Sample records for wait time strategy

  1. Sustainability: orthopaedic surgery wait time management strategies.

    PubMed

    Amar, Claudia; Pomey, Marie-Pascale; SanMartin, Claudia; De Coster, Carolyn; Noseworthy, Tom

    2015-01-01

    The purpose of this paper is to examine Canadian organizational and systemic factors that made it possible to keep wait times within federally established limits for at least 18 months. The research design is a multiple cases study. The paper selected three cases: Case 1 - staff were able to maintain compliance with requirements for more than 18 months; Case 2 - staff were able to meet requirements for 18 months, but unable to sustain this level; Case 3 - staff were never able to meet the requirements. For each case the authors interviewed persons involved in the strategies and collected documents. The paper analysed systemic and organizational-level factors; including governance and leadership, culture, resources, methods and tools. Findings indicate that the hospital that was able to maintain compliance with the wait time requirements had specific characteristics: an exclusive mandate to do only hip and knee replacement surgery; motivated staff who were not distracted by other concerns; and a strong team spirit. The authors' research highlights an important gradient between three cases regarding the factors that sustain waiting times. The paper show that the hospital factory model seems attractive in a super-specialized surgery context. However, patients are selected for simple surgeries, without complications, and so this cannot be considered a unique model.

  2. Wait time management strategies for total joint replacement surgery: sustainability and unintended consequences.

    PubMed

    Pomey, Marie-Pascale; Clavel, Nathalie; Amar, Claudia; Sabogale-Olarte, Juan Carlos; Sanmartin, Claudia; De Coster, Carolyn; Noseworthy, Tom

    2017-09-07

    In Canada, long waiting times for core specialized services have consistently been identified as a key barrier to access. Governments and organizations have responded with strategies for better access management, notably for total joint replacement (TJR) of the hip and knee. While wait time management strategies (WTMS) are promising, the factors which influence their sustainable implementation at the organizational level are understudied. Consequently, this study examined organizational and systemic factors that made it possible to sustain waiting times for TJR within federally established limits and for at least 18 months or more. The research design is a multiple case study of WTMS implementation. Five cases were selected across five Canadian provinces. Three success levels were pre-defined: 1) the WTMS maintained compliance with requirements for more than 18 months; 2) the WTMS met requirements for 18 months but could not sustain the level thereafter; 3) the WTMS never met requirements. For each case, we collected documents and interviewed key informants. We analyzed systemic and organizational factors, with particular attention to governance and leadership, culture, resources, methods, and tools. We found that successful organizations had specific characteristics: 1) management of the whole care continuum, 2) strong clinical leadership; 3) dedicated committees to coordinate and sustain strategy; 4) a culture based on trust and innovation. All strategies led to relatively similar unintended consequences. The main negative consequence was an initial increase in waiting times for TJR and the main positive consequence was operational enhancement of other areas of specialization based on the TJR model. This study highlights important differences in factors which help to achieve and sustain waiting times. To be sustainable, a WTMS needs to generate greater synergies between contextual-level strategy (provincial or regional) and organizational objectives and

  3. Public views on a wait time management initiative: a matter of communication.

    PubMed

    Bruni, Rebecca A; Laupacis, Andreas; Levinson, Wendy; Martin, Douglas K

    2010-08-05

    Many countries have tried to reduce waiting times for health care through formal wait time reduction strategies. Our paper describes views of members of the public about a wait time management initiative--the Ontario Wait Time Strategy (OWTS) (Canada). Scholars and governmental reports have advocated for increased public involvement in wait time management. We provide empirically derived recommendations for public engagement in a wait time management initiative. Two qualitative studies: 1) an analysis of all emails sent by the public to the (OWTS) email address; and 2) in-depth interviews with members of the Ontario public. Email correspondents and interview participants supported the intent of the OWTS. However they wanted more information about the Strategy and its actions. Interview participants did not feel they were sufficiently made aware of the Strategy and email correspondents requested additional information beyond what was offered on the Strategy's website. Moreover, the email correspondents believed that some of the information that was provided on the Strategy's website and through the media was inaccurate, misleading, and even dishonest. Interview participants strongly supported public involvement in the OWTS priority setting. Findings suggest the public wanted increased communication from and with the OWTS. Effective communication can facilitate successful public engagement, and in turn fair and legitimate priority setting. Based on the study's findings we developed concrete recommendations for improving public involvement in wait time management.

  4. Public views on a wait time management initiative: a matter of communication

    PubMed Central

    2010-01-01

    Background Many countries have tried to reduce waiting times for health care through formal wait time reduction strategies. Our paper describes views of members of the public about a wait time management initiative - the Ontario Wait Time Strategy (OWTS) (Canada). Scholars and governmental reports have advocated for increased public involvement in wait time management. We provide empirically derived recommendations for public engagement in a wait time management initiative. Methods Two qualitative studies: 1) an analysis of all emails sent by the public to the (OWTS) email address; and 2) in-depth interviews with members of the Ontario public. Results Email correspondents and interview participants supported the intent of the OWTS. However they wanted more information about the Strategy and its actions. Interview participants did not feel they were sufficiently made aware of the Strategy and email correspondents requested additional information beyond what was offered on the Strategy's website. Moreover, the email correspondents believed that some of the information that was provided on the Strategy's website and through the media was inaccurate, misleading, and even dishonest. Interview participants strongly supported public involvement in the OWTS priority setting. Conclusions Findings suggest the public wanted increased communication from and with the OWTS. Effective communication can facilitate successful public engagement, and in turn fair and legitimate priority setting. Based on the study's findings we developed concrete recommendations for improving public involvement in wait time management. PMID:20687952

  5. Advertising emergency department wait times.

    PubMed

    Weiner, Scott G

    2013-03-01

    Advertising emergency department (ED) wait times has become a common practice in the United States. Proponents of this practice state that it is a powerful marketing strategy that can help steer patients to the ED. Opponents worry about the risk to the public health that arises from a patient with an emergent condition self-triaging to a further hospital, problems with inaccuracy and lack of standard definition of the reported time, and directing lower acuity patients to the higher cost ED setting instead to primary care. Three sample cases demonstrating the pitfalls of advertising ED wait times are discussed. Given the lack of rigorous evidence supporting the practice and potential adverse effects to the public health, caution about its use is advised.

  6. Advertising Emergency Department Wait Times

    PubMed Central

    Weiner, Scott G.

    2013-01-01

    Advertising emergency department (ED) wait times has become a common practice in the United States. Proponents of this practice state that it is a powerful marketing strategy that can help steer patients to the ED. Opponents worry about the risk to the public health that arises from a patient with an emergent condition self-triaging to a further hospital, problems with inaccuracy and lack of standard definition of the reported time, and directing lower acuity patients to the higher cost ED setting instead to primary care. Three sample cases demonstrating the pitfalls of advertising ED wait times are discussed. Given the lack of rigorous evidence supporting the practice and potential adverse effects to the public health, caution about its use is advised. PMID:23599836

  7. Two Effective Ways to Implement Wait Time. A Symposium on Wait Time.

    ERIC Educational Resources Information Center

    Swift, J. Nathan; And Others

    The effects of instructional guides and a wait time feedback device (called a "Wait Timer") on the classroom interaction of middle school science teachers are examined. The Wait Timer, an unobtrusive indicator of wait time, is an automatic device that activates a light when a person speaks. The duration of the light at the end of a…

  8. Public involvement in the priority setting activities of a wait time management initiative: a qualitative case study.

    PubMed

    Bruni, Rebecca A; Laupacis, Andreas; Levinson, Wendy; Martin, Douglas K

    2007-11-16

    As no health system can afford to provide all possible services and treatments for the people it serves, each system must set priorities. Priority setting decision makers are increasingly involving the public in policy making. This study focuses on public engagement in a key priority setting context that plagues every health system around the world: wait list management. The purpose of this study is to describe and evaluate priority setting for the Ontario Wait Time Strategy, with special attention to public engagement. This study was conducted at the Ontario Wait Time Strategy in Ontario, Canada which is part of a Federal-Territorial-Provincial initiative to improve access and reduce wait times in five areas: cancer, cardiac, sight restoration, joint replacements, and diagnostic imaging. There were two sources of data: (1) over 25 documents (e.g. strategic planning reports, public updates), and (2) 28 one-on-one interviews with informants (e.g. OWTS participants, MOHLTC representatives, clinicians, patient advocates). Analysis used a modified thematic technique in three phases: open coding, axial coding, and evaluation. The Ontario Wait Time Strategy partially meets the four conditions of 'accountability for reasonableness'. The public was not directly involved in the priority setting activities of the Ontario Wait Time Strategy. Study participants identified both benefits (supporting the initiative, experts of the lived experience, a publicly funded system and sustainability of the healthcare system) and concerns (personal biases, lack of interest to be involved, time constraints, and level of technicality) for public involvement in the Ontario Wait Time Strategy. Additionally, the participants identified concern for the consequences (sustainability, cannibalism, and a class system) resulting from the Ontario Wait Times Strategy. We described and evaluated a wait time management initiative (the Ontario Wait Time Strategy) with special attention to public

  9. Public involvement in the priority setting activities of a wait time management initiative: a qualitative case study

    PubMed Central

    Bruni, Rebecca A; Laupacis, Andreas; Levinson, Wendy; Martin, Douglas K

    2007-01-01

    Background As no health system can afford to provide all possible services and treatments for the people it serves, each system must set priorities. Priority setting decision makers are increasingly involving the public in policy making. This study focuses on public engagement in a key priority setting context that plagues every health system around the world: wait list management. The purpose of this study is to describe and evaluate priority setting for the Ontario Wait Time Strategy, with special attention to public engagement. Methods This study was conducted at the Ontario Wait Time Strategy in Ontario, Canada which is part of a Federal-Territorial-Provincial initiative to improve access and reduce wait times in five areas: cancer, cardiac, sight restoration, joint replacements, and diagnostic imaging. There were two sources of data: (1) over 25 documents (e.g. strategic planning reports, public updates), and (2) 28 one-on-one interviews with informants (e.g. OWTS participants, MOHLTC representatives, clinicians, patient advocates). Analysis used a modified thematic technique in three phases: open coding, axial coding, and evaluation. Results The Ontario Wait Time Strategy partially meets the four conditions of 'accountability for reasonableness'. The public was not directly involved in the priority setting activities of the Ontario Wait Time Strategy. Study participants identified both benefits (supporting the initiative, experts of the lived experience, a publicly funded system and sustainability of the healthcare system) and concerns (personal biases, lack of interest to be involved, time constraints, and level of technicality) for public involvement in the Ontario Wait Time Strategy. Additionally, the participants identified concern for the consequences (sustainability, cannibalism, and a class system) resulting from the Ontario Wait Times Strategy. Conclusion We described and evaluated a wait time management initiative (the Ontario Wait Time Strategy

  10. 46 CFR 9.10 - Waiting time.

    Code of Federal Regulations, 2010 CFR

    2010-10-01

    ... 46 Shipping 1 2010-10-01 2010-10-01 false Waiting time. 9.10 Section 9.10 Shipping COAST GUARD... § 9.10 Waiting time. The same construction should be given the act when charging for waiting time as... for duty the waiting time amounts to at least one hour. ...

  11. 46 CFR 9.10 - Waiting time.

    Code of Federal Regulations, 2012 CFR

    2012-10-01

    ... 46 Shipping 1 2012-10-01 2012-10-01 false Waiting time. 9.10 Section 9.10 Shipping COAST GUARD... § 9.10 Waiting time. The same construction should be given the act when charging for waiting time as... for duty the waiting time amounts to at least one hour. ...

  12. 46 CFR 9.10 - Waiting time.

    Code of Federal Regulations, 2011 CFR

    2011-10-01

    ... 46 Shipping 1 2011-10-01 2011-10-01 false Waiting time. 9.10 Section 9.10 Shipping COAST GUARD... § 9.10 Waiting time. The same construction should be given the act when charging for waiting time as... for duty the waiting time amounts to at least one hour. ...

  13. Patient Satisfaction Is Associated With Time With Provider But Not Clinic Wait Time Among Orthopedic Patients.

    PubMed

    Patterson, Brendan M; Eskildsen, Scott M; Clement, R Carter; Lin, Feng-Chang; Olcott, Christopher W; Del Gaizo, Daniel J; Tennant, Joshua N

    2017-01-01

    Clinic wait time is considered an important predictor of patient satisfaction. The goal of this study was to determine whether patient satisfaction among orthopedic patients is associated with clinic wait time and time with the provider. The authors prospectively enrolled 182 patients at their outpatient orthopedic clinic. Clinic wait time was defined as the time between patient check-in and being seen by the surgeon. Time spent with the provider was defined as the total time the patient spent in the examination room with the surgeon. The Consumer Assessment of Healthcare Providers and Systems survey was used to measure patient satisfaction. Factors associated with increased patient satisfaction included patient age and increased time with the surgeon (P=.024 and P=.037, respectively), but not clinic wait time (P=.625). Perceived wait time was subject to a high level of error, and most patients did not accurately report whether they had been waiting longer than 15 minutes to see a provider until they had waited at least 60 minutes (P=.007). If the results of the current study are generalizable, time with the surgeon is associated with patient satisfaction in orthopedic clinics, but wait time is not. Further, the study findings showed that patients in this setting did not have an accurate perception of actual wait time, with many patients underestimating the time they waited to see a provider. Thus, a potential strategy for improving patient satisfaction is to spend more time with each patient, even at the expense of increased wait time. [Orthopedics. 2017; 40(1):43-48.]. Copyright 2016, SLACK Incorporated.

  14. Can We Predict Patient Wait Time?

    PubMed

    Pianykh, Oleg S; Rosenthal, Daniel I

    2015-10-01

    The importance of patient wait-time management and predictability can hardly be overestimated: For most hospitals, it is the patient queues that drive and define every bit of clinical workflow. The objective of this work was to study the predictability of patient wait time and identify its most influential predictors. To solve this problem, we developed a comprehensive list of 25 wait-related parameters, suggested in earlier work and observed in our own experiments. All parameters were chosen as derivable from a typical Hospital Information System dataset. The parameters were fed into several time-predicting models, and the best parameter subsets, discovered through exhaustive model search, were applied to a large sample of actual patient wait data. We were able to discover the most efficient wait-time prediction factors and models, such as the line-size models introduced in this work. Moreover, these models proved to be equally accurate and computationally efficient. Finally, the selected models were implemented in our patient waiting areas, displaying predicted wait times on the monitors located at the front desks. The limitations of these models are also discussed. Optimal regression models based on wait-line sizes can provide accurate and efficient predictions for patient wait time. Copyright © 2015 American College of Radiology. Published by Elsevier Inc. All rights reserved.

  15. Space, place and (waiting) time: reflections on health policy and politics.

    PubMed

    Sheard, Sally

    2018-02-19

    Health systems have repeatedly addressed concerns about efficiency and equity by employing trans-national comparisons to draw out the strengths and weaknesses of specific policy initiatives. This paper demonstrates the potential for explicit historical analysis of waiting times for hospital treatment to add value to spatial comparative methodologies. Waiting times and the size of the lists of waiting patients have become key operational indicators. In the United Kingdom, as National Health Service (NHS) financial pressures intensified from the 1970s, waiting times have become a topic for regular public and political debate. Various explanations for waiting times include the following: hospital consultants manipulate NHS waiting lists to maintain their private practice; there is under-investment in the NHS; and available (and adequate) resources are being used inefficiently. Other countries have also experienced ongoing tensions between the public and private delivery of universal health care in which national and trans-national comparisons of waiting times have been regularly used. The paper discusses the development of key UK policies, and provides a limited Canadian comparative perspective, to explore wider issues, including whether 'waiting crises' were consciously used by policymakers, especially those brought into government to implement new economic and managerial strategies, to diminish the autonomy and authority of the medical professional in the hospital environment.

  16. Waiting for cataract surgery--effects of a maximum waiting-time guarantee.

    PubMed

    Hanning, Marianne; Lundström, Mats

    2007-01-01

    To evaluate the effects of the Maximum Waiting-time Guarantee (MWG) policy for cataract surgery on volume, indications, waiting times and priority setting in Sweden. Comparison between 1993 and 1994, when the guarantee had been in force for one year, and 1998 and 1999, when the policy had been terminated for one year. Data from the National Cataract Registry covering 156,657 cataract operations for the years studied. The number of operations increased by 43% between the two study periods. Of this increase, 61% were patients with a visual acuity above 0.5 in the better eye, i.e. low-priority patients. Waiting times were longer for all patient categories in the later period and differences in waiting times between patients with differing priority diminished. Variations among the units in priority setting and waiting times were substantial, and increased after the Guarantee was terminated. The Guarantee with its explicit indications was an effective policy instrument to limit waiting times and improve access for patients with the greatest need. It is unlikely that the Guarantee caused any 'crowding out' of other patient groups. When the Guarantee was not in force, indications for surgery widened. This, however, resulted in longer waiting times for all patient groups. After the Guarantee was terminated, the already substantial differences in access and indications among ophthalmic units became even greater.

  17. [Has the time arrived for the management of waiting lists?].

    PubMed

    Bernal, E

    2002-01-01

    Individuals on the waiting list frequently suffer an additional risk caused by the mean time until they receive treatment; however, other individuals do not need the treatment for which they are waiting.Both arguments, which can be contrasted with empirical evidence, would be sufficient to affirm that waiting list management should be implemented, leaving aside policies that are more of less opportunistic. Opportunistic policies are understood as those providing misinformation on waiting lists or their "manipulation", and using programs of auto-coordination with the sole aim of reaching the end of the year without a waiting list of not more than six months, etc. The panorama is not completely bleak. Some management initiatives (and even Politics with a capital P) are opening the way forward and may enter the Agenda in the next few years. In this context, the application of guaranteed times of medical care or the prioritization of waiting lists according to explicit criteria should be highlighted. It is worth remembering that, except for the queues in the waiting rooms of health centers and emergency departments, waiting lists are mediated by the decision of the physician. Therefore, an essential strategy for managing waiting lists consists of attenuating the problems caused by uncertainty (or ignorance) of the patient's diagnosis or prognosis.

  18. [Dealing with Waiting Times in Health Systems - An International Comparative Overview].

    PubMed

    Finkenstädt, V

    2015-10-01

    Waiting times in the health system are a form of rationing that exists in many countries. Previous studies on this topic are mainly related to the problem of international comparability of waiting times or on the presentation of national strategies as to how they should be reduced. This review adds to this analysis and examines how the OECD countries deal with waiting times in the health-care system and investigates which information is published about waiting for what purpose. Furthermore, waiting times and the type of health system financing are compared. A systematic internet research on waiting times in the health-care system was conducted on the websites of the competent authorities (Ministry of Health or other authorities and institutions). The identified publications were then examined for the purpose of their deployment. Finally, the OECD Health Data were analysed to determine the relationship between tax and contribution financing of public health care expenditure. The primary form of financing was compared with the results of the waiting time analysis. 16 OECD countries are identified which officially collect and publish administrative data on waiting times on the Internet. The data are processed differently depending on the country. By providing this information, two main objectives are pursued: a public monitoring of waiting times in the health system (14 countries) and information for patients on waiting times (9 countries). Official statistics on waiting times exist mainly in countries with tax-financed health systems, whereas this is not the case in the majority of OECD countries with health systems that are funded through contributions. The publication of administrative waiting times data is primarily intended to inform the patient and as a performance indicator in terms of access to health care. Even if data on waiting times are published, the publication of indicators and the management of waiting lists alone will not solve the problem. Rather

  19. [Gender and age differences in waiting time on hospital waiting list.].

    PubMed

    Thornórðardóttir, Steinunn; Halldórsson, Matthías; Guðmundsson, Sigurður

    2002-09-01

    The size of waiting lists has traditionally been viewed as a fairly good measure of the quality of health care services. No statistical analysis exists in Iceland of the length of waiting times and the potential variation between groups of patients. This study was conducted within the office of the Directorate of Health in Iceland. This location was convenient since standardized information on waiting lists is collected by the office three times a year. Variations in waiting times were studied based on gender on the one hand and on age on the other. Data from the largest waiting lists, those amounting to 400 or more patients, were included in the study. The most frequently awaited operations were identified and the groups of people waiting for them analyzed. The departments and prospective operations included in the study were: Dept. of General Surgery at the University Hospital (UH) (laparoscopic gastro-oesophageal antireflux operation), Opthalmology at UH (phakoemulsification with implantation of artificial lens in posterior chamber), Orthopedic Surgery at UH (primary total prosthetic replacement of hip joint using sement), The Rehabilitation Center at Reykjalundur (rehabilitation, not specified), Ear, Nose and Throat (ENT) at UH (tonsillectomy), and Reconstructive Surgery at UH (reduction mammoplasty with transposition of areola). The lists were sorted by gender and age, with the latter consisting of two categories, older and younger patients. Every attempt was made as to ensure similar sample sizes for both age groups within each department. Finally, the median waiting time was determined and a Mann-Whitney test conducted in order to test for significance. The median waiting time for males at the General Surgery Dept. was 73 weeks as compared to 60 weeks for females. This was the only department where the median waiting time was significantly longer for males than for females (p<0.05). At three of the departments the older group had a longer median waiting time

  20. Wait time management strategies for scheduled care: what makes them succeed?

    PubMed

    Pomey, Marie-Pascale; Forest, Pierre-Gerlier; Sanmartin, Claudia; De Coster, Carolyn; Drew, Madeleine

    2010-02-01

    To assess experts' perceptions of the contextual and local factors that promote or inhibit the implementation of waiting time management strategies (WTMS) in Canadian healthcare organizations. We conducted 16 semi-structured interviews and one focus group with individuals involved in WTMS at the federal, provincial or organizational level. The most frequently cited local factor was physicians' participation. Physicians' leadership made the greatest difference in bringing resistant physicians on board. To be effective, however, local leadership had to be supported by senior management. Alignment of financial incentives between the contextual and local levels was also frequently cited, and interviewees stressed the importance of tools used to design, monitor, evaluate and prioritize WTMS. Finding the right balance between supportive resources and tools and an effective management system is a tough challenge. But achieving this balance will help reconcile contradictions between top-down and bottom-up WTMS.

  1. Wait Time Management Strategies for Scheduled Care: What Makes Them Succeed?

    PubMed Central

    Pomey, Marie-Pascale; Forest, Pierre-Gerlier; Sanmartin, Claudia; De Coster, Carolyn; Drew, Madeleine

    2010-01-01

    Objectives: To assess experts' perceptions of the contextual and local factors that promote or inhibit the implementation of waiting time management strategies (WTMS) in Canadian healthcare organizations. Methods: We conducted 16 semi-structured interviews and one focus group with individuals involved in WTMS at the federal, provincial or organizational level. Results: The most frequently cited local factor was physicians' participation. Physicians' leadership made the greatest difference in bringing resistant physicians on board. To be effective, however, local leadership had to be supported by senior management. Alignment of financial incentives between the contextual and local levels was also frequently cited, and interviewees stressed the importance of tools used to design, monitor, evaluate and prioritize WTMS. Conclusions: Finding the right balance between supportive resources and tools and an effective management system is a tough challenge. But achieving this balance will help reconcile contradictions between top-down and bottom-up WTMS. PMID:21286269

  2. Time while waiting: patients' experiences of scheduled surgery.

    PubMed

    Carr, Tracey; Teucher, Ulrich C; Casson, Alan G

    2014-12-01

    Research on patients' experiences of wait time for scheduled surgery has centered predominantly on the relative tolerability of perceived wait time and impacts on quality of life. We explored patients' experiences of time while waiting for three types of surgery with varied wait times--hip or knee replacement, shoulder surgery, and cardiac surgery. Thirty-two patients were recruited by their surgeons. We asked participants about their perceptions of time while waiting in two separate interviews. Using interpretative phenomenological analysis (IPA), we discovered connections between participant suffering, meaningfulness of time, and agency over the waiting period and the lived duration of time experience. Our findings reveal that chronological duration is not necessarily the most relevant consideration in determining the quality of waiting experience. Those findings helped us create a conceptual framework for lived wait time. We suggest that clinicians and policy makers consider the complexity of wait time experience to enhance preoperative patient care. © The Author(s) 2014.

  3. Third degree waiting time discrimination: optimal allocation of a public sector healthcare treatment under rationing by waiting.

    PubMed

    Gravelle, Hugh; Siciliani, Luigi

    2009-08-01

    In many public healthcare systems treatments are rationed by waiting time. We examine the optimal allocation of a fixed supply of a given treatment between different groups of patients. Even in the absence of any distributional aims, welfare is increased by third degree waiting time discrimination: setting different waiting times for different groups waiting for the same treatment. Because waiting time imposes dead weight losses on patients, lower waiting times should be offered to groups with higher marginal waiting time costs and with less elastic demand for the treatment.

  4. Stochastic nature of series of waiting times.

    PubMed

    Anvari, Mehrnaz; Aghamohammadi, Cina; Dashti-Naserabadi, H; Salehi, E; Behjat, E; Qorbani, M; Nezhad, M Khazaei; Zirak, M; Hadjihosseini, Ali; Peinke, Joachim; Tabar, M Reza Rahimi

    2013-06-01

    Although fluctuations in the waiting time series have been studied for a long time, some important issues such as its long-range memory and its stochastic features in the presence of nonstationarity have so far remained unstudied. Here we find that the "waiting times" series for a given increment level have long-range correlations with Hurst exponents belonging to the interval 1/2waiting time distribution. We find that the logarithmic difference of waiting times series has a short-range correlation, and then we study its stochastic nature using the Markovian method and determine the corresponding Kramers-Moyal coefficients. As an example, we analyze the velocity fluctuations in high Reynolds number turbulence and determine the level dependence of Markov time scales, as well as the drift and diffusion coefficients. We show that the waiting time distributions exhibit power law tails, and we were able to model the distribution with a continuous time random walk.

  5. Stochastic nature of series of waiting times

    NASA Astrophysics Data System (ADS)

    Anvari, Mehrnaz; Aghamohammadi, Cina; Dashti-Naserabadi, H.; Salehi, E.; Behjat, E.; Qorbani, M.; Khazaei Nezhad, M.; Zirak, M.; Hadjihosseini, Ali; Peinke, Joachim; Tabar, M. Reza Rahimi

    2013-06-01

    Although fluctuations in the waiting time series have been studied for a long time, some important issues such as its long-range memory and its stochastic features in the presence of nonstationarity have so far remained unstudied. Here we find that the “waiting times” series for a given increment level have long-range correlations with Hurst exponents belonging to the interval 1/2waiting time distribution. We find that the logarithmic difference of waiting times series has a short-range correlation, and then we study its stochastic nature using the Markovian method and determine the corresponding Kramers-Moyal coefficients. As an example, we analyze the velocity fluctuations in high Reynolds number turbulence and determine the level dependence of Markov time scales, as well as the drift and diffusion coefficients. We show that the waiting time distributions exhibit power law tails, and we were able to model the distribution with a continuous time random walk.

  6. Waiting time distributions in financial markets

    NASA Astrophysics Data System (ADS)

    Sabatelli, L.; Keating, S.; Dudley, J.; Richmond, P.

    2002-05-01

    We study waiting time distributions for data representing two completely different financial markets that have dramatically different characteristics. The first are data for the Irish market during the 19th century over the period 1850 to 1854. A total of 10 stocks out of a database of 60 are examined. The second database is for Japanese yen currency fluctuations during the latter part of the 20th century (1989-1992). The Irish stock activity was recorded on a daily basis and activity was characterised by waiting times that varied from one day to a few months. The Japanese yen data was recorded every minute over 24 hour periods and the waiting times varied from a minute to a an hour or so. For both data sets, the waiting time distributions exhibit power law tails. The results for Irish daily data can be easily interpreted using the model of a continuous time random walk first proposed by Montroll and applied recently to some financial data by Mainardi, Scalas and colleagues. Yen data show a quite different behaviour. For large waiting times, the Irish data exhibit a cut off; the Yen data exhibit two humps that could arise as result of major trading centres in the World.

  7. Electron Waiting Times of a Cooper Pair Splitter

    NASA Astrophysics Data System (ADS)

    Walldorf, Nicklas; Padurariu, Ciprian; Jauho, Antti-Pekka; Flindt, Christian

    2018-02-01

    Electron waiting times are an important concept in the analysis of quantum transport in nanoscale conductors. Here we show that the statistics of electron waiting times can be used to characterize Cooper pair splitters that create spatially separated spin-entangled electrons. A short waiting time between electrons tunneling into different leads is associated with the fast emission of a split Cooper pair, while long waiting times are governed by the slow injection of Cooper pairs from a superconductor. Experimentally, the waiting time distributions can be measured using real-time single-electron detectors in the regime of slow tunneling, where conventional current measurements are demanding. Our work is important for understanding the fundamental transport processes in Cooper pair splitters and the predictions may be verified using current technology.

  8. Improving Patient Satisfaction with Waiting Time

    ERIC Educational Resources Information Center

    Eilers, Gayleen M.

    2004-01-01

    Waiting times are a significant component of patient satisfaction. A patient satisfaction survey performed in the author's health center showed that students rated waiting time lowest of the listed categories--A ratings of 58% overall, 63% for scheduled appointments, and 41% for the walk-in clinic. The center used a quality improvement process and…

  9. Impact of Appointment Waiting Time on Attendance Rates at a Clinical Cancer Genetics Service.

    PubMed

    Shaw, Tarryn; Metras, Julie; Ting, Zoe Ang Li; Courtney, Eliza; Li, Shao-Tzu; Ngeow, Joanne

    2018-05-24

    The increase in demand for clinical cancer genetics services has impacted the ability to provide services timeously. Given limited resources, this often results in extended appointment waiting times. Over the last 3 years, the Cancer Genetics Service at the National Cancer Centre Singapore has continued to experience a steady increase in demand for its service. Nevertheless, significant no-show rates have been reported. This study sought to determine whether an association exists between appointment waiting times and attendance rates. Data was gathered for all participants meeting inclusion criteria. Attendance rates and appointment waiting times were calculated. The relationship between mean waiting times for those who did and did not attend their scheduled appointments was evaluated using Welch's t test and linear regression model. The results showed a significant difference in mean appointment waiting times between patients who did and did not attend (32.66 versus 43.50 days respectively; p < 0.0001). Furthermore, patients who waited for longer than 37 days were significantly less likely to attend. No-show rates increased as the waiting time increased, at a rate of 19.60% per 20 days and 21.40% per 30 days. In conclusion, appointment waiting time is a significant predictor for patient attendance. Strategies to ensure patients receive an appointment within the necessary timeframe at the desired setting are important to ensure that individuals at increased cancer risk attend their appointments in order to manage their cancer risks effectively.

  10. Are seismic waiting time distributions universal?

    NASA Astrophysics Data System (ADS)

    Davidsen, Jörn; Goltz, Christian

    2004-11-01

    We show that seismic waiting time distributions in California and Iceland have many features in common as, for example, a power-law decay with exponent α ~ 1.1 for intermediate and with exponent γ ~ 0.6 for short waiting times. While the transition point between these two regimes scales proportionally with the size of the considered area, the full distribution is not universal and depends in a non-trivial way on the geological area under consideration and its size. This is due to the spatial distribution of epicenters which does not form a simple mono-fractal. Yet, the dependence of the waiting time distributions on the threshold magnitude seems to be universal.

  11. Continuous-Time Finance and the Waiting Time Distribution: Multiple Characteristic Times

    NASA Astrophysics Data System (ADS)

    Fa, Kwok Sau

    2012-09-01

    In this paper, we model the tick-by-tick dynamics of markets by using the continuous-time random walk (CTRW) model. We employ a sum of products of power law and stretched exponential functions for the waiting time probability distribution function; this function can fit well the waiting time distribution for BUND futures traded at LIFFE in 1997.

  12. [Influence of waiting time on patient and companion satisfaction].

    PubMed

    Fontova-Almató, A; Juvinyà-Canal, D; Suñer-Soler, R

    2015-01-01

    To evaluate patient and companion satisfaction of a hospital Emergency Department and its relationship with waiting time. Prospective, observational study. Hospital de Figueres Emergency Department (Girona, Spain). sociodemographic characteristics, satisfaction level, real and perceived waiting time for triage and being seen by a physician. A total of 285 responses were received from patients and companions. The mean age of the patients and companions (n=257) was 54.6years (SD=18.3). The mean overall satisfaction (n=273) was 7.6 (SD=2.2). Lower perceived waiting time until nurse triage was related to higher overall satisfaction (Spearman rho (ρ)=-0.242, P<.001), and lower perceived waiting time until being seen by physician, with a higher overall satisfaction (ρ=-0.304; P<.001). Users who were informed about estimated waiting time showed higher satisfaction than those who were not informed (P=.001). Perceived waiting time and the information about estimated waiting time determined overall satisfaction. Copyright © 2014 SECA. Published by Elsevier Espana. All rights reserved.

  13. Waiting time care guarantees: necessity or nemesis?

    PubMed

    Joshi, N P; Noseworthy, F T; Noseworthy, T W

    2006-01-01

    One of the priorities of governments in Canada is to reduce long waiting times for health services. This has raised the prospect of introducing waiting time care guarantees. Such guarantees affirm the healthcare system's social contract with the public and provide an entitlement to Canadians to receive timely care. There are clinical, legal and political implications, which must be considered and well managed before introduction. Other countries have ventured down this path. They teach us that waiting time care guarantees are good policy and make good sense. Correspondingly, they remind us not to make a promise we are not ready to keep.

  14. Electron Waiting Times in Mesoscopic Conductors

    NASA Astrophysics Data System (ADS)

    Albert, Mathias; Haack, Géraldine; Flindt, Christian; Büttiker, Markus

    2012-05-01

    Electron transport in mesoscopic conductors has traditionally involved investigations of the mean current and the fluctuations of the current. A complementary view on charge transport is provided by the distribution of waiting times between charge carriers, but a proper theoretical framework for coherent electronic systems has so far been lacking. Here we develop a quantum theory of electron waiting times in mesoscopic conductors expressed by a compact determinant formula. We illustrate our methodology by calculating the waiting time distribution for a quantum point contact and find a crossover from Wigner-Dyson statistics at full transmission to Poisson statistics close to pinch-off. Even when the low-frequency transport is noiseless, the electrons are not equally spaced in time due to their inherent wave nature. We discuss the implications for renewal theory in mesoscopic systems and point out several analogies with level spacing statistics and random matrix theory.

  15. Reducing wait time in a hospital pharmacy to promote customer service.

    PubMed

    Slowiak, Julie M; Huitema, Bradley E; Dickinson, Alyce M

    2008-01-01

    The purpose of this study was to compare the effects of 2 different interventions on wait times at a hospital outpatient pharmacy: (1) giving feedback to employees about customer satisfaction with wait times and (2) giving a combined intervention package that included giving more specific feedback about actual wait times and goal setting for wait time reduction in addition to the customer satisfaction feedback. The relationship between customer satisfaction ratings and wait times was examined to determine whether wait times affected customer service satisfaction. Participants were 10 employees (4 pharmacists and 6 technicians) of an outpatient pharmacy. Wait times and customer satisfaction ratings were collected for "waiting customers." An ABCBA' within-subjects design was used to assess the effects of the interventions on both wait time and customer satisfaction, where A was the baseline (no feedback and no goal setting); B was the customer satisfaction feedback; C was the customer satisfaction feedback, the wait time feedback, and the goal setting for wait time reduction; and A' was a follow-up condition that was similar to the original baseline condition. Wait times were reduced by approximately 20%, and there was concomitant increased shift in levels of customer satisfaction, as indicated by the correlation between these variables (r = -0.57 and P < .05). Given the current prescription-filling process, we do not expect that major, additional reductions in wait times could be produced. Many variables may account for the variability in any individual customer's wait time. Data from this study may provide useful preliminary benchmarking data for standard pharmacy wait times.

  16. The Impact of One-Dose Package of Medicines on Patient Waiting Time in Dispensing Pharmacy: Application of a Discrete Event Simulation Model.

    PubMed

    Furushima, Daisuke; Yamada, Hiroshi; Kido, Michiko; Ohno, Yuko

    2018-01-01

    Improvement in patient waiting time in dispensing pharmacies is an important element for patient and pharmacists. The One-Dose Package (ODP) of medicines was implemented in Japan to support medicine adherence among elderly patients; however, it also contributed to increase in patient waiting times. Given the projected increase in ODP patients in the near future owing to rapid population aging, development of improved strategies is a key imperative. We conducted a cross-sectional survey at a single dispensing pharmacy to clarify the impact of ODP on patient waiting time. Further, we propose an improvement strategy developed with use of a discrete event simulation (DES) model. A total of 673 patients received pharmacy services during the study period. A two-fold difference in mean waiting time was observed between ODP and non-ODP patients (22.6 and 11.2 min, respectively). The DES model was constructed with input parameters estimated from observed data. Introduction of fully automated ODP (A-ODP) system was projected to reduce the waiting time for ODP patient by 0.5 times (from 23.1 to 11.5 min). Furthermore, assuming that 40% of non-ODP patients would transfer to ODP, the waiting time was predicted to increase to 56.8 min; however, introduction of the A-ODP system decreased the waiting time to 20.4 min. Our findings indicate that ODP is one of the elements that increases the waiting time and that it might become longer in the future. Introduction of the A-ODP system may be an effective strategy to improve waiting time.

  17. Parental satisfaction with paediatric care, triage and waiting times.

    PubMed

    Fitzpatrick, Nicholas; Breen, Daniel T; Taylor, James; Paul, Eldho; Grosvenor, Robert; Heggie, Katrina; Mahar, Patrick D

    2014-04-01

    The present study aims to determine parental and guardian's perceptions of paediatric emergency care and satisfaction with care, waiting times and triage category in a community ED. A structured questionnaire was provided to parents or guardians of paediatric patients presenting to emergency. The survey evaluated parent perceptions of waiting time, environment/facilities, professionalism and communication skills of staff and overall satisfaction of care. One hundred and thirty-three completed questionnaires were received from parents of paediatric patients. Responses were overall positive with respect to the multiple domains assessed. Parents generally considered waiting times to be appropriate and consistent with triage categories. Overall satisfaction was not significantly different for varying treatment or waiting times. Patients triaged as semi-urgent were of the opinion that waiting times were less appropriate than urgent, less-urgent or non-urgent patients. On the basis of the present study, patient perceptions and overall satisfaction of care does not appear to be primarily influenced by time spent waiting or receiving treatment. Attempts made at the triage process to ensure that semi-urgent patients have reasonable expectations of waiting times might provide an opportunity to improve these patients' expectations and perceptions. © 2014 Australasian College for Emergency Medicine and Australasian Society for Emergency Medicine.

  18. Identifying demand for health resources using waiting times information.

    PubMed

    Blundell, R; Windmeijer, F

    2000-09-01

    In this paper the differences in average waiting times are utilized to identify the determinants of demand for health services. The equilibrium waiting time framework is used, but the full equilibrium assumption is relaxed by selecting areas with low waiting times and by estimating a (semi-)parametric selection model. Determinants of supply are used as instruments for the endogeneity of waiting times. A model for the demand for acute services at the ward level in the UK is estimated. The model estimates, and their implications for health service allocations in the UK, are contrasted against more standard allocation models. The present results show that it is critically important to account for rationing by waiting times when identifying needs from care utilization data. Copyright 2000 John Wiley & Sons, Ltd.

  19. Non-Poissonian Distribution of Tsunami Waiting Times

    NASA Astrophysics Data System (ADS)

    Geist, E. L.; Parsons, T.

    2007-12-01

    Analysis of the global tsunami catalog indicates that tsunami waiting times deviate from an exponential distribution one would expect from a Poisson process. Empirical density distributions of tsunami waiting times were determined using both global tsunami origin times and tsunami arrival times at a particular site with a sufficient catalog: Hilo, Hawai'i. Most sources for the tsunamis in the catalog are earthquakes; other sources include landslides and volcanogenic processes. Both datasets indicate an over-abundance of short waiting times in comparison to an exponential distribution. Two types of probability models are investigated to explain this observation. Model (1) is a universal scaling law that describes long-term clustering of sources with a gamma distribution. The shape parameter (γ) for the global tsunami distribution is similar to that of the global earthquake catalog γ=0.63-0.67 [Corral, 2004]. For the Hilo catalog, γ is slightly greater (0.75-0.82) and closer to an exponential distribution. This is explained by the fact that tsunamis from smaller triggered earthquakes or landslides are less likely to be recorded at a far-field station such as Hilo in comparison to the global catalog, which includes a greater proportion of local tsunamis. Model (2) is based on two distributions derived from Omori's law for the temporal decay of triggered sources (aftershocks). The first is the ETAS distribution derived by Saichev and Sornette [2007], which is shown to fit the distribution of observed tsunami waiting times. The second is a simpler two-parameter distribution that is the exponential distribution augmented by a linear decay in aftershocks multiplied by a time constant Ta. Examination of the sources associated with short tsunami waiting times indicate that triggered events include both earthquake and landslide tsunamis that begin in the vicinity of the primary source. Triggered seismogenic tsunamis do not necessarily originate from the same fault zone

  20. The British Columbia Nephrologists' Access Study (BCNAS) - a prospective, health services interventional study to develop waiting time benchmarks and reduce wait times for out-patient nephrology consultations.

    PubMed

    Schachter, Michael E; Romann, Alexandra; Djurdev, Ognjenka; Levin, Adeera; Beaulieu, Monica

    2013-08-29

    Early referral and management of high-risk chronic kidney disease may prevent or delay the need for dialysis. Automatic eGFR reporting has increased demand for out-patient nephrology consultations and in some cases, prolonged queues. In Canada, a national task force suggested the development of waiting time targets, which has not been done for nephrology. We sought to describe waiting time for outpatient nephrology consultations in British Columbia (BC). Data collection occurred in 2 phases: 1) Baseline Description (Jan 18-28, 2010) and 2) Post Waiting Time Benchmark-Introduction (Jan 16-27, 2012). Waiting time was defined as the interval from receipt of referral letters to assessment. Using a modified Delphi process, Nephrologists and Family Physicians (FP) developed waiting time targets for commonly referred conditions through meetings and surveys. Rules were developed to weigh-in nephrologists', FPs', and patients' perspectives in order to generate waiting time benchmarks. Targets consider comorbidities, eGFR, BP and albuminuria. Referred conditions were assigned a priority score between 1-4. BC nephrologists were encouraged to centrally triage referrals to see the first available nephrologist. Waiting time benchmarks were simultaneously introduced to guide patient scheduling. A post-intervention waiting time evaluation was then repeated. In 2010 and 2012, 43/52 (83%) and 46/57 (81%) of BC nephrologists participated. Waiting time decreased from 98(IQR44,157) to 64(IQR21,120) days from 2010 to 2012 (p = <.001), despite no change in referral eGFR, demographics, nor number of office hrs/wk. Waiting time improved most for high priority patients. An integrated, Provincial initiative to measure wait times, develop waiting benchmarks, and engage physicians in active waiting time management associated with improved access to nephrologists in BC. Improvements in waiting time was most marked for the highest priority patients, which suggests that benchmarks had an

  1. Do waiting times affect health outcomes? Evidence from coronary bypass.

    PubMed

    Moscelli, Giuseppe; Siciliani, Luigi; Tonei, Valentina

    2016-07-01

    Long waiting times for non-emergency services are a feature of several publicly-funded health systems. A key policy concern is that long waiting times may worsen health outcomes: when patients receive treatment, their health condition may have deteriorated and health gains reduced. This study investigates whether patients in need of coronary bypass with longer waiting times are associated with poorer health outcomes in the English National Health Service over 2000-2010. Exploiting information from the Hospital Episode Statistics (HES), we measure health outcomes with in-hospital mortality and 28-day emergency readmission following discharge. Our results, obtained combining hospital fixed effects and instrumental variable methods, find no evidence of waiting times being associated with higher in-hospital mortality and weak association between waiting times and emergency readmission following a surgery. The results inform the debate on the relative merits of different types of rationing in healthcare systems. They are to some extent supportive of waiting times as an acceptable rationing mechanism, although further research is required to explore whether long waiting times affect other aspects of individuals' life. Copyright © 2016 Elsevier Ltd. All rights reserved.

  2. Effect of emergency physician burnout on patient waiting times.

    PubMed

    De Stefano, Carla; Philippon, Anne-Laure; Krastinova, Evguenia; Hausfater, Pierre; Riou, Bruno; Adnet, Frederic; Freund, Yonathan

    2018-04-01

    Burnout is common in emergency physicians. This syndrome may negatively affect patient care and alter work productivity. We seek to assess whether burnout of emergency physicians impacts waiting times in the emergency department. Prospective study in an academic ED. All patients who visited the main ED for a 4-month period in 2016 were included. Target waiting times are assigned by triage nurse to patients on arrival depending on their severity. The primary endpoint was an exceeded target waiting time for ED patients. All emergency physicians were surveyed by a psychologist to assess their level of burnout using the Maslach Burnout Inventory. We defined the level of burnout of the day in the ED as the mean burnout level of the physicians working that day (8:30 to the 8:30 the next day). A logistic regression model was performed to assess whether burnout level of the day was independently associated with prolonged waiting times, along with previously reported predictors. Target waiting time was exceeded in 7524 patients (59%). Twenty-six emergency physicians were surveyed. Median burnout score was 35 [Interquartile (24-49)]. A burnout level of the day higher than 35 was independently associated with an exceeded target waiting time (adjusted odds ratio 1.54, 95% confidence interval 1.39-1.70), together with previously reported predictors (i.e., day of the week, time of the day, trauma, age and daily census). Burnout of emergency physicians was independently associated with a prolonged waiting time for patients visiting the ED.

  3. Outpatient Office Wait Times and Quality of Care for Medicaid Patients

    PubMed Central

    Oostrom, Tamar; Einav, Liran; Finkelstein, Amy

    2018-01-01

    Time spent in the doctor’s waiting room captures an important aspect of the healthcare experience. We analyzed data on 21 million outpatient visits obtained from electronic health record systems, allowing us to measure time spent in the waiting room beyond the scheduled appointment time. Median wait time was just over 4 minutes. Almost one-fifth of visits had waits longer than 20 minutes, and 10% were over 30 minutes. Waits were shorter for early morning appointments, younger patients, and at larger practices. Median wait time was 4.1 minutes for privately-insured and 4.6 minutes for Medicaid patients; adjusting for patient and appointment characteristics, Medicaid patients were 20% more likely than the privately-insured to wait longer than 20 minutes (P<0.001), with most of this disparity explained by differences in practices and providers they saw. Wait time for Medicaid patients relative to the privately-insured was longer in states with relatively lower Medicaid reimbursement rates. PMID:28461348

  4. Organ Type and Waiting Time

    MedlinePlus

    ... but each organ type has its own individual distribution policy reflects reflect factors that are unique to each organ type: Kidney Waiting time Donor/recipient immune system compatibility Prior living donor ...

  5. The surgical waiting time initiative: A review of the Nigerian situation

    PubMed Central

    Abdulkareem, Imran Haruna

    2014-01-01

    SUMMARY The concept of surgical waiting time initiative (SWAT) was introduced in developed countries to reduce elective surgery waiting lists and increase efficiency of care. It was supplemented by increasing popularity of day surgery, which shortens elective waiting lists and minimises cancellations. It is established in Western countries, but not in developing countries like Nigeria where it is still evolving. A search was carried out in Pub Med, Google, African journals online (AJOL), Athens and Ovid for relevant publications on elective surgery waiting list in Nigeria, published in English language. Words include waiting/wait time, waiting time initiative, time to surgery, waiting for operations, waiting for intervention, waiting for procedures and time before surgery in Nigeria. A total of 37 articles published from Nigeria in relation to various waiting times were found from the search and fulfilled the inclusion criteria. Among them, 11 publications (29.7%) were related to emergency surgery waiting times, 10 (27%) were related to clinic waiting times, 9 (24.3%) were related to day case surgery, 2 (5.5%) were related to investigation waiting times and only 5 (13.5%) articles were specifically published on elective surgery waiting times. A total of 9 articles (24.5%) were published from obstetrics and gynaecology (OG), 7 (19%) from general surgery, 5 (13.5%) from public health, 3 (8%) from orthopaedics, 3 (8%) from general practice (GP), 3 (8%) from paediatrics/paediatric surgery, 2 (5.5%) from ophthalmology, 1 (2.7%) from ear, nose and throat (ENT), 1 (2.7%) from plastic surgery, 1 (2.7%) from urology and only 1 (2.7%) article was published from dental/maxillofacial surgery. Waiting times mean different things to different health practitioners in Nigeria. There were only 5/37 articles (13.5%) specifically related to elective surgery waiting times in Nigerian hospitals, which show that the concept of the SWAT is still evolving in Nigeria. Of the 37, 11 (24

  6. The surgical waiting time initiative: A review of the Nigerian situation.

    PubMed

    Abdulkareem, Imran Haruna

    2014-11-01

    The concept of surgical waiting time initiative (SWAT) was introduced in developed countries to reduce elective surgery waiting lists and increase efficiency of care. It was supplemented by increasing popularity of day surgery, which shortens elective waiting lists and minimises cancellations. It is established in Western countries, but not in developing countries like Nigeria where it is still evolving. A search was carried out in Pub Med, Google, African journals online (AJOL), Athens and Ovid for relevant publications on elective surgery waiting list in Nigeria, published in English language. Words include waiting/wait time, waiting time initiative, time to surgery, waiting for operations, waiting for intervention, waiting for procedures and time before surgery in Nigeria. A total of 37 articles published from Nigeria in relation to various waiting times were found from the search and fulfilled the inclusion criteria. Among them, 11 publications (29.7%) were related to emergency surgery waiting times, 10 (27%) were related to clinic waiting times, 9 (24.3%) were related to day case surgery, 2 (5.5%) were related to investigation waiting times and only 5 (13.5%) articles were specifically published on elective surgery waiting times. A total of 9 articles (24.5%) were published from obstetrics and gynaecology (OG), 7 (19%) from general surgery, 5 (13.5%) from public health, 3 (8%) from orthopaedics, 3 (8%) from general practice (GP), 3 (8%) from paediatrics/paediatric surgery, 2 (5.5%) from ophthalmology, 1 (2.7%) from ear, nose and throat (ENT), 1 (2.7%) from plastic surgery, 1 (2.7%) from urology and only 1 (2.7%) article was published from dental/maxillofacial surgery. Waiting times mean different things to different health practitioners in Nigeria. There were only 5/37 articles (13.5%) specifically related to elective surgery waiting times in Nigerian hospitals, which show that the concept of the SWAT is still evolving in Nigeria. Of the 37, 11 (24

  7. Interventions to reduce waiting times for elective procedures.

    PubMed

    Ballini, Luciana; Negro, Antonella; Maltoni, Susanna; Vignatelli, Luca; Flodgren, Gerd; Simera, Iveta; Holmes, Jane; Grilli, Roberto

    2015-02-23

    Long waiting times for elective healthcare procedures may cause distress among patients, may have adverse health consequences and may be perceived as inappropriate delivery and planning of health care. To assess the effectiveness of interventions aimed at reducing waiting times for elective care, both diagnostic and therapeutic. We searched the following electronic databases: Cochrane Effective Practice and Organisation of Care (EPOC) Group Specialised Register, the Cochrane Central Register of Controlled Trials (CENTRAL), MEDLINE (1946-), EMBASE (1947-), the Cumulative Index to Nursing and Allied Health Literature (CINAHL), ABI Inform, the Canadian Research Index, the Science, Social Sciences and Humanities Citation Indexes, a series of databases via Proquest: Dissertations & Theses (including UK & Ireland), EconLit, PAIS (Public Affairs International), Political Science Collection, Nursing Collection, Sociological Abstracts, Social Services Abstracts and Worldwide Political Science Abstracts. We sought related reviews by searching the Cochrane Database of Systematic Reviews and the Database of Abstracts of Reviews of Effectiveness (DARE). We searched trial registries, as well as grey literature sites and reference lists of relevant articles. We considered randomised controlled trials (RCTs), controlled before-after studies (CBAs) and interrupted time series (ITS) designs that met EPOC minimum criteria and evaluated the effectiveness of any intervention aimed at reducing waiting times for any type of elective procedure. We considered studies reporting one or more of the following outcomes: number or proportion of participants whose waiting times were above or below a specific time threshold, or participants' mean or median waiting times. Comparators could include any type of active intervention or standard practice. Two review authors independently extracted data from, and assessed risk of bias of, each included study, using a standardised form and the EPOC 'Risk

  8. International comparisons of waiting times in health care--limitations and prospects.

    PubMed

    Viberg, Nina; Forsberg, Birger C; Borowitz, Michael; Molin, Roger

    2013-09-01

    Long waiting times for health care is an important health policy issue in many countries, and many have introduced some form of national waiting time guarantees. International comparison of waiting times are critical for countries to improve policy and for patients to be able to make informed choices, especially in Europe, where patients have the right to seek care in other countries if there is undue delay. The objective of this study was to describe how countries measure waiting times and to assess whether waiting times can be compared internationally. Twenty-three OECD countries were included. Information was collected through scientific articles, official and unofficial documents and web pages. Fifteen of the 23 countries monitor and publish national waiting time statistics and have some form of waiting time guarantees. There are significant differences in how waiting times are measured: whether they measure the "ongoing" or "completed" waiting period what kind of care the patient is waiting for; the parameters used; and where in the patient journey the measurement begins. Current national waiting time statistics are of limited use for comparing health care availability among the various countries due to the differences in measurements and data collection. Different methodological issues must be taken into account when making such cross-country comparisons. Within the given context of national sovereignty of health systems it would be desirable if countries could collaborate in order to facilitate international comparisons. Such comparisons would be of benefit to all involved in the process of continuous improvement of health services. They would also benefit patients who seek cross-border alternatives for their care. Copyright © 2013 The Authors. Published by Elsevier Ireland Ltd.. All rights reserved.

  9. The British Columbia Nephrologists’ Access Study (BCNAS) – a prospective, health services interventional study to develop waiting time benchmarks and reduce wait times for out-patient nephrology consultations

    PubMed Central

    2013-01-01

    Background Early referral and management of high-risk chronic kidney disease may prevent or delay the need for dialysis. Automatic eGFR reporting has increased demand for out-patient nephrology consultations and in some cases, prolonged queues. In Canada, a national task force suggested the development of waiting time targets, which has not been done for nephrology. Methods We sought to describe waiting time for outpatient nephrology consultations in British Columbia (BC). Data collection occurred in 2 phases: 1) Baseline Description (Jan 18-28, 2010) and 2) Post Waiting Time Benchmark-Introduction (Jan 16-27, 2012). Waiting time was defined as the interval from receipt of referral letters to assessment. Using a modified Delphi process, Nephrologists and Family Physicians (FP) developed waiting time targets for commonly referred conditions through meetings and surveys. Rules were developed to weigh-in nephrologists’, FPs’, and patients’ perspectives in order to generate waiting time benchmarks. Targets consider comorbidities, eGFR, BP and albuminuria. Referred conditions were assigned a priority score between 1-4. BC nephrologists were encouraged to centrally triage referrals to see the first available nephrologist. Waiting time benchmarks were simultaneously introduced to guide patient scheduling. A post-intervention waiting time evaluation was then repeated. Results In 2010 and 2012, 43/52 (83%) and 46/57 (81%) of BC nephrologists participated. Waiting time decreased from 98(IQR44,157) to 64(IQR21,120) days from 2010 to 2012 (p = <.001), despite no change in referral eGFR, demographics, nor number of office hrs/wk. Waiting time improved most for high priority patients. Conclusions An integrated, Provincial initiative to measure wait times, develop waiting benchmarks, and engage physicians in active waiting time management associated with improved access to nephrologists in BC. Improvements in waiting time was most marked for the highest priority

  10. Optimal Server Scheduling to Maintain Constant Customer Waiting Times

    DTIC Science & Technology

    1988-12-01

    I I• I I I I I LCn CN OPTIMAL SERVER SCHEDUUNG TO MAINTAIN CONSTANT CUSTOMER WAITING TIMES THESIS Thomas J. Frey Captain UISAF AFIT/GOR/ENS/88D-7...hw bees appsewlf in ple rtan. cd = , ’ S 087 AFIT/GORMENS/8D-7 OPTIMAL SERVER SCHEDUUNG TO MAINTAIN~ CONSTANT CUSTOMER WAITING TIMES THESIS Thomas j...CONSTANT CUSTOMER WAITING TIMES THESIS Presented to the Faculty of the School of Engineering of the Air Force Institute of Technology Air University In

  11. Wait times for gastroenterology consultation in Canada: The patients’ perspective

    PubMed Central

    Paterson, WG; Barkun, AN; Hopman, WM; Leddin, DJ; Paré, P; Petrunia, DM; Sewitch, MJ; Switzer, C; van Zanten, S Veldhuyzen

    2010-01-01

    Long wait times for health care have become a significant issue in Canada. As part of the Canadian Association of Gastroenterology’s Human Resource initiative, a questionnaire was developed to survey patients regarding wait times for initial gastroenterology consultation and its impact. A total of 916 patients in six cities from across Canada completed the questionnaire at the time of initial consultation. Self-reported wait times varied widely, with 26.8% of respondents reporting waiting less than two weeks, 52.4% less than one month, 77.1% less than three months, 12.5% reported waiting longer than six months and 3.6% longer than one year. One-third of patients believed their wait time was too long, with 9% rating their wait time as ‘far too long’; 96.4% believed that maximal wait time should be less than three months, 78.9% believed it should be less than one month and 40.3% believed it should be less than two weeks. Of those working or attending school, 22.6% reported missing at least one day of work or school because of their symptoms in the month before their appointment, and 9.0% reported missing five or more days in the preceding month. A total of 20.2% of respondents reported being very worried about having a serious disease (ie, scored 6 or higher on 7-point Likert scale), and 17.6% and 14.8%, respectively, reported that their symptoms caused major impairment of social functioning and with the activities of daily living. These data suggest that a significant proportion of Canadians with digestive problems are not satisfied with their wait time for gastroenterology consultation. Furthermore, while awaiting consultation, many patients experience an impaired quality of life because of their gastrointestinal symptoms. PMID:20186353

  12. Enhancing outpatient clinics management software by reducing patients' waiting time.

    PubMed

    Almomani, Iman; AlSarheed, Ahlam

    The Kingdom of Saudi Arabia (KSA) gives great attention to improving the quality of services provided by health care sectors including outpatient clinics. One of the main drawbacks in outpatient clinics is long waiting time for patients-which affects the level of patient satisfaction and the quality of services. This article addresses this problem by studying the Outpatient Management Software (OMS) and proposing solutions to reduce waiting times. Many hospitals around the world apply solutions to overcome the problem of long waiting times in outpatient clinics such as hospitals in the USA, China, Sri Lanka, and Taiwan. These clinics have succeeded in reducing wait times by 15%, 78%, 60% and 50%, respectively. Such solutions depend mainly on adding more human resources or changing some business or management policies. The solutions presented in this article reduce waiting times by enhancing the software used to manage outpatient clinics services. Both quantitative and qualitative methods have been used to understand current OMS and examine level of patient's satisfaction. Five main problems that may cause high or unmeasured waiting time have been identified: appointment type, ticket numbering, doctor late arrival, early arriving patient and patients' distribution list. These problems have been mapped to the corresponding OMS components. Solutions to the above problems have been introduced and evaluated analytically or by simulation experiments. Evaluation of the results shows a reduction in patient waiting time. When late doctor arrival issues are solved, this can reduce the clinic service time by up to 20%. However, solutions for early arriving patients reduces 53.3% of vital time, 20% of the clinic time and overall 30.3% of the total waiting time. Finally, well patient-distribution lists make improvements by 54.2%. Improvements introduced to the patients' waiting time will consequently affect patients' satisfaction and improve the quality of health care services

  13. Pooled Open Blocks Shorten Wait Times for Nonelective Surgical Cases.

    PubMed

    Zenteno, Ana C; Carnes, Tim; Levi, Retsef; Daily, Bethany J; Price, Devon; Moss, Susan C; Dunn, Peter F

    2015-07-01

    Assess the impact of the implementation of a data-driven scheduling strategy that aimed to improve the access to care of nonelective surgical patients at Massachusetts General Hospital (MGH). Between July 2009 and June 2010, MGH experienced increasing throughput challenges in its perioperative environment: approximately 30% of the nonelective patients were waiting more than the prescribed amount of time to get to surgery, hampering access to care and aggravating the lack of inpatient beds. This work describes the design and implementation of an "open block" strategy: operating room (OR) blocks were reserved for nonelective patients during regular working hours (prime time) and their management centralized. Discrete event simulation showed that 5 rooms would decrease the percentage of delayed patients from 30% to 2%, assuming that OR availability was the only reason for preoperative delay. Implementation began in January 2012. We compare metrics for June through December of 2012 against the same months of 2011. The average preoperative wait time of all nonelective surgical patients decreased by 25.5% (P < 0.001), even with a volume increase of 9%. The number of bed-days occupied by nonurgent patients before surgery declined by 13.3% whereas the volume increased by 4.5%. The large-scale application of an open-block strategy significantly improved the flow of nonelective patients at MGH when OR availability was a major reason for delay. Rigorous metrics were developed to evaluate its performance. Strong managerial leadership was crucial to enact the new practices and turn them into organizational change.

  14. Numbers or apologies? Customer reactions to telephone waiting time fillers.

    PubMed

    Munichor, Nira; Rafaeli, Anat

    2007-03-01

    The authors examined the effect of time perception and sense of progress in telephone queues on caller reactions to 3 telephone waiting time fillers: music, apologies, and information about location in the queue. In Study 1, conducted on 123 real calls, call abandonment was lowest, and call evaluations were most positive with information about location in the queue as the time filler. In Study 2, conducted with 83 participants who experienced a simulated telephone wait experience, sense of progress in the queue rather than perceived waiting time mediated the relationship between telephone waiting time filler and caller reactions. The findings provide insight for the management and design of telephone queues, as well as theoretical insight into critical cognitive processes that underlie telephone waiting, opening up an important new research agenda. (c) 2007 APA, all rights reserved.

  15. Waiting times for hospital admissions: the impact of GP fundholding.

    PubMed

    Propper, Carol; Croxson, Bronwyn; Shearer, Arran

    2002-03-01

    Waiting times for hospital care are a significant issue in the UK National Health Service (NHS). The reforms of the health service in 1990 gave a subset of family doctors (GP fundholders) both the ability to choose the hospital where their patients were treated and the means to pay for some services. One of the key factors influencing family doctors' choice of hospital was patient waiting time. However, without cash inducements, hospitals would get no direct reward from giving shorter waiting times to a subset of patients. Using a unique dataset, we investigate whether GP fundholders were able to secure shorter waiting times for their patients, whether they were able to do so in cases where they had no financial rewards to offer hospitals, and whether the impact of fundholding spilled over into shorter waiting times for all patients.

  16. Waiting time effect of a GM type orifice pulse tube refrigerator

    NASA Astrophysics Data System (ADS)

    Zhu, Shaowei; Kakimi, Yasuhiro; Matsubara, Yoichi

    In a general GM type orifice pulse tube refrigerator, there are two short periods during which both the high pressure valve and the low pressure valve are closed in one cycle. We call the short period `waiting time'. The pressure differences across the high pressure valve and the low pressure valve are decreased by using long waiting time. The pressure difference loss is decreased. Thus, the cooling capacity and the efficiency are increased, and the no-load temperature is decreased. The mechanism of the waiting time is discussed with numerical analysis and verified by experiments. Experiments show that there is an optimum waiting time for the no-load temperature, the cooling capacity and the efficiency, respectively. The no-load temperature of 40.3 K was achieved with a 90° waiting time. The cooling capacity of 58 W at 80 K was achieved with a 60° waiting time. The no-load temperature of 45.1 K and the cooling capacity of 45 W at 80 K were achieved with a 1° waiting time.

  17. Wait Time for Counseling Affecting Perceived Stigma and Attitude toward the University

    ERIC Educational Resources Information Center

    Blau, Gary; DiMino, John; Sheridan, Natalie; Stein, Alexander; Casper, Steven; Chessler, Marcy; Beverly, Clyde

    2015-01-01

    A sample of 99 undergraduates in counseling was divided into two groups based on wait time from triage to intake, "less wait time" (up to two weeks) versus "more wait time" (at least two weeks). The less wait time group showed "higher willingness to recommend the university," "higher institutional…

  18. Reducing pharmacy wait time to promote customer service: a follow-up study.

    PubMed

    Slowiak, Julie M; Huitema, Bradley E

    2015-01-01

    The present study had 3 objectives: (1) to evaluate the effects of 2 different interventions (feedback regarding customer satisfaction with wait time and combined feedback and goal setting) on wait time in a hospital outpatient pharmacy; (2) to assess the extent to which the previously applied interventions maintained their effects; and (3) to evaluate the differences between the effects of the original study and those of the present follow-up study. Participants were 10 employees (4 pharmacists and 6 technicians) of an outpatient pharmacy. Wait times and customer satisfaction ratings were collected for "waiting customers." An ABCB within-subjects design was used to assess the effects of the interventions on both wait time and customer satisfaction, where A was the baseline (no feedback and no goal setting); B was the customer satisfaction feedback; and C was the customer satisfaction feedback, the wait time feedback, and the goal setting for wait time reduction. Wait time decreased after baseline when the combined intervention was introduced, and wait time increased with the reintroduction of satisfaction feedback (alone). The results of the replication study confirm the pattern of the results of the original study and demonstrate high sensitivity of levels of customer satisfaction with wait time. The most impressive result of the replication is the nearly 2-year maintenance of lower wait time between the end of the original study and the beginning (baseline) of the replication.

  19. Individual and system influences on waiting time for substance abuse treatment.

    PubMed

    Carr, Carey J A; Xu, Jiangmin; Redko, Cristina; Lane, D Timothy; Rapp, Richard C; Goris, John; Carlson, Robert G

    2008-03-01

    Waiting time is a contemporary reality of many drug abuse treatment programs, resulting in substantial problems for substance users and society. Individual and system factors that influence waiting time are diverse and may vary at different points in the treatment continuum. This study assessed waiting time preceding clinical assessment at a centralized intake unit and during the period after the assessment but before treatment entry. The present study included 577 substance abusers who were enrolled in a large clinical trial of two brief treatment interventions in a midsize metropolitan area in Ohio. Bivariate analyses identified individual and system factors that influenced preassessment and postassessment waiting time, as well as total wait to treatment services. Multivariate analyses demonstrated that longer wait time for an assessment is influenced by being court referred, less belief in having a substance abuse problem, and less desire for change. A shorter wait to actually enter treatment is predicted by having a case manager, being more ready for treatment, and having less severe employment and alcohol problems. The different influences present during the two waiting periods suggest that assessment and treatment programs need to implement system changes and entry enhancement interventions that are specific to the needs of substance abusers at each waiting period.

  20. Real waiting times for surgery. Proposal for an improved system for their management.

    PubMed

    Abásolo, Ignacio; Barber, Patricia; González López-Valcárcel, Beatriz; Jiménez, Octavio

    2014-01-01

    In Spain, official information on waiting times for surgery is based on the interval between the indication for surgery and its performance. We aimed to estimate total waiting times for surgical procedures, including outpatient visits and diagnostic tests prior to surgery. In addition, we propose an alternative system to manage total waiting times that reduces variability and maximum waiting times without increasing the use of health care resources. This system is illustrated by three surgical procedures: cholecystectomy, carpal tunnel release and inguinal/femoral hernia repair. Using data from two Autonomous Communities, we adjusted, through simulation, a theoretical distribution of the total waiting time assuming independence of the waiting times of each stage of the clinical procedure. We show an alternative system in which the waiting time for the second consultation is established according to the time previously waited for the first consultation. Average total waiting times for cholecystectomy, carpal tunnel release and inguinal/femoral hernia repair were 331, 355 and 137 days, respectively (official data are 83, 68 and 73 days, respectively). Using different negative correlations between waiting times for subsequent consultations would reduce maximum waiting times by between 2% and 15% and substantially reduce heterogeneity among patients, without generating higher resource use. Total waiting times are between two and five times higher than those officially published. The relationship between the waiting times at each stage of the medical procedure may be used to decrease variability and maximum waiting times. Copyright © 2013 SESPAS. Published by Elsevier Espana. All rights reserved.

  1. Waiting endurance time estimation of electric two-wheelers at signalized intersections.

    PubMed

    Huan, Mei; Yang, Xiao-bao

    2014-01-01

    The paper proposed a model for estimating waiting endurance times of electric two-wheelers at signalized intersections using survival analysis method. Waiting duration times were collected by video cameras and they were assigned as censored and uncensored data to distinguish between normal crossing and red-light running behavior. A Cox proportional hazard model was introduced, and variables revealing personal characteristics and traffic conditions were defined as covariates to describe the effects of internal and external factors. Empirical results show that riders do not want to wait too long to cross intersections. As signal waiting time increases, electric two-wheelers get impatient and violate the traffic signal. There are 12.8% of electric two-wheelers with negligible wait time. 25.0% of electric two-wheelers are generally nonrisk takers who can obey the traffic rules after waiting for 100 seconds. Half of electric two-wheelers cannot endure 49.0 seconds or longer at red-light phase. Red phase time, motor vehicle volume, and conformity behavior have important effects on riders' waiting times. Waiting endurance times would decrease with the longer red-phase time, the lower traffic volume, or the bigger number of other riders who run against the red light. The proposed model may be applicable in the design, management and control of signalized intersections in other developing cities.

  2. Waiting Endurance Time Estimation of Electric Two-Wheelers at Signalized Intersections

    PubMed Central

    Huan, Mei; Yang, Xiao-bao

    2014-01-01

    The paper proposed a model for estimating waiting endurance times of electric two-wheelers at signalized intersections using survival analysis method. Waiting duration times were collected by video cameras and they were assigned as censored and uncensored data to distinguish between normal crossing and red-light running behavior. A Cox proportional hazard model was introduced, and variables revealing personal characteristics and traffic conditions were defined as covariates to describe the effects of internal and external factors. Empirical results show that riders do not want to wait too long to cross intersections. As signal waiting time increases, electric two-wheelers get impatient and violate the traffic signal. There are 12.8% of electric two-wheelers with negligible wait time. 25.0% of electric two-wheelers are generally nonrisk takers who can obey the traffic rules after waiting for 100 seconds. Half of electric two-wheelers cannot endure 49.0 seconds or longer at red-light phase. Red phase time, motor vehicle volume, and conformity behavior have important effects on riders' waiting times. Waiting endurance times would decrease with the longer red-phase time, the lower traffic volume, or the bigger number of other riders who run against the red light. The proposed model may be applicable in the design, management and control of signalized intersections in other developing cities. PMID:24895659

  3. Outpatient Office Wait Times And Quality Of Care For Medicaid Patients.

    PubMed

    Oostrom, Tamar; Einav, Liran; Finkelstein, Amy

    2017-05-01

    The time patients spend in a doctor's waiting room prior to a scheduled appointment is an important component of the quality of the overall health care experience. We analyzed data on twenty-one million outpatient visits obtained from electronic health record systems, which allowed us to measure time spent in the waiting room beyond the scheduled appointment time. Median wait time was a little more than four minutes. Almost one-fifth of visits had waits longer than twenty minutes, and 10 percent were more than thirty minutes. Waits were shorter for early-morning appointments, for younger patients, and at larger practices. Median wait time was 4.1 minutes for privately insured patients and 4.6 minutes for Medicaid patients. After adjustment for patient and appointment characteristics, Medicaid patients were 20 percent more likely than the privately insured patients to wait longer than twenty minutes, with most of this disparity explained by differences in practices and providers they saw. Wait times for Medicaid patients relative to privately insured patients were longer in states with relatively lower Medicaid reimbursement rates. The study complements other work that suggests that Medicaid patients face some additional barriers in the receipt of care. Project HOPE—The People-to-People Health Foundation, Inc.

  4. Equity in specialist waiting times by socioeconomic groups: evidence from Spain.

    PubMed

    Abásolo, Ignacio; Negrín-Hernández, Miguel A; Pinilla, Jaime

    2014-04-01

    In countries with publicly financed health care systems, waiting time--rather than price--is the rationing mechanism for access to health care services. The normative statement underlying such a rationing device is that patients should wait according to need and irrespective of socioeconomic status or other non-need characteristics. The aim of this paper is to test empirically that waiting times for publicly funded specialist care do not depend on patients' socioeconomic status. Waiting times for specialist care can vary according to the type of medical specialty, type of consultation (review or diagnosis) and the region where patients' reside. In order to take into account such variability, we use Bayesian random parameter models to explain waiting times for specialist care in terms of need and non-need variables. We find that individuals with lower education and income levels wait significantly more time than their counterparts.

  5. The uncertainty room: strategies for managing uncertainty in a surgical waiting room.

    PubMed

    Stone, Anne M; Lammers, John C

    2012-01-01

    To describe experiences of uncertainty and management strategies for staff working with families in a hospital waiting room. A 288-bed, nonprofit community hospital in a Midwestern city. Data were collected during individual, semistructured interviews with 3 volunteers, 3 technical staff members, and 1 circulating nurse (n = 7), and during 40 hours of observation in a surgical waiting room. Interview transcripts were analyzed using constant comparative techniques. The surgical waiting room represents the intersection of several sources of uncertainty that families experience. Findings also illustrate the ways in which staff manage the uncertainty of families in the waiting room by communicating support. Staff in surgical waiting rooms are responsible for managing family members' uncertainty related to insufficient information. Practically, this study provided some evidence that staff are expected to help manage the uncertainty that is typical in a surgical waiting room, further highlighting the important role of communication in improving family members' experiences.

  6. Longer wait times affect future use of VHA primary care.

    PubMed

    Wong, Edwin S; Liu, Chuan-Fen; Hernandez, Susan E; Augustine, Matthew R; Nelson, Karin; Fihn, Stephan D; Hebert, Paul L

    2017-07-29

    Improving access to the Veterans Health Administration (VHA) is a high priority, particularly given statutory mandates of the Veterans Access, Choice and Accountability Act. This study examined whether patient-reported wait times for VHA appointments were associated with future reliance on VHA primary care services. This observational study examined 13,595 VHA patients dually enrolled in fee-for-service Medicare. Data sources included VHA administrative data, Medicare claims and the Survey of Healthcare Experiences of Patients (SHEP). Primary care use was defined as the number of face-to-face visits from VHA and Medicare in the 12 months following SHEP completion. VHA reliance was defined as the number of VHA visits divided by total visits (VHA+Medicare). Wait times were derived from SHEP responses measuring the usual number of days to a VHA appointment with patients' primary care provider for those seeking immediate care. We defined appointment wait times categorically: 0 days, 1day, 2-3 days, 4-7 days and >7 days. We used fractional logistic regression to examine the relationship between wait times and reliance. Mean VHA reliance was 88.1% (95% CI = 86.7% to 89.5%) for patients reporting 0day waits. Compared with these patients, reliance over the subsequent year was 1.4 (p = 0.041), 2.8 (p = 0.001) and 1.6 (p = 0.014) percentage points lower for patients waiting 2-3 days, 4-7 days and >7 days, respectively. Patients reporting longer usual wait times for immediate VHA care exhibited lower future reliance on VHA primary care. Longer wait times may reduce care continuity and impact cost shifting across two federal health programs. Copyright © 2017. Published by Elsevier Inc.

  7. Preclinic group education sessions reduce waiting times and costs at public pain medicine units.

    PubMed

    Davies, Stephanie; Quintner, John; Parsons, Richard; Parkitny, Luke; Knight, Paul; Forrester, Elizabeth; Roberts, Mary; Graham, Carl; Visser, Eric; Antill, Tracy; Packer, Tanya; Schug, Stephan A

    2011-01-01

    To assess the effects of preclinic group education sessions and system redesign on tertiary pain medicine units and patient outcomes. Prospective cohort study. Two public hospital multidisciplinary pain medicine units. People with persistent pain. A system redesign from a "traditional" model (initial individual medical appointments) to a model that delivers group education sessions prior to individual appointments. Based on Patient Triage Questionnaires patients were scheduled to attend Self-Training Educative Pain Sessions (STEPS), a two day eight hour group education program, followed by optional patient-initiated clinic appointments. Number of patients completing STEPS who subsequently requested individual outpatient clinic appointment(s); wait-times; unit cost per new patient referred; recurrent health care utilization; patient satisfaction; Global Perceived Impression of Change (GPIC); and utilized pain management strategies. Following STEPS 48% of attendees requested individual outpatient appointments. Wait times reduced from 105.6 to 16.1 weeks at one pain unit and 37.3 to 15.2 weeks at the second. Unit cost per new patient appointed reduced from $1,805 Australian Dollars (AUD) to AUD$541 (for STEPS). At 3 months, patients scored their satisfaction with "the treatment received for their pain" more positively than at baseline (change score=0.88; P=0.0003), GPIC improved (change score=0.46; P<0.0001) and mean number of active strategies utilized increased by 4.12 per patient (P=0.0004). The introduction of STEPS was associated with reduced wait-times and costs at public pain medicine units and increased both the use of active pain management strategies and patient satisfaction. Wiley Periodicals, Inc.

  8. Shorter Perceived Outpatient MRI Wait Times Associated With Higher Patient Satisfaction.

    PubMed

    Holbrook, Anna; Glenn, Harold; Mahmood, Rabia; Cai, Qingpo; Kang, Jian; Duszak, Richard

    2016-05-01

    The aim of this study was to assess differences in perceived versus actual wait times among patients undergoing outpatient MRI examinations and to correlate those times with patient satisfaction. Over 15 weeks, 190 patients presenting for outpatient MR in a radiology department in which "patient experience" is one of the stated strategic priorities were asked to (1) estimate their wait times for various stages in the imaging process and (2) state their satisfaction with their imaging experience. Perceived times were compared with actual electronic time stamps. Perceived and actual times were compared and correlated with standardized satisfaction scores using Kendall τ correlation. The mean actual wait time between patient arrival and examination start was 53.4 ± 33.8 min, whereas patients perceived a mean wait time of 27.8 ± 23.1 min, a statistically significant underestimation of 25.6 min (P < .001). Both shorter actual and perceived wait times at all points during patient encounters were correlated with higher satisfaction scores (P < .001). Patients undergoing outpatient MR examinations in an environment designed to optimize patient experience underestimated wait times at all points during their encounters. Shorter perceived and actual wait times were both correlated with higher satisfaction scores. As satisfaction surveys play a larger role in an environment of metric transparency and value-based payments, better understanding of such factors will be increasingly important. Copyright © 2016 American College of Radiology. Published by Elsevier Inc. All rights reserved.

  9. Indication criteria for cataract extraction and gender differences in waiting time.

    PubMed

    Smirthwaite, Goldina; Lundström, Mats; Albrecht, Susanne; Swahnberg, Katarina

    2014-08-01

    The purpose of this study was to investigate national indication criteria tool for cataract extraction (NIKE), a clinical tool for establishing levels of indications for cataract surgery, in relation to gender differences in waiting times for cataract extraction (CE). Data were collected by The Swedish National Cataract Register (NCR). Eye clinics report to NCR voluntarily and on regular basis (98% coverage). Comparisons regarding gender difference in waiting times were performed between NIKE-categorized and non-NIKE-categorized patients, as well as between different indication groups within the NIKE-system. All calculations were performed in spss version 20. Multivariate analyses were carried out using logistic regression, and single variable analyses were carried out by Student's t-test or chi square as appropriate. Gender, age, visual acuity and NIKE-categorization were associated with waiting time. Female patients had a longer waiting time to CE than male, both within and outside the NIKE-system. Gender difference in waiting time was somewhat larger among patients who had not been categorized by NIKE. In the non-NIKE-categorized group, women waited 0.20 months longer than men. In the group which was NIKE-categorized, women waited 0.18 months longer than men. It is reasonable to assume that prioritizing patients by means of NIKE helps to reduce the gender differences in waiting time. Gender differences in waiting time have decreased as NIKE was introduced and there may be a variety of explanations for this. However, with the chosen study design, we could not distinguish between effects related to NIKE and those due to other factors which occurred during the study period. © 2013 Acta Ophthalmologica Scandinavica Foundation. Published by John Wiley & Sons Ltd.

  10. Waiting for thyroid surgery: a study of psychological morbidity and determinants of health associated with long wait times for thyroid surgery.

    PubMed

    Eskander, Antoine; Devins, Gerald M; Freeman, Jeremy; Wei, Alice C; Rotstein, Lorne; Chauhan, Nitin; Sawka, Anna M; Brown, Dale; Irish, Jonathan; Gilbert, Ralph; Gullane, Patrick; Higgins, Kevin; Enepekides, Danny; Goldstein, David

    2013-02-01

    Patients with thyroid pathology tend have longer surgical wait times. Uncertainty during this wait can have negative psychologically impact. This study aims to determine the degree of psychological morbidity in patients waiting for thyroid surgery. Prospectively assessing patients pre- and postoperative psychological morbidity (level 2c). Patients waiting for thyroidectomy were mailed a sociodemographic and four psychological morbidity questionnaires: Impact of Events Scale-Revised (IES-R), Illness Intrusiveness Ratings Scale (IIRS), Perceived Stress Scale (PSS) and Hospital Anxiety and Depression Scale (HADS). We assessed whether anxiety was related to length of wait and a number of clinical/sociodemographic factors. We achieved a 53% response rate over a 3-year period, with 176 patients providing complete preoperative data; and 74 (42%) completed postoperative data. The average age was 53 (± 12) years; 82% were female. Respondents with a suspicious or known malignancy waited an average of 107 days while those with benign neoplastic biopsies waited an average of 218 days for thyroidectomy. Respondents reported substantial psychological morbidity with high IES-R, IIRS, PSS, and HADS scores. There was no significant association between psychological morbidity and wait times, clinical or sociodemographic factors. Postoperative anxiety decreased significantly in all psychological morbidity measures except for the IIRS. Patients waiting for thyroid surgery have mild to moderate psychological morbidity and long wait times for surgery. These appear not to be related. Psychological morbidity decreases after surgery. Reducing wait time can potentially reduce the time that patients have to live with unnecessary stress and anxiety. Copyright © 2012 The American Laryngological, Rhinological, and Otological Society, Inc.

  11. Waiting times in the ambulatory sector--the case of chronically ill patients.

    PubMed

    Sundmacher, Leonie; Kopetsch, Thomas

    2013-09-10

    First, the influence of determinants on the waiting times of chronically ill patients in the ambulatory sector is investigated. The determinants are subdivided into four groups: (1) need, (2) socio-economic factors, (3) health system and (4) patient time pressures. Next, the influence of waiting times on the annual number of consultations is examined to assess whether the existing variation in waiting times influences the frequency of medical examinations. The waiting times of chronically ill patients are analysed since regular ambulatory care for this patient group could both improve treatment outcomes and lower costs. Individual data from the 2010 Representative Survey conducted by the National Association of Statutory Health Insurance Physicians (KBV) together with regional data from the Federal Office of Construction and Regional Planning. This is a retrospective observational study. The dependent variables are waiting times in the ambulatory sector and the number of consultations of General Practitioners (GPs) and specialist physicians in the year 2010. The explanatory variables of interest are 'need' and 'health system' in the first model and 'length of waiting times' in the second. Negative binomial models with random effects are used to estimate the incidence rate ratios of increased waiting times and number of consultations. Subsequently, the models are stratified by urban and rural areas. In the pooled regression the factor 'privately insured' shortens the waiting time for treatment by a specialist by approximately 28% (about 3 days) in comparison with members of the statutory health insurance system. The category of insurance has no influence on the number of consultations of GPs. In addition, the regression results stratified by urban and rural areas show that in urban areas the factor 'privately insured' reduces the waiting time for specialists by approximately 35% (about 3.3 days) while in rural areas there is no evidence of statistical influence. In

  12. Wait-Time and Multiple Representation Levels in Chemistry Lessons

    ERIC Educational Resources Information Center

    Li, Winnie Sim Siew; Arshad, Mohammad Yusof

    2014-01-01

    Wait-time is an important aspect in a teaching and learning process, especially after the teacher has posed questions to students, as it is one of the factors in determining quality of students' responses. This article describes the practices of wait-time one after teacher's questions at multiple representation levels among twenty three chemistry…

  13. [Waiting list in general and digestive surgery: patient expectations, quality of life during waiting time and overall satisfaction].

    PubMed

    Parés, D; Duran, E; Hermoso, J; Comajuncosas, J; Gris, P; Lopez-Negre, J L; Urgellés, J; Orbeal, R; Vallverdú, H; Jimeno, J

    2013-01-01

    The structural resources of the National Health system are limited, and therefore early surgery cannot be performed on all patients. The objective was to analyse the satisfaction perceived by the patient as regards the delay of treatment by waiting list of three types of surgery. The influence of expectations on waiting times, and impaired quality of life due to the clinical symptoms during the delay, were studied. A prospective study was conducted using a postal questionnaire. We compared the expectations (scale of 1 to 5), the impact on quality of life for symptoms (scale of 1 to 5) and the level of patient satisfaction (scale of 1 to 5) with respect to time on the waitng list for cholelithiasis, inguinal hernia and haemorrhoids. The predictors of patient dissatisfaction were analysed. A total of 57 patients were included. When comparing the characteristics of patients with and without satisfaction over time on the waiting list, days on the waiting list (P=.044), the change in the quality of life due to the symptoms (P=.028), and expectations (P<.001) were significantly different between the two groups. In the multivariate analysis, the expectation was associated with patient dissatisfaction as regards the time on waiting list (OR: 3.14 95% CI: 5.91 to 220.73, P<.001). The level of patient dissatisfaction is associated with expectations about time in waiting list. Copyright © 2012 SECA. Published by Elsevier Espana. All rights reserved.

  14. Heart Surgery Waiting Time: Assessing the Effectiveness of an Action.

    PubMed

    Badakhshan, Abbas; Arab, Mohammad; Gholipour, Mahin; Behnampour, Naser; Saleki, Saeid

    2015-08-01

    Waiting time is an index assessing patient satisfaction, managerial effectiveness and horizontal equity in providing health care. Although heart surgery centers establishment is attractive for politicians. They are always faced with the question of to what extent they solve patient's problems. The objective of this study was to evaluate factors influencing waiting time in patients of heart surgery centers, and to make recommendations for health-care policy-makers for reducing waiting time and increasing the quality of services from this perspective. This cross-sectional study was performed in 2013. After searching articles on PubMed, Elsevier, Google Scholar, Ovid, Magiran, IranMedex, and SID, a list of several criteria, which relate to waiting time, was provided. Afterwards, the data on waiting time were collected by a researcher-structured checklist from 156 hospitalized patients. The data were analyzed by SPSS 16. The Kolmogorov Smirnov and Shapiro tests were used for determination of normality. Due to the non-normal distribution, non-parametric tests, such as Kruskal-Wallis and Mann-Whitney were chosen for reporting significance. Parametric tests also used reporting medians. Among the studied variables, just economic status had a significant relation with waiting time (P = 0.37). Fifty percent of participants had diabetes, whereas this estimate was 43.58% for high blood pressure. As the cause of delay, 28.2% of patients reported financial problems, 18.6% personal problem and 13.5% a delay in providing equipment by the hospital. It seems the studied hospital should review its waiting time arrangements and detach them, as far as possible, from subjective and personal (specialists) decisions. On the other hand, ministries of health and insurance companies should consider more financial support. It is also recommend that hospitals should arrange preoperational psychiatric consultation for increasing patients' emotionally readiness.

  15. Lung cancer care trajectory at a Canadian centre: an evaluation of how wait times affect clinical outcomes.

    PubMed

    Kasymjanova, G; Small, D; Cohen, V; Jagoe, R T; Batist, G; Sateren, W; Ernst, P; Pepe, C; Sakr, L; Agulnik, J

    2017-10-01

    Lung cancer continues to be one of the most common cancers in Canada, with approximately 28,400 new cases diagnosed each year. Although timely care can contribute substantially to quality of life for patients, it remains unclear whether it also improves patient outcomes. In this work, we used a set of quality indicators that aim to describe the quality of care in lung cancer patients. We assessed adherence with existing guidelines for timeliness of lung cancer care and concordance with existing standards of treatment, and we examined the association between timeliness of care and lung cancer survival. Patients with lung cancer diagnosed between 2010 and 2015 were identified from the Pulmonary Division Lung Cancer Registry at our centre. We demonstrated that the interdisciplinary pulmonary oncology service successfully treated most of its patients within the recommended wait times. However, there is still work to be done to decrease variation in wait time. Our results demonstrate a significant association between wait time and survival, supporting the need for clinicians to optimize the patient care trajectory. It would be helpful for Canadian clinicians treating patients with lung cancer to have wait time guidelines for all treatment modalities, together with standard definitions for all time intervals. Any reductions in wait times should be balanced against the need for thorough investigation before initiating treatment. We believe that our unique model of care leads to an acceleration of diagnostic steps. Avoiding any delay associated with referral to a medical oncologist for treatment could be an acceptable strategy with respect to reducing wait time.

  16. SU-F-P-20: Predicting Waiting Times in Radiation Oncology Using Machine Learning

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

    Joseph, A; Herrera, D; Hijal, T

    Purpose: Waiting times remain one of the most vexing patient satisfaction challenges facing healthcare. Waiting time uncertainty can cause patients, who are already sick or in pain, to worry about when they will receive the care they need. These waiting periods are often difficult for staff to predict and only rough estimates are typically provided based on personal experience. This level of uncertainty leaves most patients unable to plan their calendar, making the waiting experience uncomfortable, even painful. In the present era of electronic health records (EHRs), waiting times need not be so uncertain. Extensive EHRs provide unprecedented amounts ofmore » data that can statistically cluster towards representative values when appropriate patient cohorts are selected. Predictive modelling, such as machine learning, is a powerful approach that benefits from large, potentially complex, datasets. The essence of machine learning is to predict future outcomes by learning from previous experience. The application of a machine learning algorithm to waiting time data has the potential to produce personalized waiting time predictions such that the uncertainty may be removed from the patient’s waiting experience. Methods: In radiation oncology, patients typically experience several types of waiting (eg waiting at home for treatment planning, waiting in the waiting room for oncologist appointments and daily waiting in the waiting room for radiotherapy treatments). A daily treatment wait time model is discussed in this report. To develop a prediction model using our large dataset (with more than 100k sample points) a variety of machine learning algorithms from the Python package sklearn were tested. Results: We found that the Random Forest Regressor model provides the best predictions for daily radiotherapy treatment waiting times. Using this model, we achieved a median residual (actual value minus predicted value) of 0.25 minutes and a standard deviation residual of 6

  17. Decreasing Wait Times and Increasing Patient Satisfaction: A Lean Six Sigma Approach.

    PubMed

    Godley, Mary; Jenkins, Jeanne B

    2018-06-08

    Patient satisfaction scores in the vascular interventional radiology department were low, especially related to wait times in registration and for tests/treatments, with low scores for intentions to recommend. The purpose of our quality improvement project was to decrease wait times and improve patient satisfaction using Lean Six Sigma's define, measure, analyze, improve, and control (DMAIC) framework with a pre-/postintervention design. There was a statistically significant decrease in wait times (P < .0019) and an increase in patient satisfaction scores in 3 areas: registration wait times (from 17 to 99 percentiles), test/treatment (from 19 to 60 percentiles), and likelihood to recommend (from 6 to 97 percentiles). Lean Six Sigma was an effective framework for use in decreasing wait times and improving patient satisfaction.

  18. Emergency department waiting times: Do the raw data tell the whole story?

    PubMed

    Green, Janette; Dawber, James; Masso, Malcolm; Eagar, Kathy

    2014-02-01

    To determine whether there are real differences in emergency department (ED) performance between Australian states and territories. Cross-sectional analysis of 2009-10 attendances at an ED contributing to the Australian non-admitted patient ED care database. The main outcome measure was difference in waiting time across triage categories. There were more than 5.8 million ED attendances. Raw ED waiting times varied by a range of factors including jurisdiction, triage category, geographic location and hospital peer group. All variables were significant in a model designed to test the effect of jurisdiction on ED waiting times, including triage category, hospital peer group, patient socioeconomic status and patient remoteness. When the interaction between triage category and jurisdiction entered the model, it was found to have a significant effect on ED waiting times (P<0.001) and triage was also significant (P<0.001). Jurisdiction was no longer statistically significant (P=0.248 using all triage categories and 0.063 using only Australian Triage Scale 2 and 3). Although the Council of Australian Governments has adopted raw measures for its key ED performance indicators, raw waiting time statistics are misleading. There are no consistent differences in ED waiting times between states and territories after other factors are accounted for. WHAT IS KNOWN ABOUT THE TOPIC? The length of time patients wait to be treated after presenting at an ED is routinely used to measure ED performance. In national health agreements with the federal government, each state and territory in Australia is expected to meet waiting time performance targets for the five ED triage categories. The raw data indicate differences in performance between states and territories. WHAT DOES THIS PAPER ADD? Measuring ED performance using raw data gives misleading results. There are no consistent differences in ED waiting times between the states and territories after other factors are taken into account

  19. Toward systematic reviews to understand the determinants of wait time management success to help decision-makers and managers better manage wait times

    PubMed Central

    2013-01-01

    Background Long waits for core specialized services have consistently been identified as a key barrier to access. Governments and organizations at all levels have responded with strategies for better wait list management. While these initiatives are promising, insufficient attention has been paid to factors influencing the implementation and sustainability of wait time management strategies (WTMS) implemented at the organizational level. Methods A systematic review was conducted using the main electronic databases, such as CINAHL, MEDLINE, and Cochrane Database of Systematic Reviews, to identify articles published between 1990 and 2011 on WTMS for scheduled care implemented at the organizational level or higher and on frameworks for analyzing factors influencing their success. Data was extracted on governance, culture, resources, and tools. We organized a workshop with Canadian healthcare policy-makers and managers to compare our initial findings with their experience. Results Our systematic review included 47 articles: 36 related to implementation and 11 to sustainability. From these, we identified a variety of WTMS initiated at the organizational level or higher, and within these, certain factors that were specific to either implementation or sustainability and others common to both. The main common factors influencing success at the contextual level were stakeholder engagement and strong funding, and at the organizational level, physician involvement, human resources capacity, and information management systems. Specific factors for successful implementation at the contextual level were consultation with front-line actors and common standards and guidelines, and at the organizational level, financial incentives and dedicated staffing. For sustainability, we found no new factors. The workshop participants identified the same major factors as found in the articles and added others, such as information sharing between physicians and managers. Conclusions Factors

  20. Toward systematic reviews to understand the determinants of wait time management success to help decision-makers and managers better manage wait times.

    PubMed

    Pomey, Marie-Pascale; Forest, Pierre-Gerlier; Sanmartin, Claudia; Decoster, Carolyn; Clavel, Nathalie; Warren, Elaine; Drew, Madeleine; Noseworthy, Tom

    2013-06-06

    Long waits for core specialized services have consistently been identified as a key barrier to access. Governments and organizations at all levels have responded with strategies for better wait list management. While these initiatives are promising, insufficient attention has been paid to factors influencing the implementation and sustainability of wait time management strategies (WTMS) implemented at the organizational level. A systematic review was conducted using the main electronic databases, such as CINAHL, MEDLINE, and Cochrane Database of Systematic Reviews, to identify articles published between 1990 and 2011 on WTMS for scheduled care implemented at the organizational level or higher and on frameworks for analyzing factors influencing their success. Data was extracted on governance, culture, resources, and tools. We organized a workshop with Canadian healthcare policy-makers and managers to compare our initial findings with their experience. Our systematic review included 47 articles: 36 related to implementation and 11 to sustainability. From these, we identified a variety of WTMS initiated at the organizational level or higher, and within these, certain factors that were specific to either implementation or sustainability and others common to both. The main common factors influencing success at the contextual level were stakeholder engagement and strong funding, and at the organizational level, physician involvement, human resources capacity, and information management systems. Specific factors for successful implementation at the contextual level were consultation with front-line actors and common standards and guidelines, and at the organizational level, financial incentives and dedicated staffing. For sustainability, we found no new factors. The workshop participants identified the same major factors as found in the articles and added others, such as information sharing between physicians and managers. Factors related to implementation were studied

  1. Waiting time for radiotherapy in women with cervical cancer

    PubMed Central

    do Nascimento, Maria Isabel; Azevedo e Silva, Gulnar

    2016-01-01

    ABSTRACT OBJECTIVE To describe the waiting time for radiotherapy for patients with cervical cancer. METHODS This descriptive study was conducted with 342 cervical cancer cases that were referred to primary radiotherapy, in the Baixada Fluminense region, RJ, Southeastern Brazil, from October 1995 to August 2010. The waiting time was calculated using the recommended 60-day deadline as a parameter to obtaining the first cancer treatment and considering the date at which the diagnosis was confirmed, the date of first oncological consultation and date when the radiotherapy began. Median and proportional comparisons were made using the Kruskal Wallis and Chi-square tests. RESULTS Most of the women (72.2%) began their radiotherapy within 60 days from the diagnostic confirmation date. The median of this total waiting time was 41 days. This median worsened over the time period, going from 11 days (1995-1996) to 64 days (2009-2010). The median interval between the diagnostic confirmation and the first oncological consultation was 33 days, and between the first oncological consultation and the first radiotherapy session was four days. The median waiting time differed significantly (p = 0.003) according to different stages of the tumor, reaching 56 days, 35 days and 30 days for women whose cancers were classified up to IIA; from IIB to IIIB, and IVA-IVB, respectively. CONCLUSIONS Despite most of the women having had access to radiotherapy within the recommended 60 days, the implementation of procedures to define the stage of the tumor and to reestablish clinical conditions took a large part of this time, showing that at least one of these intervals needs to be improved. Even though the waiting times were ideal for all patients, the most advanced cases were quickly treated, which suggests that access to radiotherapy by women with cervical cancer has been reached with equity. PMID:26786473

  2. The relationship between educational attainment and waiting time among the elderly in Norway.

    PubMed

    Carlsen, Fredrik; Kaarboe, Oddvar Martin

    2015-11-01

    We investigate whether educational attainment affects waiting time of elderly patients in somatic hospitals. We consider three distinct pathways; that patients with different educational attainment have different disease patterns, that patients with different levels of education receive treatments at different hospitals, and that patient choice and supply of local health services within hospital catchment areas explain unequal waiting time of different educational groups. We find evidence of an educational gradient in waiting time for male patients, but not for female patients. Conditional on age, male patients with tertiary education wait 45% shorter than male patients with secondary or primary education. The first pathway is not quantitatively important as controlling for disease patters has little effect on relative waiting times. The second pathway is important. Relative to patients with primary education, variation in waiting time and education level across local hospitals contributes to higher waiting time for male patients with secondary education and female patients with secondary or tertiary education and lower waiting time for male patients with tertiary education. These effects are in the order of 15-20%. The third pathway is also quantitatively important. The educational gradients within catchment areas disappear when we control for travel distance and supply of private specialists. Copyright © 2015 Elsevier Ireland Ltd. All rights reserved.

  3. The waiting time problem in a model hominin population.

    PubMed

    Sanford, John; Brewer, Wesley; Smith, Franzine; Baumgardner, John

    2015-09-17

    Functional information is normally communicated using specific, context-dependent strings of symbolic characters. This is true within the human realm (texts and computer programs), and also within the biological realm (nucleic acids and proteins). In biology, strings of nucleotides encode much of the information within living cells. How do such information-bearing nucleotide strings arise and become established? This paper uses comprehensive numerical simulation to understand what types of nucleotide strings can realistically be established via the mutation/selection process, given a reasonable timeframe. The program Mendel's Accountant realistically simulates the mutation/selection process, and was modified so that a starting string of nucleotides could be specified, and a corresponding target string of nucleotides could be specified. We simulated a classic pre-human hominin population of at least 10,000 individuals, with a generation time of 20 years, and with very strong selection (50% selective elimination). Random point mutations were generated within the starting string. Whenever an instance of the target string arose, all individuals carrying the target string were assigned a specified reproductive advantage. When natural selection had successfully amplified an instance of the target string to the point of fixation, the experiment was halted, and the waiting time statistics were tabulated. Using this methodology we tested the effect of mutation rate, string length, fitness benefit, and population size on waiting time to fixation. Biologically realistic numerical simulations revealed that a population of this type required inordinately long waiting times to establish even the shortest nucleotide strings. To establish a string of two nucleotides required on average 84 million years. To establish a string of five nucleotides required on average 2 billion years. We found that waiting times were reduced by higher mutation rates, stronger fitness benefits, and

  4. [Waiting time for the first colposcopic examination in women with abnormal Papanicolaou test].

    PubMed

    Nascimento, Maria Isabel do; Rabelo, Irene Machado Moraes Alvarenga; Cardoso, Fabrício Seabra Polidoro; Musse, Ricardo Neif Vieira

    2015-08-01

    To evaluate the waiting times before obtaining the first colposcopic examination for women with abnormal Papanicolaou smears. Retrospective cohort study conducted on patients who required a colposcopic examination to clarify an abnormal pap test, between 2002 January and 2008 August, in a metropolitan region of Brazil. The waiting times were defined as: Total Waiting Time (interval between the date of the pap test result and the date of the first colposcopic examination); Partial A Waiting Time (interval between the date of the pap test result and the date of referral); Partial B Waiting Time (interval between the date of referral and the date of the first colposcopic examination). Means, medians, relative and absolute frequencies were calculated. The Kruskal-Wallis test and Pearson's chi-square test were used to determine statistical significance. A total of 1,544 women with mean of age of 34 years (SD=12.6 years) were analyzed. Most of them had access to colposcopic examination within 30 days (65.8%) or 60 days (92.8%) from referral. Mean Total Waiting Time, Partial A Waiting Time, and Partial B Waiting Time were 94.5 days (SD=96.8 days), 67.8 days (SD=95.3 days) and 29.2 days (SD=35.1 days), respectively. A large part of the women studied had access to colposcopic examination within 60 days after referral, but Total waiting time was long. Measures to reduce the waiting time for obtaining the first colposcopic examination can help to improve the quality of care in the context of cervical cancer control in the region, and ought to be addressed at the phase between the date of the pap test results and the date of referral to the teaching hospital.

  5. Outpatient clinic waiting time, provider communication styles and satisfaction with healthcare in India.

    PubMed

    Mehra, Payal

    2016-08-08

    Purpose - The purpose of this paper is to evaluate the impact of extended waiting time on patients' perceptions of provider communication skills and in-clinic satisfaction, in three major cities in India. Design/methodology/approach - In total, 625 patients were interviewed. The multivariate general linear model was used to determine the causality and relationship between the independent and the dependent variable. A moderation analysis was also conducted to assess waiting time role as a potential moderator in doctor-patient communication. Findings - Results show that patients with higher waiting time were less satisfied with health care quality. Male patients and patients of male providers were more affected by extended waiting time than female patients and patients of female providers. The advanced regression analysis, however, suggests weak support for waiting time and its effect on overall satisfaction with clinic quality. Waiting time did not moderate the relationship between satisfaction with dominant communication style, and overall satisfaction at the outpatient clinic. Research limitations/implications - A cross-sectional study does not easily lend itself to explaining causality with certainty. Thus, sophisticated techniques, such as structural equation modelling may also be utilized to assess the influence of extended waiting time on satisfaction with healthcare at outpatient clinics. Practical implications - Findings are relevant for providers as the onus is on them to ensure patient satisfaction. They should initiate a workable waiting time assessment model at the operational level. Originality/value - There has been a relatively lesser focus on patient waiting time in patient-provider satisfaction studies. In India, this aspect is still vastly unexplored especially in the context of outpatient clinics. Gender wise pattern of patient satisfaction and waiting time is also missing in most studies.

  6. Associations Between Waiting Times, Service Times, and Patient Satisfaction in an Endocrinology Outpatient Department: A Time Study and Questionnaire Survey

    PubMed Central

    Xie, Zhenzhen; Or, Calvin

    2017-01-01

    The issue of long patient waits has attracted increasing public attention due to the negative effects of waiting on patients’ satisfaction with health care. The present study examined the associations between actual waiting time, perceived acceptability of waiting time, actual service time, perceived acceptability of service time, actual visit duration, and the level of patient satisfaction with care. We conducted a cross-sectional time study and questionnaire survey of endocrinology outpatients visiting a major teaching hospital in China. Our results show that actual waiting time was negatively associated with patient satisfaction regarding several aspects of the care they received. Also, patients who were less satisfied with the sociocultural atmosphere and the identity-oriented approach to their care tended to perceive the amounts of time they spent waiting and receiving care as less acceptable. It is not always possible to prevent dissatisfaction with waiting, or to actually reduce waiting times by increasing resources such as increased staffing. However, several improvements in care services can be considered. Our suggestions include providing clearer, more transparent information to keep patients informed about the health care services that they may receive, and the health care professionals who are responsible for those services. We also suggest that care providers are encouraged to continue to show empathy and respect for patients, that patients are provided with private areas where they can talk with health professionals and no one can overhear, and that hospital staff treat the family members or friends who accompany patients in a courteous and friendly way. PMID:29161947

  7. Associations Between Waiting Times, Service Times, and Patient Satisfaction in an Endocrinology Outpatient Department: A Time Study and Questionnaire Survey.

    PubMed

    Xie, Zhenzhen; Or, Calvin

    2017-01-01

    The issue of long patient waits has attracted increasing public attention due to the negative effects of waiting on patients' satisfaction with health care. The present study examined the associations between actual waiting time, perceived acceptability of waiting time, actual service time, perceived acceptability of service time, actual visit duration, and the level of patient satisfaction with care. We conducted a cross-sectional time study and questionnaire survey of endocrinology outpatients visiting a major teaching hospital in China. Our results show that actual waiting time was negatively associated with patient satisfaction regarding several aspects of the care they received. Also, patients who were less satisfied with the sociocultural atmosphere and the identity-oriented approach to their care tended to perceive the amounts of time they spent waiting and receiving care as less acceptable. It is not always possible to prevent dissatisfaction with waiting, or to actually reduce waiting times by increasing resources such as increased staffing. However, several improvements in care services can be considered. Our suggestions include providing clearer, more transparent information to keep patients informed about the health care services that they may receive, and the health care professionals who are responsible for those services. We also suggest that care providers are encouraged to continue to show empathy and respect for patients, that patients are provided with private areas where they can talk with health professionals and no one can overhear, and that hospital staff treat the family members or friends who accompany patients in a courteous and friendly way.

  8. Effect of Lean Processes on Surgical Wait Times and Efficiency in a Tertiary Care Veterans Affairs Medical Center.

    PubMed

    Valsangkar, Nakul P; Eppstein, Andrew C; Lawson, Rick A; Taylor, Amber N

    2017-01-01

    There are an increasing number of veterans in the United States, and the current delay and wait times prevent Veterans Affairs institutions from fully meeting the needs of current and former service members. Concrete strategies to improve throughput at these facilities have been sparse. To identify whether lean processes can be used to improve wait times for surgical procedures in Veterans Affairs hospitals. Databases in the Veterans Integrated Service Network 11 Data Warehouse, Veterans Health Administration Support Service Center, and Veterans Information Systems and Technology Architecture/Dynamic Host Configuration Protocol were queried to assess changes in wait times for elective general surgical procedures and clinical volume before, during, and after implementation of lean processes over 3 fiscal years (FYs) at a tertiary care Veterans Affairs medical center. All patients evaluated by the general surgery department through outpatient clinics, clinical video teleconferencing, and e-consultations from October 2011 through September 2014 were included. Patients evaluated through the emergency department or as inpatient consults were excluded. The surgery service and systems redesign service held a value stream analysis in FY 2013, culminating in multiple rapid process improvement workshops. Multidisciplinary teams identified systemic inefficiencies and strategies to improve interdepartmental and patient communication to reduce canceled consultations and cases, diagnostic rework, and no-shows. High-priority triage with enhanced operating room flexibility was instituted to reduce scheduling wait times. General surgery department pilot projects were then implemented mid-FY 2013. Planned outcome measures included wait time, clinic and telehealth volume, number of no-shows, and operative volume. Paired t tests were used to identify differences in outcome measures after the institution of reforms. Following rapid process improvement workshop project rollouts, mean

  9. General practice cooperatives: long waiting times for home visits due to long distances?

    PubMed Central

    Giesen, Paul; van Lin, Nieke; Mokkink, Henk; van den Bosch, Wil; Grol, Richard

    2007-01-01

    Background The introduction of large-scale out-of-hours GP cooperatives has led to questions about increased distances between the GP cooperatives and the homes of patients and the increasing waiting times for home visits in urgent cases. We studied the relationship between the patient's waiting time for a home visit and the distance to the GP cooperative. Further, we investigated if other factors (traffic intensity, home visit intensity, time of day, and degree of urgency) influenced waiting times. Methods Cross-sectional study at four GP cooperatives. We used variance analysis to calculate waiting times for various categories of traffic intensity, home visit intensity, time of day, and degree of urgency. We used multiple logistic regression analysis to calculate to what degree these factors affected the ability to meet targets in urgent cases. Results The average waiting time for 5827 consultations was 30.5 min. Traffic intensity, home visit intensity, time of day and urgency of the complaint all seemed to affect waiting times significantly. A total of 88.7% of all patients were seen within 1 hour. In the case of life-threatening complaints (U1), 68.8% of the patients were seen within 15 min, and 95.6% of those with acute complaints (U2) were seen within 1 hour. For patients with life-threatening complaints (U1) the percentage of visits that met the time target of 15 minuts decreased from 86.5% (less than 2.5 km) to 16.7% (equals or more than 20 km). Discussion and conclusion Although home visits waiting times increase with increasing distance from the GP cooperative, it appears that traffic intensity, home visit intensity, and urgency also influence waiting times. For patients with life-threatening complaints waiting times increase sharply with the distance. PMID:17295925

  10. General practice cooperatives: long waiting times for home visits due to long distances?

    PubMed

    Giesen, Paul; van Lin, Nieke; Mokkink, Henk; van den Bosch, Wil; Grol, Richard

    2007-02-12

    The introduction of large-scale out-of-hours GP cooperatives has led to questions about increased distances between the GP cooperatives and the homes of patients and the increasing waiting times for home visits in urgent cases. We studied the relationship between the patient's waiting time for a home visit and the distance to the GP cooperative. Further, we investigated if other factors (traffic intensity, home visit intensity, time of day, and degree of urgency) influenced waiting times. Cross-sectional study at four GP cooperatives. We used variance analysis to calculate waiting times for various categories of traffic intensity, home visit intensity, time of day, and degree of urgency. We used multiple logistic regression analysis to calculate to what degree these factors affected the ability to meet targets in urgent cases. The average waiting time for 5827 consultations was 30.5 min. Traffic intensity, home visit intensity, time of day and urgency of the complaint all seemed to affect waiting times significantly. A total of 88.7% of all patients were seen within 1 hour. In the case of life-threatening complaints (U1), 68.8% of the patients were seen within 15 min, and 95.6% of those with acute complaints (U2) were seen within 1 hour. For patients with life-threatening complaints (U1) the percentage of visits that met the time target of 15 minutes decreased from 86.5% (less than 2.5 km) to 16.7% (equals or more than 20 km). Although home visits waiting times increase with increasing distance from the GP cooperative, it appears that traffic intensity, home visit intensity, and urgency also influence waiting times. For patients with life-threatening complaints waiting times increase sharply with the distance.

  11. [Improving the CMP appointment waiting time for children and adolescents].

    PubMed

    Cani, Pascale

    2014-01-01

    The increasing activity of mental health centres for children and adolescents and longer waiting times in obtaining a first appointment have led an area of child psychiatry to question the organisation of new consultation applications. Two CMP in the sector had a waiting period of over 40 days for half of the patients. Two improvement actions were implemented:the implementation of organisation and reception nurses and the development of a new applications management process. The evaluation after one year showed a decrease of half of the appointment waiting time without changing the non showed up rate.

  12. Improved estimation of commuter waiting times using headway and commuter boarding information

    NASA Astrophysics Data System (ADS)

    Ramli, Muhamad Azfar; Jayaraman, Vasundhara; Kwek, Hyen Chee; Tan, Kian Heong; Lee Kee Khoon, Gary; Monterola, Christopher

    2018-07-01

    The average amount of waiting time spent by commuters is one of the key indicators of service quality for public bus operations. While actual measurements of actual waiting time is difficult to be done en masse, models of waiting time can be derived from bus headways and these models have been adopted by transport planners in monitoring and regulating service reliability of operators. However, these models are founded on several assumptions on the patterns of commuter arrival which may not be applicable for bus services that experience high demand and heavily fluctuating commuter patterns. Given the availability of granular data on commuter boarding from automated fare collection systems, we propose a new methodology to better estimate the average waiting time of commuters. The formulation is anchored and validated using a three-month dataset from ten selected bus routes in Singapore. Finally, we discuss how our new measure allows for minimization of commuter waiting time through schedule optimization.

  13. Waiting times for cancer patients in Sweden: A nationwide population-based study.

    PubMed

    Robertson, Stephanie; Adolfsson, Jan; Stattin, Pär; Sjövall, Annika; Winnersjö, Rocio; Hanning, Marianne; Sandelin, Kerstin

    2017-05-01

    The reported long waiting times for cancer patients have mostly been related to prognostic outcome and less to patient-related experience to outcome. We assessed waiting times for patients with cancer of the breast, prostate, colon or rectum in Sweden. The median time from referral to start of treatment was assessed using data from clinical cancer registers for patients who received curative treatment during 2011, 2012 and 2013. The median overall waiting time in different counties ranged from 7 to 28 days for breast cancer, from 117 to 280 days for prostate cancer, from 27 to 64 days for colon cancer and from 48 to 80 days for rectal cancer. For the entire nation, the median time from referral to start of treatment remained unchanged from 2011 to 2013 for each cancer diagnosis. Large variations were found in waiting times between different counties in Sweden and between different types of cancer. The long waiting times identified in this study emphasize the need to improve national programmes for more rapid diagnosis and treatment.

  14. Improving equitable access to imaging under universal-access medicine: the ontario wait time information program and its impact on hospital policy and process.

    PubMed

    Kielar, Ania Z; El-Maraghi, Robert H; Schweitzer, Mark E

    2010-08-01

    In Canada, equal access to health care is the goal, but this is associated with wait times. Wait times should be fair rather than uniform, taking into account the urgency of the problem as well as the time an individual has already waited. In November 2004, the Ontario government began addressing this issue. One of the first steps was to institute benchmarks reflecting "acceptable" wait times for CT and MRI. A public Web site was developed indicating wait times at each Local Health Integration Network. Since starting the Wait Time Information Program, there has been a sustained reduction in wait times for Ontarians requiring CT and MRI. The average wait time for a CT scan went from 81 days in September 2005 to 47 days in September 2009. For MRI, the resulting wait time was reduced from 120 to 105 days. Increased patient scans have been achieved by purchasing new CT and MRI scanners, expanding hours of operation, and improving patient throughput using strategies learned from the Lean initiative, based on Toyota's manufacturing philosophy for car production. Institution-specific changes in booking procedures have been implemented. Concurrently, government guidelines have been developed to ensure accountability for monies received. The Ontario Wait Time Information Program is an innovative first step in improving fair and equitable access to publicly funded imaging services. There have been reductions in wait times for both CT and MRI. As various new processes are implemented, further review will be necessary for each step to determine their individual efficacy. Copyright 2010 American College of Radiology. Published by Elsevier Inc. All rights reserved.

  15. [Women's satisfaction with waiting times for further investigation in breast cancer screening].

    PubMed

    Molina-Barceló, Ana; Salas Trejo, Dolores; Miranda García, Josefa

    2011-01-01

    To determine the factors associated with satisfaction with waiting times for further investigation in breast cancer screening. We carried out a cross-sectional study by telephone survey of a representative sample of women (N=316) participating in the breast cancer screening program of the autonomous region of Valencia (Spain) who required additional tests to confirm the diagnosis. Descriptive analysis was performed by contingency tables (p<0.05) and multivariate association by odds ratios (OR) of logistic regression models (95%CI). Satisfaction with the waiting time was 78.6%. A higher risk of dissatisfaction was found in women from a "high" social class (OR=3.17; 95% CI: 1.10-9.14), those who perceived that the waiting time was "more than 2 weeks", both "since the notification of the need for further investigation until completion of the first test" (OR=15,54; 95%CI: 5,87-41,12) and "since the completion of the last test until notification of the final result" (OR=11.57; 95% CI: 2.96-45.19), and in women who experienced the attention as "worse than expected" (OR=15.40; 95% CI: 1.41-168.64). The maximum waiting time acceptable to the highest percentage of women was "up to 1 week" for each waiting period (n=47, 73.5%; n=14, 45.2%). Waiting times of no more than 1 week and never more than 2 weeks for each waiting period are recommended. Women should be given an approximate waiting time, paying special attention to women aged 45 to 54 years attending their initial screening. 2010 SESPAS. Published by Elsevier Espana. All rights reserved.

  16. Should I stay or should I go? Hospital emergency department waiting times and demand.

    PubMed

    Sivey, Peter

    2018-03-01

    In the absence of the price mechanism, hospital emergency departments rely on waiting times, alongside prioritisation mechanisms, to restrain demand and clear the market. This paper estimates by how much the number of treatments demanded is reduced by a higher waiting time. I use variation in waiting times for low-urgency patients caused by rare and resource-intensive high-urgency patients to estimate the relationship. I find that when waiting times are higher, more low-urgency patients are deterred from treatment and leave the hospital during the waiting period without being treated. The waiting time elasticity of demand for low-urgency patients is approximately -0.25 and is highest for the lowest-urgency patients. Copyright © 2017 John Wiley & Sons, Ltd.

  17. Survey of Access to GastroEnterology in Canada: The SAGE wait times program

    PubMed Central

    Leddin, Desmond; Bridges, Ronald J; Morgan, David G; Fallone, Carlo; Render, Craig; Plourde, Victor; Gray, Jim; Switzer, Connie; McHattie, Jim; Singh, Harminder; Walli, Eric; Murray, Iain; Nestel, Anthony; Sinclair, Paul; Chen, Ying; Irvine, E Jan

    2010-01-01

    BACKGROUND: Assessment of current wait times for specialist health services in Canada is a key method that can assist government and health care providers to plan wisely for future health needs. These data are not readily available. A method to capture wait time data at the time of consultation or procedure has been developed, which should be applicable to other specialist groups and also allows for assessment of wait time trends over intervals of years. METHODS: In November 2008, gastroenterologists across Canada were asked to complete a questionnaire (online or by fax) that included personal demographics and data from one week on at least five consecutive new consultations and five consecutive procedure patients who had not previously undergone a procedure for the same indication. Wait times were collected for 18 primary indications and results were then compared with similar survey data collected in 2005. RESULTS: The longest wait times observed were for screening colonoscopy (201 days) and surveillance of previous colon cancer or polyps (272 days). The shortest wait times were for cancer-likely based on imaging or physical examination (82 days), severe or rapidly progressing dysphagia or odynophagia (83 days), documented iron-deficiency anemia (90 days) and dyspepsia with alarm symptoms (99 days). Compared with 2005 data, total wait times in 2008 were lengthened overall (127 days versus 155 days; P<0.05) and for most of the seven individual indications that permitted data comparison. CONCLUSION: Median wait times for gastroenterology services continue to exceed consensus conference recommended targets and have significantly worsened since 2005. PMID:20186352

  18. Discrimination in a universal health system: explaining socioeconomic waiting time gaps.

    PubMed

    Johar, Meliyanni; Jones, Glenn; Keane, Micheal P; Savage, Elizabeth; Stavrunova, Olena

    2013-01-01

    One of the core goals of a universal health care system is to eliminate discrimination on the basis of socioeconomic status. We test for discrimination using patient waiting times for non-emergency treatment in public hospitals. Waiting time should reflect patients' clinical need with priority given to more urgent cases. Using data from Australia, we find evidence of prioritisation of the most socioeconomically advantaged patients at all quantiles of the waiting time distribution. These patients also benefit from variation in supply endowments. These results challenge the universal health system's core principle of equitable treatment. Copyright © 2012 Elsevier B.V. All rights reserved.

  19. Waiting time for cataract surgery and its influence on patient attitudes.

    PubMed

    Chan, Frank Wan-kin; Fan, Alex Hoi; Wong, Fiona Yan-yan; Lam, Philip Tsze-ho; Yeoh, Eng-kiong; Yam, Carrie Ho-kwan; Griffiths, Sian; Lam, Dennis Shun-chiu; Congdon, Nathan

    2009-08-01

    To characterize willingness to pay for private operations and preferred waiting time among patients awaiting cataract surgery in Hong Kong. This was a cross-sectional survey. Subjects randomly selected from cataract surgical waiting lists in Hong Kong (n = 467) underwent a telephone interview based on a structured, validated questionnaire. Data were collected on private insurance coverage, preferred waiting time, amount willing to pay for surgery, and self-reported visual function and health status. Among 300 subjects completing the interview, 144 (48.2%) were 76 years of age or older, 177 (59%) were women, and mean time waiting for surgery was 17 +/- 15 months. Among 220 subjects (73.3%) willing to pay anything for surgery, the mean amount was US$552 +/- 443. With adjustment for age, education, and monthly household income, subjects willing to pay anything were less willing to wait 12 months for surgery (OR = 4.34; P = 0.002), more likely to know someone having had cataract surgery (OR = 2.20; P = 0.03), and more likely to use their own savings to pay for the surgery (OR = 2.21; P = 0.04). Subjects considering private cataract surgery, knowing people who have had cataract surgery, using nongovernment sources to pay for surgery, and having lower visual function were willing to pay more. Many patients wait significant periods for cataract surgery in Hong Kong, and are willing to pay substantial amounts for private operations. These results may have implications for other countries with cataract waiting lists.

  20. RECONCILIATION OF WAITING TIME STATISTICS OF SOLAR FLARES OBSERVED IN HARD X-RAYS

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

    Aschwanden, Markus J.; McTiernan, James M., E-mail: aschwanden@lmsal.co, E-mail: jimm@ssl.berkeley.ed

    2010-07-10

    We study the waiting time distributions of solar flares observed in hard X-rays with ISEE-3/ICE, HXRBS/SMM, WATCH/GRANAT, BATSE/CGRO, and RHESSI. Although discordant results and interpretations have been published earlier, based on relatively small ranges (<2 decades) of waiting times, we find that all observed distributions, spanning over 6 decades of waiting times ({Delta}t {approx} 10{sup -3}-10{sup 3} hr), can be reconciled with a single distribution function, N({Delta}t) {proportional_to} {lambda}{sub 0}(1 + {lambda}{sub 0{Delta}}t){sup -2}, which has a power-law slope of p {approx} 2.0 at large waiting times ({Delta}t {approx} 1-1000 hr) and flattens out at short waiting times {Delta}t {approx}waiting times is invariant for sampling with different flux thresholds, while the mean waiting time scales reciprocically with the number of detected events, {Delta}t {sub 0} {proportional_to} 1/n {sub det}. This waiting time distribution can be modeled with a nonstationary Poisson process with a flare rate {lambda} = 1/{Delta}t that varies as f({lambda}) {proportional_to} {lambda}{sup -1}exp - ({lambda}/{lambda}{sub 0}). This flare rate distribution requires a highly intermittent flare productivity in short clusters with high rates, separated by relatively long quiescent intervals with very low flare rates.« less

  1. Posted wait times an added advantage to multi-facility systems?

    PubMed

    2011-04-01

    Methodist Le Bonheur Healthcare in Memphis, TN, is investigating whether posting ED wait times via the internet can positively impact patient flow in the six EDs the health system operates in the Memphis region. The health system began posting wait times in August 2010, resulting in increases in ED volume ranging from 6% to 10%. The health system is monitoring ED arrivals by zip code to assess any impact on load balancing between its busy EDs. One marketing challenge is that a competitor is posting ED wait times as well, but it is posting the time it takes for a patient to be placed in a bed as opposed to the door-to-provider time that Methodist Le Bonheur is posting. The approach has the most impact on lower-acuity patients, but experts worry that in the future, payers may not be reimbursed for ED care for these patients.

  2. Wait Time and Effective Social Studies Instruction: What Can Research in Science Education Tell Us?

    ERIC Educational Resources Information Center

    Atwood, Virgina A.; Wilen, William W.

    1991-01-01

    Defines wait time as the length of time teachers wait for answers from students after asking a question. Maintains that increasing wait time can stimulate reflective thinking and student involvement. Reviews the research literature on wait time studies in science education. Finds that student responses improve and participation expands with…

  3. Waiting time distribution in public health care: empirics and theory.

    PubMed

    Dimakou, Sofia; Dimakou, Ourania; Basso, Henrique S

    2015-12-01

    Excessive waiting times for elective surgery have been a long-standing concern in many national healthcare systems in the OECD. How do the hospital admission patterns that generate waiting lists affect different patients? What are the hospitals characteristics that determine waiting times? By developing a model of healthcare provision and analysing empirically the entire waiting time distribution we attempt to shed some light on those issues. We first build a theoretical model that describes the optimal waiting time distribution for capacity constraint hospitals. Secondly, employing duration analysis, we obtain empirical representations of that distribution across hospitals in the UK from 1997-2005. We observe important differences on the 'scale' and on the 'shape' of admission rates. Scale refers to how quickly patients are treated and shape represents trade-offs across duration-treatment profiles. By fitting the theoretical to the empirical distributions we estimate the main structural parameters of the model and are able to closely identify the main drivers of these empirical differences. We find that the level of resources allocated to elective surgery (budget and physical capacity), which determines how constrained the hospital is, explains differences in scale. Changes in benefits and costs structures of healthcare provision, which relate, respectively, to the desire to prioritise patients by duration and the reduction in costs due to delayed treatment, determine the shape, affecting short and long duration patients differently. JEL Classification I11; I18; H51.

  4. An investigation of the impact of prolonged waiting times on blood donors in Ireland.

    PubMed

    McKeever, T; Sweeney, M R; Staines, A

    2006-02-01

    The aim of this study was to investigate the impact of prolonged queuing times on blood donors, by measuring their satisfaction levels, and positive and negative affects. As donation times have increased over the past number of years within the Irish Blood Transfusion Service, this is an important issue to examine in a climate where voluntary donors are becoming scarce and demands on people's time are increasing. Eighty-five blood donors were sampled from one urban and one rural blood donor clinic. The respondents conducted a questionnaire by means of face-to-face interview, while waiting in the clinic. The questionnaire contained the Positive and Negative Affect Scale (PANAS), and a waiting satisfaction scale. Both actual and perceived waiting times of the donors were noted. Waiting time was found to be negatively related to satisfaction. Inexperienced donors expressed higher levels of negative affect than experienced donors. Urban donors were significantly more satisfied than rural donors. There was a significant difference in perceived waiting time between lone donors and those queuing in a group, with those waiting alone perceiving their wait as shorter. While all respondents stated that they intended to donate again, over one-third stated that prolonged waiting times would be their most likely deterrent. However, only 15% stated that long queuing times might actually prevent them from donating in the future, and almost all respondents said that they would recommend donation to a friend, despite long queuing times. Although our results show that the respondents were not satisfied with current waiting times, it did not seem to affect their future intentions to donate. These findings provide some optimism for the future of blood donation in Ireland, as they suggest a strong sense of commitment to donation within the population sampled. Future research could explore the application of 'the service industry' approach to waiting times to blood donation clinics.

  5. Waiting time distribution revealing the internal spin dynamics in a double quantum dot

    NASA Astrophysics Data System (ADS)

    Ptaszyński, Krzysztof

    2017-07-01

    Waiting time distribution and the zero-frequency full counting statistics of unidirectional electron transport through a double quantum dot molecule attached to spin-polarized leads are analyzed using the quantum master equation. The waiting time distribution exhibits a nontrivial dependence on the value of the exchange coupling between the dots and the gradient of the applied magnetic field, which reveals the oscillations between the spin states of the molecule. The zero-frequency full counting statistics, on the other hand, is independent of the aforementioned quantities, thus giving no insight into the internal dynamics. The fact that the waiting time distribution and the zero-frequency full counting statistics give a nonequivalent information is associated with two factors. Firstly, it can be explained by the sensitivity to different timescales of the dynamics of the system. Secondly, it is associated with the presence of the correlation between subsequent waiting times, which makes the renewal theory, relating the full counting statistics and the waiting time distribution, no longer applicable. The study highlights the particular usefulness of the waiting time distribution for the analysis of the internal dynamics of mesoscopic systems.

  6. Efficiency of performing pulmonary procedures in a shared endoscopy unit: procedure time, turnaround time, delays, and procedure waiting time.

    PubMed

    Verma, Akash; Lee, Mui Yok; Wang, Chunhong; Hussein, Nurmalah B M; Selvi, Kalai; Tee, Augustine

    2014-04-01

    The purpose of this study was to assess the efficiency of performing pulmonary procedures in the endoscopy unit in a large teaching hospital. A prospective study from May 20 to July 19, 2013, was designed. The main outcome measures were procedure delays and their reasons, duration of procedural steps starting from patient's arrival to endoscopy unit, turnaround time, total case durations, and procedure wait time. A total of 65 procedures were observed. The most common procedure was BAL (61%) followed by TBLB (31%). Overall procedures for 35 (53.8%) of 65 patients were delayed by ≥ 30 minutes, 21/35 (60%) because of "spillover" of the gastrointestinal and surgical cases into the time block of pulmonary procedure. Time elapsed between end of pulmonary procedure and start of the next procedure was ≥ 30 minutes in 8/51 (16%) of cases. In 18/51 (35%) patients there was no next case in the room after completion of the pulmonary procedure. The average idle time of the room after the end of pulmonary procedure and start of next case or end of shift at 5:00 PM if no next case was 58 ± 53 minutes. In 17/51 (33%) patients the room's idle time was >60 minutes. A total of 52.3% of patients had the wait time >2 days and 11% had it ≥ 6 days, reason in 15/21 (71%) being unavailability of the slot. Most pulmonary procedures were delayed due to spillover of the gastrointestinal and surgical cases into the block time allocated to pulmonary procedures. The most common reason for difficulty encountered in scheduling the pulmonary procedure was slot unavailability. This caused increased procedure waiting time. The strategies to reduce procedure delays and turnaround times, along with improved scheduling methods, may have a favorable impact on the volume of procedures performed in the unit thereby optimizing the existing resources.

  7. Quality of life in patients with unilateral vestibular schwannoma on wait and see - strategy.

    PubMed

    Klersy, P C; Arlt, F; Hofer, M; Meixensberger, J

    2018-01-01

    A 'wait and see' strategy is an option when managing patients with small vestibular schwannomas (VS). A risk of growth and worsening of hearing may influence a patient's daily quality of life (QOL). Therefore, the present study focused on QOL parameters in patients who are on a 'wait and see' strategy following magnetic resonance imaging (MRI)-based diagnosis of small unilateral VS. Sixty-five patients (mean age 64.4 years; male:female, 32:33) who suffered from a small unilateral VS (9.34 mm, range 1.5-23 mm) between 2013 and 2016 were included in a prospective single center study. During follow-up, in addition to clinical and neurological examinations and MRI imaging, all patients answered the Short Form 36 questionnaire once to characterize QOL. Additionally, the severity of tinnitus was determined by the Mini-TQ-12 from Hiller and Goebel. It was found during follow-up that there was no lowering of QOL in patients with small VS who were on 'wait and see' strategy compared with Germany's general population and no tumor growth was detected in 53 patients (81.5%). Patients with a tumor diameter larger than 10 mm did not suffer from stronger tinnitus, vertigo or unsteadiness than the group with an average tumor size, which is smaller than 10 mm. Sixty-two patients (95.4%) showed ipsilateral hearing loss and three of these reported deafness (4.6%). Severe vertigo or tinnitus is connected with lower levels of mental component scale and physical component scale. These findings reduced the QOL (p = 0.05). In our series, QOL is not influenced in patients with unilateral untreated small VS in comparison to Germany's general population. This is helpful information when advising patients during follow-up and finding out the optimal timing of individual treatment.

  8. The 2012 SAGE wait times program: Survey of Access to GastroEnterology in Canada

    PubMed Central

    Leddin, Desmond; Armstrong, David; Borgaonkar, Mark; Bridges, Ronald J; Fallone, Carlo A; Telford, Jennifer J; Chen, Ying; Colacino, Palma; Sinclair, Paul

    2013-01-01

    BACKGROUND: Periodically surveying wait times for specialist health services in Canada captures current data and enables comparisons with previous surveys to identify changes over time. METHODS: During one week in April 2012, Canadian gastroenterologists were asked to complete a questionnaire (online or by fax) recording demographics, reason for referral, and dates of referral and specialist visits for at least 10 consecutive new patients (five consultations and five procedures) who had not been seen previously for the same indication. Wait times were determined for 18 indications and compared with those from similar surveys conducted in 2008 and 2005. RESULTS: Data regarding adult patients were provided by 173 gastroenterologists for 1374 consultations, 540 procedures and 293 same-day consultations and procedures. Nationally, the median wait times were 92 days (95% CI 85 days to 100 days) from referral to consultation, 55 days (95% CI 50 days to 61 days) from consultation to procedure and 155 days (95% CI 142 days to 175 days) (total) from referral to procedure. Overall, wait times were longer in 2012 than in 2005 (P<0.05); the wait time to same-day consultation and procedure was shorter in 2012 than in 2008 (78 days versus 101 days; P<0.05), but continued to be longer than in 2005 (P<0.05). The total wait time remained longest for screening colonoscopy, increasing from 201 days in 2008 to 279 days in 2012 (P<0.05). DISCUSSION: Wait times for gastroenterology services continue to exceed recommended targets, remain unchanged since 2008 and exceed wait times reported in 2005. PMID:23472243

  9. [Reducing patient waiting time for the outpatient phlebotomy service using six sigma].

    PubMed

    Kim, Yu Kyung; Song, Kyung Eun; Lee, Won Kil

    2009-04-01

    One of the challenging issues of the outpatient phlebotomy services at most hospitals is that patients have a long wait. The outpatient phlebotomy team of Kyungpook National University Hospital applied six sigma breakthrough methodologies to reduce the patient waiting time. The DMAIC (Define, Measure, Analyze, Improve, and Control) model was employed to approach the project. Two hundred patients visiting the outpatient phlebotomy section were asked to answer the questionnaires at inception of the study to ascertain root causes. After correction, we surveyed 285 patients for same questionnaires again to follow-up the effects. A defect was defined as extending patient waiting time so long and at the beginning of the project, the performance level was 2.61 sigma. Using fishbone diagram, all the possible reasons for extending patient waiting time were captured, and among them, 16 causes were proven to be statistically significant. Improvement plans including a new receptionist, automatic specimen transport system, and adding one phlebotomist were put into practice. As a result, the number of patients waited more than 5 min significantly decreased, and the performance level reached 3.0 sigma in December 2007 and finally 3.35 sigma in July 2008. Applying the six sigma, the performance level of waiting times for blood drawing exceeding five minutes were improved from 2.61 sigma to 3.35 sigma.

  10. A Model to Study: Cannibalization, FMC, and Customer Waiting Time

    DTIC Science & Technology

    2002-02-01

    4825 Mark Center Drive • Alexandria, Virginia 22311-1850 CRM D0005957.A2/Final February 2002 A Model to Study: Cannibalization, FMC, and Customer ...numerical example In this section, we will derive the relationship between cannibaliza- tion rates, customer waiting time (CWT) for needed spare parts... relationships between the FMC given by equation 1, the mean customer wait time for spare parts, denoted µ, and the 5. According to [19], not every part can be

  11. The effect of waiting times on demand and supply for elective surgery: Evidence from Italy.

    PubMed

    Riganti, Andrea; Siciliani, Luigi; Fiorio, Carlo V

    2017-09-01

    Waiting times are a major policy concern in publicly funded health systems across OECD countries. Economists have argued that, in the presence of excess demand, waiting times act as nonmonetary prices to bring demand for and supply of health care in equilibrium. Using administrative data disaggregated by region and surgical procedure over 2010-2014 in Italy, we estimate demand and supply elasticities with respect to waiting times. We employ linear regression models with first differences and instrumental variables to deal with endogeneity of waiting times. We find that demand is inelastic to waiting times while supply is more elastic. Estimates of demand elasticity are between -0.15 to -0.24. Our results have implications on the effectiveness of policies aimed at increasing supply and their ability to reduce waiting times. Copyright © 2017 John Wiley & Sons, Ltd.

  12. Cost-effectiveness of Wait Time Reduction for Intensive Behavioral Intervention Services in Ontario, Canada.

    PubMed

    Piccininni, Caroline; Bisnaire, Lise; Penner, Melanie

    2017-01-01

    Earlier access to intensive behavioral intervention (IBI) is associated with improved outcomes for children with severe autism spectrum disorder (ASD); however, there are long waiting times for this program. No analyses have been performed modeling the cost-effectiveness of wait time reduction for IBI. To model the starting age for IBI with reduced wait time (RWT) (by half) and eliminated wait time (EWT), and perform a cost-effectiveness analysis comparing RWT and EWT with current wait time (CWT) from government and societal perspectives. Published waiting times were used to model the mean starting age for IBI for CWT, RWT, and EWT in children diagnosed with severe ASD who were treated at Ontario's Autism Intervention Program. Inputs were loaded into a decision analytic model, with an annual discount rate of 3% applied. Incremental cost-effectiveness ratios (ICERs) were determined. One-way and probabilistic sensitivity analyses were performed to assess the effect of model uncertainty. We used data from the year 2012 (January 1 through December 31) provided from the Children's Hospital of Eastern Ontario IBI center for the starting ages. Data analysis was done from May through July 2015. The outcome was independence measured in dependency-free life-years (DFLYs) to 65 years of age. To derive this, expected IQ was modeled based on probability of early (age <4 years) or late (age ≥4 years) access to IBI. Probabilities of having an IQ in the normal (≥70) or intellectual disability (<70) range were calculated. The IQ strata were assigned probabilities of achieving an independent (60 DFLYs), semidependent (30 DFLYs), or dependent (0 DFLYs) outcome. Costs were calculated for provincial government and societal perspectives in Canadian dollars (Can$1 = US$0.78). The mean starting ages for IBI were 5.24 years for CWT, 3.89 years for RWT, and 2.71 years for EWT. From the provincial government perspective, EWT was the dominant strategy, generating the most DFLYs for

  13. Waiting time for cancer treatment and mental health among patients with newly diagnosed esophageal or gastric cancer: a nationwide cohort study.

    PubMed

    Song, Huan; Fang, Fang; Valdimarsdóttir, Unnur; Lu, Donghao; Andersson, Therese M-L; Hultman, Christina; Ye, Weimin; Lundell, Lars; Johansson, Jan; Nilsson, Magnus; Lindblad, Mats

    2017-01-03

    vulnerabilities. Our study sheds further light on the complexity of waiting time management, and calls for a comprehensive strategy that takes into account different domains of patient well-being in addition to the overall survival.

  14. Reducing waiting time and raising outpatient satisfaction in a Chinese public tertiary general hospital-an interrupted time series study.

    PubMed

    Sun, Jing; Lin, Qian; Zhao, Pengyu; Zhang, Qiongyao; Xu, Kai; Chen, Huiying; Hu, Cecile Jia; Stuntz, Mark; Li, Hong; Liu, Yuanli

    2017-08-22

    It is globally agreed that a well-designed health system deliver timely and convenient access to health services for all patients. Many interventions aiming to reduce waiting times have been implemented in Chinese public tertiary hospitals to improve patients' satisfaction. However, few were well-documented, and the effects were rarely measured with robust methods. We conducted a longitudinal study of the length of waiting times in a public tertiary hospital in Southern China which developed comprehensive data collection systems. Around an average of 60,000 outpatients and 70,000 prescribed outpatients per month were targeted for the study during Oct 2014-February 2017. We analyzed longitudinal time series data using a segmented linear regression model to assess changes in levels and trends of waiting times before and after the introduction of waiting time reduction interventions. Pearson correlation analysis was conducted to indicate the strength of association between waiting times and patient satisfactions. The statistical significance level was set at 0.05. The monthly average length of waiting time decreased 3.49 min (P = 0.003) for consultations and 8.70 min (P = 0.02) for filling prescriptions in the corresponding month when respective interventions were introduced. The trend shifted from baseline slight increasing to afterwards significant decreasing for filling prescriptions (P =0.003). There was a significant negative correlation between waiting time of filling prescriptions and outpatient satisfaction towards pharmacy services (r = -0.71, P = 0.004). The interventions aimed at reducing waiting time and raising patient satisfaction in Fujian Provincial Hospital are effective. A long-lasting reduction effect on waiting time for filling prescriptions was observed because of carefully designed continuous efforts, rather than a one-time campaign, and with appropriate incentives implemented by a taskforce authorized by the hospital managers. This

  15. Waiting time of inpatients before elective surgical procedures at a State Government Teaching Hospital in India.

    PubMed

    Ray, Shreyasi; Kirtania, Jyotirmay

    2017-01-01

    Abundant published literature exists addressing the issues of outpatient waiting lists before surgery. However, there is no published literature on inpatient waiting time before elective surgical procedures. This study aims to measure the inpatient waiting time, identify the factors that affect the inpatient waiting time, and recommend the ways of reducing the waiting time of inpatients before elective surgical procedures, at a state government teaching hospital in India. Descriptive research methods and quality control tools were used for this prospective observational study. Descriptive statistics, Shapiro-Wilk test of normality, Wilcoxon-Mann-Whitney Test, and Kruskal-Wallis test were used. Pareto charts were used to highlight the most important modifiable factors among the set of factors causing increased waiting time. We also applied the M/M/c model (Erlang - A model) of queue theory to analyze the traffic intensity and system congestion. The median waiting time of inpatients before elective surgery was 12 days (interquartile range = 11.5 days). The waiting time was influenced significantly (P < 0.05) by the patient's age, physical status, and the financial status. The surgical specialty, blood product booking and procurement, cross-specialty consultation before surgery, and Intensive Care Unit booking were the other important factors. Modifiable and nonmodifiable factors affecting the inpatient waiting time of surgical patients were identified. Control measures that can reduce the waiting time of inpatients before elective surgery were identified.

  16. Access to specialist gastroenterology care in Canada: The Practice Audit in Gastroenterology (PAGE) Wait Times Program

    PubMed Central

    Armstrong, David; Barkun, Alan NG; Chen, Ying; Daniels, Sandra; Hollingworth, Roger; Hunt, Richard H; Leddin, Desmond

    2008-01-01

    BACKGROUND: Canadian wait time data are available for the treatment of cancer and heart disease, as well as for joint replacement, cataract surgery and diagnostic imaging procedures. Wait times for gastroenterology consultation and procedures have not been studied, although digestive diseases pose a greater economic burden in Canada than cancer or heart disease. METHODS: Specialist physicians completed the practice audit if they provided digestive health care, accepted new patients and recorded referral dates. For patients seen for consultation or investigation over a one-week period, preprogrammed personal digital assistants were used to collect data including the main reason for referral, initial referral and consultation dates, procedure dates (if performed), personal and family history, and patient symptoms, signs and test results. Patient triaging, appropriateness of the referral and timeliness of care were noted. RESULTS: Over 10 months, 199 physicians recorded details of 5559 referrals, including 1903 visits for procedures. The distribution of total wait times (from referral to procedure) nationally was highly skewed at 91/203 days (median/75th percentile), with substantial interprovincial variation: British Columbia, 66/185 days; Alberta, 134/284 days; Ontario, 110/208 days; Quebec, 71/149 days; New Brunswick, 104/234 days; and Nova Scotia, 42/84 days. The percentage of physicians by province offering average-risk screening colonoscopy varied from 29% to 100%. DISCUSSION: Access to specialist gastroenterology care in Canada is limited by long wait times, which exceed clinically reasonable waits for specialist treatment. Although exhibiting some methodological limitations, this large practice audit sampling offers broadly generalized results, as well as a means to identify barriers to health care delivery and evaluate strategies to address these barriers, with the goals of expediting appropriate care for patients with digestive health disorders and

  17. Wait times in the emergency department for patients with mental illness

    PubMed Central

    Atzema, Clare L.; Schull, Michael J.; Kurdyak, Paul; Menezes, Natasja M.; Wilton, Andrew S.; Vermuelen, Marian J.; Austin, Peter C.

    2012-01-01

    Background: It has been suggested that patients with mental illness wait longer for care than other patients in the emergency department. We determined wait times for patients with and without mental health diagnoses during crowded and noncrowded periods in the emergency department. Methods: We conducted a population-based retrospective cohort analysis of adults seen in 155 emergency departments in Ontario between April 2007 and March 2009. We compared wait times and triage scores for patients with mental illness to those for all other patients who presented to the emergency department during the study period. Results: The patients with mental illness (n = 51 381) received higher priority triage scores than other patients, regardless of crowding. The time to assessment by a physician was longer overall for patients with mental illness than for other patients (median 82, interquartile range [IQR] 41–147 min v. median 75 [IQR 36–140] min; p < 0.001). The median time from the decision to admit the patient to hospital to ward transfer was markedly shorter for patients with mental illness than for other patients (median 74 [IQR 15–215] min v. median 152 [IQR 45–605] min; p < 0.001). After adjustment for other variables, patients with mental illness waited 10 minutes longer to see a physician compared with other patients during noncrowded periods (95% confidence interval [CI] 8 to 11), but they waited significantly less time than other patients as crowding increased (mild crowding: −14 [95% CI −12 to −15] min; moderate crowding: −38 [95% CI −35 to −42] min; severe crowding: −48 [95% CI −39 to −56] min; p < 0.001). Interpretation: Patients with mental illness were triaged appropriately in Ontario’s emergency departments. These patients waited less time than other patients to see a physician under crowded conditions and only slightly longer under noncrowded conditions. PMID:23148052

  18. Waiting time as a competitive device: an example from general medical practice.

    PubMed

    Iversen, Tor; Lurås, Hilde

    2002-09-01

    From a theoretical model we predict that only physicians with quality characteristics perceived as inferior by patients are willing to embark on waiting time reductions. Because of variation in these quality characteristics among physicians, market equilibrium is likely to show a range of waiting times for physician services. This hypothesis is supported by results from a study of Norwegian general practitioners. Since the waiting time offered by a physician influences the number of patient-initiated consultations, a policy implication of our study is that the distinction between patient-initiated and physician-initiated consultations may be less clear-cut than often assumed in the literature.

  19. An empirical analysis of the impact of choice on waiting times.

    PubMed

    Siciliani, Luigi; Martin, Steve

    2007-08-01

    Policy-makers often claim that enhancing patient choice induces more competition among hospitals and may therefore reduce waiting times. This paper tests this claim using 120 English NHS hospitals over the period 1999-2001. Several proxies for the degree of choice (or competition) are constructed including: (a) the number of hospitals within the catchment area of each hospital; (b) the number of hospitals in the catchment area of each hospital standardised by the population of the catchment area; (c) the inverse of the Herfindahl index (or 'the number of effective competitors'). Several control variables are included: the availability of doctors, junior doctors, nurses, and other personnel; the availability of acute beds; the emergency admission rate; the day-case rate; the average length of inpatient stay; an indicator of case-mix; and mortality and re-admission rates. We find that more choice is significantly associated with lower waiting times at the sample mean (five hospitals) although the quantitative effect is modest: an extra hospital in a catchment area will only reduce waiting by at most a few days (or 1-2% reduction in waiting). There is also some evidence that increases in choice can boost waiting times when the degree of choice is very high (i.e. more than 11 hospitals are included in the catchment area). Copyright 2007 John Wiley & Sons, Ltd.

  20. Waiting time disparities in breast cancer diagnosis and treatment: a population-based study in France.

    PubMed

    Molinié, F; Leux, C; Delafosse, P; Ayrault-Piault, S; Arveux, P; Woronoff, A S; Guizard, A V; Velten, M; Ganry, O; Bara, S; Daubisse-Marliac, L; Tretarre, B

    2013-10-01

    Waiting times are key indicators of a health's system performance, but are not routinely available in France. We studied waiting times for diagnosis and treatment according to patients' characteristics, tumours' characteristics and medical management options in a sample of 1494 breast cancers recorded in population-based registries. The median waiting time from the first imaging detection to the treatment initiation was 34 days. Older age, co-morbidity, smaller size of tumour, detection by organised screening, biopsy, increasing number of specimens removed, multidisciplinary consulting meetings and surgery as initial treatment were related to increased waiting times in multivariate models. Many of these factors were related to good practices guidelines. However, the strong influence of organised screening programme and the disparity of waiting times according to geographical areas were of concern. Better scheduling of diagnostic tests and treatment propositions should improve waiting times in the management of breast cancer in France. Copyright © 2013 Elsevier Ltd. All rights reserved.

  1. Methodology for Analysis, Modeling and Simulation of Airport Gate-waiting Delays

    NASA Astrophysics Data System (ADS)

    Wang, Jianfeng

    This dissertation presents methodologies to estimate gate-waiting delays from historical data, to identify gate-waiting-delay functional causes in major U.S. airports, and to evaluate the impact of gate operation disruptions and mitigation strategies on gate-waiting delay. Airport gates are a resource of congestion in the air transportation system. When an arriving flight cannot pull into its gate, the delay it experiences is called gate-waiting delay. Some possible reasons for gate-waiting delay are: the gate is occupied, gate staff or equipment is unavailable, the weather prevents the use of the gate (e.g. lightning), or the airline has a preferred gate assignment. Gate-waiting delays potentially stay with the aircraft throughout the day (unless they are absorbed), adding costs to passengers and the airlines. As the volume of flights increases, ensuring that airport gates do not become a choke point of the system is critical. The first part of the dissertation presents a methodology for estimating gate-waiting delays based on historical, publicly available sources. Analysis of gate-waiting delays at major U.S. airports in the summer of 2007 identifies the following. (i) Gate-waiting delay is not a significant problem on majority of days; however, the worst delay days (e.g. 4% of the days at LGA) are extreme outliers. (ii) The Atlanta International Airport (ATL), the John F. Kennedy International Airport (JFK), the Dallas/Fort Worth International Airport (DFW) and the Philadelphia International Airport (PHL) experience the highest gate-waiting delays among major U.S. airports. (iii) There is a significant gate-waiting-delay difference between airlines due to a disproportional gate allocation. (iv) Gate-waiting delay is sensitive to time of a day and schedule peaks. According to basic principles of queueing theory, gate-waiting delay can be attributed to over-scheduling, higher-than-scheduled arrival rate, longer-than-scheduled gate-occupancy time, and reduced gate

  2. Outpatient Waiting Time in Health Services and Teaching Hospitals: A Case Study in Iran

    PubMed Central

    Mohebbifar, Rafat; Hasanpoor, Edris; Mohseni, Mohammad; Sokhanvar, Mobin; Khosravizadeh, Omid; Isfahani, Haleh Mousavi

    2014-01-01

    Background: One of the most important indexes of the health care quality is patient’s satisfaction and it takes place only when there is a process based on management. One of these processes in the health care organizations is the appropriate management of the waiting time process. The aim of this study is the systematic analyzing of the outpatient waiting time. Methods: This descriptive cross sectional study conducted in 2011 is an applicable study performed in the educational and health care hospitals of one of the medical universities located in the north west of Iran. Since the distributions of outpatients in all the months were equal, sampling stage was used. 160 outpatients were studied and the data was analyzed by using SPSS software. Results: Results of the study showed that the waiting time for the outpatients of ophthalmology clinic with an average of 245 minutes for each patient allocated the maximum time among the other clinics for itself. Orthopedic clinic had the minimal waiting time including an average of 77 minutes per patient. The total average waiting time for each patient in the educational hospitals under this study was about 161 minutes. Conclusion: by applying some models, we can reduce the waiting time especially in the realm of time and space before the admission to the examination room. Utilizing the models including the one before admission, electronic visit systems via internet, a process model, six sigma model, queuing theory model and FIFO model, are the components of the intervention that reduces the outpatient waiting time. PMID:24373277

  3. Continuous time random walk model with asymptotical probability density of waiting times via inverse Mittag-Leffler function

    NASA Astrophysics Data System (ADS)

    Liang, Yingjie; Chen, Wen

    2018-04-01

    The mean squared displacement (MSD) of the traditional ultraslow diffusion is a logarithmic function of time. Recently, the continuous time random walk model is employed to characterize this ultraslow diffusion dynamics by connecting the heavy-tailed logarithmic function and its variation as the asymptotical waiting time density. In this study we investigate the limiting waiting time density of a general ultraslow diffusion model via the inverse Mittag-Leffler function, whose special case includes the traditional logarithmic ultraslow diffusion model. The MSD of the general ultraslow diffusion model is analytically derived as an inverse Mittag-Leffler function, and is observed to increase even more slowly than that of the logarithmic function model. The occurrence of very long waiting time in the case of the inverse Mittag-Leffler function has the largest probability compared with the power law model and the logarithmic function model. The Monte Carlo simulations of one dimensional sample path of a single particle are also performed. The results show that the inverse Mittag-Leffler waiting time density is effective in depicting the general ultraslow random motion.

  4. Waiting Time: The De-Subjectification of Children in Danish Asylum Centres

    ERIC Educational Resources Information Center

    Vitus, Kathrine

    2010-01-01

    This article analyses the relationship between time and subjectification, focusing on the temporal structures created within Danish asylum centres and politics, and on children's experiences of and reactions to open-ended waiting. Such waiting leads to existential boredom which manifests in the children as restlessness, fatigue and despair. The…

  5. Burst wait time simulation of CALIBAN reactor at delayed super-critical state

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

    Humbert, P.; Authier, N.; Richard, B.

    2012-07-01

    In the past, the super prompt critical wait time probability distribution was measured on CALIBAN fast burst reactor [4]. Afterwards, these experiments were simulated with a very good agreement by solving the non-extinction probability equation [5]. Recently, the burst wait time probability distribution has been measured at CEA-Valduc on CALIBAN at different delayed super-critical states [6]. However, in the delayed super-critical case the non-extinction probability does not give access to the wait time distribution. In this case it is necessary to compute the time dependent evolution of the full neutron count number probability distribution. In this paper we present themore » point model deterministic method used to calculate the probability distribution of the wait time before a prescribed count level taking into account prompt neutrons and delayed neutron precursors. This method is based on the solution of the time dependent adjoint Kolmogorov master equations for the number of detections using the generating function methodology [8,9,10] and inverse discrete Fourier transforms. The obtained results are then compared to the measurements and Monte-Carlo calculations based on the algorithm presented in [7]. (authors)« less

  6. Discovering the impact of preceding units' characteristics on the wait time of cardiac surgery unit from statistic data.

    PubMed

    Liu, Jiming; Tao, Li; Xiao, Bo

    2011-01-01

    Prior research shows that clinical demand and supplier capacity significantly affect the throughput and the wait time within an isolated unit. However, it is doubtful whether characteristics (i.e., demand, capacity, throughput, and wait time) of one unit would affect the wait time of subsequent units on the patient flow process. Focusing on cardiac care, this paper aims to examine the impact of characteristics of the catheterization unit (CU) on the wait time of cardiac surgery unit (SU). This study integrates published data from several sources on characteristics of the CU and SU units in 11 hospitals in Ontario, Canada between 2005 and 2008. It proposes a two-layer wait time model (with each layer representing one unit) to examine the impact of CU's characteristics on the wait time of SU and test the hypotheses using the Partial Least Squares-based Structural Equation Modeling analysis tool. Results show that: (i) wait time of CU has a direct positive impact on wait time of SU (β = 0.330, p < 0.01); (ii) capacity of CU has a direct positive impact on demand of SU (β = 0.644, p < 0.01); (iii) within each unit, there exist significant relationships among different characteristics (except for the effect of throughput on wait time in SU). Characteristics of CU have direct and indirect impacts on wait time of SU. Specifically, demand and wait time of preceding unit are good predictors for wait time of subsequent units. This suggests that considering such cross-unit effects is necessary when alleviating wait time in a health care system. Further, different patient risk profiles may affect wait time in different ways (e.g., positive or negative effects) within SU. This implies that the wait time management should carefully consider the relationship between priority triage and risk stratification, especially for cardiac surgery.

  7. Patient satisfaction with wait times at an emergency ophthalmology on-call service.

    PubMed

    Chan, Brian J; Barbosa, Joshua; Moinul, Prima; Sivachandran, Nirojini; Donaldson, Laura; Zhao, Lily; Mullen, Sarah J; McLaughlin, Christopher R; Chaudhary, Varun

    2018-04-01

    To assess patient satisfaction with emergency ophthalmology care and determine the effect provision of anticipated appointment wait time has on scores. Single-centre, randomized control trial. Fifty patients triaged at the Hamilton Regional Eye Institute (HREI) from November 2015 to July 2016. Fifty patients triaged for next-day appointments at the HREI were randomly assigned to receive standard-of-care preappointment information or standard-of-care information in addition to an estimated appointment wait time. Patient satisfaction with care was assessed postvisit using the modified Judgements of Hospital Quality Questionnaire (JHQQ). In determining how informing patients of typical wait times influenced satisfaction, the Mann-Whitney U test was performed. As secondary study outcomes, we sought to determine patient satisfaction with the intervention material using the Fisher exact test and the effect that wait time, age, sex, education, mobility, and number of health care providers seen had on satisfaction scores using logistic regression analysis. The median JHQQ response was "very good" (4/5) and between "very good" and "excellent" (4.5/5) in the intervention and control arms, respectively. There was no difference in patient satisfaction between the cohorts (Mann-Whitney U = 297.00, p = 0.964). Logistic regression analysis demonstrated that wait times influenced patient satisfaction (OR = 0.919, 95% CI 0.864-0.978, p = 0.008). Of the intervention arm patients, 92.0% (N = 23) found the preappointment information useful, whereas only 12.5% (N = 3) of the control cohort patients noted the same (p < 0.001). Provision of anticipated wait time information to patients in an emergency on-call ophthalmology clinic did not influence satisfaction with care as captured by the JHQQ. Copyright © 2018 Canadian Ophthalmological Society. Published by Elsevier Inc. All rights reserved.

  8. Transition in the waiting-time distribution of price-change events in a global socioeconomic system

    NASA Astrophysics Data System (ADS)

    Zhao, Guannan; McDonald, Mark; Fenn, Dan; Williams, Stacy; Johnson, Nicholas; Johnson, Neil F.

    2013-12-01

    The goal of developing a firmer theoretical understanding of inhomogeneous temporal processes-in particular, the waiting times in some collective dynamical system-is attracting significant interest among physicists. Quantifying the deviations between the waiting-time distribution and the distribution generated by a random process may help unravel the feedback mechanisms that drive the underlying dynamics. We analyze the waiting-time distributions of high-frequency foreign exchange data for the best executable bid-ask prices across all major currencies. We find that the lognormal distribution yields a good overall fit for the waiting-time distribution between currency rate changes if both short and long waiting times are included. If we restrict our study to long waiting times, each currency pair’s distribution is consistent with a power-law tail with exponent near to 3.5. However, for short waiting times, the overall distribution resembles one generated by an archetypal complex systems model in which boundedly rational agents compete for limited resources. Our findings suggest that a gradual transition arises in trading behavior between a fast regime in which traders act in a boundedly rational way and a slower one in which traders’ decisions are driven by generic feedback mechanisms across multiple timescales and hence produce similar power-law tails irrespective of currency type.

  9. The effect of in-office waiting time on physician visit frequency among working-age adults.

    PubMed

    Tak, Hyo Jung; Hougham, Gavin W; Ruhnke, Atsuko; Ruhnke, Gregory W

    2014-10-01

    Disparities in unmet health care demand resulting from socioeconomic, racial, and financial factors have received a great deal of attention in the United States. However, out-of-pocket costs alone do not fully reflect the total opportunity cost that patients must consider as they seek medical attention. While there is an extensive literature on the price elasticity of demand for health care, empirical evidence regarding the effect of waiting time on utilization is sparse. Using the nationally representative 2003 Community Tracking Study Household Survey, the most recent iteration containing respondents' physician office visit frequency and estimated in-office waiting time in the United States (N = 23,484), we investigated the association between waiting time and calculated time cost with the number of physician visits among a sample of working-age adults. To avoid the bias that literature suggests would result from excluding respondents with zero physician visits, we imputed waiting time for the essential inclusion of such individuals. On average, respondents visited physician offices 3.55 times, during which time they waited 28.7 min. The estimates from a negative binomial model indicated that a doubling of waiting time was associated with a 7.7 percent decrease (p-value < 0.001) in physician visit frequency. For women and unemployed respondents, who visited physicians more frequently, the decrease was even larger, suggesting a stronger response to greater waiting times. We believe this finding reflects the discretionary nature of incremental visits in these groups, and a consequent lower perceived marginal benefit of additional visits. The results suggest that in-office waiting time may have a substantial influence on patients' propensity to seek medical attention. Although there is a belief that expansions in health insurance coverage increase health care utilization by reducing financial barriers to access, our results suggest that unintended consequences

  10. Platelet function measurement-based strategy to reduce bleeding and waiting time in clopidogrel-treated patients undergoing coronary artery bypass graft surgery: the timing based on platelet function strategy to reduce clopidogrel-associated bleeding related to CABG (TARGET-CABG) study.

    PubMed

    Mahla, Elisabeth; Suarez, Thomas A; Bliden, Kevin P; Rehak, Peter; Metzler, Helfried; Sequeira, Alejandro J; Cho, Peter; Sell, Jeffery; Fan, John; Antonino, Mark J; Tantry, Udaya S; Gurbel, Paul A

    2012-04-01

    Aspirin and clopidogrel therapy is associated with a variable bleeding risk in patients undergoing coronary artery bypass graft surgery (CABG). We evaluated the role of platelet function testing in clopidogrel-treated patients undergoing CABG. One hundred eighty patients on background aspirin with/without clopidogrel therapy undergoing elective first time isolated on-pump CABG were enrolled in a prospective single-center, nonrandomized, unblinded investigation (Timing Based on Platelet Function Strategy to Reduce Clopidogrel-Associated Bleeding Related to CABG [TARGET-CABG] study) between September 2008 and January 2011. Clopidogrel responsiveness (ADP-induced platelet-fibrin clot strength [MA(ADP)]) was determined by thrombelastography; CABG was done within 1 day, 3-5 days, and >5 days in patients with an MA(ADP) >50 mm, 35-50 mm, and <35 mm, respectively. The primary end point was 24-hour chest tube drainage and key secondary end point was total number of transfused red blood cells. Equivalence was defined as ≤25% difference between groups. ANCOVA was used to adjust for confounders. Mean 24-hour chest tube drainage in clopidogrel-treated patients was 93% (95% confidence interval, 81-107%) of the amount observed in clopidogrel-naive patients, and the total amount of red blood cells transfused did not differ between groups (1.80 U versus 2.08 U, respectively, P=0.540). The total waiting period in clopidogrel-treated patients was 233 days (mean, 2.7 days per patient). A strategy based on preoperative platelet function testing to determine the timing of CABG in clopidogrel-treated patients was associated with the same amount of bleeding observed in clopidogrel-naive patients and ≈50% shorter waiting time than recommended in the current guidelines. URL: http://www.clinicaltrials.gov. Unique identifier: NCT00857155.

  11. Decreasing Psychiatric Admission Wait Time in the Emergency Department by Facilitating Psychiatric Discharges.

    PubMed

    Stover, Pamela R; Harpin, Scott

    2015-12-01

    Limited capacity in a psychiatric unit contributes to long emergency department (ED) admission wait times. Regulatory and accrediting agencies urge hospitals nationally to improve patient flow for better access to care for all types of patients. The purpose of the current study was to decrease psychiatric admission wait time from 10.5 to 8 hours and increase the proportion of patients discharged by 11 a.m. from 20% to 50%. The current study compared pre- and post-intervention data. Plan-Do-Study-Act cycles aimed to improve discharge processes and timeliness through initiation of new practices. Admission wait time improved to an average of 5.1 hours (t = 3.87, p = 0.006). The proportion of discharges occurring by 11 a.m. increased to 46% (odds ratio = 3.42, p < 0.0001). Improving discharge planning processes and timeliness in a psychiatric unit significantly decreased admission wait time from the ED, improving access to psychiatric care. Copyright 2015, SLACK Incorporated.

  12. CULTURAL COMPETENCE IN OUTPATIENT SUBSTANCE ABUSE TREATMENT: MEASUREMENT AND RELATIONSHIP TO WAIT TIME AND RETENTION

    PubMed Central

    Guerrero, Erick; Andrews, Christina M.

    2011-01-01

    BACKGROUND Culturally competent practice is broadly acknowledged to be an important strategy to increase the quality of services for racial/ethnic minorities in substance abuse treatment. However, few empirically derived measures of organizational cultural competence exist, and relatively little is known about how these measures affect treatment outcomes. METHOD Using a nationally representative sample of outpatient substance abuse treatment (OSAT) programs, this study used item response theory to create two measures of cultural competence-organizational practices and managers' culturally sensitive beliefs—and examined their relationship to client wait time and retention using Poisson regression modeling RESULTS The most common and precisely measured organizational practices reported by OSAT managers included matching providers and clients based on language/dialect; offering cross-cultural training; and fostering connections with community and faith-based organizations connected to racial and ethnic minority groups. The most culturally sensitive belief among OSAT managers was support for language/dialect matching for racial and ethnic minority clients. Results of regression modeling indicate that organizational practices were not related to either outcome. However, managers' culturally sensitive beliefs were negatively associated with average wait time (p < 0.05), and positively associated with average retention (p < 0.01). CONCLUSIONS Managers' culturally sensitive beliefs—considered to be influential for effective implementation of culturally competent practices—may be particularly relevant in influencing wait time and retention in OSAT organizations that treat Latinos and African American clients. PMID:21680111

  13. The ecology of the patient visit: physical attractiveness, waiting times, and perceived quality of care.

    PubMed

    Becker, Franklin; Douglass, Stephanie

    2008-01-01

    This study examined the relationship between the attractiveness of the physical environment of healthcare facilities and patient perceptions of quality, service, and waiting time through systematic observations and patient satisfaction surveys at 7 outpatient practices at Weill Cornell Medical Center. Findings indicate positive correlations between more attractive environments and higher levels of perceived quality, satisfaction, staff interaction, and reduction of patient anxiety. The comparison of actual observed time and patients' perception of time showed that patients tend to overestimate shorter waiting times and underestimate longer waiting times in both the waiting area and the examination room. Further examinations of the way outpatient-practice environments impact patient and staff perceptions and how those perceptions impact behavior and medical outcomes are suggested.

  14. Improving wait times to care for individuals with multimorbidities and complex conditions using value stream mapping.

    PubMed

    Sampalli, Tara; Desy, Michel; Dhir, Minakshi; Edwards, Lynn; Dickson, Robert; Blackmore, Gail

    2015-04-05

    Recognizing the significant impact of wait times for care for individuals with complex chronic conditions, we applied a LEAN methodology, namely - an adaptation of Value Stream Mapping (VSM) to meet the needs of people with multiple chronic conditions and to improve wait times without additional resources or funding. Over an 18-month time period, staff applied a patient-centric approach that included LEAN methodology of VSM to improve wait times to care. Our framework of evaluation was grounded in the needs and perspectives of patients and individuals waiting to receive care. Patient centric views were obtained through surveys such as Patient Assessment of Chronic Illness Care (PACIC) and process engineering based questions. In addition, LEAN methodology, VSM was added to identify non-value added processes contributing to wait times. The care team successfully reduced wait times to 2 months in 2014 with no wait times for care anticipated in 2015. Increased patient engagement and satisfaction are also outcomes of this innovative initiative. In addition, successful transformations and implementation have resulted in resource efficiencies without increase in costs. Patients have shown significant improvements in functional health following Integrated Chronic Care Service (ICCS) intervention. The methodology will be applied to other chronic disease management areas in Capital Health and the province. Wait times to care in the management of multimoribidities and other complex conditions can add a significant burden not only on the affected individuals but also on the healthcare system. In this study, a novel and modified LEAN methodology has been applied to embed the voice of the patient in care delivery processes and to reduce wait times to care in the management of complex chronic conditions. © 2015 by Kerman University of Medical Sciences.

  15. Wait Time for Treatment in Hospital Emergency Departments: 2009

    MedlinePlus

    ... on Vital and Health Statistics Annual Reports Health Survey Research Methods Conference Reports from the National Medical Care ... SOURCE: CDC/NCHS, National Hospital Ambulatory Medical Care ... with previous research, longer wait time for treatment was associated with ...

  16. Effect of socioeconomic deprivation on waiting time for cardiac surgery: retrospective cohort study

    PubMed Central

    Pell, Jill P; Pell, Alastair C H; Norrie, John; Ford, Ian; Cobbe, Stuart M

    2000-01-01

    Objective To determine whether the priority given to patients referred for cardiac surgery is associated with socioeconomic status. Design Retrospective study with multivariate logistic regression analysis of the association between deprivation and classification of urgency with allowance for age, sex, and type of operation. Multivariate linear regression analysis was used to determine association between deprivation and waiting time within each category of urgency, with allowance for age, sex, and type of operation. Setting NHS waiting lists in Scotland. Participants 26 642 patients waiting for cardiac surgery, 1 January 1986 to 31 December 1997. Main outcome measures Deprivation as measured by Carstairs deprivation category. Time spent on NHS waiting list. Results Patients who were most deprived tended to be younger and were more likely to be female. Patients in deprivation categories 6 and 7 (most deprived) waited about three weeks longer for surgery than those in category 1 (mean difference 24 days, 95% confidence interval 15 to 32). Deprived patients had an odds ratio of 0.5 (0.46 to 0.61) for having their operations classified as urgent compared with the least deprived, after allowance for age, sex, and type of operation. When urgent and routine cases were considered separately, there was no significant difference in waiting times between the most and least deprived categories. Conclusions Socioeconomically deprived patients are thought to be more likely to develop coronary heart disease but are less likely to be investigated and offered surgery once it has developed. Such patients may be further disadvantaged by having to wait longer for surgery because of being given lower priority. PMID:10617517

  17. A Cross-Sectional Survey of Population-Wide Wait Times for Patients Seeking Medical vs. Cosmetic Dermatologic Care.

    PubMed

    Yadav, Geeta; Goldberg, Hanna R; Barense, Morgan D; Bell, Chaim M

    2016-01-01

    Though previous work has examined some aspects of the dermatology workforce shortage and access to dermatologic care, little research has addressed the effect of rising interest in cosmetic procedures on access to medical dermatologic care. Our objective was to determine the wait times for Urgent and Non-Urgent medical dermatologic care and Cosmetic dermatology services at a population level and to examine whether wait times for medical care are affected by offering cosmetic services. A population-wide survey of dermatology practices using simulated calls asking for the earliest appointment for a Non-Urgent, Urgent and Cosmetic service. Response rates were greater than 89% for all types of care. Wait times across all types of care were significantly different from each other (all P < 0.05). Cosmetic care was associated with the shortest wait times (3.0 weeks; Interquartile Range (IQR) = 0.4-3.4), followed by Urgent care (9.0 weeks; IQR = 2.1-12.9), then Non-Urgent Care (12.7 weeks; IQR = 4.4-16.4). Wait times for practices offering only Urgent care were not different from practices offering both Urgent and Cosmetic care (10.3 vs. 7.0 weeks). Longer wait times and greater variation for Urgent and Non-Urgent dermatologic care and shorter wait times and less variation for Cosmetic care. Wait times were significantly longer in regions with lower dermatologist density. Provision of Cosmetic services did not increase wait times for Urgent care. These findings suggest an overall dermatology workforce shortage and a need for a more streamlined referral system for dermatologic care.

  18. Renewal processes based on generalized Mittag-Leffler waiting times

    NASA Astrophysics Data System (ADS)

    Cahoy, Dexter O.; Polito, Federico

    2013-03-01

    The fractional Poisson process has recently attracted experts from several fields of study. Its natural generalization of the ordinary Poisson process made the model more appealing for real-world applications. In this paper, we generalized the standard and fractional Poisson processes through the waiting time distribution, and showed their relations to an integral operator with a generalized Mittag-Leffler function in the kernel. The waiting times of the proposed renewal processes have the generalized Mittag-Leffler and stretched-squashed Mittag-Leffler distributions. Note that the generalizations naturally provide greater flexibility in modeling real-life renewal processes. Algorithms to simulate sample paths and to estimate the model parameters are derived. Note also that these procedures are necessary to make these models more usable in practice. State probabilities and other qualitative or quantitative features of the models are also discussed.

  19. Creation of a diagnostic wait times measurement framework based on evidence and consensus.

    PubMed

    Gilbert, Julie E; Dobrow, Mark J; Kaan, Melissa; Dobranowski, Julian; Srigley, John R; Jusko Friedman, Audrey; Irish, Jonathan C

    2014-09-01

    Public reporting of wait times worldwide has to date focused largely on treatment wait times and is limited in its ability to capture earlier parts of the patient journey. The interval between suspicion and diagnosis or ruling out of cancer is a complex phase of the cancer journey. Diagnostic delays and inefficient use of diagnostic imaging procedures can result in poor patient outcomes, both physical and psychosocial. This study was designed to develop a framework that could be adopted for multiple disease sites across different jurisdictions to enable the measurement of diagnostic wait times and diagnostic delay. Diagnostic benchmarks and targets in cancer systems were explored through a targeted literature review and jurisdictional scan. Cancer system leaders and clinicians were interviewed to validate the information found in the jurisdictional scan. An expert panel was assembled to review and, through a modified Delphi consensus process, provide feedback on a diagnostic wait times framework. The consensus process resulted in agreement on a measurement framework that identified suspicion, referral, diagnosis, and treatment as the main time points for measuring this critical phase of the patient journey. This work will help guide initiatives designed to improve patient access to health services by developing an evidence-based approach to standardization of the various waypoints during the diagnostic pathway. The diagnostic wait times measurement framework provides a yardstick to measure the performance of programs that are designed to manage and expedite care processes between referral and diagnosis or ruling out of cancer. Copyright © 2014 by American Society of Clinical Oncology.

  20. Differential Effects of Wait-Time on Textually Explicit and Implicit Responding: Interactional Explanation.

    ERIC Educational Resources Information Center

    Pond, Marlene R.; Newman, Isadore

    The effects of wait-time, the pause following a teacher question and the pause after a student response, on the length and number of student responses were analyzed at different cognitive levels. Data were obtained from 95 students in grade 4 and from 5 teachers using a wait-time of 5 seconds. Four oral discussion sessions by teachers and students…

  1. Estimating the waiting time of multi-priority emergency patients with downstream blocking.

    PubMed

    Lin, Di; Patrick, Jonathan; Labeau, Fabrice

    2014-03-01

    To characterize the coupling effect between patient flow to access the emergency department (ED) and that to access the inpatient unit (IU), we develop a model with two connected queues: one upstream queue for the patient flow to access the ED and one downstream queue for the patient flow to access the IU. Building on this patient flow model, we employ queueing theory to estimate the average waiting time across patients. Using priority specific wait time targets, we further estimate the necessary number of ED and IU resources. Finally, we investigate how an alternative way of accessing ED (Fast Track) impacts the average waiting time of patients as well as the necessary number of ED/IU resources. This model as well as the analysis on patient flow can help the designer or manager of a hospital make decisions on the allocation of ED/IU resources in a hospital.

  2. A Cross-Sectional Survey of Population-Wide Wait Times for Patients Seeking Medical vs. Cosmetic Dermatologic Care

    PubMed Central

    Yadav, Geeta; Goldberg, Hanna R.; Barense, Morgan D.; Bell, Chaim M.

    2016-01-01

    Background Though previous work has examined some aspects of the dermatology workforce shortage and access to dermatologic care, little research has addressed the effect of rising interest in cosmetic procedures on access to medical dermatologic care. Our objective was to determine the wait times for Urgent and Non-Urgent medical dermatologic care and Cosmetic dermatology services at a population level and to examine whether wait times for medical care are affected by offering cosmetic services. Methods A population-wide survey of dermatology practices using simulated calls asking for the earliest appointment for a Non-Urgent, Urgent and Cosmetic service. Results Response rates were greater than 89% for all types of care. Wait times across all types of care were significantly different from each other (all P < 0.05). Cosmetic care was associated with the shortest wait times (3.0 weeks; Interquartile Range (IQR) = 0.4–3.4), followed by Urgent care (9.0 weeks; IQR = 2.1–12.9), then Non-Urgent Care (12.7 weeks; IQR = 4.4–16.4). Wait times for practices offering only Urgent care were not different from practices offering both Urgent and Cosmetic care (10.3 vs. 7.0 weeks). Interpretation Longer wait times and greater variation for Urgent and Non-Urgent dermatologic care and shorter wait times and less variation for Cosmetic care. Wait times were significantly longer in regions with lower dermatologist density. Provision of Cosmetic services did not increase wait times for Urgent care. These findings suggest an overall dermatology workforce shortage and a need for a more streamlined referral system for dermatologic care. PMID:27632206

  3. Application of Queueing Theory to the Analysis of Changes in Outpatients' Waiting Times in Hospitals Introducing EMR

    PubMed Central

    Cho, Kyoung Won; Kim, Seong Min; Chae, Young Moon

    2017-01-01

    Objectives This research used queueing theory to analyze changes in outpatients' waiting times before and after the introduction of Electronic Medical Record (EMR) systems. Methods We focused on the exact drawing of two fundamental parameters for queueing analysis, arrival rate (λ) and service rate (µ), from digital data to apply queueing theory to the analysis of outpatients' waiting times. We used outpatients' reception times and consultation finish times to calculate the arrival and service rates, respectively. Results Using queueing theory, we could calculate waiting time excluding distorted values from the digital data and distortion factors, such as arrival before the hospital open time, which occurs frequently in the initial stage of a queueing system. We analyzed changes in outpatients' waiting times before and after the introduction of EMR using the methodology proposed in this paper, and found that the outpatients' waiting time decreases after the introduction of EMR. More specifically, the outpatients' waiting times in the target public hospitals have decreased by rates in the range between 44% and 78%. Conclusions It is possible to analyze waiting times while minimizing input errors and limitations influencing consultation procedures if we use digital data and apply the queueing theory. Our results verify that the introduction of EMR contributes to the improvement of patient services by decreasing outpatients' waiting time, or by increasing efficiency. It is also expected that our methodology or its expansion could contribute to the improvement of hospital service by assisting the identification and resolution of bottlenecks in the outpatient consultation process. PMID:28261529

  4. Application of Queueing Theory to the Analysis of Changes in Outpatients' Waiting Times in Hospitals Introducing EMR.

    PubMed

    Cho, Kyoung Won; Kim, Seong Min; Chae, Young Moon; Song, Yong Uk

    2017-01-01

    This research used queueing theory to analyze changes in outpatients' waiting times before and after the introduction of Electronic Medical Record (EMR) systems. We focused on the exact drawing of two fundamental parameters for queueing analysis, arrival rate (λ) and service rate (µ), from digital data to apply queueing theory to the analysis of outpatients' waiting times. We used outpatients' reception times and consultation finish times to calculate the arrival and service rates, respectively. Using queueing theory, we could calculate waiting time excluding distorted values from the digital data and distortion factors, such as arrival before the hospital open time, which occurs frequently in the initial stage of a queueing system. We analyzed changes in outpatients' waiting times before and after the introduction of EMR using the methodology proposed in this paper, and found that the outpatients' waiting time decreases after the introduction of EMR. More specifically, the outpatients' waiting times in the target public hospitals have decreased by rates in the range between 44% and 78%. It is possible to analyze waiting times while minimizing input errors and limitations influencing consultation procedures if we use digital data and apply the queueing theory. Our results verify that the introduction of EMR contributes to the improvement of patient services by decreasing outpatients' waiting time, or by increasing efficiency. It is also expected that our methodology or its expansion could contribute to the improvement of hospital service by assisting the identification and resolution of bottlenecks in the outpatient consultation process.

  5. In the queue for coronary artery bypass grafting: patients' perceptions of risk and 'maximal acceptable waiting time'.

    PubMed

    Llewellyn-Thomas, H; Thiel, E; Paterson, M; Naylor, D

    1999-04-01

    To elicit patients' maximal acceptable waiting times (MAWT) for non-urgent coronary artery bypass grafting (CABG), and to determine if MAWT is related to prior expectations of waiting times, symptom burden, expected relief, or perceived risks of myocardial infarction while waiting. Seventy-two patients on an elective CABG waiting list chose between two hypothetical but plausible options: a 1-month wait with 2% risk of surgical mortality, and a 6-month wait with 1% risk of surgical mortality. Waiting time in the 6-month option was varied up if respondents chose the 6-month/lower risk option, and down if they chose the 1-month/higher risk option, until the MAWT switch point was reached. Patients also reported their expected waiting time, perceived risks of myocardial infarction while waiting, current function, expected functional improvement and the value of that improvement. Only 17 (24%) patients chose the 6-month/1% risk option, while 55 (76%) chose the 1-month/2% risk option. The median MAWT was 2 months; scores ranged from 1 to 12 months (with two outliers). Many perceived high cumulative risks of myocardial infarction if waiting for 1 (upper quartile, > or = 1.45%) or 6 (upper quartile, > or = 10%) months. However, MAWT scores were related only to expected waiting time (r = 0.47; P < 0.0001). Most patients reject waiting 6 months for elective CABG, even if offered along with a halving in surgical mortality (from 2% to 1%). Intolerance for further delay seems to be determined primarily by patients' attachment to their scheduled surgical dates. Many also have severely inflated perceptions of their risk of myocardial infarction in the queue. These results suggest a need for interventions to modify patients' inaccurate risk perceptions, particularly if a scheduled surgical date must be deferred.

  6. Practical solutions for reducing container ships' waiting times at ports using simulation model

    NASA Astrophysics Data System (ADS)

    Sheikholeslami, Abdorreza; Ilati, Gholamreza; Yeganeh, Yones Eftekhari

    2013-12-01

    The main challenge for container ports is the planning required for berthing container ships while docked in port. Growth of containerization is creating problems for ports and container terminals as they reach their capacity limits of various resources which increasingly leads to traffic and port congestion. Good planning and management of container terminal operations reduces waiting time for liner ships. Reducing the waiting time improves the terminal's productivity and decreases the port difficulties. Two important keys to reducing waiting time with berth allocation are determining suitable access channel depths and increasing the number of berths which in this paper are studied and analyzed as practical solutions. Simulation based analysis is the only way to understand how various resources interact with each other and how they are affected in the berthing time of ships. We used the Enterprise Dynamics software to produce simulation models due to the complexity and nature of the problems. We further present case study for berth allocation simulation of the biggest container terminal in Iran and the optimum access channel depth and the number of berths are obtained from simulation results. The results show a significant reduction in the waiting time for container ships and can be useful for major functions in operations and development of container ship terminals.

  7. Sci-Fri AM: Quality, Safety, and Professional Issues 04: Predicting waiting times in Radiation Oncology using machine learning

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

    Joseph, Ackeem; Herrera, David; Hijal, Tarek

    We describe a method for predicting waiting times in radiation oncology. Machine learning is a powerful predictive modelling tool that benefits from large, potentially complex, datasets. The essence of machine learning is to predict future outcomes by learning from previous experience. The patient waiting experience remains one of the most vexing challenges facing healthcare. Waiting time uncertainty can cause patients, who are already sick and in pain, to worry about when they will receive the care they need. In radiation oncology, patients typically experience three types of waiting: Waiting at home for their treatment plan to be prepared Waiting inmore » the waiting room for daily radiotherapy Waiting in the waiting room to see a physician in consultation or follow-up These waiting periods are difficult for staff to predict and only rough estimates are typically provided, based on personal experience. In the present era of electronic health records, waiting times need not be so uncertain. At our centre, we have incorporated the electronic treatment records of all previously-treated patients into our machine learning model. We found that the Random Forest Regression model provides the best predictions for daily radiotherapy treatment waiting times (type 2). Using this model, we achieved a median residual (actual minus predicted value) of 0.25 minutes and a standard deviation residual of 6.5 minutes. The main features that generated the best fit model (from most to least significant) are: Allocated time, median past duration, fraction number and the number of treatment fields.« less

  8. Decreasing laboratory turnaround time and patient wait time by implementing process improvement methodologies in an outpatient oncology infusion unit.

    PubMed

    Gjolaj, Lauren N; Gari, Gloria A; Olier-Pino, Angela I; Garcia, Juan D; Fernandez, Gustavo L

    2014-11-01

    Prolonged patient wait times in the outpatient oncology infusion unit indicated a need to streamline phlebotomy processes by using existing resources to decrease laboratory turnaround time and improve patient wait time. Using the DMAIC (define, measure, analyze, improve, control) method, a project to streamline phlebotomy processes within the outpatient oncology infusion unit in an academic Comprehensive Cancer Center known as the Comprehensive Treatment Unit (CTU) was completed. Laboratory turnaround time for patients who needed same-day lab and CTU services and wait time for all CTU patients was tracked for 9 weeks. During the pilot, the wait time from arrival to CTU to sitting in treatment area decreased by 17% for all patients treated in the CTU during the pilot. A total of 528 patients were seen at the CTU phlebotomy location, representing 16% of the total patients who received treatment in the CTU, with a mean turnaround time of 24 minutes compared with a baseline turnaround time of 51 minutes. Streamlining workflows and placing a phlebotomy station inside of the CTU decreased laboratory turnaround times by 53% for patients requiring same day lab and CTU services. The success of the pilot project prompted the team to make the station a permanent fixture. Copyright © 2014 by American Society of Clinical Oncology.

  9. Effects of Wait Time When Communicating with Children Who Have Sensory and Additional Disabilities

    ERIC Educational Resources Information Center

    Johnson, Nicole; Parker, Amy T.

    2013-01-01

    Introduction: This study utilized wait-time procedures to determine if they are effective in helping children with deafblindness or multiple disabilities that include a visual impairment communicate in their home. Methods: A single subject with an alternating treatment design was used for the study. Zero- to one-second wait time was utilized…

  10. The impact of diagnostic imaging wait times on the prognosis of lung cancer.

    PubMed

    Byrne, Suzanne C; Barrett, Brendan; Bhatia, Rick

    2015-02-01

    This study was performed to determine whether gaps in patient flow from initial lung imaging to computed tomography (CT) guided lung biopsy in patients with non-small cell lung cancer (NSCLC) was associated with a change in tumour size, stage, and thus prognosis. All patients who had a CT-guided lung biopsy in 2009 (phase I) and in 2011 (phase II) with a pathologic diagnosis of primary lung cancer (NSCLC) at Eastern Health, Newfoundland, were identified. Dates of initial abnormal imaging, confirmatory CT (if performed), and CT-guided biopsy were recorded, along with tumour size and resulting T stage at each time point. In 2010, wait times for diagnostic imaging at Eastern Health were reduced. The stage and prognosis of NSCLC in 2009 was compared with 2011. In phase 1, there was a statistically significant increase in tumour size (mean difference, 0.67 cm; P < .0001) and stage (P < .0001) from initial image to biopsy. There was a moderate correlation between the time (in days) between the images and change in size (r = 0.33, P = .008) or stage (r = 0.26, P = .036). In phase II, the median wait time from initial imaging to confirmatory CT was reduced to 7.5 days (from 19 days). At this reduced wait time, there was no statistically significant increase in tumour size (mean difference, 0.02; P > .05) or stage (P > .05) from initial imaging to confirmatory CT. Delays in patient flow through diagnostic imaging resulted in an increase in tumour size and stage, with a negative impact on prognosis of NSCLC. This information contributed to the hiring of additional CT technologists and extended CT hours to decrease the wait time for diagnostic imaging. With reduced wait times, the prognosis of NSCLC was not adversely impacted as patients navigated through diagnostic imaging. Copyright © 2015 Canadian Association of Radiologists. All rights reserved.

  11. Using Queuing Theory and Simulation Modelling to Reduce Waiting Times in An Iranian Emergency Department

    PubMed Central

    Haghighinejad, Hourvash Akbari; Kharazmi, Erfan; Hatam, Nahid; Yousefi, Sedigheh; Hesami, Seyed Ali; Danaei, Mina; Askarian, Mehrdad

    2016-01-01

    Background: Hospital emergencies have an essential role in health care systems. In the last decade, developed countries have paid great attention to overcrowding crisis in emergency departments. Simulation analysis of complex models for which conditions will change over time is much more effective than analytical solutions and emergency department (ED) is one of the most complex models for analysis. This study aimed to determine the number of patients who are waiting and waiting time in emergency department services in an Iranian hospital ED and to propose scenarios to reduce its queue and waiting time. Methods: This is a cross-sectional study in which simulation software (Arena, version 14) was used. The input information was extracted from the hospital database as well as through sampling. The objective was to evaluate the response variables of waiting time, number waiting and utilization of each server and test the three scenarios to improve them. Results: Running the models for 30 days revealed that a total of 4088 patients left the ED after being served and 1238 patients waited in the queue for admission in the ED bed area at end of the run (actually these patients received services out of their defined capacity). The first scenario result in the number of beds had to be increased from 81 to179 in order that the number waiting of the “bed area” server become almost zero. The second scenario which attempted to limit hospitalization time in the ED bed area to the third quartile of the serving time distribution could decrease the number waiting to 586 patients. Conclusion: Doubling the bed capacity in the emergency department and consequently other resources and capacity appropriately can solve the problem. This includes bed capacity requirement for both critically ill and less critically ill patients. Classification of ED internal sections based on severity of illness instead of medical specialty is another solution. PMID:26793727

  12. Influence of nurse navigation on wait times for breast cancer care in a Canadian regional cancer center.

    PubMed

    Baliski, Christopher; McGahan, Colleen E; Liberto, Caitlyn M; Broughton, Sandra; Ellard, Susan; Taylor, Marianne; Bates, Janet; Lai, Anky

    2014-05-01

    The wait times for breast cancer care in our region do not meet acceptable benchmarks. We implemented the Interior Breast Rapid Access Investigation and Diagnosis (IB-RAPID) nurse navigation program to address this issue. The IB-RAPID prospective database was reviewed for patients entering the program between April 1, 2011 and April 30, 2012 (2011/2012 cohort), and was compared with patients from the same area in 2010. The main end point was the time between the 1st diagnostic imaging test and the surgery. Multiple linear regression was performed to investigate factors influencing the wait times. The wait times decreased with the introduction of IB-RAPID (59 vs 48 days; median). Stage of disease, total number of biopsies, and magnetic resonance imaging (MRI) use influenced wait times. MRI significantly delayed surgical intervention in both groups with those not having an MRI having a shorter wait time to surgery (68.5 vs 57.6 days; mean) in 2011/2012. The implementation of nurse navigation for patients with breast cancer appears to be effective at reducing the wait times for surgical treatment. Copyright © 2014 Elsevier Inc. All rights reserved.

  13. Parallel-Batch Scheduling and Transportation Coordination with Waiting Time Constraint

    PubMed Central

    Gong, Hua; Chen, Daheng; Xu, Ke

    2014-01-01

    This paper addresses a parallel-batch scheduling problem that incorporates transportation of raw materials or semifinished products before processing with waiting time constraint. The orders located at the different suppliers are transported by some vehicles to a manufacturing facility for further processing. One vehicle can load only one order in one shipment. Each order arriving at the facility must be processed in the limited waiting time. The orders are processed in batches on a parallel-batch machine, where a batch contains several orders and the processing time of the batch is the largest processing time of the orders in it. The goal is to find a schedule to minimize the sum of the total flow time and the production cost. We prove that the general problem is NP-hard in the strong sense. We also demonstrate that the problem with equal processing times on the machine is NP-hard. Furthermore, a dynamic programming algorithm in pseudopolynomial time is provided to prove its ordinarily NP-hardness. An optimal algorithm in polynomial time is presented to solve a special case with equal processing times and equal transportation times for each order. PMID:24883385

  14. Lévy walks with variable waiting time: A ballistic case

    NASA Astrophysics Data System (ADS)

    Kamińska, A.; Srokowski, T.

    2018-06-01

    The Lévy walk process for a lower interval of an excursion times distribution (α <1 ) is discussed. The particle rests between the jumps, and the waiting time is position-dependent. Two cases are considered: a rising and diminishing waiting time rate ν (x ) , which require different approximations of the master equation. The process comprises two phases of the motion: particles at rest and in flight. The density distributions for them are derived, as a solution of corresponding fractional equations. For strongly falling ν (x ) , the resting particles density assumes the α -stable form (truncated at fronts), and the process resolves itself to the Lévy flights. The diffusion is enhanced for this case but no longer ballistic, in contrast to the case for the rising ν (x ) . The analytical results are compared with Monte Carlo trajectory simulations. The results qualitatively agree with observed properties of human and animal movements.

  15. Strategy as active waiting.

    PubMed

    Sull, Donald N

    2005-09-01

    Successful executives who cut their teeth in stable industries or in developed countries often stumble when they face more volatile markets. They falter, in part, because they assume they can gaze deep into the future and develop a long-term strategy that will confer a sustainable competitive advantage. But visibility into the future of volatile markets is sharply limited because so many different variables are in play. Factors such as technological innovation, customers' evolving needs, government policy, and changes in the capital markets interact with one another to create unexpected outcomes. Over the past six years, Donald Sull, an associate professor at London Business School, has led a research project examining some of the world's most volatile markets, from national markets like China and Brazil to industries like enterprise software, telecommunications, and airlines. One of the most striking findings from this research is the importance of taking action during comparative lulls in the storm. Huge business opportunities are relatively rare; they come along only once or twice in a decade. And, for the most part, companies can't manufacture those opportunities; changes in the external environment converge to make them happen. What managers can do is prepare for these golden opportunities by managing smart during the comparative calm of business as usual. During these periods of active waiting, leaders must probe the future and remain alert to anomalies that signal potential threats or opportunities; exercise restraint to preserve their war chests; and maintain discipline to keep the troops battle ready. When a golden opportunity or"sudden death"threat emerges, managers must have the courage to declare the main effort and concentrate resources to seize the moment.

  16. Impact of waiting time on the quality of life of patients awaiting coronary artery bypass grafting.

    PubMed

    Sampalis, J; Boukas, S; Liberman, M; Reid, T; Dupuis, G

    2001-08-21

    A lack of resources has created waiting lists for many elective surgical procedures within Canada's universal health care system. Coronary artery bypass grafting (CABG) for the treatment of atherosclerotic ischemic heart disease is one of these affected surgical procedures. We studied the impact of waiting times on the quality of life of patients awaiting CABG. A prospective cohort of 266 patients from 3 hospitals in Montreal was used. Patients who gave informed consent were followed from the time they were registered for CABG until 6 months after surgery; recruitment began in November 1993, and the last follow-up was completed in July 1995. Patient groups were classified according to the duration of the wait for CABG (< or = 97 days or > 97 days). We measured the following outcomes: quality of life (using the Medical Outcomes Study 36-item Short Form [SF-36]), incidence of chest pain (using the New York Heart Association angina classification), frequency of symptoms (using the Cardiac Symptom Inventory) and rates of complications and death before and after surgery. There were no differences in quality of life at baseline between the 2 groups. Immediately before surgery, compared with patients who waited 97 days or less, those who waited longer had significantly reduced physical functioning (change from baseline SF-36 score 0 v. -4 respectively, p = 0.001), vitality (change from baseline score -0.1 v. -1.3, p = 0.01), social functioning (change from baseline score 0.4 v. -0.4, p = 0.03) and general health (change from baseline score 1.1 v. -1.7, p = 0.001). At 6 months after surgery, compared with patients who waited 97 days or less for CABG, those who waited longer had reduced physical functioning (change from baseline SF-36 score 4.0 v. -0.1 respectively, p = 0.001), physical role (change from baseline score 0.8 v. 0.0, p = 0.001), vitality (change from baseline score 2.2 v. 0.9, p = 0.001), mental health (change from baseline score 1.2 v. 0.0, p = 0.001) and

  17. Impact of pharmacy automation on patient waiting time: an application of computer simulation.

    PubMed

    Tan, Woan Shin; Chua, Siang Li; Yong, Keng Woh; Wu, Tuck Seng

    2009-06-01

    This paper aims to illustrate the use of computer simulation in evaluating the impact of a prototype automated dispensing system on waiting time in an outpatient pharmacy and its potential as a routine tool in pharmacy management. A discrete event simulation model was developed to investigate the impact of a prototype automated dispensing system on operational efficiency and service standards in an outpatient pharmacy. The simulation results suggest that automating the prescription-filing function using a prototype that picks and packs at 20 seconds per item will not assist the pharmacy in achieving the waiting time target of 30 minutes for all patients. Regardless of the state of automation, to meet the waiting time target, 2 additional pharmacists are needed to overcome the process bottleneck at the point of medication dispense. However, if the automated dispensing is the preferred option, the speed of the system needs to be twice as fast as the current configuration to facilitate the reduction of the 95th percentile patient waiting time to below 30 minutes. The faster processing speed will concomitantly allow the pharmacy to reduce the number of pharmacy technicians from 11 to 8. Simulation was found to be a useful and low cost method that allows an otherwise expensive and resource intensive evaluation of new work processes and technology to be completed within a short time.

  18. Appointment Wait Time, Primary Care Provider Status, and Patient Demographics are Associated With Nonattendance at Outpatient Gastroenterology Clinic.

    PubMed

    Shrestha, Manish P; Hu, Chengcheng; Taleban, Sasha

    2016-09-22

    We intended to identify the factors associated with missed appointments at a gastroenterology (GI) clinic in an academic setting. Missed clinic appointments reduce clinic efficiency, waste resources, and increase costs. Limited data exist on subspecialty clinic attendance. We performed a case-control study using data from the electronic health record of patients scheduled for an appointment at the adult GI clinic at the Banner University Medical Center between March and October of 2014. Patients who missed their appointment during the study period served as cases. Controls were randomly selected from patients who completed their appointment during the study period. Analysis included univariate and multivariate logistic regression analysis. Of 2331 scheduled clinic appointments, 195 (8.4%) were missed appointments. Longer waiting time from referral to scheduled appointment was significantly associated with missed appointment (AOR=1.014; 95% CI, 1.01-1.02; P<0.001). Patients with primary care providers (PCPs) were less likely to miss their appointment than those without PCPs (AOR=0.35; 95% CI, 0.18-0.66; P=0.001). Among patient demographic characteristics, ethnicity and marital status were associated with missed appointment. Wait time, ethnicity, marital status, and PCP status were associated with missed GI clinic appointments. Further investigations are needed to assess the effects of intervention strategies directed at reducing appointment wait time and increasing PCP-based care.

  19. Concordance between partners in desired waiting time to birth for newlyweds in India

    PubMed Central

    Singh, Abhishek; Becker, Stan

    2014-01-01

    Examining waiting time to birth among newlywed couples is likely to provide insights into the desire for spacing births among newlywed husbands and wives. Data from the Indian National Family Health Survey of 2005-06 is used to examine the desired waiting time (DWT) to birth among newlywed couples. The dependent variable is spousal concordance on desired times. Overall 65 % of couples have concordant DWTs. Among discordant couples, wives were more likely to want to wait longer than their husbands. Couples from richer wealth quintiles were more likely than couples from poorest quintile to have a concordant DWTs. Muslims were less likely than Hindus to have concordant desires. There is a need for spacing methods among newlyweds. This may have implications for the Indian Family Planning Programme which to date has largely focused on sterilization. Programmes need to include newlywed husbands to promote use of spacing methods. PMID:21933466

  20. Public health care and private insurance demand: the waiting time as a link.

    PubMed

    Jofre-Bonet, M

    2000-01-01

    This paper analyzes the effect of waiting times in the Spanish public health system on the demand for private health insurance. Expected utility maximization determines whether or not individuals buy a private health insurance. The decision depends not only on consumer's covariates such as income, socio-demographic characteristics and health status, but also on the quality of the treatment by the public provider. We interpret waiting time as a qualitative attribute of the health care provision. The empirical analysis uses the Spanish Health Survey of 1993. We cope with the absence of income data by using the Spanish Family Budget Survey of 1990-91 as a complementary data set, following the Arellano-Meghir method [4]. Results indicate that a reduction in the waiting time lowers the probability of buying private health insurance. This suggests the existence of a crowd-out in the health care provision market.

  1. Act-and-wait time-delayed feedback control of autonomous systems

    NASA Astrophysics Data System (ADS)

    Pyragas, Viktoras; Pyragas, Kestutis

    2018-02-01

    Recently an act-and-wait modification of time-delayed feedback control has been proposed for the stabilization of unstable periodic orbits in nonautonomous dynamical systems (Pyragas and Pyragas, 2016 [30]). The modification implies a periodic switching of the feedback gain and makes the closed-loop system finite-dimensional. Here we extend this modification to autonomous systems. In order to keep constant the phase difference between the controlled orbit and the act-and-wait switching function an additional small-amplitude periodic perturbation is introduced. The algorithm can stabilize periodic orbits with an odd number of real unstable Floquet exponents using a simple single-input single-output constraint control.

  2. Access to specialist gastroenterology care in Canada: Comparison of wait times and consensus targets

    PubMed Central

    Leddin, Desmond; Armstrong, David; Barkun, Alan NG; Chen, Ying; Daniels, Sandra; Hollingworth, Roger; Hunt, Richard H; Paterson, William G

    2008-01-01

    BACKGROUND: Monitoring wait times and defining targets for care have been advocated to improve health care delivery related to cancer, heart, diagnostic imaging, joint replacements and sight restoration. There are few data on access to care for digestive diseases, although they pose a greater economic burden than cancer or heart disease in Canada. The present study compared wait times for specialist gastroenterology care with recent, evidence-based, consensus-defined benchmark wait times for a range of digestive diseases. METHODS: Total wait times from primary care referral to investigation were measured for seven digestive disease indications by using the Practice Audit in Gastroenterology program, and were benchmarked against consensus recommendations. RESULTS: Total wait times for 1903 patients who were undergoing investigation exceeded targets for those with probable cancer (median 26 days [25th to 75th percentiles eight to 56 days] versus target of two weeks); probable inflammatory bowel disease (101 days [35 to 209 days] versus two weeks); documented iron deficiency anemia (71 days [19 to 142 days] versus two months); positive fecal occult blood test (73 days [36 to 148 days] versus two months); dyspepsia with alarm symptoms (60 days [23 to 140 days] versus two months); refractory dyspepsia without alarm symptoms (126 days [42 to 225 days] versus two months); and chronic constipation and diarrhea (141 days [68 to 264 days] versus two months). A minority of patients were seen within target times: probable cancer (33% [95% CI 20% to 47%]); probable inflammatory bowel disease (12% [95% CI 1% to 23%]); iron deficiency anemia (46% [95% CI 37% to 55%]); positive occult blood test (41% [95% CI 28% to 54%]); dyspepsia with alarm symptoms (51% [95% CI 41% to 60%]); refractory dyspepsia without alarm symptoms (33% [95% CI 19% to 47%]); and chronic constipation and diarrhea (21% [95% CI 14% to 29%]). DISCUSSION: Total wait times for the seven indications exceeded the

  3. Minimizing patient waiting time in emergency department of public hospital using simulation optimization approach

    NASA Astrophysics Data System (ADS)

    Ibrahim, Ireen Munira; Liong, Choong-Yeun; Bakar, Sakhinah Abu; Ahmad, Norazura; Najmuddin, Ahmad Farid

    2017-04-01

    Emergency department (ED) is the main unit of a hospital that provides emergency treatment. Operating 24 hours a day with limited number of resources invites more problems to the current chaotic situation in some hospitals in Malaysia. Delays in getting treatments that caused patients to wait for a long period of time are among the frequent complaints against government hospitals. Therefore, the ED management needs a model that can be used to examine and understand resource capacity which can assist the hospital managers to reduce patients waiting time. Simulation model was developed based on 24 hours data collection. The model developed using Arena simulation replicates the actual ED's operations of a public hospital in Selangor, Malaysia. The OptQuest optimization in Arena is used to find the possible combinations of a number of resources that can minimize patients waiting time while increasing the number of patients served. The simulation model was modified for improvement based on results from OptQuest. The improvement model significantly improves ED's efficiency with an average of 32% reduction in average patients waiting times and 25% increase in the total number of patients served.

  4. WAITING TIMES OF QUASI-HOMOLOGOUS CORONAL MASS EJECTIONS FROM SUPER ACTIVE REGIONS

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

    Wang Yuming; Liu Lijuan; Shen Chenglong

    Why and how do some active regions (ARs) frequently produce coronal mass ejections (CMEs)? These are key questions for deepening our understanding of the mechanisms and processes of energy accumulation and sudden release in ARs and for improving our space weather prediction capability. Although some case studies have been performed, these questions are still far from fully answered. These issues are now being addressed statistically through an investigation of the waiting times of quasi-homologous CMEs from super ARs in solar cycle 23. It is found that the waiting times of quasi-homologous CMEs have a two-component distribution with a separation atmore » about 18 hr. The first component is a Gaussian-like distribution with a peak at about 7 hr, which indicates a tight physical connection between these quasi-homologous CMEs. The likelihood of two or more occurrences of CMEs faster than 1200 km s{sup -1} from the same AR within 18 hr is about 20%. Furthermore, the correlation analysis among CME waiting times, CME speeds, and CME occurrence rates reveals that these quantities are independent of each other, suggesting that the perturbation by preceding CMEs rather than free energy input is the direct cause of quasi-homologous CMEs. The peak waiting time of 7 hr probably characterizes the timescale of the growth of the instabilities triggered by preceding CMEs. This study uncovers some clues from a statistical perspective for us to understand quasi-homologous CMEs as well as CME-rich ARs.« less

  5. What Are We Waiting For Customer Wait Time, Fill Rate, And Marine Corps Equipment Operational Availability

    DTIC Science & Technology

    2016-12-01

    managed by an RIP. SECREPs are typically critical repair assemblies that require consistently high fill- rates to satisfy maintenance customers ...fill-rate is potentially misreporting performance and areas where short customer wait times could potentially suffice for inventory management . A...supply. Inventory forecasting and management should focus on parts with CWTs that do not satisfy the maintenance customer and 100% fill-rates should

  6. Effect of self-triage on waiting times at a walk-in sexual health clinic.

    PubMed

    Hitchings, Samantha; Barter, Janet

    2009-10-01

    Lengthy waiting times can be a major problem in walk-in sexual health clinics. They are stressful for both patients and staff and may lead to clients with significant health issues leaving the department before being seen by a clinician. A self-triage system may help reduce waiting times and duplication of work, improve patient pathways and decrease wasted visits. This paper describes implementation of a self-triage system in two busy sexual and reproductive health clinics. Patients were asked to complete a self-assessment form on registration to determine the reason for attendance. This then enabled patients to be directed to the most appropriate specialist or clinical service. The benefits of this approach were determined by measuring patient waiting times, reduction in unnecessary specialist review together with patient acceptability as tested by a patient satisfaction survey. The ease of comprehension of the triage form was also assessed by an independent readers' panel. A total of 193 patients were recruited over a 4-month period from November 2004 to February 2005. Patients from the November and December clinics were assigned to the 'traditional treatment' arm, with patients at subsequent clinics being assigned to the 'self-triage' system. Waiting times were collected by the receptionist and clinic staff. Ninety six patients followed the traditional route, 97 the new self-triage system. Sixty-nine (35.8%) patients completed the satisfaction survey. The self-triage system significantly reduced waiting time from 40 (22, 60) to 23 (10, 40) minutes [results expressed as median (interquartile range)]. There was a non-significant reduction in the proportion of patients seeing two clinicians from 21% to 13% (p = 0.17). Satisfaction levels were not significantly altered (95% compared to 97% satisfied, p = 0.64). The readers' panel found the triage form both easy to understand and to complete. Self-triage can effectively reduce clinic waiting times and allow better

  7. The influence of insurance status on waiting times in German acute care hospitals: an empirical analysis of new data

    PubMed Central

    2009-01-01

    Background There is an ongoing debate in Germany about the assumption that patients with private health insurance (PHI) benefit from better access to medical care, including shorter waiting times (Lüngen et al. 2008), compared to patients with statutory health insurance (SHI). Problem Existing analyses of the determinants for waiting times in Germany are a) based on patient self-reports and b) do not cover the inpatient sector. This paper aims to fill both gaps by (i) generating new primary data and (ii) analyzing waiting times in German hospitals. Methods We requested individual appointments from 485 hospitals within an experimental study design, allowing us to analyze the impact of PHI versus SHI on waiting times (Asplin et al. 2005). Results In German acute care hospitals patients with PHI have significantly shorter waiting times than patients with SHI. Conclusion Discrimination in waiting times by insurance status does occur in the German acute hospital sector. Since there is very little transparency in treatment quality in Germany, we do not know whether discrimination in waiting times leads to discrimination in the quality of treatment. This is an important issue for future research. PMID:20025744

  8. Modeling of waiting times and price changes in currency exchange data

    NASA Astrophysics Data System (ADS)

    Repetowicz, Przemysław; Richmond, Peter

    2004-11-01

    A theory which describes the share price evolution at financial markets as a continuous-time random walk (Physica A 287 (2000) 468, Physica A 314 (2002) 749, Eur. Phys. J. B 27 (2002) 273, Physica A 376 (2000) 284) has been generalized in order to take into account the dependence of waiting times t on price returns x. A joint probability density function (pdf) φ(x,t) which uses the concept of a Lévy stable distribution is worked out. The theory is fitted to high-frequency US $/Japanese Yen exchange rate and low-frequency 19th century Irish stock data. The theory has been fitted both to price return and to waiting time data and the adherence to data, in terms of the χ2 test statistic, has been improved when compared to the old theory.

  9. Specific timely appointments for triage reduced waiting lists in an outpatient physiotherapy service.

    PubMed

    Harding, K E; Bottrell, J

    2016-12-01

    Waiting lists with triage systems are commonly used in outpatient physiotherapy but may not be effective. Could an alternative model of access and triage reduce waiting times over a sustained period with no additional resources? Observational study comparing retrospective data for 11 months prior to the introduction of a new model of access compared with data for the equivalent 11 months afterwards. Patients referred to a physiotherapy outpatient department at an outer metropolitan hospital before (n=721) and after (n=707) the introduction of the new model. A model of access and triage known as 'specific timely appointments for triage' (STAT), in which appointment slots are preserved in advance specifically for new patients based on calculation of average demand. Time from referral to first assessment, number of appointments per patient, occasions of non-attendance and total length of stay in the service. Median time from referral to first appointment was 18 days [interquartile range (IQR) 11 to 33 days] in the pre-intervention group, compared with 14 days (IQR 9 to 21 days) in the post-intervention group (P<0.01). The number of physiotherapy appointments also reduced (IQR 2 to 6 vs IQR 1 to 4; P<0.01). There were no changes in non-attendance rates or total time in the service. Waiting time for outpatient physiotherapy was 22% lower in the year following the introduction of the STAT model. While acknowledging the limitations of a pre- and post-measurement design, this model may have potential for reducing waiting times for outpatient physiotherapy without additional resources. Copyright © 2015 Chartered Society of Physiotherapy. Published by Elsevier Ltd. All rights reserved.

  10. The effects of publishing emergency department wait time on patient utilization patterns in a community with two emergency department sites: a retrospective, quasi-experiment design.

    PubMed

    Xie, Bin; Youash, Sabrina

    2011-06-14

    Providing emergency department (ED) wait time information to the public has been suggested as a mechanism to reduce lengthy ED wait times (by enabling patients to select the ED site with shorter wait time), but the effects of such a program have not been evaluated. We evaluated the effects of such a program in a community with two ED sites. Descriptive statistics for wait times of the two sites before and after the publication of wait time information were used to evaluate the effects of the publication of wait time information on wait times. Multivariate logistical regression was used to test whether or not individual patients used published wait time to decide which site to visit. We found that the rates of wait times exceeding 4 h, and the 95th percentile of wait times in the two sites decreased after the publication of wait time information, even though the average wait times experienced a slight increase. We also found that after controlling for other factors, the site with shorter wait time had a higher likelihood of being selected after the publication of wait time information, but there was no such relationship before the publication. These findings were consistent with the hypothesis that the publication of wait time information leads to patients selecting the site with shorter wait time. While publishing ED wait time information did not improve average wait time, it reduced the rates of lengthy wait times.

  11. Impact of wait times on the effectiveness of transcatheter aortic valve replacement in severe aortic valve disease: a discrete event simulation model.

    PubMed

    Wijeysundera, Harindra C; Wong, William W L; Bennell, Maria C; Fremes, Stephen E; Radhakrishnan, Sam; Peterson, Mark; Ko, Dennis T

    2014-10-01

    There is increasing demand for transcatheter aortic valve replacement (TAVR) as the primary treatment option for patients with severe aortic stenosis who are high-risk surgical candidates or inoperable. We used mathematical simulation models to estimate the hypothetical effectiveness of TAVR with increasing wait times. We applied discrete event modelling, using data from the Placement of Aortic Transcatheter Valves (PARTNER) trials. We compared TAVR with medical therapy in the inoperable cohort, and compared TAVR to conventional aortic valve surgery in the high-risk cohort. One-year mortality and wait-time deaths were calculated in different scenarios by varying TAVR wait times from 10 days to 180 days, while maintaining a constant wait time for surgery at a mean of 15.6 days. In the inoperable cohort, the 1-year mortality for medical therapy was 50%. When the TAVR wait time was 10 days, the TAVR wait-time mortality was 1.9% with a 1-year mortality of 31.5%. TAVR wait-time deaths increased to 28.9% with a 180-day wait, with a 1-year mortality of 41.4%. In the high-risk cohort, the wait-time deaths and 1-year mortality for the surgical patients were 2.5% and 27%, respectively. The TAVR wait-time deaths increased from 2.2% with a 10-day wait to 22.4% with a 180-day wait, and a corresponding increase in 1-year mortality from 24.5% to 32.6%. Mortality with TAVR exceeded surgery when TAVR wait times exceeded 60 days. Modest increases in TAVR wait times have a substantial effect on the effectiveness of TAVR in inoperable patients and high-risk surgical candidates. Copyright © 2014 Canadian Cardiovascular Society. Published by Elsevier Inc. All rights reserved.

  12. [Patients' satisfaction and waiting time in oncology day care centers in Champagne-Ardenne].

    PubMed

    Debreuve-Theresette, A; Jovenin, N; Stona, A C; Kraïem-Leleu, M; Burde, F; Parent, D; Hettler, D; Rey, J B

    2015-12-01

    Quality of life of patients suffering from cancer may be influenced by the way healthcare is organized and by patient experiences. Nowadays, chemotherapy is often provided in day care centers. This study aimed to assess patient waiting time and satisfaction in oncology day care centers in Champagne-Ardenne, France. This cross-sectional survey involved all patients receiving ambulatory chemotherapy during a one-week period in day care centers of Champagne-Ardenne public and private healthcare institutions participating in the study. Sociodemographic, medical and outpatient data were collected. Patient satisfaction was measured using the Out-Patsat35 questionnaire. Eleven (out of 16) oncology day care centers and 441 patients participated in the study. Most of the patients were women (n=252, 57.1%) and the mean age was 61±12 years. The mean satisfaction score was 82±14 (out of 100) and the mean waiting time between the assigned appointment time and administration of chemotherapy was 97±60 min. This study has shown that waiting times are important. However, patients are satisfied with the healthcare organization, especially regarding nursing support. Early preparation of chemotherapy could improve these parameters. Copyright © 2015 Elsevier Masson SAS. All rights reserved.

  13. Reading and Television Viewing Habits of American Adults during Time Spent in Waiting Rooms.

    ERIC Educational Resources Information Center

    Spirn, Sharon L.

    In order to determine the reading and television viewing habits of American adults during time spent in waiting rooms, a study observed 100 adults waiting outside the Emergency Treatment Room of John F. Kennedy Hospital in Edison, New Jersey, over a four-week period. Results revealed that more of these adults chose to watch television as an…

  14. In the queue for total joint replacement: patients' perspectives on waiting times. Ontario Hip and Knee Replacement Project Team.

    PubMed

    Llewellyn-Thomas, H A; Arshinoff, R; Bell, M; Williams, J I; Naylor, C D

    1998-02-01

    We assessed patients on the waiting lists of a purposive sample of orthopaedic surgeons in Ontario, Canada, to determine patients' attitudes towards time waiting for hip or knee replacement. We focused on 148 patients who did not have a definite operative date, obtaining complete information on 124 (84%). Symptom severity was assessed with the Western Ontario/McMaster Osteoarthritis Index and a disease-specific standard gamble was used to elicit patients' overall utility for their arthritic state. Next, in a trade-off task, patients considered a hypothetical choice between a 1-month wait for a surgeon who could provide a 2% risk of post-operative mortality, or a 6-month wait for joint replacement with a 1% risk of post-operative mortality. Waiting times were then shifted systematically until the patient abandoned his/her initial choice, generating a conditional maximal acceptable wait time. Patients were divided in their attitudes, with 57% initially choosing a 6-month wait with a 1% mortality risk. The overall distribution of conditional maximum acceptable wait time scores ranged from 1 to 26 months, with a median of 7 months. Utility values were independently but weakly associated with patients' tolerance of waiting times (adjusted R-square = 0.059, P = 0.004). After splitting the sample along the median into subgroups with a relatively 'low' and 'high' tolerance for waiting, the subgroup with the apparently lower tolerance for waiting reported lower utility scores (z = 2.951; P = 0.004) and shorter times since their surgeon first advised them of the need for surgery (z = 3.014; P = 0.003). These results suggest that, in the establishment and monitoring of a queue management system for quality-of-life-enhancing surgery, patients' own perceptions of their overall symptomatic burden and ability to tolerate delayed relief should be considered along with information derived from clinical judgements and pre-weighted health status instruments.

  15. Poisson-process generalization for the trading waiting-time distribution in a double-auction mechanism

    NASA Astrophysics Data System (ADS)

    Cincotti, Silvano; Ponta, Linda; Raberto, Marco; Scalas, Enrico

    2005-05-01

    In this paper, empirical analyses and computational experiments are presented on high-frequency data for a double-auction (book) market. Main objective of the paper is to generalize the order waiting time process in order to properly model such empirical evidences. The empirical study is performed on the best bid and best ask data of 7 U.S. financial markets, for 30-stock time series. In particular, statistical properties of trading waiting times have been analyzed and quality of fits is evaluated by suitable statistical tests, i.e., comparing empirical distributions with theoretical models. Starting from the statistical studies on real data, attention has been focused on the reproducibility of such results in an artificial market. The computational experiments have been performed within the Genoa Artificial Stock Market. In the market model, heterogeneous agents trade one risky asset in exchange for cash. Agents have zero intelligence and issue random limit or market orders depending on their budget constraints. The price is cleared by means of a limit order book. The order generation is modelled with a renewal process. Based on empirical trading estimation, the distribution of waiting times between two consecutive orders is modelled by a mixture of exponential processes. Results show that the empirical waiting-time distribution can be considered as a generalization of a Poisson process. Moreover, the renewal process can approximate real data and implementation on the artificial stocks market can reproduce the trading activity in a realistic way.

  16. 'Waiting for' and 'waiting in' public and private hospitals: a qualitative study of patient trust in South Australia.

    PubMed

    Ward, Paul R; Rokkas, Philippa; Cenko, Clinton; Pulvirenti, Mariastella; Dean, Nicola; Carney, A Simon; Meyer, Samantha

    2017-05-05

    Waiting times for hospital appointments, treatment and/or surgery have become a major political and health service problem, leading to national maximum waiting times and policies to reduce waiting times. Quantitative studies have documented waiting times for various types of surgery and longer waiting times in public vs private hospitals. However, very little qualitative research has explored patient experiences of waiting, how this compares between public and private hospitals, and the implications for trust in hospitals and healthcare professionals. The aim of this paper is to provide a deep understanding of the impact of waiting times on patient trust in public and private hospitals. A qualitative study in South Australia, including 36 in-depth interviews (18 from public and 18 from private hospitals). Data collection occurred in 2012-13, and data were analysed using pre-coding, followed by conceptual and theoretical categorisation. Participants differentiated between experiences of 'waiting for' (e.g. for specialist appointments and surgery) and 'waiting in' (e.g. in emergency departments and outpatient clinics) public and private hospitals. Whilst 'waiting for' public hospitals was longer than private hospitals, this was often justified and accepted by public patients (e.g. due to reduced government funding), therefore it did not lead to distrust of public hospitals. Private patients had shorter 'waiting for' hospital services, increasing their trust in private hospitals and distrust of public hospitals. Public patients also recounted many experiences of longer 'waiting in' public hospitals, leading to frustration and anxiety, although they rarely blamed or distrusted the doctors or nurses, instead blaming an underfunded system and over-worked staff. Doctors and nurses were seen to be doing their best, and therefore trustworthy. Although public patients experienced longer 'waiting for' and 'waiting in' public hospitals, it did not lead to widespread distrust

  17. Impact of Lean on patient cycle and waiting times at a rural district hospital in KwaZulu-Natal

    PubMed Central

    Naidoo, Logandran

    2016-01-01

    Background Prolonged waiting time is a source of patient dissatisfaction with health care and is negatively associated with patient satisfaction. Prolonged waiting times in many district hospitals result in many dissatisfied patients, overworked and frustrated staff, and poor quality of care because of the perceived increased workload. Aim The aim of the study was to determine the impact of Lean principles techniques, and tools on the operational efficiency in the outpatient department (OPD) of a rural district hospital. Setting The study was conducted at the Catherine Booth Hospital (CBH) – a rural district hospital in KwaZulu-Natal, South Africa. Methods This was an action research study with pre-, intermediate-, and post-implementation assessments. Cycle and waiting times were measured by direct observation on two occasions before, approximately two-weekly during, and on two occasions after Lean implementation. A standardised data collection tool was completed by the researcher at each of the six key service nodes in the OPD to capture the waiting times and cycle times. Results All six service nodes showed a reduction in cycle times and waiting times between the baseline assessment and post-Lean implementation measurement. Significant reduction was achieved in cycle times (27%; p < 0.05) and waiting times (from 11.93 to 10 min; p = 0.03) at the Investigations node. Although the target reduction was not achieved for the Consulting Room node, there was a significant reduction in waiting times from 80.95 to 74.43 min, (p < 0.001). The average efficiency increased from 16.35% (baseline) to 20.13% (post-intervention). Conclusion The application of Lean principles, tools and techniques provides hospital managers with an evidence-based management approach to resolving problems and improving quality indicators. PMID:27543283

  18. Impact of Lean on patient cycle and waiting times at a rural district hospital in KwaZulu-Natal.

    PubMed

    Naidoo, Logandran; Mahomed, Ozayr H

    2016-07-26

    Prolonged waiting time is a source of patient dissatisfaction with health care and is negatively associated with patient satisfaction. Prolonged waiting times in many district hospitals result in many dissatisfied patients, overworked and frustrated staff, and poor quality of care because of the perceived increased workload. The aim of the study was to determine the impact of Lean principles techniques, and tools on the operational efficiency in the outpatient department (OPD) of a rural district hospital. The study was conducted at the Catherine Booth Hospital (CBH) - a rural district hospital in KwaZulu-Natal, South Africa. This was an action research study with pre-, intermediate-, and post-implementation assessments. Cycle and waiting times were measured by direct observation on two occasions before, approximately two-weekly during, and on two occasions after Lean implementation. A standardised data collection tool was completed by the researcher at each of the six key service nodes in the OPD to capture the waiting times and cycle times. All six service nodes showed a reduction in cycle times and waiting times between the baseline assessment and post-Lean implementation measurement. Significant reduction was achieved in cycle times (27%; p < 0.05) and waiting times (from 11.93 to 10 min; p = 0.03) at the Investigations node. Although the target reduction was not achieved for the Consulting Room node, there was a significant reduction in waiting times from 80.95 to 74.43 min, (p < 0.001). The average efficiency increased from 16.35% (baseline) to 20.13% (post-intervention). The application of Lean principles, tools and techniques provides hospital managers with an evidence-based management approach to resolving problems and improving quality indicators.

  19. Influence of Waiting Time on the Levitation Force Between a Permanent Magnet and a Superconductor

    NASA Astrophysics Data System (ADS)

    Zhang, Xing-Yi; Zhou, You-He; Zhou, Jun

    This paper describes the experimental results of the levitation force of single-grained YBaCuO bulk superconductors preparing by the top-seeded melt-growth method with different waiting time tw below an NdFeB permanent magnet. It was found that waiting time has large effects on the zero-field-cooled (ZFC) and field-cooled (FC) levitation force, and the levitation force shows aging characteristics at the liquid nitrogen temperature.

  20. The unethical focus on access: a study of medical ethics and the waiting-time guarantee.

    PubMed

    Karlberg, H I; Brinkmo, B-M

    2009-03-01

    All civilized societies favour ethical principles of equity. In healthcare, these principles generally focus on needs for medical care. Methods for establishing priorities among such needs are instrumental in this process. In this study, we analysed whether rules on access to healthcare, waiting-time guarantees, conflict with ethical principles of distributive justice. We interviewed directors, managers and other decision-makers of various healthcare providers of hospitals, primary care organizations and purchasing offices. We also conducted focus group interviews with professionals from a number of distinct medical areas. Our informants and their co-workers were reasonably familiar with the ethical platforms for priority-setting established by the Swedish parliament, giving the sickest patients complete priority. However, to satisfy the waiting-time guarantees, the informants often had to make priority decisions contrary to the ethical principles by favouring access before needs to keep waiting times within certain limits. The common opinion was that the waiting-time guarantee leads to crowding-out effects, overruling the ethical principles based on needs. For more than a decade, the interpretation in Sweden of the equitable principle based on medical needs has been distorted through political decisions, leading to healthcare providers giving priority to access rather than needs for care.

  1. Gender and socioeconomic status as determinants of waiting time for inpatient surgery in a system with implicit queue management.

    PubMed

    Arnesen, Kjell E; Erikssen, Jan; Stavem, Knut

    2002-12-01

    In a system with implicit queue management, to examine gender and socioeconomic status as determinants of waiting time for inpatient surgery, after adjusting for other potential predictors. A cohort of 452 subjects was examined in outpatient clinics of a general hospital and referred to inpatient surgery. They were followed until scheduled hospital admission (n=396) or until the requested procedure no longer was relevant (n=56). We compared waiting time between groups from referral date until hospital admission, using Kaplan-Meier estimates of waiting times and log rank test. A Cox proportional hazards model was used for assessing the risk ratio (RR) of hospital admission for scheduled surgery. Gender and socioeconomic status could not explain variations in waiting time. However, patients with suspected/verified neoplastic disease or a risk of serious deterioration without treatment had markedly shorter waiting times than the reference groups, with adjusted RR (95% confidence intervals (95%CI)) of time to receiving in-patient surgery of 2.3 (1.7-3.0) and 2.0 (1.3-3.0), respectively. Being on sick leave was associated with shorter waiting time, adjusted RR of 1.7 (1.2-2.5). Referrals from within the hospital or other hospitals had also shorter waiting times than referrals from primary health care physicians, adjusted RR=1.4 (1.1-1.8). There was no evidence of bias against women or people in lower socioeconomic classes in this implicit queue management system. However, patients' access to inpatient surgery was associated with malignancy, prognosis, sick leave status, physician experience, referral pattern and the major diagnosis category.

  2. Discrimination in waiting times by insurance type and financial soundness of German acute care hospitals.

    PubMed

    Schwierz, Christoph; Wübker, Achim; Wübker, Ansgar; Kuchinke, Björn A

    2011-10-01

    This paper shows that patients with private health insurance (PHI) are being offered significantly shorter waiting times than patients with statutory health insurance (SHI) in German acute hospital care. This behavior may be driven by the higher expected profitability of PHI relative to SHI holders. Further, we find that hospitals offering private insurees shorter waiting times when compared with SHI holders have a significantly better financial performance than those abstaining from or with less discrimination.

  3. Conceptualising time before surgery: The experience of patients waiting for hip replacement

    PubMed Central

    Johnson, Emma C.; Horwood, Jeremy; Gooberman-Hill, Rachael

    2014-01-01

    Interpretations of time underlie patients' experiences of illness and the way in which the National Health Service (NHS) is organised. In the NHS, achieving short waiting times for treatment is seen as important, and this is particularly evident in relation to chronic conditions where the time waiting in care from onset of symptoms to successful management can last months and years. One example of a chronic condition with high prevalence is osteoarthritis, estimated to affect 10% of people aged over 55 years in the UK. Osteoarthritis of the hip is particularly common, and treatments include exercise and medication. If these options do not provide enough relief from pain and functional difficulties, then joint replacement may be considered. With over 70,000 such operations conducted every year in England and Wales, processes relating to waiting times impact on many patients. This article explores how 24 patients with osteoarthritis experience time during the lead up to hip replacement surgery. We draw on data collected during longitudinal in-depth interviews with patients a median of 9.5 days before surgery and at two to four weeks post-operatively. Transcripts of audio-recorded interviews were imported into Atlas.ti® and inductive thematic analysis undertaken. Increasing pain and deterioration in function altered the experience of time during the journey towards hip replacement. Patients made essential changes to how they filled their days. They experienced lost and wasted time and faced disruption to the temporal order of their lives. A surgical date marked in the calendar became their focus. However, this date was not static, moving because of changing perceptions of duration and real-time alterations by the healthcare system. Findings highlight that patients' experience of time is complex and multi-dimensional and does not reflect the linear, monochronic conceptualisation of time embedded in the healthcare system. PMID:24997442

  4. WAITING TIME DISTRIBUTION OF SOLAR ENERGETIC PARTICLE EVENTS MODELED WITH A NON-STATIONARY POISSON PROCESS

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

    Li, C.; Su, W.; Fang, C.

    2014-09-10

    We present a study of the waiting time distributions (WTDs) of solar energetic particle (SEP) events observed with the spacecraft WIND and GOES. The WTDs of both solar electron events (SEEs) and solar proton events (SPEs) display a power-law tail of ∼Δt {sup –γ}. The SEEs display a broken power-law WTD. The power-law index is γ{sub 1} = 0.99 for the short waiting times (<70 hr) and γ{sub 2} = 1.92 for large waiting times (>100 hr). The break of the WTD of SEEs is probably due to the modulation of the corotating interaction regions. The power-law index, γ ∼more » 1.82, is derived for the WTD of the SPEs which is consistent with the WTD of type II radio bursts, indicating a close relationship between the shock wave and the production of energetic protons. The WTDs of SEP events can be modeled with a non-stationary Poisson process, which was proposed to understand the waiting time statistics of solar flares. We generalize the method and find that, if the SEP event rate λ = 1/Δt varies as the time distribution of event rate f(λ) = Aλ{sup –α}exp (– βλ), the time-dependent Poisson distribution can produce a power-law tail WTD of ∼Δt {sup α} {sup –3}, where 0 ≤ α < 2.« less

  5. An Estimation Method of Waiting Time for Health Service at Hospital by Using a Portable RFID and Robust Estimation

    NASA Astrophysics Data System (ADS)

    Ishigaki, Tsukasa; Yamamoto, Yoshinobu; Nakamura, Yoshiyuki; Akamatsu, Motoyuki

    Patients that have an health service by doctor have to wait long time at many hospitals. The long waiting time is the worst factor of patient's dissatisfaction for hospital service according to questionnaire for patients. The present paper describes an estimation method of the waiting time for each patient without an electronic medical chart system. The method applies a portable RFID system to data acquisition and robust estimation of probability distribution of the health service and test time by doctor for high-accurate waiting time estimation. We carried out an health service of data acquisition at a real hospital and verified the efficiency of the proposed method. The proposed system widely can be used as data acquisition system in various fields such as marketing service, entertainment or human behavior measurement.

  6. Factors Associated with Waiting Time for Access to Mental Health Services for Children and Adolescents in Norway

    ERIC Educational Resources Information Center

    Andersson, Helle Wessel

    2004-01-01

    The present study addresses the question of equality of access, as it relates to waiting time for specialised mental health treatment for children and adolescents. The aim was to investigate whether demographic, clinical factors and service-related factors were associated with waiting time. Data was based on a documentation system in which all…

  7. Waiting times before dental care under general anesthesia in children with special needs in the Children's Hospital of Casablanca.

    PubMed

    Badre, Bouchra; Serhier, Zineb; El Arabi, Samira

    2014-01-01

    Oral diseases may have an impact on quality of children's life. The presence of severe disability requires the use of care under general anesthesia (GA). However, because of the limited number of qualified health personnel, waiting time before intervention can be long. To evaluate the waiting time before dental care under general anesthesia for children with special needs in Morocco. A retrospective cohort study was carried out in pediatric dentistry unit of the University Hospital of Casablanca. Data were collected from records of patients seen for the first time between 2006 and 2011. The waiting time was defined as the time between the date of the first consultation and intervention date. 127 children received dental care under general anesthesia, 57.5% were male and the average age was 9.2 (SD = 3.4). Decay was the most frequent reason for consultation (48%), followed by pain (32%). The average waiting time was 7.6 months (SD = 4.2 months). The average number of acts performed per patient was 13.5. Waiting times were long, it is necessary to take measures to reduce delays and improve access to oral health care for this special population.

  8. A modelling framework for mitigating customers' waiting time at a vehicle inspection centre

    NASA Astrophysics Data System (ADS)

    Ahmad, Norazura; Abidin, Norhaslinda Zainal; Ilyas, Khibtiyah; Abduljabbar, Waleed Khalid

    2017-11-01

    In Malaysia, an agency that is entrusted by the Government to perform mandatory vehicle inspection for public, commercial and private vehicles, receive many customers daily. Often complaints of problems received from the customers are associated with waiting time that leads to lost of business and dissatisfied customers. To address this issue, we propose a framework for modelling a vehicle inspection system using an integration of simulation and optimization approaches. The strengths of simulation and optimization are reviewed briefly that is hoped to reveal the synergy between the established methods in determining an appropriate customer's waiting time for inspection at a vehicle inspection centre. Relevant concepts and preliminary results are also presented and discussed in this paper.

  9. Ramsey waits: allocating public health service resources when there is rationing by waiting.

    PubMed

    Gravelle, Hugh; Siciliani, Luigi

    2008-09-01

    The optimal allocation of a public health care budget across treatments must take account of the way in which care is rationed within treatments since this will affect their marginal value. We investigate the optimal allocation rules for public health care systems where user charges are fixed and care is rationed by waiting. The optimal waiting time is higher for treatments with demands more elastic to waiting time, higher costs, lower charges, smaller marginal welfare loss from waiting by treated patients, and smaller marginal welfare losses from under-consumption of care. The results hold for a wide range of welfarist and non-welfarist objective functions and for systems in which there is also a private health care sector. They imply that allocation rules based purely on cost effectiveness ratios are suboptimal because they assume that there is no rationing within treatments.

  10. Reducing Wait Time for Lung Cancer Diagnosis and Treatment: Impact of a Multidisciplinary, Centralized Referral Program.

    PubMed

    Common, Jessica L; Mariathas, Hensley H; Parsons, Kaylah; Greenland, Jonathan D; Harris, Scott; Bhatia, Rick; Byrne, SuzanneC

    2018-06-04

    A multidisciplinary, centralized referral program was established at our institution in 2014 to reduce delays in lung cancer diagnosis and treatment following diagnostic imaging observed with the traditional, primary care provider-led referral process. The main objectives of this retrospective cohort study were to determine if referral to a Thoracic Triage Panel (TTP): 1) expedites lung cancer diagnosis and treatment initiation; and 2) leads to more appropriate specialist consultation. Patients with a diagnosis of lung cancer and initial diagnostic imaging between March 1, 2015, and February 29, 2016, at a Memorial University-affiliated tertiary care centre in St John's, Newfoundland, were identified and grouped according to whether they were referred to the TTP or managed through a traditional referral process. Wait times (in days) from first abnormal imaging to biopsy and treatment initiation were recorded. Statistical analysis was performed using the Wilcoxon rank-sum test. A total of 133 patients who met inclusion criteria were identified. Seventy-nine patients were referred to the TTP and 54 were managed by traditional means. There was a statistically significant reduction in median wait times for patients referred to the TTP. Wait time from first abnormal imaging to biopsy decreased from 61.5 to 36.0 days (P < .0001). Wait time from first abnormal imaging to treatment initiation decreased from 118.0 to 80.0 days (P < .001). The percentage of specialist consultations that led to treatment was also greater for patients referred to the TTP. A collaborative, centralized intake and referral program helps to reduce wait time for diagnosis and treatment of lung cancer. Copyright © 2018 Canadian Association of Radiologists. Published by Elsevier Inc. All rights reserved.

  11. The effect of a positive reappraisal coping intervention and problem-solving skills training on coping strategies during waiting period of IUI treatment: An RCT.

    PubMed

    Ghasemi, Marzieh; Kordi, Masoumeh; Asgharipour, Negar; Esmaeili, Habibollah; Amirian, Maliheh

    2017-11-01

    Waiting period of fertility treatment is stressful, therefore it is necessary to use effective coping strategies to cope with waiting period of intrauterine insemination (IUI) treatment. The aim of this study was comparing the effect of the positive reappraisal coping intervention (PRCI) with the problem-solving skills training (PSS) on the coping strategies of IUI waiting period, in infertile women referred to Milad Infertility Center in Mashhad. In this randomized clinical trial, 108 women were evaluated into three groups. The control group received the routine care, but in PRCI group, two training sessions were held and they were asked to review the coping thoughts cards and fill out the daily monitoring forms during the waiting period, and in PSS group problem-solving skill were taught during 3 sessions. The coping strategies were compared between three groups on the 10 th day of IUI waiting period. Results showed that the mean score for problem-focused were significantly different between the control (28.54±9.70), PSS (33.71±9.31), and PRCI (30.74±10.96) (p=0.025) groups. There were significant differences between the PSS group and others groups, and mean emotion-focused were significantly different between the control (32.09±11.65), PSS (29.20±9.88), and PRCI (28.74±7.96) (p=0.036) groups. There were significant differences between the PRCI and the control group (p=0.047). PSS was more effective to increase problem-focused coping strategies than PRCI, therefore it is recommended that this intervention should be used in infertility treatment centers.

  12. The effect of a positive reappraisal coping intervention and problem-solving skills training on coping strategies during waiting period of IUI treatment: An RCT

    PubMed Central

    Ghasemi, Marzieh; Kordi, Masoumeh; Asgharipour, Negar; Esmaeili, Habibollah; Amirian, Maliheh

    2017-01-01

    Background: Waiting period of fertility treatment is stressful, therefore it is necessary to use effective coping strategies to cope with waiting period of intrauterine insemination (IUI) treatment. Objective: The aim of this study was comparing the effect of the positive reappraisal coping intervention (PRCI) with the problem-solving skills training (PSS) on the coping strategies of IUI waiting period, in infertile women referred to Milad Infertility Center in Mashhad. Materials and Methods: In this randomized clinical trial, 108 women were evaluated into three groups. The control group received the routine care, but in PRCI group, two training sessions were held and they were asked to review the coping thoughts cards and fill out the daily monitoring forms during the waiting period, and in PSS group problem-solving skill were taught during 3 sessions. The coping strategies were compared between three groups on the 10th day of IUI waiting period. Results: Results showed that the mean score for problem-focused were significantly different between the control (28.54±9.70), PSS (33.71±9.31), and PRCI (30.74±10.96) (p=0.025) groups. There were significant differences between the PSS group and others groups, and mean emotion-focused were significantly different between the control (32.09±11.65), PSS (29.20±9.88), and PRCI (28.74±7.96) (p=0.036) groups. There were significant differences between the PRCI and the control group (p=0.047). Conclusion: PSS was more effective to increase problem-focused coping strategies than PRCI, therefore it is recommended that this intervention should be used in infertility treatment centers. PMID:29404530

  13. Performance Contracting and Quality Improvement in Outpatient Treatment: Effects on Waiting Time and Length of Stay

    PubMed Central

    Stewart, Maureen T.; Horgan, Constance M.; Garnick, Deborah W.; Ritter, Grant; McLellan, A. Thomas

    2012-01-01

    We evaluate effects of a performance contract (PC) implemented in Delaware in 2001 and participation in quality improvement (QI) programs on waiting time for treatment and length of stay (LOS) using client treatment episode level data from Delaware (n = 12,368) and Maryland (n = 147,151) for 1998 – 2006. Results of difference-in-difference analyses indicate waiting time declined 13 days following the PC, after controlling for client characteristics and historical trends. Participation in the PC and a formal QI program was associated with a decrease of 20 days. LOS increased 22 days under the PC and 24 days under the PC and QI programs, after controlling for client characteristics. The PC and QI program were associated with improvements in LOS and waiting time, although we cannot determine which aspects of the programs (incentives, training, monitoring) resulted in these changes. PMID:22445031

  14. Waiting times before dental care under general anesthesia in children with special needs in the Children's Hospital of Casablanca

    PubMed Central

    Badre, Bouchra; Serhier, Zineb; El Arabi, Samira

    2014-01-01

    Introduction Oral diseases may have an impact on quality of children's life. The presence of severe disability requires the use of care under general anesthesia (GA). However, because of the limited number of qualified health personnel, waiting time before intervention can be long. Aim: To evaluate the waiting time before dental care under general anesthesia for children with special needs in Morocco. Methods A retrospective cohort study was carried out in pediatric dentistry unit of the University Hospital of Casablanca. Data were collected from records of patients seen for the first time between 2006 and 2011. The waiting time was defined as the time between the date of the first consultation and intervention date. Results 127 children received dental care under general anesthesia, 57.5% were male and the average age was 9.2 (SD = 3.4). Decay was the most frequent reason for consultation (48%), followed by pain (32%). The average waiting time was 7.6 months (SD = 4.2 months). The average number of acts performed per patient was 13.5. Conclusion Waiting times were long, it is necessary to take measures to reduce delays and improve access to oral health care for this special population. PMID:25328594

  15. Conceptualising time before surgery: the experience of patients waiting for hip replacement.

    PubMed

    Johnson, Emma C; Horwood, Jeremy; Gooberman-Hill, Rachael

    2014-09-01

    Interpretations of time underlie patients' experiences of illness and the way in which the National Health Service (NHS) is organised. In the NHS, achieving short waiting times for treatment is seen as important, and this is particularly evident in relation to chronic conditions where the time waiting in care from onset of symptoms to successful management can last months and years. One example of a chronic condition with high prevalence is osteoarthritis, estimated to affect 10% of people aged over 55 years in the UK. Osteoarthritis of the hip is particularly common, and treatments include exercise and medication. If these options do not provide enough relief from pain and functional difficulties, then joint replacement may be considered. With over 70,000 such operations conducted every year in England and Wales, processes relating to waiting times impact on many patients. This article explores how 24 patients with osteoarthritis experience time during the lead up to hip replacement surgery. We draw on data collected during longitudinal in-depth interviews with patients a median of 9.5 days before surgery and at two to four weeks post-operatively. Transcripts of audio-recorded interviews were imported into Atlas.ti(®) and inductive thematic analysis undertaken. Increasing pain and deterioration in function altered the experience of time during the journey towards hip replacement. Patients made essential changes to how they filled their days. They experienced lost and wasted time and faced disruption to the temporal order of their lives. A surgical date marked in the calendar became their focus. However, this date was not static, moving because of changing perceptions of duration and real-time alterations by the healthcare system. Findings highlight that patients' experience of time is complex and multi-dimensional and does not reflect the linear, monochronic conceptualisation of time embedded in the healthcare system. Copyright © 2014 The Authors. Published by

  16. A study on the impact of prioritising emergency department arrivals on the patient waiting time.

    PubMed

    Van Bockstal, Ellen; Maenhout, Broos

    2018-05-03

    In the past decade, the crowding of the emergency department has gained considerable attention of researchers as the number of medical service providers is typically insufficient to fulfil the demand for emergency care. In this paper, we solve the stochastic emergency department workforce planning problem and consider the planning of nurses and physicians simultaneously for a real-life case study in Belgium. We study the patient arrival pattern of the emergency department in depth and consider different patient acuity classes by disaggregating the arrival pattern. We determine the personnel staffing requirements and the design of the shifts based on the patient arrival rates per acuity class such that the resource staffing cost and the weighted patient waiting time are minimised. In order to solve this multi-objective optimisation problem, we construct a Pareto set of optimal solutions via the -constraints method. For a particular staffing composition, the proposed model minimises the patient waiting time subject to upper bounds on the staffing size using the Sample Average Approximation Method. In our computational experiments, we discern the impact of prioritising the emergency department arrivals. Triaging results in lower patient waiting times for higher priority acuity classes and to a higher waiting time for the lowest priority class, which does not require immediate care. Moreover, we perform a sensitivity analysis to verify the impact of the arrival and service pattern characteristics, the prioritisation weights between different acuity classes and the incorporated shift flexibility in the model.

  17. Fruit fate, seed germination and growth of an invasive vine - an experimental test of 'sit and wait' strategy

    Treesearch

    Cathryn H. Greenberg; Lindsay M. Smith; Douglas J. Levey

    2001-01-01

    Oriental bittersweet (Celastrus orbiculutus Thunb.) is a non-indigenous, invasive woody vine in North America that proliferates in disturbed open sites. Unlike most invasive species, C. orbiculatus exhibits a `sit and wait' strategy by establishing and persisting indefinitely in undisturbed, closed canopy forest and...

  18. Graph transformation method for calculating waiting times in Markov chains.

    PubMed

    Trygubenko, Semen A; Wales, David J

    2006-06-21

    We describe an exact approach for calculating transition probabilities and waiting times in finite-state discrete-time Markov processes. All the states and the rules for transitions between them must be known in advance. We can then calculate averages over a given ensemble of paths for both additive and multiplicative properties in a nonstochastic and noniterative fashion. In particular, we can calculate the mean first-passage time between arbitrary groups of stationary points for discrete path sampling databases, and hence extract phenomenological rate constants. We present a number of examples to demonstrate the efficiency and robustness of this approach.

  19. Experience of being a low priority patient during waiting time at an emergency department

    PubMed Central

    Dahlen, Ingrid; Westin, Lars; Adolfsson, Annsofie

    2012-01-01

    Background Work in the emergency department is characterized by fast and efficient medical efforts to save lives, but can also involve a long waiting time for patients. Patients are given a priority rating upon their arrival in the clinic based on the seriousness of their problem, and nursing care for lower priority patients is given a lower prioritization. Regardless of their medical prioritization, all patients have a right to expect good nursing care while they are waiting. The purpose of this study was to illustrate the experience of the low prioritized patient during their waiting time in the emergency department. Methods A phenomenological hermeneutic research method was used to analyze an interview transcript. Data collection consisted of narrative interviews. The interviewees were 14 patients who had waited more than three hours for surgical, orthopedic, or other medical care. Results The findings resulted in four different themes, ie, being dependent on care, being exposed, being vulnerable, and being secure. Lower priority patients are not paid as much attention by nursing staff. Patients reported feeling powerless, insulted, and humiliated when their care was delayed without their understanding what was happening to them. Not understanding results in exposure that violates self-esteem. Conclusion The goal of the health care provider must be to minimize and prevent suffering, prevent feelings of vulnerability, and to create conditions for optimal patient well being. PMID:22334799

  20. Emergency Department Waiting Times (EDWaT): A Patient Flow Management and Quality of Care Rating mHealth Application.

    PubMed

    Househ, Mowafa; Yunus, Faisel

    2014-01-01

    Saudi hospital emergency departments (ED) have suffered from long waiting times, which have led to a delay in emergency patient care. The increase in the population of Saudi Arabia is likely to further stretch the healthcare services due to overcrowding leading to decreased healthcare quality, long patient waits, patient dissatisfaction, ambulance diversions, decreased physician productivity, and increased frustration among medical staff. This will ultimately put patients at risk for poor health outcomes. Time is of the essence in emergencies and to get to an ED that has the shortest waiting time can mean life or death for a patient, especially in cases of stroke and myocardial infarction. In this paper, we present our work on the development of a mHealth Application - EDWaT - that will: provide patient flow information to the emergency medical services staff, help in quick routing of patients to the nearest hospital, and provide an opportunity for patients to review and rate the quality of care received at an ED, which will then be forwarded to ED services administrators. The quality ratings will help patients to choose between two EDs with the same waiting time and distance from their location. We anticipate that the use of EDWaT will help improve ED wait times and the quality of care provision in Saudi hospitals EDs.

  1. The effect of waiting times from general practitioner referral to MRI or orthopaedic consultation for the knee on patient-based outcomes.

    PubMed

    Brealey, S; Andronis, L; Dale, V; Gibbon, A J; Gilbert, F J; Hendry, M; Hood, K; King, D; Wilkinson, C

    2012-11-01

    The purpose of this study was to test for the effect of waiting time from general practitioner (GP) referral to MRI or to orthopaedic consultation on outcomes of patients with knee problems, and to test whether any characteristics of trial participants predicted waiting time to MRI or orthopaedics. We undertook secondary analyses of data on 553 participants from a randomised trial who were recruited from 163 general practices during November 2002 to October 2004. Of the patients allocated to MRI, 263 (94%) had an MRI, and of those referred to orthopaedics, 236 (86%) had an orthopaedic consultation. The median (interquartile range) waiting time in days from randomisation to MRI was 41.0 (21.0-71.0) and to orthopaedic appointment was 78.5 (54.5-167.5). Waiting time was found to have no significant effect on patient outcome for both the Short Form 36-item (SF-36) physical functioning score (p=0.570) and the Knee Quality of Life 26-item (KQoL-26) physical functioning score (p=0.268). There was weak evidence that males waited less time for their MRI (p=0.049) and older patients waited longer for their orthopaedic referral (p=0.049). For patients who resided in the catchment areas of some centres there were significantly longer waiting times for both MRI and orthopaedic appointment. Where patients reside is a strong predictor of waiting time for access to services such as MRI or orthopaedics. There is no evidence to suggest, however, that this has a significant effect on physical well-being in the short term for patients with knee problems.

  2. Waiting-time distributions of magnetic discontinuities: clustering or Poisson process?

    PubMed

    Greco, A; Matthaeus, W H; Servidio, S; Dmitruk, P

    2009-10-01

    Using solar wind data from the Advanced Composition Explorer spacecraft, with the support of Hall magnetohydrodynamic simulations, the waiting-time distributions of magnetic discontinuities have been analyzed. A possible phenomenon of clusterization of these discontinuities is studied in detail. We perform a local Poisson's analysis in order to establish if these intermittent events are randomly distributed or not. Possible implications about the nature of solar wind discontinuities are discussed.

  3. Waiting-time distributions of magnetic discontinuities: Clustering or Poisson process?

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

    Greco, A.; Matthaeus, W. H.; Servidio, S.

    2009-10-15

    Using solar wind data from the Advanced Composition Explorer spacecraft, with the support of Hall magnetohydrodynamic simulations, the waiting-time distributions of magnetic discontinuities have been analyzed. A possible phenomenon of clusterization of these discontinuities is studied in detail. We perform a local Poisson's analysis in order to establish if these intermittent events are randomly distributed or not. Possible implications about the nature of solar wind discontinuities are discussed.

  4. Reduction of admit wait times: the effect of a leadership-based program.

    PubMed

    Patel, Pankaj B; Combs, Mary A; Vinson, David R

    2014-03-01

    Prolonged admit wait times in the emergency department (ED) for patients who require hospitalization lead to increased boarding time in the ED, a significant cause of ED congestion. This is associated with decreased quality of care, higher morbidity and mortality, decreased patient satisfaction, increased costs for care, ambulance diversion, higher numbers of patients who leave without being seen (LWBS), and delayed care with longer lengths of stay (LOS) for other ED patients. The objective was to assess the effect of a leadership-based program to expedite hospital admissions from the ED. This before-and-after observational study was undertaken from 2006 through 2011 at one community hospital ED. A team of ED and hospital leaders implemented a program to reduce admit wait times, using a computerized hospital-wide tracking system to monitor inpatient and ED bed status. The team collaboratively and consistently moved ED patients to their inpatient beds within an established goal of 60 minutes after an admission decision was reached. Top leadership actively intervened in real time by contacting staff whenever delays occurred to expedite immediate solutions to achieve the 60-minute goal. The primary outcome measures were the percentage of ED patients who were admitted to inpatient beds within 60 minutes from the time the beds were requested and ED boarding time. LOS, patient satisfaction, LWBS rate, and ambulance diversion hours were also measured. After ED census, hospital admission rates, and ED bed capacity were controlled for using a multivariable linear regression analysis, the admit wait time reduction program contributed to an increase in patients being admitted to the hospital within 60 minutes by 16 percentage points (95% confidence intervals [CI] = 10 to 22 points; p < 0.0001) and a decrease in boarding time per admission of 46 minutes (95% CI = 63 to 82 minutes; p < 0.0001). LOS decreased for admitted patients by 79 minutes (95% CI = 55 to 104 minutes; p < 0

  5. A Study to Determine Patient Waiting Time at the Outpatient Pharmacy at Wilford Hall USAF Medical Center

    DTIC Science & Technology

    1988-06-01

    at Wilford Hall USAF Medical Center significantly reduced the patient wait time at the main outpatient pharmacy. Satellite pharmacies have been ).’l...PRESENTING TO WINDOW 1, 19 MAR 88. 47 C:. A’.’E-:A: -ESCRIRTIONS PER PATIENT ...........48 H. WILFORD HALL MEDICAL CENTER OUTPATIENT QUESTIONNAIRE...that wait times at tne outpatient pharmacy were excessive. It was this concern that motivated the Medical Center Administrator to request that patient

  6. Does Wait-List Size at Registration Influence Time to Surgery? Analysis of a Population-Based Cardiac Surgery Registry

    PubMed Central

    Sobolev, Boris; Levy, Adrian; Hayden, Robert; Kuramoto, Lisa

    2006-01-01

    Objective To determine whether the probability of undergoing coronary bypass surgery within a certain time was related to the number of patients on the wait list at registration for the operation in a publicly funded health system. Methods A prospective cohort study comparing waiting times among patients registered on wait lists at the hospitals delivering adult cardiac surgery. For each calendar week, the list size, the number of new registrations, and the number of direct admissions immediately after angiography characterized the demand for surgery. Results The length of delay in undergoing treatment was associated with list size at registration, with shorter times for shorter lists (log-rank test 1,198.3, p<.0001). When the list size at registration required clearance time over 1 week patients had 42 percent lower odds of undergoing surgery compared with lists with clearance time less than 1 week (odds ratio [OR] 0.58 percent, 95 percent, confidence interval [CI] 0.53–0.63), after adjustment for age, sex, comorbidity, period, and hospital. The weekly number of new registrations exceeding weekly service capacity had an independent effect toward longer service delays when the list size at registration required clearance time less than 1 week (OR 0.56 percent, 95 percent CI 0.45–0.71), but not for longer lists. Every time the operation was performed for a patient requiring surgery without registration on wait lists, the odds of surgery for listed patients were reduced by 6 percent (OR 0.94, CI 0.93–0.95). Conclusion For wait-listed patients, time to surgery depends on the list size at registration, the number of new registrations, as well as on the weekly number of patients who move immediately from angiography to coronary bypass surgery without being registered on a wait list. Hospital managers may use these findings to improve resource planning and to reduce uncertainty when providing advice on expected treatment delays. PMID:16430599

  7. Strategic attention deployment for delay of gratification in working and waiting situations.

    PubMed

    Peake, Philip K; Mischel, Walter; Hebl, Michelle

    2002-03-01

    Two studies examined whether the detrimental effects of attention to rewards on delay of gratification in waiting situations holds-or reverses-in working situations. In Study 1, preschoolers waited or worked for desired delayed rewards. Delay times increased when children worked in the presence of rewards but, as predicted, this increase was due to the distraction provided by the work itself. not because attention to rewards motivated children to sustain work. Analysis of spontaneous attention deployment showed that attending to rewards reduces delay time regardless of the working or waiting nature of the task. Fixing attention on rewards was a particularly detrimental strategy regardless of the type of task. Study 2 showed that when the work is not engaging, however, attention to rewards can motivate instrumental work and facilitate delay of gratification as long as attention deployment does not become fixed on the rewards.

  8. The impact of travel distance, travel time and waiting time on health-related quality of life of diabetes patients: An investigation in six European countries.

    PubMed

    Konerding, Uwe; Bowen, Tom; Elkhuizen, Sylvia G; Faubel, Raquel; Forte, Paul; Karampli, Eleftheria; Mahdavi, Mahdi; Malmström, Tomi; Pavi, Elpida; Torkki, Paulus

    2017-04-01

    The effects of travel distance and travel time to the primary diabetes care provider and waiting time in the practice on health-related quality of life (HRQoL) of patients with type 2 diabetes are investigated. Survey data of 1313 persons with type 2 diabetes from six regions in England (274), Finland (163), Germany (254), Greece (165), the Netherlands (354), and Spain (103) were analyzed. Various multiple linear regression analyses with four different EQ-5D-3L indices (English, German, Dutch and Spanish index) as target variables, with travel distance, travel time, and waiting time in the practice as focal predictors and with control for study region, patient's gender, patient's age, patient's education, time since diagnosis, thoroughness of provider-patient communication were computed. Interactions of regions with the remaining five control variables and the three focal predictors were also tested. There are no interactions of regions with control variables or focal predictors. The indices decrease with increasing travel time to the provider and increasing waiting time in the provider's practice. HRQoL of patients with type 2 diabetes might be improved by decreasing travel time to the provider and waiting time in the provider's practice. Copyright © 2017 Elsevier B.V. All rights reserved.

  9. Efficiency of colorectal cancer care among veterans: analysis of treatment wait times at Veterans Affairs Medical Centers.

    PubMed

    Merkow, Ryan P; Bilimoria, Karl Y; Sherman, Karen L; McCarter, Martin D; Gordon, Howard S; Bentrem, David J

    2013-07-01

    Timeliness of cancer treatment is an important aspect of health care quality. Veterans Affairs Medical Centers (VAMCs) are expected to treat a growing number of patients with cancer. Our objectives were to examine treatment times from diagnosis to first-course therapy for patients with colon and rectal cancers and assess factors associated with prolonged wait times. From the VA Central Cancer Registry, patients who underwent colon or rectal resection for cancer from 1998 to 2008 were identified. Time from diagnosis to definitive cancer-directed therapy was measured, and multivariable regression methods were used to determine predictors of prolonged wait times for colon (≥ 45 days) and rectal (≥ 60 days) cancers. From 124 VAMCs, 14,097 patients underwent colectomy, and 3,390 underwent rectal resection for cancer. For colon cancer, the median time to treatment increased by 68% over time (P < .001). From 2007 to 2008, the median time to colectomy was 32 days. Predictors of prolonged wait times included age ≥ 55 years (v < 55 years), time period (2007 to 2008 v 1998 to 2000), black race (v white), marriage status (married v unmarried), high-volume center status (v low volume), and treatment at a different hospital (v same hospital as initial diagnosis; all P < .05). For rectal cancer, the overall median time to first-course treatment increased by 74% (P < .001). From 2007 to 2008, the median time to proctectomy was 47 days. Similar predictors of prolonged wait times were identified for rectal cancer. Time to first treatment has increased for patients with colon and rectal cancers at VAMCs. Patient, tumor, and hospital factors are associated with prolonged time to treatment.

  10. Wait times for physical and occupational therapy in the public system for people with arthritis in quebec.

    PubMed

    Delaurier, Ashley; Bernatsky, Sasha; Raymond, Marie-Hélène; Feldman, Debbie Ehrmann

    2013-01-01

    Although arthritis is the leading cause of pain and disability in Canada, and physical therapy (PT) and occupational therapy (OT) are beneficial both for chronic osteoarthritis (OA) and for inflammatory arthritis such as rheumatoid arthritis (RA), there appear to be problems with access to such services. The aim of this study was to document wait times from referral by physician to consultation with PT or OT in the public health care system for people with arthritis in Quebec, Canada. Appointments were requested by telephone, using hypothetical case scenarios; wait times were defined as the time between initial request and appointment date. Descriptive statistics were used to examine the wait times in relation to diagnosis, service provider and geographic area. For both scenarios (OA and RA) combined, 13% were offered an appointment within 6 months, 13% offered given an appointment within 6-12 months, 24% were told they would need to wait longer than 12 months, and 22% were refused services. The remaining 28% were told they would require an evaluation appointment for functional assessment before being given an appointment for therapy. No difference was found between RA and OA diagnoses. Our study suggests that most people with arthritis living in the province of Quebec are not receiving publicly accessible PT or OT intervention in a timely manner.

  11. Impact of co-located general practitioner (GP) clinics and patient choice on duration of wait in the emergency department.

    PubMed

    Sharma, Anurag; Inder, Brett

    2011-08-01

    To empirically model the determinants of duration of wait of emergency (triage category 2) patients in an emergency department (ED) focusing on two questions: (i) What is the effect of enhancing the degree of choice for non-urgent (triage category 5) patients on duration of wait for emergency (category 2) patients in EDs; and (ii) What is the effect of co-located GP clinics on duration of wait for emergency patients in EDs? The answers to these questions will help in understanding the effectiveness of demand management strategies, which are identified as one of the solutions to ED crowding. The duration of wait for each patient (difference between arrival time and time first seen by treating doctor) was modelled as a function of input factors (degree of choice, patient characteristics, weekend admission, metro/regional hospital, concentration of emergency (category 2) patients in hospital service area), throughput factors (availability of doctors and nurses) and output factor (hospital bed capacity). The unit of analysis was a patient episode and the model was estimated using a survival regression technique. The degree of choice for non-urgent (category 5) patients has a non-linear effect: more choice for non-urgent patients is associated with longer waits for emergency patients at lower values and shorter waits at higher values of degree of choice. Thus more choice of EDs for non-urgent patients is related to a longer wait for emergency (category 2) patients in EDs. The waiting time for emergency patients in hospital campuses with co-located GP clinics was 19% lower (1.5 min less) on average than for those waiting in campuses without co-located GP clinics. These findings suggest that diverting non-urgent (category 5) patients to an alternative model of care (co-located GP clinics) is a more effective demand management strategy and will reduce ED crowding.

  12. Consumer behaviour in the waiting area.

    PubMed

    Mobach, Mark P

    2007-02-01

    To determine consumer behaviour in the pharmacy waiting area. The applied methods for data-collection were direct observations. Three Dutch community pharmacies were selected for the study. The topics in the observation list were based on available services at each waiting area (brochures, books, illuminated new trailer, children's play area, etc.). Per patient each activity was registered, and at each pharmacy the behaviour was studied for 2 weeks. Most patients only waited during the waiting time at the studied pharmacies. Few consumers obtained written information during their wait. The waiting area may have latent possibilities to expand the information function of the pharmacy and combine this with other activities that distract the consumer from the wait. Transdisciplinary research, combining knowledge from pharmacy practice research with consumer research, has been a useful approach to add information on queueing behaviour of consumers.

  13. Self-similarity of waiting times in fracture systems

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

    Niccolini, G.; Bosia, F.; Carpinteri, A.

    2009-08-15

    Experimental and numerical results are presented for a fracture experiment carried out on a fiber-reinforced element under flexural loading, and a statistical analysis is performed for acoustic emission waiting-time distributions. By an optimization procedure, a recently proposed scaling law describing these distributions for different event magnitude scales is confirmed by both experimental and numerical data, thus reinforcing the idea that fracture of heterogeneous materials has scaling properties similar to those found for earthquakes. Analysis of the different scaling parameters obtained for experimental and numerical data leads us to formulate the hypothesis that the type of scaling function obtained depends onmore » the level of correlation among fracture events in the system.« less

  14. Deficient neural activity subserving decision-making during reward waiting time in intertemporal choice in adult attention-deficit hyperactivity disorder.

    PubMed

    Todokoro, Ayako; Tanaka, Saori C; Kawakubo, Yuki; Yahata, Noriaki; Ishii-Takahashi, Ayaka; Nishimura, Yukika; Kano, Yukiko; Ohtake, Fumio; Kasai, Kiyoto

    2018-04-24

    Impulsivity, which significantly affects social adaptation, is an important target behavioral characteristic in interventions for attention-deficit hyperactivity disorder (ADHD). Typically, people are willing to wait longer to acquire greater rewards. Impulsivity in ADHD may be associated with brain dysfunction in decision-making involving waiting behavior under such situations. We tested the hypothesis that brain circuitry during a period of waiting (i.e., prior to the acquisition of reward) is altered in adults with ADHD. The participants included 14 medication-free adults with ADHD and 16 healthy controls matched for age, sex, IQ, and handedness. The behavioral task had participants choose between a delayed, larger monetary reward and an immediate, smaller monetary reward, where the reward waiting time actually occurred during functional magnetic resonance imaging measurement. We tested for group differences in the contrast values of blood-oxygen-level dependent signals associated with the length of waiting time, calculated using the parametric modulation method. While the two groups did not differ in the time discounting rate, the delay-sensitive contrast values were significantly lower in the caudate and visual cortex in individuals with ADHD. The higher impulsivity scores were significantly associated with lower delay-sensitive contrast values in the caudate and visual cortex. These results suggest that deficient neural activity affects decision-making involving reward waiting time during intertemporal choice tasks, and provide an explanation for the basis of impulsivity in adult ADHD. © 2018 The Author. Psychiatry and Clinical Neurosciences © 2018 Japanese Society of Psychiatry and Neurology.

  15. Waiting time for coronal preparation and the influence of different cements on tensile strength of metal posts.

    PubMed

    Oliveira, Ilione Kruschewsky Costa Sousa; Arsati, Ynara Bosco de Oliveira Lima; Basting, Roberta Tarkany; França, Fabiana Mantovani Gomes

    2012-01-01

    This study aimed to assess the effect of post-cementation waiting time for core preparation of cemented cast posts and cores had on retention in the root canal, using two different luting materials. Sixty extracted human canines were sectioned 16 mm from the root apex. After cast nickel-chromium metal posts and cores were fabricated and luted with zinc phosphate (ZP) cement or resin cement (RC), the specimens were divided into 3 groups (n = 10) according to the waiting time for core preparation: no preparation (control), 15 minutes, or 1 week after the core cementation. At the appropriate time, the specimens were subjected to a tensile load test (0.5 mm/min) until failure. Two-way ANOVA (time versus cement) and the Tukey tests (P < 0.05) showed significantly higher (P < 0.05) tensile strength values for the ZP cement groups than for the RC groups. Core preparation and post-cementation waiting time for core recontouring did not influence the retention strength. ZP was the best material for intraradicular metal post cementation.

  16. Managing patients' wait time in specialist out-patient clinic using real-time data from existing queue management and ADT systems.

    PubMed

    Ju, John Chen; Gan, Soon Ann; Tan Siew Wee, Justine; Huang Yuchi, Peter; Mei Mei, Chan; Wong Mei Mei, Sharon; Fong, Kam Weng

    2013-01-01

    In major cancer centers, heavy patients load and multiple registration stations could cause significant wait time, and can be result in patient complains. Real-time patient journey data and visual display are useful tools in hospital patient queue management. This paper demonstrates how we capture patient queue data without deploying any tracing devices; and how to convert data into useful patient journey information to understand where interventions are likely to be most effective. During our system development, remarkable effort has been spent on resolving data discrepancy and balancing between accuracy and system performances. A web-based dashboard to display real-time information and a framework for data analysis were also developed to facilitate our clinics' operation. Result shows our system could eliminate more than 95% of data capturing errors and has improved patient wait time data accuracy since it was deployed.

  17. Which factors influence patients' maximum acceptable waiting time for cataract surgery? - a questionnaire survey.

    PubMed

    Weingessel, Birgit; Richter-Mueksch, Sibylla; Vécsei-Marlovits, Pia V

    2011-05-01

    To evaluate patients’ maximum acceptable waiting time (MAWT) and to assess the determinants of patient perceptions of MAWT. A total of 500 consecutive patients with cataract were asked to fill out a preoperative questionnaire, addressing patients’ MAWT to undergo cataract surgery. Patients’ visual impairment (VF-14 score), education, profession and social status were evaluated, and an ophthalmologic examination was performed. Univariate analysis included Spearman’s correlation test, unpaired Student’s t-test and the Mann–Whitney U test. Univariate and multivariate associations were calculated using unconditional logistic regression. The mean MAWT was 3.17 ± 2.12 months. The mean VF-14 score was 72.10 ± 22.54. Between VF-14 score and MAWT, there was a significant correlation (r = 0.180, p = 0.004). Patients with higher education (high school, university) accepted significantly longer MAWT (3.92 ± 2.38 months versus 3.02 ± 2.00 months, p = 0.009). Patients who had self-noticed visual impairment were nearly four times (OR: 3.88, 95% CI = 2.07–7.28, p < 0.001) more likely to accept only MAWT of <3 months. Patients with low tolerance for waiting had greater self-reported difficulty with vision. Patients’ acceptance of waiting was not associated with clinical visual acuity measures. Education, ability to work, living independently and taking care of dependents were also strong predictors from patients’ perspective. Considering the implementation of standards for waiting lists, these facts should be taken into account. © 2010 The Authors. Journal compilation © 2010 Acta Ophthalmol.

  18. Average waiting time in FDDI networks with local priorities

    NASA Technical Reports Server (NTRS)

    Gercek, Gokhan

    1994-01-01

    A method is introduced to compute the average queuing delay experienced by different priority group messages in an FDDI node. It is assumed that no FDDI MAC layer priorities are used. Instead, a priority structure is introduced to the messages at a higher protocol layer (e.g. network layer) locally. Such a method was planned to be used in Space Station Freedom FDDI network. Conservation of the average waiting time is used as the key concept in computing average queuing delays. It is shown that local priority assignments are feasable specially when the traffic distribution is asymmetric in the FDDI network.

  19. Socioeconomic status and waiting times for health services: An international literature review and evidence from the Italian National Health System.

    PubMed

    Landi, Stefano; Ivaldi, Enrico; Testi, Angela

    2018-04-01

    In the absence of priority criteria, waiting times are an implicit rationing instrument where the absence or limited use of prices creates an excess of demand. Even in the presence of priority criteria, waiting times may be unfair because they reduce health care demand of patients in lower socio-economic conditions due to high opportunity costs of time or a decay in their health level. Significant evidence has shown a relationship between socioeconomic status and the length of waiting time. The first phase of the study involved an extensive review of the existent literature for the period of 2002-2016 in the main databases (Scopus, PubMed and Science Direct). Twenty-eight met the eligibility criteria. The 27 papers were described and classified. The e mpirical objective of this study was to determine whether socioeconomic characteristics affect waiting time for different health services in the Italian national health system. The services studied were specialist visits, diagnostics tests and elective surgeries. A classification tree and logistic regression models were implemented. Data from the 2013 Italian Health National Survey were used. The analysis found heterogeneous results for different types of service. Individuals with lower education and economic resources have a higher risk of experiencing excessive waiting times for diagnostic and specialist visits. For elective surgery, socioeconomic inequalities are present but appear to be lower. Copyright © 2018 Elsevier B.V. All rights reserved.

  20. Maternity waiting homes in Ethiopia--three decades experience.

    PubMed

    Gaym, Asheber; Pearson, Luwei; Soe, Khynn Win Win

    2012-07-01

    Access to comprehensive emergency obstetric care is limited in Ethiopia. Maternity waiting homes are part of the strategies utilized to improve access to hard to reach rural populations. Despite long years of existence of this service in Ethiopia, the practice has not been adequately assessed so far. Describe the current status of maternity waiting home services in Ethiopia All facilities in Ethiopia that have a maternity waiting home were identified from FMOH data as well as personal contacts with focal persons at Regional Health Bureaus in the nine regions and UNICEF regional offices. A standardized data collection tool for facility assessment was developed by the quality referral team, Health Section, UNICEF. Data collection included site visits and documentation of infrastructural related issues through a facility checklist. Service related issues were also collected from log books and other documents as well as through interview with relevant staff Focus group discussions were held with all MWHs attendants who were found admitted at the time of the review at Attat, Wolisso and Gidole hospital maternity waiting homes on major thematic areas identified by the review team regarding MWH care The practice of maternity waiting homes in Ethiopia spans more than three decades. Nine facilities located in five Regional States had maternity waiting home services. All except one were located in hospitals. Admission capacity ranged from 4 up to 44 mothers at a time. Seven of the maternity waiting homes required the clients to cater for their own food, firewood and clothing supply providing only kitchen space and few kitchen utensils. Clients came from as far as 400 kms away to obtain services. Medical care and documentation of services were not standardized Duration of stay varied from 3-90 days. Monthly admission rates varied from 0-84 mothers at different institutions. Major indications for admission were previous caesarean section 34%; previous fistula repair 12

  1. Waiting room time: An opportunity for parental oral health education.

    PubMed

    Soussou, Randa; Aleksejūnienė, Jolanta; Harrison, Rosamund

    2017-09-14

    The UBC Children's Dental Program (CDP) has provided free dental treatments to underserved low-income children, but its preventive component needs to be enhanced. The study aims were: 1) to develop a "waiting-room based" dental education program engaging caregivers of these children, and 2) to assess the program's feasibility, acceptability and effectiveness. In preparation, a situational analysis (SA) included structured interviews with caregivers, and with various stakeholders (e.g., dental students, instructors, health authority) involved in the CDP program. Based on the SA, caregiver-centered education was designed using an interactive power point presentation; after the presentation, each caregiver set personalized goals for modifying his/her child's dental behaviours. Evaluation of the program was done with follow-up telephone calls; the program's effectiveness was assessed by comparing before/after proportions of caregivers brushing their child's teeth, children brushing teeth in the morning and evening, children eating sugar-containing snacks, and children drinking sugar-containing drinks. The program proved to be easy to implement (feasible) and the recruitment rate was 99% (acceptable). The follow-up rate was 81%. The SA identified that the caregivers' knowledge about caries etiology and prevention was limited. All recruited caregivers completed the educational session and set goals for their family. The evaluation demonstrated an increase in caregiver-reported short-term diet and oral self-care behaviours of their children. A dental education program engaging caregivers in the waiting room was a feasible, acceptable and promising strategy for improving short-term dental behaviours of children.

  2. Waiting for the right time: how and why young Thai women manage to avoid heterosexual intercourse.

    PubMed

    Supametaporn, Pinhatai; Stern, Phyllis Noerager; Rodcumdee, Branom; Chaiyawat, Waraporn

    2010-08-01

    Nineteen young Thai women were purposively selected from networks of nongovernmental organizations involving children and youths in Bangkok. Our grounded theory findings indicated that these young women used the basic social process they called "waiting for the right time" in order to maintain heterosexual abstinence. Waiting for the right time involved one overarching condition, honoring parental love, and included three overlapping properties: learning rules, planning life path, and ways of preserving virginity. The findings provide information that may lead to the development of culturally competent interventions for middle-class Thai youths to remain healthy and avoid pregnancy.

  3. Wait times to rheumatology care for patients with rheumatic diseases: a data linkage study of primary care electronic medical records and administrative data.

    PubMed

    Widdifield, Jessica; Bernatsky, Sasha; Thorne, J Carter; Bombardier, Claire; Jaakkimainen, R Liisa; Wing, Laura; Paterson, J Michael; Ivers, Noah; Butt, Debra; Lyddiatt, Anne; Hofstetter, Catherine; Ahluwalia, Vandana; Tu, Karen

    2016-01-01

    The Wait Time Alliance recently established wait time benchmarks for rheumatology consultations in Canada. Our aim was to quantify wait times to primary and rheumatology care for patients with rheumatic diseases. We identified patients from primary care practices in the Electronic Medical Record Administrative data Linked Database who had referrals to Ontario rheumatologists over the period 2000-2013. To assess the full care pathway, we identified dates of symptom onset, presentation in primary care and referral from electronic medical records. Dates of rheumatologist consultations were obtained by linking with physician service claims. We determined the duration of each phase of the care pathway (symptom onset to primary care encounter, primary care encounter to referral, and referral to rheumatologist consultation) and compared them with established benchmarks. Among 2430 referrals from 168 family physicians, 2015 patients (82.9%) were seen by 146 rheumatologists within 1 year of referral. Of the 2430 referrals, 2417 (99.5%) occurred between 2005 and 2013. The main reasons for referral were osteoarthritis (32.4%) and systemic inflammatory rheumatic diseases (30.6%). Wait times varied by diagnosis and geographic region. Overall, the median wait time from referral to rheumatologist consultation was 74 (interquartile range 27-101) days; it was 66 (interquartile range 18-84) days for systemic inflammatory rheumatic diseases. Wait time benchmarks were not achieved, even for the most urgent types of referral. For systemic inflammatory rheumatic diseases, most of the delays occurred before referral. Rheumatology wait times exceeded established benchmarks. Targeted efforts are needed to promote more timely access to both primary and rheumatology care. Routine linkage of electronic medical records with administrative data may help fill important gaps in knowledge about waits to primary and specialty care.

  4. Radiotherapy wait times for patients with a diagnosis of invasive cancer, 1992-2000.

    PubMed

    Johnston, Grace M; MacGarvie, Vicki L; Elliott, David; Dewar, Ron A D; MacIntyre, Maureen M; Nolan, Maureen C

    2004-06-01

    To study the wait times for cancer patients from the time of diagnosis to consultation with a radiation oncologist (T1), from consultation to radiotherapy (T2) and from diagnosis to radiotherapy (T3) in the context of treatment practices and measurement issues. From 1992 to 2000, we studied 6585 Nova Scotian patients over the age of 24 years with a diagnosis of breast, lung, colorectal or prostate cancer who received radiotherapy within 1 year of diagnosis. Multivariate analyses examined associations between wait time and diagnosis year, age, sex, median household income (MHI), distance to the cancer centre and extent of disease. Univariate findings reported are median times and interquartile ranges. The T3 was 16 weeks for breast and colorectal cancer, 6 weeks for lung cancer and 18 weeks for prostate cancer. The greatest T1 decrease over time was for prostate cancer: 13-8 weeks (hazards ratio [HR] = 1.07, 95% confidence interval [CI] = 1.05-1.10). The T2 increased for all cancers, and the T3 increased from 5 to 7 weeks for lung cancer, from 17 to 22 weeks for prostate cancer and from 10 to 18 weeks for breast cancer, with no change for colorectal cancer. The T3 decreased by age for breast cancer (HR = 1.12, CI = 1.10-1.14) and prostate cancer (HR = 1.07, CI = 1.02-1.11), showed no consistent association with distance to a cancer centre and varied by extent of disease. Patients with localized lung disease had a longer T3 than those with distant disease, but the opposite results were noted for patients with breast cancer. The T3 was greater for regional than distant disease in lung and breast cancers. Sex and MHI had no effect. Wait times reflected clinical practice, and there were no adverse patterns related to age, sex, income or distance from a cancer centre.

  5. Influence of positive distractions on children in two clinic waiting areas.

    PubMed

    Pati, Debajyoti; Nanda, Upali

    2011-01-01

    To examine the influence of positive distraction on the behavior and activity of children in two clinic waiting areas. People spend a considerable proportion of time waiting in hospitals. Studies show that the quality of waiting environments influences the perception of quality of care and caregivers, that perception of waiting time is a better indicator of patient satisfaction than actual waiting time, and that the waiting environment contributes to the perception of wait time. In fact, the attractiveness of the physical environment in waiting areas has been shown to be significantly associated with higher perceived quality of care, less anxiety, and higher reported positive interaction with staff. Can positive distractions in waiting areas improve the waiting experience, as indicated by the behavior and activities of children waiting for treatment? Five distraction conditions were randomly introduced in the waiting area of the dental and cardiac clinics of a major pediatric tertiary care center through a single plasma screen intervention. The attention, behavior, and activities of waiting children were recorded. Data on 158 pediatric patients were collected over 12 days during December 2008 and January 2009. Data analysis shows that the introduction of distraction conditions was associated with more calm behavior and less fine and gross movement, suggesting significant calming effects associated with the distraction conditions. Data also suggest that positive distraction conditions are significant attention grabbers and could be an important contributor to improving the waiting experience for children in hospitals by improving environmental attractiveness.

  6. Anomalous transport in fluid field with random waiting time depending on the preceding jump length

    NASA Astrophysics Data System (ADS)

    Zhang, Hong; Li, Guo-Hua

    2016-11-01

    Anomalous (or non-Fickian) transport behaviors of particles have been widely observed in complex porous media. To capture the energy-dependent characteristics of non-Fickian transport of a particle in flow fields, in the present paper a generalized continuous time random walk model whose waiting time probability distribution depends on the preceding jump length is introduced, and the corresponding master equation in Fourier-Laplace space for the distribution of particles is derived. As examples, two generalized advection-dispersion equations for Gaussian distribution and lévy flight with the probability density function of waiting time being quadratic dependent on the preceding jump length are obtained by applying the derived master equation. Project supported by the Foundation for Young Key Teachers of Chengdu University of Technology, China (Grant No. KYGG201414) and the Opening Foundation of Geomathematics Key Laboratory of Sichuan Province, China (Grant No. scsxdz2013009).

  7. Interest of waiting time for spontaneous early reconnection after cavotricuspid isthmus ablation: A monocentric randomized trial.

    PubMed

    Marchandise, Sébastien; Scavée, Christophe; Barbraud, Cynthia; de Meester de Ravenstein, Christophe; Balola Bagalwa, Mittérand; Goesaert, Cédric; Reis-Pinheiro, Ivone; le Polain de Waroux, Jean-Benoit

    2017-12-01

    The aim of this study was to determine the rate of recurrent atrial flutter (AFl) after isolated cavotricuspid isthmus (CTI) ablation and to evaluate the impact of a waiting period with the search for early resumption of the CTI block on the long-term outcome. Three hundred and nineteen consecutive patients referred for typical AFl ablation were randomly assigned to CTI ablation with continuous reevaluation of the CTI block during 30 minutes and early reablation if needed (waiting time [WT] + group, n  =  155) or to CTI ablation with no waiting period after proven bidirectional CTI block (WT - group, n  =  164). All patients were regularly followed-up. In the WT+ group, 10 patients (6%) presented a recovery across the CTI (time to recovery: 17 ± 7') and were reablated at the end of the waiting period. After a median follow-up of 21 months, the rate of recurrent AFl was significantly higher in the WT - group as compared to the WT+ group (11.6% [19/164] vs 2.5% [4/155], respectively; P  =  0.007). However, no significant differences in the subsequent rate of AF were observed between the two groups (29% [WT -] vs 32% [WT+], P  =  0.66). During the follow-up, 28 patients from the WT - group underwent a second ablation procedure (16 AFl redo and 12 AF ablation) versus 10 patients form the WT+ group (three AFl redo and seven AF ablation). Waiting 30 minutes after CTI ablation to check for early resumption and early reablation allows for decreasing significantly the rate of recurrent atrial flutter. © 2017 Wiley Periodicals, Inc.

  8. Why wait so long for child care? An analysis of waits, queues and work in a South African urban health centre.

    PubMed

    Bachmann, M O; Barron, P

    1997-01-01

    Long waits at large urban clinics obstruct primary care delivery, imposing time costs on patients, deterring appropriate utilization and causing patient dissatisfaction. This paper reports on an innovative attempt by staff in a large South African urban health centre to analyse a system of queues and preventive and curative services for pre-school children, and thereafter to evaluate changes. The study had a cross-sectional work study design, with repeated measurement of waiting times after 13 months. At baseline the preventive clinic was found to have several inessential processes and waits; these were eliminated or overlapped, and clinic sessions per week were increased. A year later median waiting times had decreased substantially in the preventive clinic, but had increased in the curative clinic. Simple research can explain long waits, inform and measure changes, and provide evidence to justify primary care integration and would be useful in health centres and hospital outpatient departments in developing countries.

  9. Evaluation of a New Equation for Calculating the Maximum Wait Time for Pilots That Have Used an Impairing Medication

    DTIC Science & Technology

    2013-08-01

    were treating pre-existing medical conditions using over-the-counter (OTC) medications ( Aspirin ™, Tylenol™, antihistamines, etc.), provided blood...time would decrease to 45 hours for a 150-lb person taking the same dose; a 300 -lb individual taking a 25-mg dose would only need to wait 31 hours...If the 300 -lb person mentioned above had taken a 50- mg dose, the wait time would have been 45 hours, which is approximately the same as the 49

  10. Effect of electronically delivered prescriptions on compliance and pharmacy wait time among emergency department patients.

    PubMed

    Fernando, Tasha J; Nguyen, Duy D; Baraff, Larry J

    2012-01-01

    The primary objectives were to assess whether electronically delivered prescriptions lead to reduced pharmacy wait time, improved patient satisfaction, and improved compliance with prescriptions. Secondary objectives included determining other reasons for noncompliance and if there was an association between prescription noncompliance and subsequent physician and emergency department (ED) visits. In this prospective study, patients discharged from the Ronald Reagan UCLA Medical Center ED with prescriptions for nonnarcotic medications were randomized to a control group who were discharged with standard written prescriptions or an intervention group who had their prescriptions electronically delivered to the pharmacy of their choice. All study participants were contacted 7 to 31 days after ED discharge for a structured telephone interview. Of the 454 patients enrolled, follow-up was successful for 224 patients (52.4%). Twenty-eight patients did not fill their prescriptions (12.5% noncompliance rate). The top three reasons patients stated for not picking up their medications were perceiving their prescription as unnecessary (n = 11), medication affordability (n = 5), and lack of time (n = 4). There was no difference in primary prescription noncompliance between the two study groups (p = 0.58). However, electronically delivered prescriptions significantly reduced the median pharmacy wait time, from 15 to 0 minutes (p = 0.001), and improved patient satisfaction at the pharmacy (p = 0.034). Neither subsequent physician nor ED visits were increased by primary prescription noncompliance. Electronically delivered prescriptions significantly minimized pharmacy wait time and improved patient satisfaction at the pharmacy, but did not improve primary compliance with prescriptions. © 2011 by the Society for Academic Emergency Medicine.

  11. The effect of waiting: A meta-analysis of wait-list control groups in trials for tinnitus distress.

    PubMed

    Hesser, Hugo; Weise, Cornelia; Rief, Winfried; Andersson, Gerhard

    2011-04-01

    The response rates and effects of being placed on a wait-list control condition are well documented in psychiatric populations. Despite the usefulness of such estimates and the frequent use of no-treatment controls in clinical trials for tinnitus, the effect of waiting in a tinnitus trial has not been investigated systematically. The aim of the present study was to quantify the overall effect of wait-list control groups on tinnitus distress. Studies were retrieved via a systematic review of randomised controlled trials of cognitive behaviour therapy for tinnitus distress. Outcomes of psychometrically robust tinnitus-specific measures (Tinnitus Handicap Inventory, Tinnitus Questionnaire, Tinnitus Reaction Questionnaire) from wait-list control groups were quantified using meta-analytic techniques. Percentage of change and standard mean difference effect sizes were calculated using the pre and post wait period. Eleven studies involving 314 wait-list subjects with tinnitus were located. The analysis for a waiting period of 6 to 12 weeks revealed a mean decrease in scores on tinnitus-specific measures of 3% to 8%. Across studies, a statically significant small mean within-group effect size was obtained (Hedges' g=.17). The effects were moderated by methodological quality of the trial, sample characteristics (i.e., age, tinnitus duration), time of the wait-list and how diagnosis was established. Subjects in a tinnitus trial improve in tinnitus distress over a short waiting phase. The effects of waiting are highly variable and depend on the characteristics of the sample and of the trial. Copyright © 2011 Elsevier Inc. All rights reserved.

  12. Effect of match-run frequencies on the number of transplants and waiting times in kidney exchange.

    PubMed

    Ashlagi, Itai; Bingaman, Adam; Burq, Maximilien; Manshadi, Vahideh; Gamarnik, David; Murphey, Cathi; Roth, Alvin E; Melcher, Marc L; Rees, Michael A

    2018-05-01

    Numerous kidney exchange (kidney paired donation [KPD]) registries in the United States have gradually shifted to high-frequency match-runs, raising the question of whether this harms the number of transplants. We conducted simulations using clinical data from 2 KPD registries-the Alliance for Paired Donation, which runs multihospital exchanges, and Methodist San Antonio, which runs single-center exchanges-to study how the frequency of match-runs impacts the number of transplants and the average waiting times. We simulate the options facing each of the 2 registries by repeated resampling from their historical pools of patient-donor pairs and nondirected donors, with arrival and departure rates corresponding to the historical data. We find that longer intervals between match-runs do not increase the total number of transplants, and that prioritizing highly sensitized patients is more effective than waiting longer between match-runs for transplanting highly sensitized patients. While we do not find that frequent match-runs result in fewer transplanted pairs, we do find that increasing arrival rates of new pairs improves both the fraction of transplanted pairs and waiting times. © 2017 The American Society of Transplantation and the American Society of Transplant Surgeons.

  13. How to report and monitor the performance of waiting list management.

    PubMed

    Torkki, Markus; Linna, Miika; Seitsalo, Seppo; Paavolainen, Pekka

    2002-01-01

    Potential problems concerning waiting list management are often monitored using mean waiting times based on empirical samples. However, the appropriateness of mean waiting time as an indicator of access can be questioned if a waiting list is not managed well, e.g., if the queue discipline is violated. This study was performed to find out about the queue discipline in waiting lists for elective surgery to reveal potential discrepancies in waiting list management. There were 1,774 waiting list patients for hallux valgus or varicose vein surgery or sterilization. The waiting time distributions of patients receiving surgery and of patients still waiting for an operation are presented in column charts. The charts are compared with two model charts. One model chart presents a high queue discipline (first in-first out) and another a poor queue discipline (random) queue. There were significant differences in waiting list management across hospitals and patient categories. Examples of a poor queue discipline were found in queues for hallux valgus and varicose vein operations. A routine waiting list reporting should be used to guarantee the quality of waiting list management and to pinpoint potential problems in access. It is important to monitor not only the number of patients in the waiting list but also the queue discipline and the balance between demand and supply of surgical services. The purpose for this type of reporting is to ensure that the priority setting made at health policy level also works in practise.

  14. [Surgery for colorectal cancer since the introduction of the Netherlands national screening programmeInvestigations into changes in number of resections and waiting times for surgery].

    PubMed

    de Neree Tot Babberich, M P M; van der Willik, E M; van Groningen, J T; Ledeboer, M; Wiggers, T; Wouters, M W J M

    2017-01-01

    To investigate the impact of the Netherlands national colorectal cancer screening programme on the number of surgical resections for colorectal carcinoma and on waiting times for surgery. Descriptive study. Data were extracted from the Dutch Surgical Colorectal Audit. Patients with primary colorectal cancer surgery between 2011-2015 were included. The volume and median waiting times for the years 2011-2015 are described. Waiting times from first tumor positive biopsy until the operation (biopsy-operation) and first preoperative visit to the surgeon until the operation (visit-operation) are analyzed with a univariate and multivariate linear regression analysis. Separate analysis was done for visit-operation for academic and non-academic hospitals and for screening compared to non-screening patients. In 2014 there was an increase of 1469 (15%) patients compared to 2013. In 2015 this increase consisted of 1168 (11%) patients compared to 2014. In 2014 and 2015, 1359 (12%) and 3111 (26%) patients were referred to the surgeon through screening, respectively. The median waiting time of biopsy-operation significantly decreased (ß: 0.94, 95%BI) over the years 2014-2015 compared to 2011-2013. In non-academic hospitals, the waiting time visit-operation also decreased significantly (ß: 0.89, 95%BI 0.87-0.90) over the years 2014-2015 compared to 2011-2013. No difference was found in waiting times between patients referred to the surgeon through screening compared to non-screening. There is a clear increase in volume since the introduction of the colorectal cancer screening programme without an increase in waiting time until surgery.

  15. Who breaches the four-hour emergency department wait time target? A retrospective analysis of 374,000 emergency department attendances between 2008 and 2013 at a type 1 emergency department in England.

    PubMed

    Bobrovitz, Niklas; Lasserson, Daniel S; Briggs, Adam D M

    2017-11-02

    The four-hour target is a key hospital emergency department performance indicator in England and one that drives the physical and organisational design of the ED. Some studies have identified time of presentation as a key factor affecting waiting times. Few studies have investigated other determinants of breaching the four-hour target. Therefore, our objective was to describe patterns of emergency department breaches of the four-hour wait time target and identify patients at highest risk of breaching. This was a retrospective cohort study of a large type 1 Emergency department at an NHS teaching hospital in Oxford, England. We analysed anonymised individual level patient data for 378,873 emergency department attendances, representing all attendances between April 2008 and April 2013. We examined patient characteristics and emergency department presentation circumstances associated with the highest likelihood of breaching the four-hour wait time target. We used 374,459 complete cases for analysis. In total, 8.3% of all patients breached the four-hour wait time target. The main determinants of patients breaching the four-hour wait time target were hour of arrival to the ED, day of the week, patient age, ED referral source, and the types of investigations patients receive (p < 0.01 for all associations). Patients most likely to breach the four-hour target were older, presented at night, presented on Monday, received multiple types of investigation in the emergency department, and were not self-referred (p < 0.01 for all associations). Patients attending from October to February had a higher odds of breaching compared to those attending from March to September (OR 1.63, 95% CI 1.59 to 1.66). There are a number of independent patient and circumstantial factors associated with the probability of breaching the four-hour ED wait time target including patient age, ED referral source, the types of investigations patients receive, as well as the hour, day, and month of

  16. Headache Education Active-Waiting Directive: A Program to Enhance Well-Being During Long Referral Wait Times.

    PubMed

    Lagman-Bartolome, Ana Marissa; Lawler, Valerie; Lay, Christine

    2018-01-01

    The aim of this initiative was to evaluate the clinical impact, patient acceptability, and sustainability of implementing a newly developed evidence-guided migraine education program in an academic headache center. Headache is the fifth most common reason for emergency department (ED) visits and accounts for more than 10 million physician visits annually. Successful management of headaches presents a challenge to both primary care providers and neurologists. The catchment area for an academic headache specialty center in a large metropolitan area is over 6 million with an average wait time of over 15 months. This delays diagnosis and impacts patients, thus a Headache Education Active-Waiting Directive (HEAD) was developed to improve patient knowledge and self-care skills among migraine patients awaiting an initial appointment. This was a prospective pre- and post-intervention study. English-speaking adults, aged 18-65 years, referred to the Center for Headache at the University of Toronto for headache consultation between May and December 2012, who had not previously been assessed by a headache specialist, were consented and enrolled. Data on Migraine Disability Assessment (MIDAS) with additional questions on emergency visits, lifestyle, and self-efficacy were collected premigraine and postmigraine education program session. Two hundred and forty-eight patients attended the HEAD program and 177 (71%) consented to the study. Detailed predata and postdata were available for 152 participants (mean age 42.5 ± 11.9 years, 86% females); 117/150 (78%) presented with depressive symptoms and 90/146 (62%) presented with anxiety symptoms. One hundred and thirty-seven of 143 (96%) were using headache treatment. Eighty of 137 (58%) were overusing over-the-counter medications and only 21/137 (15%) were on preventative treatment.  There was a decrease in the MIDAS scores of participants at postsession testing prior to neurological consultation (pre-MIDAS mean 50.0

  17. Investigating the Relationship between Customer Wait Time and Operational Availability through Simulation Modeling

    DTIC Science & Technology

    2012-12-01

    STATEMENT Approved for public release; distribution is unlimited 12b. DISTRIBUTION CODE A 13. ABSTRACT (maximum 200 words) Customer Wait Time ( CWT ...inventory level, thereby increasing the material readiness of the operating forces. Intuitively, decreasing CWT increases operational availability (Ao...and CWT has led to arbitrary stock policies that do not account for the cost and benefit they provide. This project centers on monetizing the

  18. Interior effects on comfort in healthcare waiting areas.

    PubMed

    Bazley, C; Vink, P; Montgomery, J; Hedge, A

    2016-07-21

    This study compared the effects of pre-experience and expectations on participant comfort upon waking, arrival to, and after an appointment, as well as the assessment of properly placed Feng Shui elements in three healthcare waiting rooms. Participants assessed comfort levels using self-report surveys. The researcher conducted 'intention interviews' with each doctor to assess the goals of each waiting area design, and conducted a Feng Shui assessment of each waiting area for properly placed Feng Shui elements. The waiting area designed by the Feng Shui expert rated 'most comfortable', followed by the waiting area design by a doctor, and the lowest comfort rating for the conventional waiting room design. Results show a sufficiently strong effect to warrant further research. Awareness of the external environment, paired with pre-experience and expectation, influences comfort for people over time. Fostering and encouraging a holistic approach to comfort utilizing eastern and western concepts and ergonomic principles creates a sense of "placeness" and balance in the design for comfort in built environments. This is new research information on the influences of the comfort experience over time, to include pre-experience, expectations and the placement of elements in the external environment.

  19. What factors influence cataract waiting list time?

    PubMed Central

    Churchill, A.; Vize, C.; Stewart, O.; Backhouse, O.

    2000-01-01

    AIMS—To determine whether there were any specific factors that influenced waiting list time (WLT) for patients undergoing cataract surgery.
METHODS—70 preoperative cataract patients were interviewed by one of the authors using a questionnaire to score visual acuity, coexisting ocular pathology and disabilities, threat to independent living/employment, and perceived visual handicap for detailed, gross, and driving vision. Individuals were analysed separately according to whether it was their first or second cataract operation.
RESULTS—The median WLT for first eye surgery was 9 months (n = 31) and 13 months for second eye surgery (n = 36). The WLT ranged from 2 to 25 months for first eyes and 0.25-18 months for second eyes. Where there was a perceived threat to independent living or employment the WLT was found to be significantly shorter than the median. A high overall score correlated with a shorter WLT. Surgical priority was also given to individuals with anisometropia >3 dioptres.
CONCLUSION—This study has demonstrated that there are specific factors that influence clinicians when prioritising patients for cataract surgery.

 PMID:10729304

  20. Analyzing patient's waiting time in emergency & trauma department in public hospital - A case study

    NASA Astrophysics Data System (ADS)

    Roslan, Shazwa; Tahir, Herniza Md; Nordin, Noraimi Azlin Mohd; Zaharudin, Zati Aqmar

    2014-09-01

    Emergency and Trauma Department (ETD) is an important element for a hospital. It provides medical service, which operates 24 hours a day in most hospitals. However overcrowding is not exclusion for ETD. Overflowing occurs due to affordable services provided by public hospitals, since it is funded by the government. It is reported that a patient attending ETD must be treated within 90 minutes, in accordance to achieve the Key Performance Indicator (KPI). However, due to overcrowd situations, most patients have to wait longer than the KPI standard. In this paper, patient's average waiting time is analyzed. Using Chi-Square Test of Goodness, patient's inter arrival per hour is also investigated. As conclusion, Monday until Wednesday was identified as the days that exceed the KPI standard while Chi-Square Test of Goodness showed that the patient's inter arrival is independent and random.

  1. Delay decomposition at a single server queue with constant service time and multiple inputs. [Waiting time on computer network

    NASA Technical Reports Server (NTRS)

    Ziegler, C.; Schilling, D. L.

    1977-01-01

    Two networks consisting of single server queues, each with a constant service time, are considered. The external inputs to each network are assumed to follow some general probability distribution. Several interesting equivalencies that exist between the two networks considered are derived. This leads to the introduction of an important concept in delay decomposition. It is shown that the waiting time experienced by a customer can be decomposed into two basic components called self delay and interference delay.

  2. Waiting time distribution for continuous stochastic systems

    NASA Astrophysics Data System (ADS)

    Gernert, Robert; Emary, Clive; Klapp, Sabine H. L.

    2014-12-01

    The waiting time distribution (WTD) is a common tool for analyzing discrete stochastic processes in classical and quantum systems. However, there are many physical examples where the dynamics is continuous and only approximately discrete, or where it is favourable to discuss the dynamics on a discretized and a continuous level in parallel. An example is the hindered motion of particles through potential landscapes with barriers. In the present paper we propose a consistent generalization of the WTD from the discrete case to situations where the particles perform continuous barrier crossing characterized by a finite duration. To this end, we introduce a recipe to calculate the WTD from the Fokker-Planck (Smoluchowski) equation. In contrast to the closely related first passage time distribution (FPTD), which is frequently used to describe continuous processes, the WTD contains information about the direction of motion. As an application, we consider the paradigmatic example of an overdamped particle diffusing through a washboard potential. To verify the approach and to elucidate its numerical implications, we compare the WTD defined via the Smoluchowski equation with data from direct simulation of the underlying Langevin equation and find full consistency provided that the jumps in the Langevin approach are defined properly. Moreover, for sufficiently large energy barriers, the WTD defined via the Smoluchowski equation becomes consistent with that resulting from the analytical solution of a (two-state) master equation model for the short-time dynamics developed previously by us [Phys. Rev. E 86, 061135 (2012), 10.1103/PhysRevE.86.061135]. Thus, our approach "interpolates" between these two types of stochastic motion. We illustrate our approach for both symmetric systems and systems under constant force.

  3. 24 CFR 982.204 - Waiting list: Administration of waiting list.

    Code of Federal Regulations, 2010 CFR

    2010-04-01

    ... size (number of bedrooms for which family qualifies under PHA occupancy standards); (3) Date and time... list. (d) Family size. (1) The order of admission from the waiting list may not be based on family size, or on the family unit size for which the family qualifies under the PHA occupancy policy. (2) If the...

  4. Waiting Online: A Review and Research Agenda.

    ERIC Educational Resources Information Center

    Ryan, Gerard; Valverde, Mireia

    2003-01-01

    Reviews 21 papers based on 13 separate empirical studies on waiting on the Internet, drawn from the areas of marketing, system response time, and quality of service studies. The article proposes an agenda for future research, including extending the range of research methodologies, broadening the definition of waiting on the Internet, and…

  5. Do new workforce roles reduce waiting times in ED? A difference-in-difference evaluation using hospital administrative data.

    PubMed

    Scott, Anthony; Yong, Jongsay

    2015-04-01

    This paper evaluates the effect of introducing two new workforce roles under a pilot program conducted in Victoria, Australia. The trial took place at a regional hospital's emergency department (ED) between 1 July 2008 and 30 June 2009. The evaluation is based on three outcome measures: waiting time (in minutes) at ED before treatment; proportion of presentations with waiting time on target; and length of stay (in days), for ED presentations that led to in-patient admissions. The technique of difference-in-differences analysis is used. A total of 142,980 patient records from the pilot hospital and three comparison hospitals were extracted from the Victorian Emergency Minimum Dataset (VEMD). Further, 21,925 records of patients whose ED presentations led to in-patient admissions were extracted from the Victorian Admitted Episodes Dataset (VAED). The evaluation finds the piloted roles have lowered waiting time and raised the proportion of on-target presentations. These effects were found to be the strongest for less urgent triage categories. However, the evidence on in-patient length of stay was mixed. The results provide positive evidence that new workforce roles can be effective in improving the efficiency of emergency care delivery. Copyright © 2014 Elsevier Ireland Ltd. All rights reserved.

  6. Factors Associated with Waiting Time for Breast Cancer Treatment in a Teaching Hospital in Ghana

    ERIC Educational Resources Information Center

    Dedey, Florence; Wu, Lily; Ayettey, Hannah; Sanuade, Olutobi A.; Akingbola, Titilola S.; Hewlett, Sandra A.; Tayo, Bamidele O.; Cole, Helen V.; de-Graft Aikins, Ama; Ogedegbe, Gbenga; Adanu, Richard

    2016-01-01

    Background: Breast cancer is the leading cause of cancer-related mortality among women in Ghana. Data are limited on the predictors of poor outcomes in breast cancer patients in low-income countries; however, prolonged waiting time has been implicated. Among breast cancer patients who received treatment at Korle Bu Teaching Hospital, this study…

  7. Waiting for coronary angiography: is there a clinically ordered queue?

    PubMed

    Hemingway, H; Crook, A M; Feder, G; Dawson, J R; Timmis, A

    2000-03-18

    Among over 3000 patients undergoing coronary angiography in the absence of a formal queue-management system, we found that a-priori urgency scores were strongly associated with waiting times, prevalence of coronary-artery disease, rate of revascularisation, and mortality. These data challenge the widely held assumption that such waiting lists are not clinically ordered; however, the wide variation in waiting times within urgency categories suggests the need for further improvements in clinical queueing.

  8. Public reporting on quality, waiting times and patient experience in 11 high-income countries.

    PubMed

    Rechel, Bernd; McKee, Martin; Haas, Marion; Marchildon, Gregory P; Bousquet, Frederic; Blümel, Miriam; Geissler, Alexander; van Ginneken, Ewout; Ashton, Toni; Saunes, Ingrid Sperre; Anell, Anders; Quentin, Wilm; Saltman, Richard; Culler, Steven; Barnes, Andrew; Palm, Willy; Nolte, Ellen

    2016-04-01

    This article maps current approaches to public reporting on waiting times, patient experience and aggregate measures of quality and safety in 11 high-income countries (Australia, Canada, England, France, Germany, Netherlands, New Zealand, Norway, Sweden, Switzerland and the United States). Using a questionnaire-based survey of key national informants, we found that the data most commonly made available to the public are on waiting times for hospital treatment, being reported for major hospitals in seven countries. Information on patient experience at hospital level is also made available in many countries, but it is not generally available in respect of primary care services. Only one of the 11 countries (England) publishes composite measures of overall quality and safety of care that allow the ranking of providers of hospital care. Similarly, the publication of information on outcomes of individual physicians remains rare. We conclude that public reporting of aggregate measures of quality and safety, as well as of outcomes of individual physicians, remain relatively uncommon. This is likely to be due to both unresolved methodological and ethical problems and concerns that public reporting may lead to unintended consequences. Copyright © 2016 The Authors. Published by Elsevier Ireland Ltd.. All rights reserved.

  9. Impact of the single point of access referral system to reduce waiting times and improve clinical outcomes in an assistive technology service.

    PubMed

    Hosking, Jonathan; Gibson, Colin

    2016-07-01

    The introduction of a single point referral system that prioritises clients depending on case complexity and overcomes the need for re-admittance to a waiting list via a review system has been shown to significantly reduce maximum waiting times for a Posture and Mobility (Special Seating) Service from 102.0 ± 24.33 weeks to 19.2 ± 8.57 weeks (p = 0.015). Using this service model linear regression revealed a statistically significant improvement in the performance outcome of prescribed seating solutions with shorter Episode of Care completion times (p = 0.023). In addition, the number of Episodes of Care completed per annum was significantly related to the Episode of Care completion time (p = 0.019). In conclusion, it is recommended that it may be advantageous to apply this service model to other assistive technology services in order to reduce waiting times and to improve clinical outcomes.

  10. Timed Transfer : An Evaluation of Its Structure, Performance and Cost

    DOT National Transportation Integrated Search

    1983-08-01

    Timed transfer is a transit operating strategy in which vehicles from different routes are routed and scheduled to meet simultaneously at common stops to facilitate no-wait or minimum-wait passenger transfers. Timed transfers are being used primarily...

  11. Seasonality of service provision in hip and knee surgery: a possible contributor to waiting times? A time series analysis.

    PubMed

    Upshur, Ross E G; Moineddin, Rahim; Crighton, Eric J; Mamdani, Muhammad

    2006-03-01

    The question of how best to reduce waiting times for health care, particularly surgical procedures such as hip and knee replacements is among the most pressing concern of the Canadian health care system. The objective of this study was to test the hypothesis that significant seasonal variation exists in the performance of hip and knee replacement surgery in the province of Ontario. We performed a retrospective, cross-sectional time series analysis examining all hip and knee replacement surgeries in people over the age of 65 in the province of Ontario, Canada between 1992 and 2002. The main outcome measure was monthly hospitalization rates per 100,000 population for all hip and knee replacements. There was a marked increase in the rate of hip and knee replacement surgery over the 10-year period as well as an increasing seasonal variation in surgeries. Highly significant (Fisher Kappa = 16.05, p < 0.01; Bartlett-Kolmogorov-Smirnov Test = 0.31, p < 0.01) and strong (R2Autoreg = 0.85) seasonality was identified in the data. Holidays and utilization caps appear to exert a significant influence on the rate of service provision. It is expected that waiting times for hip and knee replacement could be reduced by reducing seasonal fluctuations in service provision and benchmarking services to peak delivery. The results highlight the importance of system behaviour in seasonal fluctuation of service delivery.

  12. The effect of early education on patient anxiety while waiting for elective cardiac catheterization.

    PubMed

    Harkness, Karen; Morrow, Lydia; Smith, Kelly; Kiczula, Michele; Arthur, Heather M

    2003-07-01

    A supply-demand mismatch with respect to cardiac catheterization (CATH) often results in patients experiencing waiting times that vary from a few weeks to several months. Long delays can impose both physical and psychological distress for patients. The purpose of this study was to examine the effect of a psychoeducational nursing intervention at the beginning of the waiting period on patient anxiety during the waiting time for elective CATH. This was a 2-group randomized controlled trial. Intervention patients received a nurse-delivered, detailed information/education session within 2 weeks of being placed on the waiting list for elective CATH. Control group patients received usual care. The mean waiting time for CATH was 13.4+/-7.2 weeks, which did not differ between groups (P=0.509). Anxiety increased in both groups over the waiting time (P=0.028). Health-related quality of life deteriorated over the waiting time in both groups (P<0.05). On a visual analogue scale, there was a significant difference (P=0.002) between the intervention (4.0+/-2.7) and control (5.2+/-3.0) groups in self-reported anxiety 2 weeks prior to CATH. The waiting period prior to elective CATH has a negative impact on patients' perceived anxiety and quality of life and a simple intervention, provided at the beginning of the waiting period, may positively affect the experience of waiting.

  13. Waiting for hip arthroplasty: economic costs and health outcomes.

    PubMed

    Fielden, Jann M; Cumming, J M; Horne, J G; Devane, P A; Slack, A; Gallagher, L M

    2005-12-01

    This prospective cohort study of 153 patients aimed to determine the economic and health costs of waiting for total hip arthroplasty (THA). Health-related quality of life, using self-completed WOMAC and EQ-5D questionnaires, was assessed monthly from enrolment preoperatively to 6 months postsurgery. Monthly cost diaries were used to record costs. The mean waiting time was 5.1 months and mean total cost of waiting for surgery was NZ 4305 dollars(US 2876 dollars) per person (pp) (NZ 1 dollar = US 0.668 dollar). Waiting more than 6 months was associated with a higher total mean cost (NZ 4278 dollars/US 2858 dollars pp) than waiting less than 6 months (NZ 2828 dollars/US 1889 dollars pp; P < .01). Improvements from preoperative to postoperative WOMAC and EQ-5D scores were identified (P < or = .01). Waiting longer led to poorer physical function preoperatively (P < or = .01). Those with poor initial health status showed greater improvement in WOMAC (P = .0001) and EQ-5D (P = .003) measures by 6 months after surgery. Longer waits for total hip arthroplasty incur greater economic costs and deterioration in physical function while waiting.

  14. On the gap between an empirical distribution and an exponential distribution of waiting times for price changes in a financial market

    NASA Astrophysics Data System (ADS)

    Sazuka, Naoya

    2007-03-01

    We analyze waiting times for price changes in a foreign currency exchange rate. Recent empirical studies of high-frequency financial data support that trades in financial markets do not follow a Poisson process and the waiting times between trades are not exponentially distributed. Here we show that our data is well approximated by a Weibull distribution rather than an exponential distribution in the non-asymptotic regime. Moreover, we quantitatively evaluate how much an empirical data is far from an exponential distribution using a Weibull fit. Finally, we discuss a transition between a Weibull-law and a power-law in the long time asymptotic regime.

  15. Colour Consideration for Waiting areas in hospitals

    NASA Astrophysics Data System (ADS)

    Zraati, Parisa

    2012-08-01

    Colour is one the most important factors in the nature that can have some affects on human behaviour. Many years ago, it was proven that using colour in public place can have some affect on the users. Depend of the darkness and lightness; it can be vary from positive to negative. The research will mainly focus on the colour and psychological influences and physical factors. The statement of problem in this research is what is impact of colour usually applied to waiting area? The overall aim of the study is to explore the visual environment of hospitals and to manage the colour psychological effect of the hospital users in the waiting area by creating a comfortable, pleasant and cozy environment for users while spend their time in waiting areas. The analysisconcentrate on satisfaction and their interesting regarding applied colour in two private hospital waiting area in Malaysia.

  16. Delays in Prior Living Kidney Donors Receiving Priority on the Transplant Waiting List

    PubMed Central

    Klassen, David K.; Kucheryavaya, Anna Y.; Stewart, Darren E.

    2016-01-01

    Background and objectives Prior living donors (PLDs) receive very high priority on the Organ Procurement and Transplantation Network (OPTN) kidney waiting list. Program delays in adding PLDs to the waiting list, setting their status to active, and submitting requests for PLD priority can affect timely access to transplantation. Design, setting, participants, & measurements We used the OPTN and the Centers for Medicare and Medicaid Services data to examine timing of (1) listing relative to start of dialysis, (2) activation on the waiting list, and (3) requests for PLD priority relative to listing date. There were 210 PLDs (221 registrations) added to the OPTN kidney waiting list between January 1, 2010 and July 31, 2015. Results As of September 4, 2015, 167 of the 210 PLDs received deceased donor transplants, six received living donor transplants, two died, five were too sick to transplant, and 29 were still waiting. Median waiting time to deceased donor transplant for PLDs was 98 days. Only 40.7% of 221 PLD registrations (n=90) were listed before they began dialysis; 68.3% were in inactive status for <90 days, 17.6% were in inactive status for 90–365 days, 8.6% were in inactive status for 1–2 years, and 5.4% were in inactive status for >2 years. Median time of PLDs waiting in active status before receiving PLD priority was 2 days (range =0–1450); 67.4% of PLDs received PLD priority within 7 days after activation, but 15.4% waited 8–30 days, 8.1% waited 1–3 months, 4.1% waited 3–12 months, and 5.0% waited >1 year in active status for PLD priority. After receiving priority, most were transplanted quickly. Median time in active status with PLD priority before deceased donor transplant was 23 days. Conclusions Fewer than one half of listed PLDs were listed before starting dialysis. Most listed PLDs are immediately set to active status and receive PLD priority quickly, but a substantial number spends time in active status without PLD priority or a large

  17. What's in a wait? Contrasting management science and economic perspectives on waiting for emergency care.

    PubMed

    Morton, Alec; Bevan, Gwyn

    2008-02-01

    The current paper reviews and contrasts a management science view of waiting for healthcare, which centres on queues as devices for buffering demand, with an economic view, which stresses the role of the incentive structure, in the context of English Accident and Emergency Departments. We demonstrate that the management science view provides insight into waiting time performance within a single facility but is limited in its ability to shed light on variations in performance across facilities. We argue, with reference to supporting data, that such variations may be explainable by a proper understanding of the incentive structure in A&E Departments.

  18. Seasonality of service provision in hip and knee surgery: A possible contributor to waiting times? A time series analysis

    PubMed Central

    Upshur, Ross EG; Moineddin, Rahim; Crighton, Eric J; Mamdani, Muhammad

    2006-01-01

    Background The question of how best to reduce waiting times for health care, particularly surgical procedures such as hip and knee replacements is among the most pressing concern of the Canadian health care system. The objective of this study was to test the hypothesis that significant seasonal variation exists in the performance of hip and knee replacement surgery in the province of Ontario. Methods We performed a retrospective, cross-sectional time series analysis examining all hip and knee replacement surgeries in people over the age of 65 in the province of Ontario, Canada between 1992 and 2002. The main outcome measure was monthly hospitalization rates per 100 000 population for all hip and knee replacements. Results There was a marked increase in the rate of hip and knee replacement surgery over the 10-year period as well as an increasing seasonal variation in surgeries. Highly significant (Fisher Kappa = 16.05, p < 0.01; Bartlett-Kolmogorov-Smirnov Test = 0.31, p < 0.01) and strong (R2Autoreg = 0.85) seasonality was identified in the data. Conclusion Holidays and utilization caps appear to exert a significant influence on the rate of service provision. It is expected that waiting times for hip and knee replacement could be reduced by reducing seasonal fluctuations in service provision and benchmarking services to peak delivery. The results highlight the importance of system behaviour in seasonal fluctuation of service delivery. PMID:16509992

  19. Waiting Narratives of Lung Transplant Candidates

    PubMed Central

    Yelle, Maria T.; Stevens, Patricia E.; Lanuza, Dorothy M.

    2013-01-01

    Before 2005, time accrued on the lung transplant waiting list counted towards who was next in line for a donor lung. Then in 2005 the lung allocation scoring system was implemented, which meant the higher the illness severity scores, the higher the priority on the transplant list. Little is known of the lung transplant candidates who were listed before 2005 and were caught in the transition when the lung allocation scoring system was implemented. A narrative analysis was conducted to explore the illness narratives of seven lung transplant candidates between 2006 and 2007. Arthur Kleinman's concept of illness narratives was used as a conceptual framework for this study to give voice to the illness narratives of lung transplant candidates. Results of this study illustrate that lung transplant candidates expressed a need to tell their personal story of waiting and to be heard. Recommendation from this study calls for healthcare providers to create the time to enable illness narratives of the suffering of waiting to be told. Narrative skills of listening to stories of emotional suffering would enhance how healthcare providers could attend to patients' stories and hear what is most meaningful in their lives. PMID:23476760

  20. Waiting narratives of lung transplant candidates.

    PubMed

    Yelle, Maria T; Stevens, Patricia E; Lanuza, Dorothy M

    2013-01-01

    Before 2005, time accrued on the lung transplant waiting list counted towards who was next in line for a donor lung. Then in 2005 the lung allocation scoring system was implemented, which meant the higher the illness severity scores, the higher the priority on the transplant list. Little is known of the lung transplant candidates who were listed before 2005 and were caught in the transition when the lung allocation scoring system was implemented. A narrative analysis was conducted to explore the illness narratives of seven lung transplant candidates between 2006 and 2007. Arthur Kleinman's concept of illness narratives was used as a conceptual framework for this study to give voice to the illness narratives of lung transplant candidates. Results of this study illustrate that lung transplant candidates expressed a need to tell their personal story of waiting and to be heard. Recommendation from this study calls for healthcare providers to create the time to enable illness narratives of the suffering of waiting to be told. Narrative skills of listening to stories of emotional suffering would enhance how healthcare providers could attend to patients' stories and hear what is most meaningful in their lives.

  1. Evaluation of a new equation for calculating the maximum wait time for pilots that have used an impairing medication.

    DOT National Transportation Integrated Search

    2013-08-01

    Pilots thatuse an impairing medication to treat a medicalcondition are required to wait an appropriate amount of time after completing the treatment before returning to duty.However, toxicology findings for pilots involved in fatal aviation accidents...

  2. An Efficient Wait-Free Vector

    DOE PAGES

    Feldman, Steven; Valera-Leon, Carlos; Dechev, Damian

    2016-03-01

    The vector is a fundamental data structure, which provides constant-time access to a dynamically-resizable range of elements. Currently, there exist no wait-free vectors. The only non-blocking version supports only a subset of the sequential vector API and exhibits significant synchronization overhead caused by supporting opposing operations. Since many applications operate in phases of execution, wherein each phase only a subset of operations are used, this overhead is unnecessary for the majority of the application. To address the limitations of the non-blocking version, we present a new design that is wait-free, supports more of the operations provided by the sequential vector,more » and provides alternative implementations of key operations. These alternatives allow the developer to balance the performance and functionality of the vector as requirements change throughout execution. Compared to the known non-blocking version and the concurrent vector found in Intel’s TBB library, our design outperforms or provides comparable performance in the majority of tested scenarios. Over all tested scenarios, the presented design performs an average of 4.97 times more operations per second than the non-blocking vector and 1.54 more than the TBB vector. In a scenario designed to simulate the filling of a vector, performance improvement increases to 13.38 and 1.16 times. This work presents the first ABA-free non-blocking vector. Finally, unlike the other non-blocking approach, all operations are wait-free and bounds-checked and elements are stored contiguously in memory.« less

  3. No-waiting dentine self-etch concept-Merit or hype.

    PubMed

    Huang, Xue-Qing; Pucci, César R; Luo, Tao; Breschi, Lorenzo; Pashley, David H; Niu, Li-Na; Tay, Franklin R

    2017-07-01

    A recently-launched universal adhesive, G-Premio Bond, provides clinicians with the alternative to use the self-etch technique for bonding to dentine without waiting for the adhesive to interact with the bonding substrate (no-waiting self-etch; Japanese brochure), or after leaving the adhesive undisturbed for 10s (10-s self-etch; international brochure). The present study was performed to examine in vitro performance of this new universal adhesive bonded to human coronal dentine using the two alternative self-etch modes. One hundred and ten specimens were bonded using two self-etch application modes and examined with or without thermomechanical cycling (10,000 thermal cycles and 240,000 mechanical cycles) to simulate one year of intraoral functioning. The bonded specimens were sectioned for microtensile bond testing, ultrastructural and nanoleakage examination using transmission electron microscopy. Changes in the composition of mineralised dentine after adhesive application were examined using Fourier transform infrared spectroscopy. Both reduced application time and thermomechanical cycling resulted in significantly lower bond strengths, thinner hybrid layers, and significantly more extensive nanoleakage after thermomechanical cycling. Using the conventional 10-s application time improved bonding performance when compared with the no-waiting self-etch technique. Nevertheless, nanoleakage was generally extensive under all testing parameters employed for examining the adhesive. Although sufficient bond strength to dentine may be achieved using the present universal adhesive in the no-waiting self-etch mode that does not require clinicians to wait prior to polymerisation of the adhesive, this self-etch concept requires further technological refinement before it can be recommended as a clinical technique. Although the surge for cutting application time to increase user friendliness remains the most frequently sought conduit for advancement of dentine bonding

  4. Improvements in medical quality and patient safety through implementation of a case bundle management strategy in a large outpatient blood collection center.

    PubMed

    Zhao, Shuzhen; He, Lujia; Feng, Chenchen; He, Xiaoli

    2018-06-01

    Laboratory errors in blood collection center (BCC) are most common in the preanalytical phase. It is, therefore, of vital importance for administrators to take measures to improve healthcare quality and patient safety.In 2015, a case bundle management strategy was applied in a large outpatient BCC to improve its medical quality and patient safety.Unqualified blood sampling, complications, patient waiting time, largest number of patients waiting during peak hours, patient complaints, and patient satisfaction were compared over the period from 2014 to 2016.The strategy reduced unqualified blood sampling, complications, patient waiting time, largest number of patients waiting during peak hours, and patient complaints, while improving patient satisfaction.This strategy was effective in improving BCC healthcare quality and patient safety.

  5. The Effect of 5S-Continuous Quality Improvement-Total Quality Management Approach on Staff Motivation, Patients’ Waiting Time and Patient Satisfaction with Services at Hospitals in Uganda

    PubMed Central

    Take, Naoki; Byakika, Sarah; Tasei, Hiroshi; Yoshikawa, Toru

    2015-01-01

    This study aimed at analyzing the effect of 5S practice on staff motivation, patients’ waiting time and patient satisfaction with health services at hospitals in Uganda. Double-difference estimates were measured for 13 Regional Referral Hospitals and eight General Hospitals implementing 5S practice separately. The study for Regional Referral Hospitals revealed 5S practice had the effect on staff motivation in terms of commitment to work in the current hospital and waiting time in the dispensary in 10 hospitals implementing 5S, but significant difference was not identified on patient satisfaction. The study for General Hospitals indicated the effect of 5S practice on patient satisfaction as well as waiting time, but staff motivation in two hospitals did not improve. 5S practice enables the hospitals to improve the quality of services in terms of staff motivation, waiting time and patient satisfaction and it takes as least four years in Uganda. The fourth year since the commencement of 5S can be a threshold to move forward to the next step, Continuous Quality Improvement. PMID:28299136

  6. The Effect of 5S-Continuous Quality Improvement-Total Quality Management Approach on Staff Motivation, Patients' Waiting Time and Patient Satisfaction with Services at Hospitals in Uganda.

    PubMed

    Take, Naoki; Byakika, Sarah; Tasei, Hiroshi; Yoshikawa, Toru

    2015-03-31

    This study aimed at analyzing the effect of 5S practice on staff motivation, patients' waiting time and patient satisfaction with health services at hospitals in Uganda. Double-difference estimates were measured for 13 Regional Referral Hospitals and eight General Hospitals implementing 5S practice separately. The study for Regional Referral Hospitals revealed 5S practice had the effect on staff motivation in terms of commitment to work in the current hospital and waiting time in the dispensary in 10 hospitals implementing 5S, but significant difference was not identified on patient satisfaction. The study for General Hospitals indicated the effect of 5S practice on patient satisfaction as well as waiting time, but staff motivation in two hospitals did not improve. 5S practice enables the hospitals to improve the quality of services in terms of staff motivation, waiting time and patient satisfaction and it takes as least four years in Uganda. The fourth year since the commencement of 5S can be a threshold to move forward to the next step, Continuous Quality Improvement.

  7. The relationship between waiting times and 'adherence' to the Scottish Intercollegiate Guidelines Network 98 guideline in autism spectrum disorder diagnostic services in Scotland.

    PubMed

    McKenzie, Karen; Forsyth, Kirsty; O'Hare, Anne; McClure, Iain; Rutherford, Marion; Murray, Aja; Irvine, Linda

    2016-05-01

    The aim of this study was to explore the extent to which the Scottish Intercollegiate Guidelines Network 98 guidelines on the assessment and diagnosis of autism spectrum disorder were adhered to in child autism spectrum disorder diagnostic services in Scotland and whether there was a significant relationship between routine practice which more closely reflected these recommendations (increased adherence) and increased waiting times. Retrospective, cross-sectional case note analysis was applied to data from 80 case notes. Adherence ranged from a possible 0 (no adherence) to 19 (full adherence). Overall, 17/22 of the recommendations were adhered to in over 50 of the 80 cases and in 70 or more cases for 11/22 of the recommendations, with a mean adherence score of 16 (standard deviation = 1.9). No significant correlation was found between adherence and total wait time for untransformed (r = 0.15, p = 0.32) or transformed data (r = 0.12, p = 0.20). The results indicated that the assessment and diagnostic practices were consistent with the relevant Scottish Intercollegiate Guidelines Network 98 guideline recommendations. Increased adherence to the 19 included recommendations was not significantly related to increased total waiting times, indicating that the Scottish Intercollegiate Guidelines Network 98 recommendations have generally been integrated into practice, without a resultant increase in patient waits. © The Author(s) 2015.

  8. The effectiveness of interventions aimed at reducing anxiety in health care waiting spaces: a systematic review of randomized and nonrandomized trials.

    PubMed

    Biddiss, Elaine; Knibbe, Tara Joy; McPherson, Amy

    2014-08-01

    Reducing waiting anxiety is an important objective of patient-centered care. Anxiety is linked to negative health outcomes, including longer recovery periods, lowered pain thresholds, and for children in particular, resistance to treatment, nightmares, and separation anxiety. The goals of this study were (1) to systematically review published research aimed at reducing preprocedural waiting anxiety, and (2) to provide directions for future research and development of strategies to manage preprocedural waiting anxiety in health care environments. We performed a systematic review of the literature via ISI Web of Knowledge, PubMed, PsycINFO, EMBASE, CINAHL, and Medline. Included in this review were studies describing measurable outcomes in response to interventions specifically intended to improve the waiting experience of patients in health care settings. Primary outcomes of interest were stress and anxiety. Exclusion criteria included (a) studies aimed at reducing wait times and management of waiting lists only, (b) waiting in non-health care settings, (c) design of health care facilities with nonspecific strategies pertaining to waiting spaces, (d) strategies to reduce pain or anxiety during the course of medical procedures, and (e) interventions such as massage, acupuncture, or hypnosis that require dedicated staff and/or private waiting environments to administer. We identified 8690 studies. Forty-one articles met the inclusion criteria. In adult populations, 33 studies were identified, wherein the effects of music (n = 25), aromatherapy (n = 6), and interior design features (n = 2) were examined. Eight pediatric studies were identified investigating play opportunities (n = 2), media distractions (n = 2), combined play opportunities and media distractions (n = 3), and music (n = 1). Based on results from 1129 adult participants in the 14 studies that evaluated music and permitted meta-analysis, patients who listened to music before a medical procedure exhibited a

  9. Patient perceptions regarding physician reimbursements, wait times, and out-of-pocket payments for anterior cruciate ligament reconstruction in Ontario.

    PubMed

    Memon, Muzammil; Ginsberg, Lydia; de Sa, Darren; Nashed, Andrew; Simunovic, Nicole; Phillips, Mark; Denkers, Matthew; Ogilvie, Rick; Peterson, Devin; Ayeni, Olufemi R

    2017-12-01

    Currently, there is a lack of knowledge regarding patient perceptions surrounding physician reimbursements, appropriate wait times, and out-of-pocket payment options for anterior cruciate ligament reconstruction (ACLR). Our objective was to determine the current state of these perceptions in an Ontario setting. A survey was developed and pretested to address patient perceptions about physician reimbursements, appropriate wait times, and out-of-pocket payment options for ACLR using a focus group of experts and by reviewing prior surveys. The survey was administered to patients in a waiting room setting. Two hundred and fifty completed surveys were obtained (79.9% response rate). Participants responded that an appropriate physician reimbursement for ACLR was $1000.00 and that the Ontario Health Insurance Plan (OHIP) reimbursed physicians $700.00 for ACLR. Seventy-four percent of participants responded that the OHIP reimbursement of $615.20 for the procedure was either lower or much lower than what they considered to be an appropriate reimbursement for ACLR. Over 90% of participants responded that an ACLR should occur within 90 days of injury. Thirty-five percent of participants were willing to pay $750.00 out-of-pocket to have an ACLR done sooner, while 16.4% of participants were willing to pay $2500.00 out-of-pocket to travel outside of Canada for expedited surgery. This survey study demonstrates that patients' estimates of both appropriate and actual physician reimbursements were greater than the current reimbursement for ACLR. Further, the majority of individuals report that the surgical fee for ACLR is lower than what they consider to be an appropriate amount of compensation for the procedure. Additionally, nearly all respondents believe that a ruptured ACL should be reconstructed within 90 days of injury. Consequently, a number of patients are willing to pay out-of-pocket for expedited surgery either in Canada or abroad. However, patients' preferences for

  10. Handgun waiting periods reduce gun deaths

    PubMed Central

    Luca, Michael; Malhotra, Deepak

    2017-01-01

    Handgun waiting periods are laws that impose a delay between the initiation of a purchase and final acquisition of a firearm. We show that waiting periods, which create a “cooling off” period among buyers, significantly reduce the incidence of gun violence. We estimate the impact of waiting periods on gun deaths, exploiting all changes to state-level policies in the Unites States since 1970. We find that waiting periods reduce gun homicides by roughly 17%. We provide further support for the causal impact of waiting periods on homicides by exploiting a natural experiment resulting from a federal law in 1994 that imposed a temporary waiting period on a subset of states. PMID:29078268

  11. Not all waits are equal: an exploratory investigation of emergency care patient pathways.

    PubMed

    Swancutt, Dawn; Joel-Edgar, Sian; Allen, Michael; Thomas, Daniel; Brant, Heather; Benger, Jonathan; Byng, Richard; Pinkney, Jonathan

    2017-06-24

    Increasing pressure in the United Kingdom (UK) urgent care system has led to Emergency Departments (EDs) failing to meet the national requirement that 95% of patients are admitted, discharged or transferred within 4-h of arrival. Despite the target being the same for all acute hospitals, individual Trusts organise their services in different ways. The impact of this variation on patient journey time and waiting is unknown. Our study aimed to apply the Lean technique of Value Stream Mapping (VSM) to investigate care processes and delays in patient journeys at four contrasting hospitals. VSM timing data were collected for patients accessing acute care at four hospitals in South West England. Data were categorised according to waits and activities, which were compared across sites to identify variations in practice from the patient viewpoint. We included Public and Patient Involvement (PPI) to fully interpret our findings; observations and initial findings were considered in a PPI workshop. One hundred eight patients were recruited, comprising 25,432 min of patient time containing 4098 episodes of care or waiting. The median patient journey was 223 min (3 h, 43 min); just within the 4-h target. Although total patient journey times were similar between sites, the stage where the greatest proportion of waiting occurred varied. Reasons for waiting were dominated by waits for beds, investigations or results to be available. From our sample we observed that EDs without a discharge/clinical decision area exhibited a greater proportion of waiting time following an admission or discharge decision. PPI interpretation indicated that patients who experience waits at the beginning of their journey feel more anxious because they are 'not in the system yet'. The novel application of VSM analysis across different hospitals, coupled with PPI interpretation, provides important insight into the impact of care provision on patient experience. Measures that could reduce patient

  12. Use of HCV+ Donors Does Not Affect HCV Clearance With Directly Acting Antiviral Therapy But Shortens the Wait Time to Kidney Transplantation.

    PubMed

    Sawinski, Deirdre; Patel, Nikunjkumar; Appolo, Brenda; Bloom, Roy

    2017-05-01

    Hepatitis C virus (HCV) infection is prevalent in the renal transplant population but direct acting antiviral agents (DAA) provide an effective cure of HCV infection without risk of allograft rejection. We report our experience treating 43 renal transplant recipients with 4 different DAA regimens. One hundred percent achieved a sustained viral response by 12 weeks after therapy, and DAA regimens were well tolerated. Recipients transplanted with a HCV+ donor responded equally well to DAA therapy those transplanted with a kidney from an HCV- donor, but recipients of HCV+ organs experienced significantly shorter wait times to transplantation, 485 days (interquartile range, 228-783) versus 969 days (interquartile range, 452-2008; P = 0.02). On this basis, we advocate for a strategy of early posttransplant HCV eradication to facilitate use of HCV+ organs whenever possible. Additional studies are needed to identify the optimal DAA regimen for kidney transplant recipients, accounting for efficacy, timing relative to transplant, posttransplant clinical outcomes, and cost.

  13. Cardiac EASE (Ensuring Access and Speedy Evaluation) – the impact of a single-point-of-entry multidisciplinary outpatient cardiology consultation program on wait times in Canada

    PubMed Central

    Bungard, Tammy J; Smigorowsky, Marcie J; Lalonde, Lucille D; Hogan, Terry; Doliszny, Katharine M; Gebreyesus, Ghirmay; Garg, Sipi; Archer, Stephen L

    2009-01-01

    BACKGROUND: Universal access to health care is valued in Canada but increasing wait times for services (eg, cardiology consultation) raise safety questions. Observations suggest that deficiencies in the process of care contribute to wait times. Consequently, an outpatient clinic was designed for Ensuring Access and Speedy Evaluation (Cardiac EASE) in a university group practice, providing cardiac consultative services for northern Alberta. Cardiac EASE has two components: a single-point-of-entry intake service (prospective testing using physician-approved algorithms and previsit triage) and a multidisciplinary clinic (staffed by cardiologists, nurse practitioners and doctoral-trained pharmacists). OBJECTIVES: It was hypothesized that Cardiac EASE would reduce the time to initial consultation and a definitive diagnosis, and also increase the referral capacity. METHODS: The primary and secondary outcomes were time from referral to initial consultation, and time to achieve a definitive diagnosis and management plan, respectively. A conventionally managed historical control group (three-month pre-EASE period in 2003) was compared with the EASE group (2004 to 2006). The conventional referral mechanism continued concurrently with EASE. RESULTS: A comparison between pre-EASE (n=311) and EASE (n=3096) revealed no difference in the mean (± SD) age (60±16 years), sex (55% and 52% men, respectively) or reason for referral, including chest pain (31% and 40%, respectively) and arrhythmia (27% and 29%, respectively). Cardiac EASE reduced the time to initial cardiac consultation (from 71±45 days to 33±19 days) and time to a definitive diagnosis (from 120±86 days to 51±58 days) (P<0.0001). The annual number of new referrals increased from 1512 in 2002 to 2574 in 2006 due to growth in the Cardiac EASE clinic. The number of patients seen through the conventional referral mechanism and their wait times remained constant during the study period. CONCLUSIONS: Cardiac EASE reduced

  14. Handgun waiting periods reduce gun deaths.

    PubMed

    Luca, Michael; Malhotra, Deepak; Poliquin, Christopher

    2017-11-14

    Handgun waiting periods are laws that impose a delay between the initiation of a purchase and final acquisition of a firearm. We show that waiting periods, which create a "cooling off" period among buyers, significantly reduce the incidence of gun violence. We estimate the impact of waiting periods on gun deaths, exploiting all changes to state-level policies in the Unites States since 1970. We find that waiting periods reduce gun homicides by roughly 17%. We provide further support for the causal impact of waiting periods on homicides by exploiting a natural experiment resulting from a federal law in 1994 that imposed a temporary waiting period on a subset of states. Copyright © 2017 the Author(s). Published by PNAS.

  15. Poster - 26: Electronic Waiting Room Management for a busy Cancer Centre

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

    Kildea, John; Hijal, Tarek

    We describe an electronic waiting room management system that we have developed and deployed in our cancer centre. Our system connects with our electronic medical records systems, gathers data for a machine learning algorithm to predict future patient waiting times, and is integrated with a mobile phone app. The system has been in operation for over nine months and has led to reduced lines, calmer waiting rooms and overwhelming patient and staff satisfaction.

  16. ED adds business center to wait area.

    PubMed

    2007-10-01

    Providing your patients with Internet access in the waiting area can do wonders for their attitudes and make them much more understanding of long wait times. What's more, it doesn't take a fortune to create a business center. The ED at Florida Hospital Celebration (FL) Health made a world of difference with just a couple of computers and a printer. Have your information technology staff set the computers up to preserve the privacy of your internal computer system, and block out offensive sites. Access to medical sites can help reinforce your patient education efforts.

  17. Reward Sensitivity and Waiting Impulsivity: Shift towards Reward Valuation away from Action Control

    PubMed Central

    Mechelmans, Daisy J; Strelchuk, Daniela; Doñamayor, Nuria; Banca, Paula; Robbins, Trevor W; Baek, Kwangyeol

    2017-01-01

    Abstract Background Impulsivity and reward expectancy are commonly interrelated. Waiting impulsivity, measured using the rodent 5-Choice Serial Reaction Time task, predicts compulsive cocaine seeking and sign (or cue) tracking. Here, we assess human waiting impulsivity using a novel translational task, the 4-Choice Serial Reaction Time task, and the relationship with reward cues. Methods Healthy volunteers (n=29) performed the monetary incentive delay task as a functional MRI study where subjects observe a cue predicting reward (cue) and wait to respond for high (£5), low (£1), or no reward. Waiting impulsivity was tested with the 4-Choice Serial Reaction Time task. Results For high reward prospects (£5, no reward), greater waiting impulsivity on the 4-CSRT correlated with greater medial orbitofrontal cortex and lower supplementary motor area activity to cues. In response to high reward cues, greater waiting impulsivity was associated with greater subthalamic nucleus connectivity with orbitofrontal cortex and greater subgenual cingulate connectivity with anterior insula, but decreased connectivity with regions implicated in action selection and preparation. Conclusion These findings highlight a shift towards regions implicated in reward valuation and a shift towards compulsivity away from higher level motor preparation and action selection and response. We highlight the role of reward sensitivity and impulsivity, mechanisms potentially linking human waiting impulsivity with incentive approach and compulsivity, theories highly relevant to disorders of addiction. PMID:29020291

  18. How Long Are Cancer Patients Waiting for Oncological Therapy in Poland?

    PubMed

    Osowiecka, Karolina; Rucinska, Monika; Nowakowski, Jacek J; Nawrocki, Sergiusz

    2018-03-23

    The five-year relative survival rate in Poland is approximately 10% lower compared with the average for Europe. One of the factors that may contribute to the inferior treatment results in Poland could be the long time between cancer suspicion and the beginning of treatment. The aim of the study was to determine the real waiting time for cancer diagnosis and treatment in Poland. The study was carried out in six cancer centers on a group of 1373 patients, using a questionnaire to interview patients. The median waiting time was estimated as follows: (A) from suspicion (the date of the first visit, with symptoms, to a doctor or a preventive or screening test) until histopathological diagnosis; (B) from suspicion until initial treatment; and (C) from diagnosis until initial treatment. The median times from suspicion to treatment, from suspicion to diagnosis, and from diagnosis to treatment, were 10.6, 5.6, and 5.0 weeks, respectively. Using multivariate analysis, the strongest influence was estimated, in a case of tumor localization, to be the method of initial treatment and facilities. The waiting time for cancer treatment in Poland is too long. The highest influence on waiting time was determined, in the case of tumors, as the type of cancer and factors related to the health care system.

  19. Applying the Lean principles of the Toyota Production System to reduce wait times in the emergency department.

    PubMed

    Ng, David; Vail, Gord; Thomas, Sophia; Schmidt, Nicki

    2010-01-01

    In recognition of patient wait times, and deteriorating patient and staff satisfaction, we set out to improve these measures in our emergency department (ED) without adding any new funding or beds. In 2005 all staff in the ED at Hôtel-Dieu Grace Hospital began a transformation, employing Toyota Lean manufacturing principles to improve ED wait times and quality of care. Lean techniques such as value-stream mapping, just-in-time delivery techniques, workplace organization, reduction of systemic wastes, use of the worker as the source of quality improvement and ongoing refinement of our process steps formed the basis of our project. Our ED has achieved major improvements in departmental flow without adding any additional ED or inpatient beds. The mean registration to physician time has decreased from 111 minutes to 78 minutes. The number of patients who left without being seen has decreased from 7.1% to 4.3%. The length of stay (LOS) for discharged patients has decreased from a mean of 3.6 to 2.8 hours, with the largest decrease seen in our patients triaged at levels 4 or 5 using the Canadian Emergency Department Triage and Acuity Scale. We noted an improvement in ED patient satisfaction scores following the implementation of Lean principles. Lean manufacturing principles can improve the flow of patients through the ED, resulting in greater patient satisfaction along with reduced time spent by the patient in the ED.

  20. Lean-driven improvements slash wait times, drive up patient satisfaction scores.

    PubMed

    2012-07-01

    Administrators at LifePoint Hospitals, based in Brentwood, TN, used lean manufacturing techniques to slash wait times by as much as 30 minutes and achieve double-digit increases in patient satisfaction scores in the EDs at three hospitals. In each case, front-line workers took the lead on identifying opportunities for improvement and redesigning the patient-flow process. As a result of the new efficiencies, patient volume is up by about 25% at all three hospitals. At each hospital, the improvement process began with Kaizen, a lean process that involves bringing personnel together to flow-chart the current system, identify problem areas, and redesign the process. Improvement teams found big opportunities for improvement at the front end of the flow process. Key to the approach was having a plan up front to deal with non-compliance. To sustain improvements, administrators gather and disseminate key metrics on a daily basis.

  1. Spin-resolved electron waiting times in a quantum-dot spin valve

    NASA Astrophysics Data System (ADS)

    Tang, Gaomin; Xu, Fuming; Mi, Shuo; Wang, Jian

    2018-04-01

    We study the electronic waiting-time distributions (WTDs) in a noninteracting quantum-dot spin valve by varying spin polarization and the noncollinear angle between the magnetizations of the leads using the scattering matrix approach. Since the quantum-dot spin valve involves two channels (spin up and down) in both the incoming and outgoing channels, we study three different kinds of WTDs, which are two-channel WTD, spin-resolved single-channel WTD, and cross-channel WTD. We analyze the behaviors of WTDs in short times, correlated with the current behaviors for different spin polarizations and noncollinear angles. Cross-channel WTD reflects the correlation between two spin channels and can be used to characterize the spin-transfer torque process. We study the influence of the earlier detection on the subsequent detection from the perspective of cross-channel WTD, and define the influence degree quantity as the cumulative absolute difference between cross-channel WTDs and first-passage time distributions to quantitatively characterize the spin-flip process. We observe that influence degree versus spin-transfer torque for different noncollinear angles as well as different polarizations collapse into a single curve showing universal behaviors. This demonstrates that cross-channel WTDs can be a pathway to characterize spin correlation in spintronics system.

  2. Development of an Information Model for Kidney Transplant Wait List.

    PubMed

    Bircan, Hüseyin Yüce; Özçelik, Ümit; Uysal, Nida; Demirağ, Alp; Haberal, Mehmet

    2015-11-01

    Deceased-donor kidney transplant is unique among surgical procedures that are an urgent procedure performed in an elective population. It has not been possible to accurately determine when a given patient will be called for transplant. Patients on the active transplant list can be called for a transplant at any time. As a result, every effort must be made to optimize their health according to best practices and published clinical practice guidelines. Once the patient is placed on the transplant wait list after undergoing an initial extensive evaluation, continued surveillance is required. Therefore, we developed a kidney transplant wait list surveillance software program that alerts organ transplant coordinator on time regarding which patients need a work-up. The new designed software has a database of our waiting patients with their completed and pending controls. The software also has built-in functions to warn the responsible staff with an E-mail. If one of the controls of a recipient delayed, the software sends an automated E-mail to the staff regarding the patients delayed controls. The software is a Web application that works on any platform with a Web browser and Internet connection and allows access by multiple users. The software has been developed with NET platform. The database is SQL server. The software has the following functions: patient communication info, search, alert list, alert E-mail, control entry, and system management. As of January 2014, a total of 21 000 patients were registered on the National Kidney Transplant wait list in Turkey and the kidney transplant wait list had been expanding by 2000 to 3000 patients each year. Therefore computerized wait list programs are crucial to help to transplant centers to keep their patients up-to-date on time.

  3. Reducing Patient Waiting Times for Radiation Therapy and Improving the Treatment Planning Process: a Discrete-event Simulation Model (Radiation Treatment Planning).

    PubMed

    Babashov, V; Aivas, I; Begen, M A; Cao, J Q; Rodrigues, G; D'Souza, D; Lock, M; Zaric, G S

    2017-06-01

    We analysed the radiotherapy planning process at the London Regional Cancer Program to determine the bottlenecks and to quantify the effect of specific resource levels with the goal of reducing waiting times. We developed a discrete-event simulation model of a patient's journey from the point of referral to a radiation oncologist to the start of radiotherapy, considering the sequential steps and resources of the treatment planning process. We measured the effect of several resource changes on the ready-to-treat to treatment (RTTT) waiting time and on the percentage treated within a 14 calendar day target. Increasing the number of dosimetrists by one reduced the mean RTTT by 6.55%, leading to 84.92% of patients being treated within the 14 calendar day target. Adding one more oncologist decreased the mean RTTT from 10.83 to 10.55 days, whereas a 15% increase in arriving patients increased the waiting time by 22.53%. The model was relatively robust to the changes in quantity of other resources. Our model identified sensitive and non-sensitive system parameters. A similar approach could be applied by other cancer programmes, using their respective data and individualised adjustments, which may be beneficial in making the most effective use of limited resources. Copyright © 2017 The Royal College of Radiologists. Published by Elsevier Ltd. All rights reserved.

  4. Prototype of a Questionnaire and Quiz System for Supporting Increase of Health Awareness During Wait Time in Dispensing Pharmacy

    NASA Astrophysics Data System (ADS)

    Toda, Takeshi; Chen, Poa-Min; Ozaki, Shinya; Ideguchi, Naoko; Miyaki, Tomoko; Nanbu, Keiko; Ikeda, Keiko

    For quit-smoking clinic and its campaign, there was a need for pharmacists to investigate pediatric patient's parent consciousness to tobacco harm utilizing wait time in a pediatric dispensing pharmacy. In this research, we developed the questionnaire and quiz total system using the tablet for user interface, in which people can easily answer the questionnaire/quiz and quickly see the total results on the spot in order to enhance their consciousness to the tobacco harm. The system also provides their tobacco dependence level based on the questionnaire results and some advice for their health and dietary habits due to the tobacco dependence level. From a field trial with one hundred four examinees in the pediatric dispensing pharmacy, the user interface was useful compared to conventional questionnaire form. The system could enhance their consciousness to tobacco harm and make their beneficial use of waiting time in dispensing pharmacy. Some interesting suggestions for improvement and new services were also obtained.

  5. Spatial structure increases the waiting time for cancer

    PubMed Central

    Martens, Erik A.; Kostadinov, Rumen; Maley, Carlo C.; Hallatschek, Oskar

    2012-01-01

    Cancer results from a sequence of genetic and epigenetic changes which lead to a variety of abnormal phenotypes including increased proliferation and survival of somatic cells, and thus, to a selective advantage of pre-cancerous cells. The notion of cancer progression as an evolutionary process has been experiencing increasing interest in recent years. Many efforts have been made to better understand and predict the progression to cancer using mathematical models; these mostly consider the evolution of a well-mixed cell population, even though pre-cancerous cells often evolve in highly structured epithelial tissues. In this study, we propose a novel model of cancer progression that considers a spatially structured cell population where clones expand via adaptive waves. This model is used to assess two different paradigms of asexual evolution that have been suggested to delineate the process of cancer progression. The standard scenario of periodic selection assumes that driver mutations are accumulated strictly sequentially over time. However, when the mutation supply is sufficiently high, clones may arise simultaneously on distinct genetic backgrounds, and clonal adaptation waves interfere with each other. We find that in the presence of clonal interference, spatial structure increases the waiting time for cancer, leads to a patchwork structure of non-uniformly sized clones, decreases the survival probability of virtually neutral (passenger) mutations, and that genetic distance begins to increase over a characteristic length scale Lc. These characteristic features of clonal interference may help to predict the onset of cancers with pronounced spatial structure and to interpret spatially-sampled genetic data obtained from biopsies. Our estimates suggest that clonal interference likely occurs in the progression of colon cancer, and possibly other cancers where spatial structure matters. PMID:22707911

  6. Spatial structure increases the waiting time for cancer

    NASA Astrophysics Data System (ADS)

    Martens, Erik A.; Kostadinov, Rumen; Maley, Carlo C.; Hallatschek, Oskar

    2011-11-01

    Cancer results from a sequence of genetic and epigenetic changes that lead to a variety of abnormal phenotypes including increased proliferation and survival of somatic cells and thus to a selective advantage of pre-cancerous cells. The notion of cancer progression as an evolutionary process has been attracting increasing interest in recent years. A great deal of effort has been made to better understand and predict the progression to cancer using mathematical models; these mostly consider the evolution of a well-mixed cell population, even though pre-cancerous cells often evolve in highly structured epithelial tissues. In this study, we propose a novel model of cancer progression that considers a spatially structured cell population where clones expand via adaptive waves. This model is used to assess two different paradigms of asexual evolution that have been suggested to delineate the process of cancer progression. The standard scenario of periodic selection assumes that driver mutations are accumulated strictly sequentially over time. However, when the mutation supply is sufficiently high, clones may arise simultaneously on distinct genetic backgrounds, and clonal adaptation waves interfere with each other. We find that in the presence of clonal interference, spatial structure increases the waiting time for cancer, leads to a patchwork structure of non-uniformly sized clones and decreases the survival probability of virtually neutral (passenger) mutations, and that genetic distance begins to increase over a characteristic length scale Lc. These characteristic features of clonal interference may help us to predict the onset of cancers with pronounced spatial structure and to interpret spatially sampled genetic data obtained from biopsies. Our estimates suggest that clonal interference likely occurs in the progression of colon cancer and possibly other cancers where spatial structure matters.

  7. A comparison of control modes for time-delayed remote manipulation

    NASA Technical Reports Server (NTRS)

    Starr, G. P.

    1982-01-01

    Transmission time delay in the communication channel of a manual control system is investigated. A time delay can exist in remote manipulation systems, caused by long communication distances or bandwidth limitations. Ferrell 1 conducted the first research in time-delayed manipulation using a two degree-of-freedom manipulator. His subjects, working at time delays of 1.0, 2.1, and 3.2 s, could accomplish tasks even requiring great accuracy. The subjects spontaneously adopted a pattern of moving cautiously, then waiting to see the results of their actions. In experiments with a six degree-of-freedom master-slave manipulator system and time delays of 1.0 to 6 s, Black 2 saw that subjects tried to use the move-and-wait strategy; but there were often difficulties. The subjects seemed to have a problem in holding the master arm stationary while waiting for feedback. Any undesired drifting of the master arm introduced a discrepancy between the positions of the master and slave. This discrepancy was not perceived because of the time delay. The subject would then begin his next move with an inherent error. The difficulty of effectively using the move-and-wait strategy with a master-slave manipulator suggested that rate control might be a more effective control mode with time delay.

  8. Optimizing the arrival, waiting, and NPO times of children on the day of pediatric endoscopy procedures.

    PubMed

    Smallman, Bettina; Dexter, Franklin

    2010-03-01

    Research in predictive variability of operating room (OR) times has been performed using data from multidisciplinary, tertiary hospitals with mostly adult patients. In this article, we discuss case-duration prediction for children receiving general anesthesia for endoscopy. We critique which of the several types of OR management decisions dependent on accuracy of prediction are relevant to series (lists) of brief pediatric anesthetics. OR information system data were obtained for all children (aged 18 years and younger) undergoing a gastroenterology procedure with an anesthesiologist over 21 months. Summaries of data were used for a qualitative, systematic review of prior studies to learn which apply to brief pediatric cases. Patient arrival times were changed to be based on the statistical method relating actual and scheduled start times (Wachtel and Dexter, Anesth Analg 2007;105:127-40). Even perfect case-duration prediction would not affect whether a brief case was performed on a certain date and/or in a certain OR. There was no evidence of usefulness in calculating the probability that one case would last longer than another or in resequencing cases to influence postanesthesia care unit staffing or patient waiting from scheduled start times. The only decision for which the accuracy of case-duration prediction mattered was for the shortest time that preceding cases in the OR may take. Knowledge of the preceding procedures in the OR was not useful for that purpose because there were hundreds of combinations of preceding procedures and some cases cancelled. Instead, patient ready times were chosen based on 5% lower prediction bounds for ratios of actual to scheduled OR times. The approach was useful based on a 30% reduction in patient waiting times from scheduled start times with corresponding expected reductions in average and peak numbers of patients in the holding area. For brief pediatric OR anesthetics, predictive variability of case durations matters

  9. Protocol to Exploit Waiting Resources for UASNs.

    PubMed

    Hung, Li-Ling; Luo, Yung-Jeng

    2016-03-08

    The transmission speed of acoustic waves in water is much slower than that of radio waves in terrestrial wireless sensor networks. Thus, the propagation delay in underwater acoustic sensor networks (UASN) is much greater. Longer propagation delay leads to complicated communication and collision problems. To solve collision problems, some studies have proposed waiting mechanisms; however, long waiting mechanisms result in low bandwidth utilization. To improve throughput, this study proposes a slotted medium access control protocol to enhance bandwidth utilization in UASNs. The proposed mechanism increases communication by exploiting temporal and spatial resources that are typically idle in order to protect communication against interference. By reducing wait time, network performance and energy consumption can be improved. A performance evaluation demonstrates that when the data packets are large or sensor deployment is dense, the energy consumption of proposed protocol is less than that of existing protocols as well as the throughput is higher than that of existing protocols.

  10. Weight-ing: the experience of waiting on weight loss.

    PubMed

    Glenn, Nicole M

    2013-03-01

    Perhaps we want to be perfect, strive for health, beauty, and the admiring gaze of others. Maybe we desire the body of our youth, the "healthy" body, the body that has just the right fit. Regardless of the motivation, we might find ourselves striving, wanting, and waiting on weight loss. What is it to wait on weight loss? I explore the meaning of this experience-as-lived using van Manen's guide to phenomenological reflection and writing. Weight has become an increasing focus of contemporary culture, demonstrated, for example, by a growing weight-loss industry and global obesity "epidemic." Weight has become synonymous with health status, and weight loss with "healthier." I examine the weight wait through experiences of the common and uncommon, considering relations to time, body, space, and the other with the aim of evoking a felt, embodied, emotive understanding of the meaning of waiting on weight loss. I also discuss the implications of the findings.

  11. Using lean manufacturing principles to evaluate wait times for HIV-positive patients in an urban clinic in Kenya.

    PubMed

    Monroe-Wise, Aliza; Reisner, Elizabeth; Sherr, Kenneth; Ojakaa, David; Mbau, Lilian; Kisia, Paul; Muhula, Samuel; Farquhar, Carey

    2017-12-01

    As human immunodeficiency virus (HIV) treatment programs expand in Africa, delivery systems must be strengthened to support patient retention. Clinic characteristics may affect retention, but a relationship between clinic flow and attrition is not established. This project characterized HIV patient experience and flow in an urban Kenyan clinic to understand how these may affect retention. We used Toyota's lean manufacturing principles to guide data collection and analysis. Clinic flow was evaluated using value stream mapping and time and motion techniques. Clinic register data were analyzed. Two focus group discussions were held to characterize HIV patient experience. Results were shared with clinic staff. Wait times in the clinic were highly variable. We identified four main barriers to patient flow: inconsistent patient arrivals, inconsistent staffing, filing system defects, and serving patients out of order. Focus group participants explained how clinic operations affected their ability to engage in care. Clinic staff were eager to discuss the problems identified and identified numerous low-cost potential solutions. Lean manufacturing methodologies can guide efficiency interventions in low-resource healthcare settings. Using lean techniques, we identified bottlenecks to clinic flow and low-cost solutions to improve wait times. Improving flow may result in increased patient satisfaction and retention.

  12. Evaluating wait times from screening to breast cancer diagnosis among women undergoing organised assessment vs usual care

    PubMed Central

    Chiarelli, Anna M; Muradali, Derek; Blackmore, Kristina M; Smith, Courtney R; Mirea, Lucia; Majpruz, Vicky; O'Malley, Frances P; Quan, May Lynn; Holloway, Claire MB

    2017-01-01

    Background: Timely coordinated diagnostic assessment following an abnormal screening mammogram reduces patient anxiety and may optimise breast cancer prognosis. Since 1998, the Ontario Breast Screening Program (OBSP) has offered organised assessment through Breast Assessment Centres (BACs). For OBSP women seen at a BAC, an abnormal mammogram is followed by coordinated referrals through the use of navigators for further imaging, biopsy, and surgical consultation as indicated. For OBSP women seen through usual care (UC), further diagnostic imaging is arranged directly from the screening centre and/or through their physician; results must be communicated to the physician who is then responsible for arranging any necessary biopsy and/or surgical consultation. This study aims to evaluate factors associated with diagnostic wait times for women undergoing assessment through BAC and UC. Methods: Of the 2 147 257 women aged 50–69 years screened in the OBSP between 1 January 2002 and 31 December 2009, 155 866 (7.3%) had an abnormal mammogram. A retrospective design identified two concurrent cohorts of women diagnosed with screen-detected breast cancer at a BAC (n=4217; 47%) and UC (n=4827; 53%). Multivariable logistic regression analyses examined associations between wait times and assessment and prognostic characteristics by pathway. A two-sided 5% significance level was used. Results: Screened women with breast cancer were two times more likely to be diagnosed within 7 weeks when assessed through a BAC vs UC (OR=1.91, 95% CI=1.73–2.10). In addition, compared with UC, women assessed through a BAC were significantly more likely to have their first assessment procedure within 3 weeks of their abnormal mammogram (OR=1.25, 95% CI=1.12–1.39), ⩽3 assessment procedures (OR=1.54, 95% CI=1.41–1.69), ⩽2 assessment visits (OR=1.86, 95% CI=1.70–2.05), and ⩾2 procedures per visit (OR=1.41, 95% CI=1.28–1.55). Women diagnosed through a BAC were also more likely

  13. Evaluating wait times from screening to breast cancer diagnosis among women undergoing organised assessment vs usual care.

    PubMed

    Chiarelli, Anna M; Muradali, Derek; Blackmore, Kristina M; Smith, Courtney R; Mirea, Lucia; Majpruz, Vicky; O'Malley, Frances P; Quan, May Lynn; Holloway, Claire Mb

    2017-05-09

    Timely coordinated diagnostic assessment following an abnormal screening mammogram reduces patient anxiety and may optimise breast cancer prognosis. Since 1998, the Ontario Breast Screening Program (OBSP) has offered organised assessment through Breast Assessment Centres (BACs). For OBSP women seen at a BAC, an abnormal mammogram is followed by coordinated referrals through the use of navigators for further imaging, biopsy, and surgical consultation as indicated. For OBSP women seen through usual care (UC), further diagnostic imaging is arranged directly from the screening centre and/or through their physician; results must be communicated to the physician who is then responsible for arranging any necessary biopsy and/or surgical consultation. This study aims to evaluate factors associated with diagnostic wait times for women undergoing assessment through BAC and UC. Of the 2 147 257 women aged 50-69 years screened in the OBSP between 1 January 2002 and 31 December 2009, 155 866 (7.3%) had an abnormal mammogram. A retrospective design identified two concurrent cohorts of women diagnosed with screen-detected breast cancer at a BAC (n=4217; 47%) and UC (n=4827; 53%). Multivariable logistic regression analyses examined associations between wait times and assessment and prognostic characteristics by pathway. A two-sided 5% significance level was used. Screened women with breast cancer were two times more likely to be diagnosed within 7 weeks when assessed through a BAC vs UC (OR=1.91, 95% CI=1.73-2.10). In addition, compared with UC, women assessed through a BAC were significantly more likely to have their first assessment procedure within 3 weeks of their abnormal mammogram (OR=1.25, 95% CI=1.12-1.39), ⩽3 assessment procedures (OR=1.54, 95% CI=1.41-1.69), ⩽2 assessment visits (OR=1.86, 95% CI=1.70-2.05), and ⩾2 procedures per visit (OR=1.41, 95% CI=1.28-1.55). Women diagnosed through a BAC were also more likely than those in UC to have imaging (OR=1.99, 95

  14. The influence of parametric and external noise in act-and-wait control with delayed feedback.

    PubMed

    Wang, Jiaxing; Kuske, Rachel

    2017-11-01

    We apply several novel semi-analytic approaches for characterizing and calculating the effects of noise in a system with act-and-wait control. For concrete illustration, we apply these to a canonical balance model for an inverted pendulum to study the combined effect of delay and noise within the act-and-wait setting. While the act-and-wait control facilitates strong stabilization through deadbeat control, a comparison of different models with continuous vs. discrete updating of the control strategy in the active period illustrates how delays combined with the imprecise application of the control can seriously degrade the performance. We give several novel analyses of a generalized act-and-wait control strategy, allowing flexibility in the updating of the control strategy, in order to understand the sensitivities to delays and random fluctuations. In both the deterministic and stochastic settings, we give analytical and semi-analytical results that characterize and quantify the dynamics of the system. These results include the size and shape of stability regions, densities for the critical eigenvalues that capture the rate of reaching the desired stable equilibrium, and amplification factors for sustained fluctuations in the context of external noise. They also provide the dependence of these quantities on the length of the delay and the active period. In particular, we see that the combined influence of delay, parametric error, or external noise and on-off control can qualitatively change the dynamics, thus reducing the robustness of the control strategy. We also capture the dependence on how frequently the control is updated, allowing an interpolation between continuous and frequent updating. In addition to providing insights for these specific models, the methods we propose are generalizable to other settings with noise, delay, and on-off control, where analytical techniques are otherwise severely scarce.

  15. Anxiety prior to breast biopsy: Relationships with length of time from breast biopsy recommendation to biopsy procedure and psychosocial factors.

    PubMed

    Hayes Balmadrid, Melissa A; Shelby, Rebecca A; Wren, Anava A; Miller, Lauren S; Yoon, Sora C; Baker, Jay A; Wildermann, Liz A; Soo, Mary Scott

    2017-04-01

    This study investigated how time from breast biopsy recommendation to biopsy procedure affected pre-biopsy anxiety ( N = 140 women), and whether the relationship between wait time and anxiety was affected by psychosocial factors (chronic life stress, traumatic events, social support). Analyses showed a significant interaction between wait time and chronic life stress. Increased time from biopsy recommendation was associated with greater anxiety in women with low levels of life stress. Women with high levels of life stress experienced increased anxiety regardless of wait time. These results suggest that women may benefit from shorter wait times and receiving strategies for managing anxiety.

  16. No More Waits and Delays: Streamlining Workflow to Decrease Patient Time of Stay for Image-guided Musculoskeletal Procedures.

    PubMed

    Cheung, Yvonne Y; Goodman, Eric M; Osunkoya, Tomiwa O

    2016-01-01

    Long wait times limit our ability to provide the right care at the right time and are commonly products of inefficient workflow. In 2013, the demand for musculoskeletal (MSK) procedures increased beyond our department's ability to provide efficient and timely service. We initiated a quality improvement (QI) project to increase efficiency and decrease patient time of stay. Our project team included three MSK radiologists, one senior resident, one technologist, one administrative assistant/scheduler, and the lead technologist. We adopted and followed the Lean Six Sigma DMAIC (define, measure, analyze, improve, and control) approach. The team used tools such as voice of the customer (VOC), along with affinity and SIPOC (supplier, input, process, output, customer) diagrams, to understand the current process, identify our customers, and develop a project charter in the define stage. During the measure stage, the team collected data, created a detailed process map, and identified wastes with the value stream mapping technique. Within the analyze phase, a fishbone diagram helped the team to identify critical root causes for long wait times. Scatter plots revealed relationships among time variables. Team brainstorming sessions generated improvement ideas, and selected ideas were piloted via plan, do, study, act (PDSA) cycles. The control phase continued to enable the team to monitor progress using box plots and scheduled reviews. Our project successfully decreased patient time of stay. The highly structured and logical Lean Six Sigma approach was easy to follow and provided a clear course of action with positive results. (©)RSNA, 2016.

  17. Sustainability of Routine Notification and Request legislation on eye bank tissue supply and corneal transplantation wait times in Canada.

    PubMed

    Lee, Kenneth; Boimer, Corey; Hershenfeld, Samantha; Sharpen, Linda; Slomovic, Allan R

    2011-10-01

    To assess whether provinces with Routine Notification and Request (RNR) legislation have sustained increases in corneal tissue supply and decreases in wait times for corneal transplantation surgery. Cross-sectional survey of Canadian corneal transplant (CT) surgeons and eye banks. Canadian CT surgeons and representatives from the 10 Canadian eye banks. Voluntary and anonymous surveys were distributed between July and October 2009. Eligible CT surgeons were defined as ophthalmologists who practice in Canada; currently perform Penetrating keratoplasty (PKP), Deep anterior lamellar keratoplasty (DALK), Deep lamellar endothelial keratoplasty (DLEK), Descemet stripping endothelial keratoplasty (DSEK), or Descemet membrane endothelial keratoplasty (DMEK); and have obtained tissues from a Canadian eye bank. From 2006 to 2009, for provinces with RNR legislation and where data are available, mean wait times from date of diagnosis to date of CT surgery have increased: in Ontario, from 31 ± 34 weeks to 36 ± 27 weeks; in British Columbia, from 39 ± 20 weeks to 42 ± 35 weeks; in Manitoba, from 32 ± 23 weeks to 49 ± 36 weeks. In addition, the amount of corneal tissue in RNR provinces suitable for transplant, with the exception of British Columbia, has declined between 2006 and 2008: in Ontario, 1186 tissues to 999 tissues (16% decline); in Manitoba, 92 tissues to 83 tissues (10% decline); in New Brunswick, 129 tissues to 98 tissues (24% decline). Although initially effective, RNR legislation has not sustained an increase in corneal tissue availability nor has it shortened wait times in most provinces. Incorporation of community hospitals into the RNR catchment, improved enforcement, and continued education of hospital staff regarding the RNR process may be effective in making this legislation more sustainable in the long term. Copyright © 2011 Canadian Ophthalmological Society. Published by Elsevier Inc. All rights reserved.

  18. The Impact of Patient-to-Patient Interaction in Health Facility Waiting Rooms on Their Perception of Health Professionals.

    PubMed

    Willis, William Kent; Ozturk, Ahmet Ozzie; Chandra, Ashish

    2015-01-01

    Patients have to wait in waiting rooms prior to seeing the physician. But there are few studies that demonstrate what they are actually doing in the waiting room. This exploratory study was designed to investigate the types of discussions that patients in the waiting room typically engage in with other patients and how the conversations affected their opinion on general reputation of the clinic, injections/blocks as treatment procedures, waiting time, time spent with the caregiver, overall patient satisfaction, and the pain medication usage policy. The study demonstrates that patient interaction in the waiting room has a positive effect on patient opinion of the pain clinic and the caregivers.

  19. A Study to Determine Patient Waiting Time at the Outpatient Pharmacy at Wilford Hall USAF Medical Center

    DTIC Science & Technology

    1988-06-01

    Given the very nature of health care facilities, there is significant variablity in patient loads and the types of disorders treated at various...two Department of Defense health care facilities in San Antonio that best compare to WHMC) reflect WHMC’s patient wait time to be quite impressive...Med Center (If applicable) U.S. ARMY-BAYLOR UNIVERSITY GRADUATE Lackland AFB, Texas I SG-3R PIOGRAM IN HEALTH CAPE ADMINISTRATION 6c. ADDRESS (City

  20. Mentoring new nurses in stressful times.

    PubMed

    Young, Lisa E

    2009-06-01

    Meeting benchmarks of Ontario's Wait Time Strategy and the expansion of The Ottawa Hospital are key issues driving the recruitment of perioperative nurses in Ottawa and Eastern Ontario. Added pressures resulting from Canada's aging population and a nationwide nursing shortage mean perioperative nurses are overworked and understaffed. Preceptoring new members of staff raises valid concerns as many of the new recruits have little or no operating room experience. The Dreyfus Model of Skill Acquisition demonstrates the importance of time and patience in supporting the learning process. Mentoring is a valuable strategy in an effort to teach and guide new nurses, to increase nursing retention, and to promote professional growth and recognition. Building successful mentorship programs, through the creation of healthy organizational cultures, transformational leadership and staff development programs, will strengthen support for nurses in stressful times. The stress of meeting the province-wide benchmarks outlined in Ontario's Wait Time Strategy and the expansion of perioperative services at The Ottawa Hospital in Ontario are two key issues driving the need for the recruitment of nurses into the specialty of perioperative nursing. As a result of Canada's aging population and a nationwide nursing shortage, perioperative nurses are over-worked and under-staffed while being faced with the pressure to preceptor new staff members while struggling to meet the daily demands of the wait list strategy. This article discusses current trends in healthcare and the career path changes being made by many nurses in response to the demand for specialty trained nurses. It is followed by a brief explanation of the Dreyfus Model of Skill Acquisition. Mentoring is presented as an effective strategy in the guidance and teaching of new nurses with a discussion of the benefits and suggestions on how to build a successful mentorship program to support nurses in these stressful times.

  1. Bracing Later and Coping Better: Benefits of Mindfulness During a Stressful Waiting Period.

    PubMed

    Sweeny, Kate; Howell, Jennifer L

    2017-10-01

    People frequently await uncertain news, yet research reveals that the strategies people naturally use to cope with uncertainty are largely ineffective. We tested the role of mindfulness for improving the experience of a stressful waiting period. Law graduates awaiting their bar exam results either reported their trait mindfulness (Study 1; N = 150) or were instructed to practice mindfulness meditation (Study 2; N = 90). As hypothesized, participants who were naturally more mindful or who practiced mindfulness managed their expectations more effectively by bracing for the worst later in the waiting period and perceived themselves as coping better. Additionally, participants who were low in dispositional optimism and high in intolerance of uncertainty benefited most from mindfulness (relative to control) meditation. These findings point to a simple and effective way to wait better, particularly for those most vulnerable to distress.

  2. A comparative analysis of centralized waiting lists for patients without a primary care provider implemented in six Canadian provinces: study protocol.

    PubMed

    Breton, Mylaine; Green, Michael; Kreindler, Sara; Sutherland, Jason; Jbilou, Jalila; Wong, Sabrina T; Shaw, Jay; Crooks, Valorie A; Contandriopoulos, Damien; Smithman, Mélanie Ann; Brousselle, Astrid

    2017-01-21

    -provincial learning exchange approach where we propose to compare centralized waiting lists and analyze variations in strategies used to increase attachment to a regular primary care provider. Fostering inter-provincial healthcare systems connectivity to improve centralized waiting lists' practices across Canada can lever attachment to a regular provider for timely access to continuous, comprehensive and coordinated healthcare for all Canadians and particular for those who are vulnerable.

  3. The influence of question type, reasoning level, and wait time on student participation rates when using clicker questions with large classes

    NASA Astrophysics Data System (ADS)

    Hartman, K.; Koh, J.; Murty, S. A.; Ramos, R. D. P.; Goodkin, N.

    2017-12-01

    "Wait time" is defined as the length of the pause between an instructor initiating a question and either the student answering it or the instructor interjecting (Rowe, 1974). However, the nature of the question-answer dynamic changes with student response systems that allow hundreds of students to answer the same question simultaneously before displaying the results to the class. In this study, we looked at 129 student response questions asked across 240 minutes of class lectures to determine if longer wait times were associated with higher student response rates. We also examined whether the type and reasoning level of the questions were diagnostic of their response rates. 644 undergraduate science students enrolled in an interdisciplinary environmental science course. During each of the course's lessons, the instructor presented a mix of lecture content, short student response activities (clicker questions), and small group discussion opportunities. Using the recorded videos, we coded each student response question for its question type and reasoning level. We divided the question types into three categories: yes/no questions, yes/no/maybe questions, and other questions. To code for the reasoning level necessary to answer each question, we used a collapsed version of Bloom's Revised Taxonomy (Krathwohl, 2002). Questions that had a definite answer and relied on recalling facts or paraphrasing the lecture content were coded as "knowledge" questions. Questions that required students to apply what they had learned to analyze a new scenario, or come to a judgement were coded as "higher order" questions. An analysis of variance (ANOVA) using the question type and reasoning level as fixed factors and wait time as a covariate to predict student response rate indicated a strong interaction between question type and reasoning level F(6, 94) = 4.53, p<.01. In general, knowledge questions were answered by a higher percentage (M=91%) of students than higher order questions (M

  4. Impact of visual art on patient behavior in the emergency department waiting room.

    PubMed

    Nanda, Upali; Chanaud, Cheryl; Nelson, Michael; Zhu, Xi; Bajema, Robyn; Jansen, Ben H

    2012-07-01

    Wait times have been reported to be one of the most important concerns for people visiting emergency departments (EDs). Affective states significantly impact perception of wait time. There is substantial evidence that art depicting nature reduces stress levels and anxiety, thus potentially impacting the waiting experience. To analyze the effect of visual art depicting nature (still and video) on patients' and visitors' behavior in the ED. A pre-post research design was implemented using systematic behavioral observation of patients and visitors in the ED waiting rooms of two hospitals over a period of 4 months. Thirty hours of data were collected before and after new still and video art was installed at each site. Significant reduction in restlessness, noise level, and people staring at other people in the room was found at both sites. A significant decrease in the number of queries made at the front desk and a significant increase in social interaction were found at one of the sites. Visual art has positive effects on the ED waiting experience. Copyright © 2012 Elsevier Inc. All rights reserved.

  5. The impact of waiting for intervention on costs and effectiveness: the case of transcatheter aortic valve replacement.

    PubMed

    Ribera, Aida; Slof, John; Ferreira-González, Ignacio; Serra, Vicente; García-Del Blanco, Bruno; Cascant, Purificació; Andrea, Rut; Falces, Carlos; Gutiérrez, Enrique; Del Valle-Fernández, Raquel; Morís-de laTassa, César; Mota, Pedro; Oteo, Juan Francisco; Tornos, Pilar; García-Dorado, David

    2017-11-23

    The economic crisis in Europe might have limited access to some innovative technologies implying an increase of waiting time. The purpose of the study is to evaluate the impact of waiting time on the costs and benefits of transcatheter aortic valve replacement (TAVR) for the treatment of severe aortic stenosis. This is a cost-utility analysis from the perspective of the Spanish National Health Service. Results of two prospective hospital registries (158 and 273 consecutive patients) were incorporated into a probabilistic Markov model to compare quality adjusted life years (QALYs) and costs for TAVR after waiting for 3-12 months, relative to immediate TAVR. We simulated a cohort of 1000 patients, male, and 80 years old; other patient profiles were assessed in sensitivity analyses. As waiting time increased, costs decreased at the expense of lower survival and loss of QALYs, leading to incremental cost-effectiveness ratios for eliminating waiting lists of about 12,500 € per QALY. In subgroup analyses prioritization of patients for whom higher benefit was expected led to a smaller loss of QALYs. Concerning budget impact, long waiting lists reduced spending considerably and permanently. A shorter waiting time is likely to be cost-effective (considering commonly accepted willingness-to-pay thresholds in Europe) relative to 3 months or longer waiting periods. If waiting lists are nevertheless seen as unavoidable due to severe but temporary budgetary restrictions, prioritizing patients for whom higher benefit is expected appears to be a way of postponing spending without utterly sacrificing patients' survival and quality of life.

  6. Analysis of bluetooth and wi-fi technology to measure wait times of personal vehicles at Arizona-Mexico ports of entry : [executive summary].

    DOT National Transportation Integrated Search

    2015-11-01

    The Arizona Department of Transportation (ADOT), Office of P3 Initiatives and International : Affairs selected Lee Engineering to analyze the penetration rate of Anonymous Re-Identification : (ARID) technology to measure wait time of U.S. and Mexico ...

  7. Optimizing the admission time of outbound trucks entering a cross-dock with uniform arrival time by considering a queuing model

    NASA Astrophysics Data System (ADS)

    Motaghedi-Larijani, Arash; Aminnayeri, Majid

    2017-03-01

    Cross-docking is a supply-chain strategy that can reduce transportation and inventory costs. This study is motivated by a fruit and vegetable distribution centre in Tehran, which has cross-docks and a limited time to admit outbound trucks. In this article, outbound trucks are assumed to arrive at the cross-dock with a single outbound door with a uniform distribution (0,L). The total number of assigned trucks is constant and the loading time is fixed. A queuing model is modified for this situation and the expected waiting time of each customer is calculated. Then, a curve for the waiting time is calculated. Finally, the length of window time L is optimized to minimize the total cost, which includes the waiting time of the trucks and the admission cost of the cross-dock. Some illustrative examples of cross-docking are presented and solved using the proposed method.

  8. Determining prescription durations based on the parametric waiting time distribution.

    PubMed

    Støvring, Henrik; Pottegård, Anton; Hallas, Jesper

    2016-12-01

    The purpose of the study is to develop a method to estimate the duration of single prescriptions in pharmacoepidemiological studies when the single prescription duration is not available. We developed an estimation algorithm based on maximum likelihood estimation of a parametric two-component mixture model for the waiting time distribution (WTD). The distribution component for prevalent users estimates the forward recurrence density (FRD), which is related to the distribution of time between subsequent prescription redemptions, the inter-arrival density (IAD), for users in continued treatment. We exploited this to estimate percentiles of the IAD by inversion of the estimated FRD and defined the duration of a prescription as the time within which 80% of current users will have presented themselves again. Statistical properties were examined in simulation studies, and the method was applied to empirical data for four model drugs: non-steroidal anti-inflammatory drugs (NSAIDs), warfarin, bendroflumethiazide, and levothyroxine. Simulation studies found negligible bias when the data-generating model for the IAD coincided with the FRD used in the WTD estimation (Log-Normal). When the IAD consisted of a mixture of two Log-Normal distributions, but was analyzed with a single Log-Normal distribution, relative bias did not exceed 9%. Using a Log-Normal FRD, we estimated prescription durations of 117, 91, 137, and 118 days for NSAIDs, warfarin, bendroflumethiazide, and levothyroxine, respectively. Similar results were found with a Weibull FRD. The algorithm allows valid estimation of single prescription durations, especially when the WTD reliably separates current users from incident users, and may replace ad-hoc decision rules in automated implementations. Copyright © 2016 John Wiley & Sons, Ltd. Copyright © 2016 John Wiley & Sons, Ltd.

  9. Strategic Attention Deployment for Delay of Gratification in Working and Waiting Situations.

    ERIC Educational Resources Information Center

    Peake, Philip K.; Hebl, Michelle; Mischel, Walter

    2002-01-01

    Two studies examined whether effects of attention to rewards during a delay of gratification task in waiting situations affects preschoolers' ability to delay gratification in working situations. Findings show that when work provides distraction, attention on rewards reduces delay time whether working or waiting; when work is not engaging,…

  10. Watchful waiting and factors predictive of secondary treatment of localized prostate cancer.

    PubMed

    Wu, Hongyan; Sun, Leon; Moul, Judd W; Wu, Hong Yu; McLeod, David G; Amling, Christopher; Lance, Raymond; Kusuda, Leo; Donahue, Timothy; Foley, John; Chung, Andrew; Sexton, Wade; Soderdahl, Douglas

    2004-03-01

    Watchful waiting remains an important treatment option for some patients with localized prostate cancer. We defined the demographic, clinical and outcome features of men selecting watchful waiting as an initial treatment strategy, and determined factors predictive of eventual progression to secondary treatment. Of 8390 patients diagnosed with prostate cancer from 1990 to 2001 in the Department of Defense Center for Prostate Disease Research Database, 1158 patients chose watchful waiting as initial treatment. The demographic and clinical differences between patients on watchful waiting and those choosing other initial treatments were compared using the chi-square test. Secondary treatment-free survival according to various prognostic factors was plotted using the Kaplan-Meier method and differences were tested using the log rank test. A multivariate Cox proportional hazards regression analysis was performed to determine which factors were independent predictors of secondary treatment. Compared to other patients, those selecting watchful waiting were older, had lower prostate specific antigen (PSA) at diagnosis, and were more likely to have lower stage (cT1) and lower grade (Gleason sum 7 or less) cancers. Age, PSA and clinical stage were all significant and independent predictors of secondary treatment. The relative risk of secondary treatment can be expressed as EXP (-0.034 x age at diagnosis + 0.284 x LOG (diagnostic PSA) + 0.271 x clinical stage T2 + 0.264 x clinical stage T3). Men who elect watchful waiting as initial management for prostate cancer are older with lower Gleason sums and serum PSA. In these men, age at diagnosis, serum PSA and clinical stage are the most significant predictors of requiring or selecting secondary treatment.

  11. Reproductive strategies of northern geese: Why wait?

    USGS Publications Warehouse

    Ely, Craig R.; Bollinger, K.S.; Densmore, R.V.; Rothe, T.C.; Petrula, M.J.; Takekawa, John Y.; Orthmeyer, D.L.

    2007-01-01

    Migration and reproductive strategies in waterbirds are tightly linked, with timing of arrival and onset of nesting having important consequences for reproductive success. Whether migratory waterbirds are capital or income breeders is predicated by their spring migration schedule, how long they are on breeding areas before nesting, and how adapted they are to exploiting early spring foods at northern breeding areas. However, for most species, we know little about individual migration schedules, arrival times, and duration of residence on breeding areas before nesting. To document these relationships in a northern nesting goose, we radiotracked winter-marked Tule Greater White-fronted Geese (Anser albifrons elgasi; hereafter “Tule Geese”; n = 116) from the time of their arrival in Alaska through nesting. Tule Geese arrived on coastal feeding areas in mid-April and moved to nesting locations a week later. They initiated nests 15 days (range: 6–24 days) after arrival, a period roughly equivalent to the duration of rapid follicle growth. Tule Geese that arrived the earliest were more likely to nest than geese that arrived later; early arrivals also spent more time on the breeding grounds and nested earlier than geese that arrived later. The length of the prenesting period was comparable to that of other populations of this species, but longer than for goose species that initiate rapid follicle growth before arrival on the breeding grounds. We suggest that Tule Geese nesting in more temperate climates are more likely to delay breeding to exploit local food resources than Arctic-nesting species that may be constrained by short growing seasons.

  12. What happened to the no-wait hospital? A case study of implementation of operational plans for reduced waits.

    PubMed

    Hansson, Johan; Tolf, Sara; Øvretveit, John; Carlsson, Jan; Brommels, Mats

    2012-01-01

    Both research and practice show that waiting lists are hard to reduce. Implementing complex interventions for reduced waits is an intricate and challenging process that requires special attention for surrounding factors helping and hindering the implementation. This article reports a case study of a hospital implementation of operational plans for reduced waits, with an emphasis on the process of change. A case study research design, theoretically informed by the Pettigrew and Whipp model of strategic change, was applied. Data were gathered from individual and focus group interviews with informants from different organizational levels at different times and from documents and plans. The findings revealed arrangements both helping and hindering the implementation work. Helping factors were the hospital's contemporary savings requirements and experiences from similar change initiatives. Those hindering the actions to plan and agree the changes were unclear support functions and unclear task prioritization. One contribution of this study is to demonstrate the advantages, disadvantages, and challenges of a contextualized case study for increased understanding of factors influencing organizational change implementation. One lesson for current policy is to regard context factors that are critical for successful implementation.

  13. The Psychosocial Influences of Waiting Periods on Patients Undergoing Endoscopic Submucosal Dissection.

    PubMed

    Nagao, Noriko; Tsuchiya, Aya; Ando, Sae; Arita, Mizue; Toyonaga, Takashi; Miyawaki, Ikuko

    This study aimed to clarify psychosocial influences of waiting periods on patients undergoing endoscopic submucosal dissection for cancer at an advanced medical care facility in Japan. Subjects were consenting patients hospitalized from 2009 to 2010. Qualitative and quantitative data were gathered about patients' characteristics, disease and stage, and waiting period. Qualitative content analysis was used to analyze free statements and interview data. Subjects included 154 patients with an average wait period of 46.28 days for admission. Qualitative analysis revealed the following wait period perceptions. For calmness, results indicated (1) no anxiety, (2) relief based on doctors' positive judgment, (3) whatever happens/no choice, and (4) trust in doctor. For uneasiness, perceptions included (1) the sooner, the better/eagerly waiting, (2) anxiety and concern, and (3) emotional instability. Four waiting period coping types were identified: (1) making phone inquiries, (2) busy and forgot about the medical procedure, (3) relief from anxiety, and (4) unable to function well in daily life. Patients need to be educated about cancer progression and provided an estimated wait time. They also require more information about how to manage daily life such as monitoring factors from the nursing domain including physical condition, digestive symptoms, diet, and exercise.

  14. The Psychosocial Influences of Waiting Periods on Patients Undergoing Endoscopic Submucosal Dissection

    PubMed Central

    Tsuchiya, Aya; Ando, Sae; Arita, Mizue; Toyonaga, Takashi; Miyawaki, Ikuko

    2017-01-01

    This study aimed to clarify psychosocial influences of waiting periods on patients undergoing endoscopic submucosal dissection for cancer at an advanced medical care facility in Japan. Subjects were consenting patients hospitalized from 2009 to 2010. Qualitative and quantitative data were gathered about patients' characteristics, disease and stage, and waiting period. Qualitative content analysis was used to analyze free statements and interview data. Subjects included 154 patients with an average wait period of 46.28 days for admission. Qualitative analysis revealed the following wait period perceptions. For calmness, results indicated (1) no anxiety, (2) relief based on doctors' positive judgment, (3) whatever happens/no choice, and (4) trust in doctor. For uneasiness, perceptions included (1) the sooner, the better/eagerly waiting, (2) anxiety and concern, and (3) emotional instability. Four waiting period coping types were identified: (1) making phone inquiries, (2) busy and forgot about the medical procedure, (3) relief from anxiety, and (4) unable to function well in daily life. Patients need to be educated about cancer progression and provided an estimated wait time. They also require more information about how to manage daily life such as monitoring factors from the nursing domain including physical condition, digestive symptoms, diet, and exercise. PMID:26987103

  15. Fault-tolerant wait-free shared objects

    NASA Technical Reports Server (NTRS)

    Jayanti, Prasad; Chandra, Tushar D.; Toueg, Sam

    1992-01-01

    A concurrent system consists of processes communicating via shared objects, such as shared variables, queues, etc. The concept of wait-freedom was introduced to cope with process failures: each process that accesses a wait-free object is guaranteed to get a response even if all the other processes crash. However, if a wait-free object 'crashes,' all the processes that access that object are prevented from making progress. In this paper, we introduce the concept of fault-tolerant wait-free objects, and study the problem of implementing them. We give a universal method to construct fault-tolerant wait-free objects, for all types of 'responsive' failures (including one in which faulty objects may 'lie'). In sharp contrast, we prove that many common and interesting types (such as queues, sets, and test&set) have no fault-tolerant wait-free implementations even under the most benign of the 'non-responsive' types of failure. We also introduce several concepts and techniques that are central to the design of fault-tolerant concurrent systems: the concepts of self-implementation and graceful degradation, and techniques to automatically increase the fault-tolerance of implementations. We prove matching lower bounds on the resource complexity of most of our algorithms.

  16. Wait too long to talk about kidney disease and you could be waiting for a kidney.

    MedlinePlus

    ... Home Current Issue Past Issues Public Service Announcement Kidney Disease Past Issues / Summer 2006 Table of Contents ... Javascript on. Wait too long to talk about kidney disease and you could be waiting for a ...

  17. Addition of long-distance heart procurement promotes changes in heart transplant waiting list status.

    PubMed

    Atik, Fernando Antibas; Couto, Carolina Fatima; Tirado, Freddy Ponce; Moraes, Camila Scatolin; Chaves, Renato Bueno; Vieira, Nubia W; Reis, João Gabbardo

    2014-01-01

    Evaluate the addition of long-distance heart procurement on a heart transplant program and the status of heart transplant recipients waiting list. Between September 2006 and October 2012, 72 patients were listed as heart transplant recipients. Heart transplant was performed in 41 (57%), death on the waiting list occurred in 26 (36%) and heart recovery occurred in 5 (7%). Initially, all transplants were performed with local donors. Long-distance, interstate heart procurement initiated in February 2011. Thirty (73%) transplants were performed with local donors and 11 (27%) with long-distance donors (mean distance=792 km±397). Patients submitted to interstate heart procurement had greater ischemic times (212 min ± 32 versus 90 min±18; P<0.0001). Primary graft dysfunction (distance 9.1% versus local 26.7%; P=0.23) and 1 month and 12 months actuarial survival (distance 90.1% and 90.1% versus local 90% and 86.2%; P=0.65 log rank) were similar among groups. There were marked incremental transplant center volume (64.4% versus 40.7%, P=0.05) with a tendency on less waiting list times (median 1.5 month versus 2.4 months, P=0.18). There was a tendency on reduced waiting list mortality (28.9% versus 48.2%, P=0.09). Incorporation of long-distance heart procurement, despite being associated with longer ischemic times, does not increase morbidity and mortality rates after heart transplant. It enhances viable donor pool, and it may reduce waiting list recipient mortality as well as waiting time.

  18. A Critical Analysis of the Role of Wait Time in Classroom Interactions and the Effects on Student and Teacher Interactional Behaviours

    ERIC Educational Resources Information Center

    Ingram, Jenni; Elliott, Victoria

    2016-01-01

    Extending the pauses between teachers' and students' turns (wait time) has been recommended as a way of improving classroom learning. Drawing on the Conversation Analysis literature on classroom interactions alongside extracts of classroom interactions, the relationship between these pauses and the interactional behaviour of teachers and students…

  19. Attrition after Intake at a University Counseling Center: Relationship among Client Race, Problem Type, and Time on a Waiting List

    ERIC Educational Resources Information Center

    Levy, Jacob J.; Thompson-Leonardelli, Kenya; Smith, Nathan Grant; Coleman, M. Nicole

    2005-01-01

    The present study examined the relationship between attrition after intake and the length of time that clients spent on a waiting list by client race and problem type. Participants were 1,461 clients who completed an initial intake evaluation at a large, Mid-Atlantic counseling center over a 6-year period. Fullfactorial hierarchical logistic…

  20. Models of emergency departments for reducing patient waiting times.

    PubMed

    Laskowski, Marek; McLeod, Robert D; Friesen, Marcia R; Podaima, Blake W; Alfa, Attahiru S

    2009-07-02

    In this paper, we apply both agent-based models and queuing models to investigate patient access and patient flow through emergency departments. The objective of this work is to gain insights into the comparative contributions and limitations of these complementary techniques, in their ability to contribute empirical input into healthcare policy and practice guidelines. The models were developed independently, with a view to compare their suitability to emergency department simulation. The current models implement relatively simple general scenarios, and rely on a combination of simulated and real data to simulate patient flow in a single emergency department or in multiple interacting emergency departments. In addition, several concepts from telecommunications engineering are translated into this modeling context. The framework of multiple-priority queue systems and the genetic programming paradigm of evolutionary machine learning are applied as a means of forecasting patient wait times and as a means of evolving healthcare policy, respectively. The models' utility lies in their ability to provide qualitative insights into the relative sensitivities and impacts of model input parameters, to illuminate scenarios worthy of more complex investigation, and to iteratively validate the models as they continue to be refined and extended. The paper discusses future efforts to refine, extend, and validate the models with more data and real data relative to physical (spatial-topographical) and social inputs (staffing, patient care models, etc.). Real data obtained through proximity location and tracking system technologies is one example discussed.

  1. Models of Emergency Departments for Reducing Patient Waiting Times

    PubMed Central

    Laskowski, Marek; McLeod, Robert D.; Friesen, Marcia R.; Podaima, Blake W.; Alfa, Attahiru S.

    2009-01-01

    In this paper, we apply both agent-based models and queuing models to investigate patient access and patient flow through emergency departments. The objective of this work is to gain insights into the comparative contributions and limitations of these complementary techniques, in their ability to contribute empirical input into healthcare policy and practice guidelines. The models were developed independently, with a view to compare their suitability to emergency department simulation. The current models implement relatively simple general scenarios, and rely on a combination of simulated and real data to simulate patient flow in a single emergency department or in multiple interacting emergency departments. In addition, several concepts from telecommunications engineering are translated into this modeling context. The framework of multiple-priority queue systems and the genetic programming paradigm of evolutionary machine learning are applied as a means of forecasting patient wait times and as a means of evolving healthcare policy, respectively. The models' utility lies in their ability to provide qualitative insights into the relative sensitivities and impacts of model input parameters, to illuminate scenarios worthy of more complex investigation, and to iteratively validate the models as they continue to be refined and extended. The paper discusses future efforts to refine, extend, and validate the models with more data and real data relative to physical (spatial–topographical) and social inputs (staffing, patient care models, etc.). Real data obtained through proximity location and tracking system technologies is one example discussed. PMID:19572015

  2. The design and testing of interactive hospital spaces to meet the needs of waiting children.

    PubMed

    Biddiss, Elaine; McPherson, Amy; Shea, Geoffrey; McKeever, Patricia

    2013-01-01

    To design an innovative interactive media display in a pediatric hospital clinic waiting space that addresses the growing demand for accessible, contact-surface-free options for play. In healthcare settings, waiting can be anxiety provoking for children and their accompanying family members. Opportunities for positive distraction have been shown to reduce waiting anxiety, leading to positive health outcomes. An interactive media display, ScreenPlay, was created and evaluated using a participatory design approach and a combination of techniques including quality function deployment and mixed data elicitation methods (questionnaires, focus groups, and observations). The user and organizational design requirements were established and used to review contemporary strategies for positive distraction in healthcare waiting spaces and to conceptualize and test ScreenPlay. Ten staff members, 11 children/youths, and 6 parents participated in the design and evaluation of ScreenPlay. ScreenPlay provided a positive, engaging experience without the use of contact surfaces through which infections can be spread. It was accessible to children, youth, and adults of all motor abilities. All participants strongly agreed that the interactive media display would improve the healthcare waiting experience. ScreenPlay is an interactive display that is the result of a successful model for the design of healthcare waiting spaces that is collaborative, interdisciplinary, and responsive to the needs of its community. Design process, healing environments, hospital, interdisciplinary, pediatric.

  3. Statistical distributions of avalanche size and waiting times in an inter-sandpile cascade model

    NASA Astrophysics Data System (ADS)

    Batac, Rene; Longjas, Anthony; Monterola, Christopher

    2012-02-01

    Sandpile-based models have successfully shed light on key features of nonlinear relaxational processes in nature, particularly the occurrence of fat-tailed magnitude distributions and exponential return times, from simple local stress redistributions. In this work, we extend the existing sandpile paradigm into an inter-sandpile cascade, wherein the avalanches emanating from a uniformly-driven sandpile (first layer) is used to trigger the next (second layer), and so on, in a successive fashion. Statistical characterizations reveal that avalanche size distributions evolve from a power-law p(S)≈S-1.3 for the first layer to gamma distributions p(S)≈Sαexp(-S/S0) for layers far away from the uniformly driven sandpile. The resulting avalanche size statistics is found to be associated with the corresponding waiting time distribution, as explained in an accompanying analytic formulation. Interestingly, both the numerical and analytic models show good agreement with actual inventories of non-uniformly driven events in nature.

  4. Prescription duration and treatment episodes in oral glucocorticoid users: application of the parametric waiting time distribution.

    PubMed

    Laugesen, Kristina; Støvring, Henrik; Hallas, Jesper; Pottegård, Anton; Jørgensen, Jens Otto Lunde; Sørensen, Henrik Toft; Petersen, Irene

    2017-01-01

    Glucocorticoids are widely used medications. In many pharmacoepidemiological studies, duration of individual prescriptions and definition of treatment episodes are important issues. However, many data sources lack this information. We aimed to estimate duration of individual prescriptions for oral glucocorticoids and to describe continuous treatment episodes using the parametric waiting time distribution. We used Danish nationwide registries to identify all prescriptions for oral glucocorticoids during 1996-2014. We applied the parametric waiting time distribution to estimate duration of individual prescriptions each year by estimating the 80th, 90th, 95th and 99th percentiles for the interarrival distribution. These corresponded to the time since last prescription during which 80%, 90%, 95% and 99% of users presented a new prescription for redemption. We used the Kaplan-Meier survival function to estimate length of first continuous treatment episodes by assigning estimated prescription duration to each prescription and thereby create treatment episodes from overlapping prescriptions. We identified 5,691,985 prescriptions issued to 854,429 individuals of whom 351,202 (41%) only redeemed 1 prescription in the whole study period. The 80th percentile for prescription duration ranged from 87 to 120 days, the 90th percentile from 116 to 150 days, the 95th percentile from 147 to 181 days, and the 99th percentile from 228 to 259 days during 1996-2014. Based on the 80th, 90th, 95th and 99th percentiles of prescription duration, the median length of continuous treatment was 113, 141, 170 and 243 days, respectively. Our method and results may provide an important framework for future pharmacoepidemiological studies. The choice of which percentile of the interarrival distribution to apply as prescription duration has an impact on the level of misclassification. Use of the 80th percentile provides a measure of drug exposure that is specific, while the 99th percentile provides

  5. Analyzing discharge strategies during acute care: a discrete-event simulation study.

    PubMed

    Crawford, Elizabeth A; Parikh, Pratik J; Kong, Nan; Thakar, Charuhas V

    2014-02-01

    We developed a discrete-event simulation model of patient pathway through an acute care hospital that comprises an ED and several inpatient units. The effects of discharge timing on ED waiting and boarding times, ambulance diversions, leave without treatment, and readmissions were explicitly modeled. We then analyzed the impact of 1 static and 2 proactive discharge strategies on these system outcomes. Our analysis indicated that although the 2 proactive discharge strategies significantly reduced ED waiting and boarding times, and several other measures, compared with the static strategy (P < 0.01), the number of readmissions increased substantially. Further analysis indicated that these findings are sensitive to changes in patient arrival rate and conditions for ambulance diversion. Determining the appropriate time to discharge patients not only can affect individual patients' health outcomes, but also can affect various aspects of the hospital. The study improves our understanding of how individual inpatient discharge decisions can be objectively viewed in terms of their impact on other operations, such as ED crowding and readmission, in an acute care hospital.

  6. Maternity waiting homes and institutional birth in Nicaragua: policy options and strategic implications.

    PubMed

    García Prado, Ariadna; Cortez, Rafael

    2012-01-01

    With the aim of promoting institutional births and reducing the high maternal and child mortality rates in rural and poor zones, the government of Nicaragua is supporting the creation of maternity waiting homes. This study analyzes that strategy and examines the factors associated with the use of maternity waiting homes and institutional birth. To that end, we apply a quantitative approach, by means of an econometric analysis of the data extracted from surveys conducted in 2006 on a sample of women and parteras or traditional birth attendants, as well as a qualitative approach based on interviews with key informants. Results indicate that although the operation of the maternity waiting homes is usually satisfactory, there is still room for improvement along the following lines: (i) disseminating information about the homes to both women and men, as the latter frequently decide the course of women's healthcare, and to parteras, who can play an important role in referring women; (ii) strengthening the postpartum care; (iii) ensuring financial sustainability by obtaining regular financial support from the government to complement contributions from the community; and (iv) strengthening the local management and involvement of the regional government. These measures might be useful for health policy makers in Nicaragua and in other developing countries that are considering this strategy. Copyright © 2011 John Wiley & Sons, Ltd.

  7. The Watch-and-Wait Task: On the Reliability and Validity of a New Method of Assessing Self-Control in Preschool Children

    ERIC Educational Resources Information Center

    Neubauer, Anna; Gawrilow, Caterina; Hasselhorn, Marcus

    2012-01-01

    A preschooler's ability to delay gratification in the waiting task is predictive of several developmental outcomes, despite this task's relatively low reliability level. Success in this task depends on the use of distraction strategies. The new Watch-and-Wait Task (WWT) has been developed to enhance reliability and to investigate whether the…

  8. Active Surveillance Versus Watchful Waiting for Localized Prostate Cancer: A Model to Inform Decisions.

    PubMed

    Loeb, Stacy; Zhou, Qinlian; Siebert, Uwe; Rochau, Ursula; Jahn, Beate; Mühlberger, Nikolai; Carter, H Ballentine; Lepor, Herbert; Braithwaite, R Scott

    2017-12-01

    An increasing proportion of prostate cancer is being managed conservatively. However, there are no randomized trials or consensus regarding the optimal follow-up strategy. To compare life expectancy and quality of life between watchful waiting (WW) versus different strategies of active surveillance (AS). A Markov model was created for US men starting at age 50, diagnosed with localized prostate cancer who chose conservative management by WW or AS using different testing protocols (prostate-specific antigen every 3-6 mo, biopsy every 1-5 yr, or magnetic resonance imaging based). Transition probabilities and utilities were obtained from the literature. Primary outcomes were life years and quality-adjusted life years (QALYs). Secondary outcomes include radical treatment, metastasis, and prostate cancer death. All AS strategies yielded more life years compared with WW. Lifetime risks of prostate cancer death and metastasis were, respectively, 5.42% and 6.40% with AS versus 8.72% and 10.30% with WW. AS yielded more QALYs than WW except in cohorts age >65 yr at diagnosis, or when treatment-related complications were long term. The preferred follow-up strategy was also sensitive to whether people value short-term over long-term benefits (time preference). Depending on the AS protocol, 30-41% underwent radical treatment within 10 yr. Extending the surveillance biopsy interval from 1 to 5 yr reduced life years slightly, with a 0.26 difference in QALYs. AS extends life more than WW, particularly for men with higher-risk features, but this is partly offset by the decrement in quality of life since many men eventually receive treatment. More intensive active surveillance protocols extend life more than watchful waiting, but this is partly offset by decrements in quality of life from subsequent treatment. Copyright © 2017 European Association of Urology. Published by Elsevier B.V. All rights reserved.

  9. SCREEN: A simple layperson administered screening algorithm in low resource international settings significantly reduces waiting time for critically ill children in primary healthcare clinics.

    PubMed

    Hansoti, Bhakti; Jenson, Alexander; Kironji, Antony G; Katz, Joanne; Levin, Scott; Rothman, Richard; Kelen, Gabor D; Wallis, Lee A

    2017-01-01

    In low resource settings, an inadequate number of trained healthcare workers and high volumes of children presenting to Primary Healthcare Centers (PHC) result in prolonged waiting times and significant delays in identifying and evaluating critically ill children. The Sick Children Require Emergency Evaluation Now (SCREEN) program, a simple six-question screening algorithm administered by lay healthcare workers, was developed in 2014 to rapidly identify critically ill children and to expedite their care at the point of entry into a clinic. We sought to determine the impact of SCREEN on waiting times for critically ill children post real world implementation in Cape Town, South Africa. This is a prospective, observational implementation-effectiveness hybrid study that sought to determine: (1) the impact of SCREEN implementation on waiting times as a primary outcome measure, and (2) the effectiveness of the SCREEN tool in accurately identifying critically ill children when utilised by the QM and adherence by the QM to the SCREEN algorithm as secondary outcome measures. The study was conducted in two phases, Phase I control (pre-SCREEN implementation- three months in 2014) and Phase II (post-SCREEN implementation-two distinct three month periods in 2016). In Phase I, 1600 (92.38%) of 1732 children presenting to 4 clinics, had sufficient data for analysis and comprised the control sample. In Phase II, all 3383 of the children presenting to the 26 clinics during the sampling time frame had sufficient data for analysis. The proportion of critically ill children who saw a professional nurse within 10 minutes increased tenfold from 6.4% to 64% (Phase I to Phase II) with the median time to seeing a professional nurse reduced from 100.3 minutes to 4.9 minutes, (p < .001, respectively). Overall layperson screening compared to Integrated Management of Childhood Illnesses (IMCI) designation by a nurse had a sensitivity of 94.2% and a specificity of 88.1%, despite large variance

  10. Prediction of Pathological Complete Response Using Endoscopic Findings and Outcomes of Patients Who Underwent Watchful Waiting After Chemoradiotherapy for Rectal Cancer.

    PubMed

    Kawai, Kazushige; Ishihara, Soichiro; Nozawa, Hiroaki; Hata, Keisuke; Kiyomatsu, Tomomichi; Morikawa, Teppei; Fukayama, Masashi; Watanabe, Toshiaki

    2017-04-01

    Nonoperative management for patients with rectal cancer who have achieved a clinical complete response after chemoradiotherapy is becoming increasingly important in recent years. However, the definition of and modality used for patients with clinical complete response differ greatly between institutions, and the role of endoscopic assessment as a nonoperative approach has not been fully investigated. This study aimed to investigate the ability of endoscopic assessments to predict pathological regression of rectal cancer after chemoradiotherapy and the applicability of these assessments for the watchful waiting approach. This was a retrospective comparative study. This study was conducted at a single referral hospital. A total of 198 patients with rectal cancer underwent preoperative endoscopic assessments after chemoradiotherapy. Of them, 186 patients underwent radical surgery with lymph node dissection. The histopathological findings of resected tissues were compared with the preoperative endoscopic findings. Twelve patients refused radical surgery and chose watchful waiting; their outcomes were compared with the outcomes of patients who underwent radical surgery. The endoscopic criteria correlated well with tumor regression grading. The sensitivity and specificity for a pathological complete response were 65.0% to 87.1% and 39.1% to 78.3%. However, endoscopic assessment could not fully discriminate pathological complete responses, and the outcomes of patients who underwent watchful waiting were considerably poorer than the patients who underwent radical surgery. Eventually, 41.7% of the patients who underwent watchful waiting experienced uncontrollable local failure, and many of these occurrences were observed more than 3 years after chemoradiotherapy. The number of the patients treated with the watchful waiting strategy was limited, and the selection was not randomized. Although endoscopic assessment after chemoradiotherapy correlated with pathological response

  11. Predictors of fecundability and conception waits among the Dogon of Mali.

    PubMed

    Strassmann, B I; Warner, J H

    1998-02-01

    Surprisingly little is known about the mechanisms that underlie variation in female fertility in humans. Data on this topic are nonetheless vital to a number of pragmatic and theoretical enterprises, including population planning, infertility treatment and prevention, and evolutionary ecology. Here we study female fertility by focusing on one component of the interbirth interval: the waiting time to conception during menstrual cycling. Our study population is a Dogon village of 460 people in Mali, West Africa. This population is pronatalist and noncontracepting. In accordance with animist beliefs, the women spend five nights sleeping at a menstrual hut during menses. By censusing the women present at the menstrual huts in the study village on each of 736 consecutive nights, we were able to monitor women's conception waits prospectively. Hormonal profiles confirm the accuracy of the data on conception waits obtained from the menstrual hut census (Strassmann [1996], Behavioral Ecology 7:304-315). Using survival analysis, we identified significant predictors of the waiting time to conception: wife's age (years), husband's age (< 35, 35-49, > 49 years), marital duration (years), gravidity (number of prior pregnancies), and breast-feeding status. Additional variables were not significant, including duration of postpartum amenorrhea, sex of the last child, nutritional status, economic status, polygyny, and marital status (fiancée vs. married). We fit both continuous and discrete time survival models, but the former appeared to be a better choice for these data.

  12. Improving Health Care Accessibility: Strategies and Recommendations.

    PubMed

    Almorsy, Lamia; Khalifa, Mohamed

    2016-01-01

    Access time refers to the interval between requesting and actual outpatient appointment. It reflects healthcare accessibility and has a great influence on patient treatment and satisfaction. King Faisal Specialist Hospital and Research Center, Jeddah, Saudi Arabia studied the accessibility to outpatient services in order to develop useful strategies and recommendations for improvement. Utilized, unutilized and no-show appointments were analyzed. It is crucial to manage no-shows and short notice appointment cancellations by preparing a waiting list for those patients who can be called in to an appointment on the same day using an open access policy. An overlapping appointment scheduling model can be useful to minimize patient waiting time and doctor idle time in addition to the sensible use of appointment overbooking that can significantly improve productivity.

  13. The waiting room: vector for health education? The general practitioner's point of view.

    PubMed

    Gignon, Maxine; Idris, Hadjila; Manaouil, Cecile; Ganry, Oliver

    2012-09-18

    General practitioners (GPs) play a central role in disseminating information and most health policies are tending to develop this pivotal role of GPs in dissemination of health-related information to the public. The objective of this study was to evaluate use of the waiting room by GPs as a vector for health promotion. A cross-sectional study was conducted on a representative sample of GPs using semi-structured, face-to-face interviews. A structured grid was used to describe the documents. Quantitative and qualitative analysis was performed. Sixty GPs participated in the study. They stated that a waiting room had to be pleasant, but agreed that it was a useful vector for providing health information. The GPs stated that they distributed documents designed to improve patient care by encouraging screening, providing health education information and addressing delicate subjects more easily. However, some physicians believed that this information can sometimes make patients more anxious. A large number of documents were often available, covering a variety of topics. General practitioners intentionally use their waiting rooms to disseminate a broad range of health-related information, but without developing a clearly defined strategy. It would be interesting to correlate the topics addressed by waiting room documents with prevention practices introduced during the visit.

  14. Improving Patients Experience in Peadiatric Emergency Waiting Room.

    PubMed

    Ehrler, Frederic; Siebert, Johan; Wipfli, Rolf; Duret, Cyrille; Gervaix, Alain; Lovis, Christian

    2016-01-01

    When visiting the emergency department, the perception of the time spent in the waiting room before the beginning of the care, may influence patients' experience. Based on models of service evaluation, highlighting the importance of informing people about their waiting process and their place in the queue, we have developed an innovative information screen aiming at improving perception of time by patients. Following an iterative process, a group of experts including computer scientists, ergonomists and caregivers designed a solution adapted to the pediatric context. The solution includes a screen displaying five lanes representing triage levels. Patients are represented by individual avatars, drawn sequentially in the appropriate line. The interface has been designed using gamification principle, aiming at increasing acceptance, lowering learning curve and improving satisfaction. Questionnaire based evaluation results revealed high satisfaction from the 278 respondents even if the informative content was not always completely clear.

  15. Dedicated Pediatricians in Emergency Department: Shorter Waiting Times and Lower Costs

    PubMed Central

    Melo, Manuel Rocha; Ferreira-Magalhães, Manuel; Flor-Lima, Filipa; Rodrigues, Mariana; Severo, Milton; Almeida-Santos, Luis; Caldas-Afonso, Alberto; Barros, Pedro Pita; Ferreira, António

    2016-01-01

    Background Dedicated pediatricians in emergency departments (EDs) may be beneficial, though no previous studies have assessed the related costs and benefits/harms. We aimed to evaluate the net benefits and costs of dedicated emergency pediatricians in a pediatric ED. Methods Cost-consequences analysis of visits to a pediatric ED of a tertiary hospital. Two pediatric ED Medical Teams (MT) were compared: MT-A (May–September 2012), with general pediatrics physicians only; and MT-B (May–September 2013), with emergency dedicated pediatricians. The main outcomes analyzed were relevant clinical outcomes, patient throughput time and costs. Results We included 8,694 children in MT-A and 9,417 in MT-B. Medication use in the ED increased from 42.3% of the children in MT-A to 49.6% in MT-B; diagnostic tests decreased from 24.2% in MT-A to 14.3% in MT-B. Hospitalization increased from 1.3% in MT-A to 3.0% in MT-B; however, there was no significant difference in diagnosis-related group relative weight of hospitalized children in MT-A and MT-B (MT-A, 0.979; MT-B, 1.075). No differences were observed in ED readmissions or in patients leaving without being seen by a physician. The patient throughput time was significantly shorter in MT-B, with faster times to first medical observation. Within the cost domains analyzed, the total expenditures per children observed in the ED were 16% lower in MT-B: 37.87 euros in MT-A; 31.97 euros in MT-B. Conclusion The presence of dedicated emergency pediatricians in a pediatric ED was associated with significantly lower waiting times in the ED, reduced costs, and similar clinical outcomes. PMID:27564093

  16. "Wait and see" vaccinating behaviour during a pandemic: a game theoretic analysis.

    PubMed

    Bhattacharyya, Samit; Bauch, Chris T

    2011-07-26

    During the 2009 H1N1 pandemic, many individuals did not seek vaccination immediately but rather decided to "wait and see" until further information was available on vaccination costs. This behaviour implies two sources of strategic interaction: as more individuals become vaccinated, both the perceived vaccination cost and the probability that susceptible individuals become infected decline. Here we analyze the outcome of these two strategic interactions by combining game theory with a disease transmission model during an outbreak of a novel influenza strain. The model exhibits a "wait and see" Nash equilibrium strategy, with vaccine delayers relying on herd immunity and vaccine safety information generated by early vaccinators. This strategic behaviour causes the timing of the epidemic peak to be strongly conserved across a broad range of plausible transmission rates, in contrast to models without such adaptive behaviour. The model exhibits not only feedback mechanisms but also a feed-forward mechanism: a high initial perceived vaccination cost perpetuates high perceived vaccine costs (and lower vaccine coverage) throughout the remainder of the outbreak. This suggests that any effect of risk communication at the start of a pandemic outbreak will be amplified compared to the same amount of risk communication effort distributed throughout the outbreak. Copyright © 2011 Elsevier Ltd. All rights reserved.

  17. Built spaces and features associated with user satisfaction in maternity waiting homes in Malawi.

    PubMed

    McIntosh, Nathalie; Gruits, Patricia; Oppel, Eva; Shao, Amie

    2018-07-01

    To assess satisfaction with maternity waiting home built spaces and features in women who are at risk for underutilizing maternity waiting homes (i.e. residential facilities that temporarily house near-term pregnant mothers close to healthcare facilities that provide obstetrical care). Specifically we wanted to answer the questions: (1) Are built spaces and features associated with maternity waiting home user satisfaction? (2) Can built spaces and features designed to improve hygiene, comfort, privacy and function improve maternity waiting home user satisfaction? And (3) Which built spaces and features are most important for maternity waiting home user satisfaction? A cross-sectional study comparing satisfaction with standard and non-standard maternity waiting home designs. Between December 2016 and February 2017 we surveyed expectant mothers at two maternity waiting homes that differed in their design of built spaces and features. We used bivariate analyses to assess if built spaces and features were associated with satisfaction. We compared ratings of built spaces and features between the two maternity waiting homes using chi-squares and t-tests to assess if design features to improve hygiene, comfort, privacy and function were associated with higher satisfaction. We used exploratory robust regression analysis to examine the relationship between built spaces and features and maternity waiting home satisfaction. Two maternity waiting homes in Malawi, one that incorporated non-standardized design features to improve hygiene, comfort, privacy, and function (Kasungu maternity waiting home) and the other that had a standard maternity waiting home design (Dowa maternity waiting home). 322 expectant mothers at risk for underutilizing maternity waiting homes (i.e. first-time mothers and those with no pregnancy risk factors) who had stayed at the Kasungu or Dowa maternity waiting homes. There were significant differences in ratings of built spaces and features between the

  18. Which patients are not included in the English Cancer Waiting Times monitoring dataset, 2009-2013? Implications for use of the data in research.

    PubMed

    Di Girolamo, C; Walters, S; Gildea, C; Benitez Majano, S; Coleman, M P; Rachet, B; Morris, M

    2018-03-06

    Cancer waiting time targets are routinely monitored in England, but the Cancer Waiting Times monitoring dataset (CWT) does not include all eligible patients, introducing scope for bias. Data from adults diagnosed in England (2009-2013) with colorectal, lung, or ovarian cancer were linked from CWT to cancer registry, mortality, and Hospital Episode Statistics data. We present demographic characteristics and net survival for patients who were and were not included in CWT. A CWT record was found for 82% of colorectal, 76% of lung, and 77% of ovarian cancer patients. Patients not recorded in CWT were more likely to be in the youngest or oldest age groups, have more comorbidities, have been diagnosed through emergency presentation, have late or missing stage, and have much poorer survival. Researchers and policy-makers should be aware of the limitations in the completeness and representativeness of CWT, and draw conclusions with appropriate caution.

  19. Info card for surgery waiting room improves satisfaction.

    PubMed

    2015-11-01

    A hospital is reporting improved patient satisfaction from providing an information card in the surgery department. The card includes expected wait times. The card is provided by the patient transport team. Telephone numbers are included for more information. Staff update family members hourly during surgery.

  20. Abortion Patients' Experience and Perceptions of Waiting Periods: Survey Evidence before Arizona's Two-visit 24-hour Mandatory Waiting Period Law.

    PubMed

    Karasek, Deborah; Roberts, Sarah C M; Weitz, Tracy A

    2016-01-01

    More than one-half of U.S. states now have laws requiring women to wait at least 24 hours between receiving information about abortion and the actual abortion procedure, with a few requiring longer waits, and one-fourth requiring that women receive this information in person. Although public discussions of waiting periods focus on how they affect women, we know little about abortion patients' perceptions of these requirements. We collected data from 379 women seeking abortion care at an abortion facility in Arizona before Arizona's 24-hour waiting period two-visit requirement went into effect. Surveys focused on patients' experiences receiving abortion care before the waiting period and perceptions about how the additional clinic visit would affect them. Most women reported one or more financial or logistical challenges in obtaining abortion care. More than two-thirds reported difficulty paying abortion appointment-related expenses. These expenses prevented or delayed almost one-half from paying other expenses, such as rent, bills, and food, with lower income women more affected. The majority expected that the additional visit would result in additional financial and logistical hardships and delay them in having an abortion, with 90% reporting that the waiting period would lead to at least one hardship. Eight percent reported that the waiting period would have a positive effect on emotional well-being, and more than one-half reported that it would have a negative effect on emotional well-being. Only a small minority of women seeking abortion care view a two-visit waiting period law as benefiting them; the overwhelming majority expect a waiting period to have adverse consequences. Copyright © 2016 Jacobs Institute of Women's Health. Published by Elsevier Inc. All rights reserved.

  1. A research on motion design for APP's loading pages based on time perception

    NASA Astrophysics Data System (ADS)

    Cao, Huai; Hu, Xiaoyun

    2018-04-01

    Due to restrictions caused by objective reasons like network bandwidth, hardware performance and etc., waiting is still an inevitable phenomenon that appears in our using mobile-terminal products. Relevant researches show that users' feelings in a waiting scenario can affect their evaluations on the whole product and services the product provides. With the development of user experience and inter-facial design subjects, the role of motion effect in the interface design has attracted more and more scholars' attention. In the current studies, the research theory of motion design in a waiting scenario is imperfect. This article will use the basic theory and experimental research methods of cognitive psychology to explore the motion design's impact on user's time perception when users are waiting for loading APP pages. Firstly, the article analyzes the factors that affect waiting experience of loading APP pages based on the theory of time perception, and then discusses motion design's impact on the level of time-perception when loading pages and its design strategy. Moreover, by the operation analysis of existing loading motion designs, the article classifies the existing loading motions and designs an experiment to verify the impact of different types of motions on the user's time perception. The result shows that the waiting time perception of mobile's terminals' APPs is related to the loading motion types, the combination type of loading motions can effectively shorten the waiting time perception as it scores a higher mean value in the length of time perception.

  2. Spontaneous activity in the waiting brain: a marker of impulsive choice in attention-deficit/hyperactivity disorder?

    PubMed

    Hsu, Chia-Fen; Benikos, Nicholas; Sonuga-Barke, Edmund J S

    2015-04-01

    Spontaneous very low frequency oscillations (VLFO), seen in the resting brain, are attenuated when individuals are working on attention demanding tasks or waiting for rewards (Hsu et al., 2013). Individuals with attention-deficit/hyperactivity disorder (ADHD) display excess VLFO when working on attention tasks. They also have difficulty waiting for rewards. Here we examined the waiting brain signature in ADHD and its association with impulsive choice. DC-EEG from 21 children with ADHD and 21 controls (9-15 years) were collected under four conditions: (i) resting; (ii) choosing to wait; (iii) being "forced" to wait; and (iv) working on a reaction time task. A questionnaire measured two components of impulsive choice. Significant VLFO reductions were observed in controls within anterior brain regions in both working and waiting conditions. Individuals with ADHD showed VLFO attenuation while working but to a reduced level and none at all when waiting. A closer inspection revealed an increase of VLFO activity in temporal regions during waiting. Excess VLFO activity during waiting was associated with parents' ratings of temporal discounting and delay aversion. The results highlight the potential role for waiting-related spontaneous neural activity in the pathophysiology of impulsive decision-making of ADHD. Copyright © 2015 The Authors. Published by Elsevier Ltd.. All rights reserved.

  3. Waiting can be an optimal conservation strategy, even in a crisis discipline

    PubMed Central

    Possingham, Hugh P.; Bode, Michael

    2017-01-01

    Biodiversity conservation projects confront immediate and escalating threats with limited funding. Conservation theory suggests that the best response to the species extinction crisis is to spend money as soon as it becomes available, and this is often an explicit constraint placed on funding. We use a general dynamic model of a conservation landscape to show that this decision to “front-load” project spending can be suboptimal if a delay allows managers to use resources more strategically. Our model demonstrates the existence of temporal efficiencies in conservation management, which parallel the spatial efficiencies identified by systematic conservation planning. The optimal timing of decisions balances the rate of biodiversity decline (e.g., the relaxation of extinction debts, or the progress of climate change) against the rate at which spending appreciates in value (e.g., through interest, learning, or capacity building). We contrast the benefits of acting and waiting in two ecosystems where restoration can mitigate forest bird extinction debts: South Australia’s Mount Lofty Ranges and Paraguay’s Atlantic Forest. In both cases, conservation outcomes cannot be maximized by front-loading spending, and the optimal solution recommends substantial delays before managers undertake conservation actions. Surprisingly, these delays allow superior conservation benefits to be achieved, in less time than front-loading. Our analyses provide an intuitive and mechanistic rationale for strategic delay, which contrasts with the orthodoxy of front-loaded spending for conservation actions. Our results illustrate the conservation efficiencies that could be achieved if decision makers choose when to spend their limited resources, as opposed to just where to spend them. PMID:28894004

  4. Waiting can be an optimal conservation strategy, even in a crisis discipline.

    PubMed

    Iacona, Gwenllian D; Possingham, Hugh P; Bode, Michael

    2017-09-26

    Biodiversity conservation projects confront immediate and escalating threats with limited funding. Conservation theory suggests that the best response to the species extinction crisis is to spend money as soon as it becomes available, and this is often an explicit constraint placed on funding. We use a general dynamic model of a conservation landscape to show that this decision to "front-load" project spending can be suboptimal if a delay allows managers to use resources more strategically. Our model demonstrates the existence of temporal efficiencies in conservation management, which parallel the spatial efficiencies identified by systematic conservation planning. The optimal timing of decisions balances the rate of biodiversity decline (e.g., the relaxation of extinction debts, or the progress of climate change) against the rate at which spending appreciates in value (e.g., through interest, learning, or capacity building). We contrast the benefits of acting and waiting in two ecosystems where restoration can mitigate forest bird extinction debts: South Australia's Mount Lofty Ranges and Paraguay's Atlantic Forest. In both cases, conservation outcomes cannot be maximized by front-loading spending, and the optimal solution recommends substantial delays before managers undertake conservation actions. Surprisingly, these delays allow superior conservation benefits to be achieved, in less time than front-loading. Our analyses provide an intuitive and mechanistic rationale for strategic delay, which contrasts with the orthodoxy of front-loaded spending for conservation actions. Our results illustrate the conservation efficiencies that could be achieved if decision makers choose when to spend their limited resources, as opposed to just where to spend them.

  5. Failure to cope: the hidden curriculum of emergency department wait times and the implications for clinical training.

    PubMed

    Webster, Fiona; Rice, Kathleen; Dainty, Katie N; Zwarenstein, Merrick; Durant, Steve; Kuper, Ayelet

    2015-01-01

    The study explored optimal intraprofessional collaboration between physicians in the emergency department (ED) and those from general internal medicine (GIM). Prior to the study, a policy was initiated that mandated reductions in ED wait times. The researchers examined the impact of these changes on clinical practice and trainee education. In 2010-2011, an ethnographic study was undertaken to observe consults between GIM and ED at an urban teaching hospital in Ontario, Canada. Additional ad hoc interviews were conducted with residents, nurses, and faculty from both departments as well as formal one-on-one interviews with 12 physicians. Data were coded and analyzed using concepts of institutional ethnography. Participants perceived that efficiency was more important than education and was in fact the new definition of "good" patient care. The informal label "failure to cope" to describe high-needs patients suggested that in many instances, patients were experienced as a barrier to optimal efficiency. This resulted in tension during consults as well as reduced opportunities for education. The authors suggest that the emphasis on wait times resulted in more importance being placed on "getting the patient out" of the ED than on providing safe, compassionate, person-centered medical care. Resource constraints were hidden within a discourse that shifted the problem of overcrowding in the ED to patients with complex chronic conditions. The term "failure to cope" became activated when overworked physicians tried to avoid assuming care for high-needs patients, masking institutionally produced stress and possibly altering the way patients are perceived.

  6. Cumulative incidence for wait-list death in relation to length of queue for coronary-artery bypass grafting: a cohort study.

    PubMed

    Sobolev, Boris G; Kuramoto, Lisa; Levy, Adrian R; Hayden, Robert

    2006-08-24

    In deciding where to undergo coronary-artery bypass grafting, the length of surgical wait lists is often the only information available to cardiologists and their patients. Our objective was to compare the cumulative incidence for death on the wait list according to the length of wait lists at the time of registration for the operation. The study cohort included 8966 patients who registered to undergo isolated coronary-artery bypass grafting (82.4% men; 71.9% semi-urgent; 22.4% non-urgent). The patients were categorized according to wait-list clearance time at registration: either "1 month or less" or "more than 1 month". Cumulative incidence for wait-list death was compared between the groups, and the significance of difference was tested by means of regression models. Urgent patients never registered on a wait list with a clearance time of more than 1 month. Semi-urgent patients registered on shorter wait lists more often than non-urgent patients (79.1% vs. 44.7%). In semi-urgent and non-urgent patients, the observed proportion of wait-list deaths by 52 weeks was lower in category "1 month or less" than in category "more than 1 month" (0.8% [49 deaths] vs. 1.6% [39 deaths], P < 0.005). After adjustment, the odds of death before surgery were 64% higher in patients on longer lists, odds ratio [OR] = 1.64 (95% confidence interval [CI] 1.02-2.63). The observed death rate was higher in category "more than 1 month" than in category "1 month or less", 0.79 (95%CI 0.54-1.04) vs. 0.58 (95% CI 0.42-0.74) per 1000 patient-weeks, the adjusted OR = 1.60 (95%CI 1.01-2.53). Longer wait times (log-rank test = 266.4, P < 0.001) and higher death rates contributed to a higher cumulative incidence for death on the wait list with a clearance time of more than 1 month. Long wait lists for coronary-artery bypass grafting are associated with increased probability that a patient dies before surgery. Physicians who advise patients where to undergo cardiac revascularization should consider

  7. "Wait a while, my love" -- an Indonesian popular song with a family planning message.

    PubMed

    Pekerti, R; Musa, R

    1989-10-01

    "Wait a While, My Love," recorded by pop singer Irianti Emingpraja, was the first Indonesian rock sock to contain a family planning message. The album including the song has sold over 100,000 copies. The song has also been packaged as a 60-second video that can be used as an opening theme for radio and television programs. The song, aimed at encouraging Indonesian youth to postpone marriage, has the following lyrics: "Flying free like a seagull/I'll cover many places 'round the world/Give me time for study and reflection, to grow as a mature wise woman/Oh, wait a while, my love/Don't buy me a ring, reflection of your inner love/I'll climb my way up to the top of the world/And reaching our rainbow of hope." The song was produced with support from the United Nations Fund for Population Activities and the Indonesian National Family Planning Coordinating Board. Key factors to be examined in producing a popular song with a family planning message include the specific message desired, the target audience, type of music, the singer, the producer, marketing, a multimedia campaign strategy, and distribution outlets.

  8. Spatiotemporal switching signals for cancer stem cell activation in pediatric origins of adulthood cancer: Towards a watch-and-wait lifetime strategy for cancer treatment.

    PubMed

    Li, Shengwen Calvin; Kabeer, Mustafa H

    2018-02-26

    Pediatric origin of cancer stem cell hypothesis holds great promise and potential in adult cancer treatment, however; the road to innovation is full of obstacles as there are plenty of questions left unanswered. First, the key question is to characterize the nature of such stem cells (concept). Second, the quantitative imaging of pediatric stem cells should be implemented (technology). Conceptually, pediatric stem cell origins of adult cancer are based on the notion that plasticity in early life developmental programming evolves local environments to cancer. Technologically, such imaging in children is lacking as all imaging is designed for adult patients. We postulate that the need for quantitative imaging to measure space-time changes of plasticity in early life developmental programming in children may trigger research and development of the imaging technology. Such quantitative imaging of pediatric origin of adulthood cancer will help develop a spatiotemporal monitoring system to determine cancer initiation and progression. Clinical validation of such speculative hypothesis-that cancer originates in a pediatric environment-will help implement a wait-and-watch strategy for cancer treatment.

  9. Association between the Medicare Modernization Act of 2003 and patient wait times and travel distance for chemotherapy.

    PubMed

    Shea, Alisa M; Curtis, Lesley H; Hammill, Bradley G; DiMartino, Lisa D; Abernethy, Amy P; Schulman, Kevin A

    2008-07-09

    The Medicare Prescription Drug, Improvement, and Modernization Act of 2003 (MMA) altered reimbursements for outpatient chemotherapy drugs and drug administration services. Anecdotal reports suggest that these adjustments may have negatively affected access to chemotherapy for Medicare beneficiaries. To compare patient wait times and travel distances for chemotherapy before and after the enactment of the MMA. Analysis of a nationally representative 5% sample of claims from the Centers for Medicare & Medicaid Services for the period 2003 through 2006. Patients were Medicare beneficiaries with incident breast cancer, colorectal cancer, leukemia, lung cancer, or lymphoma who received chemotherapy in inpatient hospital, institutional outpatient, or physician office settings. Days from incident diagnosis to first chemotherapy visit and distance traveled for treatment, controlling for age, sex, race/ethnicity, cancer type, geographic region, comorbid conditions, and year of diagnosis and treatment. There were 5082 incident cases of breast cancer, colorectal cancer, leukemia, lung cancer, or lymphoma in 2003; 5379 cases in 2004; 5116 cases in 2005; and 5288 cases in 2006. Approximately 70% of patients received treatment in physician office settings in each year. Although the distribution of treatment settings in 2004 and 2005 was not significantly different from 2003 (P = .24 and P = .72, respectively), there was a small but significant change from 2003 to 2006 (P = .02). The proportion of patients receiving chemotherapy in inpatient settings decreased from 10.2% in 2003 to 8.8% in 2006 (P = .03), and the proportion in institutional outpatient settings increased from 21.1% to 22.5% (P = .004). The proportion in physician offices remained at 68.7% (P = .29). The median time from diagnosis to initial chemotherapy visit was 28 days in 2003, 27 days in 2004, 29 days in 2005, and 28 days in 2006. In multivariate analyses, average wait times for chemotherapy were 1.96 days

  10. Psychosocial and Patient Education Needs of Prostate Cancers Selecting Watchful Waiting

    DTIC Science & Technology

    2008-11-01

    Davis Highway, Suite 1204, Arlington, VA 22202-4302, and to the Office of Management and Budget, Paperwork Reduction Project (0704-0188), Washington...ABSTRACT (Maximum 200 Words) While watchful waiting is an accepted disease management strategy for localized prostate cancer, there is little...information available on the impact of the disease and the expectant management on men’s well-being. The few studies that have focused on these issues

  11. The effectiveness of service delivery initiatives at improving patients' waiting times in clinical radiology departments: a systematic review.

    PubMed

    Olisemeke, B; Chen, Y F; Hemming, K; Girling, A

    2014-12-01

    We reviewed the literature for the impact of service delivery initiatives (SDIs) on patients' waiting times within radiology departments. We searched MEDLINE, EMBASE, CINAHL, INSPEC and The Cochrane Library for relevant articles published between 1995 and February, 2013. The Cochrane EPOC risk of bias tool was used to assess the risk of bias on studies that met specified design criteria. Fifty-seven studies met the inclusion criteria. The types of SDI implemented included extended scope practice (ESP, three studies), quality management (12 studies), productivity-enhancing technologies (PETs, 29 studies), multiple interventions (11 studies), outsourcing and pay-for-performance (one study each). The uncontrolled pre- and post-intervention and the post-intervention designs were used in 54 (95%) of the studies. The reporting quality was poor: many of the studies did not test and/or report the statistical significance of their results. The studies were highly heterogeneous, therefore meta-analysis was inappropriate. The following type of SDIs showed promising results: extended scope practice; quality management methodologies including Six Sigma, Lean methodology, and continuous quality improvement; productivity-enhancing technologies including speech recognition reporting, teleradiology and computerised physician order entry systems. We have suggested improved study design and the mapping of the definitions of patient waiting times in radiology to generic timelines as a starting point for moving towards a situation where it becomes less restrictive to compare and/or pool the results of future studies in a meta-analysis.

  12. Impact of Optimization Strategy and Adoption Rate on V2X Technology on Environmental Impact

    DOT National Transportation Integrated Search

    2018-12-31

    This research evaluated the effects of automated vehicle control strategies on system level emissions, travel time and wait time through a series of traffic lights. The study was conducted using traffic simulation and a realistic vehicle mix. Two con...

  13. Monitoring trends in waiting periods in Canada for elective surgery: validation of a method using administrative data.

    PubMed

    Shortt, Samuel E D; Shaw, Ralph A; Elliott, David; Mackillop, William J

    2004-06-01

    Provincial governments require timely, economical methods to monitor surgical waiting periods. Although use of prospective procedure-specific registers would be the ideal method, a less elaborate system has been proposed that is based on physician billing data. This study assessed the validity of using the date of the last service billed prior to surgery as a proxy for the beginning of the post-referral, pre-surgical waiting period. We examined charts for 31,824 elective surgical encounters between 1992 and 1996 at an Ontario teaching hospital. The date of the last service before surgery (the last billing date) was compared with the date of the consultant's letter indicating a decision to book surgery (i.e., to begin waiting). Several surgical specialties (but excluding cardiac, orthopedic and gynecologic) had a close correlation between the dates of the last pre-surgery visit and those of the actual decision to place the patient on the waiting list. Similar results were found for 12 of 15 individually studied procedures, including some orthopedic and gynecological procedures. Used judiciously, billing data is a timely, inexpensive and generally accurate method by which provincial governments could monitor trends in waiting times for appropriately selected surgical procedures.

  14. An improved global dynamic routing strategy for scale-free network with tunable clustering

    NASA Astrophysics Data System (ADS)

    Sun, Lina; Huang, Ning; Zhang, Yue; Bai, Yannan

    2016-08-01

    An efficient routing strategy can deliver packets quickly to improve the network capacity. Node congestion and transmission path length are inevitable real-time factors for a good routing strategy. Existing dynamic global routing strategies only consider the congestion of neighbor nodes and the shortest path, which ignores other key nodes’ congestion on the path. With the development of detection methods and techniques, global traffic information is readily available and important for the routing choice. Reasonable use of this information can effectively improve the network routing. So, an improved global dynamic routing strategy is proposed, which considers the congestion of all nodes on the shortest path and incorporates the waiting time of the most congested node into the path. We investigate the effectiveness of the proposed routing for scale-free network with different clustering coefficients. The shortest path routing strategy and the traffic awareness routing strategy only considering the waiting time of neighbor node are analyzed comparatively. Simulation results show that network capacity is greatly enhanced compared with the shortest path; congestion state increase is relatively slow compared with the traffic awareness routing strategy. Clustering coefficient increase will not only reduce the network throughput, but also result in transmission average path length increase for scale-free network with tunable clustering. The proposed routing is favorable to ease network congestion and network routing strategy design.

  15. Waiting Points in Nova and X-ray Burst Nucleosynthesis

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

    Sunayama, Tomomi; Smith, Michael Scott; Lingerfelt, Eric J

    2008-01-01

    In nova and X-ray burst nucleosynthesis, waiting points are nuclei in the reaction path which interrupt the nuclear flow towards heavier nuclei, typically because of a weak proton capture reaction and a long beta+ lifetime. Waiting points can influence the energy generation and final abundances synthesized in these explosions. We have constructed a systematic, quantitative set of criteria to identify rp-process waiting points, and use them to search for waiting points in post-processing simulations of novae and X-ray bursts. These criteria have been incorporated into the Computational Infrastructure for Nuclear Astrophysics, online at nucastrodata.org, to enable anyone to run customizedmore » searches for waiting points.« less

  16. Waiting Points in Nova and X-ray burst Nucleosynthesis

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

    Sunayama, Tomomi; Oak Ridge Institute for Science Education, Oak Ridge, Tennessee 37831-0117; Smith, Michael S.

    2008-05-21

    In nova and X-ray burst nucleosynthesis, waiting points are nuclei in the reaction path which delay the nuclear flow towards heavier nuclei, typically because of a weak proton capture reaction and a long {beta}{sup +} lifetime. Waiting points can influence the energy generation and final abundances synthesized in these explosions. We have constructed a systematic, quantitative set of criteria to identify rp-process waiting points, and use them to search for waiting points in post-processing simulations of novae and X-ray bursts. These criteria have been incorporated into the Computational Infrastructure for Nuclear Astrophysics, online at nucastrodata.org, to enable anyone to runmore » customized searches for waiting points.« less

  17. Which patients are not included in the English Cancer Waiting Times monitoring dataset, 2009–2013? Implications for use of the data in research

    PubMed Central

    Di Girolamo, C; Walters, S; Gildea, C; Benitez Majano, S; Coleman, M P; Rachet, B; Morris, M

    2018-01-01

    Background: Cancer waiting time targets are routinely monitored in England, but the Cancer Waiting Times monitoring dataset (CWT) does not include all eligible patients, introducing scope for bias. Methods: Data from adults diagnosed in England (2009–2013) with colorectal, lung, or ovarian cancer were linked from CWT to cancer registry, mortality, and Hospital Episode Statistics data. We present demographic characteristics and net survival for patients who were and were not included in CWT. Results: A CWT record was found for 82% of colorectal, 76% of lung, and 77% of ovarian cancer patients. Patients not recorded in CWT were more likely to be in the youngest or oldest age groups, have more comorbidities, have been diagnosed through emergency presentation, have late or missing stage, and have much poorer survival. Conclusions: Researchers and policy-makers should be aware of the limitations in the completeness and representativeness of CWT, and draw conclusions with appropriate caution. PMID:29348489

  18. Ultrasound waiting lists: rational queue or extended capacity?

    PubMed

    Brasted, Christopher

    2008-06-01

    The features and issues regarding clinical waiting lists in general and general ultrasound waiting lists in particular are reviewed, and operational aspects of providing a general ultrasound service are also discussed. A case study is presented describing a service improvement intervention in a UK NHS hospital's ultrasound department, from which arises requirements for a predictive planning model for an ultrasound waiting list. In the course of this, it becomes apparent that a booking system is a more appropriate way of describing the waiting list than a conventional queue. Distinctive features are identified from the literature and the case study as the basis for a predictive model, and a discrete event simulation model is presented which incorporates the distinctive features.

  19. Transformation of a Pediatric Primary Care Waiting Room: Creating a Bridge to Community Resources.

    PubMed

    Henize, Adrienne W; Beck, Andrew F; Klein, Melissa D; Morehous, John; Kahn, Robert S

    2018-06-01

    Introduction Children and families living in poverty frequently encounter social risks that significantly affect their health and well-being. Physicians' near universal access to at-risk children and their parents presents opportunities to address social risks, but time constraints frequently interfere. We sought to redesign our waiting room to create a clinic-to-community bridge and evaluate the impact of that redesign on family-centered outcomes. Methods We conducted a pre-post study of a waiting room redesign at a large, academic pediatric primary care center. Design experts sought input about an optimal waiting room from families, community partners and medical providers. Family caregivers were surveyed before and after redesign regarding perceived availability of help with social needs and access to community resources, and hospitality and feelings of stress. Pre-post differences were assessed using the Chi square or Wilcoxon rank sum test. Results The key redesign concepts that emerged included linkages to community organizations, a welcoming environment, and positive distractions for children. A total of 313 caregiver surveys were completed (pre-160; post-153). Compared to pre-redesign, caregivers surveyed post-redesign were significantly more likely to perceive the waiting room as a place to obtain help connecting to community resources and find information about clinical and educational resources (both p < 0.05). Families were also significantly more likely to report the waiting room as more welcoming and relaxing, with sufficient privacy and space (all p < 0.05). Discussion Waiting rooms, typically a place of wasted time and space, can be redesigned to enhance families' engagement and connection to community resources.

  20. The willingness to pay for wait reduction: the disutility of queues for cataract surgery in Canada, Denmark, and Spain.

    PubMed

    Bishai, D M; Lang, H C

    2000-03-01

    We estimate demand curves for a one month reduction in waiting time for cataract surgery based on survey data collected in 1992 in Manitoba, Barcelona, and Denmark. Patients answered, "Would you be willing to pay [Bid, B] to reduce your waiting time for cataract surgery to less than one month?" Controlling for SES and visual status, Barcelonan patients have greater WTP for shortened waiting time than the Danes and Manitobans. We estimate the value (in 1992 $) of lost consumer surplus due to the cataract surgery queue at $128 per patient in Manitoba, $160 in Denmark, and $243 in Barcelona.

  1. Monitoring trends in waiting periods in Canada for elective surgery: validation of a method using administrative data

    PubMed Central

    Shortt, Samuel E.D.; Shaw, Ralph A.; Elliott, David; Mackillop, William J.

    2004-01-01

    Background Provincial governments require timely, economical methods to monitor surgical waiting periods. Although use of prospective procedure-specific registers would be the ideal method, a less elaborate system has been proposed that is based on physician billing data. This study assessed the validity of using the date of the last service billed prior to surgery as a proxy for the beginning of the post-referral, pre-surgical waiting period. Method We examined charts for 31 824 elective surgical encounters between 1992 and 1996 at an Ontario teaching hospital. The date of the last service before surgery (the last billing date) was compared with the date of the consultant's letter indicating a decision to book surgery (i.e., to begin waiting). Results Several surgical specialties (but excluding cardiac, orthopedic and gynecologic) had a close correlation between the dates of the last pre-surgery visit and those of the actual decision to place the patient on the waiting list. Similar results were found for 12 of 15 individually studied procedures, including some orthopedic and gynecological procedures. Conclusion Used judiciously, billing data is a timely, inexpensive and generally accurate method by which provincial governments could monitor trends in waiting times for appropriately selected surgical procedures. PMID:15264378

  2. Obesity Surgery Score (OSS) for Prioritization in the Bariatric Surgery Waiting List: a Need of Public Health Systems and a Literature Review.

    PubMed

    Casimiro Pérez, José Antonio; Fernández Quesada, Carlos; Del Val Groba Marco, María; Arteaga González, Iván; Cruz Benavides, Francisco; Ponce, Jaime; de Pablos Velasco, Pedro; Marchena Gómez, Joaquín

    2018-04-01

    In the last decades, we have experienced an increase in the prevalence of obesity in western countries with a higher demand for bariatric surgery and consequently prolonged waiting times. Currently, in many public hospitals, the only criterion that establishes priority for bariatric surgery is waiting time regardless of obesity severity. We propose a new, simple, and homogeneous clinical prioritization system, the Obesity Surgery Score (OSS), which takes into account simultaneously and equitably the time on surgical waiting list and the obesity severity based on three variables: body mass index, obesity-related comorbidities, and functional limitations. We have reviewed the current literature related to obesity clinical staging systems, and we have carried out an analysis of our patients in waiting list and divided their characteristics according to their degree of severity (A, B, or C) in the OSS. Patients with OSS grade C have a higher mean BMI, greater severity in comorbidities, and greater socio-labor impact. The current surgery waiting time of our series is of 26 months. Currently, 27 patients (51.9%) with OSS grade B and 15 patients (51.7%) with OSS grade C have been on our waiting list for more than 1 year. Since the obesity severity, the waiting time and its clinical consequences are associated with an increase in morbidity and mortality, it is important to apply a structured prioritization system for bariatric surgery waiting list. This allows prioritization of patients at greater risk, improves patient prognosis, and optimizes costs and available health resources.

  3. Optimization of BEV Charging Strategy

    NASA Astrophysics Data System (ADS)

    Ji, Wei

    This paper presents different approaches to optimize fast charging and workplace charging strategy of battery electric vehicle (BEV) drivers. For the fast charging analysis, a rule-based model was built to simulate BEV charging behavior. Monte Carlo analysis was performed to explore to the potential range of congestion at fast charging stations which could be more than four hours at the most crowded stations. Genetic algorithm was performed to explore the theoretical minimum waiting time at fast charging stations, and it can decrease the waiting time at the most crowded stations to be shorter than one hour. A deterministic approach was proposed as a feasible suggestion that people should consider to take fast charging when the state of charge is approaching 40 miles. This suggestion is hoped to help to minimize potential congestion at fast charging stations. For the workplace charging analysis, scenario analysis was performed to simulate temporal distribution of charging demand under different workplace charging strategies. It was found that if BEV drivers charge as much as possible and as late as possible at workplace, it could increase the utility of solar-generated electricity while relieve grid stress of extra intensive electricity demand at night caused by charging electric vehicles at home.

  4. The Religious Meaning in "Waiting for Godot"

    ERIC Educational Resources Information Center

    Wang, Jing

    2011-01-01

    "Waiting for Godot" is one of the classic works of theater of the absurd. The play seems absurd but with a deep religious meaning. This text tries to explore the theme in four parts of God and man, breaking the agreement, repentance and imprecation and waiting for salvation.

  5. Outcomes of videotape instruction in clinic waiting area.

    PubMed

    Oermann, Marilyn H; Webb, Sue A; Ashare, Jo Ann

    2003-01-01

    The purpose of our study was to examine the effectiveness of general health-promotion teaching for patients in the waiting room of a clinic, using focused videotape instruction. An experimental design was used. Subjects were patients (N = 215) in the waiting rooms of clinics in a university medical center in the Midwest. Patients were randomly assigned to two groups: focused videotape instruction in the clinic (n = 106) and control (no instruction in the clinic waiting area) (n = 109). The outcome measures included patient learning about a health education topic and patient satisfaction with overall care, explanations by the provider, and education received during the clinic visit. There was a significant gain in knowledge for patients who viewed the videotape in the waiting room (t = 5.43, df = 213, p < .0001), and they were more satisfied with their education compared with the control group (t = 4.73, df = 213, p < .0001). This study supports focused video instruction as an effective and efficient teaching intervention for disseminating health information in the waiting area.

  6. Sustaining change: the imperative for patient access strategies.

    PubMed

    Glynn, Peter A R

    2006-01-01

    The paper by Trypuc, MacLeod and Hudson provides a timely and important overview of methods to sustain provincial wait time strategies. The emphasis on accountability for patient access to timely care throughout the healthcare system comes through strongly--as it should. These accountabilities are made "real" through purchase service agreements. Physician-hospital relationships are a fundamental aspect of this accountability. This commentary suggests the inclusion of two additional supporting tools in addition to those cited by the authors of the lead paper--quality monitoring and the use of industrial engineering techniques for queue management and patient flow analysis. Strong and persistent leadership of patient access strategies will ensure sustainable change.

  7. The association between waiting for psychological therapy and therapy outcomes as measured by the CORE-OM.

    PubMed

    Beck, Alison; Burdett, Mark; Lewis, Helen

    2015-06-01

    To investigate the impact of waiting for psychological therapy on client well-being as measured by the Clinical Outcomes in Routine Evaluation-Outcome Measure (CORE-OM) global distress (GD) score. Global distress scores were retrieved for all clients referred for psychological therapy in a secondary care mental health service between November 2006 and May 2013 and who had completed a CORE-OM at assessment and first session. GD scores for a subgroup of 103 clients who had completed a CORE-OM during the last therapy session were also reviewed. The study sample experienced a median wait of 41.14 weeks between assessment and first session. The relationship between wait time from referral acceptance to assessment, and assessment GD score was not significant. During the period between assessment and first session no significant difference in GD score was observed. Nevertheless 29.1% of the sample experienced reliable change; 16.0% of clients reliably improved and 13.1% reliably deteriorated whilst waiting for therapy. Demographic factors were not found to have a significant effect on the change in GD score between assessment and first session. Waiting time was associated with post-therapy outcomes but not to a degree which was meaningful. The majority of individuals (54.4%), regardless of whether they improved or deteriorated whilst waiting for therapy, showed reliable improvement at end of therapy as measured by the CORE-OM. The majority of GD scores remained stable while waiting for therapy; however, 29.1% of secondary care clients experienced either reliable improvement or deterioration. Irrespective of whether they improved, deteriorated or remained unchanged whilst waiting for therapy, most individuals who had a complete end of therapy assessment showed reliable improvements following therapy. There was no significant difference in GD score between assessment and first session recordings. A proportion of clients (29.1%) showed reliable change, either improvement or

  8. Exploring implications of Medicaid participation and wait times for colorectal screening on early detection efforts in Connecticut--a secret-shopper survey.

    PubMed

    Patel, Vatsal B; Nahar, Richa; Murray, Betty; Salner, Andrew L

    2013-04-01

    Routine colorectal screening, decreases in incidence, and advances in treatment have lowered colorectal cancer mortality rates over the past three decades. Nevertheless, it remains the second most common cause of cancer death amongst men and women combined in U.S. Most cases of colon cancer are diagnosed at a late stage leading to poor survival outcomes for patients. After extensive research of publically available data, it would appear that the state of Connecticut does not have available state-wide data on patient wait times for routine colonoscopy screening. Furthermore, there are no publicly available, or Connecticut-specific, reports on Medicaid participation rates for colorectal screening amongst gastroenterologists (GI) in Connecticut. In 2012, the American Cancer Society report on Colorectal Cancer Screening Rates confirmed barriers to health-care access and disparities in health outcomes and survival rates for colon cancer patients based on race, ethnicity, and low socioeconomic status. Given this information, one could conjecture that low Medicaid participation rates among GIs could potentially have a more severe impact on health-care access and outcomes for underserved populations. At present, funding and human resources are being employed across the state of Connecticut to address bottlenecks in colorectal cancer screening. More specifically, patient navigation and outreach programs are emerging and expanding to address the gaps in services for hard-to-reach populations and the medically underserved. Low Medicaid participation rates and increased wait times for colonoscopy screening may impair the efficacy of colorectal cancer patient navigation and outreach efforts and potentially funding for future interventions. In this study, we report the results of our secret-shopper telephone survey comprising of 93 group and independent gastroenterologist (GI) practices in different counties of Connecticut. Reviewing online resources and yellow pages

  9. Hunger, waiting time and transport costs: time to confront challenges to ART adherence in Africa.

    PubMed

    Hardon, A P; Akurut, D; Comoro, C; Ekezie, C; Irunde, H F; Gerrits, T; Kglatwane, J; Kinsman, J; Kwasa, R; Maridadi, J; Moroka, T M; Moyo, S; Nakiyemba, A; Nsimba, S; Ogenyi, R; Oyabba, T; Temu, F; Laing, R

    2007-05-01

    Adherence levels in Africa have been found to be better than those in the US. However around one out of four ART users fail to achieve optimal adherence, risking drug resistance and negative treatment outcomes. A high demand for 2nd line treatments (currently ten times more expensive than 1st line ART) undermines the sustainability of African ART programs. There is an urgent need to identify context-specific constraints to adherence and implement interventions to address them. We used rapid appraisals (involving mainly qualitative methods) to find out why and when people do not adhere to ART in Uganda, Tanzania and Botswana. Multidisciplinary teams of researchers and local health professionals conducted the studies, involving a total of 54 semi-structured interviews with health workers, 73 semi-structured interviews with ARTusers and other key informants, 34 focus group discussions, and 218 exit interviews with ART users. All the facilities studied in Botswana, Tanzania and Uganda provide ARVs free of charge, but ART users report other related costs (e.g. transport expenditures, registration and user fees at the private health facilities, and lost wages due to long waiting times) as main obstacles to optimal adherence. Side effects and hunger in the initial treatment phase are an added concern. We further found that ART users find it hard to take their drugs when they are among people to whom they have not disclosed their HIV status, such as co-workers and friends. The research teams recommend that (i) health care workers inform patients better about adverse effects; (ii) ART programmes provide transport and food support to patients who are too poor to pay; (iii) recurrent costs to users be reduced by providing three-months, rather than the one-month refills once optimal adherence levels have been achieved; and (iv) pharmacists play an important role in this follow-up care.

  10. Developmental changes in anger expression and attention focus: Learning to wait

    PubMed Central

    Cole, Pamela M.; Tan, Patricia Z.; Hall, Sarah E.; Zhang, Yiyun; Crnic, Keith A.; Blair, Clancy B.; Li, Runze

    2011-01-01

    Being able to wait is an essential part of self-regulation. The present study examined the developmental course of changes in the latency to and duration of target waiting behaviors by following 65 boys and 55 girls from rural and semi-rural economically strained homes from ages 18 to 48 months. Age-related changes in latency to and duration of children’s anger expressions and attention focus (e.g., self-initiated distraction) during an eight minute wait for a gift were found. On average, at 18 and 24 months of age, children were quick to react angrily and slower to shift attention away from the desired object than they were at later ages. Over time, children were quicker to distract themselves. By 36 months, distractions occurred before children expressed anger, and anger expressions were briefer. At 48 months, children typically made a quick bid to mother about demands of waiting before distracting themselves; on average, they did not appear angry until the latter half of the wait. Unexpectedly, children bid to their mothers as much at age 48 months as they had at 18 months; however bids became less angry as children got older. Developmental changes in distraction and bidding predicted age-related changes in the latency to anger. Findings are discussed in terms of the neurocognitive control of attention around age 30 months, the limitations of children’s self-regulatory efforts at age 48 months, and the importance of fostering children’s ability to forestall, as well as modulate, anger. PMID:21639619

  11. Waiting for transplant: physical, psychosocial, and nutritional status considerations for pediatric candidates and implications for care.

    PubMed

    Anthony, Samantha J; Annunziato, Rachel A; Fairey, Elise; Kelly, Vicky L; So, Stephanie; Wray, Jo

    2014-08-01

    The waiting period for an organ transplant has been described as a time of tremendous uncertainty and vulnerability, posing unique challenges and stressors for pediatric transplant candidates and their families. It has been identified as the most stressful stage of the transplant journey, yet little attention has been given to the physical, psychological, or social impact of the waiting period in the literature. In this review, we discuss the physical, nutritional, and psychosocial implications of the waiting period for child and adolescent transplant candidates and the impact on their parents and siblings. We identify areas for future research and provide recommendations for clinical practice to support children, adolescents, and families during the waiting period. © 2014 John Wiley & Sons A/S. Published by John Wiley & Sons Ltd.

  12. Emotion Regulation Strategies and Childhood Obesity in High Risk Preschoolers

    PubMed Central

    Power, Thomas G.; Olivera, Yadira A.; Hill, Rachael A.; Beck, Ashley D.; Hopwood, Veronica; Garcia, Karina Silva; Ramos, Guadalupe G.; Fisher, Jennifer Orlet; O’Connor, Teresia M.; Hughes, Sheryl O.

    2016-01-01

    The current study examined the relationships between the specific strategies that preschool children use to regulate their emotions and childhood weight status to see if emotion regulation strategies would predict childhood weight status over and above measures of eating self-regulation. 185 4- to 5-year-old Latino children were recruited through Head Start centers in a large city in the southeastern U.S. Children completed both a delay of gratification task (emotion regulation) and an eating in the absence of hunger task (eating regulation). Eating regulation also was assessed by maternal reports. Four emotion regulation strategies were examined in the delay of gratification task: shut out stimuli, prevent movement, distraction, and attention to reward. Hierarchical linear regressions predicting children’s weight status showed that both measures of eating regulation negatively predicted child obesity, and the use of prevent movement negatively predicted child obesity. Total wait time during the delay of gratification tasks was not a significant predictor. The current findings are consistent with studies showing that for preschool children, summary measures of emotion regulation (e.g., wait time) are not concurrently associated with child obesity. In contrast, the use of emotion regulation strategies was a significant predictor of lower child weight status. These findings help identify emotion regulation strategies that prevention programs can target for helping children regulate their emotions and decrease their obesity risk. PMID:27620645

  13. Coping with worry while waiting for diagnostic results: a qualitative study of the experiences of pregnant couples following a high-risk prenatal screening result.

    PubMed

    Lou, Stina; Nielsen, Camilla P; Hvidman, Lone; Petersen, Olav B; Risør, Mette B

    2016-10-21

    It is well documented that pregnant women experience increased worry and uncertainty following a high-risk prenatal screening result. While waiting for diagnostic results this worry continues to linger. It has been suggested that high-risk women put the pregnancy mentally 'on hold' during this period, however, not enough is known about how high-risk women and their partners cope while waiting for diagnostic results. The aim of this study was to identify the strategies employed to cope with worry and uncertainty. Qualitative, semi-structured interviews with 16 high-risk couples who underwent diagnostic testing. The couples were recruited at a university hospital fetal medicine unit in Denmark. Data were analysed using thematic analysis. All couples reported feeling worried and sad upon receiving a high-risk screening result. While waiting for diagnostic results, the couples focused on coming to their own understanding of the situation and employed both social withdrawal and social engagement as strategies to prevent worry from escalating. Additionally, couples used gratitude, reassuring reasoning and selective memory as means to maintain hopes for a good outcome. Discussions about what to do in case of an abnormal test result were notably absent in the accounts of waiting. This bracketing of the potential abnormal result allowed the couples to hold on to a 'normal' pregnancy and to employ an 'innocent-till-proven-guilty' approach to their worries about the fetus's health. None of the interviewed couples regretted having prenatal screening and all of them expected to have prenatal screening in a future pregnancy. The couples in this study did not put the pregnancy mentally 'on hold'. Worry and uncertainty must be understood as managed through a diverse range of practical and emotional strategies that change and overlap in the process of waiting. Clinicians may support appropriate ways of coping with worry and waiting through empathetic and empowering clinical

  14. Toward Implementing Patient Flow in a Cancer Treatment Center to Reduce Patient Waiting Time and Improve Efficiency.

    PubMed

    Suss, Samuel; Bhuiyan, Nadia; Demirli, Kudret; Batist, Gerald

    2017-06-01

    Outpatient cancer treatment centers can be considered as complex systems in which several types of medical professionals and administrative staff must coordinate their work to achieve the overall goals of providing quality patient care within budgetary constraints. In this article, we use analytical methods that have been successfully employed for other complex systems to show how a clinic can simultaneously reduce patient waiting times and non-value added staff work in a process that has a series of steps, more than one of which involves a scarce resource. The article describes the system model and the key elements in the operation that lead to staff rework and patient queuing. We propose solutions to the problems and provide a framework to evaluate clinic performance. At the time of this report, the proposals are in the process of implementation at a cancer treatment clinic in a major metropolitan hospital in Montreal, Canada.

  15. Plants are not sitting ducks waiting for herbivores to eat them.

    PubMed

    Lev-Yadun, Simcha

    2016-05-03

    There is a common attitude toward plants, accordingly, plants are waiting around to be found and eaten by herbivores. This common approach toward plants is a great underestimation of the huge and variable arsenal of defensive plant strategies. Plants do everything evolution has allowed them to do in order not to be eaten. Therefore, plants are not sitting ducks and many plants outsmart and even exploit many invertebrate and vertebrate herbivores and carnivores for pollination and for seed dispersal, and even carnivores and parasitoids for defense.

  16. Children's preferences concerning ambiance of dental waiting rooms.

    PubMed

    Panda, A; Garg, I; Shah, M

    2015-02-01

    Despite many advances in paediatric dentistry, the greatest challenge for any paediatric dentist is to remove the anxiety related to a dental visit and have a child patient to accept dental treatment readily. Minor changes made in the waiting room design can have a major effect on the way any child perceives the upcoming dental experience. This study was carried out to determine children's preferences regarding the dental waiting area so as to improve their waiting experience and reduce their preoperative anxiety before a dental appointment. This was a cross-sectional descriptive study using survey methodology. A questionnaire designed to evaluate children's preferences regarding the waiting room was distributed to new paediatric patients, aged between 6 and 11 years of age, attending an outpatient dental facility and was completed by 212 children (127 males, 85 females). The analyses were carried out on cross-tables using Phi (for 2×2 tables) or Cramer's V (for larger than 2×2 tables) to assess responses to the questionnaire items across age groups and gender. A majority of children preferred music and the ability to play in a waiting room. They also preferred natural light and walls with pictures. They preferred looking at an aquarium or a television and sitting on beanbags and chairs and also preferred plants and oral hygiene posters Repetious. The results obtained from this study may help the dental team decide on an appropriate design of their paediatric waiting room so as to make children comfortable in the dental environment and improve delivery of health care.

  17. Waiting for Radiology Test Results: Patient Expectations and Emotional Disutility.

    PubMed

    Woolen, Sean; Kazerooni, Ella A; Wall, Amber; Parent, Kelly; Cahalan, Shannon; Alameddine, Mitchell; Davenport, Matthew S

    2018-02-01

    To measure patient willingness to wait and emotional disutility of waiting for outpatient imaging test results. A prospective HIPAA-compliant multicenter outpatient quality improvement survey was administered by a trained interviewer to 218 outpatients from November 1, 2016, to February 1, 2017. The survey was vetted by patient- and family-centered care advocates with experience in survey design and underwent precognitive testing for readability. Six clinical scenarios were tested. Descriptive statistics were calculated. The response (93% [202 of 218]) and completion (93% [188 of 202]) rates were excellent. Anxiety (28% [57 of 202]), depression (26% [53 of 202]), and cancer (23% [46 of 202]) histories were common. Median stated expectations for imaging test results receipt were 3 days after a screening examination (interquartile range [IQR] 5 days); 2 days after chest x-ray for chest pain (IQR 3) or MRI or CT for back pain (IQR 2); and 1 day after chest x-ray for pneumonia (IQR 2), MRI or CT for brain tumor (IQR 2), or CT for cancer treatment (IQR 3). If imaging results are not received, the median time patients stated they would wait to call their provider was 1 to 5 days (varied by indication). Waiting for imaging results exerts an emotional change in 45% (91 of 202) of individuals, with the majority (85% [77 of 91]) experiencing anxiety (minimal 28%, mild 45%, moderate 22%, severe 4%, extreme 1%). Patients expect outpatient imaging results within 1 to 3 days and will call providers by 1 to 5 days. Waiting for test results commonly induces anxiety. Copyright © 2017 American College of Radiology. Published by Elsevier Inc. All rights reserved.

  18. Interior view; Street Car Waiting House North Philadelphia Station, ...

    Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey

    Interior view; Street Car Waiting House - North Philadelphia Station, Street Car Waiting House, 2900 North Broad Street, on northwest corner of Broad Street & Glenwood Avenue, Philadelphia, Philadelphia County, PA

  19. Towards decision support for waiting lists: an operations management view.

    PubMed

    Vissers, J M; Van Der Bij, J D; Kusters, R J

    2001-06-01

    This paper considers the phenomenon of waiting lists in a healthcare setting, which is characterised by limitations on the national expenditure, to explore the potentials of an operations management perspective. A reference framework for waiting list management is described, distinguishing different levels of planning in healthcare--national, regional, hospital and process--that each contributes to the existence of waiting lists through managerial decision making. In addition, different underlying mechanisms in demand and supply are distinguished, which together explain the development of waiting lists. It is our contention that within this framework a series of situation specific models should be designed to support communication and decision making. This is illustrated by the modelling of the demand for cataract treatment in a regional setting in the south-eastern part of the Netherlands. An input-output model was developed to support decisions regarding waiting lists. The model projects the demand for treatment at a regional level and makes it possible to evaluate waiting list impacts for different scenarios to meet this demand.

  20. Detail; Street Car Waiting House window, north wall North ...

    Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey

    Detail; Street Car Waiting House window, north wall - North Philadelphia Station, Street Car Waiting House, 2900 North Broad Street, on northwest corner of Broad Street & Glenwood Avenue, Philadelphia, Philadelphia County, PA

  1. West view; Street Car Waiting House, east elevation North ...

    Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey

    West view; Street Car Waiting House, east elevation - North Philadelphia Station, Street Car Waiting House, 2900 North Broad Street, on northwest corner of Broad Street & Glenwood Avenue, Philadelphia, Philadelphia County, PA

  2. Cost-Effectiveness of Reduced Waiting Time for Head and Neck Cancer Patients due to a Lean Process Redesign.

    PubMed

    Simons, Pascale A M; Ramaekers, Bram; Hoebers, Frank; Kross, Kenneth W; Marneffe, Wim; Pijls-Johannesma, Madelon; Vandijck, Dominique

    2015-07-01

    Compared with new technologies, the redesign of care processes is generally considered less attractive to improve patient outcomes. Nevertheless, it might result in better patient outcomes, without further increasing costs. Because early initiation of treatment is of vital importance for patients with head and neck cancer (HNC), these care processes were redesigned. This study aimed to assess patient outcomes and cost-effectiveness of this redesign. An economic (Markov) model was constructed to evaluate the biopsy process of suspicious lesion under local instead of general anesthesia, and combining computed tomography and positron emission tomography for diagnostics and radiotherapy planning. Patients treated for HNC were included in the model stratified by disease location (larynx, oropharynx, hypopharynx, and oral cavity) and stage (I-II and III-IV). Probabilistic sensitivity analyses were performed. Waiting time before treatment start reduced from 5 to 22 days for the included patient groups, resulting in 0.13 to 0.66 additional quality-adjusted life-years. The new workflow was cost-effective for all the included patient groups, using a ceiling ratio of €80,000 or €20,000. For patients treated for tumors located at the larynx and oral cavity, the new workflow resulted in additional quality-adjusted life-years, and costs decreased compared with the regular workflow. The health care payer benefited €14.1 million and €91.5 million, respectively, when individual net monetary benefits were extrapolated to an organizational level and a national level. The redesigned care process reduced the waiting time for the treatment of patients with HNC and proved cost-effective. Because care improved, implementation on a wider scale should be considered. Copyright © 2015 International Society for Pharmacoeconomics and Outcomes Research (ISPOR). Published by Elsevier Inc. All rights reserved.

  3. Improving Customer Waiting Time at a DMV Center Using Discrete-Event Simulation

    NASA Technical Reports Server (NTRS)

    Arnaout, Georges M.; Bowling, Shannon

    2010-01-01

    Virginia's Department of Motor Vehicles (DMV) serves a customer base of approximately 5.6 million licensed drivers and ID card holders and 7 million registered vehicle owners. DMV has more daily face-to-face contact with Virginia's citizens than any other state agency [1]. The DMV faces a major difficulty in keeping up with the excessively large customers' arrival rate. The consequences are queues building up, stretching out to the entrance doors (and sometimes even outside) and customers complaining. While the DMV state employees are trying to serve at their fastest pace, the remarkably large queues indicate that there is a serious problem that the DMV faces in its services, which must be dealt with rapidly. Simulation is considered as one of the best tools for evaluating and improving complex systems. In this paper, we use it to model one of the DMV centers located in Norfolk, VA. The simulation model is modeled in Arena 10.0 from Rockwell systems. The data used is collected from experts of the DMV Virginia headquarter located in Richmond. The model created was verified and validated. The intent of this study is to identify key problems causing the delays at the DMV centers and suggest possible solutions to minimize the customers' waiting time. In addition, two tentative hypotheses aiming to improve the model's design are tested and validated.

  4. North view; Street Car Waiting House, south (front) elevation ...

    Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey

    North view; Street Car Waiting House, south (front) elevation - North Philadelphia Station, Street Car Waiting House, 2900 North Broad Street, on northwest corner of Broad Street & Glenwood Avenue, Philadelphia, Philadelphia County, PA

  5. A web-based appointment system to reduce waiting for outpatients: a retrospective study.

    PubMed

    Cao, Wenjun; Wan, Yi; Tu, Haibo; Shang, Fujun; Liu, Danhong; Tan, Zhijun; Sun, Caihong; Ye, Qing; Xu, Yongyong

    2011-11-22

    Long waiting times for registration to see a doctor is problematic in China, especially in tertiary hospitals. To address this issue, a web-based appointment system was developed for the Xijing hospital. The aim of this study was to investigate the efficacy of the web-based appointment system in the registration service for outpatients. Data from the web-based appointment system in Xijing hospital from January to December 2010 were collected using a stratified random sampling method, from which participants were randomly selected for a telephone interview asking for detailed information on using the system. Patients who registered through registration windows were randomly selected as a comparison group, and completed a questionnaire on-site. A total of 5641 patients using the online booking service were available for data analysis. Of them, 500 were randomly selected, and 369 (73.8%) completed a telephone interview. Of the 500 patients using the usual queuing method who were randomly selected for inclusion in the study, responses were obtained from 463, a response rate of 92.6%. Between the two registration methods, there were significant differences in age, degree of satisfaction, and total waiting time (P<0.001). However, gender, urban residence, and valid waiting time showed no significant differences (P>0.05). Being ignorant of online registration, not trusting the internet, and a lack of ability to use a computer were three main reasons given for not using the web-based appointment system. The overall proportion of non-attendance was 14.4% for those using the web-based appointment system, and the non-attendance rate was significantly different among different hospital departments, day of the week, and time of the day (P<0.001). Compared to the usual queuing method, the web-based appointment system could significantly increase patient's satisfaction with registration and reduce total waiting time effectively. However, further improvements are needed for broad

  6. Emotion regulation strategies and childhood obesity in high risk preschoolers.

    PubMed

    Power, Thomas G; Olivera, Yadira A; Hill, Rachael A; Beck, Ashley D; Hopwood, Veronica; Garcia, Karina Silva; Ramos, Guadalupe G; Fisher, Jennifer Orlet; O'Connor, Teresia M; Hughes, Sheryl O

    2016-12-01

    The current study examined the relationships between the specific strategies that preschool children use to regulate their emotions and childhood weight status to see if emotion regulation strategies would predict childhood weight status over and above measures of eating self-regulation. 185 4- to 5-year-old Latino children were recruited through Head Start centers in a large city in the southeastern U.S. Children completed both a delay of gratification task (emotion regulation) and an eating in the absence of hunger task (eating regulation). Eating regulation also was assessed by maternal reports. Four emotion regulation strategies were examined in the delay of gratification task: shut out stimuli, prevent movement, distraction, and attention to reward. Hierarchical linear regressions predicting children's weight status showed that both measures of eating regulation negatively predicted child obesity, and the use of prevent movement negatively predicted child obesity. Total wait time during the delay of gratification tasks was not a significant predictor. The current findings are consistent with studies showing that for preschool children, summary measures of emotion regulation (e.g., wait time) are not concurrently associated with child obesity. In contrast, the use of emotion regulation strategies was a significant predictor of lower child weight status. These findings help identify emotion regulation strategies that prevention programs can target for helping children regulate their emotions and decrease their obesity risk. Copyright © 2016 Elsevier Ltd. All rights reserved.

  7. [Criteria for prioritising patients on surgical waiting lists in the National Health System].

    PubMed

    Allepuz, A; Espallargues, M; Martínez, O

    2009-01-01

    To survey the importance of previously proposed criteria for prioritising patients on surgical waiting lists and to analyse their use in daily practice. Cross-sectional study through a self-auto-administered postal questionnaire to hospital managers, medical directors, admissions managers, and department heads of general surgery, ophthalmology, orthopaedics and traumatology surgery and vascular surgery from 139 centres. The questionnaire comprised 3 sections: a) 3 to 5 of the most important criteria had to be selected and their use in daily practice had to be assessed; b) new criteria were proposed, c) socio-demographic data. The mean and its standard deviation of each criterion of importance were calculated. The proposed criteria were categorised and their frequency was calculated. The questionnaire was answered by the 22% of those surveyed. Disease severity, speed of progression, waiting time and pain were the criteria considered as most important and were the most used. The current clinical situation and the professional environment were the two most common categories defined from the criteria proposed by those surveyed. The surgical priority should be determined by other criteria related to surgery necessity besides waiting time. Establishing prioritisation criteria could enable current implicit criteria to be used explicitly.

  8. Where does the waiting list begin? A short review of the dynamics and organization of modern waiting lists.

    PubMed

    Rotstein, Dalia L; Alter, David A

    2006-06-01

    Waiting for medical care is the by-product of system rationing, where demand exceeds supply. In this short report we expand on the conventional concept of the queue, by focusing on the regulation of demand and by incorporating a funnel and spout analogy. Real-world examples are used to illustrate the infancy of funnel or demand-side reform initiatives targeting the queue, and the suggestion is made that policy needs to address the concept of 'waiting' much earlier in the treatment cycle.

  9. Waiting for Merlot: anticipatory consumption of experiential and material purchases.

    PubMed

    Kumar, Amit; Killingsworth, Matthew A; Gilovich, Thomas

    2014-10-01

    Experiential purchases (money spent on doing) tend to provide more enduring happiness than material purchases (money spent on having). Although most research comparing these two types of purchases has focused on their downstream hedonic consequences, the present research investigated hedonic differences that occur before consumption. We argue that waiting for experiences tends to be more positive than waiting for possessions. Four studies demonstrate that people derive more happiness from the anticipation of experiential purchases and that waiting for an experience tends to be more pleasurable and exciting than waiting to receive a material good. We found these effects in studies using questionnaires involving a variety of actual planned purchases, in a large-scale experience-sampling study, and in an archival analysis of news stories about people waiting in line to make a purchase. Consumers derive value from anticipation, and that value tends to be greater for experiential than for material purchases. © The Author(s) 2014.

  10. Waiting for Water

    ERIC Educational Resources Information Center

    Lamson-Nussbaum, Jorie

    2013-01-01

    The author waits in the hot and oppressive air while dust devils are born and die over the newly plowed field. It is a dry spring and she prays for rain. The lupine beans withered to dry threads last week and the corn that sprouted in a green haze over the north field is turning to brown paper. However, driving north, the author discovers the Rum…

  11. Mice plan decision strategies based on previously learned time intervals, locations, and probabilities

    PubMed Central

    Tosun, Tuğçe; Gür, Ezgi; Balcı, Fuat

    2016-01-01

    Animals can shape their timed behaviors based on experienced probabilistic relations in a nearly optimal fashion. On the other hand, it is not clear if they adopt these timed decisions by making computations based on previously learnt task parameters (time intervals, locations, and probabilities) or if they gradually develop their decisions based on trial and error. To address this question, we tested mice in the timed-switching task, which required them to anticipate when (after a short or long delay) and at which of the two delay locations a reward would be presented. The probability of short trials differed between test groups in two experiments. Critically, we first trained mice on relevant task parameters by signaling the active trial with a discriminative stimulus and delivered the corresponding reward after the associated delay without any response requirement (without inducing switching behavior). During the test phase, both options were presented simultaneously to characterize the emergence and temporal characteristics of the switching behavior. Mice exhibited timed-switching behavior starting from the first few test trials, and their performance remained stable throughout testing in the majority of the conditions. Furthermore, as the probability of the short trial increased, mice waited longer before switching from the short to long location (experiment 1). These behavioral adjustments were in directions predicted by reward maximization. These results suggest that rather than gradually adjusting their time-dependent choice behavior, mice abruptly adopted temporal decision strategies by directly integrating their previous knowledge of task parameters into their timed behavior, supporting the model-based representational account of temporal risk assessment. PMID:26733674

  12. Waiting Lists for Radiation Therapy: A Case Study

    PubMed Central

    2001-01-01

    Background Why waiting lists arise and how to address them remains unclear, and an improved understanding of these waiting list "dynamics" could lead to better management. The purpose of this study is to understand how the current shortage in radiation therapy in Ontario developed; the implications of prolonged waits; who is held accountable for managing such delays; and short, intermediate, and long-term solutions. Methods A case study of the radiation therapy shortage in 1998-99 at Princess Margaret Hospital, Toronto, Ontario, Canada. Relevant documents were collected; semi-structured, face-to-face interviews with ten administrators, health care workers, and patients were conducted, audio-taped and transcribed; and relevant meetings were observed. Results The radiation therapy shortage arose from a complex interplay of factors including: rising cancer incidence rates; broadening indications for radiation therapy; human resources management issues; government funding decisions; and responsiveness to previous planning recommendations. Implications of delays include poorer cancer control rates; patient suffering; and strained doctor-patient relationships. An incompatible relationship exists between moral responsibility, borne by government, and legal liability, borne by physicians. Short-term solutions include re-referral to centers with available resources; long-term solutions include training and recruiting health care workers, improving workload standards, increasing compensation, and making changes to the funding formula. Conclusion Human resource planning plays a critical role in the causes and solutions of waiting lists. Waiting lists have harsh implications for patients. Accountability relationships require realignment. PMID:11319944

  13. Wait watchers: the application of a waiting list active management program in ambulatory care.

    PubMed

    de Belvis, Antonio Giulio; Marino, Marta; Avolio, Maria; Pelone, Ferruccio; Basso, Danila; Dei Tos, Gian Antonio; Cinquetti, Sandro; Ricciardi, Walter

    2013-04-01

    This study describes and evaluates the application of a waiting list management program in ambulatory care. Waiting list active management survey (telephone call and further contact); before and after controlled trial. Local Health Trust in Veneto Region (North-East of Italy) in 2008-09. Five hundred and one people on a 554 waiting list for C Class ambulatory care diagnostic and/or clinical investigations (electrocardiography plus cardiology ambulatory consultation, eye ambulatory consultation, carotid vessels Eco-color-Doppler, legs Eco-color-Doppler or colonoscopy, respectively). Active list management program consisting of a telephonic interview on 21 items to evaluate socioeconomic features, self-perceived health status, social support, referral physician, accessibility and patients' satisfaction. A controlled before-and-after study was performed to evaluate anonymously the overall impact on patients' self-perceived quality of care. The rate of patients with deteriorating healthcare conditions; rate of dropout; interviewed degree of satisfaction about the initiative; overall impact on citizens' perceived quality of care. 95.4% patients evaluated the initiative as useful. After the intervention, patients more likely to have been targeted with the program showed a statistically significant increase in self-reported quality of care. Positive impact of the program on some dimensions of ambulatory care quality (health status, satisfaction, willingness to remain in the queue), thus confirming the outstanding value of 'not to leave people alone' and 'not to leave them feeling themselves alone' in healthcare delivery.

  14. Traffic-related air pollution in the community of San Ysidro, CA, in relation to northbound vehicle wait times at the US-Mexico border Port of Entry

    NASA Astrophysics Data System (ADS)

    Quintana, Penelope J. E.; Dumbauld, Jill J.; Garnica, Lynelle; Chowdhury, M. Zohir; Velascosoltero, José; Mota-Raigoza, Arturo; Flores, David; Rodríguez, Edgar; Panagon, Nicolas; Gamble, Jamison; Irby, Travis; Tran, Cuong; Elder, John; Galaviz, Vanessa E.; Hoffman, Lisa; Zavala, Miguel; Molina, Luisa T.

    2014-05-01

    The San Diego/Tijuana US-Mexico border crossing at the San Ysidro Port of Entry (POE) is the world's busiest international land border crossing (GSA, 2013). San Ysidro, California, is the US community immediately adjacent to the border crossing. More than 90% of San Ysidro residents are Hispanic, and the average household income is less than 60% of the San Diego regional average. This study investigated the San Ysidro POE as a source of traffic-related air pollutants in San Ysidro, especially in relation to wind direction and northbound vehicle wait times. The pollutants ultrafine particulate matter (UFP), black carbon (BC), and particulate matter <2.5 μm in diameter (PM2.5) were periodically sampled through the course of 2010 at four rooftop locations: one commercial establishment near the POE, two elementary schools in San Ysidro, and a coastal estuary reference site. Weather data from two nearby sites and northbound border wait times were also collected. Results indicate consistently higher daytime BC and UFP concentrations at the measurement sites near the POE. Pollution concentrations were higher during low wind speeds or when wind was blowing from the POE towards San Ysidro. In February, March and November measurements, black carbon pollution appeared to be significantly positively associated with the POE northbound wait times when the wind direction was blowing from the POE towards San Ysidro or during low wind speeds, but not when the wind direction was from the west/northwest towards the POE. This pilot study is the first to investigate the potential effect of the POE, especially the long northbound traffic delays, on the nearby community of San Ysidro. Disparities in traffic exposures are an environmental justice issue and this should be taken into account during planning and operation of POEs.

  15. ON THE BRIGHTNESS AND WAITING-TIME DISTRIBUTIONS OF A TYPE III RADIO STORM OBSERVED BY STEREO/WAVES

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

    Eastwood, J. P.; Hudson, H. S.; Krucker, S.

    2010-01-10

    Type III solar radio storms, observed at frequencies below {approx}16 MHz by space-borne radio experiments, correspond to the quasi-continuous, bursty emission of electron beams onto open field lines above active regions. The mechanisms by which a storm can persist in some cases for more than a solar rotation whilst exhibiting considerable radio activity are poorly understood. To address this issue, the statistical properties of a type III storm observed by the STEREO/WAVES radio experiment are presented, examining both the brightness distribution and (for the first time) the waiting-time distribution (WTD). Single power-law behavior is observed in the number distribution asmore » a function of brightness; the power-law index is {approx}2.1 and is largely independent of frequency. The WTD is found to be consistent with a piecewise-constant Poisson process. This indicates that during the storm individual type III bursts occur independently and suggests that the storm dynamics are consistent with avalanche-type behavior in the underlying active region.« less

  16. Waiting in the surgery.

    PubMed

    Fry, F

    1994-07-01

    The concise Oxford English Dictionary defines 'dilemma' as an argument forcing one to choose one of two alternatives, both of which are unfavourable. This is a situation that frequently confronts the general practitioner. This paper will present one practitioner's view on the subject of patients waiting to see the doctor.

  17. Local Recurrence After Complete Clinical Response and Watch and Wait in Rectal Cancer After Neoadjuvant Chemoradiation: Impact of Salvage Therapy on Local Disease Control

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

    Habr-Gama, Angelita, E-mail: gamange@uol.com.br; University of São Paulo School of Medicine, São Paulo; Gama-Rodrigues, Joaquim

    Purpose: To review the risk of local recurrence and impact of salvage therapy after Watch and Wait for rectal cancer with complete clinical response (cCR) after chemoradiation therapy (CRT). Methods and Materials: Patients with cT2-4N0-2M0 distal rectal cancer treated with CRT (50.4-54 Gy + 5-fluorouracil-based chemotherapy) and cCR at 8 weeks were included. Patients with cCR were enrolled in a strict follow-up program with no immediate surgery (Watch and Wait). Local recurrence-free survival was compared while taking into account Watch and Wait strategy alone and Watch and Wait plus salvage. Results: 90 of 183 patients experienced cCR at initial assessment after CRT (49%). Whenmore » early tumor regrowths (up to and including the initial 12 months of follow-up) and late recurrences were considered together, 28 patients (31%) experienced local recurrence (median follow-up time, 60 months). Of those, 26 patients underwent salvage therapy, and 2 patients were not amenable to salvage. In 4 patients, local re-recurrence developed after Watch and Wait plus salvage. The overall salvage rate for local recurrence was 93%. Local recurrence-free survival at 5 years was 69% (all local recurrences) and 94% (after salvage procedures). Thirteen patients (14%) experienced systemic recurrence. The 5-year cancer-specific overall survival and disease-free survival for all patients (including all recurrences) were 91% and 68%, respectively. Conclusions: Local recurrence may develop in 31% of patients with initial cCR when early regrowths (≤12 months) and late recurrences are grouped together. More than half of these recurrences develop within 12 months of follow-up. Salvage therapy is possible in ≥90% of recurrences, leading to 94% local disease control, with 78% organ preservation.« less

  18. Assessing the performance of centralized waiting lists for patients without a regular family physician using clinical-administrative data.

    PubMed

    Breton, Mylaine; Smithman, Mélanie Ann; Brousselle, Astrid; Loignon, Christine; Touati, Nassera; Dubois, Carl-Ardy; Nour, Kareen; Boivin, Antoine; Berbiche, Djamal; Roberge, Danièle

    2017-01-05

    priority to vulnerable patients and attaching of a large number of patients. Results also showed heterogeneity in the performance of centralized waiting lists across Quebec. Finally, our findings suggest it is critical that similar mechanisms should use available data to identify the best strategies for reducing variations and improving performance.

  19. Mass Measurements beyond the Major r-Process Waiting Point {sup 80}Zn

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

    Baruah, S.; Herlert, A.; Schweikhard, L.

    2008-12-31

    High-precision mass measurements on neutron-rich zinc isotopes {sup 71m,72-81}Zn have been performed with the Penning trap mass spectrometer ISOLTRAP. For the first time, the mass of {sup 81}Zn has been experimentally determined. This makes {sup 80}Zn the first of the few major waiting points along the path of the astrophysical rapid neutron-capture process where neutron-separation energy and neutron-capture Q-value are determined experimentally. The astrophysical conditions required for this waiting point and its associated abundance signatures to occur in r-process models can now be mapped precisely. The measurements also confirm the robustness of the N=50 shell closure for Z=30.

  20. Traffic pollutant exposures experienced by pedestrians waiting to enter the U.S. at a major U.S.-Mexico border crossing

    NASA Astrophysics Data System (ADS)

    Galaviz, V. E.; Yost, M. G.; Simpson, C. D.; Camp, J. E.; Paulsen, M. H.; Elder, J. P.; Hoffman, L.; Flores, D.; Quintana, P. J. E.

    2014-05-01

    Pedestrians waiting to cross into the US from Mexico at Ports of Entry experience long wait times near idling vehicles. The near-road environment is associated with elevated pollutant levels and adverse health outcomes. This is the first exposure assessment conducted to quantify northbound pedestrian commuter exposure to traffic-related air pollutants at the U.S.-Mexico border San Ysidro Port of Entry (SYPOE). Seventy-three persons who regularly crossed the SYPOE in the pedestrian line and 18 persons who did not cross were recruited to wear personal air monitors for 24-h to measure traffic pollutants particulate matter less than 2.5 μm (PM2.5), 1-nitropyrene (1-NP) - a marker for diesel exhaust - and carbon monoxide (CO). Fixed site concentrations were collected at SYPOE and occurred during the time subjects were crossing northbound to approximate their exposure to 1-NP, ultrafine particles (UFP), PM2.5, CO, and black carbon (BC) while standing in line during their border wait. Subjects who crossed the border in pedestrian lanes had a 6-fold increase in exposure to 1-NP, a 3-fold increase in exposure to CO, and a 2-fold increase in exposure to gravimetric PM2.5, vs. non-border commuters. Univariate regression analysis for UFP (median 40,000 # cm-3) found that border wait time for vehicles explained 21% of variability and relative humidity 13%, but when modeled together neither predictor remained significant. Concentrations at the SYPOE of UFP, PM2.5, CO, and BC are similar to those in other near-roadway studies that show associations with acute and chronic adverse health effects. Although results are limited by small sample numbers, these findings warrant concern for adverse health effects experienced by pedestrian commuters waiting in a long northbound queue at SYPOE and demonstrates a potential health benefit of reduced wait times at the border.

  1. Time management strategies for research productivity.

    PubMed

    Chase, Jo-Ana D; Topp, Robert; Smith, Carol E; Cohen, Marlene Z; Fahrenwald, Nancy; Zerwic, Julie J; Benefield, Lazelle E; Anderson, Cindy M; Conn, Vicki S

    2013-02-01

    Researchers function in a complex environment and carry multiple role responsibilities. This environment is prone to various distractions that can derail productivity and decrease efficiency. Effective time management allows researchers to maintain focus on their work, contributing to research productivity. Thus, improving time management skills is essential to developing and sustaining a successful program of research. This article presents time management strategies addressing behaviors surrounding time assessment, planning, and monitoring. Herein, the Western Journal of Nursing Research editorial board recommends strategies to enhance time management, including setting realistic goals, prioritizing, and optimizing planning. Involving a team, problem-solving barriers, and early management of potential distractions can facilitate maintaining focus on a research program. Continually evaluating the effectiveness of time management strategies allows researchers to identify areas of improvement and recognize progress.

  2. Detail; Street Car Waiting House, support for exterior light fixture ...

    Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey

    Detail; Street Car Waiting House, support for exterior light fixture - North Philadelphia Station, Street Car Waiting House, 2900 North Broad Street, on northwest corner of Broad Street & Glenwood Avenue, Philadelphia, Philadelphia County, PA

  3. The cost-effectiveness of immediate treatment or watch and wait with deferred chemotherapy for advanced asymptomatic follicular lymphoma.

    PubMed

    Prettyjohns, Matthew; Hoskin, Peter; McNamara, Christopher; Linch, David

    2018-01-01

    Recent evidence has shown that immediate treatment with rituximab induction, with and without maintenance, substantially reduces the need for further treatment in patients with advanced asymptomatic follicular lymphoma. This analysis estimates the cost-effectiveness of immediate treatment approaches in comparison to a watch and wait approach from the perspective of the UK National Health Service. A Markov decision model was developed to estimate the cost-effectiveness of treatment strategies in patients with asymptomatic follicular lymphoma. The model was populated using effectiveness data from a systematic literature review with the key clinical data sourced from a randomised trial, in which the treatment strategies were compared. Costs were estimated using UK national sources. In comparison to watchful waiting, both rituximab strategies were found to be more effective and cost saving. In comparison to rituximab induction, the addition of rituximab maintenance marginally increased effectiveness but substantially increased costs, resulting in an incremental cost-effectiveness ratio (ICER) of £69 406 per quality-adjusted life year (QALY). In probabilistic sensitivity analysis, rituximab induction was found to have a 68% probability of being cost-effective at a threshold of £20 000 per QALY. In conclusion, active treatment with rituximab induction is a cost-effective strategy to adopt in patients with asymptomatic follicular lymphoma. © 2017 John Wiley & Sons Ltd.

  4. 8 CFR 207.5 - Waiting lists and priority handling.

    Code of Federal Regulations, 2010 CFR

    2010-01-01

    ... REFUGEES § 207.5 Waiting lists and priority handling. Waiting lists are maintained for each designated refugee group of special humanitarian concern. Each applicant whose application is accepted for filing by... filing is the priority date for purposes of case control. Refugees or groups of refugees may be selected...

  5. Challenging "Waiting for Superman"

    ERIC Educational Resources Information Center

    Bruhn, Molle

    2014-01-01

    A group of New York City public school teachers, angry about the depiction of public schools in 'Waiting for Superman," decide to make their own film about the realities of the current education reform movement. They persevered even though they had no budget when they started and lacked a background in filmmaking. "The Inconvenient Truth…

  6. Financial cost to institutions on patients waiting for gall bladder disease surgery.

    PubMed

    Waqas, Ahmed; Qasmi, Shahzad Ahmed; Kiani, Faran; Raza, Ahmed; Khan, Khizar Ishtiaque; Manzoor, Shazia

    2014-01-01

    The aim of this study was to determine the financial costs to institution on patients waiting for gall bladder disease surgery and suggest measures to reduce them. This multi-centre prospective descriptive survey was performed on all patients who underwent an elective cholecystectomy by three consultants at secondary care hospitals in Pakistan between Jan 2010 to Jan 2012. Data was collected on demographics, the duration of mean waiting time, specific indications and nature of disease for including the patients in the waiting list, details of emergency re-admissions while awaiting surgery, investigations done, treatment given and expenditures incurred on them during these episodes. A total of 185 patients underwent elective open cholecystectomy. The indications for listing the patients for surgery were biliary colic in 128 patients (69%), acute cholecystitis in 43 patients (23%), obstructive jaundice in 8 patients (4.5%) and acute pancreatitis in 6 patients (3.2%). 146 (78.9%) and 39 (21.1%) of patients were listed as outdoor electives and indoor emergencies respectively. Of the 185 patients, 54 patients (29.2%) were re-admitted. Financial costs in Pakistani rupees per episode of readmission were 23050 per episode in total and total money spent on all readmissions was Rs. 17,05,700/-. Financial costs on health care institutions due to readmissions in patients waiting for gall bladder disease surgery are high. Identifying patients at risk for these readmissions and offering them early laparoscopic cholecystectomy is very important.

  7. Abuse Pattern of Toluene Exposure Alters Mouse Behavior in a Waiting-for-Reward Operant Task

    PubMed Central

    Bowen, Scott E.; McDonald, Phillip

    2009-01-01

    Inhaling solvents for recreational purposes continues to be a world-wide public health concern. Toluene, a volatile solvent in many abused products, adversely affects the central nervous system. However, the long-term neurobehavioral effects of exposure to high-concentration, binge patterns typical of toluene abuse remain understudied. We studied the behavioral effects of repeated toluene exposure on cognitive function following binge toluene exposure on behavioral impulse control in Swiss Webster mice using a “wait-for-reward” operant task. Mice were trained on a fixed-ratio (FR) schedule using sweetened milk as a reward. Upon achieving FR15, a wait component was added which delivered free rewards in the absence of responses at increasing time intervals (2 sec, 4 sec, 6 sec, etc…). Mice continued to receive free rewards until they pressed a lever that reinstated the FR component (FR Reset). Once proficient in the FR-Wait task, mice were exposed to either 1,000 ppm, 3,600 ppm or 6,000 ppm toluene, or 0 ppm (air controls) for 30 min per day for 40 days. To avoid acute effects of toluene exposure, behavior was assessed 23 hours later. Repeated toluene exposure decreased response rates, the number of FR resets, and increased mean wait time, resulting in a higher response-to-reinforcer ratio than exhibited by controls. Mice receiving the higher exposure level (6,000 ppm) showed a dramatic decrease in the number of rewards received, which was reversed when toluene exposure ceased. Mice receiving the lower exposure level (1,000 ppm) showed little change in the number of rewards. These results indicate that repeated binge exposures to high concentrations of toluene can significantly interfere with performance as measured by a waiting-for-reward task, suggesting a significant impact on cognitive and/or psychomotor function. PMID:18832024

  8. Respiratory viral RNA on toys in pediatric office waiting rooms.

    PubMed

    Pappas, Diane E; Hendley, J Owen; Schwartz, Richard H

    2010-02-01

    Toys in pediatric office waiting rooms may be fomites for transmission of viruses. Eighteen samples were taken from office objects on 3 occasions. Samples were tested for presence of picornavirus (either rhinovirus or enterovirus) on all 3 sample days; in addition, January samples were tested for respiratory syncytial virus and March samples were tested for influenza A and B. In addition, 15 samples were obtained from the sick waiting room before and after cleaning. Polymerase chain reaction was used to detect picornavirus, respiratory syncytial virus, and influenza A or B virus. Finally, 20 samples were obtained from the fingers of a researcher after handling different toys in the sick waiting room, and samples were then obtained from all the same toys; all samples were tested for picornavirus by polymerase chain reaction. Viral RNA was detected on 11 of 52 (21%) of toys sampled. Ten of the positives were picornavirus; 1 was influenza B virus. Three (30%) of 10 toys from the new toy bag, 6 of 30 (20%) in the sick child waiting room, and 2 of 12 (17%) in the well child waiting room were positive. Six (40%) of 15 toys in the sick waiting room were positive for picornaviral RNA before cleaning; after cleaning, 4 (27%) of 15 were positive in spite of the fact that RNA was removed from 4 of 6 of the original positives. Three (15%) of 20 toys in the sick waiting room were positive for picornaviral RNA, but RNA was not transferred to the fingers of the investigator who handled these toys. About 20% of the objects in a pediatric office may be contaminated with respiratory viral RNA, most commonly picornavirus RNA. Cleaning with a disinfectant cloth was only modestly effective in removing the viral RNA from the surfaces of toys, but transfer of picornaviral RNA from toys to fingers was inefficient.

  9. Efficient priority queueing routing strategy on networks of mobile agents

    NASA Astrophysics Data System (ADS)

    Wu, Gan-Hua; Yang, Hui-Jie; Pan, Jia-Hui

    2018-03-01

    As a consequence of their practical implications for communications networks, traffic dynamics on complex networks have recently captivated researchers. Previous routing strategies for improving transport efficiency have paid little attention to the orders in which the packets should be forwarded, just simply used first-in-first-out queue discipline. Here, we apply a priority queuing discipline and propose a shortest-distance-first routing strategy on networks of mobile agents. Numerical experiments reveal that the proposed scheme remarkably improves both the network throughput and the packet arrival rate and reduces both the average traveling time and the rate of waiting time to traveling time. Moreover, we find that the network capacity increases with an increase in both the communication radius and the number of agents. Our work may be helpful for the design of routing strategies on networks of mobile agents.

  10. High Emergency Lung Transplantation: dramatic decrease of waiting list death rate without relevant higher post-transplant mortality.

    PubMed

    Roux, Antoine; Beaumont-Azuar, Laurence; Hamid, Abdul Monem; De Miranda, Sandra; Grenet, Dominique; Briend, Guillaume; Bonnette, Pierre; Puyo, Philippe; Parquin, François; Devaquet, Jerome; Trebbia, Gregoire; Cuquemelle, Elise; Douvry, Benoit; Picard, Clément; Le Guen, Morgan; Chapelier, Alain; Stern, Marc; Sage, Edouard

    2015-09-01

    Many candidates for lung transplantation (LT) die on the waiting list, raising the question of graft availability and strategy for organ allocation. We report the experience of the new organ allocation program, "High Emergency Lung Transplantation" (HELT), since its implementation in our center in 2007. Retrospective analysis of 201 lung transplant patients, of whom 37 received HELT from 1st July 2007 to 31th May 2012. HELT candidates had a higher impairment grade on respiratory status and higher Lung Allocation Score (LAS). HELT patients had increased incidence of perioperative complications (e.g., perioperative bleeding) and extracorporeal circulatory assistance (75% vs. 36.6%, P = 0.0005). No significant difference was observed between HELT and non-HELT patients in mechanical ventilation duration (15.5 days vs. 11 days, P = 0.27), intensive care unit length of stay (15 days vs. 10 days, P = 0.22) or survival rate at 12 (81% vs. 80%), and 24 months post-LT (72.9% vs. 75.0%). Lastly, mortality on the waiting list was spectacularly reduced from 19% to 2% when compared to the non-HELT 2004-2007 group. Despite a more severe clinical status of patients on the waiting list, HELT provided similar results to conventional LT. These results were associated with a dramatic reduction in the mortality rate of patients on the waiting list. © 2015 Steunstichting ESOT.

  11. Two facets of patience in young children: Waiting with and without an explicit reward.

    PubMed

    Barragan-Jason, Gladys; Atance, Cristina; Kopp, Leia; Hopfensitz, Astrid

    2018-07-01

    Patience, or the ability to tolerate delay, is typically studied using delay of gratification (DoG) tasks. However, among other factors (e.g., type of reward), the use of a reward to test patience is affected by an individual's motivation to obtain the reward (e.g., degree of preference for the small vs. large reward). In addition, DoG tasks do not assess the extent to which an individual can wait in the absence of an explicit reward-or what we term "patience as a virtue." Accordingly, the current study used a new measure of patience-the "pure waiting paradigm"-in which 3- to 5-year-old children waited 3 min with nothing to do and with no explicit reward. We then examined the relation between performance on this task (as assessed by children's spontaneous patient behaviors) and performance on two DoG tasks (candy and video rewards). Significant correlations were found between DoG performance and patient behaviors in the pure waiting paradigm, especially when controlling for motivation. These results and methodology show for the first time a direct link between patience as a virtue and DoG performance and also provide new insights about the study of patience in children. Copyright © 2018 Elsevier Inc. All rights reserved.

  12. Activity in children with ADHD during waiting situations in the classroom: a pilot study.

    PubMed

    Antrop, Inge; Buysse, Ann; Roeyers, Herbert; Van Oost, Paulette

    2005-03-01

    According to the optimal stimulation theory and the delay aversion hypothesis, children with attention deficit hyperactivity disorder (ADHD) experience difficulties when they are confronted with low levels of stimulation and delay, respectively. This study investigated the activity level of children with ADHD during waiting situations in the classroom. Three series of hypothesis were made: (1) with respect to the comparison between waiting and non-waiting intervals, (2) with respect to the effects of non-temporal stimulation, and (3) with respect to the effects of temporal stimulation on behaviour during waiting. The activity level of 14 children with ADHD and 14 control children between the ages of 6 and 11 years was observed during two non-waiting class situations and three waiting situations: without any stimulation, in the presence of nontemporal stimulation and in the presence of temporal stimulation. Both groups of children obtained higher activity scores for all behavioural dimensions during waiting compared with non-waiting situations. The results further revealed additive effects of waiting and diagnostic group on behaviour. Additional nontemporal stimulation during waiting affected the behaviour of all children for most behavioural characteristics. For noisiness, additive effects were also found for diagnostic group and either non-temporal stimulation or temporal stimulation. For restlessness, a trend for an interaction effect between diagnostic group and nontemporal stimulation was found. The findings have clear implications for school observations within an assessment protocol.

  13. The Application of Waiting Lines System in Improving Customer Service Management: The Examination of Malaysia Fast Food Restaurants Industry

    NASA Astrophysics Data System (ADS)

    Ismail, Zurina; Shokor, Shahrul Suhaimi AB

    2016-03-01

    Rapid life time change of the Malaysian lifestyle had served the overwhelming growth in the service operation industry. On that occasion, this paper will provide the idea to improve the waiting line system (WLS) practices in Malaysia fast food chains. The study will compare the results in between the single server single phase (SSSP) and the single server multi-phase (SSMP) which providing Markovian Queuing (MQ) to be used for analysis. The new system will improve the current WLS, plus intensifying the organization performance. This new WLS were designed and tested in a real case scenario and in order to develop and implemented the new styles, it need to be focusing on the average number of customers (ANC), average number of customer spending time waiting in line (ACS), and the average time customers spend in waiting and being served (ABS). We introduced new WLS design and there will be prompt discussion upon theories of benefits and potential issues that will benefit other researchers.

  14. Waiting lists and elective surgery: ordering the queue.

    PubMed

    Curtis, Andrea J; Russell, Colin O H; Stoelwinder, Johannes U; McNeil, John J

    2010-02-15

    In the Australian public health system, access to elective surgery is rationed through the use of waiting lists in which patients are assigned to broad urgency categories. Surgeons are principally responsible for referring patients to waiting lists, deciding on the appropriate urgency category, and selecting patients from the waiting list to receive surgery. There are few agreed-upon criteria to help surgeons make these decisions, leading to striking differences between institutions in proportions of patients allocated to urgency categories. In other countries with publicly funded health systems, programs have been developed that aim to make prioritisation more consistent and access to surgery more equitable. As demand for health care increases, similar programs should be established in Australia using relevant clinical and psychosocial factors. Prioritisation methodology adapted for elective surgery may have a role in prioritising high-demand procedures in other areas of health care.

  15. Traffic pollutants measured inside vehicles waiting in line at a major US-Mexico Port of Entry.

    PubMed

    Quintana, Penelope J E; Khalighi, Mehdi; Castillo Quiñones, Javier Emmanuel; Patel, Zalak; Guerrero Garcia, Jesus; Martinez Vergara, Paulina; Bryden, Megan; Mantz, Antoinette

    2018-05-01

    At US-Mexico border Ports of Entry, vehicles idle for long times waiting to cross northbound into the US. Long wait times at the border have mainly been studied as an economic issue, however, exposures to emissions from idling vehicles can also present an exposure risk. Here we present the first data on in-vehicle exposures to driver and passengers crossing the US-Mexico border at the San Ysidro, California Port of Entry (SYPOE). Participants were recruited who regularly commuted across the border in either direction and told to drive a scripted route between two border universities, one in the US and one in Mexico. Instruments were placed in participants' cars prior to commute to monitor-1-minute average levels of the traffic pollutants ultrafine particles (UFP), black carbon (BC) and carbon monoxide (CO) in the breathing zone of drivers and passengers. Location was determined by a GPS monitor. Results reported here are for 68 northbound participant trips. The highest median levels of in-vehicle UFP were recorded during the wait to cross at the SYPOE (median 29,692particles/cm 3 ) significantly higher than the portion of the commute in the US (median 20,508particles/cm 3 ) though not that portion in Mexico (median 22, 191particles/cm 3 ). In-vehicle BC levels at the border were significantly lower than in other parts of the commute. Our results indicate that waiting in line at the SYPOE contributes a median 62.5% (range 15.5%-86.0%) of a cross-border commuter's exposure to UFP and a median 44.5% (range (10.6-79.7%) of exposure to BC inside the vehicle while traveling in the northbound direction. Reducing border wait time can significantly reduce in-vehicle exposures to toxic air pollutants such as UFP and BC, and these preventable exposures can be considered an environmental justice issue. Copyright © 2017 Elsevier B.V. All rights reserved.

  16. Waiting list randomized controlled trial within a case-finding design: methodological considerations.

    PubMed

    Ronaldson, Sarah; Adamson, Joy; Dyson, Lisa; Torgerson, David

    2014-10-01

    Randomized controlled trials (RCTs) are widely used in health care research to provide high-quality evidence of effectiveness of an intervention. However, sometimes a study does not require an RCT in order to answer its primary objective; a case-finding design may be more appropriate. The aim of this paper was to introduce a new study design that nests a waiting list RCT within a case-finding study. An example of the new study design is the DOC Study, which primarily aims to determine the diagnostic accuracy of lung function tests for chronic obstructive pulmonary disease. It also investigates the impact of lung function tests on smoking behaviour through use of a waiting list design. The first step of the study design is to obtain participants' consent. Individuals are then randomized to one of two groups; either the 'intervention now' group or the 'intervention later' group, that is, participants are placed on a waiting list. All participants receive the same intervention; the only difference between the groups is the timing of the intervention. The design addresses patient preference issues and recruitment issues that can arise in other trial designs. Potential limitations include differential attrition between study groups and potential demoralization for the 'intervention later' group. The 'waiting list case-finding trial' design is a valuable method that could be applied to case-finding studies; the design enables the case-finding component of a study to be maintained while simultaneously exploring additional hypotheses through conducting a trial. © 2014 John Wiley & Sons, Ltd.

  17. The Role of Diverse Strategies in Sustainable Knowledge Production

    PubMed Central

    Wu, Lingfei; Baggio, Jacopo A.; Janssen, Marco A.

    2016-01-01

    Online communities are becoming increasingly important as platforms for large-scale human cooperation. These communities allow users seeking and sharing professional skills to solve problems collaboratively. To investigate how users cooperate to complete a large number of knowledge-producing tasks, we analyze Stack Exchange, one of the largest question and answer systems in the world. We construct attention networks to model the growth of 110 communities in the Stack Exchange system and quantify individual answering strategies using the linking dynamics on attention networks. We identify two answering strategies. Strategy A aims at performing maintenance by doing simple tasks, whereas strategy B aims at investing time in doing challenging tasks. Both strategies are important: empirical evidence shows that strategy A decreases the median waiting time for answers and strategy B increases the acceptance rate of answers. In investigating the strategic persistence of users, we find that users tends to stick on the same strategy over time in a community, but switch from one strategy to the other across communities. This finding reveals the different sets of knowledge and skills between users. A balance between the population of users taking A and B strategies that approximates 2:1, is found to be optimal to the sustainable growth of communities. PMID:26934733

  18. The Role of Diverse Strategies in Sustainable Knowledge Production.

    PubMed

    Wu, Lingfei; Baggio, Jacopo A; Janssen, Marco A

    2016-01-01

    Online communities are becoming increasingly important as platforms for large-scale human cooperation. These communities allow users seeking and sharing professional skills to solve problems collaboratively. To investigate how users cooperate to complete a large number of knowledge-producing tasks, we analyze Stack Exchange, one of the largest question and answer systems in the world. We construct attention networks to model the growth of 110 communities in the Stack Exchange system and quantify individual answering strategies using the linking dynamics on attention networks. We identify two answering strategies. Strategy A aims at performing maintenance by doing simple tasks, whereas strategy B aims at investing time in doing challenging tasks. Both strategies are important: empirical evidence shows that strategy A decreases the median waiting time for answers and strategy B increases the acceptance rate of answers. In investigating the strategic persistence of users, we find that users tends to stick on the same strategy over time in a community, but switch from one strategy to the other across communities. This finding reveals the different sets of knowledge and skills between users. A balance between the population of users taking A and B strategies that approximates 2:1, is found to be optimal to the sustainable growth of communities.

  19. Tactical AI in Real Time Strategy Games

    DTIC Science & Technology

    2015-03-26

    TACTICAL AI IN REAL TIME STRATEGY GAMES THESIS Donald A. Gruber, Capt, USAF AFIT-ENG-MS-15-M-021 DEPARTMENT OF THE AIR FORCE AIR UNIVERSITY AIR FORCE...protection in the United States. AFIT-ENG-MS-15-M-021 TACTICAL AI IN REAL TIME STRATEGY GAMES THESIS Presented to the Faculty Department of Electrical...DISTRIBUTION STATEMENT A APPROVED FOR PUBLIC RELEASE; DISTRIBUTION UNLIMITED. AFIT-ENG-MS-15-M-021 TACTICAL AI IN REAL TIME STRATEGY GAMES THESIS Donald A

  20. SouthWest view, Street Car Waiting House, north and east elevations ...

    Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey

    South-West view, Street Car Waiting House, north and east elevations - North Philadelphia Station, Street Car Waiting House, 2900 North Broad Street, on northwest corner of Broad Street & Glenwood Avenue, Philadelphia, Philadelphia County, PA

  1. NorthEast view; Street Car Waiting House, south (front) and west ...

    Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey

    North-East view; Street Car Waiting House, south (front) and west elevations - North Philadelphia Station, Street Car Waiting House, 2900 North Broad Street, on northwest corner of Broad Street & Glenwood Avenue, Philadelphia, Philadelphia County, PA

  2. Informing Healthcare Waiting Area Design Using Transparency Attributes: A Comparative Preference Study.

    PubMed

    Jiang, Shan; Powers, Matthew; Allison, David; Vincent, Ellen

    2017-07-01

    This study aimed to explore people's visual preference for waiting areas in general hospital environments designed with transparency attributes that fully integrate nature. Waiting can be a tedious and frustrating experience among people seeking healthcare treatments and negatively affect their perception of the quality of care. Positive distractions and supportive designs have gained increasing attraction to improve people's waiting experience. Nature, which has shown therapeutic effects according to a growing amount of evidence, could be a distinguished positive distraction in waiting areas. Additionally, the theory of transparency was operationalized to indicate a spatial continuity between the external nature and the built interiors in general healthcare waiting area design. A survey method was adopted in the study. Twenty-one images of general healthcare waiting areas depicting three design typologies were preselected following a strict procedure, including designs with (a) no window views, (b) limited window views to nature, and (c) transparent spaces with maximum natural views. Ninety-five student participants rated the images based on their visual preference using a Likert-type scale. The results showed that transparent waiting areas were significantly preferred. A significant positive relationship existed between the level of transparency and people's preference scores. The factor analysis indicated additional supportive features that may affect people's preferences, including daylight, perceived warmth, noninstitutional furniture arrangement, visual orientation, and the use of natural materials for interior design. However, these tentative results need to be furthered tested with the real patient population as the next step of this study.

  3. A state of limbo: the politics of waiting in neo-liberal Latvia.

    PubMed

    Ozoliņa-Fitzgerald, Liene

    2016-09-01

    This article presents an ethnographic study of politics of waiting in a post-Soviet context. While activation has been explored in sociological and anthropological literature as a neo-liberal governmental technology and its application in post-socialist context has also been compellingly documented, waiting as a political artefact has only recently been receiving increased scholarly attention. Drawing on ethnographic fieldwork at a state-run unemployment office in Riga, this article shows how, alongside activation, state welfare policies also produce passivity and waiting. Engaging with the small but developing field of sociological literature on the politics of waiting, I argue that, rather than interpreting it as a clash between 'neo-liberal' and 'Soviet' regimes, we should understand the double-move of activation and imposition of waiting as a key mechanism of neo-liberal biopolitics. This article thus extends the existing theorizations of the temporal politics of neo-liberalism. © London School of Economics and Political Science 2016.

  4. Six Differentiated Strategies for ESL Literacy for Birth to Third Grade Developmentally Disabled and Normal Students of Hispanic Heritage

    ERIC Educational Resources Information Center

    Jaramillo, James; Jaramillo, Olga

    2013-01-01

    When one effectively employs the strategies of exploratory-learning, wait-time, intervention, guided reading, meaning, and phonological-morphological-syntactical awareness-for infants and on up-to 3rd grade students-all-in a Montessori-like-learning-literacy-setting replete with semantical interactions with phonology, syllabology, morphology, and…

  5. Transient probabilities for queues with applications to hospital waiting list management.

    PubMed

    Joy, Mark; Jones, Simon

    2005-08-01

    In this paper we study queuing systems within the NHS. Recently imposed government performance targets lead NHS executives to investigate and instigate alternative management strategies, thereby imposing structural changes on the queues. Under such circumstances, it is most unlikely that such systems are in equilibrium. It is crucial, in our opinion, to recognise this state of affairs in order to make a balanced assessment of the role of queue management in the modern NHS. From a mathematical perspective it should be emphasised that measures of the state of a queue based upon the assumption of statistical equilibrium (a pervasive methodology in the study of queues) are simply wrong in the above scenario. To base strategic decisions around such ideas is therefore highly questionable and it is one of the purposes of this paper to offer alternatives: we present some (recent) research whose results generate performance measures and measures of risk, for example, of waiting-times growing unacceptably large; we emphasise that these results concern the transient behaviour of the queueing model-there is no asssumption of statistical equilibrium. We also demonstrate that our results are computationally tractable.

  6. Randomized control trial: Online parent program and waiting period for unmarried parents in Title IV-D court.

    PubMed

    Rudd, Brittany N; Holtzworth-Munroe, Amy; Reyome, Jason G; Applegate, Amy G; D'Onofrio, Brian M

    2015-10-01

    Despite a lack of research on parent education programs for unmarried parents, many judicial officers mandate participation. We recruited an understudied sample likely at high risk for negative outcomes-182 court cases involving unmarried parents on government assistance in which paternity was contested and then established via genetic testing ordered by the court. This 2 × 2 randomized controlled trial evaluated the impact on initial litigation outcomes of two factors: (a) participation in an online parent education program or not and (b) having a waiting period between the establishment of paternity and the court hearing concerning child-related issues or not. Using an intent-to-treat framework, we found that among cases not assigned to the program, there was no difference in the rate of full agreement on child-related issues (e.g., child support, custody, parenting time) when comparing cases assigned to a waiting period and cases not assigned to a waiting period. In contrast, for cases assigned to the program, cases also assigned a waiting period were less likely to reach a full agreement than cases that had their hearing on the same day. In addition, cases in the "program and waiting period" condition were less likely to return to court for their hearing than cases in the "no program and waiting period" condition. In exploratory analyses of the subsample of cases in which both parents were present at the court hearing, the pattern of results remained the same, although the findings were no longer statistically significant. (c) 2015 APA, all rights reserved).

  7. The experience of adults who choose watchful waiting or active surveillance as an approach to medical treatment: a qualitative systematic review.

    PubMed

    Rittenmeyer, Leslie; Huffman, Dolores; Alagna, Michael; Moore, Ellen

    2016-02-01

    "Watchful waiting" or "active surveillance" is an alternative approach in the medical management of certain diseases. Most often considered appropriate as an approach to treatment for low-risk prostate cancer, it is also found in the literature in breast cancer surveillance, urinary lithiasis, lymphocytic leukemia, depression and small renal tumors. This systematic review sought to:Identify and synthesize the best available international evidence on the experience of adults who choose watchful waiting or active surveillance as an approach to medical treatment. To this end the questions addressed in this review were:1. How do patients who have chosen watchful waiting or active surveillance describe the process of coming to the decision?2. What were the factors that influenced their decision to choose?3. How do patients who have chosen watchful waiting or active surveillance describe the experience? Male or female patients, 18 years or older, who experience the phenomenon of choosing or not choosing watchful waiting or active surveillance as a treatment approach.The phenomena of interest were accounts of the experiences of adult patients who choose watchful waiting or active surveillance as an approach to medical treatment.This review considered studies that focused on qualitative data including, but not limited to, designs such as phenomenology, grounded theory, ethnography, action research and critical theory. Mixed method studies with narrative description and patient voice were also considered. Grey literature such as research reports and dissertations were also included. The search strategy aimed to find both published and unpublished studies through electronic databases, reference lists, and the World Wide Web. Extensive searches were undertaken of relevant databases to include CINAHL, PubMed, SCOPUS and PsycINFO. A three-step search strategy was used in each component of the review. Studies were limited to English language papers. The search considered papers

  8. Tutorial in medical decision modeling incorporating waiting lines and queues using discrete event simulation.

    PubMed

    Jahn, Beate; Theurl, Engelbert; Siebert, Uwe; Pfeiffer, Karl-Peter

    2010-01-01

    In most decision-analytic models in health care, it is assumed that there is treatment without delay and availability of all required resources. Therefore, waiting times caused by limited resources and their impact on treatment effects and costs often remain unconsidered. Queuing theory enables mathematical analysis and the derivation of several performance measures of queuing systems. Nevertheless, an analytical approach with closed formulas is not always possible. Therefore, simulation techniques are used to evaluate systems that include queuing or waiting, for example, discrete event simulation. To include queuing in decision-analytic models requires a basic knowledge of queuing theory and of the underlying interrelationships. This tutorial introduces queuing theory. Analysts and decision-makers get an understanding of queue characteristics, modeling features, and its strength. Conceptual issues are covered, but the emphasis is on practical issues like modeling the arrival of patients. The treatment of coronary artery disease with percutaneous coronary intervention including stent placement serves as an illustrative queuing example. Discrete event simulation is applied to explicitly model resource capacities, to incorporate waiting lines and queues in the decision-analytic modeling example.

  9. Algorithm for prioritization of patients on the waiting list for liver transplantation.

    PubMed

    Gambato, M; Senzolo, M; Canova, D; Germani, G; Tomat, S; Masier, A; Russo, F P; Perissinotto, E; Zanus, G; Cillo, U; Burra, P

    2007-01-01

    Prioritization of patients on the waiting list (WL) for OLT is still a critical issue. Numerous models have been developed to predict mortality before and after OLT. The aim of the study was to prospectively evaluate cirrhotics with and without hepatocellular carcinoma (HCC) undergoing orthotopic liver transplantation (OLT) severity of liver disease on the WL and at transplant, mortality on the WL and after OLT, and their correlations. An algorithm based on seven patient variables (MELD, CTP, UNOS, HCC, BMI, waiting time, age) was created by software dedicated to prioritize patients on the waiting list. We evaluated 118 patients including 75 men and 43 women of age range 19 to 66 years, who underwent OLT from July 2004 to June 2006. Mean CTP and MELD at listing were 8.44 (range 6-12) and 13 (range 2-24), respectively. Overall mortality on the WL at 24 months was 13%, which was significantly higher among patients with MELD > 25 compared to patients with MELD 0 to 15 (P < .0001) or MELD 16 to 25 (P = .0007) at listing. Mean MELD at OLT was 15 (range 7-36), which was significantly lower in patients with than without HCC (MELD 12 vs 16; P = .0003). Six hundred-day patient survival was significantly lower among patients with MELD > 25 compared to patients with MELD < 25 at OLT (P = .017), whereas no difference in survival was observed between patients with and without HCC. The sickest patients are characterized by high mortality both on the waiting list and after liver transplantation. Patients with HCC are transplanted in better condition compared to patients without HCC with the same survival.

  10. The ethics of waiting and anticipating life beyond.

    PubMed

    Milton, Constance L

    2014-01-01

    Waiting is a common everyday experience. It is particularly important to person(s) and families living with changing complex health patterns and may be especially vital to those who are anticipating the end-of-life and beyond. The author in this column offers a discussion of potential definitions, meanings, and straight thinking responsibilities for healthcare professionals, as they provide professional services with persons and families who may be arduously experiencing the phenomenon of waiting with ever-changing health situations. Implications for professional nurse practice are offered from a humanbecoming perspective.

  11. Family caring strategies in neutropenia.

    PubMed

    Eggenberger, Sandra K; Krumwiede, Norma; Meiers, Sonja J; Bliesmer, Mary; Earle, Patricia

    2004-12-01

    Aggressive chemotherapy protocols result in neutropenia in approximately half of all patients receiving chemotherapy. Thus, neutropenia continues to be a significant and potentially life-threatening side effect of treatment, even with use of colony-stimulating factors. Families of patients with neutropenia often provide the primary healing environment because most chemotherapy protocols are managed on an outpatient basis. To learn about the family's experience of managing chemotherapy-induced neutropenia (CIN), a grounded-theory methodology was used to analyze data from seven families. The central theme revealed by these families was "turbulent waiting with intensified connections." This meant that when families had a sense of greater vulnerability in response to the waiting after diagnosis of CIN, they connected intensely with each other and healthcare providers. Families reported that connections with nurses became more significant when neutropenia interrupted chemotherapy. Families also developed family caring strategies to manage this period of waiting for the chemotherapy to resume. These strategies included family inquiry, family vigilance, and family balancing. Nurses need to be aware of approaches to support the family's ability to manage CIN. Interventions and approaches constructed from the perspective of a family-professional partnership will enhance the family cancer experience as well as ongoing family growth and function.

  12. Review of "Waiting for Superman"

    ERIC Educational Resources Information Center

    Dutro, Elizabeth

    2011-01-01

    "Waiting for Superman" offers what appear to be straightforward, commonsense solutions to inequities in schooling. The film argues that heroic action can be taken to fix what it portrays as the disaster of public schooling. The film disregards poverty as a factor in school performance and connection--and therefore never addresses anti-poverty…

  13. High emergency organ allocation rule in lung transplantation: a simulation study.

    PubMed

    Riou, Julien; Boëlle, Pierre-Yves; Christie, Jason D; Thabut, Gabriel

    2017-10-01

    The scarcity of suitable organ donors leads to protracted waiting times and mortality in patients awaiting lung transplantation. This study aims to assess the short- and long-term effects of a high emergency organ allocation policy on the outcome of lung transplantation. We developed a simulation model of lung transplantation waiting queues under two allocation strategies, based either on waiting time only or on additional criteria to prioritise the sickest patients. The model was informed by data from the United Network for Organ Sharing. We compared the impact of these strategies on waiting time, waiting list mortality and overall survival in various situations of organ scarcity. The impact of a high emergency allocation strategy depends largely on the organ supply. When organ supply is sufficient (>95 organs per 100 patients), it may prevent a small number of early deaths (1 year survival: 93.7% against 92.4% for waiting time only) without significant impact on waiting times or long-term survival. When the organ/recipient ratio is lower, the benefits in early mortality are larger but are counterbalanced by a dramatic increase of the size of the waiting list. Consequently, we observed a progressive increase of mortality on the waiting list (although still lower than with waiting time only), a deterioration of patients' condition at transplant and a decrease of post-transplant survival times. High emergency organ allocation is an effective strategy to reduce mortality on the waiting list, but causes a disruption of the list equilibrium that may have detrimental long-term effects in situations of significant organ scarcity.

  14. High emergency organ allocation rule in lung transplantation: a simulation study

    PubMed Central

    Boëlle, Pierre-Yves; Christie, Jason D.; Thabut, Gabriel

    2017-01-01

    The scarcity of suitable organ donors leads to protracted waiting times and mortality in patients awaiting lung transplantation. This study aims to assess the short- and long-term effects of a high emergency organ allocation policy on the outcome of lung transplantation. We developed a simulation model of lung transplantation waiting queues under two allocation strategies, based either on waiting time only or on additional criteria to prioritise the sickest patients. The model was informed by data from the United Network for Organ Sharing. We compared the impact of these strategies on waiting time, waiting list mortality and overall survival in various situations of organ scarcity. The impact of a high emergency allocation strategy depends largely on the organ supply. When organ supply is sufficient (>95 organs per 100 patients), it may prevent a small number of early deaths (1 year survival: 93.7% against 92.4% for waiting time only) without significant impact on waiting times or long-term survival. When the organ/recipient ratio is lower, the benefits in early mortality are larger but are counterbalanced by a dramatic increase of the size of the waiting list. Consequently, we observed a progressive increase of mortality on the waiting list (although still lower than with waiting time only), a deterioration of patients’ condition at transplant and a decrease of post-transplant survival times. High emergency organ allocation is an effective strategy to reduce mortality on the waiting list, but causes a disruption of the list equilibrium that may have detrimental long-term effects in situations of significant organ scarcity. PMID:29181383

  15. Individual differences in intrinsic brain connectivity predict decision strategy.

    PubMed

    Barnes, Kelly Anne; Anderson, Kevin M; Plitt, Mark; Martin, Alex

    2014-10-15

    When humans are provided with ample time to make a decision, individual differences in strategy emerge. Using an adaptation of a well-studied decision making paradigm, motion direction discrimination, we probed the neural basis of individual differences in strategy. We tested whether strategies emerged from moment-to-moment reconfiguration of functional brain networks involved in decision making with task-evoked functional MRI (fMRI) and whether intrinsic properties of functional brain networks, measured at rest with functional connectivity MRI (fcMRI), were associated with strategy use. We found that human participants reliably selected one of two strategies across 2 days of task performance, either continuously accumulating evidence or waiting for task difficulty to decrease. Individual differences in decision strategy were predicted both by the degree of task-evoked activation of decision-related brain regions and by the strength of pretask correlated spontaneous brain activity. These results suggest that spontaneous brain activity constrains strategy selection on perceptual decisions.

  16. Whittling Down the Wait Time: Exploring Models to Minimize the Delay from Initial Concern to Diagnosis and Treatment of Autism Spectrum Disorder.

    PubMed

    Gordon-Lipkin, Eliza; Foster, Jessica; Peacock, Georgina

    2016-10-01

    The process from initial concerns to diagnosis of autism spectrum disorder (ASD) can be a long and complicated process. The traditional model for evaluation and diagnosis of ASD often consists of long wait-lists and evaluations that result in a 2-year difference between the earliest signs of ASD and mean age of diagnosis. Multiple factors contribute to this diagnostic bottleneck, including time-consuming evaluations, cost of care, lack of providers, and lack of comfort of primary care providers to diagnose autism. This article explores innovative clinical models that have been implemented to address this as well as future directions and opportunities. Copyright © 2016 Elsevier Inc. All rights reserved.

  17. Waiting for a kidney transplant: the experience of patients with end-stage renal disease in South Korea.

    PubMed

    Chong, Hye Jin; Kim, Hyun Kyung; Kim, Sung Reul; Lee, Sik

    2016-04-01

    To explore the experiences of Korean patients with end-stage renal disease awaiting kidney transplantation. The need for kidney transplantation has increased worldwide, while the number of kidney donors has not increased commensurately. This mismatch is a serious issue in South Korea. Prolonged waits for transplantation may cause physical and psychosocial issues and lead to poor outcomes. Nevertheless, the experience of waiting for kidney transplantation in South Korea has never been explored in depth. A qualitative descriptive design was used. The participants were eight patients diagnosed with end-stage renal disease on the waiting list for kidney transplantation in South Korea. Data were collected through individual in-depth interviews. All conversations during interviews were recorded and transcribed verbatim. Transcribed data were analysed using conventional content analysis. The experience of waiting for kidney transplantation consisted of six categories: (1) the light at the end of the tunnel, (2) being on call without any promise, (3) a tough tug of war between excitement and frustration, (4) doubts in the complexity, (5) A companion on the hard journey and (6) getting ready for D-day. Kidney transplantation candidates experience psychosocial difficulties and concerns while waiting for long periods of time without any assurance of resolution. Systematic education and psychosocial support from health care professionals and family members help patients get through what they describe as a difficult journey. Comprehensive management programs for kidney transplantation candidates are needed. Health care professionals need to recognise the psychosocial concerns of patients awaiting kidney transplantation. Clinicians should provide patients with information and support throughout the waiting period. © 2016 John Wiley & Sons Ltd.

  18. Bottlenecks and Waiting Points in Nucleosynthesis in X-ray bursts and Novae

    NASA Astrophysics Data System (ADS)

    Smith, Michael S.; Sunayama, Tomomi; Hix, W. Raphael; Lingerfelt, Eric J.; Nesaraja, Caroline D.

    2010-08-01

    To better understand the energy generation and element synthesis occurring in novae and X-ray bursts, we give quantitative definitions to the concepts of ``bottlenecks'' and ``waiting points'' in the thermonuclear reaction flow. We use these criteria to search for bottlenecks and waiting points in post-processing element synthesis explosion simulations. We have incorporated these into the Computational Infrastructure for Nuclear Astrophysics, a suite of nuclear astrophysics codes available online at nucastrodata.org, so that anyone may perform custom searches for bottlenecks and waiting points.

  19. Bottlenecks and Waiting Points in Nucleosynthesis in X-ray bursts and Novae

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

    Smith, Michael S.; Hix, W. Raphael; Nesaraja, Caroline D.

    2010-08-12

    To better understand the energy generation and element synthesis occurring in novae and X-ray bursts, we give quantitative definitions to the concepts of ''bottlenecks'' and ''waiting points'' in the thermonuclear reaction flow. We use these criteria to search for bottlenecks and waiting points in post-processing element synthesis explosion simulations. We have incorporated these into the Computational Infrastructure for Nuclear Astrophysics, a suite of nuclear astrophysics codes available online at nucastrodata.org, so that anyone may perform custom searches for bottlenecks and waiting points.

  20. Effectiveness and acceptance of a web-based depression intervention during waiting time for outpatient psychotherapy: study protocol for a randomized controlled trial.

    PubMed

    Grünzig, Sasha-Denise; Baumeister, Harald; Bengel, Jürgen; Ebert, David; Krämer, Lena

    2018-05-22

    Due to limited resources, waiting periods for psychotherapy are often long and burdening for those in need of treatment and the health care system. In order to bridge the gap between initial contact and the beginning of psychotherapy, web-based interventions can be applied. The implementation of a web-based depression intervention during waiting periods has the potential to reduce depressive symptoms and enhance well-being in depressive individuals waiting for psychotherapy. In a two-arm randomized controlled trial, effectiveness and acceptance of a guided web-based intervention for depressive individuals on a waitlist for psychotherapy are evaluated. Participants are recruited in several German outpatient clinics. All those contacting the outpatient clinics with the wish to enter psychotherapy receive study information and a depression screening. Those adults (age ≥ 18) with depressive symptoms above cut-off (CES-D scale > 22) and internet access are randomized to either intervention condition (treatment as usual and immediate access to the web-based intervention) or waiting control condition (treatment as usual and delayed access to the web-based intervention). At three points of assessment (baseline, post-treatment, 3-months-follow-up) depressive symptoms and secondary outcomes, such as quality of life, attitudes towards psychotherapy and web-based interventions and adverse events are assessed. Additionally, participants' acceptance of the web-based intervention is evaluated, using measures of intervention adherence and satisfaction. This study investigates a relevant setting for the implementation of web-based interventions, potentially improving the provision of psychological health care. The results of this study contribute to the evaluation of innovative and resource-preserving health care models for outpatient psychological treatment. This trial has been registered on 13 February 2017 in the German clinical trials register (DRKS); registration

  1. Toys are a potential source of cross-infection in general practitioners' waiting rooms.

    PubMed Central

    Merriman, Eileen; Corwin, Paul; Ikram, Rosemary

    2002-01-01

    The waiting rooms of general practitioners' surgeries usually have toys provided for children. The level of contamination of these toys and the effectiveness of toy decontamination was investigated in this study. Hard toys from general practitioners' waiting rooms had relatively low levels of contamination, with only 13.5% of toys showing any coliform counts. There were no hard toys with heavy contamination by coliforms or other bacteria. Soft toys were far more likely to be contaminated, with 20% of toys showing moderate to heavy coliform contamination and 90% showing moderate to heavy bacterial contamination. Many waiting-room toys are not cleaned routinely. Soft toys are hard to disinfect and tend to rapidly become recontaminated after cleaning. Conversely, hard toys can be cleaned and disinfected easily. Soft toys in general practitioners' waiting rooms pose an infectious risk and it is therefore recommended that soft toys are unsuitable for doctors' waiting rooms. PMID:11885823

  2. Outcome Probability versus Magnitude: When Waiting Benefits One at the Cost of the Other

    PubMed Central

    Young, Michael E.; Webb, Tara L.; Rung, Jillian M.; McCoy, Anthony W.

    2014-01-01

    Using a continuous impulsivity and risk platform (CIRP) that was constructed using a video game engine, choice was assessed under conditions in which waiting produced a continuously increasing probability of an outcome with a continuously decreasing magnitude (Experiment 1) or a continuously increasing magnitude of an outcome with a continuously decreasing probability (Experiment 2). Performance in both experiments reflected a greater desire for a higher probability even though the corresponding wait times produced substantive decreases in overall performance. These tendencies are considered to principally reflect hyperbolic discounting of probability, power discounting of magnitude, and the mathematical consequences of different response rates. Behavior in the CIRP is compared and contrasted with that in the Balloon Analogue Risk Task (BART). PMID:24892657

  3. The evaluation of a formalized queue management system for coronary angiography waiting lists.

    PubMed

    Alter, D A; Newman, Alice M; Cohen, Eric A; Sykora, Kathy; Tu, Jack V

    2005-11-01

    Lengthy waiting lists for coronary angiography have been described in many health care systems worldwide. The extent to which formal queue management systems may improve the prioritization and survival of patients in the angiography queue is unknown. To prospectively evaluate the performance of a formal queue management system for patients awaiting coronary angiography in Ontario. The coronary angiography urgency scale, a formal queue management system developed in 1993 using a modified Delphi panel, allocates recommended maximum waiting times (RMWTs) in accordance with clinical necessity. By using a provincial clinical registry, 35,617 consecutive patients referred into the coronary angiography queue between April 1, 2001, and March 31, 2002, were prospectively tracked. Cox proportional hazards models were used to examined mortality risk across urgency after adjusting for additional clinical and comorbid factors. Good agreement was determined in urgency ratings between scores from the coronary angiography urgency scale and implicit physician judgement, which was obtained independently at the time of the index referral (weighted kappa = 0.49). The overall mortality in the queue was 0.3% (0.47%, 0.26% and 0.13% for urgent, semiurgent and elective patients, respectively). Urgency, as specified by the coronary angiography urgency scale, was the strongest predictor of death in the queue (P<0.001). However, when patients were censored according to their RMWTs, mortality was similar across different levels of urgency. Consequently, up to 18.5 deaths per 10,000 patients could have potentially been averted had patients been triaged and undergone coronary angiography within the RMWT as specified by the coronary angiography urgency scale. The incorporation of the coronary angiography urgency scale as a formal queue management system may decrease mortality in the coronary angiography queue. The authors recommend its implementation in health care systems where patients

  4. Explaining How to Play Real-Time Strategy Games

    NASA Astrophysics Data System (ADS)

    Metoyer, Ronald; Stumpf, Simone; Neumann, Christoph; Dodge, Jonathan; Cao, Jill; Schnabel, Aaron

    Real-time strategy games share many aspects with real situations in domains such as battle planning, air traffic control, and emergency response team management which makes them appealing test-beds for Artificial Intelligence (AI) and machine learning. End user annotations could help to provide supplemental information for learning algorithms, especially when training data is sparse. This paper presents a formative study to uncover how experienced users explain game play in real-time strategy games. We report the results of our analysis of explanations and discuss their characteristics that could support the design of systems for use by experienced real-time strategy game users in specifying or annotating strategy-oriented behavior.

  5. G-quadruplex and G-rich sequence stimulate Pif1p-catalyzed downstream duplex DNA unwinding through reducing waiting time at ss/dsDNA junction

    PubMed Central

    Zhang, Bo; Wu, Wen-Qiang; Liu, Na-Nv; Duan, Xiao-Lei; Li, Ming; Dou, Shuo-Xing; Hou, Xi-Miao; Xi, Xu-Guang

    2016-01-01

    Alternative DNA structures that deviate from B-form double-stranded DNA such as G-quadruplex (G4) DNA can be formed by G-rich sequences that are widely distributed throughout the human genome. We have previously shown that Pif1p not only unfolds G4, but also unwinds the downstream duplex DNA in a G4-stimulated manner. In the present study, we further characterized the G4-stimulated duplex DNA unwinding phenomenon by means of single-molecule fluorescence resonance energy transfer. It was found that Pif1p did not unwind the partial duplex DNA immediately after unfolding the upstream G4 structure, but rather, it would dwell at the ss/dsDNA junction with a ‘waiting time’. Further studies revealed that the waiting time was in fact related to a protein dimerization process that was sensitive to ssDNA sequence and would become rapid if the sequence is G-rich. Furthermore, we identified that the G-rich sequence, as the G4 structure, equally stimulates duplex DNA unwinding. The present work sheds new light on the molecular mechanism by which G4-unwinding helicase Pif1p resolves physiological G4/duplex DNA structures in cells. PMID:27471032

  6. Rendering hospital budgets volume based and open ended to reduce waiting lists: does it work?

    PubMed

    van de Vijsel, Aart R; Engelfriet, Peter M; Westert, Gert P

    2011-04-01

    In the past decades fixed budgets for hospitals were replaced by reimbursement based on outputs in several countries in order to bring down waiting lists. This was also the case in the Netherlands where fixed global budgets were replaced by budgets that are to a large extent volume based and in practice open-ended. The objective of this study was to examine the effectiveness of this Dutch policy measure, which was implemented in 2001. We carried out a statistical analysis and interpretation of trends in Dutch hospital admission rates. We observed a significant turn in the development of in-patient admission rates after the abolition of budget caps in 2001: decreasing admission rates turned into an internationally exceptional increase of more than 3% per year. Day care admissions had already been rising explosively for two decades, but the pace increased after 2001. The increase in the number of admissions includes a broad range of patient categories that were not in the first place associated with long waiting times. The growth was attributable for a large part to admissions for observation of the patient and the evaluation of symptoms, not resulting in a definite medical diagnosis. We considered several factors, other than the availability of more resources, to explain the growth: the ageing of the population, making up for waiting list arrears, ditto for "under consumption" of unplanned care and, as to the growth of day care, substitution for inpatient care. However, these factors were all found to fall short as an explanation. Although waiting times have dropped since the change in the budget system, they continue to be long for several procedures. Our study indicates that making available more resources to admit patients, or otherwise an increase in hospital activity, do not in itself lead to equilibrium between demand and supply because the volume and composition of demand are partly induced by supply. We conclude that abolishing budget caps to solve waiting

  7. The Effects of Waiting for Treatment: A Meta-Analysis of Waitlist Control Groups in Randomized Controlled Trials for Social Anxiety Disorder.

    PubMed

    Steinert, Christiane; Stadter, Katja; Stark, Rudolf; Leichsenring, Falk

    2017-05-01

    Social anxiety disorder (SAD) is a highly prevalent mental disorder. However, little is known about how SAD changes in subjects who do not receive treatment. Waitlist control groups (WLCGs) are frequently included in randomized controlled trials (RCTs) on the treatment of mental disorders. Data from WLCGs are of value as they provide information on the untreated short-term course of a disorder and may serve as disorder-specific norms of change (benchmarks) against which treatment outcomes of SAD can be compared. Thus, we performed a meta-analysis focusing on the effects occurring in WLCGs of RCTs for SAD. Our study was conducted along the PRISMA guidelines. Thirty RCTs (total n = 2460) comprising 30 WLCGs and 47 treatment groups were included. Mean waiting time was 10.6 weeks. The pooled effect of waiting on SAD measures was g = 0.128 (95% CI: 0.057-0.199). Effects regarding other forms of anxiety, depression and functioning were of similarly small size. In contrast, change in the treatment groups was large, both within (g = 0.887) and between groups (g = 0.860). Our results show that for SAD, changes occurring in WLCGs of RCTs are small. The findings may serve as benchmarks in pilot studies of a new treatment or as an additional comparison in studies comparing two active treatments. For psychotherapy research in general, the small effect sizes found in WLCGs confirm that testing a treatment against a waiting list is not a very strict test. Further research on WLCGs in specific mental disorders is required, for example examining the expectancies of patients randomized to waiting. Copyright © 2016 John Wiley & Sons, Ltd. In clinical practice, patients suffering from a mental disorder often have to wait for treatment. By analyzing data from waitlist control groups we can gain estimates of symptom change that occur during waiting. It could be seen that waiting for treatment only results in a negligible effect. Thus, in the short-term (i.e., 10.6

  8. The PRCI study: design of a randomized clinical trial to evaluate a coping intervention for medical waiting periods used by women undergoing a fertility treatment.

    PubMed

    Ockhuijsen, Henrietta D L; van den Hoogen, Agnes; Macklon, Nickolas S; Boivin, Jacky

    2013-09-03

    Many medical situations necessitate a stressful period of waiting for potentially threatening test results. The medical waiting period is often associated with negative anticipatory anxiety and rumination about the outcome of treatment. Few evidence-based self-help coping interventions are available to assist individuals manage these periods. Theory and research suggest that positive reappraisal coping strategies may be particularly useful for this type of unpredictable and uncontrollable stressful context. The objective of this study is to investigate the effects of a Positive Reappraisal Coping Intervention (PRCI) on psychological well-being of women waiting for the outcome of their fertility treatment cycle. In a three-armed randomized controlled trial, the effectiveness of the PRCI will be tested. Consecutive patients undergoing in vitro fertilisation in a Dutch university hospital and meeting selection criteria will be invited to participate. Those who agree will be randomized to one of three experimental groups (N=372). The PRCI Intervention group will receive the intervention that comprises an explanatory leaflet and the 10 statements designed to promote positive reappraisal coping, to be read at least once in the morning, once in the evening. To capture the general impact of PRCI on psychological wellbeing patients will complete questionnaires before the waiting period (pre-intervention), on day ten of the 14-day waiting period (intervention) and six weeks after the start of the waiting period (post-intervention). To capture the specific effects of the PRCI during the waiting period, patients will also be asked to monitor daily their emotions and reactions during the 14-day waiting period. The primary outcome is general anxiety, measured by the Hospital Anxiety and Depression Scale. Secondary outcomes are positive and negative emotions during the waiting period, depression, quality of life, coping and treatment outcome. During recruitment for the RCT it was

  9. [Strategies for lung cancer with ischemic heart disease].

    PubMed

    Miyamoto, Nobuhiro; Kishimoto, Koji; Suehiro, Shouichi; Oda, Teiji; Tanabe, Kazuaki

    2015-04-01

    For lung cancer surgery which merged ischemic heart disease to need coronary artery treatments, the strategy is demanded on the timing of each treatment. Our department conforms to American College of Chest Physicians( ACCP) guideline and treatment strategies are decided as follows. 1) If right heart load has already occurred, we choose limited surgery for lung cancer. 2) Two-stage surgery is performed with principle. Coronary artery treatment is given priority to against left main trunk disease and unstable angina. 3) Simultaneous surgery is chosen for lung cancer more than stage II or lung cancer pressing neighboring organ and vessel not to be able to wait coronary artery treatments. Since 2007, we performed 4 simultaneous surgeries and experienced 3 pneumonia cases, 1 patient died in 5 months. We must decide a strategy in consideration of progress of the lung cancer and cardiac urgency.

  10. An exploration of the basis for patient complaints about the oldness of magazines in practice waiting rooms: cohort study

    PubMed Central

    Alrutz, Stowe; Moyes, Simon

    2014-01-01

    Objective To explore the basis for patient complaints about the oldness of most magazines in practice waiting rooms. Design Cohort study. Setting Waiting room of a general practice in Auckland, New Zealand. Participants 87 magazines stacked into three mixed piles and placed in the waiting room: this included non-gossipy magazines (Time magazine, the Economist, Australian Women’s Weekly, National Geographic, BBC History) and gossipy ones (not identified for fear of litigation). Gossipy was defined as having five or more photographs of celebrities on the front cover and most gossipy as having up to 10 such images. Interventions The magazines were marked with a unique number on the back cover, placed in three piles in the waiting room, and monitored twice weekly. Main outcome measures Disappearance of magazines less than 2 months old versus magazines 3-12 months old, the overall rate of loss of magazines, and the rate of loss of gossipy versus non-gossipy magazines. Results 47 of the 82 magazines with a visible date on the front cover were aged less than 2 months. 28 of these 47 (60%) magazines and 10 of the 35 (29%) older magazines disappeared (P=0.002). After 31 days, 41 of the 87 (47%, 95% confidence interval 37% to 58%) magazines had disappeared. None of the 19 non-gossipy magazines (the Economist and Time magazine) had disappeared compared with 26 of the 27 (96%) gossipy magazines (P<0.001). All 15 of the most gossipy magazines and all 19 of the non-gossipy magazines had disappeared by 31 days. The study was terminated at this point. Conclusions General practice waiting rooms contain mainly old magazines. This phenomenon relates to the disappearance of the magazines rather than to the supply of old ones. Gossipy magazines were more likely to disappear than non-gossipy ones. On the grounds of cost we advise practices to supply old copies of non-gossipy magazines. A waiting room science curriculum is urgently needed. PMID:25500116

  11. An exploration of the basis for patient complaints about the oldness of magazines in practice waiting rooms: cohort study.

    PubMed

    Arroll, Bruce; Alrutz, Stowe; Moyes, Simon

    2014-12-11

    To explore the basis for patient complaints about the oldness of most magazines in practice waiting rooms. Cohort study. Waiting room of a general practice in Auckland, New Zealand. 87 magazines stacked into three mixed piles and placed in the waiting room: this included non-gossipy magazines (Time magazine, the Economist, Australian Women's Weekly, National Geographic, BBC History) and gossipy ones (not identified for fear of litigation). Gossipy was defined as having five or more photographs of celebrities on the front cover and most gossipy as having up to 10 such images. The magazines were marked with a unique number on the back cover, placed in three piles in the waiting room, and monitored twice weekly. Disappearance of magazines less than 2 months old versus magazines 3-12 months old, the overall rate of loss of magazines, and the rate of loss of gossipy versus non-gossipy magazines. 47 of the 82 magazines with a visible date on the front cover were aged less than 2 months. 28 of these 47 (60%) magazines and 10 of the 35 (29%) older magazines disappeared (P=0.002). After 31 days, 41 of the 87 (47%, 95% confidence interval 37% to 58%) magazines had disappeared. None of the 19 non-gossipy magazines (the Economist and Time magazine) had disappeared compared with 26 of the 27 (96%) gossipy magazines (P<0.001). All 15 of the most gossipy magazines and none of the non-gossipy magazines [corrected] had disappeared by 31 days. The study was terminated at this point. General practice waiting rooms contain mainly old magazines. This phenomenon relates to the disappearance of the magazines rather than to the supply of old ones. Gossipy magazines were more likely to disappear than non-gossipy ones. On the grounds of cost we advise practices to supply old copies of non-gossipy magazines. A waiting room science curriculum is urgently needed. © Arroll et al 2014.

  12. Utilizing lean tools to improve value and reduce outpatient wait times in an Indian hospital.

    PubMed

    Miller, Richard; Chalapati, Nirisha

    2015-01-01

    This paper aims to demonstrate how lean tools were applied to some unique issues of providing healthcare in a developing country where many patients face challenges not found in developed countries. The challenges provide insight into how lean tools can be utilized to provide similar results across the world. This paper is based on a qualitative case study carried out by a master's student implementing lean at a hospital in India. This paper finds that lean tools such as value-stream mapping and root cause analysis can lead to dramatic reductions in waste and improvements in productivity. The problems of the majority of patients paying for their own healthcare and lacking transportation created scheduling problems that required patients to receive their diagnosis and pay for treatment within a single day. Many additional wastes were identified that were significantly impacting the hospital's ability to provide care. As a result of this project, average outpatient wait times were reduced from 1 hour to 15 minutes along with a significant increase in labor productivity. The results demonstrate how lean tools can increase value to the patients. It also provides are framework that can be utilized for healthcare providers in developed and developing countries to analyze their value streams to reduce waste. This paper is one of the first to address the unique issues of implementing lean to a healthcare setting in a developing country.

  13. Developmental Changes in Anger Expression and Attention Focus: Learning to Wait

    ERIC Educational Resources Information Center

    Cole, Pamela M.; Tan, Patricia Z.; Hall, Sarah E.; Zhang, Yiyun; Crnic, Keith A.; Blair, Clancy B.; Li, Runze

    2011-01-01

    Being able to wait is an essential part of self-regulation. In the present study, the authors examined the developmental course of changes in the latency to and duration of target-waiting behaviors by following 65 boys and 55 girls from rural and semirural economically strained homes from ages 18 months to 48 months. Age-related changes in latency…

  14. The Time of Grand Strategy

    DTIC Science & Technology

    2013-03-01

    representations of Chronos illustrate this combination: the god of time is either portrayed pushing the zodiac wheel,6 illustrating the rotation of seasons... The Time of Grand Strategy by Lieutenant-Colonel Arnaud Goujon French Army United States Army War...College Class of 2013 DISTRIBUTION STATEMENT: A Approved for Public Release Distribution is Unlimited COPYRIGHT STATEMENT: The author is not an

  15. Worth the Wait? Using Past Patterns to Determine Wait Periods for E-Books Released after Print

    ERIC Educational Resources Information Center

    Kohn, Karen

    2018-01-01

    This paper asks if there is an optimal wait period for e-books that balances libraries' desire to acquire books soon after their publication with the frequent desire to purchase books electronically whenever feasible. Analyzing 13,043 titles that Temple University Libraries received on its e-preferred approval plan in 2014-15, the author looks at…

  16. Factors Affecting the Selection of Patients on Waiting List: A Single Center Study.

    PubMed

    Can, Ö; Kasapoğlu, U; Boynueğri, B; Tuğcu, M; Çağlar Ruhi, B; Canbakan, M; Murat Gökçe, A; Ata, P; İzzet Titiz, M; Apaydın, S

    2015-06-01

    There is an increasing gap between organ supply and demand for cadaveric transplantation in our country. Our aim was to evaluate factors affecting selection of patients on waiting list at our hospital. Patients who have been waiting on list and who were transplanted were compared in order to find factors, which affected the selection of patients. Non-parametric Mann-Whitney U test was used for comparison and cox regression analysis was used to find the risk factors that decrease the probability of transplantation in this retrospective case-control study. Patients in the transplanted group were significantly younger, had relatively lower body mass index than the awaiting group. Cardiovascular diseases were more in the awaiting group than the transplanted group. There was no patient with diabetes in transplanted group, despite fifteen diabetic patients were in the awaiting group. Selected patients had lower immunologic risk with regard to peak panel reactive antibody levels. No significant difference was found for gender, hypertension, hyperlipidemia, viral serology, time spent on dialysis and on waiting list between two groups. With cox regression analysis female gender, older age, diabetes mellitus, high body mass index, positive hepatitis B serology and high levels of peak class 1-2 peak panel reactive antibody positivity were found as risk factors that decrease the probability of transplantation. A tendency for selection of low risk patients was found with this study. Time and energy consuming complications and short allograft survival after transplantation in high risk patients and the scarcity of cadaveric pool in our country may contribute to this tendency. Copyright © 2015 Elsevier Inc. All rights reserved.

  17. Mortality on the Waiting List for Lung Transplantation in Patients with Idiopathic Pulmonary Fibrosis: A Single-Centre Experience.

    PubMed

    Bennett, David; Fossi, Antonella; Bargagli, Elena; Refini, Rosa Metella; Pieroni, Maria; Luzzi, Luca; Ghiribelli, Claudia; Paladini, Piero; Voltolini, Luca; Rottoli, Paola

    2015-10-01

    Lung transplantation (LTX) is nowadays accepted as a treatment option for selected patients with end-stage pulmonary disease. Idiopathic pulmonary fibrosis (IPF) is characterized by the radiological and histologic appearance of usual interstitial pneumonia. It is associated with a poor prognosis, and LTX is considered an effective treatment to significantly modify the natural history of this disease. The aim of the present study was to analyse mortality during the waiting list in IPF patients at a single institution. A retrospective analysis on IPF patients (n = 90) referred to our Lung Transplant Program in the period 2001-2014 was performed focusing on patients' characteristics and associated risk factors. Diagnosis of IPF was associated with high mortality on the waiting list with respect to other diagnosis (p < 0.05). No differences in demographic, clinical, radiological data and time spent on the waiting list were observed between IPF patients who underwent to LTX or lost on the waiting list. Patients who died showed significant higher levels of pCO2 and needed higher flows of O2-therapy on effort (p < 0.05). Pulmonary function tests failed to predict mortality and no other medical conditions were associated with survival. Patients newly diagnosed with IPF, especially in small to medium lung transplant volume centres and in Countries where a long waiting list is expected, should be immediately referred to transplantation, delay results in increased mortality. Early identification of IPF patients with a rapid progressive phenotype is strongly needed.

  18. The Sit-and-Wait Hypothesis in Bacterial Pathogens: A Theoretical Study of Durability and Virulence.

    PubMed

    Wang, Liang; Liu, Zhanzhong; Dai, Shiyun; Yan, Jiawei; Wise, Michael J

    2017-01-01

    The intriguing sit-and-wait hypothesis predicts that bacterial durability in the external environment is positively correlated with their virulence. Since its first proposal in 1987, the hypothesis has been spurring debates in terms of its validity in the field of bacterial virulence. As a special case of the vector-borne transmission versus virulence tradeoff, where vector is now replaced by environmental longevity, there are only sporadic studies over the last three decades showing that environmental durability is possibly linked with virulence. However, no systematic study of these works is currently available and epidemiological analysis has not been updated for the sit-and-wait hypothesis since the publication of Walther and Ewald's (2004) review. In this article, we put experimental evidence, epidemiological data and theoretical analysis together to support the sit-and-wait hypothesis. According to the epidemiological data in terms of gain and loss of virulence (+/-) and durability (+/-) phenotypes, we classify bacteria into four groups, which are: sit-and-wait pathogens (++), vector-borne pathogens (+-), obligate-intracellular bacteria (--), and free-living bacteria (-+). After that, we dive into the abundant bacterial proteomic data with the assistance of bioinformatics techniques in order to investigate the two factors at molecular level thanks to the fast development of high-throughput sequencing technology. Sequences of durability-related genes sourced from Gene Ontology and UniProt databases and virulence factors collected from Virulence Factor Database are used to search 20 corresponding bacterial proteomes in batch mode for homologous sequences via the HMMER software package. Statistical analysis only identified a modest, and not statistically significant correlation between mortality and survival time for eight non-vector-borne bacteria with sit-and-wait potentials. Meanwhile, through between-group comparisons, bacteria with higher host-mortality are

  19. The Sit-and-Wait Hypothesis in Bacterial Pathogens: A Theoretical Study of Durability and Virulence

    PubMed Central

    Wang, Liang; Liu, Zhanzhong; Dai, Shiyun; Yan, Jiawei; Wise, Michael J.

    2017-01-01

    The intriguing sit-and-wait hypothesis predicts that bacterial durability in the external environment is positively correlated with their virulence. Since its first proposal in 1987, the hypothesis has been spurring debates in terms of its validity in the field of bacterial virulence. As a special case of the vector-borne transmission versus virulence tradeoff, where vector is now replaced by environmental longevity, there are only sporadic studies over the last three decades showing that environmental durability is possibly linked with virulence. However, no systematic study of these works is currently available and epidemiological analysis has not been updated for the sit-and-wait hypothesis since the publication of Walther and Ewald’s (2004) review. In this article, we put experimental evidence, epidemiological data and theoretical analysis together to support the sit-and-wait hypothesis. According to the epidemiological data in terms of gain and loss of virulence (+/-) and durability (+/-) phenotypes, we classify bacteria into four groups, which are: sit-and-wait pathogens (++), vector-borne pathogens (+-), obligate-intracellular bacteria (--), and free-living bacteria (-+). After that, we dive into the abundant bacterial proteomic data with the assistance of bioinformatics techniques in order to investigate the two factors at molecular level thanks to the fast development of high-throughput sequencing technology. Sequences of durability-related genes sourced from Gene Ontology and UniProt databases and virulence factors collected from Virulence Factor Database are used to search 20 corresponding bacterial proteomes in batch mode for homologous sequences via the HMMER software package. Statistical analysis only identified a modest, and not statistically significant correlation between mortality and survival time for eight non-vector-borne bacteria with sit-and-wait potentials. Meanwhile, through between-group comparisons, bacteria with higher host

  20. Popular Media Portrayals of Inequity and School Reform in "The Wire" and "Waiting for 'Superman'"

    ERIC Educational Resources Information Center

    Gerstl-Pepin, Cynthia

    2015-01-01

    Two popular media forms are examined--the documentary film "Waiting for 'Superman'" and the HBO television series, "The Wire"--that present distinct, and at times conflicting, depictions of how to address educational inequity. Qualitative media content analysis was used to analyze the two media documents and to situate them…

  1. Load Balancing Strategies for Multiphase Flows on Structured Grids

    NASA Astrophysics Data System (ADS)

    Olshefski, Kristopher; Owkes, Mark

    2017-11-01

    The computation time required to perform large simulations of complex systems is currently one of the leading bottlenecks of computational research. Parallelization allows multiple processing cores to perform calculations simultaneously and reduces computational times. However, load imbalances between processors waste computing resources as processors wait for others to complete imbalanced tasks. In multiphase flows, these imbalances arise due to the additional computational effort required at the gas-liquid interface. However, many current load balancing schemes are only designed for unstructured grid applications. The purpose of this research is to develop a load balancing strategy while maintaining the simplicity of a structured grid. Several approaches are investigated including brute force oversubscription, node oversubscription through Message Passing Interface (MPI) commands, and shared memory load balancing using OpenMP. Each of these strategies are tested with a simple one-dimensional model prior to implementation into the three-dimensional NGA code. Current results show load balancing will reduce computational time by at least 30%.

  2. 13. 'WAITING AT THE DRAWBRIDGE.' THE COAL SCHOONER LUCY MAY ...

    Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey

    13. 'WAITING AT THE DRAWBRIDGE.' THE COAL SCHOONER LUCY MAY WAITING AT THE DRAW, JUNE 19, 1896. Photocopy of photograph (original glass plate negative #T89 in the collection of the Annisquam Historical Society, Annisquam, Massachusetts). Photographer: Martha Harvey (1862-1949). (The handwritten legend along the top edge of the photograph is scratched in the emulsion of the original glass plate negative. Consequently it reads in reverse when printed.) - Annisquam Bridge, Spanning Lobster Cove between Washington & River Streets, Gloucester, Essex County, MA

  3. Fixation of competing strategies when interacting agents differ in the time scale of strategy updating

    NASA Astrophysics Data System (ADS)

    Zhang, Jianlei; Weissing, Franz J.; Cao, Ming

    2016-09-01

    A commonly used assumption in evolutionary game theory is that natural selection acts on individuals in the same time scale; e.g., players use the same frequency to update their strategies. Variation in learning rates within populations suggests that evolutionary game theory may not necessarily be restricted to uniform time scales associated with the game interaction and strategy adaption evolution. In this study, we remove this restricting assumption by dividing the population into fast and slow groups according to the players' strategy updating frequencies and investigate how different strategy compositions of one group influence the evolutionary outcome of the other's fixation probabilities of strategies within its own group. Analytical analysis and numerical calculations are performed to study the evolutionary dynamics of strategies in typical classes of two-player games (prisoner's dilemma game, snowdrift game, and stag-hunt game). The introduction of the heterogeneity in strategy-update time scales leads to substantial changes in the evolution dynamics of strategies. We provide an approximation formula for the fixation probability of mutant types in finite populations and study the outcome of strategy evolution under the weak selection. We find that although heterogeneity in time scales makes the collective evolutionary dynamics more complicated, the possible long-run evolutionary outcome can be effectively predicted under technical assumptions when knowing the population composition and payoff parameters.

  4. 24 CFR 982.207 - Waiting list: Local preferences in admission to program.

    Code of Federal Regulations, 2011 CFR

    2011-04-01

    ... 24 Housing and Urban Development 4 2011-04-01 2011-04-01 false Waiting list: Local preferences in... Admission to Tenant-Based Program § 982.207 Waiting list: Local preferences in admission to program. (a) Establishment of PHA local preferences. (1) The PHA may establish a system of local preferences for selection of...

  5. 24 CFR 882.513 - Public notice to low-income families; waiting list.

    Code of Federal Regulations, 2010 CFR

    2010-04-01

    ...-income families; waiting list. (a) Public notice to low-income Families. (1) If the PHA does not have a... 24 Housing and Urban Development 4 2010-04-01 2010-04-01 false Public notice to low-income families; waiting list. 882.513 Section 882.513 Housing and Urban Development Regulations Relating to...

  6. 24 CFR 882.513 - Public notice to low-income families; waiting list.

    Code of Federal Regulations, 2011 CFR

    2011-04-01

    ...-income families; waiting list. (a) Public notice to low-income Families. (1) If the PHA does not have a... 24 Housing and Urban Development 4 2011-04-01 2011-04-01 false Public notice to low-income families; waiting list. 882.513 Section 882.513 Housing and Urban Development REGULATIONS RELATING TO...

  7. 24 CFR 882.513 - Public notice to low-income families; waiting list.

    Code of Federal Regulations, 2013 CFR

    2013-04-01

    ...-income families; waiting list. (a) Public notice to low-income Families. (1) If the PHA does not have a... 24 Housing and Urban Development 4 2013-04-01 2013-04-01 false Public notice to low-income families; waiting list. 882.513 Section 882.513 Housing and Urban Development REGULATIONS RELATING TO...

  8. 24 CFR 882.513 - Public notice to low-income families; waiting list.

    Code of Federal Regulations, 2014 CFR

    2014-04-01

    ...-income families; waiting list. (a) Public notice to low-income Families. (1) If the PHA does not have a... 24 Housing and Urban Development 4 2014-04-01 2014-04-01 false Public notice to low-income families; waiting list. 882.513 Section 882.513 Housing and Urban Development REGULATIONS RELATING TO...

  9. 24 CFR 882.513 - Public notice to low-income families; waiting list.

    Code of Federal Regulations, 2012 CFR

    2012-04-01

    ...-income families; waiting list. (a) Public notice to low-income Families. (1) If the PHA does not have a... 24 Housing and Urban Development 4 2012-04-01 2012-04-01 false Public notice to low-income families; waiting list. 882.513 Section 882.513 Housing and Urban Development REGULATIONS RELATING TO...

  10. Time management strategies in nursing practice.

    PubMed

    Waterworth, Susan

    2003-09-01

    With the increasing emphasis on efficiency and effectiveness in health care, how a nurse manages her time is an important consideration. Whilst time management is recognized as an important component of work performance and professional nursing practice, the reality of this process in nursing practice has been subject to scant empirical investigation. To explore how nurses organize and manage their time. A qualitative study was carried out, incorporating narratives (22 nurses), focus groups (24 nurses) and semi-structured interviews (22 nurses). In my role as practitioner researcher I undertook observation and had informal conversations, which provided further data. Study sites were five health care organizations in the United Kingdom during 1995-1999. Time management is complex, with nurses using a range of time management strategies and a repertoire of actions. Two of these strategies, namely routinization and prioritizing, are discussed, including their implications for understanding time management by nurses in clinical practice. Ignoring the influence of 'others', the team and the organization perpetuates a rather individualistic and self-critical perspective of time management. This may lead to a failure to address problems in the organizing of work, and the co-ordinating of care involving other health care workers.

  11. Do case-generic measures of queue performance for bypass surgery accurately reflect the waiting-list experiences of those most urgent?

    PubMed

    Burstein, Jason; Lee, Douglas S; Alter, David A

    2006-02-01

    Queue performance is typically assessed using generic measures, which capture the queue in aggregate. The objective of this study was to examine whether case-generic measures of queue performance appropriately reflected the waiting-list experiences of those patients with greatest disease severity. We examined the queue for isolated coronary artery bypass grafting (CABG) in Ontario between April 1993 and March 2000 using data obtained from the Cardiac Care Network. Our primary measure of queue performance was the proportion of patients who received their bypass surgery within their recommended maximum waiting times (%RMWTs) in any given month. We compared case-generic measures of queue performance to case-specific measures of queue performance stratified by urgency level. The queue was largely comprised of elective cases ranging from 73% (1993) to 57%(1999). Urgent patients comprised the minority of the queue ranging from 14% (1993) to 20% (1999). Case-generic month-to-month variations in the percentage of cases completed within RMWTs (an aggregated waiting list measure encompassing the characteristics of all patients in the queue) closely resembled the experiences of elective patients (R2 = 0.81), but conversely, bore little relationship to the waiting-list experiences of those most urgent (R2 = 0.15). Case-generic measures of queue performance for bypass surgery in Ontario were not reflective of the waiting-list experiences of those most urgent. Our results reinforce the concept that urgency-specific waiting list monitoring systems are required to best evaluate and appropriately respond to fluctuations in queue performance.

  12. Evaluating a community saturation model of abstinence education: an application of social marketing strategies.

    PubMed

    Tanner, John F; Anne Raymond, Mary; Ladd, Stacey D

    2009-01-01

    This study examines a community saturation program, a social marketing strategy, promoting abstinence education and evaluates the effects of this strategy on adolescents' attitudes and sexual behaviors. The study also examines components of the strategy to determine which program element was most influential. The Worth the Wait program was implemented in five counties in Texas beginning in 1999 for the first county and in 2000 and 2001 for the other four counties. A total of 2007 students in grades 7 through 12 were tracked and answered an end-of-the-year post-program survey after varying time periods of school program participation. Results indicate that a saturation program can be effective in reducing teen pregnancy.

  13. Effectiveness of Multimedia for Transplant Preparation for Kidney Transplant Waiting List Patients.

    PubMed

    Charoenthanakit, C; Junchotikul, P; Sittiudomsuk, R; Saiyud, A; Pratumphai, P

    2016-04-01

    A multimedia program could effectively advise patients about preparing for transplantation while on the waiting list for a kidney transplant. This study aimed to compare knowledge about transplant preparation for patients on a kidney transplant waiting list before and after participating in a multimedia program, and to evaluate patient satisfaction with the multimedia program. Research design was quasiexperimental with the use of 1 group. Subjects were 186 patients on the kidney transplant waiting list after HLA matching in Ramathibodi Hospital. The questionnaires were developed by the researchers. The statistical tools used were basic statistics, percentage, average, standard deviation, and the difference of score between before and after participation in the multimedia program (t test). The evaluation knowledge for transplant preparation for kidney transplant waiting list patients after participating in the multimedia program averaged 85.40%, and there was an increased improvement of score by an average 3.27 out of a possible full score of 20 (P < .05). The result of patient satisfaction for the multimedia program had good average, 4.58. Copyright © 2016 Elsevier Inc. All rights reserved.

  14. The impact of nurse practitioner services on cost, quality of care, satisfaction and waiting times in the emergency department: a systematic review.

    PubMed

    Jennings, Natasha; Clifford, Stuart; Fox, Amanda R; O'Connell, Jane; Gardner, Glenn

    2015-01-01

    To provide the best available evidence to determine the impact of nurse practitioner services on cost, quality of care, satisfaction and waiting times in the emergency department for adult patients. The delivery of quality care in the emergency department is emerging as one of the most important service indicators in health delivery. Increasing service pressures in the emergency department have resulted in the adoption of service innovation models: the most common and rapidly expanding of these is emergency nurse practitioner services. The rapid uptake of emergency nurse practitioner service in Australia has outpaced the capacity to evaluate this service model in terms of outcomes related to safety and quality of patient care. Previous research is now outdated and not commensurate with the changing domain of delivering emergency care with nurse practitioner services. A comprehensive search of four electronic databases from 2006 to 2013 was conducted to identify research evaluating nurse practitioner service impact in the emergency department. English language articles were sought using MEDLINE, CINAHL, Embase and Cochrane and included two previous systematic reviews completed five and seven years ago. A three step approach was used. Following a comprehensive search, two reviewers assessed all identified studies against the inclusion criteria. From the original 1013 studies, 14 papers were retained for critical appraisal on methodological quality by two independent reviewers and data were extracted using standardised tools. Narrative synthesis was conducted to summarise and report the findings as insufficient data was available for meta-analysis of results. This systematic review has shown that emergency nurse practitioner service has a positive impact on quality of care, patient satisfaction and waiting times. There was insufficient evidence to draw conclusions regarding outcomes of a cost benefit analysis. Synthesis of the available research attempts to provide an

  15. Community-based birth waiting homes in Northern Sierra Leone: Factors influencing women's use.

    PubMed

    Kyokan, Michiko; Whitney-Long, Melissa; Kuteh, Mabel; Raven, Joanna

    2016-08-01

    to explore the factors influencing women's use of birth waiting homes in the Northern Bombali district, Sierra Leone. this was a descriptive exploratory study using qualitative research methodology, which included in depth interviews, key informant interviews, focus group discussions, document review and observations. two chiefdoms in the Northern Bombali district, Sierra Leone. eight interviews were conducted with women who had delivered in the past one year and used birth waiting homes; eight key informant interviews with a project manager, birth waiting homes hosts, and community members; thirteen women who delivered in the past year without using birth waiting homes (four interviews and two focus group discussions). there are several factors influencing the use of birth waiting homes (BWHs) including: past experience of childbirth, promotion of the birth waiting homes by traditional birth attendance, distance and costs of transport to the homes, child care and other family commitments, family's views of the importance of the homes, the costs of food during women's stay, and information given to women and families about when and how to use the homes. some barriers, especially those related to family commitments and costs of food, are challenging to solve. In order to make a BWH a user-friendly and viable option, it may be necessary to adjust ways in which BWHs are used. Good linkage with the health system is strength of the programme. However, further strengthening of community participation in monitoring and managing the BWHs is needed for the long term success and sustainability of the BWHs. Copyright © 2016 Elsevier Ltd. All rights reserved.

  16. Characteristics of Older Adults on Waiting Lists for Meals on Wheels: Identifying Areas for Intervention

    PubMed Central

    Thomas, Kali S.; Smego, Raul; Akobundu, Ucheoma; Dosa, David

    2016-01-01

    The purpose of this study was to characterize the population of seniors on Meals on Wheels’ (MOW) waiting lists and identify their rate of depression, anxiety, falls, and fear of falling. Data come from surveys of 626 seniors on waiting lists across the country and the 2013 National Health and Aging Trends Study (NHATS). Results suggest that seniors on waiting lists for MOW were more likely to be widowed, less educated, older, Black, Hispanic, and receive Medicaid than the population of community-dwelling older adults. In addition, 31% of seniors on MOW waiting lists were depressed, compared with 12% of seniors in the national population (p < .001), and 28% exhibited signs and symptoms of anxiety, compared with 10% of the national population of seniors (p < .001). Seniors on waiting lists were significantly more likely to have fallen in the last month and be fearful of falling than the national population of seniors (p < .001). Individuals on MOW waiting lists are a vulnerable and high-risk group. By seeking to better understand clients’ needs, appropriate services can be tailored to promote independent living and improve older adults’ well-being. PMID:26597791

  17. Stress-reducing effects of real and artificial nature in a hospital waiting room.

    PubMed

    Beukeboom, Camiel J; Langeveld, Dion; Tanja-Dijkstra, Karin

    2012-04-01

    This field study investigated the potential stress-reducing effects of exposure to real or artificial nature on patients in a hospital waiting room. Additionally, it was investigated whether perceived attractiveness of the room could explain these effects. In this between-patients experimental design, patients were exposed to one of the following: real plants, posters of plants, or no nature (control). These conditions were alternately applied to two waiting rooms. The location of this study was two waiting rooms at the Radiology Department of a Dutch hospital. The subjects comprised 457 patients (60% female and 40% male) who were mostly scheduled for echocardiogram, dual-energy x-ray absorptiometry, magnetic resonance imaging, computed tomography scans, or nuclear research. Patients exposed to real plants, as well as patients exposed to posters of plants, report lower levels of experienced stress compared to the control condition. Further analyses show that these small but significant effects of exposure to nature are partially mediated by the perceived attractiveness of the waiting room. Natural elements in hospital environments have the potential to reduce patients' feelings of stress. By increasing the attractiveness of the waiting room by adding either real plants or posters of plants, hospitals can create a pleasant atmosphere that positively influences patients' well-being.

  18. Trends in Wait-list Mortality in Children Listed for Heart Transplantation in the United States

    PubMed Central

    Singh, Tajinder P.; Almond, Christopher S.; Piercey, Gary; Gauvreau, Kimberlee

    2014-01-01

    We sought to evaluate trends in overall and race-specific pediatric heart transplant (HT) wait-list mortality in the United States (US) during the last 20 years. We identified all children <18 years old listed for primary HT in the US during 1989–2009 (N=8096, 62% white, 19% black, 13% Hispanic, 6% other) using the Organ Procurement and Transplant Network database. Wait-list mortality was assessed in 4 successive eras (1989–1994, 1995–1999, 2000–2004, and 2005–2009). Overall wait-list mortality declined in successive eras (26%, 23%, 18% and 13%, respectively). The decline across eras remained significant in adjusted analysis (hazard ratio [HR] 0.70 in successive eras, 95% confidence interval [CI] 0.67, 0.74) and was 67% lower for children listed during 2005–2009 vs. those listed during 1989–1994 (HR 0.33, CI 0.28, 0.39). In models stratified by race, wait-list mortality decreased in all racial groups in successive eras. In models stratified by era, minority children were not at higher risk of wait-list mortality in the most recent era. We conclude that the risk of wait-list mortality among US children listed for HT has decreased by two-thirds during the last 20 years. Racial gaps in wait-list mortality present variably in the past are not present in the current era. PMID:21883920

  19. A comparison of walk-in counselling and the wait list model for delivering counselling services.

    PubMed

    Stalker, Carol A; Riemer, Manuel; Cait, Cheryl-Anne; Horton, Susan; Booton, Jocelyn; Josling, Leslie; Bedggood, Joanna; Zaczek, Margaret

    2016-10-01

    Walk-in counselling has been used to reduce wait times but there are few controlled studies to compare outcomes between walk-in and the traditional model of service delivery. To compare change in psychological distress by clients receiving services from two models of service delivery, a walk-in counselling model and a traditional counselling model involving a wait list. Mixed-methods sequential explanatory design including quantitative comparison of groups with one pre-test and two follow-ups, and qualitative analysis of interviews with a sub-sample. Five-hundred and twenty-four participants ≥16 years were recruited from two Family Counselling Agencies; the General Health Questionnaire-12 assessed change in psychological distress. Hierarchical linear modelling revealed clients of the walk-in model improved faster and were less distressed at the four-week follow-up compared to the traditional service delivery model. Ten weeks later, both groups had improved and were similar. Participants receiving instrumental services prior to baseline improved more slowly. The qualitative data confirmed participants highly valued the accessibility of the walk-in model, and were frustrated by the lengthy waits associated with the traditional model. This study improves methodologically on previous studies of walk-in counselling, an approach to service delivery not conducive to randomized controlled trials.

  20. Impact of insurance status on heart transplant wait-list mortality for patients with left ventricular assist devices.

    PubMed

    Emani, Sitaramesh; Tumin, Dmitry; Foraker, Randi E; Hayes, Don; Smith, Sakima A

    2017-02-01

    To test the hypotheses that receipt of Medicaid or Medicare (versus private insurance or self-pay) and low socioeconomic status (SES) leads to increased mortality and lower chances of transplantation among heart transplant (HTx) candidates with bridge to transplant left ventricular assist devices (BTT LVADs). Survival while awaiting HTx has improved with the use of BTT LVADs. However, it is unknown whether benefits extend uniformly across patient groups based on insurance status. Data from the United Network of Organ Sharing (UNOS) registry between 2006 and 2015 were examined for first-time HTx candidates ≥18 and <65 years who had LVAD support while wait-listed. Multivariable survival analysis was conducted on competing outcomes of mortality and time to transplant stratified by insurance source at the time of listing. Additional covariates included demographic information and SES. A total of 4626 patients met inclusion criteria, with 3353 being used for multivariable analysis. A majority of patients (68%) underwent HTx during the study period. BTT LVAD wait-list mortality was found to be greater among Medicaid beneficiaries vs. private insurance (SHR 1.57, P<.05) and did not diminish with the inclusion of neighborhood SES. Transplantation as an outcome demonstrated no difference by insurance status. Medicaid insurance status is associated with worse survival on the HTx wait-list among patients with BTT LVADs, although access to transplant was not different among insurance groups. The disparity is not reflective of SES in general and therefore points to other barriers inherent to Medicaid beneficiaries. © 2016 John Wiley & Sons A/S. Published by John Wiley & Sons Ltd.

  1. Five Strategies of Successful Part-Time Work.

    ERIC Educational Resources Information Center

    Corwin, Vivien; Lawrence, Thomas B.; Frost, Peter J.

    2001-01-01

    Identifies commonalities in the approaches of successful part-time professionals. Discusses five strategies for success: (1) communicating work-life priorities and schedules to the organization; (2) making the business case for part-time arrangements; (3) establishing time management routines; (4) cultivating advocates in senior management; and…

  2. Time Spent in Indirect Nursing Care

    DTIC Science & Technology

    1983-09-01

    divisions of labor by nursing personnel: a) direct patient care (28-35%); b) indirect care (50-62%); and c) personal time (10-15%). In comparing the... personal time (13-18%) (Kuhn, 1983). It must be noted that, in the VA data, wait time (time waiting to render care) has been subsumed under personal time...interval during the entire eight-hour shift. The first task observed being performed by the monitored person was the activity documented. 5 U. ,. 1 -V

  3. Willingness to Pay for a Maternity Waiting Home Stay in Zambia.

    PubMed

    Vian, Taryn; White, Emily E; Biemba, Godfrey; Mataka, Kaluba; Scott, Nancy

    2017-03-01

    Complications of pregnancy and childbirth can pose serious risks to the health of women, especially in resource-poor settings. Zambia has been implementing a program to improve access to emergency obstetric and neonatal care, including expansion of maternity waiting homes-residential facilities located near a qualified medical facility where a pregnant woman can wait to give birth. Yet it is unclear how much support communities and women would be willing to provide to help fund the homes and increase sustainability. We conducted a mixed-methods study to estimate willingness to pay for maternity waiting home services based on a survey of 167 women, men, and community elders. We also collected qualitative data from 16 focus group discussions to help interpret our findings in context. The maximum willingness to pay was 5.0 Zambian kwacha or $0.92 US dollars per night of stay. Focus group discussions showed that willingness to pay is dependent on higher quality of services such as food service and suggested that the pricing policy (by stay or by night) could influence affordability and use. While Zambians seem to value and be willing to contribute a modest amount for maternity waiting home services, planners must still address potential barriers that may prevent women from staying at the shelters. These include cash availability and affordability for the poorest households. © 2016 by the American College of Nurse-Midwives.

  4. Willingness to Pay for a Maternity Waiting Home Stay in Zambia

    PubMed Central

    White, Emily E.; Biemba, Godfrey; Mataka, Kaluba; Scott, Nancy

    2016-01-01

    Introduction Complications of pregnancy and childbirth can pose serious risks to the health of women, especially in resource‐poor settings. Zambia has been implementing a program to improve access to emergency obstetric and neonatal care, including expansion of maternity waiting homes‐residential facilities located near a qualified medical facility where a pregnant woman can wait to give birth. Yet it is unclear how much support communities and women would be willing to provide to help fund the homes and increase sustainability. Methods We conducted a mixed‐methods study to estimate willingness to pay for maternity waiting home services based on a survey of 167 women, men, and community elders. We also collected qualitative data from 16 focus group discussions to help interpret our findings in context. Results The maximum willingness to pay was 5.0 Zambian kwacha or $0.92 US dollars per night of stay. Focus group discussions showed that willingness to pay is dependent on higher quality of services such as food service and suggested that the pricing policy (by stay or by night) could influence affordability and use. Discussion While Zambians seem to value and be willing to contribute a modest amount for maternity waiting home services, planners must still address potential barriers that may prevent women from staying at the shelters. These include cash availability and affordability for the poorest households. PMID:28419708

  5. Comparison of listing strategies for allosensitized heart transplant candidates requiring transplant at high urgency: a decision model analysis.

    PubMed

    Feingold, B; Webber, S A; Bryce, C L; Park, S Y; Tomko, H E; Comer, D M; Mahle, W T; Smith, K J

    2015-02-01

    Allosensitized children who require a negative prospective crossmatch have a high risk of death awaiting heart transplantation. Accepting the first suitable organ offer, regardless of the possibility of a positive crossmatch, would improve waitlist outcomes but it is unclear whether it would result in improved survival at all times after listing, including posttransplant. We created a Markov decision model to compare survival after listing with a requirement for a negative prospective donor cell crossmatch (WAIT) versus acceptance of the first suitable offer (TAKE). Model parameters were derived from registry data on status 1A (highest urgency) pediatric heart transplant listings. We assumed no possibility of a positive crossmatch in the WAIT strategy and a base-case probability of a positive crossmatch in the TAKE strategy of 47%, as estimated from cohort data. Under base-case assumptions, TAKE showed an incremental survival benefit of 1.4 years over WAIT. In multiple sensitivity analyses, including variation of the probability of a positive crossmatch from 10% to 100%, TAKE was consistently favored. While model input data were less well suited to comparing survival when awaiting transplantation across a negative virtual crossmatch, our analysis suggests that taking the first suitable organ offer under these circumstances is also favored. © Copyright 2015 The American Society of Transplantation and the American Society of Transplant Surgeons.

  6. A Tribute to Waiting Room Moms Everywhere

    ERIC Educational Resources Information Center

    Ansfield, Mara

    2008-01-01

    Waiting rooms are oases for mothers of children with special needs. They congregate in these small holding areas, sitting on musty couches, while their children receive the latest therapeutic interventions. Sometimes they sit quietly, sneaking glances at each other while pretending to read year-old "People" magazines. Sometimes they crawl under a…

  7. 19. WILEY CITY LINE STONE TROLLEY WAITING STATION ON ...

    Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey

    19. WILEY CITY LINE - STONE TROLLEY WAITING STATION ON CONGDON ORCHARD PROPERTY - Yakima Valley Transportation Company Interurban Railroad, Connecting towns of Yakima, Selah & Wiley City, Yakima, Yakima County, WA

  8. Hospital outpatient perceptions of the physical environment of waiting areas: the role of patient characteristics on atmospherics in one academic medical center

    PubMed Central

    Tsai, Chun-Yen; Wang, Mu-Chia; Liao, Wei-Tsen; Lu, Jui-Heng; Sun, Pi-hung; Lin, Blossom Yen-Ju; Breen, Gerald-Mark

    2007-01-01

    Background This study examines hospital outpatient perceptions of the physical environment of the outpatient waiting areas in one medical center. The relationship of patient characteristics and their perceptions and needs for the outpatient waiting areas are also examined. Method The examined medical center consists of five main buildings which house seventeen primary waiting areas for the outpatient clinics of nine medical specialties: 1) Internal Medicine; 2) Surgery; 3) Ophthalmology; 4) Obstetrics-Gynecology and Pediatrics; 5) Chinese Medicine; 6) Otolaryngology; 7) Orthopedics; 8) Family Medicine; and 9) Dermatology. A 15-item structured questionnaire was developed to rate patient satisfaction covering the four dimensions of the physical environments of the outpatient waiting areas: 1) visual environment; 2) hearing environment; 3) body contact environment; and 4) cleanliness. The survey was conducted between November 28, 2005 and December 8, 2005. A total of 680 outpatients responded. Descriptive, univariate, and multiple regression analyses were applied in this study. Results All of the 15 items were ranked as relatively high with a range from 3.362 to 4.010, with a neutral score of 3. Using a principal component analysis' summated scores of four constructed dimensions of patient satisfaction with the physical environments (i.e. visual environment, hearing environment, body contact environment, and cleanliness), multiple regression analyses revealed that patient satisfaction with the physical environment of outpatient waiting areas was associated with gender, age, visiting frequency, and visiting time. Conclusion Patients' socio-demographics and context backgrounds demonstrated to have effects on their satisfaction with the physical environment of outpatient waiting areas. In addition to noticing the overall rankings for less satisfactory items, what should receive further attention is the consideration of the patients' personal characteristics when

  9. Hospital outpatient perceptions of the physical environment of waiting areas: the role of patient characteristics on atmospherics in one academic medical center.

    PubMed

    Tsai, Chun-Yen; Wang, Mu-Chia; Liao, Wei-Tsen; Lu, Jui-Heng; Sun, Pi-Hung; Lin, Blossom Yen-Ju; Breen, Gerald-Mark

    2007-12-05

    This study examines hospital outpatient perceptions of the physical environment of the outpatient waiting areas in one medical center. The relationship of patient characteristics and their perceptions and needs for the outpatient waiting areas are also examined. The examined medical center consists of five main buildings which house seventeen primary waiting areas for the outpatient clinics of nine medical specialties: 1) Internal Medicine; 2) Surgery; 3) Ophthalmology; 4) Obstetrics-Gynecology and Pediatrics; 5) Chinese Medicine; 6) Otolaryngology; 7) Orthopedics; 8) Family Medicine; and 9) Dermatology. A 15-item structured questionnaire was developed to rate patient satisfaction covering the four dimensions of the physical environments of the outpatient waiting areas: 1) visual environment; 2) hearing environment; 3) body contact environment; and 4) cleanliness. The survey was conducted between November 28, 2005 and December 8, 2005. A total of 680 outpatients responded. Descriptive, univariate, and multiple regression analyses were applied in this study. All of the 15 items were ranked as relatively high with a range from 3.362 to 4.010, with a neutral score of 3. Using a principal component analysis' summated scores of four constructed dimensions of patient satisfaction with the physical environments (i.e. visual environment, hearing environment, body contact environment, and cleanliness), multiple regression analyses revealed that patient satisfaction with the physical environment of outpatient waiting areas was associated with gender, age, visiting frequency, and visiting time. Patients' socio-demographics and context backgrounds demonstrated to have effects on their satisfaction with the physical environment of outpatient waiting areas. In addition to noticing the overall rankings for less satisfactory items, what should receive further attention is the consideration of the patients' personal characteristics when redesigning more comfortable and customized

  10. Timing Recovery Strategies in Magnetic Recording Systems

    NASA Astrophysics Data System (ADS)

    Kovintavewat, Piya

    At some point in a digital communications receiver, the received analog signal must be sampled. Good performance requires that these samples be taken at the right times. The process of synchronizing the sampler with the received analog waveform is known as timing recovery. Conventional timing recovery techniques perform well only when operating at high signal-to-noise ratio (SNR). Nonetheless, iterative error-control codes allow reliable communication at very low SNR, where conventional techniques fail. This paper provides a detailed review on the timing recovery strategies based on per-survivor processing (PSP) that are capable of working at low SNR. We also investigate their performance in magnetic recording systems because magnetic recording is a primary method of storage for a variety of applications, including desktop, mobile, and server systems. Results indicate that the timing recovery strategies based on PSP perform better than the conventional ones and are thus worth being employed in magnetic recording systems.

  11. Quality management: reduction of waiting time and efficiency enhancement in an ENT-university outpatients' department

    PubMed Central

    Helbig, Matthias; Helbig, Silke; Kahla-Witzsch, Heike A; May, Angelika

    2009-01-01

    Background Public health systems are confronted with constantly rising costs. Furthermore, diagnostic as well as treatment services become more and more specialized. These are the reasons for an interdisciplinary project on the one hand aiming at simplification of planning and scheduling patient appointments, on the other hand at fulfilling all requirements of efficiency and treatment quality. Methods As to understanding procedure and problem solving activities, the responsible project group strictly proceeded with four methodical steps: actual state analysis, analysis of causes, correcting measures, and examination of effectiveness. Various methods of quality management, as for instance opinion polls, data collections, and several procedures of problem identification as well as of solution proposals were applied. All activities were realized according to the requirements of the clinic's ISO 9001:2000 certified quality management system. The development of this project is described step by step from planning phase to inauguration into the daily routine of the clinic and subsequent control of effectiveness. Results Five significant problem fields could be identified. After an analysis of causes the major remedial measures were: installation of a patient telephone hotline, standardization of appointment arrangements for all patients, modification of the appointments book considering the reason for coming in planning defined working periods for certain symptoms and treatments, improvement of telephonic counselling, and transition to flexible time planning by daily updates of the appointments book. After implementation of these changes into the clinic's routine success could be demonstrated by significantly reduced waiting times and resulting increased patient satisfaction. Conclusion Systematic scrutiny of the existing organizational structures of the outpatients' department of our clinic by means of actual state analysis and analysis of causes revealed the necessity

  12. Racial/ethnic disparities in emergency department waiting time for stroke patients in the United States.

    PubMed

    Karve, Sudeep J; Balkrishnan, Rajesh; Mohammad, Yousef M; Levine, Deborah A

    2011-01-01

    Emergency department waiting time (EDWT), the time from arrival at the ED to evaluation by an emergency physician, is a critical component of acute stroke care. We assessed racial/ethnic differences in EDWT in a national sample of patients with ischemic or hemorrhagic stroke. We identified 543 ED visits for ischemic stroke (International Classification of Diseases, Ninth Revision, Clinical Modification [ICD-9-CM] codes 433.x1, 434.xx, and 436.xx) and hemorrhagic stroke (ICD-9-CM codes 430.xx, 431.xx, and 432.xx) in persons age ≥ 18 years representing 2.1 million stroke-related ED visits in the United States using the National Hospital Ambulatory Medical Care Survey for years 1997-2000 and 2003-2005. Using linear regression (outcome, log-transformed EDWT) and logistic regression (outcome, EDWT > 10 minutes, based on National Institute of Neurological Disorders and Stroke guidelines), we adjusted associations between EDWT and race/ethnicity (non-Hispanic whites [designated whites herein], non-Hispanic blacks [blacks], and Hispanics) for age, sex, region, mode of transportation, insurance, hospital characteristics, triage status, hospital admission, stroke type, and survey year. Compared with whites, blacks had a longer EDWT in univariate analysis (67% longer, P = .03) and multivariate analysis (62% longer, P = .03), but Hispanics had a similar EDWT in both univariate analysis (31% longer, P = .65) and multivariate analysis (5% longer, P = .91). Longer EDWT was also seen with nonambulance mode of arrival, urban hospitals, or nonemergency triage. Race was significantly associated with EDWT > 10 minutes (whites, 55% [referent]; blacks, 70% [P = .03]; Hispanics, 62% [P = .53]). These differences persisted after adjustment (blacks: odds ratio [OR] = 2.08, 95% confidence interval [CI] = 1.05-4.09; Hispanics: OR = 1.07, 95% CI = 0.52-2.22). Blacks, but not Hispanics, had significantly longer EDWT than whites. The longer EDWT in black stroke patients may lead to treatment

  13. HLA Mismatching Strategies for Solid Organ Transplantation – A Balancing Act

    PubMed Central

    Zachary, Andrea A.; Leffell, Mary S.

    2016-01-01

    HLA matching provides numerous benefits in organ transplantation including better graft function, fewer rejection episodes, longer graft survival, and the possibility of reduced immunosuppression. Mismatches are attended by more frequent rejection episodes that require increased immunosuppression that, in turn, can increase the risk of infection and malignancy. HLA mismatches also incur the risk of sensitization, which can reduce the opportunity and increase waiting time for a subsequent transplant. However, other factors such as donor age, donor type, and immunosuppression protocol, can affect the benefit derived from matching. Furthermore, finding a well-matched donor may not be possible for all patients and usually prolongs waiting time. Strategies to optimize transplantation for patients without a well-matched donor should take into account the immunologic barrier represented by different mismatches: what are the least immunogenic mismatches considering the patient’s HLA phenotype; should repeated mismatches be avoided; is the patient sensitized to HLA and, if so, what are the strengths of the patient’s antibodies? This information can then be used to define the HLA type of an immunologically optimal donor and the probability of such a donor occurring. A probability that is considered to be too low may require expanding the donor population through paired donation or modifying what is acceptable, which may require employing treatment to overcome immunologic barriers such as increased immunosuppression or desensitization. Thus, transplantation must strike a balance between the risk associated with waiting for the optimal donor and the risk associated with a less than optimal donor. PMID:28003816

  14. 7. LOWER STATION, FIRST FLOOR, WAITING ROOM, LOOKING EAST. ...

    Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey

    7. LOWER STATION, FIRST FLOOR, WAITING ROOM, LOOKING EAST. - Monongahela Incline Plane, Connecting North side of Grandview Avenue at Wyoming Street with West Carson Street near Smithfield Street, Pittsburgh, Allegheny County, PA

  15. Sex and dark times' strategy: The Dark Triad and time perspective.

    PubMed

    Moraga, Fernando R G; Nima, Ali A; Garcia, Danilo

    2017-03-01

    We investigated the effect of sex on associations between dark traits and time perspective dimensions. Responses by participants (N = 338) to the Short Dark Triad Inventory and the Zimbardo Time Perspective Inventory showed that while sex was involved in different time perspective associations of Machiavellianism, psychopathy, and narcissism, it did not moderate the dark times' strategy. © 2017 The Institute of Psychology, Chinese Academy of Sciences and John Wiley & Sons Australia, Ltd.

  16. Effectiveness and cost-effectiveness of web-based treatment for phobic outpatients on a waiting list for psychotherapy: protocol of a randomised controlled trial.

    PubMed

    Kok, Robin N; van Straten, Annemieke; Beekman, Aartjan; Bosmans, Judith; de Neef, Manja; Cuijpers, Pim

    2012-08-31

    Phobic disorders are highly prevalent and constitute a considerable burden for patients and society. As patients wait for face-to-face psychotherapy for phobic disorders in outpatient clinics, this time can be used for guided self-help interventions. The aim of this study is to investigate a five week internet-based guided self-help programme of exposure therapy in terms of clinical effectiveness and impact on speed of recovery in psychiatric outpatients, as well as the cost-effectiveness of this pre-treatment waiting list intervention. A randomised controlled trial will be conducted among 244 Dutch adult patients recruited from waiting lists of outpatient clinics for face-to-face psychotherapy for phobic disorders. Patients suffering from at least one DSM-IV classified phobic disorder (social phobia, agoraphobia or specific phobia) are randomly allocated (at a 1:1 ratio) to either a five-week internet-based guided self-help program followed by face-to-face psychotherapy, or a control group followed by face-to-face psychotherapy. Waiting list status and duration are unchanged and actual need for further treatment is evaluated prior to face-to-face psychotherapy. Clinical and economic self-assessment measurements take place at baseline, post-test (five weeks after baseline) and at 3, 6, 9 and 12 months after baseline. Offering pre-treatment internet-based guided self-help efficiently uses time otherwise lost on a waiting list and may increase patient satisfaction. Patients are expected to need fewer face-to-face sessions, reducing total treatment cost and increasing speed of recovery. Internet-delivered treatment for phobias may be a valuable addition to psychotherapy as demand for outpatient treatment increases while budgets decrease. Netherlands Trial Register NTR2233.

  17. Tonsillectomy Versus Watchful Waiting for Recurrent Throat Infection: A Systematic Review.

    PubMed

    Morad, Anna; Sathe, Nila A; Francis, David O; McPheeters, Melissa L; Chinnadurai, Sivakumar

    2017-02-01

    The effectiveness of tonsillectomy or adenotonsillectomy ("tonsillectomy") for recurrent throat infection compared with watchful waiting is uncertain. To compare sleep, cognitive, behavioral, and health outcomes of tonsillectomy versus watchful waiting in children with recurrent throat infections. MEDLINE, Embase, and the Cochrane Library. Two investigators independently screened studies against predetermined criteria. One investigator extracted data with review by a second. Investigators independently assessed risk of bias and strength of evidence (SOE) and confidence in the estimate of effects. Seven studies including children with ≥3 infections in the previous 1 to 3 years addressed this question. In studies reporting baseline data, number of infections/sore throats decreased from baseline in both groups, with greater decreases in sore throat days, clinician contacts, diagnosed group A streptococcal infections, and school absences in tonsillectomized children in the short term (<12 months). Quality of life was not markedly different between groups at any time point. Few studies fully categorized infection/sore throat severity; attrition was high. Throat infections, utilization, and school absences improved in the first postsurgical year in tonsillectomized children versus children not receiving surgery. Benefits did not persist over time; longer-term outcomes are limited. SOE is moderate for reduction in short-term throat infections and insufficient for longer-term reduction. SOE is low for no difference in longer-term streptococcal infection reduction. SOE is low for utilization and missed school reduction in the short term, low for no difference in longer-term missed school, and low for no differences in quality of life. Copyright © 2017 by the American Academy of Pediatrics.

  18. PSO-MISMO modeling strategy for multistep-ahead time series prediction.

    PubMed

    Bao, Yukun; Xiong, Tao; Hu, Zhongyi

    2014-05-01

    Multistep-ahead time series prediction is one of the most challenging research topics in the field of time series modeling and prediction, and is continually under research. Recently, the multiple-input several multiple-outputs (MISMO) modeling strategy has been proposed as a promising alternative for multistep-ahead time series prediction, exhibiting advantages compared with the two currently dominating strategies, the iterated and the direct strategies. Built on the established MISMO strategy, this paper proposes a particle swarm optimization (PSO)-based MISMO modeling strategy, which is capable of determining the number of sub-models in a self-adaptive mode, with varying prediction horizons. Rather than deriving crisp divides with equal-size s prediction horizons from the established MISMO, the proposed PSO-MISMO strategy, implemented with neural networks, employs a heuristic to create flexible divides with varying sizes of prediction horizons and to generate corresponding sub-models, providing considerable flexibility in model construction, which has been validated with simulated and real datasets.

  19. The Waite Campus: Industry, Research and Educational Collaboration.

    ERIC Educational Resources Information Center

    PEB Exchange, 1997

    1997-01-01

    The Waite Campus at the University of Adelaide, South Australia, houses industrial, research, and educational organizations. One advantage of this co-location is sharing the cost of facilities and equipment. The facilities described include Plant Research Center, Wine Science Laboratory, refectory, library, conference facilities, teleteaching,…

  20. Controlling Viscous Fingering Using Time-Dependent Strategies

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

    Stone, Howard; Zheng, Zhong; Kim, Hyoungsoo

    Control and stabilization of viscous fingering of immiscible fluids impacts a wide variety of pressure-driven multiphase flows. Here, we report theoretical and experimental results on time-dependent control strategy by manipulating the gap thickness b(t) in a lifting Hele-Shaw cell in the power-law form b(t) = b 1t 1/7. Experimental results show good quantitative agreement with the predictions of linear stability analysis. Furthermore, by choosing the value of a single time-independent control parameter we can either totally suppress the viscous fingering instability or maintain a series of non-splitting viscous fingers during the fluid displacement process. Besides the gap thickness of amore » Hele-Shaw cell, in principle, time-dependent control strategies can also be placed on the injection rate, viscosity of the displaced fluid, and interfacial tensions between the two fluids.« less