Skip to main content

Advertisement

Log in

A functional inference for multivariate current status data with mismeasured covariate

  • Published:
Lifetime Data Analysis Aims and scope Submit manuscript

Abstract

Covariate measurement error problems have been recently studied for current status failure time data but not yet for multivariate current status data. Motivated by the three-hypers dataset from a health survey study, where the failure times for three-hypers (hyperglycemia, hypertension, hyperlipidemia) are subject to current status censoring and the covariate self-reported body mass index may be subject to measurement error, we propose a functional inference method under the proportional odds model for multivariate current status data with mismeasured covariates. The new proposal utilizes the working independence strategy to handle correlated current status observations from the same subject, as well as the conditional score approach to handle mismeasured covariate without specifying the covariate distribution. The asymptotic theory, together with a stable computation procedure combining the Newton–Raphson and self-consistency algorithms, is established for the proposed estimation method. We evaluate the method through simulation studies and illustrate it with three-hypers data.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

References

  • Carroll RJ, Ruppert D, Stefanski LA, Crainiceanu CM (2006) Measurement error in nonlinear models: a modern perspective, 2nd edn. Chapman & Hall, Boca Raton

    Book  Google Scholar 

  • Chang IS, Wen CC, Wu YJ (2007) A profile likelihood theory for the correlated gamma-frailty model with current status family data. Stat Sinica 17:1023–1046

    MATH  MathSciNet  Google Scholar 

  • Chen MH, Tong X, Sun J (2007) The proportional odds model for multivariate interval-censored failure time data. Stat Med 26:5147–5161

    Article  MathSciNet  Google Scholar 

  • Chen MH, Tong X, Sun J (2009) A frailty model approach for regression analysis of multivariate current status data. Stat Med 28:3424–3436

    Article  MathSciNet  Google Scholar 

  • Clayton DG (1978) A model for association in bivariate life tables and its application in epidemiological studies of familial tendency in chronic disease incidence. Biometrika 65:141–151

    Article  MATH  MathSciNet  Google Scholar 

  • Dunson DB, Dinse GE (2002) Bayesian models for multivariate current status data with informative censoring. Biometrics 58:79–88

    Article  MATH  MathSciNet  Google Scholar 

  • Goggins WB, Finkelstein DM (2000) A proportional hazards model for multivariate interval-censored failure time data. Biometrics 56:940–943

    Article  MATH  Google Scholar 

  • Gumbel EJ (1960) Bivariate exponential distributions. J Am Stat Assoc 55:698–707

    Article  MATH  MathSciNet  Google Scholar 

  • Heyde CC, Morton R (1998) Multiple roots in general estimating equations. Biometrika 85:954–959

    Article  MATH  Google Scholar 

  • Hosmer DW, Lemeshow S (2000) Applied logistic regression. Wiley, New York

    Book  MATH  Google Scholar 

  • Huang J (1996) Efficient estimation for the Cox model with interval censoring. Ann Stat 24:540–568

    Article  MATH  Google Scholar 

  • Korosok MR (2008) Introduction to empirical processes and semiparametric inference. Springer, New York

    Book  Google Scholar 

  • Ma S (2009) Cure model with current status data. Stat Sinica 19:233–249

    MATH  Google Scholar 

  • Murphy SA, van der Vaart AW (1999) Observed information in semiparametric models. Bernoulli 5:381–412

    Article  MATH  MathSciNet  Google Scholar 

  • Rudin W (1973) Functional analysis. McGraw-Hill, New York

    MATH  Google Scholar 

  • Stefanski LA, Carroll RJ (1987) Conditional scores and optimal scores in generalized linear measurement error models. Biometrika 69:331–342

    MathSciNet  Google Scholar 

  • Su PF, Chi Y (2014) Marginal regression approach for additive hazards models with clustered current status data. Stat Med 33:46–58

    Article  MathSciNet  Google Scholar 

  • Tsiatis AA, Davidian M (2001) A semiparametric estimator for the proportional hazards model with longitudinal covariates measured with error. Biometrika 88:447–458

    Article  MATH  MathSciNet  Google Scholar 

  • van der Vaart AW (1998) Asymptotic statistics. Cambridge University Press, Cambridge

    Book  MATH  Google Scholar 

  • van der Vaart AW, Wellner JA (1996) Weak convergence and empirical processes. Springer, New York

    Book  MATH  Google Scholar 

  • Wei LJ, Lin DY, Weissfeld L (1989) Regression analysis of multivariate incomplete failure time data by modeling marginal distributions. J Am Stat Assoc 84:1065–1073

    Article  MathSciNet  Google Scholar 

  • Wen CC, Chen YH (2011) Nonparametric maximum likelihood analysis of clustered current status data with the gamma frailty Cox model. Comput Stat Data Anal 55:1053–1060

    Article  MATH  MathSciNet  Google Scholar 

  • Wen CC, Chen YH (2012) Conditional score approach to errors-in-variable current status data under the proportional odds model. Scand J Stat 39:635–644

    Article  MATH  MathSciNet  Google Scholar 

  • Wen CC, Huang SYH, Chen YH (2011) Cox regression for current status data with mismeasured covariates. Can J Stat 39:73–88

    Article  MATH  MathSciNet  Google Scholar 

Download references

Acknowledgments

We wish thank the Associate Editor and two anonymous referees for their many helpful comments and suggestions that greatly improved the paper. This research was supported by the National Science of Council of Taiwan grants.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Chi-Chung Wen.

Appendix

Appendix

We use the notation \({\mathbb P}_n, P_0,\) and \(P\) for the expectations taken under the empirical distribution, the true underlying distribution, and a given model, respectively. For simplicity the assumptions and proofs for theories are presented under the simpler setting where the distribution of \((C_1,\ldots ,C_K)\) is independent of \((X,Z_1,\ldots ,Z_K),\) though the proposed method can allow the dependence case. We consider \(H_1,\ldots ,H_K\) are functions in \(\mathcal{H},\) the set of right-continuous non-decreasing functions that are uniformly bounded on the study period \([0,\tau ^*].\) Let \(\ell ({\varvec{\theta }})=\log L({\varvec{\theta }}).\) The asymptotic theories are based on the following regularity assumptions, which have been similarly made in the studies of univariate current status data (e.g. Huang 1996; van der Vaart 1998; Ma 2009; Wen and Chen 2012).

  1. (C1)

    The examination times \(C_1,\ldots ,C_K\) possess a common continuous density whose support is an interval \([\tau _1,\tau _2]\) with \(0<\tau _1<\tau _2<\tau ^*.\)

  2. (C2)

    The true parameter \(({\varvec{\beta }}_0,\sigma ^2_0)\) lies in the interior of a compact subset \(\mathcal B \times \mathcal{Q}\) of \({\mathbb R}^d \times (0,\infty )\); all \(H_{10},\ldots , H_{K0} \) are continuously differentiable on \([\tau _1,\tau _2]\) and satisfy \(-M<H_{k0}(\tau _1)<H_{k0}(\tau _2)<M\) for \(k=1,\ldots ,K.\)

  3. (C3)

    The distribution of \((X,\mathbf{Z}_1,\ldots ,\mathbf{Z}_K)\) is not concentrated on any proper subspace of \({\mathbb R}^{1+\sum _k d_k} \) and has a bounded support, where \(d_k=\dim (\mathbf{Z}_k)\).

  4. (C4)

    The functions \(\mathbf{g}_1^*,\ldots ,\mathbf{g}_K^*\) given in (6) are differentiable with bounded derivatives on \([\tau _1,\tau _2]\).

  5. (C5)

    The information matrix \(\mathcal{I}\) defined in (7) is invertible.

Proof of Theorem 1

Let \((\widehat{H}_{1,({\varvec{\beta }},\sigma ^2)}, \ldots , \widehat{H}_{K,({\varvec{\beta }},\sigma ^2)} )\) denote the maximizer of \(L_n\) with \(({\varvec{\beta }}, \sigma ^2)\) fixed. We first apply Theorem 5.7 of van der Vaart (1998) to establish the consistency of \((\widehat{H}_{1,({\varvec{\beta }}_0,\widehat{\sigma }^2)}, \ldots , \widehat{H}_{K,({\varvec{\beta }}_0,\widehat{\sigma }^2)} )\). Since the class of monotone and uniformly bounded functions is a Donsker class, by Theorem 2.10.6 of van der Vaart and Wellner (1996) and conditions (C1)-(C3), we know that the class \(\{ \ell ({\varvec{\beta }}_0,H_1,\ldots ,H_K,\) \( \sigma ^2_0) \ | \ H_1,\ldots ,H_K \in \mathcal{H} \}\) is Donsker and hence Glivenko-Cantelli. Further, by Jensen’s inequality, we have

$$\begin{aligned} P_0 \{\ell ({\varvec{\beta }}_0,H_1,\ldots ,H_K,\sigma ^2_0)-\ell ({\varvec{\theta }}_0)\} \le \log (P_0 \left\{ \frac{L({\varvec{\beta }}_0,H_1,\ldots ,H_K,\sigma ^2_0)}{L({\varvec{\theta }}_0)}\right\} ) =0, \end{aligned}$$

and the equality holds only if \(H_k=H_{k0}\) on \((\tau _1,\tau _2)\) for all \(k=1,\ldots ,K.\) This indicates that

$$\begin{aligned} \sup _{\sum _{k=1}^K\Vert H_k-H_{k0}\Vert _2 >\varepsilon }P_0 \ell ({\varvec{\beta }}_0,H_1,\ldots ,H_K,\sigma ^2_0)< P_0\ell ({\varvec{\theta }}_0). \end{aligned}$$

Furthermore, note that

$$\begin{aligned} {\mathbb P}_n \ell ({\varvec{\theta }}_0)+o_p(1)&= {\mathbb P}_n \ell ({\varvec{\beta }}_0,H_{10},\ldots ,H_{K0},\widehat{\sigma }^2)\\&\le {\mathbb P}_n \ell ({\varvec{\beta }}_0, \widehat{H}_{1,({\varvec{\beta }}_0,\widehat{\sigma }^2)}, \ldots , \widehat{H}_{K,({\varvec{\beta }}_0,\widehat{\sigma }^2)} ,\widehat{\sigma }^2 )\\&= {\mathbb P}_n \ell ({\varvec{\beta }}_0,\widehat{H}_{1,({\varvec{\beta }}_0,\widehat{\sigma }^2)}, \ldots , \widehat{H}_{K,({\varvec{\beta }}_0,\widehat{\sigma }^2)} ,\sigma ^2_0)+o_p(1), \end{aligned}$$

where the inequality follows from the definition of \((\widehat{H}_{1,({\varvec{\beta }},\sigma ^2)}, \ldots , \widehat{H}_{K,({\varvec{\beta }},\sigma ^2)} )\), and two equalities are obtained by the mean value theorem and the consistency of \(\widehat{\sigma }^2.\) Therefore, by Theorem 5.7 of citev98, we have \(\sum _{k=1}^K \Vert \widehat{H}_{k,({\varvec{\beta }}_0,\widehat{\sigma }^2)}-H_{k0}\Vert _2 {\rightarrow } \ 0,\) and so \(\widehat{H}_{k,({\varvec{\beta }}_0,\widehat{\sigma }^2)}(t) \mathop {\rightarrow }\limits ^{P} H_{k0}(t),\) \(k=1,\ldots , K,\) for every \(t\) in \([\tau _1,\tau _2]\).

By the consistency shown above and the mean value theorem, we have

$$\begin{aligned} {\mathbb P}_n {\varvec{\ell }}_0({\varvec{\beta }}_0,\widehat{H}_{1,({\varvec{\beta }}_0,\widehat{\sigma }^2)}, \ldots , \widehat{H}_{K,({\varvec{\beta }}_0,\widehat{\sigma }^2)},\widehat{\sigma }^2)={\mathbb P}_n {\varvec{\ell }}_0({\varvec{\beta }}_0,H_{10}, \ldots , H_{K0},\sigma ^2_0)+o_p(1). \end{aligned}$$

Since the right hand side of the above display set to zero is an unbiased estimating equation, a consistent sequence of solutions of \({\varvec{\beta }}\) to it exists, indicating the existence of a consistent solution of \({\varvec{\beta }}\) to the CS estimating equation \({\mathbb P}_n {\varvec{\ell }}_0({\varvec{\beta }},\widehat{H}_{1,({\varvec{\beta }},\widehat{\sigma }^2)}, \ldots , \widehat{H}_{K,({\varvec{\beta }},\widehat{\sigma }^2)},\widehat{\sigma }^2)=0.\)

Now we shall prove

$$\begin{aligned} \left( \sum _{k=1}^K\Vert \widehat{H}_{k,({\varvec{\beta }},\sigma ^2)}-H_0\Vert _2^2\right) ^{1/2}=O_P(\Vert {\varvec{\beta }}-{\varvec{\beta }}_0\Vert +\Vert \sigma ^2-\sigma ^2_0\Vert +n^{-1/3}), \end{aligned}$$

by verifying the conditions (3.5) and (3.6) in Theorem 3.2 of Murphy and van der Vaart (1999). Here \(\Vert \cdot \Vert \) is the Euclidean norm, \(\Vert H\Vert _2^2=\int H^2(u)dQ(u),\) and \(Q\) denotes the marginal distribution of \(C_k.\) The rates of convergence and consistencies of \(\widehat{H}_k, k=1,\ldots ,K,\) can then be obtained by the consistency of \((\widehat{\varvec{\beta }}, \widehat{\sigma }^2).\)

Given two functions \(l\) and \(u\), the bracket \([l,u]\) is the set of all functions \(f\) with \(l\le f \le u\). An \(\varepsilon \)-bracket in \(L_2(P)=\{f: Pf^2<\infty \}\) is a bracket \([l,u]\) with \(P(u-l)^2<\varepsilon ^2.\) For a subclass \(\mathcal{C}\) of \(L^2(P),\) the bracketing number \(N_{[ \ ]}(\varepsilon ,\mathcal{C},L_2(P))\) is the minimum number of \(\varepsilon \)-bracket need to cover \(\mathcal C\) (see van der Vaart 1998). Let \(\varPsi =\{ \ell ({\varvec{\theta }}) : {\varvec{\theta }}\in \mathcal{B}\times \mathcal{H}^K \times \mathcal{Q}\}.\) Note that each element in \(\varPsi \) is uniformly bounded and satisfies \(P_0\{ \ell ({\varvec{\theta }}) -\ell ({\varvec{\beta }},H_{10},\ldots ,H_{K0},\sigma ^2) \}^2 \preceq \sum _{k=1}^K \Vert H_k-H_{k0}\Vert _2^2+\Vert {\varvec{\beta }}-{\varvec{\beta }}_0\Vert ^2+\Vert \sigma ^2-\sigma ^2_0\Vert ^2.\) The notation \(\preceq \) means smaller than, up to a constant. Lemma 1 below gives the bracketing integral \(J(\varDelta ,{\varPsi },L_2(P))\), defined as \(\int _0^{\varDelta } \{1+\log N_{[ \ ]}(\varepsilon ,\varPsi ,L_2(P))\}^{1/2}d\varepsilon ,\) is \(O(\varDelta ^{1/2})\). It then follows from Lemma 3.3 of Murphy and van der Vaart (1999) that their condition (3.6) is satisfied for \(\phi _n(\varDelta )=\varDelta ^{1/2}.\)

A similar Taylor series argument used in Lemma 2 below gives \(P_0\{\ell ({\varvec{\theta }}_0)-\ell ({\varvec{\beta }},H_{10},\ldots ,H_{K0},\sigma ^2))\}\preceq \Vert {\varvec{\beta }}-{\varvec{\beta }}_0\Vert ^2+\Vert \sigma ^2-\sigma ^2_0\Vert ^2.\) This and Lemma 2 imply

$$\begin{aligned} P_0\{\ell ({\varvec{\theta }})\!-\!\ell ({\varvec{\beta }},H_{10},\ldots ,H_{K0},\sigma ^2)\}\preceq \!-\!\sum _{k=1}^K \Vert H_k\!-\!H_{k0}\Vert _2^2\!+\!\Vert {\varvec{\beta }}\!-\!{\varvec{\beta }}_0\Vert ^2\!+\!\Vert \sigma ^2\!-\!\sigma ^2_0\Vert ^2, \end{aligned}$$

which is condition (3.5) of Murphy and van der Vaart (1999). This completes the proof.\(\square \)

Lemma 1

\(\log N_{[ \ ]}(\varepsilon ,\varPsi ,L_2(P_0)) =O( {1/ \varepsilon })\).

Proof

For fixed \(({\varvec{\beta }}, \sigma ^2)\) and \(1\le k \le K,\) the functions in \(\varPsi \) depend on \(H_k\) monotonically for \(\varDelta _{k}=1\) and \(\varDelta _{k}=0\) separately. Thus, given a \(\varepsilon \)-bracket \(H^{L}_k\le H_k \le H^{U}_k,\) it follows from monotonicity of \(\mathcal{E}_{k} \) in \(H_k\) that we can get a bracket \((\ell ^L,\ell ^U)\) for \(\ell ({\varvec{\theta }})\) where

$$\begin{aligned} \ell ^L&\equiv \log \prod _{k=1}^K \left[ \mathcal{E}_{k} ({\varvec{\beta }},H_1^L,\ldots ,H_K^L,\sigma ^2)^{\varDelta _{k}} \{1-\mathcal{E}_{k} ({\varvec{\beta }},H_1^U,\ldots ,H_K^U,\sigma ^2)\} ^{1-\varDelta _{k}} \right] ;\\ \ell ^U&\equiv \log \prod _{k=1}^K \left[ \mathcal{E}_{k}({\varvec{\beta }},H_1^U,\ldots ,H_K^U,\sigma ^2)^{\varDelta _{k}} \{1-\mathcal{E}_{k}({\varvec{\beta }},H_1^L,\ldots ,H_K^L,\sigma ^2)\} ^{1-\varDelta _{k}} \right] . \end{aligned}$$

Further, by the mean value theorem, we have \(|\ell ^L-\ell ^U|^2 \preceq \sum _{k=1}^K (H_k^{U}-H_k^{L})^2 (C_{k}).\) Thus brackets for \(H_k\) of \(\Vert \cdot \Vert _2\)-size \(\varepsilon \) can translate into brackets for \(\ell ({\varvec{\theta }})\) of \(L_2(P_0)\)-size proportional to \(\varepsilon .\) By Example 19.11 of van der Vaart (1998), we can cover the set of all \(H_k\) by \(\exp (C/\varepsilon )\) brackets of size \(\varepsilon \) for some constant \(C.\) Next we allow \(\zeta =({\varvec{\beta }}',\sigma ^2)'\) to vary freely as well. Because \(\mathcal{B}\times \mathcal{Q}\) is finite-dimensional and \((\partial /\partial \zeta ) \ell ({\varvec{\theta }})(O)\) is uniformly bounded in \(({\varvec{\theta }},O),\) this increases the entropy only slightly. Lemma 1 is thus proved.\(\square \)

Lemma 2

For \({\varvec{\theta }}\) near \({\varvec{\theta }}_0,\) \(P_0\{ \ell ({\varvec{\theta }})-\ell ({\varvec{\theta }}_0) \} \preceq -\{\sum _{k=1}^K\Vert H_k-H_{k0}\Vert _2^2+\Vert {\varvec{\beta }}-{\varvec{\beta }}_0\Vert ^2+\Vert \sigma ^2-\sigma ^2_0\Vert ^2\}.\)

Proof

Let \({\varvec{\theta }}_1=({\varvec{\beta }},H_{1},\ldots ,H_{K},\sigma ^2_0).\) It suffices to show

$$\begin{aligned}&P_0\{ \ell ({\varvec{\theta }}_1)-\ell ({\varvec{\theta }}_0) \} \preceq -\{\sum _{k=1}^K\Vert H_k-H_{k0}\Vert _2^2+\Vert {\varvec{\beta }}-{\varvec{\beta }}_0\Vert ^2\},\end{aligned}$$
(9)
$$\begin{aligned}&P_0\{ \ell ({\varvec{\theta }})-\ell ({\varvec{\theta }}_1)\} \preceq -\Vert \sigma ^2-\sigma ^2_0\Vert ^2. \end{aligned}$$
(10)

Let \(V=\{(C_k,S_k(\beta _{10},\sigma ^2_0),Z_k), k=1,\ldots ,K \}.\) By the Kullback-Leibler inequality, \(E_{0}\{ \ell ({\varvec{\theta }}_1)(O)|V \}\) is maximized at \(({\varvec{\beta }}_0,H_{10}(C_1),\ldots ,H_{K0}(C_K)).\) So its first derivative is equal to 0 there (this can also be verified directly by the fact that \(E_0\{\varDelta _k|V\}=\mathcal{E}_k({\varvec{\theta }}_0)(O)\)). Since \((C_k,S_k,Z_k)\)’s have bounded support and the parameter spaces are compact, a Taylor’s expansion gives \(E_0\{ \ell ({\varvec{\theta }}_1)(O)-\ell ({\varvec{\theta }}_0)(O) |V \} \preceq -\{\sum _{k} (H_k(C_k)-H_{k0}(C_k))^2+\Vert {\varvec{\beta }}-{\varvec{\beta }}_0\Vert ^2\},\) which implies (9).

We now prove (10). Denote \(p_{{\varvec{\theta }}}\) the density of observed data \(O\) given the model \({\varvec{\theta }}.\) Because the densities \(p_{{\varvec{\theta }}_0}/p_{{\varvec{\theta }}_1}\) are uniformly bounded above and below by a positive constant, (10) is equivalent to

$$\begin{aligned} P_{{\varvec{\theta }}_1}\{ \ell ({\varvec{\theta }})-\ell ({\varvec{\theta }}_1) \}\preceq - \Vert \sigma ^2-\sigma ^2_0\Vert ^2. \end{aligned}$$
(11)

Let \(V_1=\{(C_k,S_k(\beta _{1},\sigma ^2_0),Z_k), k=1,\ldots ,K \}.\) By the fact that \(E_{{\varvec{\theta }}_1}\{\varDelta _k|V_1\}=\mathcal{E}_k({\varvec{\theta }}_1)(O),\) we can show directly that the first derivative of \(E_{{\varvec{\theta }}_1}\{ \ell ({\varvec{\theta }})(O)|V_1 \}\) with respect to \(\sigma ^2\) at \(\sigma ^2=\sigma ^2_0\) is equal to 0. Note that the second derivative of \(E_{{\varvec{\theta }}_1}\{ \ell ({\varvec{\theta }})(O)|V_1 \}\) with respect to \(\sigma ^2\) is given by \(-E_{{\varvec{\theta }}_1}\{\beta _1^4 \sum _k \mathcal{V}_k({\varvec{\theta }})(O)/(4m^2)|V_1 \}.\) Consequently, a Taylor’s expansion around \(\sigma ^2=\sigma ^2_0\) can thus yield \(E_{{\varvec{\theta }}_1}\{ \ell ({\varvec{\theta }})(O)-\ell ({\varvec{\theta }}_1)(O) |V_1 \} \preceq - \Vert \sigma ^2-\sigma ^2_0\Vert ^2,\) which implies (11). This completes the proof.\(\square \)

For fixed \(1\le k \le K,\) consider a parametric path \(H_{k,\varepsilon }\) in \(\mathcal H\) through \(H_k,\) that is, \(H_{k,\varepsilon } \in \mathcal{H}\) and \(H_{k,\varepsilon }=H_k\) when \(\varepsilon =0.\) Let \(\dot{\mathcal{H}}=\{g_k:({\partial }/{\partial \varepsilon })|_{\varepsilon =0} H_{k,\varepsilon }=g_k,k=1,\ldots ,K\}.\) Then the score for \(H_k\) along the direction \(g_k,\) define by \( ({\partial }/{\partial \varepsilon })|_{\varepsilon =0} \ell ({\varvec{\beta }},H_1,\ldots ,H_{k-1},H_{k,\varepsilon },H_{k+1},\ldots ,H_K,\sigma ^2),\) has the form

$$\begin{aligned} \ell _k({\varvec{\theta }})[g_k](O)= g(C_{k})\{\varDelta _{k}-\mathcal{E}_{k}({\varvec{\theta }})(O)\}. \end{aligned}$$

Also define \(\ell _{0k}({\varvec{\theta }})[g_k]=({\partial }/{\partial \varepsilon }) |_{\varepsilon =0} \ell _0({\varvec{\beta }},H_1,\ldots ,H_{k-1},H_{k,\varepsilon },H_{k+1},\ldots ,H_K,\sigma ^2)\) and \(\ell _{k'k}({\varvec{\theta }})[\tilde{g}_{k'},g_k]=({\partial }/{\partial \varepsilon }) |_{\varepsilon =0} \ell _{k'}({\varvec{\beta }},H_1,\ldots ,H_{k-1},H_{k,\varepsilon },H_{k+1},\ldots ,H_K,\sigma ^2)[\tilde{g}_{k'}],\) where \(1\le k,k'\le K,\) \(g_k\) and \(\tilde{g}_k'\) are in \(\dot{\mathcal{H}}.\) They have forms

$$\begin{aligned}&\ell _{0k}({\varvec{\theta }})[g_k]=- g_k(C_{k})\{S_{k}-{\varvec{\beta }}_1{\tilde{\sigma }^2}, Z' \}'\mathcal{V}_{k}({\varvec{\theta }})(O),\\&\ell _{kk}({\varvec{\theta }})[\tilde{g}_k, g_k]=- g_k(C_{k}){\tilde{g}}_k(C_{k})\mathcal{V}_{k}({\varvec{\theta }})(O), \end{aligned}$$

and \(\ell _{k'k}({\varvec{\theta }})[\tilde{g}_k', g_k]=0\) if \(k'\ne k.\) Following semiparametric M-estimator theories (e.g. Korosok 2008), the function \({\varvec{\ell }}^*\) given in Sect. 3 is defined as \({\varvec{\ell }}^*({\varvec{\theta }})= \ell _0({\varvec{\theta }})- \sum _{k=1}^K \ell _k({\varvec{\theta }})[\mathbf g_k^*],\) where \(\mathbf{g}_k^*\) is the \(d\)-dimensional \((d=\dim ({\varvec{\beta }}))\) vector-valued function satisfying

$$\begin{aligned} P_0 \left( \sum _{k=1}^K\ell _{0k}({\varvec{\theta }}_0)[g_k]-\sum _{k'=1}^K\sum _{k=1}^K\ell _{k'k}({\varvec{\theta }}_0)[\mathbf{g}_{k'}^*,g_k]\right) =0, \end{aligned}$$
(12)

for all \(g_k\) in \(\dot{\mathcal{H}}.\) Fixing one \(k\) \((1\le k\le K)\) and setting \(g_{k'}=0\) for all \(k' \ne k\) in (12), we have

$$\begin{aligned}&\int g_k(u)E[\{S_{k}-\beta _1{\tilde{\sigma }^2}, Z' \}'\mathcal{V}_{k}({\varvec{\theta }})(O)|C_{k}=u ]dQ(u)\\&\quad = \int g_k(u)\mathbf{g}_k^*(u)E[\mathcal{V}_{k}({\varvec{\theta }})(O)|C_{k}=u ]dQ(u), \end{aligned}$$

which implies that \(\mathbf{g}_k^*\) is given by (6). Below we establish the asymptotic theory of the CS estimator.

Proof of Theorem 2

We first verify

$$\begin{aligned} \sqrt{n}P_0 {\varvec{\ell }}^*({\varvec{\beta }}_0,\widehat{H}_1,\ldots ,\widehat{H}_K,\sigma ^2_0)=o_p(1). \end{aligned}$$
(13)

Apply a Taylor expansion to \({\varvec{\ell }}^*({\varvec{\beta }}_0, H_1,\ldots ,H_K,\sigma ^2_0)(O)\) at the point \((H_{10}(C_{1}),\ldots , \) \(H_{K0}(C_{K}))\) to get

$$\begin{aligned}&P_0 {\varvec{\ell }}^*({\varvec{\beta }}_0,H_1,\ldots ,H_K,\sigma ^2_0)- P_0 {\varvec{\ell }}^*({\varvec{\theta }}_0) \\ \nonumber&\quad =P_0\left( \sum _{k} \ell _{0k}({\varvec{\theta }}_0)[H_k-H_{k0}]-\sum _{k',k}\ell _{k'k}({\varvec{\theta }}_0)[\mathbf{g}_{k'}^*,H_k-H_{k0}]\right) \\ \nonumber&\qquad +O_p(\sum _{k=1}^K\Vert H_k-H_{k0}\Vert ^2_2). \end{aligned}$$
(14)

Using (12), the fact that \(P_0 {\varvec{\ell }}^*({\varvec{\theta }}_0)=0,\) and applying the rates of convergence on \(\widehat{H}_k, k=1,\ldots , K,\) to (14), we get (13).

Applying Theorem 2.10.6 of van der Vaart and Wellner (1996), it can be verified with condition (C4) that \(\{ {\varvec{\ell }}^*({\varvec{\theta }}) | {\varvec{\theta }}\in \mathcal{B} \times \mathcal{H}^K \times \mathcal{Q}\}\) and \(\{ \varphi (\sigma ^2) | \sigma ^2 \in \mathcal{Q} \}\) are uniformly bounded Donsker classes; the proof of which is technical and hence omitted here. Combining this with the consistency of \(\widehat{\varvec{\theta }}\) leads to

$$\begin{aligned} \sqrt{n} ({\mathbb P}_n-P_0) \left[ \begin{array}{c} {\varvec{\ell }}^*(\widehat{\varvec{\theta }})-{\varvec{\ell }}^*({\varvec{\theta }}_0)\\ \varphi (\widehat{\sigma }^2)-\varphi (\sigma ^2_0) \end{array} \right] =o_p(1). \end{aligned}$$

Adding (13) to the first row of preceding display and using the facts that \(P_0 {\varvec{\ell }}^*({\varvec{\theta }}_0)=0\) and \({\mathbb P}_n {\varvec{\ell }}^*({\widehat{\varvec{\theta }}})={\mathbb P}_n\varphi (\widehat{\sigma }^2)=0,\) it is seen that

$$\begin{aligned} -\sqrt{n}P_0 \left[ \begin{array}{c} {\varvec{\ell }}^*(\widehat{\varvec{\theta }})-{\varvec{\ell }}^*({\varvec{\beta }}_0,{\widehat{H}}_1,\ldots ,\widehat{H}_K,\sigma ^2_0)\\ \varphi (\widehat{\sigma }^2)-\varphi (\sigma ^2_0) \end{array} \right] =\sqrt{n}{\mathbb P}_n \left[ \begin{array}{c} {\varvec{\ell }}^*({\varvec{\theta }}_0)\\ \varphi (\sigma ^2_0) \end{array} \right] +o_p(1). \end{aligned}$$

By the mean value theorem, there exists \((\tilde{{\varvec{\beta }}},\tilde{\sigma ^2})\) lying between \((\widehat{{\varvec{\beta }}},\widehat{\sigma }^2)\) and \(({\varvec{\beta }}_0,\sigma ^2_0)\) such that

$$\begin{aligned}&-\sqrt{n} P_0 \left[ \begin{array}{cc} \frac{\partial }{\partial {\varvec{\beta }}} {\varvec{\ell }}^*({\tilde{{\varvec{\beta }}}} ,{\widehat{H}}_1,\ldots ,\widehat{H}_K,\tilde{\sigma ^2}) &{} \frac{\partial }{\partial \sigma ^2 } {\varvec{\ell }}^*({\tilde{{\varvec{\beta }}}},{\widehat{H}}_1,\ldots ,\widehat{H}_K,\tilde{\sigma ^2})\\ 0 &{}\frac{\partial }{\partial \sigma ^2 }\varphi (\tilde{\sigma ^2}) \end{array} \right] \left( \begin{array}{c} {\widehat{\varvec{\beta }}}-{\varvec{\beta }}_0\\ \widehat{\sigma }^2-\sigma ^2_0 \end{array} \right) \\&\quad =\sqrt{n}{\mathbb P}_n \left[ \begin{array}{c} {\varvec{\ell }}^*({\varvec{\theta }}_0)\\ \varphi (\sigma ^2_0) \end{array} \right] +o_p(1). \end{aligned}$$

By the consistency of \((\widehat{\varvec{\beta }},\widehat{\sigma }^2)\) and condition (C5), we have

$$\begin{aligned} \sqrt{n}\left[ \begin{array}{c} {\widehat{\varvec{\beta }}}-{\varvec{\beta }}_0\\ \widehat{\sigma }^2-\sigma ^2_0 \end{array} \right] = \mathcal{I}^{-1}\sqrt{n}{\mathbb P}_n \left[ \begin{array}{c} {\varvec{\ell }}^*({\varvec{\theta }}_0)\\ \varphi (\sigma ^2_0) \end{array} \right] +o_P(1)\mathop {\rightarrow }\limits ^{d} N(0,\mathcal{I}^{-1}{\varSigma }(\mathcal{I}^{-1})'). \end{aligned}$$

This completes the proof.\(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wen, CC., Huang, YH. & Wu, YJ. A functional inference for multivariate current status data with mismeasured covariate. Lifetime Data Anal 21, 379–396 (2015). https://doi.org/10.1007/s10985-014-9296-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10985-014-9296-6

Keywords

Navigation