Skip to main content
Log in

Subtleties in the interpretation of hazard contrasts

  • Published:
Lifetime Data Analysis Aims and scope Submit manuscript

Abstract

The hazard ratio is one of the most commonly reported measures of treatment effect in randomised trials, yet the source of much misinterpretation. This point was made clear by Hernán (Epidemiology (Cambridge, Mass) 21(1):13–15, 2010) in a commentary, which emphasised that the hazard ratio contrasts populations of treated and untreated individuals who survived a given period of time, populations that will typically fail to be comparable—even in a randomised trial—as a result of different pressures or intensities acting on different populations. The commentary has been very influential, but also a source of surprise and confusion. In this note, we aim to provide more insight into the subtle interpretation of hazard ratios and differences, by investigating in particular what can be learned about a treatment effect from the hazard ratio becoming 1 (or the hazard difference 0) after a certain period of time. We further define a hazard ratio that has a causal interpretation and study its relationship to the Cox hazard ratio, and we also define a causal hazard difference. These quantities are of theoretical interest only, however, since they rely on assumptions that cannot be empirically evaluated. Throughout, we will focus on the analysis of randomised experiments.

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

  • Aalen OO, Cook RJ, Røysland K (2015) Does cox analysis of a randomized survival study yield a causal treatment effect? Lifetime Data Anal 21(4):579–93

    Article  MathSciNet  Google Scholar 

  • Bartolucci F, Grilli L (2011) Modeling partial compliance through copulas in a principal stratification framework. J Am Stat Assoc 106(494):469–479

    Article  MathSciNet  Google Scholar 

  • Collett D (2015) Modelling survival data in medical research. CRC Press, Boca Raton

    Book  Google Scholar 

  • Frangakis CE, Rubin DB (2002) Principal stratification in causal inference. Biometrics 58(1):21–9

    Article  MathSciNet  Google Scholar 

  • Hernán M (2010) The hazards of hazard ratios. Epidemiology (Cambridge, Mass) 21(1):13–15

    Article  MathSciNet  Google Scholar 

  • Lederle FA, Kyriakides TC, Stroupe KT, Freischlag JA, Padberg FT, Matsumura JS, Huo Z, Johnson GR (2019) Open versus endovascular repair of abdominal aortic aneurysm. N Engl J Med 380(22):2126–2135 PMID: 31141634

    Article  Google Scholar 

  • Lin DY, Wei LJ, Ying Z (1993) Checking the Cox model with cumulative sums of martingale-based residuals. Biometrika 80:557–572

    Article  MathSciNet  Google Scholar 

  • Martinussen T, Scheike TH (2006) Dynamic regression models for survival data. Springer, New York

    MATH  Google Scholar 

  • Martinussen T, Vansteelandt S (2013) On collapsibility and confounding bias in cox and aalen regression models. Lifetime Data Anal 19(3):279–96

    Article  MathSciNet  Google Scholar 

  • Nelsen RB (2006) An introduction to copulas. Springer, New York

    MATH  Google Scholar 

  • Oakes D (1989) Bivariate survival models induced by frailties. J Am Stat Assoc 84(406):487–493

    Article  MathSciNet  Google Scholar 

  • Primrose JN, Fox RP, Palmer DH, Malik HZ, Prasad R, Mirza D, Anthony A, Corrie P, Falk S, Finch-Jones M, Wasan H, Ross P, Wall L, Wadsley J, Evans JTR, Stocken D, Praseedom R, Ma YT, Davidson B, Neoptolemos JP, Iveson T, Raftery J, Zhu S, Cunningham D, Garden OJ, Stubbs C, Valle JW, Bridgewater J, Primrose J, Fox R, Morement H, Chan O, Rees C, Ma Y, Hickish T, Falk S, Finch-Jones M, Pope I, Corrie P, Crosby T, Sothi S, Sharkland K, Adamson D, Wall L, Evans J, Dent J, Hombaiah U, Iwuji C, Anthoney A, Bridgewater J, Cunningham D, Gillmore R, Ross P, Slater S, Wasan H, Waters J, Valle J, Palmer D, Malik H, Neoptolemos J, Faluyi O, Sumpter K, Dernedde U, Maduhusudan S, Cogill G, Archer C, Iveson T, Wadsley J, Darby S, Peterson M, Mukhtar A, Thorpe J, Bateman A, Tsang D, Cummins S, Nolan L, Beaumont E, Prasad R, Mirza D, Stocken D, Praseedom R, Davidson B, Raftery J, Zhu S, Garden J, Stubbs C, Coxon F (2019) Capecitabine compared with observation in resected biliary tract cancer (bilcap): a randomised, controlled, multicentre, phase 3 study. Lancet Oncol 20(5):663–673

    Article  Google Scholar 

  • Rubin DB (2000) Causal inference without counterfactuals: comment. J Am Stat Assoc 95(450):435–438

    Google Scholar 

  • Stablein DM, Koutrouvelis IA (1985) A two-sample test sensitive to crossing hazards in uncensored and singly censored data. Biometrics 41(3):643–52

    Article  MathSciNet  Google Scholar 

  • Uno H, Claggett B, Tian L, Inoue E, Gallo P, Miyata T, Schrag D, Takeuchi M, Uyama Y, Zhao L, Skali H, Solomon S, Jacobus S, Hughes M, Packer M, Wei L-J (2014) Moving beyond the hazard ratio in quantifying the between-group difference in survival analysis. J Clin Oncol 32(22):2380–5

    Article  Google Scholar 

  • Vansteelandt S, Martinussen T, Tchetgen ET (2014) On adjustment for auxiliary covariates in additive hazard models for the analysis of randomized experiments. Biometrika 101(1):237–244

    Article  MathSciNet  Google Scholar 

  • White IR, Royston P (2009) Imputing missing covariate values for the cox model. Stat Med 28(15):1982–98

    Article  MathSciNet  Google Scholar 

  • Zhao L, Claggett B, Tian L, Uno H, Pfeffer MA, Solomon SD, Trippa L, Wei LJ (2016) On the restricted mean survival time curve in survival analysis. Biometrics 72(1):215–21

    Article  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Torben Martinussen.

Additional information

Publisher's Note

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

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (pdf 1132 KB)

Appendix A

Appendix A

1.1 A.1 Binary frailty variable

Let Z be binary, for instance \(P(Z=0.2)=0.2\) and \(P(Z=1.2)=0.8\), low and high risk groups, then \(E(Z)=1\), and similar results as in the Gamma-distribution case (Sect. 3.2) are obtained. In this binary case, the Laplace transform is

$$\begin{aligned} \phi _Z(u)=0.2e^{-0.2u}+0.8e^{-1.2u}. \end{aligned}$$

Therefore

$$\begin{aligned} \text{ HR}_Z(t)\frac{g_Z(e^{-\Lambda (t,a=1)})}{g_Z(e^{-\Lambda (t,a=0)})} =\frac{\lambda (t;a=1)}{\lambda (t;a=0)}, \end{aligned}$$

where

$$\begin{aligned} g_Z(u)=\{D\log {(\phi _Z)}\}\{\phi _Z^{-1}(u)\}, \end{aligned}$$
(17)

which is an increasing function. Hence, if we take \(\beta _1<0\) and \(\beta _2=0\), then again

$$\begin{aligned} \text{ HR}_Z(t)<1 \end{aligned}$$

for all t.

1.2 A.2 Frailty model arising by marginalization

The DGP (4) given by \(\lambda (t;A,Z)\) with \(\nu =\infty \) induces the marginal Cox model for the observed data where we do not get access to Z. It is tempting to interpret \(\text{ HR}_{Z}(t)\) as a causal hazard ratio, but this only holds under further untestable assumptions as shown in Sect. 4.1. Below, we formulate a more general DGP involving an additional frailty variable so that both the marginal Cox model and the model \(\lambda (t;A,Z)\) are correctly specified.

Rename Z to \(Z_1\). We then show that similarly we can also pick a DGP \(\lambda (t;A,Z_1,Z_2)\) so that it marginalizes to \(\lambda (t;A,Z_1)\) that further marginalizes to \(\lambda (t;A)\), the latter being the Cox model. For ease of calculations, let

$$\begin{aligned} \lambda (t;A,Z_1,Z_2)=Z_2\lambda ^*(t;A,Z_1) \end{aligned}$$

with \(Z_2\) being Gamma distributed with mean and variance equal to 1, and independent of \(Z_1\) and A. Similar calculations as those in Sect. 3.1 gives the following expression

$$\begin{aligned} \lambda (t;A,Z_1,Z_2)=Z_1Z_2\lambda _0(t)e^{\beta A} \exp {\bigl [\Lambda _0(t)e^{\beta A}+Z_1\{ \exp {\{\Lambda _0(t)e^{\beta A}}-1\}\bigr ]} \end{aligned}$$
(18)

Hence, if the DGP is governed by (18) then model (4), with \(\nu =\infty \), and the marginal Cox model, only conditioning on A, are also correctly specified.

1.3 A.3 Hazard differences

Assume (5) and (6). Then

$$\begin{aligned} \lim _{h\rightarrow 0}P(t\le T^1< t+h|T^0\ge t,T^1\ge t)=\psi (t) +\frac{E\{\omega (t,Z)e^{-2\int _0^t\omega (s,Z)\, ds}\}}{E\{e^{-2\int _0^t\omega (s,Z)\, ds}\}} \end{aligned}$$

and

$$\begin{aligned} \lim _{h\rightarrow 0}P(t\le T^0< t+h|T^0\ge t,T^1\ge t) =\frac{E\{\omega (t,Z)e^{-2\int _0^t\omega (s,Z)\, ds}\}}{E\{e^{-2\int _0^t\omega (s,Z)\, ds}\}} \end{aligned}$$

and therefore

$$\begin{aligned} \psi (t)= & {} \lim _{h\rightarrow 0}P(t\le T^1< t+h|T^0\ge t,T^1\ge t)\\&-\lim _{h\rightarrow 0}P(t\le T^0< t+h|T^0\ge t,T^1\ge t). \end{aligned}$$

1.4 A.4 \(\text{ HR }(t)\) under copulas not obeying (11)

We outline now how to generate samples from the copula \(C^{*}\) to gain further insight, Nelsen (2006), pp. 40–41,

  1. 1.

    Generate a standard uniformly distributed random variable u;

  2. 2.

    Set

    $$\begin{aligned} v = \left\{ \begin{array}{ll} u&{} \quad \text {if}\ u\le \xi { or}u\ge 1-\xi \\ 1-u &{} \quad \text {if}\ \xi<u<1-\xi \end{array}\right. \end{aligned}$$
  3. 3.

    Set \(T^0=\Lambda _0^{-1}\{-\log {(u)}\}\) and \(T^1=\Lambda _0^{-1}\{-\log {(v)}/\phi \}\) (assuming \(\Lambda _0\) is strictly increasing).

As an illustration, we took \(\Lambda _0(t)=t\), \(\phi =2/3\) and \(\xi =0.251\) and generated \(n=2,000,000\) samples. For these simulated data we obtained an estimated Cox HR of 0.666 and Spearman’s \(\rho \) was estimated to equal 0.753, which fits nicely with the theoretical values. As mentioned, if we follow the path \((\exp {\{-\Lambda _0(t)\}},\exp {\{-\phi \Lambda _0(t)\}})\) (starting at (1,1) when \(t=0\)) in \([0,1]^2\) in such a way that we pass through the upper triangle of \(J_2\) then \(\text{ HR }(t)\) will be equal to 0 before we reach \(J_2\) but then it jumps to \(k(u,v)\phi \), where \(k(u,v)>1\). To illustrate this, we estimated \(\text{ HR }(t)\) non-parametrically in time-points t until we cross the line \(v=1-u\) in \(J_2\). This was done by taking the observed frequency of \(T^1\)-failures in the time interval \([t,t+\delta ]\), \(\delta =0.01\), among those where both \(T^0\ge t\) and \(T^1\ge t\); and similarly for \(T^0\); and then taking the ratio of the two. This is seen in Fig. 3, where the theoretical values \( \phi \exp {\{-t(\phi -1) \}}\) are superimposed. The dashed line corresponds to \(\phi =2/3\). We see that the theoretical values fit nicely to the estimated ones. For this copula it can be shown that \(\text{ HR }(t)\) is always below 1, but as we have just seen it can also be larger than \(\phi \)! Hence, to have that \(\text{ HR }(t)<\phi \) we need more than just positive correlation between \(T^0\) and \(T^1\) as this example illustrates.

1.5 A.5 Monotonicity and \(\text{ HR }(t)\)

Assume that \(T^1\ge T^0\) a.s. and that the distribution of \((T^0,T^1)\) is absolutely continuous with density \(f(t_0,t_1)\) with respect to the Lebesgue measure. Monotonicity is attained if \(f(t_0,t_1)=0\) for \(t_0>t_1\). We now look at

$$\begin{aligned} z(h)&=P(T^0\ge t, t\le T^1\le t+h)\\&=P(T^0\ge t, t\le T^1\le t+h, T^0\le T^1)\\&=\int _t^{t+h}\int _t^{t_1}f(t_0,t_1)dt_0dt_1. \end{aligned}$$

Hence

$$\begin{aligned} \lim _{h\rightarrow 0}\frac{z(h)}{h}=\lim _{h\rightarrow 0}\frac{G(t+h) -G(t)}{h}=G^{\prime }(t)=\int _t^{t}f(t_0,t)dt_0=0, \end{aligned}$$

where

$$\begin{aligned} G(u)=\int _0^{u}\left\{ \int _t^{t_1}f(t_0,t_1)dt_0\right\} dt_1. \end{aligned}$$

But this means \(\lambda _{T^1}(t|T^0\ge t,T^1\ge t)=0\), and therefore \(\text{ HR }(t)=0\). To avoid having \(\lambda _{T^1}(t|T^0\ge t,T^1\ge t)\) equal to zero we need for instance \(T^0\) given \(T^1\) to have a discrete distribution, but this seems unappealing.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Martinussen, T., Vansteelandt, S. & Andersen, P.K. Subtleties in the interpretation of hazard contrasts. Lifetime Data Anal 26, 833–855 (2020). https://doi.org/10.1007/s10985-020-09501-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10985-020-09501-5

Keywords

Navigation