Skip to main content
Log in

Dynamic Prediction of Renal Failure Using Longitudinal Biomarkers in a Cohort Study of Chronic Kidney Disease

  • Published:
Statistics in Biosciences Aims and scope Submit manuscript

Abstract

In longitudinal studies, prognostic biomarkers are often measured longitudinally. It is of both scientific and clinical interest to predict the risk of clinical events, such as disease progression or death, using these longitudinal biomarkers as well as other time-dependent and time-independent information about the patient. The prediction is dynamic in the sense that it can be made at any time during the follow-up, adapting to the changing at-risk population and incorporating the most recent longitudinal data. One approach is to build a joint model of longitudinal predictor variables and time to the clinical event, and draw predictions from the posterior distribution of the time to event conditional on longitudinal history. Another approach is to use the landmark model, which is a system of prediction models that evolve with the follow-up time. We review the pros and cons of the two approaches and present a general analytical framework using the landmark approach. The proposed framework allows the measurement times of longitudinal data to be irregularly spaced and differ between subjects. We propose a unified kernel weighting approach for estimating the model parameters, calculating the predicted probabilities, and evaluating the prediction accuracy through double time-dependent receiver operating characteristic curves. We illustrate the proposed analytical framework using the African American study of kidney disease and hypertension to develop a landmark model for dynamic prediction of end-stage renal diseases or death among patients with chronic kidney 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

Similar content being viewed by others

References

  1. Anderson JR, Cain KC, Gelber RD (1983) Analysis of survival by tumor response. J Clin Oncol 1(11):710–719

    Article  Google Scholar 

  2. Blanche P, Dartigues JF, Jacqmin-Gadda H (2013) Review and comparison of ROC curve estimators for a time-dependent outcome with marker-dependent censoring. Biom J 55:687–704

    Article  MathSciNet  MATH  Google Scholar 

  3. Blanche P, Proust-Lima C, Loubere L, Berr C, Dartigues JF, Jacqmin-Gadda H (2015) Quantifying and comparing dynamic predictive accuracy of joint models for longitudinal marker and time-to-event in presence of censoring and competing risks. Biometrics 71(1):102–113

    Article  MathSciNet  MATH  Google Scholar 

  4. Buonaccorsi JP (1995) Prediction in the presence of measurement error: general discussion and an example predicting defoliation. Biometrics 51(4):1562–1569

    Article  MathSciNet  MATH  Google Scholar 

  5. Copas JB, Corbett P (2002) Overestimation of the receiver operating characteristic curve for logistic regression. Biometrika 89(2):315–331

    Article  MathSciNet  MATH  Google Scholar 

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

    MATH  Google Scholar 

  7. Dafni U (2011) Landmark analysis at the 25-year landmark point. Circ Cardiovasc Qual Outcomes 4(3):363–371

    Article  MathSciNet  Google Scholar 

  8. Echouffo-Tcheugui JB, Kengne AP (2012) Risk models to predict chronic kidney disease and its progression: a systematic review. PLoS Med 9(11):e1001344

    Article  Google Scholar 

  9. Feuer EJ, Hankey BF, Gaynor JJ, Wesley MN, Baker SG, Meyer JS (1992) Graphical representation of survival curves associated with a binary non-reversible time dependent covariate. Stat Med 11:455–474

    Article  Google Scholar 

  10. Gassman JJ, Greene T, Wright JT Jr et al (2003) Design and statistical aspects of the African American Study of Kidney Disease and Hypertension (AASK). J Am Soc Nephrol 14(7):S154–S165

    Article  Google Scholar 

  11. Gerds TA, Schumacher M (2006) Consistent estimation of the expected Brier score in general survival models with right-censored event times. Biom J 48(6):1029–1040

    Article  MathSciNet  Google Scholar 

  12. Gong Q, Schaubel DE (2013) Partly conditional estimation of the effect of a time-dependent factor in the presence of dependent censoring. Biometrics 69(2):338–347

    Article  MathSciNet  MATH  Google Scholar 

  13. Heagerty PJ, Lumley T, Pepe MS (2000) Time-dependent ROC curves for censored survival data and a diagnostic marker. Biometrics 56:337–344

    Article  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  15. Hsieh F, Tseng YK, Wang JL (2006) Joint modeling of survival and longitudinal data: likelihood approach revisited. Biometrics 62(4):1037–1043

    Article  MathSciNet  MATH  Google Scholar 

  16. Kidney Disease: Improving Global Outcomes (KDIGO) CKD Work Group (2012) KDIGO 2012 clinical practice guideline for the evaluation and management of chronic kidney disease. Kidney Int Suppl 2013(3):1–150

  17. Lee EW, Wei LJ, Amato DA (1992) Cox-type regression analysis for large numbers of small groups of correlated failure time observations. In: Klein JP, Goel PK (eds) In survival analysis: state of the art. Kluwer Academic, Norwell, MA, pp 237–247

    Chapter  Google Scholar 

  18. Li L, Astor BC, Lewis J, Hu B, Appel LJ, Lipkowitz MS, Toto RD, Wang X, Wright JT Jr, Greene TH (2012) Longitudinal progression trajectory of GFR among patients with CKD. Am J Kidney Dis 59(4):504–512

    Article  Google Scholar 

  19. Li L, Greene T (2008) Varying coefficients model with measurement error. Biometrics 64(2):519–526

    Article  MathSciNet  MATH  Google Scholar 

  20. Li L, Hu B, Greene T (2015) A simple method to estimate the time-dependent ROC curve under right censoring. COBRA preprint series # 1167

  21. Li L, Hu B, Greene T (2009) A semiparametric joint model for longitudinal and survival data with application to hemodialysis study. Biometrics 65(3):737–745

    Article  MathSciNet  MATH  Google Scholar 

  22. Lin DY (1994) Cox regression analysis of multivariate failure time data: the marginal approach. Stat Med 13(21):2233–2247

    Article  Google Scholar 

  23. Lin DY (2007) On the Breslow estimator. Lifetime Data Anal 13:471–480

    Article  MathSciNet  MATH  Google Scholar 

  24. Marks A, Fluck N, Black C (2014) Chronic kidney disease where next? Predicting outcomes and planning care pathways. Eur Med J Nephrol 1:67–75

    Google Scholar 

  25. National Kidney Foundation (2002) K/DOQI clinical practice guidelines for chronic kidney disease: evaluation, classification, and stratification. Am J Kidney Dis 39(2 Suppl 1): S1-266

  26. Parast L, Cheng SC, Cai T (2012) Landmark prediction of long term survival incorporating short term event time information. J Am Stat Assoc 107(500):1492–1501

    Article  MathSciNet  MATH  Google Scholar 

  27. Pullenayegum EM, Lim LS (2014). Longitudinal data subject to irregular observation: a review of methods with a focus on visit processes, assumptions, and study design. Stat Methods Med Res. doi:10.1177/0962280214536537

  28. Rizopoulos D (2011) Dynamic predictions and prospective accuracy in joint models for longitudinal and time-to-event data. Biometrics 67:819–829

    Article  MathSciNet  MATH  Google Scholar 

  29. Rizopoulos D (2012) Joint models for longitudinal and time-to-event data: with applications in R. Chapman & Hall/CRC, Boca Raton

    Book  MATH  Google Scholar 

  30. Rizopoulos D, Verbeke G, Lesaffre E (2009) Fully exponential Laplace approximations for the joint modelling of survival and longitudinal data. J R Stat Soc Ser B 71(3):637–654

    Article  MathSciNet  MATH  Google Scholar 

  31. Rosansky SJ, Glassock RJ (2014) Is a decline in estimated GFR an appropriate surrogate end point for renoprotection drug trials? Kidney Int 85(4):723–727

    Article  Google Scholar 

  32. Spiekerman CF, Lin DY (1998) Marginal regression models for multivariate failure time data. J Am Stat Assoc 93(443):1164–1175

    Article  MathSciNet  MATH  Google Scholar 

  33. Sun J, Sun L, Liu D (2007) Regression analysis of longitudinal data in the presence of informative observation and censoring times. J Am Stat Assoc 102(480):1397–1406

    Article  MathSciNet  MATH  Google Scholar 

  34. Tangri N, Kitsios GD, Inker LA, Griffith J, Naimark DM, Walker S, Rigatto C, Uhlig K, Kent DM, Levey AS (2013) Risk prediction models for patients with chronic kidney disease: a systematic review. Ann Intern Med 158(8):596–603

    Article  Google Scholar 

  35. Taylor JM, Park Y, Ankerst DP, Proust-Lima C, Williams S, Kestin L, Bae K, Pickles T, Sandler H (2013) Real-time individual predictions of prostate cancer recurrence using joint models. Biometrics 69(1):206–213

    Article  MathSciNet  MATH  Google Scholar 

  36. Tseng YK, Hsieh F, Wang JL (2005) Joint modelling of accelerated failure time and longitudinal data. Biometrika 92(3):587603

    Article  MathSciNet  MATH  Google Scholar 

  37. Tsiatis AA, Davidian M (2004) An overview of joint modeling of longitudinal and time-to-event data. Stat Sin 14:793818

    MATH  Google Scholar 

  38. United States Renal Data System (2014) USRDS 2014 annual data report: atlas of chronic kidney disease and end-stage renal disease in the United States. National Institute of Diabetes and Digestive and Kidney Diseases, Bethesda, MD

  39. Van Houwelingen HC (2007) Dynamic prediction by landmarking in event history analysis. Scand J Stat 34(1):70–85

    Article  MathSciNet  MATH  Google Scholar 

  40. Van Houwelingen HC, Putter H (2012) Dynamic prediction in clinical survival analysis. Chapman & Hall/CRC, Boca Raton

    MATH  Google Scholar 

  41. Vickers AJ, Till C, Tangen CM, Lilja H, Thompson IM (2011) An empirical evaluation of guidelines on prostate-specific antigen velocity in prostate cancer detection. J Natl Cancer Inst 103(6):462–469

    Article  Google Scholar 

  42. Wu L, Liu W, Yi GY, Huang Y (2012) Analysis of longitudinal and survival data: joint modeling, inference methods, and issues. J Probab Stat, Article ID 640153

  43. Zhang W, Lee S-Y, Song X (2002) Local polynomial fitting in semi-varying coefficient model. J Multivar Anal 82:166–188

    Article  MATH  Google Scholar 

  44. Zheng YY, Heagerty PJ (2005) Partly conditional survival models for longitudinal data. Biometrics 61(2):379–391

    Article  MathSciNet  MATH  Google Scholar 

  45. Zheng Y, Heagerty PJ (2007) Prospective accuracy for longitudinal markers. Biometrics 63(2):332–341

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Liang Li.

Additional information

This research was supported by NIH grants P30CA016672 and 5R01DK090046. The authors thank the Guest Editor and two anonymous reviewers for their insightful comments that greatly improved this paper.

Appendices

Appendix 1

We prove the consistency of the sensitivity estimator (8). The proof for the specificity estimator (9) is similar and omitted. The proof makes use of the following regularity conditions:

  1. (A1)

    The kernel function K(.) is a symmetric density function centered around zero with compact support.

  2. (A2)

    The variable U has bounded support; its density function \(f_U(.)\) is Lipschitz continuous and \(f_U(s) > 0\), a.s.

  3. (A3)

    \(h_1 \rightarrow 0\), \(nh_1^2 \rightarrow \infty \), and \(h_1\) satisfies the conditions in Lemma 1 below.

According to [6], the kernel weighted Kaplan–Meier estimator is consistent under standard regularity conditions: \(\hat{S}_{\tilde{T}}(t| U_i, Q_i ) = S_{\tilde{T}}(t|U_i, Q_i) + o_p(1)\). Therefore,

$$\begin{aligned} \hat{W}(U_i, Q_i, \tilde{Y}_i, \tilde{\delta }_i) =&1 - (1-\tilde{\delta }_i) \dfrac{ S_{\tilde{T}}( \tau _1 | U_i, Q_i ) + o_p(1) }{ S_{\tilde{T}}( \tilde{Y}_i | U_i, Q_i ) + o_p(1) } \\ =&W(U_i, Q_i, \tilde{Y}_i, \tilde{\delta }_i) + o_p(1) \end{aligned}$$
$$\begin{aligned} \hat{Se}(s, c) =&\dfrac{ \sum _{i \in \mathcal {R}(s)} K_{h_1}(U_i, s) W( U_i, Q_i, \tilde{Y}_i, \tilde{\delta }_i)1(Q_i> c) + o_p(1) }{ \sum _{i \in \mathcal {R}(s)} K_{h_1}(U_i, s) W( U_i, Q_i, \tilde{Y}_i, \tilde{\delta }_i) + o_p(1) } \\ =&\dfrac{ \sum _{i \in \mathcal {R}(s)} K_{h_1}(U_i, s) W( U_i, Q_i, \tilde{Y}_i, \tilde{\delta }_i)1(Q_i> c) }{ \sum _{i \in \mathcal {R}(s)} K_{h_1}(U_i, s) W( U_i, Q_i, \tilde{Y}_i, \tilde{\delta }_i) } + o_p(1) \\ =&\dfrac{ \sum _{i=1}^n K_{h_1}(U_i, s) W( U_i, Q_i, \tilde{Y}_i, \tilde{\delta }_i)1(Q_i> c)1(\tilde{Y}_i> s) }{ \sum _{i=1}^n K_{h_1}(U_i, s) W( U_i, Q_i, \tilde{Y}_i, \tilde{\delta }_i) 1(\tilde{Y}_i > s) } + o_p(1). \end{aligned}$$

Using Lemma 2 below, the equation above equals

$$\begin{aligned} \hat{Se}(s, c) =&~ \dfrac{ E\left\{ W(U_i, Q_i, \tilde{Y}_i, \tilde{\delta }_i) 1(Q_i> c) 1(\tilde{Y}_i> s) ~|~ s \right\} f_U(s) }{ E\left\{ W(U_i, Q_i, \tilde{Y}_i, \tilde{\delta }_i) 1(\tilde{Y}_i > s) ~|~ s \right\} f_U(s) } + o_p(1), \end{aligned}$$

where \(f_U(.)\) is the density function of \(U_i\). Replacing \(W(U_i, Q_i, \tilde{Y}_i, \tilde{\delta }_i)\) by \( E\left( 1(\tilde{T}_i \le \tau _1) ~|~ U_i, Q_i, \tilde{Y}_i, \tilde{\delta }_i \right) \) in the equation above, we have

$$\begin{aligned} \hat{Se}(s, c) =&\dfrac{ E\left\{ 1(\tilde{T}_i \le \tau _1) 1(Q_i> c) 1(\tilde{Y}_i> s) ~|~ s \right\} }{ E\left\{ 1(\tilde{T}_i \le \tau _1) 1(\tilde{Y}_i> s) ~|~ s \right\} } + o_p(1) \\ =&P( Q_i> c | \tilde{T}_i \in (s, s + \tau _1] , \tilde{Y}_i > s, s ) + o_p(1) \nonumber . \end{aligned}$$

This completes the proof of consistency.

Lemma 1

Let \(\{(X_i, Y_i), i=1,\ldots ,n\}\) be independent and identically distributed random vectors, where the \(Y_i\)’s are scalar random variables. Assume that \(E|Y|^s<\infty \) and \(\text {sup}_x\int |Y|^s f(X, Y)dY < \infty \), where f denotes the joint density of (XY). K is a bounded positive function with a bounded support, satisfying a Lipschitz condition. Then

$$\begin{aligned} \underset{x}{\mathbf {sup}} \Bigl |\dfrac{1}{n}\sum _{i=1}^{n} \bigl [ K_h(X_i-x)Y_i - E\{ K_h(X_i-x)Y_i \} \bigr ] \Bigr | = O_p\Bigl ( \Bigl \{ \dfrac{\log (1/h)}{nh} \Bigr \}^{1/2} \Bigr ) \end{aligned}$$

provided that \(n^{2\epsilon -1}h \rightarrow \infty \) for some \(\epsilon < 1-s^{-1}\). (This lemma was used in [43]).

Lemma 2

Suppose \(\{(Y_i, T_i), i=1,\ldots ,n\}\) satisfy the conditions in Lemma 1; \(T_i\)’s satisfy the regularity conditions (A1)–(A3); E(Y|T) is Lipschitz continuous. Then as \(n\rightarrow \infty \):

$$\begin{aligned} n^{-1}\sum _{i=1}^{n}K_h(T_i - t_0)Y_i = E(Y|t_0)f(t_0) + O_p(h^2) + O_p\Bigl ( \Bigl \{ \dfrac{\log (1/h)}{nh} \Bigr \}^{1/2} \Bigr ) . \nonumber \end{aligned}$$

The proof of this result uses Lemma 1 and a Taylor series expansion of E(Y|T)f(T) around \(t_0\). (This lemma was used in [19]).

Appendix 2

With irregularly spaced measurement times, some subjects may not have an adequate number of measurements within the history window \(( s-\tau _2, s)\) for calculating the slope, or rate of change, of a biomarker. We outline a solution to this problem. Suppose we want to make a prediction on a subject at time s. The repeated biomarker measurements and the corresponding measurement times within the history window are denoted by vectors \(\mathbf {Z}_{0}\) and \(\mathbf {t}_0\), respectively. Here we focus on the case when s equals the last measurement time in \(\mathbf {t}_{0}\). The case when s is not a measurement time has been discussed in Sect. 2.2. Let \(\mathcal {R}(s)\) denote the set of subjects in the training dataset that are at risk at time s, i.e., \(\mathcal {R}(s) = \{ i | i=1,2,\ldots ,n; Y_i > s \}\). For the i-th subject in \(\mathcal {R}(s)\), let \(t_{ij}\) be the measurement time that is the closest to s, i.e., \(| t_{ij} - s | \le | t_{ij^{'}} - s | \) for any \(j^{'} \in \{ 1,2,\ldots ,n_i \}\) (in case of ties, make a random choice). Let \(\mathbf {Z}_{i}\) and \(\mathbf {t}_{i}\) be the corresponding vectors of biomarker measurements and measurement times in the history window, respectively (by definition, the last element in \(\mathbf {t}_{i}\) is \(t_{ij}\)). Fit a random intercept and slope model to the data \(\{ \mathbf {Z}_i, \mathbf {t}_i, i \in \mathcal {R}(s) \}\), with the i-th subject being weighted by the kernel weight \(W_i = h^{-1}K\left( (t_{ij} - s)/h \right) \). This is done by maximizing the following weighted multivariate Gaussian log-likelihood with respect to \(\mathbf {\alpha }(s)\), \(\mathbf {\Sigma }(s)\), and \(\sigma (s)\):

$$\begin{aligned}&\underset{i \in \mathcal {R}(s) }{\sum } \left( -\dfrac{W_i}{2} \right) \log \left| \mathbf {D}_i \mathbf {\Sigma }(s) \mathbf {D}_i^T + \sigma (s)^2 \mathbf {I} \right| \nonumber \\&\quad + \underset{i \in \mathcal {R}(s) }{\sum } \left( -\dfrac{W_i}{2} \right) \left( \mathbf {Z}_i - \mathbf {D}_i\mathbf {\alpha }(s) \right) ^T \left( \mathbf {D}_i \mathbf {\Sigma }(s) \mathbf {D}_i^T + \sigma (s)^2 \mathbf {I} \right) ^{-1} \left( \mathbf {Z}_i - \mathbf {D}_i\mathbf {\alpha }(s) \right) ,\qquad \quad \end{aligned}$$
(10)

where \(\mathbf {D}_i\) is a two-column matrix consisting of a vector of 1’s and \(\mathbf {t}_i\); \(\mathbf {\alpha }(s)\) is a \(2\times 1\) vector of the fixed effect coefficients for intercept and slope; and \(\mathbf {\Sigma }(s)\) is the \(2\times 2\) variance matrix of the random effects. Given the estimated parameters, the rate of change may be defined to be the second element of the following \(2 \times 1\) vector:

$$\begin{aligned} \hat{\mathbf {\alpha }}(s) + \hat{\mathbf {\Sigma }}(s) \hat{D}_0^T \left( \mathbf {D}_0 \hat{\mathbf {\Sigma }}(s) \mathbf {D}_0^T + \sigma (s)^2 \mathbf {I} \right) ^{-1} \left( \mathbf {Z}_0 - \mathbf {D}_0 \hat{\mathbf {\alpha }}(s) \right) , \end{aligned}$$
(11)

where \(\mathbf {D}_0\) is a two-column matrix consisting of a vector of 1’s and \(\mathbf {t}_0\). Equation (11) is the best linear unbiased prediction of the random intercept and slope. Since \(\mathbf {\alpha }(s)\), \(\mathbf {\Sigma }(s)\), and \(\sigma (s)\) can be estimated at any landmark time s, their trajectories characterize the temporal change of the at-risk population.

We emphasize that the procedure above is not intended to provide a consistent or unbiased estimate of the slope of the biomarker trajectory at time s for each subject, or the mean slope of the at-risk subjects at time s, because the trajectory may be non-linear and \(\tau _2\) does not go to 0 as n increases. The procedure should be viewed as a principled way of defining biomarker slope other than using the least squares estimate. It causes the individual least squares slope to shrink toward the mean slope of the at-risk subjects and thus increases the stability of individual slopes. When the individual biomarker trajectory is non-linear and smooth, estimating the slope as a smooth time-varying function is difficult in both the landmark modeling and joint modeling frameworks, unless strong assumptions are made for the trajectory shape or large numbers of repeated measures per subject are available. In clinical practice, the concept of a time-varying first-order derivative of a non-linear smooth “true” trajectory is rarely used in prognostic evaluation. Physicians often use the least square slope in the history window to describe, in broad strokes, the overall trend of the biomarker during that time, even when in reality most of the biomarker trajectories are non-linear to certain extent and their first-order derivatives are never constant within the history window. Examples include the prostate-specific antigen (PSA) velocity [41] and GFR slope [31]. The procedure above is an algorithmic formulation of the concept of rate of change used in clinical practice. Therefore, this linear mixed model with time-varying parameters should be viewed as a working assumption.

Even when the “true” individual trajectory is indeed linear within the history window with a constant underlying “true” slope, the BLUP is not the true slope and their difference is often viewed as a manifestation of measurement error in the biomarker [21]. The magnitude of the error is of order \(\sigma _i/\sqrt{m_i}\), where \(\sigma _i\) is the residual variance of subject i and \(m_i\) is the number of eGFR measurements within the history window. The measurement error model literature showed that covariate measurement error does not cause bias in prediction if the distribution of the observed data is the same in the training and testing datasets [4]. Therefore, despite the measurement error, we can still use the BLUP for prediction purpose. This is the justification for the proposed biomarker slope definition. However, the assumption that the training and testing datasets have the same distribution implies that the distribution of the measurement times \(t_{ij}\) must also be the same, as the measurement times are treated as random in this paper (Sect. 2.1 ). If, for example, we develop a landmark model out of a training dataset where every subject follows a six-month visit schedule with some random deviation, and apply it to a new dataset where every patient follows a three-month visit schedule, then it would not be appropriate to use the BLUP with a time-varying linear mixed model developed from the training data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, L., Luo, S., Hu, B. et al. Dynamic Prediction of Renal Failure Using Longitudinal Biomarkers in a Cohort Study of Chronic Kidney Disease . Stat Biosci 9, 357–378 (2017). https://doi.org/10.1007/s12561-016-9183-7

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12561-016-9183-7

Keywords

Navigation