Response and nonresponse

Filter results by

Search Help
Currently selected filters that can be removed

Keyword(s)

Geography

1 facets displayed. 0 facets selected.

Content

1 facets displayed. 0 facets selected.
Sort Help
entries

Results

All (136)

All (136) (20 to 30 of 136 results)

  • Articles and reports: 12-001-X201600214677
    Description:

    How do we tell whether weighting adjustments reduce nonresponse bias? If a variable is measured for everyone in the selected sample, then the design weights can be used to calculate an approximately unbiased estimate of the population mean or total for that variable. A second estimate of the population mean or total can be calculated using the survey respondents only, with weights that have been adjusted for nonresponse. If the two estimates disagree, then there is evidence that the weight adjustments may not have removed the nonresponse bias for that variable. In this paper we develop the theoretical properties of linearization and jackknife variance estimators for evaluating the bias of an estimated population mean or total by comparing estimates calculated from overlapping subsets of the same data with different sets of weights, when poststratification or inverse propensity weighting is used for the nonresponse adjustments to the weights. We provide sufficient conditions on the population, sample, and response mechanism for the variance estimators to be consistent, and demonstrate their small-sample properties through a simulation study.

    Release date: 2016-12-20

  • Articles and reports: 12-001-X201500114172
    Description:

    When a random sample drawn from a complete list frame suffers from unit nonresponse, calibration weighting to population totals can be used to remove nonresponse bias under either an assumed response (selection) or an assumed prediction (outcome) model. Calibration weighting in this way can not only provide double protection against nonresponse bias, it can also decrease variance. By employing a simple trick one can estimate the variance under the assumed prediction model and the mean squared error under the combination of an assumed response model and the probability-sampling mechanism simultaneously. Unfortunately, there is a practical limitation on what response model can be assumed when design weights are calibrated to population totals in a single step. In particular, the choice for the response function cannot always be logistic. That limitation does not hinder calibration weighting when performed in two steps: from the respondent sample to the full sample to remove the response bias and then from the full sample to the population to decrease variance. There are potential efficiency advantages from using the two-step approach as well even when the calibration variables employed in each step is a subset of the calibration variables in the single step. Simultaneous mean-squared-error estimation using linearization is possible, but more complicated than when calibrating in a single step.

    Release date: 2015-06-29

  • Articles and reports: 12-001-X201500114173
    Description:

    Nonresponse is present in almost all surveys and can severely bias estimates. It is usually distinguished between unit and item nonresponse. By noting that for a particular survey variable, we just have observed and unobserved values, in this work we exploit the connection between unit and item nonresponse. In particular, we assume that the factors that drive unit response are the same as those that drive item response on selected variables of interest. Response probabilities are then estimated using a latent covariate that measures the will to respond to the survey and that can explain a part of the unknown behavior of a unit to participate in the survey. This latent covariate is estimated using latent trait models. This approach is particularly relevant for sensitive items and, therefore, can handle non-ignorable nonresponse. Auxiliary information known for both respondents and nonrespondents can be included either in the latent variable model or in the response probability estimation process. The approach can also be used when auxiliary information is not available, and we focus here on this case. We propose an estimator using a reweighting system based on the previous latent covariate when no other observed auxiliary information is available. Results on its performance are encouraging from simulation studies on both real and simulated data.

    Release date: 2015-06-29

  • Articles and reports: 11-522-X201300014262
    Description:

    Measurement error is one source of bias in statistical analysis. However, its possible implications are mostly ignored One class of models that can be especially affected by measurement error are fixed-effects models. By validating the survey response of five panel survey waves for welfare receipt with register data, the size and form of longitudinal measurement error can be determined. It is shown, that the measurement error for welfare receipt is serially correlated and non-differential. However, when estimating the coefficients of longitudinal fixed effect models of welfare receipt on subjective health for men and women, the coefficients are biased only for the male subpopulation.

    Release date: 2014-10-31

  • Articles and reports: 11-522-X201300014263
    Description:

    Collecting information from sampled units over the Internet or by mail is much more cost-efficient than conducting interviews. These methods make self-enumeration an attractive data-collection method for surveys and censuses. Despite the benefits associated with self-enumeration data collection, in particular Internet-based data collection, self-enumeration can produce low response rates compared with interviews. To increase response rates, nonrespondents are subject to a mixed mode of follow-up treatments, which influence the resulting probability of response, to encourage them to participate. Factors and interactions are commonly used in regression analyses, and have important implications for the interpretation of statistical models. Because response occurrence is intrinsically conditional, we first record response occurrence in discrete intervals, and we characterize the probability of response by a discrete time hazard. This approach facilitates examining when a response is most likely to occur and how the probability of responding varies over time. The nonresponse bias can be avoided by multiplying the sampling weight of respondents by the inverse of an estimate of the response probability. Estimators for model parameters as well as for finite population parameters are given. Simulation results on the performance of the proposed estimators are also presented.

    Release date: 2014-10-31

  • Articles and reports: 11-522-X201300014277
    Description:

    This article gives an overview of adaptive design elements introduced to the PASS panel survey in waves four to seven. The main focus is on experimental interventions in later phases of the fieldwork. These interventions aim at balancing the sample by prioritizing low-propensity sample members. In wave 7, interviewers received a double premium for completion of interviews with low-propensity cases in the final phase of the fieldwork. This premium was restricted to a random half of the cases with low estimated response propensity and no final status after four months of prior fieldwork. This incentive was effective in increasing interviewer effort, however, led to no significant increase in response rates.

    Release date: 2014-10-31

  • Articles and reports: 12-001-X201300211885
    Description:

    Web surveys are generally connected with low response rates. Common suggestions in textbooks on Web survey research highlight the importance of the welcome screen in encouraging respondents to take part. The importance of this screen has been empirically proven in research, showing that most respondents breakoff at the welcome screen. However, there has been little research on the effect of the design of this screen on the level of the breakoff rate. In a study conducted at the University of Konstanz, three experimental treatments were added to a survey of the first-year student population (2,629 students) to assess the impact of different design features of this screen on the breakoff rates. The methodological experiments included varying the background color of the welcome screen, varying the promised task duration on this first screen, and varying the length of the information provided on the welcome screen explaining the privacy rights of the respondents. The analyses show that the longer stated length and the more attention given to explaining privacy rights on the welcome screen, the fewer respondents started and completed the survey. However, the use of a different background color does not result in the expected significant difference.

    Release date: 2014-01-15

  • Articles and reports: 82-003-X201300511792
    Geography: Canada
    Description:

    This article describes implementation of the indoor air component of the 2009 to 2011 Canadian Health Measures Survey and presents information about response rates and results of field quality control samples.

    Release date: 2013-05-15

  • Articles and reports: 12-001-X201200211753
    Description:

    Nonresponse in longitudinal studies often occurs in a nonmonotone pattern. In the Survey of Industrial Research and Development (SIRD), it is reasonable to assume that the nonresponse mechanism is past-value-dependent in the sense that the response propensity of a study variable at time point t depends on response status and observed or missing values of the same variable at time points prior to t. Since this nonresponse is nonignorable, the parametric likelihood approach is sensitive to the specification of parametric models on both the joint distribution of variables at different time points and the nonresponse mechanism. The nonmonotone nonresponse also limits the application of inverse propensity weighting methods. By discarding all observed data from a subject after its first missing value, one can create a dataset with a monotone ignorable nonresponse and then apply established methods for ignorable nonresponse. However, discarding observed data is not desirable and it may result in inefficient estimators when many observed data are discarded. We propose to impute nonrespondents through regression under imputation models carefully created under the past-value-dependent nonresponse mechanism. This method does not require any parametric model on the joint distribution of the variables across time points or the nonresponse mechanism. Performance of the estimated means based on the proposed imputation method is investigated through some simulation studies and empirical analysis of the SIRD data.

    Release date: 2012-12-19

  • Articles and reports: 12-001-X201200211754
    Description:

    The propensity-scoring-adjustment approach is commonly used to handle selection bias in survey sampling applications, including unit nonresponse and undercoverage. The propensity score is computed using auxiliary variables observed throughout the sample. We discuss some asymptotic properties of propensity-score-adjusted estimators and derive optimal estimators based on a regression model for the finite population. An optimal propensity-score-adjusted estimator can be implemented using an augmented propensity model. Variance estimation is discussed and the results from two simulation studies are presented.

    Release date: 2012-12-19
Data (0)

Data (0) (0 results)

No content available at this time.

Analysis (135)

Analysis (135) (50 to 60 of 135 results)

  • Articles and reports: 11-522-X200800010994
    Description:

    The growing difficulty of reaching respondents has a general impact on non-response in telephone surveys, especially those that use random digit dialling (RDD), such as the General Social Survey (GSS). The GSS is an annual multipurpose survey with 25,000 respondents. Its aim is to monitor the characteristics of and major changes in Canada's social structure. GSS Cycle 21 (2007) was about the family, social support and retirement. Its target population consisted of persons aged 45 and over living in the 10 Canadian provinces. For more effective coverage, part of the sample was taken from a follow-up with the respondents of GSS Cycle 20 (2006), which was on family transitions. The remainder was a new RDD sample. In this paper, we describe the survey's sampling plan and the random digit dialling method used. Then we discuss the challenges of calculating the non-response rate in an RDD survey that targets a subset of a population, for which the in-scope population must be estimated or modelled. This is done primarily through the use of paradata. The methodology used in GSS Cycle 21 is presented in detail.

    Release date: 2009-12-03

  • Articles and reports: 11-522-X200800010996
    Description:

    In recent years, the use of paradata has become increasingly important to the management of collection activities at Statistics Canada. Particular attention has been paid to social surveys conducted over the phone, like the Survey of Labour and Income Dynamics (SLID). For recent SLID data collections, the number of call attempts was capped at 40 calls. Investigations of the SLID Blaise Transaction History (BTH) files were undertaken to assess the impact of the cap on calls.The purpose of the first study was to inform decisions as to the capping of call attempts, the second study focused on the nature of nonresponse given the limit of 40 attempts.

    The use of paradata as auxiliary information for studying and accounting for survey nonresponse was also examined. Nonresponse adjustment models using different paradata variables gathered at the collection stage were compared to the current models based on available auxiliary information from the Labour Force Survey.

    Release date: 2009-12-03

  • Articles and reports: 11-522-X200800010999
    Description:

    The choice of number of call attempts in a telephone survey is an important decision. A large number of call attempts makes the data collection costly and time-consuming; and a small number of attempts decreases the response set from which conclusions are drawn and increases the variance. The decision can also have an effect on the nonresponse bias. In this paper we study the effects of number of call attempts on the nonresponse rate and the nonresponse bias in two surveys conducted by Statistics Sweden: The Labour Force Survey (LFS) and Household Finances (HF).

    By use of paradata we calculate the response rate as a function of the number of call attempts. To estimate the nonresponse bias we use estimates of some register variables, where observations are available for both respondents and nonrespondents. We also calculate estimates of some real survey parameters as functions of varying number of call attempts. The results indicate that it is possible to reduce the current number of call attempts without getting an increased nonresponse bias.

    Release date: 2009-12-03

  • Articles and reports: 11-522-X200800011000
    Description:

    The present report reviews the results of a mailing experiment that took place within a large scale demonstration project. A postcard and stickers were sent to a random group of project participants in the period between a contact call and a survey. The researchers hypothesized that, because of the additional mailing (the treatment), the response rates to the upcoming survey would increase. There was, however, no difference between the response rates of the treatment group that received the additional mailing and the control group. In the specific circumstances of the mailing experiment, sending project participants a postcard and stickers as a reminder of the upcoming survey and of their participation in the pilot project was not an efficient way to increase response rates.

    Release date: 2009-12-03

  • Articles and reports: 11-522-X200800011001
    Description:

    Currently underway, the Québec Population Health Survey (EQSP), for which collection will wrap up in February 2009, provides an opportunity, because of the size of its sample, to assess the impact that sending out introductory letters to respondents has on the response rate in a controlled environment. Since this regional telephone survey is expected to have more than 38,000 respondents, it was possible to use part of its sample for this study without having too great an impact on its overall response rate. In random digit dialling (RDD) surveys such as the EQSP, one of the main challenges in sending out introductory letters is reaching the survey units. Doing so depends largely on our capacity to associate an address with the sample units and on the quality of that information.

    This article describes the controlled study proposed by the Institut de la statistique du Québec to measure the effect that sending out introductory letters to respondents had on the survey's response rate.

    Release date: 2009-12-03

  • Articles and reports: 11-536-X200900110804
    Description:

    This paper deals with calibration estimation for surveys with nonresponse. Efficient weighting adjustment for unit nonresponse requires powerful auxiliary information. The weights in the calibration estimator are computed on information about a specified auxiliary vector. Even with the "best possible" auxiliary vector, some bias remains in the estimator. An indicator of the remaining bias is presented and analyzed.

    The many potential auxiliary variables allow the statistician to compose a wide variety of possible auxiliary vectors. The need arises to compare these vectors to assess their effectiveness for bias reduction. To this end we examine an indicator useful for ranking alternative auxiliary vectors in regard to their ability to reduce the bias. The indicator is computed on the auxiliary vector values for the sampled units, responding and nonresponding. An advantage is its independence of the study variables, of which there are many in a large survey.

    The properties of the indicator are examined in empirical studies. A synthetic population is constructed and potential auxiliary vectors are ranked with the aid of the indicator. Another empirical illustration illustrates how the indicator is used for selecting auxiliary variables in a large survey at Statistics Sweden.

    Release date: 2009-08-11

  • Articles and reports: 11-536-X200900110805
    Description:

    The estimation of a finite population distribution function is considered in the presence of nonresponse. An imputation approach is discussed which may also be interpreted as a form of weighted estimation. It is assumed that there are complete measurements on at least one auxiliary variable which is strongly related to the variable of interest. The paper is motivated by an application to the estimation of the distribution of hourly pay using data from the Labour Force Survey in the United Kingdom. In this case the main auxiliary variable is a proxy measure of the variable of interest. Techniques discussed include predictive mean matching, nearest neighbour imputation, fractional imputation and propensity score matching. Some theoretical and numerical properties of alternative procedures will be discussed.

    Release date: 2009-08-11

  • Articles and reports: 12-001-X200900110887
    Description:

    Many survey organisations focus on the response rate as being the quality indicator for the impact of non-response bias. As a consequence, they implement a variety of measures to reduce non-response or to maintain response at some acceptable level. However, response rates alone are not good indicators of non-response bias. In general, higher response rates do not imply smaller non-response bias. The literature gives many examples of this (e.g., Groves and Peytcheva 2006, Keeter, Miller, Kohut, Groves and Presser 2000, Schouten 2004).

    We introduce a number of concepts and an indicator to assess the similarity between the response and the sample of a survey. Such quality indicators, which we call R-indicators, may serve as counterparts to survey response rates and are primarily directed at evaluating the non-response bias. These indicators may facilitate analysis of survey response over time, between various fieldwork strategies or data collection modes. We apply the R-indicators to two practical examples.

    Release date: 2009-06-22

  • Articles and reports: 12-001-X200800210756
    Description:

    In longitudinal surveys nonresponse often occurs in a pattern that is not monotone. We consider estimation of time-dependent means under the assumption that the nonresponse mechanism is last-value-dependent. Since the last value itself may be missing when nonresponse is nonmonotone, the nonresponse mechanism under consideration is nonignorable. We propose an imputation method by first deriving some regression imputation models according to the nonresponse mechanism and then applying nonparametric regression imputation. We assume that the longitudinal data follow a Markov chain with finite second-order moments. No other assumption is imposed on the joint distribution of longitudinal data and their nonresponse indicators. A bootstrap method is applied for variance estimation. Some simulation results and an example concerning the Current Employment Survey are presented.

    Release date: 2008-12-23

  • Articles and reports: 75F0002M1992009
    Description:

    There are many issues to consider when developing and conducting a survey. Length, complexity and timing of the survey are all factors that may influence potential respondents' likelihood to participate in a survey. One important issue that affects this decision is the extent to which a questionnaire appears to be an invasion of privacy. Information on income and finances is one type of information that many people are reluctant to share but that is important for policy and research purposes.

    Collecting such information for the Survey of Consumer Finances (SCF) has proven difficult, and has resulted in higher than average non-response rate for a supplemental survey to the Labour Force Survey. Given the similarity between the SCF and an upcoming survey, the Survey of Labour and Income Dynamics (SLID), it is important to examine the reasons behind the SCF's higher non-response rate and obtain suggestions for increasing response rate and gaining commitment from respondents to the 6-year SLID.

    Statistics Canada asked Price Waterhouse to conduct focus groups and in-depth interviews with respondents and non-respondents to the SCF. The objectives of these focus groups and in-depth interviews were to explore reasons for response and non-response, issues of privacy and confidentiality and understanding of the terms used in the survey, and to test reactions to the appearance of a draft SLID package.

    Release date: 2008-10-21
Reference (1)

Reference (1) ((1 result))

  • Surveys and statistical programs – Documentation: 75-005-M2023001
    Description: This document provides information on the evolution of response rates for the Labour Force Survey (LFS) and a discussion of the evaluation of two aspects of data quality that ensure the LFS estimates continue providing an accurate portrait of the Canadian labour market.
    Release date: 2023-10-30
Date modified: