Skip to main content

Advertisement

Log in

An additive hazards cure model with informative interval censoring

  • Published:
Lifetime Data Analysis Aims and scope Submit manuscript

Abstract

The existence of a cured subgroup happens quite often in survival studies and many authors considered this under various situations (Farewell in Biometrics 38:1041–1046, 1982; Kuk and Chen in Biometrika 79:531–541, 1992; Lam and Xue in Biometrika 92:573–586, 2005; Zhou et al. in J Comput Graph Stat 27:48–58, 2018). In this paper, we discuss the situation where only interval-censored data are available and furthermore, the censoring may be informative, for which there does not seem to exist an established estimation procedure. For the analysis, we present a three component model consisting of a logistic model for describing the cure rate, an additive hazards model for the failure time of interest and a nonhomogeneous Poisson model for the observation process. For estimation, we propose a sieve maximum likelihood estimation procedure and the asymptotic properties of the resulting estimators are established. Furthermore, an EM algorithm is developed for the implementation of the proposed estimation approach, and extensive simulation studies are conducted and suggest that the proposed method works well for practical situations. Also the approach is applied to a cardiac allograft vasculopathy study that motivated this investigation.

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

Similar content being viewed by others

References

  • Cai C, Zou Y, Peng Y, Zhang J (2012) smcure: An R-package for estimating semiparametric mixture cure models. Comput Methods Prog Biomed 108:1255–1260

    Article  Google Scholar 

  • Chen X, Fan Y, Tsyrennikov V (2006) Efficient estimation of semiparametric multivariate Copula models. J Am Stat Assoc 101:1228–1240

    Article  MathSciNet  Google Scholar 

  • Chen M, Tong X, Sun J (2009) A frailty model approach for regression analysis of multivariate current status data. Stat Med 28:3424–3426

    Article  MathSciNet  Google Scholar 

  • Cook R, Lawless J (2007) The statistical analysis of recurrent events. Springer, Berlin

    MATH  Google Scholar 

  • Efron B (1979) Bootstrap methods: another look at the jackknife. Ann Stat 7:1–26

    Article  MathSciNet  Google Scholar 

  • Farewell V (1982) The use of mixture models for the analysis of survival data with long-term survivors. Biometrics 38:1041–1046

    Article  Google Scholar 

  • Finkelstein D (1986) A proportional hazards model for interval-censored failure time data. Biometrics 42:845–854

    Article  MathSciNet  Google Scholar 

  • He X, Tong X, Sun J (2009) Semiparametric analysis of panel count data with correlated observation and follow-up times. Lifetime Data Anal 15:177–196

    Article  MathSciNet  Google Scholar 

  • Hu T, Xiang L (2013) Efficient estimation for semiparametric cure models with interval-censored data. J Multivar Anal 121:139–151

    Article  MathSciNet  Google Scholar 

  • Huang C, Qin J, Wang M (2010) Semiparametric analysis for recurrent event data with time-dependent covariates and informative censoring. Biometircs 66:39–49

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  • Kalbfleisch J, Prentice R (2002) The statistical analysis of failure time data, 2nd edn. Wiley, New York

    Book  Google Scholar 

  • Kuk A, Chen C (1992) A mixture model combining logistic regression with proportional hazards regression. Biometrika 79:531–541

    Article  Google Scholar 

  • Lam K, Xue H (2005) A semiparametric regression cure model with current status data. Biometrika 92:573–586

    Article  MathSciNet  Google Scholar 

  • Liu H, Shen Y (2009) A semiparametric regression cure model for interval-censored data. J Am Stat Assoc 104:1168–1178

    Article  MathSciNet  Google Scholar 

  • Liu Y, Hu T, Sun J (2016) Regression analysis of current status data in the presence of a cured subgroup and dependent censoring. Lifetime Data Anal 23:626–650

    Article  MathSciNet  Google Scholar 

  • Li S, Hu T, Wang P, Sun J (2017) Regression analysis of current status data in the presence of dependent censoring with applications to tumorigenicity experiments. Comput Stat Data Anal 110:75–86

    Article  MathSciNet  Google Scholar 

  • Li S, Hu T, Zhao X, Sun J (2019) A class of semiparametric transformation cure models for interval-censored failure time data. Comput Stat Data Anal 133:153–165

    Article  MathSciNet  Google Scholar 

  • Lorentz G (1986) Bernstein polynomials, 2nd edn. University Toronto Press, Toronto

    MATH  Google Scholar 

  • Ma L, Hu T, Sun J (2015) Sieve maximum likelihood regression analysis of dependent current status data. Biometrika 102:731–738

    Article  MathSciNet  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  • Ma S (2010) Mixed case interval censored data with a cured subgroup. Stat Sin 20:1165–1181

    MathSciNet  MATH  Google Scholar 

  • Pollard D (1984) Convergence of stochastic process. Springer, New York

    Book  Google Scholar 

  • Shao F, Li J, Ma S, Lee MLT (2014) Semiparametric varying-coefficient model for interval censored data with a cured proportion. Stat Med 33:1700–1712

    Article  MathSciNet  Google Scholar 

  • Sharples LD, Jackson CH, Parameshwar J, Wallwork J, Large SR (2003) Diagnostic accuracy of coronary angiography and risk factors for postCheart-transplant cardiac allograft vasculopathy. Transplantation 76:679–682

    Article  Google Scholar 

  • Shen X (1997) On the methods of sieves and penalization. Ann Stat 25:2555–2591

    Article  MathSciNet  Google Scholar 

  • Shen X, Wong W (1994) Convergence rate of sieve estimates. Ann Stat 22:580–615

    Article  MathSciNet  Google Scholar 

  • Sun J (2006) The statistical analysis of interval-censored failure time data. Springer, New York

    MATH  Google Scholar 

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

    Book  Google Scholar 

  • Wang P, Zhao H, Sun J (2016) Regression analysis of case k interval-censored failure time data in the presence of informative censoring. Biometrics 72:1103–1112

    Article  MathSciNet  Google Scholar 

  • Wang S, Wang C, Wang P, Sun J (2018) Semiparametric analysis of the additive hazards model with informatively interval-censored failure time data. Comput Stat Data Anal 125:1–9

    Article  MathSciNet  Google Scholar 

  • Wang S, Wang C, Wang P, Sun J (2020) Estimation of the additive hazards model with case K interval-censored failure time data in the presence of informative censoring. Comput Stat Data Anal 144:1–15

    MathSciNet  MATH  Google Scholar 

  • Xiang L, Ma X, Yau K (2011) Mixture cure model with random effects for clustered interval-censored survival data. Stat Med 30:995–1006

    Article  MathSciNet  Google Scholar 

  • Xu L, Zhang J (2010) Multiple imputation method for the semiparametric accelerated failure time mixture cure model. Comput Stat Data Anal 54:1808–1816

    Article  MathSciNet  Google Scholar 

  • Yu B, Peng Y (2008) Mixture cure models for multivariate survival data. Comput Stat Data Anal 52:1524–1532

    Article  MathSciNet  Google Scholar 

  • Zhou J, Zhang J, Lu W (2018) Computationally efficient estimation for the generalized odds rate mixture cure model with interval-censored data. J Comput Graph Stat 27:48–58

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The authors wish to thank the Associate Editor and three reviewers for their many helpful and useful comments and suggestions that greatly improved the paper. This work was partly supported by the National Natural Science Foundation of China Grant Nos. 11901054, 11671054, 11671168 and the Tian Yuan Mathematical Foundation of National Natural Science Foundation of China(Nos. 11926340, 11926341).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Chunjie Wang.

Additional information

Publisher's Note

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

Appendix

Appendix

In this appendix, we will sketch the proofs for Theorems 13 and for this, we will mainly use some results about the empirical processes given in Van der Vaart and Wellner (1996). Throughout the following proofs, we denote \(Pf=\int f(x)dP(X)\) and \(P_{n}f=n^{-1}\sum _{i=1}^{n}f(X_{i})\) for a function f and a random variable X with the distribution P and let J represent a generic constant that may vary from place to place. We first present the required regularity conditions.

  1. (C1)

    (i) \((\beta ^{\top }_{1}, \beta _{2}, \alpha ^{\top }, \eta ^{\top }, \sigma ^{2}) \in {\mathcal {B}}\), where \({\mathcal {B}}\) is a compact subset of \({\mathbb {R}}^{2p+q+2}\);(ii) The rth derivative of \(\varLambda _{0}(\cdot )(\varLambda _{0h}(\cdot ))\) is bounded and continuous.

  2. (C2)

    (i) There exists a positive \(\eta ^{*}\) such that \(P(U_{i,j}-U_{i,j-1}\ge \eta ^{*})=1\) for subject i, where \(j=2,\ldots ,K_{i}\); (ii) For subject i, the union of the support of \(U_{ij}\) is contained in an interval [ab], where \(0<a<b<\infty \) and \(j=1,\ldots ,K_{i}\).

  3. (C3)

    The covariates X and Z have bounded supports in \({\mathbb {R}}^{p}\) and \({\mathbb {R}}^{q}\), respectively, where p and q are the dimension of X and Z;

  4. (C4)

    Let \(l(\theta ;O)=\log L_{O}(\theta )\) with \(L_{O}(\theta )\) given in Sect. 2. For any \(\theta ^{1},\theta ^{2}\in \varTheta \), if \(l(\theta ^{1};O)=l(\theta ^{2};O)\), then \(\theta ^{1}=\theta ^{2}\). In particular, O denotes the single observation data.

  5. (C5)

    For every \(\theta \) in a neighbourhood of \(\theta _{0}\), \(P(l(\theta ;O)-l(\theta _{0};O))\preceq -d^{2}(\theta ,\theta _{0})\), where \(\preceq \) means “smaller than, up to a constant”.

  6. (C6)

    (i) \(0< E[(\dot{l}(\theta _{0};O)[\upsilon ])^{2}] < \infty \) for \(\upsilon \ne 0, \upsilon \in V\), where \(\dot{l}(\theta _{0};O)[\upsilon ]\) is defined in (E5) below; (ii) For \(\theta \in \{\theta : d(\theta ,\theta _{0})=O(\delta _{n})\}\), \(P\{\ddot{l}(\theta ;O)[\theta -\theta _{0},\theta -\theta _{0}]- \ddot{l}(\theta _{0};O)[\theta -\theta _{0},\theta -\theta _{0}]\}=O(\delta _{n}^{3})\) and \(\delta _{n}^{3} = O(n^{-1})\), where \(\ddot{l}(\theta ;O)[\upsilon ,{\tilde{\upsilon }}]\) is defined in (E8)

Proof of Theorem 1

We first define the covering number of the class \({\mathcal {L}}_{n}= \{l(\theta ;O):\theta \in \varTheta _{n}\}\). In particular, O denotes the single observation data. For any \(\epsilon >0\), define the covering number \(N(\epsilon , {\mathcal {L}}_{n},L_{1}(P_{n}) )\) as the smallest value of \(\kappa \) for which there exists \(\{ \theta ^{(1)}, \ldots , \theta ^{(\kappa )}\}\) such that

$$\begin{aligned} \min _{j\in \{1, \ldots , \kappa \}} \frac{1}{n}\sum _{i=1}^{n}| l(\theta ;O_{i})-l(\theta ^{(j)};O_{i})| < \epsilon \end{aligned}$$

for all \(\theta \in \varTheta _{n}\), where \(\theta ^{(j)}=(\beta _{1}^{(j)},\beta _{2}^{(j)}, \alpha ^{(j)},\eta ^{(j)},\sigma ^{2(j)}, \varLambda _{n}^{(j)}, \varLambda _{nh}^{(j)})^{\top }\in \varTheta _{n}\), \(j=1,\ldots ,\kappa \). We will define \(N(\epsilon , {\mathcal {L}}_{n}, L_{1}(P_{n}))= \infty \) if no such \(\kappa \) exists.

Let \(l(\theta ;O)\) denote the log likelihood function based on the single observation as defined in Sect. 2. Then the covering number(definition 2.1.5 in Van der Vaart and Wellner 1996) of the class \({\mathcal {L}}_{n}=\{l(\theta ;O):\theta \in \varTheta _{n}\}\) satisfies

$$\begin{aligned} N(\epsilon , {\mathcal {L}}_{n}, L_{1}(P_{n})) \le JM^{2p+q+2}M_{n}^{2(m+1)}\epsilon ^{-p_{m}}, \end{aligned}$$

where \(p_{m}=2p+q+2+2(m+1)\), \(M_{n}=O(n^{a})\) with \(0< a < \frac{1}{2}\). Adopting similar proofs of inequality (31) in Pollard (1984, p. 34), we have

$$\begin{aligned} \sup _{\theta \in \varTheta _{n}}|P_{n}l(\theta ;O)-Pl(\theta ;O)|\rightarrow 0. \hbox {a.s.} \end{aligned}$$
(E1)

Let \(M(\theta ;O)=-l(\theta ;O)\),

$$\begin{aligned} \zeta _{1n}= & {} \sup _{\theta \in \varTheta _{n}}|P_{n}M(\theta ;O)-PM(\theta ;O)|, \\ \zeta _{2n}= & {} P_{n}M(\theta _{0};O)-PM(\theta _{0};O). \end{aligned}$$

Denote \(C_{\epsilon }=\{\theta ;d(\theta ,\theta _{0})\ge \epsilon ,\theta \in \varTheta _{n}\}\).

$$\begin{aligned}&\inf _{C_{\epsilon }}PM(\theta ;O)= \inf _{C_{\epsilon }}\{PM(\theta ;O)- P_{n}M(\theta ;O)+P_{n}M(\theta ;O)\} \\&\quad \le \zeta _{1n}+\inf _{C_{\epsilon }}P_{n}M(\theta ;O) \end{aligned}$$
(E2)

If \({\hat{\theta }}_{n}\in C_{\epsilon }\), we have

$$\begin{aligned} \inf _{C_{\epsilon }}P_{n}M(\theta ;O)&=P_{n}M({\hat{\theta }}_{n};O) \le P_{n}M(\theta _{0};O)=P_{n}M(\theta _{0};O)-PM(\theta _{0};O)\\&\quad +PM(\theta _{0};O) \\&=\zeta _{2n}+PM(\theta _{0};O) \end{aligned}$$
(E3)

By identification condition(C4), we obtain that

$$\begin{aligned} \inf _{C_{\epsilon }}PM(\theta ;O)-PM(\theta _{0};O)=\delta _{\epsilon }>0. \end{aligned}$$
(E4)

By (E2) and (E3), we have

$$\begin{aligned} \inf _{C_{\epsilon }}PM(\theta ;O)\le \zeta _{1n}+\zeta _{2n}+ PM(\theta _{0};O)=\zeta _{n}+PM(\theta _{0};O) \end{aligned}$$
(E5)

with \(\zeta _{n}=\zeta _{1n}+\zeta _{2n}\). By (E4) and (E5), we get \(\zeta _{n}\ge \delta _{\epsilon }\). Here, we have \(\{{\hat{\theta }}_{n}\in C_{\epsilon }\}\subseteq \{\zeta _{n}\ge \delta _{\epsilon }\}\). By (E1) and Strong Law of Large Numbers, we have \(\zeta _{1n}=o(1),\zeta _{2n}=o(1)\), a.s. Therefore, by \(\cup _{k=1}^{\infty }\cap _{n=k}^{\infty }\{{\hat{\theta }}_{n}\in C_{\epsilon }\} \subseteq \cup _{k=1}^{\infty }\cap _{n=k}^{\infty } \{\zeta _{n}\ge \delta _{\epsilon }\}\), we complete the proof. \(\square \)

Proof of Theorem 2

To establish the convergence rate, note that by Theorem 1.6.2 of Lorentz (1986), there exists Bernstein polynomials \(\varLambda _{n0}\) and \(\varLambda _{nh0}\) such that \(\parallel \varLambda _{0}- \varLambda _{n0} \parallel _{\infty }=O(n^{-\frac{r\nu }{2}}) \) and \(\parallel \varLambda _{0h}- \varLambda _{nh0} \parallel _{\infty }=O(n^{-\frac{r\nu }{2}})\), respectively. For any \(\chi >0\), define the class \({\mathcal {F}}_{\chi }= \{l(\theta ;O)-l(\theta _{n0};O):\theta \in \varTheta _{n},d(\theta ,\theta _{n0})\le \chi \}\) with \(\theta _{n0}=(\beta _{1n0}^{\top },\beta _{2n0},\alpha _{n0}^{\top },\eta _{n0}^{\top }, \sigma ^{2}_{n0},\varLambda _{1n0},\varLambda _{2n0})^{\top }\). Following the calculation of Shen and Wong (1994, p. 597), we can establish that \(\log N_{[~]}(\epsilon ,{\mathcal {F}}_{\chi }, \parallel \cdot \parallel )\le JN\log (\chi /\epsilon )\) with \(N=2(m+1)\), where \(N_{[~]}(\epsilon ,{\mathcal {F}},d)\) denotes the bracketing number(definition 2.1.6 in Van der Vaart and Wellner 1996) with respect to the metric or semi-metric d of a function class \({\mathcal {F}}\). Condition (C5) lead to \(\parallel l(\theta ;O)-l(\theta _{n0};O)\parallel ^{2}_{2} \le J\chi ^{2}\) for any \(l(\theta ;O)-l(\theta _{n0};O) \in {\mathcal {F}}_{\chi }\). Therefore, by Lemma 3.4.2 of Van der Vaart and Wellner (1996), we obtain

$$\begin{aligned}&E_{P}\parallel n^{\frac{1}{2}}(P_{n}-P)\parallel _{{\mathcal {F}}_{\chi }} \le CH_{\chi }(\epsilon ,{\mathcal {F}}_{\chi },\parallel \cdot \parallel _{2}) \left\{ 1+\frac{H_{\chi }(\epsilon ,{\mathcal {F}}_{\chi },\parallel \cdot \parallel _{2})}{\chi ^{2}n^{\frac{1}{2}}}\right\} ,\\&\quad \le C\cdot CN^{\frac{1}{2}}\chi \left\{ 1+\frac{CN^{\frac{1}{2}\chi }}{\chi ^{2}N^{\frac{1}{2}}}\right\} =C(N^{\frac{1}{2}}+N/n^{\frac{1}{2}})\doteq \phi _{n}(\chi ). \end{aligned}$$
(E6)

where \(H_{\chi }(\epsilon ,{\mathcal {F}}_{\chi },\parallel \cdot \parallel _{2})= \int _{0}^{\chi }\{1+\log N_{[~]}(\epsilon ,{\mathcal {F}}_{\chi }, \parallel \cdot \parallel _{2})\}^{1/2}d\epsilon \). It is easy to see that \(\phi _{n}(\chi )/\chi \) decreasing in \(\chi \), and \(r^{2}_{n}\phi _{n}(1/r_{n})=r_{n}N^{\frac{1}{2}}+r^{2}_{n}N/n^{\frac{1}{2}}<2n^{\frac{1}{2}}\), where \(r_{n}=N^{-\frac{1}{2}}n^{\frac{1}{2}}=n^{(-\nu +1)/2},0<\nu <\frac{1}{2}\). Here, \(n^{(1-\nu )/2}\cdot \) \(d({\hat{\theta }},\theta _{n0})=O_{P}(1)\) by Theorem 3.2.5 of var der Van der Vaart and Wellner (1996). This, together with \(d(\theta _{n0},\theta _{0})=O_{p}(n^{-\frac{r\nu }{2}})\) (Theorem 1.6.2 of Lorentz 1986),yields that \(d({\hat{\theta }},\theta _{0})=O_{P}(n^{-(1-\nu )/2}+n^{-\frac{r\nu }{2}})\). This completes the proof. \(\square \)

Proof of Theorem 3

To complete the proof of Theorem 3, we need following notations. Define V as the linear span of \(\varTheta -\theta _{0}\), where \(\theta _{0}=(\beta _{10}^{\top },\beta _{20},\alpha _{0}^{\top }, \eta _{0}^{\top }, \sigma ^{2}_{0},\varLambda _{0},\varLambda _{0h})^{\top }\) denotes the true value of \(\theta \) and \(\varTheta _{0}\) the true parameter space. Let \(l(\theta ;O)\) be the log-likelihood function and \(\delta _{n}=n^{-r\upsilon /2}+n^{-(1-\upsilon )/2}\). For any \(\theta \in \{\theta \in \varTheta _{0}:d(\theta ,\theta _{0})=O(\delta _{n})\}\), define the first order directional derivative of \(l(\theta ;O)\) at the direction \(\upsilon \in V\) as

$$\begin{aligned} \dot{l}(\theta ;O)[\upsilon ] = \frac{dl(\theta +s\upsilon ;O)}{ds}| _{s=0}, \end{aligned}$$
(E7)

and the second order directional derivative as

$$\begin{aligned} \ddot{l}(\theta ;O)[\upsilon ,{\tilde{\upsilon }}]= \frac{d^{2}l(\theta +s\upsilon +{\tilde{s}}{\tilde{\upsilon }};O)}{d{\tilde{s}}}\mid _{s=0,{\tilde{s}}=0} =\frac{d\dot{l}(\theta +{\tilde{s}}{\tilde{\upsilon }};O)}{d{\tilde{s}}}\mid _{{\tilde{s}}=0}. \end{aligned}$$
(E8)

Define the Fisher inner product on the space V as \(\langle \upsilon , {\tilde{\upsilon }}\rangle = P\{\dot{l}(\theta ;O)[\upsilon ] \dot{l}(\theta ;O)[{\tilde{\upsilon }}]\}\) and the Fisher norm for \(\upsilon \in V\) as \(\parallel \upsilon \parallel ^{2}= \langle \upsilon ,\upsilon \rangle \). Let \({\bar{V}}\) be the closed linear span of V under the Fisher norm. Then \(({\bar{V}},\parallel \cdot \parallel )\) is a Hilbert space.

Furthermore, define the smooth functional of \(\theta \) as \(\gamma (\theta )=h_{1}^{\top }\beta _{1}+h_{2}\beta _{2}+h_{3}^{\top }\alpha + h_{4}^{\top }\eta + h_{5}\sigma ^{2}\), where \(h=(h_{1}^{\top },h_{2},h_{3}^{\top },h_{4}^{\top }, h_{5})^{\top }\) is any vector of \(2p+q+2\) dimension with \(\parallel h\parallel \le 1\). Let \(\vartheta = (\beta _{1}^{\top },\beta _{2},\alpha ^{\top },\eta ^{\top },\sigma ^{2})^{\top }\). Also for any \(\upsilon =(\upsilon _{\vartheta },b_{1},b_{2}) \in V\), we denote

$$\begin{aligned} {\dot{\gamma }}(\theta _{0})[\upsilon ]=\frac{d\gamma (\theta _{0}+s\upsilon )}{ds}\mid _{s=0}= h^{\top }\upsilon _{\vartheta }. \end{aligned}$$

Note that \(\gamma (\theta )-\gamma (\theta _{0})={\dot{\gamma }}(\theta _{0})[\theta -\theta _{0}]\). By the Riesz representation theorem, there exists \(\upsilon ^{*} \in {\bar{V}}\) such that \({\dot{\gamma }}(\theta _{0})[\upsilon ]=\langle \upsilon ^{*},\upsilon \rangle \) for all \(\upsilon \in {\bar{V}}\) and \(\parallel \upsilon ^{*}\parallel ^{2}=\parallel {\dot{\gamma }}(\theta _{0})\parallel \). Thus it follows from the proof method of Theorem 3 of Ma et al. (2015), we can show that

$$\begin{aligned} \sqrt{n}\langle {\hat{\theta }}-\theta _{0},\upsilon ^{*}\rangle + o_{p}(1) \rightarrow N(0,\parallel \upsilon ^{*}\parallel ^{2}) \end{aligned}$$

in distribution.

Then we will proof that \(\parallel \upsilon ^{*} \parallel ^{2}=h^{\top }\varSigma h\). For each component \(\vartheta _{k}, k=1,2,\ldots ,2p+q+2\), we denote by \(\varphi _{k}^{*}=(b_{1k}^{*},b_{2k}^{*})\) the solution to

$$\begin{aligned} \inf _{\varphi _{k}^{*}}E\{l_{\vartheta }\cdot e_{k}-l_{b_{1k}^{*}}[b_{1k}^{*}]-l_{b_{2k}^{*}}[b_{2k}^{*}]\}^{2}, \end{aligned}$$
(E9)

where \(l_{\vartheta }=(l_{\beta _{1}}^{\top },l_{\beta _{2}},l_{\alpha }^{\top }, l_{\eta }^{\top },l_{\sigma ^{2}})^{\top }\), \(l_{\vartheta }\) is the derivatives of \(l(\theta ;O)\) with respect to \(\vartheta \), \(l_{b_{1k}^{*}}[b_{1k}^{*}]\) and \(l_{b_{2k}^{*}}[b_{2k}^{*}]\) are the direction derivatives of \(l(\theta ;O)\) with respect to \(\varLambda _{0},\varLambda _{0h}\) at the direction \(b_{1k}^{*},b_{2k}^{*}\), respectively. \(e_{k}\) is a \((2p+q+2)\)-dimensional vector of zeros expect the k-th element being equal to 1. Define the k-th element of \(S_{\vartheta }\) as \(l_{\vartheta }\cdot e_{k}-l_{b_{1k}^{*}}[b_{1k}^{*}]-l_{b_{2k}^{*}}[b_{2k}^{*}]\), \(k=1,2,\ldots ,2p+q+2\).

Thus, we can show that

$$\begin{aligned} \sup _{\upsilon \in {\bar{V}}:\parallel \upsilon \parallel>0}\frac{\mid {\dot{\gamma }}(\theta _{0})[\upsilon ]\mid ^{2}}{\parallel \upsilon \parallel ^{2}} =\sup _{\upsilon \in {\bar{V}}:\parallel \upsilon \parallel>0} \frac{\mid h^{\top }\upsilon _{\vartheta } \mid ^{2}}{\parallel \upsilon \parallel ^{2}} =\sup _{\upsilon \in {\bar{V}}:\parallel \upsilon \parallel >0} \frac{\mid h^{\top }\upsilon _{\vartheta } \mid ^{2}}{P\{\dot{l}(\theta ;O)[\upsilon ]\}^{2}}. \end{aligned}$$

From the minimization procedure defined in (E9), using similar arguments to Sect. 3.2 in Chen et al. (2006), we obtain

$$\begin{aligned} \sup _{\upsilon \in {\bar{V}}:\parallel \upsilon \parallel >0}\frac{\mid {\dot{\gamma }}(\theta _{0})[\upsilon ]\mid ^{2}}{\parallel \upsilon \parallel ^{2}} =\sup _{(b_{1},b_{2})\ne 0} \frac{\mid h^{\top }\upsilon _{\vartheta } \mid ^{2}}{E(l_{\vartheta }[\upsilon _{\vartheta }]+l_{b_{1}}[b_{1k}]+ l_{b_{2}}[b_{2k}])^{2}}\\ =h^{\top }[E(S_{\vartheta }S_{\vartheta }^{\top })]^{-1}h^{\top }=h^{\top }\varSigma h. \end{aligned}$$

Thus the conclusion of the theorem follows by \(h^{\top }\left( (\hat{\beta _{1}}-\beta _{10})^{\top },(\hat{\beta _{2}}-\beta _{20}),\right. \left. ({\hat{\alpha }}-\alpha _{0})^{\top }, ({\hat{\eta }}-\eta _{0})^{\top }, (\hat{\sigma ^{2}}-\sigma _{0}^{2}) \right) =\gamma ({\hat{\theta }})-\gamma (\theta _{0}) ={\dot{\gamma }}(\theta _{0}) [{\hat{\theta }}-\theta _{0}]=\langle {\hat{\theta }}-\theta _{0},\upsilon ^{*}\rangle \) and the Cramér–Wold device. The semiparametric efficiency can be established by applying the result of Theorem 4 in Shen (1997). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, S., Wang, C. & Sun, J. An additive hazards cure model with informative interval censoring. Lifetime Data Anal 27, 244–268 (2021). https://doi.org/10.1007/s10985-021-09515-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10985-021-09515-7

Keywords

Navigation