• 沒有找到結果。

XiaowenZhu andClementA.Stone BayesianComparisonofAlternativeGradedResponseModelsforPerformanceAssessmentApplications

N/A
N/A
Protected

Academic year: 2022

Share "XiaowenZhu andClementA.Stone BayesianComparisonofAlternativeGradedResponseModelsforPerformanceAssessmentApplications"

Copied!
26
0
0

加載中.... (立即查看全文)

全文

(1)

Ó The Author(s) 2012 Reprints and permission:

sagepub.com/journalsPermissions.nav DOI: 10.1177/0013164411434638 http://epm.sagepub.com

Alternative Graded Response Models for

Performance Assessment Applications

Xiaowen Zhu1and Clement A. Stone2

Abstract

This study examined the relative effectiveness of Bayesian model comparison meth- ods in selecting an appropriate graded response (GR) model for performance assess- ment applications. Three popular methods were considered: deviance information criterion (DIC), conditional predictive ordinate (CPO), and posterior predictive model checking (PPMC). Using these methods, several alternative GR models were compared with Samejima’s unidimensional GR model, including the one-parameter GR model, the rating scale model, simple- and complex-structure two-dimensional GR models, and the GR model for testlets. Results from a simulation study indicated that these methods appeared to be equally accurate in selecting the preferred model.

However, CPO and PPMC can be used to compare models at the item level, and PPMC can also be used to compare both the relative and absolute fit of different models.

Keywords

Bayesian model comparison, DIC, CPO, posterior predictive model checking, graded response models, multidimensionality, local independence

1Xi’an Jiaotong University, Xi’an, China

2University of Pittsburgh, Pittsburgh, PA, USA Corresponding author:

Xiaowen Zhu, Department of Sociology & Institute for Empirical Social Science Research (IESSR), Xi’an Jiaotong University, Xi’an, Shaanxi, 710049, China

Email: xwzhu@mail.xjtu.edu.cn

(2)

With the increased popularity of including performance-based items in large-scale assessments, standard unidimensional polytomous item response theory (IRT) mod- els are commonly used to analyze performance assessment data. However, the under- lying assumptions (e.g., unidimensionality and local item independence) may be violated for performance assessment responses (Lane & Stone, 2006), and thus, more complex polytomous models may be needed for analysis. For example, a multidi- mensional polytomous model may be more appropriate for responses exhibiting mul- tidimensionality, or a polytomous model for testlets may be more suitable for assessments that involve a subset of items with a common stimulus. To choose an appropriate model for a particular testing application, model comparison techniques can be employed.

For more complex IRT models, it has become increasingly common to estimate models using Bayesian methods with Markov Chain Monte Carlo (MCMC) algo- rithms. For these methods, different types of Bayesian approaches have been dis- cussed for comparing competing models: (a) Bayesian information-based criteria, such as deviance information criterion (DIC; Spiegelhalter, Best, Carlin, & van der Linde, 2002) and (b) indices based on Bayes factors (BF; Kass & Raftery, 1995), such as the pseudo-Bayes factor (PsBF; Geisser & Eddy, 1979; Gelfand, Dey, &

Chang, 1992) and the conditional predictive ordinate (CPO; Kim & Bolt, 2007).

Also, the posterior predictive model checking (PPMC) method can be used for model comparisons, though it is essentially a Bayesian tool for evaluating model data-fit (e.g., Be´guin & Glas, 2001; Li, Bolt, & Fu, 2006; Sinharay, 2005).

Research comparing different model comparison methods in IRT applications has been limited. Li et al. (2006) investigated the performance of three Bayesian methods (DIC, PsBF, and PPMC) in selecting a testlet model for dichotomous item responses.

They found that PsBF and PPMC were effective in identifying the data-generating testlet models, but DIC tended to favor more complex testlet models. Kang and Cohen (2007) examined the accuracy of the G2 likelihood test statistic, Akaike’s information criterion (AIC; Akaike, 1974), Bayesian information criterion (BIC;

Schwarz, 1978), DIC, and PsBF indices in selecting the most appropriate dichoto- mous IRT model. The PsBF index was found to be the more accurate index. The DIC index tended to select a more complex model, and conversely, the BIC index tended to select a simpler model. Kang, Cohen, and Sung (2009) later compared the performance of the AIC, BIC, DIC, and PsBF indices with respect to the selection of unidimensional polytomous IRT models. They found that BIC was accurate and con- sistent in selecting the true or simulated polytomous model. Performance of AIC was similar to BIC whereas DIC and PsBF were less accurate in some conditions when the true model was the graded response (GR) model.

The above findings indicate that model comparison methods perform inconsistently in Bayesian IRT applications. Their performance has been dependent on specific con- ditions as well as specific models that were compared. The purpose of this study was to extend previous research specifically to performance assessment applications and to investigate the effectiveness of different Bayesian methods in selecting a model

(3)

from a set of alternative models that are applicable to performance assessments. The set of models included Samejima’s (1969) unidimensional two-parameter GR (2P- GR) model, a one-parameter version of the GR model (1P-GR), the rating scale (RS) model (Muraki, 1990), the simple- and complex-structure two-dimensional GR mod- els (De Ayala, 1994), and the GR model for testlets (Wang, Bradlow, & Wainer, 2002). All models were estimated in WinBUGS (Spiegelhalter, Thomas, Best, &

Lunn, 2003) using MCMC methods, and compared using three Bayesian model com- parison approaches: the DIC index, the CPO index, and the PPMC method.

Model Comparison Criteria Deviance Information Criterion

DIC (Spiegelhalter et al., 2002) is an information-based index similar to other infor- mation criteria indices in that it weights both model fit and model complexity in iden- tifying a preferred model. The DIC index is widely used with Bayesian estimation with MCMC methods and defined as

DIC = D hð Þ + pD: ð1Þ

The first term, D hð Þ, is the posterior mean of the deviance between the data and a model and defined as

D hð Þ = Ehjy½D hð Þ = Ehjy½2 log p yjhð Þ, ð2Þ where y represents the response data, h denotes the model parameters, and p yjhð Þ is the likelihood function. The second term in Equation (1), pD, represents the effective number of parameters in the model or model complexity. It is defined as the differ- ence between the posterior mean of the deviance and the deviance at the posterior mean of model parameters ^h:

pD= D hð Þ  D ^ð Þ:h ð3Þ Note that DIC is a Bayesian generalization (Speigelhalter et al., 2002) of AIC and is related to BIC. DIC was developed since MCMC estimation uses prior information and the actual number of parameters cannot be clearly identified as required for the AIC and BIC indices. As for AIC and BIC, smaller values of DIC indicate better fit.

DIC was the focus of this study since it is more commonly used for comparing mod- els estimated with Bayesian methods and is more readily accessible to researchers using WinBUGS.

Conditional Predictive Ordinate

Bayes factors (BF; Kass & Raftery, 1995) have traditionally been used with Bayesian methods to compare competing models based on the observed data. The BF for

(4)

comparing Model 1 (M1) with Model 2 (M2) is defined by the posterior odds of M1

to M2divided by the prior odds of M1to M2. Using Bayes theorem, the BF reduces to the ratio of marginal likelihoods of the data (y) under each model:

BF = P Mð 1jyÞ=P Mð 2jyÞ

P Mð 1Þ=P Mð 2Þ =P yjMð 1Þ

P yjMð 2Þ: ð4Þ

A BF larger than 1 supports selection of M1and a value less than 1 supports selec- tion of M2. Research has shown that there are several issues with using this method in practice. For instance, the calculation of BF from MCMC output becomes difficult for complex models, and it is not well defined for improper priors (Li et al., 2006).

As a result, PsBF (Geisser & Eddy, 1979; Gelfand et al., 1992) and CPO (Kim &

Bolt, 2007) indices have been discussed as surrogates for BF.

Both PsBF and CPO indices require calculation of cross-validation predictive den- sities. Let y(r), obsdenote the set of observations yobswith the rthobservation omitted, and let h denote all the parameters under the assumed model. The cross-validation predictive density is defined as

f y rjy(r)

= ð

f y rjh, y(r)

f hjy (r)

dh: ð5Þ

The density fðyrjy(r)Þ indicates the values of yr that are likely when the model is estimated from all observations except yr. In the context of item response data, yrrep- resents a single examinee’s response to an individual item. The product of fðyrjy(r)Þ across all observations can be used as an estimate of the marginal likelihood in Equation (4). The PsBF of Model 1 against Model 2 is defined as

PsBF = QR

r = 1

f yr, obsjy(r), obs, M1

 

QR

r = 1

f yr, obsjy(r), obs, M2

  , ð6Þ

where R denotes the total number of item responses from all examinees. When com- paring the models at the item level, R equals the number of examinees N. When comparing the models at the test level, R equals the number of the responses for all examinees to all items (i.e., R = N 3 I, where I is the total number of items). As for the BF index, a PsBF greater than 1 supports selection of Model 1 and a value less than 1 supports selection of Model 2.

The CPO index for a model can be defined as

CPOM= YR

r = 1

f y r, obsjy(r), obs, M

: ð7Þ

(5)

The model with larger CPO value is preferred. It is easy to see that the same con- clusions will be obtained using either CPO or PsBF indices, so only the CPO index was used in the current study.

Posterior Predictive Model Checking

PPMC (Rubin, 1984) is a flexible Bayesian model-checking tool that has recently proved useful in assessing different aspects of fit for IRT models (e.g., Fu, Bolt, &

Li, 2005; Levy, Mislevy, & Sinharay, 2009; Sinharay, 2005, 2006; Sinharay, Johnson, & Stern, 2006; Zhu & Stone, 2011). PPMC involves simulating data using posterior distributions for model parameters and comparing features of simulated data against observed data. A discrepancy measure computed on both observed and simulated data is used to evaluate whether differences reflect potential misfit of a model (Gelman, Carlin, Stern, & Rubin, 2003). The comparison of realized (or observed) and posterior predictive (or simulated) values for the discrepancy measure can be performed using graphical displays as well as by computing posterior predic- tive p (PPP) values. PPP values provide numerical summaries of model fit. They reflect the relative occurrence of a value for a discrepancy measure D y, hð Þ based on the observed data y in the distribution of discrepancy values from replicated values D yð rep, hÞ:

PPP = P D yð ð rep, hÞ  D y, hð ÞjyÞ = ðð

D yðrep, hÞ  D y, hð Þ

p yð repjhÞp hjyð Þdyrepdh, ð8Þ

where y represents the response data, yreprepresents the replicated data based on the model, and h is a vector of model parameters. PPP values near 0.5 indicate that there is no systematic difference between the realized and predictive discrepancies, and thus indicate adequate fit of the model. PPP values near 0 or 1 (e.g., values \.05 or ..95) suggest conversely that the realized discrepancies are inconsistent with the posterior predictive discrepancy values, and thus indicate inadequate model-data-fit (e.g., Gelman et al., 2003; Levy et al., 2009; Sinharay, 2005).

PPMC has been found useful in model comparisons when Bayesian estimation is used (e.g., Be´guin & Glas, 2001; Li et al., 2006; Sinharay, 2005). The relative fit of a set of alternative candidate models can be evaluated by comparing the numbers of extreme PPP values across items (or item pairs). A model with fewer numbers of items (or item pairs) with extreme PPP values is considered to fit the data better than an alternative model with larger numbers of items (or item pairs) with extreme PPP values.

Selecting appropriate discrepancy measures is an important consideration in appli- cations of the PPMC method. Discrepancy measures should be chosen to reflect sources of potential misfit most relevant to a testing application. In this study, the proposed measures included four item-level measures: the item score distribution, the item–total score correlation, Yen’s Q1statistic (Yen, 1981), and the pseudo-count fit (e.g., Stone, 2000); and two pairwise measures: global odds ratios (global ORs;

(6)

Agresti, 2002), and Yen’s Q3statistic (Yen, 1993). These particular measures were selected from a broader set based on their usefulness to detect IRT model-data-fit in previous testing applications (cf., Fu et al., 2005; Levy et al., 2009; Sinharay, 2005, 2006; Sinharay et al., 2006; Zhu & Stone, 2011).

1. Item-level measures. The item score distribution is an intuitive measure for assessing item-level fit. It represents the number of examinees responding to each response category for each item. The difference between observed and posterior pre- dictive item score distributions can be summarized using a goodness-of-fit statistic (x2j). For polytomous items, this statistic can be defined as

x2j =XMj

k = 0

Ojk Ejk

 2

Ejk

, ð9Þ

where Mjis the highest score on Item j, and Ojkand Ejkare the observed and poster- ior predicted number of examinees scoring in response category k on Item j, respec- tively. Ejkis calculated by summing the probabilities of responding to category k on Item j across all N examinees: Ejk= PN

i = 1

pijkð Þ:ui

The item–total score correlation measure is the correlation between examinees’

total test scores and their scores on a particular item. This measure reflects item dis- crimination and therefore should be sensitive to any violations of the equal discrimi- nation assumption. Sinharay et al. (2006) found that this correlation was effective in detecting misfit of 1PL models to data based on 2PL or 3PL models. Therefore, it was expected that this measure would be also useful in discriminating between the 1P-GR and 2P-GR models to be compared in this study. Pearson correlations were used to estimate associations between the five-category items and total test scores.

A number of other statistics have been proposed to assess the fit of IRT models at the item level (e.g., Orlando & Thissen, 2000; Stone, 2000; Yen, 1981). All these statistics compare observed response score distributions and model predictions at dis- crete levels of ability for each item. These item-fit statistics performed well for detecting item misfit in the frequentist framework, and several were also found effec- tive with PPMC (e.g., Sinharay, 2006; Zhu & Stone, 2011). In this study, two of these statistics were used with PPMC: Yen’s Q1(Yen, 1981) and the pseudo-count fit statistic (e.g., Stone, 2000). The former is a traditional item-fit index and the latter is an alternative index developed specifically for performance assessment applications.

Yen’s Q1item-fit statistic is a goodness-of-fit statistic defined as

x2= X10

j = 1

XK

k = 1

NjOjk Ejk2

Ejk , ð10Þ

where Njis the number of examinees within ability subgroup j, Ojkand Ejkare the observed and predicted proportion of responses to category k for ability subgroup j, respectively. In Yen’s statistic, examinees are divided into 10 ability subgroups of

(7)

approximately equal size after they are rank-ordered by their ability estimates. The expected proportion to a response category for a subgroup is the mean of the prob- abilities of responses to that category for all the examinees in that subgroup.

The pseudo-count fit statistic is a goodness-of-fit statistic developed to account for the imprecision in ability estimation for short tests, such as performance assessments.

Whereas Yen’s Q1cross-classifies examinees into only one cell of the item-fit table based on each item response and point estimate of ability, the pseudo-count fit statis- tic assigns each examinee to multiple ability groups based on posterior expectations (posterior probabilities for each discrete ability level). Summing the posterior expec- tations across all examinees provides a pseudo-observed score distribution. Model- based predictions for score categories at each ability level are calculated in the same way as in Yen’s Q1statistic.

It may be worthy to note that a difference between these statistics and the item- score distribution index is that the observed and model-based predictions are compared at different discrete levels of the ability parameter in IRT models. Thus, summations in the equations are across response categories and discrete ability levels.

2. Pairwise measures. Yen’s Q3index (Yen, 1981) and an OR measure were used as possible pairwise measures in this study. These measures are classic local- dependence indices and reflect the association between responses to item pairs.

Yen’s Q3statistic measures the correlations between pairs of items after accounting for the latent ability. For Items j and j0, Q3is defined as the correlation of deviation scores across all examinees: Q3jj9= r d j, dj9

, where dj is the deviation between observed and expected performance on Item j. Alternatively, Chen and Thissen (1997) used OR to evaluate local dependence for dichotomous items. The OR for dichotomous item pairs (j and j*) are computed from 2 3 2 tables by n00n11=n01n10, where npq is the observed number of examinees having response p (0 or 1) on Item j and response q (0 or 1) on Item j*.

In contrast with dichotomously scored items, multiple ORs can be computed with polytomous items since the contingency table is R 3 C (R . 2 and C . 2). For this study, a global OR (Agresti, 2002) was used as a possible discrepancy measure. For any two items, the R 3 C contingency table may be reduced to a 2 3 2 contingency table by dichotomizing the response categories of each item. The global OR was then defined as the cross-ratio of this pooled 2 3 2 table. In this study, the dichotomization was based on score rubrics often used with performance assessments. For items with 5 response categories (0-4), Categories 3 and 4 were treated as ‘‘correct’’ responses, and Categories 0, 1, and 2 were treated as ‘‘incorrect’’ responses.

Both Yen’s Q3and OR measures have been found to be effective in detecting mul- tidimensionality and local dependence among item responses with PPMC (e.g., Levy et al., 2009; Sinharay et al., 2006; Zhu & Stone, 2011). Therefore, they were expected to be useful in comparing unidimensional models with multidimensional IRT models or testlet models considered in this study. Although Q3has been found to be more effective than the OR measure when evaluating the fit of a single IRT model with PPMC (e.g., Levy et al., 2009; Zhu & Stone, 2011), both measures were used in the

(8)

present study to evaluate their respective usefulness in the context of a model com- parison approach.

Alternative Graded Response Models for Performance Assessments

Samejima’s (1969) 2P-GR model is a commonly used polytomous model for perfor- mance assessment applications. For Item j with (mj + 1) response categories (0, 1, 2, . . ., mj), the probability that an examinee receives a category score x (x = 1, 2, . . ., mj) or higherðPijxð ÞÞ is modeled by the logistic deviate eui Z, where z = Da jui bjx

, D is the scaling constant (1.7 or 1), uiis examinee ability, ajis the discrimination (slope) parameter for Item j, and bjxis a threshold parameter for Category x of Item j. The probability of a particular response in Category x for an examinee on Item jðPijxð ÞÞui is then defined as the difference between the cumulative probabilities for two adjacent categories: Pij(x1)ð Þ and Pui ijxð Þ, with two constraints Pui ij0ð Þ = 1 andui

Pij(mi+ 1)ð Þ = 0.ui

For the standard 2P-GR model, one slope parameter is estimated for each item.

Under the assumption that all items have the same discrimination, the 2P-GR model is reduced to a 1P-GR model with z = Da u  i bjx

. Therefore, the 1P model is a restricted version of the standard GR model with a common slope parameter.

Another restricted case of the 2P-GR model is Muraki’s (1990) RS model. In this model, z = Da jui (bj cx)

and the threshold parameters (bjx) in the 2P-GR model are partitioned into two terms: a location parameter (bj) for each item, and one set of category threshold parameters (cx) that applies to all items. The RS model is a restricted version of the 2P-GR model since category threshold parameters are assumed to be equal across all items in the RS model, whereas they are free to vary across items in the 2P-GR model. As a result, the number of parameters in the RS model is reduced greatly as compared with the standard 2P-GR model. The RS model was originally developed for analyzing responses to items with a rating-scale type response format. However, Lane and Stone (2006) pointed out that the RS model may be appropriate for performance assessments. When a general rubric is used as the basis for developing specific item rubrics, the response scales and the differences between score levels may be the same across the set of items.

The above models are appropriate for analyzing item responses that are assumed to be determined by one latent trait (i.e., unidimensional). When a performance assessment is designed to measure more than one ability, the item responses exhibit a multidimensional structure. De Ayala (1994) discussed a multidimensional version of the GR model, where z = DP

h

ajhuh bjx

and uhis the ability level on dimen- sion h, ajhis the discrimination (slope) parameter of Item j on dimension h, and bjxis the threshold parameter for Category x of Item j.

Finally, in performance assessments, a single stimulus (e.g., passage) may be used with several items (Yen, 1993). The responses to these items are therefore likely to

(9)

be locally dependent, and standard IRT models may be inappropriate. A modified GR model for testlets was proposed by Wang, Bradlow, and Wainer (2002), where z = Da jui bjx gid( j)

. In this model, a random testlet effect (gid( j)) is intro- duced to reflect an interaction for Person i with testlet d( j). Ability ui is typically assumed to have a N(0, 1) distribution, and gid( j)is assumed to be distributed as N(0, s2d( j)). The variance of gid( j) varies across testlets, and its value indicates the amount of local dependence in each testlet. As s2d( j)increases, the amount of local dependence increases. When s2d( j)= 0, the items within the testlet are conditionally independent.

Simulation Study

Design of the Simulation Study

A simulation study was conducted to evaluate the use of Bayesian model comparison methods for selecting a model among alternative or competing GR models for per- formance assessment applications. The set of alternative models reflected sources of potential misfit for standard GR models and reflected models that may be theoreti- cally more appropriate for some types of performance assessments.

Table 1 presents the design and specific conditions used in this simulation, and Table 2 presents item parameters for each GR model. For each condition, 20 response data sets were generated based on the data-generating model (Mg) with each data set containing responses for 2,000 simulated examinees to 15 polytomous items with 5 response categories. The test length and the number of response categories were fixed at typical values in performance assessment applications. Note that performance assessments are comprised commonly of forms with fewer items than multiple-choice type tests (Lane & Stone, 2006). A large sample size of 2,000 was used to ensure that model parameters were estimated precisely, and model selection results would not be affected by any inaccuracy in model parameter estimation.

Condition 1 was used to evaluate the effectiveness of the three model comparison methods in discriminating between the 2P-GR, 1P-GR, and RS models for the unidi- mensional responses simulated under the 2P-GR model. As shown in Table 2, the Table 1. Design and Conditions in this Simulation Study

Data generating model (Mg) Data analysis model (Ma) Condition Number Two-parameter unidimensional GR

(2P-GR)

2P-GR vs. 1P-GR vs. RS 1

Two-dimensional simple-structure GR (2D simple-GR)

2P-GR vs. 2D simple-GR 2

Two-dimensional complex-structure GR (2D complex-GR)

2P-GR vs. 2D complex-GR 3

Testlet GR 2P-GR vs. testlet GR 4

Note. GR = graded response; 1P-GR = 1-parameter unidimensional GR model; RS = rating scale model.

(10)

configuration of item parameters for the 2P-GR model involved a combination of three slope parameters: 1.0, 1.7, and 2.4 (reflecting low, average, and high discrimi- nation) and five threshold parameter configurations (reflecting different difficulty levels): (a) 22.0, 21.0, 0.0, 1.0; (b) 21.5, 20.5, 0.5, 1.5; (c) 21.0, 0.0, 1.0, 2.0; (d) 23.0, 21.5, 20.5, 1.0; (e) 21.0, 0.5, 1.5, 3.0. Ability parameters were randomly simulated from a N(0, 1) distribution.

In Condition 2, the simulated test measured two dimensions but each item mea- sured only one dimension (i.e., simple-structure case). Specifically, the first eight items were designed to measure one dimension, and the remaining seven items mea- sured the second dimension. As shown in Table 2, the slope parameters on Dimension 1 (a1) ranged from 1.0 to 2.4 for Items 1 to 8, and 0s for Items 9 to 15.

The slope parameters on Dimension 2 (a2) were 0 for Items 1 to 8 and ranged from 1.0 to 2.4 for Items 9 to 15. The threshold parameters for these 15 items were the same as for the 2P-GR model. Ability parameters on two dimensions were randomly selected from a bivariate normal distribution (0, 1) with a fixed correlation of 0.60, a correlation that may represent a value typical for large-scale performance assess- ments (cf., Lane, Stone, Ankenmann, & Liu, 1995). The responses were simulated based on the multidimensional GR model, and the three methods were used to com- pare the fit of the unidimensional 2P-GR model and the two-dimensional simple- structure (2D simple) GR model.

For Condition 3, two-dimensional complex-structure (2D complex case) responses were simulated to reflect a performance assessment that not only measures a Table 2. Item Parameters for Alternative Graded Response (GR) Models Under Conditions 1 to 4

2P-GR and testlet GR 2D simple-GR 2D complex-GR

Item a b1 b2 b3 b4 a1 a2 a1 a2

1 1.0 22.0 21.0 0.0 1.0 1.0 0.0 1.0 0.5

2 1.0 21.5 20.5 0.5 1.5 1.7 0.0 1.0 0.5

3 1.0 21.0 0.0 1.0 2.0 2.4 0.0 1.0 0.5

4 1.0 23.0 21.5 20.5 1.0 1.0 0.0 1.0 0.5

5 1.0 21.0 0.5 1.5 3.0 1.7 0.0 1.0 0.5

6 1.7 22.0 21.0 0.0 1.0 2.4 0.0 1.7 0.0

7 1.7 21.5 20.5 0.5 1.5 1.0 0.0 1.7 0.0

8 1.7 21.0 0.0 1.0 2.0 1.7 0.0 1.7 0.0

9 1.7 23.0 21.5 20.5 1.0 0.0 2.4 1.7 0.0

10 1.7 21.0 0.5 1.5 3.0 0.0 1.0 1.7 0.0

11 2.4 22.0 21.0 0.0 1.0 0.0 1.7 2.4 0.0

12 2.4 21.5 20.5 0.5 1.5 0.0 2.4 2.4 0.0

13 2.4 21.0 0.0 1.0 2.0 0.0 1.0 2.4 0.0

14 2.4 23.0 21.5 20.5 1.0 0.0 1.7 2.4 0.0

15 2.4 21.0 0.5 1.5 3.0 0.0 2.4 2.4 0.0

(11)

dominant ability (e.g., math word problems) but also consists of items that measure a nuisance or construct-irrelevant dimension (e.g., reading). All 15 items measured the dominant dimension, but the first 5 items were designed to also measure a nuisance dimension (see Table 2). The degree to which item performance depended on the nui- sance ability was captured by the ratio of the slope (a1) for the dominant ability to the slope (a2) for the nuisance ability. The ratio of a2to a1was set to 0.5 for the first five items. The thresholds (b1-b4) and slope parameters for the dominant ability (a1) were the same as for the 2P-GR model (see Table 2). The values of a2were 0.5 for Items 1 to 5 and 0.0 for other items. Ability parameters for two dimensions were randomly selected from a bivariate normal distribution (0, 1) with a correlation of 0.3. A low level of correlation was used because the second dimension was assumed to be either a nuisance dimension or construct-irrelevant dimension. The responses were simu- lated based on the multidimensional GR model, and the three methods were used to compare the fit of the unidimensional 2P-GR model and the 2D complex GR model.

In Condition 4, responses to a test consisting of a testlet were generated under the modified GR model for testlets. As shown in Table 2, item parameters for this testlet GR model were the same as that for the 2P-GR model, except one testlet was simu- lated (Items 6, 7, and 8). Ability parameters were randomly selected from N(0, 1), and the testlet effect gjd(i) was randomly selected from a N(0, s2d(i)) for the items in the testlet. The variance of the testlet effect s2d(i) was specified at 1.0 reflecting a moderate degree of dependence among the testlet items (Bradlow, Wainer, & Wang, 1999; Li et al., 2006; Wang et al., 2002).

For each condition, 20 data sets of 2,000 item responses were generated using SAS based on the Mg model described in Table 1. Using the specified model para- meters (i.e., item and ability), the probabilities of each simulated examinee responding to each item response category were calculated, and these probabilities were used to obtain simulated discrete item responses (see Zhu, 2009, for com- plete results). For each of the generated data sets within each condition, different analysis models (Ma) were estimated, and the three model comparison methods (DIC, CPO, and PPMC) were then used to compare those models. A preferred model was selected based on each method. The effectiveness of these three meth- ods was measured by the number of times the Mg was selected as the preferred model across the 20 replications.

In addition to test-level model comparisons within replications, item-level results for both CPO and the discrepancy measures used with the PPMC method were exam- ined. While test-level results are useful in the model comparison process, item-level results may provide diagnostic information regarding the specific misfitting items, which in turn may suggest alternative models.

Bayesian Estimation of Different Models

All the GR models in this study were estimated using MCMC methods and WinBUGS 1.4 (Spiegelhalter et al., 2003). For the standard 2P-GR model, the

(12)

following priors were specified: ui;Normal(0, 1) for Person i, and aj;Log normal(0, 1), bj1;Normal(0, 0:25), and bj(k + 1);Normal(0, 0:25)I (bjk) for Item j, where the notation I (bjk) indicates that the threshold bj(k + 1)was sampled to be larger than bjk as required under the GR model. For the 1P-GR model, all priors were the same as for the 2P-GR model except that one slope was estimated. For the RS model, the priors were ui;Normal(0, 1) for Person i, aj;Log normal(0, 1) and bj;Normal(0, 0:25) for Item j, and c1;Normal(0, 0:25) and c(k + 1);Normal (0, 0:25)I (ck) with a constraint ofP

k

ck= 0. It should be noted that, consistent with WinBUGS, precision parameters that are the inverse of variance parameters were used in these prior distribution specifications. These prior specifications were similar to those used in previous research (e.g., Kang et al., 2009).

For the 2D simple-GR model, the prior distributions for item parameters were specified as for the 2P-GR model. The abilities on two dimensions were assigned multivariate normal priors, with means of 0 and variances of 1, and the correlation between two abilities was assigned a normal prior with mean equal to the true corre- lation of 0.6 and variance of 0.25. This approach was used to address the metric inde- terminacy problem in a manner similar to that used by Yao and Boughton (2007).

For estimating the 2D complex-GR model, it is important to solve both metric indeterminacy and rotational indeterminacy. As for 2D simple-GR model, the metric indeterminacy problem was addressed by assigning the abilities on two dimension means of 0 and variances of 1. To solve the rotational indeterminacy issue, the two ability axes were constrained to be orthogonal, and the slope parameters for the nui- sance dimension were fixed at 0 for the last 10 items (items measuring the dominant dimension only). Thus, any relationship between the dimensions was derived from the items that were common to both dimensions rather than from a specified correla- tion between dimensions (see Table 2, Condition 3). Priors for abilities followed a multivariate normal distribution with means 0 and a variance–covariance matrix equal to the identity matrix. These approaches were similar to those used by Bolt and Lall (2003). Prior distributions for other item parameters were the same as defined for the 2P-GR model. It should be noted that although use of an orthogonal solution affects the evaluation of item parameter recovery, model-based expectations (i.e., response category probabilities) are invariant to rotations (e.g., orthogonal vs. obli- que). Thus, the solution for rotational indeterminacy should not affect results evaluat- ing model fit and model comparisons.

For the testlet GR model, prior distributions for the item parameters and the exam- inees’ abilities were specified as for the GR models. The testlet effect was assigned a normal prior with mean of 0 and random variance of s2d( j). As in Bradlow et al.

(1999) and Li et al. (2006), the hyperparameter s2d( j)was specified as an inverse chi- square distribution with degrees of freedom equal to 0.5, indicating a lack of informa- tion about this parameter.

Convergence of the Markov chain for each GR model was evaluated using multi- ple diagnostics, such as history plots, autocorrelation plots, and multiple chains to ensure that target posterior distributions were being sampled (Kim & Bolt, 2007).

(13)

Based on the results (see Zhu, 2009, for details), one chain of 5,000 iterations was run for the 2P-GR, 1P-GR, RS, and testlet GR models, and the first 3,000 iterations were discarded. The remaining 2,000 iterations were thinned by taking every other iteration to reduce serial dependencies across iterations. The resulting 1,000 itera- tions formed the posterior sample and were used in estimating model parameters and conducting model comparisons. For the 2D simple-GR and 2D complex-GR models, a longer chain of 8,000 iterations was run. The first 5,000 iterations were discarded, and the remaining 3,000 iterations were thinned by taking every third iteration to obtain a posterior sample of size 1,000.

Convergence was also assessed by examining parameter recovery for each model using the root mean square error (RMSE) between the true (i.e., generating) and pos- terior estimates or mean values from the posterior distributions across the 20 replica- tions. Thus, the RMSEs for a specific model were obtained when the data-generating model was the same as the analysis model (i.e., Mg = Ma). RMSE results indicated that the parameters were recovered well (see Zhu, 2009, for complete results). For example, the average RMSE across all 15 items was 0.07 for both slope and thresh- old parameters for the 2P-GR model; for the 2D simple-GR model, the RMSE for the interdimensional correlation was 0.016; and for the testlet GR model, the RMSE for the testlet variance was 0.037. The low RMSE values indicated that a posterior sample of 1,000 was adequate for accurate recovery of item parameters for each GR model.

Computation of Model Comparison Criteria

Estimates of the DIC index for different models were directly available from WinBUGS. The computation of the CPO index was implemented by first computing the CPO at the level of an individual item response (CPOij) in WinBUGS by using,

CPOij= 1 T

XT

t = 1

1 f y ijjht

!1

, ð11Þ

where yijis the response of an examinee i on a particular Item j, T is the total number of MCMC draws after the chain has converged, and f y ijjht

is the likelihood of the observed item response yijbased on the sampled parameter values at draw t. An item- level CPO index for a model was obtained in SAS by taking the product of the values of CPOij across all examinees: CPOj= QN

i = 1

CPOij, where N is the total number of examinees. In addition, a CPO index for the overall test was computed by taking the product of the item-level CPOjacross all items: CPO = QI

j = 1

CPOj, where I is the total number of items on the test. The logarithm of the CPO index was used for comparing

(14)

models in this study. Recall that the preferred model was the model with the smaller DIC value or the larger CPO value.

The different models in each condition were also compared using the PPMC method. All six discrepancy measures discussed previously were used with PPMC for Condition 1. However, for Conditions 2 to 4, only the two pair-wise discrepancy mea- sures (global OR and Yen’s Q3index) were used since these types of measures have been found most useful for detecting multidimensionality or local dependence (e.g., Levy et al., 2009; Sinharay et al., 2006; Zhu & Stone, 2011). To compare different models using PPMC, the frequency of extreme PPP values (i.e., values \0.05 or .0.95 as discussed earlier) for the selected measures was computed for each model across items or item pairs. When Ma = Mg, it was expected that no or few extreme PPP values would be observed. When the alternative model was estimated, however, more items or item pairs with extreme PPP values would be expected. The preferred model was the model with the fewest number of items and item pairs with extreme PPP values. Note that among all the six discrepancy measures used in this study, three measures (the item score distribution, the global OR, and Yen’s Q3) and their PPP values were computed in WinBUGS, and the other three measures (the item–total score correlation, Yen’s Q1, and the pseudo-count fit statistic) were computed in SAS. The interested reader can consult (Zhu, 2009) for the SAS and WinBUGS code to compute these various measures.

Results

Condition 1 (2P-GR vs. 1P-GR vs. RS Models)

Table 3 presents the average values (Mean) for the model comparison indices as well as the frequencies (Freq) that each model was selected as the preferred model across the 20 replications. For the item-level discrepancy measures (i.e., item score distribution, item–total score correlation, Yen’s Q1, pseudo-count fit statistic), the means represent the average numbers of extreme PPP values (i.e., \0.05 or .0.95) across the 15 items and the 20 replications. For the pairwise measures (global OR and Yen’s Q3), the means represent the average number of extreme PPP values across the 105 item pairs and repli- cations. Note that these values were all rounded to the nearest integer.

As shown in this table, the mean DIC value was the smallest for the true model (i.e., 2P-GR) and the largest for the 1P-GR model. Similarly, the mean test-level CPO value was the largest for the true model and the smallest for the 1P-GR model.

Thus, both DIC and CPO results indicated that the 2P-GR model fit the data better than the RS model, which in turn fit the data better than the 1P-GR model. Moreover, the generating or true model was consistently selected as the preferred model across the 20 replications using the DIC and CPO indices.

Regarding the PPMC results in Table 3, the 2P-GR model had the lowest average values for all discrepancy measures, providing evidence that the 2P-GR model was preferred. For example, for the item-total correlation measure, no item had extreme PPP values for the true model (Mean = 0). However, across the 15 items, there were

(15)

Table3.ModelSelectionResultsAcrossItemsorItemPairs(OverallTest):Condition1 DICCPOItemscore distributionItem–total correlationYen’s Q1Pseudo-count fitstatisticGlobal ORYen’ Q ModelMeanFreqMeanFreqMeanFreqMeanFreqMeanFreqMeanFreqMeanFreqMean 2P-GRa 7396020216076200200200200206207 1P-GR754980216405002012010010079048 RS74484021619001401308013031013 Note.DIC=devianceinformationcriterion;CPO=conditionalpredictiveordinate;GR=gradedresponse;RS=ratingscalemodel;Freq=frequency;OR ratio;2P-GR=two-parameterGRmodel;1P-GR=one-parameterversionoftheGRmodel. a.Themodelwasthedata-generating(i.e.,true)model.

(16)

an average of 12 items with extreme PPP values for the 1P-GR model, and an average of 13 items with extreme PPP values for the RS model. For Yen’s Q3measure, across 105 item pairs an average of 7 pairs had extreme PPP values for the true model, but an average of 48 and 13 item pairs had extreme PPP values for the 1P-GR and RS models, respectively. In addition, all measures, except the item score distribution measure, consistently identified the true model as the preferred model over the 20 replications. Using the item score distribution measure, both the 2P-GR and 1P-GR models were identified as preferred models since no items with extreme PPP values were observed for either model. This result indicated that this measure may be inef- fective in discriminating between 2P- and 1P-GR models. Since this measure does not compare observed and expected values at different discrete values of ability, as in Yen’s Q1and the pseudo-count fit statistic, it is not sensitive to discrepancies between observed and expected values that may exist at specific ability levels. Furthermore, positive and negative discrepancies may cancel each other when the discrepancies are collapsed across ability levels as in the item score distribution statistic.

The previous comparisons focused on the selection of a preferred model for the overall test. Although overall model comparisons are useful, comparing models at the item level can provide additional diagnostic information. For example, informa- tion regarding item-level fit could be used to identify a subset of items that should be reexamined, modified, or replaced. Among the Bayesian comparison methods con- sidered, CPO and PPMC can be used to compare the models at both test- and item- levels. For Condition 1, the item-level CPO results (not shown here) indicated that the 2P-GR model was preferred for each of the 15 items. With respect to the item- level PPMC results, median PPP values for all four item-level discrepancy measures were around 0.50 for the 2P-GR model. In contrast, extreme PPP values were observed for a majority of items when the 1P-GR or RS models were estimated with all but the item score distribution discrepancy measure. The differences in the num- bers of extreme PPP values between the models indicated that the 2P-GR was preferred.

It is worthwhile to note that both the DIC and CPO indices are used to compare the relative fit of different models. Thus, they may be used to identify which model among a set of candidate models provides a better fit to the data, but they may not be used to evaluate the degree of fit in an absolute sense. When one or more models among the set of candidate models are appropriate, the better fitting model can be selected by using either DIC or CPO index. However, when the models to be com- pared are not appropriate or do not fit the data, using these indices may be mislead- ing. For example, if only the 1P-GR and RS models were compared using the DIC or CPO indices for Condition 1, the RS model would be preferred over the 1P-GR model based on the results in Table 3. However, the RS model is not really appropri- ate since the true model was the 2P-GR model.

In contrast to the DIC and CPO indices, the PPMC method can be also used to evaluate the absolute fit of different models. As shown in Table 3, all effective discre- pancy measures had a number of extreme PPP values for the 1P-GR and RS models,

(17)

indicating that neither of these models fit the simulated 2P-GR data. As expected, no or few PPP values were extreme for the estimated 2P-GR model. This indicated that the 2P model was not only preferred over the other two models, but it also fit the data.

Condition 2 (2P-GR vs. 2D Simple-GR Models)

Table 4 (see Condition 2) presents the results comparing the estimation of a unidi- mensional 2P-GR model with a 2D simple-GR model (true or generating model) for the simulated 2D simple-structure data. The table provides mean values for each model comparison index, as well as the frequencies the true model (i.e., 2D simple- GR) was preferred across the 20 replications.

The lower DIC value and the higher CPO value for the 2D simple-GR model indi- cated that the true model was preferred for the overall test. For the PPMC results, on average, only 7 out of 105 item pairs with extreme PPP values for the global OR mea- sure (or 4 for Yen’s Q3index) were observed when the true model was used to ana- lyze the data. However, when the unidimensional 2P-GR model was estimated, there were a large number of item pairs with extreme PPP values—75 and 102 pairs for the global OR and Yen’s Q3index, respectively. Thus, the PPMC results indicated that the two-dimensional model was preferred over the unidimensional GR model and also indicated that this model provided adequate model fit with respect to associations among item-pair responses. Also, it is apparent that the three methods appeared to perform equally well with regard to the frequency of selecting the two-dimensional GR model as the preferred model at the overall test level. All the indices identified the true model for each of the 20 replications.

As for Condition 1, item-level results were examined to diagnose any particular source of model misfit. The item-level CPO results are presented in Table 5 (see Table 4. Model Selection Results Across Items (Overall Test): Conditions 2 to 4

DIC CPO

PPMC (global OR)

PPMC (Yen’s Q3)

Condition Model Mean Freq Mean Freq Mean Freq Mean Freq

2 2D simple-GRa 75434 20 216430 20 7 20 4 20

2P-GR 78854 0 217143 0 75 0 102 0

3 2D complex-GRa 71905 20 215645 20 4 18 4 20

2P-GR 72093 0 215670 0 8 2 15 0

4 Testlet GRa 74170 20 216145 20 6 18 5 20

2P-GR 74924 0 216287 0 10 2 21 0

Note. DIC = deviance information criterion; CPO = conditional predictive ordinate; PPMC = posterior predictive model checking; OR = odds ratio; GR = graded response; 2P = two parameter; 2D = two- dimensional; Freq = frequency.

a. The model was the data-generating (i.e., true) model for each condition.

(18)

Condition 2), including the mean CPO values for each estimated model as well as the frequencies the true model was selected as the preferred model across the 20 replications. For each item, the average CPO value (across the 20 replications) for the two-dimensional model was larger than the value for the unidimensional model.

This indicated that the two-dimensional model, which was preferred for the overall test, was also preferred for each item. The large differences for the item-level CPO values between the two models provided strong evidence that the two-dimensional model fit each item significantly better than the unidimensional model. In addition, the two-dimensional model was chosen as the preferred model for each of the 20 replications.

With regard to the PPMC results, pie plots were used rather than tables of PPP values to examine and explore patterns in item-level results. As an example, the pie plots in Figure 1 (Condition 2) provide median PPP values for Yen’s Q3measure when items responses were simulated under a 2D simple-GR model and two compet- ing models were estimated: the 2P-GR model and the 2D simple-GR model. In each plot, there is one pie for each item pair, and the proportion of a circle that is filled corresponds to the magnitude of median PPP value across the 20 replications. An empty pie reflects a PPP-value of 0.0, whereas a fully filled pie reflects a PPP value of 1.0. As shown from the pie plots, when a unidimensional GR model was estimated (upper left plot), median PPP values for all item pairs were extreme and the items fell into two clusters—Items 1 to 8 in one, and Items 9 to 15 in another. This pattern Table 5. Item-Level CPO Index Results for Each Item: Conditions 2 to 4

Condition 2 Condition 3 Condition 4

Item

2D

Simple-GRa 2P-GR Freq

2D

Complex-GRa 2P-GR Freq Testlet

GRa 2P-GR Freq

1 21274 21291 20 21201 21205 19 21269 21269 15

2 21162 21209 20 21238 21242 19 21293 21294 15

3 2998 21077 20 21209 21212 19 21266 21267 16

4 21214 21231 20 21076 21081 20 21119 21120 14

5 21001 21044 20 21071 21076 20 21204 21204 16

6 2999 21075 20 21114 21115 14 21121 21162 20

7 21301 21319 20 21140 21139 10 21151 21193 20

8 21133 21180 20 21113 21114 13 21124 21166 20

9 2848 2923 20 2983 2984 17 2986 2987 18

10 21209 21227 20 2979 2978 15 2986 2987 16

11 21133 21184 20 2961 2959 14 2968 2970 18

12 21032 21114 20 2988 2989 14 2991 2993 16

13 21277 21297 20 2958 2959 14 2962 2965 19

14 21002 21049 20 2807 2808 18 2811 2814 20

15 2845 2923 20 2806 2808 12 2814 2816 20

Note. CPO = conditional predictive ordinate; GR = graded response; 2P = two-parameter; 2D = two- dimensional; Freq = frequency.

a. The model was the data-generating (i.e., true) model for each condition.

(19)

indicated that a unidimensional model did not fit the data and a two-dimensional model might be considered more appropriate to model the data. As expected, when a two-dimensional model (i.e., true model) was estimated (upper right plot), median PPP values were around 0.5 for all pairs, suggesting this model fit the data. Pie plots based on the global OR measure exhibited a similar pattern.

Figure 1. Display of median PPP values for item pairs for Yen’s Q3for Conditions 2 to 4 Note. PPP = posterior predictive p; 2P = two-parameter; GR = graded response; 2D = two-dimensional.

(20)

Condition 3 (2P-GR vs. 2D Complex-GR Models)

In Condition 3, the generated data were 2D with complex structure. Items 1 to 5 mea- sured a dominant dimension as well as a nuisance dimension, and Items 6 to 15 only measured the dominant dimension (see Table 2). A unidimensional 2P-GR model and a two-dimensional GR model were estimated and compared.

The model comparison results for the overall test are presented in Table 4 (see Condition 3). The lower DIC value and higher CPO value for the 2D complex-GR model (true model) indicated that this multidimensional model was preferred over the unidimensional GR model. For the PPMC indices, 4 out of 105 item pairs demon- strated extreme PPP values for both pair-wise measures when the two-dimensional model was estimated. In contrast, when the unidimensional GR model was estimated, more item pairs demonstrated extreme PPP values—8 and 15 pairs for the global OR measure and Yen’s Q3index, respectively.

The DIC index, the CPO index, and the PPMC method using Yen’s Q3measure appeared to perform equally well with regard to the frequency of selecting the two- dimensional GR model as the preferred model. Based on the model comparisons for these indices, the true model was selected as the preferred model every time (20) over the 20 replications. However, when the global OR measure was used, the unidi- mensional GR model was selected incorrectly as the preferred model for 2 of the 20 replications. This example again illustrates that the choice of the discrepancy mea- sure may affect the performance of the PPMC application in comparing different models.

The item-level CPO results (means and frequencies) for this condition are pre- sented in Table 5 (see Condition 3). For Items 1 to 5, which measured both the domi- nant and nuisance dimensions, the item-level CPO identified the generating two-dimensional model as the preferred model for 19 to 20 times over the 20 replica- tions. However, for the items that only measured the dominant dimension (Items 6-15), the two-dimensional model was identified as the preferred model with lower frequencies (10 to 18 over the 20 replications). This would be expected since the uni- dimensional GR model should be appropriate for these items.

It may be worthy to note that the differences between the item-level CPO values were quite small for the two models, particularly for the items reflecting one dimen- sion (Items 6-15). Spiegelhalter et al. (2003) discussed that any difference in DIC less than 5 units for two models may not indicate sufficient evidence to favor one model over another. Though there are no discussed guidelines available for CPO, the item-level CPO results for this condition may indicate that a difference of less than 3 units may not provide sufficient evidence supporting one model over another.

However, the amount of difference in CPO necessary to suggest a significant differ- ence between models needs further investigation.

The pie plots in Figure 1 (Condition 3) display median PPP values for Yen’s Q3

measure for this condition. All the PPP values were around 0.5 when the true 2D complex-GR model was estimated (middle right plot) and therefore provided evi- dence of model fit. In contrast, when the unidimensional GR model was estimated

(21)

(middle left plot), all the PPP values were extreme for the item pairs involving the first five items, but around 0.5 for the other item pairs. This pattern indicated that the unidimensional GR model was not appropriate for Items 1 to 5, but was appropriate for Items 6 to 15. Additionally, the close to 0 PPP values for the item pairs among Items 1 to 5 indicated that the realized correlations among these five items were con- sistently larger than the predicted correlations under the unidimensional GR model.

Condition 4 (2P-GR vs. Testlet GR Models)

In Condition 4, the responses to a test with locally dependent items (testlet) were simulated. Items 6, 7, and 8 were designed to reflect a testlet with moderate depen- dence among the items. Table 4 (see Condition 4) presents the results for comparing a GR model for testlets with a unidimensional GR model. The lower DIC value, the higher CPO value, and the fewer item pairs with extreme PPP values (i.e., values

\0.05 or .0.95) for the testlet GR model all indicated that this complex model fit the overall test better than the unidimensional GR model. Furthermore, the DIC index, the CPO index, and the PPMC method using Yen’s Q3measure appeared to perform equally well. Based on these measures, the testlet GR model was selected as the preferred model every time (20) over the 20 replications. However, when the glo- bal OR measure was used with PPMC, the testlet GR model was chosen as the pre- ferred model for 18 of the 20 replications. As for Condition 3, use of Yen’s Q3as a discrepancy measure appeared to be slightly more effective than use of the global OR measure.

Table 5 (Condition 4) presents results for the item-level model comparisons. For the items in the testlet (Items 6, 7, and 8), the mean CPO values for the testlet GR model were much larger (~42 units) than the values for the unidimensional model, and the testlet GR model was chosen as the preferred model for all the 20 replica- tions. For the simulated locally independent items, the mean CPO values for the two models were very close. The mean CPO values for the testlet GR model were less than 3 units larger than the values for the unidimensional GR model. Based on the CPO results for Condition 3, less than 3 units in difference may indicate insufficient evidence for selecting the complex testlet GR model as the preferred model.

Nonetheless, if these small differences were considered evidence in favor of the test- let model, the testlet GR model would be selected as the preferred model for these items for 14 to 20 replications. Thus, more research is needed to determine what dif- ferences in CPO values should be used to select one model over another.

The two pie plots at the bottom of Figure 1 (see Condition 4) illustrate the median PPP values for Yen’s Q3 measure for the testlet and unidimensional GR models, respectively. As can be observed, when the testlet GR model was estimated (bottom right plot), all PPP values were around 0.5, suggesting model-data-fit for the testlet GR model. In contrast, when the unidimensional GR model was estimated (bottom left plot), all PPP values were extreme for item pairs reflecting local dependence (Items 6, 7, and 8), but around 0.5 for pairs among simulated locally independent

(22)

items. Additionally, PPP values of nearly 0 for item pairs comprising the testlet indi- cated that the realized correlations among these items were consistently larger than predicted correlations under the unidimensional GR model. These results indicated that the unidimensional GR model was not appropriate for Items 6, 7, and 8, but was appropriate for the other items.

It should be noted that results for the 2D complex-structure model (Condition 3) and the testlet model (Condition 4) were very similar. As noted by a reviewer, this was because of the similarity between the 2D complex-structure model used herein and the testlet model. In both models all items measure a single primary dimension while a subset of items in both the 2D complex-structure and testlets models are related further because of another source of shared variance.

Discussion

The purpose of this study was to evaluate the relative performance of three Bayesian model comparison methods (DIC, CPO, and PPMC) in selecting a preferred model among a set of alternative GR models for performance assessment applications. The alternative models considered in this study reflected sources of potential misfit and reflected more complex IRT models that might be theoretically more appropriate for some types of performance assessment data but require Bayesian estimation methods.

For the conditions examined in this study, the results of this study indicated that these three methods appeared to be equally accurate in selecting the true model as the pre- ferred model when considering all items simultaneously (test level). However, CPO and PPMC were found to be more informative than DIC since information about model fit was also available at the item level.

Consistent with previous studies (Li et al., 2006; Sinharay, 2005), the PPMC approach was found to be effective for performing model comparisons in the perfor- mance assessment context. Moreover, an advantage of PPMC applications is that they can be used to compare both the relative and absolute fit of different models. In con- trast, the DIC and CPO model comparison indices only consider the relative fit of dif- ferent models. If the true model is among the candidate models, DIC or CPO can be used to select the preferred model that should also fit the data. However, in practice, it is not known whether the true model is included in the set of candidate models.

Thus, it is important that the assessment of absolute model fit be combined with model comparison methods. Ideally, model selection (e.g., DIC/CPO) and model fit (PPMC) methods should both be used for real testing applications. As a result of their simplicity, DIC or CPO could be first used to choose a preferred model from a set of candidate models. Using multiple discrepancy measures, the PPMC method could then be applied to evaluate the fit of the selected model at the test and item levels. In this sense, these methods are not competing in nature, but are instead complementary.

Contrary to previous studies, DIC, CPO, and PPMC appeared to perform equally well in this study, whereas their performance has been found in previous research to depend on specific conditions as well as models compared. This finding may be due

參考文獻

相關文件

• an algorithm for robust fitting of models in the presence of many data outliers. • Compare to

For machine learning applications, no need to accurately solve the optimization problem Because some optimal α i = 0, decomposition methods may not need to update all the

Hence, Buddhism has become even more adapted to contemporary society for its realistic and universal values, as well as making outstanding contributions to the contemporary

Writing texts to convey information, ideas, personal experiences and opinions on familiar topics with elaboration. Writing texts to convey information, ideas, personal

Language arts materials which deal with universal issues can be used as resources for simulating activities to enable students to develop positive values, think from

The Moral, Civic and National Education (MCNE) Section 2 is responsible for curriculum development related to values education as well as organising professional development

The left panel shows boxplots showing the 100 posterior predictive p values (PPP-values) for each observed raw score across the 100 simulated data sets generated from

We define Flat Direction Hybrid Inflation (FDHI) models as those motivated by the properties of moduli fields or flat directions of the standard model. For moduli fields with no