Skip to main content

Non-parametric Maximum Likelihood Estimation for Case-Cohort and Nested Case-Control Designs with Competing Risks Data

  • Chapter
  • First Online:
Statistical Modeling in Biomedical Research

Part of the book series: Emerging Topics in Statistics and Biostatistics ((ETSB))

  • 1101 Accesses

Abstract

Assuming cause-specific hazards given by Cox’s regression model, we provide non-parametric maximum likelihood estimator (NPMLEs) in the nested case-control or case-cohort design with competing risks data. We propose an iterative algorithm based on self-consistency equations derived from score functions to compute NPMLE and compute the predicted cumulative incidence function with their corresponding confidence intervals and bands. Consistency and asymptotic normality are established, together with a consistent estimator of the asymptotic variance based on the observed profile likelihood. Simulation studies show that the numerical performance of NPMLE approach is satisfactory and compares well with that of weighted partial likelihood. Our method is applied to the Taiwan National Health Insurance Research Database (NHIRD) to analyze the occurrences of liver and lung cancers in type 2 diabetic mellitus patients.

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 74.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 99.00
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
Price excludes VAT (USA)
  • Durable hardcover 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. Borgan, ø., & Keogh, R. H. (2015). Nested case-control: Should one break the matching? Lifetime Data Analysis, 21, 517–541.

    Google Scholar 

  2. Chang, I. S., Hsiung, C. A., Wang, M. C., & Wen, C. C. (2005). An asymptotic theory for the nonparametric maximum likelihood estimator in the Cox-gene model. Bernoulli, 11, 863–892.

    Article  MathSciNet  Google Scholar 

  3. Chang, I. S., Hsiung, C. A., Wen, C. C., Wu, Y. J., & Yang, C. C. (2007). Non-parametric maximum-likelihood estimation in a semiparametric mixture model for competing-risks data. Scandinavian Journal of Statistics, 34, 870–895.

    MathSciNet  MATH  Google Scholar 

  4. Chen, H. Y. (2002). Double-semiparametric method for missing covariates in Cox regression models. Journal of the American Statistical Association, 97, 565–576.

    Article  MathSciNet  Google Scholar 

  5. Chen, K. N. (2001). Generalized case-cohort sampling. Journal of the Royal Statistical Society: Series B, 63, 791–809.

    Article  MathSciNet  Google Scholar 

  6. Fine, J. P., & Gray, R. J. (1999). A proportional hazards model for the subdistribution of a competing risk. Journal of the American Statistical Association, 94, 496–509.

    Article  MathSciNet  Google Scholar 

  7. Keogh, R. H., & White, I. R. (2013). Using full-cohort data in nested case–control and case–cohort studies by multiple imputation. Statistics in Medicine, 32, 4021–4043.

    Article  MathSciNet  Google Scholar 

  8. Kim, R. S. (2013). Lesser known facts about nested case-control designs. Journal of Translational Medicine and Epidemiology, 1, 1007.

    Google Scholar 

  9. Kulathinal, S., & Arjas, E. (2006). Bayesian inference from case-cohort data with multiple end-points. Scandinavian Journal of Statistics, 33, 25–36.

    Article  MathSciNet  Google Scholar 

  10. Lu, W., & Peng, L. (2008). Semiparametric analysis of mixture regression models with competing risks data. Lifetime Data Analysis, 14, 231–252.

    Article  MathSciNet  Google Scholar 

  11. Murphy, S. A., & van der Vaart, A. W. (1999). Observed information in semi-parametric models. Bernoulli, 5, 381–412.

    Article  MathSciNet  Google Scholar 

  12. Murphy, S. A., & van der Vaart, A. W. (2000). On profile likelihood. Journal of the American Statistical Association, 95, 449–465.

    Article  MathSciNet  Google Scholar 

  13. Prentice, R. L. (1986). A case-cohort design for epidemiologic cohort studies and disease prevention trials. Biometrika, 73, 1–11.

    Article  MathSciNet  Google Scholar 

  14. Saarela, O., Kulathinal, S., Arjas, E., & Läärä, E. (2008). Nested case-control data utilized for multiple outcomes: A likelihood approach and alternatives. Statistics in Medicine, 27, 5991–6008.

    Article  MathSciNet  Google Scholar 

  15. Samuelsen, S. O. (1997). A pseudolikelihood approach to analysis of nested case-control studies. Biometrika, 84, 379–394.

    Article  MathSciNet  Google Scholar 

  16. Scheike, T. H., & Juul, A. (2004). Maximum likelihood estimation for Cox’s regression model under nested case-control sampling. Biostatistics, 5, 193–206.

    Article  Google Scholar 

  17. Scheike, T. H., & Martinussen, T. (2004). Maximum likelihood estimation for Cox’s regression model under case-cohorts sampling. Scandinavian Journal of Statistics, 31, 283–293.

    Article  MathSciNet  Google Scholar 

  18. Sørensen, P., & Andersen, P. K. (2000). Competing risks analysis of the case-cohort design. Biometrika, 87, 49–59.

    Article  MathSciNet  Google Scholar 

  19. Støer, N. C., & Samuelsen, S. O. (2012). Comparison of estimators in nested case-control studies with multiple outcomes. Lifetime Data Analysis, 18, 261–283.

    Article  MathSciNet  Google Scholar 

  20. Thomas, D. C. (1977). Addendum to “methods of cohort analysis: appraisal by application to asbestos mining,” by F. D. K. Liddell, J. C. McDonald and D. C. Thomas. Journal of the Royal Statistical Society. Series A, 140, 483–485.

    Google Scholar 

  21. van der Vaart, A. W. (1998). Asymptotic statistics. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  22. van der Vaart, A. W., & Wellner, J. A. (1996). Weak convergence and empirical processes with application to statistics. New York: Springer.

    Book  Google Scholar 

  23. Vigneri, P., Frasca, L., Sciacca, L., Pandini, G., & Vigneri, R. (2009). Diabetes and cancer. Endocrine-Related Cancer, 16, 1103–1123.

    Article  Google Scholar 

Download references

Acknowledgements

We are very grateful to the Editors and referees for their very valuable comments that helped to improve the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yi-Hau Chen .

Editor information

Editors and Affiliations

Appendices

Appendix 1: Proof of Lemma 1

\(P_{M} l_{2,\left (\boldsymbol {\beta }_{1}^{\left (q\right )},\cdots ,\boldsymbol {\beta }_{K}^{\left (q\right )},\boldsymbol {p},\varLambda _{1}^{\left (q\right )},\cdots ,\varLambda _{K}^{\left (q\right )} \right )} \left [\boldsymbol {h}_{2} \right ]=0\) means \(\sum _{i=1}^{M}\left \{I_{\left (o_{i}=1\right )} \left (\frac {\boldsymbol {\mu }_{i}^{T} \boldsymbol {h}_{2} }{\boldsymbol {\mu }_{i}^{T} \boldsymbol {p}} \right )+I_{\left (o_{i}=0\right )}\right .\) \( \left .\left (\frac {\boldsymbol {\phi }_{i}^{T} \boldsymbol {h}_{2} }{\boldsymbol {\phi }_{i}^{T} \boldsymbol {p}} \right )\right \}=0\). Here \(\boldsymbol {\mu }_{i}=\left (I_{\boldsymbol {Z}_{i}=\boldsymbol {W}_{1} },I_{\boldsymbol {Z}_{i}=\boldsymbol {W}_{2}},\cdots ,I_{\boldsymbol {Z}_{i}=\boldsymbol {W}_{J}} \right )^{T}\)and

$$\displaystyle \begin{aligned} \begin{array}{rcl}&\displaystyle &\displaystyle \boldsymbol{\phi}_{i}=\left(\exp \left(-\sum_{k=1}^{K}\varLambda_{k}^{\left(q\right)} \left(X_{i} \right)\exp \left(\boldsymbol{W}_{1}^{T} \boldsymbol{\beta}_{k}^{\left(q\right)} \right) \right),\right.\\ &\displaystyle &\displaystyle \qquad \qquad \left.\cdots,\exp \left(-\sum_{k=1}^{K}\varLambda_{k}^{\left(q\right)} \left(X_{i} \right)\exp \left(\boldsymbol{W}_{J}^{T} \boldsymbol{\beta}_{k}^{\left(q\right)} \right) \right)\right)^{T}.\end{array} \end{aligned} $$

Setting \(\boldsymbol {h}_{2}=\left (0,\cdots ,0,1,0,\cdots ,-1\right )^{T}\), having the lth and Jth coordinate equal to 1 and −1 respecting and all the other coordinates being 0, we know

$$\displaystyle \begin{aligned}\sum_{i=1}^{M}\left\{I_{\left(o_{i}=1\right)} \left(\frac{\mu_{i,l}}{\boldsymbol{\mu}_{i}^{T} \boldsymbol{p}} \right){+}I_{\left(o_{i}=0\right)} \left(\frac{\phi_{i,l}}{\boldsymbol{\phi}_{i}^{T} p} \right)\right\}{=}\sum_{i=1}^{M}\left\{I_{\left(o_{i}=1\right)} \left(\frac{\mu_{i,J}}{\boldsymbol{\mu}_{i}^{T} \boldsymbol{p}} \right){+}I_{\left(o_{i}=0\right)} \left(\frac{\phi_{i,J}}{\boldsymbol{\phi}_{i}^{T} \boldsymbol{p}} \right)\right\}. \end{aligned}$$

Then

$$\displaystyle \begin{aligned}\sum _{i=1}^{M}\left[I_{\left(o_{i}=1\right)} I_{\boldsymbol{Z}_{i}=\boldsymbol{W}_{l} }+I_{\left(o_{i}=0\right)} \alpha _{il} \right] =p_{l} \sum_{i=1}^{M}\left\{I_{\left(o_{i} =1\right)} \left(\frac{I_{\boldsymbol{Z}_{i} =\boldsymbol{W}_{J} } }{p_{J} } \right)+I_{\left(o_{i} =0\right)} \left(\frac{\alpha _{iJ} }{p_{J} } \right)\right\}.\end{aligned}$$

Straight-forward simplification shows that

$$\displaystyle \begin{aligned}M{=}\sum_{l=1}^{J}\sum_{i=1}^{M}\left[I_{\left(o_{i}=1\right)} I_{\boldsymbol{Z}_{i}=\boldsymbol{W}_{l} }+I_{\left(o_{i}=0\right)} \alpha_{il} \right]{=}\sum_{i=1}^{M}\left\{I_{\left(o_{i} =1\right)} \left(\frac{I_{\boldsymbol{Z}_{i}=\boldsymbol{W}_{J}}}{p_{J}} \right){+}I_{\left(o_{i}=0\right)} \left(\frac{\alpha_{iJ}}{p_{J}} \right)\right\},\end{aligned}$$

then we get that

$$\displaystyle \begin{aligned}p_{l}=\frac{1}{M} \sum_{i=1}^{M}\left[I_{\left(o_{i}=1\right)} I_{\boldsymbol{Z}_{i}=\boldsymbol{W}_{l}}+I_{\left(o_{i}=0\right)} \alpha_{il} \right].\end{aligned}$$

This shows that

$$\displaystyle \begin{aligned}\boldsymbol{p}=\left(\frac{1}{M} \boldsymbol{1}^{T} \boldsymbol{\alpha} \left(\boldsymbol{p}\right)\right)^{T},\end{aligned}$$

and the proof is complete.

Making use of Lemma 1, we define

$$\displaystyle \begin{aligned}\boldsymbol{p}^{\left(q+1\right)}=\left(\frac{1}{M} \boldsymbol{1}^{T} \boldsymbol{\alpha} \left(\boldsymbol{p}^{\left(q\right)} \right)\right)^{T}.\end{aligned}$$

Appendix 2: Proof of Lemma 6

For any h ∈ H p. If we can show that ϕ θ,h is continuous at θ 0, then we have, for arbitrary ε > 0, there is some \(\delta =\delta \left (\varepsilon ,\boldsymbol {\theta }_{0}\right )>0\) such that \(\left \| \phi _{\boldsymbol {\theta },\boldsymbol {h}}-\phi _{\boldsymbol {\theta }_{0},\boldsymbol {h}} \right \| <\varepsilon \) whenever \(\left \| \boldsymbol {\theta }-\boldsymbol {\theta }_{0} \right \| <\delta \). Hence, whenever \(\left \| \boldsymbol {\theta }-\boldsymbol {\theta }_{0} \right \|<\delta \),

$$\displaystyle \begin{aligned}\left|E\left(\phi_{\boldsymbol{\theta},\boldsymbol{h}}-\phi_{\boldsymbol{\theta}_{0},\boldsymbol{h}} \right)^{2} \right|\le E\left\| \phi_{\boldsymbol{\theta},\boldsymbol{h}}-\phi_{\boldsymbol{\theta}_{0},\boldsymbol{h}} \right\|{}^{2} \le \varepsilon^{2},\end{aligned}$$

which means that \(E\left (\phi _{\boldsymbol {\theta },\boldsymbol {h}}-\phi _{\boldsymbol {\theta }_{0},\boldsymbol {h}} \right )^{2} \to 0\) as θ →θ 0 for any h ∈ H p. Thus,

$$\displaystyle \begin{aligned}\sup_{\boldsymbol{h}\in H_{p}} E\left(\phi_{\boldsymbol{\theta},\boldsymbol{h}}-\phi_{\boldsymbol{\theta}_{0},\boldsymbol{h}} \right)^{2} \to 0\end{aligned}$$

as θ →θ 0.

Now, we want to show that ϕ θ,h is continuous at θ 0, for any h ∈ H p. It is clearly that the following functions \(\exp \left (\boldsymbol {Z}_{1}^{T} \boldsymbol {\beta }_{k} \right )\), \(\exp \left (\boldsymbol {W}_{j}^{T} \boldsymbol {\beta }_{k}\right )\), \(\exp \left (-\sum _{k=1}^{2}\varLambda _{k} \left (X_{1} \right )\exp \left (\boldsymbol {W}_{j}^{T} \boldsymbol {\beta }_{k} \right ) \right )\) and \(pr\left (\boldsymbol {Z}_{1}=\boldsymbol {W}_{j} |T_{1} \ge X_{1} \right )\) are continuous and we know \(\sum _{j=1}^{J}\exp \left (-\sum _{k=1}^{2}\varLambda _{k} \left (X_{1} \right )\exp \left (\boldsymbol {W}_{j}^{T} \boldsymbol {\beta }_{k} \right ) \right )p_{j}\) is bounded and bounded away from zero. Finally, we may see that \(\int _{0}^{X_{1}}h\left (t\right )d\varLambda _{k} \left (t\right )\) is continuous at Λ k0, since

$$\displaystyle \begin{aligned}{\mathop{\sup}\limits_{h\in BV\left[0,\tau \right]}} \int_{0}^{\tau}\left|h\left(t\right)\right|d\left|\varLambda_{k} \left(t\right)-\varLambda_{k0} \left(t\right)\right| \le c\int_{0}^{\tau}d\left|\varLambda_{k} \left(t\right)-\varLambda_{k0} \left(t\right)\right|,\end{aligned}$$

we know \(\left \| \varLambda _{k}-\varLambda _{k0} \right \|{ }_{V} \to 0\), which means

$$\displaystyle \begin{aligned}\int_{0}^{\tau}d\left|\varLambda_{k} \left(t\right)-\varLambda_{k0} \left(t\right)\right|=\int_{0}^{\tau}\left|\left(\varLambda_{k} \left(t\right)-\varLambda_{k0} \left(t\right)\right)^{/} \right|dt \to 0.\end{aligned}$$

This completes the proof.

Appendix 3: Σ is Positive Definite and Symmetric

  1. (i)

    Σ is positive definite.

  2. (ii)

    Σ is symmetric.

Proof of (i)

Note that \(\varSigma ^{-1}{=}\left (\sigma _{12}^{-1} \left (\boldsymbol {e}_{1},0,0,0\right ),\sigma _{12}^{-1} \left (\boldsymbol {e}_{2},0,0,0\right ),\cdots ,\sigma _{12}^{-1} \right .\) \(\left .\left (\boldsymbol {e}_{D},0,0,0\right )\right )\). Here we want to show Σ −1 is positive definite. Let \(\boldsymbol {h}_{12}=\left (y_{1},y_{2},\cdots ,y_{D} \right )^{T}\) be given. Consider

$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle &\displaystyle \boldsymbol{h}_{12}^{T} \varSigma^{-1} \boldsymbol{h}_{12} \\ &\displaystyle =&\displaystyle y_{1} \sigma_{12}^{-1} \left(\boldsymbol{e}_{1},0,0,0\right)_{1} y_{1}+y_{2} \sigma_{12}^{-1} \left(\boldsymbol{e}_{1},0,0,0\right)_{2} y_{1}+\cdots+y_{D} \sigma_{12}^{-1} \left(\boldsymbol{e}_{1},0,0,0\right)_{D} y_{1} \\ &\displaystyle &\displaystyle + y_{1} \sigma_{12}^{-1} \left(\boldsymbol{e}_{2},0,0,0\right)_{1} y_{2}+y_{2} \sigma_{12}^{-1} \left(\boldsymbol{e}_{2},0,0,0\right)_{2} y_{2}+\cdots+y_{D} \sigma_{12}^{-1} \left(\boldsymbol{e}_{2},0,0,0\right)_{D} y_{2} \\ &\displaystyle &\displaystyle + \cdots \\ &\displaystyle &\displaystyle + y_{1} \sigma_{12}^{-1} \left(\boldsymbol{e}_{D},0,0,0\right)_{1} y_{D}+y_{2} \sigma_{12}^{-1} \left(\boldsymbol{e}_{D},0,0,0\right)_{2} y_{D}+\cdots+y_{D} \sigma_{12}^{-1} \left(\boldsymbol{e}_{D},0,0,0\right)_{D} y_{D} \\ &\displaystyle =&\displaystyle y_{1} \sigma_{12}^{-1} \left(\left(y_{1},y_{2},\cdots,y_{D} \right),0,0,0\right)_{1}+y_{2} \sigma_{12}^{-1} \left(\left(y_{1},y_{2},\cdots,y_{D} \right),0,0,0\right)_{2} \\ &\displaystyle &\displaystyle + \cdots \\ &\displaystyle &\displaystyle + y_{D} \sigma_{12}^{-1} \left(\left(y_{1},y_{2},\cdots,y_{D} \right),0,0,0\right)_{D} \\ &\displaystyle =&\displaystyle \boldsymbol{h}_{12}^{T} \sigma_{12}^{-1} \left(h_{12},0,0,0\right)>0, \end{array} \end{aligned} $$

for h 12 nonzero. This completes the proof.

Proof of (ii)

We want to show that Σ −1 is symmetric. Fix i. Choose h such that Let \(\boldsymbol {h}_{1}^{\prime }=\sigma _{1}^{-1} \left (\boldsymbol {e}_{i},0,0,0\right )\), \(\boldsymbol {h}_{2}^{\prime }=\sigma _{2}^{-1} \left (\boldsymbol {e}_{i},0,0,0\right )\), \(\boldsymbol {h}_{3}^{\prime }=\sigma _{3}^{-1} \left (\boldsymbol {e}_{i},0,0,0\right )\), \(\boldsymbol {h}_{4}^{\prime }=\sigma _{4}^{-1} \left (\boldsymbol {e}_{i},0,0,0\right )\) and \(\boldsymbol {h}_{5}^{\prime }=\sigma _{5}^{-1} \left (\boldsymbol {e}_{i},0,0,0\right )\). Choose h such that \(\boldsymbol {h}_{1}=\sigma _{1}^{-1} \left (\boldsymbol {e}_{k},0,0,0\right )\), \(\boldsymbol {h}_{2}=\sigma _{2}^{-1} \left (\boldsymbol {e}_{k},0,0,0\right )\), \(\boldsymbol {h}_{3}=\sigma _{3}^{-1} \left (\boldsymbol {e}_{k},0,0,0\right )\), \(\boldsymbol {h}_{4}=\sigma _{4}^{-1} \left (\boldsymbol {e}_{k},0,0,0\right )\) and \(\boldsymbol {h}_{5}=\sigma _{5}^{-1} \left (\boldsymbol {e}_{k},0,0,0\right )\). It remains to check the following case from (8), we can obtain

$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle &\displaystyle \sigma_{1} \left(\boldsymbol{h}\right)^{T} \boldsymbol{h}_{1}^{\prime}+\sigma_{2} \left(\boldsymbol{h}\right)^{T} \boldsymbol{h}_{2}^{\prime}+\sigma_{3} \left(\boldsymbol{h}\right)^{T} \boldsymbol{h}_{3}^{\prime} \\ &\displaystyle &\displaystyle +\int_{0}^{\tau }\sigma_{4} \left(\boldsymbol{h}\right)\left(t\right)\boldsymbol{h}_{4}^{\prime} \left(t\right)d\varLambda_{10} \left(t\right) +\int_{0}^{\tau}\sigma_{5} \left(\boldsymbol{h}\right)\left(t\right)\boldsymbol{h}_{5}^{\prime} \left(t\right)d\varLambda_{20} \left(t\right)\\ &\displaystyle &\displaystyle =\boldsymbol{e}_{k}^{T} \sigma_{12}^{-1} \left(\boldsymbol{e}_{i} ,0,0,0\right).\\ &\displaystyle &\displaystyle \sigma_{1} \left(\boldsymbol{h}^{\prime} \right)^{T} \boldsymbol{h}_{1} +\sigma _{2} \left(\boldsymbol{h}^{\prime} \right)^{T} \boldsymbol{h}_{2} +\sigma_{3} \left(\boldsymbol{h}^{\prime} \right)^{T} \boldsymbol{h}_{3} \\ &\displaystyle &\displaystyle +\int _{0}^{\tau }\sigma _{4} \left(\boldsymbol{h}^{\prime} \right)\left(t\right)\boldsymbol{h}_{4} \left(t\right)d\varLambda _{10} \left(t\right) +\int_{0}^{\tau }\sigma_{5} \left(\boldsymbol{h}^{\prime} \right)\left(t\right)\boldsymbol{h}_{5} \left(t\right)d\varLambda_{20} \left(t\right)\\ &\displaystyle &\displaystyle =\boldsymbol{e}_{i}^{T} \sigma_{12}^{-1} \left(\boldsymbol{e}_{k} ,0,0,0\right). \end{array} \end{aligned} $$

Hence \(\boldsymbol {e}_{k}^{T} \sigma _{12}^{-1} \left (\boldsymbol {e}_{i},0,0,0\right )=\boldsymbol {e}_{i}^{T} \sigma _{12}^{-1} \left (\boldsymbol {e}_{k},0,0,0\right )\). This completes the proof.

Appendix 4: \(\sqrt {M} \left (\widehat {\boldsymbol {\beta }}_{M}-\boldsymbol {\beta }_{0} \right )\) Has Asymptotic Variance Σ −1

$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle &\displaystyle var\left(\sqrt{M} \left(\widehat{\boldsymbol{\beta}}_{M}-\boldsymbol{\beta}_{0} \right)\right) \\ &\displaystyle &\displaystyle =var\left(\varSigma^{-1} \sqrt{M} \left(P_{M}-P_{0} \right)\tilde{l}_{0} \right) \\ &\displaystyle &\displaystyle =\varSigma^{-1} var\left(\left(\sum_{i=1}^{M} \tilde{l}_{0} \left(X_{i},E_{i},\boldsymbol{Z}_{i} \cdot O_{i},O_{i} \right) /\sqrt{M}\right)\right)\varSigma^{-1} \\ &\displaystyle &\displaystyle =\varSigma^{-1} var\left(\tilde{l}_{0} \right)\varSigma^{-1} \\ &\displaystyle &\displaystyle =\varSigma^{-1} E\left(\tilde{l}_{0} \tilde{l}_{0}^{T} \right)\varSigma^{-1}-\varSigma^{-1} \left\{E\left(\tilde{l}_{0} \right)E\left(\tilde{l}_{0} \right)^{T} \right\}\varSigma^{-1} \\ &\displaystyle &\displaystyle =\varSigma^{-1}. \end{array} \end{aligned} $$

Because \(E\left (\tilde {l}_{0} \tilde {l}_{0}^{T} \right )=\varSigma \) and

$$\displaystyle \begin{aligned} \begin{array}{rcl} &\displaystyle &\displaystyle \boldsymbol{e}_{i}^{T} \varSigma^{-1} \left\{E\left(\tilde{l}_{0} \right)E\left(\tilde{l}_{0} \right)^{T} \right\}\varSigma^{-1} \boldsymbol{e}_{k} \\ &\displaystyle &\displaystyle =\left\{\boldsymbol{e}_{i}^{T} \varSigma^{-1} E\left(\tilde{l}_{0} \right)\left(l_{3,\boldsymbol{\theta}_{0}} \left[\boldsymbol{h}_{3} \right]+l_{4,\boldsymbol{\theta}_{0}} \left[\boldsymbol{h}_{4} \right]+l_{5,\boldsymbol{\theta}_{0} } \left[\boldsymbol{h}_{5} \right]\right)\right\} \\ &\displaystyle &\displaystyle \times \left\{\left(l_{3,\boldsymbol{\theta}_{0} } \left[\boldsymbol{h}_{3} \right]+l_{4,\boldsymbol{\theta}_{0}} \left[\boldsymbol{h}_{4} \right]+l_{5,\boldsymbol{\theta}_{0}} \left[\boldsymbol{h}_{5} \right]\right)^{-1} E\left(\tilde{l}_{0} \right)^{T} \varSigma^{-1} \boldsymbol{e}_{k} \right\} \\ &\displaystyle &\displaystyle =0. \end{array} \end{aligned} $$

This completes the proof.

Appendix 5: Two Results Due to V&W [22]

We quote two important results used in this paper.

Theorem 3.3.1 of V&W [22]

Let Ψ n and Ψ be random maps and a fixed map, respectively, from Θ into a Banach space such that

$$\displaystyle \begin{aligned}\sqrt{n} \left(\varPsi_{n}-\varPsi \right)\left(\widehat{\theta}_{n} \right)-\sqrt{n} \left(\varPsi_{n}-\varPsi \right)\left(\theta_{0} \right)=o_{p^{*}} \left(1+\sqrt{n} \left\| \widehat{\theta}_{n}-\theta_{0} \right\| \right),\end{aligned}$$

and such that the sequence \(\sqrt {n} \left (\varPsi _{n}-\varPsi \right )\left (\theta _{0} \right )\) converges in distribution to a tight random element Z. Let \(\theta \to \varPsi \left (\theta \right )\) be Fréchet differentiable at θ 0 with a continuously invertible derivative \(\dot {\varPsi }_{0}\). If \(\varPsi \left (\theta _{0} \right )=0\), \(\widehat {\theta }_{n}\) satisfies \(\varPsi _{n} \left (\widehat {\theta }_{n} \right )=o_{p^{*} } \left (n^{-1/2} \right )\), and converges in outer probability to θ 0, then

$$\displaystyle \begin{aligned}\sqrt{n} \dot{\varPsi}_{\theta_{0} } \left(\widehat{\theta}_{n}-\theta_{0} \right)=-\sqrt{n} \left(\varPsi_{n}-\varPsi \right)\left(\theta_{0} \right)+o_{p^{*}} \left(1\right).\end{aligned}$$

Consequently, \(\sqrt {n} \left (\widehat {\theta }_{n}-\theta _{0} \right )\to -\dot {\varPsi }_{\theta _{0} }^{-1} Z\).

In the case of independent and identically distributed observations, the theorem may be applied with \(\varPsi _{n} \left (\theta \right )h=P_{n} \phi _{\theta ,h}\) and \(\varPsi \left (\theta \right )h=P\phi _{\theta ,h} \) for given measurable functions ϕ θ,h, indexed by Θ and arbitrary index set H. In this case, \(\sqrt {n} \left (\varPsi _{n}-\varPsi \right )\left (\theta \right )=\left \{G_{n} \phi _{\theta ,h}:h\in H\right \}\) is the empirical process indexed by the classes of functions \(\left \{\phi _{\theta ,h}:h\in H\right \}\).

Lemma 3.3.5 of V&W [22]

Suppose that the class of functions

$$\displaystyle \begin{aligned}\left\{\phi_{\theta,h}-\phi_{\theta_{0},h}:\left\| \theta-\theta_{0} \right\| <\delta,h\in H\right\}\end{aligned}$$

is P-Donsker for some δ > 0 and

$$\displaystyle \begin{aligned}{\mathop{\sup}\limits_{h\in H}} P\left(\phi_{\theta,h}-\phi_{\theta_{0},h} \right)^{2} \to 0\end{aligned}$$

as θ → θ 0.

If \(\widehat {\theta }_{n}\) converges in outer probability to θ 0, then

$$\displaystyle \begin{aligned}\left\| G_{n} \left(\phi_{\widehat{\theta}_{n},h}-\phi_{\theta_{0},h} \right)\right\|{}_{H}=o_{p^{*}} \left(1+\sqrt{n} \left\| \widehat{\theta}_{n}-\theta_{0} \right\| \right).\end{aligned}$$

Appendix 6: The Conditional Distribution of Z Given Y

In modeling the conditional distribution of Z given Y when Y is of discrete type, denote the stratified levels of Y as \(SY=\left (1,\cdots ,S\right )\) and describe Y by using dummy variable according to stratified levels.

Assume that there are J s distinct values among the observed covariates under Y -stratum s and they are denoted by \(\left (\boldsymbol {W}_{s1},\boldsymbol {W}_{s2},\cdots ,\boldsymbol {W}_{sJ_{s}} \right )\). Let 0 ≤ p sj ≤ 1 and \(\sum _{j=1}^{J_{s}}p_{sj}=1\) be given. Then the conditional distribution \(f_{e}\left (\boldsymbol {Z}|SY \right )\) is given by

$$\displaystyle \begin{aligned}f_{e} \left(\boldsymbol{Z}|SY\right)=\prod_{s=1}^{S}\left[\sum_{j=1}^{J_{s} }p_{sj} I_{\boldsymbol{Z}=\boldsymbol{W}_{sj}} \right]^{I_{\left(SY=s\right)}},\end{aligned}$$

and the corresponding marginal survival function G of T is given by

$$\displaystyle \begin{aligned}G_{e} \left(t|SY\right)=\prod_{s=1}^{S}\left[\sum_{j=1}^{J_{s}}\left(\exp \left(-\sum_{k=1}^{K}\varLambda_{k} \left(t\right)\exp \left(\boldsymbol{Y}^{T} \boldsymbol{\eta}_{k}+\boldsymbol{W}_{sj}^{T} \boldsymbol{\beta}_{k} \right) \right)\right)p_{sj} \right]^{I_{\left(SY=s\right)}}.\end{aligned}$$

This approach to modeling the covariate distribution is also taken by Scheike and Juul [16]. The proposed NPMLE then applies with the conditional covariate distribution and the marginal survival function of T.

When Y is of continuous type, we can apply kernel smoothing techniques to obtain the estimate for the conditional density of Z given Y . But this is beyond the scope of this paper and deserves further research.

Rights and permissions

Reprints and permissions

Copyright information

© 2020 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Wang, JH., Pan, CH., Chen, YH., Chang, IS. (2020). Non-parametric Maximum Likelihood Estimation for Case-Cohort and Nested Case-Control Designs with Competing Risks Data. In: Zhao, Y., Chen, DG. (eds) Statistical Modeling in Biomedical Research. Emerging Topics in Statistics and Biostatistics . Springer, Cham. https://doi.org/10.1007/978-3-030-33416-1_17

Download citation

Publish with us

Policies and ethics