Skip to main content
Log in

Function-based hypothesis testing in censored two-sample location-scale models

  • Published:
Lifetime Data Analysis Aims and scope Submit manuscript

Abstract

Function-based hypothesis testing in two-sample location-scale models has been addressed for uncensored data using the empirical characteristic function. A test of adequacy in censored two-sample location-scale models is lacking, however. A plug-in empirical likelihood approach is used to introduce a test statistic, which, asymptotically, is not distribution free. Hence for practical situations bootstrap is necessary for performing the test. A multiplier bootstrap and a model appropriate resampling procedure are given to approximate critical values from the null asymptotic distribution. Although minimum distance estimators of the location and scale are deployed for the plug-in, any consistent estimators can be used. Numerical studies are carried out that validate the proposed testing method, and real example illustrations are given.

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
Fig. 4

Similar content being viewed by others

References

  • Ahmed N, Subramanian S (2016) Semiparametric simultaneous confidence bands for the difference of survival functions. Lifetime Data Anal 22:504–530

    Article  MathSciNet  Google Scholar 

  • Andersen PK, Borgan O, Gill RD, Keiding N (1993) Statistical models based on counting processes. Springer, New York

    Book  Google Scholar 

  • Bhattacharya R, Subramanian SS (2014) Two-sample location-scale estimation from semiparametric random censorship models. J Multivar Anal 132:25–38

    Article  MathSciNet  Google Scholar 

  • Bickel P, Götze F, van Zwet W (1997) Resampling fewer than n observations: gains, losses and remedies for losses. Stat Sin 7:1–31

    MathSciNet  MATH  Google Scholar 

  • Bickel PJ, Sakov A (2008) On the choice of \(m\) in the \(m\) out of \(n\) bootstrap and confidence bounds for extrema. Stat Sin 18:967–985

    MathSciNet  MATH  Google Scholar 

  • Chen S, Dahl DB, Khan S (2005) Nonparametric identifiaction and estimation of a censored location-scale regression model. J Am Stat Assoc 100:212–221

    Article  Google Scholar 

  • Dette H, Heuchenne C (2012) Scale checks in censored regression. Scand J Stat 39:323–339

    Article  MathSciNet  Google Scholar 

  • Doksum K (1974) Empirical probability plots and statistical inference for nonlinear models in the two-sample case. Ann Stat 2:267–277

    Article  MathSciNet  Google Scholar 

  • Doksum K, Sievers GL (1976) Plotting with confidence: graphical comparisons of two populations. Biometrika 63:421–434

    Article  MathSciNet  Google Scholar 

  • Epps TW, Singleton KJ (1986) An omnibus test for the two-sample problem using the empirical characteristic function. J Stat Comput Simul 26:177–2003

    Article  MathSciNet  Google Scholar 

  • Fleming TR, Harrington DP (1991) Counting processes and survival analysis. Wiley, New York

    MATH  Google Scholar 

  • Gill RD, Johansen S (1990) A survey of product-integration with a view toward application in survival analysis. Ann Stat 15:1501–1555

    Article  MathSciNet  Google Scholar 

  • Hall P, Lombard F, Potgieter CJ (2013) A new approach to function based hypothesis testing in location-scale families. Technometrics 55:215–223

    Article  MathSciNet  Google Scholar 

  • Heuchenne C, Van Keilegom I (2010) Estimation in nonparametric location-scale regression models with censored data. Ann Inst Stat Math 62:439–463

    Article  MathSciNet  Google Scholar 

  • Hjort NL, McKeague IW, Van Keilegom I (2009) Extending the scope of empirical likelihood. Ann Stat 37:1079–1111

    Article  MathSciNet  Google Scholar 

  • Hsieh F (1996) Empirical process approach in a two-sample location-scale model with censored data. Ann Stat 24:2705–2719

    Article  MathSciNet  Google Scholar 

  • Kalbfleisch JD, Prentice RL (2002) The statistical analysis of failure time data. Wiley, New Jersey

    Book  Google Scholar 

  • Klein JP, Moeschberger ML (2003) Survival analysis. Techniques for censored and truncated data. Springer, New York

    MATH  Google Scholar 

  • Kyle R, Therneau T, Rajkumar V, Offord J, Larson D, Plevak M, Melton LJ III (2002) A long-term study of prognosis in monoclonal gammopathy of undetermined significance. New Engl J Med 346:564–569

    Article  Google Scholar 

  • Li G (1995) On nonparametric likelihood ratio estimation of survival probabilities for censored data. Stat Probab Lett 25:95–104

    Article  MathSciNet  Google Scholar 

  • Li G (2003) Nonparametric likelihood ratio goodness-of-fit tests for survival data. J Multivar Anal 86:166–182

    Article  MathSciNet  Google Scholar 

  • Li G, Wang Q-H (2003) Empirical likelihood regression analysis for right censored data. Stat Sin 13:51–68

    MathSciNet  MATH  Google Scholar 

  • Lv X, Zhang G, Ren G (2017) Gini index estimation for lifetime data. Lifetime Data Anal 23:275–304

    Article  MathSciNet  Google Scholar 

  • Lin DY, Wei LJ, Ying Z (1993) Checking the Cox model with cumulative sums of martingale-based residuals. Biometrika 80:557–572

    Article  MathSciNet  Google Scholar 

  • Lu HHS, Wells MT, Tiwari RC (1994) Inference for shift functions in the two-sample problem with right-censored data: with applications. J Am Stat Assoc 89:1017–1026

    Article  MathSciNet  Google Scholar 

  • McKeague IW, Zhao Y (2002) Simultaneous confidence bands for ratios of survival functions via empirical likelihood. Stat Probab Lett 60:405–415

    Article  MathSciNet  Google Scholar 

  • McKeague IW, Zhao Y (2005) Comparing distribution functions via empirical likelihood. Int J Biostat. https://doi.org/10.2202/1557-4679.1007

  • Mondal S, Subramanian S (2016) Simultaneous confidence bands for Cox regression from semiparametric random censorship. Lifetime Data Anal 22:122–144

    Article  MathSciNet  Google Scholar 

  • O’Brien PC (1988) Comparing two samples: extensions of the t, rank sum, and log-rank tests. J Am Stat Assoc 83:52–61

    Google Scholar 

  • Patil PN (1993) On the least squares cross-validation bandwidth in hazard rate estimation. Ann Stat 21:1792–1810

    Article  MathSciNet  Google Scholar 

  • Politis D, Romano J (1994) Large sample confidence regions on subsamples under minimal assumptions. Ann Stat 22:2031–2050

    Article  MathSciNet  Google Scholar 

  • Potgieter CJ, Lombard F (2016) Nonparametric two-sample estimation of location and scale parameters from empirical characteristic functions. J Stat Comput Simul 86:3225–3242

    Article  MathSciNet  Google Scholar 

  • Qin G, Tsao M (2003) Empirical likelihood inference for median regression models for censored survival data. J Multivar Anal 85:416–430

    Article  MathSciNet  Google Scholar 

  • Satten G, Datta S (2001) The Kaplan–Meier estimator as an inverse-probability-of-censoring weighted average. Am Stat 55:207–210

    Article  MathSciNet  Google Scholar 

  • Thomas DR, Grunkemeier GL (1975) Confidence interval estimation of survival probabilities for censored data. J Am Stat Assoc 70:865–871

    Article  MathSciNet  Google Scholar 

  • Ying Z, Jung S-H, Wei LJ (1995) Survival analysis with median regression models. J Am Stat Assoc 90:178–184

    Article  MathSciNet  Google Scholar 

  • Zhang Z, Li G (1996) A simple quantile approach to the two-sample location-scale problem with random censorship. J Nonparametric Stat 6:323–335

    Article  MathSciNet  Google Scholar 

  • Zhang Z, Yu Q (2002) A minimum distance estimation approach to the two-sample location-scale problem. Lifetime Data Anal 8:289–305

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The author expresses his sincere thanks to an Associate Editor and two reviewers whose comments improved the overall quality of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sundarraman Subramanian.

Appendix

Appendix

1.1 Review of the MDE of \(\varvec{\theta }\)

See also Bhattacharya and Subramanian (2014), where semiparametric estimation in two-sample location-scale models is investigated. Let \(\mathcal{Q}_{\varvec{\theta }}(s) = {\hat{Q}}_2(s)-\varphi _{\varvec{\theta }}({\hat{Q}}_1(s))\), where \({\hat{Q}}_i(s), i=1,2\), are the Kaplan–Meier quantiles. Let G(s) denote a positive measure on (0, 1). To obtain the MDE \({\hat{\varvec{\theta }}}\), one minimizes \(\mathcal{S}(\varvec{\theta })\), a Cramér-von Mises type discrepancy (Zhang and Yu 2002), defined as

$$\begin{aligned} \mathcal{S}(\varvec{\theta })= & {} \int \mathcal{Q}^2_{\varvec{\theta }}(s)dG(s):= \langle \mathcal{Q}^2_{\varvec{\theta }},G\rangle . \end{aligned}$$
(A.1)

Define \({\varvec{d}}=(d_1,d_2)^T\) where \(d_l=\langle {\hat{Q}}_1^{l-1}{\hat{Q}}_2, G\rangle \), \(l=1,2\), and let \({\varvec{A}}=(A_{il})_{2\times 2}\) where \(A_{il}=\langle {\hat{Q}}_1^{i-1}{\hat{Q}}_1^{l-1}, G\rangle \). Minimizing \(\mathcal{S}(\varvec{\theta })\) given by Eq. (A.1) yields \({\varvec{A}}\varvec{\theta }= {\varvec{d}}\). When \({\hat{Q}}_1\) is not a constant, A is non-singular (Zhang and Yu 2002). Then \({\hat{\varvec{\theta }}}={\varvec{A}}^{-1}{\varvec{d}}\). More specifically, let

$$\begin{aligned} {\hat{\varvec{U}}}\ =\ ({\hat{U}}_1, {\hat{U}}_2, {\hat{U}}_3, {\hat{U}}_4)^T= & {} \left( \langle {\hat{Q}}_1, G\rangle , \langle {\hat{Q}}_1^2, G\rangle , \langle {\hat{Q}}_2, G\rangle , \langle {\hat{Q}}_1{\hat{Q}}_2, G\rangle \right) ^T.\nonumber \\ \end{aligned}$$
(A.2)

Then \({\hat{\varvec{\theta }}}\equiv {\varvec{g}}({\hat{\varvec{U}}})=(g_1({\hat{\varvec{U}}}),g_2({\hat{\varvec{U}}}))^T\), where \(g_1\) and \(g_2\) are functions of \({\varvec{u}}=(u_1, u_2,u_3,u_4)^T\):

$$\begin{aligned} g_1({\varvec{u}}) = \frac{u_2u_3-u_1u_4}{cu_2-u_1^2}; \quad g_2({\varvec{u}}) = \frac{cu_4-u_1u_3}{cu_2-u_1^2}, \end{aligned}$$

where \(c=\langle 1, G\rangle \). Zhang and Yu (2002) derived a large-sample representation for \({\hat{\varvec{\theta }}}\). To develop an equivalent martingale representation for \({\hat{\varvec{\theta }}}\), let \({\varvec{D}}=[D_{il}]_{2\times 4}=\frac{d{\varvec{g}}}{d{\varvec{u}}}\vert _{{\varvec{u}}={\varvec{U}}}\), where

$$\begin{aligned} {\varvec{U}}= & {} (U_1, U_2, U_3, U_4)^T = \left( \langle Q_1, G\rangle , \langle Q_1^2, G\rangle , \langle Q_2, G\rangle , \langle Q_1Q_2, G\rangle \right) ^T; \\ \frac{d{\varvec{g}}}{d{\varvec{u}}}= & {} \begin{pmatrix} \frac{-u_4(cu_2-{u_1}^2)+2u_1(u_2u_3-u_1u_4)}{(cu_2-{u_1}^2)^2} &{}\quad \frac{u_3(cu_2-{u_1}^2)-c(u_2u_3-u_1u_4)}{(cu_2-{u_1}^2)^2} &{}\quad \frac{u_2}{cu_2-{u_1}^2} &{}\quad \frac{-u_1}{cu_2-{u_1}^2} \\ \frac{-u_3(cu_2-{u_1}^2)+2u_1(-u_1u_3+cu_4)}{(cu_2-{u_1}^2)^2} &{}\quad \frac{c(u_1u_3-cu_4)}{(cu_2-{u_1}^2)^2} &{}\quad \frac{-u_1}{cu_2-{u_1}^2} &{}\quad \frac{c}{cu_2-{u_1}^2} \end{pmatrix}. \end{aligned}$$

Then \({\hat{\varvec{\theta }}}-\varvec{\theta }_0 = {\varvec{D}}({\hat{\varvec{U}}}-{\varvec{U}}) + o_{{\mathbb {P}}}(n^{-1/2})\). Let \(\tau _{i,l}\) and \(\tau _{i,u}\) denote the infimum and supremum of the support set for \(G\circ F_i(\cdot )\equiv G(F_i(\cdot ))\). To deal with \({\hat{\varvec{U}}}-{\varvec{U}}\), for \(s \in (\tau _{i,l}, \tau _{i,u})\), define

$$\begin{aligned} e_i(s)= & {} \left\langle S_iI_{(s,\tau _{i,u})}/f_i, G\circ F_i\right\rangle ,\quad i=1,2, \\ {\tilde{e}}_{il}(s)= & {} \left\langle (Q_l\circ F_i)S_iI_{(s,\tau _{i,u})}/f_i, G\circ F_i\right\rangle ,\quad i=1,2,\;\; l=1,2. \end{aligned}$$

For the first component of \({\hat{\varvec{U}}}-{\varvec{U}}\), say, apply Eq. (2.4), then apply Theorem II.6.2 of Andersen et al. (1993) for \({\hat{F}}_1-F_1\), and finally interchange the order of integration to obtain

$$\begin{aligned} \langle {\hat{Q}}_1 - Q_1, G\rangle= & {} -\int e_1(s)d\left( {\hat{\Lambda }}_1(s) - \Lambda _1(s)\right) + o_{{\mathbb {P}}}(n^{-1/2}). \end{aligned}$$

The other components of \({\hat{\varvec{U}}}-{\varvec{U}}\) can be treated likewise. Let \(\varvec{W}(s)\equiv [W_{kl}(s)]_{4\times 2}\), where \(W_{11}(s)=e_1(s), W_{21}(s)=2{\tilde{e}}_{11}(s)\), \(W_{32}(s)=e_2(s)\), \(W_{41}(s)={\tilde{e}}_{12}(s)\), \(W_{42}(s)={\tilde{e}}_{21}(s)\), and \(W_{12}(s)=W_{22}(s)=W_{31}(s)=0\). It follows that \({\hat{\varvec{\theta }}}- \varvec{\theta }_0\) has the large-sample representation

$$\begin{aligned} {\hat{\varvec{\theta }}}-\varvec{\theta }_0= & {} -{\varvec{D}}\int \varvec{W}(s)\, d \begin{pmatrix} \hat{\Lambda }_1(s) - \Lambda _1(s) \\ \hat{\Lambda }_2(s) - \Lambda _2(s) \end{pmatrix} + o_{{\mathbb {P}}}(n^{-1/2}) \nonumber \\= & {} -{\varvec{D}}\int \varvec{W}(s)\, \begin{pmatrix} dM_{1\cdot }(s)/(n_1y_1(s)) \\ dM_{2\cdot }(s)/(n_2y_2(s)) \end{pmatrix} + o_{{\mathbb {P}}}(n^{-1/2}) . \end{aligned}$$
(A.3)

Note that the last step above is obtained by applying Eq. (2.1). We will exploit Eq. (A.3) when we derive the large-sample null distribution of the plug-in EL.

Remark A.1

For a purely location model, \(\theta _2=1\). Then \(\theta _1\equiv \theta \) is the only parameter and its MDE is \({\hat{\varvec{\theta }}}\equiv {\hat{\theta }} = \langle {\hat{Q}}_2 - {\hat{Q}}_1, G\rangle \). We can write \(\hat{\theta }- \theta _0={\varvec{D}}^T({\hat{\varvec{U}}}-{\varvec{U}})\), where \(D=(-1,1)^T\) and \({\hat{\varvec{U}}}-{\varvec{U}}=(\langle {\hat{Q}}_1 - Q_1, G\rangle , \langle {\hat{Q}}_2 - Q_2, G\rangle )^T\). Then \({\hat{\theta }} - \theta _0\) has the large-sample representation

$$\begin{aligned} {\hat{\theta }}-\theta _0= & {} \frac{1}{n_1}\sum _{j=1}^{n_1} \int \frac{e_1(s)}{y_1(s)}dM_{1j}(s) - \frac{1}{n_2}\sum _{j=1}^{n_2} \int \frac{e_2(s)}{y_2(s)}dM_{2j}(s) + o_{{\mathbb {P}}}(n^{-1/2}),\nonumber \\ \end{aligned}$$
(A.4)

which one can exploit when testing the adequacy of a location model, see Remark 7.

Recall that \({\tilde{\kappa }}_i(t)\) is the number of distinct uncensored lifetimes by time t for each sample. In the notation to handle ties, the two cumulative hazard estimators are given by

$$\begin{aligned} \hat{\Lambda }_i(t)= & {} \sum _{j=1}^{\tilde{\kappa }_i(t)}\left( \frac{d_{ij} }{r_{ij}}\right) ;\quad {\hat{\zeta }}_i(t)\ =\ - \sum _{j=1}^{\tilde{\kappa }_i(t)}\log \left( 1 - \frac{d_{ij}}{r_{ij}}\right) ,\;\; i=1,2. \end{aligned}$$
(A.5)

Let \(\Vert h\Vert _{\alpha _1}^{\alpha _2} = \sup _{t\in [\alpha _1,\alpha _2]}|h(t)|\). Lemma 1 gives the asymptotic equivalence between \({\hat{\Lambda }}_i\) and \({\hat{\zeta }}_i\).

Lemma 1

Assume that \(S_i(t), i=1,2\), are continuous. Then \(\Vert \hat{\Lambda }_i-{\hat{\zeta }}_i\Vert _{\alpha _1}^{\alpha _2} = o_{{\mathbb {P}}}(n^{-1/2})\).

Proof

Following McKeague and Zhao (2005), see their Eq. (A.6), we can show that

$$\begin{aligned} n_i^{1/2}|\hat{\Lambda }_i(t) - {\hat{\zeta }}_i(t)|\le & {} \left\{ \max _{j \le \tilde{\kappa }_i(t)}\left( \frac{n_i^{1/2}d_{ij}}{r_{ij}}\right) \right\} \times \left\{ \sum _{j=1}^{\tilde{\kappa }_i(t)}\left( \frac{d_{ij}}{r_{ij}}\right) \right\} . \end{aligned}$$
(A.6)

By continuity of \(S_i\), the first quantity on the right hand side (RHS) of inequality (A.6) is

$$\begin{aligned} \max _{j \le \tilde{\kappa }_i(t)}\left( n_i^{1/2}/r_{ij}\right)\le & {} n_i^{-1/2} \Vert n_i/Y_{i\cdot }\Vert _{0}^{\alpha _2}\ =\ O\left( n_i^{-1/2}\right) O_{{\mathbb {P}}}(1) \ =\ o_{{\mathbb {P}}}(1). \end{aligned}$$
(A.7)

From Eq. (A.5), the second quantity on the RHS of inequality (A.6) is bounded above by \(\hat{\Lambda }_i(\alpha _2)=O_{{\mathbb {P}}}(1)\). Combined with inequalities (A.6) and (A.7), this completes the proof. \(\square \)

Lemma 2

The Lagrange multiplier, \({\hat{\lambda }}_{\varvec{\theta }}(t)\), solving Eq. (2.10), satisfies

$$\begin{aligned} |{\hat{\lambda }}_{\varvec{\theta }}(t)|\le & {} \frac{n_2({\hat{\zeta }}_1(t) - \hat{\Lambda }_2(\varphi _{\varvec{\theta }}(t)))}{\hat{\Lambda }_2(\varphi _{\varvec{\theta }}(t)))},\quad \mathrm{when\ }{\hat{\lambda }}_{\varvec{\theta }}(t) < 0. \end{aligned}$$
(A.8)
$$\begin{aligned} |{\hat{\lambda }}_{\varvec{\theta }}(t)|\le & {} \frac{n_1({\hat{\zeta }}_2(\varphi _{\varvec{\theta }}(t)) - \hat{\Lambda }_1(t))}{\hat{\Lambda }_1(t)},\quad \mathrm{when\ }{\hat{\lambda }}_{\varvec{\theta }}(t) > 0. \end{aligned}$$
(A.9)

Proof

From Eq. (2.10), \({\hat{\lambda }}_{\varvec{\theta }}(t)\) satisfies the equation \(I_1(t)-I_2(t)=0\), where

$$\begin{aligned} I_1(t)= \sum _{j=1}^{{\tilde{\kappa }}_1(t)}\log \left( 1- \frac{d_{1j}}{r_{1j}-{\hat{\lambda }}_{\varvec{\theta }}(t)}\right) ; \quad I_2(t)=\sum _{j=1}^{{\tilde{\kappa }}_2(\varphi _{\varvec{\theta }}(t))}\log \left( 1-\frac{d_{2j}}{r_{2j}+{\hat{\lambda }}_{\varvec{\theta }}(t)}\right) . \end{aligned}$$

Consider \({\hat{\lambda }}_{\varvec{\theta }}(t)<0\). Follow Li (1995), leading to his Eq. (2.12) on p. 101, to obtain

$$\begin{aligned} -I_2(t)\ge & {} \sum _{j=1}^{{\tilde{\kappa }}_2(\varphi _{\varvec{\theta }}(t))}\frac{d_{2j}}{r_{2j}}\left( \frac{n_2}{n_2-|{\hat{\lambda }}_{\varvec{\theta }}(t)|}\right) \ =\ \hat{\Lambda }_2(\varphi _{\varvec{\theta }}(t))\left( \frac{1}{1-|{\hat{\lambda }}_{\varvec{\theta }}(t)|/n_2}\right) . \end{aligned}$$

Applying some elementary calculations (cf. Ahmed and Subramanian, 2015), we can show that, for large enough \(n_2\), \(0<|{\hat{\lambda }}_{\varvec{\theta }}(t)|/n_2<1\) almost surely. Now, as in McKeague and Zhao (2005), use the fact that \(1/(1-x)\ge 1+x\) for \(0<x<1\), to obtain

$$\begin{aligned} -I_2(t)\ge & {} \hat{\Lambda }_2(\varphi _{\varvec{\theta }}(t)) + \frac{\hat{\Lambda }_2(\varphi _{\varvec{\theta }}(t))|{\hat{\lambda }}_{\varvec{\theta }}(t)|}{n_2}. \end{aligned}$$
(A.10)

Since \(-{\hat{\lambda }}_{\varvec{\theta }}(t)>0\), follow Li (1995) leading to his Eq. (2.13) on p. 101, and obtain

$$\begin{aligned} I_1(t)\ge & {} -\sum _{j=1}^{{\tilde{\kappa }}_1(t)}\frac{d_{1j}}{r_{1j}}\left( \frac{n_1}{n_1+|{\hat{\lambda }}_{\varvec{\theta }}(t)|}\right) + \sum _{j=1}^{{\tilde{\kappa }}_1(t)}\left( \log \left( 1-\frac{d_{1j}}{r_{1j}}\right) +\frac{d_{1j}}{r_{1j}}\right) \nonumber \\= & {} -\hat{\Lambda }_1(t)\left( \frac{n_1}{n_1+|{\hat{\lambda }}_{\varvec{\theta }}(t)|}\right) - {\hat{\zeta }}_1(t) + \hat{\Lambda }_1(t) \nonumber \\= & {} \hat{\Lambda }_1(t)\left( \frac{|{\hat{\lambda }}_{\varvec{\theta }}(t)|}{n_1+|{\hat{\lambda }}_{\varvec{\theta }}(t)|}\right) - {\hat{\zeta }}_1(t)\ \ge \ - {\hat{\zeta }}_1(t). \end{aligned}$$
(A.11)

Combine inequalities (A.10) and (A.11) to obtain inequality (A.8) from

$$\begin{aligned} 0= & {} I_1(t)-I_2(t) \ \ge \ - {\hat{\zeta }}_1(\varphi _{\varvec{\theta }}(t)) + \hat{\Lambda }_2(t) + \frac{\hat{\Lambda }_2(t)|{\hat{\lambda }}_{\varvec{\theta }}(t)|}{n_2}. \end{aligned}$$

When \({\hat{\lambda }}_{\varvec{\theta }}(t)>0\), the above approach can be repeated by applying Li’s (1995) technique in reverse order. First obtain the inequality

$$\begin{aligned} I_2(t)\ge & {} -\hat{\Lambda }_2(\varphi _{\varvec{\theta }}(t))\left( \frac{n_2}{n_2+|{\hat{\lambda }}_{\varvec{\theta }}(t)|}\right) - {\hat{\zeta }}_2(\varphi _{\varvec{\theta }}(t)) + \hat{\Lambda }_2(\varphi _{\varvec{\theta }}(t)) \nonumber \\= & {} \hat{\Lambda }_2(\varphi _{\varvec{\theta }}(t))\left( \frac{|{\hat{\lambda }}_{\varvec{\theta }}(t)|}{n_2+|{\hat{\lambda }}_{\varvec{\theta }}(t)|}\right) - {\hat{\zeta }}_2(\varphi _{\varvec{\theta }}(t))\ \ge \ -{\hat{\zeta }}_2(\varphi _{\varvec{\theta }}(t)). \end{aligned}$$
(A.12)

Then obtain the inequality

$$\begin{aligned} -I_1(t)\ge & {} \hat{\Lambda }_1(t)\left( \frac{n_1}{n_1-|{\hat{\lambda }}_{\varvec{\theta }}(t)|}\right) \ \ge \ \hat{\Lambda }_1(t)\left( 1+|{\hat{\lambda }}_{\varvec{\theta }}(t)|/n_1\right) . \end{aligned}$$
(A.13)

Combine inequalities (A.12) and (A.13) to obtain inequality (A.9) from

$$\begin{aligned} 0= & {} I_2(t)-I_1(t) \ge \ -{\hat{\zeta }}_2(\varphi _{\varvec{\theta }}(t)) + \hat{\Lambda }_1(t) + \frac{\hat{\Lambda }_1(t)|{\hat{\lambda }}_{\varvec{\theta }}(t)|}{n_1}. \end{aligned}$$

\(\square \)

Remark A.3

Define \(\mu (\cdot )=-\log (\cdot )\), which is a continuously differentiable function. Then \({\hat{\zeta }}_2(t)=\mu ({\hat{S}}_2(t))\). It follows by Lemma 1 of Ying et al. (1995) that

$$\begin{aligned} \sup _{t \in [\alpha _1,\alpha _2]}\left| {\hat{\zeta }}_2(\varphi _{{\hat{\varvec{\theta }}}}(t))+\log S_2(\varphi _{{\hat{\varvec{\theta }}}}(t))-{\hat{\zeta }}_2(\varphi _{\varvec{\theta }_0}(t)) - \log S_2(\varphi _{\varvec{\theta }_0}(t))\right|= & {} o_{{\mathbb {P}}}(n^{-1/2}).\nonumber \\ \end{aligned}$$
(A.14)

Lemma 3

The estimate \({\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t)\), solving Eq. (2.10) with \(\varvec{\theta }={\hat{\varvec{\theta }}}\), satisfies \(\Vert {\hat{\lambda }}_{{\hat{\varvec{\theta }}}}\Vert _{\alpha _1}^{\alpha _2} = O_{{\mathbb {P}}}(n^{1/2})\).

Proof

The denominators on the RHS of inequalities (A.8) and (A.9) are each \(O_{{\mathbb {P}}}(1)\), uniformly for \(t\in [\alpha _1,\alpha _2]\). This follows from \(\Vert \hat{\Lambda }_i-\Lambda _i\Vert _0^{\tau _i}=o_{{\mathbb {P}}}(1)\), where \(\tau _1=\alpha _2\) and \(\tau _2=\varphi _{\varvec{\theta }_0}(\alpha _2)\), see Sect. 2.2. Furthermore, the denominators on the RHS of inequalities (A.8) and (A.9) are bounded away from 0 for \(t\in [\alpha _1, \alpha _2]\). It remains to show that the numerators on the RHS of inequalities (A.8) and (A.9) are each \(O_{{\mathbb {P}}}(n^{1/2})\) uniformly for \(t\in [\alpha _1,\alpha _2]\). Consider the numerator on the RHS of inequality (A.8). Note that under the null hypothesis, \(-\log (S_2(\varphi _{\varvec{\theta }_0}(t))=\Lambda _2(\varphi _{\varvec{\theta }_0}(t))=\Lambda _1(t)\). Therefore,

$$\begin{aligned} {\hat{\zeta }}_1(t) - \hat{\Lambda }_2(\varphi _{{\hat{\varvec{\theta }}}}(t))= & {} \left[ {\hat{\zeta }}_1(t)-{\hat{\Lambda }}_1(t)\right] +\left[ {\hat{\Lambda }}_1(t)-\Lambda _1(t)\right] \\&-\,\left[ \hat{\Lambda }_2(\varphi _{{\hat{\varvec{\theta }}}}(t))-{\hat{\zeta }}_2(\varphi _{{\hat{\varvec{\theta }}}}(t))\right] \\&-\,\left[ {\hat{\zeta }}_2(\varphi _{{\hat{\varvec{\theta }}}}(t)) + \log S_2(\varphi _{{\hat{\varvec{\theta }}}}(t)) - {\hat{\zeta }}_2(\varphi _{\varvec{\theta }_0}(t)) - \log S_2(\varphi _{\varvec{\theta }_0}(t))\right] \\&+ \left[ \log S_2(\varphi _{{\hat{\varvec{\theta }}}}(t)) - \log S_2(\varphi _{\varvec{\theta }_0}(t))\right] \\&-\,\left[ {\hat{\zeta }}_2(\varphi _{\varvec{\theta }_0}(t)) + \log S_2(\varphi _{\varvec{\theta }_0}(t))\right] . \end{aligned}$$

By Lemma 1, \(\Vert {\hat{\zeta }}_1-{\hat{\Lambda }}_1\Vert _{\alpha _1}^{\alpha _2}=o_{{\mathbb {P}}}(n^{-1/2})\). Also, it is standard that \(\Vert {\hat{\Lambda }}_1-\Lambda _1\Vert _{\alpha _1}^{\alpha _2}=O_{{\mathbb {P}}}(n^{-1/2})\). By Remark A.2, the third expression within square brackets is \(o_{{\mathbb {P}}}(n^{-1/2})\) uniformly for \(t \in [\alpha _1, \alpha _2]\). By Remark A.3, the fourth expression within square brackets is \(o_{{\mathbb {P}}}(n^{-1/2})\) uniformly for \(t \in [\alpha _1, \alpha _2]\). By a Taylor expansion around \(\varphi _{\varvec{\theta }_0}(t)\) and the \(n^{1/2}\) consistency of the MDE \({\hat{\varvec{\theta }}}\) of \(\varvec{\theta }\), the fifth expression within square brackets is seen to be \(O_{{\mathbb {P}}}(n^{-1/2})\) uniformly for \(t\in [\alpha _1, \alpha _2]\). By a Taylor expansion, the last expression, which equals \(-\log {\hat{S}}_2(\varphi _{\varvec{\theta }_0}(t)) + \log S_2(\varphi _{\varvec{\theta }_0}(t))\), is also \(O_p(n^{-1/2})\) uniformly for \(t\in [\alpha _1, \alpha _2]\). Likewise, it can be shown that the numerator on the RHS of inequality (A.9) is also \(O_{{\mathbb {P}}}(n^{-1/2})\) uniformly for \(t\in [\alpha _1, \alpha _2]\). \(\square \)

In the proof of Theorem 1, we will show that \(-2R_{{\hat{\varvec{\theta }}}}(t)\) is asymptotically equivalent to the square of \(n^{1/2}{\hat{{\mathbb {V}}}}(t)\) scaled by the reciprocal of \(\sigma _{\mathrm{c}}^2(t)\) defined by Eq. (2.14),

Lemma 4

Under the null hypothesis, \({\hat{{\mathbb {V}}}}(t)\) is asymptotically linear with influence function \(\sum _{i=1}^2 J_{i1}(t)/\sqrt{\rho _i}\), where \(J_{11}(t)\) and \(J_{21}(t)\) are given by Eqs. (2.15) and (2.16) respectively.

Proof

First apply a Taylor’s expansion for \(\log {\hat{S}}_1(t)\) about \(S_1(t)\). Then apply the Duhamel equation (Andersen et al. 1993) and Eq. (2.1) to yield an asymptotic representation for the first term of \({\hat{{\mathbb {V}}}}(t)\) as shown below:

$$\begin{aligned} -{\hat{\zeta }}_1(t)= & {} \log S_1(t) + \left( \log {\hat{S}}_1(t) - \log S_1(t)\right) \nonumber \\= & {} \log S_1(t) + \frac{{\hat{S}}_1(t) - S_1(t)}{S_1(t)} + o_{{\mathbb {P}}}(n^{-1/2}) \nonumber \\= & {} -\Lambda _1(t) - \frac{1}{n_1}\sum _{j=1}^{n_1} \int _0^t \frac{1}{y_1(u)}dM_{1j}(u) + o_{{\mathbb {P}}}(n^{-1/2}). \end{aligned}$$
(A.15)

For the second term of \({\hat{{\mathbb {V}}}}(t)\), applying Eq. (A.14) of Remark A.3, it follows that

$$\begin{aligned} {\hat{\zeta }}_2(\varphi _{{\hat{\varvec{\theta }}}}(t))= & {} -\log S_2(\varphi _{{\hat{\varvec{\theta }}}}(t)) - \log {\hat{S}}_2(\varphi _{\varvec{\theta }_0}(t)) + \log (S_2(\varphi _{\varvec{\theta }_0}(t)) + o_{{\mathbb {P}}}(n^{-1/2}) \nonumber \\= & {} \Lambda _2(\varphi _{{\hat{\varvec{\theta }}}}(t)) - \Lambda _2(\varphi _{\varvec{\theta }_0}(t)) - \log {\hat{S}}_2(\varphi _{\varvec{\theta }_0}(t)) + o_{{\mathbb {P}}}(n^{-1/2}). \end{aligned}$$
(A.16)

Applying the delta method to the first and second terms on the RHS of Eq. (A.16) yields

$$\begin{aligned} \Lambda _2(\varphi _{{\hat{\varvec{\theta }}}}(t)) - \Lambda _2(\varphi _{\varvec{\theta }_0}(t))= & {} \lambda _2(\varphi _{\varvec{\theta }_0}(t))\, (\varphi _{{\hat{\varvec{\theta }}}}(t)-\varphi _{\varvec{\theta }_0}(t)) + o_{{\mathbb {P}}}(n^{-1/2}) \nonumber \\= & {} \lambda _2(\varphi _{\varvec{\theta }_0}(t))\,{\varvec{c}}_t^T \left( {\hat{\varvec{\theta }}}-\varvec{\theta }_0\right) + o_{{\mathbb {P}}}(n^{-1/2}) \nonumber \\= & {} -\lambda _2(\varphi _{\varvec{\theta }_0}(t))\,{\varvec{c}}_t^T {\varvec{D}}\int \varvec{W}(u)\,\begin{pmatrix} dM_{1\cdot }(u)/(n_1y_1(u)) \\ dM_{2\cdot }(u)/(n_2y_2(u)) \end{pmatrix} \nonumber \\&+\, o_{{\mathbb {P}}}(n^{-1/2}),\quad \end{aligned}$$
(A.17)

the last step that leads to Eq. (A.17) following from Eq. (A.3), where \(\lambda _2\) is the hazard function associated with \(F_2\) and \({\varvec{c}}_t=(1,t)^T\). The third term on the RHS of Eq. (A.16) can be treated exactly the way the first term of \({\hat{{\mathbb {V}}}}(t)\) yielded Eq. (A.15):

$$\begin{aligned} - \log {\hat{S}}_2(\varphi _{\varvec{\theta }_0}(t))= & {} -\log S_2(\varphi _{\varvec{\theta }_0}(t)) - \log {\hat{S}}_2(\varphi _{\varvec{\theta }_0}(t)) + \log S_2(\varphi _{\varvec{\theta }_0}(t)) \nonumber \\= & {} \Lambda _2(\varphi _{\varvec{\theta }_0}(t)) -\frac{{\hat{S}}_2(\varphi _{\varvec{\theta }_0}(t)) - S_2(\varphi _{\varvec{\theta }_0}(t))}{S_2(\varphi _{\varvec{\theta }_0}(t))} + o_{{\mathbb {P}}}(n^{-1/2})\nonumber \\= & {} \Lambda _2(\varphi _{\varvec{\theta }_0}(t)) + \frac{1}{n_2}\sum _{j=1}^{n_2} \int _0^{\varphi _{\varvec{\theta }_0}(t)} \frac{1}{y_2(u)}dM_{2j}(u) + o_{{\mathbb {P}}}(n^{-1/2}).\nonumber \\ \end{aligned}$$
(A.18)

Since \(\Lambda _1(t)=\Lambda _2(\varphi _{\varvec{\theta }_0}(t))\), from Eqs. (2.13)–(A.18) we obtain

$$\begin{aligned} n^{1/2}{\hat{{\mathbb {V}}}}(t)=\sum _{i=1}^2 \left\{ n_i^{-1/2} \sum _{j=1}^{n_i}\frac{J_{ij}(t)}{\sqrt{\rho _i}}\right\} + o_{{\mathbb {P}}}(1). \end{aligned}$$

\(\square \)

1.2 Proof of Theorem 1

Let \(\breve{\kappa }_1(t)={\tilde{\kappa }}_1(t)\), \(\breve{\kappa }_2(t)={\tilde{\kappa }}_2(\varphi _{{\hat{\varvec{\theta }}}}(t))\). From Eq. (2.10), \(h_1(-{\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t))-h_2({\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t))=0\), where

$$\begin{aligned} h_i(\lambda )= & {} \sum _{j=1}^{\breve{\kappa }_i(t)}\log \left( 1 - \frac{d_{ij}}{r_{ij} + \lambda }\right) . \end{aligned}$$
(A.19)

Note that \(h_1(0) = - {\hat{\zeta }}_1(t)\) and \(h_2(0)=- {\hat{\zeta }}_2(\varphi _{{\hat{\varvec{\theta }}}}(t))\) [cf. Eq. (A.5)]. Note also that

$$\begin{aligned} h_i^{'}(\lambda )= & {} \sum _{j=1}^{\breve{\kappa }_i(t)}\frac{d_{ij}}{(r_{ij}+\lambda )(r_{ij}+\lambda - d_{ij})},\quad h_i^{''}(\lambda )= \sum _{j=1}^{\breve{\kappa }_i(t)}\frac{d_{ij}(2(r_{ij} +\lambda )- d_{ij} )}{(r_{ij}+\lambda )^2(r_{ij}+\lambda -d_{ij})^2}. \end{aligned}$$

Clearly, \(n_1h_1^{'}(0)={\hat{\sigma }}_1^2(t)\) and \(n_2h_2^{'}(0)={\hat{\sigma }}_2^2(\varphi _{{\hat{\varvec{\theta }}}}(t))\), see Eq. (2.2). Let \(|{\hat{\xi }}_i|\le |{\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t)|, i=1,2\). Taylor’s expansion about 0 yields

$$\begin{aligned} h_1(-{\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t))= & {} - {\hat{\zeta }}_1(t) -\frac{{\hat{\sigma }}_1^2(t){\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t)}{n_1} + \frac{1}{2}h^{''}_1({\hat{\xi }}_1)({\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t))^2, \end{aligned}$$
(A.20)
$$\begin{aligned} h_2\left( {\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t)\right)= & {} - {\hat{\zeta }}_2(\varphi _{{\hat{\varvec{\theta }}}}(t)) + \frac{{\hat{\sigma }}_2^2(\varphi _{{\hat{\varvec{\theta }}}}(t)){\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t)}{n_2} + \frac{1}{2}h^{''}_1({\hat{\xi }}_2)\left( {\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t)\right) ^2. \end{aligned}$$
(A.21)

Applying the Glivenko–Cantelli lemma to \(r_{ij}\), we have \(\Vert h_i''({\hat{\xi }}_i)\Vert _{\alpha _1}^{\alpha _2}=O_{{\mathbb {P}}}(n_i^{-2})\). Therefore, by Lemma 3, it follows that \(\Vert h^{''}_i({\hat{\xi }}_i)({\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t))^2\Vert _{\alpha _1}^{\alpha _2}=O_{{\mathbb {P}}}(n_i^{-1})=O_{{\mathbb {P}}}(n^{-1})\). Recalling \({\hat{{\mathbb {V}}}}(t)\) defined by Eq. (2.13), it follows from Eqs. (A.20), (A.21), and (2.14) that

$$\begin{aligned} 0= & {} -{\hat{\zeta }}_1(t) + {\hat{\zeta }}_2(\varphi _{{\hat{\varvec{\theta }}}}(t)) - {\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t)\left( \frac{{\hat{\sigma }}_1^2(t)}{n_1} +\frac{{\hat{\sigma }}_2^2(\varphi _{{\hat{\varvec{\theta }}}}(t))}{n_2}\right) + O_{{\mathbb {P}}}\left( \frac{1}{n}\right) \\= & {} -{\hat{{\mathbb {V}}}}(t) - \frac{{\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t) {\hat{\sigma }}^2_{\mathrm{c}}(t)}{n} + O_{{\mathbb {P}}}\left( \frac{1}{n}\right) . \end{aligned}$$

Solving for \({\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t)\), we obtain

$$\begin{aligned} {\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t)= & {} \frac{n}{{\hat{\sigma }}^2_{\mathrm{c}}(t)}\left\{ -{\hat{{\mathbb {V}}}}(t)+O_{{\mathbb {P}}}\left( \frac{1}{n}\right) \right\} . \end{aligned}$$
(A.22)

To complete the proof of Theorem 1, consider Eq. (2.12). Using Taylor expansions of \(\log (1+x)\) and \(\log (1-x)\) about 0 and applying Eq. (A.22), the leading term of \(-2R_{{\hat{\varvec{\theta }}}}(t)\) equals

$$\begin{aligned} \left( {\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t)\right) ^2\left[ \sum _{i=1}^2\sum _{j=1}^{\breve{\kappa }_i(t)}\frac{d_{ij}}{r_{ij}(r_{ij}-d_{ij})}\right]= & {} \left( {\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t)\right) ^2\left( \frac{{\hat{\sigma }}_1^2(t)}{n_1} +\frac{{\hat{\sigma }}_2^2(\varphi _{{\hat{\varvec{\theta }}}}(t))}{n_2}\right) \nonumber \\= & {} \frac{n}{\left( {\hat{\sigma }}^2_{\mathrm{c}}(t)\right) ^2}\left\{ {\hat{{\mathbb {V}}}}(t)\right. \nonumber \\&\left. +O_{{\mathbb {P}}}\left( \frac{1}{n}\right) \right\} ^2\left( \frac{{\hat{\sigma }}_1^2(t)}{n_1/n} +\frac{{\hat{\sigma }}_2^2(\varphi _{{\hat{\varvec{\theta }}}}(t))}{n_2/n}\right) \nonumber \\= & {} \frac{1}{\left( {\hat{\sigma }}^2_{\mathrm{c}}(t)\right) ^2}\left\{ n^{1/2}{\hat{{\mathbb {V}}}}(t)+o_{{\mathbb {P}}}(1)\right\} ^2 {\hat{\sigma }}^2_{\mathrm{c}}(t) + o_{{\mathbb {P}}}(1) \nonumber \\= & {} \frac{1}{\sigma ^2_{\mathrm{c}}(t)}\left\{ n^{1/2}{\hat{{\mathbb {V}}}}(t)\right\} ^2 + o_{{\mathbb {P}}}(1), \end{aligned}$$
(A.23)

uniformly over \([\alpha _1,\alpha _2]\). The subsequent terms of \(-2R(t)\) are proportional to

$$\begin{aligned} \left( {\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t)\right) ^l\left[ \sum _{i=1}^2\sum _{j=1}^{\breve{\kappa }_i(t)}\left( \frac{1}{(r_{ij}-d_{ij})^{l-1}} - \frac{1}{r_{ij}^{l-1}}\right) \right] , \quad l=3, 4, \ldots , \end{aligned}$$

each of which is \(o_{{\mathbb {P}}}(1)\), uniformly for \(t\in [\alpha _1,\alpha _2]\). For example, when \(l=3\), we have

$$\begin{aligned} \frac{2}{3}\left( {\hat{\lambda }}_{{\hat{\varvec{\theta }}}}(t)\right) ^3\sum _{j=1}^{\breve{\kappa }_i(t)}\left( \frac{1}{r_{ij}-d_{ij}} - \frac{1}{r_{ij}}\right) \left( \frac{1}{r_{ij} - d_{ij}} + \frac{1}{r_{ij}}\right) ,\quad i=1,2, \end{aligned}$$

which, uniformly for \(t\in [\alpha _1,\alpha _2]\), equals

$$\begin{aligned} O_{{\mathbb {P}}}(n^{3/2})O_{{\mathbb {P}}}(n_i^{-1})\sum _{j=1}^{\breve{\kappa }_1(t)}\left( \frac{1}{r_{1j} - d_{1j}} - \frac{1}{r_{1j}}\right)= & {} O_{{\mathbb {P}}}(n^{1/2})O_{{\mathbb {P}}}(n_1^{-1})\ =\ o_{{\mathbb {P}}}(1); \\ O_{{\mathbb {P}}}(n^{3/2})O_{{\mathbb {P}}}(n_i^{-1})\sum _{j=1}^{\breve{\kappa }_2(t)}\left( \frac{1}{r_{2j} - d_{2j}} - \frac{1}{r_{2j}}\right)= & {} O_{{\mathbb {P}}}(n^{1/2})O_{{\mathbb {P}}}(n_2^{-1})\ =\ o_{{\mathbb {P}}}(1). \end{aligned}$$

From Eq. (A.23) and Lemma 4 the proof of Theorem 1 is completed. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Subramanian, S. Function-based hypothesis testing in censored two-sample location-scale models. Lifetime Data Anal 26, 183–213 (2020). https://doi.org/10.1007/s10985-018-09456-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10985-018-09456-8

Keywords

Navigation