Next Article in Journal
Symmetric and Asymmetric Crisscross Augmented Ladder Multilevel Inverter with Fuzzy Logic Controller Optimization
Next Article in Special Issue
A New Asymmetric Modified Topp–Leone Distribution: Classical and Bayesian Estimations under Progressive Type-II Censored Data with Applications
Previous Article in Journal
A Soliton Solution for the Kadomtsev–Petviashvili Model Using Two Novel Schemes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Generalized Fiducial Inference for the Stress–Strength Reliability of Generalized Logistic Distribution

1
School of Mathematics and Statistics, Hebei University of Economics and Business, Shijiazhuang 050061, China
2
School of Science, Hebei University of Science and Technology, Shijiazhuang 050018, China
3
Department of Science, Karume Institute of Science and Technology, Zanzibar P.O. Box 467, Tanzania
*
Author to whom correspondence should be addressed.
Symmetry 2023, 15(7), 1365; https://doi.org/10.3390/sym15071365
Submission received: 14 June 2023 / Revised: 1 July 2023 / Accepted: 3 July 2023 / Published: 5 July 2023
(This article belongs to the Special Issue Symmetry in Probability Theory and Statistics)

Abstract

:
Generalized logistic distribution, as the generalized form of the symmetric logistic distribution, plays an important role in reliability analysis. This article focuses on the statistical inference for the stress–strength parameter R = P ( Y < X ) of the generalized logistic distribution with the same and different scale parameters. Firstly, we use the frequentist method to construct asymptotic confidence intervals, and adopt the generalized inference method for constructing the generalized point estimators as well as the generalized confidence intervals. Then the generalized fiducial method is applied to construct the fiducial point estimators and the fiducial confidence intervals. Simulation results demonstrate that the generalized fiducial method outperforms other methods in terms of the mean square error, average length, and empirical coverage. Finally, three real datasets are used to illustrate the proposed methods.

1. Introduction

The stress–strength, which was initially proposed by Birnbaum [1] and developed by Birnbaum and McCarty [2], plays an important role in reliability analysis. For two independent random variables, X and Y, the stress–strength parameter is defined as R = P ( Y < X ) . If stress Y is greater than strength X, it may result in component failure or system malfunction. The stress–strength parameter is originally used in the industrial field to calculate the reliability of the products [3,4]. It is also increasingly used to estimate the probability that one variable exceeds another [5,6], which is of great significance in practical application and has been widely used in various fields, such as electrical cable failure analysis, leukemia treatment, and jute fiber testing. See more details for [5,7,8,9,10].
In the literature, there are many life distributions that can be used to estimate R, such as Weibull [5], Pareto [6,11], generalized Pareto [12], exponential [8,13], generalized exponential [14], Lomax [15], unit-half-normal [16], unit-Gompertz [17], and generalized logistic (GL) [18,19,20,21] distributions. The logistic distribution is a symmetric heavy-tailed distribution. However, it is not suitable for handling asymmetric or thin-tailed data. Therefore, it is necessary to further extend the logistic distribution according to practical problems, which can handle the data including symmetric, heavy-tailed, asymmetric, and thin-tailed. The GL distribution, as defined by Balakrishnan and Leung [22], is one of the generalized forms of the standard logistic distribution. By introducing a shape parameter to the distribution, the GL distribution expands the range of values for the skewness coefficient and tail index, which allows a wider range of data fitting capabilities. It has attracted extensive attention and is widely used in various fields, including demography, biology, finance, and neural network, as detailed in [23]. Therefore, we select the GL distribution with the following probability density function (PDF)
f ( x ; σ , α ) = α e x σ σ ( 1 + e x σ ) α + 1 , < x < ,
and the corresponding cumulative distribution function (CDF) is
F ( x ; σ , α ) = 1 ( 1 + e x σ ) α , < x < ,
where σ > 0 and α > 0 are the scale and shape parameters, respectively. The GL distribution exhibits a negative skew when α > 1 and a positive skew when 0 < α < 1 , and it becomes the standard logistic distribution (it is symmetric) when α = 1 . Meanwhile, the PDF of the GL distribution is unimodal and log-concave, making it suitable for modeling data with both left and right skewness [18]. The expectation and variance of X can be calculated from the moment-generating function of the GL distribution [24]; that is,
E ( X ) = σ ( ψ ( α ) ψ ( 1 ) ) and V a r ( X ) = σ 2 ( ψ ( 1 ) + ψ ( α ) ) ,
where ψ ( z ) = Γ ( z ) / Γ ( z ) is the digamma function and ψ ( z ) = d ψ ( z ) / d z is the trigamma function, with the gamma function Γ ( z ) = 0 x z 1 e x d x for z > 0 . From Formula (3), the coefficient of skewness for X, corresponding to the third standardized moment, is expressed as
S k e w ( X ) = ψ ( α ) ψ ( 1 ) ( ψ ( α ) + ψ ( 1 ) ) 3 / 2 ,
which implies that the expression does not depend on the parameter σ .
Statisticians have conducted numerous kinds of research on R based on the GL distribution, most focus on frequentist and Bayesian inference. For the single component, Asgharzadeh et al. [18], Babayi et al. [19], and Okasha [20] considered the estimation almost at the same time. Asgharzadeh et al. [18] considered the estimation of R for GL distribution under three different cases, and obtained the estimators and confidence intervals based on maximum likelihood (ML), bootstrap, and Bayesian methods. Babayi et al. [19] used ML and Bayes methods to obtain the point estimations and confidence intervals of R for GL distribution with the same and different scale parameters. When the scale parameters were the same, Okasha [20] obtained the point and interval estimations of R using ML and Bayes methods. For the multicomponent stress–strength reliability, Rasekhi et al. [21] discussed the point and interval estimations under Bayesian and ML methods.
Based on the above research, it was found that the empirical coverage of ML estimation sometimes fails to reach the nominal level, while the choice of the prior distribution is improper or subjective in Bayesian inference. Furthermore, Tao [25] stated that the Jeffreys prior and reference prior are improper in the GL distribution, which leads to the improper posterior distribution of the parameter. When the exact pivotal quantity is not available, the generalized inference (GI) proposed by Weerahandi [26] provides us with another way of thinking, and Wang et al. [27] have successfully estimated the R of the generalized exponential distribution based on the GI method. Moreover, Hannig et al. [28] stated that the posterior of generalized fiducial distribution (GFD) is always proper and the confidence intervals of generalized fiducial inference (GFI) intend to maintain stated coverage (or be conservative) while having an average length comparable to or shorter than other methods. Yan and Liu [29], Yan et al. [30], and Cai et al. [31] used the above fiducial approach to consider the estimation of the parameters of the generalized exponential distribution, Lomax distribution, and Weibull distribution, respectively, where GFI often provides better estimation results than the traditional methods. See [32,33] for more applications of the GFI method. For the above reasons, the research objective of this article is to find a more appropriate method among the existing methods to estimate the stress–strength of the GL distribution with the same and different scale parameters. Our original contribution is mainly to introduce the GI and GFI methods to the estimation of R and compare their performance with the frequentist method. Furthermore, we show the advantages of the GFI method in terms of mean square error, average length, and empirical coverage.
The structure of the rest paper is as follows. For the hypothesis of the same and different scale parameters, Section 2 and Section 3 develop the point and interval estimations of R based on the ML, GI, and GFI methods. Section 4 simulates and compares the above methods. Section 5 demonstrates the proposed estimations by providing three real data examples. The implications of our findings are discussed in Section 6. The conclusions based on the research results are drawn in Section 7.

2. Estimation of R with the Same Scale and Different Shape Parameters

Suppose X G L ( σ , α 1 ) and Y G L ( σ , α 2 ) are independent random variables with the same scale parameter σ , then R = P ( Y < X ) can be calculated as follows
R = P ( Y < X ) = x f Y ( y ) f X ( x ) d y d x = 1 ( 1 + e x σ ) α 2 · α 1 e x σ σ ( 1 + e x σ ) α 1 + 1 d x = α 1 α 1 + α 2 .

2.1. Maximum Likelihood Estimation of R

Given the observed data, x = ( x 1 , , x n ) T and y = ( y 1 , , y m ) T , the log-likelihood function of GL distribution is
L ( σ , α 1 , α 2 | x , y ) = n log α 1 + m log α 2 ( n + m ) log σ i = 1 n x i + j = 1 m y j σ ( α 1 + 1 ) i = 1 n log ( 1 + e x i σ ) ( α 2 + 1 ) j = 1 m log ( 1 + e y j σ ) .
The corresponding ML estimators of σ , α 1 , and α 2 can be derived from
L σ = n + m σ + i = 1 n x i + j = 1 m y j σ 2
( α 1 + 1 ) i = 1 n x i e x i σ σ 2 ( 1 + e x i σ ) ( α 2 + 1 ) j = 1 m y j e y j σ σ 2 ( 1 + e y j σ ) = 0 ,
L α 1 = n α 1 i = 1 n log ( 1 + e x i σ ) = 0 ,
L α 2 = m α 2 j = 1 m log ( 1 + e y j σ ) = 0 .
From Formulas (8) and (9), the ML estimators of α 1 and α 2 as the functions of σ , say α ^ 1 ( σ ) and α ^ 2 ( σ ) , respectively, can be obtained as
α ^ 1 ( σ ) = n i = 1 n log ( 1 + e x i σ ) and α ^ 2 ( σ ) = m j = 1 m log ( 1 + e y j σ ) .
From Formula (7), the ML estimator of σ can be determined by the following nonlinear equation
h ( σ ) = σ ,
where
h ( σ ) = 1 n + m [ i = 1 n x i + j = 1 m y j ( n i = 1 n log ( 1 + e x i σ ) + 1 ) i = 1 n x i e x i σ 1 + e x i σ ( m j = 1 m log ( 1 + e y j σ ) + 1 ) j = 1 m y j e y j σ 1 + e y j σ ] .
Since σ ^ is a fixed point solution of Equation (11), it can be obtained by the iterative algorithm h ( σ ( k ) ) = σ ( k + 1 ) , where σ ( k ) is the kth iteration of σ ^ . The iteration procedure will stop when | σ ( k ) σ ( k + 1 ) | is small enough. Substituting σ ^ into (10), we can have α ^ 1 and α ^ 2 . Accordingly, the ML estimator of R is
R ^ M L = α ^ 1 α ^ 1 + α ^ 2 ,
where α ^ 1 and α ^ 2 are the ML estimators of α 1 and α 2 , respectively.
The confidence interval of R can be derived by the following asymptotic distribution; that is,
( σ ^ , α ^ 1 , α ^ 2 ) T L N ( σ , α 1 , α 2 ) T , I 0 1 ,
where I 0 is the observed Fisher information matrix, i.e.,
I 0 1 = 2 L σ 2 2 L σ α 1 2 L σ α 2 2 L α 1 σ 2 L α 1 2 2 L α 1 α 2 2 L α 2 σ 2 L α 2 α 1 2 L α 2 2 | ( σ , α 1 , α 2 ) T = ( σ ^ , α ^ 1 , α ^ 2 ) T 1 V a r ( σ ^ ) C o v ( σ ^ , α ^ 1 ) C o v ( σ ^ , α ^ 2 ) C o v ( α ^ 1 , σ ^ ) V a r ( α ^ 1 ) C o v ( α ^ 1 , α ^ 2 ) C o v ( α ^ 2 , σ ^ ) C o v ( α ^ 2 , α ^ 1 ) V a r ( α ^ 2 ) .
By using the Delta method, the asymptotic variance of R ^ M L is given by
V a r ( R ^ M L ) = R σ , R α 1 , R α 2 I 0 1 R σ , R α 1 , R α 2 T .
Consequently, the 100 ( 1 γ ) % asymptotic confidence interval of R is
R ^ M L z 1 γ / 2 V a r ( R ^ M L ) R R ^ M L + z 1 γ / 2 V a r ( R ^ M L ) ,
where z 1 γ / 2 is the 1 γ / 2 quantile of standard normal distribution.

2.2. Generalized Inference of R

Since Wang et al. [27] successfully estimated the generalized exponential distribution by the GI method, we introduce the GI method to estimate R under GL distribution, which is formally similar to the generalized exponential distribution.
Lemma 1 
(Wang et al. [27] and Yu et al. [34]). Let Z 1 , , Z n be a random sample from the exponential distribution with mean θ and Z ( 1 ) < Z ( 2 ) < ··· < Z ( n ) be the corresponding order statistics. Let
S i = j = 1 i Z ( j ) + ( n i ) Z ( i ) , i = 1 , , n , T = 2 i = 1 n 1 log ( S n / S i ) .
Then ( 1 ) T and S n are independent; ( 2 ) T χ 2 ( 2 n 2 ) and 2 S n / θ χ 2 ( 2 n ) .
Lemma 2. 
Let
f ( σ ) = log ( 1 + e b σ ) log ( 1 + e a σ ) ,
where b > a > 0 are constants. Thus, f ( σ ) is strictly decreasing on ( 0 , + ) .
Proof of Lemma 2. 
From the function
f ( σ ) = log ( 1 + e b σ ) log ( 1 + e a σ ) , b > a > 0 ,
we can calculate that
f ( σ ) = b ( e a σ + 1 ) log ( 1 + e a σ ) a ( e b σ + 1 ) log ( 1 + e b σ ) σ 2 ( e a σ + 1 ) ( e b σ + 1 ) ( log ( 1 + e a σ ) ) 2 .
It is obvious that the denominator of f ( σ ) is greater than 0, so we mainly focus on the numerator. Let
g ( σ ) = b ( e a σ + 1 ) log ( 1 + e a σ ) a ( e b σ + 1 ) log ( 1 + e b σ ) ,
then
g ( σ ) = a b σ 2 log ( 1 + e b σ ) e b σ log ( 1 + e a σ ) e a σ .
Because log ( 1 + x ) x is strictly decreasing on ( 0 , + ) and the e x σ is strictly increasing in x > 0 for σ > 0 , we have that log ( 1 + e x σ ) e x σ is strictly decreasing in x > 0 for σ > 0 . Thus g ( σ ) < 0 on ( 0 , + ) and g ( σ ) is strictly decreasing on ( 0 , + ) . Therefore, for σ > 0 , we can obtain
g ( σ ) < lim σ 0 + g ( σ ) = 0 .
Finally, we have that f ( σ ) < 0 on ( 0 , + ) ; thus, f ( σ ) is strictly decreasing on ( 0 , + ) . □
Let X 1 , , X n be a random sample from G L ( σ , α 1 ) and X = ( X ( 1 ) , , X ( n ) ) T be the corresponding order statistics. If a random variable X follows the standard uniform distribution, then log X follows the standard exponential distribution E x p ( 1 ) . Obviously, we can find that ( 1 + e X ( 1 ) σ ) α 1 ,…, ( 1 + e X ( n ) σ ) α 1 are the order statistics from the standard uniform distribution. Thus, Z ( i ) = log ( 1 + e X ( n i + 1 ) σ ) , i = 1 , , n are the order statistics from the exponential distribution with mean 1 / α 1 . Similarly, we can easily obtain the order statistics Z ( j ) = log ( 1 + e Y ( m j + 1 ) σ ) , j = 1 , , m from the exponential distribution with mean 1 / α 2 , where the random sample Y 1 , , Y m follows G L ( σ , α 2 ) . Let
T ( σ ) = 2 i = 1 n 1 log ( S n / S i ) + j = 1 m 1 log ( S m / S j ) ,
where S i = u = 1 i Z ( u ) + ( n i ) Z ( i ) and S j = v = 1 j Z ( v ) + ( m j ) Z ( j ) . Then from Lemma 1, we have T ( σ ) χ 2 ( 2 n + 2 m 4 ) . Together, Lemma 2 with
S n S i = 1 + S n S i S i = 1 + Z ( i + 1 ) Z ( i ) + + Z ( n ) Z ( i ) ( n i ) Z ( 1 ) Z ( i ) + + Z ( i 1 ) Z ( i ) + ( n i + 1 ) ,
S m S j = 1 + S m S j S j = 1 + Z ( j + 1 ) Z ( j ) + + Z ( m ) Z ( j ) ( m j ) Z ( 1 ) Z ( j ) + + Z ( j 1 ) Z ( j ) + ( m j + 1 ) ,
T ( σ ) is strictly increasing on ( 0 , + ) . Notice that
lim σ 0 + T ( σ ) = 0 and lim σ + T ( σ ) = + .
Therefore, equation T ( σ ) = T has the unique solution g ( T , X , Y ) when n and m are given. The solution of equation T ( σ ) = T can be obtained by the bisection method.
According to Lemma 1, we find that U 1 = 2 α 1 S n χ 2 ( 2 n ) , then α 1 = U 1 / ( 2 S n ) . α 2 = U 2 / ( 2 S m ) can be calculated in the same way, so the generalized pivotal quantity is
R G I = U 2 / s m U 1 / s n + U 2 / s m ,
where s n = i = 1 n log 1 + e x ( i ) / g ( T , x , y ) , s m = j = 1 m log 1 + e y ( j ) / g ( T , x , y ) , x = ( x ( 1 ) , , x ( n ) ) T and y = ( y ( 1 ) , , y ( m ) ) T are the observed values of X = ( X ( 1 ) , , X ( n ) ) T and Y = ( Y ( 1 ) , , Y ( m ) ) T , respectively. The generalized point estimation and generalized confidence interval of R can be obtained by using the following algorithm.
  • Generate a realization t of T from χ 2 ( 2 n + 2 m 4 ) . Then for given samples x and y , one can obtain a realization of g ( T , x , y ) from the equation T ( σ ) = t .
  • Derive a realization of U 1 and U 2 from χ 2 ( 2 n ) and χ 2 ( 2 m ) , respectively. Compute R ^ G I ( 1 ) on the basis of (20).
  • Perform Step 1 and Step 2 for N times, iteratively. The value of N is equal to 1000.
  • The generalized point estimator of R is R ^ G I = 1 N l = 1 N R ^ G I ( l ) . If R ^ G I , γ / 2 and R ^ G I , 1 γ / 2 denote the γ / 2 and 1 γ / 2 percentile of R ^ G I , the generalized confidence interval of R is [ R ^ G I , γ / 2 , R ^ G I , 1 γ / 2 ] .

2.3. Generalized Fiducial Inference of R

Let the data-generating equation be
x = G ( U , θ ) ,
where x denotes the data, θ is the parameter vector, and U is a random vector with 0–1 uniform distribution U ( 0 , 1 ) in each dimension. Under some differentiability conditions, Hannig et al. [28] provided a user-friendly formula to compute the GFD of θ , i.e.,
f F ( θ ) = f ( x | θ ) J ( x , θ ) f ( x | θ ) J ( x , θ ) d θ ,
where f ( x | θ ) denotes the joint density function of x and the function J ( x , θ ) is a Jacobian determinant. We usually take J ( x , θ ) as the infinite norm as follows
J ( x , θ ) = D e t d d θ G ( u , θ ) | u = G 1 ( x , θ ) .
In practice, Hannig et al. [28] recommended using D e t ( A ) = i = ( i 1 , , i p ) | d e t ( A ) i | and the above sum goes over n p of p-tuples of indexes i = ( 1 i 1 < < i p n ) . For any n × p matrix A, the sub-matrix ( A ) i is the p × p matrix consisting of the rows i = ( i 1 , , i p ) of A.
Regarding our concern, we have that
U i = F ( x i ; θ ) , i = 1 , , n ,
where θ = ( σ , α 1 ) T , F ( x i ; σ , α 1 ) ( 1 + e x i σ ) α 1 is the CDF of GL distribution and U i follows U ( 0 , 1 ) . According to (24), we can obtain the data generating equation x = G ( U , θ ) and the i-th x i = G i ( u i , θ ) is
x i = σ log ( U i 1 α 1 1 ) .
Then, we have
G i σ | u i = ( 1 + e x i σ ) α 1 = x i σ and G i α 1 | u i = ( 1 + e x i σ ) α 1 = σ α 1 ( 1 + e x i σ ) log ( 1 + e x i σ ) .
Substituting (26) into (23), it follows that
J ( x , σ , α 1 ) = 1 α 1 i j x i ( 1 + e x j σ ) log ( 1 + e x j σ ) x j ( 1 + e x i σ ) log ( 1 + e x i σ ) .
The function J ( y , σ , α 2 ) can be obtained through a similar process. Finally, we can derive the following GFD for ( σ , α 1 , α 2 ) ; that is,
f F ( σ , α 1 , α 2 | x , y ) = f ( x , y | σ , α 1 , α 2 ) J ( x , y , σ , α 1 , α 2 ) 0 0 0 f ( x , y | σ , α 1 , α 2 ) J ( x , y , σ , α 1 , α 2 ) d σ d α 1 d α 2 ,
where
f ( x , y | σ , α 1 , α 2 ) = f ( x | σ , α 1 ) f ( y | σ , α 2 ) = i = 1 n f i ( x i ; σ , α 1 ) j = 1 m f j ( y j ; σ , α 2 ) = i = 1 n α 1 e x i σ σ ( 1 + e x i σ ) α 1 + 1 · j = 1 m α 2 e y j σ σ ( 1 + e y j σ ) α 2 + 1 , J ( x , y , σ , α 1 , α 2 ) = w 1 J ( x , σ , α 1 ) + w 2 J ( y , σ , α 2 ) = n 2 n 2 + m 2 J ( x , σ , α 1 ) + m 2 n 2 + m 2 J ( y , σ , α 2 ) .
Specifically,
f F ( σ , α 1 , α 2 | x , y ) α 1 n e i = 1 n x i σ σ n i = 1 n ( 1 + e x i σ ) α 1 + 1 · α 2 m e j = 1 m y j σ σ m j = 1 m ( 1 + e y j σ ) α 2 + 1 × n 2 n 2 + m 2 · 1 α 1 i j | q ( x i , x j , σ ) | + m 2 n 2 + m 2 · 1 α 2 i j | q ( y i , y j , σ ) | ,
where
q ( x i , x j , σ ) = x i ( 1 + e x j σ ) log ( 1 + e x j σ ) x j ( 1 + e x i σ ) log ( 1 + e x i σ ) , q ( y i , y j , σ ) = y i ( 1 + e y j σ ) log ( 1 + e y j σ ) y j ( 1 + e y i σ ) log ( 1 + e y i σ ) .
On the one hand, the conditional fiducial density function of σ given α 1 and α 2 can be obtained from (29) and it is given by
f F ( σ | α 1 , α 2 , x , y ) e i = 1 n x i σ σ n i = 1 n ( 1 + e x i σ ) α 1 + 1 · e j = 1 m y j σ σ m j = 1 m ( 1 + e y j σ ) α 2 + 1 × n 2 n 2 + m 2 · i j | q ( x i , x j , σ ) | + m 2 n 2 + m 2 · i j | q ( y i , y j , σ ) | .
On the other hand, we can obtain
f F ( σ , α 1 | x ) = f ( x | σ , α 1 ) J ( x , σ , α 1 ) 0 0 f ( x | σ , α 1 ) J ( x , σ , α 1 ) d σ d α 1 α 1 n e i = 1 n x i σ σ n i = 1 n ( 1 + e x i σ ) α 1 + 1 · 1 α 1 i j | q ( x i , x j , σ ) | .
Therefore, the conditional fiducial density functions of α 1 given σ can be obtained as
f F ( α 1 | σ , x ) α 1 n 1 e α 1 i = 1 n log ( 1 + e x i σ ) ,
similarly, the conditional fiducial density functions of α 2 given σ are
f F ( α 2 | σ , y ) α 2 m 1 e α 2 j = 1 m log ( 1 + e y j σ ) ,
which implies that the conditional density of α 1 and α 2 are G a n , i = 1 n log ( 1 + e x i σ ) and G a m , j = 1 m log ( 1 + e y j σ ) , respectively, where G a stands for the Gamma distribution.
Using the Gibbs sampler to estimate the GFD requires being able to sample from the full conditional distribution for each quantity involved, so this is the case for α 1 and α 2 , but not for σ . Consequently, we introduce the standard Metropolis–Hastings steps into the Gibbs sampler to update σ in (30) while updating α 1 and α 2 from their exact conditional distribution. To reduce the autocorrelation of the Monte Carlo Markov Chains, we introduce a thin parameter T, which is an integer specifying the number of steps between each saved sample. The detailed steps of the algorithm are as follows.
  • Obtain three starting values of σ ( 0 ) , α 1 ( 0 ) and α 2 ( 0 ) .
  • Let σ ( l ) , α 1 ( l ) , and α 2 ( l ) denote the values of the lth iteration. Sample a candidate σ ( l + 1 ) from f F ( σ | α 1 , α 2 , x , y ) by using the Metropolis–Hastings method [35]. Sample the candidate α 1 ( l + 1 ) and α 2 ( l + 1 ) from G a n , i = 1 n log ( 1 + e x i σ ) and G a m , j = 1 m log ( 1 + e y j σ ) , respectively. R ^ G F I ( l ) can be obtained by plugging the values of σ ( l ) , α 1 ( l ) and α 2 ( l ) into Formula (5).
  • Conduct Step 2 for M + T N times, iteratively, where M is the burn-in period. The values of M and N are both equal to 1000, and the value of T is equal to 10.
  • The generalized fiducial point estimator of R is R ^ G F I = 1 N l = M + 1 M + N R ^ G F I ( l ) . Select the N γ / 2 th and N ( 1 γ / 2 ) th of the permutation as R ^ G F I , γ / 2 and R ^ G F I , 1 γ / 2 , respectively. Then, the 100 ( 1 γ ) % fiducial confidence interval of R is [ R ^ G F I , γ / 2 , R ^ G F I , 1 γ / 2 ] .

3. Estimation of R with Different Scale and Shape Parameters

Suppose X G L ( σ 1 , α 1 ) and Y G L ( σ 2 , α 2 ) are independently distributed under different scale parameters, σ 1 and σ 2 , then R = P ( Y < X ) , it can be easily seen that
R = P ( Y < X ) = x f Y ( y ) f X ( x ) d y d x = 0 1 1 + ( t 1 α 1 1 ) σ 1 σ 2 α 2 d t .

3.1. Maximum Likelihood Estimation of R

Let x = ( x 1 , , x n ) T be a random sample from G L ( σ 1 , α 1 ) and let y = ( y 1 , , y m ) T be another independent random sample from G L ( σ 2 , α 2 ) . The log-likelihood function is
L ( σ 1 , σ 2 , α 1 , α 2 | x , y ) = n log α 1 + m log α 2 n log σ 1 m log σ 2 1 σ 1 i = 1 n x i 1 σ 2 j = 1 m y j ( α 1 + 1 ) i = 1 n log ( 1 + e x i σ 1 ) ( α 2 + 1 ) j = 1 m log ( 1 + e y j σ 2 ) .
Similarly, the ML estimators of α 1 as a function of σ 1 and α 2 as a function of σ 2 are
α ^ 1 ( σ 1 ) = n i = 1 n log ( 1 + e x i σ 1 ) and α ^ 2 ( σ 2 ) = m j = 1 m log ( 1 + e y j σ 2 ) .
The ML estimators of σ 1 and σ 2 can be solved from h 1 ( σ 1 ) = σ 1 and h 2 ( σ 2 ) = σ 2 . Therefore, the ML estimator of R is
R ^ M L = 0 1 1 + ( t 1 α ^ 1 1 ) σ ^ 1 σ ^ 2 α ^ 2 d t ,
and the asymptotic 100 ( 1 γ ) % confidence intervals of R can also be obtained.

3.2. Generalized Inference of R

Let
T 1 ( σ 1 ) = 2 i = 1 n 1 log ( S n / S i ) and T 2 ( σ 2 ) = 2 j = 1 m 1 log ( S m / S j ) .
Then we have T 1 ( σ 1 ) χ 2 ( 2 n 2 ) and T 2 ( σ 2 ) χ 2 ( 2 m 2 ) from Lemma 1. In addition, we can prove that T 1 ( σ 1 ) and T 2 ( σ 2 ) are strictly increasing on ( 0 , + ) , and
lim σ i 0 + T i ( σ i ) = 0 and lim σ i + T i ( σ i ) = + , i = 1 , 2 .
Furthermore, when T 1 ( σ 1 ) χ 2 ( 2 n 2 ) and T 2 ( σ 2 ) χ 2 ( 2 m 2 ) are given, both T 1 ( σ 1 ) = T 1 and T 2 ( σ 2 ) = T 2 have unique solutions denoted by σ 1 = g 1 ( T 1 , X ) and σ 2 = g 2 ( T 2 , Y ) .
Since U 1 = 2 α 1 S n χ 2 ( 2 n ) and U 2 = 2 α 2 S m χ 2 ( 2 m ) , we find that α 1 = U 1 / ( 2 S n ) and α 2 = U 2 / ( 2 S m ) . The generalized pivotal quantity of R is
R G I = 0 1 1 + ( t 2 s n U 1 1 ) g 1 ( T 1 , X ) g 2 ( T 2 , Y ) U 2 2 s m d t ,
where s n = i = 1 n log 1 + e x ( i ) / g 1 ( T 1 , x ) and s m = j = 1 m log 1 + e y ( j ) / g 2 ( T 2 , y ) . The steps to calculate the generalized point and interval estimations of R are similar to those in Section 2.2.

3.3. Generalized Fiducial Inference of R

For the observed value x = ( x 1 , , x n ) T , we have
U i = F ( x i ; η ) , i = 1 , , n ,
where η = ( σ 1 , α 1 ) T and F ( x i ; σ 1 , α 1 ) ( 1 + e x i σ 1 ) α 1 is the CDF of GL distribution. Based on (41), we can obtain the i-th x i = G i ( u i , θ ) is x i = σ 1 log ( U i 1 α 1 1 ) . Then, we have
G i σ 1 | u i = ( 1 + e x i σ 1 ) α 1 = x i σ 1 and G i α 1 | u i = ( 1 + e x i σ 1 ) α 1 = σ 1 α 1 ( 1 + e x i σ 1 ) log ( 1 + e x i σ 1 ) .
It can be calculated that
J ( x , σ 1 , α 1 ) = 1 α 1 i j x i ( 1 + e x j σ 1 ) log ( 1 + e x j σ 1 ) x j ( 1 + e x i σ 1 ) log ( 1 + e x i σ 1 ) .
Finally, we obtain the following GFD for ( σ 1 , α 1 ) , i.e.,
f F ( σ 1 , α 1 | x ) = f ( x | σ 1 , α 1 ) J ( x , σ 1 , α 1 ) 0 0 f ( x | σ 1 , α 1 ) J ( x , σ 1 , α 1 ) d σ 1 d α 1 ,
where f ( x | σ 1 , α 1 ) = i = 1 n f i ( x i ; σ 1 , α 1 ) . Specifically,
f F ( σ 1 , α 1 | x ) α 1 n e i = 1 n x i σ 1 σ 1 n i = 1 n ( 1 + e x i σ 1 ) α 1 + 1 · 1 α 1 i j | q ( x i , x j , σ 1 ) | ,
where
q ( x i , x j , σ 1 ) = x i ( 1 + e x j σ 1 ) log ( 1 + e x j σ 1 ) x j ( 1 + e x i σ 1 ) log ( 1 + e x i σ 1 ) .
From Formula (45), the conditional fiducial density function of σ 1 given α 1 is given by
f F ( σ 1 | α 1 , x ) σ 1 n e i = 1 n x i σ 1 ( α 1 + 1 ) i = 1 n log ( 1 + e x i σ 1 ) i = 1 n | q ( x i , x j , σ 1 ) | .
Then, the conditional fiducial density function of α 1 given σ 1 is
f F ( α 1 | σ 1 , x ) α 1 n 1 e α 1 i = 1 n log ( 1 + e x i σ 1 ) ,
which implies that the conditional density of α 1 is G a n , i = 1 n log ( 1 + e x i σ 1 ) . Based on the same method, the conditional fiducial density function of σ 2 and α 2 can be obtained as follows
f F ( σ 2 | α 2 , y ) σ 2 m e j = 1 m y j σ 2 ( α 2 + 1 ) j = 1 m log ( 1 + e y j σ 2 ) j = 1 m | q ( y i , y j , σ 2 ) | ,
f F ( α 2 | σ 2 , y ) α 2 m 1 e α 2 j = 1 m log ( 1 + e y j σ 2 ) ,
which means the conditional density of α 2 is G a m , j = 1 m log ( 1 + e y j σ 2 ) .
We still introduce standard Metropolis–Hastings steps into the Gibbs sampler to update σ 1 and σ 2 while updating α 1 and α 2 from their exact conditional distributions. The detailed steps of the algorithm are similar to Section 2.3.

4. Simulations

Let R ^ M L represent the ML estimators, R ^ G I denote the point estimators via the GI method, and R ^ G F I denotes the point estimators by the GFI method. ACI refers to the asymptotic confidence interval, GCI denotes the generalized confidence interval, and FCI is the fiducial confidence interval. To compare the above point estimators, 1000 simulations are conducted by using the mean square error (MSE) and relative mean square error (RMSE). The RMSE is calculated as the MSE obtained from ML and GI methods divided by the MSE of the GFI method. For example, the RMSE of R ^ M L is given by the MSE of R ^ M L divided by the MSE of R ^ G F I , where the GFI method is always the benchmark method. Meanwhile, we calculate the performance of the above confidence intervals with average length and empirical coverage. The relative length is the ratio of the average length gained by the ML and GI methods to the average length obtained by the GFI method. Different combinations of ( n , m , σ , α 1 , α 2 ) and ( n , m , σ 1 , σ 2 , α 1 , α 2 ) are provided at a nominal level 1 γ = 0.95 . We have the following conclusions.

4.1. Analysis of Point Estimates

  • The case with the same scale parameter.
Table 1 provides the MSEs of R for different parameter combinations, and Figure 1 shows the boxplots of RMSEs of R. The detailed information is shown as follows.
From Table 1 and Figure 1, the RMSEs of R ^ M L are larger than 1 while the RMSEs of R ^ G I are close to 1. Specifically, the MSEs of R ^ G F I are often smaller than those of R ^ M L , and the gap is significant in small and moderate samples, such as n , m 25 . Meanwhile, the difference between the MSEs of R ^ G F I and R ^ G I is trivial.
  • The case with different scale parameters.
The MSEs of R and the boxplots of RMSEs of R under different parameter combinations are shown in Table 2 and Figure 2.
In Figure 2, it is shown that the RMSEs of R ^ M L and R ^ G I are often larger than 1. From Table 2, the MSEs of R ^ G F I are smaller than those of R ^ M L , and the MSEs of R ^ G I exhibit lower stability. At the same time, the MSEs of the three methods decrease with the increase in the sample size.

4.2. Analysis of Interval Estimates

  • The case with the same scale parameter.
Table 3 provides the average length and empirical coverage of R, and Figure 3 shows the boxplots of relative length and empirical coverage. The details are as follows.
Table 3 and Figure 3 show that the relative lengths of ACIs are greater than 1 and the ACIs are too liberal. The difference between GCIs and FCIs is small and both of them are conservative. When the sample size is small, GCIs and FCIs are better than ACIs. Meanwhile, the average lengths of the three methods tend to decrease with the increase in sample size.
  • The case with different scale parameters.
The average length and empirical coverage of R, the boxplots of relative length, and empirical coverage are shown in Table 4 and Figure 4.
Figure 4 demonstrates that the relative lengths of ACIs are greater than 1 while those of GCIs are smaller than 1. FCIs are close to the nominal level while ACIs and GCIs are obviously liberal. To be specific, Table 4 shows that the average lengths of ACIs are long while the empirical coverages of ACIs are often less than 0.95. The average lengths of GCIs are short, but the empirical coverages of GCIs exhibit instability. The average lengths of FCIs and GCIs are comparable when the sample size is large, and FCIs can reach the nominal level.

5. Real Data Example

5.1. The Breaking Strengths of Jute Fibers

The first dataset was originally introduced by Xia et al. [36]. It consists of the breaking strengths of jute fibers at 4 different gauge lengths: 5 mm, 10 mm, 15 mm, and 20 mm. The breaking strengths of jute fibers at 10 mm and 20 mm are presented in Table 5.
The breaking strengths of jute fibers at two different gauge lengths are fitted with GL distribution, respectively. The estimated scale parameters, shape parameters, Kolmogorov–Smirnov (K-S) distances, and the corresponding p-values are shown in Table 6.
Referring to Table 6, the p-values obtained from the K-S test indicate that the GL distribution shows good agreement with the jute fiber data. Since the difference is significant between the scale parameters estimated at the two gauge lengths, it is reasonable to assume the scale parameters are different. The point and interval estimations of R are shown in Table 7, which implies that the ACI is the shortest while the GCI is the longest. Because the empirical coverage of ACI tends to be liberal in the simulation, we prefer to recommend the GFI method.

5.2. The Sulfur Dioxide Concentration Data

To illustrate the methods developed in Section 2 and Section 3, the second real dataset provided by Roberts [37] is considered. It consists of data on the monthly and annual maxima of one-hour mean concentrations of sulfur dioxide (pphm) for Long Beach, California from 1956 to 1974. In this paper, the average hourly concentrations of sulfur dioxide in January, March, and August are shown in Table 8.
The sulfur dioxide concentration data of three months are fitted with GL distributions separately. We present the estimated scale parameters, shape parameters, K-S distances, and corresponding p-values in Table 9.
From Table 9, the p-values of the K-S test are pretty good (p-values of 0.995, 0.804, and 0.978, respectively), which means that the GL distribution fits well with the sulfur dioxide concentration data. In addition, the p-values of the K-S test for the GL distribution are larger than those of the Weibull and generalized exponential distributions, which means the GL distribution provides a better fit than other distributions. Hence, the GL distribution is adopted in this real dataset and we consider the following two cases.
  • The case with the same scale parameter.
In this case, the average hourly concentrations of sulfur dioxide in January and March are chosen. Since the two estimated scale parameters are not very different, it is natural to assume that the two scale parameters are equal. The ML estimations for the parameters σ , α 1 , and α 2 are given by σ ^ = 8.247 , α ^ 1 = 10.593 , and α ^ 2 = 7.179 , respectively. Using the three methods in Section 2, the point and interval estimations for R are shown in Table 10.
From Table 10, it can be concluded that the difference between the point estimates of R is small while the FCI is shorter than the ACI and GCI.
  • The case with different scale parameters.
In this case, the average hourly concentration of sulfur dioxide in January and August are selected. Table 9 shows that the differences between the two estimated scale and shape parameters are significant, so it is reasonable to assume the parameters are different. The ML estimations for the parameters σ 1 , α 1 , σ 2 , and α 2 are given by σ ^ 1 = 8.070 , α ^ 1 = 11.009 , σ ^ 2 = 7.376 , and α ^ 2 = 8.154 , respectively. The point and interval estimations for R are shown in Table 10 by using the three methods in Section 3. It is found that the difference between the point estimates of R is small, and the ACI is the shortest while the GCI is the longest.
In general, the FCI performs better in the first case while the ACI performs better in the second case. However, the FCI is recommended, considering that the empirical coverage of ACI in the simulation is often lower than the nominal level.

5.3. The Insulating Fluid Data

The Ln times to breakdown of insulating fluid in an accelerated test reported by Nelson [38] is chosen as the third real data example. The Ln times to breakdown for insulating fluid were reported at different voltages of 26, 28, 30, 32, 34, 36, and 38 kV. The Ln times to breakdown 32, 34, and 36 kV are demonstrated in Table 11.
The data of Ln times to breakdown at three different voltages are fitted with GL distributions separately. Table 12 demonstrates the estimated scale parameters, shape parameters, K-S distances, and the corresponding p-values.
Based on Table 12, the GL distribution fits quite well with the data of Ln times to breakdown 32, 34, and 36 kV. Therefore, it is reasonable for us to apply the GL distribution to this real dataset, and we still consider the following two cases.
  • The case with the same scale parameter.
The data of Ln times to breakdown at 32 and 34 kV are selected in this case. If we suppose the two scale parameters are equal, the ML estimations for the parameters σ , α 1 and α 2 are given by σ ^ = 1.424 , α ^ 1 = 2.769 , and α ^ 2 = 2.882 , respectively. The point and interval estimations for R are shown in Table 13, illustrating that the ACI is the shortest while the GCI is the longest.
  • The case with different scale parameters.
In the second case, the data of Ln times to breakdown at 34 and 36 kV are considered. When we assume all the parameters are different, the ML estimations for the parameters σ 1 , α 1 , σ 2 , and α 2 can be given by σ ^ 1 = 1.2106 , α ^ 1 = 3.0672 , σ ^ 2 = 0.7552 , and α ^ 2 = 2.3542 , respectively. Table 13 states the point and interval estimations for R, and it shows that the ACI is the shortest and the GCI is the longest.
According to the above two cases, it seems that the effect of ACI is better. Since the simulation results show that the empirical coverage of ACI is often lower than the nominal level, we still prefer to select FCI with higher reliability.

6. Discussion

The estimation of R in the GL distribution is an important research problem. Most of the existing literature studies focus on the ML estimation and Bayesian inference. However, the ML estimation cannot obtain the exact pivotal quantity and its empirical coverage sometimes fails to reach the nominal level. In Bayesian inference, the choice of the prior distribution is improper or subjective. Therefore, we introduce two novel methods to estimate R in the GL distribution.
On the one hand, there are two theoretical implications worth noting. First, the GFI method is applied to estimate R. The prior of the GFI is based on actual data, which makes the posterior distribution more objective. In addition, the weighted prior is applied when the scale parameters are the same. Our findings suggest that this approach of constructing the prior is suitable for estimating R in the two-parameter GL distribution and can be extended to other distributions as well. Second, the GI method offers another way when the conventional pivotal quantity is not available. By developing two lemmas, the generalized point estimation and generalized confident interval of R can be given.
On the other hand, this article has three practical implications. First, the simulation results indicate that the generalized fiducial method is better for the point estimation of R with the comparisons of the MSE. Moreover, it can be concluded that the GFI method often outperforms the ML and GI methods for the interval estimation of R, which presents more advantages in average length and empirical coverage. Second, the results of the three real data example state that the estimation of R can be applied in many different fields. Third, the two-parameter GL distribution without a location parameter is particularly useful in estimating R, where the dataset contains values less than zero. This characteristic expands its applicability to a wider range of datasets and deserves more attention in the scale-shape life distribution.
There are some limitations in our study. Due to encountering censored data in numerous survival analyses, such as the research of Rao [39], Babayi and Khorram [40], and Wang et al. [41], the statistical inference of parameters, reliability, and stress–strength based on censored samples under the GL distribution would be an interesting direction for future works.

7. Conclusions

This article considers the statistical inference of R for the generalized logistic distribution with either the same or different scale parameters. Based on the simulation of the point estimation, the MSE of the GI and GFI methods is often smaller for the same scale parameter. However, the GFI has the smallest MSE when the scale parameters are different. According to the simulation of the interval estimation, both the GI and GFI methods exhibit shorter average lengths and more conservative empirical coverage for the same scale parameter. When the scale parameters are different, the GFI method performs better in the length and coverage. Therefore, we believe that the GFI method is more suitable for estimating R in the GL distribution and many other issues related to it.

Author Contributions

Conceptualization, L.Y. and X.C.; methodology, M.L., L.Y., Y.Q. and X.C.; software, M.L. and L.Y.; validation, M.L. and L.Y.; formal analysis, M.L.; investigation, Y.Q. and X.C.; resources, L.Y., Y.Q. and X.C.; data curation, M.L.; writing—original draft preparation, M.L.; writing—review and editing, K.K.S.; visualization, M.L.; supervision, L.Y. and K.K.S.; project administration, L.Y. and X.C.; funding acquisition, L.Y. and X.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Natural Science Foundation of Hebei Province under grant No. A2020207006 and No. A2022208001, the Foundation of Hebei Educational Department under grant No. CXZZSS2023103, and the National Natural Science Foundation of China under grant No. 12001155.

Data Availability Statement

No data were used to support this study.

Acknowledgments

The work is based on research supported by the Natural Science Foundation of Hebei Province, the Foundation of Hebei Educational Department, and the National Natural Science Foundation of China. Any opinion, finding, conclusion, or recommendation expressed in this material is that of the authors, and the NRF does not accept any liability in this regard.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Birnbaum, Z.W. On a use of the mann-whitney statistics. In Proceedings of the Third Berkeley Symposium on Mathematical Statistics and Probability, Berkeley, CA, USA, 26–31 December 1954; Volume 1, pp. 13–17. [Google Scholar]
  2. Birnbaum, Z.W.; McCarty, R.C. A distribution-free upper confidence bound for Pr(Y<X), based on independent samples of X and Y. Ann. Math. Stat. 1958, 29, 558–562. [Google Scholar]
  3. Hall, I.J. Approximate one-sided tolerance limits for the difference or sum of two independent normal variates. J. Qual. Technol. 1984, 16, 15–19. [Google Scholar] [CrossRef]
  4. Guttman, I.; Johnson, R.A.; Bhattacharyya, G.K.; Reiser, B. Confidence limits for stress-strength models with explanatory variables. Technometrics 1988, 30, 161–168. [Google Scholar] [CrossRef]
  5. Krishnamoorthy, K.; Lin, Y. Confidence limits for stress–strength reliability involving Weibull models. J. Stat. Plan. Infer. 2010, 140, 1754–1764. [Google Scholar] [CrossRef]
  6. Nooghabi, M.J.; Naderi, M. Stress–strength reliability inference for the Pareto distribution with outliers. J. Comput. Appl. Math. 2022, 404, 113911. [Google Scholar] [CrossRef]
  7. Surles, J.G.; Padgett, W.J. Inference for P(Y<X) in the Burr Type X model. J. Appl. Stat. Sci. 1998, 7, 225–238. [Google Scholar]
  8. Jafari, A.A.; Bafekri, S. Inference on stress-strength reliability for the two-parameter exponential distribution based on generalized order statistics. Math. Popul. Stud. 2021, 28, 201–227. [Google Scholar] [CrossRef]
  9. Yousef, M.M.; Almetwally, E.M. Multi stress-strength reliability based on progressive first failure for Kumaraswamy model: Bayesian and non-Bayesian estimation. Symmetry 2021, 13, 2120. [Google Scholar] [CrossRef]
  10. Yousef, M.M.; Fayomi, A.; Almetwally, E.M. Simulation techniques for strength component partially accelerated to analyze stress–strength model. Symmetry 2023, 15, 1183. [Google Scholar] [CrossRef]
  11. Gunasekera, S. Generalized inferences of R = P(X>Y) for Pareto distribution. Stat. Pap. 2014, 56, 333–351. [Google Scholar] [CrossRef]
  12. Rezaei, S.; Tahmasbi, R.; Mahmoodi, M. Estimation of P(Y<X) for generalized Pareto distribution. J. Stat. Plan. Infer. 2010, 140, 480–494. [Google Scholar]
  13. Krishnamoorthy, K.; Mukherjee, S.; Guo, H. Inference on reliability in two-parameter exponential stress–strength model. Metrika 2007, 65, 261–273. [Google Scholar] [CrossRef] [Green Version]
  14. Kundu, D.; Gupta, R.D. Estimation of P(Y<X) for generalized exponential distribution. Metrika 2005, 61, 291–308. [Google Scholar]
  15. Mahmoud, M.A.W.; El-Sagheer, R.M.; Soliman, A.A.; Ellah, A.H.A. Bayesian estimation of P(Y<X) based on record values from the Lomax distribution and MCMC technique. J. Mod. Appl. Stat. Meth. 2016, 15, 488–510. [Google Scholar]
  16. Cruz, R.D.L.; Salinas, H.S.; Meza, C. Reliability estimation for stress-strength model based on Unit-half-normal distribution. Symmetry 2022, 14, 837. [Google Scholar] [CrossRef]
  17. Alsadat, N.; Hassan, A.S.; Elgarhy, M.; Chesneau, C.; Mohamed, R.E. An efficient stress–Strength reliability estimate of the Unit gompertz distribution using ranked set sampling. Symmetry 2023, 15, 1121. [Google Scholar] [CrossRef]
  18. Asgharzadeh, A.; Valiollahi, R.; Raqab, M.Z. Estimation of the stress–strength reliability for the generalized logistic distribution. Stat. Methodol. 2013, 15, 73–94. [Google Scholar] [CrossRef]
  19. Babayi, S.; Khorram, E.; Tondro, F. Inference of R = P(X<Y) for generalized logistic distribution. J. Theor. Appl. Stat. 2014, 48, 862–871. [Google Scholar]
  20. Okasha, H.M. Estimation of P(Y<X) for generalized logistic distribution. J. Appl. Stat. Sci. 2014, 19, 43–56. [Google Scholar]
  21. Rasekhi, M.; Saber, M.M.; Yousof, H.M. Bayesian and classical inference of reliability in multicomponent stress-strength under the generalized logistic model. Commun. Stat. Theory Meth. 2020, 50, 5114–5125. [Google Scholar] [CrossRef]
  22. Balakrishnan, N.; Leung, M.Y. Order statistics from the type I generalized logistic distribution. Commun. Stat. Simul. Comput. 1988, 17, 25–50. [Google Scholar] [CrossRef]
  23. Balakrishnan, N. Handbook of the Logistic Distribution, 2nd ed.; Marcel Dekker: New York, NY, USA, 2010; pp. 50–77. [Google Scholar]
  24. Lagos-Álvarez, B.; Jerez-Lillo, N.; Navarrete, J.P.; Figueroa-Zúñiga, J.; Leiva, V. A type I generalized logistic distribution: Solving its estimation problems with a bayesian approach and numerical applications based on simulated and engineering data. Symmetry 2022, 14, 655. [Google Scholar] [CrossRef]
  25. Tao, M. Objective Bayesian Analysis for the Generalized Logistic Distribution and Doubly Accelerated Degradation Model. Master’s Thesis, Anhui Normal University, Wuhu, China, 2019. [Google Scholar]
  26. Weerahandi, S. Generalized confidence intervals. J. Am. Stat. Assoc. 1993, 88, 899–905. [Google Scholar] [CrossRef]
  27. Wang, B.; Geng, Y.; Zhou, J. Inference for the generalized exponential stress-strength model. Appl. Math. Modell. 2017, 53, 267–275. [Google Scholar] [CrossRef]
  28. Hannig, J.; Iyer, H.; Lai, R.C.S.; Lee, T.C.M. Generalized fiducial inference: A review and new results. J. Am. Stat. Assoc. 2016, 111, 1346–1361. [Google Scholar] [CrossRef]
  29. Yan, L.; Liu, X. Generalized fiducial inference for generalized exponential distribution. J. Stat. Comput. Sim. 2018, 88, 1369–1381. [Google Scholar] [CrossRef]
  30. Yan, L.; Geng, J.; Wang, L.; He, D. Generalized fiducial inference for the Lomax distribution. J. Stat. Comput. Sim. 2021, 91, 2402–2413. [Google Scholar] [CrossRef]
  31. Cai, X.; Feng, S.; Yan, L. Generalized fiducial inference for the lower confidence limit of reliability based on Weibull distribution. Commun. Stat. Simul. Comput. 2022, 1–11. [Google Scholar] [CrossRef]
  32. Wang, X.; Zou, C.; Yi, Y.; Wang, J.; Li, X. Fiducial inference for gamma distributions: Two-sample problems. Commun. Stat. Simul. Comput. 2021, 50, 811–821. [Google Scholar] [CrossRef]
  33. Wang, X.; Li, M.; Sun, W.; Gao, Z.; Li, X. Confidence intervals for zero-inflated gamma distribution. Commun. Stat. Simul. Comput. 2022, 1–18. [Google Scholar] [CrossRef]
  34. Yu, K.; Wang, B.; Patilea, V. New estimating equation approaches with application in lifetime data analysis. Ann. Inst. Stat. Math. 2013, 65, 589–615. [Google Scholar] [CrossRef]
  35. Chib, S.; Greenberg, E. Understanding the Metropolis-Hastings algorithm. Am. Stat. 1995, 49, 327–335. [Google Scholar]
  36. Xia, Z.P.; Yu, J.Y.; Cheng, L.D.; Liu, L.F.; Wang, W.M. Study on the breaking strength of jute fibres using modified Weibull distribution. Compos. Part A Appl. Sci. Manuf. 2009, 40, 54–59. [Google Scholar] [CrossRef]
  37. Roberts, E.M. Review of statistics of extreme values with applications to air quality data: Part II. Applications. J. Air. Pollut. Control Assoc. 1979, 29, 733–740. [Google Scholar] [CrossRef]
  38. Nelson, W. Analysis of accelerated life test Data-Least squares methods for the inverse power law model. IEEE. Trans. Reliab. 1979, 24, 103–107. [Google Scholar] [CrossRef]
  39. Rao, G.S. Estimation of stress-strength reliability from truncated type-I generalised logistic distribution. Int. J. Math. Oper. Res. 2015, 7, 372–382. [Google Scholar] [CrossRef]
  40. Babayi, S.; Khorram, E. Inference of stress-strength for the Type-II generalized logistic distribution under progressively Type-II censored samples. Commun. Stat. Simul. Comput. 2017, 47, 1975–1995. [Google Scholar] [CrossRef]
  41. Wang, X.; Li, X.; Zhang, L.; Liu, Z.; Li, M. Fiducial inference on gamma distributions: Two-sample problems with multiple detection limits. Environ. Ecol. Stat. 2022, 29, 453–475. [Google Scholar] [CrossRef]
Figure 1. Summary of RMSE of the point estimations for R with the same scale parameter.
Figure 1. Summary of RMSE of the point estimations for R with the same scale parameter.
Symmetry 15 01365 g001
Figure 2. Summary of RMSEs of the point estimations for R with different scale parameters.
Figure 2. Summary of RMSEs of the point estimations for R with different scale parameters.
Symmetry 15 01365 g002
Figure 3. Summary of the relative length and empirical coverage for R with the same scale parameter. (a) Display of the relative length; (b) display of the empirical coverage.
Figure 3. Summary of the relative length and empirical coverage for R with the same scale parameter. (a) Display of the relative length; (b) display of the empirical coverage.
Symmetry 15 01365 g003
Figure 4. Summary of relative length and empirical coverage for R with different scale parameters. (a) Display of the relative length; (b) display of the empirical coverage.
Figure 4. Summary of relative length and empirical coverage for R with different scale parameters. (a) Display of the relative length; (b) display of the empirical coverage.
Symmetry 15 01365 g004
Table 1. The MSE of the point estimations for R with the same scale parameter.
Table 1. The MSE of the point estimations for R with the same scale parameter.
( σ , α 1 , α 2 ) nmMSE for R
R ^ ML R ^ GI R ^ GFI
(1.0, 1.5, 2.0)15150.0081580.0071700.007172
15250.0063600.0057310.005707
15500.0051740.0047610.004743
25150.0063920.0058270.005788
25250.0051200.0047490.004730
25500.0035600.0033640.003337
50150.0052240.0048950.004892
50250.0033720.0032310.003230
50500.0023530.0022690.002250
(1.0, 2.0, 1.5)15150.0090410.0079480.007925
15250.0070530.0063880.006369
15500.0050250.0046390.004625
25150.0062050.0055950.005577
25250.0051550.0047420.004749
25500.0038720.0036170.003631
50150.0051040.0046930.004661
50250.0036760.0034680.003457
50500.0023160.0022320.002221
Table 2. The MSEs of the point estimations for R with different scale parameters.
Table 2. The MSEs of the point estimations for R with different scale parameters.
( σ 1 , α 1 , σ 2 , α 2 ) nmMSE for R
R ^ ML R ^ GI R ^ GFI
(1.0, 1.5, 2.0, 2.0)15150.0099010.0187220.008245
15250.0069400.0141360.006285
15500.0046040.0092510.004256
25150.0092010.0186260.007926
25250.0056620.0131870.005149
25500.0034310.0078220.003238
50150.0077470.0166160.006800
50250.0049200.0116900.004507
50500.0028450.0076150.002739
(2.0, 2.0, 1.0, 1.5)15150.0101450.0079370.008674
15250.0087580.0075020.007744
15500.0080420.0070360.007151
25150.0070370.0054600.006330
25250.0055630.0046690.005115
25500.0051550.0047100.004744
50150.0045250.0033730.004208
50250.0034110.0028110.003236
50500.0031050.0027650.002963
Table 3. The average length and empirical coverage of 95 % two-sided confidence intervals for R with the same scale parameter.
Table 3. The average length and empirical coverage of 95 % two-sided confidence intervals for R with the same scale parameter.
( σ , α 1 , α 2 ) nmAverage LengthEmpirical Coverage
ACIGCIFCIACIGCIFCI
(1.0, 1.5, 2.0)15150.3500.3380.3380.9110.9570.958
15250.3170.3050.3050.9300.9570.960
15500.2830.2750.2750.9500.9600.962
25150.3120.3050.3050.8990.9540.955
25250.2770.2650.2660.9110.9540.953
25500.2400.2310.2310.9490.9570.959
50150.2870.2770.2770.8850.9520.951
50250.2470.2320.2330.9190.9550.953
50500.1970.1900.1900.9490.9570.961
(1.0, 2.0, 1.5)15150.3480.3360.3370.9120.9490.953
15250.3100.3050.3050.9420.9610.958
15500.2780.2770.2770.9350.9580.958
25150.3130.3040.3040.9440.9640.962
25250.2730.2650.2650.9390.9530.955
25500.2330.2310.2320.9360.9560.957
50150.2840.2750.2760.9470.9540.955
50250.2360.2310.2310.9470.9560.958
50500.1920.1900.1900.9480.9580.955
Table 4. The average length and empirical coverage of 95 % two-sided confidence intervals for R with different scale parameters.
Table 4. The average length and empirical coverage of 95 % two-sided confidence intervals for R with different scale parameters.
( σ 1 , α 1 , σ 2 , α 2 ) nmAverage LengthEmpirical Coverage
ACIGCIFCIACIGCIFCI
(1.0, 1.5, 2.0, 2.0)15150.3650.3520.3590.8990.8180.957
15250.3130.3040.3090.9280.8040.959
15500.2580.2540.2590.9230.8210.953
25150.3460.3320.3410.9000.7780.947
25250.2890.2780.2840.9290.7650.962
25500.2300.2220.2280.9410.7930.961
50150.3290.3170.3280.9130.7650.955
50250.2680.2610.2650.9320.7610.957
50500.2060.2000.2040.9400.7290.954
(2.0, 2.0, 1.0, 1.5)15150.3680.3600.3610.9010.9690.957
15250.3490.3430.3430.9180.9590.949
15500.3300.3280.3280.9080.9570.952
25150.3120.3060.3080.9370.9660.956
25250.2890.2840.2850.9370.9650.956
25500.2690.2670.2670.9240.9440.940
50150.2590.2570.2600.9370.9750.953
50250.2310.2270.2290.9390.9710.949
50500.2060.2030.2040.9220.9460.937
Table 5. The breaking strengths of jute fibers at different gauge lengths.
Table 5. The breaking strengths of jute fibers at different gauge lengths.
Gauge LengthsDataSample Size
10 mm693.73, 704.66, 323.83, 778.17, 123.06, 637.66, 383.43, 151.48, 108.94, 50.16, 671.49, 183.16, 257.44, 727.23, 291.27, 101.15, 376.42, 163.40, 141.38, 700.74, 262.90, 353.24, 422.11, 43.93, 590.48, 212.13, 303.90, 506.60,530.55,177.2530
20 mm71.46, 419.02, 284.64, 585.57, 456.60, 113.85, 187.85, 688.16, 662.66, 45.58, 578.62, 756.70, 594.29, 166.49, 99.72, 707.36, 765.14, 187.13, 145.96, 350.70, 547.44, 116.99, 375.81, 581.60, 119.86, 48.01, 200.16, 36.75, 244.53, 83.5530
Table 6. The scale parameter, shape parameter, K-S, and p-values of the breaking strengths of jute fibers.
Table 6. The scale parameter, shape parameter, K-S, and p-values of the breaking strengths of jute fibers.
Gauge LengthsScale ParameterShape ParameterK–Sp-Value
10 mm170.8265.1900.1180.756
20 mm178.2784.2100.1610.376
Table 7. The result of R for the breaking strengths of jute fibers.
Table 7. The result of R for the breaking strengths of jute fibers.
10 mm and 20 mm
PointIntervalLength
ACI0.535[0.390, 0.680]0.290
GCI0.526[0.338, 0.702]0.364
FCI0.531[0.374, 0.691]0.317
Table 8. The sulfur dioxide concentration data under different months.
Table 8. The sulfur dioxide concentration data under different months.
MonthsDataSample Size
January47, 22, 15, 20, 22, 25, 20, 12, 16, 16, 27, 30, 51, 37, 23, 22, 30, 10, 819
March44, 20, 20, 20, 23, 20, 15, 27, 3, 9, 25, 32, 18, 55, 10, 20, 18, 8, 919
August21, 16, 20, 15, 9, 10, 10, 4, 25, 18, 18, 26, 25, 17, 40, 55, 19, 16, 919
Table 9. The scale parameter, shape parameter, K-S, and p-values of sulfur dioxide concentration data.
Table 9. The scale parameter, shape parameter, K-S, and p-values of sulfur dioxide concentration data.
MonthsScale ParameterShape ParameterK–Sp-Value
January8.07011.0090.0960.995
March8.4176.9910.1470.807
August7.3768.1540.1090.978
Table 10. The results of R of sulfur dioxide concentration data.
Table 10. The results of R of sulfur dioxide concentration data.
January and MarchJanuary and August
PointIntervalLengthPointIntervalLength
ACI0.596[0.444, 0.748]0.3040.626[0.453, 0.799]0.346
GCI0.586[0.412, 0.731]0.3190.621[0.327, 0.840]0.513
FCI0.590[0.431, 0.726]0.2950.615[0.401, 0.808]0.407
Table 11. The insulating fluid data at different voltages.
Table 11. The insulating fluid data at different voltages.
VoltagesDataSample Size
32 kV−1.3094, −0.9163, −0.3711, −0.2358, 1.0116, 1.3635, 2.2905, 2.6354, 2.7682, 3.3250, 3.9748, 4.4170, 4.4918, 4.6109, 5.371115
34 kV−1.6608, −0.2485, −0.0409, 0.2700, 1.0224, 1.1505, 1.4231, 1.5411, 1.5789, 1.8718, 1.9947, 2.0806, 2.1126, 2.4898, 3.4578, 3.4818, 3.5237, 3.6030, 4.288919
36 kV−1.0499, −0.5277, −0.0409, −0.0101, 0.5247, 0.6780, 0.7275, 0.9477, 0.9969, 1.0647, 1.3001, 1.3837, 1.6770, 2.6224, 3.238615
Table 12. The scale parameter, shape parameter, K-S, and p-values of insulating fluid data.
Table 12. The scale parameter, shape parameter, K-S, and p-values of insulating fluid data.
VoltagesScale ParameterShape ParameterK–Sp-Value
32 kV1.6902.6800.1410.888
34 kV1.2113.0670.1240.900
36 kV0.7552.3540.1190.967
Table 13. The result of R of insulating fluid data.
Table 13. The result of R of insulating fluid data.
32 kV and 34 kV34 kV and 36 kV
PointIntervalLengthPointIntervalLength
ACI0.490[0.329, 0.651]0.3220.674[0.499, 0.848]0.349
GCI0.489[0.323, 0.661]0.3380.666[0.447, 0.829]0.382
FCI0.487[0.319, 0.650]0.3310.652[0.466, 0.817]0.351
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, M.; Yan, L.; Qiao, Y.; Cai, X.; Said, K.K. Generalized Fiducial Inference for the Stress–Strength Reliability of Generalized Logistic Distribution. Symmetry 2023, 15, 1365. https://doi.org/10.3390/sym15071365

AMA Style

Li M, Yan L, Qiao Y, Cai X, Said KK. Generalized Fiducial Inference for the Stress–Strength Reliability of Generalized Logistic Distribution. Symmetry. 2023; 15(7):1365. https://doi.org/10.3390/sym15071365

Chicago/Turabian Style

Li, Menghan, Liang Yan, Yaru Qiao, Xia Cai, and Khamis K. Said. 2023. "Generalized Fiducial Inference for the Stress–Strength Reliability of Generalized Logistic Distribution" Symmetry 15, no. 7: 1365. https://doi.org/10.3390/sym15071365

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop