Next Article in Journal
A Modified Regularization Method for a Spherically Symmetric Inverse Heat Conduction Problem
Next Article in Special Issue
Reliability Inferences of the Inverted NH Parameters via Generalized Type-II Progressive Hybrid Censoring with Applications
Previous Article in Journal
A Class-Incremental Detection Method of Remote Sensing Images Based on Selective Distillation
Previous Article in Special Issue
Robust Sparse Reduced-Rank Regression with Response Dependency
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Confidence Intervals for Common Coefficient of Variation of Several Birnbaum–Saunders Distributions

Department of Applied Statistics, Faculty of Applied Science, King Mongkut’s University of Technology North Bangkok, Bangkok 10800, Thailand
*
Author to whom correspondence should be addressed.
Symmetry 2022, 14(10), 2101; https://doi.org/10.3390/sym14102101
Submission received: 5 September 2022 / Revised: 30 September 2022 / Accepted: 7 October 2022 / Published: 9 October 2022

Abstract

:
The Birnbaum–Saunders (BS) distribution, also known as the fatigue life distribution, is right-skewed and used to model the failure times of industrial components. It has received much attention due to its attractive properties and its relationship to the normal distribution (which is symmetric). Furthermore, the coefficient of variation (CV) is commonly used to analyze variation within a dataset. In some situations, the independent samples are collected from different instruments or laboratories. Consequently, it is of importance to make inference for the common CV. To this end, confidence intervals based on the generalized confidence interval (GCI), method of variance estimates recovery (MOVER), large-sample (LS), Bayesian credible interval (BayCrI), and highest posterior density interval (HPDI) methods are proposed herein to estimate the common CV of several BS distributions. Their performances in terms of their coverage probabilities and average lengths were investigated by using Monte Carlo simulation. The simulation results indicate that the HPDI-based confidence interval outperformed the others in all of the investigated scenarios. Finally, the efficacies of the proposed confidence intervals are illustrated by applying them to real datasets of PM10 (particulate matter ≤ 10 μm) concentrations from three pollution monitoring stations in Chiang Mai, Thailand.

1. Introduction

The original idea behind the Birnbaum–Saunders (BS) distribution lies in an investigation of vibrations in commercial aircraft that cause material fatigue. Fatigue is a type of structural deterioration that happens when a material is subjected to fluctuating stress and tension [1]. To address these problems, Birnbaum and Saunders [2] proposed the fatigue life distribution, which is commonly known as the BS distribution to describe the failure time of materials and equipment subjected to dynamic loads where failure is caused by the initiation and growth of a dominant fracture. The BS distribution is positively asymmetric and unimodal with two positive parameters: α , the shape parameter, and β , which is both the scale parameter and the median of the distribution. In addition, it has many attractive properties and has a close relationship with the normal distribution. The BS distribution is very effective for fitting data that are all positive. Despite its origins in materials science, the BS distribution has recently been applied to various other fields, including the environment, business, industry, finance, and medical sciences [3,4,5,6].
The coefficient of variation (CV) is an important descriptive statistic for analyzing the variability of data. In particular, it is a measure of variability relative to the mean. The CV is defined as a ratio of the standard deviation ( σ ) to the mean ( μ ), namely CV = σ / μ . It is free from the unit of measurement, and, thus, it has been preferentially used for comparing relative variability between two or more populations rather than the variance or standard deviation [7]. In many situations, independent samples are collected from methods involving different instruments, methodologies, and/or laboratories, and so estimating the common CV of these related populations is of great interest. Many researchers have developed confidence intervals for estimating the common CV of several populations from various distributions using several methods. For example, Tian [8] used the concept of the generalized confidence interval (GCI) to construct the confidence interval for the common CV of several independent normal samples. Verrill and Johnson [9] proposed a likelihood ratio-based confidence interval for a common CV of several normal distributions. Behboodian and Jafari [10] utilized the concept of generalized p-values and GCI to develop a new method for estimating the confidence interval for the common CV of several normal populations. Ng [11] suggested a method for estimating the confidence interval for the common CV of several lognormal samples by utilizing the concept of the generalized variable. Thangjai and Niwitpong [12] developed the adjusted method of variance estimates recovery (MOVER) for constructing the confidence interval for the common CV of two-parameter exponential distributions and then compared its performance with GCI and large-sample (LS) confidence intervals. Liu and Xu [13] introduced a new confidence interval for the common CV of several normal distributions based on the concept of the confidence distribution interval. Recently, Yosboonruang et al. [14] constructed confidence intervals for the common CV of delta-lognormal distributions using the fiducial GCI (FGCI), equal-tailed Bayesian credible intervals (BayCrI) based on the independent Jeffreys or uniform priors, and MOVER.
Estimating the parameters of a BS distribution is of significant interest to many researchers and has recently garnered much attention in the literature. For instance, the maximum likelihood estimation (MLE) of α and β were introduced in Birnbaum and Saunders [15] and Engelhardt et al. [16]. Ng et al. [17] presented modified moment estimators (MMEs) for α and β and a bias reduction method with Jackknife resampling to reduce the biases of the MMEs and MLEs. Wu and Wong [18] improved the confidence interval for the two-parameter BS distribution based on a high-order likelihood asymptotic procedure. Xu and Tang [19] explored Bayesian estimators for α and β under the reference prior by using Lindley’s method and Gibbs’ sampling to obtain approximate Bayesian estimators for these two parameters. Wang [20] examined GCI for α , as well as some important reliability quantities, such as mean, quantiles, and a reliability function. Wang et al. [21] considered Bayesian estimators under inverse-gamma priors for α and β to compute their Bayesian estimates and credible intervals. Guo et al. [22] applied a hybrid of the generalized inference method and the LS theory for interval estimation and hypothesis testing of the common mean of several BS distributions. Puggard et al. [23] proposed confidence intervals for the CV and the difference between the CVs of BS distributions based on GCI, the bootstrap confidence interval, BayCrI, and the highest posterior density interval (HPDI). Recently, Puggard et al. [24] presented confidence intervals for the ratio of the variances of two independent BS distributions using the generalized fiducial confidence interval, BayCrI, and HPDI based on a prior distribution with partial information and a proper prior with known hyperparameters. However, estimating the common CV of two or more independent BS distributions has not previously been reported. Therefore, the goal of this study is to estimate confidence intervals for the common CV of several BS distributions based on the concepts of GCI, MOVER, LS, BayCrI, and HPDI.
The remainder of this study is organized as follows. Section 2 provides the methodologies for constructing confidence intervals for the common CV of several BS distributions. Section 3 covers the methodology and results of an extensive Monte Carlo simulation study to compare the performances of the proposed confidence intervals. An illustration of the proposed confidence intervals with datasets of PM10 (particulate matter (PM) ≤ 10 μm) concentrations collected in March 2019 from three pollution monitoring stations in Chiang Mai, Thailand, is presented in Section 4. Finally, conclusions are covered in Section 5.

2. Methods

Let X i j = ( X i 1 , X i 2 , , X i n i ) be random samples of size n i drawn from a BS distribution, where i = 1 , 2 , , k and j = 1 , 2 , , n i . The cumulative distribution function (cdf) of random variable X i j can be written as:
F ( x i j ) = Φ 1 α i x i j β i β i x i j , x i j > 0 , α i , β i > 0 ,
where Φ ( · ) is the standard normal cdf and α i and β i are the shape and the scale parameters, respectively. Thus, the probability density function (pdf) of X i j is given by:
f ( x i j , α i , β i ) = 1 2 α i β i 2 π β i x i j 1 2 + β i x i j 3 2 e x p 1 2 α i 2 x i j β + β i x i j 2 .
The expected value and variance of X i j are defined as:
E ( X i j ) = β i ( 1 + 1 2 α i 2 )
and
V ( X i j ) = ( α i β i ) 2 ( 1 + 5 4 α i 2 ) ,
respectively. Therefore, the CV of X i j can be easily obtained as:
λ i = α i 1 + 5 4 α i 2 1 + 1 2 α i 2 .
According to Ng et al. [17], the MMEs of ( α i , β i ) are given by:
α ^ i = 2 x ¯ i j = 1 n i x i j 1 / n i 1 / 2 1 1 / 2 and β ^ i = x ¯ i j = 1 n i x i j 1 / n i 1 1 / 2 ,
where x ¯ i = j = 1 n i x i j / n i . In addition, it has been shown in the study of Ng et al. [17] that the asymptotic joint distribution of α i and β i is bivariate normal, which is given by:
α ^ i β ^ i N α i β i , α i 2 2 n i 0 0 ( α i β i ) 2 n i 1 + 3 4 α i 2 ( 1 + 1 2 α i 2 ) 2 .
By applying the delta method, it follows that:
n i ( λ ^ i λ i ) d N 0 , α i 2 2 8 ( 2 α i 2 + 1 ) ( α i 2 + 2 ) 2 5 α i 2 + 4 2 ,
where λ ^ i = α ^ i 1 + 5 4 α ^ i 2 1 + 1 2 α ^ i 2 . By applying Equation (7), the variance of λ ^ i becomes:
δ i = V ( λ ^ i ) = α i 2 2 n i 8 ( 2 α i 2 + 1 ) ( α i 2 + 2 ) 2 5 α i 2 + 4 2 .
According to Thangjai and Niwitpong [12] and Yosboonruang et al. [14], the common CV of several BS distributions can be written as:
λ = i = 1 k w i λ i i = 1 k w i ,
where w i = 1 / V ( λ ^ i ) . The following proposed methods are used to construct the confidence intervals for the common CV of several BS distributions.

2.1. The GCI Approach

Weerahandi [25] introduced the concept of the generalized pivotal quantity (GPQ) and deduced the GCI as an extension of the classical confidence interval. In contrast to a traditional pivotal quantity, the GPQ can be a function of the nuisance parameters and has a distribution that is independent of the unknown parameter and an observed value that is independent of the nuisance parameters. Therefore, the GCI is useful in situations when the traditional pivot quantity is either unavailable or difficult to obtain. A full detailed discussion, as well as several applications can be found in Weerahandi [25,26], Tian [8], Behboodian and Jafari [10], Chen and Ye [27,28,29], and Luo et al. [30].
Consider k independent random samples X i 1 , X i 2 , , X i n i from BS distributions. According to Sun [31] and Wang [20], the GPQ for β i can be defined as:
T β i : = T β i ( x i j ; T i ) = m a x ( β i 1 , β i 2 ) , i f T i 0 m i n ( β i 1 , β i 2 ) , i f T i > 0 ,
where x i j = ( x i 1 , x i 2 , , x i n ) are the observed values of X i j and T i follow a t-distribution with n i 1 degrees of freedom (denoted as T i t ( n i 1 ) ). By applying Equation (10), β i 1 and β i 2 are the two solutions for:
[ ( n i 1 ) B i 2 1 n i D i T i 2 ] β i 2 2 ( n i 1 ) A i B i ( 1 A i B i ) T i 2 β i + ( n i 1 ) A i 2 1 n i C i T i 2 = 0 ,
where A i = n i 1 j = 1 n i X i j , B i = n i 1 j = 1 n i 1 / X i j , C i = j = 1 n i ( X i j A i ) 2 and D i = j = 1 n i ( 1 / X i j B i ) 2 . Subsequently, Wang [20] also established the GPQ for α i which is derived as:
T α i : = T α i ( x i j ; υ i , T i ) = S i 2 T β i 2 2 n i T β i + S i 1 T β i υ i 1 / 2 ,
where S i 1 = j = 1 n i X i j , S i 2 = j = 1 n i 1 / X i j and υ i follow a Chi-squared distribution with n i degrees of freedom (denoted as υ i χ ( n i ) 2 ). By substituting T α i into Equations (5) and (8), the respective GPQs of λ i and the variance of λ ^ i become:
T λ i = T α i 1 + 5 4 T α i 2 1 + 1 2 T α i 2
and
T δ i = T α i 2 2 n i 8 ( 2 T α i 2 + 1 ) ( T α i 2 + 2 ) 2 5 T α i 2 + 4 2 .
Consequently, the GPQ for the common CV of several BS distributions is the weighted average of GPQ T λ i based on k individual samples as follows:
T λ = i = 1 k T w i T λ i i = 1 k T w i ,
where T w i = 1 / T δ i . It follows that the 100 ( 1 γ ) % GCI for λ can be constructed as [ T λ ( γ / 2 ) , T λ ( 1 γ / 2 ) ] , where T λ ( γ / 2 ) and T λ ( 1 γ / 2 ) denote the 100 ( γ / 2 ) th and 100 ( 1 γ / 2 ) th percentiles of T λ , respectively. Algorithm 1 summarizes the computational steps for constructing GCI.
Algorithm 1: GCI approach
  • Generate datasets x i j , for i = 1 , 2 , , k ; j = 1 , 2 , , n i from a BS distribution.
  • Compute A i , B i , C i , D i , S i 1 and S i 2 , respectively.
  • For m = 1 to M
  • Generate T i t ( n i 1 ) , and then compute T β i by using Equation (10).
  • If T β i < 0 , regenerate T i t ( n i 1 ) .
  • Generate υ i χ ( n i ) 2 , and then compute T α i by using Equation (12).
  • Compute T λ i and T w i to obtain T λ .
  • (End M loops)
  • Compute T λ ( γ / 2 ) and T λ ( 1 γ / 2 ) .

2.2. The MOVER Approach

The original concept behind MOVER is to estimate a closed-form confidence interval for the sum or difference between two independent parameters based on the confidence intervals of the individual parameters [32,33]. The MOVER technique was recently applied to a linear combination of parameters θ 1 , θ 2 , , θ k [34]. Suppose i = 1 k c i θ i is a linear combination of parameters θ 1 , θ 2 , , θ k , where c i are known constants. Assume that θ ^ i is an unbiased estimate of θ i . In addition, let ( l i , u i ) denote the 100 ( 1 γ ) % confidence interval for θ i , for i = 1 , 2 , , k . Hence, the 100 ( 1 γ ) % MOVER confidence interval for i = 1 k c i θ i can be written as:
L = i = 1 k c i θ ^ i i = 1 k c i 2 ( θ ^ i l i * ) 2 ; l i * = l i if c i > 0 u i if c i < 0
and
U = i = 1 k c i θ ^ i + i = 1 k c i 2 ( θ ^ i u i * ) 2 , ; u i * = u i if c i > 0 l i if c i < 0 .
By applying Equation (13), the 100 ( 1 γ ) % confidence interval for λ i based on the GPQs becomes
[ L i , U i ] = [ T λ i ( γ / 2 ) , T λ i ( 1 γ / 2 ) ] ,
where T λ i ( γ / 2 ) and T λ i ( 1 γ / 2 ) denote the 100 ( γ / 2 ) th and 100 ( 1 γ / 2 ) th percentiles of T λ i , respectively. Therefore, the 100 ( 1 γ ) % MOVER confidence interval for the common CV of several BS distributions can be expressed as
L = i = 1 k c i * λ ^ i i = 1 k c i * 2 ( λ ^ i L i * ) 2 ; L i * = L i if c i * > 0 U i if c i * < 0
and
U = i = 1 k c i * λ ^ i + i = 1 k c i * 2 ( λ ^ i U i * ) 2 ; U i * = U i if c i * > 0 L i if c i * < 0 ,
where c i * = w i / j = 1 k w j . The confidence interval based on MOVER can be easily constructed using Algorithm 2.
Algorithm 2: MOVER approach
  • Generate datasets x i j , for i = 1 , 2 , , k ; j = 1 , 2 , , n i from a BS distribution.
  • Compute c i * and λ ^ i .
  • Compute the 100 ( 1 γ ) % GCI for λ i by applying Equation (18).
  • Compute L and U, by using Equations (19) and (20), respectively, leading to obtain the 95% confidence interval based on MOVER.

2.3. The LS Approach

A large sample is a set of values that are used to estimate the true value of a population parameter. For the BS distribution, the LS estimate of the CV is a pooled estimate of it, as defined in Equation (9). Therefore, the 100 ( 1 γ ) % LS confidence interval for the common CV can be derived as:
[ L L S , U L S ] = λ ^ z 1 γ 2 1 / i = 1 k w i , λ ^ + z 1 γ 2 1 / i = 1 k w i .
Algorithm 3 was applied to obtain the LS confidence interval.
Algorithm 3: LS approach
  • Generate datasets x i j , for i = 1 , 2 , , k ; j = 1 , 2 , , n i from a BS distribution.
  • Compute λ i to obtain λ ^ .
  • Compute the 95% LS confidence interval for λ by using Equation (21).

2.4. The BayCrI Approach

The Bayesian method involves making statistical inferences about a parameter based on two sources of information: experimental data via its likelihood function and judgment based on previous knowledge via its prior distribution. Combining these data sources results in uncovering the posterior distribution.
For the BS distribution, the likelihood function for the parameters ( α i , β i ) from random sample x i j = ( x i 1 , x i 2 , , x i n i ) can be written as:
L ( x i j | α i , β i ) 1 α i n i β i n i j = 1 n i β i x i j 1 2 + β i x i j 3 2 e x p j = 1 n i 1 2 α i 2 x i j β i + β i x i j 2 .
The reference (independent Jeffreys’) prior of a BS distribution can lead to an improper posterior distribution [35], so a suitable prior with known hyperparameters is needed to ensure that a proper one is obtained. By utilizing useful reparameterization η i = α i 2 , an inverse-gamma (IG) distribution with parameters a i and b i is a suitable prior for η i (denoted as I G ( η i | a i , b i ) ). In addition, an IG distribution with parameters c i and d i is a suitable prior for β i (denoted as I G ( β i | c i , d i ) ) [21]. Hence, the joint posterior density function of ( η i , β i ) can be obtained by combining the likelihood function from Equation (22) with the IG prior distributions for η i and β i as follows:
p ( η i , β i | x i j ) L ( x i j | α i , β i ) π ( η i | a i , b i ) π ( β i | c i , d i ) 1 ( η i ) n i 2 β i n i j = 1 n i β i x i j 1 2 + β i x i j 3 2 e x p j = 1 n i 1 2 η i x i j β i + β i x i j 2 × ( η i ) a i 1 e x p b i η i β i c i 1 e x p d i β i .
Subsequently, the marginal posterior distribution of β i can be written as:
π ( β i | x i j ) β i ( n i + c i + 1 ) e x p d i β i j = 1 n i β i x i j 1 2 + β i x i j 3 2 × j = 1 n i 1 2 x i j β i + β i x i j 2 + b i ( n i + 1 ) 2 a i .
Moreover, the conditional posterior distribution of η i given β i can be derived as:
η i | β i x i j I G n i 2 + a i , 1 2 j = 1 n i x i j β i + β i x i j 2 + b i .
Since the marginal posterior in Equation (24) is mathematically intractable, the Markov Chain–Monte Carlo method can be utilized to draw posterior samples to be used for inference. According to Wang et al. [21], the posterior sample of β i ( β i * ) can be generated by applying the generalized ratio-of-uniforms method [36] as follows.
Let
A ( r i ) = ( u i , v i ) : 0 < u i π v i u i r i | x i j 1 / ( r i + 1 ) ,
where π ( · | x i j ) is defined as in Equation (24) and r i 0 is a constant value. If ( u i , v i ) is a random vector uniformly distributed over A ( r i ) , then β i = v i / u i r i has probability density function π ( β i | x i j ) / π ( β i | x i j ) d β i . In general, directly generating ( u i , v i ) uniformly over A ( r i ) is not possible, so the accept–reject method from minimal bounding rectangle [ 0 , a ( r i ) ] × [ b ( r i ) , b + ( r i ) ] is applied, where
a ( r i ) = sup β i > 0 { [ π ( β i | x i j ) ] 1 / ( r i + 1 ) } ,
b ( r i ) = inf β i > 0 { β i [ π ( β i | x i j ) ] r i / ( r i + 1 ) } ,
and
b + ( r i ) = sup β i > 0 { β i [ π ( β i | x i j ) ] r i / ( r i + 1 ) } .
As in Wang et al. [21], a ( r i ) and b + ( r i ) are finite, whereas b ( r i ) = 0 . The principal steps of the generalized ratio-of-uniforms method for generating the posterior sample of β i from Equation (24) can be summarized as follows:
1. Calculate a ( r i ) and b + ( r i ) .
2. Generate u i and v i from U ( 0 , a ( r i ) ) and U ( 0 , b + ( r i ) ) , where U ( v , w ) is a uniform distribution with parameters v and w.
3. Calculate ρ i = v i / u i r i .
4. If u i [ π ( ρ i | x i j ) ] 1 / ( r i + 1 ) , set β i * = ρ i ; otherwise repeat the procedure.
For the posterior sample of α i (denoted as α i * ), a value for η i from Equation (25) is generated by using the L e a r n B a y e s package from the R software, then α i * = η i . By Equations (5) and (8), the Bayesian estimator for the CV and variance of CV become
λ i * = α i * 1 + 5 4 α i * 2 1 + 1 2 α i * 2
and
δ i * = α i * 2 2 n i 8 ( 2 α i * 2 + 1 ) ( α i * 2 + 2 ) 2 5 α i * 2 + 4 2 ,
respectively. Consequently, the Bayesian estimator for the common CV of several BS distributions can be derived as
λ * = i = 1 k w i * λ i * i = 1 k w i * ,
where w i * = 1 / δ i * . Finally, the 100 ( 1 γ ) % BayCrI for λ can be constructed as [ λ * ( γ / 2 ) , λ * ( 1 γ / 2 ) ] , where λ * ( γ / 2 ) and λ * ( 1 γ / 2 ) denote the 100 ( γ / 2 ) th and 100 ( 1 γ / 2 ) th percentiles of λ * , respectively. Therefore, BayCrI for λ can be estimated via Algorithm 4.
Algorithm 4: BayCrI approach
  • Generate datasets x i j , for i = 1 , 2 , , k ; j = 1 , 2 , , n i from a BS distribution.
  • Set the values for a i , b i , c i , d i , and r i .
  • Compute a ( r i ) and b + ( r i ) .
  • At the hth step,
    (a)
    Generate u i U ( 0 , a ( r i ) ) and v i U ( 0 , b + ( r i ) ) , independently, and then compute ρ i = v i / u i r i .
    (b)
    If u i [ π ( ρ i | x i j ) ] 1 / ( r i + 1 ) , accept ρ i and set β i , ( h ) * = ρ i ; otherwise, repeat step (a).
    (c)
    Generate η ˜ i , ( h ) I G n i 2 + a i , 1 2 j = 1 n i x i j β i , ( h ) * + β i , ( h ) * x i j 2 + b i and then α i , ( h ) * = η ˜ i , ( h ) .
    (d)
    Compute λ i , ( h ) * and w i , ( h ) * to obtain λ ( h ) * .
  • Repeat step (4) H times.
  • Compute the 100 ( 1 γ ) % BayCrI for λ .

2.5. The HPDI Approach

The Bayesian estimation has already been produced in the previous subsection, but in most cases, we have to construct an interval containing the estimated values of parameters with a high probability. HPDI has the property that the probability density of each point inside the interval is higher than that of every point outside it, and so the intervals of the former are the shortest given probability level ( 1 γ ) [37]. The HDInterval package (version 0.2.2) from the R software was applied at step (6) in Algorithm 4 to calculate the HPDI for λ .

3. Simulation Study and Results

Since a theoretical comparison of the confidence intervals is not possible, a Monte Carlo simulation study was conducted to assess their performances by comparing their coverage probabilities and average lengths. Throughout the simulation study, the nominal confidence level was set at 0.95. The best-performing method for a particular scenario is the one with a coverage probability greater than or close to the nominal confidence level and the shortest average length. Since β i , i = 1 , 2 , , k is the scale parameter, its value was fixed as β i = 1.0 without losing any generality. The settings for the sample size and shape parameter are provided in Table 1. The number of simulation runs was 1000 replications with 3000 pivotal quantities for GCI. The following settings were used for BayCrI and HPDI: H = 1000 ; hyperparameters a i = b i = c i = d i = 10 4 ; and r i = 2 [21].
The simulation results for k = 3 , 5 , and 10 are reported in Table 2, Table 3 and Table 4, respectively. It can be seen that they are similar for these three scenarios, and, thus, we can draw the following conclusions. The coverage probabilities of the GCI, BayCrI, and HPDI confidence intervals were greater than or close to the nominal confidence level of 0.95 under most circumstances whereas those for the MOVER and LS confidence intervals were under in all of the scenarios. As the sample sizes were increased, the coverage probabilities of the MOVER and LS confidence intervals performed better but were still under the nominal confidence level of 0.95. Note that both are based on the MME of α i , which is highly biased when the sample size is small and α i is large [17]. When considering the average lengths, those of the LS confidence interval were the shortest under most circumstances, followed by MOVER. However, the coverage probabilities of these two confidence intervals were lower than the nominal confidence level of 0.95 for all cases, and so they failed to meet the requirements. Among the remainder, the average lengths of HPDI were the shortest in all of the circumstances tested whereas those of GCI were the longest. When the sample sizes were increased, the average lengths of all of the confidence intervals became shorter, whereas when the shape parameter was increased, the average lengths of all of the confidence intervals became longer. Overall, HPDI performed the best in the simulation study because it fulfilled the requirements for both criteria.

4. Application of the Confidence Interval Methods with Real Data

Air pollution is currently one of the most important public health concerns since it causes mortality and morbidity. Of the various air pollutants, PM10 and PM2.5 (PM ≤ 2.5 μm) are widely considered to be the most damaging and important. In Chiang Mai, agricultural burning and forest fires during the dry season caused a haze of predominantly PM10 and PM2.5 each year. It begins in early February, peaks in March, and subsides by the end of April. During this time period, the population is significantly impacted by PM2.5 and PM10 pollution, with concentrations substantially above the World Health Organization’s recommended levels. The average daily PM10 concentrations from three pollution monitoring stations located in Chiang Mai province: (1) Chang Phueak, (2) Si Phum, and (3) Changkerng were obtained from the Pollution Control Department [38] and selected to assess the performances of the proposed confidence intervals. Since the concentrations of PM10 are always positive and vary depending on factors, such as source, local topography, and local meteorology, they are positively skewed and suitable for fitting to a lognormal, BS, exponential, gamma, or Weibull distribution. It is important to check the suitability of the distribution for the datasets, and so minimum Akaike information criterion (AIC) and Bayesian information criterion (BIC) analyses were conducted.
As reported in Table 5 and Table 6, it can be concluded that the BS distribution is suitable for fitting these datasets. The summary statistics for the PM10 concentrations data from the three pollution monitoring stations located in Chiang Mai are provided in Table 7. The estimated common CV was 0.4453. Note that we set r i = 2 and a i = b i = c i = d i = 10 4 ; i = 1 , 2 , , k for BayCrI and HPDI. Table 8 reports the 95% confidence intervals for the common CV of PM10 concentration data from three pollution monitoring stations in Chiang Mai, Thailand. Similar to the simulation results when ( n 1 , n 2 , n 3 ) = ( 30 , 30 , 30 ) , the average length of the LS confidence interval was the shortest, followed by MOVER. However, their coverage probabilities were under the nominal confidence level of 0.95, and so they are not recommended for constructing the confidence interval for the common CV of these datasets. When comparing GCI, BayCrI, and HPDI, although all three provided coverage probabilities greater than or close to the nominal confidence level of 0.95, the latter provided the shortest average length. Hence, HPDI is the most suitable method when considering the coverage probability and the average length together.

5. Conclusions

Herein, we propose confidence intervals for the common CV of several BS distributions constructed by using the GCI, MOVER, LS, BayCrI, and HPDI approaches. Their performances were studied numerically through Monte Carlo simulation in terms of their coverage probabilities and average lengths. The simulation results indicate that the coverage probabilities for GCI, BayCrI, and HPDI were greater than or close to the nominal confidence level, while HPDI produced the shortest average length for all cases. Therefore, HPDI is appropriate for constructing the confidence interval for the common CV of several BS distributions. Meanwhile, the coverage probabilities of the MOVER and LS confidence intervals were under the nominal confidence level, and so neither can be recommended as a solution for this scenario. Furthermore, when applying the methods to analyze PM10 concentrations from three pollution monitoring stations in Chiang Mai, Thailand, the results are in accordance with those from the simulation study.

Author Contributions

Conceptualization, S.N.; Data curation, W.P.; Formal analysis, W.P. and S.N.; Funding acquisition, S.N.; Investigation, S.-A.N. and S.N.; Methodology, S.-A.N. and S.N.; Project administration, S.-A.N.; Resources, S.-A.N.; Software, W.P.; Supervision, S.-A.N. and S.N.; Visualization, S.-A.N.; Writing original draft, W.P.; Writing–review and editing, S.-A.N. and S.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Science, Research and Innovation Fund (NSRF), and King Mongkut’s University of Technology North Bangkok with Contract no. KMUTNB-FF-65-23.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The real datasets of PM10 concentration were obtained from the Pollution Control Department [38].

Acknowledgments

The first author wishes to express gratitude for financial support provided by the Thailand Science Achievement Scholarship (SAST).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vassilopoulos, A.P. Fatigue life modeling and prediction methods for composite materials and structures—Past, present, and future prospects. In Fatigue Life Prediction of Composites and Composite Structures; Vassilopoulos, A.P., Ed.; Woodhead Publishing: Sawston, UK, 2020; pp. 1–43. [Google Scholar]
  2. Birnbaum, Z.W.; Saunders, S.C. A new family of life distributions. J. Appl. Probab. 1969, 6, 319–327. [Google Scholar] [CrossRef]
  3. Jin, X.; Kawczak, J. Birnbaum–Saunders and lognormal kernel estimators for modelling durations in high frequency financial data. Ann. Econ. Financ. 2003, 4, 103–124. [Google Scholar]
  4. Leao, J.; Leiva, V.; Saulo, H.; Tomazella, V. Birnbaum–Saunders frailty regression models: Diagnostics and application to medical data. Biom. J. 2017, 59, 291–314. [Google Scholar] [CrossRef]
  5. Lio, Y.L.; Tsai, T.R.; Wu, S.J. Acceptance sampling plans from truncated life tests based on the Birnbaum–Saunders distribution for percentiles. Commun. Stat. Simul. Comput. 2010, 39, 119–136. [Google Scholar] [CrossRef]
  6. Saulo, H.; Leiva, V.; Ziegelmann, F.A.; Marchant, C. A nonparametric method for estimating asymmetric densities based on skewed Birnbaum–Saunders distributions applied to environmental data. Stoch. Environ. Res. Risk. Assess. 2013, 27, 1479–1491. [Google Scholar] [CrossRef]
  7. Rice, J.A. Mathematical Statistics and Data Analysis; Duxbury Press: Belmont, CA, USA, 2006. [Google Scholar]
  8. Tian, L. Inferences on the common coefficient of variation. Stat. Med. 2005, 24, 2213–2220. [Google Scholar] [CrossRef]
  9. Verrill, S.; Johnson, R.A. Confidence bounds and hypothesis tests for normal distribution coefficients of variation. Commun. Statist. Theor. Meth. 2007, 36, 2187–2206. [Google Scholar] [CrossRef]
  10. Behboodian, J.; Jafari, A.A. Generalized confidence interval for the common coefficient of variation. J. Stat. Theory Appl. 2008, 7, 349–363. [Google Scholar]
  11. Ng, C.K. Inference on the common coefficient of variation when populations are lognormal: A simulation-based approach. J. Stat. Adv. Theory Appl. 2014, 11, 117–134. [Google Scholar]
  12. Thangjai, W.; Niwitpong, S.A. Confidence intervals for the weighted coefficients of variation of two-parameter exponential distributions. Cogent Math. 2017, 4, 1315880. [Google Scholar] [CrossRef]
  13. Liu, X.; Xu, X. A note on combined inference on the common coefficient of variation using confidence distributions. Electron. J. Stat. 2015, 9, 219–233. [Google Scholar] [CrossRef]
  14. Yosboonruang, N.; Niwitpong, S.A.; Niwitpong, S. Bayesian computation for the common coefficient of variation of delta-lognormal distributions with application to common rainfall dispersion in Thailand. PeerJ 2022, 10, e12858. [Google Scholar] [CrossRef] [PubMed]
  15. Birnbaum, Z.W.; Saunders, S.C. Estimation for a family of life distributions with applications to fatigue. J. Appl. Probab. 1969, 6, 328–347. [Google Scholar] [CrossRef]
  16. Engelhardt, M.; Bain, L.J.; Wright, F.T. Inference on the parameters of the Birnbaum–Saunders fatigue life distribution based on maximum likelihood estimation. Technometrics 1981, 23, 251–255. [Google Scholar] [CrossRef]
  17. Ng, H.K.T.; Kundu, D.; Balakrishnan, N. Modified moment estimation for the two-parameter Birnbaum–Saunders distribution. Comput. Stat. Data Anal. 2003, 43, 283–298. [Google Scholar] [CrossRef]
  18. Wu, J.; Wong, A.C.M. Improved interval estimation for the two-parameter Birnbaum-Saunders distribution. Comput. Stat. Data Anal. 2004, 47, 809–821. [Google Scholar] [CrossRef]
  19. Xu, A.; Tang, Y. Reference analysis for Birnbaum-Saunders distribution. Comput. Stat. Data Anal. 2010, 54, 185–192. [Google Scholar] [CrossRef]
  20. Wang, B.X. Generalized interval estimation for the Birnbaum-Saunders distribution. Comput. Stat. Data Anal. 2012, 56, 4320–4326. [Google Scholar] [CrossRef]
  21. Wang, M.; Sun, X.; Park, C. Bayesian analysis of Birnbaum-Saunders distribution via the generalized ratio-of-uniforms method. Comput. Stat. 2016, 31, 207–225. [Google Scholar] [CrossRef]
  22. Guo, X.; Wu, H.; Li, G.; Li, Q. Inference for the common mean of several Birnbaum–Saunders populations. J. Appl. Stat. 2017, 44, 941–954. [Google Scholar] [CrossRef]
  23. Puggard, W.; Niwitpong, S.A.; Niwitpong, S. Bayesian estimation for the coefficients of variation of Birnbaum–Saunders distributions. Symmetry 2021, 13, 2130. [Google Scholar] [CrossRef]
  24. Puggard, W.; Niwitpong, S.A.; Niwitpong, S. Confidence intervals for comparing the variances of two independent Birnbaum–Saunders distributions. Symmetry 2022, 14, 1492. [Google Scholar] [CrossRef]
  25. Weerahandi, S. Generalized confidence intervals. J. Am. Stat. Assoc. 1993, 88, 899–905. [Google Scholar] [CrossRef]
  26. Weerahandi, S. Generalized Inference in Repeated Measures: Exact Methods in MANOVA and Mixed Models; Wiley: Boston, NY, USA, 2004. [Google Scholar]
  27. Chen, P.; Ye, Z.-S. Estimation of field reliability based on aggregate lifetime data. Technometrics 2017, 59, 115–125. [Google Scholar] [CrossRef]
  28. Chen, P.; Ye, Z.-S. Approximate statistical limits for a gamma distribution. J. Qaul. Tech. 2017, 49, 64–77. [Google Scholar] [CrossRef]
  29. Chen, P.; Ye, Z.-S. Uncertainty quantification for monotone stochastic degradation models. J. Qaul. Tech. 2018, 50, 207–219. [Google Scholar] [CrossRef]
  30. Luo, C.; Shen, L.; Xu, A. Modelling and estimation of system reliability under dynamic operating environments and lifetime ordering constraints. Rel. Eng. Sys. Safe. 2022, 218, 108136. [Google Scholar] [CrossRef]
  31. Sun, Z.L. The confidence intervals for the scale parameter of the Birnbaum–Saunders fatigue life distribution. Acta Armamentarii 2009, 30, 1558–1561. [Google Scholar]
  32. Zou, G.Y.; Donner, A. Construction of confidence limits about effect measures: A general approach. Stat. Med. 2008, 27, 1693–1702. [Google Scholar] [CrossRef] [PubMed]
  33. Zou, G.Y.; Taleban, J.; Hao, C.Y. Confidence interval estimation for lognormal data with application to health economics. Comput. Stat. Data Anal. 2009, 53, 3755–3764. [Google Scholar] [CrossRef]
  34. Krishnamoorthy, K.; Oral, E. Standardized likelihood ratio test for comparing several log-normal means and confidence interval for the common mean. Stat. Methods Med. Res. 2015, 26, 2919–2937. [Google Scholar] [CrossRef] [PubMed]
  35. Xu, A.; Tang, Y. Bayesian analysis of Birnbaum–Saunders distribution with partial information. Comput. Stat. Data Anal. 2011, 55, 2324–2333. [Google Scholar] [CrossRef]
  36. Wakefield, J.C.; Gelfand, A.E.; Smith, A.F.M. Efficient generation of random variates via the ratio-of-uniforms method. Stat. Comput. 1991, 1, 129–133. [Google Scholar] [CrossRef]
  37. Box, G.E.P.; Tiao, G.C. Bayesian Inference in Statistical Analysis; Wiley: New York, NY, USA, 1992. [Google Scholar]
  38. Reports on Smog Situation in the North Home Page. Available online: http://www.pcd.go.th/ (accessed on 25 July 2022). (In Thai)
Table 1. The parameter settings for k = 3 , 5 , and 10.
Table 1. The parameter settings for k = 3 , 5 , and 10.
Scenarios ( n 1 , n 2 , , n k ) ( α 1 , α 2 , , α k )
k = 3
1–6 ( 30 3 ) ( 0.5 3 ) , ( 0.5 , 1.0 2 ) , ( 1.0 3 ) , ( 1.0 2 , 2.0 ) , (1.0, 1.5, 2.0), ( 1.5 , 2.0 2 )
7–12 ( 30 2 , 50 ) ( 0.5 3 ) , ( 0.5 , 1.0 2 ) , ( 1.0 3 ) , ( 1.0 2 , 2.0 ) , (1.0, 1.5, 2.0), ( 1.5 , 2.0 2 )
13–18 ( 50 3 ) ( 0.5 3 ) , ( 0.5 , 1.0 2 ) , ( 1.0 3 ) , ( 1.0 2 , 2.0 ) , (1.0, 1.5, 2.0), ( 1.5 , 2.0 2 )
19–24 ( 50 2 , 100 ) ( 0.5 3 ) , ( 0.5 , 1.0 2 ) , ( 1.0 3 ) , ( 1.0 2 , 2.0 ) , (1.0, 1.5, 2.0), ( 1.5 , 2.0 2 )
25–30 ( 100 3 ) ( 0.5 3 ) , ( 0.5 , 1.0 2 ) , ( 1.0 3 ) , ( 1.0 2 , 2.0 ) , (1.0, 1.5, 2.0), ( 1.5 , 2.0 2 )
k = 5
31–36 ( 30 2 , 50 3 ) ( 0.5 3 , 1.0 , 2.0 ) , ( 0.5 2 , 1.0 2 , 1.5 ) , ( 0.5 , 1.0 3 , 1.5 ) , (0.5, 1.0 2 , 2.0 2 ), ( 1.0 3 , 1.5 2 ), (1.0, 1.5, 2.0 3 )
37–42 ( 30 2 , 50 2 , 100 ) ( 0.5 3 , 1.0 , 2.0 ) , ( 0.5 2 , 1.0 2 , 1.5 ) , ( 0.5 , 1.0 3 , 1.5 ) , (0.5, 1.0 2 , 2.0 2 ), ( 1.0 3 , 1.5 2 ), (1.0, 1.5, 2.0 3 )
43–48 ( 30 , 50 2 , 100 2 ) ( 0.5 3 , 1.0 , 2.0 ) , ( 0.5 2 , 1.0 2 , 1.5 ) , ( 0.5 , 1.0 3 , 1.5 ) , (0.5, 1.0 2 , 2.0 2 ), ( 1.0 3 , 1.5 2 ), (1.0, 1.5, 2.0 3 )
49–54 ( 50 5 ) ( 0.5 3 , 1.0 , 2.0 ) , ( 0.5 2 , 1.0 2 , 1.5 ) , ( 0.5 , 1.0 3 , 1.5 ) , (0.5, 1.0 2 , 2.0 2 ), ( 1.0 3 , 1.5 2 ), (1.0, 1.5, 2.0 3 )
55–60 ( 50 2 , 100 3 ) ( 0.5 3 , 1.0 , 2.0 ) , ( 0.5 2 , 1.0 2 , 1.5 ) , ( 0.5 , 1.0 3 , 1.5 ) , (0.5, 1.0 2 , 2.0 2 ), ( 1.0 3 , 1.5 2 ), (1.0, 1.5, 2.0 3 )
k = 10
61–66( 30 5 , 50 5 )( 0.5 3 , 1.0 7 ) , ( 0.5 3 , 1.0 4 , 1.5 3 ) , ( 0.5 3 , 1.0 2 , 1.5 3 , 2.0 2 ) , ( 1.0 4 , 1.5 3 , 2.0 3 ) , ( 1.0 3 , 1.5 3 , 2.0 4 ), ( 1.0 2 , 1.5 2 , 2.0 6 )
67–72( 30 5 , 50 3 , 100 2 )( 0.5 3 , 1.0 7 ) , ( 0.5 3 , 1.0 4 , 1.5 3 ) , ( 0.5 3 , 1.0 2 , 1.5 3 , 2.0 2 ) , ( 1.0 4 , 1.5 3 , 2.0 3 ) , ( 1.0 3 , 1.5 3 , 2.0 4 ), ( 1.0 2 , 1.5 2 , 2.0 6 )
73–78( 30 3 , 50 4 , 100 3 )( 0.5 3 , 1.0 7 ) , ( 0.5 3 , 1.0 4 , 1.5 3 ) , ( 0.5 3 , 1.0 2 , 1.5 3 , 2.0 2 ) , ( 1.0 4 , 1.5 3 , 2.0 3 ) , ( 1.0 3 , 1.5 3 , 2.0 4 ), ( 1.0 2 , 1.5 2 , 2.0 6 )
79–84( 50 6 , 100 4 )( 0.5 3 , 1.0 7 ) , ( 0.5 3 , 1.0 4 , 1.5 3 ) , ( 0.5 3 , 1.0 2 , 1.5 3 , 2.0 2 ) , ( 1.0 4 , 1.5 3 , 2.0 3 ) , ( 1.0 3 , 1.5 3 , 2.0 4 ), ( 1.0 2 , 1.5 2 , 2.0 6 )
Table 2. The coverage probabilities and the average lengths of the 95% confidence intervals for the common CV of several BS distributions when k = 3 .
Table 2. The coverage probabilities and the average lengths of the 95% confidence intervals for the common CV of several BS distributions when k = 3 .
ScenariosCoverage ProbabilityAverage Length
GCIMOVERLSBayCrIHPDIGCIMOVERLSBayCrIHPDI
10.9310.9360.8510.9220.9120.15980.15880.14470.15900.1566
20.9560.8480.7990.9530.9480.28870.21050.19480.28730.2835
30.9480.9420.8780.9490.9430.28170.26570.25160.27950.2761
40.9570.8160.7270.9510.9460.40240.25830.24780.39020.3844
50.9610.8720.8120.9580.9550.38150.26150.25350.37030.3661
60.9470.9100.8790.9450.9500.30900.25400.25010.29860.2957
70.9360.9350.8770.9330.9220.14290.14200.13140.14240.1404
80.9490.8260.7740.9470.9420.27880.19760.18430.27750.2743
90.9520.9440.8870.9510.9430.25100.23940.22910.24880.2461
100.9520.8370.7810.9540.9500.34680.23080.22290.34160.3381
110.9460.8400.8060.9470.9430.32220.23280.22700.31650.3138
120.9530.9250.8960.9500.9480.26700.22700.22420.26020.2579
130.9500.9510.8970.9490.9440.12090.12060.11420.12050.1191
140.9530.8370.7890.9520.9430.22370.15970.15270.22310.2207
150.9460.9450.9070.9460.9390.21090.20380.19750.20940.2073
160.9410.8340.7830.9400.9360.29800.19810.19350.29250.2889
170.9490.8430.8060.9500.9430.28000.20080.19740.27580.2733
180.9370.9060.8940.9370.9320.22540.19480.19320.22140.2195
190.9260.9300.8890.9240.9180.10420.10370.09950.10370.1025
200.9480.8270.7950.9500.9520.21010.14600.14040.20910.2072
210.9510.9420.9250.9460.9410.18050.17600.17200.17930.1776
220.9520.8150.7650.9480.9460.24830.16850.16550.24630.2441
230.9460.8550.8140.9440.9420.22860.17020.16800.22630.2244
240.9400.9170.9030.9410.9370.18540.16610.16520.18320.1817
250.9400.9430.9090.9360.9260.08400.08370.08160.08380.0828
260.9480.8340.8070.9490.9460.15810.11170.10930.15720.1558
270.9590.9590.9330.9570.9540.14550.14320.14110.14460.1433
280.9460.8240.8120.9530.9430.20360.13890.13740.20180.1999
290.9510.8470.8370.9500.9460.19190.14100.13980.18970.1881
300.9540.9320.9240.9480.9480.15190.13680.13630.15040.1491
Table 3. The coverage probabilities and the average lengths of the 95% confidence intervals for the common CV of several BS distributions when k = 5 .
Table 3. The coverage probabilities and the average lengths of the 95% confidence intervals for the common CV of several BS distributions when k = 5 .
ScenariosCoverage ProbabilityAverage Length
GCIMOVERLSBayCrIHPDIGCIMOVERLSBayCrIHPDI
310.9530.7060.6120.9480.9420.23750.12250.11470.23340.2293
320.9450.7500.6640.9430.9320.23650.14170.13250.23470.2323
330.9440.7490.6840.9470.9480.25690.15600.14750.25500.2526
340.9590.6310.5900.9570.9500.34810.14980.14270.34310.3399
350.9540.9000.8400.9510.9480.21650.17670.17110.21430.2124
360.9580.8910.8580.9600.9630.21980.16290.16020.21590.2140
370.9590.6090.5440.9530.9430.25110.11470.10800.24830.2453
380.9310.7170.6340.9270.9240.23760.13270.12480.23670.2346
390.9530.7260.6860.9520.9450.24770.14380.13680.24600.2441
400.9450.5680.5290.9490.9380.31380.13610.13040.31090.3084
410.9470.8790.8280.9500.9470.19260.15940.15510.19100.1893
420.9540.8970.8580.9540.9510.18740.14550.14350.18460.1830
430.9490.6320.5800.9480.9310.21530.10290.09800.21370.2111
440.9610.7480.7040.9630.9550.19980.11570.11040.19890.1970
450.9490.7470.6880.9420.9380.21180.12920.12400.21050.2086
460.9450.5910.5560.9460.9410.27460.12210.11780.27260.2703
470.9410.8910.8510.9430.9410.16790.14130.13840.16680.1654
480.9560.8880.8680.9560.9600.16050.12910.12770.15860.1573
490.9410.7410.6780.9390.9370.19640.10790.10280.19310.1897
500.9420.7870.7220.9400.9320.18700.12010.11480.18600.1839
510.9580.7870.7400.9510.9460.22180.13570.13040.22010.2181
520.9530.6340.5990.9580.9520.30620.13130.12680.30140.2982
530.9390.8780.8310.9370.9340.19300.16080.15680.19110.1893
540.9420.8390.8060.9470.9480.21120.15090.14900.20730.2054
550.9580.6900.6510.9540.9520.17040.08810.08520.16870.1666
560.9610.7440.6820.9570.9520.17460.10360.10000.17400.1725
570.9480.7650.7280.9460.9390.18500.11330.11010.18410.1826
580.9500.5970.5730.9500.9490.24800.10790.10520.24640.2445
590.9440.8970.8660.9500.9450.15020.12730.12520.14960.1483
600.9510.8760.8490.9520.9440.14700.11680.11590.14560.1444
Table 4. The coverage probabilities and the average lengths of the 95% confidence intervals for the common CV of several BS distributions when k = 10 .
Table 4. The coverage probabilities and the average lengths of the 95% confidence intervals for the common CV of several BS distributions when k = 10 .
ScenariosCoverage ProbabilityAverage Length
GCIMOVERLSBayCrIHPDIGCIMOVERLSBayCrIHPDI
610.9210.7360.6340.9210.9100.15940.10600.09920.15870.1573
620.9320.6540.5460.9250.9230.19750.10710.10040.19540.1938
630.9450.5340.4560.9360.9300.25080.10560.09940.24710.2448
640.9510.7740.6950.9480.9430.18390.12290.11970.18000.1783
650.9580.8220.7530.9550.9580.17730.12130.11860.17360.1721
660.9370.8400.7810.9440.9480.16790.11850.11640.16370.1622
670.9440.7470.6170.9380.9340.14810.09760.09200.14700.1459
680.9330.6230.5340.9310.9230.19080.09930.09370.18990.1882
690.9450.5160.4540.9420.9370.23690.09590.09090.23480.2328
700.9510.8120.7600.9520.9470.15370.10810.10580.15190.1506
710.9530.8430.7870.9570.9550.14820.10680.10490.14600.1447
720.9550.8700.8240.9600.9580.14110.10490.10340.13790.1367
730.9190.7360.6480.9200.9160.13830.09130.08670.13740.1363
740.9310.6310.5390.9290.9230.17850.09330.08860.17770.1762
750.9550.5820.4990.9540.9440.21470.09060.08640.21330.2115
760.9560.8280.7880.9570.9550.13810.09920.09750.13680.1356
770.9530.8200.7740.9480.9450.13240.09850.09700.13080.1297
780.9410.8740.8400.9400.9370.12420.09640.09530.12240.1214
790.9400.7510.6890.9380.9370.11900.07970.07670.11850.1174
800.9560.6520.5850.9560.9520.14940.08070.07780.14890.1476
810.9410.5570.5150.9370.9340.18570.07910.07660.18430.1827
820.9490.8350.7860.9470.9490.12880.09160.09030.12750.1265
830.9440.8110.7880.9420.9370.12330.09020.08910.12210.1210
840.9540.8680.8360.9580.9600.11700.08880.08790.11550.1145
Table 5. AIC results for the fitting of five tested distributions.
Table 5. AIC results for the fitting of five tested distributions.
DistributionsLognormalBSExponentialGammaWeibull
Chang Phueak334.4613333.9956366.2775335.8974339.0671
Si Phum326.8568326.2713351.1782328.6437331.6056
Changkerng301.5002301.1981337.6783303.2755307.9442
Table 6. BIC results for the fitting of five tested distributions.
Table 6. BIC results for the fitting of five tested distributions.
DistributionsLognormalBSExponentialGammaWeibull
Chang Phueak337.3293336.8636367.7115338.7653341.9351
Si Phum329.7248329.1392352.6122331.5116334.4735
Changkerng304.3682304.1661339.1123306.1435310.8121
Table 7. Summary statistics for the PM10 data.
Table 7. Summary statistics for the PM10 data.
AreanMin.MedianMeanMax.VarianceCV
Chang Phueak3161122131.03232823310.5560.4391
Si Phum314285102.70972482798.6800.5151
Changkerng31438182.61291821191.0450.4177
Table 8. The 95% confidence interval for the common CV of PM10 data from three pollution monitoring stations in Chiang Mai, Thailand.
Table 8. The 95% confidence interval for the common CV of PM10 data from three pollution monitoring stations in Chiang Mai, Thailand.
MethodsIntervalLength
GCI0.3788–0.51630.1375
MOVER0.3860–0.52120.1352
LS0.3698–0.49720.1274
BayCrI0.3796–0.51600.1364
HPDI0.3727–0.50590.1332
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Puggard, W.; Niwitpong, S.-A.; Niwitpong, S. Confidence Intervals for Common Coefficient of Variation of Several Birnbaum–Saunders Distributions. Symmetry 2022, 14, 2101. https://doi.org/10.3390/sym14102101

AMA Style

Puggard W, Niwitpong S-A, Niwitpong S. Confidence Intervals for Common Coefficient of Variation of Several Birnbaum–Saunders Distributions. Symmetry. 2022; 14(10):2101. https://doi.org/10.3390/sym14102101

Chicago/Turabian Style

Puggard, Wisunee, Sa-Aat Niwitpong, and Suparat Niwitpong. 2022. "Confidence Intervals for Common Coefficient of Variation of Several Birnbaum–Saunders Distributions" Symmetry 14, no. 10: 2101. https://doi.org/10.3390/sym14102101

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