Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 16.259
Filtrar
1.
Trials ; 25(1): 527, 2024 Aug 06.
Artículo en Inglés | MEDLINE | ID: mdl-39107853

RESUMEN

BACKGROUND: Mediation analysis, often completed as secondary analysis to estimating the main treatment effect, investigates situations where an exposure may affect an outcome both directly and indirectly through intervening mediator variables. Although there has been much research on power in mediation analyses, most of this has focused on the power to detect indirect effects. Little consideration has been given to the extent to which the strength of the mediation pathways, i.e., the intervention-mediator path and the mediator-outcome path respectively, may affect the power to detect the total effect, which would correspond to the intention-to-treat effect in a randomized trial. METHODS: We conduct a simulation study to evaluate the relation between the mediation pathways and the power of testing the total treatment effect, i.e., the intention-to-treat effect. Consider a sample size that is computed based on the usual formula for testing the total effect in a two-arm trial. We generate data for a continuous mediator and a normal outcome using the conventional mediation models. We estimate the total effect using simple linear regression and evaluate the power of a two-sided test. We explore multiple data generating scenarios by varying the magnitude of the mediation paths whilst keeping the total effect constant. RESULTS: Simulations show the estimated total effect is unbiased across the considered scenarios as expected, but the mean of its standard error increases with the magnitude of the mediator-outcome path and the variability in the residual error of the mediator, respectively. Consequently, this affects the power of testing the total effect, which is always lower than planned when the mediator-outcome path is non-trivial and a naive sample size was employed. Analytical explanation confirms that the intervention-mediator path does not affect the power of testing the total effect but the mediator-outcome path. The usual effect size consideration can be adjusted to account for the magnitude of the mediator-outcome path and its residual error. CONCLUSIONS: The sample size calculation for studies with efficacy and mechanism evaluation should account for the mediator-outcome association or risk the power to detect the total effect/intention-to-treat effect being lower than planned.


Asunto(s)
Simulación por Computador , Ensayos Clínicos Controlados Aleatorios como Asunto , Proyectos de Investigación , Tamaño de la Muestra , Humanos , Ensayos Clínicos Controlados Aleatorios como Asunto/métodos , Análisis de Mediación , Análisis de Intención de Tratar , Resultado del Tratamiento , Interpretación Estadística de Datos , Modelos Lineales , Modelos Estadísticos
2.
PLoS One ; 19(8): e0301301, 2024.
Artículo en Inglés | MEDLINE | ID: mdl-39110741

RESUMEN

Interrupted time series (ITS) designs are increasingly used for estimating the effect of shocks in natural experiments. Currently, ITS designs are often used in scenarios with many time points and simple data structures. This research investigates the performance of ITS designs when the number of time points is limited and with complex data structures. Using a Monte Carlo simulation study, we empirically derive the performance-in terms of power, bias and precision- of the ITS design. Scenarios are considered with multiple interventions, a low number of time points and different effect sizes based on a motivating example of the learning loss due to COVID school closures. The results of the simulation study show the power of the step change depends mostly on the sample size, while the power of the slope change depends on the number of time points. In the basic scenario, with both a step and a slope change and an effect size of 30% of the pre-intervention slope, the required sample size for detecting a step change is 1,100 with a minimum of twelve time points. For detecting a slope change the required sample size decreases to 500 with eight time points. To decide if there is enough power researchers should inspect their data, hypothesize about effect sizes and consider an appropriate model before applying an ITS design to their research. This paper contributes to the field of methodology in two ways. Firstly, the motivation example showcases the difficulty of employing ITS designs in cases which do not adhere to a single intervention. Secondly, models are proposed for more difficult ITS designs and their performance is tested.


Asunto(s)
COVID-19 , Análisis de Series de Tiempo Interrumpido , Método de Montecarlo , Pandemias , Instituciones Académicas , COVID-19/epidemiología , COVID-19/prevención & control , Humanos , SARS-CoV-2/aislamiento & purificación , Aprendizaje , Simulación por Computador , Tamaño de la Muestra
3.
Biom J ; 66(6): e202300271, 2024 Sep.
Artículo en Inglés | MEDLINE | ID: mdl-39132909

RESUMEN

Many clinical trials assess time-to-event endpoints. To describe the difference between groups in terms of time to event, we often employ hazard ratios. However, the hazard ratio is only informative in the case of proportional hazards (PHs) over time. There exist many other effect measures that do not require PHs. One of them is the average hazard ratio (AHR). Its core idea is to utilize a time-dependent weighting function that accounts for time variation. Though propagated in methodological research papers, the AHR is rarely used in practice. To facilitate its application, we unfold approaches for sample size calculation of an AHR test. We assess the reliability of the sample size calculation by extensive simulation studies covering various survival and censoring distributions with proportional as well as nonproportional hazards (N-PHs). The findings suggest that a simulation-based sample size calculation approach can be useful for designing clinical trials with N-PHs. Using the AHR can result in increased statistical power to detect differences between groups with more efficient sample sizes.


Asunto(s)
Modelos de Riesgos Proporcionales , Tamaño de la Muestra , Humanos , Ensayos Clínicos como Asunto , Biometría/métodos
4.
Trials ; 25(1): 532, 2024 Aug 12.
Artículo en Inglés | MEDLINE | ID: mdl-39128997

RESUMEN

OBJECTIVE: To assess the cost-effectiveness of using cheaper-but-noisier outcome measures, such as a short questionnaire, for large simple clinical trials. BACKGROUND: To detect associations reliably, trials must avoid bias and random error. To reduce random error, we can increase the size of the trial and increase the accuracy of the outcome measurement process. However, with fixed resources, there is a trade-off between the number of participants a trial can enrol and the amount of information that can be collected on each participant during data collection. METHODS: To consider the effect on measurement error of using outcome scales with varying numbers of categories, we define and calculate the variance from categorisation that would be expected from using a category midpoint; define the analytic conditions under which such a measure is cost-effective; use meta-regression to estimate the impact of participant burden, defined as questionnaire length, on response rates; and develop an interactive web-app to allow researchers to explore the cost-effectiveness of using such a measure under plausible assumptions. RESULTS: An outcome scale with only a few categories greatly reduced the variance of non-measurement. For example, a scale with five categories reduced the variance of non-measurement by 96% for a uniform distribution. We show that a simple measure will be more cost-effective than a gold-standard measure if the relative increase in variance due to using it is less than the relative increase in cost from the gold standard, assuming it does not introduce bias in the measurement. We found an inverse power law relationship between participant burden and response rates such that a doubling the burden on participants reduces the response rate by around one third. Finally, we created an interactive web-app ( https://benjiwoolf.shinyapps.io/cheapbutnoisymeasures/ ) to allow exploration of when using a cheap-but-noisy measure will be more cost-effective using realistic parameters. CONCLUSION: Cheaper-but-noisier questionnaires containing just a few questions can be a cost-effective way of maximising power. However, their use requires a judgement on the trade-off between the potential increase in risk of information bias and the reduction in the potential of selection bias due to the expected higher response rates.


Asunto(s)
Ensayos Clínicos como Asunto , Análisis Costo-Beneficio , Proyectos de Investigación , Humanos , Encuestas y Cuestionarios , Proyectos de Investigación/normas , Ensayos Clínicos como Asunto/economía , Ensayos Clínicos como Asunto/normas , Reproducibilidad de los Resultados , Tamaño de la Muestra , Resultado del Tratamiento , Modelos Económicos , Determinación de Punto Final
5.
PLoS One ; 19(8): e0296207, 2024.
Artículo en Inglés | MEDLINE | ID: mdl-39088468

RESUMEN

Polygenic risk scores (PRS) are instrumental in genetics, offering insights into an individual level genetic risk to a range of diseases based on accumulated genetic variations. These scores rely on Genome-Wide Association Studies (GWAS). However, precision in PRS is often challenged by the requirement of extensive sample sizes and the potential for overlapping datasets that can inflate PRS calculations. In this study, we present a novel methodology, Meta-Reductive Approach (MRA), that was derived algebraically to adjust GWAS results, aiming to neutralize the influence of select cohorts. Our approach recalibrates summary statistics using algebraic derivations. Validating our technique with datasets from Alzheimer disease studies, we showed that the summary statistics of the MRA and those derived from individual-level data yielded the exact same values. This innovative method offers a promising avenue for enhancing the accuracy of PRS, especially when derived from meta-analyzed GWAS data.


Asunto(s)
Enfermedad de Alzheimer , Predisposición Genética a la Enfermedad , Estudio de Asociación del Genoma Completo , Herencia Multifactorial , Estudio de Asociación del Genoma Completo/métodos , Humanos , Enfermedad de Alzheimer/genética , Herencia Multifactorial/genética , Polimorfismo de Nucleótido Simple , Tamaño de la Muestra , Factores de Riesgo
6.
BMC Med Res Methodol ; 24(1): 155, 2024 Jul 19.
Artículo en Inglés | MEDLINE | ID: mdl-39030495

RESUMEN

BACKGROUND: There is increasing interest in the capacity of adaptive designs to improve the efficiency of clinical trials. However, relatively little work has investigated how economic considerations - including the costs of the trial - might inform the design and conduct of adaptive clinical trials. METHODS: We apply a recently published Bayesian model of a value-based sequential clinical trial to data from the 'Hydroxychloroquine Effectiveness in Reducing symptoms of hand Osteoarthritis' (HERO) trial. Using parameters estimated from the trial data, including the cost of running the trial, and using multiple imputation to estimate the accumulating cost-effectiveness signal in the presence of missing data, we assess when the trial would have stopped had the value-based model been used. We used re-sampling methods to compare the design's operating characteristics with those of a conventional fixed length design. RESULTS: In contrast to the findings of the only other published retrospective application of this model, the equivocal nature of the cost-effectiveness signal from the HERO trial means that the design would have stopped the trial close to, or at, its maximum planned sample size, with limited additional value delivered via savings in research expenditure. CONCLUSION: Evidence from the two retrospective applications of this design suggests that, when the cost-effectiveness signal in a clinical trial is unambiguous, the Bayesian value-adaptive design can stop the trial before it reaches its maximum sample size, potentially saving research costs when compared with the alternative fixed sample size design. However, when the cost-effectiveness signal is equivocal, the design is expected to run to, or close to, the maximum sample size and deliver limited savings in research costs.


Asunto(s)
Teorema de Bayes , Análisis Costo-Beneficio , Osteoartritis , Proyectos de Investigación , Humanos , Análisis Costo-Beneficio/métodos , Análisis Costo-Beneficio/estadística & datos numéricos , Osteoartritis/economía , Osteoartritis/tratamiento farmacológico , Osteoartritis/terapia , Hidroxicloroquina/uso terapéutico , Hidroxicloroquina/economía , Ensayos Clínicos como Asunto/métodos , Ensayos Clínicos como Asunto/economía , Ensayos Clínicos como Asunto/estadística & datos numéricos , Tamaño de la Muestra
7.
J Med Internet Res ; 26: e52998, 2024 Jul 09.
Artículo en Inglés | MEDLINE | ID: mdl-38980711

RESUMEN

BACKGROUND: In-depth interviews are a common method of qualitative data collection, providing rich data on individuals' perceptions and behaviors that would be challenging to collect with quantitative methods. Researchers typically need to decide on sample size a priori. Although studies have assessed when saturation has been achieved, there is no agreement on the minimum number of interviews needed to achieve saturation. To date, most research on saturation has been based on in-person data collection. During the COVID-19 pandemic, web-based data collection became increasingly common, as traditional in-person data collection was possible. Researchers continue to use web-based data collection methods post the COVID-19 emergency, making it important to assess whether findings around saturation differ for in-person versus web-based interviews. OBJECTIVE: We aimed to identify the number of web-based interviews needed to achieve true code saturation or near code saturation. METHODS: The analyses for this study were based on data from 5 Food and Drug Administration-funded studies conducted through web-based platforms with patients with underlying medical conditions or with health care providers who provide primary or specialty care to patients. We extracted code- and interview-specific data and examined the data summaries to determine when true saturation or near saturation was reached. RESULTS: The sample size used in the 5 studies ranged from 30 to 70 interviews. True saturation was reached after 91% to 100% (n=30-67) of planned interviews, whereas near saturation was reached after 33% to 60% (n=15-23) of planned interviews. Studies that relied heavily on deductive coding and studies that had a more structured interview guide reached both true saturation and near saturation sooner. We also examined the types of codes applied after near saturation had been reached. In 4 of the 5 studies, most of these codes represented previously established core concepts or themes. Codes representing newly identified concepts, other or miscellaneous responses (eg, "in general"), uncertainty or confusion (eg, "don't know"), or categorization for analysis (eg, correct as compared with incorrect) were less commonly applied after near saturation had been reached. CONCLUSIONS: This study provides support that near saturation may be a sufficient measure to target and that conducting additional interviews after that point may result in diminishing returns. Factors to consider in determining how many interviews to conduct include the structure and type of questions included in the interview guide, the coding structure, and the population under study. Studies with less structured interview guides, studies that rely heavily on inductive coding and analytic techniques, and studies that include populations that may be less knowledgeable about the topics discussed may require a larger sample size to reach an acceptable level of saturation. Our findings also build on previous studies looking at saturation for in-person data collection conducted at a small number of sites.


Asunto(s)
COVID-19 , Entrevistas como Asunto , Humanos , Tamaño de la Muestra , Entrevistas como Asunto/métodos , Investigación Cualitativa , SARS-CoV-2 , Pandemias , Recolección de Datos/métodos , Internet
8.
Pharmacoepidemiol Drug Saf ; 33(7): e5864, 2024 Jul.
Artículo en Inglés | MEDLINE | ID: mdl-39013838

RESUMEN

PURPOSE: To compare the performance (covariate balance, effective sample size [ESS]) of stable balancing weights (SBW) versus propensity score weighting (PSW). Two applied cases were used to compare performance: (Case 1) extreme imbalance in baseline covariates between groups and (Case 2) substantial discrepancy in sample size between groups. METHODS: Using the Premier Healthcare Database, we selected patients who (Case 1) underwent a surgical procedure with one of two different bipolar forceps between January 2000 and June 2020, or (Case 2) a neurological procedure using one of two different nonabsorbable surgical sutures between January 2000 and March 2020. Average treatment effects on the treated (ATT) weights were generated based on selected covariates. SBW was implemented using two techniques: (1) "grid search" to find weights of minimum variance at the lowest target absolute standardized mean difference (SMD); (2) finding weights of minimum variance at prespecified SMD tolerance. PSW and SBW methods were compared on postweighting SMDs, the number of imbalanced covariates, and ESS of the ATT-weighted control group. RESULTS: In both studies, improved covariate balance was achieved with both SBW techniques. All methods suffered from postweighting ESS that was lower than the unweighted control group's original sample size; however, SBW methods achieved higher ESS for the control groups. Sensitivity analyses using SBW to apply variable-specific SMD thresholds increased ESS, outperforming PSW. CONCLUSIONS: In this applied example, the optimization-based SBW method provided ample flexibility with respect to prespecification of covariate balance goals and resulted in better postweighting covariate balance and larger ESS as compared with PSW.


Asunto(s)
Puntaje de Propensión , Humanos , Tamaño de la Muestra , Bases de Datos Factuales , Femenino , Masculino , Persona de Mediana Edad
9.
Pharm Stat ; 23(4): 557-569, 2024.
Artículo en Inglés | MEDLINE | ID: mdl-38992978

RESUMEN

Biomarkers are key components of personalized medicine. In this paper, we consider biomarkers taking continuous values that are associated with disease status, called case and control. The performance of such a biomarker is evaluated by the area under the curve (AUC) of its receiver operating characteristic curve. Oftentimes, two biomarkers are collected from each subject to test if one has a larger AUC than the other. We propose a simple non-parametric statistical test for comparing the performance of two biomarkers. We also present a simple sample size calculation method for this test statistic. Our sample size formula requires specification of AUC values (or the standardized effect size of each biomarker between cases and controls together with the correlation coefficient between two biomarkers), prevalence of cases in the study population, type I error rate, and power. Through simulations, we show that the testing on two biomarkers controls type I error rate accurately and the proposed sample size closely maintains specified statistical power.


Asunto(s)
Área Bajo la Curva , Biomarcadores , Simulación por Computador , Curva ROC , Humanos , Tamaño de la Muestra , Biomarcadores/análisis , Estudios de Casos y Controles , Medicina de Precisión/métodos , Medicina de Precisión/estadística & datos numéricos , Modelos Estadísticos , Proyectos de Investigación/estadística & datos numéricos , Interpretación Estadística de Datos
10.
Nan Fang Yi Ke Da Xue Xue Bao ; 44(6): 1182-1187, 2024 Jun 20.
Artículo en Chino | MEDLINE | ID: mdl-38977349

RESUMEN

OBJECTIVE: To explore the applicable conditions of the Cox-TEL (Cox PH-Taylor expansion adjustment for long-term survival data) method for analysis of survival data that contain cured patients. METHODS: The simulated survival data method based on Weibull distribution was used to simulate and generate the survival data with different cure rates, censored rates, and cure rate differences. The Cox-TEL method was used for analysis of the generated simulation data, and its performance was evaluated by calculating its type Ⅰ error and power. RESULTS: Almost all the type Ⅰ error of the hazard ratios (HRs) obtained by the Cox-TEL method under different conditions were slightly greater than 0.05, and this method showed a good test power for estimating the HRs for data with a large sample size and a large difference in proportions (DPs). For the data of cured patients, the type Ⅰ error of the DPs obtained by the Cox-TEL method was well around 0.05, and its test power was robust in most of the scenarios. CONCLUSION: The Cox-TEL method is effective for analyzing data of uncured patients and obtaining reliable HRs for most of the survival data with a sample size, a low censored rates, and a large difference in cure rates. The method is capable of accurately estimating the DPs regardless of the sample size, censored rates, or the cure rates.


Asunto(s)
Simulación por Computador , Modelos de Riesgos Proporcionales , Humanos , Reproducibilidad de los Resultados , Análisis de Supervivencia , Tamaño de la Muestra
11.
BMC Med Res Methodol ; 24(1): 146, 2024 Jul 10.
Artículo en Inglés | MEDLINE | ID: mdl-38987715

RESUMEN

BACKGROUND: Risk prediction models are routinely used to assist in clinical decision making. A small sample size for model development can compromise model performance when the model is applied to new patients. For binary outcomes, the calibration slope (CS) and the mean absolute prediction error (MAPE) are two key measures on which sample size calculations for the development of risk models have been based. CS quantifies the degree of model overfitting while MAPE assesses the accuracy of individual predictions. METHODS: Recently, two formulae were proposed to calculate the sample size required, given anticipated features of the development data such as the outcome prevalence and c-statistic, to ensure that the expectation of the CS and MAPE (over repeated samples) in models fitted using MLE will meet prespecified target values. In this article, we use a simulation study to evaluate the performance of these formulae. RESULTS: We found that both formulae work reasonably well when the anticipated model strength is not too high (c-statistic < 0.8), regardless of the outcome prevalence. However, for higher model strengths the CS formula underestimates the sample size substantially. For example, for c-statistic = 0.85 and 0.9, the sample size needed to be increased by at least 50% and 100%, respectively, to meet the target expected CS. On the other hand, the MAPE formula tends to overestimate the sample size for high model strengths. These conclusions were more pronounced for higher prevalence than for lower prevalence. Similar results were drawn when the outcome was time to event with censoring. Given these findings, we propose a simulation-based approach, implemented in the new R package 'samplesizedev', to correctly estimate the sample size even for high model strengths. The software can also calculate the variability in CS and MAPE, thus allowing for assessment of model stability. CONCLUSIONS: The calibration and MAPE formulae suggest sample sizes that are generally appropriate for use when the model strength is not too high. However, they tend to be biased for higher model strengths, which are not uncommon in clinical risk prediction studies. On those occasions, our proposed adjustments to the sample size calculations will be relevant.


Asunto(s)
Modelos Estadísticos , Humanos , Tamaño de la Muestra , Medición de Riesgo/métodos , Medición de Riesgo/estadística & datos numéricos , Simulación por Computador , Algoritmos
12.
Trials ; 25(1): 458, 2024 Jul 05.
Artículo en Inglés | MEDLINE | ID: mdl-38970042

RESUMEN

Despite progress in reducing the infant mortality in India, the neonatal mortality decline has been slower, necessitating concerted efforts to achieve Sustainable Development Goal-3. A promising strategy aiming to prevent neonatal sepsis in high-risk, vulnerable, low birth weight neonates through an innovative intervention includes probiotic supplementation. This article communicates the decision by the ProSPoNS trial investigators to establish a Central Endpoint Adjudication Committee (CEAC) as an addendum to the protocol published in Trials in 2021 for the purpose of clarifying the primary outcome. In the published protocol, the study hypothesis and primary objective are based on "sepsis," the primary outcome has been specified as sepsis/PSBI, whereas the sample size estimation was performed based on the "physician diagnosed sepsis." To align all the three above, the investigators meeting, held on 17th-18th August 2023, at MGIMS Sevagram, Wardha, deliberated and unanimously agreed that "physician diagnosed sepsis" is the primary study outcome which includes sepsis/PSBI. The CEAC, chaired by an external subject expert and members including trial statistician, a microbiologist, and all site principal investigators will employ four criteria to determine "physician diagnosed sepsis": (1) blood culture status, (2) sepsis screen status, (3) PSBI/non-PSBI signs and symptoms, and (4) the clinical course for each sickness event. Importantly, this clarification maintains consistency with the approved study protocol (Protocol No. 5/7/915/2012 version 3.1 dated 14 Feb 2020), emphasizing the commitment to methodological transparency and adherence to predefined standards. The decision to utilize the guidance of a CEAC is recommended as the gold standard in multicentric complex clinical trials to achieve consistency and accuracy in assessment of outcomes.Trial registrationClinical Trial Registry of India (CTRI) CTRI/2019/05/019197. Registered on 16 May 2019.


Asunto(s)
Sepsis Neonatal , Humanos , Recién Nacido , Sepsis Neonatal/diagnóstico , Sepsis Neonatal/tratamiento farmacológico , Ensayos Clínicos Controlados Aleatorios como Asunto , Determinación de Punto Final/normas , India , Probióticos/uso terapéutico , Probióticos/efectos adversos , Resultado del Tratamiento , Mortalidad Infantil , Proyectos de Investigación , Tamaño de la Muestra
13.
Hum Brain Mapp ; 45(10): e26768, 2024 Jul 15.
Artículo en Inglés | MEDLINE | ID: mdl-38949537

RESUMEN

Structural neuroimaging data have been used to compute an estimate of the biological age of the brain (brain-age) which has been associated with other biologically and behaviorally meaningful measures of brain development and aging. The ongoing research interest in brain-age has highlighted the need for robust and publicly available brain-age models pre-trained on data from large samples of healthy individuals. To address this need we have previously released a developmental brain-age model. Here we expand this work to develop, empirically validate, and disseminate a pre-trained brain-age model to cover most of the human lifespan. To achieve this, we selected the best-performing model after systematically examining the impact of seven site harmonization strategies, age range, and sample size on brain-age prediction in a discovery sample of brain morphometric measures from 35,683 healthy individuals (age range: 5-90 years; 53.59% female). The pre-trained models were tested for cross-dataset generalizability in an independent sample comprising 2101 healthy individuals (age range: 8-80 years; 55.35% female) and for longitudinal consistency in a further sample comprising 377 healthy individuals (age range: 9-25 years; 49.87% female). This empirical examination yielded the following findings: (1) the accuracy of age prediction from morphometry data was higher when no site harmonization was applied; (2) dividing the discovery sample into two age-bins (5-40 and 40-90 years) provided a better balance between model accuracy and explained age variance than other alternatives; (3) model accuracy for brain-age prediction plateaued at a sample size exceeding 1600 participants. These findings have been incorporated into CentileBrain (https://centilebrain.org/#/brainAGE2), an open-science, web-based platform for individualized neuroimaging metrics.


Asunto(s)
Envejecimiento , Encéfalo , Imagen por Resonancia Magnética , Humanos , Adolescente , Femenino , Anciano , Adulto , Niño , Adulto Joven , Masculino , Encéfalo/diagnóstico por imagen , Encéfalo/anatomía & histología , Encéfalo/crecimiento & desarrollo , Anciano de 80 o más Años , Preescolar , Persona de Mediana Edad , Envejecimiento/fisiología , Imagen por Resonancia Magnética/métodos , Neuroimagen/métodos , Neuroimagen/normas , Tamaño de la Muestra
14.
Contemp Clin Trials ; 144: 107620, 2024 Sep.
Artículo en Inglés | MEDLINE | ID: mdl-38977178

RESUMEN

We propose a Cross-validated ADaptive ENrichment design (CADEN) in which a trial population is enriched with a subpopulation of patients who are predicted to benefit from the treatment more than an average patient (the sensitive group). This subpopulation is found using a risk score constructed from the baseline (potentially high-dimensional) information about patients. The design incorporates an early stopping rule for futility. Simulation studies are used to assess the properties of CADEN against the original (non-enrichment) cross-validated risk scores (CVRS) design which constructs a risk score at the end of the trial. We show that when there exists a sensitive group of patients, CADEN achieves a higher power and a reduction in the expected sample size compared to the CVRS design. We illustrate the application of the design in two real clinical trials. We conclude that the new design offers improved statistical efficiency over the existing non-enrichment method, as well as increased benefit to patients. The method has been implemented in an R package caden.


Asunto(s)
Proyectos de Investigación , Humanos , Tamaño de la Muestra , Medición de Riesgo/métodos , Simulación por Computador , Selección de Paciente
15.
BMC Med Res Methodol ; 24(1): 151, 2024 Jul 16.
Artículo en Inglés | MEDLINE | ID: mdl-39014324

RESUMEN

The test-negative design (TND) is an observational study design to evaluate vaccine effectiveness (VE) that enrolls individuals receiving diagnostic testing for a target disease as part of routine care. VE is estimated as one minus the adjusted odds ratio of testing positive versus negative comparing vaccinated and unvaccinated patients. Although the TND is related to case-control studies, it is distinct in that the ratio of test-positive cases to test-negative controls is not typically pre-specified. For both types of studies, sparse cells are common when vaccines are highly effective. We consider the implications of these features on power for the TND. We use simulation studies to explore three hypothesis-testing procedures and associated sample size calculations for case-control and TND studies. These tests, all based on a simple logistic regression model, are a standard Wald test, a continuity-corrected Wald test, and a score test. The Wald test performs poorly in both case-control and TND when VE is high because the number of vaccinated test-positive cases can be low or zero. Continuity corrections help to stabilize the variance but induce bias. We observe superior performance with the score test as the variance is pooled under the null hypothesis of no group differences. We recommend using a score-based approach to design and analyze both case-control and TND. We propose a modification to the TND score sample size to account for additional variability in the ratio of controls over cases. This work enhances our understanding of the data generating mechanism in a test-negative design (TND) and how it is distinct from that of a case-control study due to its passive recruitment of controls.


Asunto(s)
Proyectos de Investigación , Humanos , Tamaño de la Muestra , Estudios de Casos y Controles , Eficacia de las Vacunas/estadística & datos numéricos , Modelos Logísticos , Simulación por Computador , Oportunidad Relativa , Vacunación/estadística & datos numéricos , Estudios Observacionales como Asunto/métodos , Estudios Observacionales como Asunto/estadística & datos numéricos
16.
PLoS One ; 19(7): e0305654, 2024.
Artículo en Inglés | MEDLINE | ID: mdl-39024199

RESUMEN

Even though deep learning shows impressive results in several applications, its use on problems with High Dimensions and Low Sample Size, such as diagnosing rare diseases, leads to overfitting. One solution often proposed is feature selection. In deep learning, along with feature selection, network sparsification is also used to improve the results when dealing with high dimensions low sample size data. However, most of the time, they are tackled as separate problems. This paper proposes a new approach that integrates feature selection, based on sparsification, into the training process of a deep neural network. This approach uses a constrained biobjective gradient descent method. It provides a set of Pareto optimal neural networks that make a trade-off between network sparsity and model accuracy. Results on both artificial and real datasets show that using a constrained biobjective gradient descent increases the network sparsity without degrading the classification performances. With the proposed approach, on an artificial dataset, the feature selection score reached 0.97 with a sparsity score of 0.92 with an accuracy of 0.9. For the same accuracy, none of the other methods reached a feature score above 0.20 and sparsity score of 0.35. Finally, statistical tests validate the results obtained on all datasets.


Asunto(s)
Aprendizaje Profundo , Redes Neurales de la Computación , Tamaño de la Muestra , Humanos , Algoritmos
17.
Invest Ophthalmol Vis Sci ; 65(8): 7, 2024 Jul 01.
Artículo en Inglés | MEDLINE | ID: mdl-38958969

RESUMEN

Purpose: To describe and demonstrate sample size and power calculation for ophthalmic studies with a binary outcome from one or both eyes. Methods: We describe sample size and power calculation for four commonly used eye designs: (1) one-eye design or person-design: one eye per subject or outcome is at person-level; (2) paired design: two eyes per subject and two eyes are in different treatment groups; (3) two-eye design: two eyes per subject and both eyes are in the same treatment group; and (4) mixture design: mixture of one eye and two eyes per subject. For each design, we demonstrate sample size and power calculations in real ophthalmic studies. Results: Using formulas and commercial or free statistical packages including SAS, STATA, R, and PS, we calculated sample size and power. We demonstrated that different statistical packages require different parameters and provide similar, yet not identical, results. We emphasize that studies using data from two eyes of a subject need to account for the intereye correlation for appropriate sample size and power calculations. We demonstrate the gain in efficiency in designs that include two eyes of a subject compared to one-eye designs. Conclusions: Ophthalmic studies use different eye designs that include one or both eyes in the same or different treatment groups. Appropriate sample size and power calculations depend on the eye design and should account for intereye correlation when two eyes from some or all subjects are included in a study. Calculations can be executed using formulas and commercial or free statistical packages.


Asunto(s)
Bioestadística , Oftalmología , Humanos , Tamaño de la Muestra , Bioestadística/métodos , Proyectos de Investigación , Oftalmopatías/diagnóstico
18.
J Assoc Physicians India ; 72(7): 34-40, 2024 Jul.
Artículo en Inglés | MEDLINE | ID: mdl-38990585

RESUMEN

AIM: This study illustrates parameters, procedures, and calculations for the statistical determination of sample size for different clinical study designs. MATERIALS AND METHODS: In any research process, the sample size is an important consideration for the implementation of the planned study. From time to time, literature on sample size has been documented in the medical literature. However, the situations covered under them lack comprehensiveness in terms of different study designs, demonstration of calculations, and overreliance on statistical software. RESULTS: The present study provides various facets of sample size determination, such as prerequisite parameters, mathematical formulation, and calculations for clinical study designs [descriptive studies, randomized controlled trials (RCT), correlational studies, comparison of multiple outcomes, survival analysis, sensitivity, and specificity], which will be quite useful. CONCLUSION: This communication will be a good education and learning source for medical professionals to pick and choose a specific scenario and estimate the sample size.


Asunto(s)
Proyectos de Investigación , Tamaño de la Muestra , Humanos , Ensayos Clínicos Controlados Aleatorios como Asunto , Estudios Clínicos como Asunto
19.
PLoS One ; 19(7): e0306334, 2024.
Artículo en Inglés | MEDLINE | ID: mdl-38959247

RESUMEN

OBJECTIVE: While statistical analysis plays a crucial role in medical science, some published studies might have utilized suboptimal analysis methods, potentially undermining the credibility of their findings. Critically appraising analytical approaches can help elevate the standard of evidence and ensure clinicians and other stakeholders have trustworthy results on which to base decisions. The aim of the present study was to examine the statistical characteristics of original articles published in Peruvian medical journals in 2021-2022. DESIGN AND SETTING: We performed a methodological study of articles published between 2021 and 2022 from nine medical journals indexed in SciELO-Peru, Scopus, and Medline. We included original articles that conducted analytical analyses (i.e., association between variables). The statistical variables assessed were: statistical software used for analysis, sample size, and statistical methods employed (measures of effect), controlling for confounders, and the method employed for confounder control or epidemiological approaches. RESULTS: We included 313 articles (ranging from 11 to 77 across journals), of which 67.7% were cross-sectional studies. While 90.7% of articles specified the statistical software used, 78.3% omitted details on sample size calculation. Descriptive and bivariate statistics were commonly employed, whereas measures of association were less common. Only 13.4% of articles (ranging from 0% to 39% across journals) presented measures of effect controlling for confounding and explained the criteria for selecting such confounders. CONCLUSION: This study revealed important statistical deficiencies within analytical studies published in Peruvian journals, including inadequate reporting of sample sizes, absence of measures of association and confounding control, and suboptimal explanations regarding the methodologies employed for adjusted analyses. These findings highlight the need for better statistical reporting and researcher-editor collaboration to improve the quality of research production and dissemination in Peruvian journals.


Asunto(s)
Publicaciones Periódicas como Asunto , Perú , Publicaciones Periódicas como Asunto/estadística & datos numéricos , Humanos , Tamaño de la Muestra , Edición/estadística & datos numéricos , Proyectos de Investigación
20.
AAPS J ; 26(4): 77, 2024 Jul 03.
Artículo en Inglés | MEDLINE | ID: mdl-38960976

RESUMEN

Dose-scale pharmacodynamic bioequivalence is recommended for evaluating the consistency of generic and innovator formulations of certain locally acting drugs, such as orlistat. This study aimed to investigate the standard methodology for sample size determination and the impact of study design on dose-scale pharmacodynamic bioequivalence using orlistat as the model drug. A population pharmacodynamic model of orlistat was developed using NONMEM 7.5.1 and utilized for subsequent simulations. Three different study designs were evaluated across various predefined relative bioavailability ratios of test/reference (T/R) formulations. These designs included Study Design 1 (2×1 crossover with T1 60 mg, R1 60 mg, and R2 120 mg), Study Design 2 (2×1 crossover with T2 120 mg, R1 60 mg, and R2 120 mg), and Study Design 3 (2×2 crossover with T1 60 mg, T2 120 mg, R1 60 mg, and R2 120 mg). Sample sizes were determined using a stochastic simulation and estimation approach. Under the same T/R ratio and power, Study Design 3 required the minimum sample size for bioequivalence, followed by Study Design 1, while Study Design 2 performed the worst. For Study Designs 1 and 3, a larger sample size was needed on the T/R ratio < 1.0 side for the same power compared to that on the T/R ratio > 1.0 side. The opposite asymmetry was observed for Study Design 2. We demonstrated that Study Design 3 is most effective for reducing the sample size for orlistat bioequivalence studies, and the impact of T/R ratio on sample size shows asymmetry.


Asunto(s)
Estudios Cruzados , Orlistat , Equivalencia Terapéutica , Orlistat/farmacocinética , Orlistat/administración & dosificación , Humanos , Tamaño de la Muestra , Proyectos de Investigación , Disponibilidad Biológica , Modelos Biológicos , Fármacos Antiobesidad/farmacocinética , Fármacos Antiobesidad/administración & dosificación , Lactonas/farmacocinética , Lactonas/administración & dosificación , Simulación por Computador , Relación Dosis-Respuesta a Droga
SELECCIÓN DE REFERENCIAS
DETALLE DE LA BÚSQUEDA