Skip to content
Publicly Available Published by De Gruyter April 6, 2019

Simulation Extrapolation Method for Cox Regression Model with a Mixture of Berkson and Classical Errors in the Covariates using Calibration Data

  • Jean de Dieu Tapsoba EMAIL logo , Edward C. Chao and Ching-Yun Wang

Abstract

Many biomedical or epidemiological studies often aim to assess the association between the time to an event of interest and some covariates under the Cox proportional hazards model. However, a problem is that the covariate data routinely involve measurement error, which may be of classical type, Berkson type or a combination of both types. The issue of Cox regression with error-prone covariates has been well-discussed in the statistical literature, which has focused mainly on classical error so far. This paper considers Cox regression analysis when some covariates are possibly contaminated with a mixture of Berkson and classical errors. We propose a simulation extrapolation-based method to address this problem when two replicates of the mismeasured covariates are available along with calibration data for some subjects in a subsample only. The proposed method places no assumption on the mixture percentage. Its finite-sample performance is assessed through a simulation study. It is applied to the analysis of data from an AIDS clinical trial study.

1 Introduction

Many epidemiological or biomedical studies often aim to investigate the association between the time to an event of interest and some covariates. The examination of the association is usually performed under the popular Cox proportional hazards model [1] that stipulates proportional covariate effects on the hazard function for the time-to-event. A commonly encountered problem in such an examination is that some covariates potentially involve measurement error. This may be due to imprecision of the data measuring instrument, high costs to obtain precise measurements or the nature of the covariates themselves. For example, CD4 counts in HIV studies [2], dietary protein intake in nutrition studies and radiation dose in radiation epidemiology studies [3] are well-known to be subject to measurement error. Naively omitting the measurement error and simply using the error-affected covariate data in the survival analysis could lead to invalid inference [4].

There are two different types of measurement errors, namely the classical type and the Berkson type [3]. The main difference between these two types of errors lies in their relationship with the unobserved covariates. The classical error is assumed to be independent of the unobserved covariates whereas the Berkson error is positively correlated with the unobserved covariates. Self-reported data, collected through questionnaire typically involve recall error that is usually assumed to be of the classical type [5]. Also, grouped-exposure data obtained by assigning the same value of the unobserved exposure variable (e. g. air pollution or radiation dose) to individuals sharing some characteristics are often assumed to be contaminated with Berkson error [6]. In many other situations, the covariate data may be affected by a mixture of errors of both classical and Berkson types. For example, DS02 radiation dose estimates for atomic-bomb survivors, who were followed up by the Radiation Effects Research Foundation (RERF) are contaminated with both errors [7]. The classical error is thought to originate from the survivors’ individual recollections of location and shielding at the time of the bombings and the Berkson error is believed to be associated with averaging radiation dose estimates among survivors with common location and shielding type. Taken individually and if omitted, measurement error of each of these two types could affect the estimation of the survival model parameters. The impact of the classical error in the covariates of a Cox regression model has been well-documented and includes the attenuation of the covariate effect [8, 9]. Moreover, a few references have reported that the Berkson error may also have a similar effect [6, 10]. However, the effect of a mixture of classical and Berkson errors in the covariates of a proportional hazards model still remains to be well investigated. The presence of errors of both types in the covariates of the regression model is expected to worsen the attenuation effect as will become clear in the simulation study section. It is therefore important to simultaneously account for both types of errors when there are present in the covariates of a Cox proportional hazards model.

Several methods have been developed to deal with covariate measurement error in a Cox proportional hazards model and extensive attention has been given to the classical error. Popular methods adjusting for classical error in survival analysis are the regression calibration (RC) method [11, 12], simulation extrapolation (SIMEX) method [3, 13, 14], parametric corrected score [15], non-parametric corrected score [8, 16], conditional score [17, 18], maximum likelihood method [19, 20] and non-parametric maximum likelihood method [21]. On the other hand, methods addressing the issue of Berkson error in the covariates in Cox regression model are very scarce and this problem still has to be well-discussed in the literature. Among the very limited number of references, [22] discussed an extension to a Berkson error of a corrected profile likelihood method for Cox model with classical error in the covariates. It is important to note that methods that simply deal with classical error or Berkson error exclusively may not work well when errors of both types are involved in the covariates. Moreover, the existing methods accounting for a mixture of errors of both types are developed for generalized linear models [23, 24] and may not be directly applicable in the context of survival analysis. Therefore, new statistical methods that jointly adjust for both classical and Berkson errors need to be developed for a reliable inference in this situation.

In this paper, we address the problem of parameter estimation in Cox regression analysis when some covariates are subject to a combination of both Berkson and classical errors. We are not aware of an existing reference in the statistical literature that has properly dealt with this problem. We assume that two replicates of the error-prone covariates are available for all individuals. Also, we assume that data on an instrumental variable for the mis-measured covariate are available only for some individuals in a subset of the study cohort that is here referred to as calibration subsample. For illustration, we consider the AIDS Clinical Trials Group (ACTG) 175 study, a randomized double-blind clinical trial that was conducted to compare four antiretroviral treatments with either a single nucleoside or two nucleosides in (HIV-1) infected adult participants [25]. In addition to sociodemographic data, the study also collected data on survival time and virologic variables such as CD4 cell count and CD8 cell count. An important interest is in assessing the effect of the true baseline CD4 cell count on time to AIDS disease development or death under a Cox proportional hazards model adjusting for the treatment effect. A complication related to the use of standard survival analysis methods for this assessment is the possible contamination of the CD4 cell count data with measurement error, which has been assumed to be of the classical type in previous applications [8, 16]. Here we allow the measurement error to incorporate features of both classical and Berkson error types. We use a mixture of classical and Berkson errors model, which is more general than the classical model. Moreover, we treat the two CD4 cell counts taken before the start of the treatment and within three weeks of randomization as replicates. Also, the CD8 cell count measured at randomization is likely correlated with the baseline CD4 cell count and may be used as instrumental variable for the true baseline CD4 count. There were 2441 participants with two replicates for the true baseline CD4 cell count and only 1459 of them had CD8 cell count measured at randomization. We propose a SIMEX-based new approach to adjusting for both classical error and Berkson error under these settings. It places no assumption on the mixture proportion of the error variances.

In the next section, we provide the model formulations for the survival data and the mixture of classical and Berkson errors. In Section 3, we present the naive and RC methods and develop the SIMEX method for this problem in Section 4. We present a simulation study in Section 5 and apply the developed method to the data from the ACTG 175 study in Section 6. Finally, we conclude with some discussions in Section 7.

2 Model formulations

Suppose that there are n study subjects, which are followed-up over a study period of length τ. For subject i, let Ti0 denote the unobserved survival time, which is assumed to be right-censored and Ci be the censoring time. The observed survival data are (Ti,δi), where Ti=min(Ti0,Ci) is the observed failure time and δi=I(Ti0Ci) is the non-censoring indicator, i = 1,…, n. Also, let Xi denote a vector of time-independent covariates that are subject to measurement error, and Zi represent a vector of other time-independent covariates that can be measured accurately. It is assumed that replicates Wij, j=1,,ki(2) of Xi are available for all subjects and data on an instrumental variable Qi for Xi are available only for some subjects in a calibration subsample. Letting ηi indicate whether Qi is observed (ηi=1) or not (ηi=0) and Wi=(Wi1,,Wiki), the observed data are (Ti,δi,Wi,Zi,ηiQi), i = 1,…, n, which are assumed to be independent and identically distributed random vectors. In the ACTG 175 data example, δi is the indicator of the event of interest, which is AIDS disease progression or death during the study follow-up and Ti is the time to the event if δi=1 or end of follow-up (due to lost to follow up or end of the study observation period) if δi=0. Also, Xi is a single covariate representing the log-transformation of baseline CD4 counts, Zi is a vector of indicator variables of randomized treatment groups, Wij (j = 1,2) are the log-transformation of two CD4 count measurements taken before treatment initiation and Qi is the log-transformation of baseline CD8.

We consider the following mixture of Berkson and classical errors model for the relation between Wij and Xi:

(1)Xi=Li+Ub,i,Wij=Li+Uc,ij,

where Li is a latent variable with mean μl and variance Σl, and Ub,i is a zero-mean error with variance Σb. Also, Uc,ij, j=1,,ki, are independent and identically normally distributed with mean 0 and variance Σc. Moreover, Uc,ij is independent of Ub,i, j=1,,ki, (Uc,i1,,Uc,iki) and Ub,i are independent of (Ti0,Ci,Li,Zi), i = 1,…, n. Model (1), which was also studied by [23, 26] and [27] in different contexts encompasses features of both classical and Berkson error models. It reduces to a Berkson measurement error model when Σc=0 and a classical error model when Σb=0. Furthermore, Qi is assumed to be associated with Xi according to the following model:

(2)Qi=α0+α1Xi+ϵi,

where α=(α0,α1) is a vector of unknown parameters and ϵi is a zero mean normally distributed random variable with variance σϵ2 and independent of (Ti0,Ci,Li,Zi,Ub,i,Uc,i1,,Uc,iki), i = 1,…, n. The time-to-event is associated with Xi and Zi via the Cox proportional hazards model, which is a commonly used tool for survival data analysis. Assuming that Ti0 is conditionally independent of Ci, Wi and Qi given Xi and Zi, the model formulates the hazard function as follows.

(3)λ(u|Xi,Zi,Wi,Qi)=λ0(u)exp(βxXi+βzZi),

where λ0(.) is an unspecified baseline hazards function and β=(βx,βz) is a vector of unknown parameters. Our interest lies in the estimation of β based on the observed data. In what follows, we let Yi(u)I(Tiu) denote the at-risk process and define the counting process Ni(u)I(Tiu,δi=1), where I(.) is the indicator function.

3 Naive and RC methods

The naive and RC estimations are two simple imputation-based methods that could be used to estimate the parameter β although, they might not handle correctly the measurement errors, especially when the error magnitudes are not small.

3.1 Naive method

The estimation of β would be based on the usual Cox partial likelihood score if Xi were observed. In such a situation, the partial likelihood estimator for β would solve Un(β,X,Z)=0, where

(4)Un(β,X,Z)=1ni=1n0τXiZiS1(u,β,X,Z)S0(u,β,X,Z)dNi(u),

and Sk(u,β,X,Z)=n1i=1nYi(u){(Xi,Zi)}kexp(βxXi+βzZi), k = 0,1. Since Xi cannot be observed in our context, a naive estimator of β is obtained by replacing Xi by Wˉi=ki1j=1kiWij in (4) and solving Un(β,Wˉ,Z)=0. It has been well documented in the literature pertaining to Cox regression with covariates affected by classical measurement error that the naive estimator is often downwardly biased [9, 28]. Also, it has been reported in [10] that the naive approach could lead to inconsistent estimation of the baseline hazard as well as incorrect tests for the proportional hazards assumption when the error is of Berkson type. Hence, it is clear that ignoring the mixture of Berkson and classical errors in our context will also result in a biased estimation and expectedly a more severe attenuation effect compared with the situation when the error is simply of Berkson or classical type.

3.2 Regression calibration method

The RC method is a simple and well-known method that can be applied to reduce the bias caused by the naive estimation. It has been extensively studied by [11, 29] and [12] in the context of classical error in the covariates of a Cox model. One implementation (RC1) of RC for our problem is to use E(Xi|Wˉi) in place of Xi in (4). Another RC approach (RC2) that makes use of the calibration data consists to replace Xi in (4) with E(Xi|Wˉi,ηiQi) which is E(Xi|Wˉi,Qi) if i is in the calibration sample (ηi=1) and E(Xi|Wˉi) otherwise (ηi=0), i = 1,…, n. In general, RC methods perform well in linear regression when the error is classical, Berkson or a mixture of Berkson and classical errors. They may also work satisfactorily in nonlinear regression when the measurement error or the covariate effect is small. However, they do not lead to consistent estimation for generalized linear models when the link function is different from the identity function and the error is of classical type [3], Chapter 6], Berkson type [30] or a combination of both classical and Berkson types [24]. In Cox regression with covariates affected by a classical error, [11, 28] noted that the bias of RC estimator is large when the covariate effect or the error variance is large. As will be seen in our simulation study, RC may not perform well in Cox regression with a mixture of classical and Berkson errors in the covariates. A conditional expectation of the form E{h(Xi)|Wˉi,ηiQi} can be evaluated based on the conditional density function of (Li,Ub,i) given (Wˉi,ηiQi) as follows.

E{h(Xi)|Wˉi,ηiQi}=lubh(Xi=l+ub)L(Qi|Xi=l+ub)ηiL(Wˉi|Li=l)L(Li=l)L(Ub,i=ub)dldublubL(Qi|Xi=l+ub)ηiL(Wˉi|Li=l)L(Li=l)L(Ub,i=b)dldub,

where L(.) denotes density function. Integrals that are involved in the conditional expectations can be computed by means of the Gauss-Hermite quadrature.

4 SIMEX approach

The SIMEX method is a two-step procedure consisting of a simulation step and a subsequent extrapolation step. It was originally developed by [13] to deal with classical error in the covariates of a linear or nonlinear regression model and later studied by [3, 14, 31] and [32]. In addition, applications of this method to a mixture of errors of both types in the covariates of a generalized linear models are discussed in [3] and [24]. Here we provide an extension of the method to the Cox proportional hazards model when some covariates are subject to both classical and Berkson errors. We describe the implementation of the two steps of the procedure for our problem in what follows. The simulation step requires an estimating function that would yield a consistent estimator of β if there were no classical error (Σc=0) and Xi were substituted with Wˉi in that estimating function. The partial likelihood score in (4) would have been a suitable choice for this purpose if the error were simply of classical type. However, its use in our context is complicated by the presence of Berkson-type error in addition to classical error. We instead use estimation equations based on a maximum likelihood function.

It should be noted that if Xi were observed and λ0(.) known, the maximum likelihood estimator of β would solve the following equation:

i=1n0τXiZidNi(u)Yi(u)XiZiexp(βxXi+βzZi)λ0(u)du=0.

Also, it is useful to note that Li=Wˉi if the error were simply of Berkson type (Σc=0) under Model (1). Assuming that λ0(.) and the distribution of (Li,Zi) do not involve β, a consistent estimator for β in the presence of simply Berkson error (Σc=0) would be obtained by solving

(5)i=1n0τEXiZidNi(u)Yi(u)XiZiexp(βxXi+βzZi)λ0(u)du|Ni,Yi,Li,Zi=0.

Also, a Breslow type estimator for the baseline hazards function λ0(.) could be expressed as λˆ0(u)=i=1ndNi(u)/i=1nYi(u)Eexp(βxXi+βzZi)|Ni,Yi,Li,Zi. However, Li cannot be observed due to the presence of classical error (in addition to the Berkson error) and simply replacing Li with Wˉi in (5) will potentially lead to a biased estimator for β in our situation. In the simulation step of the proposed SIMEX-based method, a number of R (with R>1) naive estimates of β are obtained by solving (5), in which Li is imputed with Wζ,r,i=Wˉi+ζUr,i, where ζ is a non-negative scalar and Ur,iN(0,ki1Σc), i = 1,…, n, r=1,,R. Let βˆr(ζj) denote the naive estimator for  β using Wζj,r,i in place of Li in (5) and denote  βˆ(ζj)=R1r=1Rβˆr(ζj) the pseudo-estimate of β using ζj, where 0=ζ0<ζ1<<ζJ, and J>1 is the number of pseudo-estimates. The extrapolation step consists in fitting a regression model of each component of  βˆ(ζj) on the ζj’s, j = 1,…, J, using the ordinary least squares estimation method. The SIMEX estimate of each component of β is then obtained by extrapolating back to the case when ζ=1, which represents the situation of no classical error. Common extrapolation functions are polynomial in ζ of the form P(ζ,a)=a0+a1ζ++aqζq, where q1 is the degree of the polynomial. For a particular component of β, an estimator aˆ=(aˆ0,,aˆq) for the parameter a=(a0,,aq) is obtained via the regression of the corresponding components of the βˆ(ζj)s on the ζj’s, j = 1,…, J. For example, letting βˆx(ζj) represent the component of βˆ(ζj) that corresponds to βx, and βˆx(ζ)={βˆx(ζ1),,βˆx(ζJ)}, an estimator for a would be aˆ={D(ζ)D(ζ)}1D(ζ)βˆx(ζ), where D(ζ) is the J × q-matrix whose jth row is (1,ζj,,ζjq). The SIMEX estimate of that particular component of β is then obtained as P(1,aˆ)=aˆ0aˆ1++aˆq(1)q. Here R, J and q are assumed known. The standard errors can be obtained based on the bootstraps method. The SIMEX method leads to a consistent estimator for β if the true expression of the extrapolation function is used. In practice, approximations to the form of the extrapolation function are necessary as its closed form expression is unknown. It is important to note that the performance of the method could be affected by the choice of the extrapolation function.

A conditional expectation of the form E{h(Xi)|Ni,Yi,Li,Zi} involved in (5) is given as follows.

E{h(Xi)|Ni,Yi,Li,Zi}=ubh(Xi=Li+ub)L(Ni,Yi|Xi=Li+ub,Zi)L(Li,Zi)L(Ub,i=ub)dububL(Ni,Yi|Xi=Li+ub,Zi)L(Li,Zi)L(Ub,i=ub)dub,

where

L(Ni,Yi|Xi,Zi)=u[0,τ]{Yi(u)λ0(u)}dNi(u)expdNi(u)(βxXi+βzZi)eβxXi+βzZi0τYi(u)λ0(u)du.

The nuisance parameters μ, α, Σl, Σb, Σc and σϵ2 are needed for the evaluation of the conditional expectations involved in the implementation of the RC and SIMEX methods. They may be estimated from the data if unknown. For the single covariate case for example, the vector of nuisance parameters is ν=(μ,α0,α1,σl2,σb2,σc2,σϵ2), where σl2=Σl, σb2=Σb and σc2=Σc. All the components of ν can be identified using the observations (Wi1,,Wiki,ηiQi,Ti), i = 1,…, n. Moreover, a consistent estimator of ν can be obtained based on the method of moments, which leads to the following estimating equations:

i=1nj=1ki(Wijμl)=0;i=1nj=1ki1j=j+1ki{(Wijμl)(Wijμl)σl2}=0;i=1nj=1ki1j=j+1ki{(WijWij)22σc2}=0;i=1nηi(Qiα0α1μl)=0;i=1nj=1kiηi{(Qiα0α1μl)(Wijμl)α1σl2}=0;i=1nηi{(Qiα0α1μl)2α12(σl2+σb2)σϵ2}=0;i=1nj=1kiηi(TiTˉ){(Qiα0α1μl)σl2α1(Wijμl)(σl2+σb2)}=0,

where Tˉ=n1i=1nTi.

5 Simulation study

We performed a simulation study to examine the finite-sample performance of the proposed SIMEX method and compare it to the naive and RC methods for the case of a single covariate with a mixture of Berkson and classical measurement errors. The latent variable Li was generated from N(μl,σl2) with μl=0 and σl2=1. The unobserved covariate Xi was simulated as Xi=Li+Ub,i, where the Berkson error, Ub,i followed a zero-mean normal distribution with variance σb2. Two replicates, Wij, j=1,2=ki, were generated from the classical error model Wij=Li+Uc,ij with Uc,ij generated from N(0,σc2). We set σb2=0.3 or 0.5, and σc2=0.3 or 0.5 to study the separate and combined effects of the Berkson error and classical error on the performances of the various estimators. The survival times were simulated following the Cox proportional hazards model with unit baseline hazard λ(u|Xi,Zi)=exp(βxXi+βzZi), where Zi followed a Bernoulli trial with a success probability of 0.5, βx=log(2) and βz=0. A common censoring was applied to subjects inducing a censoring rate of 50%. The instrumental variable was simulated following the model Qi=α0+α1Xi+ϵi, where ϵi was normal with mean 0 and variance σϵ2. We took α0=1, α1=0.8 and ϵi=0.5. The variable ηi, indicating whether Qi is available or not was generated from the Bernoulli distribution with probability of success P(ηi=1)=c. We set c = 0.5 or 0.7 to explore how the proportion of the calibration data influences the results.

A total of 500 Monte Carlo samples of size n = 500 or 1000 were generated in the simulations. We estimated the parameter of interest β=(βx,βz) based on the naive method (Naive), which replaces Xi by Wˉi=(Wi1+Wi2)/2, RC method (RC1) that uses E(Xi|Wˉi) in place of Xi, RC method (RC2) that substitutes Xi by E(Xi|Wˉi,Qi) if ηi=1 and E(Xi|Wˉi) otherwise and the SIMEX method with quadratic polynomial (SIM1) or polynomial of degree 4 (SIM2) as extrapolation function in extrapolation step. For the SIMEX method, we created R=50 additional data sets of measurement error at each point ζ0,0.25,0.5,0.75,1 in the simulation step. In addition, the vector of nuisance parameters ν=(μl,α0,α1,σl2,σb2,σc2,σϵ2) was estimated by the method of moments. Integrals involved in the evaluation of condition expectations for the implementation of RC and SIMEX methods were computed using Gauss-Hermite integration techniques with 20 quadrature points. The estimators were evaluated with regard to their biases (Bias), sample standard deviation of the estimates (SD), average of the estimated standard erros (ASE) of the estimators, mean squared errors (MSE=Bias2+SD2) and coverage probabilities (CP) of their 95% Wald confidence intervals. The standard errors for RC1, RC2 estimators were computed using the sandwich method, whereas the estimation of the standard error for the SIMEX estimators was based on the bootstrap method, re-sampling 30 times.

Table 1 shows the results of the simulations pertaining to the estimation of the parameter β for 500. For βx, the naive estimator noticeably shows the worst performance among all the methods with respect to biases and coverage probabilities. Its biases and MSE’s are substantially large and increase as the variance of the Berkson error or that of the classical error becomes large. The attenuation effect seems to be more serious with the classical error than the Berkson error. The bias problem is severely accentuated by the presence of errors of both types. Also, it can be seen that its coverage probabilities are well below the nominal level of 95%. Hence, it is important to correct for both types of errors when they are present in the covariates of a Cox regression model. RC methods appear to have reduced the bias from the naive estimator although not completely. RC2, which uses the data in the calibration subsample generally exhibits smaller biases (except when σb2=0.3 and P(η = 1) = 0.5) and larger coverage probabilities than RC1. Both SIM1 and SIM2 display smaller biases than the RC-based estimators. SIM1 works well in correcting the bias from the Berkson error and performs acceptably when the magnitude of the classical error is small. However, its bias becomes noticeable when the variance of the classical error is large. This highlights the importance of the choice of the extrapolation function for SIMEX method, which may perform well in terms of bias reduction when the choice is appropriate. Meanwhile, SIM2 appears to show very small biases and good coverage probabilities, although its MSE is larger than the RC and SIM1 counterparts. It can also be noted that the biases and ASE’s of RC2, SIM1 and SIM2 decrease to some extent as the proportion of the calibration data increases. Moreover, all methods, including the naive one perform equally well with respect to biases for the estimation of the parameter βz. The results of the simulations for the sample size n = 1000 are shown in Table 2. The performances of the methods are similar to those observed in Table 1. Also, the biases and ASE’s are smaller than those presented in Table 1 for all the methods.

Table 1:

Simulation results for the estimation of β: n = 500; λ(u)=exp(βxX+βzZ), X=L+Ub; L ~ N(0,1); UbN(0,σb2) is the Berkson error; Wj=L+Uc,j; Uc,jN(0,σc2) is the classical error, j = 1,2; β=(log(2),0); Q=1+0.8X+ϵ; ε ~ N(0,0.5); p(η = 1) is the proportion of the calibration data; Z ~ Bernoulli(0.5); censoring rate is 50%; Naive, “naive” estimator; RC1, RC replacing X by E(X|Wˉ); RC2, RC replacing X by {E(X|Wˉ,Q)}η{E(X|Wˉ)}1η; SIM1, SIMEX using quadratic extrapolation function; SIM2, SIMEX using polynomial of degree 4 as extrapolation function.

σb2βσc2=0.3σc2=0.5
NaiveRC1RC2SIM1SIM2NaiveRC1RC2SIM1SIM2
P(η = 1) = 0.5
0.3βxBias13.0754.5514.5800.3250.44218.0975.1355.2181.6890.392
SD6.2387.2287.8128.77612.5065.9027.5188.3189.00313.858
ASE6.3277.1018.0818.85912.3126.0277.3258.6509.19413.747
MSE2.0990.7300.8200.7711.5663.6230.8290.9640.8391.922
CP44.49089.78093.58795.19093.78816.63387.17492.98694.79094.188
βzBias0.3020.3020.3380.3270.2790.3250.3250.3640.3550.284
SD6.7976.7976.7607.3867.8736.7976.7976.7617.4668.217
ASE6.3466.3386.3536.9847.6886.3466.3366.3547.0758.069
MSE0.4630.4630.4580.5470.6210.4630.4630.4580.5590.676
CP93.18692.58793.58792.58594.18893.78892.98693.78893.18694.389
0.5βxBias14.6686.3866.0180.5020.30719.5226.9236.5671.9960.163
SD6.1557.1217.9639.15013.0285.8177.3888.5219.29414.309
ASE6.3017.0708.4629.27012.8236.0067.3019.0919.57114.244
MSE2.5300.9150.9960.8401.6984.1491.0251.1570.9042.048
CP35.20083.80091.00094.60094.00010.80082.60091.00094.00095.400
βzBias0.3520.3520.3850.3880.3490.3540.3540.3880.3930.340
SD6.7246.7246.6787.5428.0406.7146.7146.6727.5928.344
ASE6.3466.3366.3567.1857.8576.3466.3366.3587.2558.212
MSE0.4530.4530.4470.5700.6480.4520.4520.4470.5780.697
CP93.60093.60093.60093.20095.20094.00094.20093.20093.40095.000
P(η = 1) = 0.7
0.3βxBias13.0504.5254.0660.5750.16218.0925.1304.5822.0100.009
SD6.2257.2198.0518.58412.5825.8977.5118.6048.80513.914
ASE6.3267.1007.9938.69312.1286.0277.3268.5678.99913.515
MSE2.0900.7260.8130.7401.5833.6210.8270.9500.8161.936
CP45.09089.78092.98694.38993.38716.60087.20093.80094.00094.188
βzBias0.3190.3190.2910.3310.2690.3170.3170.2880.3360.257
SD6.8046.8046.7427.3267.8136.7936.7936.7327.3948.139
ASE6.3466.3366.3526.9177.6176.3466.3366.3536.9957.982
MSE0.4640.4640.4550.5380.6110.4620.4620.4540.5480.663
CP93.18692.58593.78892.98693.78893.80093.00093.60093.20093.600
0.5βxBias14.6686.3865.1270.8220.03819.5226.9235.5432.3530.249
SD6.1557.1218.2598.93913.0705.8177.3888.8419.09814.372
ASE6.3017.0708.3289.06312.6196.0067.3018.9569.33013.985
MSE2.5300.9150.9450.8061.7084.1491.0251.0890.8832.066
CP35.20083.80091.80095.40093.60010.80082.60093.00093.80093.600
βzBias0.3520.3520.3000.3610.2980.3540.3540.3020.3670.281
SD6.7246.7246.6537.4607.9596.7146.7146.6487.5138.268
ASE6.3466.3366.3527.1077.7766.3466.3366.3547.1688.117
MSE0.4530.4530.4440.5580.6340.4520.4520.4430.5660.684
CP93.60093.60093.00093.20095.20094.00094.20093.60093.00095.000
  1. Note: SD denotes the sample standard deviation of the estimates; ASE is the average of the estimated standard errors; MSE is the mean squared errors; CP represents the coverage probability of the 95% confidence intervals. The results are presented in the 102 scale.

Table 2:

Simulation results for the estimation of β: n = 1000; λ(u)=exp(βxX+βzZ), X=L+Ub; L ~ N(0,1); UbN(0,σb2) is the Berkson error; Wj=L+Uc,j; Uc,jN(0,σc2) is the classical error, j = 1,2; β=(log(2),0); Q=1+0.8X+ϵ; ε ~ N(0,0.5); p(η = 1) is the proportion of the calibration data; Z ~ Bernoulli(0.5); censoring rate is 50%; Naive, “naive” estimator; RC1, RC replacing X by E(X|Wˉ); RC2, RC replacing X by {E(X|Wˉ,Q)}η{E(X|Wˉ)}1η; SIM1, SIMEX using quadratic extrapolation function; SIM2, SIMEX using polynomial of degree 4 as extrapolation function.

σb2βσc2=0.3σc2=0.5
NaiveRC1RC2SIM1SIM2NaiveRC1RC2SIM1SIM2
P(η = 1) = 0.5
0.3βxBias12.8514.3643.9880.2860.41217.8594.9564.4841.7030.272
SD4.3355.0235.3426.1528.6754.1715.3195.8406.3729.693
ASE4.4414.9515.4536.2058.6414.2335.1065.8236.4109.627
MSE1.8390.4430.4440.3790.7543.3640.5290.5420.4350.940
CP17.23485.77285.77294.38994.9901.60082.00087.00093.20094.400
βzBias0.1190.1190.1280.1570.2340.1130.1130.1200.1640.251
SD4.4554.4554.4904.7895.1344.4704.4704.5044.8425.351
ASE4.4824.4844.4924.8825.3624.4814.4834.4924.9305.610
MSE0.1990.1990.2020.2300.2640.2000.2000.5050.2350.287
CP94.79094.18894.18894.38995.79294.40094.00094.20094.80096.000
0.5βxBias14.5106.2735.3160.4180.35519.3396.8065.7271.9420.193
SD4.3555.0475.5976.5099.1854.1975.3516.1446.72210.230
ASE4.4234.9245.6646.5259.0434.2185.0836.0776.71110.028
MSE2.2950.6480.5960.4250.8453.9160.7500.7050.4901.047
CP9.00072.80086.20095.00095.4000.60069.80086.40092.80094.800
βzBias0.1030.1030.1200.1410.2110.0940.0940.1080.1470.229
SD4.4734.4734.5084.9525.2684.4864.4864.5214.9925.465
ASE4.4824.4834.4915.0195.4644.4814.4824.4915.0554.695
MSE0.2000.2000.2030.2450.2780.2010.2010.2050.2490.299
CP94.60095.00094.60094.80095.60095.00094.60094.40095.60095.800
P(η = 1) = 0.7
0.3βxBias12.8374.3513.5400.4000.28117.8594.9563.9581.8390.110
SD4.3335.0225.3135.9658.5534.1715.3195.8016.1969.561
ASE4.4424.9525.4716.0578.4924.2335.1065.8406.2509.453
MSE1.8360.4410.4080.3570.7323.3640.5290.4930.4180.914
CP17.23485.77290.58194.58994.9901.60082.00089.80092.40094.400
βzBias0.1220.1220.0890.1610.2330.1130.1130.0820.1650.249
SD4.4574.4574.4424.7635.1094.4704.4704.4524.8165.326
ASE4.4824.4844.4924.8635.3464.4814.4834.4914.9105.590
MSE0.1990.1990.1970.2270.2620.2000.2000.1980.2320.284
CP94.79094.18895.19094.79095.79294.40094.00094.80095.20096.400
0.5βxBias14.5106.2734.6260.5280.23519.3396.8074.9452.0600.060
SD4.3555.0465.5156.2778.9944.1975.3516.0516.50010.024
ASE4.4234.9245.7026.3448.8664.2185.0836.1136.5109.819
MSE2.2950.6480.5180.3970.8103.9160.7500.6110.4551.005
CP9.00072.80089.60095.60095.6000.60069.60088.80093.40094.800
βzBias0.1030.1030.0520.1420.2090.0940.0940.0450.1460.225
SD4.4734.4734.4764.9195.2384.4864.4864.4884.9585.435
ASE4.4824.4834.4924.9995.4474.4814.4824.4915.0345.674
MSE0.2000.2000.2000.2420.2750.2010.2010.2010.2460.296
CP94.60095.00095.00094.80095.60095.00094.60094.80094.80096.000
  1. Note: SD denotes the sample standard deviation of the estimates; ASE is the average of the estimated standard errors; MSE is the mean squared errors; CP represents the coverage probability of the 95% confidence intervals. The results are presented in the 102 scale.

The simulation results for the estimation of the nuisance parameters by the method of moments are presented in Table 3 for the setting with σb2=0.5 and σc2=0.5. The estimators generally show small bias and acceptable coverage probabilities. Also, it can be noted that the standard errors for the nuisance parameters involved in the distribution of the instrumental variable get smaller as the size of the calibration subsample increases. The biases and ASE’s of the estimators decrease as the sample size increases.

Table 3:

Simulation results for the estimation of the nuisance parameters: λ(u)=exp(βxX+βzZ), X=L+Ub; L ~ N(0,1); UbN(0,σb2) is the Berkson error; Wj=L+Uc,j; Uc,jN(0,σc2) is the classical error, j = 1,2; σb2=0.5; σc2=0.5; β=(log(2),0); Q=1+0.8X+ϵ; ε ~ N(0,0.5); p(η = 1) is the proportion of the calibration data; Z ~ Bernoulli(0.5); censoring rate is 50%.

n = 500n = 1000
μlα0α1σl2σb2σc2σϵ2μlα0α1σl2σb2σc2σϵ2
P(η = 1) = 0.5
Bias0.2750.3450.2310.4228.8410.0103.1300.1190.2960.3590.0882.4190.0830.854
SD4.8626.8979.7838.28936.2353.25916.5553.4094.9846.7815.68224.9742.22012.437
ASE4.9897.0039.5418.01736.4593.15717.6863.5334.9556.7625.69523.8002.22611.876
MSE0.2370.4770.9580.68913.9110.1062.8390.1160.2490.4610.3236.2950.0491.554
CP95.00096.20093.00094.80094.00095.00097.00096.20094.40095.80094.20094.00095.60094.400
P(η = 1) = 0.7
Bias0.2750.0520.1720.4185.3140.0092.7450.1190.0140.1280.0851.5620.0840.844
SD4.8625.7247.8508.29030.9773.25914.6273.4094.0595.3245.68220.0412.22110.285
ASE4.9895.6607.6638.01728.9253.15714.8433.5334.0105.4035.69519.2692.22610.012
MSE0.2370.3280.6160.6899.8780.1062.2150.1160.1650.2840.3234.0410.0491.065
CP95.00093.40094.80094.80094.20095.00096.60096.20094.20094.20094.20094.80095.60095.000
  1. Note: SD denotes the sample standard deviation of the estimates; ASE is the average of the estimated standard errors; MSE is the mean squared errors; CP represents the coverage probability of the 95% confidence intervals. The results are presented in the 102 scale.

In addition, we conducted some simulations to investigate the performances of the methods when Xi is not normally distributed. The simulation settings were similar to those in Table 1 with the exception that Li was simulated from gamma distribution G(1,1), Ub,i followed the uniform U(σb3,σb3), σb2=0.5, σc2=0.5 and P(ηi=1)=0.5. We estimated the parameter β by all the methods under the correctly specified or misspecified models for Li and Ub,i to gain insight into potential effects of models misspecification on the performances of the methods. The misspecified model for Ub,i assumed normality for Ub,i even though this assumption did not hold. The vector of nuisance parameters was estimated using the method of moments. The results of these simulations are reported in Table 4. The performances of the methods under the correctly specified models are similar to those observed in Table 1 and Table 2 with the SIMEX methods showing smaller biases and better coverage probabilities than the RC methods. Also, the SIMEX estimators show similar performances under both correct and misspecified models situations, suggesting that the method may not be very affected by a misspecification of the distribution for the latent variable or the measurement error.

Table 4:

Results of simulation assessing the effects of a misspecified distribution for the Berkson error: λ(u)=exp(βxX+βzZ), X=L+Ub; L ~ G(1,1); UbU(σb3,σb3) is the Berkson error; Wj=L+Uc,j; Uc,jN(0,σc2) is the classical error, j = 1,2; β=(log(2),0); Q=1+0.8X+ϵ; ε ~ N(0,0.5); Z ~ Bernoulli(0.5); censoring rate is 50%; σb2=0.5; σc2=0.5; p(η = 1) = 0.5; Naive, “naive” estimator; RC1, RC replacing X by E(X|Wˉ); RC2, RC replacing X by {E(X|Wˉ,Q)}η{E(X|Wˉ)}1η; SIM1, SIMEX using quadratic extrapolation function; SIM2, SIMEX using polynomial of degree 4 as extrapolation function.

nβCorrect modelMisspecified model
NaiveRC1RC2SIM1SIM2NaiveRC1RC2SIM1SIM2
500βxBias16.2336.9885.9310.3190.54316.2332.5692.6480.2020.823
SD5.9236.9457.9169.20213.3335.9237.5558.9629.25813.589
ASE5.5235.9896.5228.38912.4095.5236.5068.4968.27912.198
MSE2.9860.9710.9780.8481.7812.9860.6370.8730.8571.853
CP15.10274.69479.38891.22493.87815.10288.16394.89890.61293.673
βzBias0.1380.1420.1000.1260.0241.380.1380.1040.1290.033
SD6.3346.3696.2997.0808.0946.3346.3346.2907.0638.084
ASE6.3596.3776.4657.2598.2056.3596.3676.4107.2338.181
MSE0.4010.4060.3970.5010.6550.4010.4010.3960.4990.653
CP95.30695.10294.49095.91894.49095.30695.51094.49096.12294.490
1000βxBias16.4847.4625.3760.9150.23716.4842.9932.0410.8990.214
SD4.0164.7565.2406.0679.1754.0245.1755.8946.0669.179
ASE3.8764.1554.3405.9728.7203.8764.5305.6885.9578.705
MSE2.8780.7830.5640.3760.8422.8790.3570.3890.3760.843
CP1.00053.60070.80093.20093.0001.00485.14193.37392.97293.173
βzBias0.0070.0110.0670.0180.1530.0070.0070.0930.0030.177
SD4.4854.5244.5054.9325.6614.4854.4854.4624.9345.672
ASE4.4824.4884.4905.0045.6344.4824.4854.4934.9945.625
MSE0.2010.2050.2030.2430.3210.2010.2010.1990.2430.322
CP94.80094.60094.40094.20094.40094.77994.77994.98094.37893.976
  1. Note: SD denotes the sample standard deviation of the estimates; ASE is the average of the estimated standard errors; MSE is the mean squared errors; CP represents the coverage probability of the 95% confidence intervals. The results are presented in the 102 scale.

6 Application

In this section, we apply the estimation methods to the analysis of the data from the ACTG 175 study, which was designed to compare Zidovudine (ZDV) alone, Zidovudine combined with Didanosine (ZDV+ddI), Zidovudine combined with Zalcitabine (ZDV+ddC) and Didanosine (ddI) alone in HIV-1 infected patients. A detailed description of the data can be found in [25]. These data have been previously analyzed by a number of authors under different settings [16, 17]. In this application, we considered the subset of participants with two CD4 cell counts taken prior to the start of treatment and within three weeks of randomization. There were in total 2441 subjects in the analyzed dataset and only 1459 of them had available CD8 cell counts taken at randomization. Also, the median follow-up time for the participants was 144.3 weeks with an interquartile range of 42.2 weeks. There was a total of 306 events, consisting of death or AIDS disease development during follow-up. The two CD4 cell counts, measured before the beginning of treatment and within three weeks of randomization were treated as replicates for the baseline CD4 cell counts. Furthermore, CD4 and CD8 measurements were log-transformed for variance stabilization and approximate normality achievement. The scatter plot of the first replicate versus the second replicate of the baseline log10(1+CD4) and the scatter plot of the log10(1+CD8) against each of these replicates in the calibration subsample are shown in the upper panel of Figure 1, which also displays the qq-plots of these variables in its lower panel.

Figure 1: Scatter plots and qq-plots of duplicates of baseline log10(1+CD4)$\text{log}_{10}(1+CD4)$ and log10(1+CD8)$\text{log}_{10}(1+CD8)$ in calibration sample. (a) scatter plot of first replicate versus second replicate for baseline log10(1+CD4)$\text{log}_{10}(1+CD4)$. (b) Scatter plot of first log10(1+CD4)$\text{log}_{10}(1+CD4)$ replicate versus log10(1+CD8)$\text{log}_{10}(1+CD8)$ in calibration. (c) Scatter plot of second log10(1+CD4)$\text{log}_{10}(1+CD4)$ replicate versus log10(1+CD8)$\text{log}_{10}(1+CD8)$ in calibration. (d) qq-plot of first log10(1+CD4)$\text{log}_{10}(1+CD4)$ replicate. (e) qq-plot of second log10(1+CD4)$\text{log}_{10}(1+CD4)$ replicate. (f) qq-plot of log10(1+CD8)$\text{log}_{10}(1+CD8)$ in calibration subsample.
Figure 1:

Scatter plots and qq-plots of duplicates of baseline log10(1+CD4) and log10(1+CD8) in calibration sample. (a) scatter plot of first replicate versus second replicate for baseline log10(1+CD4). (b) Scatter plot of first log10(1+CD4) replicate versus log10(1+CD8) in calibration. (c) Scatter plot of second log10(1+CD4) replicate versus log10(1+CD8) in calibration. (d) qq-plot of first log10(1+CD4) replicate. (e) qq-plot of second log10(1+CD4) replicate. (f) qq-plot of log10(1+CD8) in calibration subsample.

Our interest in this application was in assessing the association between the true baseline log10(1+CD4) and the time to progression to AIDS disease or death adjusting for the treatment effect in the following proportional hazards model:

λ(u)=λ0(u)exp(βXXi+βZZi),

where Xi is the true baseline log10(1+CD4), Zi=(Z1i,Z2i,Z3i), Z1i, Z2i and Z3i represent the indicator variables for ZDV+ddI, ZDV+ddC and ddI, respectively, for subject i, λ0(.) is an unspecified baseline hazards function, βX and βZ=(βZ1,βZ2,βZ3) are the regression coefficients to be estimated. Moreover, we coded the first and second replicates for baseline log10(1+CD4) as Wi1 and Wi2, respectively, and denoted the log10(1+CD8) measured at randomization by Qi for subject i. The measurement error and Qi were modeled as in (1) and (2), respectively. We estimated the main parameter β=(βX,βZ) by the RC and SIMEX methods. The implementations of the methods were done similarly as in the simulation study and assumed normality for the measurement errors. The vector of nuisance parameters ν was estimated using the method of moments.

The results of the analysis are reported in Table 5. The estimates of the nuisance parameters by the method of moments are shown in the lower portion of this table. They suggest that the error involved in the measurements of the true baseline log10(1+CD4) is essentially of the classical type. Furthermore, the results of the estimation of βX by all the methods in the upper part of the table indicate that the log10(1+CD4) is significantly associated with the time to AIDS disease or death. Larger values of CD4 cell counts lead to lower risk of death or AIDS development. The estimates obtained based on the RC and SIMEX methods show stronger log10(1+CD4) effect than the naive method. In addition, it appears from the estimation of βZ that Didanosine and the combination of Zidovudine and Didanosine treatments are significantly different from Zidovudine in terms of their effects on time to progression to AIDS or death.

Table 5:

Results of the ACTG 175 data analysis: Naive, “naive” estimator; RC1, RC not using data in calibration subsample; RC2, RC making use of calibration subsample data; SIM1, SIMEX using quadratic extrapolation function; SIM2, SIMEX using polynomial of degree 4 as extrapolation function.

βRCSIMEX
NaiveRC1RC2SIM1SIM2
βXEstimate4.53105.50575.41795.40115.1889
SE0.40860.46180.66900.61630.8546
p-value<0.0001<0.0001<0.0001<0.0001<0.0001
βZ1Estimate0.51140.51140.50950.53300.5282
SE0.16130.16120.16160.17830.1907
p-value0.00150.00150.00160.00280.0056
βZ2Estimate0.29500.29500.29340.30440.2978
SE0.15400.15390.15370.17880.1933
p-value0.05540.05520.05620.08860.1234
βZ3Estimate0.43210.43210.43430.44240.4427
SE0.15760.15750.15780.18880.1981
p-value0.00610.00610.00590.01910.0254
νμlα0α1σl2σb2σc2σϵ2
Estimate2.52431.80350.45260.01501050.00650.0386
SE0.00270.21410.08470.00050.01360.00030.0030
p-value<0.0001<0.0001<0.0001<0.00010.9996<0.0001<0.0001
  1. βX is the coefficient of log10(1+CD4); βZ1, βZ2 and βZ3 are the coefficients of the indicator variables for ZDV+ddI, ZDV+ddC and ddI, respectively; SE means standard error and p-value is the Wald test-based p-value.

7 Conclusion

We have developed a new estimation method for Cox proportional hazards models with some covariates subject to a mixture of Berkson and classical measurement errors. The proposed method is based on the simulation extrapolation algorithm and leads to a consistent estimator of the vector of parameters of the proportional hazards model provided that the true extrapolation function is known. In practice, polynomial functions can be used to approximate this function. In addition, the variances of the classical and Berkson errors are estimated using the method of moments with no assumption about the mixture percentage of the error variances. The method makes use of two replicates of the error-prone covariates and data on an instrumental variable that are available for some subjects in a calibration subsample only.

Extensive simulations have shown that the proposed SIMEX method outperforms the regression calibration method with regard to biases and coverage probabilities in many finite sample size situations. A sensitivity analysis has revealed that the method may perform satisfactorily when the distribution of the Berkson error is misspecified.

Acknowledgements

This research was supported in parts by the National Institute of Health grants R01ES017030 (Wang and Tapsoba), HL121347 (Wang and Tapsoba) and GM100573 (Chao, Wang and Tapsoba).

References

[1] Cox DR. Regression models and life tables (with discussion). J R Stat Soc Ser B. 1972;34:187–220.Search in Google Scholar

[2] Tsiatis AA, DeGruttola V, Wulfsohm MS. Modeling the relationship of survival to longitudinal data measured with error. Applications to survival and CD4 counts in patients with AIDS. J Am Stat Assoc. 1995;90:27–37.10.1080/01621459.1995.10476485Search in Google Scholar

[3] Carroll RJ, Ruppert D, Stefanski LA, Crainiceanu CM. Nonlinear measurement error models, a modern perspective, 2nd ed. London: Chapman and Hall, 2006.10.1201/9781420010138Search in Google Scholar

[4] Prentice RL. Covariate measurement errors and parameter estimates in failure time regression. Biometrika. 1982;69:331–42.10.1093/biomet/69.2.331Search in Google Scholar

[5] Wang CY. Robust best linear estimation for regression analysis using surrogate and instrumental variables. Biostatistics. 2012;13:326–40. DOI: 10.1093/biostatistics/kxr051.Search in Google Scholar

[6] Kim HY, Yasui Y, Burtyn I. Attenuation in rsik-estimation in logistic and Cox proportional hazards models due to grouped-based exposure assessment strategy. Ann Occup Hyg. 2006;50:623–35.Search in Google Scholar

[7] Schafer DW, Gilbert ES. Some statistical implications of dose uncertainty in radiation dose response analyses. Radiat Res. 2006;166:303–12.10.1667/RR3358.1Search in Google Scholar

[8] Huang Y, Wang CY. Cox regression with accurate covariates unascertainable: a nonparametric correction approach. J Am Stat Assoc. 2000;95:1209–1910.1080/01621459.2000.10474321Search in Google Scholar

[9] Huges MD. Regression dilution in the proportional hazards model. Biometrics. 1993;49:1056–66.10.2307/2532247Search in Google Scholar

[10] Kuchenhoff H, Bender R, Langner I. Effect of Berkson measurement error on parameter estimates in Cox regression models. Lifetime Data Anal. 2007;13:261–272.10.1007/s10985-007-9036-2Search in Google Scholar

[11] Wang CY. Robust sandwich covariance estimation for regression estimator in Cox regression with measurement error. Stat Probab Lett. 1999;45:371–8.10.1016/S0167-7152(99)00079-6Search in Google Scholar

[12] Wang CY, Wang N, Wang S. Regression analysis when covariates are regression parameters of a random effects model for observed longitudinal measurements. Biometrics. 2000;56:487–95.10.1111/j.0006-341X.2000.00487.xSearch in Google Scholar PubMed

[13] Cook JR, Stefanski LA. Simulation-extrapolation estimation in parametric measurement error models. J Am Stat Assoc. 1994;89:1314–28.10.1080/01621459.1994.10476871Search in Google Scholar

[14] Wang CY, Huang Y. Error in timing regression with observed longitudinal measurements. Stat Med. 2003;22:2577–90. DOI: 10.1002/sim.1435.Search in Google Scholar PubMed

[15] Nakamura T. Proportional hazards model with covariates subject to measurement error. Biometrics. 1992;48:829–38.10.2307/2532348Search in Google Scholar

[16] Song X, Wang CY. Proportional hazards model with covariates measurement error and instrumental variables. J Am Stat Assoc. 2014. DOI: 10.1080/01621459.2014.896805.Search in Google Scholar PubMed PubMed Central

[17] Tapsoba JD, Wang CY, Lee SM. Joint modeling of survival time and longitudinal data with subject-specific change points in the covariates. Stat Med. 2011;30:232–49. DOI: 10.1002/sim.4107.Search in Google Scholar PubMed PubMed Central

[18] Tsiatis AA, Davidian M. A semiparametric estimator for the proportional hazards model with longitudinal covariates measured with error. Biometrika. 2001;88:447–58.10.1093/biomet/88.2.447Search in Google Scholar

[19] Hu P, Tsiatis AA, Davidian M. Estimating the parameters in the Cox model when covariate variables are measured with error. Biometrics. 1998;54:1407–19.10.2307/2533667Search in Google Scholar

[20] Song X, Davidian M, Tsiatis AA. A semiparametric likelihood approach to joint modeling of longitudinal and time-to-event data. Biometrics. 2002;58:742–53.10.1111/j.0006-341X.2002.00742.xSearch in Google Scholar PubMed

[21] Wang CY. Nonparametric maximum likelihood estimator for Cox regression with subject-specific measurement error. Scand J Stat. 2008;35:613–28.10.1111/j.1467-9469.2008.00605.xSearch in Google Scholar

[22] Yan Y, Yi GY. A corrected profile likelihood method for survival data with covariate measurement error under the Cox model. Canad J Stat. 2015;43:454–80.10.1002/cjs.11258Search in Google Scholar

[23] Li Y, Guolo A, Hoffman FO, Carroll RJ. Shared uncertainty in measurement error problems, with application to Nevada Test Site fallout data. Biometrics. 2007;63:1226–36. DOI: 10.1111/j.1541-0420.2007.00810.x.Search in Google Scholar PubMed

[24] Tapsoba JD, Wang CY, Lee SM. Expected estimating equation using calibration data for generalized linear models with a mixture of Berkson and classical errors in covariates. Stat Med. 2014;33:675–92. DOI: 10.1002/sim.5966.Search in Google Scholar PubMed PubMed Central

[25] Hammer SM, Katzenstein DA, Huges MD, Gundacker H, Schooley RT, Haubrich MR, Henry WK, Lederman MM, Phair JP, Niu M, Hirch MS, Merigan TC. A trial comparing nucleoside monotherapy with combination therapy in HIV-infected adults with CD4 cell counts from 200 to 500 per cubic millimeter. N Engl J Med. 1996;335:1081–90.10.1056/NEJM199610103351501Search in Google Scholar PubMed

[26] Mallick B, Hoffman FO, Carroll RJ. Semiparametric regression modeling with mixtures of Berkson and classical error, with application to fallout from the Nevada test site. Biometrics. 2002;58:13–20.10.1111/j.0006-341X.2002.00013.xSearch in Google Scholar PubMed

[27] Carroll RJ, Delaigle A, Hall P. Non-parametric regression estimation from data contaminated by a mixture of Berkson and classical errors. J R Stat Soc., Ser B. 2007;69:859–78. DOI: 10.1111/j.1467-9868.2007.00614.x.Search in Google Scholar PubMed PubMed Central

[28] Wang CY. Corrected score estimator for joint modeling of longitudinal and failure time data. Stat Sin. 2006;16:235–53.Search in Google Scholar

[29] Wang CY, Hsu L, Feng ZD, Prentice RL. Regression calibration in failure time regression. Biometrics. 1997;53:131–45.10.2307/2533103Search in Google Scholar

[30] Whitemore AS, Keller JB. Approximations for regression with covariate measurement error. J Am Stat Assoc. 1988;83:1057–66.10.1080/01621459.1988.10478701Search in Google Scholar

[31] Apanasovich TV, Carroll RJ, Maity A. SIMEX and standard error estimation in semiparametric measurement error models. Electron J Stat. 2009;3:318–48. DOI: 10.1214/08-EJS341.Search in Google Scholar PubMed PubMed Central

[32] Carroll RJ, Kuchenhoff H, Lombard F, Stefanski LA. Asymptotics for SIMEX estimator in nonlinear measurement error models. J Am Stat Assoc. 1996;91:242–50. DOI: 10.1111/j.1467-9868.2007.00614.x.Search in Google Scholar

Received: 2018-03-05
Revised: 2019-01-13
Accepted: 2019-03-20
Published Online: 2019-04-06

© 2019 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/ijb-2018-0028/html
Scroll to top button