Hostname: page-component-7bb8b95d7b-5mhkq Total loading time: 0 Render date: 2024-09-28T12:10:06.596Z Has data issue: false hasContentIssue false

On some properties of distributions possessing a bathtub-shaped failure rate average

Published online by Cambridge University Press:  04 January 2023

Ruhul Ali Khan*
Affiliation:
Indian Institute of Engineering Science and Technology, Shibpur
Dhrubasish Bhattacharyya*
Affiliation:
Indian Statistical Institute, Kolkata
Murari Mitra*
Affiliation:
Indian Institute of Engineering Science and Technology, Shibpur
*
*Postal address: Department of Mathematics, Indian Institute of Engineering Science and Technology, Shibpur, P.O. -Botanic Garden, Howrah- 711103, West Bengal, India.
****Postal address: Applied Statistics Unit, Indian Statistical Institute, 203, B. T. Road, Kolkata 700108, India. Email address: dhrubasish018@gmail.com
*Postal address: Department of Mathematics, Indian Institute of Engineering Science and Technology, Shibpur, P.O. -Botanic Garden, Howrah- 711103, West Bengal, India.
Rights & Permissions [Opens in a new window]

Abstract

The life distribution of a device subject to shocks governed by a homogeneous Poisson process is shown to have a bathtub failure rate average (BFRA) when the probabilities $\bar{P}_k$ of surviving k shocks possess the corresponding discrete property. We prove closure under the formation of weak limits for BFRA distributions and explore related moment convergence issues within the BFRA family. Similar results for increasing and decreasing failure rate average distributions are obtained either independently or as consequences of our results. We also establish some results outlining the positions of various non-monotonic ageing classes such as bathtub failure rate, increasing initially then decreasing mean residual life, new worse then better than used in expectation, and increasing initially then decreasing mean time to failure in the hierarchy. Finally, an open problem is posed and a partial solution provided.

Type
Original Article
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of Applied Probability Trust

1. Introduction

Usually, ageing is the process of ‘growing up’ and ‘growing old’ for mechanical devices or biological systems. It begins at birth and ends with death. In reliability theory, the term ageing of a system with a lifetime distribution implies that the residual life of the system is affected by its age in some probabilistic sense. This interpretation allows cases in which a system may typically face positive ageing, negative ageing, or no ageing.

In ‘positive ageing’, the age of the system has an adverse effect on the residual lifetime, i.e. the residual lifetime of the system decreases in some probabilistic sense with increase in age. This type of ageing pattern is the most common, since the majority of systems experience gradual wear and tear under usual operating conditions. ‘Negative ageing’ describes the opposite beneficial effect on the life of the unit as age progresses. In fact, negative ageing is the dual of positive ageing. ‘No ageing’ is equivalent to saying that the age of the system has no effect on the distribution of its residual lifetime.

Monotonic ageing is the phenomenon where the pattern of ageing remains the same during the entire lifespan of a system. Interestingly enough, in many practical situations the ageing pattern happens to be non-monotonic, typically characterized by a ‘burn-in phase’ (negative ageing), a useful life phase (no ageing), and finally a ‘wear-out phase’ (positive ageing). A well-known example is that of human beings, where ageing is beneficial during infancy and young adulthood, i.e. a person becomes stronger with increase in age, and subsequently the reverse phenomenon of gradual degeneration is observed. This is also the case with mechanical systems that may have a high initial failure probability due to design or manufacturing errors, but after some point of time the system improves its performance due to work hardening (see [Reference Aarset1, Reference Wang58]). Moreover, in this context, [Reference Gupta and Warren25] gave two real-life examples: (i) [Reference Langlands, Pocock, Kerr and Gore42] found in a study of curability of breast cancer that the peak of mortality occurred after about three years and then declines slowly; (ii) [Reference Bennett11] analyzed lung cancer data from the Veterans Administration, presented in [Reference Prentice54], and showed that the empirical failure rates for both low and high potassium sulfide groups are non-monotonic. So the modelling of these types of phenomena necessitates the use of various non-monotonic ageing families of life distributions. The specific time point where the ageing profile changes trend is called a change point or turning point of that family. Attempts to model such scenarios have led to several non-parametric classes of distributions such as bathtub failure rate (BFR) [Reference Glaser23], increasing initially then decreasing mean residual life (IDMRL) [Reference Guess, Hollander and Proschan24], new worse then better than used in expectation (NWBUE) [Reference Mitra and Basu47], increasing initially then decreasing mean time to failure (IDMTTF) [Reference Izadi, Sharafi and Khaledi27] etc. [Reference Mi46] observed a trend change in the failure rate average which results in the family of bathtub failure rate average (BFRA) distributions, subsequently developed and studied in detail in [Reference Bhattacharyya, Ghosh and Mitra12].

At this juncture, we recall some basic notation and definitions in the context of BFRA distributions. Suppose that the lifetime of a system (or a component thereof) is represented by a random variable X with cumulative distribution function F(x) and survival function $\bar{F}(x)\;:\!=\;1-F(x)$ .

Definition 1.1. The failure rate average function of a life distribution F is defined by $r(x) = -({\ln{\bar{F}(x)}})/{x}$ , $x>0$ .

If F is absolutely continuous with respect to the Lebesgue measure on the real line, then $r(x)=({1}/{x})\int_0^x h(u) \, \mathrm{d} u$ , where $h(u)={f(u)}/{\bar{F}(u)}$ is called the hazard rate of F and f denotes the probability density function of F.

Definition 1.2. A life distribution F is said to be a BFRA distribution if there exists a point $x_0 \geq 0$ such that the failure rate average function r(x) is non-increasing on $(0, x_0)$ and non-decreasing on $[x_0, \infty)$ . The point $x_0$ is referred to as a turning point (or change point) of F in the BFRA sense.

In this case, we write F is BFRA $(x_0)$ . For $x_0=0$ , the BFRA class reduces to increasing failure rate average (IFRA). We also include all decreasing failure rate average (DFRA) distributions in the BFRA class by adopting the notion that a BFRA distribution having ‘a change point at infinity’ is in fact DFRA. The upside-down bathtub failure rate average (UBFRA) class, which is the dual of the BFRA class, can be obtained in an obvious way by reversing the monotonicity in the above definition. Note that $r(x)=R(x)/x$ , where $R(x) = {-}\log{\bar{F}(x)}$ is the hazard function of F. Singpurwalla [Reference Singpurwalla56] referred to this hazard function as ‘hazard potential’ and provided an innovative interpretation of it. Moreover, another equivalent condition for BFRA distributions can be written as ‘ $\bar{F}(x)^{1/x}$ is non-decreasing on $(0, x_0)$ and non-increasing on $[x_0, \infty)$ ’.

Failure rate average (FRA) is a very important fundamental concept in reliability and survival analysis (see [Reference Barlow and Proschan8, Reference Lai and Xie41]). The IFRA class of distributions, first introduced in [Reference Birnbaum, Esary and Marshall16], generalizes the increasing failure rate (IFR) family and is equivalent to the class of distribution functions where R(x) is star-shaped. The class of IFRA distributions is the smallest ageing class that is closed under the formation of coherent structures and contains the limiting case of no wear (see [Reference Birnbaum, Esary and Marshall16]). Further, this class arises from the cumulative damage shock model when a device is subjected to shocks driven by a Poisson process (see [Reference Esary and Marshall21, Reference Esary, Marshall and Proschan22]). Tests of exponentiality against nonexponential IFRA distributions have been proposed by several authors [Reference Bandyopadhyay and Basu5, Reference Klefsjö37, Reference Kochar39, Reference Link43, Reference Tiwari, Rao Jammalamadaka and Zalkikar57, Reference Wells and Tiwari59]. [Reference Mi46] and [Reference Bhattacharyya, Ghosh and Mitra12] both showed that if a twice-differentiable life distribution function F is BFR $(t_0)$ , then F is BFRA $(x_0)$ where $x_0\geq t_0$ . The relationship between the discrete versions of BFR and BFRA was also given in [Reference Mi45]. [Reference Bhattacharyya, Ghosh and Mitra12] proposed a non-parametric test of exponentiality against BFRA alternatives by assuming that the proportion of early failures is known, and investigated the closure of BFRA property under some reliability operations.

Throughout the paper, we follow the convention that ‘increasing’ (‘decreasing’) means ‘non-decreasing’ (‘non-increasing’).

The rest of the paper is organized as follows. Section 2 extends the results of [Reference Esary, Marshall and Proschan22], specifically in the context of the BFRA class of life distributions. Shock model theory is developed when the shock survival probabilities $\bar{P}_k$ are discrete BFRA and the shock arrivals follow a homogeneous Poisson process. The BFRA property of the discrete failure distribution $P_k$ is shown to be reflected as the BFRA property of $\bar{H}(t)$ . Section 3 deals with closure under weak convergence within the BFRA family and, as a consequence, we obtain closure under the formation of weak limits within the IFRA (DFRA) family. Moreover, the relation between convergence of moment sequences and weak convergence within BFRA and IFRA classes is explored. Further, we show that these connections are not meaningful in the context of UBFRA and DFRA distributions. In Section 4, we prove that the BFRA class of distributions contains all BFR distributions. Further, we establish some results to investigate the interrelationships among well-known non-monotonic ageing families. Finally, we conclude by suggesting possible avenues of future work, and also pose an interesting open problem in this area.

2. Shock model theory

We focus on a system where failure is caused by a sequence of shocks occurring randomly in time according to a counting process $\lbrace N(t)\colon t\geq 0\rbrace$ . Let the system have a probability $\bar{P}_k$ of surviving the first k shocks, $k=0, 1, 2, \dots$ , where the $\bar{P}_k$ satisfy the inheritance condition $1= \bar{P}_0 \geq \bar{P}_1 \geq \bar{P}_2 \geq \cdots$ , and $\sum_{i=0}^\infty \bar{P}_{i} = \gamma < \infty$ , i.e. the expected number of shocks required to cause failure of the system is finite. Assume that only the shocks are responsible for failure, and constant wear and tear is completely absent. Then, using a conditioning argument, the probability $\bar{H}(t)$ that the system will survive beyond time t may be written as $\bar{H}(t) = \sum_{k=0}^\infty \mathbb{P}[N(t)=k]\bar{P}_k$ . Now, this can be written as

(2.1) \begin{equation} \bar{H}(t)= \sum_{k=0}^\infty \mathrm{e}^{-\lambda t}\frac{(\lambda t)^k}{k!}\bar{P}_k\end{equation}

when the shock arrivals follow a homogeneous Poisson process with intensity $\lambda>0$ . A shock model of this kind was considered in [Reference Esary, Marshall and Proschan22] for the first time in the context of ageing scenarios. ‘Shocks’ can be interpreted as claims, the technical reasons for failure in a running machine, or the cause of deterioration of a biological organ. Thus, the Poisson shock model has various applications in several fields, including risk, survival, and reliability analysis (see [Reference Nakagawa51] and the references therein). In the vast literature concerning shock model theory, [Reference Anderson3, Reference Boland and Proschan17, Reference Ebrahimi19, Reference El-Neweihi, Proschan and Sethuraman20, Reference Kochar40, Reference Pellerey53, Reference Shanthikumar and Sumita55] and others are important. Thus, it is relevant to investigate the origin of the ageing class under consideration via discrete survival probabilities.

In reliability shock models, for the discrete distribution $P_k$ , $k=0,1,2,\dots$ , most research tries to establish that properties of $P_k$ are reflected in corresponding properties of the continuous life distribution H(t). This is shown in [Reference Esary, Marshall and Proschan22] for ILR, IFR, IFRA, DMRL, NBU, and NBUE classes, [Reference Klefsjö35, Reference Klefsjö38] for HNBUE and $\mathcal{L}$ classes, [Reference Deshpande, Kochar and Singh18] for strong increasing failure rate and SNBU, and [Reference Abouammoh, Hindi and Ahmed2] for NBUFR and NBAFR. For non-monotonic ageing scenarios, [Reference Mitra and Basu49] first proved such results in the context of NWBUE and BFR classes. Analogous results have been established in [Reference Anis4, Reference Khan, Bhattacharyya and Mitra32] in the context of the IDMRL and IDMTTF distribution families. For a more detailed overview, see [Reference Mitra and Khan50] and the references therein.

In this section we establish the corresponding results for BFRA classes of life distributions, which remains unexplored. The following definition introduces the notion of the discrete version of the BFRA property.

Definition 2.1. The sequence $\lbrace \bar{P}_k, k=0,1,\dots \rbrace$ is said to have the discrete BFRA (D-BFRA) property if there exists $k_0 \geq 0$ such that $\bar{P_k}^{1/k}$ is increasing in k for $k<k_0$ and decreasing in k for $k\geq k_0$ .

At the outset, we recapitulate some results of total positivity theory and establish two lemmas before going on to prove the main result in this section.

Lemma 2.1. Consider a function $\phi\colon [a, b) \rightarrow (0, 1]$ where $a, b\in [0, \infty)$ . Then $\phi(s)^{1/s}$ is increasing (decreasing) in s if, for every $c\geq 0$ , $\phi(s)-\mathrm{e}^{-cs}$ has at most one change of sign, and if one change occurs, it occurs in the order $-,+$ ( $+,-$ ).

Proof. Let $s_1, s_2 \in [a, b)$ , $s_1<s_2$ . We want to show that $\phi(s_1)^{1/s_1}\leq \phi(s_2)^{1/s_2}$ , i.e. $({1}/{s_1})\log{(\phi(s_1))}\leq ({1}/{s_2})\log{(\phi(s_2))}$ , i.e. $g(s_1)/s_1\geq g(s_2)/s_2$ where $g(s)=-\log{(\phi(s))}$ . Let $c_1=g(s_1)/s_1$ and $c_2=g(s_2)/s_2$ . Suppose that, for every $c\geq 0$ , $\phi(s)-\mathrm{e}^{-cs}$ has at most one change of sign, from $-$ to $+$ if one occurs. So, in particular, this happens for $c_1$ and $c_2$ . Thus, $\log{(\phi(s))}+c_i s$ has at most one change of sign from $-$ to $+$ for $i=1,2$ , i.e. $c_i s -g(s)$ has at most one change of sign $-$ to $+$ if one occurs for $i=1,2$ . Now, by choice of $c_1$ and $c_2$ , we have $g(s) \leq c_i s$ for $s\leq s_i$ and $g(s) \geq c_i s$ for $s > s_i$ . If possible, let $c_1=g(s_1)/s_1 < g(s_2)/s_2=c_2$ . Then, for $s_1<s<s_2$ , $g(s) \leq c_1 s$ as $s>s_1$ , $g(s) < c_2 s$ as $c_1<c_2$ , and $g(s) \leq g(s)$ as $s<s_2$ , which is a contradiction. So $\phi(s)^{1/s}$ is increasing in s. The proof for the ‘decreasing’ case is analogous.

Lemma 2.1 shows the characterization of monotonicity of $\phi(s)^{1/s}$ in terms of the sign changes. Before proceeding further, we recall the definition of a totally positive function and an important theorem of Karlin ([Reference Karlin29, p. 21]).

Definition 2.2. Let $X, Y \subseteq \mathbb{R}$ . A function $L\colon X\times Y \rightarrow \mathbb{R}$ is called a totally positive function of order n ( $TP_n$ ) if, for every $i=1,2,\dots , n$ ,

\begin{align*} \begin{vmatrix} L(s_1, t_1) & L(s_1, t_2) & \dots & L(s_1, t_i) \\[5pt] L(s_2, t_1) & L(s_2, t_2) & \dots & L(s_2, t_i) \\[5pt] \vdots & \vdots & \ddots &\vdots \\[5pt] L(s_i, t_1) & L(s_i, t_2) & \dots & L(s_i, t_i) \end{vmatrix} \geq 0 , \end{align*}

where $ s_1<s_2<\cdots <s_i$ ( $s_j \in X$ , $j=1,2,\dots, i$ ) and $t_1<t_2<\cdots <t_i$ ( $t_j \in Y$ , $j=1,2,\dots, i$ ).

Definition 2.3. A function L is called totally positive (TP) when L is $TP_n$ for every $n\geq1$ .

Consider a subset $\lbrace x_i\mid 1\leq i \leq m \rbrace$ of $\mathbb{R}$ and let $S^-(x_1, x_2, \dots x_m)$ be the number of sign changes in $x_1, x_2, \dots x_m$ ignoring zero terms. Let Q be an ordered subset of $\mathbb{R}$ and $g\colon Q\rightarrow \mathbb{R}$ be a function. Define $S^-(g)\;:\!=\;\sup_{y_1<y_2<\cdots <y_m \in Q} S^-( g(y_1), g(y_2), \dots , g(y_m))$ .

Let L(x, y) defined on $X\times Y$ be Borel measurable, and assume for simplicity that the integral $\int_Y L(x, y) \, \mathrm{d} \mu (y)$ exists for every x in X. Here, $\mu$ represents a fixed sigma-finite regular measure defined on Y such that $\mu(U)>0$ for each open set U for which $U\cap Y$ is non-empty.

Theorem 2.1. Let L(x, y) be $TP_r$ on $X\times Y$ . Let g be a bounded Borel-measurable function on Y. Let the transformation $f(x)=\int_Y L(x, y) g(y) \, \mathrm{d} \mu (y)$ be finite for each x on X. Then $S^-(f) \leq S^-(g)$ provided $S^-(g)\leq r-1$ . Moreover, if $S^-(f) = S^-(g) \leq r-1$ , then g and f exhibit the same sequence of signs when their respective arguments traverse the domain of definition from left to right.

Theorem 2.1 is called the variation diminishing property (VDP) of TP functions; see [Reference Karlin29, Chapter 5] for a proof. The following main theorem demonstrates that $\bar{H}(t)$ inherits the BFRA property if the sequence $\lbrace\bar{P}_k,\,k\geq 0\rbrace$ possesses the discrete BFRA property.

Theorem 2.2. Let the sequence $\lbrace\bar{P}_k,\,k\geq 0\rbrace$ possess the D-BFRA property. Then $\bar{H}(t)$ defined in (2.1) is BFRA.

Proof. Let $\eta\in[0,1]$ . Now, from Definition 2.1, as ${\bar{P}_k}^{1/k}$ is increasing in k for $k<k_0$ and ${\bar{P}_k}^{1/k}$ is decreasing in k for $k\geq k_0$ for some $k_0\in \mathbb{N}$ , $\bar{P}_k - \eta^{k}$ has at most two sign changes in the direction $-,+,-$ . Now,

\begin{equation*} \bar{H}(t) - \mathrm{e}^{-(1- \eta)\lambda t } = \sum_{k=0}^\infty \mathrm{e}^{-\lambda t}\frac{(\lambda t)^k}{k!}\bar{P}_k - \sum_{k=0}^\infty \mathrm{e}^{-\lambda t} \frac{{(\lambda \eta t)}^k}{k!} \nonumber = \sum_{k=0}^\infty (\bar{P}_k-\eta^k) \mathrm{e}^{-\lambda t} \frac{(\lambda t)^k}{k!}. \nonumber \end{equation*}

In the first case, suppose $\bar{P}_k - \eta^{k}$ has two sign changes in the direction $-,+,-$ . Note that $\mathrm{e}^{-\lambda t}{(\lambda t)^k}/{k!}$ is a $TP_3$ function for $s>0$ and $k\in \mathbb{N}\cup \lbrace 0\rbrace$ , since $\mathrm{e}^{-\lambda t}{(\lambda t)^k}/{k!}$ is a TP function for any $\lambda>0$ . Then, by the VDP, $S^-(\bar{H}(t) -\mathrm{e}^{-(1- \eta)\lambda t })=0$ , 1, or 2. If $S^-(\bar{H}(t) -\mathrm{e}^{-(1- \eta)\lambda t })=2$ , then, by the VDP, $\bar{H}(t) -\mathrm{e}^{-(1- \eta)\lambda t }$ has the same sign change property in t. Note that $\bar{H}(t)\geq \mathrm{e}^{-\lambda t}$ since the hazard rate $h(t)\leq \lambda$ for all $t>0$ (see [Reference Esary, Marshall and Proschan22, p. 629]). Hence, for any $c\leq \lambda$ , $\bar{H}(t) -\mathrm{e}^{-c t }$ has the same sign change property in t, and the choice for $\eta$ is $1-c/\lambda$ . Moreover, for any $c> \lambda$ , $\bar{H}(t) -\mathrm{e}^{-c t }$ is always positive for $t>0$ . Thus, there exists $d\in [0, \infty)$ such that, for any $c\geq 0$ , $\bar{H}(t) -\mathrm{e}^{-c t }$ has at most one change of sign from $-$ to $+$ for $t\leq d$ , and $\bar{H}(t) -\mathrm{e}^{-c t }$ has at most one change of sign from $+$ to $-$ for $t> d$ . Applying Lemma 2.1 on [0, d) and $[d, \infty)$ , we get the result that $\bar{H}(t)$ is BFRA. If $\bar{H}(t) -\mathrm{e}^{-(1- \eta)\lambda t }$ has at most one change of sign, then the sign of $\bar{H}(t) -\mathrm{e}^{-(1- \eta)\lambda t }$ can change from $+$ to $-$ or $-$ to $+$ . Again, an application of Lemma 2.1 together with the fact that $\bar{H}(t)\geq \mathrm{e}^{-\lambda t}$ yields the result.

In the second case, suppose $\bar{P}_k - \eta^{k} $ has one sign change from $+,-$ or $-,+$ . Then, by the VDP, $\bar{H}(t) -\mathrm{e}^{-(1- \eta)\lambda t }$ has the same sign change property in t, if sign change occurs in $\bar{P}_k - \eta^{k} $ . Again, using the fact that $\bar{H}(t)\geq \mathrm{e}^{-\lambda t}$ and arguing as in [Reference Esary, Marshall and Proschan22], we can conclude that, for any $c\geq 0$ , $\bar{H}(t) -\mathrm{e}^{-c t }$ has at most one sign change in t, if one occurs. Again, Lemma 2.1 yields the result that $\bar{H}(t)$ is IFRA or DFRA.

Corollary 2.1. $\bar{H}(t)$ belongs to the IFRA (DFRA) class if $\lbrace\bar{P}_k,\,k\geq 0\rbrace$ satisfies the discrete IFRA (DFRA) property.

Remark 2.1. It is important to note that [Reference Esary, Marshall and Proschan22, (3.4) of Theorem 3.1] is identical to Corollary 2.1.

3. Weak convergence issues within BFRA class

During the last few decades, the topic of weak convergence in various ageing classes has generated substantial interest among reliability experts. Weak convergence issues within a class of distributions has been addressed in [Reference Basu, Simons, Sen and Holland10] for IFR, [Reference Basu and Bhattacharjee9] for HNBUE, [Reference Khan, Bhattacharyya and Mitra33] for IMIT, [Reference Mitra and Basu47, Reference Mitra and Basu48] for NWBUE and BFR, [Reference Anis4] for IDMRL, [Reference Khan, Bhattacharyya and Mitra32] for IDMTTF distributions. Let $F_n$ , $n=1,2, \dots $ be a sequence of BFRA distributions with $F_n$ converging to F in distribution. In this section, we first try to establish that F belongs to the BFRA class, i.e. closure under the formation of weak limits within the BFRA family.

Theorem 3.1. Suppose $\lbrace F_n,\,n\geq 1 \rbrace$ is a sequence of BFRA distributions. Let $x_{0n}$ be a change point of $F_n$ , and suppose that $F_n$ converges to F in distribution, where F is assumed to be continuous. Then F is BFRA.

Proof. Let $r(x)=\Lambda_{F_n}(x)=R_{F_n}(x)/x$ be the corresponding FRA function. By the BFRA property, $\Lambda_{F_n}$ is decreasing on $(0, x_{0n})$ and increasing on $[ x_{0n}, \infty)$ , where $R_{F_n}(x)=-\ln{\bar{F}_n(x)}$ . Now, $F_n \stackrel{\mathcal{L}}{\rightarrow} F$ implies $R_{F_n}(x) \rightarrow R_{F}(x)$ , and consequently $\Lambda_{F_n} \rightarrow \Lambda_{F}$ pointwise, where $\stackrel{\mathcal{L}}{\rightarrow}$ denotes convergence in distribution. For any BFRA distribution H, we define $C_H = \lbrace x_0\colon \Lambda_{H}(x)$ is decreasing on $(0, x_{0})$ and increasing on $[ x_{0}, \infty) \rbrace$ . Now, $x_{0n}\in C_{F_n}$ , $n=1,2, \dots$ . Two cases may arise: (i) $\lbrace x_{0n}\rbrace$ is bounded; (ii) $\lbrace x_{0n}\rbrace$ is unbounded.

For case (i), $\lbrace x_{0n}\rbrace_{n=1}^{\infty}$ is bounded, an application of the Bolzano–Weierstrass theorem yields a subsequence $\lbrace x_{0n_k}\rbrace$ of $\lbrace x_{0n}\rbrace$ such that $x_{0n_k}\rightarrow l$ (finite) as $k\rightarrow\infty$ . Thus, for any $\epsilon>0$ , there exists $k_0\in \mathbb{N}$ such that $l - \epsilon < \lbrace x_{0n_k}\rbrace < l + \epsilon$ for all $k\geq k_0$ . Now, for any $x_1, x_2 > 0$ such that $x_1<x_2<l -\epsilon$ , we have $\Lambda_{F_{n_k}}(x_1) \geq \Lambda_{F_{n_k}}(x_2)$ since $\Lambda_{F_{n_k}}(x)$ is decreasing on $(0, x_{0n_k})$ . Consequently, using the fact that $\Lambda_{F_n}\rightarrow \Lambda_{F}$ , we get $\Lambda_{F}(x_1) \geq \Lambda_{F}(x_2)$ as $k\rightarrow \infty$ . Thus, $\Lambda_F(x)$ is decreasing on $(0, l - \epsilon)$ . Similarly, we can show that $\Lambda_F(x)$ is increasing on $(l + \epsilon, \infty)$ . Since $\epsilon>0$ is arbitrary, F is BFRA with l as a change point.

For case (ii), $\lbrace x_{0n}\rbrace_{n=1}^{\infty}$ is unbounded, there exists a subsequence $\lbrace x_{0n_k}\rbrace$ such that $x_{0n_k}\rightarrow \infty$ as $k\rightarrow \infty$ . Now, for any $0<x_1<x_2<\infty$ there exists $k_0\in\mathbb{N}$ sufficiently large that $x_1 < x_2 < x_{0n_k}$ for all $k\geq k_0$ . Thus, $\Lambda_{F_{n_k}}(x_1) \geq \Lambda_{F_{n_k}}(x_2)$ , as $\Lambda_{F_{n_k}}(x)$ is decreasing on $(0, x_{0n_k})$ . Consequently, using the fact that $\Lambda_{F_{n_k}}\rightarrow \Lambda_{F}$ pointwise as $k\rightarrow \infty$ , we get $\Lambda_{F}(x_1) \geq \Lambda_{F}(x_2)$ . Hence, $\Lambda_F(x)$ is DFRA, i.e. BFRA with a change point at $\infty$ .

The weak limit of a sequence of IFRA and DFRA distributions is not known so far. However, as a consequence of Theorem 3.1, we obtain the following corollary which shows closure under the formation of weak limits within the IFRA (DFRA) family of distributions.

Corollary 3.1. Let $\lbrace F_n,\,n\geq 1\rbrace$ be a sequence of IFRA (DFRA) distributions. Let F be a continuous distribution function, and $F_n$ converge to F in distribution. Then F is IFRA (DFRA).

The next theorem deals with convergence of the sequence of change points of $F_n$ .

Theorem 3.2. Suppose $\lbrace F_n,\,n\geq 1\rbrace$ is a sequence of BFRA distributions having unique change points $x_{0n}$ . Let F be a continuous distribution function with unique change point $x_0$ in the BFRA sense, $F_n$ converging to F in distribution. Then $\lim_{n\rightarrow \infty}x_{0n}=x_0$ ( $\leq \infty$ ).

Proof. Note that the unique change point means the FRA function is not constant in any neighborhood of its change. We again prove this theorem by considering two cases: (i) $\lbrace x_{0n}\rbrace$ is bounded; (ii) $\lbrace x_{0n}\rbrace$ is unbounded.

For case (i), $\lbrace x_{0n}\rbrace_{n=1}^{\infty}$ is bounded, following case (i) of the proof of Theorem 3.1, it can be easily shown that there exists a convergent subsequence of $\lbrace x_{0n}\rbrace$ whose limit is a change point of F. Then, by the uniqueness of the change point of F, it follows that $\lim_{n\rightarrow \infty}x_{0n}=x_0$ ( $< \infty$ ).

For case (ii), $\lbrace x_{0n}\rbrace_{n=1}^{\infty}$ is unbounded, case (ii) of the proof of Theorem 3.1 implies that F is necessarily a DFRA distribution, i.e. a BFRA distribution with unique change point $x_0=\infty$ . Now suppose that $\lbrace x_{0n}\rbrace \nrightarrow \infty$ as $n\rightarrow\infty$ . Then there exists $M>0$ such that $x_{0n}\leq M$ infinitely often, so that it should be possible to find a subsequence $\lbrace x_{0n_k}\rbrace$ converging to a finite limit $\alpha$ , say. But then, by case (i) in the proof of Theorem 3.1, along with the uniqueness of the change point of F, we have the contradiction that $\infty=\alpha <\infty$ . Thus, $x_{0n} \rightarrow x_0=\infty$ as $n\rightarrow \infty$ .

In the next theorem we explore the convergence of the corresponding moment sequences of all orders.

Theorem 3.3. Let $\lbrace F_n,\,n\geq 1\rbrace$ be a sequence of absolutely continuous BFRA distributions having a unique change point $x_{0n}$ . Assume that $\lbrace x_{0n}\rbrace_{n=1}^{\infty}$ is bounded. If $F_n$ converges to F in distribution where F is continuous, then, for every $r>0$ , $\mathbb{E}_{F_n}(X^r) \rightarrow \mathbb{E}_F(X^r)$ as $n\rightarrow \infty$ .

Proof. Using [Reference Bhattacharyya, Ghosh and Mitra12, Theorem 3.2] we get finiteness of moments of $F_n$ . Now, following Theorem 3.1 we conclude that F is BFRA. If F is BFRA with a finite change point then again an application of [Reference Bhattacharyya, Ghosh and Mitra12, Theorem 3.2] yields finiteness of moments of F. Now we consider the case when F is BFRA with a change point at infinity, i.e. DFRA. Note that there exists B such that $x_{0n}\leq B$ for all n since $\lbrace x_{0n}\rbrace_{n=1}^{\infty}$ is bounded. Take $B<x<y$ . Then $\Lambda_{F_n}(x)\leq \Lambda_{F_n}(y)$ , since the FRA function of $F_n$ is increasing on $[x_{0n}, \infty)$ . Consequently, taking limits of both sides as $n\rightarrow\infty$ , we get $\Lambda_{F}(x)\leq \Lambda_{F}(y)$ . Now, the DFRA property of F leads to $\Lambda_{F}(x)\geq \Lambda_{F}(y)$ . Thus, $\Lambda_{F}(x)$ is constant for all $x>B$ . Suppose $\Lambda_{F}(x)=c$ for all $x>B$ . Thus $\bar{F}(x) \leq 1$ for $x\leq B$ and $\bar{F}(x) = \mathrm{e}^{-{c x}}$ for $x>B$ , and hence $\int_0^{\infty} x^{r-1} \bar{F}(x) \, \mathrm{d} x <\infty$ . Consequently, it suffices to show that, for all $r>0$ , $\lim_{n\rightarrow\infty}\int_0^{\infty} x^{r-1} \bar{F}_n(x) \, \mathrm{d} x = \int_0^{\infty} x^{r-1} \bar{F}(x) \, \mathrm{d} x$ . Suppose $\epsilon>0$ is such that $\epsilon + \bar{F}(M) = \mathrm{e}^{-\theta}$ for some $\theta>0$ , where $M\in \mathbb{R}^+$ such that $M>B$ . Now, $F_n$ converges to F in distribution implies that there exists $N\in\mathbb{N}$ depending on $\epsilon$ such that $\bar{F}_n(M) < \mathrm{e}^{-\theta}$ for all $n\geq N$ . Note that the FRA function of $F_n$ increasing on $[x_{0n}, \infty)$ implies that $\lbrace \bar{F}_n(x)\rbrace^{{1}/{x}}$ is decreasing on $[x_{0n},\infty)$ . Consequently, $\bar{F}_n(x)\leq V(x)$ , where $V(x) = 1$ for $x\leq M$ , and $V(x) = \mathrm{e}^{-{\theta x}/{M}}$ for $x>M$ . Thus, $x^{r-1} \bar{F}_n(x)$ is bounded by the integrable function $x^{r-1} V(x)$ on $(0,\infty)$ . Hence, the theorem follows in view of the dominated convergence theorem, since $F_n(x)\rightarrow F(x)$ pointwise.

In the next theorem we explore the convergence of moment sequences for the IFRA family. Here we even relax the condition of absolute continuity.

Theorem 3.4. Let $\lbrace F_n,\,n\geq 1\rbrace$ be a sequence of continuous distributions which are IFRA with finite means. If $F_n$ converges to F in distribution where F is continuous with finite mean, then, for every $r\geq 1$ , $\mathbb{E}_{F_n}(X^r) \rightarrow \mathbb{E}_F(X^r)$ as $n\rightarrow \infty$ .

Proof. For an IFRA distribution G with mean $\mu$ , from [Reference Barlow and Proschan8] we get that

(3.1) \begin{equation} \mathbb{E}_G(X^r)\leq \Gamma(r+1)\mu^r \quad \text{if } r\geq 1 , \end{equation}

and there exists $\zeta_p$ such that $F(\zeta_p)=p$ , $0<p<1$ , for which

(3.2) \begin{equation} \bar{G}(x)\leq \begin{cases} 1 & \text{ for } x\leq \zeta_p , \\[5pt] \mathrm{e}^{-\alpha x } & \text{ for } x>\zeta_p , \end{cases} \end{equation}

where $\alpha=-({1}/{\zeta_p})\ln{(1-p)}$ . Corollary 3.1 ensures that F is IFRA, and (3.1) implies that $\mathbb{E}_F(X^r)<\infty$ for all $r>1$ . Now, using (3.1) and (3.2), and following the argument of the proof of Theorem 3.3, we get the result.

Theorem 3.5. Suppose F is an absolutely continuous BFRA distribution with a finite change point at $x_0$ . Then F is uniquely determined by its moment sequence.

Proof. [Reference Bhattacharyya, Ghosh and Mitra12, Theorem 3.2] shows that, for all $r\geq 1$ ,

\begin{equation*} \mathbb{E}_F(X^r) \;:\!=\; \int_0^{\infty} x^r \, \mathrm{d} F(x) \leq \frac{\Gamma (r+1)} {\big(\min_{u\in[0,t_0]}{\lambda (u)}\big)^r} , \end{equation*}

from which it can be shown that the power series $\sum_{r=0}^{\infty} ({u^r}/{r!}) \mathbb{E}_F(X^r)$ has a non-null radius of convergence. The theorem now follows easily using [Reference Loève44, p. 217].

Using (3.1) together with [Reference Loève44, p. 217], the following theorem holds.

Theorem 3.6. Suppose F is a continuous IFRA distribution with finite mean. Then F is uniquely determined by its moment sequence.

The following theorem is related to the converse of Theorem 3.3.

Theorem 3.7. Let $\lbrace F_n,\,n\geq 1\rbrace$ be a sequence of absolutely continuous BFRA distributions with a unique change point $x_{0n}<\infty$ for all $n\geq 1$ , and suppose that F is an absolutely continuous BFRA with unique finite change point $x_0$ such that, for all integers $r>0$ ,

(3.3) \begin{equation} \lim_{n\rightarrow\infty}\int_0^{\infty} x^r \, \mathrm{d} F_n(x) = \int_0^{\infty} x^r \, \mathrm{d} F(x). \end{equation}

Then $F_n$ converges to F in distribution.

Proof. As a consequence of Theorem 3.5 and (3.3), every weakly convergent subsequence of $\lbrace F_n,\,n\geq 1\rbrace$ necessarily converges to the distribution F. This concludes the proof.

The converse of Theorem 3.4 also holds for IFRA family following an argument analogous to the proof of Theorem 3.7, using Theorem 3.6 instead of Theorem 3.5.

Theorem 3.8. Let $F_n,\, n=1, 2, \dots ,$ be a sequence of continuous IFRA distributions. Suppose that F is continuous and IFRA such that, for all integers $r\geq 1$ , $\mathbb{E}_{F_n}(X^r) \rightarrow \mathbb{E}_F(X^r)$ as $n\rightarrow \infty$ . Then $F_n$ converges to F in distribution.

Example 3.1. Consider a survival function $\bar{F_1}(x)=(x^2+1)^{-1}$ , $x\geq 0$ , of a life distribution. It is easily seen that $F_1$ is a UBFR distribution with a change point at 1. Then, from [Reference Mi46, Theorem 1] or [Reference Bhattacharyya, Ghosh and Mitra12, Theorem 2.1], it can be concluded that $F_1$ is a UBFRA distribution since $F_1$ is a twice-differentiable function. Further, it can also be shown that the change point of $F_1$ in UBFRA sense is finite. Note that $\mathbb{E}_F(X)=\pi/2<\infty$ and $\mathbb{E}_F(X^r)=\infty$ for all $r>1$ . Consequently, the versions of Theorems 3.3, 3.5, and 3.7 are not meaningful in the context of UBFRA distributions.

Remark 3.1. The versions of Theorems 3.4, 3.6, and 3.8 are not meaningful in the context of DFRA distributions.

4. Interrelationships among non-monotonic ageing classes

Interrelationships among ageing classes have received widespread attention in the literature from the very beginning [Reference Barlow and Proschan7, Reference Joe and Proschan28, Reference Klefsjö36, Reference Lai and Xie41]. In this section, we study the interrelationships between the BFR family of [Reference Glaser23], the IDMRL family of [Reference Guess, Hollander and Proschan24], the NWBUE family of [Reference Mitra and Basu47], and the IDMTTF family of [Reference Izadi, Sharafi and Khaledi27]. In the context of the abovementioned non-monotonic ageing classes, see the recent works [Reference Bhattacharyya, Khan and Mitra13, Reference Bhattacharyya, Khan and Mitra14, Reference Bhattacharyya, Khan and Mitra15, Reference Izadi and Manesh26, Reference Khan, Bhattacharyya and Mitra30, Reference Khan, Bhattacharyya and Mitra31, Reference Khan and Mitra34]. Exploiting ‘total time on test’ (TTT) transform characterizations, [Reference Mitra and Basu47] established that the NWBUE family of life distributions contains both IDMRL and BFR classes. [Reference Mi46] proved that BFR implies both BFRA and IDMRL for a twice-differentiable distribution function. [Reference Izadi, Sharafi and Khaledi27] established that $\{\mathrm{BFR}\} \subset \{\mathrm{IDMTTF}\} \subset \{\mathrm{NWBUE}\}$ . [Reference Khan, Bhattacharyya and Mitra32] proved that a BFR life distribution F with change point $t_0$ implies F is IDMRL with a change point $\tau$ where $\tau\leq t_0$ , and that the $\{\mathrm{IDMRL}\}$ and $\{\mathrm{IDMTTF}\}$ families intersect each other. Moreover, [Reference Khan, Bhattacharyya and Mitra32] summarized the following interrelationships among the abovementioned classes of life distributions:

Here, $x^\prime$ , $\tau$ , $t_0$ , and $t_0^\prime$ denote the change points of F in BFRA, IDMRL, IDMTTF, and NWBUE senses respectively. If each of the classes in the preceding chain is replaced by its dual, the corresponding structure also holds. [Reference Mi46] and [Reference Bhattacharyya, Ghosh and Mitra12] both proved that a BFR family of twice-differentiable distributions is necessarily BFRA. Here we establish a general result that a BFRA family of distributions contains a BFR class of distributions, where we do not consider any assumptions like ‘twice differentiability of distribution F’. The following lemma plays a key role in establishing the main result.

Lemma 4.1. Suppose $y^*,y^{**}\in [0,\infty)$ and $g\colon[y^*,y^{**}]\rightarrow \mathbb{R}^+$ is a convex function. Let $A_g= \lbrace \alpha \mid g(y) = \alpha y \text{ has at least one solution in } [y^*,y^{**}] \rbrace$ . Now define

\begin{align*} \tilde{y}_1 & = \min \lbrace y \mid y \text{ is a solution of } g(y) = \tilde{\alpha}y \text{ in } [y^*,y^{**}] \rbrace , \\[5pt] \tilde{y}_2 & = \max \lbrace y \mid y \text{ is a solution of } g(y) = \tilde{\alpha}y \text{ in } [y^*,y^{**}] \rbrace , \end{align*}

where $\tilde{a}=\min A_g$ . Then:

  1. (i) ${g(y)}/{y}$ is decreasing in y for all $y\in [y^*,y^{**}]$ if $\tilde{y}_2=y^{**}$ .

  2. (ii) ${g(y)}/{y}$ is increasing in y for all $y\in [y^*,y^{**}]$ if $\tilde{y}_1=y^*$ .

  3. (iii) ${f(y)}/{y}$ is decreasing in y for all $y\in [y^*,\tilde{y}]$ and increasing in y for all $y\in [\tilde{y},y^{**}]$ , where $\tilde{y}\in [\tilde{y}_1, \tilde{y}_2]$ if $y^*<\tilde{y_1}\leq \tilde{y_2}<y^{**}$ .

Proof. The proof is similar to the proof of [Reference Izadi, Sharafi and Khaledi27, Lemma 1].

The following theorem is the main result of this section.

Theorem 4.1. If a life distribution F is BFR $(x^\prime)$ then F is BFRA $(x_0)$ , where $x^\prime\leq x_0$ .

Proof. There are three separate cases to consider.

Case (i): $x^\prime=0$ . In this case, F is IFR. [Reference Barlow6] showed that IFR implies IFRA. Thus, F is BFRA with change point $x_0 =0$ .

Case (ii): $0<x^\prime<\infty$ . If F is BFR $(x^\prime)$ , then $R(x)=-\ln{\bar{F}(x)}$ is a positive concave function on $(0, x^\prime]$ . Hence, for all $x\in (0, x^\prime]$ and for $0< \beta< 1$ , $\ln{\bar{F}(\beta x)} = \ln{\bar{F}(\beta x+ \bar{\beta} 0)}$ , where $\bar{\beta}=1-\beta$ , i.e. $\ln{\bar{F}(\beta x)} \leq \beta \ln{\bar{F}(x)}$ , i.e. $\lbrace \bar{F}(\beta x)\rbrace^{{1}/{\beta x}} \leq \lbrace \bar{F}(x)\rbrace^{{1}/{x}}$ . Thus, $\lbrace \bar{F}(x)\rbrace^{{1}/{x}}$ is increasing in $x\in (0, x^\prime]$ , i.e. the FRA of F is decreasing on $(0, x^\prime]$ . Again, note that $R(x)=-\ln{\bar{F}(x)}$ is a positive convex function on $[x^\prime, \infty)$ , since F is a BFR distribution with a change point at $x^\prime$ . By using Lemma 4.1, the result follows immediately.

Case (iii): $x^\prime$ is a change point at $\infty$ . In this case, F is DFR. [Reference Barlow6] showed that DFR implies DFRA. Thus, F is BFRA with a change point at infinity.

Note that [Reference Bhattacharyya, Ghosh and Mitra12, Example 2.1] indicates that the BFR class of life distributions is strictly smaller than the BFRA class. We now investigate the relationships between BFRA, IDMRL, IDMTTF, and NWBUE, since IDMRL, NWBUE and IDMTTF are larger classes of distributions than the BFR class.

In the next example we try to find the connection between the BFRA and IDMRL distribution classes.

Example 4.1. Consider the distribution function given by

\begin{equation*} F_2(x)= \begin{cases} 1 - \mathrm{e}^{-x}, & 0 \leq x < 1 , \\[3pt] 1 - (\mathrm{e} x)^{-1}, & 1 \leq x < 2 , \\[3pt] 1 - \frac{1}{2}\mathrm{e}^{-2}\exp\big[{-}\frac{1}{2}x(x-3)\big], & 2 \leq x < 3 , \\[3pt] 1 - \frac{1}{2}\mathrm{e}^{\frac{19}{4}}\exp\big[{-}\frac{1}{4}x(12-x)\big], & 3 \leq x < 3.5 , \\[3pt] 1 - \frac{1}{2}\mathrm{e}^{-\frac{1}{2}}\exp\big[{-}\frac{5}{28}x^2\big], & x \geq 3.5. \end{cases} \end{equation*}

It was shown in [Reference Bhattacharyya, Ghosh and Mitra12] that $F_2$ is a BFRA distribution. The change point $x_0$ in the BFRA sense turns out to be $2.320\,84$ . Note that the values $e_{F_2}(2.7) = 0.715\,849$ , $e_{F_2}(2.84) = 0.701\,676$ , and $e_{F_2}(3.1) = 0.704\,291$ clearly indicate that $F_2$ does not have the IDMRL property. A plot of the mean residual life (MRL) function $e_{F_2}(x)$ of $F_2$ is given in Figure 1.

Figure 1. Mean residual life function of $F_2$ .

Example 4.2. [Reference Anis4] considered an IDMRL distribution with change point 2 given by

\begin{equation*} F_3(x)= \begin{cases} 1 - (1+x)^{-2}, & 0 < x \leq 1 , \\[3pt] 1 - 4x^{-2}{\mathrm{e}^{(1-x)/2x}}, & 1 \leq x < 2 , \\[3pt] 1 - ({18-x})/({2^8 \mathrm{e}^{1/4}}), & 2 \leq x < 10 , \\[3pt] 1 - {\big(6-\frac{x}{5}\big)^4}/({2^{13} \mathrm{e}^{1/4}}), & 10 \leq x < 25 , \\[3pt] 1 - {2^{-13}}{\mathrm{e}^{3/4}}\exp[x-25-\mathrm{e}^{x-25}], & x \geq 25 . \end{cases} \end{equation*}

Let $r_{F_3}(t)$ be the FRA function of $F_3$ . The fact that $F_3$ is not BFRA is clear from the values $r_{F_3}(0.5) = 1.621\,86$ , $r_{F_3}(1) = 1.386\,29$ , $r_{F_3}(1.5) = 1.575\,93$ , and $r_{F_3}(3) = 1.029\,04$ . Figure 2 displays the FRA function of $F_3$ .

Figure 2. Failure rate average function of $F_3$ .

Examples 4.1 and 4.2 show that neither of the BFRA and IDMRL classes is contained in the other. Next, we investigate the connection between BFRA and NWBUE distributions.

Example 4.3. Consider the distribution function given by

\begin{equation*} F_4(x)= \begin{cases} 1 - ({1+2 x})^{-2}, & 0 \leq x < 2 , \\[3pt] \frac{24}{25}, & 2 \leq x < 3 , \\[3pt] 1 - \frac{9}{25}(2 x-3)^{-2}, & 3 \leq x < 4 , \\[3pt] 1 - \frac{9}{2500}x\exp[{({16-x^2})/{20}}], & x \geq 4. \end{cases} \end{equation*}

The MRL function of $F_4$ , given by

\begin{equation*} e_{F_4}(x)= \begin{cases} \frac{1}{2}({1+2 x}), & 0 \leq x < 2 , \\[5pt] \frac{1}{2}(9-2 x), & 2 \leq x < 3 , \\[5pt] \frac{1}{2}({x-3}), & 3 \leq x < 4 , \\[5pt] {10}{x}^{-1}, & x \geq 4, \end{cases} \end{equation*}

shows that $F_4$ is NWBUE with change point $t_0^\prime = 20$ in the NWBUE sense. The values $r_{F_4}(3) = 1.072\,96$ , $r_{F_4}(3.4) = 1.085\,78$ , and $r_{F_4}(4) = 1.060\,13$ illustrate that $F_4$ does not possess the BFRA property. A plot of the FRA function $r_{F_4}$ of $F_4$ is given in Figure 3.

Figure 3. Failure rate average function of $F_4$ .

Remark 4.1. Examples 4.1 and 4.2 show that neither of the BFRA and IDMRL classes is contained in the other. Moreover, Theorem 4.1 and [Reference Khan, Bhattacharyya and Mitra32, Theorem 16] show that BFRA and IDMRL classes of distributions contain all BFR distributions.

Remark 4.2. Theorem 4.1 and [Reference Mitra and Basu47, Proposition 2.1] show that BFRA and NWBUE classes of distributions contain the BFR family of distributions. Now, from Example 4.3 we can conclude that either $\{\mathrm{BFRA}\} \subset \{\mathrm{NWBUE}\}$ or neither of the classes is contained in the other.

Remark 4.3. Theorem 4.1 and [Reference Izadi, Sharafi and Khaledi27, Theorem 4] show that both BFRA and IDMTTF classes of distributions contain the BFR family.

From the remarks above it can be easily seen that the BFRA, IDMRL, NWBUE, and IDMTTF classes of distributions contain BFR distributions. Moreover, we are unable to establish any clear connection between IDMTTF and BFRA distributions. But at this juncture we would like to pose an open problem:

(4.1) \begin{equation}{If \, a \, life \, distribution} \, F \, {is \, BFRA}(x_0) \, {then} \, F \, {is \, IDMTTF}(t_0), \, {where} \, t_0 \geq x_0.\end{equation}

In the next theorem we have tried to provide a partial answer to (4.1) by following an argument similar to [Reference Neath and Samaniego52] that attempts to show that if F is IFRA then $\tau(u)/u$ is decreasing in u, where

(4.2) \begin{equation} \tau(u)=\int_0^{F^{-1}(u)}\bar{F}(x) \, \mathrm{d} x, \qquad u\in[0,1] ,\end{equation}

is the total-time-on-test (TTT) transform of F, and $F^{-1}(u)=\inf\lbrace x \colon F(x) \geq u \rbrace$ .

Theorem 4.2. Suppose an absolutely continuous lifetime distribution F is BFRA $(x_0)$ . Then the mean time to failure (MTTF) function is increasing on $(0,x_0]$ .

Proof. In order to prove this theorem we first assume that F is absolutely continuous with respect to Lebesgue measure, and later we extend the argument to the ‘continuous’ case. Note that the MTTF function $M_F(t)$ of F can be written as

\begin{align*} M_F(t) = \frac{\int_0^t \bar{F}(x) \, \mathrm{d} x}{\int_0^t \lambda(x)\bar{F}(x) \, \mathrm{d} x}, \end{align*}

since F is absolutely continuous with respect to Lebesgue measure. Now the numerator of $\frac{\mathrm{d}}{\mathrm{d} t} M_F(t)$ is given by

\begin{equation*} Q(t) = \bar{F}(t) \int_0^t \lambda(x)\bar{F}(x) \, \mathrm{d} x - \lambda(t)\bar{F}(t) \int_0^t \bar{F}(x) \, \mathrm{d} x \nonumber = \bar{F}(t) \int_0^t [ \lambda(x) - \lambda(t) ] \bar{F}(x) \, \mathrm{d} x . \nonumber \end{equation*}

Thus, to prove this theorem it is enough to show that $S(t)=\int_0^t[ \lambda(x) - \lambda(t)] \bar{F}(x) \, \mathrm{d} x \geq 0$ for $t \leq x_0$ . Using integration by parts, S(t) can be written as

\begin{equation*} S(t) = \bigg[\frac{-\ln{\bar{F}(t)}}{t} - \lambda(t) \bigg]t\bar{F}(t) + \int_0^t\bigg[\frac{-\ln{\bar{F}(x)}}{x} - \lambda(t) \bigg] x \, \mathrm{d} \bar{F}(x) . \end{equation*}

Note that, for all $t\leq x_0$ ,

\begin{equation*} S(t) \geq \bigg[\frac{-\ln{\bar{F}(t)}}{t} - \lambda(t) \bigg] \bigg[ t\bar{F}(t) + \int_0^t x \, \mathrm{d} \bar{F}(x) \bigg] \nonumber \geq 0 , \end{equation*}

since $({-\!\ln{\bar{F}(t)}})/{t}$ is decreasing on $(0, x_0]$ . Thus, $M_F(t)$ is increasing on $(0,x_0]$ .

At this stage, the following hierarchy represents the updated interrelationships among the non-monotonic ageing classes.

5. Discussion

The main contributions of this paper center around the homogeneous Poisson shock model in the framework of BFRA distributions, weak convergence issues within the BFRA class, and interrelationships among non-monotonic ageing classes. However, there is substantial scope for future work, for example, non-homogeneous Poisson shock models and pure birth shock models for the BFRA class remain to be explored. Further, in this scenario we can also consider shock models in a more general setup where failure occurs due to shocks in the presence of continuous wear and tear.

In Section 3 we posed the open problem (4.1). The TTT transform, defined in (4.2), plays an important role in characterizing ageing classes of life distributions (see [Reference Klefsjö36]). Thus, (4.1) has the following equivalent formulation in terms of the TTT transform: ‘If F is BFRA then there exists $\tilde{u}\in [0,1]$ such that $\tau(x)/x$ is increasing on $[0,\tilde{u}]$ and decreasing on $(\tilde{u},1]$ .’ In this regard, the following comments would be in order. We believe that the result contained in (4.1) is true. In fact, the distribution in Example 4.1 (due to [Reference Bhattacharyya, Ghosh and Mitra12]) is BFRA with change point $2.320\,84$ and IDMTTF with change point $2.404\,25$ $(\!>\!2.320\,84)$ , as can be seen from Figure 4.

Figure 4. Mean time to failure function of $F_2$ .

If, indeed, that result does hold, then the hierarchy would turn out as follows:

Acknowledgements

The authors would like to thank the reviewers and concerned Editor for their valuable comments and suggestions, which have led to a substantial improvement of an earlier version of this paper.

Funding information

There are no funding bodies to thank relating to the creation of this article.

Competing interests

There were no competing interests to declare which arose during the preparation or publication process of this article.

References

Aarset, M. V. (1987). How to identify a bathtub hazard rate. IEEE Trans. Reliab. 36, 106108.CrossRefGoogle Scholar
Abouammoh, A., Hindi, M. and Ahmed, A. (1988). Shock models with NBUFR and NBAFR survivals. Trabajos de Estadistica 3, 97.CrossRefGoogle Scholar
Anderson, K. K. (1987). Limit theorems for general shock models with infinite mean intershock times. J. Appl. Prob. 24, 449456.CrossRefGoogle Scholar
Anis, M. (2012). On some properties of the IDMRL class of life distributions. J. Statist. Planning Infer. 142, 30473055.CrossRefGoogle Scholar
Bandyopadhyay, D. and Basu, A. P. (1989). A note on tests for exponentiality by Deshpande. Biometrika 76, 403405.10.1093/biomet/76.2.403CrossRefGoogle Scholar
Barlow, R. E. (1979). Geometry of the total time on test transform. Naval Res. Logistics Quart. 26, 393402.CrossRefGoogle Scholar
Barlow, R. E. and Proschan, F. (1965). Mathematical Theory of Reliability. Wiley, New York.Google Scholar
Barlow, R. E. and Proschan, F. (1981). Statistical Theory of Reliability and Life Testing. To Begin With, Silver Spring, MD.Google Scholar
Basu, S. K. and Bhattacharjee, M. C. (1984). On weak convergence within the HNBUE family of life distributions. J. Appl. Prob. 21, 654660.CrossRefGoogle Scholar
Basu, S. K. and Simons, G. (1982). Moment spaces for IFR distributions: Applications and related material. In Contributions to Statistics. Essays in Honour of Norman L. Johnson, ed Sen, P. K.. Holland, North, Amsterdam.Google Scholar
Bennett, S. (1983). Log-logistic regression models for survival data. J. R. Statist. Soc. C [Appl. Statist.] 32, 165171.Google Scholar
Bhattacharyya, D., Ghosh, S. and Mitra, M. (2020). On a non-monotonic ageing class based on the failure rate average. Commun. Statist. Theory Meth. 51, 48074826.CrossRefGoogle Scholar
Bhattacharyya, D., Khan, R. A. and Mitra, M. (2020). A test of exponentiality against DMTTF alternatives via L-statistics. Statist. Prob. Lett. 165, 108853.CrossRefGoogle Scholar
Bhattacharyya, D., Khan, R. A. and Mitra, M. (2021). A goodness of fit test for mean time to failure function in age replacement. J. Statist. Comput. Simul. 91, 36373652.CrossRefGoogle Scholar
Bhattacharyya, D., Khan, R. A. and Mitra, M. (2021). Two-sample nonparametric test for comparing mean time to failure functions in age replacement. J. Statist. Planning Infer. 212, 3444.CrossRefGoogle Scholar
Birnbaum, Z. W., Esary, J. D. and Marshall, A. (1966). A stochastic characterization of wear-out for components and systems. Ann. Math. Statist. 37. 816–825.CrossRefGoogle Scholar
Boland, P. J. and Proschan, F. (1983). Optimum replacement of a system subject to shocks. Operat. Res. 31, 697704.CrossRefGoogle Scholar
Deshpande, J. V., Kochar, S. C. and Singh, H. (1986). Aspects of positive ageing. J. Appl. Prob. 23. 748–758.CrossRefGoogle Scholar
Ebrahimi, N. (1999). Stochastic properties of a cumulative damage threshold crossing model. J. Appl. Prob. 36, 720732.CrossRefGoogle Scholar
El-Neweihi, E., Proschan, F. and Sethuraman, J. (1983). A multivariate new better than used class derived from a shock model. Operat. Res. 31, 177183.CrossRefGoogle Scholar
Esary, J. D. and Marshall, A. (1974). Families of components, and systems, exposed to a compound Poisson damage process. In Reliability and Biometry: Statistical Analysis of Life Length, eds F. Proschan and R. J. Serfling. SIAM, Philadelphia, PA, pp. 3146.Google Scholar
Esary, J., Marshall, A. and Proschan, F. (1973). Shock models and wear processes. Ann. Prob. 1, 627649.CrossRefGoogle Scholar
Glaser, R. E. (1980). Bathtub and related failure rate characterizations. J. Amer. Statist. Assoc. 75, 667672.CrossRefGoogle Scholar
Guess, F., Hollander, M. and Proschan, F. (1986). Testing exponentiality versus a trend change in mean residual life. Ann. Statist. 14, 13881398.10.1214/aos/1176350165CrossRefGoogle Scholar
Gupta, R. C. and Warren, R. (2001). Determination of change points of non-monotonic failure rates. Commun. Statist. Theory Meth. 30, 19031920.CrossRefGoogle Scholar
Izadi, M. and Manesh, S. F. (2019). Testing exponentiality against a trend change in mean time to failure in age replacement. Commun. Statist. Theory Meth. 50, 33583370.CrossRefGoogle Scholar
Izadi, M., Sharafi, M. and Khaledi, B.-E. (2018). New nonparametric classes of distributions in terms of mean time to failure in age replacement. J. Appl. Prob. 55, 12381248.CrossRefGoogle Scholar
Joe, H. and Proschan, F. (1984). Percentile residual life functions. Operat. Res. 32, 668678.CrossRefGoogle Scholar
Karlin, S. (1968). Total Positivity, Vol. 1. Stanford University Press.Google Scholar
Khan, R. A., Bhattacharyya, D. and Mitra, M. (2020). A change point estimation problem related to age replacement policies. Operat. Res. Lett. 48, 105108.CrossRefGoogle Scholar
Khan, R. A., Bhattacharyya, D. and Mitra, M. (2021). Exact and asymptotic tests of exponentiality against nonmonotonic mean time to failure type alternatives. Statist. Papers 62, 30153045.CrossRefGoogle Scholar
Khan, R. A., Bhattacharyya, D. and Mitra, M. (2021). On classes of life distributions based on the mean time to failure function. J. Appl. Prob. 58, 289313.CrossRefGoogle Scholar
Khan, R. A., Bhattacharyya, D. and Mitra, M. (2021). On some properties of the mean inactivity time function. Statist. Prob. Lett. 170, 108993.CrossRefGoogle Scholar
Khan, R. A. and Mitra, M. (2019). Sharp bounds for survival probability when ageing is not monotone. Prob. Eng. Inf. Sci. 33, 205219.CrossRefGoogle Scholar
Klefsjö, B. (1981). HNBUE survival under some shock models. Scand. J. Statist., 8, 3947.Google Scholar
Klefsjö, B. (1982). On aging properties and total time on test transforms. Scand. J. Statist., 9, 3741.Google Scholar
Klefsjö, B. (1983). Some tests against aging based on the total time on test transform. Commun. Statist. Theory Meth. 12, 907927.CrossRefGoogle Scholar
Klefsjö, B. (1983). A useful ageing property based on the Laplace transform. J. Appl. Prob. 20, 615626.CrossRefGoogle Scholar
Kochar, S. C. (1985). Testing exponentiality against monotone failure rate average. Commun. Statist. Theory Meth. 14, 381392.CrossRefGoogle Scholar
Kochar, S. C. (1990) On preservation of some partial orderings under shock models. Adv. Appl. Prob. 22, 508509.CrossRefGoogle Scholar
Lai, C. D. and Xie, M. (2006). Stochastic Ageing and Dependence for Reliability. Springer, New York.Google Scholar
Langlands, A. O., Pocock, S. J., Kerr, G. and Gore, S. M. (1979). Long term survival of patients with breast cancer: A study of the curability of the disease. British Med. J., 2, 12471251.CrossRefGoogle ScholarPubMed
Link, W. A. (1989). Testing for exponentiality against monotone failure rate average alternatives. Commun. Statist. Theory Meth. 18. 3009–3017.CrossRefGoogle Scholar
Loève, M. (1963). Probability Theory, 3rd edn. Van Nostrand, New York.Google Scholar
Mi, J. (1993). Discrete bathtub failure rate and upside-down bathtub mean residual life. Naval Res. Logistics 40, 361371.10.1002/1520-6750(199304)40:3<361::AID-NAV3220400306>3.0.CO;2-I3.0.CO;2-I>CrossRefGoogle Scholar
Mi, J. (1995). Bathtub failure rate and upside-down bathtub mean residual life. IEEE Trans. Reliab. 44, 388391.Google Scholar
Mitra, M. and Basu, S. K. (1994). On a nonparametric family of life distributions and its dual. J. Statist. Planning Infer. 39, 385397.CrossRefGoogle Scholar
Mitra, M. and Basu, S. K. (1996). On some properties of the bathtub failure rate family of life distributions. Microelectron. Reliab. 36, 679684.CrossRefGoogle Scholar
Mitra, M. and Basu, S. K. (1996). Shock models leading to non-monotonic ageing classes of life distributions. J. Statist. Planning Infer. 55, 131138.CrossRefGoogle Scholar
Mitra, M. and Khan, R. A. (2021). Reliability shock models: A brief excursion. In Applied Advanced Analytics, ed A. K. Laha. Springer, Singapore, pp. 1942.CrossRefGoogle Scholar
Nakagawa, T. (2007). Shock and Damage Models in Reliability Theory. Springer, New York.Google Scholar
Neath, A. A. and Samaniego, F. J. (1992). On the total time on test transform of an IFRA distribution. Statist. Prob. Lett. 14, 289291.CrossRefGoogle Scholar
Pellerey, F. (1993). Partial orderings under cumulative damage shock models. Adv. Appl. Prob. 25, 939946.CrossRefGoogle Scholar
Prentice, R. L. (1973). Exponential survivals with censoring and explanatory variables. Biometrika 60, 279288.CrossRefGoogle Scholar
Shanthikumar, J. G. and Sumita, U. (1983). General shock models associated with correlated renewal sequences. J. Appl. Prob. 20, 600614.CrossRefGoogle Scholar
Singpurwalla, N. D. (2006). The hazard potential: Introduction and overview. J. Amer. Statist. Assoc. 101, 17051717.CrossRefGoogle Scholar
Tiwari, R. C., Rao Jammalamadaka, S. and Zalkikar, J. N. (1989). Testing an increasing failure rate average distribution with censored data. Statistics 20, 279286.Google Scholar
Wang, F. (2000). A new model with bathtub-shaped failure rate using an additive Burr XII distribution. Reliab. Eng. Syst. Saf. 70, 305312.CrossRefGoogle Scholar
Wells, M. T. and Tiwari, R. C. (1991). A class of tests for testing an increasing failure-rate-average distribution with randomly right-censored data. IEEE Trans. Reliab. 40, 152156.CrossRefGoogle Scholar
Figure 0

Figure 1. Mean residual life function of $F_2$.

Figure 1

Figure 2. Failure rate average function of $F_3$.

Figure 2

Figure 3. Failure rate average function of $F_4$.

Figure 3

Figure 4. Mean time to failure function of $F_2$.