477
Views
1
CrossRef citations to date
0
Altmetric
Articles

Semiparametric estimation for accelerated failure time mixture cure model allowing non-curable competing risk

, &
Pages 97-108 | Received 16 Jan 2019, Accepted 24 Mar 2019, Published online: 11 Apr 2019

Abstract

The mixture cure model is the most popular model used to analyse the major event with a potential cure fraction. But in the real world there may exist a potential risk from other non-curable competing events. In this paper, we study the accelerated failure time model with mixture cure model via kernel-based nonparametric maximum likelihood estimation allowing non-curable competing risk. An EM algorithm is developed to calculate the estimates for both the regression parameters and the unknown error densities, in which a kernel-smoothed conditional profile likelihood is maximised in the M-step, and the resulting estimates are consistent. Its performance is demonstrated through comprehensive simulation studies. Finally, the proposed method is applied to the colorectal clinical trial data.

1. Introduction

In some medical studies, like prostate cancer or breast cancer, it is often observed that a substantial proportion of study patients never experience the event of interest, which are treated as cured or non-susceptible subjects. A number of mixture cure models have been proposed in the literature for analysing such survival data. The two-component mixture cure model can capture cured and uncured patients simultaneously. The early approach of univariate mixture cure model was used to model the survival data by Boag (Citation1949), which assumed that a fraction of the patients are cured from cancer and will never experience the event. Since then, parametric and semiparametric mixture cure models have been proposed (Li & Taylor, Citation2002; Peng & Dear, Citation2000; Peng, Dear, Denham, Citation1998; Sy & Taylor, Citation2000). The proportional hazards (PHMC) model was further discussed by Sy Taylor (Citation2000) and Peng (Citation2003). Besides, Li Taylor (Citation2002), Zhang Peng (Citation2007), Xu Zhang (Citation2009) and Lu (Citation2010) developed the accelerated failure time (AFTMC) model.

Although PHMC model specifies that the effects of the covariates act multipicatively on the hazard function, the AFTMC model regresses the logarithm of the failure time over the covariates, assuming a direct relationship between failure time over the covariates. Li and Taylor (Citation2002) and Zhang and Peng (Citation2007) estimated parameters through EM for the AFTMC model with semiparametric methods, and the large sample properties was discussed in Xu and Zhang (Citation2009). Lu (Citation2010) developed the AFTMC model and the error density was estimated by the kernel method.

The competing risk model is the main tool in explaining more than one event in survival analysis (Crowder, Citation2001; David & Moeschberge, Citation1978; Kalbfleisch & Prentice, Citation2002; Kleinbaum & Klein, Citation2006). There two common approaches for competing risk data, one is the cumulative incidence function (CIF) (Fine & Gray, Citation1999; Kalbfleisch & Prentice, Citation2002; Tai, Machin, White, & Gebski, Citation2001), the other one is cause-specified hazard or subhazard (Fusaro, Bacchetti, & Jewell, Citation1996; Gaynor et al., Citation1993; Klein, Citation2006; Ohneberg & Schumacher, Citation2017; Pintilie, Citation2007). Usually, the marginal hazards equals to the cause-specified hazards when the independence among events of interest are assumed (Gamel, Weller, Wesley, & Feuer, Citation2000; Kuk, Citation1992; Larson & Dinse, Citation1985; Ng & McLachlan, Citation2003). In mixture model, the marginal hazards are the hazards of single risk (Lambert, Thompson, Weston, & Dickman, Citation2006; Yu & Tiwari, Citation2007; Yu, Tiwari, Cronin, & Feuer, Citation2004).

Due to the scarcity of efficient and reliable computational methods, there is little literature dealing with AFTMC model and competing risk data. In this paper, we assume all patients may experience either event 1 or event 2, where event 1 is the primary event with a potential cure rate and event 2 is all other possible events of interest, which are non-curable. Different from the standard competing risk model, there is a cured fraction existed in event 1. For the patients being cured from event 1, they will only experience event 2; while for the patients being uncured from event 1, they will experience either event 1 or event 2. The rest of the paper is organised as follows. In Section 2, the accelerated failure time mixture cure model allowing non-curable competing risk is developed. An EM algorithm is developed to implement the estimation in Section 3, we maximise a kernel-smoothed conditional profile likelihood in the M-step, which is motivated by Zeng and Lin (Citation2007) in efficient estimation for the accelerated failure time model without cure fraction. We show that the estimates are consistent and efficient in the appendix. In addition, the proposed model and method are evaluated by the comprehensive simulation study in Section 4 and illustrated by the real data analysis in Section 5. Conclusions are made in Section 6.

2. Accelerated failure time mixture cure model allowing non-curable competing risk

Consider a study with n subjects. For subject i, let T1,i denote the event time undergoing the risk of primary interest and T2,i the event time under other risks. In addition, let Xi denote the covariates and Ci the censoring time for subject i. εi is used as event indicator (εi=1 for event of interest; εi=2 for other risk), and define Ti=Tj,i when εi=j,j=1,2, T~i=min(Ti,Ci), δi=I(TiCi). The observed data then consist of O={(t~i,δi,εi,Xi),i=1,,n}. Yi is the indicator of the uncured patient i subject to the event of primary interest. That is, Yi=1 means that subject i is subjected to the event of primary interest with uncured probability P(Yi=1). Note that Yi is not directly observable. As usual, we assume that Ci is given or independent of (T1,i,T2,i,Yi) given Xi.

As commonly done in the mixture cure models, we assume a logistic regression for Yi, i.e. πα(Xi)P(Yi=1|Xi)=exp(αX~i)1+exp(αX~i), where X~i=(1,Xi).

To model the competing risk data, we adopt a similar mixture regression approach as studied in Lu and Peng (Citation2008). Note that when Yi=0, εi2. When Yi=1, we assume a logistic regression model πθ(Xi)P(εi=1|Xi,Yi=1)=exp(θX~i)1+exp(θX~i). Then, we have P(εi=1|Xi)=πα(Xi)πθ(Xi) and P(εi=2|Xi)=1πα(Xi)+πα(Xi){1πθ(Xi)}. Here, we use a logistic link for both πα and πθ, but other link functions, such as the log-log and probit links, can also be used to model πα and πθ.

In addition, we consider a accelerated failure time regression model for Ti given Xi and ϵj,i=j, i.e. (1) logTj,i=βjTXi+ϵj,i,j=1,2,(1) (2) fj(t|Xi)=exp(βjTXi)fϵjelog(t)βjTXi or Sj(t|Xi)=Sϵjelog(t)βjTXi, j=1,2,(2) where βj is a row vector of unknown parameters, ϵj,i denotes the error term for the event j with an unknown survival function, fϵj and Sϵj are, respectively, the density and survival function of eϵj,i.

Define the cumulative incidence functions (CIFs) as (3) Fj(t,Xi)=P(Tit,εi=j|Xi),j=1,2.(3) Then, we have F1(t,Xi)=πα(Xi)πθ(Xi){1S1(t|Xi)} and F2(t,Xi)=[1πα(Xi)πθ(Xi)]{1S2(t|Xi)}. Moreover, (4) S(t|Xi)P(Ti>t|Xi)=j=12P(Ti>t|εi=j,Xi)P(εi=j|Xi)={1πα(Xi)}S2(t|Xi)+πα(Xi)πθ(Xi)S1(t|Xi)+{1πθ(Xi)}S2(t|Xi)=πα(Xi)πθ(Xi)S1(t|Xi)+1πα(Xi)πθ(Xi)S2(t|Xi).(4) We see that it is just a mixture model with weights πα(Xi)πθ(Xi) and 1πα(Xi)πθ(Xi) on the model of event 1 (primary) and event 2 (others). When only one risk was considered, i.e. πθ(Xi)=1 and S2(t|Xi)=1, it becomes the simple mixture cure model S(t|Xi)=1πα(Xi)+πα(Xi)S1(t|Xi).

Define Θ=(α,θ,β1,β2,fϵ1(),fϵ2()). With known εi and δi, i=1,2,,n, the likelihood function can be written as (5) Lo=i=1nπα(Xi)πθ(Xi)exp(β1TXi)fϵ1(eRi(β1))I(εi=1)δi×exp(β2TXi)fϵ2(eRi(β2))[1πα(Xi)+πα(Xi){1πθ(Xi)}]exp(β2TXi)fϵ2(eRi(β2))I(εi=2)δi{S(t~i|Xi)}1δi,(5) where Ri(βj)=log(t~i)βjTXi,j=1,2. Direct maximisation of the above observed likelihood function is very challenging due to its complex structure. In the next Section, we derive an EM algorithm to maximise the complete likelihood function.

3. EM algorithm

Similar to the standard mixture cure mode, if Yi could be observed, then the complete likelihood function is written as (6) Lc=i=1nexp(β1TXi)fϵ1(eRi(β1))πα(Xi)πθ(Xi)exp(β1TXi)fϵ1(eRi(β1))I(εi=1)δi×exp(β1TXi)fϵ1(eRi(β1))1παXi+παXi1πθXiexp(β2TXi)fϵ2(eRi(β2))I(εi=2)δi×παXiπθXiSϵ1(eRi(β1))I(εi=1)(1δi)×exp(β1TXi)fϵ1(eRi(β1))1παXi+παXi1πθXiSϵ2(eRi(β2))I(εi=2)(1δi)=i=1nexp(β1TXi)fϵ1(eRi(β1))πα(Xi)πθ(Xi)exp(β1TXi)fϵ1eRi(β1)I(εi=1)δi×exp(β1TXi)fϵ1(eRi(β1))1παXi1yiπαXi1πθXiyiexp(β2TXi)fϵ2eRi(β2)}I(εi=2)δi×παXiπθXiSϵ1eRi(β1)I(εi=1)(1δi)×exp(β1TXi)fϵ1(eRi(β1))1παXi1yiπαXi1πθXiyiSϵ2eRi(β2)I(εi=2)(1δi).(6) It is easy to write the logarithm of the complete likelihood function into four components regarding to their unknown parameters. lc(α,θ,β1,β2|O,y)=lc1(α|O,y)+lc2(θ|O,y)+lc3(β1|O,y)+lc4(β2|O,y), where (7) lc1(α|O,y)=i=1n[I(εi=1)+I(εi=2)yi]logπα(Xi)+i=1nI(εi=2)(1yi)log1πα(Xi),(7) (8) lc2(θ|O,y)=i=1nI(εi=1)logπθ(Xi)+i=1nI(εi=2)yilog1πθ(Xi),(8) (9) lc3(β1|O,y)=i=1nI(εi=1)δiloghϵ1eRi(β1)β1TXiI(εi=1)Hϵ1eRi(β1),(9) (10) lc4(β2|O,y)=i=1nI(εi=2)δiloghϵ2eRi(β2)β2TXiI(εi=2)Hϵ2eRi(β2),(10) and fϵj=hϵjSϵj, here hϵj and Hϵj are the hazard function and cumulative functions of eϵj, j=1,2 respectively.

Take Yi as auxiliary variables, and use EM algorithm to find the maximum likelihood estimates of the unknown parameters. The E-step in the EM algorithm computes the conditional expectation of the complete log-likelihood function with respect to three unobserved probabilities P(Yi=1,εi=1|O,Θ(m)), the probability that uncured patients die from the events of primary interest; P(Yi=1,εi=2|O,Θ(m)), the probability that uncured patients die from other risks; and P(Yi=0,εi=2|O,Θ(m)), the probability that cured patients die from other risks, respectively, where m indicates the m-th step in the EM algorithm. These three probabilities sum to 1 and can, respectively, be given by (11) P(Yi=1,εi=1|O,Θ(m))=δI(εi=1)P(Yi=1|εi=1,δi=1)Θ(m)+(1δi)P(Yi=1,εi=1|δi=0)Θ(m)=δI(εi=1)+(1δi)πα(Xi)πθ(Xi)S1(t~i|Xi)S(t~i|Xi)Θ(m),(11) (12) P(Yi=1,εi=2|O,Θ(m))=δI(εi=2)P(Yi=1|εi=2,δi=1)Θ(m)+(1δi)P(Yi=1,εi=2|δi=0)Θ(m)=δI(εi=2)πα(Xi)[1πθ(Xi)]1πα(Xi)πθ(Xi)Θ(m)+(1δi)πα(Xi)1πθ(Xi)S2(t~i|Xi)S(t~i|Xi)Θ(m),(12) (13) P(Yi=0,εi=2|O,Θ(m))=δI(εi=2)P(Yi=0|εi=2,δi=1)Θ(m)+(1δi)P(Yi=0,εi=2|δi=0)Θ(m)=δI(εi=2)1πα(Xi)1πα(Xi)πθ(Xi)Θ(m)+(1δi)1πα(Xi)S2(t~i|Xi)S(t~i|Xi)Θ(m).(13)

Let pˆ11,i(m)=P(Yi=1,εi=1|O,Θ(m)) and pˆε,1i(m)=P(εi=1|O,Θ(m)), then pˆε,1i(m)=pˆ11,i(m). Let pˆ12,i(m)=P(Yi=1,εi=2|O,Θ(m)) and pˆ02,i(m)=P(Yi=0,εi=2|O,Θ(m)) and pˆε,2i(m)=P(εi=2|O,Θ(m)), then pˆε,2i(m)=pˆ12,i(m)+pˆ02,i(m). The expectations of (Equation7), (Equation8), (Equation9) and (Equation10) can be written as (14) E(lc1)=i=1n([pˆε,1i(m)+pˆ12,i(m)]log[πα(Xi)]+pˆ02,i(m)log[1πα(Xi)]),(14) (15) E(lc2)=i=1n(pˆε,1i(m)log[πθ(Xi)]+pˆ12,i(m)log[1πθ(Xi)]),(15) (16) E(lc3)=i=1nI(εi=1)δi[loghϵ1eRi(β1)β1TXi]+pˆε,1i(m)Hϵ1eRi(β1),(16) (17) E(lc4)=i=1nI(εi=2)δi[loghϵ2eRi(β2)β2TXi]+pˆε,2i(m)Hϵ2eRi(β2).(17) The M-step in the EM algorithm is to maximise (Equation14), (Equation15), (Equation16) and (Equation17) with respect to the unknown parameters α, θ, β1, β2, hϵ1 and hϵ2. Maximising (Equation14) and (Equation15) with respect to α and θ can be easily carried out using the Newton–Raphson algorithm, and maximising lc3 and lc4 in (Equation16) and (Equation17)with respect to β1, β2, hϵ1 and hϵ2 can be carried out utilising the approach proposed by Zeng and Lin (Citation2007).

To find a smooth estimator for Hϵj, or hϵj, for simplify, we take Hϵ1, or hϵ1 as example, Hϵ2 and hϵ2 are similar. We start with the simplest case of a piecewise constant hϵ1. To be specific, we partition an interval containing all elogt~iβ1Txi into Jn equally spaced intervals, 0t0<t1<<tJnM, where M denotes an upper bound for the elogt~iβ1Txi over all possible β1's in a bounded set. A piecewise constant hϵ1 takes the form (18) hϵ1(t)=k=1JnckIt~i[tk1,tk).(18) Then, for any t, (19) Hϵ1(t)=k=1Jnck(t~itk)Itk1t~i<tk+MJnk=1JnckIt~itk.(19) As log{k=1JnckI(elogt~iβ1Txi[tk1,tk))}=k=1JnlogckI(elogt~iβ1Txi[tk1,tk)), then (Equation9) can be rewritten as (20) i=1nδiI(εi=1)β1TXi+k=1Jnlogcki=1nδiI(εi=1)IeRi(β1)[tk1,tk)k=1Jncki=1npε,1i(m)(eRi(β1)tk)Itk1eRi(β1)<tk+MJni=1npε,1i(m)IeRi(β1)tk.(20) By differentiating with respect to ck, we see that the solution to the score equation of ck is (21) ck=i=1nδiI(εi=1)IeRi(β1)[tk1,tk)i=1npε,1i(m)(eRi(β1)tk)Itk1eRi(β1)<tk+MJni=1npε,1i(m)IeRi(β1)tk.(21) After plugging the equations for the ck into (Equation20) and discarding the irrelevant component, we obtain the following sieve profile function: (22) lc3p(β1)=i=1nδiI(εi=1)β1TXi+k=1Jnlogi=1nδiI(εi=1)IeRi(β1)[tk1,tk)i=1npε,1i(m)(eRi(β1)tk)Itk1eRi(β1)<tk+MJni=1npε,1i(m)IeRi(β1)tk×i=1nδiI(εi=1)IeRi(β1)[tk1,tk)k=1Jni=1nδiI(εi=1)IeRi(β1)[tk1,tk)i=1npε,1i(m)(eRi(β1)tk)Itk1eRi(β1)<tk+MJni=1npε,1i(m)IeRi(β1)tk×i=1npε,1i(m)(eRi(β1)tk)Itk1eRi(β1)<tk+MJni=1npε,1i(m)IeRi(β1)tk=i=1nδiI(εi=1)β1TXi+k=1Jni=1nδiI(εi=1)IeRi(β1)[tk1,tk)×logJnnMi=1nδiI(εi=1)IeRi(β1)[tk1,tk)k=1Jni=1nδiI(εi=1)IeRi(β1)[tk1,tk)×logJnnMi=1npε,1i(m)(eRi(β1)tk)Itk1eRi(β1)<tk+1ni=1npε,1i(m)IeRi(β1)tk.(22) Note lc3p(β1) is not smooth and may have multiple local maxima. Following Zeng and Lin (Citation2007), we further seek a smooth approximation of lc3p(β1) by the empirical measure. The kernel smoothed approximation of lc3p(β1) is (23) lsc3(β1)=i=1nδiI(εi=1)log1nanj=1nδiI(εj=1)KRj(β1)Ri(β1)ani=1nδiI(εi=1)log1nj=1npε,1i(m)Rj(β1)Ri(β1)/anKsds.(23) where K() is the kernel function and an is the bandwidth. The detail of this derivation can be found in the appendix. The selection of the kernel function and bandwidth can be found in Zeng and Lin (Citation2007). We propose to maximise lsc3(β1) over β1 and denote the resulting estimator as β1ˆ. Because K() is a smooth kernel function, we can use the Newton–Raphson algorithm or other gradient-based search algorithms to calculate β1ˆ. Given β1ˆ, we estimate hϵ1(t) by the following kernel-smoothed estimator: (24) hˆϵ1(t)=1nanti=1nδiI(εi=1)KRi(β1ˆ)logtan1ni=1npε,1i(m)Ri(β1ˆ)logt/anK(μ)dμ.(24) The corresponding estimator of Hϵ1(t) is (25) Hˆϵ1(t)=logt1nani=1nδiI(εi=1)KRi(β1ˆ)san1ni=1npε,1i(m)Ri(β1ˆ)logt/anK(μ)dμds.(25) Similar to hϵ1 and Hϵ1, the estimator of hϵ2 and Hϵ2 are (26) hˆϵ2(t)=1nanti=1nδiI(εi=2)KRi(β2ˆ)logtan1ni=1npε,2i(m)Ri(β2ˆ)logt/anK(μ)dμ,(26) and (27) Hˆϵ2(t)=logt1nani=1nδiI(εi=2)KRi(β2ˆ)san1ni=1npε,2i(m)Ri(β2ˆ)logt/anK(μ)dμds.(27) The M-step in the EM algorithm is to maximise (Equation14), (Equation15), (Equation16) and (Equation17) with respect to the unknown parameters Θ=(α,θ,β1,β2,Hϵ1(),Hϵ2()), which can be easily estimated by the Newton–Raphson method using ‘optim’ function in R.

The EM algorithm is described as follows:

Step 0: Given initial values of α(0), θ(0), β1(0), β2(0) and pˆε,ji(0)=1. hϵ1(0)(t~i|Xi) and hϵ2(0)(t~i|Xi) can be estimated by (Equation24) and (Equation26) based on the value of β1(0) and β2(0).

Step 1: Update α(m+1) and θ(m+1) by maximising (Equation14) and (Equation15). Update β1(m+1) and β2(m+1) by maxi-mising (Equation23).

Step 2: Update hϵ1(m+1)() and hϵ2(m+1)() via (Equation24) and (Equation26).

Step 3: In the (m+1)th iteration, calculate pˆε,1i(m), pˆ12,i(m), pˆ02,i(m) based on α(m), θ(m), β1(m), β2(m), hϵ1(m)(t~i|Xi) and hϵ2(m)(t~i|Xi) from (Equation11), (Equation12), (Equation13).

Step 4: Repeat Steps 1 and 3 until convergence attained. For convergence, we use the criterion, max{(α(m)α(m1))2,(θ(m)θ(m1))2,(β1(m)β1(m1))2,(β2mβ2m1)2}<0.0001.

The algorithm always converges in our simulation.

4. Simulation

We conducted numerous simulation studies to examine the proposed inference procedures. We generated failure times from the following model: (28) logTj=βj,iX1+βj,iX2+ϵj,i,j=1,2,(28) where X1 is standard normal distribution, and X2 is Bernoulli with 0.5 success probability. We considered three distribution of eϵj,i: Weibull distribution with different shape parameters, denoted by Weibull(0.5,1) and Weibull(0.1,1); lognormal distribution with different parameters, denoted by lognormal(0,1) and lognormal(1,1); Weibull distribution for the event of interest and lognormal distribution for the other event, which is Weibull(0.5,1) and lognormal (1,1). β1=(1,1) for the event of interest for uncured patient and β2=(2,2) for other risk event. We assume α=(2,1,1) in the uncure component πα, θ=(0.5,0.5,0.5) for the component in the competing risk πθ, We generated censoring times from the uniform[0,τ] distribution, where τ was chosen to produce a 25% censoring rate. We set n to 100, 500 and 1000.

Following Zeng and Lin (Citation2007), we choose the kernel function K() to be the standard normal density for convenience and tractability. We used the optimal bandwidths 41/3σn1/3, where σ is the sample standard deviations of logT.

Table  reports the biases, mean squared errors (MSE), empirical standard errors (SE), average estimated standard deviations (ESD) with bootstrap sample size 500 and 95% coverage probabilities (CP). From Table , we can see that all biases are relatively small, the SEs and ESDs are close to each other, and the 95% CPs are close to their nominal levels. With sample size increases, the biases, MSE and SE become smaller.

Table 1. Estimates of parameters for three baseline survivals.

Figures –  show the estimates of CIFs of the proposed model along with the true CIFs. We compare the CIFs of the event of primary interest (F1) and other risks (F2) with their 95% pointwise confidence intervals (CI). The estimated CIFs are close to the true values, and it becomes closer as sample sizes increase. All these evidences show that the proposed method performs well, even for small and medium size samples.

Figure 1. Estimated CIFs with Weibull distributions (solid lines), their 95% pointwise confidence intervals (dashed and dotted lines), and the true CIFs (dotdash lines). (a) n=100,F1. (b) n=100,F2. (c) n=500,F1. (d) n=500,F2. (e) n=1000,F1. (f) n=1000,F2.

Figure 1. Estimated CIFs with Weibull distributions (solid lines), their 95% pointwise confidence intervals (dashed and dotted lines), and the true CIFs (dotdash lines). (a) n=100,F1. (b) n=100,F2. (c) n=500,F1. (d) n=500,F2. (e) n=1000,F1. (f) n=1000,F2.

Figure 2. Estimated CIFs with lognormal distributions (solid lines), their 95% pointwise confidence intervals (dashed and dotted lines), and the true CIFs dotdash lines). (a) n=100,F1. (b) n=100,F2. (c) n=500,F1. (d) n=500,F2. (e) n=1000,F1. (f) n=1000,F2.

Figure 2. Estimated CIFs with lognormal distributions (solid lines), their 95% pointwise confidence intervals (dashed and dotted lines), and the true CIFs dotdash lines). (a) n=100,F1. (b) n=100,F2. (c) n=500,F1. (d) n=500,F2. (e) n=1000,F1. (f) n=1000,F2.

Figure 3. Estimated CIFs with Weibull and lognormal distributions (solid lines), their 95% pointwise confidence intervals (dashed and dotted lines), and the true CIFs (dotdash lines). (a) n=100,F1. (b) n=100,F2. (c) n=500,F1. (d) n=500,F2. (e) n=1000,F1. (f) n=1000,F2.

Figure 3. Estimated CIFs with Weibull and lognormal distributions (solid lines), their 95% pointwise confidence intervals (dashed and dotted lines), and the true CIFs (dotdash lines). (a) n=100,F1. (b) n=100,F2. (c) n=500,F1. (d) n=500,F2. (e) n=1000,F1. (f) n=1000,F2.

5. Colorectal cancer clinical trial data

To illustrate the proposed estimation method for the AFT mixture cure model with competing risks data, we consider the colorectal cancer clinical trial data from González et al. (Citation2005), which contains rehospitalisation and death data of patients diagnosed with colorectal cancer between January 1996 and December 1998. The data includes calendar time (in days) of the successive hospitalisations after surgical procedure. Gray test with P=0.0006 provides the evidence for considering the competing risk model in this data set, and González et al. (Citation2005) considered death to be a competing risk to rehospitalisation. However, some patients may never have rehospitalisation after discharge. Thus, we consider time to rehospitalisation as the primary interest and death as other possible event which is non-curable, and analyse the data via the proposed model.

This study included 523 patients. The first hospital readmission time was considered as the day between date of surgery and the first rehospitalisation after discharge related to colorectal cancer. There are 23% patients do not experience both events till the end of study and being treated as the right censoring. We considered drug treatment group (chem=1 for thiotepa drug and 0 for control group) and readmissions for comorbidity (Charlson's index:0, 1–2, 3) for possible risk factors for two events of interest. Thiotepa is a member of the class of alkylating agents, which were among the first anticancer drugs used. Alkylating agents are highly reactive and bind to certain chemical groups found in nucleic acids. These compounds inhibit proper synthesis of DNA and RNA, which leads to apoptosis or cell death. However, since alkylating agents cannot discriminate between cancerous and normal cells, both types of cells will be affected by this therapy. For example, normal cells can become cancerous due to alklyating agents. Thus, thiotepa is a highly cytotoxic compound and can potentially have adverse effects. Consequently, the effects of thiotepa on cancer recurrence and death are not obvious.

For the proposed model, all variables were included in the competing risk event with cured and without cured component. The estimated coefficients and their standard deviations for proposed model are listed in Table . Based on the uncure part and competing risk part, both Charlson's index (0) and Charlson's index(1−2) have significant impact. Similarly, both treatment and Charlson's index(1−2) show significant impact on whether patients will experience the rehospitalisation or death.

Table 2. Results for colorectal cancer clinical trial data.

After using the (Equation3), we calculated and plotted the cumulative incidence curves for rehospitalisation and death of colorectal cancer in the Figure . Figure  shows the competing relationship between the rehospitalisaion and death clearly, and the rehospita-lisaion has a higher rate than death.

Figure 4. Proposed model.

Figure 4. Proposed model.

6. Discussions and conclusions

In this paper, we developed a new accelerated failure time mixture cure model allowing non-curable competing risk. Comparing with the existing models, this model can better capture the cure rate of disease than the traditional mixture cure model. The semiparametric estimation based on the EM algorithm can be easily got with the help of existing popular R packages. Although the variance estimation is based on the bootstrap due to the complex structure of model, the comprehensive simulation shows that the performance is reasonable even when the resampling size is small.

Acknowledgements

The research is supported by the Natural Science Foundation of China (Nos. 11271136, 81530086) and the 111 Project of China (No. B14019).

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Notes on contributors

Yijun Wang

Yijun Wang is a PhD candidate in the College of Statistics, East China Normal University, Shanghai, China. Her research interests include Bayesian statistics, reliability statistics and biostatistics.

Jiajia Zhang

Jiajia Zhang is a professor of biostatistics in the Department of Epidemiology and Biostatistics, University of South Carolina, USC. She received her PhD degree from Memorial University. Her professional publications and research interests have focused on survival analysis,semiparametric estimation methods and spatial survival analysis.

Yincai Tang

Yincai Tang is professor of statistics in the College of Statistics, East China Normal University, Shanghai, China. He received his PhD degree from East China Normal University. His professional publications and research interests have focused on lifetime data analysis, degradation data analysis, big data analysis and Bayesian inference.

References

  • Boag, J.W. (1949). Maximum likelihood estimates of the proportion of patients cured by cancer therapy. Journal of the Royal Statistical Society. Series B (Methodological), 11(1), 15–53. doi: 10.1111/j.2517-6161.1949.tb00020.x
  • Crowder, M. (2001). Classical competing risk. London: Chapman and Hall/CRC.
  • David, H., & Moeschberge, M. (1978). The theory of competing risks. London: Griffn.
  • Fine, J., & Gray, R. (1999). A proportional hazards model for the subdistribution of a competing risk. Journal of the American Statistical Association, 94(446), 496–509. doi: 10.1080/01621459.1999.10474144
  • Fusaro, R., Bacchetti, P., & Jewell, N. (1996). A competing risks analysis of presenting aids diagnoses trends. Biometrics, 52, 211–225. doi: 10.2307/2533157
  • Gamel, J., Weller, E., Wesley, M., & Feuer, E. (2000). Parametric cure models of relative and cause-specific survival for grouped survival times. Computer Methods and Programs in Biomedicine, 61(2), 99–110. doi: 10.1016/S0169-2607(99)00022-X
  • Gaynor, J., Feuer, E., Tan, C., Wu, D., Little, C., Straus, D., …Brennan, M. (1993). On the use of cause-specific failure and conditional failure probabilities: Examples from clinical oncology data. Journal of the American Statistical Association, 88(422), 400–409. doi: 10.1080/01621459.1993.10476289
  • González, J. R., Fernandez, E., Moreno, V., Ribes, J., Peris, M., Navarro, M., …Borrás, J. M. (2005). Sex differences in hospital readmission among colorectal cancer patients. Journal of Epidemiology & Community Health, 59(6), 506–511. doi: 10.1136/jech.2004.028902
  • Kalbfleisch, J.D., & Prentice, R.L. (2002). The statistical analysis of failure time data. New York: John Wiley & Sons.
  • Klein, J. (2006). Modelling competing risks in cancer studies. Statistics in Medicine, 25(6), 1015–1034. doi: 10.1002/sim.2246
  • Kleinbaum, D., & Klein, M. (2006). Survival analysis: A self-learning text. New York: Springer Science & Business Media.
  • Kuk, A. (1992). A semiparametric mixture model for the analysis of competing risks data. Australin Jounal of Statistics, 34(2), 169–180.
  • Lambert, P., Thompson, J., Weston, C., & Dickman, P. (2006). Estimating and modeling the cure fraction in population-based cancer survival analysis. Biostatistics (Oxford, England), 8(3), 576–594. doi: 10.1093/biostatistics/kxl030
  • Larson, M., & Dinse, G. (1985). A mixture model for the regression analysis of competing risks data. Applied Statistics, 34(3), 201–211. doi: 10.2307/2347464
  • Li, C., & Taylor, J. (2002). A semi-parametric accelerated failure time cure model. Statistics in Medicine, 21(21), 3235–3247. doi: 10.1002/sim.1260
  • Lu, W. (2010). Efficient estimation for an accelerated failure time model with a cure fraction. Statistica Sinica, 20, 661–674.
  • Lu, W., & Peng, L. (2008). Semiparametric analysis of mixture regression models with competing risks data. Lifetime Data Analysis, 14(3), 231–252. doi: 10.1007/s10985-007-9077-6
  • Ng, S., & McLachlan, G. (2003). An em-based semi-parametric mixture model approach to the regression analysis of competing-risks data. Statistics in Medicine, 22(7), 1097–1111. doi: 10.1002/sim.1371
  • Ohneberg, K., Schumacher, M., & Beyersmann, J. (2017). Modelling two cause-specific hazards of competing risks in one cumulative proportional odds model?. Statistics in Medicine, 36(27), 4353–4363. doi: 10.1002/sim.7437
  • Peng, Y. (2003). Fitting semiparametric cure models. Computational Statistics & Data Analysis, 41(3), 481–490. doi: 10.1016/S0167-9473(02)00184-6
  • Peng, Y., & Dear, K. B. (2000). A nonparametric mixture model for cure rate estimation. Biometrics, 56(1), 237–243. doi: 10.1111/j.0006-341X.2000.00237.x
  • Peng, Y., Dear, K. B., & Denham, J. (1998). A generalized f mixture model for cure rate estimation. Statistics in Medicine, 17(8), 813–830. doi: 10.1002/(SICI)1097-0258(19980430)17:8<813::AID-SIM775>3.0.CO;2-#
  • Pintilie, M. (2007). Analysing and interpreting competing risk data. Statistics in Medicine, 26(6), 1360–1367. doi: 10.1002/sim.2655
  • Sy, J., & Taylor, J. (2000). Estimation in a cox proportional hazards cure model. Biometrics, 56(1), 227–236. doi: 10.1111/j.0006-341X.2000.00227.x
  • Tai, B., Machin, D., White, I., & Gebski, V. (2001). Competing risks analysis of patients with osteosarcoma: A comparison of four different approaches. Statistics in Medicine, 20(5), 661–684. doi: 10.1002/sim.711
  • Van Der VaartJon, A., & Wellner, J. (1996). Weak convergence and empirical processes. New York, NY: Springer.
  • Xu, L., & Zhang, J. (2009). An alternative estimation method for the semiparametric accelerated failure time mixture cure model. Communications in Statistics-Simulation and Computation, 38(9), 1980–1990. doi: 10.1080/03610910903180657
  • Yu, B., & Tiwari, R. (2007). Application of em algorithm to mixture cure model for grouped relative survival data. Journal of Data Science, 5, 41–51.
  • Yu, B., Tiwari, R., Cronin, K., & Feuer, E. (2004). Cure fraction estimation from the mixture cure models for grouped survival data. Statistics in Medicine, 23, 1733–1747. doi: 10.1002/sim.1774
  • Zeng, D., & Lin, D. (2007). Efficient estimation for the accelerated failure time model. Journal of the American Statistical Association, 102(480), 1387–1396. doi: 10.1198/016214507000001085
  • Zhang, J., & Peng, Y. (2007). A new estimation method for the semiparametric accelerated failure time mixture cure model. Statistics in Medicine, 26(16), 3157–3171. doi: 10.1002/sim.2748

Appendix

Brief description of approximation for the lc3(β1). When n, Jn, and Jn/n0, according to the Donsker theorem, we obtain 1ni=1nδiI(εi=1)β1TXiE{δI(ε=1)β1TX},1ni=1nδiI(εi=1)IeRi(β1)[tk1,tk)Pδ=1,ε=1,eRi(β1)[tk1,tk),JnnMi=1nδiI(εi=1)Itk1eRi(β1)<tkdP(δ=1,ε=1,eRi(β1)s)dss=tk1. Using the multiplier central limit theorem of Van Der VaartJon and Wellner (Citation1996), we have max1ni=1npε,1i(m)(eRi(β1)tk)Itk1eRi(β1)<tkEpε,1i(m)(eR(β1)tk)Itk1eR(β1)<tk=Op1n.max1ni=1npε,1i(m)IeRi(β1)tkEpε,1i(m)IeR(β1)tk=Op1n. Since Jn/n0 as n, we can obtain maxJnnMi=1npε,1i(m)(eRi(β1)tk)Itk1eRi(β1)<tk+1ni=1npε,1i(m)IeRi(β1)tkEJnMpε,1i(m)(eR(β1)tk)Itk1eR(β1)<tk+pε,1i(m)IeR(β1)tk0. Uniformly in β and tk. Note, for the last term, we have EJnMpε,1i(m)(eR(β1)tk)Itk1eR(β1)<tk+pε,1i(m)IeR(β1)tk=O(Jn1)+Epε,1i(m)IeR(β1)tk=op(1)+Epε,1i(m)IeR(β1)tk. Then, suplc3p(β1)/nE(δI(ε=1)β1TX)+k=1JnPδ=1,ε=1,eR(β1)[tk1,tk)×logdP(δ=1,ε=1,eR(β1)s)dss=tk1k=1JnPδ=1,ε=1,eR(β1)[tk1,tk)×logEpε,1i(m)IeR(β1)tk0. Similar to Zeng and Lin (Citation2007), we choose a kernel function K() with bandwidth an. The theory of kernel estimation indicates that under suitable regularity conditions, 1nani=1nδiI(εi=1)KRi(β1)logtandP(δ=1,ε=1,R(β1)s)dss=logt=dP(δ=1,ε=1,eR(β1)t)dtt and 1nani=1npε,1i(m)logtKRi(β1)sandsE(pε,1i(m),I(eR(β1)t)). Thus we approximate dP(δ=1,ε=1,eR(β1)t)dt/E(pε,1i(m),I(eR(β1)t)) by 1t1nani=1nδiI(εi=1)KRi(β1)log tan1nani=1npε,1i(m)log tKRi(β1)sands. The Kernel-smoothed approximation of the likelihood function is lsc3(β1)=ndP(δ=1,ε=1,eR(β1)t)E[δI(ε=1)β1TX]+0logdP(δ=1,ε=1,eR(β1)t)/dtEpε,1i(m)I(eR(β1)t)dP(δ=1,ε=1,eR(β1)t)=i=1nδiI(εi=1)β1TXii=1nδiI(εi=1)Ri(β1)

+i=1nδiI(εi=1)log1nanj=1nδiI(εj=1)KRj(β1)Ri(β1)ani=1nδiI(εi=1)log1nj=1npε,1i(m)Rj(β1)Ri(β1)/anKsds. Discarding the constant term, we obtain the following log-likelihood function, lsc3(β1)=i=1nδiI(εi=1)log1nanj=1nδiI(εj=1)KRj(β1)Ri(β1)an i=1nδiI(εi=1)log1nj=1npε,1i(m)Rj(β1)Ri(β1)/anKsds.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.