017-09-19

    To verify the influence of sleep quality on the voice. Descriptive and analytical cross-sectional study. Data were collected by an online or printed survey divided in three parts: (1) demographic data and vocal health aspects; (2) self-assessment of sleep and vocal quality, and the influence that sleep has on voice; and (3) sleep and voice self-assessment inventories-the Epworth Sleepiness Scale (ESS), the Pittsburgh Sleep Quality Index (PSQI), and the Voice Handicap Index reduced version (VHI-10). A total of 862 people were included (493 women, 369 men), with a mean age of 32 years old (maximum age of 79 and minimum age of 18 years old). The perception of the influence that sleep has on voice showed a difference (P < 0.050) between measures of sleep quality and vocal self-assessment. There were higher scores on the ESS, PSQI, and VHI-10 protocols if sleep and vocal self-assessment were poor. The results indicate that the greater the effect that sleep has on voice, the greater the perceived voice handicap. The aspects that influence a voice handicap are vocal self-assessment, ESS total score, and self-assessment of the influence that sleep has on voice. The absence of daytime sleepiness is a protective factor (odds ratio [OR] > 1) against perceived voice handicap; the presence of daytime sleepiness is a damaging factor (OR < 1). Sleep quality influences voice. Perceived poor sleep quality is related to perceived poor vocal quality. Individuals with a voice handicap observe a greater influence of sleep on voice than those without. Copyright © 2017 The Voice Foundation. Published by Elsevier Inc. All rights reserved.

  3. Pharmacotherapy for residual excessive sleepiness and cognition in CPAP-treated patients with obstructive sleep apnea syndrome: A systematic review and meta-analysis.

    PubMed

    Avellar, Ariane B C C; Carvalho, Luciane B C; Prado, Gilmar F; Prado, Lucila B F

    2016-12-01

    Pharmacotherapy has been used as an adjunct to CPAP for treatment of residual excessive sleepiness in patients with a diagnosis of obstructive sleep apnea syndrome (OSAS). However, no studies with a high level of evidence have been conducted to support this practice and confirm its effectiveness. We conducted a meta-analysis to summarize and quantify the effects of pharmacological treatment in adults with OSAS who experience residual excessive sleepiness despite adequate CPAP use. We reviewed clinical trials that compared medications to placebo and evaluated the outcomes residual excessive sleepiness, cognition, and quality of life, as well as treatment effectiveness and safety. The MEDLINE, EMBASE, LILACS, Cochrane Central Register of Controlled Trials - CENTRAL, and PsycINFO electronic databases were searched using highly sensitive search strategies. Trials were only included if measures were taken to ensure effective CPAP treatment. Eight randomized clinical trials were included. Pharmacotherapy with modafinil and armodafinil led to improvement of excessive daytime sleepiness, attention/alertness, and clinical condition as measured with the CGI-C. No improvements in quality of life or other cognitive domains (including memory, executive function, and language) could be confirmed. Pharmacotherapy did not cause any severe adverse effects, but was associated with significant dropout rates as compared with placebo. In conclusion, although our results demonstrate the effectiveness of pharmacological treatment as an adjunct to CPAP, further investigation is necessary to improve confidence in its effects. Many findings on the impact of pharmacotherapy on cognition and quality of life were evaluated through analysis of single studies, with heterogeneity in tests and absence of standardization, which reduced certainty as to whether actual improvement occurred in these outcomes. Copyright © 2015 Elsevier Ltd. All rights reserved.

  4. The role of pseudoephedrine on daytime somnolence in patients suffering from perennial allergic rhinitis (PAR).

    PubMed

    Sherkat, Amir A; Sardana, Niti; Safaee, Sahar; Lehman, Erik B; Craig, Timothy J

    2011-02-01

    Allergic rhinitis is one of several inflammatory diseases affecting the nasal mucosa. Cellular inflammation of nasal mucosa is a hallmark of this disease and is characterized by the accumulation of eosinophils and the release of various chemical messengers such as chemokines, cytokines, and histamine. This inflammation of the nose leads to nasal congestion and a reduction in sleep quality, resulting in daytime somnolence. Drugs that significantly reduce the symptoms of nasal congestion also may help in alleviating sleep-related symptoms of allergic rhinitis. Pseudoephedrine is a sympathomimetic amine that is indicated for treatment of nasal congestion associated with allergic rhinitis. Despite relieving nasal congestion, we speculated that, because of pseudoephedrine's well-known stimulant profile, sleep would not be improved. Fourteen subjects who met the inclusion criteria were enrolled into a double-blind, placebo-controlled, randomized study to either pseudoephedrine or placebo once per day in the morning, using the traditional crossover design. Skin testing test was performed to ensure a positive response to a relevant perennial allergen and a negative response to a seasonal allergen. Several questionnaires were used to evaluate the patients' sleep-related symptoms, allergic rhinitis symptoms, and quality of life. Our results showed that pseudoephedrine did not have a positive or negative effect on quality of sleep, daytime sleepiness, or daytime fatigue as compared with placebo. Pseudoephedrine did show a statistical significance in improving stuffy nose (P = .0172). With respect to quality of life, pseudoephedrine led to a statistically significant decrease in intimate relationships and sexual activity as compared with the placebo group (P = .0310). Our research suggests that sleep quality is not significantly affected by pseudoephedrine. As expected, congestion is reduced, but side effects such as a decline of intimate relationships and sexual activity may

  5. Concurrent respiratory resistance training and changes in respiratory muscle strength and sleep in an individual with spinal cord injury: case report

    PubMed Central

    Russian, Chris; Litchke, Lyn; Hudson, John

    2011-01-01

    Context Quality sleep possesses numerous benefits to normal nighttime and daytime functioning. High-level spinal cord injury (SCI) often impacts the respiratory muscles that can lead to poor respiratory function during sleep and negatively affect sleep quality. The impact of respiratory muscle training (RMT) on sleep quality, as assessed by overnight polysomnography (PSG), is yet to be determined among the spinal cord-injured population. This case report describes the effects of 10 weeks of RMT on the sleep quality of a 38-year-old male with cervical SCI. Methods Case report. Findings/results The subject completed overnight PSG, respiratory muscle strength assessment, and subjective sleepiness assessment before and after 10 weeks of RMT. The post-test results indicated improvements in sleep quality (e.g. fewer electroencephalographic (EEG) arousals during sleep) and daytime sleepiness scores following RMT. Conclusion/clinical relevance Respiratory activity has been proven to impact EEG arousal activity during sleep. Arousals during sleep lead to a fragmented sleeping pattern and affect sleep quality and daytime function. Our subject presented with a typical sleep complaint of snoring and excessive sleepiness. The subject's pre-test PSG demonstrated a large number of arousals during sleep. It is important for all individuals complaining of problems during sleep or daytime problems associated with sleep (i.e. excessive daytime sleepiness) to seek medical attention and proper evaluation. PMID:21675365

  6. Effects of Instant Messaging on School Performance in Adolescents.

    PubMed

    Grover, Karan; Pecor, Keith; Malkowski, Michael; Kang, Lilia; Machado, Sasha; Lulla, Roshni; Heisey, David; Ming, Xue

    2016-06-01

    Instant messaging may compromise sleep quality and school performance in adolescents. We aimed to determine associations between nighttime messaging and daytime sleepiness, self-reported sleep parameters, and/or school performance. Students from 3 high schools in New Jersey completed anonymous questionnaires assessing sleep duration, daytime sleepiness, messaging habits, and academic performance. Of the 2,352 students sampled, 1,537 responses were contrasted among grades, sexes, and messaging duration, both before and after lights out. Students who reported longer duration of messaging after lights out were more likely to report a shorter sleep duration, higher rate of daytime sleepiness, and poorer academic performance. Messaging before lights out was not associated with higher rates of daytime sleepiness or poorer academic performance. Females reported more messaging, more daytime sleepiness, and better academic performance than males. There may be an association between text messaging and school performance in this cohort of students. © The Author(s) 2016.

  7. Factors Associated with Sleep Quality in Maxillectomy Patients.

    PubMed

    Li, Na; Otomaru, Takafumi; Said, Mohamed Moustafa; Kanazaki, Ayako; Yeerken, Yesiboli; Taniguchi, Hisashi

    To investigate factors affecting sleep quality in maxillectomy patients after prosthetic rehabilitation and to determine the association between defect status and sleep quality. A total of 57 patients participated in this study. Sleep quality, general health, and oral health-related quality of life (OHRQoL) were evaluated. Of the total sample, 89% had poor sleep quality. Early morning awakening and daytime sleepiness were the most common complaints. Defect status and the extent of neck dissection could affect sleep quality in these patients. Improvement of OHRQoL in patients with dentomaxillary prostheses may help improve sleep.

  8. A Multi-Step Pathway Connecting Short Sleep Duration to Daytime Somnolence, Reduced Attention, and Poor Academic Performance: An Exploratory Cross-Sectional Study in Teenagers

    PubMed Central

    Perez-Lloret, Santiago; Videla, Alejandro J.; Richaudeau, Alba; Vigo, Daniel; Rossi, Malco; Cardinali, Daniel P.; Perez-Chada, Daniel

    2013-01-01

    Background: A multi-step causality pathway connecting short sleep duration to daytime somnolence and sleepiness leading to reduced attention and poor academic performance as the final result can be envisaged. However this hypothesis has never been explored. Objective: To explore consecutive correlations between sleep duration, daytime somnolence, attention levels, and academic performance in a sample of school-aged teenagers. Methods: We carried out a survey assessing sleep duration and daytime somnolence using the Pediatric Daytime Sleepiness Scale (PDSS). Sleep duration variables included week-days' total sleep time, usual bedtimes, and absolute weekdayto-weekend sleep time difference. Attention was assessed by d2 test and by the coding subtest from the WISC-IV scale. Academic performance was obtained from literature and math grades. Structural equation modeling was used to assess the independent relationships between these variables, while controlling for confounding effects of other variables, in one single model. Standardized regression weights (SWR) for relationships between these variables are reported. Results: Study sample included 1,194 teenagers (mean age: 15 years; range: 13-17 y). Sleep duration was inversely associated with daytime somnolence (SWR = -0.36, p < 0.01) while sleepiness was negatively associated with attention (SWR = -0.13, p < 0.01). Attention scores correlated positively with academic results (SWR = 0.18, p < 0.01). Daytime somnolence correlated negatively with academic achievements (SWR = -0.16, p < 0.01). The model offered an acceptable fit according to usual measures (RMSEA = 0.0548, CFI = 0.874, NFI = 0.838). A Sobel test confirmed that short sleep duration influenced attention through daytime somnolence (p < 0.02), which in turn influenced academic achievements through reduced attention (p < 0.002). Conclusions: Poor academic achievements correlated with reduced attention, which in turn was related to daytime somnolence. Somnolence

  9. 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

  10. Driver sleepiness on YouTube: A content analysis.

    PubMed

    Hawkins, A N; Filtness, A J

    2017-02-01

    Driver sleepiness is a major contributor to severe crashes and fatalities on our roads. Many people continue to drive despite being aware of feeling tired. Prevention relies heavily on education campaigns as it is difficult to police driver sleepiness. The video sharing social media site YouTube is extremely popular, particularly with at risk driver demographics. Content and popularity of uploaded videos can provide insight into the quality of publicly accessible driver sleepiness information. The purpose of this research was to answer two questions; firstly, how prevalent are driver sleepiness videos on YouTube? And secondly, what are the general characteristics of driver sleepiness videos in terms of (a) outlook on driver sleepiness, (b) tone, (c) countermeasures to driver sleepiness, and, (d) driver demographics. Using a keywords search, 442 relevant videos were found from a five year period (2nd December 2009-2nd December 2014). Tone, outlook, and countermeasure use were thematically coded. Driver demographic and video popularity data also were recorded. The majority of videos portrayed driver sleepiness as dangerous. However, videos that had an outlook towards driver sleepiness being amusing were viewed more often and had more mean per video comments and likes. Humorous videos regardless of outlook, were most popular. Most information regarding countermeasures to deal with driver sleepiness was accurate. Worryingly, 39.8% of videos with countermeasure information contained some kind of ineffective countermeasure. The use of humour to convey messages about the dangers of driver sleepiness may be a useful approach in educational interventions. Copyright © 2015 Elsevier Ltd. All rights reserved.

  11. The effect of intermittent fasting during Ramadan on sleep, sleepiness, cognitive function, and circadian rhythm.

    PubMed

    Qasrawi, Shaden O; Pandi-Perumal, Seithikurippu R; BaHammam, Ahmed S

    2017-09-01

    Studies have shown that experimental fasting can affect cognitive function, sleep, and wakefulness patterns. However, the effects of experimental fasting cannot be generalized to fasting during Ramadan due to its unique characteristics. Therefore, there has been increased interest in studying the effects of fasting during Ramadan on sleep patterns, daytime sleepiness, cognitive function, sleep architecture, and circadian rhythm. In this review, we critically discuss the current research findings in those areas during the month of Ramadan. Available data that controlled for sleep/wake schedule, sleep duration, light exposure, and energy expenditure do not support the notion that Ramadan intermittent fasting increases daytime sleepiness and alters cognitive function. Additionally, recent well-designed studies showed no effect of fasting on circadian rhythms. However, in non-constrained environments that do not control for lifestyle changes, studies have demonstrated sudden and significant delays in bedtime and wake time. Studies that controlled for environmental factors and sleep/wake schedule reported no significant disturbances in sleep architecture. Nevertheless, several studies have consistently reported that the main change in sleep architecture during fasting is a reduction in the proportion of REM sleep.

  12. Risk factors for fatigue in patients with epilepsy.

    PubMed

    Yan, Song; Wu, Yuanbin; Deng, Yanchun; Liu, Yonghong; Zhao, Jingjing; Ma, Lei

    2016-11-01

    Fatigue is highly prevalent in patients with epilepsy and has a major impact on quality of life, but little data is available on its effects and management in epilepsy. To identify the incidence and risk factors of fatigue in patients with epilepsy, 105 epilepsy patients (45 women and 60 men) were enrolled in our study. Demographic and clinical data were collected and psychological variables including fatigue, sleep quality, excess daytime sleepiness, anxiety, and depression were measured by Fatigue Severity Scale, Pittsburgh Sleep Quality Index, Epworth Sleepiness Scale, and Hospital Anxiety and Depression Scale, respectively. Of 105 patients, 29.5% exhibited fatigue (FSS score ⩾4). We found no correlation between the occurrence of fatigue and any of our demographic or clinical variables. Fatigue is correlated with low sleep quality, anxiety, and depression, but not with excess daytime sleepiness. Thus, we concluded that fatigue is highly prevalent in patients with epilepsy, and that low sleep quality, anxiety, and depression are significantly correlated with fatigue in epileptics, while excess daytime sleepiness not. Copyright © 2016. Published by Elsevier Ltd.

  13. Correlating Subjective and Objective Sleepiness: Revisiting the Association Using Survival Analysis

    PubMed Central

    Aurora, R. Nisha; Caffo, Brian; Crainiceanu, Ciprian; Punjabi, Naresh M.

    2011-01-01

    clinical practice. Given the ease of administering the ESS, it represents a relatively simple and cost-effective method for identifying individuals at risk for daytime sleepiness. Citation: Aurora RN; Caffo B; Crainiceanu C; Punjabi NM. Correlating subjective and objective sleepiness: revisiting the association using survival analysis. SLEEP 2011;34(12):1707-1714. PMID:22131609

  14. Sleep and sleepiness during an ultra long-range flight operation between the Middle East and United States.

    PubMed

    Holmes, Alexandra; Al-Bayat, Soha; Hilditch, Cassie; Bourgeois-Bougrine, Samira

    2012-03-01

    This study provides a practical example of fatigue risk management in aviation. The sleep and sleepiness of 44 pilots (11 trips × 4 pilot crew) working an ultra long-range (ULR; flight time >16 h) round-trip operation between Doha and Houston was assessed. Sleep was assessed using activity monitors and self-reported sleep diaries. Mean Karolinska Sleepiness Scores (KSS) for climb and descent did not exceed 5 ("neither alert nor sleepy"). Mean daily sleep duration was maintained above 6.3h throughout the operation. During in-flight rest periods, 98% of pilots obtained sleep and sleepiness was subsequently reduced. On layover (49.5h) crew were advised to sleep on Doha or Universal Co-ordinated Time (UTC), but 64% slept during the local (social) night time. Pilots originating from regions with a siesta culture were more likely to nap and made particularly effective use of their daytime in-flight rest periods. The results indicate that the operation is well designed from a fatigue management perspective. Copyright © 2011 Elsevier Ltd. All rights reserved.

  15. [Sleep quality and hormone levels in the morning and evening hours under chemical pollution].

    PubMed

    Budkevich, R O; Budkevich, E V

    To evaluate self-assessment of sleep and the level of hormones in the morning and evening in chemical pollution conditions. Three hundred adolescent and adult men living in the regions with low and high levels of chemical pollution were examined using questionnaires for self-assessment of quality of sleep, sleep hygiene, daytime sleepiness. Levels of cortisol and testosterone in the saliva were determined in the morning and evening hours by ELISA. In areas with low pollution level, there were normal changes in hormone levels with an increase in the morning and decrease in the evening. In high pollution conditions, the average levels of hormones increased, the morning-evening gradient disappeared. These conditions were also associated with an increase in daytime sleepiness and disturbances in the sleep-wake cycle and the endocrine regulation system that indicate the possibility of the development of internal desynchronosis.

  16. Sleep Patterns and Daytime Sleepiness in Adolescents and Young Adults with Williams Syndrome

    ERIC Educational Resources Information Center

    Goldman, S. E.; Malow, B. A.; Newman, K. D.; Roof, E.; Dykens, E. M.

    2009-01-01

    Background: Sleep disorders are common in individuals with neurodevelopmental disorders and may adversely affect daytime functioning. Children with Williams syndrome have been reported to have disturbed sleep; however, no studies have been performed to determine if these problems continue into adolescence and adulthood. Methods: This study…

  17. Aerobic exercise improves self-reported sleep and quality of life in older adults with insomnia.

    PubMed

    Reid, Kathryn J; Baron, Kelly Glazer; Lu, Brandon; Naylor, Erik; Wolfe, Lisa; Zee, Phyllis C

    2010-10-01

    To assess the efficacy of moderate aerobic physical activity with sleep hygiene education to improve sleep, mood and quality of life in older adults with chronic insomnia. Seventeen sedentary adults aged >or=55 years with insomnia (mean age 61.6 [SD±4.3] years; 16 female) participated in a randomized controlled trial comparing 16 weeks of aerobic physical activity plus sleep hygiene to non-physical activity plus sleep hygiene. Eligibility included primary insomnia for at least 3 months, habitual sleep duration <6.5h and a Pittsburgh Sleep Quality Index (PSQI) score >5. Outcomes included sleep quality, mood and quality of life questionnaires (PSQI, Epworth Sleepiness Scale [ESS], Short-form 36 [SF-36], Center for Epidemiological Studies Depression Scale [CES-D]). The physical activity group improved in sleep quality on the global PSQI (p<.0001), sleep latency (p=.049), sleep duration (p=.04), daytime dysfunction (p=.027), and sleep efficiency (p=.036) PSQI sub-scores compared to the control group. The physical activity group also had reductions in depressive symptoms (p=.044), daytime sleepiness (p=.02) and improvements in vitality (p=.017) compared to baseline scores. Aerobic physical activity with sleep hygiene education is an effective treatment approach to improve sleep quality, mood and quality of life in older adults with chronic insomnia.

  18. 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…

  19. 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.

  20. Emotional Content of Dreams in Obstructive Sleep Apnea Hypopnea Syndrome Patients and Sleepy Snorers attending a Sleep-Disordered Breathing Clinic

    PubMed Central

    Fisher, Samantha; Lewis, Keir E.; Bartle, Iona; Ghosal, Robin; Davies, Lois; Blagrove, Mark

    2011-01-01

    Study Objectives: To assess prospectively the emotional content of dreams in individuals with the obstructive sleep apnea hypopnea syndrome (OSAHS) and sleepy snorers. Methods: Prospective observational study. Forty-seven patients with sleepiness and snoring attending a sleep-disordered breathing clinic, completed a morning diary concerning pleasantness/unpleasantness of their dreams for 10 days, and then had AHI assessed by a limited-channel home sleep study. Participants and groups: Sleepy snorers, AHI < 5: n = 12 (mean age = 51.00 years [SD 7.01], 7 males); AHI 5 −14.9, n = 14 (mean age = 49.71 y [9.73], 12 males); AHI ≥ 15, n = 21 (mean age = 56.33 [11.24], 16 males). Results: All groups reported similar numbers of dreams and nightmares during the diary period. The AHI ≥ 15 group were significantly higher on dream unpleasantness than were the sleepy snorers (p < 0.05); and when only males were analyzed, this difference was also significant (p = 0.01). As AHI increased across the 3 groups, there was a significant decrease in variability of dream emotions (Levene test for homogeneity of variance between the 3 groups, p = 0.018). Mean daytime anxiety and daytime depression were significantly correlated with mean dream unpleasantness and with mean number of nightmares over the diary period. Conclusions: Patients with AHI ≥ 15 had more emotionally negative dreams than patients with AHI < 5. The variation in mean dream emotion decreased with increasing AHI, possibly because sleep fragmentation with increasing AHI results in fewer and shorter dreams, in which emotions are rarer. Citation: Fisher S; Lewis KE; Bartle I; Ghosal R; Davies L Blagrove M. Emotional content of dreams in obstructive sleep apnea hypopnea syndrome patients and sleepy snorers attending a sleep-disordered breathing clinic. J Clin Sleep Med 2011;7(1):69-74. PMID:21344048

  1. Emotional content of dreams in obstructive sleep apnea hypopnea syndrome patients and sleepy snorers attending a sleep-disordered breathing clinic.

    PubMed

    Fisher, Samantha; Lewis, Keir E; Bartle, Iona; Ghosal, Robin; Davies, Lois; Blagrove, Mark

    2011-02-15

    To assess prospectively the emotional content of dreams in individuals with the obstructive sleep apnea hypopnea syndrome (OSAHS) and sleepy snorers. Prospective observational study. Forty-seven patients with sleepiness and snoring attending a sleep-disordered breathing clinic, completed a morning diary concerning pleasantness/unpleasantness of their dreams for 10 days, and then had AHI assessed by a limited-channel home sleep study. Participants and groups: Sleepy snorers, AHI < 5: n = 12 (mean age = 51.00 years [SD 7.01], 7 males); AHI 5 -14.9, n = 14 (mean age = 49.71 y [9.73], 12 males); AHI ≥ 15, n = 21 (mean age = 56.33 [11.24], 16 males). All groups reported similar numbers of dreams and nightmares during the diary period. The AHI ≥ 15 group were significantly higher on dream unpleasantness than were the sleepy snorers (p < 0.05); and when only males were analyzed, this difference was also significant (p = 0.01). As AHI increased across the 3 groups, there was a significant decrease in variability of dream emotions (Levene test for homogeneity of variance between the 3 groups, p = 0.018). Mean daytime anxiety and daytime depression were significantly correlated with mean dream unpleasantness and with mean number of nightmares over the diary period. Patients with AHI ≥ 15 had more emotionally negative dreams than patients with AHI < 5. The variation in mean dream emotion decreased with increasing AHI, possibly because sleep fragmentation with increasing AHI results in fewer and shorter dreams, in which emotions are rarer.

  2. Comparison of the effects of continuous positive airway pressure and mandibular advancement devices on sleepiness in patients with obstructive sleep apnoea: a network meta-analysis.

    PubMed

    Bratton, Daniel J; Gaisl, Thomas; Schlatzer, Christian; Kohler, Malcolm

    2015-11-01

    Excessive daytime sleepiness is the most important symptom of obstructive sleep apnoea and can affect work productivity, quality of life, and the risk of road traffic accidents. We aimed to quantify the effects of the two main treatments for obstructive sleep apnoea (continuous positive airway pressure and mandibular advancement devices) on daytime sleepiness and to establish predictors of response to continuous positive airway pressure. We searched MEDLINE and the Cochrane Library from inception to May 31, 2015, to identify randomised controlled trials comparing the effects of continuous positive airway pressure, mandibular advancement devices or an inactive control (eg, placebo or no treatment) on the Epworth Sleepiness Scale (ESS, range 0-24 points) in patients with obstructive sleep apnoea. We did a network meta-analysis using multivariate random-effects meta-regression to assess the effect of each treatment on ESS. We used meta-regression to assess the association of the reported effects of continuous positive airway pressure versus inactive controls with the characteristics of trials and their risk of bias. We included 67 studies comprising 6873 patients in the meta-analysis. Compared with an inactive control, continuous positive airway pressure was associated with a reduction in ESS score of 2·5 points (95% CI 2·0-2·9) and mandibular advancement devices of 1·7 points (1·1-2·3). We estimated that, on average, continuous positive airway pressure reduced the ESS score by a further 0·8 points compared with mandibular advancement devices (95% CI 0·1-1·4; p=0·015). However, there was a possibility of publication bias in favour of continuous positive airway pressure that might have resulted in this difference. We noted no evidence that studies reporting higher continuous positive airway pressure adherence also reported larger treatment effects (p=0·70). Continuous positive airway pressure and mandibular advancement devices are effective treatments for

  3. Cardiac autonomic modulation and sleepiness: physiological consequences of sleep deprivation due to 40 h of prolonged wakefulness.

    PubMed

    Glos, Martin; Fietze, Ingo; Blau, Alexander; Baumann, Gert; Penzel, Thomas

    2014-02-10

    The autonomic nervous system (ANS) is modulated by sleep and wakefulness. Noninvasive assessment of cardiac ANS with heart rate variability (HRV) analysis is a window for monitoring malfunctioning of cardiovascular autonomic modulation due to sleep deprivation. This study represents the first investigation of dynamic ANS effects and of electrophysiological and subjective sleepiness, in parallel, during 40 h of prolonged wakefulness under constant routine (CR) conditions. In eleven young male healthy subjects, ECG, EEG, EOG, and EMG chin recordings were performed during baseline sleep, during 40 h of sleep deprivation, and during recovery sleep. After sleep deprivation, slow-wave sleep and sleep efficiency increased, whereas HRV - global variability and HRV sympathovagal balance - was reduced (all p<0.05). Sleep-stage-dependent analysis revealed reductions in the sympathovagal balance only for NREM sleep stages (all p<0.05). Comparison of the daytime pattern of CR day one (CR baseline) with that of CR day two (CR sleep deprivation) disclosed an increase in subjective sleepiness, in the amount of unintended sleep, and in HRV sympathovagal balance, with accompaniment by increased EEG alpha attenuation (all p<0.05). Circadian rhythm analysis revealed the strongest influence on heart rate, with less influence on HRV sympathovagal balance. Hour-by-hour analysis disclosed the difference between CR sleep deprivation and CR baseline for subjective sleepiness at almost every single hour and for unintended sleep particularly in the morning and afternoon (both p<0.05). These findings indicate that 40 h of prolonged wakefulness lead in the following night to sleep-stage-dependent reduction in cardiac autonomic modulation. During daytime, an increased occurrence of behavioral and physiological signs of sleepiness was accompanied by diminished cardiac autonomic modulation. The observed changes are an indicator of autonomic stress due to sleep deprivation - which, if chronic

  4. Effects of 6/6 and 4/8 watch systems on sleepiness among bridge officers.

    PubMed

    Härmä, Mikko; Partinen, Markku; Repo, Risto; Sorsa, Matti; Siivonen, Pertti

    2008-04-01

    During the last ten years, severe sleepiness or falling asleep by watch keeping officers has been a direct or a contributing factor in a number of maritime accidents. This study examined the relationship between two watch systems and its impact on fatigue and sleepiness in bridge officers. A questionnaire and a sleep/work diary were sent to a representative sample of the Finnish Maritime Officer Association. In all, 185 bridge officers answered the questionnaire on sleep, work hours, and safety, including the Skogby Excessive Daytime Sleepiness index (SEDS); 42% of the bridge officers worked two 4 h watches (4/8) per day, while 26% worked two 6 h watches per day (6/6). Ninety-five of the participants completed a sleep diary for seven consecutive days while at sea. The timing of the watch duties and sleep was recorded, as was subjective sleepiness every 2 h using the Karolinska Sleepiness Scale (KSS). 17.6% of the participants had fallen asleep at least once while on duty during their career. Compared to the 4/8 watch system, the officers working the 6/6 watch system reported shorter sleep durations, more frequent nodding-off on duty (7.3% vs. 1.5%), and excessive sleepiness (32% vs. 16% with SEDS>14). Based on a logistic regression analysis, high SEDS was significantly related with probable obstructive sleep apnea (OR 5.7), the 6/6 watch system (OR 4.0), and morningness-eveningness while controlling simultaneously several individual and sleep-related factors. Subjective sleepiness (KSS) was highest at 04:00 and 06:00 h. In a multivariate analysis, the KSS was significantly related to time of day, time after awaking, sleep length, and interactions of the watch systems with age, morningness-eveningness, and Epworth sleepiness scale (ESS) score. Severe sleepiness at 04:00-06:00 h was especially problematic in the 6/6 watch system among evening types and among the bridge officers with high ESS. The results suggest the 6/6 watch system is related to a higher risk of

  5. Later school start time is associated with improved sleep and daytime functioning in adolescents.

    PubMed

    Boergers, Julie; Gable, Christopher J; Owens, Judith A

    2014-01-01

    Chronic insufficient sleep is a growing concern among adolescents and is associated with a host of adverse health consequences. Early school start times may be an environmental contributor to this problem. The purpose of this study was to examine the impact of a delay in school start time on sleep patterns, sleepiness, mood, and health-related outcomes. Boarding students (n = 197, mean age = 15.6 yr) attending an independent high school completed the School Sleep Habits Survey before and after the school start time was experimentally delayed from 8:00 a.m. to 8:25 a.m. The delay in school start time was associated with a significant (29 min) increase in sleep duration on school nights. The percentage of students receiving 8 or more hours of sleep on a school night increased to more than double, from 18% to 44%. Students in 9th and 10th grade and those with lower baseline sleep amounts were more likely to report improvements in sleep duration after the schedule change. Daytime sleepiness, depressed mood, and caffeine use were all significantly reduced after the delay in school start time. Sleep duration reverted to baseline levels when the original (earlier) school start time was reinstituted. A modest (25 min) delay in school start time was associated with significant improvements in sleep duration, daytime sleepiness, mood, and caffeine use. These findings have important implications for public policy and add to research suggesting the health benefits of modifying school schedules to more closely align with adolescents' circadian rhythms and sleep needs.

  6. Daytime somnolence in adult sleepwalkers.

    PubMed

    Desautels, Alex; Zadra, Antonio; Labelle, Marc-Antoine; Dauvilliers, Yves; Petit, Dominique; Montplaisir, Jacques

    2013-11-01

    Sleepwalkers often complain of excessive daytime somnolence (EDS). Our retrospective study aimed to document the presence of EDS in a substantial sample of sleepwalkers and to explore the contribution of other sleep disorders, nocturnal sleep disruption, and sleep depth to the alteration of their daytime vigilance. Seventy adult sleepwalkers and 70 control subjects completed the Epworth Sleepiness Scale (ESS). Sleepwalkers also were studied for one night in the sleep laboratory. We compared the sleep profiles of 32 somnolent vs 38 nonsomnolent sleepwalkers and investigated the relationship between ESS scores and sleep-related variables. No differences were found in polysomnographic (PSG) parameters. Slow-wave activity (SWA) also was similar in the two subgroups. Sleepwalkers' ESS scores were not correlated with their body mass index (BMI) or periodic limb movements during sleep (PLMS) index, but they tended to be negatively correlated with indices of respiratory events. The EDS reported by adult sleepwalkers does not appear to be explained by the presence of concomitant sleep disorders or PSG signs of nocturnal sleep disruption. These results raise the possibility that EDS is part of the sleepwalking phenotype and that it is linked to its underlying pathophysiology. Copyright © 2013 Elsevier B.V. All rights reserved.

  7. Acute Versus Chronic Partial Sleep Deprivation in Middle-Aged People: Differential Effect on Performance and Sleepiness

    PubMed Central

    Philip, Pierre; Sagaspe, Patricia; Prague, Mélanie; Tassi, Patricia; Capelli, Aurore; Bioulac, Bernard; Commenges, Daniel; Taillard, Jacques

    2012-01-01

    Study Objective: To evaluate the effects of acute sleep deprivation and chronic sleep restriction on vigilance, performance, and self-perception of sleepiness. Design: Habitual night followed by 1 night of total sleep loss (acute sleep deprivation) or 5 consecutive nights of 4 hr of sleep (chronic sleep restriction) and recovery night. Participants: Eighteen healthy middle-aged male participants (age [(± standard deviation] = 49.7 ± 2.6 yr, range 46-55 yr). Measurements: Multiple sleep latency test trials, Karolinska Sleepiness Scale scores, simple reaction time test (lapses and 10% fastest reaction times), and nocturnal polysomnography data were recorded. Results: Objective and subjective sleepiness increased immediately in response to sleep restriction. Sleep latencies after the second and third nights of sleep restriction reached levels equivalent to those observed after acute sleep deprivation, whereas Karolinska Sleepiness Scale scores did not reach these levels. Lapse occurrence increased after the second day of sleep restriction and reached levels equivalent to those observed after acute sleep deprivation. A statistical model revealed that sleepiness and lapses did not progressively worsen across days of sleep restriction. Ten percent fastest reaction times (i.e., optimal alertness) were not affected by acute or chronic sleep deprivation. Recovery to baseline levels of alertness and performance occurred after 8-hr recovery night. Conclusions: In middle-aged study participants, sleep restriction induced a high increase in sleep propensity but adaptation to chronic sleep restriction occurred beyond day 3 of restriction. This sleepiness attenuation was underestimated by the participants. One recovery night restores daytime sleepiness and cognitive performance deficits induced by acute or chronic sleep deprivation. Citation: Philip P; Sagaspe P; Prague M; Tassi P; Capelli A; Bioulac B; Commenges D; Taillard J. Acute versus chronic partial sleep deprivation in

  8. Acute versus chronic partial sleep deprivation in middle-aged people: differential effect on performance and sleepiness.

    PubMed

    Philip, Pierre; Sagaspe, Patricia; Prague, Mélanie; Tassi, Patricia; Capelli, Aurore; Bioulac, Bernard; Commenges, Daniel; Taillard, Jacques

    2012-07-01

    To evaluate the effects of acute sleep deprivation and chronic sleep restriction on vigilance, performance, and self-perception of sleepiness. Habitual night followed by 1 night of total sleep loss (acute sleep deprivation) or 5 consecutive nights of 4 hr of sleep (chronic sleep restriction) and recovery night. Eighteen healthy middle-aged male participants (age [(± standard deviation] = 49.7 ± 2.6 yr, range 46-55 yr). Multiple sleep latency test trials, Karolinska Sleepiness Scale scores, simple reaction time test (lapses and 10% fastest reaction times), and nocturnal polysomnography data were recorded. Objective and subjective sleepiness increased immediately in response to sleep restriction. Sleep latencies after the second and third nights of sleep restriction reached levels equivalent to those observed after acute sleep deprivation, whereas Karolinska Sleepiness Scale scores did not reach these levels. Lapse occurrence increased after the second day of sleep restriction and reached levels equivalent to those observed after acute sleep deprivation. A statistical model revealed that sleepiness and lapses did not progressively worsen across days of sleep restriction. Ten percent fastest reaction times (i.e., optimal alertness) were not affected by acute or chronic sleep deprivation. Recovery to baseline levels of alertness and performance occurred after 8-hr recovery night. In middle-aged study participants, sleep restriction induced a high increase in sleep propensity but adaptation to chronic sleep restriction occurred beyond day 3 of restriction. This sleepiness attenuation was underestimated by the participants. One recovery night restores daytime sleepiness and cognitive performance deficits induced by acute or chronic sleep deprivation. Philip P; Sagaspe P; Prague M; Tassi P; Capelli A; Bioulac B; Commenges D; Taillard J. Acute versus chronic partial sleep deprivation in middle-aged people: differential effect on performance and sleepiness. SLEEP 2012;35(7):997-1002.

  9. Sleep quality, sleepiness and the influence of workplace breaks: A cross-sectional survey of health-care workers in two US hospitals.

    PubMed

    Wilson, Marian; Riedy, Samantha M; Himmel, Maddy; English, Ashley; Burton, Joshua; Albritton, Sandra; Johnson, Kelsey; Morgan, Patricia; Van Dongen, Hans P A

    2018-05-08

    This study assessed sleep quality, sleepiness and use of workplace break opportunities in 1285 health-care workers via an online questionnaire. Two hospitals were surveyed - one with and one without a fatigue mitigation policy. Across all respondents, 68.9% reported generally taking breaks of at least 30 min and 21.7% had access to a quiet place to rest, with no significant differences between hospitals. The presence of a fatigue mitigation policy was not associated with reduced sleepiness. However, accounting for hospital and shift characteristics, employees with access to a quiet place to rest while on break had significantly lower self-reported sleepiness scores.

  10. Effect of Melatonin Administration on Sleep Quality in Sulfur Mustard Exposed Patients with Sleep Disorders

    PubMed Central

    Mousavi, Seyyedeh Soghra; Shohrati, Majid; Vahedi, Ensieh; Abdollahpour-Alitappeh, Meghdad; Panahi, Yunes

    2018-01-01

    Sulfur mustard (SM) is a toxic agent that targets several tissues. It is the leading cause of persistent lung disease, progressive deterioration in lung function, and mortality among injured patients. Disturbed sleep and poor quality of sleep are common in SM-exposed patients with chronic respiratory problems. Melatonin is an alternative medication that has been widely used to treat poor sleep quality caused by several specific conditions. This study aimed to evaluate the efficacy of melatonin administration in improvement of sleep quality in SM-injured patients. In this randomized, double-blind and placebo-controlled trial study a total of 30 SM-exposed male patients were recruited. Patients received 3 mg melatonin (N = 15) or placebo (N = 15), orally in a single dose, 1 h before bedtime for 56 consecutive days. Sleep quality was evaluated by Pittsburgh Sleep Quality Index (PSQI); daytime sleepiness was measured by Epworth Sleepiness Scale (ESS), and the risk of obstructive sleep apnea was determined by STOP-Bang questionnaire. Compared with placebo, melatonin administration significantly improved global PSQI score, particularly sleep latency (P = 0.03) and subjective sleep quality (P = 0.004). Mean of global PSQI score was declined significantly (P = 0.01) from 10.13 ± 3.44 to 6.66 ± 3.08 in melatonin group. No differences in ESS and STOP-Bang scores were observed between two groups. Melatonin was effective in improving global PSQI score and sleep latency, but not daytime sleepiness and obstructive sleep apnea in SM-exposed patients. Further long-term studies involving larger number of patients are needed before melatonin can be safely recommended for the management of sleep disturbances in these patients.

  11. Leisure-Time Physical Activity and Sedentary Behavior and Their Cross-Sectional Associations with Excessive Daytime Sleepiness in the French SU.VI.MAX-2 Study.

    PubMed

    Andrianasolo, Roland M; Menai, Mehdi; Galan, Pilar; Hercberg, Serge; Oppert, Jean-Michel; Kesse-Guyot, Emmanuelle; Andreeva, Valentina A

    2016-04-01

    The potential benefit of physical activity in terms of decreasing excessive daytime sleepiness (EDS) prevalence is unclear, especially in aging adults. We aimed to elucidate the associations among physical activity, sedentariness, and EDS in middle-aged and older adults. We conducted a cross-sectional analysis using data from a subsample of participants in the SU.VI.MAX-2 observational study (2007-2009; N = 4179; mean age = 61.9 years). EDS was defined as a score >10 on the Epworth Sleepiness Scale. Leisure-time physical activity and different types of sedentary behavior were assessed with the Modifiable Activity Questionnaire. The associations were examined with multivariable logistic regression models. In the adjusted multivariable model, total leisure-time physical activity (modeled in quartiles, Q) was significantly, inversely associated with EDS (odds ratios (OR)Q4 vs Q1 = 0.70, 95 % confidence interval (CI) = 0.54-0.89). The association persisted in analyses restricted to individuals not taking sleep medication (ORQ4 vs Q1 = 0.72, 95 % CI = 0.54-0.95). In turn, time spent watching television and time spent reading appeared protective against EDS (ORQ4 vs Q1 = 0.73, 95 % CI = 0.57-0.94; ORQ4 vs Q1 = 0.76, 95 % CI = 0.60-0.97, respectively), whereas time spent on a computer appeared to confer an increased risk for EDS (ORQ4 vs Q1 = 1.30, 95 % CI = 1.05-1.62). When physical activity and sedentariness were modeled jointly, using WHO recommendation-based cutoffs for high/low levels, no significant associations were observed in the fully adjusted models. The findings reinforce public health recommendations promoting behavior modification and specifically moderate-intensity exercise in middle-aged and older adults. The association of high physical activity/low sedentariness with EDS, which was not supported by the data, merits further investigation before firm conclusions could be drawn.

  12. Obstructive sleep apnea with excessive daytime sleepiness is associated with non-alcoholic fatty liver disease regardless of visceral fat

    PubMed Central

    Yu, Ji Hee; Ahn, Jae Hee; Yoo, Hye Jin; Seo, Ji A; Kim, Sin Gon; Choi, Kyung Mook; Baik, Sei Hyun; Choi, Dong Seop; Shin, Chol; Kim, Nan Hee

    2015-01-01

    Background/Aims: Obstructive sleep apnea (OSA) is associated with an increased risk of obesity and non-alcoholic fatty liver disease (NAFLD), but it remains unclear whether the risk of NAFLD is independently related to OSA regardless of visceral obesity. Thus, the aim of the present study was to examine whether OSA alone or in combination with excessive daytime sleepiness (EDS) or short sleep duration was associated with NAFLD independent of visceral fat in Korean adults. Methods: A total of 621 participants were selected from the Korean Genome and Epidemiology Study (KoGES). The abdominal visceral fat area (VFA) and hepatic fat components of the participants were assessed using computed tomography scans and they were then categorized into four groups depending on the presence of OSA and EDS. Results: The proportions of NAFLD were 21.1%, 18.5%, 32.4%, and 46.7% in participants without OSA/EDS, with only EDS, with only OSA, and with both OSA and EDS, respectively. A combination of OSA and EDS increased the odds ratio (OR) for developing NAFLD (OR, 2.75; 95% confidence interval [CI], 1.21 to 6.28) compared to those without OSA/EDS, and this association remained significant (OR, 2.38; 95% CI, 1.01 to 5.59) even after adjusting for VFA. In short sleepers (< 5 hours) with OSA, the adjusted OR for NAFLD was 2.50 (95% CI, 1.08 to 5.75) compared to those sleeping longer than 5 hours without OSA. Conclusions: In the present study, OSA was closely associated with NAFLD in Korean adults. This association was particularly strong in those with EDS or short sleep duration regardless of VFA. PMID:26552460

  13. Obstructive sleep apnea with excessive daytime sleepiness is associated with non-alcoholic fatty liver disease regardless of visceral fat.

    PubMed

    Yu, Ji Hee; Ahn, Jae Hee; Yoo, Hye Jin; Seo, Ji A; Kim, Sin Gon; Choi, Kyung Mook; Baik, Sei Hyun; Choi, Dong Seop; Shin, Chol; Kim, Nan Hee

    2015-11-01

    Obstructive sleep apnea (OSA) is associated with an increased risk of obesity and non-alcoholic fatty liver disease (NAFLD), but it remains unclear whether the risk of NAFLD is independently related to OSA regardless of visceral obesity. Thus, the aim of the present study was to examine whether OSA alone or in combination with excessive daytime sleepiness (EDS) or short sleep duration was associated with NAFLD independent of visceral fat in Korean adults. A total of 621 participants were selected from the Korean Genome and Epidemiology Study (KoGES). The abdominal visceral fat area (VFA) and hepatic fat components of the participants were assessed using computed tomography scans and they were then categorized into four groups depending on the presence of OSA and EDS. The proportions of NAFLD were 21.1%, 18.5%, 32.4%, and 46.7% in participants without OSA/EDS, with only EDS, with only OSA, and with both OSA and EDS, respectively. A combination of OSA and EDS increased the odds ratio (OR) for developing NAFLD (OR, 2.75; 95% confidence interval [CI], 1.21 to 6.28) compared to those without OSA/EDS, and this association remained significant (OR, 2.38; 95% CI, 1.01 to 5.59) even after adjusting for VFA. In short sleepers (< 5 hours) with OSA, the adjusted OR for NAFLD was 2.50 (95% CI, 1.08 to 5.75) compared to those sleeping longer than 5 hours without OSA. In the present study, OSA was closely associated with NAFLD in Korean adults. This association was particularly strong in those with EDS or short sleep duration regardless of VFA.

  14. Armodafinil in the treatment of excessive sleepiness.

    PubMed

    Bogan, Richard K

    2010-04-01

    Excessive sleepiness causes impaired quality of life and increases the risk of poor health and accidents. Armodafinil is a wake-promoting agent approved in 2007 by the US Food and Drug Administration for the treatment of excessive sleepiness arising from narcolepsy, obstructive sleep apnea (OSA; even after optimal treatment for the underlying obstruction) and shift-work disorder (SWD). It is the R-enantiomer of modafinil, which is a racemic mixture of R- and S-enantiomers. This review summarizes the recent primary data on the pharmacokinetics, clinical efficacy and safety of armodafinil using literature published since 2005 that was identified from PubMed. The review describes recent advances in the understanding of the pharmacokinetic profile of the drug and why this may improve wakefulness later after dosing compared with modafinil. It also describes the recent efficacy and safety data supporting the use of armodafinil to treat excessive sleepiness in indicated patients. Armodafinil is a useful therapy for the treatment of excessive sleepiness arising from a number of clinical conditions. It is generally well tolerated and has a low potential for abuse or tolerance.

  15. Associations between sleep duration, sleep quality and diabetic retinopathy.

    PubMed

    Tan, Nicholas Y Q; Chew, Merwyn; Tham, Yih-Chung; Nguyen, Quang Duc; Yasuda, Masayuki; Cheng, Ching-Yu; Wong, Tien Yin; Sabanayagam, Charumathi

    2018-01-01

    Abnormal durations of sleep have been associated with risk of diabetes. However, it is not clear if sleep duration is associated with diabetic retinopathy (DR). In a cross-sectional study, we included 1,231 (Malay, n = 395; Indian, n = 836) adults (mean age 64.4 ± 9.0 years, 50.4% female) with diabetes from the second visit of two independent population-based cohort studies (2011-15) in Singapore. Self-reported habitual sleep duration was categorized as short (<6 h), normal (6≤ h <8), and long (≥8 h). Questionnaires were administered to detect risk of obstructive sleep apnea (OSA), excessive daytime sleepiness, and insomnia, all of which may indicate poor quality of sleep. The associations between sleep-related characteristics with moderate DR and vision-threatening DR (VTDR) were analysed using logistic regression models adjusted for potential confounders. Prevalence of moderate DR and VTDR in the study population were 10.5% and 6.3% respectively. The mean duration of sleep was 6.4 ± 1.5 h. Compared to normal sleep duration, both short and long sleep durations were associated with moderate DR with multivariable odds ratio (95% confidence interval) of 1.73 (1.03-2.89) and 2.17 (1.28-3.66) respectively. Long sleep duration (2.37 [1.16-4.89]), high risk of OSA (2.24 [1.09-4.75]), and excessive daytime sleepiness (3.27 [1.02-10.30]) were separately associated with VTDR. Sleep duration had a U-shaped association with moderate DR; long sleep duration, excessive daytime sleepiness and high risk of OSA were positively associated with VTDR.

  16. The Sleep-Time Cost of Parenting: Sleep Duration and Sleepiness Among Employed Parents in the Wisconsin Sleep Cohort Study

    PubMed Central

    Hagen, Erika W.; Mirer, Anna G.; Palta, Mari; Peppard, Paul E.

    2013-01-01

    Insufficient sleep is associated with poor health and increased mortality. Studies on whether parenthood (including consideration of number and ages of children) is associated with sleep duration or sleep problems are scant and inconclusive. Using data collected in the Wisconsin Sleep Cohort Study (n = 4,809) between 1989 and 2008, we examined cross-sectional associations of number and ages of children with self-reported parental sleep duration, daytime sleepiness, and dozing among employed adults. Longitudinal change in sleep duration over 19 years was examined to evaluate changes in parental sleep associated with children transitioning into adulthood (n = 833). Each child under age 2 years was associated with 13 fewer minutes of parental sleep per day (95% confidence interval (CI): 5, 21); each child aged 2–5 years was associated with 9 fewer minutes of sleep (95% CI: 5, 13); and each child aged 6–18 years was associated with 4 fewer minutes (95% CI: 2, 6). Adult children were not associated with shorter parental sleep duration. Parents of children over age 2 years were significantly more likely to experience daytime sleepiness and dozing during daytime activities. Parents of minor children at baseline had significantly greater increases in sleep duration over 19 years of follow-up. Parenting minor children is associated with shorter sleep duration. As children age into adulthood, the sleep duration of parents with more children approaches that of parents with fewer children. PMID:23378502

  17. The interplay between sleep and mood in predicting academic functioning, physical health and psychological health: a longitudinal study.

    PubMed

    Wong, Mark Lawrence; Lau, Esther Yuet Ying; Wan, Jacky Ho Yin; Cheung, Shu Fai; Hui, C Harry; Mok, Doris Shui Ying

    2013-04-01

    Existing studies on sleep and behavioral outcomes are mostly correlational. Longitudinal data is limited. The current longitudinal study assessed how sleep duration and sleep quality may be causally linked to daytime functions, including physical health (physical well-being and daytime sleepiness), psychological health (mood and self-esteem) and academic functioning (school grades and study effort). The mediation role of mood in the relationship between sleep quality, sleep duration and these daytime functions is also assessed. A sample of 930 Chinese students (aged 18-25) from Hong Kong/Macau completed self-reported questionnaires online across three academic semesters. Sleep behaviors are assessed by the sleep timing questionnaire (for sleep duration and weekday/weekend sleep discrepancy) and the Pittsburgh sleep quality index (sleep quality); physical health by the World Health Organization quality of life scale-brief version (physical well-being) and Epworth Sleepiness Scale (daytime sleepiness); psychological health by the depression anxiety stress scale (mood) and Rosenberg Self-esteem Scale (self-esteem) and academic functioning by grade-point-average and the college student expectation questionnaire (study effort). Structural equation modeling with a bootstrap resample of 5000 showed that after controlling for demographics and participants' daytime functions at baseline, academic functions, physical and psychological health were predicted by the duration and quality of sleep. While some sleep behaviors directly predicted daytime functions, others had an indirect effect on daytime functions through negative mood, such as anxiety. Sleep duration and quality have direct and indirect (via mood) effects on college students' academic function, physical and psychological health. Our findings underscore the importance of healthy sleep patterns for better adjustment in college years. Copyright © 2012 Elsevier Inc. All rights reserved.

  18. Attention following traumatic brain injury: Neuropsychological and driving simulator data, and association with sleep, sleepiness, and fatigue.

    PubMed

    Beaulieu-Bonneau, Simon; Fortier-Brochu, Émilie; Ivers, Hans; Morin, Charles M

    2017-03-01

    The objectives of this study were to compare individuals with traumatic brain injury (TBI) and healthy controls on neuropsychological tests of attention and driving simulation performance, and explore their relationships with participants' characteristics, sleep, sleepiness, and fatigue. Participants were 22 adults with moderate or severe TBI (time since injury ≥ one year) and 22 matched controls. They completed three neuropsychological tests of attention, a driving simulator task, night-time polysomnographic recordings, and subjective ratings of sleepiness and fatigue. Results showed that participants with TBI exhibited poorer performance compared to controls on measures tapping speed of information processing and sustained attention, but not on selective attention measures. On the driving simulator task, a greater variability of the vehicle lateral position was observed in the TBI group. Poorer performance on specific subsets of neuropsychological variables was associated with poorer sleep continuity in the TBI group, and with a greater increase in subjective sleepiness in both groups. No significant relationship was found between cognitive performance and fatigue. These findings add to the existing evidence that speed of information processing is still impaired several years after moderate to severe TBI. Sustained attention could also be compromised. Attention seems to be associated with sleep continuity and daytime sleepiness; this interaction needs to be explored further.

  19. Longitudinal Evaluation of Sleep-Disordered Breathing and Sleep Symptoms with Change in Quality of Life: The Sleep Heart Health Study (SHHS)

    PubMed Central

    Silva, Graciela E.; An, Ming-Wen; Goodwin, James L.; Shahar, Eyal; Redline, Susan; Resnick, Helaine; Baldwin, Carol M.; Quan, Stuart F.

    2009-01-01

    Study Objectives: Findings from population studies evaluating the progression and incidence of sleep disordered breathing have shown evidence of a longitudinal increase in the severity of sleep disordered breathing. The present study evaluates the association among changes in sleep disordered breathing, sleep symptoms, and quality of life over time. Design: Prospective cohort study. Data were from the Sleep Heart Health Study. Setting: Multicenter study. Participants: Three thousand seventy-eight subjects aged 40 years and older from the baseline and follow-up examination cycles were included. Measurements: The primary outcomes were changes in the Physical Component Summary and Mental Component Summary scales obtained from the Medical Outcomes Study Short-Form Health Survey. The primary exposure was change in the respiratory disturbance index obtained from unattended overnight polysomnograms performed approximately 5 years apart. Other covariates included measures of excessive daytime sleepiness and difficulty initiating and maintaining sleep. Results: Mean respiratory disturbance index increased from 8.1 ± 11 SD at baseline to 10.9 ± 14 (P < 0.0001) at follow-up. The mean Physical Component Summary and Mental Component Summary scores were 48.5 and 54.1 at baseline and 46.3 and 54.8 at follow-up. No associations between change in respiratory disturbance index and changes in Physical Component Summary or Mental Component Summary scores were seen. However, worsening of difficulty initiating and maintaining sleep and excessive daytime sleepiness were significantly associated with lower quality of life. Conclusions: A slight increase in severity of sleep disordered breathing was seen over 5 years; this was not associated with worsening of quality of life. However, subjective symptoms of quality of sleep and daytime sleepiness were associated with declining quality of life. Citation: Silva GE; An MW; Goodwin JL; Shahar E; Redline S; Resnick H; Baldwin CM; Quan SF

  20. Workplace lighting for improving alertness and mood in daytime workers.

    PubMed

    Pachito, Daniela V; Eckeli, Alan L; Desouky, Ahmed S; Corbett, Mark A; Partonen, Timo; Rajaratnam, Shantha Mw; Riera, Rachel

    2018-03-02

    Exposure to light plays a crucial role in biological processes, influencing mood and alertness. Daytime workers may be exposed to insufficient or inappropriate light during daytime, leading to mood disturbances and decreases in levels of alertness. To assess the effectiveness and safety of lighting interventions to improve alertness and mood in daytime workers. We searched the Cochrane Central Register of Controlled Trials (CENTRAL), MEDLINE, Embase, seven other databases; ClinicalTrials.gov and the World Health Organization trials portal up to January 2018. We included randomised controlled trials (RCTs), and non-randomised controlled before-after trials (CBAs) that employed a cross-over or parallel-group design, focusing on any type of lighting interventions applied for daytime workers. Two review authors independently screened references in two stages, extracted outcome data and assessed risk of bias. We used standardised mean differences (SMDs) and 95% confidence intervals (CI) to pool data from different questionnaires and scales assessing the same outcome across different studies. We combined clinically homogeneous studies in a meta-analysis. We used the GRADE system to rate quality of evidence. The search yielded 2844 references. After screening titles and abstracts, we considered 34 full text articles for inclusion. We scrutinised reports against the eligibility criteria, resulting in the inclusion of five studies (three RCTs and two CBAs) with 282 participants altogether. These studies evaluated four types of comparisons: cool-white light, technically known as high correlated colour temperature (CCT) light versus standard illumination; different proportions of indirect and direct light; individually applied blue-enriched light versus no treatment; and individually applied morning bright light versus afternoon bright light for subsyndromal seasonal affective disorder.We found no studies comparing one level of illuminance versus another.We found two CBA

  1. Person-directed, non-pharmacological interventions for sleepiness at work and sleep disturbances caused by shift work.

    PubMed

    Slanger, Tracy E; Gross, J Valérie; Pinger, Andreas; Morfeld, Peter; Bellinger, Miriam; Duhme, Anna-Lena; Reichardt Ortega, Rosalinde Amancay; Costa, Giovanni; Driscoll, Tim R; Foster, Russell G; Fritschi, Lin; Sallinen, Mikael; Liira, Juha; Erren, Thomas C

    2016-08-23

    Shift work is often associated with sleepiness and sleep disorders. Person-directed, non-pharmacological interventions may positively influence the impact of shift work on sleep, thereby improving workers' well-being, safety, and health. To assess the effects of person-directed, non-pharmacological interventions for reducing sleepiness at work and improving the length and quality of sleep between shifts for shift workers. We searched CENTRAL, MEDLINE Ovid, Embase, Web of Knowledge, ProQuest, PsycINFO, OpenGrey, and OSH-UPDATE from inception to August 2015. We also screened reference lists and conference proceedings and searched the World Health Organization (WHO) Trial register. We contacted experts to obtain unpublished data. Randomised controlled trials (RCTs) (including cross-over designs) that investigated the effect of any person-directed, non-pharmacological intervention on sleepiness on-shift or sleep length and sleep quality off-shift in shift workers who also work nights. At least two authors screened titles and abstracts for relevant studies, extracted data, and assessed risk of bias. We contacted authors to obtain missing information. We conducted meta-analyses when pooling of studies was possible. We included 17 relevant trials (with 556 review-relevant participants) which we categorised into three types of interventions: (1) various exposures to bright light (n = 10); (2) various opportunities for napping (n = 4); and (3) other interventions, such as physical exercise or sleep education (n = 3). In most instances, the studies were too heterogeneous to pool. Most of the comparisons yielded low to very low quality evidence. Only one comparison provided moderate quality evidence. Overall, the included studies' results were inconclusive. We present the results regarding sleepiness below. Bright light Combining two comparable studies (with 184 participants altogether) that investigated the effect of bright light during the night on sleepiness during a shift

  2. Assessment of sleepiness, fatigue, and depression among Gulf Cooperation Council commercial airline pilots.

    PubMed

    Aljurf, Tareq M; Olaish, Awad H; BaHammam, Ahmed S

    2018-05-01

    No studies have assessed the prevalence of fatigue, depression, sleepiness, and the risk of obstructive sleep apnea (OSA) among commercial airlines pilots in the Gulf Cooperation Council (GCC). This was a quantitative cross-sectional study conducted among pilots who were on active duty and had flown during the past 6 months for one of three commercial airline companies. We included participants with age between 20 and 65 years. Data were collected using a predesigned electronic questionnaire composed of questions related to demographic information in addition to the Fatigue Severity Scale (FSS), the Berlin Questionnaire, the Epworth Sleepiness Scale (ESS), and the Hospital Anxiety and Depression Scale (HADS). The study included 328 pilots with a mean age ± standard deviation of 41.4 ± 9.7 years. Overall, 224 (68.3%) pilots had an FSS score ≥ 36 indicating severe fatigue and 221 (67.4%) reported making mistakes in the cockpit because of fatigue. One hundred and twelve (34.1%) pilots had an ESS score ≥ 10 indicating excessive daytime sleepiness and 148 (45.1%) reported falling asleep at the controls at least once without previously agreeing with their colleagues. One hundred and thirteen (34.5%) pilots had an abnormal HADS depression score (≥ 8), and 96 (29.3%) pilots were at high risk for OSA requiring further assessment. Fatigue, sleepiness, risk of OSA, and depression are prevalent among GCC commercial airline pilots. Regular assessment by aviation authorities is needed to detect and treat these medical problems.

  3. Sleepiness and sleep-disordered breathing in truck drivers : risk analysis of road accidents.

    PubMed

    Catarino, Rosa; Spratley, Jorge; Catarino, Isabel; Lunet, Nuno; Pais-Clemente, Manuel

    2014-03-01

    Portugal has one of the highest road traffic fatality rates in Europe. A clear association between sleep-disordered breathing (SDB) and traffic accidents has been previously demonstrated. This study aimed to determine prevalence of excessive daytime sleepiness (EDS) and other sleep disorder symptoms among truck drivers and to identify which individual traits and work habits are associated to increased sleepiness and accident risk. We evaluated a sample of 714 truck drivers using a questionnaire (244 face-to-face interviews, 470 self-administered) that included sociodemographic data, personal habits, previous accidents, Epworth Sleepiness Scale (ESS), and the Berlin questionnaire (BQ). Twenty percent of drivers had EDS and 29 % were at high risk for having obstructive sleep apnea syndrome (OSAS). Two hundred sixty-one drivers (36.6 %) reported near-miss accidents (42.5 % sleep related) and 264 (37.0 %), a driving accident (16.3 % sleep related). ESS score ≥ 11 was a risk factor for both near-miss accidents (odds ratio (OR)=3.84, p<0.01) and accidents (OR=2.25, p<0.01). Antidepressant use was related to accidents (OR=3.30, p=0.03). We found an association between high Mallampati score (III-IV) and near misses (OR=1.89, p=0.04). In this sample of Portuguese truck drivers, we observed a high prevalence of EDS and other sleep disorder symptoms. Accident risk was related to sleepiness and antidepressant use. Identifying drivers at risk for OSAS should be a major priority of medical assessment centers, as a public safety policy.

  4. Use of CPAP to reduce arterial stiffness in moderate-to-severe obstructive sleep apnoea, without excessive daytime sleepiness (STIFFSLEEP): an observational cohort study protocol

    PubMed Central

    Mineiro, Maria Alexandra; Marques da Silva, Pedro; Alves, Marta; Virella, Daniel; Marques Gomes, Maria João; Cardoso, João

    2016-01-01

    Introduction Sleepiness is a cardinal symptom in obstructive sleep apnoea (OSA) but most patients have unspecific symptoms. Arterial stiffness, evaluated by pulse wave velocity (PWV), is related to atherosclerosis and cardiovascular (CV) risk. Arterial stiffness was reported to be higher in patients with OSA, improving after treatment with continuous positive airway pressure (CPAP). This study aims to assess whether the same effect occurs in patients with OSA and without sleepiness. Methods and analysis This observational study assesses the CV effect of CPAP therapy on a cohort of patients with moderate-to-severe OSA; the effect on the subcohorts of sleepy and non-sleepy patients will be compared. A systematic and consecutive sample of patients advised CPAP therapy will be recruited from a single outpatient sleep clinic (Centro Hospitalar de Lisboa Central—CHLC, Portugal). Eligible patients are male, younger than 65 years, with confirmed moderate-to-severe OSA and apnoea–hypopnea index (AHI) above 15/hour. Other sleep disorders, diabetes or any CV disease other than hypertension are exclusion criteria. Clinical evaluation at baseline includes Epworth Sleepiness Scale (ESS), and sleepiness is defined as ESS above 10. OSA will be confirmed by polygraphic study (cardiorespiratory, level 3). Participants are advised to undertake an assessment of carotid-femoral PWV (cf-PWV) and 24 hours evaluation of ambulatory blood pressure monitoring (ABPM), at baseline and after 4 months of CPAP therapy. Compliance and effectiveness of CPAP will be assessed. The main outcome is the variation of cf-PWV over time. Ethics and dissemination This protocol was approved by the Ethics Committees of CHLC (reference number 84/2012) and NOVA Medical School (number36/2014/CEFCM), Lisbon. Informed, written consent will be obtained. Its results will be presented at conferences and published in peer-reviewed journals. Trial registration number NCT02273089; Pre-results. PMID:27406645

  5. Effects of the road environment on the development of driver sleepiness in young male drivers.

    PubMed

    Ahlström, Christer; Anund, Anna; Fors, Carina; Åkerstedt, Torbjörn

    2018-03-01

    Latent driver sleepiness may in some cases be masked by for example social interaction, stress and physical activity. This short-term modulation of sleepiness may also result from environmental factors, such as when driving in stimulating environments. The aim of this study is to compare two road environments and investigate how they affect driver sleepiness. Thirty young male drivers participated in a driving simulator experiment where they drove two scenarios: a rural environment with winding roads and low traffic density, and a suburban road with higher traffic density and a more built-up roadside environment. The driving task was essentially the same in both scenarios, i.e. to stay on the road, without much interaction with other road users. A 2 × 2 design, with the conditions rural versus suburban, and daytime (full sleep) versus night-time (sleep deprived), was used. The results show that there were only minor effects of the road environment on subjective and physiological indicators of sleepiness. In contrast, there was an increase in subjective sleepiness, longer blink durations and increased EEG alpha content, both due to time on task and to night-time driving. The two road environments differed both in terms of the demand on driver action and of visual load, and the results indicate that action demand is the more important of the two factors. The notion that driver fatigue should be countered in a more stimulating visual environment such as in the city is thus more likely due to increased task demand rather than to a richer visual scenery. This should be investigated in further studies. Copyright © 2018 Elsevier Ltd. All rights reserved.

  6. [Sleepiness, safety on the road and management of risk].

    PubMed

    Garbarino, S; Traversa, F; Spigno, F

    2012-01-01

    Public health studies have shown that sleepiness at the wheel and other risks associated with sleep are responsible for 5% to 30% of road accidents, depending on the type of driver and/or road. In industrialized countries one-fifth of all traffic accidents can be ascribed to sleepiness behind the wheel. Sleep disorders and various common acute and chronic medical conditions together with lifestyles, extended work hours and prolonged wakefulness directly or indirectly affect the quality and quantity of one's sleep increasing the number of workers with sleep debt and staggered hours. These conditions may increase the risk of road accidents. Strategies to reduce this risk of both commercial and non-commercial drivers related to sleepiness include reliable diagnosis and treatment of sleep disorders, management of chronobiological conflicts, adequate catch-up sleep, and countermeasures against sleepiness at the wheel. Road transport safety requires the adoption of occupational health measures, including risk assessment, health education, technical-environmental prevention and health surveillance.

  7. 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.

  8. Global Assessment of the Impact of Type 2 Diabetes on Sleep through Specific Questionnaires. A Case-Control Study

    PubMed Central

    Lecube, Albert; Sánchez, Enric; Gómez-Peralta, Fernando; Abreu, Cristina; Valls, Joan; Mestre, Olga; Romero, Odile; Martínez, María Dolores; Sampol, Gabriel; Ciudin, Andreea; Hernández, Cristina; Simó, Rafael

    2016-01-01

    Abstract Type 2 diabetes (T2D) is an independent risk factor for sleep breathing disorders. However, it is unknown whether T2D affects daily somnolence and quality of sleep independently of the impairment of polysomnographic parameters. Material and Methods A case-control study including 413 patients with T2D and 413 non-diabetic subjects, matched by age, gender, BMI, and waist and neck circumferences. A polysomnography was performed and daytime sleepiness was evaluated using the Epworth Sleepiness Scale (ESS). In addition, 135 subjects with T2D and 45 controls matched by the same previous parameters were also evaluated through the Pittsburgh Sleep Quality Index (PSQI) to calculate sleep quality. Results Daytime sleepiness was higher in T2D than in control subjects (p = 0.003), with 23.9% of subjects presenting an excessive daytime sleepiness (ESS>10). Patients with fasting plasma glucose (FPG ≥13.1 mmol/l) were identified as the group with a higher risk associated with an ESS>10 (OR 3.9, 95% CI 1.8–7.9, p = 0.0003). A stepwise regression analyses showed that the presence of T2D, baseline glucose levels and gender but not polysomnographic parameters (i.e apnea-hyoapnea index or sleeping time spent with oxigen saturation lower than 90%) independently predicted the ESS score. In addition, subjects with T2D showed higher sleep disturbances [PSQI: 7.0 (1.0–18.0) vs. 4 (0.0–12.0), p<0.001]. Conclusion The presence of T2D and high levels of FPG are independent risk factors for daytime sleepiness and adversely affect sleep quality. Prospective studies addressed to demonstrate whether glycemia optimization could improve the sleep quality in T2D patients seem warranted. PMID:27315083

  9. Global Assessment of the Impact of Type 2 Diabetes on Sleep through Specific Questionnaires. A Case-Control Study.

    PubMed

    Lecube, Albert; Sánchez, Enric; Gómez-Peralta, Fernando; Abreu, Cristina; Valls, Joan; Mestre, Olga; Romero, Odile; Martínez, María Dolores; Sampol, Gabriel; Ciudin, Andreea; Hernández, Cristina; Simó, Rafael

    2016-01-01

    Type 2 diabetes (T2D) is an independent risk factor for sleep breathing disorders. However, it is unknown whether T2D affects daily somnolence and quality of sleep independently of the impairment of polysomnographic parameters. A case-control study including 413 patients with T2D and 413 non-diabetic subjects, matched by age, gender, BMI, and waist and neck circumferences. A polysomnography was performed and daytime sleepiness was evaluated using the Epworth Sleepiness Scale (ESS). In addition, 135 subjects with T2D and 45 controls matched by the same previous parameters were also evaluated through the Pittsburgh Sleep Quality Index (PSQI) to calculate sleep quality. Daytime sleepiness was higher in T2D than in control subjects (p = 0.003), with 23.9% of subjects presenting an excessive daytime sleepiness (ESS>10). Patients with fasting plasma glucose (FPG ≥13.1 mmol/l) were identified as the group with a higher risk associated with an ESS>10 (OR 3.9, 95% CI 1.8-7.9, p = 0.0003). A stepwise regression analyses showed that the presence of T2D, baseline glucose levels and gender but not polysomnographic parameters (i.e apnea-hyoapnea index or sleeping time spent with oxigen saturation lower than 90%) independently predicted the ESS score. In addition, subjects with T2D showed higher sleep disturbances [PSQI: 7.0 (1.0-18.0) vs. 4 (0.0-12.0), p<0.001]. The presence of T2D and high levels of FPG are independent risk factors for daytime sleepiness and adversely affect sleep quality. Prospective studies addressed to demonstrate whether glycemia optimization could improve the sleep quality in T2D patients seem warranted.

  10. Excessive daytime sleepiness at work and subjective work performance in the general population and among heavy snorers and patients with obstructive sleep apnea.

    PubMed

    Ulfberg, J; Carter, N; Talbäck, M; Edling, C

    1996-09-01

    To evaluate excessive daytime sleepiness (EDS) at work and effects on reported work performance among men in the general population and male patients suffering from snoring and obstructive sleep apnea syndrome (OSAS). A cross-sectional study of Swedish men between the ages of 30 and 64 years in the county of Kopparberg, in mid-Sweden. A random sample of the general population (n = 285) and consecutive patients referred to a sleep laboratory who fulfilled objective diagnostic criteria (snorers = 289, OSAS = 62) responded to a questionnaire. Responders from the general population were divided into 2 groups, nonsnorers (n = 223) and snorers (n = 62). To validate a question on snoring in the questionnaire, 50 men, randomly selected from the sample of the general population, underwent sleep apnea screening in their homes. The specificity of the questions about snoring was 83% and the sensitivity was 42%. The risk ratios for reporting EDS at work were 4-fold for snorers in the general population, 20-fold for snoring patients, and 40-fold for patients with OSAS as compared with nonsnoring men in the general population. Patients with OSAS and snoring patients both showed increased ratios on measures of difficulties with concentration, learning new tasks, and performing monotonous tasks when compared with nonsnorers. Snoring and sleep apnea were highly associated with excessive EDS at work and subjective work performance problems. The results provide additional evidence that snoring is not merely a nuisance.

  11. Adherence to positive airway pressure therapy in U.S. military personnel with sleep apnea improves sleepiness, sleep quality, and depressive symptoms.

    PubMed

    Mysliwiec, Vincent; Capaldi, Vincent F; Gill, Jessica; Baxter, Tristin; O'Reilly, Brian M; Matsangas, Panagiotis; Roth, Bernard J

    2015-04-01

    Obstructive sleep apnea (OSA) is frequently diagnosed in U.S. military personnel. OSA is associated with sleepiness, poor sleep quality, and service-related illnesses of insomnia, depression, post-traumatic stress disorder, and traumatic brain injury. Observational study of active duty military personnel with OSA and adherence to positive airway pressure (PAP) assessed with smart chip technology. 58 men with mean age 36.2 ± 7.7 years, mean body mass index 31.4 ± 3.7 with mean apnea-hypopnea index (AHI) 19.1 ± 19.0 are reported. 23 (39.7%) participants were adherent to PAP, and 35 (60.3%) were nonadherent. No significant differences in baseline demographics, apnea-hypopnea index, service-related illnesses, or clinical instrument scores. Military personnel adherent to PAP had significantly improved sleepiness (p = 0.007), sleep quality (p = 0.013), depressive symptoms (p = 0.01), energy/fatigue (p = 0.027), and emotional well-being (p = 0.024). Participants with moderate-severe OSA were more likely to be in the adherent group when compared with participants diagnosed with mild OSA. Military personnel with OSA have low adherence to PAP. Adherence is associated with improved depressive symptoms, sleepiness, sleep quality, energy/fatigue, emotional well-being, and social functioning. Future research should focus on interventions to improve the management of OSA in military personnel. Reprint & Copyright © 2015 Association of Military Surgeons of the U.S.

  12. Continuous Positive Airway Pressure Improves Quality of Life in Women with Obstructive Sleep Apnea. A Randomized Controlled Trial.

    PubMed

    Campos-Rodriguez, Francisco; Queipo-Corona, Carlos; Carmona-Bernal, Carmen; Jurado-Gamez, Bernabe; Cordero-Guevara, Jose; Reyes-Nuñez, Nuria; Troncoso-Acevedo, Fernanda; Abad-Fernandez, Araceli; Teran-Santos, Joaquin; Caballero-Rodriguez, Julian; Martin-Romero, Mercedes; Encabo-Motiño, Ana; Sacristan-Bou, Lirios; Navarro-Esteva, Javier; Somoza-Gonzalez, Maria; Masa, Juan F; Sanchez-Quiroga, Maria A; Jara-Chinarro, Beatriz; Orosa-Bertol, Belen; Martinez-Garcia, Miguel A

    2016-11-15

    Continuous positive airway pressure (CPAP) is the treatment of choice in patients with symptomatic obstructive sleep apnea (OSA). CPAP treatment improves quality of life (QoL) in men with OSA, but its role in women has not yet been assessed. To investigate the effect of CPAP on QoL in women with moderate to severe OSA. We conducted a multicenter, open-label randomized controlled trial in 307 consecutive women diagnosed with moderate to severe OSA (apnea-hypopnea index, ≥15) in 19 Spanish sleep units. Women were randomized to receive effective CPAP therapy (n = 151) or conservative treatment (n = 156) for 3 months. The primary endpoint was the change in QoL based on the Quebec Sleep Questionnaire. Secondary endpoints included changes in daytime sleepiness, mood state, anxiety, and depression. Data were analyzed on an intention-to-treat basis with adjustment for baseline values and other relevant clinical variables. The women in the study had a mean (SD) age of 57.1 (10.1) years and a mean (SD) Epworth Sleepiness Scale score of 9.8 (4.4), and 77.5% were postmenopausal. Compared with the control group, the CPAP group achieved a significantly greater improvement in all QoL domains of the Quebec Sleep Questionnaire (adjusted treatment effect between 0.53 and 1.33; P < 0.001 for all domains), daytime sleepiness (-2.92; P < 0.001), mood state (-4.24; P  = 0.012), anxiety (-0.89; P = 0.014), depression (-0.85; P = 0.016), and the physical component summary of the 12-item Short Form Health Survey (2.78; P = 0.003). In women with moderate or severe OSA, 3 months of CPAP therapy improved QoL, mood state, anxiety and depressive symptoms, and daytime sleepiness compared with conservative treatment. Clinical trial registered with www.clinicaltrials.gov (NCT02047071).

  13. Effect of hydrotherapy on quality of life, functional capacity and sleep quality in patients with fibromyalgia.

    PubMed

    Silva, Kyara Morgana Oliveira Moura; Tucano, Silvia Jurema Pereira; Kümpel, Claudia; Castro, Antonio Adolfo Mattos de; Porto, Elias Ferreira

    2012-12-01

    Fibromyalgia affects 8% of the population over the age of 40 years, and 75% of the patients with fibromyalgia have poor sleep quality. To assess the effects of hydrotherapy on the physical function and sleep quality of patients with fibromyalgia. Patients were under clinical care at the UNASP Outpatient Clinic. This study assessed 60 female patients with fibromyalgia aged between 30 and 65 years. Out of the 60 patients assessed, 20 were excluded and 10 left the study because they could not comply with the time schedule. All patients completed the following questionnaires: Fibromyalgia Impact Questionnaire (FIQ); Pittsburgh Sleep Quality Index, and Epworth Sleepiness Scale. Training sessions were performed twice a week for two months, each session lasting 60 minutes. Patients' mean age was 45 years, 66% were active workers, and 34% had quit work. Right after the hydrotherapy program, the patients improved the following aspects assessed by use of the FIQ: physical function, work absenteeism, ability to do job, pain intensity, fatigue, morning tiredness, stiffness (P < 0.0001), anxiety (P = 0,0013), and depression (P < 0.0001). Sleep quality (P < 0.0001) and daytime sleepiness (P = 0.0003) also improved. Hydrotherapy improves sleep quality, physical function, professional status, psychological disorders and physical symptoms in patients with fibromyalgia.

  14. Use of CPAP to reduce arterial stiffness in moderate-to-severe obstructive sleep apnoea, without excessive daytime sleepiness (STIFFSLEEP): an observational cohort study protocol.

    PubMed

    Mineiro, Maria Alexandra; Marques da Silva, Pedro; Alves, Marta; Virella, Daniel; Marques Gomes, Maria João; Cardoso, João

    2016-07-12

    Sleepiness is a cardinal symptom in obstructive sleep apnoea (OSA) but most patients have unspecific symptoms. Arterial stiffness, evaluated by pulse wave velocity (PWV), is related to atherosclerosis and cardiovascular (CV) risk. Arterial stiffness was reported to be higher in patients with OSA, improving after treatment with continuous positive airway pressure (CPAP). This study aims to assess whether the same effect occurs in patients with OSA and without sleepiness. This observational study assesses the CV effect of CPAP therapy on a cohort of patients with moderate-to-severe OSA; the effect on the subcohorts of sleepy and non-sleepy patients will be compared. A systematic and consecutive sample of patients advised CPAP therapy will be recruited from a single outpatient sleep clinic (Centro Hospitalar de Lisboa Central-CHLC, Portugal). Eligible patients are male, younger than 65 years, with confirmed moderate-to-severe OSA and apnoea-hypopnea index (AHI) above 15/hour. Other sleep disorders, diabetes or any CV disease other than hypertension are exclusion criteria. Clinical evaluation at baseline includes Epworth Sleepiness Scale (ESS), and sleepiness is defined as ESS above 10. OSA will be confirmed by polygraphic study (cardiorespiratory, level 3). Participants are advised to undertake an assessment of carotid-femoral PWV (cf-PWV) and 24 hours evaluation of ambulatory blood pressure monitoring (ABPM), at baseline and after 4 months of CPAP therapy. Compliance and effectiveness of CPAP will be assessed. The main outcome is the variation of cf-PWV over time. This protocol was approved by the Ethics Committees of CHLC (reference number 84/2012) and NOVA Medical School (number36/2014/CEFCM), Lisbon. Informed, written consent will be obtained. Its results will be presented at conferences and published in peer-reviewed journals. NCT02273089; Pre-results. Published by the BMJ Publishing Group Limited. For permission to use (where not already granted under a

  15. Caregiving-Related Sleep Problems and Their Relationship to Mental Health and Daytime Function in Female Veterans.

    PubMed

    Song, Yeonsu; Washington, Donna L; Yano, Elizabeth M; McCurry, Susan M; Fung, Constance H; Dzierzewski, Joseph M; Rodriguez, Juan Carlos; Jouldjian, Stella; Mitchell, Michael N; Alessi, Cathy A; Martin, Jennifer L

    2018-01-01

    To identify caregiving-related sleep problems and their relationship to mental health and daytime function in female Veterans. Female Veterans (N = 1,477) from cross-sectional, nationwide, postal survey data. The survey respondent characteristics included demographics, comorbidity, physical activity, health, use of sleep medications, and history of sleep apnea. They self-identified caregiving- related sleep problems (i.e., those who had trouble sleeping because of caring for a sick adult, an infant/child, or other respondents). Patient Health Questionnaire (PHQ-4) was used to assess mental health, and daytime function was measured using 11 items of International Classification of Sleep Disorders-2 (ICSD-2). Female Veterans with self-identified sleep problems due to caring for a sick adult (n = 59) experienced significantly more symptoms of depression and anxiety (p < 0.001) and impairment in daytime function (e.g., fatigue, daytime sleepiness, loss of concentration, p < 0.001) than those with self-identified sleep problems due to caring for an infant or child (n = 95) or all other respondents (n = 1,323) after controlling for the respondent characteristics. Healthcare providers should pay attention to assessing sleep characteristics of female Veterans with caregiving responsibilities, particularly those caregiving for a sick adult.

  16. Sleep quality and motor vehicle crashes in adolescents.

    PubMed

    Pizza, Fabio; Contardi, Sara; Antognini, Alessandro Baldi; Zagoraiou, Maroussa; Borrotti, Matteo; Mostacci, Barbara; Mondini, Susanna; Cirignotta, Fabio

    2010-02-15

    Sleep-related complaints are common in adolescents, but their impact on the rate of motor vehicle crashes accidents is poorly known. We studied subjective sleep quality, driving habits, and self-reported car crashes in high-school adolescents. Self-administered questionnaires (with items exploring driving habits) were distributed to 339 students who had a driver's license and attended 1 of 7 high schools in Bologna, Italy. Statistical analysis were performed to describe lifestyle habits, sleep quality, sleepiness, and their relationship with the binary dependent variable (presence or absence of car crashes) to identify the factors significantly affecting the probability of car crashes in a multivariate binary logistic regression model. Nineteen percent of the sample reported bad sleep, 64% complained of daytime sleepiness, and 40% reported sleepiness while driving. Eighty students (24%), 76% of which were males, reported that they had already crashed at least once, and 15% considered sleepiness to have been the main cause of their crash. As compared with adolescents who had not had a crash, those who had at least 1 previous crash reported that they more frequently used to drive (79% vs 62%), drove at night (25% vs 9%), drove while sleepy (56% vs 35%), had bad sleep (29% vs 16%), and used stimulants such as caffeinated soft drinks (32% vs 19%), tobacco (54% vs 27%), and drugs (21% vs 7%). The logistic procedure established a significant predictive role of male sex (p < 0.0001; odds ratio = 3.3), tobacco use (p < 0.0001; odds ratio = 3.2), sleepiness while driving (p = 0.010; odds ratio = 2.1), and bad sleep (p = 0.047; odds ratio = 1.9) for the crash risk. Our results confirm the high prevalence of sleep-related complaints among adolescents and highlight their independent role on self-reported crash risk.

  17. Sleep Disturbances in Individuals With Phelan-McDermid Syndrome: Correlation With Caregivers' Sleep Quality and Daytime Functioning.

    PubMed

    Bro, Della; O'Hara, Ruth; Primeau, Michelle; Hanson-Kahn, Andrea; Hallmayer, Joachim; Bernstein, Jonathan A

    2017-02-01

    The aims of this study were to document sleep disturbances in individuals with Phelan-McDermid syndrome (PMS), to assess whether these individuals had been evaluated for sleep disorders, and to examine relationships between the sleep behavior of these individuals and the sleep behavior and daytime functioning of their caregivers. Participants were 193 caregivers of individuals with PMS recruited by the Phelan-McDermid Syndrome Foundation. Data were collected through a survey comprising 2 questionnaires: the Children's Sleep Habits Questionnaire (CSHQ) and the Parents' Sleep Habits Questionnaire. Data were analyzed using multiple linear regression analyses, Pearson correlation analyses, and independent-samples t-tests. Ninety percent of individuals with PMS showed evidence of marked sleep disturbance based on caregiver responses to the CSHQ. However, only 22% of individuals had undergone a formal sleep assessment. Reported increased sleep disturbance in individuals with PMS was a statistically significant predictor of reported increased sleep disturbance and daytime sleepiness in their caregivers. Sleep disturbance may be present in a substantial proportion of individuals with PMS and is negatively associated with caregivers' well-being. However, most individuals with PMS have not been evaluated for sleep disorders. When properly diagnosed, many sleep disorders can be alleviated with intervention. Thus, routine screening for and evaluation of sleep disturbances in individuals with PMS may have long-term positive impacts on the well-being of these individuals and their caregivers. © 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.

  18. An interventional approach for patient and nurse safety: a fatigue countermeasures feasibility study.

    PubMed

    Scott, Linda D; Hofmeister, Nancee; Rogness, Neal; Rogers, Ann E

    2010-01-01

    Studies indicate that extended shifts worked by hospital staff nurses are associated with higher risk of errors. Long work hours coupled with insufficient sleep and fatigue are even riskier. Although other industries have developed programs to reduce fatigue-related errors and injury, fatigue countermeasures program for nurses (FCMPN) are lacking. The objective of this study was to evaluate the feasibility of an FCMPN for improving sleep duration and quality while reducing daytime sleepiness and patient care errors. Selected sleep variables, errors and drowsy driving, were evaluated among hospital staff nurses (n = 47) before and after FCMPN implementation. A one-group pretest-posttest repeated-measures approach was used. Participants provided data 2 weeks before the FCMPN, 4 weeks after receiving the intervention, and again at 3 months after intervention. Most of the nurses experienced poor sleep quality, severe daytime sleepiness, and decreased alertness at work and while operating a motor vehicle. After the FCMPN, significant improvements were noted in sleep duration, sleep quality, alertness, and error prevention. Although significant improvements were not found in daytime sleepiness scores, severity of daytime sleepiness appeared to decrease. Despite improvements in fatigue management, nurses reported feelings of guilt when engaging in FCMPN activities, especially strategic naps and relieved breaks. Initial findings support the feasibility of using an FCMPN for mitigating fatigue, improving sleep, and reducing errors among hospital staff nurses. In future investigations, the acceptability, efficacy, and effectiveness of FCMPNs can be examined.

  19. Sleep and Sleepiness among First-Time Postpartum Parents: A Field- and Laboratory-Based Multimethod Assessment

    PubMed Central

    Insana, Salvatore P.; Montgomery-Downs, Hawley E.

    2012-01-01

    The study aim was to compare sleep, sleepiness, fatigue, and neurobehavioral performance among first-time mothers and fathers during their early postpartum period. Participants were 21 first-time postpartum mother-father dyads (N=42) and seven childless control dyads (N=14). Within their natural environment, participants completed one week of wrist actigraphy monitoring, along with multi-day self-administered sleepiness, fatigue, and neurobehavioral performance measures. The assessment week was followed by an objective laboratory based test of sleepiness. Mothers obtained more sleep compared to fathers, but mothers’ sleep was more disturbed by awakenings. Fathers had greater objectively measured sleepiness than mothers. Mothers and fathers did not differ on subjectively measured sleep quality, sleepiness, or fatigue; however, mothers had worse neurobehavioral performance than fathers. Compared to control dyads, postpartum parents experienced greater sleep disturbance, sleepiness, and sleepiness associated impairments. Study results inform social policy, postpartum sleep interventions, and research on postpartum family systems and mechanisms that propagate sleepiness. PMID:22553114

  20. The on-road experiences and awareness of sleepiness in a sample of Australian highway drivers: A roadside driver sleepiness study.

    PubMed

    Watling, Christopher N; Armstrong, Kerry A; Smith, Simon S; Wilson, Adrian

    2016-01-01

    Driver sleepiness contributes substantially to road crash incidents. Simulator and on-road studies clearly reveal an impairing effect from sleepiness on driving ability. However, the degree to which drivers appreciate the dangerousness of driving while sleepy is somewhat unclear. This study sought to determine drivers' on-road experiences of sleepiness, their prior sleep habits, and personal awareness of the signs of sleepiness. Participants were a random selection of 92 drivers traveling on a major highway in the state of Queensland, Australia, who were stopped by police as part of routine drink driving operations. Participants completed a brief questionnaire that included demographic information, sleepy driving experiences (signs of sleepiness and on-road experiences of sleepiness), and prior sleep habits. A modified version of the Karolinska Sleepiness Scale (KSS) was used to assess subjective sleepiness in the 15 min prior to being stopped by police. Participants' ratings of subjective sleepiness were quite low, with 90% reporting being alert to extremely alert on the KSS. Participants were reasonably aware of the signs of sleepiness, with many signs of sleepiness associated with on-road experiences of sleepiness. Additionally, the number of hours spent driving was positively correlated with the drivers' level of sleep debt. The results suggest that participants had moderate experiences of driving while sleepy and many were aware of the signs of sleepiness. The relationship between driving long distances and increased sleep debt is a concern for road safety. Increased education regarding the dangers of sleepy driving seems warranted.

  1. Tailored educational supportive care programme on sleep quality and psychological distress in patients with heart failure: A randomised controlled trial.

    PubMed

    Chang, Yia-Ling; Chiou, Ai-Fu; Cheng, Shu-Meng; Lin, Kuan-Chia

    2016-09-01

    Up to 74% of patients with heart failure report poor sleep in Taiwan. Poor symptom management or sleep hygiene may affect patients' sleep quality. An effective educational programme was important to improve patients' sleep quality and psychological distress. However, research related to sleep disturbance in patients with heart failure is limited in Taiwan. To examine the effects of a tailored educational supportive care programme on sleep disturbance and psychological distress in patients with heart failure. randomised controlled trial. Eighty-four patients with heart failure were recruited from an outpatient department of a medical centre in Taipei, Taiwan. Patients were randomly assigned to the intervention group (n=43) or the control group (n=41). Patients in the intervention group received a 12-week tailored educational supportive care programme including individualised education on sleep hygiene, self-care, emotional support through a monthly nursing visit at home, and telephone follow-up counselling every 2 weeks. The control group received routine nursing care. Data were collected at baseline, the 4th, 8th, and 12th weeks after patients' enrollment. Outcome measures included sleep quality, daytime sleepiness, anxiety, and depression. The intervention group exhibited significant improvement in the level of sleep quality and daytime sleepiness after 12 weeks of the supportive nursing care programme, whereas the control group exhibited no significant differences. Anxiety and depression scores were increased significantly in the control group at the 12th week (p<.001). However, anxiety and depression scores in the intervention group remained unchanged after 12 weeks of the supportive nursing care programme (p>.05). Compared with the control group, the intervention group had significantly greater improvement in sleep quality (β=-2.22, p<.001), daytime sleepiness (β=-4.23, p<.001), anxiety (β=-1.94, p<.001), and depression (β=-3.05, p<.001) after 12 weeks of

  2. Sleep Quality and Motor Vehicle Crashes in Adolescents

    PubMed Central

    Pizza, Fabio; Contardi, Sara; Antognini, Alessandro Baldi; Zagoraiou, Maroussa; Borrotti, Matteo; Mostacci, Barbara; Mondini, Susanna; Cirignotta, Fabio

    2010-01-01

    Study Objectives: Sleep-related complaints are common in adolescents, but their impact on the rate of motor vehicle crashes accidents is poorly known. We studied subjective sleep quality, driving habits, and self-reported car crashes in high-school adolescents. Methods: Self-administered questionnaires (with items exploring driving habits) were distributed to 339 students who had a driver's license and attended 1 of 7 high schools in Bologna, Italy. Statistical analysis were performed to describe lifestyle habits, sleep quality, sleepiness, and their relationship with the binary dependent variable (presence or absence of car crashes) to identify the factors significantly affecting the probability of car crashes in a multivariate binary logistic regression model. Results: Nineteen percent of the sample reported bad sleep, 64% complained of daytime sleepiness, and 40% reported sleepiness while driving. Eighty students (24%), 76% of which were males, reported that they had already crashed at least once, and 15% considered sleepiness to have been the main cause of their crash. As compared with adolescents who had not had a crash, those who had at least 1 previous crash reported that they more frequently used to drive (79% vs 62%), drove at night (25% vs 9%), drove while sleepy (56% vs 35%), had bad sleep (29% vs 16%), and used stimulants such as caffeinated soft drinks (32% vs 19%), tobacco (54% vs 27%), and drugs (21% vs 7%). The logistic procedure established a significant predictive role of male sex (p < 0.0001; odds ratio = 3.3), tobacco use (p < 0.0001; odds ratio = 3.2), sleepiness while driving (p = 0.010; odds ratio = 2.1), and bad sleep (p = 0.047; odds ratio = 1.9) for the crash risk. Conclusions: Our results confirm the high prevalence of sleep-related complaints among adolescents and highlight their independent role on self-reported crash risk. Citation: Pizza F; Contardi S; Baldi Antognini A; Zagoraiou M; Borrotti M; Mostacci B; Mondini S; Cirignotta F

  3. To Assess Sleep Quality among Pakistani Junior Physicians (House Officers): A Cross-sectional Study.

    PubMed

    Surani, A A; Surani, A; Zahid, S; Ali, S; Farhan, R; Surani, S

    2015-01-01

    Sleep deprivation among junior physicians (house officers) is of growing concern. In developed countries, duty hours are now mandated, but in developing countries, junior physicians are highly susceptible to develop sleep impairment due to long working hours, on-call duties and shift work schedule. We undertook the study to assess sleep quality among Pakistani junior physicians. A cross-sectional study was conducted at private and public hospitals in Karachi, Pakistan, from June 2012 to January 2013. The study population comprised of junior doctors (house physicians and house surgeons). A consecutive sample of 350 physicians was drawn from the above-mentioned study setting. The subject underwent two validated self-administered questionnaires, that is, Pittsburgh Sleep Quality Index (PSQI) and Epworth Sleepiness Scale (ESS). A total of 334 physicians completely filled out the questionnaire with a response rate of 95.4% (334/350). Of 334 physicians, 36.8% (123/334) were classified as "poor sleepers" (global PSQI score > 5). Poor sleep quality was associated with female gender (P = 0.01), excessive daytime sleepiness (P < 0.01), lower total sleep time (P < 0.001), increased sleep onset latency (P < 0.001), and increased frequency of sleep disturbances (P < 0.001). Abnormal ESS scores (ESS > 10) were more prevalent among poor sleepers (P < 0.01) signifying increased level of daytime hypersomnolence. Sleep quality among Pakistani junior physicians is significantly poor. Efforts must be directed towards proper sleep hygiene education. Regulations regarding duty hour limitations need to be considered.

  4. Sleepiness and Sleep Disordered Breathing in Prader-Willi Syndrome: Relationship to Genotype, Growth Hormone Therapy, and Body Composition

    PubMed Central

    Williams, Korwyn; Scheimann, Ann; Sutton, Vernon; Hayslett, Elizabeth; Glaze, Daniel G.

    2008-01-01

    Study Objectives: Patients with Prader-Willi syndrome (PWS) suffer from excessive sleepiness and sleep disordered breathing (SDB). We reviewed the polysomnograms (PSGs) and multiple sleep latency tests (MSLTs) in a cohort of PWS patients to determine the relationship of BMIz scores, daytime sleepiness, growth hormone (GH) treatments, and SDB. Methods: Attended overnight PSGs were performed for PWS patients referred for concern for SDB between January 2000 and January 2005. Age at time of study, genotype, use and dose of GH, sleepiness scale, normalized body-mass index (BMIz), total sleep time, latency to stage I and REM sleep, sleep stage percentages, apnea-hypopnea index (AHI), central apnea (CA) frequency, oxygen saturation nadir, maximum carbon dioxide tension, periodic limb movement index, presence of snoring, normality of EEG, and, in several patients, mean sleep latency testing were determined. Results: All patients exhibited some form of SDB. There was a positive correlation between the BMIz and AHI. The BMIz was significantly different between GH–treated and –untreated groups, but there was not a significant difference between AHI, CA, oxygen nadir, or maximum carbon dioxide tension of the GH–treated and –untreated groups. There was no significant correlation between the MSLT and the sleepiness scale or AHI. There was also no significant difference between the AHIs of patients with different genetic defects. Conclusions: There should be a low threshold for obtaining PSG to evaluate SDB, but the type and severity of SDB were not predictable based on a sleepiness scale score, BMIz, or underlying genetic defect. Citation: Williams K; Scheimann A; Sutton V; Hayslett E; Glaze DG. Sleepiness and sleep disordered breathing in Prader-Willi syndrome: relationship to genotype, growth hormone therapy, and body composition. J Clin Sleep Med 2007;4(2):111–118. PMID:18468308

  5. Exposure to Radiofrequency Electromagnetic Fields and Sleep Quality: A Prospective Cohort Study

    PubMed Central

    Mohler, Evelyn; Frei, Patrizia; Fröhlich, Jürg; Braun-Fahrländer, Charlotte; Röösli, Martin

    2012-01-01

    Background There is persistent public concern about sleep disturbances due to radiofrequency electromagnetic field (RF-EMF) exposure. The aim of this prospective cohort study was to investigate whether sleep quality is affected by mobile phone use or by other RF-EMF sources in the everyday environment. Methods We conducted a prospective cohort study with 955 study participants aged between 30 and 60 years. Sleep quality and daytime sleepiness was assessed by means of standardized questionnaires in May 2008 (baseline) and May 2009 (follow-up). We also asked about mobile and cordless phone use and asked study participants for consent to obtain their mobile phone connection data from the mobile phone operators. Exposure to environmental RF-EMF was computed for each study participant using a previously developed and validated prediction model. In a nested sample of 119 study participants, RF-EMF exposure was measured in the bedroom and data on sleep behavior was collected by means of actigraphy during two weeks. Data were analyzed using multivariable regression models adjusted for relevant confounders. Results In the longitudinal analyses neither operator-recorded nor self-reported mobile phone use was associated with sleep disturbances or daytime sleepiness. Also, exposure to environmental RF-EMF did not affect self-reported sleep quality. The results from the longitudinal analyses were confirmed in the nested sleep study with objectively recorded exposure and measured sleep behavior data. Conclusions We did not find evidence for adverse effects on sleep quality from RF-EMF exposure in our everyday environment. PMID:22624036

  6. Sleep quantity, quality, and insomnia symptoms of medical students during clinical years

    PubMed Central

    Alsaggaf, Mohammed A.; Wali, Siraj O.; Merdad, Roah A.; Merdad, Leena A.

    2016-01-01

    Objectives: To determine sleep habits and sleep quality in medical students during their clinical years using validated measures; and to investigate associations with academic performance and psychological stress. Methods: In this cross-sectional study, medical students (n=320) were randomly selected from a list of all enrolled clinical-year students in a Saudi medical school from 2011-2012. Students filled a questionnaire including demographic and lifestyle factors, Pittsburgh Sleep Quality Index, Epworth Sleepiness Scale, and Perceived Stress Scale. Results: Students acquired on average, 5.8 hours of sleep each night, with an average bedtime at 01:53. Approximately 8% reported acquiring sleep during the day, and not during nighttime. Poor sleep quality was present in 30%, excessive daytime sleepiness (EDS) in 40%, and insomnia symptoms in 33% of students. Multivariable regression models revealed significant associations between stress, poor sleep quality, and EDS. Poorer academic performance and stress were associated with symptoms of insomnia. Conclusion: Sleep deprivation, poor sleep quality, and EDS are common among clinical years medical students. High levels of stress and the pressure of maintaining grade point averages may be influencing their quality of sleep. PMID:26837401

  7. Cyclic Alternating Pattern in Obstructive Sleep Apnea Patients with versus without Excessive Sleepiness.

    PubMed

    Korkmaz, Selda; Bilecenoglu, Nedime Tugce; Aksu, Murat; Yoldas, Tahir Kurtulus

    2018-01-01

    One of the main hypotheses on the development of daytime sleepiness (ES) is increased arousal in obstructive sleep apnea (OSA). Cyclic alternating pattern (CAP) is considered to be the main expression of sleep microstructure rather than arousal. Therefore, we aimed to investigate whether there is any difference between OSA patients with versus without ES in terms of the parameters of sleep macro- and microstructure and which variables are associated with Epworth Sleepiness Scale (ESS) score. Thirty-eight male patients with moderate to severe OSA were divided into two subgroups by having been used to ESS as ES or non-ES. There was no difference between two groups in clinical characteristics and macrostructure parameters of sleep. However, ES group had significantly higher CAP rate, CAP duration, number of CAP cycles, and duration and rate of the subtypes A2 ( p = 0.033, 0.019, 0.013, and 0.019, respectively) and lower mean phase B duration ( p = 0.028) compared with non-ES group. In correlation analysis, ESS score was not correlated with any CAP measure. OSA patients with ES have increased CAP measures rather than those without ES.

  8. Sleep quantity, quality, and insomnia symptoms of medical students during clinical years. Relationship with stress and academic performance.

    PubMed

    Alsaggaf, Mohammed A; Wali, Siraj O; Merdad, Roah A; Merdad, Leena A

    2016-02-01

    To determine sleep habits and sleep quality in medical students during their clinical years using validated measures; and to investigate associations with academic performance and psychological stress. In this cross-sectional study, medical students (n=320) were randomly selected from a list of all enrolled clinical-year students in a Saudi medical school from 2011-2012. Students filled a questionnaire including demographic and lifestyle factors, Pittsburgh Sleep Quality Index, Epworth Sleepiness Scale, and Perceived Stress Scale. Students acquired on average, 5.8 hours of sleep each night, with an average bedtime at 01:53. Approximately 8% reported acquiring sleep during the day, and not during nighttime. Poor sleep quality was present in 30%, excessive daytime sleepiness (EDS) in 40%, and insomnia symptoms in 33% of students. Multivariable regression models revealed significant associations between stress, poor sleep quality, and EDS. Poorer academic performance and stress were associated with symptoms of insomnia. Sleep deprivation, poor sleep quality, and EDS are common among clinical years medical students. High levels of stress and the pressure of maintaining grade point averages may be influencing their quality of sleep.

  9. Prevalence and risk factors of excessive daytime sleepiness in a community sample of young children: the role of obesity, asthma, anxiety/depression, and sleep.

    PubMed

    Calhoun, Susan L; Vgontzas, Alexandros N; Fernandez-Mendoza, Julio; Mayes, Susan D; Tsaoussoglou, Marina; Basta, Maria; Bixler, Edward O

    2011-04-01

    We investigated the prevalence and association of excessive daytime sleepiness (EDS) with a wide range of factors (e.g., medical complaints, obesity, objective sleep [including sleep disordered breathing], and parent-reported anxiety/depression and sleep difficulties) in a large general population sample of children. Few studies have researched the prevalence and predictors of EDS in young children, none in a general population sample of children, and the results are inconsistent. Cross-sectional Population -based. 508 school-aged children from the general population. N/A. Children underwent a 9-hour polysomnogram (PSG), physical exam, and parent completed health, sleep and psychological questionnaires. Children were divided into 2 groups: those with and without parent reported EDS. The prevalence of subjective EDS was approximately 15%. Significant univariate relationships were found between children with EDS and BMI percentile, waist circumference, heartburn, asthma, and parent reported anxiety/depression, and sleep difficulties. The strongest predictors of EDS were waist circumference, asthma, and parent-reported symptoms of anxiety/depression and trouble falling asleep. All PSG sleep variables including apnea/hypopnea index, caffeine consumption, and allergies were not significantly related to EDS. It appears that the presence of EDS is more strongly associated with obesity, asthma, parent reported anxiety/depression, and trouble falling asleep than with sleep disordered breathing (SDB) or objective sleep disruption per se. Our findings suggest that children with EDS should be thoroughly assessed for anxiety/depression, nocturnal sleep difficulties, asthma, obesity, and other metabolic factors, whereas objective sleep findings may not be as clinically useful.

  10. Poor precompetitive sleep habits, nutrients' deficiencies, inappropriate body composition and athletic performance in elite gymnasts.

    PubMed

    Silva, M-R G; Paiva, T

    2016-09-01

    This study aimed to evaluate body composition, sleep, precompetitive anxiety and dietary intake on the elite female gymnasts' performance prior to an international competition. Sixty-seven rhythmic gymnasts of high performance level were evaluated in relation to sport and training practice, body composition, sleep duration, daytime sleepiness by the Epworth Sleepiness Scale (ESS), sleep quality by the Pittsburgh Sleep Quality Index (PSQI), precompetitive anxiety by the Sport Competition Anxiety Test form A (SCAT-A) and detailed dietary intake just before an international competition. Most gymnasts (67.2%) suffered from mild daytime sleepiness, 77.6% presented poor sleep quality and 19.4% presented high levels of precompetitive anxiety. The majority of gymnasts reported low energy availability (EA) and low intakes of important vitamins including folate, vitamins D, E and K; and minerals, including calcium, iron, boron and magnesium (p < .05). Gymnasts' performance was positively correlated with age (p = .001), sport practice (p = .024), number of daily training hours (p = .000), number of hours of training/week (p = .000), waist circumference (WC) (p = .008) and sleep duration (p = .005). However, it was negatively correlated with WC/hip circumference (p = .000), ESS (p = .000), PSQI (p = .042), SCAT-A (p = .002), protein g/kg (p = .028), EA (p = .002) and exercise energy expenditure (p = .000). High performance gymnasts presented poor sleep habits with consequences upon daytime sleepiness, sleep quality and low energy availability.

  11. Measuring subjective sleepiness at work in hospital nurses: validation of a modified delivery format of the Karolinska Sleepiness Scale.

    PubMed

    Geiger Brown, Jeanne; Wieroney, Margaret; Blair, Lori; Zhu, Shijun; Warren, Joan; Scharf, Steven M; Hinds, Pamela S

    2014-12-01

    Sleepiness during the work shift is common and can be hazardous to workers and, in the case of nurses, to patients under their care. Thus, measuring sleepiness in occupational studies is an important component of workplace health and safety. The Karolinska Sleepiness Scale (KSS) is usually used as a momentary assessment of a respondent's state of sleepiness; however, end-of-shift measurement is sometimes preferred based on the study setting. We assessed the predictive validity of the KSS as an end-of-shift recall measurement, asking for "average" sleepiness over the shift and "highest" level of sleepiness during the shift. Hospital registered nurses (N=40) working 12-h shifts completed an end-of-shift diary over 4 weeks that included the National Aeronautical and Space Administration Task Load Index (NASA-TLX) work intensity items and the KSS (498 shifts over 4 weeks). Vigilant attention was assessed by measuring reaction time, lapses, and anticipations using a 10-min performance vigilance task (PVT) at the end of the shift. The Horne-Ostberg Questionnaire, Epworth Sleepiness Scale, General Sleep Disturbance Scale, and Cleveland Sleep Habits Questionnaire were also collected at baseline to assess factors that could be associated with higher sleepiness. We hypothesized that higher KSS scores would correlate with vigilant attention parameters reflective of sleepiness (slower reaction times and more lapses and anticipations on a performance vigilance task) and also with those factors known to produce higher sleepiness. These factors included the following: (1) working night shifts, especially for those with "morningness" trait; (2) working sequential night shifts; (3) having low physical and mental work demands and low time pressure; (4) having concomitant organic sleep disorders; and (5) having greater "trait" sleepiness (Epworth Sleepiness Scale). Linear mixed models and generalized linear mixed models were used to test associations that could assess the predictive

  12. Validated Measures of Insomnia, Function, Sleepiness, and Nasal Obstruction in a CPAP Alternatives Clinic Population.

    PubMed

    Lam, Austin S; Collop, Nancy A; Bliwise, Donald L; Dedhia, Raj C

    2017-08-15

    Although efficacious in the treatment of obstructive sleep apnea (OSA), continuous positive airway pressure (CPAP) can be difficult to tolerate, with long-term adherence rates approaching 50%. CPAP alternatives clinics specialize in the evaluation and treatment of CPAP-intolerant patients; yet this population has not been studied in the literature. To better understand these patients, we sought to assess insomnia, sleep-related functional status, sleepiness, and nasal obstruction, utilizing data from validated instruments. After approval from the Emory University Institutional Review Board, a retrospective chart review was performed from September 2015 to September 2016 of new patient visits at the Emory CPAP alternatives clinic. Patient demographics and responses were recorded from the Insomnia Severity Index, Functional Outcomes of Sleep Questionnaire-10 (FOSQ-10), Epworth Sleepiness Scale, and Nasal Obstruction Symptom Evaluation questionnaires. A total of 172 patients were included, with 81% having moderate-severe OSA. Most of the patients demonstrated moderate-severe clinical insomnia and at least moderate nasal obstruction. FOSQ-10 scores indicated sleep-related functional impairment in 88%. However, most patients did not demonstrate excessive daytime sleepiness. This patient population demonstrates significant symptomatology and functional impairment. Because of the severity of their OSA, they are at increased risk of complications. In order to mitigate the detrimental effects of OSA, these significantly impacted patients should be identified and encouraged to seek CPAP alternatives clinics that specialize in the treatment of this population. © 2017 American Academy of Sleep Medicine

  13. Epworth Sleepiness Scale may be an indicator for blood pressure profile and prevalence of coronary artery disease and cerebrovascular disease in patients with obstructive sleep apnea.

    PubMed

    Feng, Jing; He, Quan-ying; Zhang, Xi-long; Chen, Bao-yuan

    2012-03-01

    This study seeks to determine whether scores of a short questionnaire assessing subjective daytime sleepiness (Epworth Sleepiness Scale [ESS]) are associated with blood pressure (BP) level, BP profile, and prevalence of related coronary artery disease (CAD) and cerebrovascular disease (CVD) in obstructive sleep apnea (OSA) patients diagnosed by polysomnography (PSG). Twenty university hospital sleep centers in China mainland were organized by the Chinese Medical Association to participate in this study. Between January 2004 and April 2006, 2,297 consecutive patients (aged 18-85 years; 1,981 males and 316 females) referred to these centers were recruited. BP assessments were evaluated at four time points (daytime, evening, nighttime, and morning) under standardized conditions. Anthropometric measurements, medical history of hypertension, CAD, and CVD were collected. ESS score was calculated for each participant and at the night of BP assessment, nocturnal PSG was performed and subjects were classified into four groups based on the apnea-hypopnea index (AHI) from PSG as follows: control group (control, n = 213) with AHI < 5; mild sleep apnea (mild, n = 420) with AHI ≥ 5 and <15; moderate sleep apnea (moderate, n = 460) with AHI ≥ 15 and <30; and severe sleep apnea (severe, n = 1,204) with AHI ≥ 30. SPSS 11.5 software package was used for the relationships between ESS and BP profile and prevalence of CAD and CVD. ESS is correlated positively with average daytime, nighttime, evening, and morning BP before and even after controlling for confounding effects of age, sex, BMI, AHI, and nadir nocturnal oxygen saturation (before--r = 0.182, 0.326, 0.245, and 0.329, respectively, all P values < 0.001; after--r = 0.069, 0.212, 0.137, and 0.208, respectively, all P values < 0.001). In the severe group, nighttime, evening, morning average BPs (ABPs), the ratio of nighttime/daytime average BP (ratio of nighttime average BP to daytime average BP), and prevalence of

  14. 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

  15. Sleep Disturbances and Health-Related Quality of Life in Adults with Steady-State Bronchiectasis

    PubMed Central

    Lin, Zhiya; Tang, Yan; Lin, Zhimin; Li, Huimin; Gao, Yang; Luo, Qun; Zhong, Nanshan; Chen, Rongchang

    2014-01-01

    Background Sleep disturbances are common in patients with chronic lung diseases, but little is known about the prevalence in patients with bronchiectasis. A cross sectional study was conducted to investigate the prevalence and determinants associated with sleep disturbances, and the correlation between sleep disturbances and quality of life (QoL) in adults with steady-state bronchiectasis. Methods One hundred and forty-four bronchiectasis patients and eighty healthy subjects were enrolled. Sleep disturbances, daytime sleepiness, and QoL were measured by utilizing the Pittsburgh Sleep Quality Index (PSQI), Epworth Sleepiness Scale (ESS) and St. George Respiratory Questionnaire (SGRQ), respectively. Demographic, clinical indices, radiology, spirometry, bacteriology, anxiety and depression were also assessed. Results Adults with steady-state bronchiectasis had a higher prevalence of sleep disturbances (PSQI>5) (57% vs. 29%, P<0.001), but not daytime sleepiness (ESS≥10) (32% vs. 30%, P = 0.76), compared with healthy subjects. In the multivariate model, determinants associated with sleep disturbances in bronchiectasis patients included depression (OR, 10.09; 95% CI, 3.46–29.37; P<0.001), nocturnal cough (OR, 1.89; 95% CI, 1.13–3.18; P = 0.016), aging (OR, 1.04; 95% CI, 1.01–1.07; P = 0.009) and increased 24-hour sputum volume (OR, 2.01; 95% CI, 1.22–3.33; P = 0.006). Patients with sleep disturbances had more significantly impaired QoL affecting all domains than those without. Only 6.2% of patients reported using a sleep medication at least weekly. Conclusions In adults with steady-state bronchiectasis, sleep disturbances are more common than in healthy subjects and are related to poorer QoL. Determinants associated with sleep disturbances include depression, aging, nighttime cough and increased sputum volume. Assessment and intervention of sleep disturbances are warranted and may improve QoL. PMID:25036723

  16. Sleep disturbances and health-related quality of life in adults with steady-state bronchiectasis.

    PubMed

    Gao, Yonghua; Guan, Weijie; Xu, Gang; Lin, Zhiya; Tang, Yan; Lin, Zhimin; Li, Huimin; Gao, Yang; Luo, Qun; Zhong, Nanshan; Chen, Rongchang

    2014-01-01

    Sleep disturbances are common in patients with chronic lung diseases, but little is known about the prevalence in patients with bronchiectasis. A cross sectional study was conducted to investigate the prevalence and determinants associated with sleep disturbances, and the correlation between sleep disturbances and quality of life (QoL) in adults with steady-state bronchiectasis. One hundred and forty-four bronchiectasis patients and eighty healthy subjects were enrolled. Sleep disturbances, daytime sleepiness, and QoL were measured by utilizing the Pittsburgh Sleep Quality Index (PSQI), Epworth Sleepiness Scale (ESS) and St. George Respiratory Questionnaire (SGRQ), respectively. Demographic, clinical indices, radiology, spirometry, bacteriology, anxiety and depression were also assessed. Adults with steady-state bronchiectasis had a higher prevalence of sleep disturbances (PSQI>5) (57% vs. 29%, P<0.001), but not daytime sleepiness (ESS≥10) (32% vs. 30%, P = 0.76), compared with healthy subjects. In the multivariate model, determinants associated with sleep disturbances in bronchiectasis patients included depression (OR, 10.09; 95% CI, 3.46-29.37; P<0.001), nocturnal cough (OR, 1.89; 95% CI, 1.13-3.18; P = 0.016), aging (OR, 1.04; 95% CI, 1.01-1.07; P = 0.009) and increased 24-hour sputum volume (OR, 2.01; 95% CI, 1.22-3.33; P = 0.006). Patients with sleep disturbances had more significantly impaired QoL affecting all domains than those without. Only 6.2% of patients reported using a sleep medication at least weekly. In adults with steady-state bronchiectasis, sleep disturbances are more common than in healthy subjects and are related to poorer QoL. Determinants associated with sleep disturbances include depression, aging, nighttime cough and increased sputum volume. Assessment and intervention of sleep disturbances are warranted and may improve QoL.

  17. Prevalence of Sleep Disorders and Their Impacts on Occupational Performance: A Comparison between Shift Workers and Nonshift Workers

    PubMed Central

    Yazdi, Zohreh; Sadeghniiat-Haghighi, Khosro; Loukzadeh, Ziba; Elmizadeh, Khadijeh; Abbasi, Mahnaz

    2014-01-01

    The consequences of sleep deprivation and sleepiness have been noted as the most important health problem in our modern society among shift workers. The objective of this study was to investigate the prevalence of sleep disorders and their possible effects on work performance in two groups of Iranian shift workers and nonshift workers. This study was designed as a cross-sectional study. The data were collected by PSQI, Berlin questionnaire, Epworth Sleepiness Scale, Insomnia Severity Index, and RLS Questionnaire. Occupational impact of different sleep disorders was detected by Occupational Impact of Sleep Disorder questionnaire. These questionnaires were filled in by 210 shift workers and 204 nonshift workers. There was no significant difference in the age, BMI, marital status, and years of employment in the two groups. Shift workers scored significantly higher in the OISD. The prevalence of insomnia, poor sleep quality, and daytime sleepiness was significantly higher in shift workers. Correlations between OISD scores and insomnia, sleep quality, and daytime sleepiness were significant. We concluded that sleep disorders should receive more attention as a robust indicator of work limitation. PMID:24977041

  18. Predicting sleepiness during an awake craniotomy.

    PubMed

    Itoi, Chihiro; Hiromitsu, Kentaro; Saito, Shoko; Yamada, Ryoji; Shinoura, Nobusada; Midorikawa, Akira

    2015-12-01

    An awake craniotomy is a safe neurological surgical technique that minimizes the risk of brain damage. During the course of this surgery, the patient is asked to perform motor or cognitive tasks, but some patients exhibit severe sleepiness. Thus, the present study investigated the predictive value of a patient's preoperative neuropsychological background in terms of sleepiness during an awake craniotomy. Thirty-seven patients with brain tumor who underwent awake craniotomy were included in this study. Prior to craniotomy, the patient evaluated cognitive status, and during the surgery, each patient's performance and attitude toward cognitive tasks were recorded by neuropsychologists. The present findings showed that the construction and calculation abilities of the patients were moderately correlated with their sleepiness. These results indicate that the preoperative cognitive functioning of patients was related to their sleepiness during the awake craniotomy procedure and that the patients who exhibited sleepiness during an awake craniotomy had previously experienced reduced functioning in the parietal lobe. Copyright © 2015 Elsevier B.V. All rights reserved.

  19. Sleepiness and Motor Vehicle Crashes in a Representative Sample of Portuguese Drivers: The Importance of Epidemiological Representative Surveys.

    PubMed

    Gonçalves, M; Peralta, A R; Monteiro Ferreira, J; Guilleminault, Christian

    2015-01-01

    Sleepiness is considered to be a leading cause of crashes. Despite the huge amount of information collected in questionnaire studies, only some are based on representative samples of the population. Specifics of the populations studied hinder the generalization of these previous findings. For the Portuguese population, data from sleep-related car crashes/near misses and sleepiness while driving are missing. The objective of this study is to determine the prevalence of near-miss and nonfatal motor vehicle crashes related to sleepiness in a representative sample of Portuguese drivers. Structured phone interviews regarding sleepiness and sleep-related crashes and near misses, driving habits, demographic data, and sleep quality were conducted using the Pittsburgh Sleep Quality Index and sleep apnea risk using the Berlin questionnaire. A multivariate regression analysis was used to determine the associations with sleepy driving (feeling sleepy or falling asleep while driving) and sleep-related near misses and crashes. Nine hundred subjects, representing the Portuguese population of drivers, were included; 3.1% acknowledged falling asleep while driving during the previous year and 0.67% recalled sleepiness-related crashes. Higher education, driving more than 15,000 km/year, driving more frequently between 12:00 a.m. and 6 a.m., fewer years of having a driver's license, less total sleep time per night, and higher scores on the Epworth Sleepiness Scale (ESS) were all independently associated with sleepy driving. Sleepiness-related crashes and near misses were associated only with falling asleep at the wheel in the previous year. Sleep-related crashes occurred more frequently in drivers who had also had sleep-related near misses. Portugal has lower self-reported sleepiness at the wheel and sleep-related near misses than most other countries where epidemiological data are available. Different population characteristics and cultural, social, and road safety specificities may

  20. 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.

  1. Poor sleep quality and nightmares are associated with non-suicidal self-injury in adolescents.

    PubMed

    Liu, Xianchen; Chen, Hua; Bo, Qi-Gui; Fan, Fang; Jia, Cun-Xian

    2017-03-01

    Non-suicidal self-injury (NSSI) is prevalent and is associated with increased risk of suicidal behavior in adolescents. This study examined which sleep variables are associated with NSSI, independently from demographics and mental health problems in Chinese adolescents. Participants consisted of 2090 students sampled from three high schools in Shandong, China and had a mean age of 15.49 years. Participants completed a sleep and health questionnaire to report their demographic and family information, sleep duration and sleep problems, impulsiveness, hopelessness, internalizing and externalizing problems, and NSSI. A series of regression analyses were conducted to examine the associations between sleep variables and NSSI. Of the sample, 12.6 % reported having ever engaged in NSSI and 8.8 % engaged during the last year. Univariate logistic analyses demonstrated that multiple sleep variables including short sleep duration, insomnia symptoms, poor sleep quality, sleep insufficiency, unrefreshed sleep, sleep dissatisfaction, daytime sleepiness, fatigue, snoring, and nightmares were associated with increased risk of NSSI. After adjusting for demographic and mental health variables, NSSI was significantly associated with sleeping <6 h per night, poor sleep quality, sleep dissatisfaction, daytime sleepiness, and frequent nightmares. Stepwise logistic regression model demonstrated that poor sleep quality (OR = 2.18, 95 % CI = 1.37-3.47) and frequent nightmares (OR = 2.88, 95 % CI = 1.45-5.70) were significantly independently associated with NSSI. In conclusion, while multiple sleep variables are associated with NSSI, poor sleep quality and frequent nightmares are independent risk factors of NSSI. These findings may have important implications for further research of sleep self-harm mechanisms and early detection and prevention of NSSI in adolescents.

  2. Factors associated with self-reported driver sleepiness and incidents in city bus drivers

    PubMed Central

    ANUND, Anna; IHLSTRÖM, Jonas; FORS, Carina; KECKLUND, Göran; FILTNESS, Ashleigh

    2016-01-01

    Driver fatigue has received increased attention during recent years and is now considered to be a major contributor to approximately 15–30% of all crashes. However, little is known about fatigue in city bus drivers. It is hypothesized that city bus drivers suffer from sleepiness, which is due to a combination of working conditions, lack of health and reduced sleep quantity and quality. The overall aim with the current study is to investigate if severe driver sleepiness, as indicated by subjective reports of having to fight sleep while driving, is a problem for city based bus drivers in Sweden and if so, to identify the determinants related to working conditions, health and sleep which contribute towards this. The results indicate that driver sleepiness is a problem for city bus drivers, with 19% having to fight to stay awake while driving the bus 2–3 times each week or more and nearly half experiencing this at least 2–4 times per month. In conclusion, severe sleepiness, as indicated by having to fight sleep during driving, was common among the city bus drivers. Severe sleepiness correlated with fatigue related safety risks, such as near crashes. PMID:27098307

  3. Factors associated with self-reported driver sleepiness and incidents in city bus drivers.

    PubMed

    Anund, Anna; Ihlström, Jonas; Fors, Carina; Kecklund, Göran; Filtness, Ashleigh

    2016-08-05

    Driver fatigue has received increased attention during recent years and is now considered to be a major contributor to approximately 15-30% of all crashes. However, little is known about fatigue in city bus drivers. It is hypothesized that city bus drivers suffer from sleepiness, which is due to a combination of working conditions, lack of health and reduced sleep quantity and quality. The overall aim with the current study is to investigate if severe driver sleepiness, as indicated by subjective reports of having to fight sleep while driving, is a problem for city based bus drivers in Sweden and if so, to identify the determinants related to working conditions, health and sleep which contribute towards this. The results indicate that driver sleepiness is a problem for city bus drivers, with 19% having to fight to stay awake while driving the bus 2-3 times each week or more and nearly half experiencing this at least 2-4 times per month. In conclusion, severe sleepiness, as indicated by having to fight sleep during driving, was common among the city bus drivers. Severe sleepiness correlated with fatigue related safety risks, such as near crashes.

  4. Sleepy driver near-misses may predict accident risks.

    PubMed

    Powell, Nelson B; Schechtman, Kenneth B; Riley, Robert W; Guilleminault, Christian; Chiang, Rayleigh Ping-ying; Weaver, Edward M

    2007-03-01

    To quantify the prevalence of self-reported near-miss sleepy driving accidents and their association with self-reported actual driving accidents. A prospective cross-sectional internet-linked survey on driving behaviors. Dateline NBC News website. Results are given on 35,217 (88% of sample) individuals with a mean age of 37.2 +/- 13 years, 54.8% women, and 87% white. The risk of at least one accident increased monotonically from 23.2% if there were no near-miss sleepy accidents to 44.5% if there were > or = 4 near-miss sleepy accidents (P < 0.0001). After covariate adjustments, subjects who reported at least one near-miss sleepy accident were 1.13 (95% CI, 1.10 to 1.16) times as likely to have reported at least one actual accident as subjects reporting no near-miss sleepy accidents (P < 0.0001). The odds of reporting at least one actual accident in those reporting > or = 4 near-miss sleepy accidents as compared to those reporting no near-miss sleepy accidents was 1.87 (95% CI, 1.64 to 2.14). Furthermore, after adjustments, the summary Epworth Sleepiness Scale (ESS) score had an independent association with having a near-miss or actual accident. An increase of 1 unit of ESS was associated with a covariate adjusted 4.4% increase of having at least one accident (P < 0.0001). A statistically significant dose-response was seen between the numbers of self-reported sleepy near-miss accidents and an actual accident. These findings suggest that sleepy near-misses may be dangerous precursors to an actual accident.

  5. 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.

  6. 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.

  7. Sleep disorders among French anaesthesiologists and intensivists working in public hospitals: a self-reported electronic survey.

    PubMed

    Richter, Elisa; Blasco, Valery; Antonini, François; Rey, Marc; Reydellet, Laurent; Harti, Karim; Nafati, Cyril; Albanèse, Jacques; Leone, Marc

    2015-02-01

    Sleep disorders can affect the health of physicians and patient outcomes. To determine the prevalence of sleep disorders among French anaesthesiologists and intensivists working in a public hospital. A cross-sectional survey. Anaesthesiologists and intensivists working in French public hospitals. Sleep quality was assessed using the Pittsburgh Sleep Quality Index (PSQI) and the Epworth Sleepiness Scale (ESS) was used to assess the degree of excessive daytime sleepiness. Among 1504 responders, 677 (45%) physicians reported sleep disorders. The independent factors associated with sleep disorders were reporting of sleep disorders [odds ratio (OR) 12.04, 95% CI (95% confidence interval) 8.89 to 16.46], sleep time less than 7 h (OR 8.86, 95% CI 6.50 to 12.20), work stress (OR 2.04, 95% CI 1.49 to 2.83), stress at home (OR 1.77, 95% CI 1.24 to 2.53), anxiolytic use (OR 3.69, 95% CI 2.23 to 6.25), psychotropic drug use (OR 3.91, 95% CI 1.51 to 11.52) and excessive daytime sleepiness (OR 1.81, 95% CI 1.34 to 2.45). Six hundred and seventy-six (44%) responders reported excessive daytime sleepiness during their professional activity. The independent factors associated with excessive daytime sleepiness were female sex (OR 1.86, 95% CI 1.49 to 2.34), tea consumption (OR 1.47, 95% CI 1.14 to 1.91), regular practice of nap (OR 1.68, 95% CI 1.34 to 2.09), stress at home (OR 1.31, 95% CI 1.02 to 1.68), more than four extended work shifts monthly (OR 1.25, 95% CI 1.01 to 1.56) and sleep disorders (OR 1.73, 95% CI 1.31 to 2.29). Reporting sleep disorder duration and a sleep time less than 7 h were the two major risk factors for sleep disorders. Female sex was the major risk factor for excessive daytime sleepiness. French anaesthesiologists did not report more sleep disorders than the general population, but their alertness is impaired by a factor of two.

  8. Comment on short-term variation in subjective sleepiness.

    PubMed

    Eriksen, Claire A; Akerstedt, Torbjörn; Kecklund, Göran; Akerstedt, Anna

    2005-12-01

    Subjective sleepiness at different times is often measured in studies on sleep loss, night work, or drug effects. However, the context at the time of rating may influence results. The present study examined sleepiness throughout the day at hourly intervals and during controlled activities [reading, writing, walking, social interaction (discussion), etc.] by 10-min. intervals for 3 hr. This was done on a normal working day preceded by a scheduled early rising (to invite sleepiness) for six subjects. Analysis showed a significant U-shaped pattern across the day with peaks in the early morning and late evening. A walk and social interaction were associated with low sleepiness, compared to sedentary and quiet office work. None of this was visible in the hourly ratings. There was also a pronounced afternoon increase in sleepiness, that was not observable with hourly ratings. It was concluded that there are large variations in sleepiness related to time of day and also to context and that sparse sampling of subjective sleepiness may miss much of this variation.

  9. Sleep apnea, daytime somnolence, and idiopathic dizziness--a novel association.

    PubMed

    Sowerby, Leigh J; Rotenberg, Brian; Brine, Meggan; George, Charles F P; Parnes, Lorne S

    2010-06-01

    To determine if an association exists between sleep apnea, daytime somnolence, and chronic idiopathic dizziness. Case-control study of new patients presenting to a tertiary neuro-otologic practice. A total of 46 subjects with idiopathic dizziness (ID), 20 positive controls with dizziness (benign paroxysmal positional vertigo [BPV]), and 69 negative controls with hearing loss (HL) but no dizziness were enrolled. Participants who were patients diagnosed with the above conditions and who met all other inclusion criteria completed a sleep questionnaire and had a complete physical exam and investigations to establish or exclude a neuro-otologic diagnosis. They were subsequently evaluated for risk of symptomatic sleep disturbance based on the Epworth Sleepiness Scale (ESS), the Berlin Questionnaire, and the Multivariable Apnea Risk Index (MAP). Statistical analysis was carried out using SPSS (SPSS Inc., Chicago, IL). There was no significant demographic difference among the groups in terms of age, sex, body mass index, neck size, alcohol consumption, or smoking. Using a cutoff of both 10 and 12 on the ESS, the ID were more likely to have significant daytime somnolence than the HL group, with a likelihood ratio (LR) of 7.8 for the ESS 12 score (P = .021) and 7.1 for the ESS 10 score (P = .029). Using the MAP score, a statistically significant difference between the ID group and both the BPV group (LR 3.99, P = .046) and the HL group (LR 5.46, P = .019) was found. This study suggests that a previously undescribed link between idiopathic dizziness, daytime somnolence, and sleep apnea might exist. Prospective investigation is warranted to determine whether treatment of any sleep issues resolves symptoms of idiopathic dizziness.

  10. Ready-Made Versus Custom-Made Mandibular Repositioning Devices in Sleep Apnea: A Randomized Clinical Trial.

    PubMed

    Johal, Ama; Haria, Priya; Manek, Seema; Joury, Easter; Riha, Renata

    2017-02-15

    To compare the effectiveness of a custom-made (MRDc) versus ready-made (MRDr) mandibular repositioning devices (MRD) in the management of obstructive sleep apnea (OSA). A randomized crossover trial design was adopted in which patients with a confirmed diagnosis of OSA were randomly allocated to receive either a 3-month period of ready-made or custom-made MRD, with an intervening washout period of 2 weeks, prior to crossover. Treatment outcomes included both objective sleep monitoring and patient-centered measures (daytime sleepiness, partner snoring and quality of life). Twenty-five patients, with a mild degree of OSA (apnea-hypopnea index of 13.3 [10.9-25] events/h) and daytime sleepiness (Epworth Sleepiness Scale of 11 [6-16]), completed both arms of the trial. The MRDc achieved a complete treatment response in 64% of participants, compared with 24% with the MRDr (p < 0.001). A significant difference was observed in treatment failures, when comparing the MRDr (36%) with the MRDc (4%). Excessive daytime sleepiness (Epworth Sleepiness Scale ≥ 10) persisted in 33% (MRDc) and 66% (MRDr) of OSA subjects, following treatment. A statistically significant improvement was observed in quality of life scales following MRDc therapy only. Significant differences were observed in relation to both the number of nights per week (p = 0.004) and hours per night (p = 0.006) between the two different designs of device. The study demonstrates the significant clinical effectiveness of a custom-made mandibular repositioning device, particularly in terms of patient compliance and tolerance, in the treatment of OSA. © 2017 American Academy of Sleep Medicine

  11. Modification of the Epworth Sleepiness Scale in Central China.

    PubMed

    Zhang, Jin Nong; Peng, Bo; Zhao, Ting Ting; Xiang, Min; Fu, Wei; Peng, Yi

    2011-12-01

    The well-known excessive daytime sleepiness (EDS) assessment, Epworth Sleepiness Scale (ESS), is not consistently qualified for patients with diverse living habits. This study is aimed to build a modified ESS (mESS) and then to verify its feasibility in the assessment of EDS for patients with suspected sleep-disordered breathing (SDB) in central China. A Ten-item Sleepiness Questionnaire (10-ISQ) was built by adding two backup items to the original ESS. Then the 10-ISQ was administered to 122 patients in central China with suspected SDB [among them, 119 cases met the minimal diagnostic criteria for obstructive sleep apnea by sleep study, e.g., apnea and hypopnea index (AHI) ≥ 5 h(-1)] and 117 healthy central Chinese volunteers without SDB. Multivariate exploratory techniques were used for item validation. The unreliable item in the original ESS was replaced by the eligible backup item, thus a modified ESS (mESS) was built, and then verified. Item 8 proved to be the only unreliable item in central Chinese patients, with the least factor loading on the main factor and the lowest item-total correlation both in the 10-ISQ and in the original ESS, deletion of it would increase the Cronbach's alpha (from 0.86 to 0.87 in the 10-ISQ; from 0.83 to 0.85 in the original ESS). The mESS was subsequently built by replacing item 8 in the original ESS with item 10 in the 10-ISQ. Verification with patients' responses revealed that the mESS was a single-factor questionnaire with good internal consistency (Cronbach's alpha = 0.86). The sum score of the mESS not only correlated with AHI (P < 0.01) but was also able to discriminate the severity of obstructive apnea (P < 0.01). Nasal CPAP treatment for severe OSA reduced the score significantly (P < 0.001). The performance of the mESS was poor in evaluating normal subjects. The mESS improves the validity of ESS for our patients. Therefore, it is justified to use it instead of the original one in assessment of EDS for patients with SDB

  12. [THE EMPIRICAL DISTINCTIVENESS OF WORK ENGAGEMENT AND WORKAHOLISM AMONG HOSPITAL NURSES IN JAPAN : THE EFFECT ON SLEEP QUALITY AND JOB PERFORMANCE].

    PubMed

    Kubota, Kazumi; Shimazu, Akihito; Kawakami, Norito; Takahashi, Masaya; Nakata, Akinori; Schaufeli, Wilmar B

    2011-01-01

    The aim of the present study is to demonstrate the distinctiveness of work engagement and workaholism by examining their relationships with sleep quality and job performance. A total of 447 nurses from 3 hospitals in Japan were surveyed using a self-administrated questionnaire including Utrecht Work Engagement Scale (UWES), the Dutch Workaholism Scale (DUWAS), questions on sleep quality (7 items) regarding (1) difficulty initiating sleep, (2) difficulty maintaining sleep, (3) early morning awakening, (4) dozing off or napping in daytime, (5) excessive daytime sleepiness at work, (6) difficulty awakening in the morning, and (7) tiredness awakening in the morning, and the World Health Organization Health Work Performance Questionnaire. The Structural Equation Modeling showed that, work engagement was positively related to sleep quality and job performance whereas workaholism negatively to sleep quality and job performance. The findings suggest that work engagement and workaholism are conceptually distinctive and that the former is positively and the latter is negatively related to well-being (i.e., good sleep quality and job performance).

  13. Quality of life among untreated sleep apnea patients compared with the general population and changes after treatment with positive airway pressure.

    PubMed

    Bjornsdottir, Erla; Keenan, Brendan T; Eysteinsdottir, Bjorg; Arnardottir, Erna Sif; Janson, Christer; Gislason, Thorarinn; Sigurdsson, Jon Fridrik; Kuna, Samuel T; Pack, Allan I; Benediktsdottir, Bryndis

    2015-06-01

    Obstructive sleep apnea leads to recurrent arousals from sleep, oxygen desaturations, daytime sleepiness and fatigue. This can have an adverse impact on quality of life. The aims of this study were to compare: (i) quality of life between the general population and untreated patients with obstructive sleep apnea; and (ii) changes of quality of life among patients with obstructive sleep apnea after 2 years of positive airway pressure treatment between adherent patients and non-users. Propensity score methodologies were used in order to minimize selection bias and strengthen causal inferences. The enrolled obstructive sleep apnea subjects (n = 822) were newly diagnosed with moderate to severe obstructive sleep apnea who were starting positive airway pressure treatment, and the general population subjects (n = 742) were randomly selected Icelanders. The Short Form 12 was used to measure quality of life. Untreated patients with obstructive sleep apnea had a worse quality of life when compared with the general population. This effect remained significant after using propensity scores to select samples, balanced with regard to age, body mass index, gender, smoking, diabetes, hypertension and cardiovascular disease. We did not find significant overall differences between full and non-users of positive airway pressure in improvement of quality of life from baseline to follow-up. However, there was a trend towards more improvement in physical quality of life for positive airway pressure-adherent patients, and the most obese subjects improved their physical quality of life more. The results suggest that co-morbidities of obstructive sleep apnea, such as obesity, insomnia and daytime sleepiness, have a great effect on life qualities and need to be taken into account and addressed with additional interventions. © 2014 European Sleep Research Society.

  14. Relationships between self-reported sleep quality components and cognitive functioning in breast cancer survivors up to 10 years following chemotherapy.

    PubMed

    Henneghan, Ashley M; Carter, Patricia; Stuifbergan, Alexa; Parmelee, Brennan; Kesler, Shelli

    2018-04-23

    Links have been made between aspects of sleep quality and cognitive function in breast cancer survivors (BCS), but findings are heterogeneous. The objective of this study is to examine relationships between specific sleep quality components (latency, duration, efficiency, daytime sleepiness, sleep disturbance, use of sleep aids) and cognitive impairment (performance and perceived), and determine which sleep quality components are the most significant contributors to cognitive impairments in BCS 6 months to 10 years post chemotherapy. Women 21 to 65 years old with a history of non-metastatic breast cancer following chemotherapy completion were recruited. Data collection included surveys to evaluate sleep quality and perceived cognitive impairments, and neuropsychological testing to evaluate verbal fluency and memory. Descriptive statistics, bivariate correlations, and hierarchical multiple regression were calculated. 90 women (mean age 49) completed data collection. Moderate significant correlations were found between daytime dysfunction, sleep efficiency, sleep latency, and sleep disturbance and perceived cognitive impairment (Rs = -0.37 to -0.49, Ps<.00049), but not objective cognitive performance of verbal fluency, memory or attention. After accounting for individual and clinical characteristics, the strongest predictors of perceived cognitive impairments were daytime dysfunction, sleep efficiency, and sleep disturbance. Findings support links between sleep quality and perceived cognitive impairments in BCS and suggest specific components of sleep quality (daytime dysfunction, sleep efficiency, and sleep disturbance) are associated with perceived cognitive functioning in this population. Findings can assist clinicians in guiding survivors to manage sleep and cognitive problems and aid in the design of interventional research. This article is protected by copyright. All rights reserved.

  15. Sleep disorders in patients with multiple sclerosis.

    PubMed

    Čarnická, Zuzana; Kollár, Branislav; Šiarnik, Pavel; Krížová, Lucia; Klobučníková, Katarína; Turčáni, Peter

    2015-04-15

    Poor sleep is a frequent symptom in patients with multiple sclerosis (MS). The objective of the study was to assess the relationship between nocturnal polysomnographic (PSG) findings and quality of sleep, fatigue, and increased daytime sleepiness among patients with MS. Clinical characteristics were collected. Pittsburgh Sleep Quality Index (PSQI), Fatigue Severity Scale (FSS), Epworth Sleepiness Scale (ESS), and International Restless Legs Syndrome Rating Scale were used to assess quality of sleep, fatigue, excessive daytime sleepiness, and the presence of restless legs syndrome (RLS). All patients underwent nocturnal diagnostic PSG examination. Fifty patients with MS were enrolled into the study. Age was the only independent variable significantly determining apnea-hypopnea index and desaturation index (DI) (beta = 0.369, p = 0.010, beta 0.301, p = 0.040). PSQI and ESS score were significantly higher in a population with RLS (p = 0.004, p = 0.011). FSS significantly correlated with DI (r = 0.400, p = 0.048). Presence of RLS was the only independent variable significantly determining PSQI and ESS (p = 0.005, p = 0.025). DI and presence of RLS were independent variables determining FSS (p = 0.015, p = 0.024). Presence of RLS seems to be the main factor determining poor sleep, fatigue, and daytime somnolence. Sleep disordered breathing and its severity influences only fatigue in patients with MS. © 2015 American Academy of Sleep Medicine.

  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. Functional Impairment in Adult Sleepwalkers: A Case-Control Study

    PubMed Central

    Lopez, Regis; Jaussent, Isabelle; Scholz, Sabine; Bayard, Sophie; Montplaisir, Jacques; Dauvilliers, Yves

    2013-01-01

    Study Objectives: To investigate the restorative quality of sleep and daytime functioning in sleepwalking adult patients in comparison with controls. Design: Prospective case-control study. Setting: Data were collected at the Sleep Disorders Center, Hôpital-Gui-de Chauliac, Montpellier, France between June 2007 and January 2011. Participants: There were 140 adult sleepwalkers (100 (median age 30 y, 55% male) in whom primary SW was diagnosed) who underwent 1 night of video polysomnography. All patients participated in a standardized clinical interview and completed a battery of questionnaires to assess clinical characteristics of parasomnia, daytime sleepiness, fatigue, insomnia, depressive and anxiety symptoms, and health-related quality of life. Results were compared with those of 100 sex- and age-matched normal controls. Interventions: N/A. Measurements and Results: Of the sleepwalkers, 22.3% presented with daily episodes and 43.5% presented with weekly episodes. Median age at sleepwalking onset was 9 y. Familial history of sleepwalking was reported in 56.6% of sleepwalkers and violent sleep related behaviors in 57.9%, including injuries requiring medical care for at least one episode in 17%. Significant associations were found between sleepwalking and daytime sleepiness, fatigue, insomnia, depressive and anxiety symptoms, and altered quality of life. Early-onset sleepwalkers had higher frequency of violent behaviors and injuries. Sleepwalkers with violent behaviors had higher frequency of sleep terrors and triggering factors, with greater alteration in health-related quality of life. Conclusion: Adult sleepwalking is a potentially serious condition that may induce violent behaviors, self-injury or injury to bed partners, sleep disruption, excessive daytime sleepiness, fatigue, and psychological distress, all of which affect health-related quality of life. Citation: Lopez R; Jaussent I; Scholz S; Bayard S; Montplaisir J; Dauvilliers Y. Functional impairment in

  18. Functional impairment in adult sleepwalkers: a case-control study.

    PubMed

    Lopez, Regis; Jaussent, Isabelle; Scholz, Sabine; Bayard, Sophie; Montplaisir, Jacques; Dauvilliers, Yves

    2013-03-01

    To investigate the restorative quality of sleep and daytime functioning in sleepwalking adult patients in comparison with controls. Prospective case-control study. Data were collected at the Sleep Disorders Center, Hôpital-Gui-de Chauliac, Montpellier, France between June 2007 and January 2011. There were 140 adult sleepwalkers (100 (median age 30 y, 55% male) in whom primary SW was diagnosed) who underwent 1 night of video polysomnography. All patients participated in a standardized clinical interview and completed a battery of questionnaires to assess clinical characteristics of parasomnia, daytime sleepiness, fatigue, insomnia, depressive and anxiety symptoms, and health-related quality of life. Results were compared with those of 100 sex- and age-matched normal controls. N/A. Of the sleepwalkers, 22.3% presented with daily episodes and 43.5% presented with weekly episodes. Median age at sleepwalking onset was 9 y. Familial history of sleepwalking was reported in 56.6% of sleepwalkers and violent sleep related behaviors in 57.9%, including injuries requiring medical care for at least one episode in 17%. Significant associations were found between sleepwalking and daytime sleepiness, fatigue, insomnia, depressive and anxiety symptoms, and altered quality of life. Early-onset sleepwalkers had higher frequency of violent behaviors and injuries. Sleepwalkers with violent behaviors had higher frequency of sleep terrors and triggering factors, with greater alteration in health-related quality of life. Adult sleepwalking is a potentially serious condition that may induce violent behaviors, self-injury or injury to bed partners, sleep disruption, excessive daytime sleepiness, fatigue, and psychological distress, all of which affect health-related quality of life. Lopez R; Jaussent I; Scholz S; Bayard S; Montplaisir J; Dauvilliers Y. Functional impairment in adult sleepwalkers: a case-control study. SLEEP 2013;36(3):345-351.

  19. Aging reduces the association between sleepiness and sleep-disordered breathing

    PubMed Central

    Morrell, Mary; Finn, Laurel; McMillian, Alison; Peppard, Paul E.

    2013-01-01

    Aim To investigate age-related changes in sleepiness symptoms associated with sleep disordered breathing (SDB). Methods Wisconsin Sleep Cohort participants were assessed with polysomnography, Epworth Sleepiness Scale (ESS) and Multiple Sleep Latency Test (MSLT). SDB was defined as an apnea/hypopnea index≥15 events/hour, sleepiness as ESS≥10 and MSLT≤5 minutes. Odds ratios were calculated using generalized estimating equations associating sleepiness with SDB, and conditional logistic regression examining changes in longitudinal sleepiness status (ESS only). Models were, a priori, stratified by gender. Results ESS was measured in 1281 participants and MSLT in 998, at multiple time points (ESS n=3695; MSLT n=1846). Significant interactions were found between SDB and age in men, but not women. The odds ratios (OR) modeled for sleepiness in a 40 year old male with SDB were significant, compared to a male without SDB (OR: ESS 2.1; MSLT 2.9); however, these associations were not significant at 60 years. The within-subject odds ratio for sleepiness was also significant at 40 years (OR: 3.4), but not at 60 years. Conclusion The age-related reductions in the association between sleepiness and SDB may have clinical implications for the diagnosis and treatment of SDB in older people since sleepiness is often used as a therapeutic marker. PMID:22241742

  20. Central Sleep Apnea - Mayo Clinic

    MedlinePlus

    ... up Difficulty staying asleep (insomnia) Excessive daytime sleepiness (hypersomnia) Chest pain at night Difficulty concentrating Mood changes ... chronically fatigued, sleepy and irritable. Excessive daytime drowsiness (hypersomnia) may be due to other disorders, such as ...

  1. 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

  2. Using Latent Sleepiness to Evaluate an Important Effect of Promethazine

    NASA Technical Reports Server (NTRS)

    Feiveson, Alan H.; Hayat, Matthew; Vksman, Zalman; Putcha, Laksmi

    2007-01-01

    Astronauts often use promethazine (PMZ) to counteract space motion sickness; however PMZ may cause drowsiness, which might impair cognitive function. In a NASA ground study, subjects received PMZ and their cognitive performance was then monitored over time. Subjects also reported sleepiness using the Karolinska Sleepiness Score (KSS), which ranges from 1 - 9. A problem arises when using KSS to establish an association between true sleepiness and performance because KSS scores tend to overly concentrate on the values 3 (fairly awake) and 7 (moderately tired). Therefore, we defined a latent sleepiness measure as a continuous random variable describing a subject s actual, but unobserved true state of sleepiness through time. The latent sleepiness and observed KSS are associated through a conditional probability model, which when coupled with demographic factors, predicts performance.

  3. Circadian phase, dynamics of subjective sleepiness and sensitivity to blue light in young adults complaining of a delayed sleep schedule.

    PubMed

    Moderie, Christophe; Van der Maren, Solenne; Dumont, Marie

    2017-06-01

    To assess factors that might contribute to a delayed sleep schedule in young adults with sub-clinical features of delayed sleep phase disorder. Two groups of 14 young adults (eight women) were compared: one group complaining of a delayed sleep schedule and a control group with an earlier bedtime and no complaint. For one week, each subject maintained a target bedtime reflecting their habitual sleep schedule. Subjects were then admitted to the laboratory for the assessment of circadian phase (dim light melatonin onset), subjective sleepiness, and non-visual light sensitivity. All measures were timed relative to each participant's target bedtime. Non-visual light sensitivity was evaluated using subjective sleepiness and salivary melatonin during 1.5-h exposure to blue light, starting one hour after target bedtime. Compared to control subjects, delayed subjects had a later circadian phase and a slower increase of subjective sleepiness in the late evening. There was no group difference in non-visual sensitivity to blue light, but we found a positive correlation between melatonin suppression and circadian phase within the delayed group. Our results suggest that a late circadian phase, a slow build-up of sleep need, and an increased circadian sensitivity to blue light contribute to the complaint of a delayed sleep schedule. These findings provide targets for strategies aiming to decreasing the severity of a sleep delay and the negative consequences on daytime functioning and health. Copyright © 2017 Elsevier B.V. All rights reserved.

  4. A meta-analysis of the effect of media devices on sleep outcomes

    PubMed Central

    Carter, Ben; Rees, Philippa; Hale, Lauren; Bhattacharjee, Darsharna; Paradkar, Mandar

    2017-01-01

    Importance Sleep is vital to children’s bio-psycho-social development. Inadequate sleep quantity and quality is a public health concern with an array of detrimental health outcomes. Portable mobile and media device have become a ubiquitous part of children’s lives and may impact children’s sleep duration and quality. Objective This systematic review was conducted to examine the effect of portable media devices (e.g. mobile phones, and tablet devices) on sleep outcomes Data Sources A search strategy was developed and searches of the published and grey literature were conducted across 12 databases from January 1st 2011 to June 15th, 2015. No language restriction was applied. Study Selection We included randomized controlled trials; cohort; and cross sectional study designs. Of 467 studies identified, 20 cross-sectional studies were assessed for quality Data Extraction and Synthesis Data extraction and quality assessment was independently carried out by two reviewers and disagreements resolved by a third. Data was pooled in a random-effects meta-analysis, and an individual participant meta-analysis was carried out where possible. Main Outcomes and Measures The primary outcomes were: inadequate sleep quantity; poor sleep quality; and excessive daytime sleepiness, carried out following an a priori protocol. Results Twenty studies were included and quality assessed, involving 125,198 children, 50.1% were male. There was a strong and consistent association between bedtime media device use and: inadequate sleep quantity (OR =2.17; 95%CI 1.42-3.32); poor sleep quality (OR=1.46; 95%CI 1.14-1.88); and excessive daytime sleepiness (OR=2.72; 95%CI 1.32-5.61). Additionally, children who had access to (but did not use) media devices at night were more likely to have inadequate: sleep quantity (OR=1.79; 95%CI 1.39-2.31); sleep quality (OR=1.53; 95%CI 1.11-2.10); and daytime sleepiness (OR=2.27; 95%CI 1.54-3.35). Conclusions and relevance This was the first meta-analysis of

  5. Self-evaluated and Close Relative-Evaluated Epworth Sleepiness Scale vs. Multiple Sleep Latency Test in Patients with Obstructive Sleep Apnea

    PubMed Central

    Li, Yun; Zhang, Jihui; Lei, Fei; Liu, Hong; Li, Zhe; Tang, Xiangdong

    2014-01-01

    Objectives: The aims of this study were to determine (1) the agreement in Epworth Sleepiness Scale (ESS) evaluated by patients and their close relatives (CRs), and (2) the correlation of objective sleepiness as measured by multiple sleep latency test (MSLT) with self-evaluated and close relative-evaluated ESS. Methods: A total of 85 consecutive patients with obstructive sleep apnea (OSA) (70 males, age 46.7 ± 12.9 years old) with an apnea-hypopnea index (AHI) > 5 events per hour (mean 38.9 ± 26.8/h) were recruited into this study. All participants underwent an overnight polysomnographic assessment (PSG), MSLT, and ESS rated by both patients and their CRs. Mean sleep latency < 8 min on MSLT was considered objective daytime sleepiness. Results: Self-evaluated global ESS score (ESSG) was closely correlated with evaluation by CRs (r = 0.79, p < 0.001); the mean ESSG score evaluated by patients did not significantly differ from that evaluated by CRs (p > 0.05). However, Bland- Altman plot showed individual differences between self-evaluated and CR-evaluated ESS scores, with a 95%CI of -9.3 to 7.0. The mean sleep latency on MSLT was significantly associated with CR-evaluated ESSG (r = -0.23, p < 0.05); significance of association with self-evaluated ESSG was marginal (r = -0.21, p = 0.05). Conclusions: CR-evaluated ESS has a good correlation but also significant individual disagreement with self-evaluated ESS in Chinese patients with OSA. CR-evaluated ESS performs as well as, if not better than, self-evaluated ESS in this population when referring to MSLT. Citation: Li Y; Zhang J; Lei F; Liu H; Li Z; Tang X. Self-evaluated and close relative-evaluated Epworth Sleepiness Scale vs. multiple sleep latency test in patients with obstructive sleep apnea. J Clin Sleep Med 2014;10(2):171-176. PMID:24533000

  6. Sleep Apnea Crash Risk Study (Report)

    DOT National Transportation Integrated Search

    2004-09-01

    Sleep apnea is a condition in which a narrowing or closure of the upper airway during sleep causes repeated sleep disturbances, and possible complete awakenings, leading to poor sleep quality and excessive daytime sleepiness. The primary objectives o...

  7. Sleep, sleepiness and school start times: a preliminary study.

    PubMed

    Dexter, Donn; Bijwadia, Jagdeep; Schilling, Dana; Applebaugh, Gwendolyn

    2003-01-01

    High school students are reported to be excessively sleepy, resulting in decreased academic performance, increased psycho-social problems and increased risk of morbidity and mortality from accidents. Early school start times have been noted to contribute to this problem. This report attempts to confirm the relationship of early school start times with decreased sleep and increased sleepiness. We examined sophomore and junior students in 2 local high schools with different start times and measured the amount of time slept and sleepiness. We found that students at the early start school reported reduced sleep time and more sleepiness than their counterparts at the later starting school. Early school start times are associated with student reports of less sleep and increased sleepiness. Further studies in larger groups are recommended in view of the potential significant impact of sleep deprivation in this age group.

  8. Self-reported sleep lengths ≥ 9 hours among Swedish patients with stress-related exhaustion: Associations with depression, quality of sleep and levels of fatigue.

    PubMed

    Grossi, Giorgio; Jeding, Kerstin; Söderström, Marie; Osika, Walter; Levander, Maria; Perski, Aleksander

    2015-05-01

    Insomnia-type sleep disturbances are frequent among patients suffering from stress-related exhaustion disorder. However, clinical observations indicate that a subgroup suffer from sleep lengths frequently exceeding 9 hours, coupled with great daytime sleepiness. The aim of the present study was to investigate differences in socio-demographic variables, use of medications, sleep parameters, anxiety, depression and fatigue, between individuals with varying sleep lengths, in a sample of 420 Swedish patients (mean age 42 ± 9 years; 77% women) referred to treatment for exhaustion disorder. Patients were allocated to the groups: "never/seldom ≥ 9 hours" (n = 248), "sometimes ≥ 9 hours" (n = 115) and "mostly/always ≥ 9 hours" (n = 57), based on their self-rated frequency of sleep lengths ≥ 9 hours. The design was cross-sectional and data was collected by means of questionnaires at pre-treatment. Univariate analyses showed that patients in the "mostly/always ≥ 9 hours" group were more often on sick leave, and reported more depression and fatigue, better sleep quality and more daytime sleepiness, than patients in the other groups. Multivariate analyses showed that these patients scored higher on measures of fatigue than the rest of the sample independently of gender, use of antidepressants, sick leave, depression and quality of sleep. Patients suffering from exhaustion disorder and reporting excessive sleep seem to have a generally poorer clinical picture but better quality of sleep than their counterparts with shorter sleep lengths. The mechanisms underlying these differences, together with their prognostic value and implications for treatment remain to be elucidated in future studies.

  9. Sleep patterns and sleep disturbances across pregnancy.

    PubMed

    Mindell, Jodi A; Cook, Rae Ann; Nikolovski, Janeta

    2015-04-01

    This study sought to characterize sleep patterns and sleep problems in a large sample of women across all months of pregnancy. A total of 2427 women completed an Internet-based survey that included the Pittsburgh Sleep Quality Index (PSQI), Epworth Sleepiness Scale, vitality scale of the Short Form 36 Health Survey (SF-36), Insomnia Severity Index (ISI), Berlin questionnaire, International Restless Legs Syndrome (IRLS) question set, and a short version of the Pregnancy Symptoms Inventory (PSI). Across all months of pregnancy, women experienced poor sleep quality (76%), insufficient nighttime sleep (38%), and significant daytime sleepiness (49%). All women reported frequent nighttime awakenings (100%), and most women took daytime naps (78%). Symptoms of insomnia (57%), sleep-disordered breathing (19%), and restless legs syndrome (24%) were commonly endorsed, with no difference across the month of pregnancy for insomnia, sleep-disorder breathing, daytime sleepiness, or fatigue. In addition, high rates of pregnancy-related symptoms were found to disturb sleep, especially frequent urination (83%) and difficulty finding a comfortable sleep position (79%). Women experience significant sleep disruption, inadequate sleep, and high rates of symptoms of sleep disorder throughout pregnancy. These results suggest that all women should be screened and treated for sleep disturbances throughout pregnancy, especially given the impact of inadequate sleep and sleep disorders on fetal, pregnancy, and postpartum outcomes. Copyright © 2014 Elsevier B.V. All rights reserved.

  10. Epidemiology of insomnia in college students: relationship with mental health, quality of life, and substance use difficulties.

    PubMed

    Taylor, Daniel J; Bramoweth, Adam D; Grieser, Emily A; Tatum, Jolyn I; Roane, Brandy M

    2013-09-01

    The purpose of this study was to evaluate the prevalence and correlates of insomnia using rigorous diagnostic criteria and a comprehensive assessment battery. In a large sample (N=1,074) of college students (mean age 20.39years), participants were asked to complete a week-long sleep diary and comprehensive questionnaire packet assessing recommended daytime functioning domains (i.e., fatigue, quality of life, depression, anxiety, stress, academic performance, substance use) during the academic year. A significant portion of this sample of college students met proposed DSM-5 criteria for chronic insomnia (9.5%). The chronic insomnia group reported significantly worse sleep, fatigue, depression, anxiety, stress, and quality of life, and greater hypnotic and stimulant use for sleep problems. There were no differences between groups on excessive daytime sleepiness, academic performance, or substance use. This was a rigorous and comprehensive assessment of the prevalence and psychosocial correlates of insomnia. Insomnia is a significant problem in college students and should be regularly assessed. More research is also needed to guide treatment in this population. Copyright © 2012. Published by Elsevier Ltd.

  11. Improvement of sleep architecture parameters in cirrhotic patients with recurrent hepatic encephalopathy with the use of rifaximin.

    PubMed

    Bruyneel, Marie; Sersté, Thomas; Libert, Walter; van den Broecke, Sandra; Ameye, Lieveke; Dachy, Bernard; Mulkay, Jean-Pierre; Moreno, Christophe; Gustot, Thierry

    2017-03-01

    Sleep disorders are frequently reported in patients with cirrhosis and hepatic encephalopathy (HE). This study assessed the effect of rifaximin on sleep architecture parameters in patients with recurrent HE. This sequential, prospective, and exploratory study involved all patients with cirrhosis and recurrent HE admitted between June 2014 and September 2015. HE was assessed according to the West-Haven Classification. Patients underwent 24-h polysomnography (PSG) and 7-day actigraphy. Rapid eye movement (REM) sleep was considered to be an indicator of good sleep quality. Patients completed questionnaires assessing the quality of sleep and sleepiness. After a 28-day course of rifaximin, the same assessment was repeated. Fifteen patients were included (nine men, mean age: 57±11 years). Child-Pugh scores ranged from B7 to C15. Before rifaximin, the mean HE score was 2.7±0.7. Data from PSG analysis indicated long total sleep time (TST): 571±288 min, and limited REM sleep: 2.5% TST (0-19). Seven-day actigraphy showed an impaired number of steps: 1690/24 h (176-6945). Questionnaires indicated that patients experienced impaired sleep quality and excessive daytime sleepiness. After rifaximin, HE scores decreased to 1.7±0.6 (P<0.001). REM sleep increased to 8.5% TST (0-25) (P=0.003). No changes were observed for TST, number of steps, and on questionnaires. Patients with recurrent HE suffer from poor sleep quality and excessive daytime sleepiness. On 24-h PSG, rifaximin improves objective sleep architecture parameters with no changes in the subjective quality of sleep and sleepiness.

  12. Sleep characteristics, body mass index, and risk for hypertension in young adolescents.

    PubMed

    Peach, Hannah; Gaultney, Jane F; Reeve, Charlie L

    2015-02-01

    Inadequate sleep has been identified as a risk factor for a variety of health consequences. For example, short sleep durations and daytime sleepiness, an indicator of insufficient sleep and/or poor sleep quality, have been identified as risk factors for hypertension in the adult population. However, less evidence exists regarding whether these relationships hold within child and early adolescent samples and what factors mediate the relationship between sleep and risk for hypertension. Using data from the Study of Early Child Care and Youth Development, the present study examined body mass index (BMI) as a possible mediator for the effects of school-night sleep duration, weekend night sleep duration, and daytime sleepiness on risk for hypertension in a sample of sixth graders. The results demonstrated gender-specific patterns. Among boys, all three sleep characteristics predicted BMI and yielded significant indirect effects on risk for hypertension. Oppositely, only daytime sleepiness predicted BMI among girls and yielded a significant indirect effect on risk for hypertension. The findings provide clarification for the influence of sleep on the risk for hypertension during early adolescence and suggest a potential need for gender-specific designs in future research and application endeavors.

  13. Countermeasures for sleep loss and deprivation.

    PubMed

    Kushida, Clete A

    2006-09-01

    Sleep deprivation is ubiquitous and carries profound consequences in terms of personal and public health and safety. There is no substitute for a good night's sleep. Sleep that is optimal in quality and quantity for individuals, factoring in their age and personal sleep requirements, will minimize sleep debt and maximize daytime performance. Therefore, setting aside an adequate amount of time for sleep should be a priority; sleep should not be sacrificed at the expense of other activities of daily living. Nevertheless, there are certain therapeutic countermeasures available for individuals who are unable to obtain adequate sleep because of medical or sleep-related conditions (eg, narcolepsy, obstructive sleep apnea) when excessive daytime sleepiness is the main feature of the condition, or residual sleepiness despite treatment for the main conditions is present. These therapeutic countermeasures may also be considered in situations in which occupational constraints (eg, rotating shift work, military duty) dictate that constant or heightened vigilance is important or critical to work performance, crucial decision making, and/or survival. Exploration of the causes of sleep loss or deprivation, whether it is voluntary, or work or family induced, and/or the effects of a medical or sleep disorder, is a necessary first step in the evaluation of a patient who has significant daytime fatigue or sleepiness. Wake-promoting substances and medications such as caffeine, modafinil, methylphenidate, and dextroamphetamine may be considered in situations in which sleep loss is unavoidable or persists despite treatment of an underlying disorder that is characterized by or associated with daytime fatigue or sleepiness.

  14. A pathway underlying the impact of CPAP adherence on intimate relationship with bed partner in men with obstructive sleep apnea.

    PubMed

    Lai, Agnes Y K; Ip, Mary S M; Lam, Jamie C M; Weaver, Terri E; Fong, Daniel Y T

    2016-05-01

    Our aim was to determine the pathway underlying the effects of continuous positive airway pressure (CPAP) adherence on intimate relationship with bed partner in men with obstructive sleep apnea (OSA). We hypothesized that CPAP with good adherence affected the intimate relationship with bed partner directly and indirectly, and it was mediated through daytime sleepiness and activity level in men with OSA. Data were obtained from an education program for enhancing CPAP adherence. Men who were newly diagnosed of OSA and CPAP therapy naïve were recruited in a tertiary teaching hospital. Self-reported quality of life [Functional Outcomes of Sleep Questionnaire], daytime sleepiness [Epworth Sleepiness Scale (ESS)], and negative emotion symptoms [depression, anxiety, stress scale] were assessed before and after CPAP treatment at 1-year assessment. Seventy-three men were included in the data analysis, with a mean ± SD age of 52 ± 10 years, body mass index of 29.0 ± 5.2 kg/m(2), ESS of 9.5 ± 5.6, and median [interquartile range(IR)] apnea and hypopnea index of 31 (21, 56) events/h. The median (IR) CPAP daily usage was 4.3(0, 6.1) h/day. From the path analysis, CPAP therapy was shown to improve intimate relationship directly (ß = 0.185) and indirectly (ß = 0.050) by reducing daytime sleepiness and increasing activity level. However, negative emotion symptoms were not the mediators between CPAP adherence and the intimate relationship. CPAP therapy with good adherence is related directly and indirectly to a better intimate relationship with bed partner in men with OSA. It was possibly attributed to reduced daytime sleepiness and increased activity level.

  15. 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.

  16. Daytime napping associated with increased symptom severity in fibromyalgia syndrome.

    PubMed

    Theadom, Alice; Cropley, Mark; Kantermann, Thomas

    2015-02-07

    Previous qualitative research has revealed that people with fibromyalgia use daytime napping as a coping strategy for managing symptoms against clinical advice. Yet there is no evidence to suggest whether daytime napping is beneficial or detrimental for people with fibromyalgia. The purpose of this study was to explore how people use daytime naps and to determine the links between daytime napping and symptom severity in fibromyalgia syndrome. A community based sample of 1044 adults who had been diagnosed with fibromyalgia syndrome by a clinician completed an online questionnaire. Associations between napping behavior, sleep quality and fibromyalgia symptoms were explored using Spearman correlations, with possible predictors of napping behaviour entered into a logistic regression model. Differences between participants who napped on a daily basis and those who napped less regularly, as well as nap duration were explored. Daytime napping was significantly associated with increased pain, depression, anxiety, fatigue, memory difficulties and sleep problems. Sleep problems and fatigue explained the greatest amount of variance in napping behaviour, p < 0.010. Those who engaged in daytime naps for >30 minutes had higher memory difficulties (t = -3.45) and levels of depression (t = -2.50) than those who napped for shorter periods (<30 mins) (p < 0.010). Frequent use and longer duration of daytime napping was linked with greater symptom severity in people with fibromyalgia. Given the common use of daytime napping in people with fibromyalgia evidence based guidelines on the use of daytime napping in people with chronic pain are urgently needed.

  17. Sex difference in the association between habitual daytime napping and prevalence of diabetes: a population-based study.

    PubMed

    Sun, Kan; Li, Feng; Qi, Yiqin; Lin, Diaozhu; Ren, Meng; Xu, Mingtong; Li, Fangping; Li, Yan; Yan, Li

    2016-05-01

    Our objective was to evaluate the associations between habitual daytime napping and diabetes and whether it varies by sex, menopause, and sleep quality. We conducted a population-based cross-sectional study in 8621 eligible individuals aged 40 years or older. Information on daytime napping hours, night-time sleep duration, history of menstruation, and sleep quality was self-reported. Diabetes was diagnosed according to the 1999 World Health Organization diagnostic criteria. The prevalence of diabetes was 19.4 % in men and 15.6 % in women. Increased daytime napping hours were positively associated with parameters of glycometabolism in women, such as fasting plasma glucose, oral glucose tolerance test (OGTT) 2-h plasma glucose, and Hemoglobin A1c (HbA1c, all P for trend <0.05). In women, the prevalence of diabetes in no-habitual daytime napping group, 0-1-h daytime napping group, and more than 1-h daytime napping group were 14.5, 15.6, and 20.8 %, respectively (P for trend = 0.0004). A similar trend was detected in postmenopausal women (P for trend = 0.002). In multivariate logistic regression analysis, compared with no-habitual daytime napping postmenopausal women, those with daytime napping more than 1 h had higher prevalent diabetes (odds ratios 1.36, 95 % confidence interval, 1.04-1.77). In subgroup analysis of postmenopausal women, associations of daytime napping levels and prevalent diabetes were detected in older, overweight participants with good sleep quality who have not retired from work. In conclusion, our study suggests that habitual daytime napping is associated with prevalence of diabetes in postmenopausal women.

  18. Who is sleepier on the night shift? The influence of bio-psycho-social factors on subjective sleepiness of female nurses during the night shift.

    PubMed

    Zion, Nataly; Drach-Zahavy, Anat; Shochat, Tamar

    2018-07-01

    Sleepiness is a common complaint during the night shift and may impair performance. The current study aims to identify bio-psycho-social factors associated with subjective sleepiness during the night shift. Ninety-two female nurses working rotating shifts completed a sociodemographic questionnaire, the Munich ChronoType Questionaire for shift workers, the Pittsburg Sleep Quality Index, and the Pre-sleep Arousal Scale. Subjective sleepiness was measured hourly during two night shifts using the Karolinska Sleepiness Scale, and activity monitors assessed sleep duration 24-h before each shift. Findings showed that increased sleepiness was associated with increased age in nurses with early chronotypes and with more children. High cognitive pre-sleep arousal, but not sleep, was associated with increased sleepiness, especially in late chronotypes. The impact of bio-psycho-social factors on night shift sleepiness is complex, and depends on mutual interactions between these factors. Nurses most prone to increased sleepiness must develop personal strategies for maintaining vigilance on the night shift. Practitioner Summary: This study aims to identify bio-psycho-social factors associated with subjective sleepiness of female nurses during the night shift. Increasing sleepiness was associated with increased age in nurses with early chronotypes and with more children. Increased cognitive pre-sleep arousal, but not sleep, was associated with increased sleepiness, especially in late chronotypes.

  19. Do night naps impact driving performance and daytime recovery sleep?

    PubMed

    Centofanti, Stephanie A; Dorrian, Jillian; Hilditch, Cassie J; Banks, Siobhan

    2017-02-01

    Short, nighttime naps are used as a fatigue countermeasure in night shift work, and may offer protective benefits on the morning commute. However, there is a concern that nighttime napping may impact upon the quality of daytime sleep. The aim of the current project was to investigate the influence of short nighttime naps (<30min) on simulated driving performance and subsequent daytime recovery sleep. Thirty-one healthy subjects (aged 21-35 y; 18 females) participated in a 3-day laboratory study. After a 9-h baseline sleep opportunity (22:00h-07:00h), subjects were kept awake the following night with random assignment to: a 10-min nap ending at 04:00h plus a 10-min nap at 07:00h; a 30-min nap ending at 04:00h; or a no-nap control. A 40-min driving simulator task was administered at 07:00h and 18:30h post-recovery sleep. All conditions had a 6-h daytime recovery sleep opportunity (10:00h-16:00h) the next day. All sleep periods were recorded polysomnographically. Compared to control, the napping conditions did not significantly impact upon simulated driving lane variability, percentage of time in a safe zone, or time to first crash on morning or evening drives (p>0.05). Short nighttime naps did not significantly affect daytime recovery total sleep time (p>0.05). Slow wave sleep (SWS) obtained during the 30-min nighttime nap resulted in a significant reduction in SWS during subsequent daytime recovery sleep (p<0.05), such that the total amount of SWS in 24-h was preserved. Therefore, short naps did not protect against performance decrements during a simulated morning commute, but they also did not adversely affect daytime recovery sleep following a night shift. Further investigation is needed to examine the optimal timing, length or combination of naps for reducing performance decrements on the morning commute, whilst still preserving daytime sleep quality. Copyright © 2015 Elsevier Ltd. All rights reserved.

  20. Sleepiness at the wheel across Europe: a survey of 19 countries.

    PubMed

    Gonçalves, Marta; Amici, Roberto; Lucas, Raquel; Åkerstedt, Torbjörn; Cirignotta, Fabio; Horne, Jim; Léger, Damien; McNicholas, Walter T; Partinen, Markku; Téran-Santos, Joaquín; Peigneux, Philippe; Grote, Ludger

    2015-06-01

    The European Sleep Research Society aimed to estimate the prevalence, determinants and consequences of falling asleep at the wheel. In total, 12 434 questionnaires were obtained from 19 countries using an anonymous online questionnaire that collected demographic and sleep-related data, driving behaviour, history of drowsy driving and accidents. Associations were quantified using multivariate logistic regression. The average prevalence of falling asleep at the wheel in the previous 2 years was 17%. Among respondents who fell asleep, the median prevalence of sleep-related accidents was 7.0% (13.2% involved hospital care and 3.6% caused fatalities). The most frequently perceived reasons for falling asleep at the wheel were poor sleep in the previous night (42.5%) and poor sleeping habits in general (34.1%). Falling asleep was more frequent in the Netherlands [odds ratio = 3.55 (95% confidence interval: 1.97; 6.39)] and Austria [2.34 (1.75; 3.13)], followed by Belgium [1.52 (1.28; 1.81)], Portugal [1.34 (1.13, 1.58)], Poland [1.22 (1.06; 1.40)] and France [1.20 (1.05; 1.38)]. Lower odds were found in Croatia [0.36 (0.21; 0.61)], Slovenia [0.62 (0.43; 0.89)] and Italy [0.65 (0.53; 0.79)]. Individual determinants of falling asleep were younger age; male gender [1.79 (1.61; 2.00)]; driving ≥20 000 km year [2.02 (1.74; 2.35)]; higher daytime sleepiness [7.49 (6.26; 8.95)] and high risk of obstructive sleep apnea syndrome [3.48 (2.78; 4.36) in men]. This Pan European survey demonstrates that drowsy driving is a major safety hazard throughout Europe. It emphasizes the importance of joint research and policy efforts to reduce the burden of sleepiness at the wheel for European drivers.

  1. Association between maternal symptoms of sleep disordered breathing and fetal telomere length.

    PubMed

    Salihu, Hamisu M; King, Lindsey; Patel, Priyanshi; Paothong, Arnut; Pradhan, Anupam; Louis, Judette; Naik, Eknath; Marty, Phillip J; Whiteman, Valerie

    2015-04-01

    Our investigation aims to assess the impact of symptoms of maternal sleep-disordered breathing, specifically sleep apnea risk and daytime sleepiness, on fetal leukocyte telomere length. Pregnant women were recruited upon hospital delivery admission. Sleep exposure outcomes were measured using the Berlin Questionnaire to quantify sleep apnea and the Epworth Sleepiness Scale to measure daytime sleepiness. Participants were classified as "High Risk" or "Low Risk" for sleep apnea based on responses to the Berlin, while "Normal" or "Abnormal" daytime sleepiness was determined based on responses to the Epworth. Neonatal umbilical cord blood samples (N = 67) were collected and genomic DNA was isolated from cord blood leukocytes using Quantitative PCR. A ratio of relative telomere length was derived by telomere repeat copy number and single copy gene copy number (T/S ratio) and used to compare telomere lengths. Bootstrap and ANOVA statistical procedures were employed. On the Berlin, 68.7% of participants were classified as Low Risk while 31.3% were classified as High Risk for sleep apnea. According to the Epworth scale, 80.6% were determined to have Normal daytime sleepiness, and 19.4% were found to have Abnormal daytime sleepiness. The T/S ratio among pregnant women at High Risk for sleep apnea was significantly shorter than for those at Low Risk (P value < 0.05), and the T/S ratio among habitual snorers was significantly shorter than among non-habitual snorers (P value < 0.05). Although those with Normal Sleepiness had a longer T/S ratio than those with Abnormal Sleepiness, the difference was not statistically significant. Our results provide the first evidence demonstrating shortened telomere length among fetuses exposed to maternal symptoms of sleep disordered breathing during pregnancy, and suggest sleep disordered breathing as a possible mechanism of accelerated chromosomal aging. © 2015 Associated Professional Sleep Societies, LLC.

  2. Excessive sleepiness and self-reported shift work disorder: an Internet survey of shift workers.

    PubMed

    Lieberman, Joseph A; Sylvester, Lauren; Paik, Sharon

    2013-05-01

    To compare excessive sleepiness and quality of life (QoL) scores in shift workers who report having a diagnosis of shift work disorder (SWD) with those who report having no such diagnosis. An Internet-based survey was conducted between March and April 2009 that included shift workers with or without a self-reported diagnosis of SWD. Participation required working ≥ 21 hours/week for 2 weeks prior, a diagnosis of SWD or a score of ≥ 10 on the Epworth Sleepiness Scale, and a score of ≥ 5 on any subscale of the Sheehan Disability Scale. Surveys included 260 shift workers (103 with an SWD diagnosis and 157 without an SWD diagnosis). Diagnosed and undiagnosed respondents demonstrated similar Epworth Sleepiness Scale (13.7 vs 13.6, respectively) and Karolinska Sleepiness Scale (6.0 vs 5.5, respectively) scores. Sheehan Disability Scale social life and family life scores were similar between the 2 groups, although diagnosed respondents had a greater mean Sheehan Disability Scale work disability score compared with undiagnosed respondents (6.7 vs 5.5; P < 0.0001). Quality of life was more impaired in diagnosed patients in terms of ability to drive safely, propensity for accidents, work performance, and anxiety (P ≤ 0.039 vs undiagnosed). Work-related accidents (16% vs 5%; P = 0.0076) and injuries at work (17% vs 7%; P = 0.0233) were also reported by more diagnosed respondents than by undiagnosed respondents. Many respondents used caffeine and 57% of diagnosed respondents received prescription medication to treat symptoms of SWD. Individuals with diagnosed SWD demonstrated impairment in QoL and reported more work-related accidents and injuries, although many measures of QoL and prescription drug use were similar between groups. Shift work disorder is underrecognized by clinicians and patients, resulting in undertreatment, despite the availability of several behavioral and therapeutic treatment options.

  3. Are drivers aware of sleepiness and increasing crash risk while driving?

    PubMed

    Williamson, Ann; Friswell, Rena; Olivier, Jake; Grzebieta, Raphael

    2014-09-01

    Drivers are advised to take breaks when they feel too tired to drive, but there is question over whether they are able to detect increasing fatigue and sleepiness sufficiently to decide when to take a break. The aim of this study was to investigate the extent to which drivers have access to cognitive information about their current state of sleepiness, likelihood of falling asleep, and the implications for driving performance and the likelihood of crashing. Ninety drivers were recruited to do a 2h drive in a driving simulator. They were divided into three groups: one made ratings of their sleepiness, likelihood of falling asleep and likelihood of crashing over the next few minutes at prompts occurring at 200s intervals throughout the drive, the second rated sleepiness and likelihood of falling asleep at prompts but pressed a button on the steering wheel at any time if they felt they were near to crashing and the third made no ratings and only used a button-press if they felt a crash was likely. Fatigue and sleepiness was encouraged by monotonous driving conditions, an imposed shorter than usual sleep on the night before and by afternoon testing. Drivers who reported that they were possibly, likely or very likely to fall asleep in the next few minutes, were more than four times more likely to crash subsequently. Those who rated themselves as sleepy or likely to fall asleep had a more than 9-fold increase in the hazards of a centerline crossing compared to those who rated themselves as alert. The research shows clearly that drivers can detect changes in their levels of sleepiness sufficiently to make a safe decision to stop driving due to sleepiness. Therefore, road safety policy needs to move from reminding drivers of the signs of sleepiness and focus on encouraging drivers to respond to obvious indicators of fatigue and sleepiness and consequent increased crash risk. Copyright © 2014 Elsevier Ltd. All rights reserved.

  4. The napping behaviour of Australian university students.

    PubMed

    Lovato, Nicole; Lack, Leon; Wright, Helen

    2014-01-01

    The purpose of this study was to evaluate the self-reported sleep and napping behaviour of Australian university students and the relationship between napping and daytime functioning. A sample of 280 university first-year psychology students (median age  = 19.00 years) completed a 6-item napping behaviour questionnaire, a 12-item Daytime Feelings and Functioning Scale, the Pittsburg Sleep Quality Index and the Epworth Sleepiness Scale. Results indicated that 53.6% of students reported napping with 34% napping at least 1-2 times per week, and 17% napping three or more occasions per week. Long naps, those over 30 minutes, were taken by 77% of the napping students. Sixty-one percent of students reported they took long naps during the post-lunch dip period, from 2-4 pm. Students who nap at least once per week reported significantly more problems organizing their thoughts, gaining motivation, concentrating, and finishing tasks than students who did not nap. Students who napped also felt significantly more sleepy and depressed when compared to students who did not nap. The results also indicated that nap frequency increased with daytime sleepiness. The majority of students (51%) reported sleeping 6-7 hours per night or less. Overall, the results from this study suggest that among this population of Australian first-year university students habitual napping is common and may be used in an attempt to compensate for the detrimental effects of excessive sleepiness.

  5. The Napping Behaviour of Australian University Students

    PubMed Central

    Lovato, Nicole; Lack, Leon; Wright, Helen

    2014-01-01

    The purpose of this study was to evaluate the self-reported sleep and napping behaviour of Australian university students and the relationship between napping and daytime functioning. A sample of 280 university first-year psychology students (median age  = 19.00 years) completed a 6-item napping behaviour questionnaire, a 12-item Daytime Feelings and Functioning Scale, the Pittsburg Sleep Quality Index and the Epworth Sleepiness Scale. Results indicated that 53.6% of students reported napping with 34% napping at least 1–2 times per week, and 17% napping three or more occasions per week. Long naps, those over 30 minutes, were taken by 77% of the napping students. Sixty-one percent of students reported they took long naps during the post-lunch dip period, from 2–4pm. Students who nap at least once per week reported significantly more problems organizing their thoughts, gaining motivation, concentrating, and finishing tasks than students who did not nap. Students who napped also felt significantly more sleepy and depressed when compared to students who did not nap. The results also indicated that nap frequency increased with daytime sleepiness. The majority of students (51%) reported sleeping 6–7 hours per night or less. Overall, the results from this study suggest that among this population of Australian first-year university students habitual napping is common and may be used in an attempt to compensate for the detrimental effects of excessive sleepiness. PMID:25412257

  6. Factors associated with sleep duration in Brazilian high school students.

    PubMed

    Gomes Felden, Érico Pereira; Barbosa, Diego Grasel; Junior, Geraldo Jose Ferrari; Santos, Manoella De Oliveira; Pelegrini, Andreia; Silva, Diego Augusto Santos

    2017-01-01

    The aim of this study was to investigate the factors associated with short sleep duration on southern Brazilian high school students. Our study was comprised of 1,132 adolescents aged 14 to 19 years, enrolled in public high schools in São José, Brazil. The students answered a questionnaire about working (work and workload), health perception, smoking, school schedule, sleep (duration and daytime sleepiness), and socio-demographics data. The results showed that more than two thirds of adolescent workers had short sleep duration (76.7%), and those with a higher workload (more than 20 hours) had a shorter sleep duration (7.07 hours) compared to non-workers (7.83 hours). In the analysis of factors associated with short sleep duration, adolescents who worked (OR = 2.12, 95% CI 1.53 to 2.95) were more likely to have short sleep duration compared to those who did not work. In addition, older adolescents (17-19 years) and students with poor sleep quality were 40% and 55% more likely to have short sleep duration compared to younger adolescents (14-16 years) and students with good sleep quality, respectively. Adolescents with daytime sleepiness were more likely to have short sleep duration (OR = 1.49, 95% CI 1.06 to 2.07) compared to those without excessive daytime sleepiness. In addition students of the morning shift (OR = 6.02, 95% CI 4.23 to 8.57) and evening shift (OR = 2.16, 95% CI 1.45 to 3.22) were more likely to have short sleep duration compared to adolescents of the afternoon shift. Thereby adolescents who are workers, older, attended morning and evening classes and have excessive daytime sleepiness showed risk factors for short sleep duration. In this sense, it is pointed out the importance of raising awareness of these risk factors for short sleep duration of students from public schools from São José, located in southern Brazil.

  7. The impact of meal timing on performance, sleepiness, gastric upset, and hunger during simulated night shift.

    PubMed

    Grant, Crystal Leigh; Dorrian, Jillian; Coates, Alison Maree; Pajcin, Maja; Kennaway, David John; Wittert, Gary Allen; Heilbronn, Leonie Kaye; Vedova, Chris Della; Gupta, Charlotte Cecilia; Banks, Siobhan

    2017-10-07

    This study examined the impact of eating during simulated night shift on performance and subjective complaints. Subjects were randomized to eating at night (n=5; 23.2 ± 5.5 y) or not eating at night (n=5; 26.2 ± 6.4 y). All participants were given one sleep opportunity of 8 h (22:00 h-06:00 h) before transitioning to the night shift protocol. During the four days of simulated night shift participants were awake from 16:00 h-10:00 h with a daytime sleep of 6 h (10:00 h-16:00 h). In the simulated night shift protocol, meals were provided at ≈0700 h, 1900 h and 0130 h (eating at night); or ≈0700 h, 0930 h, 1410 h and 1900 h (not eating at night). Subjects completed sleepiness, hunger and gastric complaint scales, a Digit Symbol Substitution Task and a 10-min Psychomotor Vigilance Task. Increased sleepiness and performance impairment was evident in both conditions at 0400 h (p<0.05). Performance impairment at 0400 h was exacerbated when eating at night. Not eating at night was associated with elevated hunger and a small but significant elevation in stomach upset across the night (p<0.026). Eating at night was associated with elevated bloating on night one, which decreased across the protocol. Restricting food intake may limit performance impairments at night. Dietary recommendations to improve night-shift performance must also consider worker comfort.

  8. The impact of meal timing on performance, sleepiness, gastric upset, and hunger during simulated night shift

    PubMed Central

    GRANT, Crystal Leigh; DORRIAN, Jillian; COATES, Alison Maree; PAJCIN, Maja; KENNAWAY, David John; WITTERT, Gary Allen; HEILBRONN, Leonie Kaye; DELLA VEDOVA, Chris; GUPTA, Charlotte Cecilia; BANKS, Siobhan

    2017-01-01

    This study examined the impact of eating during simulated night shift on performance and subjective complaints. Subjects were randomized to eating at night (n=5; 23.2 ± 5.5 y) or not eating at night (n=5; 26.2 ± 6.4 y). All participants were given one sleep opportunity of 8 h (22:00 h-06:00 h) before transitioning to the night shift protocol. During the four days of simulated night shift participants were awake from 16:00 h-10:00 h with a daytime sleep of 6 h (10:00 h-16:00 h). In the simulated night shift protocol, meals were provided at ≈0700 h, 1900 h and 0130 h (eating at night); or ≈0700 h, 0930 h, 1410 h and 1900 h (not eating at night). Subjects completed sleepiness, hunger and gastric complaint scales, a Digit Symbol Substitution Task and a 10-min Psychomotor Vigilance Task. Increased sleepiness and performance impairment was evident in both conditions at 0400 h (p<0.05). Performance impairment at 0400 h was exacerbated when eating at night. Not eating at night was associated with elevated hunger and a small but significant elevation in stomach upset across the night (p<0.026). Eating at night was associated with elevated bloating on night one, which decreased across the protocol. Restricting food intake may limit performance impairments at night. Dietary recommendations to improve night-shift performance must also consider worker comfort. PMID:28740034

  9. Moderate exercise plus sleep education improves self-reported sleep quality, daytime mood, and vitality in adults with chronic sleep complaints: a waiting list-controlled trial.

    PubMed

    Gebhart, Carmen; Erlacher, Daniel; Schredl, Michael

    2011-01-01

    Research indicates that physical exercise can contribute to better sleep quality. This study investigates the six-week influence of a combined intervention on self-rated sleep quality, daytime mood, and quality of life. A nonclinical sample of 114 adults with chronic initiating and the maintaining of sleep complaints participated in the study. The intervention group of 70 adults underwent moderate physical exercise, conducted weekly, plus sleep education sessions. Improvements among participants assigned to the intervention group relative to the waiting-list control group (n = 44) were noted for subjective sleep quality, daytime mood, depressive symptoms and vitality. Derived from PSQI subscores, the intervention group reported increased sleep duration, shortened sleep latency, fewer awakenings after sleep onset, and overall better sleep efficiency compared to controls. The attained scores were well sustained and enhanced over a time that lasted through to the follow-up 18 weeks later. These findings have implications in treatment programs concerning healthy lifestyle approaches for adults with chronic sleep complaints.

  10. Moderate Exercise Plus Sleep Education Improves Self-Reported Sleep Quality, Daytime Mood, and Vitality in Adults with Chronic Sleep Complaints: A Waiting List-Controlled Trial

    PubMed Central

    Gebhart, Carmen; Erlacher, Daniel; Schredl, Michael

    2011-01-01

    Research indicates that physical exercise can contribute to better sleep quality. This study investigates the six-week influence of a combined intervention on self-rated sleep quality, daytime mood, and quality of life. A nonclinical sample of 114 adults with chronic initiating and the maintaining of sleep complaints participated in the study. The intervention group of 70 adults underwent moderate physical exercise, conducted weekly, plus sleep education sessions. Improvements among participants assigned to the intervention group relative to the waiting-list control group (n = 44) were noted for subjective sleep quality, daytime mood, depressive symptoms and vitality. Derived from PSQI subscores, the intervention group reported increased sleep duration, shortened sleep latency, fewer awakenings after sleep onset, and overall better sleep efficiency compared to controls. The attained scores were well sustained and enhanced over a time that lasted through to the follow-up 18 weeks later. These findings have implications in treatment programs concerning healthy lifestyle approaches for adults with chronic sleep complaints. PMID:23471095

  11. Sleep Restriction Therapy for Insomnia is Associated with Reduced Objective Total Sleep Time, Increased Daytime Somnolence, and Objectively Impaired Vigilance: Implications for the Clinical Management of Insomnia Disorder

    PubMed Central

    Kyle, Simon D.; Miller, Christopher B.; Rogers, Zoe; Siriwardena, A. Niroshan; MacMahon, Kenneth M.; Espie, Colin A.

    2014-01-01

    Study Objectives: To investigate whether sleep restriction therapy (SRT) is associated with reduced objective total sleep time (TST), increased daytime somnolence, and impaired vigilance. Design: Within-subject, noncontrolled treatment investigation. Setting: Sleep research laboratory. Participants: Sixteen patients [10 female, mean age = 47.1 (10.8) y] with well-defined psychophysiological insomnia (PI), reporting TST ≤ 6 h. Interventions: Patients were treated with single-component SRT over a 4-w protocol, sleeping in the laboratory for 2 nights prior to treatment initiation and for 3 nights (SRT night 1, 8, 22) during the acute interventional phase. The psychomotor vigilance task (PVT) was completed at seven defined time points [day 0 (baseline), day 1,7,8,21,22 (acute treatment) and day 84 (3 mo)]. The Epworth Sleepiness Scale (ESS) was completed at baseline, w 1-4, and 3 mo. Measurement and results: Subjective sleep outcomes and global insomnia severity significantly improved before and after SRT. There was, however, a robust decrease in PSG-defined TST during acute implementation of SRT, by an average of 91 min on night 1, 78 min on night 8, and 69 min on night 22, relative to baseline (P < 0.001; effect size range = 1.60-1.80). During SRT, PVT lapses were significantly increased from baseline (at three of five assessment points, all P < 0.05; effect size range = 0.69-0.78), returning to baseline levels by 3 mo (P = 0.43). A similar pattern was observed for RT, with RTs slowing during acute treatment (at four of five assessment points, all P < 0.05; effect size range = 0.57-0.89) and returning to pretreatment levels at 3 mo (P = 0.78). ESS scores were increased at w 1, 2, and 3 (relative to baseline; all P < 0.05); by 3 mo, sleepiness had returned to baseline (normative) levels (P = 0.65). Conclusion: For the first time we show that acute sleep restriction therapy is associated with reduced objective total sleep time, increased daytime sleepiness, and

  12. A Model of Fatigue Following Traumatic Brain Injury.

    PubMed

    Ponsford, Jennie; Schönberger, Michael; Rajaratnam, Shantha M W

    2015-01-01

    Fatigue is one of the most frequent sequelae of traumatic brain injury (TBI), although its causes are poorly understood. This study investigated the interrelationships between fatigue and sleepiness, vigilance performance, depression, and anxiety, using a structural equation modeling approach. Seventy-two participants with moderate to severe TBI (78% males) were recruited a median of 305 days postinjury. They completed the Fatigue Severity Scale, a vigilance task, the Epworth Sleepiness Scale, and Hospital Anxiety and Depression Scale. A model of the interrelationships between the study variables was developed, tested, and modified with path analysis. The modified model had a good overall fit (χ2 = 1.3, P = .54; comparative fit index = 1.0; root-mean square error of approximation = 0.0; standardized root-mean square residual = 0.02). Most paths in this model were significant (P < .05). Fatigue predicted anxiety, depression, and daytime sleepiness. Depression predicted daytime sleepiness and poor vigilance, whereas anxiety tended to predict reduced daytime sleepiness. This model confirms the complexity of fatigue experience. It supports the hypothesis that fatigue after TBI is a cause, not a consequence, of anxiety, depression, and daytime sleepiness, which, in turn (especially depression), may exacerbate fatigue by affecting cognitive functioning. These findings suggest that to alleviate fatigue, it is important to address each of these factors. However, the findings need to be confirmed with a longitudinal research design.

  13. A comparison of idiopathic hypersomnia and narcolepsy-cataplexy using self report measures and sleep diary data.

    PubMed Central

    Bruck, D; Parkes, J D

    1996-01-01

    Eighteen patients with idiopathic hypersomnia (IH) were compared with 50 patients with the narcoleptic syndrome of cataplexy and daytime sleepiness (NLS) using self report questionnaires and a diary of sleep/wake patterns. The IH group reported more consolidated nocturnal sleep, a lower propensity to nap, greater refreshment after naps, and a greater improvement in excessive daytime sleepiness since onset than the NLS group. In IH, the onset of excessive daytime sleepiness was predominantly associated with familial inheritance or a viral illness. Two variable--number of reported awakenings during nocturnal sleep and the reported change in sleepiness since onset--provided maximum discrimination between the IH and NLS groups. Confusional arousals, extended naps or nocturnal sleep, autonomic nervous system dysfunction, low ratings of medication effectiveness, or side effects of medication were not associated differentially with either IH or NLS. PMID:8778267

  14. Experience and limited lighting may affect sleepiness of tunnel workers.

    PubMed

    Lykouras, Dimosthenis; Karkoulias, Kiriakos; Patouchas, Dimitrios; Lakoumentas, John; Sampsonas, Fotis; Tranou, Magdalini-Konstantina; Faliagka, Evanthia; Tsakalidis, Athanasios; Spiropoulos, Kostas

    2014-07-03

    Working on shifts, especially on a night shift, influences the endogenous sleep regulation system leading to diminished sleep time and increased somnolence. We attempted to evaluate the impact of shifts on sleepiness and correlate the sleepiness score to the experience in a shift schedule. This cross-sectional study consists of 42 male and 2 female workers involved in a tunnel construction. They underwent spirometry, pulse oximetry and were asked to complete the Epworth Sleepiness Scale questionnaire. Statistical analysis revealed that workers of lower Epworth had a mean age of 43.6 years, compared to the mean age of 36.4 years of workers with higher Epworth. Furthermore, workers of lower Epworth were characterized by a mean number of shift years equal to 14.8, while those of higher Epworth possessed a mean number of shift years equal to 8. The shift schedule did not reveal any statistically significant correlation. Workers employed for a longer time had diminished sleepiness. However, there is no relationship between night shifts and sleepiness, possibly because of exposure to artificial lighting in the construction site.

  15. Commonly used stimulants: Sleep problems, dependence and psychological distress.

    PubMed

    Ogeil, Rowan P; Phillips, James G

    2015-08-01

    Caffeine and nicotine are commonly used stimulants that enhance alertness and mood. Discontinuation of both stimulants is associated with withdrawal symptoms including sleep and mood disturbances, which may differ in males and females. The present study examines changes in sleep quality, daytime sleepiness and psychological distress associated with use and dependence on caffeine and nicotine. An online survey comprising validated tools to assess sleep quality, excessive daytime sleepiness and psychological distress was completed by 166 participants (74 males, 96 females) with a mean age of 28 years. Participants completed the study in their own time, and were not offered any inducements to participate. Sleep quality was poorer in those dependent upon caffeine or nicotine, and there were also significant interaction effects with gender whereby females reported poorer sleep despite males reporting higher use of both stimulants. Caffeine dependence was associated with poorer sleep quality, increased daytime dysfunction, and increased levels of night time disturbance, while nicotine dependence was associated with poorer sleep quality and increased use of sleep medication and sleep disturbances. There were strong links between poor sleep and diminished affect, with psychological distress found to co-occur in the context of disturbed sleep. Stimulants are widely used to promote vigilance and mood; however, dependence on commonly used drugs including caffeine and nicotine is associated with decrements in sleep quality and increased psychological distress, which may be compounded in female dependent users. Copyright © 2015 Elsevier Ireland Ltd. All rights reserved.

  16. Observer Rated Sleepiness and Real Road Driving: An Explorative Study

    PubMed Central

    Anund, Anna; Fors, Carina; Hallvig, David; Åkerstedt, Torbjörn; Kecklund, Göran

    2013-01-01

    The aim of the present study was to explore if observer rated sleepiness (ORS) is a feasible method for quantification of driver sleepiness in field studies. Two measures of ORS were used: (1) one for behavioural signs based on facial expression, body gestures and body movements labelled B-ORS, and (2) one based on driving performance e.g. if swerving and other indicators of impaired driving occurs, labelled D-ORS. A limited number of observers sitting in the back of an experimental vehicle on a motorway about 2 hours repeatedly 3 times per day (before lunch, after lunch, at night) observed 24 participant’s sleepiness level with help of the two observer scales. At the same time the participant reported subjective sleepiness (KSS), EOG was recorded (for calculation of blink duration) and several driving measure were taken and synchronized with the reporting. Based on mixed model Anova and correlation analysis the result showed that observer ratings of sleepiness based on drivers’ impaired performance and behavioural signs are sensitive to extend the general pattern of time awake, circadian phase and time of driving. The detailed analysis of the subjective sleepiness and ORS showed weak correspondence on an individual level. Only 16% of the changes in KSS were predicted by the observer. The correlation between the observer ratings based on performance (D-ORS) and behavioural signs (B-ORS) are high (r = .588), and the B-ORS shows a moderately strong association (r = .360) with blink duration. Both ORS measures show an association (r>0.45) with KSS, whereas the association with driving performance is weak. The results show that the ORS-method detects the expected general variations in sleepy driving in field studies, however, sudden changes in driver sleepiness on a detailed level as 5 minutes is usually not detected; this holds true both when taking into account driving behaviour or driver behavioural signs. PMID:23724094

  17. Anti-sleepiness sensor systems for sober mental condition

    NASA Astrophysics Data System (ADS)

    Han, Won Heum; Jung, Hyung Sik; Lee, Hyo Gun

    2011-05-01

    The anti-sleepiness sensor systems have been devised for soldier's sober mental condition. These systems judge whether the soldier is sleepy or not, on one hand by monitoring open or closed eyes, on the other hand by measuring the heart blood beat and rate on the carotid of human's neck. They reasonably adopt one of the following methods such as optical, mechanical, magnetic impedance and piezoelectric sensor and so on. In this paper, the characteristics of those sensors are compared to one another and subsequently the suitable ones are proposed from the viewpoint of measurement and judgment reliability.; as a sensor to directly monitor the soldier's open/closed eyes the IR (Infrared) sensor is recommended, which is equipped on glasses (so called the anti-sleepiness glasses), and as a sensor to measure the heart beat and rate of blood vein, the piezoelectric PMN-PT crystal sensor mounted on a necklace turns out to be the most suitable owing to its high sensitivity (i.e. the anti-sleepiness necklace). These systems and relevant ideas are also applicable to the civilian usage, namely to the student preparing an examination as well as to the car-driver for safety.

  18. Risk Factors for Cardiovascular Disease, Metabolic Syndrome and Sleepiness in Truck Drivers.

    PubMed

    Mansur, Antonio de Padua; Rocha, Marcos Abs; Leyton, Vilma; Takada, Julio Yashio; Avakian, Solange Desirée; Santos, Alexandre J; Novo, Gisele C; Nascimento, Arledson Lima; Muñoz, Daniel Romero; Rohlfs, Waldo J C

    2015-12-01

    Truck driver sleepiness is a primary cause of vehicle accidents. Several causes are associated with sleepiness in truck drivers. Obesity and metabolic syndrome (MetS) are associated with sleep disorders and with primary risk factors for cardiovascular diseases (CVD). We analyzed the relationship between these conditions and prevalence of sleepiness in truck drivers. We analyzed the major risk factors for CVD, anthropometric data and sleep disorders in 2228 male truck drivers from 148 road stops made by the Federal Highway Police from 2006 to 2011. Alcohol consumption, illicit drugs and overtime working hours were also analyzed. Sleepiness was assessed using the Epworth Sleepiness Scale. Mean age was 43.1 ± 10.8 years. From 2006 to 2011, an increase in neck (p = 0.011) and abdominal circumference (p < 0.001), total cholesterol (p < 0.001), triglyceride plasma levels (p = 0.014), and sleepiness was observed (p < 0.001). In addition, a reduction in hypertension (39.6% to 25.9%, p < 0.001), alcohol consumption (32% to 23%, p = 0.033) and overtime hours (52.2% to 42.8%, p < 0.001) was found. Linear regression analysis showed that sleepiness correlated closely with body mass index (β = 0.19, Raj2 = 0.659, p = 0.031), abdominal circumference (β = 0.24, Raj2 = 0.826, p = 0.021), hypertension (β = -0.62, Raj2 = 0.901, p = 0.002), and triglycerides (β = 0.34, Raj2 = 0.936, p = 0.022). Linear multiple regression indicated that hypertension (p = 0.008) and abdominal circumference (p = 0.025) are independent variables for sleepiness. Increased prevalence of sleepiness was associated with major components of the MetS.

  19. Validation of the Karolinska sleepiness scale against performance and EEG variables.

    PubMed

    Kaida, Kosuke; Takahashi, Masaya; Akerstedt, Torbjörn; Nakata, Akinori; Otsuka, Yasumasa; Haratani, Takashi; Fukasawa, Kenji

    2006-07-01

    The Karolinska sleepiness scale (KSS) is frequently used for evaluating subjective sleepiness. The main aim of the present study was to investigate the validity and reliability of the KSS with electroencephalographic, behavioral and other subjective indicators of sleepiness. Participants were 16 healthy females aged 33-43 (38.1+/-2.68) years. The experiment involved 8 measurement sessions per day for 3 consecutive days. Each session contained the psychomotor vigilance task (PVT), the Karolinska drowsiness test (KDT-EEG alpha & theta power), the alpha attenuation test (AAT-alpha power ratio open/closed eyes) and the KSS. Median reaction time, number of lapses, alpha and theta power density and the alpha attenuation coefficients (AAC) showed highly significant increase with increasing KSS. The same variables were also significantly correlated with KSS, with a mean value for lapses (r=0.56). The KSS was closely related to EEG and behavioral variables, indicating a high validity in measuring sleepiness. KSS ratings may be a useful proxy for EEG or behavioral indicators of sleepiness.

  20. Experience and limited lighting may affect sleepiness of tunnel workers

    PubMed Central

    2014-01-01

    Background Working on shifts, especially on a night shift, influences the endogenous sleep regulation system leading to diminished sleep time and increased somnolence. We attempted to evaluate the impact of shifts on sleepiness and correlate the sleepiness score to the experience in a shift schedule. Materials and methods This cross-sectional study consists of 42 male and 2 female workers involved in a tunnel construction. They underwent spirometry, pulse oximetry and were asked to complete the Epworth Sleepiness Scale questionnaire. Results Statistical analysis revealed that workers of lower Epworth had a mean age of 43.6 years, compared to the mean age of 36.4 years of workers with higher Epworth. Furthermore, workers of lower Epworth were characterized by a mean number of shift years equal to 14.8, while those of higher Epworth possessed a mean number of shift years equal to 8. The shift schedule did not reveal any statistically significant correlation. Conclusions Workers employed for a longer time had diminished sleepiness. However, there is no relationship between night shifts and sleepiness, possibly because of exposure to artificial lighting in the construction site. PMID:24993796

  1. Sleep disturbance, stroke, and heart disease events: evidence from the Caerphilly cohort

    PubMed Central

    Elwood, Peter; Hack, Melissa; Pickering, Janet; Hughes, Janie; Gallacher, John

    2006-01-01

    Objective To test the hypothesis that sleep disorders are relevant to the risk of ischaemic stroke and ischaemic heart disease events in older men. Design A cohort study. Setting The Caerphilly cohort, a representative population sample of older men in South Wales, UK. Participants 1986 men aged 55–69 years completed a questionnaire on sleep patterns with help from their partners. This asked about symptoms of disturbed sleep: insomnia, snoring, restless legs, obstructive sleep apnoea, and about daytime sleepiness. During the following 10 years 107 men experienced an ischaemic stroke and 213 had an ischaemic heart disease event. Main results Up to one third of the men reported at least one symptom suggestive of sleep disturbance, and one third reported daytime sleepiness. Compared with men who reported no such symptoms, the adjusted relative odds of an ischaemic stroke were significantly increased in men with any sleep disturbance, the strongest association being with sleep apnoea (relative odds 1.97; 1.26 to 3.09). The association with daytime sleepiness was not significant for stroke. Relations with ischaemic heart disease events were all raised in men with symptoms of sleep disturbance, but none was significant, other than daytime sleepiness (relative odds: 1.41; 1.04 to 1.92). There were no significant relations with blood pressure. Conclusion The risk of an ischaemic stroke is increased in men whose sleep is frequently disturbed, and daytime sleepiness is associated with a significant increase in ischaemic heart disease events. PMID:16361457

  2. Empiric auto-titrating CPAP in people with suspected obstructive sleep apnea.

    PubMed

    Drummond, Fitzgerald; Doelken, Peter; Ahmed, Qanta A; Gilbert, Gregory E; Strange, Charlie; Herpel, Laura; Frye, Michael D

    2010-04-15

    Efficient diagnosis and treatment of obstructive sleep apnea (OSA) can be difficult because of time delays imposed by clinic visits and serial overnight polysomnography. In some cases, it may be desirable to initiate treatment for suspected OSA prior to polysomnography. Our objective was to compare the improvement of daytime sleepiness and sleep-related quality of life of patients with high clinical likelihood of having OSA who were randomly assigned to receive empiric auto-titrating continuous positive airway pressure (CPAP) while awaiting polysomnogram versus current usual care. Serial patients referred for overnight polysomnography who had high clinical likelihood of having OSA were randomly assigned to usual care or immediate initiation of auto-titrating CPAP. Epworth Sleepiness Scale (ESS) scores and the Functional Outcomes of Sleep Questionnaire (FOSQ) scores were obtained at baseline, 1 month after randomization, and again after initiation of fixed CPAP in control subjects and after the sleep study in auto-CPAP patients. One hundred nine patients were randomized. Baseline demographics, daytime sleepiness, and sleep-related quality of life scores were similar between groups. One-month ESS and FOSQ scores were improved in the group empirically treated with auto-titrating CPAP. ESS scores improved in the first month by a mean of -3.2 (confidence interval -1.6 to -4.8, p < 0.001) and FOSQ scores improved by a mean of 1.5, (confidence interval 0.5 to 2.7, p = 0.02), whereas scores in the usual-care group did not change (p = NS). Following therapy directed by overnight polysomnography in the control group, there were no differences in ESS or FOSQ between the groups. No adverse events were observed. Empiric auto-CPAP resulted in symptomatic improvement of daytime sleepiness and sleep-related quality of life in a cohort of patients awaiting polysomnography who had a high pretest probability of having OSA. Additional studies are needed to evaluate the applicability

  3. Hypnotic Relaxation and Yoga to Improve Sleep and School Functioning

    ERIC Educational Resources Information Center

    Perfect, Michelle M.; Smith, Bradley

    2016-01-01

    Sleep insufficiency, defined as inadequate sleep duration, poor sleep quality, and daytime sleepiness, has been linked with students' learning and behavioral outcomes at school. However, there is limited research on interventions designed to improve the sleep of school-age children. In order to promote more interest on this critical topic, we…

  4. [The Effects of Mobile Social Networking Service-Based Cognitive Behavior Therapy on Insomnia in Nurses].

    PubMed

    Kim, Ji Eun; Kim, Suk Sun

    2017-08-01

    This study aimed to examine the effects of cognitive behavior therapy for insomnia (CBT-I) based on the mobile social networking service (SNS) on dysfunctional beliefs and attitudes about sleep, sleep quality, daytime sleepiness, depression, and quality of life among rotating-shift nurses in a hospital in Korea. A nonequivalent control group pre-post test design was used. The participants included 55 nurses with rotating three-shift work (25 in the experimental group and 30 in the control group). For the experimental group, CBT-I using mobile SNS was provided once a week for 60 minutes over six weeks. Data were analyzed using descriptive statistics, χ²-test, independent samples t-test, and Mann-whitney U test with the SPSS 21.0 program. In the homogeneity test of the general characteristics and study variables, there were no significant differences between the two groups. Nurses in the experimental group had significantly lower scores on dysfunctional beliefs and attitudes regarding sleep and sleepiness than nurses in the control group. Nurses in the experimental group had significantly higher scores on sleep quality and quality of life than nurses in the control group. These findings indicate that using the mobile SNS-based CBT-I is feasible and has significant and positive treatment-related effects on rotating-shift nurses' irrational thoughts and beliefs in association with sleep, sleep quality, daytime sleepiness, and quality of life. These contribute to expanding our knowledge of rotating-shift nurses' sleep issues and their preferences for intervention. © 2017 Korean Society of Nursing Science

  5. Stop and revive? The effectiveness of nap and active rest breaks for reducing driver sleepiness.

    PubMed

    Watling, Christopher N; Smith, Simon S; Horswill, Mark S

    2014-11-01

    The purpose of this study was to compare the effects of two commonly utilized sleepiness countermeasures: a nap break and an active rest break. The effects of the countermeasures were evaluated by physiological (EEG), subjective, and driving performance measures. Participants completed 2 h of simulated driving, followed by a 15-min nap break or a 15-min active rest break, then completed the final hour of simulated driving. The nap break reduced EEG and subjective sleepiness. The active rest break did not reduce EEG sleepiness, with sleepiness levels eventually increasing, and resulted in an immediate reduction of subjective sleepiness. No difference was found between the two breaks for the driving performance measure. The immediate reduction of subjective sleepiness after the active rest break could leave drivers with erroneous perceptions of their sleepiness, particularly with increases of physiological sleepiness after the break. Copyright © 2014 Society for Psychophysiological Research.

  6. Traffic crash accidents in Tehran, Iran: Its relation with circadian rhythm of sleepiness.

    PubMed

    Sadeghniiat-Haghighi, Khosro; Yazdi, Zohreh; Moradinia, Mohsen; Aminian, Omid; Esmaili, Alireza

    2015-01-01

    Road traffic accidents are one of main problems in Iran. Multiple factors cause traffic accidents and the most important one is sleepiness. This factor, however, is given less attention in our country. Road traffic accidents relevant to sleepiness are studied. In this cross-sectional study, all road traffic accidents relevant to sleepiness, which were reported by police, were studied in Tehran province in 2009. The risk of road traffic accidents due to sleepiness was increased by more than sevenfold (odds ratio = 7.33) in low alertness hours (0:00-6:00) compared to other time of day. The risk of road traffic accidents due to sleepiness was decreased by 0.15-fold (odds ratio = 0.15) in hours with maximum of alertness (18:00-22:00) of circadian rhythm compared to other time of day. The occurrence of road traffic accidents due to sleepiness has significant statistical relations with driving during lowest point of alertness of circadian rhythm.

  7. A pictorial Sleepiness and Sleep Apnoea Scale to recognize individuals with high risk for obstructive sleep apnea syndrome.

    PubMed

    Edelmann, Cathrin; Ghiassi, Ramesh; Vogt, Deborah R; Partridge, Martyn R; Khatami, Ramin; Leuppi, Jörg D; Miedinger, David

    2017-01-01

    The aim of this study was to evaluate the validity of a new pictorial form of a screening test for obstructive sleep apnea syndrome (OSAS) - the pictorial Sleepiness and Sleep Apnoea Scale (pSSAS). Validation was performed in a sample of patients admitted to sleep clinics in the UK and Switzerland. All study participants were investigated with objective sleep tests such as full-night-attended polysomnography or polygraphy. The pSSAS was validated by taking into account the individual result of the sleep study, sleep-related questionnaires and objective parameters such as body mass index (BMI) or neck circumference. Different scoring schemes of the pSSAS were evaluated, and an internal validation was undertaken. The full data set consisted of 431 individuals (234 patients from the UK, 197 patients from Switzerland). The pSSAS showed good predictive performance for OSAS with an area under the curve between 0.77 and 0.81 depending on which scoring scheme was used. The subscores of the pSSAS had a moderate-to-strong correlation with widely used screening questionnaires for OSAS or excessive daytime sleepiness as well as with BMI and neck circumference. The pSSAS can be used to select patients with a high probability of having OSAS. Due to its simple pictorial design with short questions, it might be suitable for screening in populations with low health literacy and in non-native English or German speakers.

  8. Effects of various factors on sleep disorders and quality of life in Parkinson's disease.

    PubMed

    Telarovic, Srdjana; Mijatovic, Dragana; Telarovic, Irma

    2015-12-01

    In Parkinson's disease (PD), sleep disorders (SD) occur as a result of the neurochemical changes in sleep centres, neurodegenerative changes in dopaminergic neurons, and other factors. The most common SD include excessive daytime sleepiness, insomnia, restless legs syndrome and nocturia. The aim of the study was to compare quality of sleep, as a factor that greatly impacts quality of life (QoL), between PD patients and a control group and to further examine SD in the PD group with focus on incidence and SD types as well as on effects various factors (age, sex, PD characteristics, medication usage) have on these disorders. The study included 110 patients who met the criteria for the diagnosis of PD and 110 age-matched healthy controls. We used the Pittsburgh Sleep Quality Index, PD Sleep Scale, Epworth Sleepiness Scale, PD QoL Questionnaire-8 and PD Questionnaire-39 (items 30 and 33). In the group with PD, we considered the duration of the disease, the stage of disease according to the Hoehn and Yahr scale, medications and their impact on the SD. The average duration of the disease was 6 years and the mean stage was 2.44. The result showed significant differences in the sleep quality between groups. In the PD group, SD differences were also found according to gender, duration of the disease and medication usage. The most common SD were fragmented sleep, insomnia and nocturia. To improve the QoL of PD patients, it is necessary to pay more attention to detecting and solving SD.

  9. 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.

  10. Subjective sleepiness is a sensitive indicator of insufficient sleep and impaired waking function.

    PubMed

    Akerstedt, Torbjörn; Anund, Anna; Axelsson, John; Kecklund, Göran

    2014-06-01

    The main consequence of insufficient sleep is sleepiness. While measures of sleep latency, continuous encephalographical/electro-oculographical (EEG/EOG) recording and performance tests are useful indicators of sleepiness in the laboratory and clinic, they are not easily implemented in large, real-life field studies. Subjective ratings of sleepiness, which are easily applied and unobtrusive, are an alternative, but whether they measure sleepiness sensitively, reliably and validly remains uncertain. This review brings together research relevant to these issues. It is focused on the Karolinska Sleepiness Scale (KSS), which is a nine-point Likert-type scale. The diurnal pattern of sleepiness is U-shaped, with high KSS values in the morning and late evening, and with great stability across years. KSS values increase sensitively during acute total and repeated partial sleep deprivation and night work, including night driving. The effect sizes range between 1.5 and 3. The relation to driving performance or EEG/EOG indicators of sleepiness is highly significant, strongly curvilinear and consistent across individuals. High (>6) KSS values are associated particularly with impaired driving performance and sleep intrusions in the EEG. KSS values are also increased in many clinical conditions such as sleep apnea, depression and burnout. The context has a strong influence on KSS ratings. Thus, physical activity, social interaction and light exposure will reduce KSS values by 1-2 units. In contrast, time-on-task in a monotonous context will increase KSS values by 1-2 units. In summary, subjective ratings of sleepiness as described here is as sensitive and valid an indicator of sleepiness as objective measures, and particularly suitable for field studies. © 2014 European Sleep Research Society.

  11. Characteristics of Sleep Disturbances in Patients with Gastroesophageal Reflux Disease.

    PubMed

    Iwakura, Narika; Fujiwara, Yasuhiro; Shiba, Masatsugu; Ochi, Masahiro; Fukuda, Takashi; Tanigawa, Tetsuya; Yamagami, Hirokazu; Tominaga, Kazunari; Watanabe, Toshio; Arakawa, Tetsuo

    2016-01-01

    Objective Gastroesophageal reflux disease (GERD) is strongly associated with sleep disturbances; however, the detailed differences in the characteristics of sleep disturbances between GERD and non-GERD patients are unknown. The aim of the present study was to analyze the clinical characteristics as well as health-related quality of life in GERD and non-GERD patients with sleep disturbances. Methods Three hundred and fifty patients, including 124 patients with GERD and 226 patients without GERD, completed a self-administered questionnaire that evaluated clinical information. The Pittsburg Sleep Quality Index (PSQI), Hospital Anxiety and Depression Scale (HADS), Insomnia Severity Index (ISI), Epworth Sleepiness Scale (ESS) and 8-item Short-Form Health Survey (SF-8) were also used. Sleep disturbance was considered to be present if the PSQI was >5.5. Results The prevalence of sleep disturbances was significantly higher in the GERD patients (66/124, 53.9%) than in the non-GERD patients (89/226, 39.3%). Depression and anxiety were significantly more common in the subjects with sleep disturbances than in those without sleep disturbances, although there were no differences between the GERD and non-GERD patients. Among the subjects with sleep disturbances, daytime sleepiness was more common in the GERD patients than in the non-GERD patients. The subjects with sleep disturbances had a poorer health-related quality of life. The physical components of quality of life were impaired, particularly in the GERD patients with sleep disturbances. Conclusion GERD patients with sleep disturbances commonly experience daytime sleepiness and an impaired health-related quality of life, especially in terms of physical components.

  12. The effect of bright light on sleepiness among rapid-rotating 12-hour shift workers.

    PubMed

    Sadeghniiat-Haghighi, Khosro; Yazdi, Zohreh; Jahanihashemi, Hassan; Aminian, Omid

    2011-01-01

    About 20% of workers in industrialized countries are shift workers and more than half of them work on night or rotating shifts. Most night workers complain of sleepiness due to lack of adjustment of the circadian rhythm. In simulated night-work experiments, scheduled exposure to bright light has been shown to reduce these complaints. Our study assessed the effects of bright light exposure on sleepiness during night work in an industrial setting. In a cross-over design, 94 workers at a ceramic factory were exposed to either bright (2500 lux) or normal light (300 lux) during breaks on night shifts. We initiated 20-minute breaks between 24.00 and 02.00 hours. Sleepiness ratings were determined using the Stanford Sleepiness Scale at 22.00, 24.00, 02.00 and 04.00 hours. Under normal light conditions, sleepiness peaked at 02:00 hours. A significant reduction (22% compared to normal light conditions) in sleepiness was observed after workers were exposed to bright light. Exposure to bright light may be effective in reducing sleepiness among night workers.

  13. Youth internalizing symptoms, sleep-related problems, and disordered eating attitudes and behaviors: A moderated mediation analysis.

    PubMed

    Chardon, Marie L; Janicke, David M; Carmody, Julia K; Dumont-Driscoll, Marilyn C

    2016-04-01

    Internalizing symptoms increase the risk for disordered eating; however, the mechanism through which this relationship occurs remains unclear. Sleep-related problems may be a potential link as they are associated with both emotional functioning and disordered eating. The present study aims to evaluate the mediating roles of two sleep-related problems (sleep disturbance and daytime sleepiness) in the relationship between youth internalizing symptoms and disordered eating, and to explore if age moderates these relations. Participants were 225 youth (8-17years) attending a primary care appointment. Youth and legal guardians completed questionnaires about youth disordered eating attitudes and behaviors, internalizing symptoms, sleep disturbance, and daytime sleepiness. Mediation and moderated mediation analyses were utilized. The mediation model revealed both youth sleep disturbance and daytime sleepiness independently mediated the association between internalizing symptoms and disordered eating attitudes and behaviors, and explained 18% of the variance in disordered eating. The moderated mediation model including youth age accounted for 21% of the variance in disordered eating; youth age significantly interacted with sleep disturbance, but not with daytime sleepiness, to predict disordered eating. Sleep disturbance only mediated the relationship between internalizing symptoms and disordered eating in youth 12years old and younger, while daytime sleepiness was a significant mediator regardless of age. As sleep-related problems are frequently improved with the adoption of health behaviors conducive to good sleep, these results may suggest a relatively modifiable and cost-effective target to reduce youth risk for disordered eating. Copyright © 2016 Elsevier Ltd. All rights reserved.

  14. 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

  15. Effect of Adenotonsillectomy on Parent-Reported Sleepiness in Children with Obstructive Sleep Apnea.

    PubMed

    Paruthi, Shalini; Buchanan, Paula; Weng, Jia; Chervin, Ronald D; Mitchell, Ronald B; Dore-Stites, Dawn; Sadhwani, Anjali; Katz, Eliot S; Bent, John; Rosen, Carol L; Redline, Susan; Marcus, Carole L

    2016-11-01

    To describe parental reports of sleepiness and sleep duration in children with polysomnography (PSG)-confirmed obstructive sleep apnea (OSA) randomized to early adenotonsillectomy (eAT) or watchful waiting with supportive care (WWSC) in the ChildHood Adenotonsillectomy Trial (CHAT). We hypothesized children with OSA would have a larger improvement in sleepiness 6 mo following eAT compared to WWSC. Parents of children aged 5.0-9.9 y completed the Epworth Sleepiness Scale modified for children (mESS) and the Pediatric Sleep Questionnaire-Sleepiness Subscale (PSQ-SS). PSG was performed at baseline and at 7-mo endpoint. Children underwent early adenotonsillectomy or WWSC. The mESS and PSQ-SS classified 24% and 53% of the sample as excessively sleepy, respectively. At baseline, mean mESS score was 7.4 ± 5.0 (SD) and mean PSQ-SS score was 0.44 ± 0.30. Sleepiness scores were higher in African American children; children with shorter sleep duration; older children; and overweight children. At endpoint, mean mESS score decreased by 2.0 ± 4.2 in the eAT group versus 0.3 ± 4.0 in the WWSC group (P < 0.0001); mean PSQ-SS score decreased 0.29 ± 0.40 in eAT versus 0.08 ± 0.40 in the WWSC group (P < 0.0001). Despite higher baseline sleepiness, African American children experienced similar improvement with adenotonsillectomy than other children. Improvement in sleepiness was weakly associated with improved apnea-hypopnea index or oxygen desaturation indices, but not with change in other polysomnographic measures. Sleepiness assessed by parent report was prevalent; improved more after eAT than after WWSC; and was not strongly predicted by sleep disturbances identified by PSG. Childhood Adenotonsillectomy Study for Children with OSA (CHAT). ClinicalTrials.gov Identifier #NCT00560859. © 2016 Associated Professional Sleep Societies, LLC.

  16. Risk factors for precompetitive sleep behavior in elite female athletes.

    PubMed

    Silva, Maria-Raquel; Paiva, Teresa

    2018-06-07

    Sleep is of major importance for the athletes' short and long-term health, performance and recovery; however, published studies on athletes' sleep and even fewer before competition are scarce. This study evaluated the risk factors of sleep in young female gymnasts before an international competition. Sixty-seven rhythmic gymnasts (M=18.7,SD=2.9yrs.) of high performance (M=36.6,SD=7.6h/week) were evaluated regarding training and sport practice, body composition, sleep duration, daytime sleepiness by the Epworth Sleepiness Scale, sleep quality by the Pittsburgh Sleep Quality Index, precompetitive anxiety by the Sport Competition Anxiety Test form A, and detailed dietary intake just prior to a world competition. The majority of the participants (83.6%) presented reduced body fat levels (M=9.1,SD=2.1%) and low energy availability (M=31.5,SD=11.9kcal/kgFFM/day). They slept 8h10±1h30/night on weekdays and 8h40±00h40/night on weekends, 67% suffered from mild daytime sleepiness and 78% had a reduced sleep quality. Precompetitive anxiety was on average moderate (M=22.7,SD=3.2). The risk factors for short sleep duration were: 1.92 for a training volume>30hours/week (95%CI 0.84-4.39), 4.57 for menstruation (95%CI 1.17- 17.77), 6.62 for bedtime≥11:00PM (95%CI 1.74-25.10), 1.40 for BF<12%(95%CI 1.03- 1.92), 2.19 for FFM<30kg (95%IC 0.94-4.94), 2.37 for BMR<1100kcal/day (95%CI 1.06- 5.32), 1.90 for EEE≥700kcal (95%CI 0.92-3.93) and 3.17 for EA<45kcal/kgFFM/day (95%CI 0.84-6.59). Age, performance, sleep duration on weekdays and precompetitive stress were also predictors for a reduced sleep quality and/or abnormal daytime sleepiness. Age, training regime, menstruation, individual preferences for bedtime, body composition and energy were important predictors of gymnasts' precompetitive sleep with consequences upon their sleep duration, SQ and DS.

  17. The effectiveness of istradefylline for the treatment of gait deficits and sleepiness in patients with Parkinson's disease.

    PubMed

    Matsuura, Keita; Kajikawa, Hiroyuki; Tabei, Ken-Ichi; Satoh, Masayuki; Kida, Hirotaka; Nakamura, Naoko; Tomimoto, Hidekazu

    2018-01-01

    Istradefylline is useful in treating the wearing-off state in Parkinson's disease (PD). We investigated the effectiveness of istradefylline (ISD) in improving arousal, sleep, and gait deficits in patients with PD. We examined 14 patients with PD treated with ISD. We assessed the patients using the Unified Parkinson's Disease Rating Scale, Parkinson's Disease Questionnaire, Timed Up-and-Go test (TUG), Freezing of Gait Questionnaire (FOG-Q), Epworth Sleepiness Scale (ESS), and Parkinson's Disease Sleep Scale (PDSS) before and 1 month after ISD use. ESS scores were significantly lower 1 month after the start of ISD treatment (6.79±6.50) than before the intervention (8.14±6.15, Wilcoxon signed-rank test, p=0.0033). PDSS scores were not significantly different 1 month after beginning the treatment (112±23mm) when compared to those before the intervention (110±27mm, Wilcoxon signed-rank test, p=0.40). TUG scores were not changed after 1 month of ISD use (14.9±8.3s) when compared to those before the intervention (21.3±30.0s, Wilcoxon signed-rank test, p=0.59). Although these measures were not significantly affected by ISD treatment, some patients remarkably improved after the treatment. FOG-Q scores were significantly lower 1 month after the beginning of treatment (9.79±7.16) than those before the intervention (12.14±5.82, Wilcoxon signed-rank test, p=0.030). ISD may improve daytime sleepiness and FOG in patients with PD. Copyright © 2017 Elsevier B.V. All rights reserved.

  18. Driver sleepiness and risk of motor vehicle crash injuries: a population-based case control study in Fiji (TRIP 12).

    PubMed

    Herman, Josephine; Kafoa, Berlin; Wainiqolo, Iris; Robinson, Elizabeth; McCaig, Eddie; Connor, Jennie; Jackson, Rod; Ameratunga, Shanthi

    2014-03-01

    Published studies investigating the role of driver sleepiness in road crashes in low and middle-income countries have largely focused on heavy vehicles. We investigated the contribution of driver sleepiness to four-wheel motor vehicle crashes in Fiji, a middle-income Pacific Island country. The population-based case control study included 131 motor vehicles involved in crashes where at least one person died or was hospitalised (cases) and 752 motor vehicles identified in roadside surveys (controls). An interviewer-administered questionnaire completed by drivers or proxies collected information on potential risks for crashes including sleepiness while driving, and factors that may influence the quantity or quality of sleep. Following adjustment for confounders, there was an almost six-fold increase in the odds of injury-involved crashes for vehicles driven by people who were not fully alert or sleepy (OR 5.7, 95%CI: 2.7, 12.3), or those who reported less than 6 h of sleep during the previous 24 h (OR 5.9, 95%CI: 1.7, 20.9). The population attributable risk for crashes associated with driving while not fully alert or sleepy was 34%, and driving after less than 6 h sleep in the previous 24 h was 9%. Driving by people reporting symptoms suggestive of obstructive sleep apnoea was not significantly associated with crash risk. Driver sleepiness is an important contributor to injury-involved four-wheel motor vehicle crashes in Fiji, highlighting the need for evidence-based strategies to address this poorly characterised risk factor for car crashes in less resourced settings. Copyright © 2013 The Authors. Published by Elsevier Ltd.. All rights reserved.

  19. Sleep onset insomnia, daytime sleepiness and sleep duration in relationship to Toxoplasma gondii IgG seropositivity and serointensity

    PubMed Central

    Ahmad, Zaki; Moustafa, Yara W.; Stiller, John W.; Pavlovich, Mary A.; Raheja, Uttam K.; Gragnoli, Claudia; Snitker, Soren; Nazem, Sarra; Dagdag, Aline; Fang, Beverly; Fuchs, Dietmar; Lowry, Christopher A.; Postolache, Teodor T.

    2018-01-01

    Toxoplasma gondii (T. gondii) infects central nervous tissue and is kept in relative dormancy by a healthy immune system. Sleep disturbances have been found to precipitate mental illness, suicidal behavior and car accidents, which have been previously linked to T. gondii as well. We speculated that if sleep disruption, particularly insomnia, would mediate, at least partly, the link between T. gondii infection and related behavioral dysregulation, then we would be able to identify significant associations between sleep disruption and T. gondii. The mechanisms for such an association may involve dopamine (DA) production by T. gondii, or collateral effects of immune activation necessary to keep T. gondii in check. Sleep questionnaires from 2031 Old Order Amish were analyzed in relationship to T. gondii-IgG antibodies measured by enzyme-linked immunosorbent assay (ELISA). Toxoplasma gondii seropositivity and serointensity were not associated with any of the sleep latency variables or Epworth Sleepiness Scale (ESS). A secondary analysis identified, after adjustment for age group, a statistical trend toward shorter sleep duration in seropositive men (p = 0.07). In conclusion, it is unlikely that sleep disruption mediates links between T. gondii and mental illness or behavioral dysregulation. Trending gender differences in associations between T. gondii and shorter sleep need further investigation. PMID:29657364

  20. Evaluation of restless legs syndrome in fibromyalgia syndrome: an analysis of quality of sleep and life.

    PubMed

    Civelek, Gul Mete; Ciftkaya, Pinar Oztop; Karatas, Metin

    2014-01-01

    The aim of this study is to find prevalence and severity of restless legs syndrome (RLS) in patients with fibromyalgia syndrome (FMS) and detect effect of FMS and RLS coexistance on quality of sleep and life. In this study, presence and severity of RLS were detected in patients with FMS and Pitsburgh Sleep Quality Index (PSQI), Epworth Sleepiness Scale (ESS) and Fibromyalgia Impact Questionnaire (FIQ) scores of all patients were measured. One hundred and fifteen female patients with median age 49 (39.0-57.0)[median (25-75{\\%} interquartile range)] were included in the study. In 42.6% of patients RLS coexisting with FMS was found. RLS was classified as moderate in 42.9% of patients and as severe in 49.0% of patients. In patients with FMS ans RLS sleep quality, daytime sleepiness and quality of life were more severely impaired (PSQI scores were 9.0 ± 4.4 vs 7.8 ± 4.3, p=0.003; ESS scores were 5.0(3.0-7.5) vs 3.0(1.0-4.3), p=0.036 and FIQ scores were 68.1 ± 9.8 vs 59.4 ± 16.9, p=0.027) compared to patients with only FMS. Prevalence of RLS was found higher in FMS than normal population and quality of sleep and quality of life were worse in patients with RLS. Presence of RLS should be investigated in every patient with FMS and treatment plans should also cover RLS in case of coexistance with FMS. Prospective cohort studies are needed for better explanation of FMS and RLS coexistance.

  1. Sleep quality during pregnancy: associations with depressive and anxiety symptoms.

    PubMed

    Polo-Kantola, Päivi; Aukia, Linda; Karlsson, Hasse; Karlsson, Linnea; Paavonen, E Juulia

    2017-02-01

    Sleep disturbances are common during pregnancy, yet underdiagnosed and under-investigated. We evaluated sleep quality during pregnancy and assessed associated factors, especially depressive and anxiety symptoms. A total of 78 healthy pregnant women from the FinnBrain Birth Cohort Study were studied twice prospectively during pregnancy (in mid-pregnancy and late pregnancy). Sleep quality was evaluated by the Basic Nordic Sleep Questionnaire, depressive symptoms by the Edinburgh Postnatal Depression Scale, and anxiety symptoms by the State-Trait Anxiety Inventory. Poor general sleep quality, difficulty falling asleep, the number of nocturnal awakenings per night, and too-early morning awakenings increased in late pregnancy compared with mid-pregnancy (all p-values < 0.020). The total insomnia score (p < 0.001) and sleep latency increased (p = 0.005), but sleep duration and preferred sleep duration did not change. Women tended to snore more often in late pregnancy, but apneas remained rare. Almost one-fourth of the women reported both morning and daytime sleepiness, but the frequencies did not increase during the follow up. In late pregnancy, depressive and anxiety symptoms were cross-sectionally related to sleep disturbances, but depressive or anxiety symptoms in mid-pregnancy were not associated with sleep disturbances in late pregnancy. We found deterioration in sleep quality across pregnancy. However, no increase in negative daytime consequences was found, presumably indicating a compensatory capacity against sleep impairment. Additionally, depressive and anxiety symptoms and sleep disturbances were only cross-sectionally associated. Our study calls for further research on the factors that influence sleep disturbances during pregnancy. © 2016 Nordic Federation of Societies of Obstetrics and Gynecology.

  2. The Effect of Low-Frequency Road Noise on Driver Sleepiness and Performance

    PubMed Central

    Anund, Anna; Lahti, Eva; Fors, Carina; Genell, Anders

    2015-01-01

    It is a well-known fact today that driver sleepiness is a contributory factor in crashes. Factors considered as sleepiness contributor are mostly related to time of the day, hours being awake and hours slept. Factors contributing to active and passive fatigue are mostly focusing on the level of cognitive load. Less is known what role external factors, e.g. type of road, sound/noise, vibrations etc., have on the ability to stay awake both under conditions of sleepiness and under active or passive fatigue. The aim of this moving base driving simulator study with 19 drivers participating in a random order day and night time, was to evaluate the effect of low-frequency road noise on driver sleepiness and performance, including both long-term and short-term effects. The results support to some extent the hypothesis that road-induced interior vehicle sound affects driving performance and driver sleepiness. Increased low-frequency noise helps to reduce speed during both day- and night time driving, but also contributes to increase the number of lane crossings during night time. PMID:25874883

  3. The effect of low-frequency road noise on driver sleepiness and performance.

    PubMed

    Anund, Anna; Lahti, Eva; Fors, Carina; Genell, Anders

    2015-01-01

    It is a well-known fact today that driver sleepiness is a contributory factor in crashes. Factors considered as sleepiness contributor are mostly related to time of the day, hours being awake and hours slept. Factors contributing to active and passive fatigue are mostly focusing on the level of cognitive load. Less is known what role external factors, e.g. type of road, sound/noise, vibrations etc., have on the ability to stay awake both under conditions of sleepiness and under active or passive fatigue. The aim of this moving base driving simulator study with 19 drivers participating in a random order day and night time, was to evaluate the effect of low-frequency road noise on driver sleepiness and performance, including both long-term and short-term effects. The results support to some extent the hypothesis that road-induced interior vehicle sound affects driving performance and driver sleepiness. Increased low-frequency noise helps to reduce speed during both day- and night time driving, but also contributes to increase the number of lane crossings during night time.

  4. Effect of Adenotonsillectomy on Parent-Reported Sleepiness in Children with Obstructive Sleep Apnea

    PubMed Central

    Paruthi, Shalini; Buchanan, Paula; Weng, Jia; Chervin, Ronald D.; Mitchell, Ronald B.; Dore-Stites, Dawn; Sadhwani, Anjali; Katz, Eliot S.; Bent, John; Rosen, Carol L.; Redline, Susan; Marcus, Carole L.

    2016-01-01

    Study Objectives: To describe parental reports of sleepiness and sleep duration in children with polysomnography (PSG)-confirmed obstructive sleep apnea (OSA) randomized to early adenotonsillectomy (eAT) or watchful waiting with supportive care (WWSC) in the ChildHood Adenotonsillectomy Trial (CHAT). We hypothesized children with OSA would have a larger improvement in sleepiness 6 mo following eAT compared to WWSC. Methods: Parents of children aged 5.0–9.9 y completed the Epworth Sleepiness Scale modified for children (mESS) and the Pediatric Sleep Questionnaire-Sleepiness Subscale (PSQ-SS). PSG was performed at baseline and at 7-mo endpoint. Children underwent early adenotonsillectomy or WWSC. Results: The mESS and PSQ-SS classified 24% and 53% of the sample as excessively sleepy, respectively. At baseline, mean mESS score was 7.4 ± 5.0 (SD) and mean PSQ-SS score was 0.44 ± 0.30. Sleepiness scores were higher in African American children; children with shorter sleep duration; older children; and overweight children. At endpoint, mean mESS score decreased by 2.0 ± 4.2 in the eAT group versus 0.3 ± 4.0 in the WWSC group (P < 0.0001); mean PSQ-SS score decreased 0.29 ± 0.40 in eAT versus 0.08 ± 0.40 in the WWSC group (P < 0.0001). Despite higher baseline sleepiness, African American children experienced similar improvement with adenotonsillectomy than other children. Improvement in sleepiness was weakly associated with improved apnea-hypopnea index or oxygen desaturation indices, but not with change in other polysomnographic measures. Conclusions: Sleepiness assessed by parent report was prevalent; improved more after eAT than after WWSC; and was not strongly predicted by sleep disturbances identified by PSG. Clinical Trial Registration: Childhood Adenotonsillectomy Study for Children with OSA (CHAT). ClinicalTrials.gov Identifier #NCT00560859. Citation: Paruthi S, Buchanan P, Weng J, Chervin RD, Mitchell RB, Dore-Stites D, Sadhwani A, Katz ES, Bent J, Rosen

  5. High risk for obstructive sleep apnea and other sleep disorders among overweight and obese pregnant women.

    PubMed

    Rice, Jayne R; Larrabure-Torrealva, Gloria T; Luque Fernandez, Miguel Angel; Grande, Mirtha; Motta, Vicky; Barrios, Yasmin V; Sanchez, Sixto; Gelaye, Bizu; Williams, Michelle A

    2015-09-02

    Obstructive sleep apnea (OSA), a common and serious disorder in which breathing repeatedly stops during sleep, is associated with excess weight and obesity. Little is known about the co-occurrence of OSA among pregnant women from low and middle-income countries. We examined the extent to which maternal pre-pregnancy overweight or obesity status are associated with high risk for OSA, poor sleep quality, and excessive daytime sleepiness in 1032 pregnant women in Lima, Peru. The Berlin questionnaire was used to identify women at high risk for OSA. The Pittsburgh Sleep Quality Index (PSQI) and Epworth Sleepiness Scale (ESS) were used to examine sleep quality and excessive daytime sleepiness, respectively. Multinomial logistic regression procedures were employed to estimate odds ratios (aOR) and 95% confidence intervals (CI) adjusted for putative confounding factors. Compared with lean women (<25 kg/m(2)), overweight women (25-29.9 kg/m(2)) had 3.69-fold higher odds of high risk for OSA (95% CI 1.82-7.50). The corresponding aOR for obese women (≥30 kg/m(2)) was 13.23 (95% CI: 6.25-28.01). Obese women, as compared with their lean counterparts had a 1.61-fold higher odds of poor sleep quality (95% CI: 1.00-2.63). Overweight or obese pregnant women have increased odds of sleep disorders, particularly OSA. OSA screening and risk management may be indicated among pregnant women in low and middle income countries, particularly those undergoing rapid epidemiologic transitions characterized by increased prevalence of excessive adult weight gain.

  6. French consensus. Idiopathic hypersomnia: Investigations and follow-up.

    PubMed

    Leu-Semenescu, S; Quera-Salva, M-A; Dauvilliers, Y

    Idiopathic hypersomnia is a rare, central hypersomnia, recently identified and to date of unknown physiopathology. It is characterised by a more or less permanent, excessive daytime sleepiness, associated with long and unrefreshing naps. Night-time sleep is of good quality, excessive in quantity, associated with sleep inertia in the subtype previously described as "with long sleep time". Diagnosis of idiopathic hypersomnia is complex due to the absence of a quantifiable biomarker, the heterogeneous symptoms, which overlap with the clinical picture of type 2 narcolepsy, and its variable evolution over time. Detailed evaluation enables other frequent causes of somnolence, such as depression or sleep deprivation, to be eliminated. Polysomnography and multiple sleep latency tests (MSLT) are essential to rule out other sleep pathologies and to objectify excessive daytime sleepiness. Sometimes the MSLT do not show excessive sleepiness, hence a continued sleep recording of at least 24hours is necessary to show prolonged sleep (>11h/24h). In this article, we propose recommendations for the work-up to be carried out during diagnosis and follow-up for patients suffering from idiopathic hypersomnia. Copyright © 2016 Elsevier Masson SAS. All rights reserved.

  7. Awareness of driving while sleepy and road traffic accidents: prospective study in GAZEL cohort.

    PubMed

    Nabi, Hermann; Guéguen, Alice; Chiron, Mireille; Lafont, Sylviane; Zins, Marie; Lagarde, Emmanuel

    2006-07-08

    To examine the association between self assessed driving while sleepy and the risk of serious road traffic accidents (RTAs). Prospective cohort study. France. 13 299 of the 19 894 living members of the GAZEL cohort, workers and recent retirees of a French national utility company followed up since 1989. Frequency of driving while sleepy in the previous 12 months, reported in 2001; rate ratios for serious RTAs in 2001-3, estimated by using generalised linear Poisson regression models with time dependent covariates. The risk of serious RTAs increased proportionally with the frequency of self reported driving while sleepy. After adjustment for sociodemographic characteristics, driving behaviour variables, work conditions, retirement, medical conditions and treatments, depressive symptoms, and sleep disorders, the adjusted rate ratios of serious RTAs for participants who reported driving while sleepy in the previous 12 months "a few times" or "once a month or more often" were 1.5 (95% confidence interval 1.2 to 2.0) and 2.9 (1.3 to 6.3) respectively compared with those who reported not driving while sleepy over the same period. These associations were not explained by any reported sleep disorders. Self assessed driving while sleepy was a powerful predictor of serious RTAs, suggesting that drivers' awareness of their sleepiness while driving is not sufficient to prevent them from having RTAs. Messages on prevention should therefore focus on convincing sleepy drivers to stop driving and sleep before resuming their journey.

  8. Intermittent fasting during Ramadan: does it affect sleep?

    PubMed

    Bahammam, Ahmed S; Almushailhi, Khalid; Pandi-Perumal, Seithikurippu R; Sharif, Munir M

    2014-02-01

    Islamic intermittent fasting is distinct from regular voluntary or experimental fasting. We hypothesised that if a regimen of a fixed sleep-wake schedule and a fixed caloric intake is followed during intermittent fasting, the effects of fasting on sleep architecture and daytime sleepiness will be minimal. Therefore, we designed this study to objectively assess the effects of Islamic intermittent fasting on sleep architecture and daytime sleepiness. Eight healthy volunteers reported to the Sleep Disorders Centre on five occasions for polysomnography and multiple sleep latency tests: (1) during adaptation; (2) 3 weeks before Ramadan, after having performed Islamic fasting for 1 week (baseline fasting); (3) 1 week before Ramadan (non-fasting baseline); (4) 2 weeks into Ramadan (Ramadan); and (5) 2 weeks after Ramadan (non-fasting; Recovery). Daytime sleepiness was assessed using the Epworth Sleepiness Scale and the multiple sleep latency test. The participants had a mean age of 26.6 ± 4.9 years, a body mass index of 23.7 ± 3.5 kg m(-2) and an Epworth Sleepiness Scale score of 7.3 ± 2.7. There was no change in weight or the Epworth Sleepiness Scale in the four study periods. The rapid eye movement sleep percentage was significantly lower during fasting. There was no difference in sleep latency, non-rapid eye movement sleep percentage, arousal index and sleep efficiency. The multiple sleep latency test analysis revealed no difference in the sleep latency between the 'non-fasting baseline', 'baseline fasting', 'Ramadan' and 'Recovery' time points. Under conditions of a fixed sleep-wake schedule and a fixed caloric intake, Islamic intermittent fasting results in decreased rapid eye movement sleep with no impact on other sleep stages, the arousal index or daytime sleepiness. © 2013 European Sleep Research Society.

  9. Cross-sectional Internet-based survey of Japanese permanent daytime workers' sleep and daily rest periods.

    PubMed

    Ikeda, Hiroki; Kubo, Tomohide; Sasaki, Takeshi; Liu, Xinxin; Matsuo, Tomoaki; So, Rina; Matsumoto, Shun; Yamauchi, Takashi; Takahashi, Masaya

    2018-05-25

    This study aimed to describe the sleep quantity, sleep quality, and daily rest periods (DRPs) of Japanese permanent daytime workers. Information about the usual DRP, sleep quantity, and sleep quality (Japanese version of the Pittsburgh Sleep Quality Index: PSQI-J) of 3,867 permanent daytime workers in Japan was gathered through an Internet-based survey. This information was analyzed and divided into the following eight DRP groups: <10, 10, 11, 12, 13, 14, 15, and ≥16 h. The sleep durations for workers in the <10, 10, 11, 12, 13, 14, 15, and ≥16 h DRP groups were found to be 5.3, 5.9, 6.1, 6.3, 6.5, 6.7, 6.7, and 6.9 h, respectively. The trend analysis revealed a significant linear trend as the shorter the DRP, the shorter was the sleep duration. The PSQI-J scores for the <10, 10, 11, 12, 13, 14, 15, and ≥16 h DRP groups were 7.1, 6.7, 6.7, 6.3, 6.0 (5.999), 5.6, 5.2, and 5.2, respectively. The trend analysis revealed a significant linear trend as the shorter the DRP, the lower was the sleep quality. This study described sleep quantity, sleep quality, and DRP in Japanese daytime workers. It was found that a shorter DRP was associated with poorer sleep quantity as well as quality.

  10. 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.

  11. Sleep apnoea is a common occurrence in females.

    PubMed

    Franklin, Karl A; Sahlin, Carin; Stenlund, Hans; Lindberg, Eva

    2013-03-01

    Obstructive sleep apnoea (OSA) is primarily regarded as a male disorder, presenting with snoring, daytime sleepiness and cardiovascular disease. We aimed to determine the frequency of sleep apnoea among females in the general population. We investigated 400 females from a population-based random sample of 10,000 females aged 20-70 yrs. They answered a questionnaire and performed overnight polysomnography. OSA (apnoea/hypopnoea index (AHI) ≥5) was found in 50% (95% CI 45-55%) of females aged 20-70 yrs. Sleep apnoea was related to age, obesity and hypertension, but not to daytime sleepiness. Severe sleep apnoea (AHI ≥30) was present in 14% (95% CI 8.1-21%) of females aged 55-70 yrs and in 31% (95% CI 12-50%) of obese females with a body mass index of ≥30 kg·m(-2) aged 55-70 yrs. Sleep apnoea with daytime sleepiness and sleep apnoea with hypertension were observed as two different phenotypes of OSA. OSA occurs in 50% of females aged 20-70 yrs. 20% of females have moderate and 6% severe sleep apnoea. Sleep apnoea in females is related to age, obesity and hypertension, but not to daytime sleepiness. When searching for sleep apnoea in females, females with hypertension or obesity should be investigated.

  12. Social Media Use, Social Media Stress, and Sleep: Examining Cross-Sectional and Longitudinal Relationships in Adolescents.

    PubMed

    van der Schuur, Winneke A; Baumgartner, Susanne E; Sumter, Sindy R

    2018-01-09

    There are concerns that social media (SM) use and SM stress may disrupt sleep. However, evidence on both the cross-sectional and longitudinal relationships is limited. Therefore, the main aim of this study is to address this gap in the literature by examining the cross-sectional and longitudinal relationships between SM use, SM stress, and sleep (i.e., sleep latency and daytime sleepiness) in adolescents. In total, 1,441 adolescents 11-15 years, 51% boys) filled out a survey in at least one of three waves that were three to four months apart (N Wave1  = 1,241; N Wave2  = 1,216; N Wave3  = 1,103). Cross-sectionally, we found that SM use and SM stress were positively related to sleep latency and daytime sleepiness. However, when examined together, SM use was not a significant predictor of sleep latency and daytime sleepiness above the effects of SM stress. The longitudinal findings showed that SM stress was positively related to subsequent sleep latency and daytime sleepiness, but only among girls. Our findings stress that it is important to focus on how adolescents perceive and cope with their SM use, instead of focusing on the mere frequency of SM use.

  13. Self-reported sleep duration and daytime napping are associated with renal hyperfiltration in general population.

    PubMed

    Lin, Miao; Su, Qing; Wen, Junping; Wei, Shichao; Yao, Jin; Huang, Huibin; Liang, Jixing; Li, Liantao; Lin, Wei; Lin, Lixiang; Lu, Jieli; Bi, Yufang; Wang, Weiqing; Ning, Guang; Chen, Gang

    2018-03-01

    Renal hyperfiltration (RHF) has emerged as a novel marker of early renal damage in various conditions such as diabetes and metabolic syndrome. Aberrant sleep duration and excessive daytime napping may affect the development of chronic kidney disease (CKD). In this study, the association between sleep duration, daytime napping, and renal hyperfiltration was assessed. This study was conducted in three communities in China. A total of 16,119 community volunteers (5735 males and 10,384 females) aged 40-65 years without CKD were included for the study. Participants with short sleep duration (<6 h/day) or long sleep duration (≥10 h/day) were at a significantly increased risk of renal hyperfiltration. The fully adjusted ORs (95% CI) were 2.112 (1.107, 4.031) and 2.071 (1.504, 2.853), respectively (P < 0.05). In addition, those who took naps longer than 1.5 h per day had a higher risk of renal hyperfiltration compared with those without napping (OR 1.400, 95% CI 1.018-1.924). Further joint analysis indicated that participants with long sleep duration (≥10 h/day) had a more than twofold increased risk of RHF regardless of nap status compared with those who slept 8-9 h per day without daytime napping. The association between sleep duration or daytime napping and RHF could not be explained by the influence of sleep quality. Additional subgroup analysis showed long sleep duration (≥9 h/day) and long daytime napping (≥1.5 h) were associated with an increased risk of RHF among individuals with good sleep quality. Sleep duration less than 6 h/day or more than 10 h/day and long daytime napping tend to be associated with an increased risk of renal hyperfiltration in middle-aged general population, and this relationship was independent of diabetes, hypertension, obesity, or poor sleep quality.

  14. Sleep restriction therapy for insomnia is associated with reduced objective total sleep time, increased daytime somnolence, and objectively impaired vigilance: implications for the clinical management of insomnia disorder.

    PubMed

    Kyle, Simon D; Miller, Christopher B; Rogers, Zoe; Siriwardena, A Niroshan; Macmahon, Kenneth M; Espie, Colin A

    2014-02-01

    To investigate whether sleep restriction therapy (SRT) is associated with reduced objective total sleep time (TST), increased daytime somnolence, and impaired vigilance. Within-subject, noncontrolled treatment investigation. Sleep research laboratory. Sixteen patients [10 female, mean age = 47.1 (10.8) y] with well-defined psychophysiological insomnia (PI), reporting TST ≤ 6 h. Patients were treated with single-component SRT over a 4-w protocol, sleeping in the laboratory for 2 nights prior to treatment initiation and for 3 nights (SRT night 1, 8, 22) during the acute interventional phase. The psychomotor vigilance task (PVT) was completed at seven defined time points [day 0 (baseline), day 1,7,8,21,22 (acute treatment) and day 84 (3 mo)]. The Epworth Sleepiness Scale (ESS) was completed at baseline, w 1-4, and 3 mo. Subjective sleep outcomes and global insomnia severity significantly improved before and after SRT. There was, however, a robust decrease in PSG-defined TST during acute implementation of SRT, by an average of 91 min on night 1, 78 min on night 8, and 69 min on night 22, relative to baseline (P < 0.001; effect size range = 1.60-1.80). During SRT, PVT lapses were significantly increased from baseline (at three of five assessment points, all P < 0.05; effect size range = 0.69-0.78), returning to baseline levels by 3 mo (P = 0.43). A similar pattern was observed for RT, with RTs slowing during acute treatment (at four of five assessment points, all P < 0.05; effect size range = 0.57-0.89) and returning to pretreatment levels at 3 mo (P = 0.78). ESS scores were increased at w 1, 2, and 3 (relative to baseline; all P < 0.05); by 3 mo, sleepiness had returned to baseline (normative) levels (P = 0.65). For the first time we show that acute sleep restriction therapy is associated with reduced objective total sleep time, increased daytime sleepiness, and objective performance impairment. Our data have important implications for implementation guidelines

  15. Pupillographic assessment of sleepiness in sleep-deprived healthy subjects.

    PubMed

    Wilhelm, B; Wilhelm, H; Lüdtke, H; Streicher, P; Adler, M

    1998-05-01

    Spontaneous pupillary-behavior in darkness provides information about a subject's level of sleepiness. In the present work, pupil measurements in complete darkness and quiet have been recorded continuously over 11-minute period with infrared video pupillography at 25 Hz. The data have been analyzed to yield three parameters describing pupil behavior; the power of diameter variation at frequencies below 0.8 Hz (slow changes in pupil size), the pupillary unrest index, and the average pupil size. To investigate the changes of these parameters in sleep deprivation, spontaneous pupillary behavior in darkness was recorded every 2 hours in 13 healthy subjects from 19:00 to 07:00 during forced wakefulness. On each occasion, comparative subjective sleepiness was assessed with a self-rating scale (Stanford Sleepiness Scale, SSS). The power of slow pupillary oscillations (< or = 0.8 Hz) increased significantly and so did the values of SSS, while basic pupil diameter decreased significantly. Slow pupillary oscillations and SSS did not correlate well in general but high values of pupil parameters were always associated with high values in subjective rating. Our results demonstrate a strong relationship between ongoing sleep deprivation and typical changes in the frequency profiles of spontaneous pupillary oscillations and the tendency to instability in pupil size in normals. These findings suggest that the results of pupil data analysis permit an objective measurement of sleepiness.

  16. Sleep and sleepiness of fishermen on rotating schedules.

    PubMed

    Gander, Philippa; van den Berg, Margo; Signal, Leigh

    2008-04-01

    Seafaring is a hazardous occupation with high death and injury rates, but the role of seafarer fatigue in these events is generally not well documented. The International Maritime Organization has identified seafarer fatigue as an important health and safety issue. Most research to date has focused on more regularly scheduled types of operations (e.g., merchant vessels, ferries), but there is relatively little information on commercial fishing, which often involves high day-to-day and seasonal variability in work patterns and workload. The present study was designed to monitor the sleep and sleepiness of commercial fishermen at home and during extended periods at sea during the peak of the hoki fishing season, with a view to developing better fatigue management strategies for this workforce. Sleep (wrist actigraphy and sleep diaries) and sleepiness (Karolinska Sleepiness Scale [KSS] before and after each sleep period) of 20 deckhands were monitored for 4-13 days at home and for 5-9 days at sea while working a nominal 12 h on/6 h off schedule. On the 12 h on/6 hoff schedule, there was still a clear preference for sleep at night. Comparing the last three days at home and the first three days at sea showed that fishermen were more likely to have split sleep at sea (Wilcoxon signed ranks p < 0.001), but the median sleep/24 h did not differ significantly by location (5.9 h at sea vs. 6.7 h at home). However, on 23% of days at sea, fishermen obtained < 4 h total sleep/24 h, compared to 3% of days at home ( p(chi 2) < 0.01). Sleep efficiency, mean activity counts/min sleep, and subjective ratings of sleep quality did not differ significantly between the last three days at home and the first three days at sea. However, sleepiness ratings remained higher after sleep at sea (Wilcoxon signed ranks p < 0.05), with fishermen having post-sleep KSS ratings >or= 7 on 24% of days at sea vs. 9% of days at home (Wilcoxon signed ranks p < 0.01). This work adds to the limited number of

  17. The sleepy teenager - diagnostic challenges.

    PubMed

    Landtblom, Anne-Marie; Engström, Maria

    2014-01-01

    The sleepy teenager puts the doctor in a, often tricky, situation where it must be decided if we deal with normal physiology or if we should suspect pathological conditions. What medical investigations are proper to consider? What differential diagnoses should be considered in the first place? And what tools do we actually have? The symptoms and problems that usually are presented at the clinical visit can be both of medical and psychosocial character - and actually they are often a mixture of both. Subsequently, the challenge to investigate the sleepy teenager often includes the examination of a complex behavioral pattern. It is important to train and develop diagnostic skills and to realize that the physiological or pathological conditions that can cause the symptoms may have different explanations. Research in sleep disorders has shown different pathological mechanisms congruent with the variations in the clinical picture. There are probably also different patterns of involved neuronal circuits although common pathways may exist. The whole picture remains to be drawn in this interesting and challenging area.

  18. Effects of Ayurvedic Oil-Dripping Treatment with Sesame Oil vs. with Warm Water on Sleep: A Randomized Single-Blinded Crossover Pilot Study

    PubMed Central

    Yorifuji, Takashi; Tsuda, Toshihide; Doi, Hiroyuki

    2016-01-01

    Abstract Objectives: Ayurvedic oil-dripping treatment (Shirodhara) is often used for treating sleep problems. However, few properly designed studies have been conducted, and the quantitative effect of Shirodhara is unclear. This study sought to quantitatively evaluate the effect of sesame oil Shirodhara (SOS) against warm water Shirodhara (WWS) on improving sleep quality and quality of life (QOL) among persons reporting sleep problems. Methods: This randomized, single-blinded, crossover study recruited 20 participants. Each participant received seven 30-minute sessions within 2 weeks with either liquid. The washout period was at least 2 months. The Shirodhara procedure was conducted by a robotic oil-drip system. The outcomes were assessed by the Pittsburgh Sleep Quality Index (PSQI) for sleep quality, Epworth Sleepiness Scale (ESS) for daytime sleepiness, World Health Organization Quality of Life 26 (WHO-QOL26) for QOL, and a sleep monitor instrument for objective sleep measures. Changes between baseline and follow-up periods were compared between the two types of Shirodhara. Analysis was performed with generalized estimating equations. Results: Of 20 participants, 15 completed the study. SOS improved sleep quality, as measured by PSQI. The SOS score was 1.83 points lower (95% confidence interval [CI], −3.37 to −0.30) at 2-week follow-up and 1.73 points lower (95% CI, −3.84 to 0.38) than WWS at 6-week follow-up. Although marginally significant, SOS also improved QOL by 0.22 points at 2-week follow-up and 0.19 points at 6-week follow-up compared with WWS. After SOS, no beneficial effects were observed on daytime sleepiness or objective sleep measures. Conclusions: This pilot study demonstrated that SOS may be a safe potential treatment to improve sleep quality and QOL in persons with sleep problems. PMID:26669255

  19. Effects of Ayurvedic Oil-Dripping Treatment with Sesame Oil vs. with Warm Water on Sleep: A Randomized Single-Blinded Crossover Pilot Study.

    PubMed

    Tokinobu, Akiko; Yorifuji, Takashi; Tsuda, Toshihide; Doi, Hiroyuki

    2016-01-01

    Ayurvedic oil-dripping treatment (Shirodhara) is often used for treating sleep problems. However, few properly designed studies have been conducted, and the quantitative effect of Shirodhara is unclear. This study sought to quantitatively evaluate the effect of sesame oil Shirodhara (SOS) against warm water Shirodhara (WWS) on improving sleep quality and quality of life (QOL) among persons reporting sleep problems. This randomized, single-blinded, crossover study recruited 20 participants. Each participant received seven 30-minute sessions within 2 weeks with either liquid. The washout period was at least 2 months. The Shirodhara procedure was conducted by a robotic oil-drip system. The outcomes were assessed by the Pittsburgh Sleep Quality Index (PSQI) for sleep quality, Epworth Sleepiness Scale (ESS) for daytime sleepiness, World Health Organization Quality of Life 26 (WHO-QOL26) for QOL, and a sleep monitor instrument for objective sleep measures. Changes between baseline and follow-up periods were compared between the two types of Shirodhara. Analysis was performed with generalized estimating equations. Of 20 participants, 15 completed the study. SOS improved sleep quality, as measured by PSQI. The SOS score was 1.83 points lower (95% confidence interval [CI], -3.37 to -0.30) at 2-week follow-up and 1.73 points lower (95% CI, -3.84 to 0.38) than WWS at 6-week follow-up. Although marginally significant, SOS also improved QOL by 0.22 points at 2-week follow-up and 0.19 points at 6-week follow-up compared with WWS. After SOS, no beneficial effects were observed on daytime sleepiness or objective sleep measures. This pilot study demonstrated that SOS may be a safe potential treatment to improve sleep quality and QOL in persons with sleep problems.

  20. 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.

  1. Links between sleep and daytime behaviour problems in children with Down syndrome.

    PubMed

    Esbensen, A J; Hoffman, E K; Beebe, D W; Byars, K C; Epstein, J

    2018-02-01

    In the general population, sleep problems have an impact on daytime performance. Despite sleep problems being common among children with Down syndrome, the impact of sleep problems on daytime behaviours in school-age children with Down syndrome is an understudied topic. Our study examined the relationship between parent-reported and actigraphy-measured sleep duration and sleep quality with parent and teacher reports of daytime behaviour problems among school-age children with Down syndrome. Thirty school-age children with Down syndrome wore an actigraph watch for a week at home at night. Their parent completed ratings of the child's sleep during that same week. Their parent and teacher completed a battery of measures to assess daytime behaviour. Parent reports of restless sleep behaviours on the Children's Sleep Habits Questionnaire, but not actigraph-measured sleep efficiency, was predictive of parent and teacher behavioural concerns on the Nisonger Child Behaviour Rating Form and the Vanderbilt ADHD Rating Scales. Actigraph-measured sleep period and parent-reported sleep duration on the Children's Sleep Habits Questionnaire was predictive of daytime parent-reported inattention. Actigraph-measured sleep period was predictive of parent-reported hyperactivity/impulsivity. The study findings suggest that sleep problems have complex relationships to both parent-reported and teacher-reported daytime behaviour concerns in children with Down syndrome. These findings have implications for understanding the factors impacting behavioural concerns and their treatment in school-age children with Down syndrome. © 2017 MENCAP and International Association of the Scientific Study of Intellectual and Developmental Disabilities and John Wiley & Sons Ltd.

  2. Empiric Auto-Titrating CPAP in People with Suspected Obstructive Sleep Apnea

    PubMed Central

    Drummond, Fitzgerald; Doelken, Peter; Ahmed, Qanta A.; Gilbert, Gregory E.; Strange, Charlie; Herpel, Laura; Frye, Michael D.

    2010-01-01

    Objective: Efficient diagnosis and treatment of obstructive sleep apnea (OSA) can be difficult because of time delays imposed by clinic visits and serial overnight polysomnography. In some cases, it may be desirable to initiate treatment for suspected OSA prior to polysomnography. Our objective was to compare the improvement of daytime sleepiness and sleep-related quality of life of patients with high clinical likelihood of having OSA who were randomly assigned to receive empiric auto-titrating continuous positive airway pressure (CPAP) while awaiting polysomnogram versus current usual care. Methods: Serial patients referred for overnight polysomnography who had high clinical likelihood of having OSA were randomly assigned to usual care or immediate initiation of auto-titrating CPAP. Epworth Sleepiness Scale (ESS) scores and the Functional Outcomes of Sleep Questionnaire (FOSQ) scores were obtained at baseline, 1 month after randomization, and again after initiation of fixed CPAP in control subjects and after the sleep study in auto-CPAP patients. Results: One hundred nine patients were randomized. Baseline demographics, daytime sleepiness, and sleep-related quality of life scores were similar between groups. One-month ESS and FOSQ scores were improved in the group empirically treated with auto-titrating CPAP. ESS scores improved in the first month by a mean of −3.2 (confidence interval −1.6 to −4.8, p < 0.001) and FOSQ scores improved by a mean of 1.5, (confidence interval 0.5 to 2.7, p = 0.02), whereas scores in the usual-care group did not change (p = NS). Following therapy directed by overnight polysomnography in the control group, there were no differences in ESS or FOSQ between the groups. No adverse events were observed. Conclusion: Empiric auto-CPAP resulted in symptomatic improvement of daytime sleepiness and sleep-related quality of life in a cohort of patients awaiting polysomnography who had a high pretest probability of having OSA. Additional

  3. [Sleep quality in an adult population exposed to the noise of El Dorado Airport, Bogotá, 2012].

    PubMed

    Callejas, Lina María; Sarmiento, Rodrigo; Medina, Katalina; Sepúlveda, Henry; Deluque, Dayana; Escobar-Córdoba, Franklin E

    2015-08-01

    The airport of Bogotá lies within the city and its expansion could produce an increase in adverse effects on the health of the inhabitants of Fontibón and Engativá districts due to the noise it generates. To determine the prevalence of sleep disturbances and associated factors among residents of Fontibón exposed to this noise. A cross-sectional study design was used, involving a sample of 205 people aged 18 to 65, selected by means of stratified random sampling. Sleep quality was evaluated using the Pittsburgh Sleep Quality Index (PSQI) and the Epworth Sleepiness Scale (ESS). Descriptive statistics were carried out, as well as correlation tests between the different scales. A total of 60% of the residents reported poor quality sleep (PSQI>5), with a mean PSQI of 7.19 (SD=3.931), and the following pathological interruptions were found: subjective sleep quality, 27%; sleep latency, 39%; sleep duration, 33%; habitual sleep efficiency, 37%; sleep alterations, 30%; diurnal dysfunction, 40%, and use of sleeping medication, 5%. According to the Epworth Sleepiness Scale (ESS>10), 28% of residents reported daytime somnolence. Regarding the prevalence of poor quality sleep according to the Pittsburgh Sleep Quality Index, 17% of those who reported not being able to sleep because of noise associated this with air traffic. A correlation was observed between the index and the scale ( r =0.329, CI 95%: 0.20-0.44). Inhabitants of the district reported poor sleep quality due to exposure to noise, airport operations being one of the main generating sources. Noise mitigation strategies in the district need to be reviewed and the public health implications of the El Dorado Airport expansion should be considered.

  4. Driver sleepiness and risk of serious injury to car occupants: population based case control study

    PubMed Central

    Connor, Jennie; Norton, Robyn; Ameratunga, Shanthi; Robinson, Elizabeth; Civil, Ian; Dunn, Roger; Bailey, John; Jackson, Rod

    2002-01-01

    Objectives To estimate the contribution of driver sleepiness to the causes of car crash injuries. Design Population based case control study. Setting Auckland region of New Zealand, April 1998 to July 1999. Participants 571 car drivers involved in crashes where at least one occupant was admitted to hospital or killed (“injury crash”); 588 car drivers recruited while driving on public roads (controls), representative of all time spent driving in the study region during the study period. Main outcome measures Relative risk for injury crash associated with driver characteristics related to sleep, and the population attributable risk for driver sleepiness. Results There was a strong association between measures of acute sleepiness and the risk of an injury crash. After adjustment for major confounders significantly increased risk was associated with drivers who identified themselves as sleepy (Stanford sleepiness score 4-7 v 1-3; odds ratio 8.2, 95% confidence interval 3.4 to 19.7); with drivers who reported five hours or less of sleep in the previous 24 hours compared with more than five hours (2.7, 1.4 to 5.4); and with driving between 2 am and 5 am compared with other times of day (5.6, 1.4 to 22.7). No increase in risk was associated with measures of chronic sleepiness. The population attributable risk for driving with one or more of the acute sleepiness risk factors was 19% (15% to 25%). Conclusions Acute sleepiness in car drivers significantly increases the risk of a crash in which a car occupant is injured or killed. Reductions in road traffic injuries may be achieved if fewer people drive when they are sleepy or have been deprived of sleep or drive between 2 am and 5 am. What is already known on this topicDriver sleepiness is considered a potentially important risk factor for car crashes and related injuries but the association has not been reliably quantifiedPublished estimates of the proportion of car crashes attributable to driver sleepiness vary from

  5. J-curve relation between daytime nap duration and type 2 diabetes or metabolic syndrome: A dose-response meta-analysis

    PubMed Central

    Yamada, Tomohide; Shojima, Nobuhiro; Yamauchi, Toshimasa; Kadowaki, Takashi

    2016-01-01

    Adequate sleep is important for good health, but it is not always easy to achieve because of social factors. Daytime napping is widely prevalent around the world. We performed a meta-analysis to investigate the association between napping (or excessive daytime sleepiness: EDS) and the risk of type 2 diabetes or metabolic syndrome, and to quantify the potential dose-response relation using cubic spline models. Electronic databases were searched for articles published up to 2016, with 288,883 Asian and Western subjects. Pooled analysis revealed that a long nap (≥60 min/day) and EDS were each significantly associated with an increased risk of type 2 diabetes versus no nap or no EDS (odds ratio 1.46 (95% CI 1.23–1.74, p < 0.01) for a long nap and 2.00 (1.58–2.53) for EDS). In contrast, a short nap (<60 min/day) was not associated with diabetes (p = 0.75). Dose-response meta-analysis showed a J-curve relation between nap time and the risk of diabetes or metabolic syndrome, with no effect of napping up to about 40 minutes/day, followed by a sharp increase in risk at longer nap times. In summary, longer napping is associated with an increased risk of metabolic disease. Further studies are needed to confirm the benefit of a short nap. PMID:27909305

  6. Sleep Disorders in Patients with Bronchial Asthma

    PubMed Central

    Cukic, Vesna; Lovre, Vladimir; Dragisic, Dejan

    2011-01-01

    Respiratory disturbances during sleep are recognized as extremely common disorders with important clinical consequences. Breathing disorders during sleep can result in broad range of clinical manifestations, the most prevalent of which are unrefreshing sleep, daytime sleepiness and fatigue, and cognitive impairmant. There is also evidence that respiratory-related sleep disturbances can contribute to several common cardiovascular and metabolic disorders, including systemic hypertension, cardiac dysfunction, and insulin-resistance. Correlations are found between asthma-related symptoms and sleep disturbances. Difficulties inducing sleep, sleep fragmentation on polysomnography, early morning awakenings and daytime sleepiness are more common in asthmatics compared with subjects without asthma. The “morning deep” in asthma is relevant for the characterization of asthma severity, and impact drugs’ choices. Sleep and night control of asthma could be relevant to evaluate disease’s control. Appropriate asthma control recovering is guarantor for better sleep quality in these patients and less clinical consequences of respiratory disturbances during sleep. PMID:23678304

  7. Male and female ecstasy users: differences in patterns of use, sleep quality and mental health outcomes.

    PubMed

    Ogeil, Rowan P; Rajaratnam, Shantha M W; Broadbear, Jillian H

    2013-09-01

    Ecstasy users report a number of adverse effects following use including mood and sleep disturbances. The present study examined differences in characteristics of ecstasy use (amount, frequency of use, reported harm resulting from use) between males and females and assessed relationships between ecstasy use, sleep quality and mental health outcomes. An online survey of 268 ecstasy users (54.1% male, 45.9% female) was conducted. Validated sleep instruments assessing sleep quality and excessive daytime sleepiness, as well as questionnaires regarding physical and mental health (measured using the short-form health survey 12 (SF-12) and details of drug use were included. Male ecstasy users reported taking larger amounts of ecstasy, but were not more frequent users compared to females. Female ecstasy users were more likely to report increased harm following ecstasy including: feelings of guilt and remorse; failing to do what was normally expected of them; and having been told by others to cut down their ecstasy use. There were interactions between amount and gender and frequency and gender in predicting use of sleep medication and daytime dysfunction. There was a positive correlation between poorer sleep quality and negative mood, although this relationship was not moderated by sex. There is a significant association between sleep quality and mood disturbance in ecstasy users suggesting that these negative outcomes are co-morbid. These findings have implications for the treatment and advice given to ecstasy users who are experiencing sleep and/or mood related complaints. Copyright © 2013 Elsevier Ireland Ltd. All rights reserved.

  8. A systematic review of the sleep, sleepiness, and performance implications of limited wake shift work schedules.

    PubMed

    Short, Michelle A; Agostini, Alexandra; Lushington, Kurt; Dorrian, Jillian

    2015-09-01

    The aim of this review was to identify which limited wake shift work schedules (LWSW) best promote sleep, alertness, and performance. LWSW are fixed work/rest cycles where the time-at-work does is ≤8 hours and there is >1 rest period per day, on average, for ≥2 consecutive days. These schedules are commonly used in safety-critical industries such as transport and maritime industries. Literature was sourced using PubMed, Embase, PsycInfo, Scopus, and Google Scholar databases. We identified 20 independent studies (plus a further 2 overlapping studies), including 5 laboratory and 17 field-based studies focused on maritime watch keepers, ship bridge officers, and long-haul train drivers. The measurement of outcome measures was varied, incorporating subjective and objective measures of sleep: sleep diaries (N=5), actigraphy (N=4), and polysomnography, (N=3); sleepiness: Karolinska Sleepiness Scale (N=5), visual analog scale (VAS) alertness (N=2) and author-derived measures (N=2); and performance: Psychomotor Vigilance Test (PVT) (N=5), Reaction Time or Vigilance tasks (N=4), Vector and Letter Cancellation Test (N=1), and subjective performance (N=2). Of the three primary rosters examined (6 hours-on/6 hours-off, 8 hours-on/8 hours-off and 4 hours-on/8 hours-off), the 4 hours-on/8 hours-off roster was associated with better sleep and lower levels of sleepiness. Individuals working 4 hours-on/8 hours-off rosters averaged 1 hour more sleep per night than those working 6 hours-on/6 hours-off and 1.3 hours more sleep than those working 8 hours-on/8 hours-off (P<0.01). More broadly, findings indicate that LWSW schedules were associated with better sleep and lower sleepines in the case of (i) shorter time-at-work, (ii) more frequent rest breaks, (iii) shifts that start and end at the same clock time every 24 hours, and (iv) work shifts commencing in the daytime (as opposed to night). The findings for performance remain incomplete due to the small number of studies

  9. Association Between Portable Screen-Based Media Device Access or Use and Sleep Outcomes: A Systematic Review and Meta-analysis.

    PubMed

    Carter, Ben; Rees, Philippa; Hale, Lauren; Bhattacharjee, Darsharna; Paradkar, Mandar S

    2016-12-01

    Sleep is vital to children's biopsychosocial development. Inadequate sleep quantity and quality is a public health concern with an array of detrimental health outcomes. Portable mobile and media devices have become a ubiquitous part of children's lives and may affect their sleep duration and quality. To conduct a systematic review and meta-analysis to examine whether there is an association between portable screen-based media device (eg, cell phones and tablet devices) access or use in the sleep environment and sleep outcomes. A search strategy consisting of gray literature and 24 Medical Subject Headings was developed in Ovid MEDLINE and adapted for other databases between January 1, 2011, and June 15, 2015. Searches of the published literature were conducted across 12 databases. No language restriction was applied. The analysis included randomized clinical trials, cohort studies, and cross-sectional study designs. Inclusion criteria were studies of school-age children between 6 and 19 years. Exclusion criteria were studies of stationary exposures, such as televisions or desktop or personal computers, or studies investigating electromagnetic radiation. Of 467 studies identified, 20 cross-sectional studies were assessed for methodological quality. Two reviewers independently extracted data. The primary outcomes were inadequate sleep quantity, poor sleep quality, and excessive daytime sleepiness, studied according to an a priori protocol. Twenty studies were included, and their quality was assessed. The studies involved 125 198 children (mean [SD] age, 14.5 [2.2] years; 50.1% male). There was a strong and consistent association between bedtime media device use and inadequate sleep quantity (odds ratio [OR], 2.17; 95% CI, 1.42-3.32) (P < .001, I2 = 90%), poor sleep quality (OR, 1.46; 95% CI, 1.14-1.88) (P = .003, I2 = 76%), and excessive daytime sleepiness (OR, 2.72; 95% CI, 1.32-5.61) (P = .007, I2 = 50%). In addition, children who had

  10. Job stress, burnout, and job satisfaction in sleep apnea patients.

    PubMed

    Guglielmi, Ottavia; Jurado-Gámez, Bernabé; Gude, Francisco; Buela-Casal, Gualberto

    2014-09-01

    To assess job stress, burnout, and job satisfaction in patients with obstructive sleep apnea syndrome (OSAS). A total of 182 patients with OSAS and 71 healthy individuals completed the Job Content Questionnaire, the Maslach Burnout Inventory - General Survey, the Index of Job Satisfaction, the Epworth Sleepiness Scale, and the Pittsburgh Sleep Quality Index. All participants were assessed with full-night polysomnography. Survey scores of patients diagnosed with OSAS only differed from those of the control group in the emotional exhaustion dimension (P = 0.015). According to a multivariate analysis, the apnea-hypopnea index (AHI) was only correlated with perceived support at work (β coefficient = 0.142; P = 0.048). Associations were found between subjective sleep quality, perceived support from coworkers, and supervisors (β = 0.157; P = 0.025), psychological demands (β = 0.226; P = 0.001), emotional exhaustion (β = 0,405; P = 0.000), and cynicism (β = 0.224; P = 0.002). The study also revealed associations between excessive daytime sleepiness and the burnout dimensions emotional exhaustion (β = 0.232; P = 0.000) and cynicism (β = 0.139; P = 0.048). Objective parameters of OSAS such as the AHI seem to have limited influence on the psychosocial aspects of the occupational life of patients with OSAS. There is evidence of significant associations between the subjective symptoms of the disease, such as daytime sleepiness, subjective sleep quality, job stress, and burnout. Copyright © 2014 Elsevier B.V. All rights reserved.

  11. A novel sleep optimisation programme to improve athletes' well-being and performance.

    PubMed

    Van Ryswyk, Emer; Weeks, Richard; Bandick, Laura; O'Keefe, Michaela; Vakulin, Andrew; Catcheside, Peter; Barger, Laura; Potter, Andrew; Poulos, Nick; Wallace, Jarryd; Antic, Nick A

    2017-03-01

    To improve well-being and performance indicators in a group of Australian Football League (AFL) players via a six-week sleep optimisation programme. Prospective intervention study following observations suggestive of reduced sleep and excessive daytime sleepiness in an AFL group. Athletes from the Adelaide Football Club were invited to participate if they had played AFL senior-level football for 1-5 years, or if they had excessive daytime sleepiness (Epworth Sleepiness Scale [ESS] >10), measured via ESS. An initial education session explained normal sleep needs, and how to achieve increased sleep duration and quality. Participants (n = 25) received ongoing feedback on their sleep, and a mid-programme education and feedback session. Sleep duration, quality and related outcomes were measured during week one and at the conclusion of the six-week intervention period using sleep diaries, actigraphy, ESS, Pittsburgh Sleep Quality Index, Profile of Mood States, Training Distress Scale, Perceived Stress Scale and the Psychomotor Vigilance Task. Sleep diaries demonstrated an increase in total sleep time of approximately 20 min (498.8 ± 53.8 to 518.7 ± 34.3; p < .05) and a 2% increase in sleep efficiency (p < 0.05). There was a corresponding increase in vigour (p < 0.001) and decrease in fatigue (p < 0.05). Improvements in measures of sleep efficiency, fatigue and vigour indicate that a sleep optimisation programme may improve athletes' well-being. More research is required into the effects of sleep optimisation on athletic performance.

  12. A quasi-experimental study of the impact of school start time changes on adolescent sleep.

    PubMed

    Owens, Judith A; Dearth-Wesley, Tracy; Herman, Allison N; Oakes, J Michael; Whitaker, Robert C

    2017-12-01

    To determine whether simultaneous school start time changes (delay for some schools; advance for others) impact adolescents' sleep. Quasi-experimental study using cross-sectional surveys before and after changes to school start times in September 2015. Eight middle (grades 7-8), 3 secondary (grades 7-12), and 8 high (grades 9-12) schools in Fairfax County (Virginia) public schools. A total of 2017 (6% of ~34,900) students were surveyed before start time changes, and 1180 (3% of ~35,300) were surveyed after. A 50-minute delay (7:20 to 8:10 am) in start time for high schools and secondary schools and a 30-minute advance (8:00 to 7:30 am) for middle schools. Differences before and after start time changes in self-reported sleep duration and daytime sleepiness. Among respondents, 57.5% were non-Hispanic white, and 10.3% received free or reduced-priced school meals. Before start time changes, high/secondary and middle school students slept a mean (SD) of 7.4 (1.2) and 8.4 (1.0) hours on school nights, respectively, and had a prevalence of daytime sleepiness of 78.4% and 57.2%, respectively. Adjusted for potential confounders, students with a 50-minute delay slept 30.1 minutes longer (95% confidence interval [CI], 24.3-36.0) on school nights and had less daytime sleepiness (-4.8%; 95% CI, -8.5% to -1.1%), whereas students with a 30-minute advance slept 14.8 minutes less (95% CI, -21.6 to -8.0) and had more daytime sleepiness (8.0%; 95% CI, 2.5%-13.5%). Both advances and delays in school start times are associated with changes in adolescents' school-night sleep duration and daytime sleepiness. Larger changes might occur with later start times. Copyright © 2017 National Sleep Foundation. Published by Elsevier Inc. All rights reserved.

  13. The efficacy of a brief motivational enhancement education program on CPAP adherence in OSA: a randomized controlled trial.

    PubMed

    Lai, Agnes Y K; Fong, Daniel Y T; Lam, Jamie C M; Weaver, Terri E; Ip, Mary S M

    2014-09-01

    Poor adherence to CPAP treatment in OSA adversely affects the effectiveness of this therapy. This randomized controlled trial (RCT) examined the efficacy of a brief motivational enhancement education program in improving adherence to CPAP treatment in subjects with OSA. Subjects with newly diagnosed OSA were recruited into this RCT. The control group received usual advice on the importance of CPAP therapy and its care. The intervention group received usual care plus a brief motivational enhancement education program directed at enhancing the subjects' knowledge, motivation, and self-efficacy to use CPAP through the use of a 25-min video, a 20-min patient-centered interview, and a 10-min telephone follow-up. Self-reported daytime sleepiness adherence-related cognitions and quality of life were assessed at 1 month and 3 months. CPAP usage data were downloaded at the completion of this 3-month study. One hundred subjects with OSA (mean ± SD, age 52 ± 10 years; Epworth Sleepiness Scales [ESS], 9 ± 5; median [interquartile range] apnea-hypopnea index, 29 [20, 53] events/h) prescribed CPAP treatment were recruited. The intervention group had better CPAP use (higher daily CPAP usage by 2 h/d [Cohen d = 1.33, P < .001], a fourfold increase in the number using CPAP for ≥ 70% of days with ≥ 4 h/d [P < .001]), and greater improvements in daytime sleepiness (ESS) by 2.2 units (P = .001) and treatment self-efficacy by 0.2 units (P = .012) compared with the control group. Subjects with OSA who received motivational enhancement education in addition to usual care were more likely to show better adherence to CPAP treatment, with greater improvements in treatment self-efficacy and daytime sleepiness. ClinicalTrials.gov; No.: NCT01173406; URL: www.clinicaltrials.gov.

  14. 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

  15. 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).

  16. Prevalence of major obstructive sleep apnea syndrome symptoms in coal miners and healthy adults.

    PubMed

    Kart, Levent; Dutkun, Yalçın; Altın, Remzi; Ornek, Tacettin; Kıran, Sibel

    2010-01-01

    Obstructive sleep apnea syndrome obstructive sleep apnea syndrome is associated with symptoms including habitual snoring, witness apnea and excessive daytime sleepiness. Also obstructive sleep apnea syndrome is related to some occupations which are needed attention for work accident. We aimed to determine the prevalence of snoring, witnessed apnea and excessive daytime sleepiness in coal workers and healthy adults in Zonguldak city center, and also evaluate the differences between these groups. This study consisted of 423 underground coal workers and 355 individuals living in centre of Zonguldak. Study and comparison group were chosen by nonstratified randomized sampling method. Data were collected by a questionnaire that included information regarding snoring, witnessed apnea and excessive daytime sleepiness. Mean age was 43.3 ± 6.05 years in miners and 44.3 ± 11.8 years in comparison group. In miners, snoring frequency was determined as 42.6%, witnessed apneas were 4.0%, and daytime sleepiness were 4.7%. In comparison group, these symptoms were 38.6%, 4.8% and 2.8% respectively. There were no statistical differences between coal workers and comparison group in these symptoms. Also snoring prevalence was higher in smoker miners. We found that major symptoms of obstructive sleep apnea syndrome in coal workers are similar to general population in Zonguldak. Further studies that constucted higher populations and with polysomnography are needed to evaluate these findings.

  17. Treatment of a Circadian Rhythm Disturbance in a 2-Year-Old Blind Child.

    ERIC Educational Resources Information Center

    Mindell, J. A.; And Others

    1996-01-01

    The use of sleep scheduling and a daytime routine for the treatment of circadian rhythm disorder was found helpful in decreasing a blind 2-year old's nighttime wake periods and daytime sleepiness. (DB)

  18. Assessing sleepiness and sleep disorders in Australian long-distance commercial vehicle drivers: self-report versus an "at home" monitoring device.

    PubMed

    Sharwood, Lisa N; Elkington, Jane; Stevenson, Mark; Grunstein, Ronald R; Meuleners, Lynn; Ivers, Rebecca Q; Haworth, Narelle; Norton, Robyn; Wong, Keith K

    2012-04-01

    As obstructive sleep apnea (OSA) is associated with a higher risk of motor vehicle crashes, there is increasing regulatory interest in the identification of commercial motor vehicle (CMV) drivers with this condition. This study aimed to determine the relationship between subjective versus objective assessment of OSA in CMV drivers. Cross-sectional survey. Heavy vehicle truck stops located across the road network of 2 large Australian states. A random sample of long distance commercial vehicle drivers (n = 517). None. Drivers were interviewed regarding their driving experience, personal health, shift schedules, payments, and various questions on sleep and tiredness in order to describe their sleep health across a range of variables. In addition, home recordings using a flow monitor were used during one night of sleep. Only 4.4% of drivers reported a previous diagnosis of sleep apnea, while our at home diagnostic test found a further 41% of long-distance heavy vehicle drivers likely to have sleep apnea. The multivariable apnea prediction index, based on self-report measures, showed poor agreement with the home-monitor detected sleep apnea (AUC 0.58, 95%CI = 0.49-0.62), and only 12% of drivers reported daytime sleepiness (Epworth Sleepiness Scale score > 10). Thirty-six percent of drivers were overweight and a further 50% obese; 49% of drivers were cigarette smokers. Sleep apnea remains a significant and unrecognized problem in CMV drivers, who we found to have multiple health risks. Objective testing for this sleep disorder needs to be considered, as symptom reports and self-identification appear insufficient to accurately identify those at risk.

  19. Daytime Aspect Camera for Balloon Altitudes

    NASA Technical Reports Server (NTRS)

    Dietz, Kurt L.; Ramsey, Brian D.; Alexander, Cheryl D.; Apple, Jeff A.; Ghosh, Kajal K.; Swift, Wesley R.

    2002-01-01

    We have designed, built, and flight-tested a new star camera for daytime guiding of pointed balloon-borne experiments at altitudes around 40 km. The camera and lens are commercially available, off-the-shelf components, but require a custom-built baffle to reduce stray light, especially near the sunlit limb of the balloon. This new camera, which operates in the 600- to 1000-nm region of the spectrum, successfully provides daytime aspect information of approx. 10 arcsec resolution for two distinct star fields near the galactic plane. The detected scattered-light backgrounds show good agreement with the Air Force MODTRAN models used to design the camera, but the daytime stellar magnitude limit was lower than expected due to longitudinal chromatic aberration in the lens. Replacing the commercial lens with a custom-built lens should allow the system to track stars in any arbitrary area of the sky during the daytime.

  20. Hypnotic efficacy of zaleplon for daytime sleep in rested individuals.

    PubMed

    Whitmore, Jeffrey N; Fischer, Joseph R; Storm, William F

    2004-08-01

    The primary objective of this study was to determine whether zaleplon (10 mg) effectively promoted sleep during the daytime in well-rested individuals when compared to placebo. A secondary objective was to see if, while not expected, the use of zaleplon impacted the performance of well-rested individuals upon awakening. Repeated measures with 2 within-subject factors: drug (placebo/zaleplon) and trial (hourly testing during waking hours). Polysomnographic variables were recorded during a 3.5-hour nap following drug administration. Performance measures and subjective reports were collected during every waking trial of each session. The study was conducted at the Chronobiology and Sleep Laboratory located at Brooks Air Force Base. Twelve participants, 6 men and 6 women. 10-mg zaleplon or placebo capsules, single afternoon dose. Drug administration was counterbalanced and double-blinded. Zaleplon allowed participants to obtain significantly more slow-wave sleep than under placebo. There was also a trend for participants under zaleplon to accomplish a greater amount of sleep than under placebo. Performance was not adversely impacted following a 3.5-hour daytime sleep under zaleplon, nor were any undesirable symptoms induced. Zaleplon improves sleep quality when used by rested individuals to accomplish daytime sleep.

  1. Mindfulness Meditation and Improvement in Sleep Quality and Daytime Impairment Among Older Adults With Sleep Disturbances

    PubMed Central

    Black, David S.; O’Reilly, Gillian A.; Olmstead, Richard; Breen, Elizabeth C.; Irwin, Michael R.

    2015-01-01

    IMPORTANCE Sleep disturbances are most prevalent among older adults and often go untreated. Treatment options for sleep disturbances remain limited, and there is a need for community-accessible programs that can improve sleep. OBJECTIVE To determine the efficacy of a mind-body medicine intervention, called mindfulness meditation, to promote sleep quality in older adults with moderate sleep disturbances. DESIGN, SETTING, AND PARTICIPANTS Randomized clinical trial with 2 parallel groups conducted from January 1 to December 31, 2012, at a medical research center among an older adult sample (mean [SD] age, 66.3 [7.4] years) with moderate sleep disturbances (Pittsburgh Sleep Quality Index [PSQI] >5). INTERVENTIONS A standardized mindful awareness practices (MAPs) intervention (n = 24) or a sleep hygiene education (SHE) intervention (n = 25) was randomized to participants, who received a 6-week intervention (2 hours per week) with assigned homework. MAIN OUTCOMES AND MEASURES The study was powered to detect between-group differences in moderate sleep disturbance measured via the PSQI at postintervention. Secondary outcomes pertained to sleep-related daytime impairment and included validated measures of insomnia symptoms, depression, anxiety, stress, and fatigue, as well as inflammatory signaling via nuclear factor (NF)–κB. RESULTS Using an intent-to-treat analysis, participants in the MAPs group showed significant improvement relative to those in the SHE group on the PSQI. With the MAPs intervention, the mean (SD) PSQIs were 10.2 (1.7) at baseline and 7.4 (1.9) at postintervention. With the SHE intervention, the mean (SD) PSQIs were 10.2 (1.8) at baseline and 9.1 (2.0) at postintervention. The between-group mean difference was 1.8 (95%CI, 0.6–2.9), with an effect size of 0.89. The MAPs group showed significant improvement relative to the SHE group on secondary health outcomes of insomnia symptoms, depression symptoms, fatigue interference, and fatigue severity (P

  2. [Sleep disorders and epilepsy].

    PubMed

    Aoki, Ryo; Ito, Hiroshi

    2014-05-01

    It has been reported that patients with epilepsy often have insomnia and/or daytime sleepiness; the symptomatologic features differ in seizure types. Not only the administration of anti-epileptics, but also inappropriate sleep hygiene cause daytime sleepiness. In subjective assessment of sleepiness, we need to pay attention if it can correctly assess or not. The prevalence of obstructive sleep apnea in patients with epilepsy is approximately 10-30%. Sleep apnea deteriorates the seizure control because of worsen sleep condition by sleep apnea, especially in elderly patients. Some researchers report that continuous positive airway pressure was effective for seizure control. Patients with epilepsy occasionally have REM sleep behavior disorder as comorbidity. Examination using polysomnography is required for differential diagnosis.

  3. Sleep Disordered Breathing, Fatigue, and Sleepiness in HIV-Infected and -Uninfected Men

    PubMed Central

    Patil, Susheel P.; Brown, Todd T.; Jacobson, Lisa P.; Margolick, Joseph B.; Laffan, Alison; Johnson-Hill, Lisette; Godfrey, Rebecca; Johnson, Jacquett; Reynolds, Sandra; Schwartz, Alan R.; Smith, Philip L.

    2014-01-01

    Study Objectives We investigated the association of HIV infection and highly active antiretroviral therapy (HAART) with sleep disordered breathing (SDB), fatigue, and sleepiness. Methods HIV-uninfected men (HIV−; n = 60), HIV-infected men using HAART (HIV+/HAART+; n = 58), and HIV-infected men not using HAART (HIV+/HAART−; n = 41) recruited from two sites of the Multicenter AIDS cohort study (MACS) underwent a nocturnal sleep study, anthropometric assessment, and questionnaires for fatigue and the Epworth Sleepiness Scale. The prevalence of SDB in HIV- men was compared to that in men matched from the Sleep Heart Health Study (SHHS). Results The prevalence of SDB was unexpectedly high in all groups: 86.7% for HIV−, 70.7% for HIV+/HAART+, and 73.2% for HIV+/HAART−, despite lower body-mass indices (BMI) in HIV+ groups. The higher prevalence in the HIV− men was significant in univariate analyses but not after adjustment for BMI and other variables. SDB was significantly more common in HIV− men in this study than those in SHHS, and was common in participants with BMIs <25 kg/m2. HIV+ men reported fatigue more frequently than HIV− men (25.5% vs. 6.7%; p = 0.003), but self-reported sleepiness did not differ among the three groups. Sleepiness, but not fatigue, was significantly associated with SDB. Conclusions SDB was highly prevalent in HIV− and HIV+ men, despite a normal or slightly elevated BMI. The high rate of SDB in men who have sex with men deserves further investigation. Sleepiness, but not fatigue, was related to the presence of SDB. Clinicians caring for HIV-infected patients should distinguish between fatigue and sleepiness when considering those at risk for SDB, especially in non-obese men. PMID:24991815

  4. 47 CFR 73.157 - Antenna testing during daytime.

    Code of Federal Regulations, 2011 CFR

    2011-10-01

    ... 47 Telecommunication 4 2011-10-01 2011-10-01 false Antenna testing during daytime. 73.157 Section... BROADCAST SERVICES AM Broadcast Stations § 73.157 Antenna testing during daytime. (a) The licensee of a station using a directional antenna during daytime or nighttime hours may, without further authority...

  5. 47 CFR 73.157 - Antenna testing during daytime.

    Code of Federal Regulations, 2013 CFR

    2013-10-01

    ... 47 Telecommunication 4 2013-10-01 2013-10-01 false Antenna testing during daytime. 73.157 Section... BROADCAST SERVICES AM Broadcast Stations § 73.157 Antenna testing during daytime. (a) The licensee of a station using a directional antenna during daytime or nighttime hours may, without further authority...

  6. 47 CFR 73.157 - Antenna testing during daytime.

    Code of Federal Regulations, 2014 CFR

    2014-10-01

    ... 47 Telecommunication 4 2014-10-01 2014-10-01 false Antenna testing during daytime. 73.157 Section... BROADCAST SERVICES AM Broadcast Stations § 73.157 Antenna testing during daytime. (a) The licensee of a station using a directional antenna during daytime or nighttime hours may, without further authority...

  7. 47 CFR 73.157 - Antenna testing during daytime.

    Code of Federal Regulations, 2010 CFR

    2010-10-01

    ... 47 Telecommunication 4 2010-10-01 2010-10-01 false Antenna testing during daytime. 73.157 Section... BROADCAST SERVICES AM Broadcast Stations § 73.157 Antenna testing during daytime. (a) The licensee of a station using a directional antenna during daytime or nighttime hours may, without further authority...

  8. 47 CFR 73.157 - Antenna testing during daytime.

    Code of Federal Regulations, 2012 CFR

    2012-10-01

    ... 47 Telecommunication 4 2012-10-01 2012-10-01 false Antenna testing during daytime. 73.157 Section... BROADCAST SERVICES AM Broadcast Stations § 73.157 Antenna testing during daytime. (a) The licensee of a station using a directional antenna during daytime or nighttime hours may, without further authority...

  9. Electrodermal lability as an indicator for subjective sleepiness during total sleep deprivation.

    PubMed

    Michael, Lars; Passmann, Sven; Becker, Ruth

    2012-08-01

    The present study addresses the suitability of electrodermal lability as an indicator of individual vulnerability to the effects of total sleep deprivation. During two complete circadian cycles, the effects of 48h of total sleep deprivation on physiological measures (electrodermal activity and body temperature), subjective sleepiness (measured by visual analogue scale and tiredness symptom scale) and task performance (reaction time and errors in a go/no go task) were investigated. Analyses of variance with repeated measures revealed substantial decreases of the number of skin conductance responses, body temperature, and increases for subjective sleepiness, reaction time and error rates. For all changes, strong circadian oscillations could be observed as well. The electrodermal more labile subgroup reported higher subjective sleepiness compared with electrodermal more stable participants, but showed no differences in the time courses of body temperature and task performance. Therefore, electrodermal lability seems to be a specific indicator for the changes in subjective sleepiness due to total sleep deprivation and circadian oscillations, but not a suitable indicator for vulnerability to the effects of sleep deprivation per se. © 2011 European Sleep Research Society.

  10. Nightmares affect the experience of sleep quality but not sleep architecture: an ambulatory polysomnographic study.

    PubMed

    Paul, Franc; Schredl, Michael; Alpers, Georg W

    2015-01-01

    Nightmares and bad dreams are common in people with emotional disturbances. For example, nightmares are a core symptom in posttraumatic stress disorder and about 50% of borderline personality disorder patients suffer from frequent nightmares. Independent of mental disorders, nightmares are often associated with sleep problems such as prolonged sleep latencies, poorer sleep quality, and daytime sleepiness. It has not been well documented whether this is reflected in objectively quantifiable physiological indices of sleep quality. Questionnaires regarding subjective sleep quality and ambulatory polysomnographic recordings of objective sleep parameters were collected during three consecutive nights in 17 individuals with frequent nightmares (NM) and 17 healthy control participants (HC). NM participants reported worse sleep quality, more waking problems and more severe insomnia compared to HC group. However, sleep measures obtained by ambulatory polysomnographic recordings revealed no group differences in (a) overall sleep architecture, (b) sleep cycle duration as well as REM density and REM duration in each cycle and (c) sleep architecture when only nights with nightmares were analyzed. Our findings support the observation that nightmares result in significant impairment which is independent from disturbed sleep architecture. Thus, these specific problems require specific attention and appropriate treatment.

  11. Can variations in visual behavior measures be good predictors of driver sleepiness? A real driving test study.

    PubMed

    Wang, Yonggang; Xin, Mengyang; Bai, Han; Zhao, Yangdong

    2017-02-17

    The primary purpose of this study was to examine the association between variations in visual behavior measures and subjective sleepiness levels across age groups over time to determine a quantitative method of measuring drivers' sleepiness levels. A total of 128 volunteer drivers in 4 age groups were asked to finish 2-, 3-, and 4-h continuous driving tasks on expressways, during which the driver's fixation, saccade, and blink measures were recorded by an eye-tracking system and the subjective sleepiness level was measured through the Stanford Sleepiness Scale. Two-way repeated measures analysis of variance was then used to examine the change in visual behavior measures across age groups over time and compare the interactive effects of these 2 factors on the dependent visual measures. Drivers' visual behavior measures and subjective sleepiness levels vary significantly over time but not across age groups. A statistically significant interaction between age group and driving duration was found in drivers' pupil diameter, deviation of search angle, saccade amplitude, blink frequency, blink duration, and closure duration. Additionally, change in a driver's subjective sleepiness level is positively or negatively associated with variation in visual behavior measures, and such relationships can be expressed in regression models for different period of driving duration. Driving duration affects drivers' sleepiness significantly, so the amount of continuous driving time should be strictly controlled. Moreover, driving sleepiness can be quantified through the change rate of drivers' visual behavior measures to alert drivers of sleepiness risk and to encourage rest periods. These results provide insight into potential strategies for reducing and preventing traffic accidents and injuries.

  12. Sleepiness and recovery in schedule change and the eighty-four hour workweek.

    PubMed

    Nordin, M; Knutsson, A

    2001-12-01

    The aims were to evaluate sleepiness and recovery during a schedule change, and during an 84-hours workweek. The control group (16 men) stayed on a six-week schedule, whereas the intervention group (12 men) transferred to a seven-week schedule. Sleepiness was estimated, using the KSS-scale, four times during the first and the third night in the fifth or sixth shift week. Recovery was assessed through four estimations on days one, three and five during the week off. Statistical testing was carried out using repeated measurement ANOVA. Sleepiness at night was affected by night (F = 4.90, p < 0.05) and hour (F = 33.64, p < 0.001) in both groups. The intervention group was sleepier during the first recovery day compared to the control group (F = 4.02, p < 0.05). Analysis of the 84-hour-week showed an effect of night (F = 8.98, p < 0.05) and hour (F = 71.60, p < 0.001) on night work, and day (F = 22.49, p < 0.01) and hour (F = 6.66, p < 0.05) on recovery. Sleepiness was more pronounced on the first recovery day (F = 23.08, p < 0.01). The seven-week schedule showed no effect that differed from that of the control group on sleepiness during the night shift. After the 84-hour workweek the workers recovered in about three days. The new schedules may affect the first recovery day negatively.

  13. Validation of a Hindi version of the Epworth Sleepiness Scale (ESS) at AIIMS, New Delhi in sleep-disordered breathing.

    PubMed

    Kanabar, K; Sharma, S K; Sreenivas, V; Biswas, A; Soneja, M

    2016-12-01

    The Epworth Sleepiness Scale (ESS) is one of the most widely used questionnaire for the assessment of excessive daytime sleepiness (EDS) in sleep-disordered breathing (SDB). This study was conducted to assess the validity of ESS in the Hindi language. The Hindi version was developed by translation and back translation by independent translators. The English and Hindi versions were administered to 115 bilingual subjects who presented with symptoms of SDB, of whom 98 underwent a polysomnography at a tertiary care hospital in North India. The questionnaire had a high level of internal consistency as measured by Cronbach's alpha (α = 0.84). There was no significant difference between the mean ESS scores of Hindi and English versions (11.65 ± 5.47 vs 11.70 ± 5.49, respectively; p = 0.80). The Hindi version of ESS showed a strong correlation with the English version (Spearman's correlation ρ = 0.98 and weighted kappa = 0.94). Each of the 8 individual questions of Hindi ESS demonstrated a good agreement with the corresponding English version. The Hindi ESS score was significantly higher in subjects with OSA compared to those without OSA (12.67 ± 5.29 vs 7.76 ± 5.44, respectively; p = 0.002). However, there was no difference in ESS score between mild and moderate OSA or between moderate and severe OSA. The Hindi version of the ESS showed a good internal consistency and a strong correlation with the English version and can be used in the Hindi-speaking population.

  14. Differential Sleep, Sleepiness, and Neurophysiology in the Insomnia Phenotypes of Shift Work Disorder

    PubMed Central

    Gumenyuk, Valentina; Belcher, Ren; Drake, Christopher L.; Roth, Thomas

    2015-01-01

    Study Objectives: To characterize and compare insomnia symptoms within two common phenotypes of Shift Work Disorder. Design: Observational laboratory and field study. Setting: Hospital sleep center. Participants: 34 permanent night workers. Subjects were classified by Epworth Sleepiness Scale and Insomnia Severity Index into 3 subgroups: asymptomatic controls, alert insomniacs (AI), and sleepy insomniacs (SI). Measurements: Sleep parameters were assessed by sleep diary. Circadian phase was evaluated by dim-light salivary melatonin onset (DLMO). Objective sleepiness was measured using the multiple sleep latency test (MSLT). Brain activity was measured using the N1 event-related potential (ERP). A tandem repeat in PER3 was genotyped from saliva DNA. Results: (1) AI group showed normal MSLT scores but elevated N1 amplitudes indicating cortical hyperarousal. (2) SI group showed pathologically low MSLT scores but normal N1 amplitudes. (3) AI and SI groups were not significantly different from one another in circadian phase, while controls were significantly phase-delayed relative to both SWD groups. (4) AI showed significantly longer sleep latencies and lower sleep efficiency than controls during both nocturnal and diurnal sleep. SI significantly differed from controls in nocturnal sleep parameters, but differences during diurnal sleep periods were smaller and not statistically significant. (5) Genotype × phenotype χ2 analysis showed significant differences in the PER3 VNTR: 9 of 10 shift workers reporting sleepiness in a post hoc genetic substudy were found to carry the long tandem repeat on PER3, while 4 of 14 shift workers without excessive sleepiness carried the long allele. Conclusions: Our results suggest that the sleepy insomnia phenotype is comprehensively explained by circadian misalignment, while the alert insomnia phenotype resembles an insomnia disorder precipitated by shift work. Citation: Gumenyuk V, Belcher R, Drake CL, Roth T. Differential sleep

  15. The Sleepy Teenager – Diagnostic Challenges

    PubMed Central

    Landtblom, Anne-Marie; Engström, Maria

    2014-01-01

    The sleepy teenager puts the doctor in a, often tricky, situation where it must be decided if we deal with normal physiology or if we should suspect pathological conditions. What medical investigations are proper to consider? What differential diagnoses should be considered in the first place? And what tools do we actually have? The symptoms and problems that usually are presented at the clinical visit can be both of medical and psychosocial character – and actually they are often a mixture of both. Subsequently, the challenge to investigate the sleepy teenager often includes the examination of a complex behavioral pattern. It is important to train and develop diagnostic skills and to realize that the physiological or pathological conditions that can cause the symptoms may have different explanations. Research in sleep disorders has shown different pathological mechanisms congruent with the variations in the clinical picture. There are probably also different patterns of involved neuronal circuits although common pathways may exist. The whole picture remains to be drawn in this interesting and challenging area. PMID:25136329

  16. Effect of levothyroxine on prolonged nocturnal sleep time and excessive daytime somnolence in patients with idiopathic hypersomnia.

    PubMed

    Shinno, Hideto; Ishikawa, Ichiro; Yamanaka, Mami; Usui, Ai; Danjo, Sonoko; Inami, Yasushi; Horiguchi, Jun; Nakamura, Yu

    2011-06-01

    This study aims to examine the effect of levothyroxine, a thyroid hormone, on a prolonged nocturnal sleep and excessive daytime somnolence (EDS) in patients with idiopathic hypersomnia. In a prospective, open-label study, nine patients were enrolled. All subjects met criteria for idiopathic hypersomnia with long sleep time defined by the International Classification of Sleep Disorders, 2nd edition (ICSD-2). Subjects with sleep apnea syndrome, obesity or hypothyroidism were excluded. Sleep architecture and subjective daytime somnolence were estimated by polysomnography (PSG) and Epworth Sleepiness Scale (ESS), respectively. After baseline examinations, levothyroxine (25μg/day) was orally administered every day. Mean total sleep time, ESS score at baseline were compared with those after treatment (2, 4 and 8 weeks). Mean age of participants was 23.8±13.7 years old. At baseline, mean total sleep time (hours) and ESS score were 12.9±0.3 and 17.8±1.4, respectively. Mean total sleep times after treatment were 9.1±0.7 and 8.5±1.0h at 4 and 8 treatment weeks, respectively. Mean ESS scores were 8.8±2.3 and 7.4±2.8 at 4 and 8 treatment weeks, respectively. One patient dropped out at the 2nd week due to poor effect. No adverse effects were noted. After treatment with levothyroxine for over 4 weeks, prolonged sleep time and EDS were improved. Levothyroxine was effective for hypersomnia and well tolerated. Copyright © 2011 Elsevier B.V. All rights reserved.

  17. Association between Daytime Napping and Chronic Diseases in China.

    PubMed

    Zhou, Junmin; Kessler, Asia Sikora; Su, Dejun

    2016-03-01

    To explore the relationship between daytime napping and incidence of chronic diseases over the past 6 months among adults in China. Based on data collected from 13,469 respondents over age 40 in the Chinese Family Panel Studies in 2010, logistic regression models were estimated to examine the association between daytime napping and the incidence of any chronic diseases and 3 specific chronic diseases (hypertension, diabetes, and heart disease) after adjusting for confounders. Differences of risks by sex and age were also investigated. In the sample, 50.8% were women and 32.2% were over 60 years old. Adjusted estimates show respondents with daytime napping had elevated odds of developing any chronic diseases, hypertension, and diabetes compared to those who did not nap; having over 60 minutes of daytime napping had weaker association compared with shorter duration of daytime napping. The association between daytime napping and hypertension was found in women but not in men. Daytime napping appears to be associated with elevated risk of incidence of any chronic diseases, hypertension, and diabetes.

  18. Effects of Acute Sleep Deprivation Resulting from Night Shift Work on Young Doctors.

    PubMed

    Sanches, Inês; Teixeira, Fátima; dos Santos, José Moutinho; Ferreira, António Jorge

    2015-01-01

    To evaluate sleep deprivation and its effects on young physicians in relation to concentration capacity and psychomotor performance. Eighteen physicians aged 26 - 33 years were divided into 2 groups: non-sleep deprived group (with no night work) and sleep deprived group (minimum 12 hour of night work/week). We applied Pittsburgh Sleep Quality Index to screen the presence of sleep pathology and Epworth Sleepiness Scale to evaluate subjective daytime sleepiness; we used actigraphy and sleep diary to assess sleep hygiene and standard sleep-wake cycles. To demonstrate the effects of sleep deprivation, we applied Toulouse-Piéron's test (concentration test) and a battery of three reaction time tasks after the night duty. Sleep deprived group had higher daytime sleepiness on Epworth Sleepiness Scale (p < 0.05) and during week sleep deprivation was higher (p < 0.010). The mean duration of sleep during the period of night duty was 184.2 minutes to sleep deprived group and 397.7 minutes to non-sleep deprived group (p < 0.001). In the Toulouse-Piéron's test, the sleep deprived group had more omissions (p < 0.05) with a poorer result in concentration (p < 0.05). Psychomotor tests that evaluated response to simple stimuli revealed longer response latency (p < 0.05) and more errors (p < 0.05) in Sleep deprived group; in reaction to instruction test the sleep deprived group showed worse perfection index (p < 0.05); in the fine movements test there was no statistically significant difference between the groups. Acute sleep deprivation resulting from nocturnal work in medical professions is associated with a reduction in attention and concentration and delayed response to stimuli. This may compromise patient care as well as the physician's health and quality of life. It is essential to study the effects of acute sleep deprivation on the cognitive abilities and performance of health professionals.

  19. Road accidents caused by sleepy drivers: Update of a Norwegian survey.

    PubMed

    Phillips, Ross Owen; Sagberg, Fridulv

    2013-01-01

    The current study tests, updates and expands a model of factors associated with sleepy driving, originally based on a 1997 survey of accident-involved Norwegian drivers (Sagberg, F., 1999. Road accidents caused by drivers falling asleep. Accident Analysis & Prevention 31, 639-649). The aim is to establish a robust model to inform measures to tackle sleepy driving. The original questions on (i) tiredness-related accidents and (ii) incidents of sleep behind the wheel in the last 12 months were again posed in 2003 and 2008, in independent surveys of Norwegian drivers involved in accidents reported to a large insurance company. According to those drivers at-fault for the accident, tiredness or sleepiness behind the wheel contributed to between 1.9 and 3.9 per cent of all types of accident reported to the insurance company across these years. Accident-involved drivers not at fault for the accident reported a reduction in the incidence of sleep behind the wheel for the preceding year, decreasing from 8.3 per cent in 1997 to 2.9 per cent in 2008. The reasons for this are not clear. According to logistic regression analysis of survey responses, the following factors were robustly associated with road accidents involving sleepy driving: driving off the road; good road conditions; longer distance driven since the start of the trip; and fewer years with a driving licence. The following factors are consistently associated with reports of sleep behind the wheel, whether or not it leads to an accident: being male; driving further per year; being younger; and having sleep-related health problems. Taken together these findings suggest that young, inexperienced male drivers who drive long distances may be a suitable target for road safety campaigns aimed at tackling sleepy driving. Copyright © 2012 Elsevier Ltd. All rights reserved.

  20. [Epidemiology of Internet Use by an Adolescent Population and its Relation with Sleep Habits].

    PubMed

    Ferreira, Carla; Ferreira, Helena; Vieira, Maria João; Costeira, Mónica; Branco, Liliana; Dias, Ângela; Macedo, Liliana

    2017-08-31

    In the last decades, the great technological development increased Internet popularity, emerging the concern about its overuse. The objectives of this study were to assess and characterize Internet use in adolescence, determine Internet addiction and clarify its association with sleep disorders and excessive daytime sleepiness. It was performed an observational, cross sectional and community-based study. The target were students attending 7th and 8th grades, to whom was applied an online self-report questionnaire to assess sociodemographic features, Internet use, Internet dependence, sleep characteristics and excessive daytime sleepiness. A total of 727 adolescents were included with a mean age 13 ± 0.9 years. Three-quarters of teenagers use Internet daily and 41% do it for three or more hours/day, mainly at home. The phone and laptop were the main devices used. Online games and social networks use were the main activities performed. Internet dependence was observed in 19% of adolescents, and it was associated with male gender, social networks use, mainly Twitter and Instagram use, self-perceived sleep problems, initial and middle insomnia and excessive daytime sleepiness (p < 0.05). The results confirm the highlight that Internet has in adolescents routine, who prioritize in their use access to social networks and online games, using single devices, less subject to parental control. The Internet addiction rate observed and its association with sleep alterations and daytime sleepiness emphasizes the importance of this issue.

  1. Nocturnal snoring decreases daytime baroreceptor sensitivity.

    PubMed

    Schöbel, Christoph; Fietze, Ingo; Glos, Martin; Schary, Inett; Blau, Alexander; Baumann, Gert; Penzel, Thomas

    2014-07-01

    In patients with obstructive sleep apnea heart rate variability and baroreceptor sensitivity during night and daytime are impaired. Snoring without obstructive sleep apnea may already influence heart rate variability and baroreceptor sensitivity during daytime. Cardiovascular daytime testing was performed in 11 snorers and age, BMI, and gender matched controls. Sleep apnea and snoring were quantified by sleep recordings. Paced breathing was performed during daytime with ECG, non-invasive blood pressure, and respiration recorded. Heart rate variability and blood pressure variability were analyzed in the time and frequency domain. Baroreceptor sensitivity (alpha gain) was calculated. In snorers a significant increase in high frequency systolic blood pressure variability (SBPV-HF) compared to control group (0.37 mm Hg(2) vs. 0.11 mm Hg(2) for 12 breaths and 0.35 mm Hg(2) vs. 0.10 mm Hg(2) for 15 breaths) was demonstrated. Furthermore a lower baroreceptor sensitivity was found in snorers compared to controls (9.2 ms/mm Hg vs. 16.2 ms/mm Hg for 12 breaths and 8.5 ms/mm Hg vs. 17.4 ms/mm Hg for 15 breaths per minute) using the paced breathing protocol. Mean heart rate was elevated in snorers as well. Snorers may have a reduced parasympathetic tone during daytime rather than an increased sympathetic tone. Copyright © 2014 Elsevier Ltd. All rights reserved.

  2. [Sleep health education for elderly people].

    PubMed

    Miyazaki, Soichiro; Nishiyama, Akiko

    2015-06-01

    Successful aging is characterized by minimal age-associated loss of the physiological functions of sleep and circadian clock. Sleep health education is necessary to have normal, quality nighttime sleep and full daytime alertness. Elderly people show changes of sleep parameters, accompanied by increased napping. Many studies have reported that daytime sleepiness or napping in elderly people could have potentially serious effects such as dementia and life-style related diseases. The main topics of sleep health education for elderly people are as follows: Right knowledge of sleep mechanism, understanding the bad influence of excessive napping, the effects of light on the circadian rhythm and negative effects of caffeine, alcohol and television.

  3. Differential sleep, sleepiness, and neurophysiology in the insomnia phenotypes of shift work disorder.

    PubMed

    Gumenyuk, Valentina; Belcher, Ren; Drake, Christopher L; Roth, Thomas

    2015-01-01

    To characterize and compare insomnia symptoms within two common phenotypes of Shift Work Disorder. Observational laboratory and field study. Hospital sleep center. 34 permanent night workers. Subjects were classified by Epworth Sleepiness Scale and Insomnia Severity Index into 3 subgroups: asymptomatic controls, alert insomniacs (AI), and sleepy insomniacs (SI). Sleep parameters were assessed by sleep diary. Circadian phase was evaluated by dim-light salivary melatonin onset (DLMO). Objective sleepiness was measured using the multiple sleep latency test (MSLT). Brain activity was measured using the N1 event-related potential (ERP). A tandem repeat in PER3 was genotyped from saliva DNA. (1) AI group showed normal MSLT scores but elevated N1 amplitudes indicating cortical hyperarousal. (2) SI group showed pathologically low MSLT scores but normal N1 amplitudes. (3) AI and SI groups were not significantly different from one another in circadian phase, while controls were significantly phase-delayed relative to both SWD groups. (4) AI showed significantly longer sleep latencies and lower sleep efficiency than controls during both nocturnal and diurnal sleep. SI significantly differed from controls in nocturnal sleep parameters, but differences during diurnal sleep periods were smaller and not statistically significant. (5) Genotype × phenotype χ² analysis showed significant differences in the PER3 VNTR: 9 of 10 shift workers reporting sleepiness in a post hoc genetic substudy were found to carry the long tandem repeat on PER3, while 4 of 14 shift workers without excessive sleepiness carried the long allele. Our results suggest that the sleepy insomnia phenotype is comprehensively explained by circadian misalignment, while the alert insomnia phenotype resembles an insomnia disorder precipitated by shift work. © 2014 Associated Professional Sleep Societies, LLC.

  4. How to Alleviate Jet Lag/The Chronobiotic Substances (Comment reduire les effects du decalage horaire: les substances chronobioktiques)

    DTIC Science & Technology

    2002-11-01

    physical performance after a transmeridian flight. Medicine and Science in Sports and Exercise 2001; 33(4):628-634. 43. Reid K, Dawson D. Correlation...hour round-the-clock capability so that they need a high level of performances overnight and a good quality of sleep during short rest periods before...results in sleep disturbances, daytime sleepiness and performance impairment that tend to increase the hazard of mission failure and of casualties

  5. Internet addiction, sleep and health-related life quality among obese individuals: a comparison study of the growing problems in adolescent health.

    PubMed

    Eliacik, Kayi; Bolat, Nurullah; Koçyiğit, Cemil; Kanik, Ali; Selkie, Ellen; Yilmaz, Huseyin; Catli, Gonul; Dundar, Nihal Olgac; Dundar, Bumin Nuri

    2016-12-01

    The rapid rise in the global prevalence of obesity suggests that environmental factors may be responsible. The increased use of technology is associated with increased rates of obesity due to declines in physical activity and significant sedentary life style. Internet addiction is also a growing health issue associated with diminished physical activity and poor sleep quality as well as various health problems. The purpose of this study was to determine associations between Internet addiction and adolescent obesity-related problems. In this case-control study, 71 adolescents with obesity were recruited from the outpatient clinic at Tepecik Teaching Hospital and Katip Celebi University Hospital, Department of Pediatric Endocrinology in Izmir, Turkey. The control group consisted of 64 non-obese adolescents that were matched with patients in the study group by age and gender. All subjects completed socio-demographic forms, an Internet addiction scale, the Pediatric Quality of Life Inventory, the Pittsburgh Sleep Quality Index, and the Epworth Sleepiness Scale. Adolescents with obesity were significantly more likely to have Internet addiction (p = 0.002), lower quality of life (p < 0.001), and higher daytime sleepiness (p = 0.008). Moreover, binary regression analysis showed that Internet addiction and less physical activity were associated with increased odds of obesity. The results indicated a significant association between Internet addiction and obesity. Health practitioners should take possible Internet addiction, online activities, and physical activities into consideration in follow-up of obese adolescents. In addition to pharmacologic therapies and dietary interventions, providing behavioral therapy targeting healthy Internet use may be promising to reduce the effects of obesity in adolescence.

  6. Formative Assessment Probes: The Daytime Moon

    ERIC Educational Resources Information Center

    Keeley, Page

    2012-01-01

    The familiar adage "seeing is believing" implies that children will recall a particular phenomenon if they had the experience of seeing it with their own eyes. If this were true, then most children would believe that one could see the Moon in both daytime and at night. However, when children are asked, "Can you see the Moon in the daytime?" many…

  7. A Daytime Aspect Camera for Balloon Altitudes

    NASA Technical Reports Server (NTRS)

    Dietz, Kurt L.; Ramsey, Brian D.; Alexander, Cheryl D.; Apple, Jeff A.; Ghosh, Kajal K.; Swift, Wesley R.; Six, N. Frank (Technical Monitor)

    2001-01-01

    We have designed, built, and flight-tested a new star camera for daytime guiding of pointed balloon-borne experiments at altitudes around 40km. The camera and lens are commercially available, off-the-shelf components, but require a custom-built baffle to reduce stray light, especially near the sunlit limb of the balloon. This new camera, which operates in the 600-1000 nm region of the spectrum, successfully provided daytime aspect information of approximately 10 arcsecond resolution for two distinct star fields near the galactic plane. The detected scattered-light backgrounds show good agreement with the Air Force MODTRAN models, but the daytime stellar magnitude limit was lower than expected due to dispersion of red light by the lens. Replacing the commercial lens with a custom-built lens should allow the system to track stars in any arbitrary area of the sky during the daytime.

  8. The Defense Committees of Sleepy Lagoon: A Convergent Struggle against Fascism, 1942-1944

    ERIC Educational Resources Information Center

    Barajas, Frank P.

    2006-01-01

    The Sleepy Lagoon Defense Committee originated as an ad hoc committee and evolved to a broad-based movement for legal justice on behalf of seventeen youth convicted of murder and assault charges in connection with the Sleepy Lagoon case in Los Angeles in January 1943. This essay chronicles the multidimensional organizing to shift public opinion in…

  9. Hypersomnia and depressive symptoms: methodological and clinical aspects

    PubMed Central

    2013-01-01

    The associations between depressive symptoms and hypersomnia are complex and often bidirectional. Of the many disorders associated with excessive sleepiness in the general population, the most frequent are mental health disorders, particularly depression. However, most mood disorder studies addressing hypersomnia have assessed daytime sleepiness using a single response, neglecting critical and clinically relevant information about symptom severity, duration and nighttime sleep quality. Only a few studies have used objective tools such as polysomnography to directly measure both daytime and nighttime sleep propensity in depression with normal mean sleep latency and sleep duration. Hypersomnia in mood disorders, rather than a medical condition per se, is more a subjective sleep complaint than an objective finding. Mood symptoms have also been frequently reported in hypersomnia disorders of central origin, especially in narcolepsy. Hypocretin deficiency could be a contributing factor in this condition. Further interventional studies are needed to explore whether management of sleep complaints improves mood symptoms in hypersomnia disorders and, conversely, whether management of mood complaints improves sleep symptoms in mood disorders. PMID:23514569

  10. The Differential Effects of Regular Shift Work and Obstructive Sleep Apnea on Sleepiness, Mood and Neurocognitive Function.

    PubMed

    Cori, Jennifer M; Jackson, Melinda L; Barnes, Maree; Westlake, Justine; Emerson, Paul; Lee, Jacen; Galante, Rosa; Hayley, Amie; Wilsmore, Nicholas; Kennedy, Gerard A; Howard, Mark

    2018-06-15

    To assess whether poor sleep quality experienced by regular shift workers and individuals with obstructive sleep apnea (OSA) affects neurobehavioral function similarly, or whether the different etiologies have distinct patterns of impairment. Thirty-seven shift workers (> 24 hours after their last shift), 36 untreated patients with OSA, and 39 healthy controls underwent assessment of sleepiness (Epworth Sleepiness Scale [ESS]), mood (Beck Depression Index, State Trait Anxiety Inventory [STAI], Profile of Mood States), vigilance (Psychomotor Vigilance Task [PVT], Oxford Sleep Resistance Test [OSLER], driving simulation), neurocognitive function (Logical Memory, Trails Making Task, Digit Span Task, Victoria Stroop Test) and polysomnography. Sleepiness (ESS score; median, interquartile range) did not differ between the OSA (10.5, 6.3-14) and shift work (7, 5-11.5) groups, but both had significantly elevated scores relative to the control group (5, 3-6). State anxiety (STAI-S) was the only mood variable that differed significantly between the OSA (35, 29-43) and shift work (30, 24-33.5) groups, however both demonstrated several mood deficits relative to the control group. The shift work and control groups performed similarly on neurobehavioral tasks (simulated driving, PVT, OSLER and neurocognitive tests), whereas the OSA group performed worse. On the PVT, lapses were significantly greater for the OSA group (3, 2-6) than both the shift work (2, 0-3.5) and control (1, 0-4) groups. Shift workers and patients with OSA had similar sleepiness and mood deficits relative to healthy individuals. However, only the patients with OSA showed deficits on vigilance and neurocognitive function relative to healthy individuals. These findings suggest that distinct causes of sleep disturbance likely result in different patterns of neurobehavioral dysfunction. © 2018 American Academy of Sleep Medicine.

  11. Participation in nighttime activities in the genesis of depression in public school teachers from the State of Pernambuco, Brazil.

    PubMed

    Correia, Francisca Maria da Silva; de Albuquerque, Rosângela Nieto; Martins, Hugo André de Lima; Lins, Luciano da Fonseca; Lima, Murilo Duarte Costa; Dias, José Marcos da Silva; da Silva, Cícera Maria; Dos Santos, Allison José; Dos Santos, Leonardo Tárcito; Ribas, Valdenilson Ribeiro

    2012-01-01

    Teachers often undertake nighttime work involving exam corrections, projects and devising lesson plans in their homes. Many present excessive daytime sleepiness (EDS) and depression. The objective of this study was to evaluate EDS and depression in teachers from public schools. 201 female teachers were evaluated in the district of Quipapá/PE, Brazil. Among the study sample, 38 working 1 shift (CONTROL 1), 40 working 2 shifts (CONTROL 2) and 123 working 3 shifts (WTeachers-3T). The subjects were submitted to evaluation by the Epworth Sleepiness Scale and Beck Depression Inventory (BDI).The EDS data were analyzed by the Kruskal-Wallis test with Dunn's multiple comparison, p<0.05and expressed in MEDIAN (MINIMUM - MAXIMUM) whereas the depression data were analyzed by the Chi-square test, with p<0.05, expressed in percentage. WTeachers-3T presented excessive daytime sleepiness and higher rates of mild (24%) and moderate (37%) depression compared to controls - Control 1: mild (8%) and moderate (11%) - Control 2: mild (5%) and moderate (15%). This study found that teachers in the Quipapá municipality of Penambuco state working three shifts showed excessive daytime sleepiness and a higher percentage of mild and moderate depression compared to teachers working only one (1) or two (2) shifts.

  12. Impact of sleep deprivation and obstructive sleep apnea syndrome on daytime vigilance and driving performance: a laboratory perspective.

    PubMed

    Pizza, F; Contardi, S; Mondini, S; Cirignotta, F

    2012-01-01

    To study the impact of sleepiness, a well-established cause of car accidents, on driving ability, we designed a 30-min monotonous simulated driving task. Our simulated driving task encompasses both primary vehicle control (standard deviation of lane position, crash occurrence) and secondary tasks (type and reaction times to divided attention tasks). Driving simulator data were correlated to subjective (state/trait) and objective (MSLT/MWT) sleepiness measures in healthy subjects undergoing sleep deprivation (SD) and in obstructive sleep apnea (OSAS) patients. SD induced severe sleepiness during nighttime, when state sleepiness increased while primary vehicle control ability worsened. After SD, driving ability decreased and was inversely correlated to subjective and objective sleepiness at MSLT. OSAS patients driving ability was well correlated to objective sleepiness, with inverse correlation to sleep propensity at the MSLT and even more strict relation with the ability to maintain wakefulness at the MWT. Sleepiness worsens driving ability in healthy subjects after SD and in OSAS patients. Driving ability correlates with subjective and objective sleepiness measures, in particular to the ability to maintain wakefulness.

  13. 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.

  14. Psychometric properties of the Epworth Sleepiness Scale: A factor analysis and item-response theory approach.

    PubMed

    Pilcher, June J; Switzer, Fred S; Munc, Alec; Donnelly, Janet; Jellen, Julia C; Lamm, Claus

    2018-04-01

    The purpose of this study is to examine the psychometric properties of the Epworth Sleepiness Scale (ESS) in two languages, German and English. Students from a university in Austria (N = 292; 55 males; mean age = 18.71 ± 1.71 years; 237 females; mean age = 18.24 ± 0.88 years) and a university in the US (N = 329; 128 males; mean age = 18.71 ± 0.88 years; 201 females; mean age = 21.59 ± 2.27 years) completed the ESS. An exploratory-factor analysis was completed to examine dimensionality of the ESS. Item response theory (IRT) analyses were used to provide information about the response rates on the items on the ESS and provide differential item functioning (DIF) analyses to examine whether the items were interpreted differently between the two languages. The factor analyses suggest that the ESS measures two distinct sleepiness constructs. These constructs indicate that the ESS is probing sleepiness in settings requiring active versus passive responding. The IRT analyses found that overall, the items on the ESS perform well as a measure of sleepiness. However, Item 8 and to a lesser extent Item 6 were being interpreted differently by respondents in comparison to the other items. In addition, the DIF analyses showed that the responses between German and English were very similar indicating that there are only minor measurement differences between the two language versions of the ESS. These findings suggest that the ESS provides a reliable measure of propensity to sleepiness; however, it does convey a two-factor approach to sleepiness. Researchers and clinicians can use the German and English versions of the ESS but may wish to exclude Item 8 when calculating a total sleepiness score.

  15. Effects of tryptophan-rich breakfast and light exposure during the daytime on melatonin secretion at night.

    PubMed

    Fukushige, Haruna; Fukuda, Yumi; Tanaka, Mizuho; Inami, Kaoru; Wada, Kai; Tsumura, Yuki; Kondo, Masayuki; Harada, Tetsuo; Wakamura, Tomoko; Morita, Takeshi

    2014-11-19

    The purpose of the present study is to investigate effects of tryptophan intake and light exposure on melatonin secretion and sleep by modifying tryptophan ingestion at breakfast and light exposure during the daytime, and measuring sleep quality (by using actigraphy and the OSA sleep inventory) and melatonin secretion at night. Thirty three male University students (mean ± SD age: 22 ± 3.1 years) completed the experiments lasting 5 days and 4 nights. The subjects were randomly divided into four groups: Poor*Dim (n = 10), meaning a tryptophan-poor breakfast (55 mg/meal) in the morning and dim light environment (<50 lx) during the daytime; Rich*Dim (n = 7), tryptophan-rich breakfast (476 mg/meal) and dim light environment; Poor*Bright (n = 9), tryptophan-poor breakfast and bright light environment (>5,000 lx); and Rich*Bright (n = 7), tryptophan-rich breakfast and bright light. Saliva melatonin concentrations on the fourth day were significantly lower than on the first day in the Poor*Dim group, whereas they were higher on the fourth day in the Rich*Bright group. Creatinine-adjusted melatonin in urine showed the same direction as saliva melatonin concentrations. These results indicate that the combination of a tryptophan-rich breakfast and bright light exposure during the daytime could promote melatonin secretion at night; further, the observations that the Rich*Bright group had higher melatonin concentrations than the Rich*Dim group, despite no significant differences being observed between the Poor*Dim and Rich*Dim groups nor the Poor*Bright and Rich*Bright groups, suggest that bright light exposure in the daytime is an important contributor to raised melatonin levels in the evening. This study is the first to report the quantitative effects of changed tryptophan intake at breakfast combined with daytime light exposure on melatonin secretion and sleep quality. Evening saliva melatonin secretion changed significantly and indicated that a tryptophan

  16. Sleep Disturbances among Pregnant Women with History of Migraines: a Cross-sectional Study

    PubMed Central

    Qiu, Chunfang; Frederick, Ihunnaya O.; Sorensen, Tanya; Aurora, Sheena K.; Gelaye, Bizu; Enquobahrie, Daniel A.; Williams, Michelle A.

    2015-01-01

    Background Migraine is associated with sleep disturbances in men and non-pregnant women. However, relatively little is known about sleep disturbances among pregnant migraineurs. We investigated sleep disturbances among pregnant women with and without history of migraine. Methods This cross-sectional study was conducted among 1,324 women who were recruited during early pregnancy. Migraine diagnoses were based on the International Classification of Headache Disorders-II criteria. Pittsburgh Sleep Quality Index (PSQI) questionnaire was used to evaluate sleep-related characteristics including sleep duration, sleep quality, excessive daytime sleepiness, and other sleep traits. Multivariable logistic regression procedures were used to estimate adjusted odds ratios (AORs) and 95% confidence intervals (CIs). Results Migraineurs were more likely than non-migraineurs to report short sleep duration (≤6 hours) (AOR=1.47, 95% CI 1.07–2.02), poor sleep quality (PSQI>5) (AOR=1.73, 95% CI 1.35–2.23), and daytime dysfunction due to sleepiness (AOR=1.51, 95% CI 1.12–2.02). Migraineurs were also more likely than non-migraineurs to report taking sleep medication during pregnancy (AOR=1.71, 95% CI 1.20–2.42). Associations were generally similar for migraine with or without aura. The odds of sleep disturbances were particularly elevated among pre-pregnancy overweight migraineurs. Conclusion Migraine headache and sleep disturbances are common co-morbid conditions among pregnant women. PMID:25633375

  17. Sleep disturbances among pregnant women with history of migraines: A cross-sectional study.

    PubMed

    Qiu, Chunfang; Frederick, Ihunnaya O; Sorensen, Tanya; Aurora, Sheena K; Gelaye, Bizu; Enquobahrie, Daniel A; Williams, Michelle A

    2015-10-01

    Migraine is associated with sleep disturbances in men and non-pregnant women. However, relatively little is known about sleep disturbances among pregnant migraineurs. We investigated sleep disturbances among pregnant women with and without history of migraine. This cross-sectional study was conducted among 1324 women who were recruited during early pregnancy. Migraine diagnoses were based on the International Classification of Headache Disorders-II criteria. The Pittsburgh Sleep Quality Index (PSQI) questionnaire was used to evaluate sleep-related characteristics including sleep duration, sleep quality, excessive daytime sleepiness, and other sleep traits. Multivariable logistic regression procedures were used to estimate adjusted odds ratios (AORs) and 95% confidence intervals (CIs). Migraineurs were more likely than non-migraineurs to report short sleep duration (<6.5 hours) (AOR = 1.47, 95% CI 1.07-2.02), poor sleep quality (PSQI>5) (AOR = 1.73, 95% CI 1.35-2.23), and daytime dysfunction due to sleepiness (AOR = 1.51, 95% CI 1.12-2.02). Migraineurs were also more likely than non-migraineurs to report taking sleep medication during pregnancy (AOR = 1.71, 95% CI 1.20-2.42). Associations were generally similar for migraine with or without aura. The odds of sleep disturbances were particularly elevated among pre-pregnancy overweight migraineurs. Migraine headache and sleep disturbances are common comorbid conditions among pregnant women. © International Headache Society 2015.

  18. Modifiable correlates of perceived cognitive function in breast cancer survivors up to 10 years after chemotherapy completion.

    PubMed

    Henneghan, Ashley; Stuifbergen, Alexa; Becker, Heather; Kesler, Shelli; King, Elisabeth

    2018-04-01

    Cognitive changes following breast cancer treatment are likely multifactorial and have been linked to emotional factors, biophysiological factors, and fatigue, among others. Little is known about the contributions of modifiable factors such as stress, loneliness, and sleep quality. The purpose of this study was to explore the direct and indirect effects of perceived stress, loneliness, and sleep quality on perceived cognitive function (PCF) in breast cancer survivors (BCS) after chemotherapy completion. In this observational study, BCS 6 months to 10 years post chemotherapy were recruited from the community. We measured perceived stress, loneliness, sleep quality, anxiety, depression, fatigue, and PCF. Data analyses included descriptive statistics, correlations, and mediation analyses utilizing ordinary least square regression. Ninety women who were on average 3 years post chemotherapy completion participated in the study. Moderate to largely negative correlations were found between PCF and the psychosocial and sleep variables (r values ranged from - 0.31 to - 0.70, p values < .0009). Mediation analyses revealed that stress and daytime sleepiness both directly and indirectly impact PCF and that loneliness and sleep quality only have indirect effects (through anxiety and fatigue). Our findings suggest that perceived cognitive changes following breast cancer treatment are multifactorial and that higher stress levels, loneliness, daytime sleepiness, and poorer sleep quality are linked to worse perceived cognitive functioning. Also, stress, loneliness, and sleep quality may affect cognitive functioning through a shared psychobiological pathway. Interventions targeting stress, loneliness, and sleep quality may improve perceived cognitive functioning in breast cancer survivors.

  19. A Sleep Questionnaire for Children with Severe Psychomotor Impairment (SNAKE)-Concordance with a Global Rating of Sleep Quality.

    PubMed

    Dreier, Larissa Alice; Zernikow, Boris; Blankenburg, Markus; Wager, Julia

    2018-02-01

    Sleep problems are a common and serious issue in children with life-limiting conditions (LLCs) and severe psychomotor impairment (SPMI). The "Sleep Questionnaire for Children with Severe Psychomotor Impairment" (Schlaffragebogen für Kinder mit Neurologischen und Anderen Komplexen Erkrankungen, SNAKE) was developed for this unique patient group. In a proxy rating, the SNAKE assesses five different dimensions of sleep(-associated) problems (disturbances going to sleep, disturbances remaining asleep, arousal and breathing disorders, daytime sleepiness, and daytime behavior disorders). It has been tested with respect to construct validity and some aspects of criterion validity. The present study examined whether the five SNAKE scales are consistent with parents' or other caregivers' global ratings of a child's sleep quality. Data from a comprehensive dataset of children and adolescents with LLCs and SPMI were analyzed through correlation coefficients and Mann-Whitney U testing. The results confirmed the consistency of both sources of information. The highest levels of agreements with the global rating were achieved for disturbances in terms of going to sleep and disturbances with respect to remaining asleep. The results demonstrate that the scales and therefore the SNAKE itself is well-suited for gathering information on different sleep(-associated) problems in this vulnerable population.

  20. Daytime symptoms of restless legs syndrome--clinical characteristics and rotigotine effectiveness.

    PubMed

    Takahashi, Masayoshi; Ikeda, Junji; Tomida, Takayuki; Hirata, Koichi; Hattori, Nobutaka; Inoue, Yuichi

    2015-07-01

    To elucidate the prevalence and clinical characteristics of daytime restless legs syndrome (RLS) among patients with idiopathic RLS and investigate the effectiveness of rotigotine for daytime RLS. In 256 enrolled RLS patients, we investigated factors associated with the presence of RLS symptoms throughout the day. We also assessed the duration of daytime RLS symptoms at hourly intervals, time of initial symptom onset during the day, and associations between duration of daytime and nighttime RLS symptoms. In addition, we compared changes in duration and frequency of RLS symptoms during daytime and nighttime after randomly assigning patients to a 13-week treatment with rotigotine, a dopamine agonist patch with 24-hour action, or placebo. Eighty-one (31.6%) patients had daytime RLS symptoms. Only the International Restless Legs Syndrome Study Group rating scale total score was significantly associated with the presence of daytime RLS symptoms (p < 0.01) on multiple logistic regression analysis. Daytime RLS symptom onset was at 6 a.m. in 44.4% of patients; symptom duration increased significantly toward nighttime. There was a significant positive association between duration of daytime and nighttime RLS symptoms (p < 0.0001) and a greater statistically significant reduction of daytime RLS symptom duration with rotigotine treatment than with placebo (p = 0.03). Daytime symptoms are frequent in patients with RLS and may be associated with increased severity of the disorder and prolonged nighttime RLS symptoms. Rotigotine could become an important treatment choice for daytime symptoms. Copyright © 2015 Elsevier B.V. All rights reserved.