Skip to main content

Advertisement

Log in

Group-sequential logrank methods for trial designs using bivariate non-competing event-time outcomes

  • Published:
Lifetime Data Analysis Aims and scope Submit manuscript

Abstract

We discuss the multivariate (2L-variate) correlation structure and the asymptotic distribution for the group-sequential weighted logrank statistics formulated when monitoring two correlated event-time outcomes in clinical trials. The asymptotic distribution and the variance–covariance for the 2L-variate weighted logrank statistic are derived as available in various group-sequential trial designs. These methods are used to determine a group-sequential testing procedure based on calendar times or information fractions. We apply the theoretical results to a group-sequential method for monitoring a clinical trial with early stopping for efficacy when the trial is designed to evaluate the joint effect on two correlated event-time outcomes. We illustrate the method with application to a clinical trial and describe how to calculate the required sample sizes and numbers of events.

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.

Similar content being viewed by others

References

  • Andersen PK, Borgan Ø, Gill RD, Keiding N (1993) Statistical models based on counting processes. Springer, New York

    MATH  Google Scholar 

  • Andrei A-C, Murray S (2005) Simultaneous group sequential analysis of rank-based and weighted Kaplan–Meier tests for paired censored survival data. Biometrics 61:715–720

    MathSciNet  MATH  Google Scholar 

  • Asakura K, Hamasaki T, Sugimoto T, Hayashi K, Evans SR, Sozu T (2014) Sample size determination in group-sequential clinical trials with two co-primary endpoints. Stat Med 33:2897–2913

    MathSciNet  Google Scholar 

  • Clayton DG (1976) A model for association in bivariate life tablesand its application in epidemiological studies of familial tendencyin chronic disease. Biometrika 65:141–151

    Google Scholar 

  • Collett D (2003) Modelling survival data in medical research, 2nd edn. Chapman & Hall/CRC, Boca Raton

    MATH  Google Scholar 

  • Cook RJ, Farewell VT (1994) Guidelines for monitoring efficacy and toxicity responses in clinical trials. Biometrics 50:1146–1152

    MATH  Google Scholar 

  • Dabrowska DM (1988) Kaplan–Meier estimate on the plane. Ann Stat 16:1475–1489

    MathSciNet  MATH  Google Scholar 

  • Dmitrienko A, Tamhane AC, Bretz F (2009) Multiple testing problems in pharmaceutical statistics. Chapman & Hall/CRC, Boca Raton

    Google Scholar 

  • Fleming TR, Harrington DP (1991) Counting process and survival analysis. Wiley, New York

    MATH  Google Scholar 

  • Glimm E, Mauer W, Bretz F (2009) Hierarchical testing of multiple endpoints in group-sequential trials. Stat Med 29:219–228

    MathSciNet  Google Scholar 

  • Gombay E (2008) Weighted logrank statistics in sequential tests. Seq Anal 27:97–104

    MathSciNet  MATH  Google Scholar 

  • Gordon LKK, Lachin JM (1990) Implementation of group sequential logrank tests in a maximum duration trial. Biometrika 46:759–770

    Google Scholar 

  • Gu MG, Lai TL (1991) Weak convergence of time-sequential censored rank statistics with applications to sequential testing in clinical trials. Ann Stat 19:1403–1433

    MathSciNet  MATH  Google Scholar 

  • Halabi S (2012) Adjustment on the type I error rate for a clinical trial monitoring for both intermediate and primary endpoints. J Biom Biostat 7:15

    Google Scholar 

  • Hamasaki T, Asakura K, Evans SR, Sugimoto T, Sozu T (2015) Group-sequential strategies in clinical trials with multiple co-primary endpoints. Stat Biopharm Res 7:36–54

    Google Scholar 

  • Herbst RS, Redman MW, Kim ES, Semrad TJ, Bazhenova L, Masters G, Oettel K, Guaglianone P, Reynolds C, Karnad A, Arnold SM, Varella-Garcia M, Moon J, Mack PC, Blanke CD, Hirsch FR, Kelly K, Gandara DR (2018) Cetuximab plus carboplatin and paclitaxel with or without bevacizumab versus carboplatin and paclitaxel with or without bevacizumab in advanced NSCLC (SWOG S0819): a randomised, phase 3 study. Lancet Oncol 19:101–114

    Google Scholar 

  • Hougaard P (1986) A class of multivariate failure time distribution. Biometrika 73:671–678

    MathSciNet  MATH  Google Scholar 

  • Hsu L, Prentice RL (1996) On assessing the strength of dependency between failure time variables. Biometrika 83:491–506

    MathSciNet  MATH  Google Scholar 

  • Huang X, Strawderman RL (2006) A note on the Breslow survival estimator. J Nonparametr Stat 18:45–56

    MathSciNet  MATH  Google Scholar 

  • Hung HMJ, Wang SJ, O’Neill RT (2007) Statistical considerations for testing multiple endpoints in group sequential or adaptive clinical trials. J Biopharm Stat 17:1201–1210

    MathSciNet  Google Scholar 

  • Jacod J, Shiryaev AN (2003) Limit theorems for stochastic processes, 2nd edn. Springer, Berlin

    MATH  Google Scholar 

  • Jennison C, Turnbull BW (2000) Group sequential methods with applications to clinical trials. Chapman & Hall/CRC, Boca Raton

    MATH  Google Scholar 

  • Jung S-H (2008) Sample size calculation for the weighted rank statistics with paired survival data. Stat Med 27:3350–3365

    MathSciNet  Google Scholar 

  • Kosorok MR, Shi Y, DeMets DL (2004) Design and analysis of group-sequential clinical trials with multiple primary endpoints. Biometrics 60:134–145

    MathSciNet  MATH  Google Scholar 

  • Lai TL, Shih M-C (2004) Power, sample size and adaptation considerations in the design of group sequential clinical trials. Biometrika 91:507–528

    MathSciNet  MATH  Google Scholar 

  • Lan KKG, DeMets DL (1983) Discrete sequential boundaries for clinical trials. Biometrika 70:659–663

    MathSciNet  MATH  Google Scholar 

  • Lennox JL, Landovitz RJ, Ribaudo HJ, Ofotokun I, Na LH, Godfrey C, Kuritzkes DR, Sagar M, Brown TT, Cohn SE, McComsey GA, Aweeka F, Fichtenbaum CJ, Presti RM, Koletar SL, Haas DW, Patterson KB, Benson CA, Baugh BP, Leavitt RY, Rooney JF, Seekins D, Currier JS (2014) A phase III comparative study of the efficacy and tolerability of three non-nucleoside reverse transcriptase inhibitor-sparing antiretroviral regimens for treatment-naïve HIV-1-infected volunteers: a randomized, controlled trial. Ann Intern Med 161:461–471

    Google Scholar 

  • Lin DY (1991) Nonparametric sequential testing in clinical trials with incomplete multivariate observations. Biometrika 78:123–131

    MathSciNet  MATH  Google Scholar 

  • Lin DY, Shen L, Ying Z, Breslow NE (1996) Group sequential designs for monitoring survival probabilities. Biometrics 52:1033–1041

    MathSciNet  MATH  Google Scholar 

  • Murray S (2000) Nonparametric rank-based methods for group sequential monitoring of paired censored survival data. Biometrics 54:984–990

    MathSciNet  MATH  Google Scholar 

  • Nishiyama Y (2011) Statistical analysis by the theory of martingales. Kindaikagakusha, Tokyo (in Japanese)

    Google Scholar 

  • O’Brien PC, Fleming TR (1979) A multiple testing procedure for clinical trials. Biometrics 35:549–556

    Google Scholar 

  • Pocock ST (1977) Group sequential methods in the design and analysis of clinical trials. Biometrika 64:191–199

    Google Scholar 

  • Pocock ST, Geller NL, Tsiatis AA (1987) The analysis of multiple endpoints in clinical trials. Biometrics 43:487–498

    MathSciNet  Google Scholar 

  • Prentice RL, Cai J (1992) Covariance and survivor function estimation using censored multivariate failure time data. Biometrika 79:495–512

    MathSciNet  MATH  Google Scholar 

  • Rauch G, Schüler S, Wirths M, Stefan E, Kieser M (2016) Adaptive designs for two candidate primary time-to-event endpoints. Stat Biopharm Res 8:207–216

    Google Scholar 

  • Slud EV, Wei LJ (1982) Two-sample repeated significance tests based on the modified Wilcoxon statistic. J Am Stat Assoc 77:862–868

    MathSciNet  MATH  Google Scholar 

  • Sugimoto T, Sozu T, Hamasaki T, Evans SR (2013) A logrank test-based method for sizing clinical trials with two co-primary time-to-event endpoints. Biostatistics 14:409–421

    Google Scholar 

  • Sugimoto T, Hamasaki T, Sozu T, Evans SR (2017) Sizing clinical trials when comparing bivariate time-to-event outcomes. Stat Med 36:1363–1382

    MathSciNet  Google Scholar 

  • Tamhane AC, Mehta CR, Liu L (2010) Testing a primary and secondary endpoint in a group sequential design. Biometrics 66:1174–1184

    MathSciNet  MATH  Google Scholar 

  • Tamhane AC, Wu Y, Mehta C (2012) Adaptive extensions of a two-stage group sequential procedure for testing primary and secondary endpoints (I): unknown correlation between the endpoints. Stat Med 31:2027–2040

    MathSciNet  Google Scholar 

  • Tang DI, Gnecco C, Geller NL (1989) Design of group sequential clinical trials with multiple endpoints. J Am Stat Assoc 84:776–779

    Google Scholar 

  • Tsiatis AA (1982) Group sequential methods for survival analysis with staggered entry. In: Crowley J, Johnson RA (eds) Survival analysis. IMS lecture notes. Hayward, California, pp 257–268

    Google Scholar 

  • Tsiatis AA, Boucher H, Kim K (1995) Sequential methods for parametric survival models. Biometrika 82:165–173

    MathSciNet  MATH  Google Scholar 

  • Wei LJ, Lachin JM (1984) Two-sample asymptotically distribution-free tests for imcomplete multivariate observations. J Am Stat Assoc 79:653–661

    MATH  Google Scholar 

  • Wei LJ, Su JQ, Latin JM (1990) Interim analyses with repeated measurements in a sequential clinical trial. Biometrika 77:359–364

    MathSciNet  MATH  Google Scholar 

  • Wu J, Xiong X (2017) Group-sequential survival trial design and monitoring using the log-rank test. Stat Biopharm Res 9:35–43

    Google Scholar 

  • Yin G (2012) Clinical trial design: Bayesian and frequentist adaptive methods. Wiley, New York

    Google Scholar 

Download references

Acknowledgements

We thank one reviewer and the Associate Editor for their comments. Research reported in this publication was supported by JSPS KAKENHI Grant Numbers JP17K00054 and JP17K00069, the Project Promoting Clinical Trials for Development of New Drugs (18lk0201061h0002/18lk0201061h0202) from the Japan Agency for Medical Research and Development (AMED) and the National Institute of Allergy and Infectious Diseases of the National Institutes of Health under Award Number UM1AI068634. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tomoyuki Sugimoto.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix

Proof of Theorem 1

Let \({M}_{ik}^{({\ell })}(t)={N}_{ik}^{({\ell })}(t)-\int _{0}^{t}{Y}_{ik}^{({\ell })}(x)d{\Lambda }_{g_i k}(x)\) and \(\{{\mathscr {F}}_{k,t}^{({\ell })}:t\ge 0 \}\) be a standard filtration generated from the history through time t for the kth outcome and the \({\ell }\)th analysis (\({\mathscr {F}}_{k,t}^{({\ell })}\) is the smallest \(\sigma \)-algebra generated by \(\{{N}_{ik}^{({\ell })}(x),{N}_{ik}^{C({\ell })}(x): 0\le x\le t,i=1,\cdots ,n_{\ell }\}\), where \({N}_{ik}^{C({\ell })}(t)={\mathbb {1}}\{T_{ik}^{({\ell })}\le t,\)\({\Delta }_{ik}^{({\ell })}=0\}\) is a censoring counting process). As is well-known, \({M}_{ik}^{({\ell })}(t)\) has the \({\mathscr {F}}_{k,t}^{({\ell })}\)-martingale property. We discuss the asymptotic behavior using the decomposition of the weighted logrank process \( U_k^{({\ell })}(t)=\hat{m}_k^{({\ell })}(t)+n_{{\ell }}^{-\frac{1}{2}}\mathcal {M}_{k}^{({\ell })}(t) \) from the definition of \(U_k^{({\ell })}\), where

$$\begin{aligned}&\mathcal {M}_{k}^{({\ell })}(t) =\int _{0}^{t}\hat{H}_k^{({\ell })}(x)n_{{\ell }}^{\frac{1}{2}}{\textstyle \sum _{i=1}^{n_L}\mathrm{d}{\widetilde{M}}_{ik}^{({\ell })}(x)} =\int _{0}^{t}\hat{H}_k^{({\ell })}(x)n_{{\ell }}^{\frac{1}{2}}\mathrm{d}{\widetilde{M}}_{\bullet k}^{({\ell })}(x),\\&\mathrm{d}{\widetilde{M}}_{\bullet k}^{({\ell })}(x) = \frac{\mathrm{d}\overline{M}_{2k}^{({\ell })}(x)}{\overline{{Y}}_{2k}^{({\ell })}(x)}-\frac{\mathrm{d}\overline{M}_{1k}^{({\ell })}(x)}{\overline{{Y}}_{1k}^{({\ell })}(x)},\,\,\,\, \mathrm{d}\overline{M}_{jk}^{({\ell })}(x)=\textstyle \sum _{i=1}^{n_{\ell }}{\mathbb {1}}\{g_i=j\}\mathrm{d}{{M}}_{ik}^{({\ell })}(x), \\&\mathrm{d}{\widetilde{M}}_{ik}^{({\ell })}(x) ={\mathbb {1}}{\{i\le n_{{\ell }}\}} \left\{ \frac{{\mathbb {1}}{\{g_{i}=2\}}}{\overline{{Y}}_{2k}^{({\ell })}(x)}-\frac{{\mathbb {1}}{\{g_{i}=1\}}}{\overline{{Y}}_{1k}^{({\ell })}(x)} \right\} \mathrm{d}{{M}}_{ik}^{({\ell })}(x), \end{aligned}$$

and \(\mathcal {M}_{k}^{({\ell })}(t)\) is \({\mathscr {F}}_{k,t}^{({\ell })}\)-martingale because \(\hat{H}_k^{({\ell })}(t)\) is \({\mathscr {F}}_{k,t}^{({\ell })}\)-predictable.

Let \(\hat{\varvec{Z}}^*=(\hat{Z}^*_{1}({\tau }_1),\ldots ,\hat{Z}^*_1({\tau }_L),\hat{Z}^*_{2}({\tau }_1),\ldots ,\hat{Z}^*_{2}({\tau }_L))^\mathrm{T}\) and let \(\hat{Z}^*_k(\tau _{\ell })\) be \(\hat{Z}_k(\tau _{\ell })\) whose denominator is replaced by the limit version,

$$\begin{aligned} \hat{Z}^*_k(\tau _{\ell }) =n_{\ell }^{\frac{1}{2}}\displaystyle \frac{U_k^{({\ell })}(\tau _{\ell })}{\sqrt{{V}_{kk}^{0({\ell })}(\tau _{\ell })}} =n_{\ell }^{\frac{1}{2}}\hat{\mu }_{k{\ell }}+\xi _{k{\ell }}\mathcal {M}_{k}^{({\ell })}(\tau _{\ell }), \end{aligned}$$

where we write \(\xi _{k{\ell }}=1/\sqrt{{V}_{kk}^{0({\ell })}(\tau _{{\ell }})}\) for simplicity. The distribution of \(\hat{\varvec{Z}}-D_{\scriptstyle {\varvec{n}}}\hat{\varvec{\mu }}\) is asymptotically equivalent to

$$\begin{aligned} \hat{\varvec{Z}}^* - D_{\scriptstyle {\varvec{n}}}\hat{\varvec{\mu }} =\left( \xi _{11}\mathcal {M}_{1}^{(1)}(\tau _1),\ldots ,\xi _{1L}\mathcal {M}_{1}^{(L)}(\tau _L), \xi _{21}\mathcal {M}_{2}^{(1)}(\tau _1),\ldots ,\xi _{2L}\mathcal {M}_{2}^{(L)}(\tau _L) \right) ^\mathrm{T} \end{aligned}$$

because the dominated convergence theorem works by the convergence of \(\hat{V}_{kk}^{0({\ell })}(\tau _{\ell })\)\({\mathop {\rightarrow }\limits ^{P}} {V}_{kk}^{0({\ell })}(\tau _{\ell })\) uniformly on \({\ell }=1,\ldots ,L\) as \(n_L\ge \cdots \ge n_1\rightarrow \infty \). We find it necessary to study the covariance of \(\mathcal {M}_{k}^{({\ell })}\)’s for characterizing the distribution of \(\hat{\varvec{Z}}^*-D_{\scriptstyle {\varvec{n}}}\hat{\varvec{\mu }}\).

In the proof hereafter, it is sufficient to consider the case of \(L=2\). As a function related to the characteristic function of \({\mathcal {M}}_{k}^{({\ell })}(t)\), define

$$\begin{aligned} G_{k}^{({\ell })}(t)=\exp \left( {\varvec{i} z_{k{\ell }}{\mathcal {M}}_{k}^{({\ell })}(t)+{\textstyle \frac{z_{k{\ell }}^2}{2}}{\langle }{\mathcal {M}}_{k}^{({\ell })},{\mathcal {M}}_{k}^{({\ell })}{\rangle }(t)} \right) \end{aligned}$$

for a real non-zero \(z_{k{\ell }}\) and \(\varvec{i}=\sqrt{-1}\), where \({\langle }m_1,m_2{\rangle }\) denotes a predictable covariance process for two martingales \(m_1\) and \(m_2\). In this case we have

$$\begin{aligned} {\langle }{\mathcal {M}}_{k}^{({\ell })},{\mathcal {M}}_{k}^{({\ell }')}{\rangle }(t) = n_{{\ell }}^{\frac{1}{2}} n_{{\ell }'}^{\frac{1}{2}} \int _{0}^{t}\hat{H}_k^{({\ell })}(x)\hat{H}_k^{({\ell }')}(x) \left\{ \frac{\mathrm{d}{\Lambda }_{1k}(x)}{\overline{Y}_{1k}^{({\ell }\vee {\ell }')}(x)} +\frac{\mathrm{d}{\Lambda }_{2k}(x)}{\overline{Y}_{2k}^{({\ell }\vee {\ell }')}(x)} \right\} , \end{aligned}$$

following the standard martingale theory of survival analysis (see Fleming and Harrington (1991)). The consistency of \(\hat{S}^{({\ell })}_{jk}\), the Glivenko-Cantelli theorem, and Conditions 1 and 3 imply \(\sup _{0\le x\le \tau _{{\ell }}}\)\(|\hat{H}_k^{({\ell })}(x)-{H}_k^{({\ell })}(x)|{\mathop {\rightarrow }\limits ^{P}}0\) and

$$\begin{aligned} \sup _{0\le x\le \tau _{{\ell }}} \left| {\hat{H}_k^{({\ell })}(x)}\bigl /n_{{\ell }}^{-1}\bar{Y}_{jk}^{({\ell })}(x) -h_{jk}^{({\ell })}(x) \right| {\mathop {\rightarrow }\limits ^{P}}0\quad \mathrm{as}\quad n_{{\ell }}\rightarrow \infty , \end{aligned}$$
(9)

where

$$\begin{aligned} h_{jk}^{({\ell })}(x) =\frac{{H}_{k}^{({\ell })}(x)}{a_{j{\ell }} y_{jk}^{({\ell })}(x)} =W_k^{({\ell })}(x) \frac{a_{j'{\ell }} S_{j'k}(x_-)}{S_{\bullet k}^{({\ell })}(x_-)},\,\,\,\,j'=3-j, \end{aligned}$$

and note that \(0\le {H}_k^{({\ell })}(x)<\infty \) for \(x\in [0,\tau _{{\ell }}]\), \({H}_k^{({\ell })}(x)=0\) for \(\tau _{{\ell }}<x\) and \(0\le h_{jk}^{({\ell })}(x)<\infty \) for all x. The univariate asymptotic result provides \(E(\mathrm{e}^{\scriptstyle {\varvec{i}} z_{k{\ell }}{\mathcal {M}}_{k}^{({\ell })}(t)})\rightarrow \exp (-\frac{z_{k{\ell }}^2}{2}V_{kk}(t,t\mid \tau _{{\ell }},\tau _{{\ell }}))\) as \(n_{{\ell }}\rightarrow \infty \), which corresponds to the following convergences,

$$\begin{aligned} E(G_{k}^{({\ell })}(t))\rightarrow 1 \quad \mathrm{and}\quad {\langle }{\mathcal {M}}_{k}^{({\ell })},{\mathcal {M}}_{k}^{({\ell }')}{\rangle }(t){\mathop {\rightarrow }\limits ^{P}} V_{kk}(t,t\mid \tau _{{\ell }},\tau _{{\ell }'}) \end{aligned}$$

(Nishiyama 2011). For different \(k\ne k'\), it is difficult to show joint normality with correlation between \({\mathcal {M}}_{k}^{({\ell })}\) and \({\mathcal {M}}_{k'}^{({\ell })}\) with standard martingale theory of counting processes (Fleming and Harrington 1991; Andersen et al. 1993). However, we overcome the challenge applying Ito’s formula. The discrete Ito’s formula (Jacod and Shiryaev 2003; Huang and Strawderman 2006) provides the decomposition of \(G_{k}^{({\ell })}(t)\),

$$\begin{aligned} G_{k}^{({\ell })}(t)-1= & {} \displaystyle \sum _{j=1,2} \int _{0}^{t}G_{k}^{({\ell })}(x-) \widetilde{H}_{jk}^{\mathrm{a}({\ell })}(x) \mathrm{d}{\overline{M}}_{jk}^{({\ell })}(x) \nonumber \\&+\displaystyle \sum _{j=1,2} \int _{0}^{t}G_{k}^{({\ell })}(x-)\widetilde{H}_{jk}^{\underline{(}{\ell })}(x)\overline{Y}_{jk}^{({\ell })}(x)\mathrm{d}{{\Lambda }}_{jk}(x), \end{aligned}$$
(10)

where, with \(\varvec{i}_1=-\varvec{i}\) and \(\varvec{i}_2=\varvec{i}\),

$$\begin{aligned} \widetilde{H}_{jk}^{\mathrm{a}({\ell })}(x)= & {} \exp \left( \varvec{i}_j z_{k{\ell }} \textstyle \frac{\sqrt{n_{{\ell }}}\hat{H}_{k}^{({\ell })}(x)}{\overline{Y}_{jk}^{({\ell })}(x)}\right) -1,\\ \widetilde{H}_{jk}^{\mathrm {b}({\ell })}(x)= & {} \exp \left( \varvec{i}_j z_{k{\ell }} \textstyle \frac{\sqrt{n_{{\ell }}}\hat{H}_{k}^{({\ell })}(x)}{\overline{Y}_{jk}^{({\ell })}(x)}\right) -1-\varvec{i}_j z_{k{\ell }} \displaystyle \frac{\sqrt{n_{{\ell }}}\hat{H}_{k}^{({\ell })}(x)}{\overline{Y}_{jk}^{({\ell })}(x)}\\&+\frac{z_{k{\ell }}^2}{2}\left( \displaystyle \frac{\sqrt{n_{{\ell }}}\hat{H}_{k}^{({\ell })}(x)}{\overline{Y}_{jk}^{({\ell })}(x)}\right) ^2. \end{aligned}$$

The expectation of the right-hand side of (10) converges to zero as \(n_{{\ell }}\rightarrow \infty \), because

$$\begin{aligned} \begin{array}{l} E\left( \textstyle \int _{0}^{t}G_{k}^{({\ell })}(x-)\widetilde{H}_{jk}^{\mathrm{a}({\ell })}(x)\mathrm{d}{\overline{M}}_{jk}^{({\ell })}(x)\right) =0\\ \mathrm{and}\quad \textstyle E\left( \int _{0}^{t}G_{k}^{({\ell })}(x-)\widetilde{H}_{jk}^{\mathrm {b}({\ell })}(x)\overline{Y}_{jk}^{({\ell })}(x)\mathrm{d}{{\Lambda }}_{jk}(x)\right) \rightarrow 0\\ \end{array} \end{aligned}$$
(11)

by the martingale property of \(\overline{M}_{jk}^{({\ell })}\) and the Lindeberg condition, respectively. In fact, using the integrable martingale property of \(G_{k}^{({\ell })}(x_{-})\) and the well-known inequality

$$\begin{aligned} \left| \exp (\varvec{i}c)-1-\varvec{i}c+\textstyle \frac{1}{2}c^{2} \right| \le {\mathbb {1}}\{|c|\le \varepsilon \}|c|^{3}+{\mathbb {1}}\{|c|>\varepsilon \}|c|^{2} \end{aligned}$$

for any real c, the latter result of (11) is obtained as

$$\begin{aligned}&E\left( \textstyle \int _0^t \left| G_{k}^{({\ell })}(x_{-})\widetilde{H}_{jk}^{\mathrm {b}({\ell })}(x) \right| \bar{Y}_{jk}^{({\ell })}(x)\mathrm{d}{{\Lambda }}_{jk}(x) \right) \\&\quad \le \exp \left( \textstyle \frac{z_{k{\ell }}^{2}}{2} {\langle }{\mathcal {M}}_{k}^{({\ell })},{\mathcal {M}}_{k}^{({\ell })}{\rangle }(t) \right) \times \\&\qquad \Biggl \{ E\left( \textstyle \int _0^t |c_{jk{\ell }}(x)|^{3} {\mathbb {1}}\{|c_{jk{\ell }}(x)|\le \varepsilon \} \bar{Y}_{jk}^{({\ell })}(x) \mathrm{d}{\Lambda }_{jk}(x) \right) \\&\qquad + E\left( \textstyle \int _0^t {|c_{jk{\ell }}(x)|^{2}{\mathbb {1}}\{|c_{jk{\ell }}(x)|>\varepsilon \}\bar{Y}_{jk}^{({\ell })}(x)} \mathrm{d}{\Lambda }_{jk}(x) \right) \Biggr \} {\rightarrow } 0 \end{aligned}$$

as \(n_{{\ell }}\rightarrow \infty \), where \(\varepsilon \) is an arbitrary positive number, \(c_{jk{\ell }}(x)=z_{k{\ell }}\sqrt{n_{{\ell }}}\hat{H}_{k}^{({\ell })}(x)/\bar{Y}_{jk}^{({\ell })}(x)\) and we have \(\sqrt{n_{{\ell }}}c_{jk{\ell }}(x){\mathop {\rightarrow }\limits ^{P}} {z_{k{\ell }}h_{jk}^{({\ell })}(x)}\) uniformly on \((0,\tau _{{\ell }}]\) from (9). Hence, we have

$$\begin{aligned} E((G_{1}^{({\ell })}(t)-1)(G_{2}^{({\ell }')}(s)-1))\rightarrow E(G_{1}^{({\ell })}(t)G_{2}^{({\ell }')}(s))-1 \end{aligned}$$
(12)

as \(n_{{\ell }},n_{{\ell }'}\rightarrow \infty \) by the univariate results of \(E(G_{k}^{({\ell })}(t))\rightarrow 1\), while using the formula (10) we can also find

$$\begin{aligned}&E((G_{1}^{({\ell })}(t)-1)(G_{2}^{({\ell }')}(s)-1)) \rightarrow \nonumber \\&\quad \displaystyle \sum _{j=1,2} \displaystyle \int _{0}^{t}\int _{0}^{s} E\left( G_{1}^{({\ell })}(x_-)G_{2}^{({\ell }')}(y_-) \widetilde{H}_{j1}^{\mathrm{a}({\ell })}(x)\widetilde{H}_{j2}^{\mathrm{a}({\ell }')}(y) \mathrm{d}{\overline{M}}_{j1}^{({\ell })}(x)\mathrm{d}{\overline{M}}_{j2}^{({\ell }')}(y) \right) \qquad \end{aligned}$$
(13)

as \(n_{{\ell }},n_{{\ell }'}\rightarrow \infty \). Similarly to showing the latter result of (11), with asymptotic equality, we can replace the terms \(\mathrm{e}^{\scriptstyle {\varvec{i}}_{j}(c_{j1{\ell }}(x)+c_{j2{\ell }'}(y))}\) and \(\mathrm{e}^{\scriptstyle {\varvec{i}}_{j}c_{jk{\ell }}(\cdot )}\) included in (13) by

$$\begin{aligned} 1+\varvec{i}_{j}\{c_{j1{\ell }}(x)+c_{j2{\ell }'}(y)\}-\textstyle \frac{1}{2}\{c_{j1{\ell }}(x)+c_{j2{\ell }'}(y)\}^{2} \quad \mathrm{and}\quad 1+\varvec{i}_{j}c_{jk{\ell }}(\cdot )-\textstyle \frac{1}{2}c_{jk{\ell }}(\cdot )^{2}, \end{aligned}$$

respectively. In fact, we can show that

$$\begin{aligned} \tilde{H}_{j1}^{\mathrm{a}({\ell })}(x)\tilde{H}_{j2}^{\mathrm{a}({\ell }')}(y)= & {} \mathrm{e}^{\scriptstyle {\varvec{i}}_{j}(c_{j1{\ell }}(x)+c_{j2{\ell }'}(y))} -\mathrm{e}^{\scriptstyle {\varvec{i}}_{j} c_{j1{\ell }}(x)}-\mathrm{e}^{\scriptstyle {\varvec{i}}_{j} c_{j2{\ell }'}(y)}+1\\= & {} -c_{j1{\ell }}(x)c_{j2{\ell }'}(y)+o_{P}(1/\sqrt{n_{{\ell }}n_{{\ell }'}}) \end{aligned}$$

from the convergence result of \(\sqrt{n_{{\ell }}}c_{jk{\ell }}(x)\). Hence, we have

$$\begin{aligned} \sqrt{n_{{\ell }}n_{{\ell }'}} \tilde{H}_{j1}^{\mathrm{a}({\ell })}(x)\tilde{H}_{j2}^{\mathrm{a}({\ell }')}(y) {\mathop {\rightarrow }\limits ^{P}} -z_{1{\ell }}z_{2{\ell }'}{h_{j1}^{({\ell })}(x)h_{j2}^{({\ell }')}(y)} \end{aligned}$$
(14)

as \(n_{{\ell }},n_{{\ell }'}\rightarrow \infty \), so that we can apply this result to (13). Also, similar to Prentice and Cai (1992) and Sugimoto et al. (2013, 2017), we can show

$$\begin{aligned} \frac{1}{\hat{a}_{j{\ell }\wedge {\ell }'}n_{{\ell }\wedge {\ell }'}} E\left( {\iint } \mathrm{d}\overline{M}_{j1}^{({\ell })}(x) \mathrm{d}\overline{M}_{j2}^{({\ell }')}(y) \right)= & {} E\left( {\iint } {\mathrm{d}}M_{i1}^{({\ell })}(x) {\mathrm{d}}M_{i2}^{({\ell }')}(y) \mid g_i=j\right) \\= & {} {\iint } C_{{\ell }\wedge {\ell }'}(x\wedge y)A_{j}({\mathrm{d}}x,{\mathrm{d}}y). \end{aligned}$$

For simplicity, let \(\phi (t,s)=E(G_{1}^{({\ell })}(t)G_{2}^{({\ell }')}(s))\). From (12), (13), (14), \(\hat{\gamma }_{{\ell }}{\mathop {\rightarrow }\limits ^{P}}{\gamma }_{{\ell }}\), \(\hat{a}_{j{\ell }}{\mathop {\rightarrow }\limits ^{P}}{a}_{j{\ell }}\) (Conditions 1–2) and the dominated convergence theorem, we have the integral equation for \(\phi (t,s)\) under \(n_{{\ell }},n_{{\ell }'}\rightarrow \infty \),

$$\begin{aligned}&\phi (t,s)-1=-z_{1{\ell }}z_{2{\ell }'}\sqrt{\displaystyle \frac{\gamma _{{\ell }\wedge {\ell }'}}{\gamma _{{\ell }\vee {\ell }'}}}\nonumber \\&\quad \times \int _{0}^{t}\int _{0}^{s} \phi (x_-,y_-) \sum _{j=1}^{2} a_{j{\ell }\wedge {\ell }'} h_{j1}^{({\ell })}(x) h_{j2}^{({\ell }')}(y) C_{{\ell }\wedge {\ell }'}(x\wedge y)A_j({\mathrm{d}}x,{\mathrm{d}}y). \end{aligned}$$
(15)

Similarly to bivariate survival function (Dabrowska 1988), the two-dimensional Volterra integral equation

$$\begin{aligned} \phi (t,s)=1 +\textstyle \int _{0}^{t}\int _{0}^{s}\phi (x_-,y_-)b_{12}({\mathrm{d}}x,{\mathrm{d}}y) \quad \mathrm{with}\quad \phi (t,0)=\phi (0,s)=1 \end{aligned}$$

is solved as \(\phi (t,s)=\exp [\int _{0}^{t}\int _{0}^{s}\{b_{12}({\mathrm{d}}x,{\mathrm{d}}y)-b_{1}({\mathrm{d}}x,y)b_{2}(x,{\mathrm{d}}y)\}]\), where

$$\begin{aligned} b_{1}({\mathrm{d}}x,y)=\phi ({\mathrm{d}}x,y)/\phi (x_-,y_-) \quad \mathrm{and}\quad b_{2}(x,{\mathrm{d}}y)=\phi (x,{\mathrm{d}}y)/\phi (x_-,y_-). \end{aligned}$$

However, note that it is difficult to obtain \(b_{k}(x,y)\), \(k=1,2\) by directly differentiating (15) because of including the expectation of non-differentiable \(M_{i1}^{({\ell })}(x)\) and \(M_{i2}^{({\ell }')}(y)\). Alternatively, we can use the formula (10) again for the purpose, so that by the discussion similar to obtaining (15), as \(n_{{\ell }},n_{{\ell }'}\rightarrow \infty \), we have

$$\begin{aligned} \int \phi ({\mathrm{d}}x,y)= & {} \int \left\{ E\left( {\mathrm{d}}G_{1}^{({\ell })}(x){\mathrm{d}}G_{2}^{({\ell }')}(y_-)\right) +E\left( {\mathrm{d}}G_{1}^{({\ell })}(x)G_{2}^{({\ell }')}(y_-)\right) \right\} \\\rightarrow & {} \int \phi (x_-,y_-) E\left( \textstyle \sum _{j} \widetilde{H}_{j1}^{\mathrm{a}({\ell })}(x) \mathrm{d}{\overline{M}}_{j1}^{({\ell })}(x) \right) =0. \end{aligned}$$

This yields \(\iint b_{1}({\mathrm{d}}x,y)b_{2}(x,{\mathrm{d}}y)=0\). Hence, the solution of (15) is

$$\begin{aligned} \phi (t,s)= \exp \left( -z_{1{\ell }}z_{2{\ell }'} V_{12}(t,s\mid \tau _{{\ell }},\tau _{{\ell }'}) \right) . \end{aligned}$$

Therefore, if \(E(\iint {\mathrm{d}}{\overline{M}}_{j1}^{({\ell })}(x){\mathrm{d}}{\overline{M}}_{j2}^{({\ell }')}(y))\ne 0\), the correlation between the two martingales works, which results in \(E(G_{1}^{({\ell })}(t)G_{2}^{({\ell }')}(s))\ne 1\) but concludes

$$\begin{aligned} E(G_{1}^{({\ell })}(t)G_{2}^{({\ell }')}(s))\phi (t,s)^{-1}\rightarrow 1 \quad \mathrm{as}\quad n_L\ge \cdots \ge n_1\rightarrow \infty . \end{aligned}$$

In summary, these results provide that the characteristic function of marginal martingale vector \(({\mathcal {M}}_{k}^{({\ell })}(t),{\mathcal {M}}_{k'}^{({\ell }')}(s))^\mathrm{T}\) converges to that of bivariate normal distribution as

$$\begin{aligned}&E\left( \mathrm{e}^{ \scriptstyle {\varvec{i}}z_{k{\ell }}{\mathcal {M}}_{k}^{({\ell })}(t) +\scriptstyle {\varvec{i}}z_{k'{\ell }'}{\mathcal {M}}_{k'}^{({\ell }')}(s) }\right) \\&\rightarrow \exp \left( -{\textstyle \frac{1}{2}}z_{k{\ell }}^2 V_{kk}(t,s\mid \tau _{\ell },\tau _{{\ell }})\right. \\&\qquad \left. -z_{k{\ell }}z_{k'{\ell }'} V_{kk'}(t,s\mid \tau _{\ell },\tau _{{\ell }'}) -{\textstyle \frac{1}{2}}z_{k'{\ell }'}^2 V_{k'k'}(t,s\mid \tau _{{\ell }'},\tau _{{\ell }'}) \right) \\&=\left\{ \begin{array}{ll} \exp \left( -2 z_{k{\ell }}^{2} V_{kk}(t,s\mid \tau _{{\ell }},\tau _{{\ell }})\right) &{} \quad \mathrm{if}\quad k=k',{\ell }={\ell }', \\ \exp \left( -\frac{1}{2}\{z_{k{\ell }} V_{kk}(t,s \mid \tau _{{\ell }},\tau _{{\ell }})^{1/2}+z_{k{\ell }'}V_{kk}(t,s\mid \tau _{{\ell }'},\tau _{{\ell }'})^{1/2}\}^{2}\right) &{}\quad \mathrm{if}\quad k=k',{\ell }\ne {\ell }', \\ \hbox {same as the above form} &{}\quad \mathrm{otherwise}.\\ \end{array} \right. \end{aligned}$$

A replication of the similar discussion provides that \(({\mathcal {M}}_{1}^{(1)}(t),{\mathcal {M}}_{1}^{(2)}(t),{\mathcal {M}}_{2}^{(1)}(t),\)\({\mathcal {M}}_{2}^{(2)}(t))\) converges in distribution to a multivariate normal distribution with zero mean vector and covarince matrix

$$\begin{aligned} \left( \begin{array}{cccc} V_{11}(t,s\mid \tau _1,\tau _{1}), &{} &{} &{} \\ V_{11}(t,s\mid \tau _2,\tau _{1}), &{}\quad V_{11}(t,s\mid \tau _2,\tau _{2}), &{} &{} \\ V_{21}(t,s\mid \tau _1,\tau _{1}), &{}\quad V_{21}(t,s\mid \tau _1,\tau _{2}), &{}\quad V_{22}(t,s\mid \tau _1,\tau _{1}), &{}\\ V_{21}(t,s\mid \tau _2,\tau _{1}), &{}\quad V_{12}(t,s\mid \tau _2,\tau _{2}), &{}\quad V_{22}(t,s\mid \tau _2,\tau _{1}), &{}\quad V_{22}(t,s\mid \tau _2,\tau _{2}) \\ \end{array} \right) . \end{aligned}$$

These results lead imidiately to the convergence of \(\hat{\varvec{Z}}^* - D_{\scriptstyle {\varvec{n}}}\hat{\varvec{\mu }}\) in distibution to \({\varvec{Z}}^* - D_{\scriptstyle {\varvec{n}}}{\varvec{\mu }}\), as summarized in Theorem 1. \(\square \)

Some additional results

Table 2 of Sect. 4 displays the results obtained under the assumption of a late time-dependent association (Clayton copula) for the joint survival distribution of the two event-time outcomes. The users may be interested in how the results change if the other types of dependency between two outcomes are assumed. In Table 3, we provide results from the design stage calculated under the same assumptions as Table 2 except that the joint survival distribution is replaced by an early time-dependent association (Gumbel copula). The pattern of the results of MSS, MEN and AEN under Gumbel copula are quite similar to Table 2, but, as the correlation is higher, their reduction rates from the values at zero correlation are slightly larger than those under Clayton copula.

Table 3 Sample sizes, number of events, and empirical powers in a group-sequential trial with two co-primary outcomes under an early time-dependent association (Gumbel copula)

As indicated by one referee, an important matter of concern is how the Type I error rates are controlled or not. In fact, the proposed design method is based on asymptotic results. To answer such a problem, we evaluate the behavior of the actual Type I error rates under sample sizes calculated by the proposed methods. Using ARDENT study, we consider three settings of \((\psi _1,\psi _2)=\) (1.0, 1.0), (0.565, 1.0) and (1.0, 0.721) (both null hypotheses and the two marginals) under the same configurations as Sect. 4, and we confirm the behavior via Monte-Carlo simulation with 1,000,000 runs. For the simulation, a trial ended at the planned follow-up duration. When the observed numbers were larger than the planned ones, the critical value at the final analysis was recalculated based on

$$\begin{aligned} 1-P\left( Z_{k1}<c_{k1},\ldots ,Z_{kL}<\tilde{c}_{kL}\mid \mathrm{H}_{0k}\right) =\alpha _k, \end{aligned}$$

where \(\tilde{c}_{kL}\) is the critical value at the final analysis, recalculated such that the above equation is satisfied to control the Type I error adequately if the planned numbers are different from the observed ones.

Table 4 Simulation assessment: probability of rejecting null hypothesis under Clayton copula
Table 5 Simulation assessment: probability of rejecting null hypothesis under Gumbel copula
Table 6 Variance, calendar time and information fraction corresponding to the other endpoint’s information fraction

Tables 4 and 5 show the results of the actual Type I error rates, which are corresponding to the situations under null hypotheses of Tables 2 and 3 under Clayton and Gumbel copulas, respectively. Where the columns “Both” and “ALO” give the probabilities to reject two null hypotheses of OC1 and OC2 jointly (Both) and at least one (ALO), respectively, and “OC1” and “OC2” provide the probabilities to reject two single hypotheses of OC1 and OC2, respectively. We observe that the results of “Joint” are well controlled at the nominal error rate 2.5% in the three cases. Those of “ALO” are less than \(2\times 2.5\%\) only at both null hypotheses and reflect the effect of multiplicity using two times testing. Also, the results of “OC1” and “OC2’ are well controlled at the nominal Type I error rate in three cases. Therefore, our method works well in controlling the nominal Type I error rate under the calculated sample size.

Table 1 of Sect. 4 displays the planning information for a group-sequential design at the fixed analysis time points (48 and 96 weeks) considered in ARDENT trial. Other group-sequential designs based on selected information fractions can be constructed. Table 6 displays the planning information for a group-sequential design for information fractions of 0.5 and 1.0.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sugimoto, T., Hamasaki, T., Evans, S.R. et al. Group-sequential logrank methods for trial designs using bivariate non-competing event-time outcomes. Lifetime Data Anal 26, 266–291 (2020). https://doi.org/10.1007/s10985-019-09470-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10985-019-09470-4

Keywords

Navigation