Skip to main content

Bayesian Functional Mixed Models for Survival Responses with Application to Prostate Cancer

  • Conference paper
  • First Online:
Applied Statistics in Biomedicine and Clinical Trials Design

Part of the book series: ICSA Book Series in Statistics ((ICSABSS))

  • 1952 Accesses

Abstract

In this chapter, we propose a flexible approach to model functional measurements for survival outcomes. Often the class of models for functional observations are assumed to be linear, which may be too restrictive in some cases. We propose an alternative model, in which the simple linear mixed model has been modified by a more flexible semiparametric spline-based functional mixed model, wherein the usage of splines simplifies parameterizations and the joint modeling framework allows synergistic benefit between the regression of functional predictors and the modeling of survival data. We explicitly model the number and location of change points such that our formulation allows for an unknown set of basis functions characterizing the population-averaged and patient-specific trajectories. In addition, we propose a novel auxiliary variable scheme for a fully Bayesian estimation of our model, which not only allows dimension reduction of the parameter space but also allows efficient sampling from the conditional distributions. We illustrate our approach with a recent prostate cancer clinical trial study.

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 129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 169.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 169.99
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

  • Baladandayuthapani V, Mallick BK, Young Hong M, Lupton JR, Turner ND, Carroll RJ (2008) Bayesian hierarchical spatially correlated functional data analysis with application to colon carcinogenesis. Biometrics 64:64–73

    Article  MATH  MathSciNet  Google Scholar 

  • Brown ER, Ibrahim JG (2003) A Bayesian semiparametric joint hierarchical model for longitudinal and survival data. Biometrics 59:221–228

    Article  MATH  MathSciNet  Google Scholar 

  • Brown ER, Ibrahim JG, DeGruttola V (2005) A flexible B-spline model for multiple longitudinal biomarkers and survival. Biometrics 61:64–73

    Article  MATH  MathSciNet  Google Scholar 

  • Bycott P, Taylor J (1998) A comparison of smoothing techniques for CD4 data measured with error in a time-dependent Cox proportional hazards model. Stat Med 17:2061–2077

    Article  Google Scholar 

  • Catalona WJ, Smith DS, Ratliff TL, Dodds KM, Coplen DE, Yuan JJ, Petros JA, Andriole GL (1991) Measurement of prostate-specific antigen in serum as a screening test for prostate cancer. N Eng J Med 324:1156–1161

    Article  Google Scholar 

  • Chen M-H, Shao Q-M, Ibrahim JG (2000) Monte Carlo methods in Bayesian computation. Springer-Verlag, New York

    Book  MATH  Google Scholar 

  • Cox CR (1972) Regression models and life tables. J R Stat Soc, Ser B 34:187–202

    MATH  Google Scholar 

  • Cox DR (1975) Partial likelihood. Biometrika 62:269–276

    Article  MATH  MathSciNet  Google Scholar 

  • Daniels M, Pourahmadi M (2002) Bayesian analysis of covariance matrices and dynamic models for longitudinal data. Biometrika 89:553–566

    Article  MATH  MathSciNet  Google Scholar 

  • DeGruttola V, Tu XM (1994) Modelling progression of CD4-lymphocyte count and its relationship to survival time. Biometrics 50:1003–1014

    Article  Google Scholar 

  • Denison DGT, Mallick BK, Smith AFM (1998) Automatic Bayesian curve fiting. J R Stat Soc B 60:333–350

    Article  MATH  MathSciNet  Google Scholar 

  • Ding J, Wang JL (2008) Modeling longitudinal data with nonparametric multiplicative random effects jointly with survival data. Biometrics 64:546–556

    Article  MATH  MathSciNet  Google Scholar 

  • Eisenberger M, O’Dwyer P, Friedman M (1986) Gonadotropin-hormone releasing hormone analogues: a new therapeutic approach for prostatic carcinoma. J Clin Oncol 4:414–424

    Google Scholar 

  • Ellerhorst J, Tu S, Amato R, Finn L, Millikan R, Pagliaro L, Jackson A, Logothetis C (1997) Phase II trial of alternating weekly chemohormonal therapy for patients with androgen-independent prostate cancer. Clin Cancer Res 3:2371–2376

    Google Scholar 

  • Faucett CJ, Thomas DC (1996) Simultaneously modelling censored survival data and repeatedly measured covariates: a Gibbs sampling approach. Stat Med 15:1663–1685

    Article  Google Scholar 

  • Greenlee R, Murray T, Bolden S, Wingo P (2000) Cancer statistics, 2000. CA Cancer J Clin 50:7–33

    Article  Google Scholar 

  • Gelfand AE, Smith AFM (1990) Sampling-based approaches to calculating marginal densities. J Am Stat Assoc 85:398–404

    Article  MATH  MathSciNet  Google Scholar 

  • Gelfand AE, Dey DK, Chang H (1992) Model determination using predictive distribution with implementation via sampling based methods. In Bernardo JM, Berger JO, Dawid AP, Smith AFM (eds) Bayesian Statistics 4, 147–167. Oxford University Press, Oxford

    Google Scholar 

  • Guo X, Carlin BP (2004) Separate and joint modelling of longitudinal and event time data using standard computer packages. Am Stat 58(1):16–24

    Article  MathSciNet  Google Scholar 

  • Hastings WK (1970) Monte Carlo sampling methods using Markov chains and their applications. Biometrika 57:97–109

    Article  MATH  Google Scholar 

  • Hogan JW, Laird NM (1997) Mixture models for the joint distribution of repeated measures and event times. Stat Med 16:239–257

    Article  Google Scholar 

  • Holmes CC, Mallick BK (2003) Generalized nonlinear modeling with multivariate free-knot regression splines. J Am Stat Assoc 98:352–368

    Article  MATH  MathSciNet  Google Scholar 

  • Hsieh F, Tseng YK, Wang JL (2006) Joint modelling of survival and longitudinal data: likelihood approach revisited. Biometrics 62:1037–1043

    Article  MATH  MathSciNet  Google Scholar 

  • Ibrahim JG, Chen M-H, Sinha D (2001) Bayesian survival analysis. Springer-Verlag, New York

    Book  MATH  Google Scholar 

  • Kalbfleisch JD, Prentice RL (2002) The statistical analysis of failure time data, 2nd edn. Wiley, Hoboken

    Book  MATH  Google Scholar 

  • Kohn R, Smith M, Chan D (2001) Nonparametric regression using linear combinations of basis functions. Stat Comput 11:313–322

    Article  MathSciNet  Google Scholar 

  • Metropolis N, Rosenbluth AW, Rosenbluth MN, Teller AH, Teller E (1953) Equation of state calculations by fast computing machines. J Chem Phys 21:1087–1092

    Article  Google Scholar 

  • Millikan R, Wen S, Pagliaro L, Brown M, Moomey B, Do K-A, Logothets C (2008) Phase III trial of androgen ablation with or without three cycles of systemic chemotherapy for advanced prostate cancer. J Clin Oncol 26:5936–5942

    Article  Google Scholar 

  • Ngo L, Wand MP (2004) Smoothing with Mixed Model Software. J Stat Softw 9:1–54

    Google Scholar 

  • Pauler DK, Finkelstein DM (2002) Predicting time to prostate cancer recurrence based on joint models for non-linear longitudinal biomarkers and event time outcomes. Stat Med 21:3897–3911

    Article  Google Scholar 

  • Prentice R (1982) Covariate measurement errors and parameter estimation in a failure time regression model. Biometrika 69:331–342

    Article  MATH  MathSciNet  Google Scholar 

  • Raboud J, Reid N, Coates RA, Farewell VT (1993) Estimating risks of progressing to AIDS when covariates are measured with error. J R Stat Soc A 156:343–406

    Article  Google Scholar 

  • Ramsay JO, Silverman BW (2005) Functional data analysis, 2nd edn. Springer-Verlag, New York

    Google Scholar 

  • Ruppert D, Wand MP, Carroll RJ (2003) Semiparametric regression. Cambridge University Press, New York

    Book  MATH  Google Scholar 

  • Shi SR, Cote RJ, Yang C, Chen C, Xu HJ, Benedict WF, Taylor CR (1996) Development of an optimal protocol for antigen retrieval: a 'test battery` approach exemplified with reference to the staining of retinoblastoma protein (pRB) in formalin-fixed paraffin sections. J Pathol 179:347–352

    Article  Google Scholar 

  • Slasor P, Laird N (2003) Joint models for efficient estimation in proportional hazards regression models. Stat Med 22:2137–2148

    Article  Google Scholar 

  • Smith M, Kohn R (1996) Nonparametric regression using Bayesian variable selection. J Econ 75:317–343

    Article  MATH  Google Scholar 

  • Spiegelhalter DJ, Best NG, Carlin BP, van der Linde A (2002) Bayesian measures of model complexity and fit. J R Stat Soc, Ser B 64:583–639

    Article  MATH  Google Scholar 

  • Tannock I, Osoba D, Stockler M, Ernst D, Neville A, Moore M, Armitage G, Wilson J, Venner P, Coppin C, Murphy K (1996) Chemotherapy with mitoxantrone plus prednisone or prednisone alone for symptomatic hormone-resistant prostate cancer: a Canadian randomized trial with palliative end points. J Clin Oncol 14:1756–1764

    Google Scholar 

  • Thompson W, Rosen O (2008) A Bayesian model for sparse functional data. Biometrics 64:54–63

    Article  MATH  MathSciNet  Google Scholar 

  • Tsiatis AA, Degruttola V, Wulfsohn MS (1995) Modeling the relationship of survival to longitudinal data measured with error. Applications to survival and Cd4 counts in patients with AIDS. J Am Stat Assoc 90:27–37

    Article  MATH  Google Scholar 

  • Tsiatis AA, Davidian M (2004) Joint modeling of longitudinal and time-to-event data: an overview. Stat Sin 14:809–834

    MATH  MathSciNet  Google Scholar 

  • Wang Y, Taylor JMG (2001) Jointly modeling longitudinal and event time data with application to acquired immunodeficiency syndrome. J Am Stat Assoc 96:895–905

    Article  MATH  MathSciNet  Google Scholar 

  • Wulfsohn MS, Tsiatis AA (1997) A joint model for survival and longitudinal data measured with error. Biometrics 53:330–339

    Article  MATH  MathSciNet  Google Scholar 

  • Yao F (2007) Functional principal component analysis for longitudinal and survival data. Stat Sin 17:965–983

    MATH  Google Scholar 

  • Ye W, Lin X, Taylor J (2008) Semiparametric modeling of longitudinal measurements and time-to-event data–two-stage regression calibration approach. Biometrics 64:1238–1246

    Article  MATH  MathSciNet  Google Scholar 

  • Zhang S, Müller P, Do K-A (2009) A Bayesian semi-parameteric survival model with longitudinal markers. Biometrics 66(2):435–443

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Veerabhadran Baladandayuthapan .

Editor information

Editors and Affiliations

Appendices

Acknowledgments

We thank Dr. Randall Millikan from the University of Texas M.D. Anderson Cancer Center for supplying the prostate cancer data set as an application example of our proposed methodology. V. Baladandayuthapani was partially supported by NIH grant R01 CA160736. Both K-A Do and V. Baladandayuthapani were partially supported by the Cancer Center Support Grant (CCSG) (P30 CA016672). Dr. K-A Do was also partially supported by the University of Texas SPORE in prostate cancer National Institute of Health grant (P50 CA140388). The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Cancer Institute or the National Institutes of Health.

Appendix

2.1 The Model Summary with Specified Prior Distributions

To summarize the hierarchical model setup, we define

$$\begin{aligned}& &\\ {\rm Random \;function} \;{\mathbf{Y}}_{i} &\sim{\rm MVN}({\mathbf{X}}_{i,{\boldsymbol{\gamma}}}{\boldsymbol{\beta}}_{i,{\boldsymbol{\gamma}}},\sigma_{\boldsymbol{\epsilon}}^2 {\mathbf{I}}_{p_i}),\\ \nonumber \sigma_{\boldsymbol{\epsilon}}^2 & \sim IG(a_{\sigma}, b_{\sigma}),\\ \nonumber{\boldsymbol{\beta}}_{i,{\boldsymbol{\gamma}}} &\sim{\rm MVN}({\boldsymbol{\beta}}_{{\boldsymbol{\gamma}}_i}, {\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}), \\ \nonumber{\boldsymbol{\beta}}_{{\boldsymbol{\gamma}}_i} & = {\mathbf{J}}_i {\boldsymbol{\beta}}_{\boldsymbol{\gamma}}, \\ \nonumber{\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}} &= {\mathbf{J}}_i {\boldsymbol{\Omega}}_{\boldsymbol{\gamma}} {\mathbf{J}}_i^{\prime} \;{\rm where} {\boldsymbol{\Omega}}_{\boldsymbol{\gamma}} = {\rm diag}({\boldsymbol{Upsigma}}, \sigma^2 {\mathbf{I}}_{K_{\boldsymbol{\gamma}}^*}), \\ \nonumber{\boldsymbol{\beta}}_{\boldsymbol{\gamma}} & \sim{\rm MVN} (0, c{\mathbf{I}}_{K_{\boldsymbol{\gamma}}}), \\ \nonumber{\boldsymbol{\Sigma}} &\sim{\rm IW}({\mathbf{A}}, b), \\ \nonumber \sigma^2 & \sim{\rm IG}(c_{\sigma}, d_{\sigma}), \\ \nonumber \gamma_k & \sim \mathit{\rm Bernoulli} (\pi_k), \;\mathit{\rm where} \pi_k=\pi {\rm for} \;{\rm all} \;k, \\ \nonumber \pi & \sim \mathit{\rm Beta}(a_{\pi}, b_{\pi}), \\ \nonumber \mathit{\rm Linear} \;\mathit{\rm predictor} \;w_{i} &\sim{\rm N}({\mathbf{B}}_{i,{\boldsymbol{\gamma}}}^{\prime}{\boldsymbol{\theta}}_{\boldsymbol{\gamma}},\tau^2), \;\mathit{\rm where} \;{\mathbf{B}}_{i,{\boldsymbol{\gamma}}}^{\prime}=[{\boldsymbol{\beta}}_{i,{\boldsymbol{\gamma}}}^{\prime}, {\mathbf{L}}_i^{\prime}],\\ \nonumber{\boldsymbol{\theta}}_{\boldsymbol{\gamma}},\tau ^{2}| {\boldsymbol{\gamma}}, {\mathbf{V}}_{\boldsymbol{\gamma}} & \sim{\rm NIG}(0,{\mathbf{V}}_{\boldsymbol{\gamma}},a_{\tau},b_{\tau}), \;\mathit{\rm where} \;{\mathbf{V}}_{\boldsymbol{\gamma}}={\rm diag}({\mathbf{h}}),\\ \nonumber h_\ell &\sim{\rm IG}(c_\ell,d_\ell), \\ \nonumber \mathit{\rm Hazard} \;\mathit{\rm function} \;h(t \mid{\mathbf{Y}}_i) & = h_0 (t) {\rm exp}(w_i), \\ \nonumber h_0(t)&= \lambda_j \;(s_{j-1} \le t<s_j), \\ \nonumber \lambda_j & \sim{\rm IG}(a_j,b_j),\end{aligned}$$

for \(i=1,\ldots,n\), \(j=1,\ldots,J\), \(k=1, \ldots, K\), and \(\ell=1,\ldots,(K_{\gamma}+m)\).

The fourth and fifth lines in the above model need special attention. Based on the fact that the ith curve may not span the complete set of selected change points, \({\boldsymbol{\beta}}_{{\boldsymbol{\gamma}}_i}\) and \({\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}\) are the subject-specific realizations of parameters \({\boldsymbol{\beta}}_{\boldsymbol{\gamma}}\) and \({\boldsymbol{\Omega}}_{\boldsymbol{\gamma}}\), where they respectively represent the population curve and its covariance corresponding to the latent variable \({\boldsymbol{\gamma}}\). The relationship can be expressed via a rectangular indicator matrix \({\mathbf{J}}_i\) as \({\boldsymbol{\beta}}_{{\boldsymbol{\gamma}}_i} = {\mathbf{J}}_i {\boldsymbol{\beta}}_{\boldsymbol{\gamma}}\) and \({\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}} = {\mathbf{J}}_i {\boldsymbol{\Omega}}_{\boldsymbol{\gamma}} {\mathbf{J}}_i^{\prime}\) with \({\boldsymbol{\Omega}}_{\boldsymbol{\gamma}} = {\rm diag}({\boldsymbol{\Sigma}}, \sigma^2 {\mathbf{I}}_{K_{\boldsymbol{\gamma}}^*})\). For example, suppose there are five change points for the population curve, and the ith individual only spans the first two change points (i.e., does not have measurements beyond the third change point). Because the basis has the quadratic polynomial segment and the change points segment, the dimensions of \({\boldsymbol{\beta}}_{\boldsymbol{\gamma}}\) and \({\boldsymbol{\Omega}}_{\boldsymbol{\gamma}}\) will be 8 and 8 by 8. However, for the ith individual, the dimensions of \({\boldsymbol{\beta}}_{{\boldsymbol{\gamma}}_i}\) and \({\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}\) are 5 and 5 by 5. Therefore, \({\boldsymbol{\beta}}_{{\boldsymbol{\gamma}}_i}\) is linked to \({\boldsymbol{\beta}}_{\boldsymbol{\gamma}}\) via a 5 by 8 rectangular index matrix:

$${\mathbf{J}}_i=\left[\begin{array}{cccccccc} 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 \\0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 \\0 & 0 & 0 & 1 & 0 & 0 & 0 & 0 \\0 & 0 & 0 & 0 & 1 & 0 & 0 & 0\end{array} \right].$$
(3.7)

The same \({\mathbf{J}}_i\) is used to link \({\boldsymbol{\Omega}}_{\boldsymbol{\gamma}}\) and \({\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}\). These expressions with \({\mathbf{J}}_i\) enable the derivation of the posterior distributions below.

2.2 Posterior Distributions

The conditional distribution for the ith regressed covariates vector \({\boldsymbol{\beta}}_{i,{\boldsymbol{\gamma}}}\) is updated using regression likelihood

$$ \nonumber{\boldsymbol{\beta}}_{i,{\boldsymbol{\gamma}}} \mid{\mathbf{X}}_{i,{\boldsymbol{\gamma}}}, {\mathbf{Y}}_i, \sigma_{\boldsymbol{\epsilon}}^2, {\boldsymbol{\Omega}}_{\boldsymbol{\gamma}}, w_i, \tau^2,{\boldsymbol{\theta}}_{\boldsymbol{\gamma}}, {\boldsymbol{\gamma}} \sim{\rm MVN}({\boldsymbol{\beta}}_{i,{\boldsymbol{\gamma}}}^*,\tau^2 {\boldsymbol{\Omega}}_{i, {\boldsymbol{\gamma}}}^*),$$

where \({\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}^*=(\tau^2({\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}^{-1}+{\mathbf{X}}_{i,{\boldsymbol{\gamma}}}^{\prime} {\mathbf{X}}_{i,{\boldsymbol{\gamma}}}/\sigma_{\boldsymbol{\epsilon}}^2)+{\boldsymbol{\theta}}_{1i,{\boldsymbol{\gamma}}}{\boldsymbol{\theta}}_{1i,{\boldsymbol{\gamma}}}^{\prime})^{-1}\) and \({\boldsymbol{\beta}}_{i,{\boldsymbol{\gamma}}}^*={\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}^*\times\) \((\tau^2({\mathbf{X}}_{i,{\boldsymbol{\gamma}}}^{\prime}{\mathbf{Y}}_i/\sigma_{\boldsymbol{\epsilon}}^2+{\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}^{-1}{\boldsymbol{\mu}}_{i,{\boldsymbol{\gamma}}})+(w_i-{\mathbf{L}}_i^{\prime}{\boldsymbol{\theta}}_2){\boldsymbol{\theta}}_{1i,{\boldsymbol{\gamma}}})\). The notation \({\boldsymbol{\theta}}_{1i,{\boldsymbol{\gamma}}}\) is the part of the coefficients corresponding to time-dependent covariates \({\boldsymbol{\beta}}_{i,{\boldsymbol{\gamma}}}\), while \({\boldsymbol{\theta}}_2\) is the part of the coefficients corresponding to time-independent covariates \({\mathbf{L}}_i\) in later posteriors. The model variance \(\sigma_{\boldsymbol{\epsilon}}^2\) is updated by

$$\nonumber \sigma_{\boldsymbol{\epsilon}}^2 | {\boldsymbol{\beta}}_{i,{\boldsymbol{\gamma}}},{\mathbf{Y}}_i,{\mathbf{X}}_{i,{\boldsymbol{\gamma}}} \sim{\rm IG}(a_{\sigma}^*, b_{\sigma}^*),$$

where \(a_{\sigma}^*=a_{\sigma}+(\sum_{i=1}^n p_i)/2\) and \(b_{\sigma}^*=b_{\sigma}+[\sum_{i=1}^n({\mathbf{Y}}_i - {\mathbf{X}}_{i,{\boldsymbol{\gamma}}} {\boldsymbol{\beta}}_{i,{\boldsymbol{\gamma}}})^{\prime}({\mathbf{Y}}_i - {\mathbf{X}}_{i,{\boldsymbol{\gamma}}} {\boldsymbol{\beta}}_{i,{\boldsymbol{\gamma}}})]/2\). The indicator vector \({\boldsymbol{\gamma}}\) can be updated elementwise using the Metropolis–Hastings algorithm with marginal posterior \(\gamma_k \mid{\boldsymbol{\gamma}}_{-k},{\mathbf{Y}}_i, {\mathbf{X}}_{i,{\boldsymbol{\gamma}}}, \sigma_{\boldsymbol{\epsilon}}^2,\) \({\boldsymbol{Upsigma}}, \sigma^2, {\boldsymbol{\theta}}_{\boldsymbol{\gamma}}, {\mathbf{V}}_{\boldsymbol{\gamma}}\) proportional to

$$\begin{aligned} & \pi(\gamma_k)\pi({\boldsymbol{\theta}}_{\boldsymbol{\gamma}})\pi({\mathbf{V}}_{\boldsymbol{\gamma}})\left[\frac{|{\boldsymbol{\Upphi}}_{\boldsymbol{\gamma}}^{-1}|}{|c{\mathbf{I}}_{\boldsymbol{\gamma}}|} \prod\nolimits_{i=1}^n \frac{|\tau^2 {\mathbf{M}}_{i,{\boldsymbol{\gamma}}}^{-1}|}{|{\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}|} \right]^{1/2} {\rm exp} \left\{ \frac{1}{2 \tau^2} \sum\nolimits_{i=1}^n {\boldsymbol{\alpha}}_{i,{\boldsymbol{\gamma}}}^{\prime} {\mathbf{M}}_{i,{\boldsymbol{\gamma}}}^{-1} {\boldsymbol{\alpha}}_{i,{\boldsymbol{\gamma}}} \right\} \\ \nonumber & \times {\rm exp} \left\{ \frac{1}{2} \left( \sum\nolimits_{i=1}^n {\boldsymbol{\alpha}}_{i,{\boldsymbol{\gamma}}}^{\prime} {\mathbf{M}}_{i,{\boldsymbol{\gamma}}}^{-1} {\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}^{-1} {\mathbf{J}}_i\right) {\boldsymbol{\Phi}}_{\boldsymbol{\gamma}}^{-1} \left( \sum\nolimits_{i=1}^n {\mathbf{J}}_i {\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}^{-1} {\mathbf{M}}_{i,{\boldsymbol{\gamma}}}^{-1} {\boldsymbol{\alpha}}_{i,{\boldsymbol{\gamma}}} \right) \right\},\end{aligned}$$

where \({\boldsymbol{\alpha}}_{i,{\boldsymbol{\gamma}}} = \tau^2 {\mathbf{X}}_{i,{\boldsymbol{\gamma}}}^{\prime} {\mathbf{Y}}_i/\sigma_{\boldsymbol{\epsilon}}^2 + (w_i-{\mathbf{L}}_i^{\prime} {\boldsymbol{\theta}}_2){\boldsymbol{\theta}}_{1i,{\boldsymbol{\gamma}}}\), \({\mathbf{M}}_{i,{\boldsymbol{\gamma}}}=\tau^2({\mathbf{X}}_{i,{\boldsymbol{\gamma}}}^{\prime} {\mathbf{X}}_{i,{\boldsymbol{\gamma}}} /\sigma_{\boldsymbol{\epsilon}}^2 +{\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}^{-1}) + {\boldsymbol{\theta}}_{1i,{\boldsymbol{\gamma}}} {\boldsymbol{\theta}}_{1i,{\boldsymbol{\gamma}}}^{\prime}\) and \({\boldsymbol{\Upphi}}_{\boldsymbol{\gamma}}=(\sum_{i=1}^n {\mathbf{J}}_i^{\prime} {\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}^{-1} {\mathbf{J}}_i) - \tau^2 (\sum_{i=1}^n {\mathbf{J}}_i^{\prime} {\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}^{-1}{\mathbf{M}}_{i,{\boldsymbol{\gamma}}}^{-1}{\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}^{-1} {\mathbf{J}}_i)+(1/c){\mathbf{I}}_{K_{\boldsymbol{\gamma}}}\). It is worth to point out that the generation of a candidate \({\boldsymbol{\gamma}}\) is done by changing one element at a time in fixed sequencing order within each iteration. The conditional distribution for the informative scalar w i follows the combination of information from both the regression and proportional hazards models. The likelihood of the PH model leads to its nonstandard form

$$\begin{aligned} w_i \mid T_i, \delta_i, h_0(t), {\mathbf{B}}_{i, {\boldsymbol{\gamma}}}, {\boldsymbol{\theta}}_{\boldsymbol{\gamma}}, \tau^2 \quad\propto &{\rm exp} \left\{-\frac{(w_i^2-2w_i{\mathbf{B}}_{i,{\boldsymbol{\gamma}}}^{\prime} {\boldsymbol{\theta}}_{i,{\boldsymbol{\gamma}}})}{2\tau^2} \right\} \times \\ & \Bigg[ h_0(T_i) {\rm exp}(w_i) \Bigg]^{\delta_i} {\rm exp}\left\{ - \int_0^{T_i} {\rm exp}(w_i)h_0(t) du \right\},\end{aligned}$$

which can be updated by a Metropolis step.

The following layer includes the regression coefficient as population mean \({\boldsymbol{\beta}}_{\boldsymbol{\gamma}}\), which can be updated as

$$\nonumber{\boldsymbol{\beta}}_{\boldsymbol{\gamma}} \mid{\boldsymbol{\beta}}_{i,{\boldsymbol{\gamma}}}, {\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}} \sim{\rm MVN}({\boldsymbol{\beta}}_{\boldsymbol{\gamma}}^*, c{\mathbf{M}}),$$

where \({\mathbf{M}}=(c \sum_{i=1}^n {\mathbf{J}}_i^{\prime} {\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}^{-1} {\mathbf{J}}_i+{\mathbf{I}}_{K_{\boldsymbol{\gamma}}})^{-1}\) and \({\boldsymbol{\beta}}_{\boldsymbol{\gamma}}^*=c{\mathbf{M}}(\sum_{i=1}^n {\mathbf{J}}_i^{\prime}{\boldsymbol{\Omega}}_{i,{\boldsymbol{\gamma}}}^{-1}{\boldsymbol{\beta}}_{i,{\boldsymbol{\gamma}}})\). The unstructured covariance matrix of the polynomial part for quadratic spline coefficients, \({\boldsymbol{Upsigma}}\), is updated as

$$\nonumber{\boldsymbol{Upsigma}} | {\boldsymbol{\beta}}_{\boldsymbol{\gamma}}, {\boldsymbol{\beta}}_{i,{\boldsymbol{\gamma}}} \sim IW({\mathbf{A}}^*, b^*),$$

where \({\mathbf{A}}^*=[{\mathbf{A}}^{-1}+ \sum_{i=1}^n ({\boldsymbol{\alpha}}_{i1} {\boldsymbol{\alpha}}_{i1}^{\prime})]^{-1}\), \({\boldsymbol{\alpha}}_i={\boldsymbol{\beta}}_{i,{\boldsymbol{\gamma}}}-{\boldsymbol{\beta}}_{{\boldsymbol{\gamma}}_i}=[{\boldsymbol{\alpha}}_{i1}^{\prime}, {\boldsymbol{\alpha}}_{i2}^{\prime}]^{\prime}\), and \(b^*=b+n\). Here, the dimensions of \({\boldsymbol{\alpha}}_{i1}\) and \({\boldsymbol{\alpha}}_{i2}\) are 3 × 1 and \(K_{{\boldsymbol{\gamma}}_i}^* \times 1\). Linking to the covariance of the change points part for the quadratic spline coefficients, σ2, is updated as

$$\nonumber \sigma^2 \mid{\boldsymbol{\beta}}_{i,{\boldsymbol{\gamma}}}'s, {\boldsymbol{\beta}}_{\boldsymbol{\gamma}} \sim IG(c_{\sigma}^*, d_{\sigma}^*),$$

where \(c_{\sigma}^*=c_{\sigma}+(\sum_{i=1}^n K_{{\boldsymbol{\gamma}}_i}^*)/2\) and \(d_{\sigma}^*=d_{\sigma}+(\sum_{i=1}^n {\boldsymbol{\alpha}}_{i2}^{\prime}{\boldsymbol{\alpha}}_{i2})/2\). The probability of being change point π can be updated as

$$\nonumber \pi \mid{\boldsymbol{\gamma}} \sim Beta(a_{\pi}^*, b_{\pi}^*),$$

where \(a_{\pi}^*=a_{\pi}+K_{\boldsymbol{\gamma}}\) and \(b_{\pi}^*=b_{\pi}+K+K_{\boldsymbol{\gamma}}\). The common coefficient vector \({\boldsymbol{\theta}}\) in the linear predictor model is updated as

$$\nonumber{\boldsymbol{\theta}}_{\boldsymbol{\gamma}} \mid{\mathbf{w}}, {\mathbf{B}}, \tau^2, {\mathbf{V}}_{\boldsymbol{\gamma}} \sim{\rm MVN}({\boldsymbol{\theta}}^*, \tau^2 {\mathbf{V}}^*),$$

where \({\mathbf{V}}^* = ({\mathbf{V}}_{\boldsymbol{\gamma}}^{-1}+\sum_{i=1}^n {\mathbf{J}}_i^{\prime} {\mathbf{B}}_{i,{\boldsymbol{\gamma}}} {\mathbf{B}}_{i,{\boldsymbol{\gamma}}}^{\prime} {\mathbf{J}}_i)^{-1}\) and \({\boldsymbol{\theta}}^* = {\mathbf{V}}^*(\sum_{i=1}^n w_i {\mathbf{J}}_i^{\prime} {\mathbf{B}}_{i,{\boldsymbol{\gamma}}})\). Here the definition of \({\mathbf{J}}_i\) is similar to its definition in (7.2) with a dimension adjustment to match \({\mathbf{B}}_{i,{\boldsymbol{\gamma}}}\). The conjugate inverse gamma prior for variance Ï„2 leads to its conditional distribution:

$$ \tau^2 \mid{\boldsymbol{\theta}}_{\boldsymbol{\gamma}}, {\mathbf{V}}_{\boldsymbol{\gamma}}, {\mathbf{w}}, {\mathbf{B}} \sim IG(a_{\tau}^*,b_{\tau}^*),$$

where \(a_{\tau}^*=a_{\tau}+(n+K_{\boldsymbol{\gamma}}+m)/2\) and \(b_{\tau}^*=b_{\tau}+\Big[{\boldsymbol{\theta}}_{\boldsymbol{\gamma}}^{\prime} {\mathbf{V}}_{\boldsymbol{\gamma}}^{-1} {\boldsymbol{\theta}}_{\boldsymbol{\gamma}} + \sum_{i=1}^n (w_i-{\mathbf{B}}_{i,{\boldsymbol{\gamma}}}^{\prime} {\boldsymbol{\theta}}_{i,{\boldsymbol{\gamma}}})^2 \Big]/2\).

The next layer includes scale parameters h k , which is updated by

$$\nonumber h_{\ell} \mid{\boldsymbol{\theta}}_{\boldsymbol{\gamma}}, \tau^2 \sim IG(c_{\ell}^*,d_{\ell}^*),$$

where \(c_{\ell}^*=c_{\ell}+1/2\) and \(d_{\ell}^*=d_{\ell}+\theta_{\ell}^2 /2 \tau^2\).

The parameters of baseline hazard step function \(h_0(t)\), λ j ’s, can be updated using the proportional hazards model:

$$\nonumber \lambda_j \mid{\mathbf{T}}, {\mathbf{w}} \sim IG(a_j^*, b_j^*),$$

where \(a_j^*=a_j+ \sum_{i=1}^n \delta_i I(s_{j-1} \le T_i < s_j)\) and \(b_j^*=b_j+\sum_{i=1}^n \Big[ I(T_i> s_{j-1}) \times \Big.\) \(\Big. \int_{s_{j-1}}^{{\rm min}(T_i,s_j)} \mathit{\rm exp} (w_i)\mathit{{\rm d}u} \Big]\).

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer International Publishing Switzerland

About this paper

Cite this paper

Baladandayuthapan, V., Wang, X., Mallick, B., Do, KA. (2015). Bayesian Functional Mixed Models for Survival Responses with Application to Prostate Cancer. In: Chen, Z., Liu, A., Qu, Y., Tang, L., Ting, N., Tsong, Y. (eds) Applied Statistics in Biomedicine and Clinical Trials Design. ICSA Book Series in Statistics. Springer, Cham. https://doi.org/10.1007/978-3-319-12694-4_3

Download citation

Publish with us

Policies and ethics