Skip to main content

Advertisement

Log in

Subgroup specific incremental value of new markers for risk prediction

  • Published:
Lifetime Data Analysis Aims and scope Submit manuscript

Abstract

In many clinical applications, understanding when measurement of new markers is necessary to provide added accuracy to existing prediction tools could lead to more cost effective disease management. Many statistical tools for evaluating the incremental value (IncV) of the novel markers over the routine clinical risk factors have been developed in recent years. However, most existing literature focuses primarily on global assessment. Since the IncVs of new markers often vary across subgroups, it would be of great interest to identify subgroups for which the new markers are most/least useful in improving risk prediction. In this paper we provide novel statistical procedures for systematically identifying potential traditional-marker based subgroups in whom it might be beneficial to apply a new model with measurements of both the novel and traditional markers. We consider various conditional time-dependent accuracy parameters for censored failure time outcome to assess the subgroup-specific IncVs. We provide non-parametric kernel-based estimation procedures to calculate the proposed parameters. Simultaneous interval estimation procedures are provided to account for sampling variation and adjust for multiple testing. Simulation studies suggest that our proposed procedures work well in finite samples. The proposed procedures are applied to the Framingham Offspring Study to examine the added value of an inflammation marker, C-reactive protein, on top of the traditional Framingham risk score for predicting 10-year risk of cardiovascular disease.

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

  • Baker S, Pinsky P (2001) A proposed design and analysis for comparing digital and analog mammography: special receiver operating characteristic methods for cancer screening. J Am Stat Assoc 96:421–428

    Article  MathSciNet  MATH  Google Scholar 

  • Bickel P, Rosenblatt M (1973) On some global measures of the deviations of density function estimates. Ann Stat 1:1071–1095

    Article  MathSciNet  MATH  Google Scholar 

  • Blumenthal R, Michos E, Nasir K (2007) Further improvements in CHD risk prediction for women. J Am Med Assoc 297:641–643

    Article  Google Scholar 

  • Cai T, Cheng S (2008) Robust combination of multiple diagnostic tests for classifying censored event times. Biostatistics 9:216–233

    Article  MATH  Google Scholar 

  • Cai T, Dodd LE (2008) Regression analysis for the partial area under the ROC curve. Stat Sin 18: 817–836

    Google Scholar 

  • Cai T, Tian L, Wei L (2005) Semiparametric Box–Cox power transformation models for censored survival observations. Biometrika 92(3):619–632

    Article  MathSciNet  MATH  Google Scholar 

  • Cai T, Tian L, Uno H, Solomon S, Wei L (2010) Calibrating parametric subject-specific risk estimation. Biometrika 97(2):389–404

    Article  MathSciNet  MATH  Google Scholar 

  • Cook N, Ridker P (2009) The use and magnitude of reclassification measures for individual predictors of global cardiovascular risk. Ann Intern Med 150(11):795–802

    Article  Google Scholar 

  • Cook N, Buring J, Ridker P (2006) The effect of including C-reactive protein in cardiovascular risk prediction models for women. Ann Intern Med 145:21–29

    Article  Google Scholar 

  • Cox D (1972) Regression models and life-tables. J R Stat Soc B 34(2):187–220

    MATH  Google Scholar 

  • Dabrowska D (1987) Non-parametric regression with censored survival time data. Scand J Stat 14(3): 181–197

    Google Scholar 

  • Dabrowska D (1989) Uniform consistency of the kernel conditional Kaplan–Meier estimate. Ann Stat 17(3):1157–1167

    Article  MathSciNet  MATH  Google Scholar 

  • Dabrowska D (1997) Smoothed Cox regression. Ann Stat 25(4):1510–1540

    Article  MathSciNet  MATH  Google Scholar 

  • D’Agostino R (2006) Risk prediction and finding new independent prognostic factors. J Hypertens 24(4):643–645

    Article  MathSciNet  Google Scholar 

  • Dodd L, Pepe M (2003) Partial AUC estimation and regression. Biometrics 59:614–623

    Article  MathSciNet  MATH  Google Scholar 

  • Du Y, Akritas M (2002) Iid representations of the conditional Kaplan–Meier process for arbitrary distributions. Math Methods Stat 11:152–182

    MathSciNet  MATH  Google Scholar 

  • Dwyer AJ (1996) In pursuit of a piece of the ROC. Radiology 201:621–625

    Google Scholar 

  • Expert Panel on Detection, Evaluation, and Treatment of High Blood Cholesterol in Adults (2001) Executive summary of the third report of the National Cholesterol Education Program (NCEP) Expert Panel on Detection, Evaluation, and Treatment of High Blood Cholesterol in Adults (Adult Treatment Panel III). J Am Med Assoc 285(19):2486–2497

    Google Scholar 

  • Fan J, Gijbels I (1995) Data-driven bandwidth selection in local polynomial regression: variable bandwidth selection and spatial adaptation. J R Stat Soc B 57:371–394

    MathSciNet  MATH  Google Scholar 

  • Gail M, Pfeiffer R (2005) On criteria for evaluating models of absolute risk. Biostatistics 6(2):227–239

    Article  MATH  Google Scholar 

  • Gilbert P, Wei L, Kosorok M, Clemens J (2002) Simultaneous inferences on the contrast of two hazard functions with censored observations. Biometrics 58(4):773–780

    Article  MathSciNet  MATH  Google Scholar 

  • Harrell F Jr, Lee K, Mark D (1996) Multivariable prognostic models: issues in developing models, evaluating assumptions and adequacy, and measuring and reducing errors. Stat Med 15(4):361–387

    Article  Google Scholar 

  • Heagerty P, Zheng Y (2005) Survival model predictive accuracy and ROC curves. Biometrics 61:92–105

    Article  MathSciNet  MATH  Google Scholar 

  • Jiang Y, Metz C, Nishikawa R (1996) A receiver operating characteristic partial area index for highly sensitive diagnostic tests. Radiology 201:745–750

    Google Scholar 

  • Jin Z, Ying Z, Wei L (2001) A simple resampling method by perturbing the minimand. Biometrika 88(2):381–390

    Article  MathSciNet  MATH  Google Scholar 

  • Korn E, Simon R (1990) Measures of explained variation for survival data. Stat Med 9(5):487–503

    Article  Google Scholar 

  • An approach to nonparametric regression for life history data using local linear fitting. Ann Stat 23:787–823

  • McIntosh M, Pepe M (2002) Combining several screening tests: optimality of the risk score. Biometrics 58(3):657–664

    Article  MathSciNet  MATH  Google Scholar 

  • Park Y, Wei L (2003) Estimating subject-specific survival functions under the accelerated failure time model. Biometrika 9:717–723

    Article  MathSciNet  Google Scholar 

  • Park B, Kim W, Ruppert D, Jones M, Signorini D, Kohn R (1997) Simple transformation techniques for improved non-parametric regression. Scand J Stat 24(2):145–163

    Article  MathSciNet  MATH  Google Scholar 

  • Paynter N, Chasman D, Pare G, Buring J, Cook N, Miletich J, Ridker P (2010) Association between a literature-based genetic risk score and cardiovascular events in women. J Am Med Assoc 303(7):631–637

    Google Scholar 

  • Pencina M, D’Agostino R (2004) Overall C as a measure of discrimination in survival analysis: model specific population value and confidence interval estimation. Stat Med 23(13):2109–2123

    Article  Google Scholar 

  • Pencina M, D’Agostino RS, D’Agostino RJ, Vasan R (2008) Evaluating the added predictive ability of a new marker: From area under the ROC curve to reclassification and beyond (with Coomentaries & Rejoinder). Stat Med 27:157–212

    Article  MathSciNet  Google Scholar 

  • Pfeiffer R, Gail M (2010) Two criteria for evaluating risk prediction models. Biometrics 67(3):1057–1065

    Article  MathSciNet  Google Scholar 

  • Pfeffer M, Jarcho J (2006) The charisma of subgroups and the subgroups of CHARISMA. N Engl J Med 354(16):1744–1746

    Article  Google Scholar 

  • Ridker P (2007) C-reactive protein and the prediction of cardiovascular events among those at intermediate risk: moving an inflammatory hypothesis toward consensus. J Am Coll Cardiol 49(21):2129–2138

    Article  Google Scholar 

  • Ridker P, Rifai N, Rose L, Buring J, Cook N (2007) Comparison of C-reactive protein and low-density lipoprotein cholesterol levels in the prediction of first cardiovascular events. N Engl J Med 347: 1557–1565

    Google Scholar 

  • Robins J, Ya’Acov R (1997) Toward a curse of dimensionality appropriate (CODA) asymptotic theory for semi-parametric models. Stat Med 16(3):285–319

    Article  Google Scholar 

  • Rothwell P (2005) Treating Individuals 1: External validity of randomised controlled trials: “To whom do the results of this trial apply?”. Lancet 365:82–93

    Google Scholar 

  • Tian L, Zucker D, Wei L (2005) On the Cox model with time-varying regression coefficients. J Am Stat Assoc 100(469):172–183

    Article  MathSciNet  MATH  Google Scholar 

  • Tian L, Cai T, Wei LJ (2009) Identifying subjects who benefit from additional information for better prediction of the outcome variables. Biometrics 65:894–902

    Article  MathSciNet  MATH  Google Scholar 

  • Tibshirani R, Hastie T (1987) Local likelihood estimation. J Am Stat Assoc 82(398):559–567

    Article  MathSciNet  MATH  Google Scholar 

  • Tice J, Cummings S, Ziv E, Kerlikowske K (2005) Mammographic breast density and the Gail model for breast cancer risk prediction in a screening population. Breast Cancer Res Treat 94(2):115–122

    Article  Google Scholar 

  • Uno H, Cai T, Tian L, Wei L (2007) Evaluating prediction rules for t-year survivors with censored regression models. J Am Stat Assoc 102:527–537

    Google Scholar 

  • Uno H, Cai T, Pencina M, D’Agostino R, Wei L (2011a) On the C-statistics for evaluating overall adequacy of risk prediction procedures with censored survival data. Stat Med 30(10):1105–1117

    MathSciNet  Google Scholar 

  • Uno H, Cai T, Tian L, Wei LJ (2011b) Graphical procedures for evaluating overall and subject-specific incremental values from new predictiors with censored event time data. Biometrics 67:1389–1396

    Article  MathSciNet  MATH  Google Scholar 

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

    Book  MATH  Google Scholar 

  • Wacholder S, Hartge P, Prentice R, Garcia-Closas M, Feigelson H, Diver W, Thun M, Cox D, Hankinson S, Kraft P et al (2010) Performance of common genetic variants in breast-cancer risk models. N Engl J Med 362(11):986–993

    Article  Google Scholar 

  • Wand M, Marron J, Ruppert D (1991) Transformation in density estimation (with comments). J Am Stat Assoc 86:343–361

    Article  MathSciNet  MATH  Google Scholar 

  • Wang T, Gona P, Larson M, Tofler G, Levy D, Newton-Cheh C, Jacques P, Rifai N, Selhub J, Robins S (2006) Multiple biomarkers for the prediction of first major cardiovascular events and death. N Engl J Med 355:2631–2639

    Article  Google Scholar 

  • Wang R, Lagakos S, Ware J, Hunter D, Drazen J (2007) Statistics in medicine-reporting of subgroup analyses in clinical trials. N Engl J Med 357(21):2189–2194

    Article  Google Scholar 

  • Wilson PW, D’Agostino RB, Levy D, Belanger AM, Silbershatz H, Kannel WB (1998) Prediction of cornary heart disease using risk factor categories. Circulation 97:1837–1847

    Article  Google Scholar 

  • Zhao L, Cai T, Tian L, Uno H, Solomon S, Wei L, Minnier J, Kohane I, Pencina M, D’Agostino R, et al (2010) Stratifying subjects for treatment selection with censored event time data from a comparative study. Harvard University Biostatistics Working Paper Series 2010: Working Paper 122

Download references

Acknowledgments

The Framingham Heart Study and the Framingham SHARe project are conducted and supported by the National Heart, Lung, and Blood Institute (NHLBI) in collaboration with Boston University. The Framingham SHARe data used for the analyses described in this manuscript were obtained through dbGaP (access number: phs000007.v3.p2). This manuscript was not prepared in collaboration with investigators of the Framingham Heart Study and does not necessarily reflect the opinions or views of the Framingham Heart Study, Boston University, or the NHLBI. The work is supported by Grants U01-CA86368, P01-CA053996, R01-GM085047, R01-GM079330, R01-AI052817 and U54-LM008748 awarded by the National Institutes of Health.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tianxi Cai.

Appendix

Appendix

1.1 Appendix A

Let \(\mathbb{P }_n\) and \(\mathbb{P }\) denote expectation with respect to (wrt) the empirical probability measure of \(\{(T_{i},\,\varDelta _{i},\, X_{i},\, Z_{i}),\,i=1,\ldots ,n\}\) and the probability measure of \((T,\,\varDelta ,\,X,\,Z),\) respectively, and  \(\mathbb{G }_{n}=\sqrt{n}(\mathbb{P }_n-\mathbb{P }).\) We use \(\dot{\mathcal{F }}(x)\) to denote \(d\mathcal F (x)/dx\) for any function \(\mathcal F ,\,\simeq \) to denote equivalence up to \(o_{p}(1),\) and \(\lesssim \) to denote being bounded above up to a universal constant. Let \(\beta _{0}\) and \(\gamma _{0}\) denote the solution to \(E[V_{i}\{Y_{i}^{\dag }-g_{1}(\beta ^{\prime }V_{i})\}]=0\) and \(E[W_{i}\{Y_{i}^{\dag }-g_{2}(\gamma ^{\prime }W_{i})\}]=0,\) respectively. Let \(\bar{ p}_{1i}=g_{1}(\beta _{0}^{\prime }V_{i}),\) and \(\bar{ p}_{2i}=g_{2}(\gamma _{0}^{\prime }W_{i}).\) Let \(\omega = \varDelta I(T\le t_{0})/G_{X,\,Z}(T)+I(T> t_{0})/G_{X,\,Z}(t_{0}),\,{\widehat{M}}_{i}(c)=I({\widehat{p}}_{2i}\ge c)\) and \(\bar{M}_{i}(c)=I(\bar{ p}_{2i}\ge c).\) For \(y=0,\,1,\) let \(f_{y}(c;\,s)\) denote the conditional density of \(\bar{ p}_{2i}\) given \(Y^{\dag }_i = y\) and \(\bar{ p}_{1i}=s\) and we assumed that \(f_{y}(c;\,s)\) is continuous and bounded away from zero uniformly in \(c\) and \(s.\) This assumption implies that \(\mathrm{{ROC}}(u;s)\) has continuous and bounded derivative \(\dot{\mathrm{{ROC}}}(u;\,s) = \partial \mathrm{{ROC}}(u;\,s)/\partial u.\) We assume that \(V\) and \(W\) are bounded, and \(\tau (y;\,s) = \partial pr[\phi \{\bar{ p}_{1}(X)\} \le s,\,Y^{\dag }= y]/\partial s,\) is continuously differentiable with bounded derivatives and bounded away from zero. Throughout, the bandwidths are assumed to be of order \(n^{-\nu }\) with \(\nu \in (1/5,\,1/2).\) For ease of presentation and without loss of generality, we assume that \(h_1 = h_0,\) denoted by \(h,\) and suppress \(h\) from the notations. Without loss of generality, we assume that \(\sup _{t,x,z}|n^{\frac{1}{2}}\{{\widehat{G}}_{X,Z}(t)-G_{X,Z}(t)\}|=O_p(1).\) When \(C\) is assumed to be independent of both \(T\) and \((X,\,Z),\) the simple Kaplan–Meier estimator satisfies this condition. When \(C\) depends on \((X,\,Z),\,{\widehat{G}}_{X,Z}\) obtained under the Cox model also satisfies this condition provided that \(W_c\) is bounded. The kernel function \(K\) is assumed to be symmetric, smooth with a bounded support on \([-1,\,1]\) and we let \(m_2 = \int K(x)^2dx.\)

1.1.1 A.1 Asymptotic expansions for \(\widehat{\mathcal{{S}}}_y(c;\,s)\)

Uniform convergence rate for \(\widehat{\mathcal{{S}}}_y(c;\,s)\) We first establish the following uniform convergence rate of \(\widehat{\mathcal{{S}}}_y(c;\,s)= g\{{\widehat{a}}_y(c;\,s)\}\):

$$\begin{aligned} \sup _{s\in \mathcal{{I}}_{h},c}|\widehat{\mathcal{S }}_{y}(c;\,s) - \mathcal S _{y}(c;\,s)|= O_{p}\{(nh)^{-\frac{1}{2}}\log (n)\} = o_{p}(1). \end{aligned}$$
(6)

To this end, we note that for any given \(c\) and \(s\),

$$\begin{aligned} {\widehat{\varvec{\zeta }}}_{y}(c;\,s) = \left[\begin{matrix}{\widehat{\zeta }}_{a_{y}}(c;\,s) \\ {\widehat{\zeta }}_{b_{y}}(c;\,s)\end{matrix}\right] =\left[\begin{matrix}{\widehat{a}}_{y}(c;\,s)-a_{y}(c;\,s) \\ {\widehat{b}}_{y}(c;\,s)-b_{y}(c;\,s) \end{matrix} \right] \end{aligned}$$

is the solution to the estimating equation \({\widehat{\varvec{\Psi }}}_{y}(\varvec{\zeta }_{y},\,c,\,s)=0,\) where \(\varvec{\zeta }_{y}=(\zeta _{a_{y}},\,\zeta _{b_{y}})^{\prime }\) and

$$\begin{aligned} {\widehat{\varvec{\Psi }}}_{y}\left(\varvec{\zeta }_{y};\,c,\,s\right)&= \left[\begin{array}{c}{\widehat{\varvec{\Psi }}}_{y1}\left(\varvec{\zeta }_{y},\,c,\,s\right) \\ {\widehat{\varvec{\Psi }}}_{y2}\left(\varvec{\zeta }_{y},\,c,\,s\right) \end{array}\right]\\&= n^{-1}\displaystyle \sum _{i:\, Y_{i}=y}\widehat{w}_{i}\left[\begin{array}{c}1 \\ h^{-1}\widehat{\mathcal{E }}_{i1}(s)\end{array}\right]K_{h}\left\{ \widehat{\mathcal{E }}_{i1}(s)\right\} \\&\times \left[{\widehat{M}}_{i}(c) - \mathcal G \left\{ \varvec{\zeta }_{y},\,c,\,s;\,\phi \left({\widehat{p}}_{1i}\right),\,h\right\} \right], \end{aligned}$$

\(a_{y}(c;\,s) = g^{-1}\{ { \mathcal S }_{y}(c;\,s)\},\,b_{y}(c;\,s)=\partial g^{-1}\{\mathcal{S }_{y}(c;\,s)\}/\partial s\) and \(\mathcal G (\varvec{\zeta }_{y},\,c,\,s;\, e,\,h)=g[a_{y}(c;\,s)+b_{y}(c;\,s)\{e-\phi (s)\} + \zeta _{a_{y}}+\zeta _{b_{y}}h^{-1}\{e-\phi (s) \}].\) We next establish the convergence rate for \(\sup _{\varvec{\zeta }_{y},\,c,\,s}|{\widehat{\varvec{\Psi }}}_{y}(\varvec{\zeta }_{y};\,c,\,s) - \varvec{\Psi }_{y}(\varvec{\zeta }_{y};\,c,\,s)|,\) where

$$\begin{aligned} \varvec{\Psi }_{y}\left(\varvec{\zeta }_{y};\,c,\,s\right)&= \left[\begin{array}{c} \varPsi _{y1}(\varvec{\zeta }_{y},\, c,\,s) \\ \varPsi _{y2}(\varvec{\zeta }_{y};\,c,\,s)\end{array}\right] \\&= \tau (y;\,s) \left[\begin{array}{c}\mathcal S _{y}(c;\,s)-\int K(t)g\{a_{y}(c;\,s)+\zeta _{a_{y}}+\zeta _{b_{y}}t\}dt\\ -\int tK(t)g\{a_{y}(c;\,s)+\zeta _{a_{y}}+\zeta _{b_{y}}t\}dt\end{array}\right]. \end{aligned}$$

We first show that

$$\begin{aligned} \sup _{s\in \mathcal{{I}}_{h},c}\left|n^{-1}\sum _{i:Y_{i}=y} {\widehat{\omega }}_{i} K_{h}\left\{ \widehat{\mathcal{E }}_{i1}(s)\right\} {\widehat{M}}_{i}(c) - \tau (y;\,s)\mathcal{{S}}_{y}(c;\,s)\right| \end{aligned}$$

and

$$\begin{aligned}&\sup _{\varvec{\zeta }_{y},s\in \mathcal{{I}}_{h},c}\left|n^{-1}\sum _{i:Y_{i}=y} {\widehat{\omega }}_{i} K_{h}\left\{ \widehat{\mathcal{E }}_{i1}(s)\right\} \mathcal G \left\{ \varvec{\zeta }_{y},\,c,\,s;\, \phi \left({\widehat{p}}_{1i}\right)\!,\,h\right\} \right. \\&\quad \quad \left.- \tau (y;\,s)\int K(t)g\left\{ a_{y}(c;\,s)+\zeta _{a_{y}}+\zeta _{b_{y}}t\right\} dt\right| \end{aligned}$$

are both \(O_{p}\{(nh)^{-\frac{1}{2}}\log (n)\}\) where \(\mathcal{{I}}_{h}=[\phi ^{-1}(\rho _{l}+h),\,\phi ^{-1}(\rho _{u}-h)]\) and \([\rho _{l},\,\rho _{u}]\) is a subset of the support of \(\phi \{g_{1}(\beta _{0}^{T}V)\}.\) To this end, we note that since \(\sup _{u} |{\widehat{G}}_{X,Z}(u)-G_{X,Z}(u)|=O_{p}(n^{-\frac{1}{2}})\) and \(|{\widehat{\beta }}-\beta _{0}|=O_{p}(n^{-\frac{1}{2}}),\)

$$\begin{aligned}&\left|n^{-1}\sum _{i:\,Y_{i}=y}\left({\widehat{\omega }}_{i} - \omega _{i}\right)K_{h}\left\{ \widehat{\mathcal{E }}_{i1}(s)\right\} \mathcal G \left\{ \varvec{\zeta }_{y},\,c,\,s;\, \phi \left({\widehat{p}}_{1i}\right),\,h\right\} \right|\\&\quad \le n^{-1}\sum _{i:\,Y_{i}=y} |{\widehat{\omega }}_{i}-\omega _{i}|K_{h}\left\{ \widehat{\mathcal{E }}_{i1}(s)\right\} \\&\quad = O_{p}\left(n^{-\frac{1}{2}}\right). \end{aligned}$$

This implies that

$$\begin{aligned}&\left|n^{-1}\sum _{i:\,Y_{i}=y} {\widehat{\omega }}_{i} K_{h}\left\{ \widehat{\mathcal{E }}_{i1}(s)\right\} \mathcal G \left\{ \varvec{\zeta }_{y},\,c,\,s;\, \phi \left({\widehat{p}}_{1i}\right),\,h\right\} - \tau (y;\,s)\int K(t)g\left\{ a_{y}(c;\,s)+\zeta _{a_{y}}+\zeta _{b_{y}}t\right\} dt\right|\\&\quad \le \left|n^{-\frac{1}{2}}\int K_{h}\{e-\phi (s)\}\mathcal G \left\{ \varvec{\zeta }_{y},\,c,\,s;\, \phi \left({\widehat{p}}_{1i}\right),\,h\right\} d \mathbb{G }_{n}\left[\omega I\left\{ \phi \left({\widehat{p}}_{i1}\right)\le e\right\} - \omega I\left\{ \phi \left(\bar{ p}_{i1}\right)\le e\right\} \right]\right|\\&\qquad + \left|\int K_{h}\{e-\phi (s)\}\mathcal G \left\{ \varvec{\zeta }_{y},\,c,\,s;\, \phi \left({\widehat{p}}_{1i}\right),\,h\right\} d\mathbb{P }\left[\omega I\left\{ \phi \left(\bar{ p}_{i1}\right)\le e\right\} \right] - \tau (y;\,s)\int K(t)g\left\{ a_{y}(c;\,s)+\zeta _{a_{y}}+\zeta _{b_{y}}t\right\} dt\right|\\&\qquad + \left|n^{-\frac{1}{2}}\int K_{h}\{e-\phi (s)\}d\mathbb{P }\left[\omega \mathcal G \left\{ \varvec{\zeta }_{y},\, c,\,s;\,\phi \left({\widehat{p}}_{1i}\right),\,h\right\} I\left\{ \phi \left(\bar{ p}_{i1}\right)\le e\right\} \right] \right| +O_{p}\left(n^{-\frac{1}{2}}\right)\lesssim n^{-\frac{1}{2}}h^{-1}\Vert \mathbb{G }_{n}\Vert _\mathcal{H _{\delta }} + \left|n^{-\frac{1}{2}}\int K_{h}\{e-\phi (s)\}d\mathbb{P }\left[\omega \mathcal G \left\{ \varvec{\zeta }_{y},\,c,\,s;\,\phi \left({\widehat{p}}_{1i}\right)\!,\,h\right\} I\left\{ \phi \left(\bar{ p}_{i1}\right)\le e\right\} \right] \right| + O_{p}\left(n^{-\frac{1}{2}}+h^{2}\right)\!, \end{aligned}$$

where \(\mathcal H _{\delta }=\{\omega I[\phi \{g_{1}(\beta ^{\prime }v)\}\le e]-\omega I[\phi \{g_{1}(\beta _{0}^{\prime }v)\}\le e]:\,|\beta -\beta _{0}|\le \delta ,\,e\}\) is a class of functions indexed by \(\beta \) and \(e.\) By the maximum inequality of Van der Vaart and Wellner (1996), we have

$$\begin{aligned} E\Vert \mathbb{G }_{n}\Vert _\mathcal{H _{\delta }} \lesssim \delta ^{\frac{1}{2}} \{|\log (\delta )|+|\log (h)|\}\left[1 + \frac{\delta ^{\frac{1}{2}}\{|\log (\delta )|+|\log (h)|\}}{\delta n^{\frac{1}{2}}}\right] \end{aligned}$$

Together with the fact that \(|{\widehat{\beta }}-\beta _{0}|=O_{p}(n^{-\frac{1}{2}})\) from Uno et al. (2007), it implies that \(n^{-\frac{1}{2}}h^{-1}\Vert \mathbb{G }_{n}\Vert _\mathcal{H _{\delta }} = O_{p}\{(nh)^{-\frac{1}{2}}(nh^{2})^{-\frac{1}{4}}\log (n)\}.\) In addition, with the standard arguments used in Bickel and Rosenblatt (1973), it can be shown that

$$\begin{aligned}&\left|n^{-\frac{1}{2}}\int K_{h}\{e-\phi (s)\}d\mathbb{P }\left[\omega \mathcal G \left\{ \varvec{\zeta }_{y},\,c,\,s;\,\phi \left({\widehat{p}}_{1i}\right)\!,\,h\right\} I\left\{ \phi \left(\bar{ p}_{i1}\right)\le e\right\} \right] \right|\\&\quad =O_{p}\left\{ (nh)^{-\frac{1}{2}}\log (n)\right\} . \end{aligned}$$

Therefore, for \(h=n^{-\nu },\,1/5 <\nu < 1/2,\)

$$\begin{aligned}&\sup _{\varvec{\zeta }_{y},s\in \mathcal{{I}}_{h},c}\left|n^{-1}\sum _{i:\,Y_{i}=y} {\widehat{\omega }}_{i} K_{h}\left\{ \widehat{\mathcal{E }}_{i1}(s)\right\} \mathcal G \left\{ \varvec{\zeta }_{y},\,c,\,s;\, \phi \left({\widehat{p}}_{1i}\right),\,h\right\} \right.\\&\quad \left. - \tau (y;\,s)\int K(t)g\left\{ a_{y}(c;\,s)+\zeta _{a_{y}}+\zeta _{b_{y}}t\right\} dt\right| \end{aligned}$$

is \(O_{p}\{(nh)^{-\frac{1}{2}}\log (n)\}.\) Following with similar arguments as given above, coupled with the fact that \(|{\widehat{\gamma }}-\gamma _{0}|=O_{p}(n^{-\frac{1}{2}}),\) we have

$$\begin{aligned} \sup _{s\in \mathcal{{I}}_{h},c \!\in \! [0,1]}\left|\!n\!^{-1}\!\sum _{i:\,Y_{i}\!=\!y} {\widehat{\omega }}_{i} K_{h}\left\{ \widehat{\mathcal{E }}_{i1}(s)\right\} {\widehat{M}}_{i}(c) \!-\! \tau (y;\,s)\mathcal{{S}}_{y}(c;\,s)\right|\!=\! O_{p}\left\{ (nh)^{-\frac{1}{2}}\log (n)\right\} \!. \end{aligned}$$

Thus, \(\sup _{\varvec{\zeta }_{y},c,s} |{\widehat{\Psi }}_{y1}(\varvec{\zeta }_{y};c,s) - \varPsi _{y1}(\varvec{\zeta }_{y};c,s)|=O_{p}\{(nh)^{-\frac{1}{2}}\log (n)\}=o_{p}(1).\) It follows from the same arguments as given above that

$$\begin{aligned} \sup _{\varvec{\zeta }_{y},c,s}\left|{\widehat{\Psi }}_{y2}\left(\varvec{\zeta }_{y};\,c,\,s\right) - \varPsi _{y2}\left(\varvec{\zeta }_{y};\,c,\,s\right)\right|=O_{p}\left\{ (nh)^{-\frac{1}{2}}\log (n)+h\right\} =o_{p}(1). \end{aligned}$$

Therefore, \(\sup _{\varvec{\zeta }_{y},c,s} |{\widehat{\varvec{\Psi }}}_{y}(\varvec{\zeta }_{y};\,c,\,s)-\varvec{\Psi }_{y}(\varvec{\zeta }_{y};\,c,\,s)|=o_{p}(1).\) In addition, we note that \(\mathbf 0 \) is the unique solution to the equation \(\varvec{\Psi }_{y}(\varvec{\zeta }_{y};\,c,\,s) = 0\) wrt \(\varvec{\zeta }_{y}.\) It suggests that \(\sup _{s,c}|{\widehat{\varvec{\zeta }}}_{a_{y}}(c;\,s)|=O_{p}\{(nh)^{-\frac{1}{2}}\log (n)\}=o_{p}(1),\) which implies the consistency of \(\mathcal S _{y}(c;\,s),\)

$$\begin{aligned} \sup _{s\in \mathcal{{I}}_{h},c\in [0,1]}\left|\widehat{\mathcal{S }}_{y}(c;\,s) - \mathcal S _{y}(c;\,s)\right|=O_{p}\left\{ (nh)^{-\frac{1}{2}}\log (n)\right\} = o_{p}(1). \end{aligned}$$

Asymptotic expansion for \(\widehat{\mathcal{{S}}}_y(c;\,s)\) Let \({\widehat{d}}_{y}(c;\,s)=\sqrt{nh}\{{\widehat{a}}_{y}(c;\,s)-a_{y}(c;\,s)\}.\) It follows from a Taylor series expansion and the convergence rate of \(\varvec{\zeta }_y(c;\,s)\) that

$$\begin{aligned} {\widehat{d}}_{y}(c;\,s) = \frac{\sqrt{nh}\mathbb{P }_{n}({\widehat{\omega }}I(Y=y)K_{h}\{\widehat{\mathcal{E }}_{1}(s)\}[{\widehat{M}}(c) - \mathcal G ^{0}_{y}\{c,\,s;\,\phi ({\widehat{p}}_{1})\}])}{\tau \{y;\,\phi (s)\} \dot{g}\{a_{y}(c;\,s)\}} + o_p(1), \nonumber \\ \end{aligned}$$
(7)

where \(\mathcal G ^{0}_{y}(c,\,s;\,e) = g[a_{y}(c;\,s)+b_{y}(c;\,s)\{e-\phi (s)\}].\) Furthermore, since \(\sup _{t\le t_{0}} |{\widehat{G}}_{X,Z}(t)-G_{X,Z}(t)|=O_{p}(n^{-1/2}),\)

$$\begin{aligned} {\widehat{d}}_{y}(c;\,s)=\frac{\sqrt{nh}\mathbb{P }_{n}(\omega I(Y=y)K_{h}\{\widehat{\mathcal{E }}_{1}(s)\}[{\widehat{M}}(c) - \mathcal G ^{0}_{y}\{c,\,s;\,\phi ({\widehat{p}}_{1})\}])}{\tau \{y;\,\phi (s)\} \dot{g}\{a_{y}(c;\,s)\}} + o_p(1). \end{aligned}$$

We next show that \(\widehat{d}_{y}(c;\,s)\) is asymptotically equivalent to

$$\begin{aligned} \widetilde{d}_{y}(c;\,s) = \frac{\sqrt{nh}\mathbb{P }_n(\omega I(Y=y)K_{h}\{\bar{\mathcal{E }}_{1}(s)\}[\bar{M}(c) - \mathcal G _{y}^{0} \{c,\,s;\,\phi (\bar{ p}_{1})\}])}{\tau \{y;\,\phi (s)\} \dot{g}\{a_{y}(c;\,s)\}}, \end{aligned}$$
(8)

where \(\bar{\mathcal{E }}_{1}(s) = \phi (\bar{ p}_{1})-\phi (s).\) From (8) and the fact that \(\tau \{y;\,\phi (s)\}\) is bounded away from 0 uniformly in \(s,\) we have

$$\begin{aligned}&\left|\widehat{d}_{y}(s)-\widetilde{d}_{y}(s)\right|\\&\quad \lesssim h^{\frac{1}{2}}\left|\int \! K_{h}\{e\!-\!\phi (s)\}d\mathbb{G }_{n}\left(\!I(Y\!=\! y)\omega \left[{\widehat{M}}(c)I\left\{ \phi \left({\widehat{p}}_{1}\right)\!\le \! e\right\} \!-\!\bar{M}(c) I\left\{ \phi \left(\bar{ p}_{1}\right)\le e\right\} \right]\,\right)\right|\\&\qquad +h^{\frac{1}{2}}\left|\!\int \!K_{h}\{e\!-\!\phi (s)\}\mathcal G _{y}(c,\,s;\,e)d\mathbb{G }_{n}\left(\!I(Y \!=\!y)\left[\omega I\left\{ \phi \left({\widehat{p}}_{1}\right)\!\le \! e\right\} \!-\!\omega I\left\{ \phi \left(\bar{ p}_{1}\right)\!\le \!e\right\} \right]\,\right)\right|\\&\qquad +\left|\!\sqrt{nh}\!\int \!K_{h}\{e\!-\!\phi (s)\} d\mathbb{P }\left(\!I(Y\!=\! y)\left[\omega {\widehat{M}}(c)I\left\{ \phi \left({\widehat{p}}_{1}\right)\!\le \! e\right\} \!-\!\omega \bar{M}(c)I\left\{ \phi \left(\bar{ p}_{1}\right)\!\le \!e\right\} \right]\,\right)\right|\\&\qquad + \left|\sqrt{nh}\int K_{h}\{e\!-\!\phi (s)\} d\mathbb{P }\left(I(Y\!=\! y)\left[\omega \mathcal G _{y}\left\{ c,\,s;\,\phi \left({\widehat{p}}_{1}\right)\right\} I\left\{ \phi \left({\widehat{p}}_{1}\right)\le e\right\} \right.\right. \right.\\&\qquad -\left. \left. \left. \omega \mathcal G _{y}\left\{ c,\,s;\,\phi \left(\bar{ p}_{1}\right)\right\} I\left\{ \phi \left(\bar{ p}_{1}\right)\le e\right\} \right]\right)\right|\\&\quad \lesssim h^{\frac{1}{2}}\left\Vert\mathbb{G }_{n}\right\Vert_\mathcal{F _{\delta }}\!+\! h^{\frac{1}{2}}\left\Vert\mathbb{G }_{n}\right\Vert_\mathcal{H _{\delta }} \!+\!O_{p}\left\{ (nh)^{1/2}\left|{\widehat{\beta }}-\beta _{0}\right|\!+\!\left|{\widehat{\gamma }}\!-\!\gamma _{0}\right|\!+\!h^{2}\right\} , \end{aligned}$$

where \(\mathcal F _{\delta }\!=\!\{\omega I\{g_{2}(\gamma ^{\prime }w)\!\ge \! c\}I[\phi \{g_{1}(\beta ^{\prime }v)\}\le e]\!-\!\omega I\{g_{2}(\gamma _{0}^{\prime }w)\!\ge \! c\} I[\phi \{g_{1}(\beta _{0}^{\prime }v)\}\!\le \! e]:\,|\gamma -\gamma _{0}|+|\beta -\beta _{0}|\le \delta ,\,e\}\) is the class of functions indexed by \(\gamma ,\,\beta \) and \(e.\) By the maximum inequality of Van der Vaart and Wellner (1996) and the fact that \(|{\widehat{\beta }}-\beta _{0}|+|{\widehat{\gamma }}-\gamma _{0}|=O_{p}(n^{-\frac{1}{2}})\) from Uno et al. (2007), we have \(h^{\frac{1}{2}}\left\Vert\mathbb{G }_{n}\right\Vert_\mathcal{F _{\delta }} = O_{p}\{h^{-\frac{1}{2}}n^{-\frac{1}{4}}\log (n)\}\) and \(h^{\frac{1}{2}}\left\Vert\mathbb{G }_{n}\right\Vert_\mathcal{H _{\delta }} = O_{p}\{h^{-\frac{1}{2}}n^{-\frac{1}{4}}\log (n)\}.\) It follows that \(\sup _{s}|\widehat{d}_{y}(s)-\widetilde{d}_{y}(s)|=o_{p}(1).\) Then, by a delta method,

$$\begin{aligned}&\widehat{\mathcal{{W}}}_{\mathcal{{S}}_y}(c;\,s)=\sqrt{nh}\left\{ \widehat{\mathcal{S }}_{y}(c;\,s)- \mathcal S _{y}(c;\,s)\right\} \simeq \sqrt{nh}\ \mathbb{P }_n\left[K_{h}\left\{ \bar{\mathcal{E }}_{1}(s)\right\} \mathcal{{D}}_{\mathcal{{S}}_y}(c;\,s) \right]\qquad \end{aligned}$$
(9)
$$\begin{aligned}&\text{ where}\qquad \mathcal{{D}}_{\mathcal{{S}}_y}(c;\,s)= \tau \{y;\,\phi (s)\}^{-1}\omega I(Y=y)\left\{ \bar{M}(c)- \mathcal S _{y}(c;\,s)\right\} \qquad \qquad \end{aligned}$$
(10)

Using the same arguments as for establishing the uniform convergence rate of conditional Kaplan–Meier estimators (Dabrowska 1989; Du and Akritas 2002), we obtain (6). Furthermore, following similar arguments as given in Dabrowska (1987, 1997), we have \(\widehat{\mathcal{{W}}}_{\mathcal{{S}}_y}(c;\,s)\) converges weakly to a Gaussian process in \(c\) for all \(s.\) Note that as for all kernel estimators, \(\widehat{\mathcal{{W}}}_{\mathcal{{S}}_y}(c;\,s)\) does not converge as a process in \(s.\)

1.1.2 A.2 Uniform consistency of \(\widehat{\text{ pAUC}}_{f}(s)\)

Next we establish the uniform convergence rate for \(\widehat{\mathrm{{ROC}}}(u;\,s).\) To this end, we write

$$\begin{aligned} \widehat{\text{ ROC}}(u;\,s) - \text{ ROC}(u;\,s) = {\widehat{\varepsilon }}_1(u;\,s)+{\widehat{\varepsilon }}_0(u;\,s), \end{aligned}$$

where \({\widehat{\varepsilon }}_1(u;\,s) = { \widehat{ \mathcal S }}_{1}\{ { \widehat{\mathcal{S }}}_{0}^{-1}(u;\,s);\,s\} -{ \mathcal S }_{1}\{ { \widehat{\mathcal{S }}}_{0}^{-1}(u;\,s);\,s\}\) and  \({\widehat{\varepsilon }}_0(u;\,s)\!=\! { \mathcal S }_{1}\{ { \widehat{\mathcal{S }}}_{0}^{-1} (u;\,s); \,s\}\!-\! \mathcal{S }_{1}\{ { \mathcal S }_{0}^{-1}(u;\,s);\,s\}.\) It follows from (6) that \(\sup _{u;s}|{\widehat{\varepsilon }}_1(u;s)|\!\le \! \sup _{c;s}|\widehat{ { \mathcal S }}_{1} (c;\,s)\!-\!\mathcal{{S}}_1(c;\,s)|.\) Let \(\widehat{\mathcal{I}}(u;\,s)\!=\!\mathcal{{S}}_0\{\widehat{\mathcal{{S}}}_0^{-1}(u;\,s);\,s\}.\) Then \({\widehat{\varepsilon }}_0(u;\,s)\!=\!\mathrm{{ROC}}\{\widehat{\mathcal{I}}(u;\,s);s\} - \mathrm{{ROC}}(u;\,s).\) Noting that \(\sup _{u}|\widehat{\mathcal{I}}(u;\,s)\!-\!u|\!=\!\sup _u |\widehat{\mathcal{I}}(u;\,s)\!-\! \widehat{\mathcal{{S}}}_0\{\widehat{\mathcal{{S}}}_0^{-1}(u;\,s);\,s\}|\!+\!n^{-1}\!\le \!\sup _c |\mathcal{{S}}_0(c;\,s)\!-\!\widehat{\mathcal{{S}}}_0(c;\,s)|\!+\!n^{-1}\!=\!O_{p}\{(nh)^{-1/2}\log n\},\) we have \({\widehat{\varepsilon }}_0(u;\,s)=O_{p}\{(nh)^{-1/2}\log n\}\) by the continuity and boundedness of \(\dot{\mathrm{{ROC}}}(u;\,s)\). Therefore,

$$\begin{aligned} \sup _{u,s}|\widehat{\text{ ROC}}(u;\,s) - \text{ ROC}(u;\,s)| = O_p\left\{ (nh)^{-1/2}\log n\right\} \end{aligned}$$

which implies

$$\begin{aligned}&\displaystyle \sup _{s \in \mathcal{{I}}_h}\left|\widehat{\mathrm{{pAUC}}}_f(s) - \mathrm{{pAUC}}_f(s)\right|\\&\qquad \lesssim \displaystyle \sup _{s \in \mathcal{{I}}_h} \int \limits _{0}^{f} \left|\widehat{\text{ ROC}}(u;\,s)-\text{ ROC}(u;\,s)\right| du =O_p\left\{ (nh)^{-\frac{1}{2}}\log n\right\} . \end{aligned}$$

and hence the uniform consistency of \(\widehat{\mathrm{{pAUC}}}_f(s).\)

1.1.3 A.3 Asymptotic distribution of \(\widehat{\mathcal{{W}}}_{\mathrm{{\tiny pAUC}}_f}(s)\)

To derive the asymptotic distribution for \(\widehat{\mathcal{{W}}}_{\mathrm{{\tiny pAUC}}_f}(s),\) we first derive asymptotic expansions for \(\widehat{\mathcal{{W}}}_{\mathrm{{\tiny ROC}}}(u;\,s)=\sqrt{nh}\{\widehat{\text{ ROC}}(u;\,s) - \text{ ROC}(u;\,s)\} = \sqrt{nh}\ {\widehat{\varepsilon }}_1(u;\,s) + \sqrt{nh}\ {\widehat{\varepsilon }}_0(u;\,s).\) From the weak convergence of \(\widehat{\mathcal{{W}}}_{S_y}(c;\,s)\) in \(c,\) the approximation in (9), and the consistency of \(\widehat{\mathcal{{S}}}_0^{-1}(c;\,s)\) given in the Appendix A.2, we have

$$\begin{aligned} \sqrt{nh}\ {\widehat{\varepsilon }}_1(u;\,s)&\simeq \, \sqrt{nh}\left[{ \widehat{ \mathcal S }}_{1}\left\{ { \mathcal S }_{0}^{-1}(u;\,s);\,s\right\} - \mathrm{{ROC}}(u;\,s) \right]\\&\simeq \, \sqrt{nh}\mathbb{P }_n\left[K_{h}\left\{ { \bar{\mathcal{E }}}_{1}(s)\right\} \mathcal{{D}}_{\mathcal{{S}}_1}\left\{ { \mathcal S }_{0}^{-1}(u;\,s);\,s\right\} \right] \end{aligned}$$

On the other hand, from the uniform convergence of \(\widehat{\mathcal{I}}_0(u;\,s) \rightarrow u\) and the weak convergence of \(\widehat{\mathcal{{D}}}_0(c;\,s)\) in \(c,\) we have

$$\begin{aligned} \sqrt{nh}\left\{ u -\widehat{\mathcal{I}}(u;\,s) \right\}&\simeq \, \sqrt{nh}\left[\widehat{\mathcal{I}}^{-1}\left\{ \widehat{\mathcal{I}}(u;\,s);\,s\right\} -\widehat{\mathcal{I}}(u;\,s) \right] \simeq \sqrt{nh}\left\{ \widehat{\mathcal{I}}^{-1}(u;\,s) -u \right\} \\&\simeq \, \sqrt{nh}\left[\widehat{\mathcal{{S}}}_0\{\mathcal{{S}}_0^{-1}(u;s);s\} - u\right] \end{aligned}$$

This, together with a Taylor series expansion and the expansion given (9), implies that

$$\begin{aligned} \sqrt{nh}\ {\widehat{\varepsilon }}_0(u;\,s) \simeq - \dot{\mathrm{{ROC}}}(u;\,s) \mathbb{P }_n\left[K_{h}\left\{ \bar{\mathcal{E }}_{1}(s)\right\} \mathcal{{D}}_{\mathcal{{S}}_0}\left\{ \mathcal S _{0}^{-1}(u;\,s);\,s\right\} \right] \end{aligned}$$

It follows that

$$\begin{aligned} \widehat{\mathcal{{W}}}_{\mathrm{{\tiny pAUC}}_f}(s)&\simeq&\sqrt{nh}\mathbb{P }_n\left[K_{h}\left\{ \bar{\mathcal{E }}_{1}(s)\right\} \mathcal{{D}}_{\mathrm{{\tiny pAUC}}_f}(s) \right] \end{aligned}$$
(11)
$$\begin{aligned} \text{ where} \quad \mathcal{{D}}_{\mathrm{{\tiny pAUC}}_f}(s)&\!=\!&\! \int \limits _0^f \!\left[ \mathcal{{D}}_{\mathcal{{S}}_1}\!\left\{ \mathcal{{S}}_0^{-1}(u;\,s);\,s \right\} \!-\! {\dot{\mathrm{{ROC}}}}(u;\,s)\mathcal{{D}}_{\mathcal{{S}}_0}\!\left\{ \mathcal{{S}}_0^{-1}(u;\,s);\,s \right\} \right] du .\nonumber \\ \end{aligned}$$
(12)

It then follows from a central limit theorem that for any fixed \(s,\,\widehat{\mathcal{{W}}}_{\mathrm{{\tiny pAUC}}_f}(s)\) converges to a normal with mean 0 and variance

$$\begin{aligned} \sigma _{\mathrm{{\tiny pAUC}}_f}^{2}(s)\!=\!m_{2}\left[\tau \{1;\,\phi (s)\}\dot{F}_{\small \phi (\bar{ p}_{1})}(s)\right]^{-1} \sigma _{1}^{2}(s)\!+\!m_{2}\left[\tau \{0;\,\phi (s)\}\dot{F}_{\small \phi (\bar{ p}_{1})}(s)\right]^{-1}\sigma _{0}^{2}(s), \end{aligned}$$

where \(\dot{F}_{\small \phi (\bar{ p}_{1})}(s)\) is the density function of \(\phi (\bar{ p}_{1}),\)

$$\begin{aligned} \sigma _{1}^{2}(s)\!&= \! \mathrm{{E}}\!\left.\left(\!G\left(T^{\dag }\right)^{-1}\!\left[\int \limits _{0}^{f} \bar{M}\left\{ { \mathcal S }_{0}^{-1}(u;\,s)\right\} du\!-\!\mathrm{{pAUC}}_{f}(s)\right]^{2} \right| \bar{ p}_{1}\!=\!s,\;Y^{\dag }\!=\!1\right)\!,\quad \text{ and} \\ \sigma _{0}^{2}(s) \!&= \! \mathrm{{E}}\!\left.\left(\!G\left(t_{0}\right)^{-1}\!\left[\int \limits _{0}^{f} \bar{M}\left\{ { \mathcal S }_{0}^{-1}(u;\,s)\right\} d\mathrm{{ROC}}(u;\,s)\!-\!\int \limits _{0}^{f}u d\mathrm{{ROC}}(u;\,s)\!\right]^{2}\right| \bar{ p}_{1}\!=\!s,\; Y^{\dag }\!=\!0\!\right)\!. \end{aligned}$$

1.1.4 A.4 Justification for the resampling methods

To justify the resampling method, we first note that \(|\beta ^{*} - {\widehat{\beta }}| + |\gamma ^{*}-{\widehat{\gamma }}| + \sup _{t \le t_0}|G_{X,Z}^{*}(t)-{\widehat{G}}_{X,Z}(t)|=O_p(n^{-\frac{1}{2}}).\) It follows from similar arguments given in the Appendix A and Appendix 1 of Cai et al. (2010) that \(\mathcal{{W}}^{*} _{S_y}(c;\,s) = \sqrt{nh}\{\mathcal{{S}}_y^{*}(c;\,s)-\widehat{\mathcal{{S}}}_y(c;\,s)\} \simeq n^{\frac{1}{2}}h^{-1/2}\sum _{i=1}^n\widehat{\mathcal{{D}}}_{\mathcal{{S}}_y i}(c;\,s)\xi _i,\) where \(\widehat{\mathcal{{D}}}_{\mathcal{{S}}_y i}(c;\,s)\) is obtained by replacing all theoretical quantities in \(\mathcal{{D}}_{\mathcal{{S}}_y}(c;\,s)\) given in (10) with the estimated counterparts for the \(i\)th subject. This, together with similar arguments as given above for the expansion of \(\widehat{\mathcal{{W}}}_{\mathrm{{\tiny ROC}}}(u;\,s),\) implies that

$$\begin{aligned} \mathcal{{W}}^{*} _{\mathrm{{\tiny pAUC}}_f}(s)&= \int \limits _0^f \sqrt{nh}\{\mathrm{{ROC}}^*(u;\,s)-\widehat{\mathrm{{ROC}}}(u;\,s)\} du\\&\simeq&n^{-\frac{1}{2}}h^{-1/2}\sum _{i=1}^nK_{h}\left\{ \widehat{\mathcal{E }}_{1}(s)\right\} \widehat{\mathcal{{D}}}_{\mathrm{{\tiny pAUC}}_f}(s)\xi _i, \end{aligned}$$

where \(\widehat{\mathcal{{D}}}_{\mathrm{{\tiny pAUC}}_f}(s) = \int _0^f [\widehat{\mathcal{{D}}}_{\mathcal{{S}}_1 i} \{ \widehat{\mathcal{{S}}}_0^{-1}(u;\,s);\,s \}-{\dot{\mathrm{{ROC}}}}(u;\,s)\widehat{\mathcal{{D}}}_{\mathcal{{S}}_0 i} \{ \widehat{\mathcal{{S}}}_0^{-1}(u;\,s);\,s \}] du.\) Conditional on the data, \(\mathcal{{W}}^{*}_{\mathrm{{\tiny pAUC}}_f} (s)\) is approximately normally distributed with mean 0 and variance

$$\begin{aligned} {\widehat{\sigma }}_{\mathrm{{\tiny pAUC}}_f}^{2}(s) = h^{-1}\sum _{i=1}^nK_{h}\left\{ \widehat{\mathcal{E }}_{1}(s)\right\} ^2\widehat{\mathcal{{D}}}_{\mathrm{{\tiny pAUC}}_f}(s)^2. \end{aligned}$$

Using the consistency of the proposed estimators along with similar arguments as given above, it is not difficult to show that the above variance converges to \(\sigma _{\mathrm{{\tiny pAUC}}_f}^{2}(s)\) as \(n\rightarrow \infty .\) Therefore, the empirical distribution obtained from the perturbed sample can be used to approximate the distribution of \(\widehat{\mathcal{{W}}}_{\mathrm{{\tiny pAUC}}_f}(s).\)

We now show that after proper standardization, the supermum type statistics \(\varGamma \) converges weakly. To this end, we first note that, similar arguments as given in the Appendix A can be used to show that \(\sup _{s\in \mathcal{{I}}_h}|{\widehat{\sigma }}_{\mathrm{{\tiny pAUC}}_f}^{2}(s)- \sigma _{\mathrm{{\tiny pAUC}}_f}^{2}(s)|= o_{p}(n^{-\delta })\) and

$$\begin{aligned} \varGamma = \sup _{s \in \mathcal{{I}}_h} \left|\frac{\sqrt{nh}\mathbb{P }_n[K_{h}\{\bar{\mathcal{E }}_{1}(s)\} \mathcal{{D}}_{\mathrm{{\tiny pAUC}}_f}(s)]}{\sigma _{\mathrm{{\tiny pAUC}}_f}(s)}\right| + o_{p}(n^{-\delta }), \end{aligned}$$

for some small positive constant \(\delta .\) Using similar arguments in Bickel and Rosenblatt (1973), we have

$$\begin{aligned} pr\left\{ a_{n}\left(\varGamma -d_{n}\right)<x\right\} \rightarrow e^{-2e^{-x}}, \end{aligned}$$

where \(a_{n} = [2\log \{(\rho _{u}-\rho _{l})/h\}]^{\frac{1}{2}}\) and \(d_{n}=a_{n}+a_{n}^{-1}\log \{\int \dot{K}(t)^{2}dt/(4m_{2}\pi )\}.\) Now to justify the resampling procedure for constructing the CI, we note that

$$\begin{aligned} \mathcal{{W}}^*_{\mathrm{{\tiny pAUC}}_f}(s) = n^{-\frac{1}{2}}h^{\frac{1}{2}}\sum _{i=1}^{n} K_{h}\left\{ \widehat{\mathcal{E }}_{1i}(s)\right\} \widehat{\mathcal{{D}}}_{\mathrm{{\tiny pAUC}}_f}(s) \left(\xi _{i}-1\right)+\varepsilon ^*(s) \end{aligned}$$

where \(pr\{\sup _{s\in \varOmega (h)}|n^{\delta }\varepsilon ^*(s)|\ge e \mid \text{ data}\} \rightarrow 0\) in probability. Therefore,

$$\begin{aligned} \varGamma ^{*} = \sup _{s \in \mathcal{{I}}_{h}} \left| \frac{n^{-\frac{1}{2}}h^{\frac{1}{2}}\sum _{i = 1}^{n}K_{h}\{\widehat{\mathcal{E }}_{1i}(s)\} \widehat{\mathcal{{D}}}_{\mathrm{{\tiny pAUC}}_f}(s) (\xi _{i}-1) }{\sigma _{\mathrm{{\tiny pAUC}}_f}(s)}\right|+ \left|\varepsilon _{\mathrm{{\tiny sup}}}^*\right|. \end{aligned}$$

where \(pr\{|n^{\delta }\varepsilon _{\mathrm{{\tiny sup}}}^*| \ge e | \text{ data}\} \rightarrow 0.\) It follows from similar arguments as given in Tian et al. (2005) and Zhao et al. (2010) that

$$\begin{aligned} \sup \left| pr\left\{ a_{n}\left(\varGamma ^{*}-d_{n}\right)<x|\text{ data}\right\} - e^{-2e^{-x}} \right| \rightarrow 0, \end{aligned}$$

in probability as \(n \rightarrow \infty .\) Thus, the conditional distribution of \(a_{n}(\varGamma ^{*}-d_{n})\) can be used to approximate the unconditional distribution of \(a_{n}(\varGamma -d_{n}).\) When \(h_0 = h_1,\) in general, the standardized \(\varGamma \) does not converge to the extreme value distribution. However, when \(h_0=h_1 = k \in (0,\,\infty ),\) the distribution of the suitable standardized version of \(\varGamma \) still can be approximated by that of the corresponding standardized \(\varGamma ^*\) conditional on the data (Gilbert et al. 2002).

1.2 Appendix B

1.2.1 B.1 Bandwidth selection for \(\mathrm{{pAUC}}_{f}(s)\)

The choice of the bandwidths \(h_{0}\) and \(h_{1}\) is important for making inference about \(\mathcal S _{y}(c;\,s)\) and consequently \(\text{ pAUC}_{f}(s).\) Here we propose a two-stage K-fold cross-validation procedure to obtain the optimal bandwidth for \(\widehat{\mathcal{S }}_{0,h_{0}}^{-1}(u;\,s)\) and \(\widehat{\mathcal{S }}_{1,h_{1}}(c;\,s)\) sequentially. Specifically, we randomly split the data into K disjoint subsets of about equal sizes denoted by \(\{\mathcal{J }_{k},\, k=1,\ldots ,K\}.\) The two-stage procedure is described as follows:

  1. (I)

    Motivated by the fact that \(\mathcal{{S}}_{0}^{-1}(u;\,s)\) is essentially the \((1-u)\)-th quantile of the conditional distribution of \(\bar{ p}_{2}(X,\,Z)\) given \(Y^{\dag }=0\) and \(\bar{ p}_{1}(X)=s,\) for each k, we use all the observations not in \(\mathcal J _{k}\) to estimate \(q_{0,1-u}(s)=\mathcal S _{0}^{-1}(u;\,s)\) by obtaining \(\{{\widehat{\alpha }}_{0}(s;\,h),\,{\widehat{\alpha }}_{1}(s;\,h)\},\) the minimizer of

    $$\begin{aligned} \displaystyle \sum _{j \in \mathcal J _{l},l\ne k} I\left(Y_{j}=0\right)\widehat{w}_{j}K_{h}\left\{ \widehat{\mathcal{E }}_{1j}(s)\right\} \rho _{1-u}\left[{\widehat{p}}_{2j}-g\left\{ \alpha _{0} + \alpha _{1}\widehat{\mathcal{E }}_{1j}(s)\right\} \right] \end{aligned}$$

    wrt \((\alpha _{0},\,\alpha _{1}),\) where \(\rho _{\tau }(e)\) is a check function defined as \(\rho _{\tau }(e)=\tau e,\) if \(e\ge 0;\,=(\tau -1)e,\) otherwise. Let \(\widehat{q}_{0,1-u}^{(-k)}(s;\,h)=g\{{\widehat{\alpha }}_{0}(s;\,h)\}\) denote the resulting estimator of \(q_{0,1-u}(s).\) With observations in \(\mathcal J _{k},\) we obtain

    $$\begin{aligned} Err^{(q_0)}_{k}(h) = \displaystyle \sum _{i\in \mathcal J _{k}} \left(1-Y_i\right)\widehat{w}_{i}\int \limits _{0}^{f} \rho _{1-u}\left[{\widehat{p}}_{2i} - \widehat{q}_{0,1-u}^{(-k)}\left({\widehat{p}}_{1i};\,h\right)\right]du. \end{aligned}$$

    Then, we let \(h_0^{\mathrm{{\tiny opt}}}=\arg \min _{h}{\sum _{k=1}^{K}Err^{(q_0)}_{k}(h)}.\)

  2. (II)

    Next, to find an optimal \(h_1\) for \(\widehat{\mathcal{{S}}}_{1,h_1}(\cdot ;\,s),\) we choose an error function that directly relates to \(\mathrm{{pAUC}}_f(s)= - \int _{\mathcal{{S}}_0^{-1}(f;\,s)}^{\infty }\mathcal{{S}}_1(c;\,s)d\mathcal{{S}}_0(c;\,s).\) Specifically, noting the fact that

    $$\begin{aligned} E\!\!\left(\,\,\!\left.\int \limits _{\!\mathcal{{S}}^{-1}_{0}(f;\,s)}^{\infty } \left[\!I\left\{ g_{2}\left(\gamma ^{\prime }W_i\!\right)\!\ge \! c\!\right\} \!-\!\mathcal{{S}}_1(c;\,s)\right] d \mathcal S _{0}(c;\,s) \right| Y^{\dag }_{i}\!=\!1,\; g_1\left(\beta ^{\prime }X_i\!\right)\!=s\!\right)\!\!=0, \end{aligned}$$

    we use the corresponding mean integrated squared error for \(I\{g_{2}(\gamma ^{\prime }W_i)\ge c\} - \mathcal{{S}}_1(c;\,s)\) as the error function. For each \(k,\) we use all the observations which are not in \(\mathcal J _{k}\) to obtain the estimate of \(\mathcal S _{1}(c;\,s),\) denoted by \(\widehat{\mathcal{S }}_{1,h}^{(-k)}(c;\,s)\) via (4). Then, with the observations in \(\mathcal J _{k},\) we calculate the prediction error

    $$\begin{aligned} Err^{(\mathcal{{S}}_1)}_{k}(h)&\!=&\displaystyle \!-\!\sum _{i \in \mathcal J _{k},Y_{i}\!=\!1} \widehat{w}_{i} \!\int \limits _{\widehat{\mathcal{{S}}}_{0,h_{0}}^{-1}(f;\,{\widehat{p}}_{1i})}^{\infty }\! \!\left\{ \!I\!\left(\!{\widehat{p}}_{2i}\!\ge \! c \!\right)\!-\! \widehat{\mathcal{S }}_{1,h}^{(-k)}\left(\!c;\,{\widehat{p}}_{1i}\!\right)\!\right\} ^{2}\! d \widehat{\mathcal{S }}_{0,h_{0}}\left(\!c;\,{\widehat{p}}_{1i}\right)\!. \end{aligned}$$

    We let \(h_{1}^{\mathrm{{\tiny opt}}} = \arg \min _{h}{\sum _{k=1}^{K}Err_{k}^{(\mathcal{{S}}_1)}(h)}.\)

Since the order of \(h_y^{\mathrm{{\tiny opt}}}\) is expected to be \(n^{-1/5}\) (Fan and Gijbels 1995), the bandwidth we use for estimation is \(h_y = h_y^{\mathrm{{\tiny opt}}} \times n^{-d_0}\) with \(0 < d_0 < 3/10\) such that \( h_y = n^{-\nu }\) with \(1/5 < \nu < 1/2.\) This ensures that the resulting functional estimator \(\mathcal{{S}}_{y,h_y}(c;\,s)\) with the data-dependent smooth parameter has the above desirable large sample properties.

1.2.2 B.2 Bandwidth selection for \(\mathrm{{IDI}}(s)\)

Same as bandwidth selection for pAUC, we also propose a K-fold cross validation procedure to choose the optimal bandwidth \(h_{1}\) for \(\text{ IS}(s)=\int _{0}^{1}\mathcal{{S}}_{1}(c;\,s)dc\) and \(h_{0}\) for \(\text{ IP}(s)=\int _{0}^{1}\mathcal{{S}}_{0}(c;\,s)dc\) separately. The procedure is described as follows: we randomly split the data into K disjoint subsets of about equal sizes denoted by \(\{ { \mathcal J }_{k},\;k=1,\ldots ,K\}.\) Motivated by the fact (3), for each \(k\), we use all the observations not in \(\mathcal J _{k}\) to estimate \(\int _{0}^{1}\mathcal{{S}}_{y}(c,\,s)dc\) by obtaining \(\{\widehat{\varphi }^{(y)}_{0}(s;\,h),\,\widehat{\varphi }_{1}^{(y)}(s;\,h)\}\) for \(y=0,\,1,\) which is the solution to the estimating equation

$$\begin{aligned} \sum _{j\in \mathcal J _{l},l\ne k} I\left(Y_{j}=y\right){\widehat{\omega }}_{j}K_{h}\left\{ \widehat{\mathcal{E }}_{1j}(s)\right\} \left[{\widehat{p}}_{2j} - g\left\{ \varphi _{0}^{(y)}+\varphi _{1}^{(y)}\widehat{\mathcal{E }}_{1j}(s)\right\} \right]=0, \end{aligned}$$

wrt \((\varphi _{0}^{(y)},\,\varphi _{1}^{(y)}).\) Let \(\widehat{\text{ IS}}^{(-k)}(s;\,h)=g\{\widehat{\varphi }_{0}^{(1)}(s;\,h)\}\) and \(\widehat{\text{ IP}}^{(-k)}(s;\,h)=g\{\widehat{\varphi }_{0}^{(0)}(s;\,h)\}.\) With observations in \(\mathcal J _{k},\) we obtain

$$\begin{aligned} Err_{k}^{(\text{ IS})}(h) = \sum _{i \in \mathcal J _{k}}Y_i {\widehat{\omega }}_{i}\left\{ {\widehat{p}}_{2i} - \widehat{\text{ IS}}^{(-k)}\left({\widehat{p}}_{1i};\,h\right)\right\} ^{2}, \end{aligned}$$

or

$$\begin{aligned} Err_{k}^{(\text{ IP})}(h) = \sum _{i \in \mathcal J _{k}}\left(1-Y_i\right) {\widehat{\omega }}_{i}\left\{ {\widehat{p}}_{2i} - \widehat{\text{ IP}}^{(-k)}\left({\widehat{p}}_{1i};\,h\right)\right\} ^{2}. \end{aligned}$$

Then, we let \(h_{1}^{opt}\!=\!\arg \min _{h} \sum _{k=1}^{K}Err_{k}^{(\text{ IS})}(h)\) and \(h_{0}^{opt}\!=\!\arg \min _{h} \sum _{k=1}^{K}Err_{k}^{(\text{ IP})}(h).\)

1.3 Appendix C

R codes for application will be available from the corresponding author upon request.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Zhou, Q.M., Zheng, Y. & Cai, T. Subgroup specific incremental value of new markers for risk prediction. Lifetime Data Anal 19, 142–169 (2013). https://doi.org/10.1007/s10985-012-9235-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10985-012-9235-3

Keywords

Navigation