Skip to main content

Semiparametric Inference on the Absolute Risk Reduction and the Restricted Mean Survival Difference

  • Conference paper
  • First Online:
Risk Assessment and Evaluation of Predictions

Part of the book series: Lecture Notes in Statistics ((LNSP,volume 215))

  • 1906 Accesses

Abstract

For time-to-event data, when the hazards may be non-proportional, in addition to the hazard ratio, the absolute risk reduction and the restricted mean survival difference can be used to describe the time-dependent treatment effect. The absolute risk reduction measures the direct impact of the treatment on event rate or survival, and the restricted mean survival difference provides a way to evaluate the cumulative treatment effect. However, in the literature, available methods are limited for flexibly estimating these measures and making inference on them. In this article, point estimates, pointwise confidence intervals and simultaneous confidence bands of the absolute risk reduction and the restricted mean survival difference are established under a semiparametric model that can be used in a sufficiently wide range of applications. These methods are motivated by and illustrated for data from the Women’s Health Initiative estrogen plus progestin clinical trial.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 139.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 179.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. Bie, O., Borgan, O., Liestøl, K.: Confidence intervals and confidence bands for the cumulative hazard rate function and their small-sample properties. Scand. J. Stat. 14, 221–233 (1987)

    MATH  Google Scholar 

  2. Chen, P., Tsiatis, A.A.: Causal inference on the difference of the restricted mean lifetime between two groups. Biometrics 57, 1030–1038 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  3. Cheng, S.C., Wei, L.J., Ying, Z.: Predicting survival probabilities with semiparametric transformation models. J. Am. Stat. Assoc. 92, 227–235 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  4. Cox, D.R.: Regression models and life-tables (with Discussion). J. R. Stat. Soc. B 34, 187–220 (1972)

    MATH  Google Scholar 

  5. Dabrowska, D.M., Doksum, K.A., Song, J.: Graphical comparison of cumulative hazards for two populations. Biometrika 76, 763–773 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  6. Kalbfleisch, J.D., Prentice, R.L.: The Statistical Analysis of Failure Time Data, 2nd edn. Wiley, New York (2002)

    Book  MATH  Google Scholar 

  7. Kaplan, E., Meier, P.: Nonparametric estimation from incomplete observations. J. Am. Stat. Assoc. 53, 457–481 (1958)

    Article  MathSciNet  MATH  Google Scholar 

  8. Lin, D.Y., Wei, L.J., Ying, Z.: Checking the Cox model with cumulative sums of martingale-based residuals. Biometrika 80, 557–572 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  9. Lin, D.Y., Fleming, T.R., Wei, L.J.: Confidence bands for survival curves under the proportional hazards model. Biometrika 81, 73–81 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  10. Manson, J.E., Hsia, J., Johnson, K.C., Rossouw, J.E., Assaf, A.R., Lasser, N.L., Trevisan, M., Black, H.R., Heckbert, S.R., Detrano, R., Strickland, O.L., Wong, N.D., Crouse, J.R., Stein, E., Cushman, M., Women’S Health Initiative Investigators: Estrogen plus progestin and the risk of coronary heart disease. N. Engl. J. Med. 349, 523–534 (2003)

    Article  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  12. Nair, V.N.: Confidence bands for survival functions with censored data: a comparative study. Technometrics 26, 265–275 (1984)

    Article  Google Scholar 

  13. Parzen, M.I., Wei, L.J., Ying, Z.: Simultaneous confidence intervals for the difference of two survival functions. Scand. J. Stat. 24, 309–314 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  14. Peng, L., Huang, Y.: Survival analysis with temporal covariate effects. Biometrika 94, 719–733 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  15. Pollard, D.: Empirical Processes: Theory and Applications. Institute of Mathematical Statistics, Hayward (1990)

    MATH  Google Scholar 

  16. Prentice, R.L., Langer, R., Stefanick, M.L., Howard, B.V., Pettinger, M., Anderson, G., Barad, D., Curb, J.D., Kotchen, J., Kuller, L., Limacher, M., Wactawski-Wende, J., Women’S Health Initiative Investigators: combined postmenopausal hormone therapy and cardiovascular disease: toward resolving the discrepancy between observational studies and the women’s health initiative clinical trial. Am. J. Epidemiol. 162, 404–414 (2005)

    Article  Google Scholar 

  17. Royston, P., Parmar, M.K.: The use of restricted mean survival time to estimate the treatment effect in randomized clinical trials when the proportional hazards assumption is in doubt. Stat. Med. 19, 2409–2421 (2011)

    Article  MathSciNet  Google Scholar 

  18. Schaubel, D.E., Wei, G.: Double inverse-weighted estimation of cumulative treatment effects under nonproportional hazards and dependent censoring. Biometrics 67, 29–38 (2011)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  20. Tong, X., Zhu, C., Sun, J.: Semiparametric regression analysis of two-sample current status data, with applications to tumorigenicity experiments. Can. J. Stat. 35, 575–584 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  21. Writing Group for the Women’s Health Initiative Investigators: Risks and benefits of estrogen plus progestin in healthy postmenopausal women: principal results from the women’s health initiative randomized controlled trial. J. Am. Med. Assoc. 288, 321–333 (2002)

    Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  23. Yang, S., Prentice, R.L.: Estimation of the 2-sample hazard ratio function using a semiparametric model. Biostatistics 12, 354–368 (2011)

    Article  Google Scholar 

  24. Zucker, D.M.: Restricted mean life with covariates: modification and extension of a useful survival analysis method. J. Am. Stat. Assoc. 93, 702–709 (1998)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The original version of this article has previously been published in Lifetime Data Analysis in 2013.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Song Yang .

Editor information

Editors and Affiliations

Appendices

Appendix 1: Consistency

Throughout the Appendices, we assume the following regularity conditions, which is a little weaker than the conditions used in Yang and Prentice [22].

  • Condition 1. \(\lim \frac{n_{1}} {n} =\rho \in (0,1).\)

  • Condition 2. The survivor function G i of C i given Z i is continuous and satisfies

    $$\displaystyle{ \frac{1} {n}\sum _{i\leq n_{1}}G_{i}(t) \rightarrow \varGamma _{1},\ \frac{1} {n}\sum _{i>n_{1}}G_{i}(t) \rightarrow \varGamma _{2},}$$

    uniformly for tτ, for some Γ 1, Γ 2, and τ < τ 0 such that Γ j (τ) > 0, j = 1, 2.

  • Condition 3. The survivor functions S C and S T are absolutely continuous and S C (τ) > 0.

Under these conditions, the strong law of large numbers implies that (3) is satisfied.

For tτ, define

$$\displaystyle\begin{array}{rcl} L(t)& =& \varGamma _{1}S_{C} +\varGamma _{2}S_{T}, {}\\ U_{j}(t;\mathbf{b})& =& \int _{0}^{t}\varGamma _{ 1}dF_{C} +\exp (-b_{j})\int _{0}^{t}\varGamma _{ 2}dF_{T},\ j = 1,2, {}\\ \varLambda _{j}(t;\mathbf{b})& =& \int _{0}^{t}\frac{dU_{j}(s;\mathbf{b})} {L(s)},\ \ j = 1,2, {}\\ P(t;\mathbf{b})& =& \exp \{-\varLambda _{2}(t;\mathbf{b})\},\ R(t;\mathbf{b}) = \frac{1} {P(t;\mathbf{b})}\int _{0}^{t}P(s;\mathbf{b})d\varLambda _{ 1}(s;\mathbf{b}), {}\\ f_{j}^{0}(t;\mathbf{b})& =& \frac{\exp (-b_{j}){R}^{j-1}(t;\mathbf{b})} {\exp (-b_{1}) +\exp (-b_{2})R(t;\mathbf{b})},\ j = 1,2, {}\\ m_{j}(\mathbf{b})& =& \{\int _{0}^{\tau }f_{ j}^{0}\varGamma _{ 2}(t)dF_{T}(t) -\int _{0}^{\tau } \frac{f_{j}^{0}\varGamma _{ 2}(t)S_{T}(t)dR(t;\mathbf{b})} {\exp (-b_{1}) +\exp (-b_{2})R(t;\mathbf{b})}\},\ j = 1,2, {}\\ \end{array}$$

and m(b) = (m 1(b), m 2(b))′. We will also assume

  • Condition 4. The function m(b) is non-zero for \(b \in \mathcal{B}-\{\boldsymbol{\beta }\}\), where \(\mathcal{B}\) is a compact neighborhood of \(\boldsymbol{\beta }\).

Theorem 1.

Suppose that Conditions 1–4 hold. Then, (i) the zero \(\hat{\boldsymbol{\beta }}\) of Q( b ) in \(\mathcal{B}\) is strongly consistent for \(\boldsymbol{\beta }\) ; (ii) \(\hat{\varPhi }(t)\) is strongly consistent for Φ(t), uniformly for t ∈ [0,τ], and \(\hat{\varPsi }(t)\) is strongly consistent for Ψ(t), uniformly on t ∈ [0,τ]; (iii) \(\hat{\varOmega }\) converges almost surely to a limiting matrix Ω .

Proof.

Under Conditions 1–3, the limit of \(\sum _{i=1}^{n}I(X_{i} \geq t)/n\) is bounded away from zero on t ∈ [0, τ]. Thus, with probability 1,

$$\displaystyle{ \frac{\sum _{i=1}^{n}\delta _{i}{e}^{-b_{j}Z_{i}}I(X_{i} = t)} {\sum _{i=1}^{n}\delta _{i}I(X_{i} \geq t)} \rightarrow 0,\ j = 1,2, }$$
(14)

uniformly for t ∈ [0, τ] and \(b \in \mathcal{B}.\) From this, one also has, with probability 1,

$$\displaystyle{ \vert \varDelta \hat{P}(t;\mathbf{b})\vert \rightarrow 0,\ \vert \varDelta \hat{R}(t;\mathbf{b})\vert \rightarrow 0, }$$
(15)

uniformly for t ∈ [0, τ] and \(b \in \mathcal{B},\) where Δ indicates the jump of the function in t.

Define the martingale residuals

$$\displaystyle{\hat{M}_{i}(t;\mathbf{b}) =\delta _{i}I(X_{i} \leq t) -\int _{0}^{t}I(X_{ i} \geq s) \frac{\hat{R}(ds;\mathbf{b})} {{e}^{-b_{1}Z_{i}} + {e}^{-b_{2}Z_{i}}\hat{R}(s;\mathbf{b})},\ 1 \leq i \leq n.}$$

From (12) and (13), and the fundamental theorem of calculus, it follows that, with probability 1,

$$\displaystyle{ Q(\mathbf{b}) =\sum _{ i=1}^{n}\int _{ 0}^{\tau }\{f_{ i}(t;\mathbf{b}) + o(1)\}\hat{M}_{i}(dt;\mathbf{b}), }$$
(16)

uniformly in tτ, \(b \in \mathcal{B}\) and in, where \(f_{i} = {(f_{1i},f_{2i})}^{T}\), with

$$\displaystyle{f_{1i}(t;\mathbf{b}) = \frac{Z_{i}{e}^{-b_{1}Z_{i}}} {{e}^{-b_{1}Z_{i}} + {e}^{-b_{2}Z_{i}}\hat{R}(t;\mathbf{b})},\ f_{2i}(t;\mathbf{b}) = \frac{Z_{i}{e}^{-b_{2}Z_{i}}\hat{R}(t;\mathbf{b})} {{e}^{-b_{1}Z_{i}} + {e}^{-b_{2}Z_{i}}\hat{R}(t;\mathbf{b})}.}$$

From the strong law of large numbers ([15], p. 41) and repeated use of Lemma A1 of Yang and Prentice [22], one obtain, with probability 1,

$$\displaystyle{ \hat{P}(t;\mathbf{b}) \rightarrow \hat{ P}(t;\mathbf{b}),\ \hat{R}(t;\mathbf{b}) \rightarrow R(t;\mathbf{b}),\ Q(\mathbf{b})/n \rightarrow m(\mathbf{b}), }$$
(17)

uniformly in tτ and \(\mathbf{b} \in \mathcal{B}\). From these results and Condition 4, one obtains the strong consistency of \(\hat{\boldsymbol{\beta }}\), \(\hat{\varPhi }(t)\) and \(\hat{\varPsi }(t)\), and almost sure convergence of \(\hat{\varOmega }\).

Appendix 2: Weak Convergence

For \(\ C(t),\ D(t),\ \mu _{1}(t),\mu _{2}(t),\nu _{1}(t),\ \nu _{2}(t)\), let C (t), D (t), etc. be their almost sure limit. In addition, let L j be the almost sure limit of \(K_{j}/n,\ j = 1,2.\) For 0 ≤ s, t < τ, let

$$\displaystyle\begin{array}{rcl} & & \sigma _{\varPhi }(s,t) \\ & =& {D}^{{\ast}T}{(s)\varOmega }^{{\ast}}{(\int _{ 0}^{\tau } \frac{\mu _{1}^{{\ast}}\mu _{ 1}^{{\ast}T}} {1 + R}L_{1}dR +\int _{ 0}^{\tau } \frac{\mu _{2}^{{\ast}}\mu _{ 2}^{{\ast}T}} {{e}^{-\beta _{1}} + {e}^{-\beta _{2}}R}L_{2}dR)\varOmega }^{{\ast}T}{D}^{{\ast}}(t) \\ & & +{C}^{{\ast}}(s){C}^{{\ast}}(t)(\int _{ 0}^{s} \frac{\nu _{1}^{{\ast}2}} {1 + R}L_{1}dR +\int _{ 0}^{s} \frac{\nu _{2}^{{\ast}2}} {{e}^{-\beta _{1}} + {e}^{-\beta _{2}}R}L_{2}dR) \\ & & +{C}^{{\ast}}(t){D}^{{\ast}T}{(s)\varOmega }^{{\ast}}(\int _{ 0}^{t} \frac{\mu _{1}^{{\ast}}\nu _{ 1}^{{\ast}}} {1 + R}L_{1}dR +\int _{ 0}^{t} \frac{\mu _{2}^{{\ast}}\nu _{ 2}^{{\ast}}} {{e}^{-\beta _{1}} + {e}^{-\beta _{2}}R}L_{2}dR) \\ & & +{C}^{{\ast}}(s){D}^{{\ast}T}{(t)\varOmega }^{{\ast}}(\int _{ 0}^{s} \frac{\mu _{1}^{{\ast}}\nu _{ 1}^{{\ast}}} {1 + R}L_{1}dR +\int _{ 0}^{s} \frac{\mu _{2}^{{\ast}}\nu _{ 2}^{{\ast}}} {{e}^{-\beta _{1}} + {e}^{-\beta _{2}}R}L_{2}dR),{}\end{array}$$
(18)

and

$$\displaystyle\begin{array}{rcl} & & \sigma _{\varPsi }(s,t) \\ & =& \int _{0}^{s}{D}^{{\ast}T}(x)d{x\varOmega }^{{\ast}}(\int _{ 0}^{\tau }\frac{\mu _{1}^{{\ast}}(w)\mu _{ 1}^{{\ast}T}(w)} {1 + R(w)} L_{1}(w)dR(w) \\ & & +\int _{0}^{\tau } \frac{\mu _{2}^{{\ast}}(w)\mu _{ 2}^{{\ast}T}(w)} {{e}^{-\beta _{1}} + {e}^{-\beta _{2}}R(w)}L_{2}(w)dR(w){)\varOmega }^{{\ast}T}\int _{ 0}^{t}{D}^{{\ast}T}(x)dx \\ & & +\int _{0}^{s} \frac{\nu _{1}^{{\ast}2}(w)} {1 + R(w)}{(\int _{w}^{s}{C}^{{\ast}}(x)dx)}^{2}L_{ 1}(w)dR(w) \\ & & +\int _{0}^{s} \frac{\nu _{2}^{{\ast}2}(w)} {{e}^{-\beta _{1}} + {e}^{-\beta _{2}}R(w)}{(\int _{w}^{s}{C}^{{\ast}}(x)dx)}^{2}L_{ 2}(w)dR(w) \\ & & +\int _{0}^{s}{D}^{{\ast}T}(x)d{x\varOmega }^{{\ast}}\int _{ 0}^{t}\frac{\mu _{1}^{{\ast}}(w)\nu _{ 1}^{{\ast}}(w)} {1 + R(w)} (\int _{w}^{t}{C}^{{\ast}}(x)dx)L_{ 1}(w)dR(w) \\ & & +\int _{0}^{s}{D}^{{\ast}T}(x)d{x\varOmega }^{{\ast}}\int _{ 0}^{t} \frac{\mu _{2}^{{\ast}}(w)\nu _{ 2}^{{\ast}}(w)} {{e}^{-\beta _{1}} + {e}^{-\beta _{2}}R(w)}(\int _{w}^{t}{C}^{{\ast}}(x)dx)L_{ 2}dR(w) \\ & & +\int _{0}^{t}{D}^{{\ast}T}(x)d{x\varOmega }^{{\ast}}\int _{ 0}^{s}\frac{\mu _{1}^{{\ast}}(w)\nu _{ 1}^{{\ast}}(w)} {1 + R(w)} (\int _{w}^{s}{C}^{{\ast}}(x)dx)L_{ 1}(w)dR(w) \\ & & +\int _{0}^{t}{D}^{{\ast}T}(x)d{x\varOmega }^{{\ast}}\int _{ 0}^{s} \frac{\mu _{2}^{{\ast}}(w)\nu _{ 2}^{{\ast}}(w)} {{e}^{-\beta _{1}} + {e}^{-\beta _{2}}R(w)}(\int _{w}^{s}{C}^{{\ast}}(x)dx)L_{ 2}(w)dR(w).{}\end{array}$$
(19)

Theorem 2.

Suppose that Conditions 1–4 hold and that the matrix Ω is non-singular. Then, (i) U n is asymptotically equivalent to the process \(\tilde{U_{n}}\) in (8) which converges weakly to a zero-mean Gaussian process U on [0,τ], with covariance function σ Φ (s,t) in (18). In addition, \(\hat{U}_{n}(s)\) given the data converges weakly to the same limiting process U . (ii) V n (t) is asymptotically equivalent to the process \(\tilde{V }_{n}\) in (11) which converges weakly to the zero-mean Gaussian process \(\int _{0}^{t}{U}^{{\ast}}(s)ds\) on t ∈ [0,τ], with covariance function σ Ψ (s,t) in (19). The process \(\int _{0}^{t}\hat{V }_{n}(s)ds\) given the data also converges weakly to the same limiting process \(\int _{0}^{t}{U}^{{\ast}}(s)ds\) .

Proof.

(i) As in the proof for Theorem A2 (ii) in Yang and Prentice [22], from the strong embedding theorem and (16), \(Q(\boldsymbol{\beta })/\sqrt{n}\) can be shown to be asymptotically normal. Now Taylor series expansion of Q(b) around \(\boldsymbol{\beta }\) and the non-singularity of Ω imply that \(\sqrt{n}(\hat{\boldsymbol{\beta }}-\boldsymbol{\beta })\) is asymptotically normal. From the \(\sqrt{n}\)- boundedness of \(\hat{\boldsymbol{\beta }}\),

$$\displaystyle\begin{array}{rcl} \sqrt{ n}(\hat{R}(t;\hat{\boldsymbol{\beta }}) -\hat{ R}(t;\boldsymbol{\beta }))& =& \frac{\partial R(t;\beta )} {\partial \beta } \sqrt{n}(\hat{\boldsymbol{\beta }}-\boldsymbol{\beta }) + o_{p}(1), {}\\ \end{array}$$

uniformly in tτ. These results, some algebra and Taylor series expansion together show that U n is asymptotically equivalent to \(\tilde{U}_{n}\). Similarly to the proof of the asymptotic normality of \(Q(\boldsymbol{\beta })/\sqrt{n}\), one can show that \(\tilde{U}_{n}\) converges weakly to a zero-mean Gaussian process. Denote the limiting process by U . From the martingale integral representation of \(\tilde{U}_{n}\), it follows that the covariation process of U is given by σ(s, t) in (18), which can be consistently estimated by \(\hat{\sigma }(s,t)\) in (9). By checking the tightness condition and the convergence of the finite-dimensional distributions, it can be shown that \(\hat{U}_{n}(s)\) given the data also converges weakly to U .

(ii) From the results in (i), the assertions on V n and \(\tilde{V }_{n}\) follow.

Rights and permissions

Reprints and permissions

Copyright information

© 2013 Springer Science+Business Media New York

About this paper

Cite this paper

Yang, S. (2013). Semiparametric Inference on the Absolute Risk Reduction and the Restricted Mean Survival Difference. In: Lee, ML., Gail, M., Pfeiffer, R., Satten, G., Cai, T., Gandy, A. (eds) Risk Assessment and Evaluation of Predictions. Lecture Notes in Statistics, vol 215. Springer, New York, NY. https://doi.org/10.1007/978-1-4614-8981-8_2

Download citation

Publish with us

Policies and ethics