Skip to main content

Advertisement

Log in

Correction for Item Response Theory Latent Trait Measurement Error in Linear Mixed Effects Models

  • Original Research
  • Published:
Psychometrika Aims and scope Submit manuscript

Abstract

When latent variables are used as outcomes in regression analysis, a common approach that is used to solve the ignored measurement error issue is to take a multilevel perspective on item response modeling (IRT). Although recent computational advancement allows efficient and accurate estimation of multilevel IRT models, we argue that a two-stage divide-and-conquer strategy still has its unique advantages. Within the two-stage framework, three methods that take into account heteroscedastic measurement errors of the dependent variable in stage II analysis are introduced; they are the closed-form marginal MLE, the expectation maximization algorithm, and the moment estimation method. They are compared to the naïve two-stage estimation and the one-stage MCMC estimation. A simulation study is conducted to compare the five methods in terms of model parameter recovery and their standard error estimation. The pros and cons of each method are also discussed to provide guidelines for practitioners. Finally, a real data example is given to illustrate the applications of various methods using the National Educational Longitudinal Survey data (NELS 88).

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. The T-score is a standardized score, which was in fact a transformation of an IRT \(\theta \) score.

  2. Originally, \({\varvec{\Sigma }}_{u}\) needs to be constrained to be nonnegative definite. However, this is not a box constraint that “optim” function can handle. We therefore impose constraints on the variance and correlation terms.

  3. We used the list-wise deletion because we wanted to create a complete data set for illustration. Our intention was to evaluate the performance of different methods without possible interference of missing data. Because we used the NELS provided item parameters and because our structural model is simple, the possible bias introduced by list-wise deletion may be ignored.

References

  • Adams, R. J., Wilson, M., & Wu, M. (1997). Multilevel item response models: An approach to errors in variables regression. Journal of Educational and Behavioral Statistics, 22(1), 47–76.

    Article  Google Scholar 

  • Anderson, J. C., & Gerbing, D. W. (1984). The effect of sampling error on convergence, improper solutions, and goodness-of-fit indices for maximum likelihood confirmatory factor analysis. Psychometrika, 49(2), 155–173.

    Article  Google Scholar 

  • Anderson, J. C., & Gerbing, D. W. (1988). Structural equation modeling in practice: A review and recommended two-step approach. Psychological Bulletin, 103, 411–423.

    Article  Google Scholar 

  • Bacharach, V. R., Baumeister, A. A., & Furr, R. M. (2003). Racial and gender science achievement gaps in secondary education. The Journal of Genetic Psychology, 164(1), 115–126.

    Article  PubMed  Google Scholar 

  • Baker, F. B., & Kim, S.-H. (2004). Item response theory: Parameter estimation techniques. NewYork: Dekker.

    Book  Google Scholar 

  • Bianconcini, S., & Cagnone, S. (2012). A general multivariate latent growth model with applications to student achievement. Journal of Educational and Behavioral Statistics, 37, 339–364.

    Article  Google Scholar 

  • Bollen, K. A. (1989). Structural equations with latent variables. New York: Wiley.

    Book  Google Scholar 

  • Broyden, C. G. (1970). The convergence of a class of double-rank minimization algorithms 1. General considerations. IMA Journal of Applied Mathematics, 6, 76.

    Article  Google Scholar 

  • Buonaccorsi, J. P. (1996). Measurement error in the response in the general linear model. Journal of the American Statistical Association, 91(434), 633–642.

    Article  Google Scholar 

  • Burt, R. S. (1973). Confirmatory factor-analytic structures and the theory construction process. Sociological Methods and Research, 2(2), 131–190.

    Article  Google Scholar 

  • Burt, R. S. (1976). Interpretational confounding of unobserved variables in structural equation models. Sociological Methods and Research, 5(1), 3–52.

    Article  Google Scholar 

  • Byrd, R. H., Lu, P., Nocedal, J., & Zhu, C. (1995). A limited memory algorithm for bound constrained optimization. SIAM Journal on Scientific Computing, 16, 1190–1208.

    Article  Google Scholar 

  • Cai, L. (2008). SEM of another flavour: Two new applications of the supplemented EM algorithm. British Journal of Mathematical and Statistical Psychology, 61(2), 309–329.

    Article  PubMed  Google Scholar 

  • Carroll, R., Ruppert, D., Stefanski, L., & Crainiceanu, C. (2006). Measurement error in nonlinear models: A modern perspective (2nd ed.). London: Chapman and Hall.

    Book  Google Scholar 

  • Chang, H., & Stout, W. (1993). The asymptotic posterior normality of the latent trait in an IRT model. Psychometrika, 58, 37–52.

    Article  Google Scholar 

  • Cohen, A. S., Bottge, B. A., & Wells, C. S. (2001). Using item response theory to assess effects of mathematics instruction in special populations. Exceptional Children, 68(1), 23–44. https://doi.org/10.1177/001440290106800102.

    Article  Google Scholar 

  • Congdon, P. (2001). Bayesian statistical modeling. Chichester: Wiley.

    Google Scholar 

  • De Boeck, P., & Wilson, M. (2004). A framework for item response models. New York: Springer.

    Book  Google Scholar 

  • De Fraine, B., Van Damme, J., & Onghena, P. (2007). A longitudinal analysis of gender differences in academic self-concept and language achievement: A multivariate multilevel latent growth approach. Contemporary Educational Psychology, 32(1), 132–150.

    Article  Google Scholar 

  • Dempster, A. P., Laird, N. M., & Rubin, D. B. (1977). Maximum likelihood from incomplete data via the EM algorithm. Journal of the Royal Statistical Society Series B (Methodological), 39, 1–38.

    Article  Google Scholar 

  • Devanarayan, V., & Stefanski, L. (2002). Empirical simulation extrapolation for measurement error models with replicate measurements. Statistics and Probability Letters, 59, 219–225.

    Article  Google Scholar 

  • Diakow, R. (2010). The use of plausible values in multilevel modeling. Unpublished masters thesis. Berkeley: University of California.

  • Diakow, R. P. (2013). Improving explanatory inferences from assessments. Unpublished doctoral dissertation. University of California-Berkley.

  • Drechsler, J. (2015). Multiple imputation of multilevel missing data—Rigor versus simplicity. Journal of Educational and Behavioral Statistics, 40(1), 69–95.

    Article  Google Scholar 

  • Fan, X., Chen, M., & Matsumoto, A. R. (1997). Gender differences in mathematics achievement: Findings from the National Education Longitudinal Study of 1988. Journal of Experimental Education, 65(3), 229–242.

    Article  Google Scholar 

  • Fletcher, R. (1970). A new approach to variable metric algorithms. The Computer Journal, 13, 317.

    Article  Google Scholar 

  • Fox, J.-P. (2010). Bayesian item response theory modeling: Theory and applications. New York: Springer.

    Book  Google Scholar 

  • Fox, J.-P., & Glas, C. A. (2001). Bayesian estimation of a multilevel IRT model using Gibbs sampling. Psychometrika, 66(2), 271–288.

    Article  Google Scholar 

  • Fox, J.-P., & Glas, C. A. (2003). Bayesian modeling of measurement error in predictor variables using item response theory. Psychometrika, 68(2), 169–191.

    Article  Google Scholar 

  • Fuller, W. (2006). Measurement error models (2nd ed.). New York, NY: Wiley.

    Google Scholar 

  • Goldfarb, D. (1970). A family of variable metric updates derived by variational means. Mathematics of Computation, 24, 23–26.

    Article  Google Scholar 

  • Goldhaber, D. D., & Brewer, D. J. (1997). Why don’t schools and teachers seem to matter? Assessing the impact of unobservables on educational productivity. The Journal of Human Resources, 32(3), 505–523.

    Article  Google Scholar 

  • Hill, H. C., Rowan, B., & Ball, D. L. (2005). Effects of teachers’ mathematical knowledge for teaching on student achievement. American Educational Research Journal, 42(2), 371–406.

    Article  Google Scholar 

  • Hong, G., & Yu, B. (2007). Early-grade retention and children’s reading and math learning in elementary years. Educational Evaluation and Policy Analysis, 29, 239–261.

    Article  Google Scholar 

  • Hsiao, Y., Kwok, O., & Lai, M. (2018). Evaluation of two methods for modeling measurement errors when testing interaction effects with observed composite scores. Educational and Psychological Measurement, 78, 181–202.

    Article  PubMed  Google Scholar 

  • Jeynes, W. H. (1999). Effects of remarriage following divorce on the academic achievement of children. Journal of Youth and Adolescence, 28(3), 385–393. https://doi.org/10.1023/A:1021641112640.

    Article  Google Scholar 

  • Kamata, A. (2001). Item analysis by the hierarchical generalized linear model. Journal of Educational Measurement, 38, 79–93.

    Article  Google Scholar 

  • Khoo, S., West, S., Wu, W., & Kwok, O. (2006). Longitudinal methods. In M. Eid & E. Diener (Eds.), Handbook of psychological measurement: A multimethod perspective (pp. 301–317). Washington, DC: APA.

    Chapter  Google Scholar 

  • Koedel, C., Leatherman, R., & Parsons, E. (2012). Test measurement error and inference from value-added models. The B. E. Journal of Economic Analysis and Policy, 12, 1–37.

    Article  Google Scholar 

  • Kohli, N., Hughes, J., Wang, C., Zopluoglu, C., & Davison, M. L. (2015). Fitting a linear–linear piecewise growth mixture model with unknown knots: A comparison of two common approaches to inference. Psychological Methods, 20(2), 259.

    Article  PubMed  Google Scholar 

  • Kolen, M. J., Hanson, B. A., & Brennan, R. L. (1992). Conditional standard errors of measurement for scale scores. Journal of Educational Measurement, 29, 285–307.

    Article  Google Scholar 

  • Lee, S., & Song, X. (2003). Bayesian analysis of structural equation models with dichotomous variables. Statistics in Medicine, 22, 3073–3088.

    Article  PubMed  Google Scholar 

  • Lindstrom, M. J., & Bates, D. (1988). Newton–Raphson and EM algorithms for linear mixed-effects models for repeated measure data. Journal of the American Statistical Association, 83, 1014–1022.

    Google Scholar 

  • Liu, Y., & Yang, J. (2018). Bootstrap-calibrated interval estimates for latent variable scores in item response theory. Psychometrika, 83, 333–354.

    Article  PubMed  Google Scholar 

  • Lockwood, L. R., & McCaffrey, D. F. (2014). Correcting for test score measurement error in ANCOVA models for estimating treatment effects. Journal of Educational and Behavioral Statistics, 39, 22–52.

    Article  Google Scholar 

  • Lu, I. R., Thomas, D. R., & Zumbo, B. D. (2005). Embedding IRT in structural equation models: A comparison with regression based on IRT scores. Structural Equation Modeling, 12(2), 263–277.

    Article  Google Scholar 

  • Magis, D., & Raiche, G. (2012). On the relationships between Jeffrey’s model and weighted likelihood estimation of ability under logistic IRT models. Psychometrika, 77, 163–169.

    Article  Google Scholar 

  • Meng, X. (1994). Multiple-imputation inferences with uncongenial sources of input. Statistical Science, 10, 538–573.

    Article  Google Scholar 

  • Meng, X., & Rubin, D. (1993). Maximum likelihood estimation via the ECM algorithm: A general framework. Biometrika, 80, 267–278.

    Article  Google Scholar 

  • Mislevy, R. J., Beaton, A. E., Kaplan, B., & Sheehan, K. M. (1992). Estimating population characteristics from sparse matrix samples of item responses. Journal of Educational Measurement, 29(2), 133–161.

    Article  Google Scholar 

  • Monseur, C., & Adams, R. J. (2009). Plausible values: How to deal with their limitations. Journal of Applied Measurement, 10(3), 320–334.

    PubMed  Google Scholar 

  • Murphy, K. (2007). Conjugate Bayesian analysis of the Gaussian distribution. Online file at https://www.cs.ubc.ca/~murphyk/Papers/bayesGauss.pdf

  • Nelder, J. A., & Mead, R. (1965). A simplex algorithm for function minimization. Computer Journal, 7, 308–313.

    Article  Google Scholar 

  • Nussbaum, E., Hamilton, L., & Snow, R. (1997). Enhancing the validity and usefulness of large-scale educational assessment: IV.NELS:88 Science achievement to 12th grade. American Educational Research Journal, 34, 151–173.

    Google Scholar 

  • Pastor, D. A., & Beretvas, N. S. (2006). Longitudinal Rasch modeling in the context of psychotherapy outcomes assessment. Applied Psychological Measurement, 30, 100–120.

    Article  Google Scholar 

  • Pinheiro, J. C., & Bates, D. M. (1995). Approximations to the log-likelihood function in the nonlinear mixed-effects model. Journal of computational and Graphical Statistics, 4(1), 12–35.

    Google Scholar 

  • Rabe-Hesketh, S., & Skrondal, A. (2008). Multilevel and longitudinal modeling using Stata. New York: STATA Press.

    Google Scholar 

  • Rabe-Hesketh, S., Skrondal, A., & Pickles, A. (2004). GLLAMM manual. Oakland/Berkeley: University of California/Berkeley Electronic Press.

    Google Scholar 

  • Raudenbush, S. W., & Bryk, A. S. (1985). Empirical Bayes meta-analysis. Journal of Educational and Behavioral Statistics, 10, 75–98.

    Article  Google Scholar 

  • Raudenbush, S. W., & Bryk, A. S. (2002). Hierarchical linear models: Applications and data analysis methods. Thousand Oaks, CA: Sage.

    Google Scholar 

  • Raudenbush, S. W., Bryk, A. S., & Congdon, R. (2004). HLM 6 for windows (computer software). Lincolnwood, IL: Scientific Software International.

    Google Scholar 

  • Raudenbush, S. W., & Liu, X. (2000). Statistical power and optimal design for multisite randomized trials. Psychological Methods, 5(2), 199.

    Article  Google Scholar 

  • Rijmen, F., Vansteelandt, K., & De Boeck, P. (2008). Latent class models for diary method data: Parameter estimation by local computations. Psychometrika, 73(2), 167–182.

    Article  PubMed  Google Scholar 

  • Rosseel, Y. (2012). Lavaan: An R package for structural equation modeling. Journal of Statistical Software, 48(2), 1–36. https://doi.org/10.18637/jss.v048.i02.

    Article  Google Scholar 

  • Shang, Y. (2012). Measurement error adjustment using the SIMEX method: An application to student growth percentiles. Journal of Educational Measurement, 49, 446–465.

    Article  Google Scholar 

  • Shanno, D. F. (1970). Conditioning of quasi-Newton methods for function minimization. Mathematics of Computation, 24, 647–656.

    Article  Google Scholar 

  • Sirotnik, K., & Wellington, R. (1977). Incidence sampling: an integrated theory for “matrix sampling”. Journal of Educational Measurement, 14, 343–399.

    Article  Google Scholar 

  • Skrondal, A., & Kuha, J. (2012). Improved regression calibration. Psychometrika, 77, 649–669.

    Article  Google Scholar 

  • Skrondal, A., & Laake, P. (2001). Regression among factor scores. Psychometrika, 66(4), 563–575.

    Article  Google Scholar 

  • Skrondal, A., & Rabe-Hesketh, S. (2004). Generalized latent variable modeling: Multilevel, longitudinal, and structural equation models. Boca Raton: CRC Press.

    Book  Google Scholar 

  • StataCorp., (2011). Stata statistical software: Release 12. College Station, TX: StataCorp LP.

  • Stoel, R. D., Garre, F. G., Dolan, C., & Van Den Wittenboer, G. (2006). On the likelihood ratio test in structural equation modeling when parameters are subject to boundary constraints. Psychological Methods, 11(4), 439.

    Article  PubMed  Google Scholar 

  • Thompson, N., & Weiss, D. (2011). A framework for the development of computerized adaptive tests. Practical Assessment, Research and Evaluation, 16(1). http://pareonline.net/getvn.asp?v=16&n=1.

  • Tian, W., Cai, L., Thissen, D., & Xin, T. (2013). Numerical differentiation methods for computing error covariance matrices in item response theory modeling: An evaluation and a new proposal. Educational and Psychological Measurement, 73(3), 412–439.

    Article  Google Scholar 

  • van der Linden, W. J., & Glas, C. A. W. (Eds.). (2010). Elements of adaptive testing (Statistics for social and behavioral sciences series). New York: Springer.

    Google Scholar 

  • Verhelst, N. (2010). IRT models: Parameter estimation, statistical testing and application in EER. In B. P. Creemers, L. Kyriakides, & P. Sammons (Eds.), Methodological advances in educational effectiveness research (pp. 183–218). New York: Routledge.

    Google Scholar 

  • von Davier, M., & Sinharay, S. (2007). An importance sampling EM algorithm for latent regression models. Journal of Educational and Behavioral Statistics, 32(3), 233–251.

    Article  Google Scholar 

  • Wang, C. (2015). On latent trait estimation in multidimensional compensatory item response models. Psychometrika, 80, 428–449.

    Article  PubMed  Google Scholar 

  • Wang, C., Kohli, N., & Henn, L. (2016). A second-order longitudinal model for binary outcomes: Item response theory versus structural equation modeling. Structural Equation Modeling: A Multidisciplinary Journal, 23, 455–465.

    Article  Google Scholar 

  • Wang, C., & Nydick, S. (2015). Comparing two algorithms for calibrating the restricted non-compensatory multidimensional IRT model. Applied Psychological Measurement, 39, 119–134.

    Article  PubMed  Google Scholar 

  • Warm, T. A. (1989). Weighted likelihood estimation of ability in item response theory. Psychometrika, 54, 427–450.

    Article  Google Scholar 

  • Ye, F. (2016). Latent growth curve analysis with dichotomous items: Comparing four approaches. British Journal of Mathematical and Statistical Psychology, 69, 43–61.

    Article  PubMed  Google Scholar 

  • Zwinderman, A. H. (1991). A generalized Rasch model for manifest predictors. Psychometrika, 56(4), 589–600.

    Article  Google Scholar 

Download references

Acknowledgements

This project is supported by IES R305D160010 and NSF SES-1659328

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Chun Wang.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Closed-form Marginal Likelihood

In this appendix, we provide detailed derivations for the closed-form marginal likelihood for a general model where the design matrices for the fixed and random effects in the latent growth curve model are different, i.e., Eq. (3) in the paper is updated as

$$\begin{aligned} \varvec{\theta }_i=\varvec{X}_i\varvec{\beta }+\varvec{Z}_i\varvec{u}_i+\varvec{e}_i. \end{aligned}$$
(1)

The subscript in \(\varvec{X}_i\) and \(\varvec{Z}_i\) indicates the model allows for unbalanced design.

Given (1) and the measurement error model, the marginal likelihood of the structural parameters, \(L(\beta , \varvec{\Sigma }_u, \sigma ^2)\), is proportional to

$$\begin{aligned}&\prod _i \int |\sigma ^2\varvec{I}_i|^{-1/2} \hbox {e}^{-\frac{1}{2\sigma ^2}(\varvec{\theta }_i-\varvec{X}_i\varvec{\beta }-\varvec{Z}_i\varvec{u}_i)^t (\varvec{\theta }_i-\varvec{X}_i\varvec{\beta }-\varvec{Z}_i\varvec{u}_i)} \hbox {e}^{-\frac{1}{2}(\varvec{\theta }^*_i-\varvec{\theta }_i)^t\varvec{\Sigma }_{\theta _i}^{-1}(\varvec{\theta }^*_i-\varvec{\theta }_i)} |\varvec{\Sigma }_u|^{-1/2} \hbox {e}^{-\frac{1}{2}\varvec{u}_i^t\varvec{\Sigma }_u^{-1}\varvec{u}_i} \mathrm d \theta _i \mathrm d u_i \\&\quad \propto \prod _i \int |\sigma ^2\varvec{\Sigma }_{\theta _i}^{-1}+\varvec{I}_i|^{-1/2} |\varvec{\Sigma }_u|^{-1/2} \hbox {e}^{-\frac{1}{2\sigma ^2}||\varvec{X}_i\varvec{\beta }+\varvec{Z}_i\varvec{u}_i||^2 +\frac{1}{2}||(\varvec{\Sigma }_{\varvec{\theta }_i}^{-1}+\sigma ^{-2}\varvec{I}_i)^{-1/2} \{\varvec{\Sigma }_{\varvec{\theta }_i}^{-1}\varvec{\theta }^*_i+\sigma ^{-2}(\varvec{X}_i\varvec{\beta }+\varvec{Z}_i\varvec{u}_i)\}||^2}\\&\qquad -\frac{1}{2}\varvec{u}_i^t\varvec{\Sigma }_u^{-1}\varvec{u}_i \mathrm d u_i, \end{aligned}$$

where \(|\cdot |\) denote the determinant of a matrix and \(||\theta ||^2 = \theta ^t\theta \).

Observing the coefficient of the squared term of \(\varvec{u}_i\) in the power of e is

$$\begin{aligned}&-\frac{1}{2}\sigma ^{-2}\varvec{Z}_i^t\varvec{Z}_i + \frac{1}{2}\sigma ^{-4}\varvec{Z}_i^t(\varvec{\Sigma }_{\theta _i}^{-1}+\sigma ^{-2}\varvec{I}_i)^{-1}\varvec{Z}_i -\frac{1}{2}\varvec{\Sigma }_u^{-1} \\&\quad = -\frac{1}{2}\left( \varvec{\Sigma }_u^{-1} + \sigma ^{-2}\varvec{Z}_i^t\varvec{Z}_i - \sigma ^{-4} \varvec{Z}_i^t(\sigma ^2\varvec{\Sigma }_{\theta _i}^{-1}+\varvec{I}_i)^{-1}\varvec{Z}_i \right) \end{aligned}$$

and the coefficient of \(u_i\) in the power of e is

$$\begin{aligned}&-\frac{1}{2}2\sigma ^{-2}\varvec{Z}_i^t(\varvec{X}_i\varvec{\beta }) + \frac{1}{2}2\varvec{Z}_i^t(\varvec{\Sigma }_{\theta _i}^{-1}+\sigma ^{-2}\varvec{I}_i)^{-1}(\varvec{\Sigma }_{\theta _i}^{-1}\varvec{\theta }^*_i + \sigma ^{-2}\varvec{X}_i\varvec{\beta }) \cdot \sigma ^{-2} \\&\quad = -\left( \sigma ^{-2}\varvec{Z}_i^t(\varvec{X}_i\varvec{\beta }) - \varvec{Z}_i^t(\sigma ^2\varvec{\Sigma }_{\theta _i}^{-1}+\varvec{I}_i)^{-1} (\varvec{\Sigma }_{\theta _i}^{-1}\varvec{\theta }^*_i + \sigma ^{-2}\varvec{X}_i\varvec{\beta })\right) \end{aligned}$$

Thus,

$$\begin{aligned} l(\beta , \varvec{\Sigma }_u, \sigma ^2) \propto&-n\log |\varvec{\Sigma }_u| - \frac{N}{\sigma ^2}||\varvec{X}_i\varvec{\beta }||^2 + \sum _{i=1}^N \left( \log |\varvec{\Sigma }_{u,i}^*| - \log |\sigma ^2\varvec{\Sigma }_{\theta _i}^{-1}+\varvec{I}_i| \right) \\&+ \sum _{i=1}^N \left( ||(\varvec{\Sigma }_{\theta _i}^{-1}+\sigma ^{-2}\varvec{I}_i)^{-1/2} \{\varvec{\Sigma }_{\theta _i}^{-1}{\theta }^*_i+\sigma ^{-2}\varvec{X}_i\varvec{\beta }\}||^2 + ||(\varvec{\Sigma }_{u,i}^*)^{1/2}\mu _{u,i}^*||^2 \right) \end{aligned}$$

where

$$\begin{aligned}&(\varvec{\Sigma }_{u,i}^*)^{-1}=\varvec{\Sigma }_u^{-1} + \sigma ^{-2}\varvec{Z}_i^t\varvec{Z}_i - \sigma ^{-4}\varvec{Z}_i^t(\sigma ^2\varvec{\Sigma }_{\theta _i}^{-1}+\varvec{I}_i)^{-1}\varvec{Z}_i\\&\varvec{\mu }_{u,i}^* = \sigma ^{-2}\varvec{Z}_i^t(\varvec{X}_i\varvec{\beta }) - \varvec{Z}_i^t(\sigma ^2\varvec{\Sigma }_{\theta _i}^{-1}+\varvec{I}_i)^{-1}(\varvec{\Sigma }_{\theta _i}^{-1}\varvec{\theta }^*_i+\sigma ^{-2}\varvec{X}_i\varvec{\beta }). \end{aligned}$$

For multivariate models, let \(\varvec{\theta }_i=(\theta _{i11},\ldots ,\theta _{i1T},\ldots ,\theta _{iD1},\ldots ,\theta _{iDT})^t\), a \(n_iD \times 1\) vector where \(n_i\) denotes the number of measurement waves for person i. Then in the general form, the associative latent growth curve model still takes the same form as in (1), but \(\varvec{X}_i\) becomes a \(n_iD \times Dp\) design matrix, and \(\varvec{\beta } = (\beta _{01}, \beta _{02},\ldots , \beta _{0D}, \beta _{11}, \beta _{12},\ldots , \beta _{1d},\ldots , \beta _{(p-1)1},\ldots , \beta _{(p-1)D})^t\) is a \(Dp \times 1\) vector. \(\varvec{Z}_i\) is \(n_iD \times Dk\) design matrix, assuming there are k random effects. In our simulation setting, \(n_i=4\), \(p=k\); hence, \(\varvec{Z}\) takes the same form as \(\varvec{X}\).

\(\varvec{u}_i\) is a \(Dk \times 1\) vector. The covariance matrix of \(\varvec{u}_i\) is \(\varvec{\Sigma }_u\). \(e_i = (e_{i11},\ldots ,e_{i1T},\ldots ,e_{iD1},\ldots ,e_{iDT})^t\) is a \(n_iD \times 1\) vector of residuals. The covariance matrix of \(e_i\), \(\varvec{\Sigma }\), is a diagonal block matrix. It has the structure of \(\begin{pmatrix} \varvec{\Sigma } &{}\quad \cdots &{}\quad \cdots &{}\quad 0 \\ 0 &{}\quad \varvec{\Sigma } &{}\quad \cdots &{}\quad 0 \\ \vdots &{}\quad \vdots &{}\quad \ddots &{}\quad \vdots \\ 0 &{}\quad \cdots &{}\quad \cdots &{}\quad \varvec{\Sigma }\\ \end{pmatrix}_{n_iD \times n_iD}\) where \(\Sigma _{D\times D} = diag((\sigma _1)^2, (\sigma _2)^2,\ldots , (\sigma _D)^2)\) and \(\Sigma \) has \(n_i\) such diagonal blocks. Then the marginal likelihood of model parameters is:

$$\begin{aligned}&L(\varvec{\beta }, \varvec{\Sigma }_u, \varvec{\Sigma }) \\&\quad \propto \prod _i \int |\varvec{\Sigma }|^{-1/2} \hbox {e}^{-\frac{1}{2} (\varvec{\theta }_i-\varvec{X}_i\varvec{\beta }-\varvec{Z}_i\varvec{u}_i)^t \varvec{\Sigma }^{-1}(\varvec{\theta }_i-\varvec{X}_i\beta -\varvec{Z}_i\varvec{u}_i)}\\&\qquad |\varvec{\Sigma }_{\varvec{\theta }_i}|^{-1/2} \hbox {e}^{-\frac{1}{2}(\varvec{\theta }^*_i-\varvec{\theta }_i)^t\varvec{\Sigma }_{\varvec{\theta }_i}^{-1}(\varvec{\theta }^*_i-\varvec{\theta }_i)} |\varvec{\Sigma }_u|^{-1/2} \hbox {e}^{-\frac{1}{2}\varvec{u}_i^t\varvec{\Sigma }_u^{-1}\varvec{u}_i} \mathrm d \varvec{\theta }_i \mathrm d \varvec{u}_i\\&\quad \propto \prod _i \int |\varvec{\Sigma }|^{-1/2} |\varvec{\Sigma }^{-1}+\varvec{\Sigma }_{\varvec{\theta }_i}^{-1}|^{-1/2}\\&\qquad \times \hbox {e}^{-\frac{1}{2}((\varvec{X}_i\varvec{\beta }+\varvec{Z}_i\varvec{u}_i)^t \varvec{\Sigma }^{-1} (\varvec{X}_i\varvec{\beta }+\varvec{Z}_i\varvec{u}_i) + \frac{1}{2} || (\varvec{\Sigma }^{-1}+\varvec{\Sigma }_{\varvec{\theta }_i}^{-1})^{-1/2} \{\varvec{\Sigma }_{\varvec{\theta }_i}^{-1}\varvec{\theta }^*_i + \varvec{\Sigma }^{-1}(\varvec{X}_i\varvec{\beta }+\varvec{Z}_i\varvec{u}_i)\} ||^2})\\&\qquad \times |\varvec{\Sigma }_u|^{-1/2} \hbox {e}^{-\frac{1}{2}\varvec{u}_i^t\varvec{\Sigma }_u^{-1}u_i} \mathrm d u_i\\&\quad \propto \prod _i |\varvec{\Sigma }|^{-1/2}|\varvec{\Sigma }^{-1}+\varvec{\Sigma }_{\varvec{\theta }_i}^{-1}|^{-1/2} |\varvec{\Sigma }_u|^{-1/2} |\varvec{\Sigma }_{u,i}^*|^{1/2}\\&\qquad \times \hbox {e}^{-\frac{1}{2}(\varvec{X}_i\varvec{\beta })^t\varvec{\Sigma }^{-1}(\varvec{X}_i\varvec{\beta }) + \frac{1}{2}|| (\varvec{\Sigma }^{-1}+\varvec{\Sigma }_{\varvec{\theta }_i}^{-1})^{-1/2} \{\varvec{\Sigma }_{\varvec{\theta }_i}^{-1}\varvec{\theta }^*_i + \varvec{\Sigma }^{-1}(\varvec{X}_i\varvec{\beta })\} ||^2 + \frac{1}{2}|| (\varvec{\Sigma }_{u,i}^*)^{1/2} \varvec{\mu }_{u,i}^* ||^2} \end{aligned}$$

where

$$\begin{aligned} (\varvec{\Sigma }_{u,i}^*)^{-1}= & {} \varvec{\Sigma }_u^{-1} + \varvec{Z}_i^t\varvec{\Sigma }^{-1}\varvec{Z}_i - \varvec{Z}_i^t\varvec{\Sigma }^{-1} (\varvec{\Sigma }^{-1} + \varvec{\Sigma }_{\varvec{\theta }_i}^{-1})^{-1} \varvec{\Sigma }^{-1} \varvec{Z}_i\\ \varvec{\mu }_{u,i}^*= & {} \varvec{Z}_i^t \varvec{\Sigma }^{-1} (\varvec{X}_i\varvec{\beta })^t - \varvec{Z}_i^t(\varvec{\Sigma }^{-1}+ \varvec{\Sigma }_{\varvec{\theta }_i}^{-1})^{-1} \varvec{\Sigma }^{-1} \{\varvec{\Sigma }_{\varvec{\theta }_i}^{-1}\varvec{\theta }^*_i+ \varvec{\Sigma }^{-1}\varvec{X}_i\varvec{\beta }\}. \end{aligned}$$

Appendix B: Computational Details of the EM Standard Error (MIRT)

The important component of computing the standard error is the complete data Fisher information matrix \({\varvec{I}}_{c} ({\hat{\varvec{\psi }}})\). Below, we present the specific forms of these components for the MIRT models. Results from the UIRT models can be considered as a special case. Assuming a Monte Carlo sampling version of the EM algorithm is used, i.e., Eq. (19), we have,

$$\begin{aligned} \frac{\partial ^{2}E(\varvec{\psi }\vert {\hat{\varvec{\psi }}}^{(m)})}{\partial \varvec{\beta }^{2}}= & {} -N{\varvec{X}}_{i}^{t} ( {\hat{{\sigma }}^{-2}\mathrm{\mathbf{I}}_{DT} } ){\varvec{X}}_{i}, \end{aligned}$$
(B1)
$$\begin{aligned} \frac{\partial ^{2}E(\varvec{\psi }\vert {\hat{\varvec{\psi }}}^{(m)})}{\partial ( {\sigma ^{2}} )^{2}}= & {} \frac{1}{2Q}\sum \limits _{i=1}^N \sum \limits _{q=1}^Q [ T{\hat{\sigma }^{ - 4}} - 2{{[ {{{({\varvec{\theta }}_i^q)}^t} - {{\varvec{X}}_i}\hat{{\varvec{\beta }}} - Z{\varvec{u}}_i^q} ]}^t}( {{{\hat{\sigma } }^{ - 6}}{{\mathbf{I}}_{DT}}} )\nonumber \\&[ {{{({\varvec{\theta }}_i^q)}^t} - {{\varvec{X}}_i}\hat{{\varvec{\beta }}} - Z{\varvec{u}}_i^q}] ], \end{aligned}$$
(B2)
$$\begin{aligned} \frac{\partial ^{2}E(\varvec{\psi }\vert {\hat{\varvec{\psi }}}^{(m)})}{\partial \varvec{\beta }\partial \sigma ^{2}}= & {} -\frac{1}{Q}\sum \limits _{i=1}^N {\sum \limits _{q=1}^Q {[ {(\varvec{\theta }_{i}^{q} )^{t}-{\varvec{X}}_{i} {\hat{\varvec{\beta }}}-Z{\varvec{u}}_{i}^{q} } ]^{t}( {\hat{{\sigma }}^{-4}\mathrm{\mathbf{I}}_{DT} } ){\varvec{X}}_{i} } }, \end{aligned}$$
(B3)

To obtain the second derivatives with respect to the elements in the covariance matrix

$$\begin{aligned} \frac{\partial ^{2}E(\varvec{\varphi }\vert \hat{{\varvec{\varphi }} }^{(m)})}{\partial x_{p}\partial x_{q}}= & {} \frac{1}{Q}\sum \limits _{i=1}^N \sum \limits _{q=1}^Q \left\{ -\frac{1}{2}\hbox {tr}\left[ \left( \hat{\Sigma }_{u}^{-1}\frac{\partial \hat{\Sigma }_{u}}{\partial x_{q}}\hat{\Sigma }_{u}^{-1} \right) \frac{\partial \hat{\Sigma }_{u}}{\partial x_{p}}\right] \right. \nonumber \\&\left. -\frac{1}{2}{({\varvec{u}}_{i}^{q})}^{t}\left[ \hat{\Sigma }_{u}^{-1}\left( \frac{\partial \hat{\Sigma }_{u}}{\partial x_{p}}\hat{\Sigma }_{u}^{-1}\frac{\partial \hat{\Sigma }_{u}}{\partial x_{q}}+\frac{\partial \hat{\Sigma }_{u}}{\partial x_{q}}\hat{\Sigma }_{u}^{-1}\frac{\partial \hat{\Sigma }_{u}}{\partial x_{p}} \right) \hat{\Sigma }_{u}^{-1} \right] {\varvec{u}}_{i}^{q} \right\} , \end{aligned}$$
(B4)

where \(x_{p}\) and \(x_{q}\) are the two elements in the covariance matrix \(\hat{\Sigma }_{u}\). For instance, using UIRT set up as an example, if taking the second derivative of log-likelihood with respect to \(\tau _{00}\), we would set \(x_{p}=x_{q}=\tau _{00}\) in (B4), and then \(\frac{\partial \hat{\Sigma }_{u}}{\partial \tau _{00}}=\left[ {\begin{array}{c@{\quad }c} 1 &{} 0\\ 0 &{} 0\\ \end{array} } \right] \). The parameters in (B1)–(B4) are final estimates upon convergence. Hence, the Fisher information matrix for the complete data has the following form as

$$\begin{aligned} {\varvec{I}}\mathrm {=-}\left[ {\begin{array}{c@{\quad }c@{\quad }c} \frac{\partial ^{2}E(\varvec{\varphi }\vert \hat{{\varvec{\varphi }} }^{(m)})}{\partial \varvec{\beta }^{2}} &{} \frac{\partial ^{2}E(\varvec{\varphi }\vert \hat{{\varvec{\varphi }} }^{(m)})}{\partial \varvec{\beta }\partial \sigma ^{2}} &{} 0\\ \frac{\partial ^{2}E(\varvec{\varphi }\vert \hat{{\varvec{\varphi }} }^{(m)})}{\partial \varvec{\beta }\partial \sigma ^{2}} &{} \frac{\partial ^{2}E(\varvec{\varphi }\vert \hat{{\varvec{\varphi }} }^{(m)})}{\partial {{\varvec{(}}\sigma ^{2}{\varvec{)}}}^{2}} &{} 0\\ 0 &{} 0 &{} \frac{\partial ^{2}E(\varvec{\varphi }\vert \hat{{\varvec{\varphi }} }^{(m)})}{\partial {(\Sigma _{{\varvec{u}}})}^{2}}\\ \end{array} } \right] . \end{aligned}$$

The information matrix can also be obtained similarly if a closed-form conditional expectation is obtained.

Appendix C: The MCMC Algorithm

The Metropolis–Hastings algorithm within Gibbs sampler is used. For the ease of exposition below, we can rewrite the linear mixed model in (3) as follows:

$$\begin{aligned} \begin{aligned} \theta _{it}&=\pi _{0i} +\pi _{1i} \left( {t-1} \right) +e_{it} \\ \pi _{0i}&=\beta _{00} +u_{0i} \\ \pi _{1i}&=\beta _{01} +u_{1i} \end{aligned}, \end{aligned}$$
(C1)

The conjugate priors are selected whenever available. Below is an outline of the sampling schemes. At the \((m+1)^{\mathrm {th}}\) iteration, we have

Step 1: Sample \({{{\varvec{\uptheta }}}}_{i}^{*} \sim N({{{\varvec{\uptheta }} }}_{i}^{(m)},\sigma _{\theta }^{2} )\) and \(\mathrm{\mathbf{u}}_{i} \sim U( {0,1} )\), and set \(\theta _{it}^{(m+1)} =\theta _{it}^{*} \) when

$$\begin{aligned} \mathrm{\mathbf{u}}_{i} <\frac{P( {{{{\varvec{\uptheta }} }}_{i}^{*} | {{\varvec{X}}_{i} {{{\varvec{\uppi }} }}_{i}^{(m)},\sigma ^{2(m)}} } )L( {{\varvec{Y}}_{i} | {{{{\varvec{\uptheta }} }}_{i}^{*} } } )}{P( {{ {{\varvec{\uptheta }} }}_{i}^{(m)} | {{\varvec{X}}_{i} {{{\varvec{\uppi }} }}_{i}^{(m)},\sigma ^{2(m)}} } )L( {{\varvec{Y}}_{i} | {{{ {\varvec{\uptheta }} }}_{i}^{(m)} } } )}, \end{aligned}$$
(C2)

where \(\varvec{\pi }_{i}={(\pi _{0i},\pi _{1i})}^{t},\)\(L({\varvec{Y}}_{i} \vert \theta _{it}^{*} )\) is the likelihood obtained from item responses, and \(P({{{\varvec{\uptheta }} }}_{i}^{*} \vert {\mathbf{X}{\varvec{\uppi }} }_{i}^{(m)},\sigma ^{2(m)})\) is the normal density with a mean of \({\varvec{X}}_{i} {{{\varvec{\uppi }} }}_{i}^{(m)} \) and variance of \(\sigma ^{2(m)}\).

Step 2: Sample \(\pi _{i}^{(m+1)} \) from the multivariate normal distribution with covariance

$$\begin{aligned} \Sigma _{{{{\varvec{\uppi }} }}_{i} \left| {{{{\varvec{\uptheta }}}}_{i} ,\sigma ^{2},} \right. \mathrm{\mathbf{X}},{{{\varvec{\upbeta }}}},{{\varvec{\Sigma }}}_{u} } =\left( {\sigma ^{-2(m)}{\varvec{X}}_{i}^{t} {\varvec{X}}_{i} +{{\varvec{\Sigma }}}_{u}^{-1(m)} } \right) ^{-1}, \end{aligned}$$
(C3)

and mean

$$\begin{aligned} {{\varvec{\upmu } }}_{{{{\varvec{\uppi }}}}_{i} | {{{{\varvec{\uptheta }}}}_{i} ,\sigma ^{2},{\mathbf {X}},{\varvec{\upbeta }},{{\varvec{\Sigma }}}_{u} } } =\Sigma _{{\varvec{\uppi }}_{i} | {{\varvec{\uptheta }}_{i} ,\sigma ^{2},{\mathbf {X}},{\varvec{\upbeta }},{{\varvec{\Sigma }}}_{u} } } \cdot ( {{{\varvec{\Sigma }}}_{u}^{-1(m)} {\varvec{\upbeta }}^{(m)}+\sigma ^{-2(m)}{\mathbf {X}}^{t}{\varvec{\uptheta }}_{i}^{(m+1)} } ). \end{aligned}$$
(C4)

Step 3: Sample \(\sigma ^{2(m+1)}\) from the full condition distribution

$$\begin{aligned} \sigma ^{2(m+1)}\sim \hbox {Inv-Gamma}\left( {\alpha _{0} +\frac{TN}{2},\alpha _{1} +\frac{1}{2}\sum \limits _i {( {{\varvec{\uptheta }}_{i}^{(m+1)} -{\varvec{X}}_{i} {\varvec{\uppi }}_{i}^{(m+1)} } )^{t}( {{\varvec{\uptheta }}_{i}^{(m+1)} -{\varvec{X}}_{i} {\varvec{\uppi }}_{i}^{(m+1)} } )} } \right) ,\nonumber \\ \end{aligned}$$
(C5)

where the prior distribution of \(\sigma ^{2}\) is Inv-Gamma (\(\alpha _{0},\alpha _{1})\). We selected \(\alpha _{0}=.0001\) and \(\alpha _{1}=1\) (Congdon, 2001) as the hyper-parameters of a non-information prior distribution.

Step 4: Sample \({\varvec{\upbeta }}^{(m+1)}\) and \({{\varvec{\Sigma }}}_{u}^{(m+1)} \) from the normal inverse Wishart distribution with parameters \(( {{{\varvec{\upmu } }}_{{\varvec{\upbeta }}| {{\varvec{\uppi }}} }^{(m+1)},\kappa _{n},{{\varvec{\Sigma }}}_{{\mathbf {u}}| {{\varvec{\uppi }}} }^{(m+1)},\nu _{n} } )\). Here

$$\begin{aligned}&\mu _{{{\varvec{\upbeta }}|\pi } }^{(m+1)} =\frac{\kappa _{0} }{\kappa _{0} +N}\mu _{{\varvec{\upbeta }},0} +\frac{N}{\kappa _{0} +N}{{\bar{\varvec{\uppi }}}}^{(m+1)}, \end{aligned}$$
(C6)
$$\begin{aligned}&{{\varvec{\Sigma }}}_{u| \pi }^{(m+1)} \nonumber \\&\quad =\Sigma _{u,0} +\sum \limits _i {( {{\varvec{\uppi }}_{i}^{(m+1)} -{\varvec{\bar{{\uppi }}}}^{(m+1)}} )} ( {{\varvec{\uppi }}_{i}^{(m+1)} -{{\bar{\varvec{\uppi }}}}^{(m+1)}} )^{t}\nonumber \\&\qquad +\frac{\kappa _{0} }{\kappa _{0} +n}( {{{\bar{\varvec{\uppi }}}}^{(m+1)}-{{\varvec{\upmu } }}_{{\varvec{\upbeta }},0} } )( {{{\bar{\varvec{\uppi }}}}^{(m+1)}-{{\varvec{\upmu } }}_{{\varvec{\upbeta }},0} } )^{t},\nonumber \\&\kappa _{n} =\kappa _{0} +N,\nonumber \\&\nu _{n} =\nu _{0} +N, \end{aligned}$$
(C7)

where \({{\bar{\varvec{\uppi }}}}^{(m+1)}=( {\sum \nolimits _i {{\pi _{0i}^{(m+1)} } / {N},} \sum \nolimits _i {{\pi _{1i}^{(m+1)} } / N} })^{t}\). And the prior distribution is a normal inverse Wishart distribution with parameters \(({{{\varvec{\upmu } }}_{{\varvec{\upbeta }},0},\kappa _{0},{{\varvec{\Sigma }}}_{u,0},\nu _{0} })\). Regarding the hyper-parameters, \({\varvec{\upmu }}_{{\beta ,0}}\) is a 2-by-1 zero vector, \(\kappa _{{0}}=0\), \({{\varvec{\Sigma }}}_{u,0} \) is a 2-by-2 identity matrix, and \(\nu _{{0}}=-1\), yielding a non-informative prior (Murphy, 2007).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, C., Xu, G. & Zhang, X. Correction for Item Response Theory Latent Trait Measurement Error in Linear Mixed Effects Models. Psychometrika 84, 673–700 (2019). https://doi.org/10.1007/s11336-019-09672-7

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11336-019-09672-7

Keywords

Navigation