Skip to main content
Log in

Semiparametric simultaneous confidence bands for the difference of survival functions

  • Published:
Lifetime Data Analysis Aims and scope Submit manuscript

Abstract

In the analysis of censored survival data, simultaneous confidence bands are useful devices to help determine the efficacy of a treatment over a control. Semiparametric confidence bands are developed for the difference of two survival curves using empirical likelihood and compared with the nonparametric counterpart. Simulation studies are presented to show that the proposed semiparametric approach is superior, with the new confidence bands giving empirical coverage closer to the nominal level. Further comparisons reveal that the semiparametric confidence bands are tighter and, hence, more informative. For censoring rates between 10 and 40 %, the semiparametric confidence bands provide a relative reduction in enclosed area amounting to between 2 and 10 % over their nonparametric bands, with increased reduction attained for higher censoring rates. The methods are illustrated using an University of Massachusetts AIDS data set.

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
Fig. 5

Similar content being viewed by others

References

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

    Book  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Claeskens G, Hjort NL (2008) Model selection and model averaging., Cambridge series in statistical and probabilistic mathematicsCambridge University Press, Cambridge

    Book  MATH  Google Scholar 

  • Collett D (2002) Modelling binary data. CRC Press, Boca Raton

    MATH  Google Scholar 

  • Cox DR, Snell EJ (1989) Analysis of binary data. Chapman and Hall, London

    MATH  Google Scholar 

  • Dabrowska DM, Doksum KA, Song J (1989) Graphical comparison of cumulative hazards for two populations. Biometrika 76:763–773

    Article  MathSciNet  MATH  Google Scholar 

  • Dikta G (1998) On semiparametric random censorship models. J Stat Plan Inference 66:253–279

    Article  MathSciNet  MATH  Google Scholar 

  • Dikta G, Kvesic M, Schmidt C (2006) Bootstrap approximations in model checks for binary data. J Am Stat Assoc 101:521–530

    Article  MathSciNet  MATH  Google Scholar 

  • Dikta G (2014) Asymptotically efficient estimation under semi-parametric random censorship models. J Multivar Anal 124:10–24

    Article  MathSciNet  MATH  Google Scholar 

  • Einmahl JHJ, McKeague IW (1999) Confidence tubes for multiple quantile plots via empirical likelihood. Ann Stat 27:1348–1367

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Hollander M, McKeague IW, Yang J (1997) Likelihood ratio-based confidence bands for survival functions. J Am Stat Assoc 92:215–226

    Article  MathSciNet  MATH  Google Scholar 

  • Hosmer DW, Lemeshow S, May S (2008) Applied survival analysis: regression modeling of time to event data. Wiley, New York

    Book  MATH  Google Scholar 

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

    Book  MATH  Google Scholar 

  • Lee J, Hyun S (2011) Confidence bands for the difference of two survival functions under the additive risk model. J Appl Stat 38:785–797

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Lin DY, Fleming TR, Wei LJ (1994) Confidence bands for survival curves under the proportional hazards model. Biometrika 81:73–81

    Article  MathSciNet  MATH  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  MATH  Google Scholar 

  • McKeague IW, Zhao Y (2005) Comparing distribution functions via empirical likelihood. Int J Biostat 1(1):Article 5

  • Mondal S, Subramanian S (2014) Model assisted Cox regression. J Multivar Anal 123:281–303

  • Mondal S, Subramanian S (2015) Simultaneous confidence bands for Cox regression from semiparametric random censorship. Lifetime Data Anal. doi:10.1007/s10985-015-9323-2

  • Owen A (1988) Empirical likelihood ratio confidence intervals for a single functional. Biometrika 75:237–249

    Article  MathSciNet  MATH  Google Scholar 

  • Owen A (1991) Empirical likelihood for linear models. Ann Stat 19:1725–1747

    Article  MathSciNet  MATH  Google Scholar 

  • Parzen MI, Wei LJ, Ying Z (1997) Simultaneous confidence intervals for the difference of two survival functions. Scand J Stat 24:309–314

    Article  MathSciNet  MATH  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  MATH  Google Scholar 

  • Shen J, He S (2006) Empirical likelihood for the difference of two survival functions under right censorship. Stat Prob Lett 76:169–181

    Article  MathSciNet  MATH  Google Scholar 

  • Subramanian S (2004) The missing censoring-indicator model of random censorship. In: Balakrishnan N, Rao CR (eds) Advances in survival analysis, handbook of statistics, vol 23. Elsevier, Amsterdam, pp 123–141

    Chapter  Google Scholar 

  • Subramanian S, Dikta G (2009) Inverse censoring weighted median regression. Stat Methodol 6:594–603

    Article  MathSciNet  MATH  Google Scholar 

  • Subramanian S (2012) Model-based likelihood confidence intervals for survival functions. Stat Probab Lett 82:626–635

    Article  MathSciNet  MATH  Google Scholar 

  • Subramanian S, Zhang P (2013) Model-based confidence bands for survival functions. J Stat Plan Inference 143:1166–1185

    Article  MathSciNet  MATH  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  MATH  Google Scholar 

  • Yang S (2013) Semiparametric inference on the absolute risk reduction and the restricted mean survival difference. Lifetime Data Anal 19:219–241

    Article  MathSciNet  MATH  Google Scholar 

  • Yang S, Prentice RL (2005) Semiparametric analysis of short-term and long-term hazard ratios with two sample survival data. Biometrika 92:1–17

    Article  MathSciNet  MATH  Google Scholar 

  • Yuan M (2005) Semiparametric censorship model with covariates. Test 14:489–514

    Article  MathSciNet  MATH  Google Scholar 

  • Zhang M-J, Klein JP (2001) Confidence bands for the difference of two survival curves under proportional hazards model. Lifetime Data Anal 7:243–254

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

We express our sincere thanks to an Associate Editor and a reviewer whose comments improved the overall quality of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sundarraman Subramanian.

Appendix

Appendix

Proof of Eq. (2.13)

Then, suppressing the dependence on t, we can write \(R = \tilde{R}_1 + \tilde{R}_2 + \tilde{R}_3 +\tilde{R}_4\), where

$$\begin{aligned} \tilde{R}_1= & {} \sum _{j=1}^{\tilde{\kappa }_1(t)}d_{1j}\hat{m}_{1j}\log \left( \frac{d_{1j}\hat{m}_{1j}}{r_{1j}+\hat{\lambda }\hat{S}_1(T_{1j})}\right) - \sum _{j=1}^{\tilde{\kappa }_1(t)}d_{1j}\hat{m}_{1j}\log \left( \frac{d_{1j}\hat{m}_{1j}}{r_{1j}}\right) , \\ \tilde{R}_2= & {} \sum _{j=1}^{\tilde{\kappa }_1(t)}(r_{1j} - d_{1j}\hat{m}_{1j})\log \left( 1 -\frac{d_{1j}\hat{m}_{1j}}{r_{1j}+\hat{\lambda }\hat{S}_1(T_{1j})}\right) \\&- \sum _{j=1}^{\tilde{\kappa }_1(t)}(r_{1j} - d_{1j}\hat{m}_{1j})\log \left( 1 -\frac{d_{1j}\hat{m}_{1j}}{r_{1j}}\right) ,\\ \tilde{R}_3= & {} \sum _{j=1}^{\tilde{\kappa }_2(t)}d_{2j}\hat{m}_{2j}\log \left( \frac{d_{2j}\hat{m}_{2j}}{r_{2j} - \hat{\lambda }\hat{S}_2(T_{2j})}\right) - \sum _{j=1}^{\tilde{\kappa }_2(t)}d_{2j}\hat{m}_{2j}\log \left( \frac{d_{2j}\hat{m}_{2j}}{r_{2j}}\right) ,\\ \tilde{R}_4= & {} \sum _{j=1}^{\tilde{\kappa }_2(t)}(r_{2j} - d_{2j}\hat{m}_{2j})\log \left( 1 -\frac{d_{2j}\hat{m}_{2j}}{r_{2j}-\hat{\lambda }\hat{S}_2(T_{2j})}\right) \\&- \sum _{j=1}^{\tilde{\kappa }_2(t)}(r_{2j} - d_{2j}\hat{m}_{2j})\log \left( 1 -\frac{d_{2j}\hat{m}_{2j}}{r_{2j}}\right) . \end{aligned}$$

However, writing \(\tilde{R}_1 = R_1 - \breve{R}_1\), where

$$\begin{aligned} \breve{R}_1= & {} \sum _{j=1}^{\tilde{\kappa }_1(t)}(r_{1j} - d_{1j}\hat{m}_{1j})\log \left( \frac{d_{1j}\hat{m}_{1j}}{r_{1j}+\hat{\lambda }\hat{S}_1(T_{1j})}\right) - \sum _{j=1}^{\tilde{\kappa }_1(t)}(r_{1j} - d_{1j}\hat{m}_{1j})\log \left( \frac{d_{1j}\hat{m}_{1j}}{r_{1j}}\right) , \end{aligned}$$

we see that

$$\begin{aligned} R_1\equiv & {} \sum _{j=1}^{\tilde{\kappa }_1(t)}r_{1j}\log \left( \frac{d_{1j} \hat{m}_{1j}}{r_{1j}+\hat{\lambda }\hat{S}_1(T_{1j})}\right) - \sum _{j=1}^{\tilde{\kappa }_1(t)}r_{1j}\log \left( \frac{d_{1j} \hat{m}_{1j}}{r_{1j}}\right) \\= & {} -\sum _{j=1}^{\tilde{\kappa }_1(t)}r_{1j}\log \left( 1+\frac{\hat{\lambda }\hat{S}_1(T_{1j})}{r_{1j}}\right) . \end{aligned}$$

Next we write \(\tilde{R}_1 + \tilde{R}_2 \equiv R_1 + R_2\), where \(R_2=\tilde{R}_2 - \breve{R}_1\). Then, after some algebra, we obtain

$$\begin{aligned} R_2= & {} \sum _{j=1}^{\tilde{\kappa }_1(t)}(r_{1j}-d_{1j}\hat{m}_{1j})\log \left( 1 + \frac{\hat{\lambda }\hat{S}_1({T_{1j}})}{r_{1j}-d_{1j}\hat{m}_{1j}}\right) . \end{aligned}$$

Applying exactly the same technique, we have that \(\tilde{R}_3+\tilde{R}_4\equiv R_3+R_4\), where

$$\begin{aligned} R_3= & {} -\sum _{j=1}^{\tilde{\kappa }_2(t)}r_{2j}\log \left( 1-\frac{\hat{\lambda }\hat{S}_2(T_{2j})}{r_{2j}}\right) ,\\ R_4= & {} \sum _{j=1}^{\hat{\kappa }_2(t)}(r_{2j}-d_{2j}\hat{m}_{2j})\log \left( 1 - \frac{\hat{\lambda }\hat{S}_2(T_{2j})}{r_{2j}-d_{2j}\hat{m}_{2j}}\right) . \end{aligned}$$

Therefore, we have shown that \(R(t)=R_1+R_2+R_3+R_4\). It now follows that

We prove a number of lemmas. Recall that \(\tilde{\kappa }_i(t)=\sum _{j=1}^{n_i}I(Z_{ij}\le t),i=1,2\). Define

$$\begin{aligned} \hat{\zeta }_i(t)= & {} \sum _{j=1}^{\tilde{\kappa }_i(t)}\left( \frac{d_{ij} \hat{m}_{ij}}{r_{ij}}\right) \hat{S}_i(T_{ij}) \equiv \int _0^t \hat{S}_i(s)d\hat{\Lambda }_i(s),\qquad i=1,2. \end{aligned}$$
(5.1)

\(\square \)

Lemma 1

The process \(n_i^{1/2}(\hat{\zeta }_i-F_i)\) is asymptotically equivalent to \(S_i\cdot n_i^{1/2}(\hat{\Lambda }_i-\Lambda _i)\).

Proof

Add and subtract \(\int _0^t S_i(s) d\hat{\Lambda }_i(s)\) and apply the Duhamel equation to get

$$\begin{aligned} n_i^{1/2}\left( \hat{\zeta }_i(t) - F_i(t) \right)= & {} n_i^{1/2}\left( \int _0^t\hat{S}_i(s)d\hat{\Lambda }_i(s)- \int _0^t S_i(s)d\Lambda _i(s) \right) \\= & {} - n_i^{1/2}\int _0^t S_i(s) \left( \int _0^s \frac{\hat{S}_i(u-)}{S_i(u)}d\left( \hat{\Lambda }_i(u) - \Lambda _i(u)\right) \right) d\Lambda _i(s)\\&+\, n_i^{1/2} \int _0^t S_i(s)d \left( \hat{\Lambda }_i(s) - \Lambda _i(s) \right) \ :=\ I_1(t)+I_2(t). \end{aligned}$$

Interchanging the order of integration, we get, uniformly for \(t\in [0,t_i]\) such that \(y_i(t_i)>0\),

$$\begin{aligned} I_1(t)= & {} - n_i^{1/2} \int _0^t \left( \int _u^t S_i(s) d\Lambda _i(s) \right) \frac{\hat{S}_i(u-)}{S_i(u)} d \left( \hat{\Lambda }_i(u) - \Lambda _i(u) \right) . \end{aligned}$$

Applying Lenglart’s inequality, it follows that

$$\begin{aligned} I_1(t)= & {} - n_i^{1/2} \int _0^t \left( S_i(u) - S_i(t) \right) d \left( \hat{\Lambda }_i(u) - \Lambda _i(u) \right) + o_p(1)\\= & {} -I_2(t) + S_i(t)n_i^{1/2}\left( \hat{\Lambda }_i(t) - \Lambda _i(t)\right) + o_p(1). \end{aligned}$$

\(\square \)

Noting that \(\int _0^t S_i^2(s) d\Lambda _i(s)=(1-S_i^2(t))/2\), next define

$$\begin{aligned} \hat{\zeta }_i^{(2)}(t)= & {} \sum _{j=1}^{\tilde{\kappa }_i(t)}\left( \frac{d_{ij} \hat{m}_{ij}}{r_{ij}}\right) \hat{S}_i^2(T_{ij})\ \equiv \ \int _0^t \hat{S}_i^2(s)d\hat{\Lambda }_i(s),\qquad i=1,2. \end{aligned}$$
(5.2)

Lemma 2

The process \(n_i^{1/2}\left( \hat{\zeta }_i^{(2)}-\frac{1-S_i^2}{2}\right) \) is asymptotically equivalent to \(S_i^2n_i^{1/2} (\hat{\Lambda }_i-\Lambda _i)\).

Proof

Add and subtract \(\int _0^t S_i^2(s) d\hat{\Lambda }_i(s)\) and apply the Duhamel equation to get

$$\begin{aligned}&n_i^{1/2}\left( \hat{\zeta }_i^{(2)}(t)-\int _0^t S_i^2(s)d\Lambda _i(s)\right) \\&\quad = - n_i^{1/2}\int _0^t 2S_i^2(s) \left( \int _0^s \frac{\hat{S}_i(u-)}{S_i(u)}d\left( \hat{\Lambda }_i(u) - \Lambda _i(u)\right) \right) d\Lambda _i(s) \\&\qquad +\, n_i^{1/2} \int _0^t S_i^2(s)d \left( \hat{\Lambda }_i(s) - \Lambda _i(s) \right) \ :=\ I_3(t)+I_4(t). \end{aligned}$$

Interchanging the order of integration, we get, uniformly for \(t\in [0,t_i]\) such that \(y_i(t_i)>0\),

$$\begin{aligned} I_3(t)= & {} - n_i^{1/2} \int _0^t \left( \int _u^t 2S_i^2(s) d\Lambda _i(s) \right) d \left( \hat{\Lambda }_i(u) - \Lambda _i(u) \right) + o_p(1). \end{aligned}$$

Note that the integrand above equals \(S_i^2(u) - S_i^2(t)\). It follows that

$$\begin{aligned} I_3(t)= & {} S_i^2(t)n_i^{1/2}\left( \hat{\Lambda }_i(t) - \Lambda _i(t)\right) - I_4(t) + o_p(1). \end{aligned}$$

\(\square \)

Recall that \(\Vert h\Vert _{\tau _1}^{\tau _2} = \sup _{t\in [\tau _1,\tau _2]}|h(t)|\). For our next result, define

$$\begin{aligned} \tilde{\zeta }_i(t)= & {} - \sum _{j=1}^{\tilde{\kappa }_i(t)}\log \left( 1 - \frac{d_{ij} \hat{m}_{ij}}{r_{ij}}\right) \hat{S}_i(T_{ij}), \qquad i=1,2. \end{aligned}$$
(5.3)

Lemma 3

For \(i=1,2\), we have \(\Vert \hat{\zeta }_i-\tilde{\zeta }_i\Vert _{\tau _1}^{\tau _2} = o_p(n^{-1/2})\).

Proof

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

$$\begin{aligned} n_i^{1/2}|\hat{\zeta }_i(t) - \tilde{\zeta }_i(t)|\le & {} \left\{ \max _{j \le \tilde{\kappa }_i(t)}\left( \frac{n_i^{1/2}d_{ij}\hat{m}_{ij}}{r_{ij}}\right) \right\} \times \left\{ \sum _{j=1}^{\tilde{\kappa }_i(t)}\left( \frac{d_{ij} \hat{m}_{ij}}{r_{ij}}\right) \hat{S}_i(T_{ij})\right\} .\quad \end{aligned}$$
(5.4)

Under continuity of \(H_i\), the first quantity on the right side of Eq. (5.4) is bounded above by

$$\begin{aligned} \max _{j \le \tilde{\kappa }_i(t)}\left( n_i^{1/2}/r_{ij}\right)\le & {} n_i^{-1/2} \Vert 1/\bar{Y}_i\Vert _{0}^{\tau _2}\ =\ O\left( n_i^{-1/2}\right) O_p(1) \ =\ o_p(1). \end{aligned}$$
(5.5)

From Eq. (5.1), the second quantity on the right side of Eq. (5.4) is bounded above by \(\hat{\zeta }_i(\tau _2)\ {\buildrel \mathrm{a.s.} \over \longrightarrow }\ F_i(\tau _2)\), see also Theorem 2.4 of Dikta (1998) and Lemma 1. Thus, \(\hat{\zeta }_i(\tau _2)=O_p(1)\) which, combined with Eqs. (5.4) and (5.5), completes the proof. \(\square \)

Lemma 4

The Lagrange multiplier, \(\hat{\lambda }(t)\equiv \hat{\lambda }\), solving Eq. (2.12), satisfies

$$\begin{aligned} |\hat{\lambda }|\le & {} \frac{n_2(\alpha (t) + \tilde{\zeta }_1(t) -\hat{\zeta }_2(t))}{\sum _{j=1}^{\tilde{\kappa }_2(t)}d_{2j}\hat{m}_{2j} \hat{S}_2^2(T_{2j})/r_{2j}}, \mathrm{when }\hat{\lambda }> 0; \end{aligned}$$
(5.6)
$$\begin{aligned} |\hat{\lambda }|\le & {} \frac{n_1(-\alpha (t) + \tilde{\zeta }_2(t) - \hat{\zeta }_1(t))}{\sum _{j=1}^{\tilde{\kappa }_1(t)} d_{1j}\hat{m}_{1j}\hat{S}_1^2(T_{1j})/r_{1j}}, \mathrm{when\ }\hat{\lambda }<0. \end{aligned}$$
(5.7)

Proof

For our semiparametric setting, we adapt the approach of McKeague and Zhao (2005), who follow Li (1995). From Eq. (2.12), write \(\alpha (t)=I_1(t)+I_2(t)\), where

$$\begin{aligned} I_i(t)= & {} (-1)^{i+1}\sum _{j=1}^{\tilde{\kappa }_i(t)}\log \left( 1 - \frac{d_{ij}\hat{m}_{ij}}{r_{ij}+(-1)^{i+1}\hat{\lambda }\hat{S}_i(T_{ij})} \right) \hat{S}_i(T_{ij}),\qquad i=1,2. \end{aligned}$$

First assume that \(\hat{\lambda }> 0\). Since \(\mathrm{log}(1-x) + x\) is decreasing for \(x \in (0,1)\), we have that

$$\begin{aligned} \log \left( 1 - \frac{d_{1j} \hat{m}_{1j}}{r_{1j} + \hat{\lambda }\hat{S}_1(T_{1j})}\right) + \frac{d_{1j} \hat{m}_{1j}}{r_{1j}+\hat{\lambda }\hat{S}_1(T_{1j})}\ge & {} \log \left( 1 - \frac{d_{1j} \hat{m}_{1j}}{r_{1j}}\right) + \frac{d_{1j} \hat{m}_{1j} }{r_{1j}}. \end{aligned}$$

It follows that

$$\begin{aligned} I_1(t)\ge & {} \sum _{j=1}^{\tilde{\kappa }_1(t)}\left[ -\frac{d_{1j} \hat{m}_{1j}}{r_{1j}+\hat{\lambda }\hat{S}_1(T_{1j})}+ \log \left( 1 - \frac{d_{1j} \hat{m}_{1j}}{r_{1j}}\right) + \frac{d_{1j} \hat{m}_{1j} }{r_{1j}} \right] \hat{S}_1(T_{1j}).\nonumber \\ \end{aligned}$$
(5.8)

Since, for \(x>0\), the inequality \(n_1/(n_1+x)\le n_2/(n_2+x)\) holds whenever \(n_1\le n_2\), we have the following lower bound for the first term on the right hand side of Eq. (5.8):

$$\begin{aligned} -\sum _{j=1}^{\tilde{\kappa }_1(t)} \frac{d_{1j} \hat{m}_{1j}}{r_{1j}+\hat{\lambda }\hat{S}_1(T_{1j})}\hat{S}_1(T_{1j})= & {} -\sum _{j=1}^{\tilde{\kappa }_1(t)}\left( \frac{d_{1j} \hat{m}_{1j}}{r_{1j}}\right) \left( \frac{r_{1j}}{r_{1j}+\hat{\lambda }\hat{S}_1(T_{1j})}\right) \hat{S}_1(T_{1j}) \nonumber \\\ge & {} -\sum _{j=1}^{\hat{\kappa }_1(t)}\left( \frac{d_{1j}\hat{m}_{1j}}{r_{1j}}\right) \left( \frac{n_1}{n_1+|\hat{\lambda }|\hat{S}_1(T_{1j})}\right) \hat{S}_1(T_{1j}).\nonumber \\ \end{aligned}$$
(5.9)

Note that the second term on the right hand side of Eq. (5.8) equals \(-\tilde{\zeta }_1(t)\), see Eq. (5.3). Using the fact that \(-\mathrm{log} (1-x) \ge x\) when \(0 \le x < 1\), we obtain

$$\begin{aligned} I_2(t)\ge & {} \sum _{j=1}^{\tilde{\kappa }_2(t)}\left( \frac{d_{2j} \hat{m}_{2j}}{r_{2j}-\hat{\lambda }\hat{S}_2(T_{2j})}\right) \hat{S}_2(T_{2j})\\= & {} \sum _{j=1}^{\tilde{\kappa }_2(t)}\left( \frac{d_{2j} \hat{m}_{2j}}{r_{2j}}\right) \left( \frac{r_{2j}}{r_{2j}-|\hat{\lambda }|\hat{S}_2(T_{2j})}\right) \hat{S}_2(T_{2j}). \end{aligned}$$

Since \(n_1/(n_1-x)\ge n_2/(n_2-x)\), when \(x>0\) and \(n_1\le n_2\), we obtain the lower bound

$$\begin{aligned} I_2(t)\ge & {} \sum _{j=1}^{\tilde{\kappa }_2(t)}\left( \frac{d_{2j} \hat{m}_{2j}}{r_{2j}}\right) \left( \frac{n_2}{n_2-|\hat{\lambda }|\hat{S}_2(T_{2j})}\right) \hat{S}_2(T_{2j}). \end{aligned}$$
(5.10)

From Eqs. (5.8)–(5.10), we obtain

$$\begin{aligned} \alpha (t)\ge & {} - \sum _{j=1}^{\tilde{\kappa }_1(t)}\left( \frac{d_{1j}\hat{m}_{1j}}{r_{1j}}\right) \left( \frac{n_1}{n_1+|\hat{\lambda }|\hat{S}_1(T_{1j})}\right) \hat{S}_1(T_{1j}) -\tilde{\zeta }_1(t)\nonumber \\&+ \sum _{j=1}^{\tilde{\kappa }_1(t)} \left( \frac{d_{1j}\hat{m}_{1j}}{r_{1j}} \right) \hat{S}_1(T_{1j}) \nonumber \\&+ \sum _{j=1}^{\tilde{\kappa }_2(t)}\left( \frac{d_{2j}\hat{m}_{2j}}{r_{2j}}\right) \left( \frac{n_2}{n_2-|\hat{\lambda }|\hat{S}_2(T_{2j})} \right) \hat{S}_2(T_{2j}). \end{aligned}$$
(5.11)

From the right hand side of Eq. (5.11), combining the first and third terms gives a nonnegative number. Since \(1/(1 - x) \ge 1 + x\) when \(x < 1\), the fourth term is bounded below by

$$\begin{aligned}&\sum _{j=1}^{\tilde{\kappa }_2(t)}\left( \frac{d_{2j}\hat{m}_{2j}}{r_{2j}}\right) \hat{S}_2(T_{2j}) \left( 1 + |\hat{\lambda }|\hat{S}_2(T_{2j})/n_2\right) \nonumber \\&\quad = \hat{\zeta }_2(t) + \frac{1}{n_2}\sum _{j=1}^{\tilde{\kappa }_2(t)} \left( \frac{d_{2j}\hat{m}_{2j}}{r_{2j}}\right) \hat{S}_2^2(T_{2j})|\hat{\lambda }|, \end{aligned}$$

provided that \(|\hat{\lambda }|\hat{S}_2(T_{2j})/n_2<1\) almost surely. To show this, assume no ties and note from Eq. (2.14) that \(0<\hat{\lambda }<\min _{j:T_{2j}\le t}\{(r_{2j}-d_{2j}\hat{m}_{2j})/\hat{S}_2(T_{2j})\}\). Also, \(\bar{Y}_2(t)/\hat{S}_2(t){\buildrel \mathrm{a.s.} \over \longrightarrow }1-G_2(t)<1\) uniformly over \([\tau _1,\tau _2]\). For large enough \(n_2\), \(\epsilon \) sufficiently small, and some \(T_{2l}\ne 0\), we have

$$\begin{aligned} \frac{\hat{\lambda }}{n_2}< & {} \min _{j:T_{2j}\le t}\frac{r_{2j}/n_2}{\hat{S}_2(T_{2j})}\le \frac{\bar{Y}_2(T_{2l})}{\hat{S}_2(T_{2l})} < 1-G_2(T_{2l})+\epsilon \ <\ 1 \end{aligned}$$

almost surely. We therefore obtain a lower bound for \(\alpha (t)\) given by

$$\begin{aligned} \alpha (t)\ge & {} - \tilde{\zeta }_1(t) + \hat{\zeta }_2(t) + \frac{1}{n_2}\sum _{j=1}^{\tilde{\kappa }_2(t)} \left( \frac{d_{2j}\hat{m}_{2j}}{r_{2j}}\right) \hat{S}_2^2(T_{2j})|\hat{\lambda }|, \end{aligned}$$

from which Eq. (5.6) follows. Proof of Eq. (5.7) can be shown by analogous techniques. \(\square \)

Lemma 5

The Lagrange multiplier, \(\hat{\lambda }(t)\equiv \hat{\lambda }\), solving Eq. (2.12), satisfies \(\Vert \hat{\lambda }\Vert _{\tau _1}^{\tau _2} = O_p(n^{1/2})\).

Proof

That the denominators of Eqs. (5.6) and (5.7) are each \(O_p(1)\), uniformly for \(t\in [\tau _1,\tau _2]\), follows from \(\Vert \hat{\Lambda }_i-\Lambda _i\Vert _0^{\tau _2}=o(1)\) almost surely (cf. Theorem 2.4 of Dikta 1998) and Lemma 2. It remains to show that the numerators of Eqs. (5.6) and (5.7) are each \(O_p(n^{-1/2})\), uniformly for \(t\in [\tau _1,\tau _2]\). By applying Lemma 3, it suffices to show that \(\alpha (t) + \hat{\zeta }_1(t) - \hat{\zeta }_2(t) = O_p(n^{-1/2})\). Let \(n_i/n\rightarrow p_i\) as \(n\rightarrow \infty \). We then have by Lemma 1 and results from Sect. 2.1 that

$$\begin{aligned} n^{1/2}\left( \alpha (t) + \hat{\zeta }_1(t) - \hat{\zeta }_2(t)\right)&= \sum _{i=1}^2 S_i(t)\frac{n_i^{1/2}(\hat{\Lambda }_i(t)-\Lambda _i(t))}{(n_i/n)^{1/2}} + o_p(1) \nonumber \\&{\buildrel {\mathcal {D}} \over \longrightarrow }\sum _{i=1}^2\frac{S_i(t)U_i(t)}{\sqrt{p}_i}\ \equiv \ W(t), \end{aligned}$$
(5.12)

where W is the zero-mean Gaussian process with covariance function given by Eq. (2.17). \(\square \)

1.1 Proof of Theorem 1

From Eq. (2.12), we can write \(\alpha (t)=f_1(\hat{\lambda })-f_2(-\hat{\lambda })\), where

$$\begin{aligned} f_i(\lambda )= & {} \sum _{j=1}^{\tilde{\kappa }_i(t)}\log \left( 1 - \frac{ d_{ij} \hat{m}_{ij}}{r_{ij} + \lambda \hat{S}_i(T_{ij})}\right) \hat{S}_i(T_{ij}). \end{aligned}$$
(5.13)

Note that \(f_i(0) = - \tilde{\zeta }_i(t)\) [cf. Eq. (5.3)]. Before applying a Taylor’s expansion, we note that

$$\begin{aligned} f_i^{'}(\lambda )= & {} \sum _{j=1}^{\tilde{\kappa }_i(t)}\frac{d_{ij} \hat{m}_{ij}}{(r_{ij}+\lambda \hat{S}_i(T_{ij}))(r_{ij}+\lambda \hat{S}_i(T_{ij})- d_{ij}\hat{m}_{ij})}\hat{S}^2_i(T_{ij}),\\ f_i^{''}(\lambda )= & {} \sum _{j=1}^{\tilde{\kappa }_i(t)}\frac{d_{ij}\hat{m}_{ij}(2(r_{ij} +\lambda \hat{S}_i(T_{ij}))- d_{ij} \hat{m}_{ij})}{(r_{ij}+\lambda \hat{S}_i(T_{ij}))^2(r_{ij}+\lambda \hat{S}_i(T_{ij})-d_{ij} \hat{m}_{ij})^2}\hat{S}^3_i(T_{ij}). \end{aligned}$$

Recall that \(\gamma _i(t)\) is defined by Eq. (2.15). Write \(n_if_i^{'}(0)=\tilde{\gamma }_i(t)\), so that

$$\begin{aligned} \tilde{\gamma }_i(t) = n_i\sum _{j:T_{ij} \le t} \frac{d_{ij}\hat{m}_{ij}}{r_{ij}(r_{ij} - d_{ij}\hat{m}_{ij})}\hat{S}^2_i(T_{ij}). \end{aligned}$$

For \(i=1,2\), let \(|\hat{\xi }_i|\le |\hat{\lambda }|\). Taylor’s expansion about 0 yields

$$\begin{aligned} f_i\left( (-1)^{i-1}\hat{\lambda }\right)= & {} - \tilde{\zeta }_i(t) + (-1)^{i-1}\frac{\tilde{\gamma }_i(t)\hat{\lambda }}{n_i} + \frac{1}{2}f^{''}_i(\hat{\xi }_i)\hat{\lambda }^2. \end{aligned}$$
(5.14)

Applying the Glivenko–Cantelli lemma to \(r_{ij}\) and Lemma 5, we have \(\Vert f_i''(\hat{\xi }_i)\Vert _{\tau _1}^{\tau _2}=O_p(n_i^{-2})\). Therefore, by Lemma 5, it follows that \(\Vert f^{''}_i(\hat{\xi }_i)\hat{\lambda }^2\Vert _{\tau _1}^{\tau _2}=O_p(n_i^{-1})=O_p(n^{-1})\). Furthermore, \(\tilde{\gamma }_i(t)\) is uniformly consistent for \(\gamma _i(t)\) over \([0,\tau _2]\). It follows from Eq. (5.14) and Eq. (2.16) that

$$\begin{aligned} \alpha (t)= & {} -\tilde{\zeta }_1(t) + \tilde{\zeta }_2(t) + \hat{\lambda }\left( \frac{\tilde{\gamma }_1(t)}{n_1} +\frac{\tilde{\gamma }_2(t)}{n_2}\right) + O_p\left( \frac{1}{n}\right) \\= & {} -\tilde{\zeta }_1(t) + \tilde{\zeta }_2(t) + \hat{\lambda }n^{-1}\sigma ^2_d(t) + O_p(n^{-1}). \end{aligned}$$

Solving for \(\hat{\lambda }\), we obtain

$$\begin{aligned} \hat{\lambda }= & {} n\sigma ^{-2}_\mathrm{d}(t)\left( \alpha (t) +\tilde{\zeta }_1(t)-\tilde{\zeta }_2(t)+O_p(n^{-1})\right) . \end{aligned}$$
(5.15)

To complete the Proof of Theorem 1, consider Eq. (2.13). Using Taylor expansions of \(\log (1+x)\) and \(\log (1-x)\) about 0, the leading term of \(-2R(t)\) is a product of \(\hat{\lambda }^2\) and a random scaling factor expressed as a double sum, and equals

$$\begin{aligned} \hat{\lambda }^2\sum _{i=1}^2\sum _{j=1}^{\tilde{\kappa }_i(t)}\frac{d_{ij}\hat{m}_{ij} \hat{S}_i^2(T_{ij})}{r_{ij}(r_{ij}-d_{ij}\hat{m}_{ij})}. \end{aligned}$$

Each single sum can be expressed as an integral as in the Proof of Lemma 1. Applying Eq. (2.6.10) of Andersen et al. (1993), it follows that the double sum equals \(\sigma _\mathrm{d}^2(t)/n +o_p(1)\), uniformly over \([0, \tau _2]\). Therefore, applying Eq. (5.15), the leading term of \(-2R(t)\) equals

$$\begin{aligned} n \hat{\sigma }^{-2}_\mathrm{d}(t)\left( \alpha (t)+\tilde{\zeta }_1(t)-\tilde{\zeta }_2(t)+O_p(n^{-1})\right) ^2 + o_p(1)= & {} \left( \frac{W(t)}{\sigma _\mathrm{d}(t)}\right) ^2 + o_p(1).\nonumber \\ \end{aligned}$$
(5.16)

The other terms of \(-2R(t)\) are proportional to

$$\begin{aligned} \hat{\lambda }^l\left[ \sum _{i=1}^2\sum _{j=1}^{\tilde{\kappa }_1(t)}\hat{S}^l_i(T_{ij}) \left( \frac{1}{(r_{ij}-d_{ij}\hat{m}_{ij})^{l-1}} - \frac{1}{r_{ij}^{l-1}}\right) \right] , \qquad l=3, 4, \ldots , \end{aligned}$$

each of which is \(o_p(1)\), uniformly for \(t\in [0,\tau _2]\). For example, when \(l=3\), we have

$$\begin{aligned} \frac{2}{3}\hat{\lambda }^3\sum _{j=1}^{\tilde{\kappa }_i(t)}\hat{S}^3_i(T_{1j}) \left( \frac{1}{r_{ij}-d_{ij}\hat{m}_{ij}} - \frac{1}{r_{ij}}\right) \left( \frac{1}{r_{ij} - d_{ij}\hat{m}_{ij}} + \frac{1}{r_{ij}}\right) ,\qquad i=1,2, \end{aligned}$$

which, uniformly for \(t\in [0,\tau _2]\), equals

$$\begin{aligned} \frac{2}{3}O_p(n^{3/2})O_p(n_i^{-1})\sum _{j=1}^{\tilde{\kappa }_i(t)} \hat{S}^2_i(T_{ij})\left( \frac{1}{r_{ij} - d_{ij}\hat{m}_{ij}} - \frac{1}{r_{ij}}\right) = O_p(n^{1/2})\tilde{\gamma }_i(t)/n_i = o_p(1). \end{aligned}$$

Now apply Lemma 3 and Eq. (5.12) to complete the Proof of Theorem 1. \(\square \)

1.2 Large sample justification of the multiplier bootstrap

Write \(\mathbb {\hat{H}}^*_i(t) = L^*_{n_i, 1}(t) + L^*_{n_i, 2}(t)\), where \(L^*_{n_i, 1}(t)\) and \(L^*_{n_i, 2}(t)\) are defined by Eqs. (2.20) and (2.21). To show that \({\mathbb {W}}^*\) defined by Eq. (2.19) has the limit distribution as that of \(W/\sigma _\mathrm{d}\), it suffices to show that \({\hat{\mathbb {H}}}^*_i(\cdot )\) has the same weak limit as \(\hat{\mathbb {H}}_i(t)=L_{n_i,1}(t)+L_{n_i,2}(t)+o_p(1)\). Let \({\mathbb {P}}_{n_i}\), \({\mathbb {E}}_{n_i}\), \(\mathrm{Cov}_{n_i}\), and \(\mathrm{Var}_{n_i}\) be the probability measure, expectation, covariance, and variance with respect to the bootstrap, that is, conditioned on the sample \(\{(Z_{ij}, \delta _{ij}), j = 1, \ldots , n_i\}\). To show that \(\hat{\mathbb {H}}_i^*(t)\) has the limiting covariance structure given by Eq. (2.1), note that

$$\begin{aligned} \mathrm{Cov}_{n_i}(\hat{\mathbb {H}}_i^*(s), \hat{\mathbb {H}}_i^*(t))= & {} {\mathbb {E}}_{n_i}(L^*_{n_i,1}(s) L^*_{n_i,1}(t)) + {\mathbb {E}}_{n_i}(L^*_{n_i,2}(s) L^*_{n_i,2}(t)) \nonumber \\&+\,{\mathbb {E}}_{n_i}(L^*_{n_i,1}(s) L^*_{n_i,2}(t)) + {\mathbb {E}}_{n_i}(L^*_{n_i,1}(t) L^*_{n_i,2}(s)).\nonumber \\ \end{aligned}$$
(5.17)

Strong consistency of \(\hat{\varvec{\theta }}_i\), assumption \(A_6\) of Dikta (1998) and the arguments in the Proof of Theorem 2.4 of Dikta (1998) imply that \(\Vert m_i(\cdot , \varvec{\hat{\theta }}_i)-m_i(\cdot , \varvec{\theta }_i)\Vert _{0}^{\tau _2}=o(1)\) almost surely. Likewise, \(\hat{\alpha }(\cdot ,\cdot )\) is strongly uniformly consistent over \([0,\tau _2]\times [0,\tau _2]\). Finally, \(\Vert \bar{Y}_i-y_i\Vert _0^{\tau _2}=o(1)\) almost surely. The first quantity on the right hand side of Eq. (5.17) can be computed to yield

$$\begin{aligned} {\mathbb {E}}_{n_i}(L^*_{n_i,1}(s), L^*_{n_i,2}(t))= & {} \frac{1}{n_i} \sum _{j=1}^{n_i}\frac{\hat{m}_{ij}^2}{\bar{Y}_i^2(Z_{ij})}I(Z_{ij} < s \wedge t)\ =\ \frac{1}{n_i} \sum _{j=1}^{n_i}\frac{m_{ij}^2}{y_i^2(Z_{ij})}\\&\qquad I(Z_{ij} < s \wedge t) + o(1). \end{aligned}$$

By the strong law of large numbers, for almost all sample sequences \(\{Z_{ij},\delta _{ij}, 1\le j\le n_i\}\), \({\mathbb {E}}_{n_i}(L^*_{n_i,1}(s), L^*_{n_i,2}(t))\) converges to the first term on the right hand side of Eq. (2.1). The second quantity on the right hand side of Eq. (5.17) can be computed to yield

$$\begin{aligned} {\mathbb {E}}_{n_i}(L^*_{n_i,2}(s), L^*_{n_i,2}(t))= & {} \frac{1}{n_i} \sum _{j=1}^{n_i} \left( \frac{\delta _{ij} - \hat{m}_{ij}}{\hat{m}_{ij}\hat{\bar{m}}_{ij}}\right) ^2 \int _0^{s} \int _0^{t} \frac{\hat{\alpha }(u, Z_{ij})}{\bar{Y}_i(u)} \frac{\hat{\alpha }(v, Z_{ij})}{\bar{Y}_i(v)}\\&\qquad d\hat{H}_i(u)d\hat{H}_i(v)\\= & {} \frac{1}{n_i} \sum _{j=1}^{n_i} \left( \frac{\delta _{ij} - m_{ij}}{m_{ij}\bar{m}_{ij}}\right) ^2 \int _0^{s} \int _0^{t} \frac{\alpha (u, Z_{ij})}{y_i(u)} \frac{\alpha (v, Z_{ij})}{y_i(v)}\\&\qquad dH_i(u)dH_i(v) +o(1). \end{aligned}$$

By the strong law of large numbers, for almost all sample sequences \(\{Z_{ij},\delta _{ij}, 1\le j\le n_i\}\), \({\mathbb {E}}_{n_i}(L^*_{n_i,2}(s), L^*_{n_i,2}(t))\) converges to the second term on the right hand side of Eq. (2.1). One of the cross-product moment terms on the right hand side of Eq. (5.17) is given by

$$\begin{aligned} E_{n_i}(L^*_{n_i,1}(t_1), L^*_{n_i,2}(t_2)) = \frac{1}{n_i} \sum _{j=1}^{n_i} \frac{(\delta _{ij} - \hat{m}_{ij}) \hat{m}_{ij}}{\hat{m}_{ij}\hat{\bar{m}}_{ij}}\frac{I(Z_{ij} < t_1)}{\bar{Y}_i(Z_{ij})} \int _0^{t_2} \frac{\hat{\alpha }(u, Z_{ij})}{\bar{Y}_i(u)}d\hat{H}_i(u),\nonumber \\ \end{aligned}$$
(5.18)

with the other term given in an analogous way. The aforementioned arguments, followed by applying iterated conditional expectation with conditioning by \(Z_{i1}\), implies that the two cross-moment terms in Eq. (5.17) are each zero.

To show that \(\hat{\mathbb {H}}^*_i(\cdot )\) converges weakly to a zero-mean Gaussian process we verify Lindeberg’s condition and tightness. We can write \(\mathbb {\hat{H}}^*_i(t) = L^*_{n_i,1}(t) + L^*_{n_i,2}(t) = \sum _{j=1}^{n_i}B_{ij}(t)G_{ij}\), where

$$\begin{aligned} B_{ij}(t)= n_i^{-1/2} \left[ \frac{\hat{m}_{ij}}{\bar{Y}_i(Z_{ij})}I(Z_{ij} \le t) + \frac{\delta _{ij} - \hat{m}_{ij}}{\hat{m}_{ij}\hat{\bar{m}}_{ij}} \int _0^t \frac{\hat{\alpha }(u, Z_{ij})}{\bar{Y}_i(u)}d\hat{H}_i(u) \right] . \end{aligned}$$

Let \(s^2_i = \sum _{j=1}^{n_i} \mathrm{Var}_{n_i}[B_{ij}(t)G_{ij}] = \sum _{j=1}^{n_i} B_{ij}^2(t)\). As in Mondal and Subramanian (2015), it can be shown that for almost all sample sequences \(\{Z_{ij},\delta _{ij}, 1\le j\le n_i\}\), for any \(\eta _i > 0\),

$$\begin{aligned} \frac{1}{s^2_i}\sum _{j=1}^{n_i}{\mathbb {E}}_{n_1} [B_{ij}^2(t)G_{ij}^2 I(|B_{ij}(t)G_{ij}| > s_i\eta _i)] \rightarrow 0 \text{ as } n \rightarrow \infty . \end{aligned}$$

Let \(K \ge 3\). To verify tightness, we follow Mondal and Subramanian (2015) to show that

$$\begin{aligned}&\underset{n\rightarrow \infty }{\overline{\lim }}[\hat{\mathbb {H}}_i^*(t) - \hat{\mathbb {H}}_i^*(s) ]^4 \le K\underset{n\rightarrow \infty }{\overline{\lim }}\left[ \sum _{j=1}^{n_i}(B_{ij}(t) - B_{ij}(s))^2\right] ^2, \quad s < t.\quad \end{aligned}$$
(5.19)

The sum on the right hand side of inequality (5.19) is given by

$$\begin{aligned} \sum _{j=1}^{n_i} (B_{ij}(t) - B_{ij}(s))^2= & {} \frac{1}{n_i} \sum _{j=1}^{n_i}\left[ \frac{\hat{m}_{ij}}{\bar{Y}_i(Z_{ij})}I(s<Z_{ij} \le t)\right. \\&\left. + \frac{\delta _{ij} - \hat{m}_{ij}}{\hat{m}_{ij}\hat{\bar{m}}_{ij}} \int _{s}^{t} \frac{\hat{\alpha }(u, Z_{ij})}{\bar{Y}_i(u)}d\hat{H}_i(u) \right] ^2, \end{aligned}$$

which can be shown equal to

$$\begin{aligned} \int _s^t\frac{m_i^{2}(u)}{y_i^2(u)}dH_i(u)+\int _s^t\int _s^t \frac{\alpha _i(u,v)}{y_i(u)y_i(v)}dH_i(u)dH_i(v) + o_p(1). \end{aligned}$$

Therefore, the left hand side of inequality (5.19) is finite and tightness is verified.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ahmed, N., Subramanian, S. Semiparametric simultaneous confidence bands for the difference of survival functions. Lifetime Data Anal 22, 504–530 (2016). https://doi.org/10.1007/s10985-015-9348-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10985-015-9348-6

Keywords

Navigation