Next Article in Journal
Retrieval of Atmospheric Temperature Profiles from FY-4A/GIIRS Hyperspectral Data Based on TPE-MLP: Analysis of Retrieval Accuracy and Influencing Factors
Previous Article in Journal
Unsupervised Domain Adaptation with Contrastive Learning-Based Discriminative Feature Augmentation for RS Image Classification
Previous Article in Special Issue
SAR Features and Techniques for Urban Planning—A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Identifying Heterogeneity in SAR Data with New Test Statistics

by
Alejandro C. Frery
1,*,
Janeth Alpala
2 and
Abraão D. C. Nascimento
2
1
School of Mathematics and Statistics, Victoria University of Wellington, Wellington 6140, New Zealand
2
Departamento de Estatística, Universidade Federal de Pernambuco, Recife 50670-901, PE, Brazil
*
Author to whom correspondence should be addressed.
Remote Sens. 2024, 16(11), 1973; https://doi.org/10.3390/rs16111973
Submission received: 6 May 2024 / Revised: 29 May 2024 / Accepted: 29 May 2024 / Published: 30 May 2024
(This article belongs to the Special Issue SAR Processing in Urban Planning)

Abstract

:
This paper presents a statistical approach to identify the underlying roughness characteristics in synthetic aperture radar (SAR) intensity data. The physical modeling of this kind of data allows the use of the Gamma distribution in the presence of fully developed speckle, i.e., when there are infinitely many independent backscatterers per resolution cell, and none dominates the return. Such areas are often called “homogeneous” or “textureless” regions. The G I 0 distribution is also a widely accepted law for heterogeneous and extremely heterogeneous regions, i.e., areas where the fully developed speckle hypotheses do not hold. We propose three test statistics to distinguish between homogeneous and inhomogeneous regions, i.e., between gamma and G I 0 distributed data, both with a known number of looks. The first test statistic uses a bootstrapped non-parametric estimator of Shannon entropy, providing a robust assessment in uncertain distributional assumptions. The second test uses the classical coefficient of variation (CV). The third test uses an alternative form of estimating the CV based on the ratio of the mean absolute deviation from the median to the median. We apply our test statistic to create maps of p-values for the homogeneity hypothesis. Finally, we show that our proposal, the entropy-based test, outperforms existing methods, such as the classical CV and its alternative variant, in identifying heterogeneity when applied to both simulated and actual data.

1. Introduction

Synthetic Aperture Radar (SAR) technology has become essential for environmental monitoring and disaster management. It provides valuable images under various conditions, including day or night and weather situations [1,2]. However, the effective use of SAR data depends on a thorough understanding of its statistical properties because it is corrupted by speckle. This noise-like interference effect is inherent in SAR data due to the coherent nature of the imaging process [3].
Speckle in intensity format is non-Gaussian. Thus, SAR data require reliable statistical models for accurate processing. The G 0 distribution, which is suitable for SAR data, includes the Gamma law as the limiting case for fully developed speckle [4] and provides flexibility with fewer parameters for analysis.
Our work aims to improve the identification of potential roughness features in SAR intensity data. The physical modeling of SAR data allows the use of the Gamma distribution in the presence of fully developed speckle, where an infinite number of independent backscatterers per resolution unit is assumed, which is commonly referred to as homogeneous regions.
In this context, we present a set of three novel test statistics that aim to distinguish between homogeneous and non-homogeneous returns, particularly between gamma and G 0 distributed data, assuming the number of looks is known. We use properties such as entropy and coefficient of variation.
Entropy is a fundamental concept in information theory with far-reaching applications in pattern recognition, statistical physics, image processing, edge detection and SAR image analysis [5,6,7,8,9]. Shannon introduced it in 1948 [10] for a random variable to measure information and uncertainty. Shannon entropy is a crucial descriptive parameter in statistics, especially for evaluating data dispersion and performing tests for normality, exponentiality and uniformity [11,12]. Entropy estimation is challenging, especially when the model is unknown; in these cases, non-parametric methods, as those based on spacings [13,14], can be used. This strategy is flexible and robust because it does not enforce a model or parametric constraints. Different roughness levels, which materialize as different models for SAR data, have different entropy values. We designed a bootstrap-improved non-parametric estimator for the Shannon entropy.
The coefficient of variation (CV), introduced in 1896 by Pearson [15], is a relative dispersion measure widely used in various fields of applied statistics, including sampling, biostatistics, medical and biological research, climatology and other fields [16,17,18,19]. It facilitates the comparison of variability between different populations and is particularly valuable for relating variables with different units. This is because when the primary purpose is to compare the variations of several variables, the standard deviation can only serve as an adequate measure of variation if all variables are expressed in the same unit of measurement and have identical means. If these conditions are not met, then the CV is the relative measure that is usually used in real applications. The variable with the highest CV value has the largest relative dispersion around the mean value [20]. The coefficient of variation is the primary measure of heterogeneity in SAR data [21,22]. We study two ways of estimating the coefficient of variation.
We devise test statistics based on these three estimators: the classical coefficient of variation, a robust version, and the Shannon entropy estimator. We apply these test statistics to generate maps of evidence of homogeneity that reveal different types of targets in the SAR data. We show that our proposed method is superior to existing approaches with simulated data and SAR images.
The article is structured as follows: Section 2 deals with statistical modeling and entropy estimation for intensity SAR data. Section 3 outlines hypothesis tests based on non-parametric entropy and coefficients of variation estimators. In Section 4, we present experimental results. Finally, we draw conclusions in Section 5.
We used Rmarkdown https://andrewbtran.github.io/NICAR/2017/reproducibility-rmarkdown/rmarkdown.html#reproducible_research (accessed on 30 April 2024) to produce this work. It is fully reproducible because it contains the code and data needed to replicate it. The Supplementary Materials information points to the repository from where it can be downloaded.

2. Background

2.1. Statistical Modeling of Intensity SAR Data

The primary models for intensity SAR data include the Gamma and G I 0 distributions [23]. The first is suitable for fully developed speckle and is a limiting case of the second model. This is interesting due to its versatility in accurately representing regions with different roughness properties [24]. We denote Z Γ SAR ( L , μ ) and Z G I 0 ( α , γ , L ) to indicate that Z follows the distributions characterized by the respective probability density functions (pdfs):
f Z ( z ; L , μ Γ SAR ) = L L Γ ( L ) μ L z L 1 exp L z / μ 1 R + ( z )
and
f Z ( z ; α , γ , L G I 0 ) = L L Γ ( L α ) γ α Γ ( α ) Γ ( L ) · z L 1 ( γ + L z ) L α 1 R + ( z ) ,
where μ > 0 is the mean, γ > 0 is the scale, α < 0 measures the roughness, L 1 is the number of looks (either nominal or estimated, thus not restricted to integer values), Γ ( · ) is the gamma function, and 1 A ( z ) is the indicator function of the set A.
The rth order moments of the G I 0 model are
E Z r G I 0 = γ L r Γ ( α r ) Γ ( α ) · Γ ( L + r ) Γ ( L ) ,
provided α < r , and they are infinite otherwise. Therefore, assuming α < 1 , its expected value is
μ = γ L Γ ( α 1 ) Γ ( α ) · Γ ( L + 1 ) γ ( L ) = γ α + 1 .
Although the G I 0 distribution is defined by the parameters α and γ , in the SAR literature [25], the texture α and the mean μ are usually used. Reparametrizing (2) with μ , and denoting this model as G I 0 , we obtain the following:
f Z z ; μ , α , L G I 0 = L L Γ ( L α ) μ ( α + 1 ) α Γ ( α ) Γ ( L ) z L 1 μ ( α + 1 ) + L z L α .

2.2. The Shannon Entropy

The parametric representation of Shannon entropy for a system described by a continuous random variable is
H ( Z ) = f ( z ) ln f ( z ) d z ,
here, f ( · ) is the pdf that characterizes the distribution of the real-valued random variable Z.
Using (6), we obtain the Shannon entropy of Γ SAR in (1) and G I 0 in (5):
H Γ SAR ( L , μ ) = L ln L + ln Γ ( L ) + ( 1 L ) ψ ( 0 ) ( L ) + ln μ ,
H G I 0 ( μ , α , L ) = L ln L + ln Γ ( L ) + ( 1 L ) ψ ( 0 ) ( L ) + ln μ ln Γ ( L α ) + ( L α ) ψ ( 0 ) ( L α ) ( 1 α ) ψ ( 0 ) ( α ) + ln ( 1 α ) + ln Γ ( α ) L ,
where ψ ( 0 ) ( · ) is the digamma function. Figure 1 shows the entropy of G I 0 as a function of μ when α , 20 , 8 , 3 . Notice that it converges to the entropy of Γ SAR when α , as expected. The more heterogeneous (large α values) the SAR region is, the larger the entropy (or degree of disorder) is.
Appendix A provides the proof of the convergence of H G I 0 ( μ , α , L ) to H Γ SAR ( L , μ ) when α .

2.3. Estimation of the Shannon Entropy

The problem of the non-parametric estimation of H ( Z ) has been studied by many authors, including [11,26,27,28]. Their proposals use estimators based on differences between order statistics: spacings.
Vasicek [26] introduced one of the first non-parametric estimators based on spacings. Under the assumption that Z = ( Z 1 , Z 2 , , Z n ) is a random sample from the distribution F ( z ) , the estimator is defined as
H ^ V ( Z ) = 1 n i = 1 n ln n 2 m Z ( i + m ) Z ( i m ) ,
where m < n / 2 is a positive integer, Z ( 1 ) Z ( 2 ) Z ( n ) are the order statistics, and Z ( i + m ) Z ( i m ) is the m-spacing, in which Z ( i ) = Z ( 1 ) if i < 1 , Z ( i ) = Z ( n ) if i > n .
Several authors have explored adaptations to Vasicek’s estimator. We consider three estimators known for their superior performance [24]:
  • Correa [27]:
    H ^ C ( Z ) = 1 n i = 1 n log j = i m i + m ( j i ) Z ( j ) Z ¯ ( i ) n j = i m i + m Z ( j ) Z ¯ ( i ) 2 ,
    where Z ¯ ( i ) = ( 2 m + 1 ) 1 j = i m i + m Z ( j ) , m < n 2 , Z ( i ) = Z ( 1 ) for i < 1 and Z ( i ) = Z ( n ) for i > n . Based on simulations, he showed that his estimator has a smaller mean square error than Vasicek’s approach.
  • Ebrahimi et al. [29]:
    H ^ E ( Z ) = 1 n i = 1 n log n c i m Z ( i + m ) Z ( i m ) ,
    where
    c i = 1 + ( i 1 ) / m if 1 i m , 2 if m + 1 i n m , 1 + ( n i ) / m if n m + 1 i n .
  • Al-Omari [30]:
    H ^ AO ( Z ) = 1 n i = 1 n log n ω i m Z ( i + m ) Z ( i m ) ,
    where
    ω i = 3 / 2 if 1 i m , 2 if m + 1 i n m , 3 / 2 if n m + 1 i n ,
    in which Z ( i m ) = Z ( 1 ) for i m , and Z ( i + m ) = Z ( n ) for i n m .
These estimators are asymptotically consistent, i.e., they converge in probability to the true value when m , n and m / n 0 . However, we will use improved bootstrap-improved versions because we need them to perform well with small samples.

2.4. Enhanced Estimators with Bootstrap

We use the bootstrap technique to refine the accuracy of non-parametric entropy estimators. In this approach, new datasets are generated by replicate sampling from an existing dataset [31].
Let us assume that the non-parametric entropy estimator H ^ = θ ^ ( Z ) is inherently biased, i.e.,
Bias θ ^ ( Z ) = E θ ^ ( Z ) θ 0 .
Our bootstrap-improved estimator is of the following form:
H ˜ = 2 θ ^ ( Z ) 1 B b = 1 B θ ^ b ( Z ( b ) ) ,
where B is the number of observations obtained by resampling from Z with replacement. Applying this methodology, the original estimators of Correa, Ebrahimi and Al-Omari are now referred to as the proposed bootstrap-improved versions: H ˜ C , H ˜ E , and H ˜ AO , respectively.
We analyzed the performance of these estimators with a Monte Carlo study: 1000 samples from the Γ SAR distribution of size n 9 , 25 , 49 , 81 , 121 with μ 1 , 10 and L = 5 . The results are consistent with other situations. We used B = 200 bootstrap samples and the heuristic spacing m = n + 0.5 , as recommended in the literature.
In Figure 2, we show the bias and mean squared error (MSE) of the original non-parametric entropy estimators and their respective bootstrap-enhanced versions. Bootstrap-enhanced estimators have smaller bias and MSE, especially for sample sizes below 81. The results of the simulation can be found in Table 1.

2.5. Coefficient of Variation and a Robust Alternative

The population CV is defined as a ratio of the population standard deviation ( σ ) to the population mean ( μ ) :
CV = σ μ , μ 0 .
The CV can be easily estimated as the ratio of the sample mean to the sample standard deviation.
We explore a robust alternative to estimate the CV, as described by [32]: the ratio between the mean absolute deviation from the median (MnAD) and the median, two well-known robust measures of scale and location, respectively. The sample version for the MnAD is defined as n 1 i = 1 n | x i Q ^ 2 | , where Q ^ 2 is an estimate for the median of the population, for example, the sample median.

3. Hypothesis Testing

We aim to test the following hypotheses:
H 0 : The data come from the Γ SAR law , H 1 : The data come from the G I 0 distribution .
We are testing the hypothesis that the data are fully developed speckle versus the alternative of data with roughness. As for the parametric problem, once it is not possible to define the hypothesis H 0 = α = , it is impossible to solve this problem with parametric inference alternatives (such as likelihood ratio, score, gradient, and Wald hypothesis test). The proposed tests to solve this physical problem in SAR systems are described below.

3.1. The Proposed Test Based on Non-Parametric Entropy

For a random sample Z = ( Z 1 , Z 2 , , Z n ) from a distribution D , a test statistic is proposed. It is based on an empirical distribution that arises from the difference between non-parametrically estimated entropies H ˜ ( Z ) and the analytical entropy of Γ SAR (7) evaluated at the logarithm of the sample mean, where L 1 is known.
Hence, the entropy-based test statistic is defined as
S ( Z ; L ) = H ˜ ( Z ) H Γ SAR ( L ) + ln Z ¯ .
This test statistic aims to assess the behavior of the data under the null hypothesis using the empirical distribution. If the data represent fully developed speckle, the density should center around zero, i.e., S ( Z ; L ) 0 . Otherwise, the empirical distribution would shift from zero under the alternative hypothesis, suggesting significant differences and heterogeneous clutter.
The comparison between the bootstrap-improved estimators is shown in Table 2, where the test accuracy under the null hypothesis is presented alongside running times. The test accuracy is evaluated through 1000 simulated samples of different sizes with each size replicated 100 times using bootstrap resampling.
The processing time is an important feature, especially considering the application of these estimators to large datasets of SAR images, as seen in Section 4.
As visible from Table 2, the accuracy of the test results across the three estimators shows similarities in specific sample sizes. However, practical scenarios in SAR image processing often involve small sample sizes, which are typically obtained over windows of size 7 × 7 . It is also noteworthy that the H ˜ AO estimator exhibited the shortest processing time, which was followed by H ˜ E and H ˜ C . Considering this aspect, we select the H ˜ AO estimator for subsequent simulations. Henceforth, the test statistical (13) will be denoted as S H ˜ AO ( Z ; L ) .
We now verify the normality of the data generated by the S H ˜ AO ( Z ; L ) test. Figure 3 shows the empirical densities obtained by applying the S H ˜ AO ( Z ; L ) test to different sample sizes drawn from the Γ SAR distribution, where L takes values 3 , 5 , 8 , 11 and μ = 1 . Additionally, Table 3 summarizes the main descriptive statistics, including mean, standard deviation (SD), variance (Var), skewness (SK), excessive kurtosis (EK) and Anderson–Darling p values for normality. Results with p values greater than 0.05 do not indicate a violation of the normality assumption. A low variance, skewness, and excessive kurtosis of almost zero indicate limited dispersion, asymmetry, and a light tail. Normal Q–Q plots confirm no evidence against a normal distribution, as shown in Figure 4.
After checking the data’s normality, we examined the proposed test’s abilities in terms of size (Type I error rate) and power (probability of correctly rejecting the null hypothesis when the alternative hypothesis is true).
Under H 0 , the distribution of the test statistic is asymptotically normal. Therefore, the p values are calculated as 2 Φ ( | ε | ) , where Φ is the standard Gaussian cumulative distribution function, and ε is the standardized test statistic given by
ε = H ˜ AO ( Z ) H Γ SAR ( L ) + ln Z ¯ σ ^ .
We conducted the test at nominal significance levels of 1%, 5%, and 10%. These levels represent the probability of rejecting the null hypothesis when it is actually true (Type I error rates). To verify that the test maintains these nominal levels, we performed 1000 simulations under the null hypothesis ( Γ SAR distribution) with different sample sizes and values of L, and μ = 1 . In all cases, the observed Type I error rates were consistent with the nominal levels, indicating that the test maintains the specified significance levels.
We also evaluated the test’s power by performing 1000 simulations under the alternative hypothesis ( G I 0 distribution) with μ = 1 and α = 2 . The power generally improves with increasing sample size and number of looks. The results are shown in Table 4.

3.2. The Proposed Test Based on Coefficient of Variation and a Robust Alternative

In addition to the S H ˜ AO ( Z ; L ) test, we also propose a test statistic based on the classical CV. This test statistic is defined as follows:
T CV = S Z ¯ ,
where S and Z ¯ are the sample standard deviation and mean, respectively.
Similarly, we use another test statistic based on the ratio of the MnAD to the median. This statistic is given by
T CV MnAD = MnAD Median .
We proceed to identify suitable models for these estimators of the CV and then form test statistics.
The situations in which the use of CV and CV MnAD may be appropriate, i.e., when the observations are positive, the log-normal (LN) and the inverse Gaussian distribution (IG) are often more appropriate than the Gamma and Weibull distributions [33,34].
It is shown that the IG distribution is well approximated by the log-normal distribution, which means that the IG distribution also does not share the problem of the non-existence of a fixed-width confidence interval with the Gaussian case [35].
The biparametric LN distribution has the following density:
f Z ( z ; μ LN , σ LN ) = 1 σ LN z 2 π exp ( ln z μ LN ) 2 2 σ LN 2 1 R + ( z ) ,
where μ LN is any real number, and σ LN is positive.

3.3. Model Selection Criterion

We used the Akaike information criterion (AIC) and the Bayesian information criterion (BIC) to select the best-fitting distribution.
The AIC deals with the trade-off between the goodness-of-fit and the model’s simplicity in terms of the number of model parameters [36]. The model or distribution with the lowest value of AIC is chosen to be the best. The BIC assesses the goodness-of-fit of a distribution or model, but it avoids overfitting by penalizing additional degrees of freedom [37]. The model with the lowest BIC value is chosen as the best.
The AIC and BIC results in Table 5, Table 6, Table 7 and Table 8 indicate that the CV and CV MnAD data from different distributed Γ SAR and G I 0 synthetic sample sizes match the properties of an LN distribution. It is important to note that this conclusion was drawn empirically based on a dictionary of analytically tractable distributions and is well-defined under biparametric, unimodal, asymmetric, and positive distributions.
Figure 5, Figure 6, Figure 7 and Figure 8 show empirical and fitted density plots, Q–Q plots, P–P plots, as well as empirical and fitted cumulative distribution functions. They provide qualitative sources that confirm that the LN distribution is the most appropriate distribution. As expected, both test statistics work well under the null hypothesis. Under the alternative hypothesis, the P–P plot shows that CV MnAD is more robust than CV, although both statistics suffer from the tail effect caused by the distributed G I 0 data.

4. Results

This section presents the simulations we performed to evaluate the proposed test statistics’ performance, which was followed by applications to SAR data.

4.1. Simulated Data

Figure 9a shows the phantom with dimensions of 500 × 500 pixels. It was proposed by Gomez et al. [38] as a tool to assess the performance of speckle-reduction filters.
Figure 9b shows the simulated image with each small phantom displaying texture variations. The observations are independent draws from the G I 0 distribution (5), with L = 5 and varying α and μ , which are annotated in the image for each quadrant. Light regions correspond to textured observations (heterogeneous), while darker regions represent textureless areas (homogeneous).
The α parameter of the G I 0 distribution is essential for interpreting texture characteristics. Values near zero greater than 3 suggest extremely textured targets, such as urban zones [39]. As the value decreases, it indicates regions with moderate texture (in the 6 , 3 region), related to forest zones, while values below 6 correspond to textureless regions, such as pasture, agricultural fields, and water bodies [40].
We applied the three test statistics, namely S H ˜ AO ( Z ; L ) , T CV , and T CV MnAD , to the simulated image using local sliding windows of size 7 × 7 , as shown in Figure 10a–c.
The resulting p-values for each test are shown in Figure 11a–c. In Figure 12a–c, maps are depicted using a color table between black, gray levels, and white. All p-values above 0.05 are represented in white (indicating no evidence to reject the null hypothesis), while those below 0.05 are shown in black (indicating evidence to reject the hypothesis). We notice that the S H ˜ AO ( Z ; L ) performs significantly better than the other tests in identifying heterogeneous areas in the simulated image.

4.2. SAR Data

We evaluated the proposed test statistics using four SAR images: one of Rotterdam, Netherlands, acquired by the TerraSAR-X satellite with Single-look Slant-range Complex (SSC) and High-resolution Spotlight-mode (HS); another of the Coast of Jalisco, Mexico; and two of Illinois, USA, acquired by the Sentinel-1B in Ground Range Detected (GRD) format, with Extra Wide-swath mode (EW) and Strip Map mode (SM). Table 9 provides detailed parameters of the selected SAR images. They contain mountainous areas, agricultural regions, water bodies, and urban areas, as shown in Figure 13a–d.
We used the number of looks L informed in the SNAP (Sentinel Application Platform) metadata: the product of azimuth and range looks. This value is close to the equivalent number of looks calculated as the ratio of the squared sample mean to the sample variance in areas with fully developed speckle.
Figure 14a–d show the optical images obtained from Sentinel-2 L2A in true colors (composition of RGB channels) for the corresponding SAR images with acquisition dates as follows: Rotterdam (20-12-2018), Coast of Jalisco (05-09-2021), Illinois-Region 1 (25-06-2020), and Illinois-Region 2 (01-08-2021). These optical images facilitate the interpretation of the results, making it easier to distinguish between homogeneous and heterogeneous regions.
The three statistical tests are applied to the SAR images using 7 × 7 local sliding windows. This window size balances detailed local information and a sufficient number of data points, as illustrated in Figure 15, Figure 16, Figure 17 and Figure 18.
The p-values obtained for each test are presented in Figure 19, Figure 20, Figure 21 and Figure 22, respectively.
In Figure 23, Figure 24, Figure 25 and Figure 26, the maps of p-values composed of a linear gradient of black and white colors represent the decisions at a 5 significance level. Dark areas represent values below 0.05 , indicating evidence to reject the null hypothesis and suggesting heterogeneity in these regions. In contrast, values above 0.05 are represented as white areas, indicating no evidence to reject the fully developed speckle hypothesis.
Using Shannon entropy is more meaningful than using the original and robust CV to capture heterogeneity, particularly for Sentinel-1B images with a high number of looks. It is justified that the dark areas of the maps based on the T CV and T CV MnAD show coverage patterns similar to those reported for the S H ˜ AO ( Z ; L ) map. This suggests that although CV-based tests may produce slightly less pronounced results than the entropy-based test, they still demonstrate a comparable ability to detect heterogeneity within SAR images.
It is noticeable that the entropy and CV-based tools predicted heterogeneity regions and boundaries where the statistical properties of texture vary. The T CV MnAD test was shown to be an effective edge detector. It emerges as a robust alternative to the classical CV test, making it less susceptible to the influence of outliers and allowing it to produce more precise edges. Considering a higher significance level may increase the sensitivity to edge detection but also increase the risk of detecting false heterogeneous regions.
Additionally, assuming a 5 threshold for p-values, in most cases, the heterogeneous regions detected by the S H ˜ AO ( Z ; L ) test were more extensive than those detected by the T CV and T CV MnAD tests. This was mainly observable in Figure 12a, Figure 24a and Figure 25a.
Conversely, when applying the same tests to a high-resolution TerraSAR-X image of Rotterdam, with a single look, the S H ˜ AO ( Z ; L ) test detected less heterogeneity compared to the CV-based tests. This indicates that the performance of the entropy test depends significantly on the number of looks, performing better with a higher number of looks as in the Sentinel-1B images.

4.3. Quantitative Analysis of Test Statistics

In order to conduct a more quantitative analysis of the results, we compared the mean and standard deviation values of the different tests among selected reference areas in the Illinois-Region 2 image. These reference areas represent water, vegetation, and urban land covers, as shown in Figure 27.
We selected labeled polygons (water, vegetation, urban) and conducted 100 sampling iterations with replacement from each category, using window sizes of 11 × 11 pixels. For each sample, we applied the corresponding test statistic. Finally, we computed the mean and standard deviation of the results obtained from the 100 samples.
Table 10 presents the comparison of the mean and standard deviation values for the three tests across the reference areas.
Based on the analysis, S H ˜ AO ( Z ; L ) appears to be the most sensitive test for classification purposes, because it shows the largest distinction in mean values between the different land cover types, indicating it is better at differentiating between homogeneous and heterogeneous areas. However, the CV and CV MnAD test are also reliable with a lower variability, making them good candidates depending on the specific application.

5. Conclusions

This article provides a practical and theoretical answer to the following physical question: how does on detect heterogeneity in SAR images, assuming that the SAR intensity follows the Γ SAR model? To this end, we proposed three novel hypothesis tests: one from the Shannon entropy and two from the variation coefficient variants. The performance of our proposals was evaluated using a Monte Carlo study. The results showed that they were conservative in estimating the probability of a Type I error (false alarm rate) and the test power (probability of detection), which increases with sample size. An application to four recent SAR images was performed. The results showed that the Shannon entropy-based test was more robust than the CV-based tests for Sentinel-1B images. For high-resolution images with a single look, such as the TerraSAR-X image, the classical CV test was more effective in detecting heterogeneous regions. In addition, all tests could recognize images with different textures and identify edges where the texture type changes.
Despite presenting the results of using the estimators as local filters, we suggest using them in specific samples that visibly contain only one type of land use or land cover. This type of analysis can assist users in selecting appropriate training samples for classification purposes.

Supplementary Materials

This article was written in Rmarkdown and is fully reproducible. The code and data are accessible at https://github.com/rjaneth/identifying-heterogeneity-in-sar-data-with-new-test-statistics (accessed on 30 April 2024).

Author Contributions

All authors (A.C.F., J.A. and A.D.C.N.) discussed the results and contributed to the research. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The Sentinel-1 images are made publicly available by the European Space Agency via Copernicus Data Space Ecosystem at https://dataspace.copernicus.eu/ (accessed on 24 April 2024).

Acknowledgments

The author J.A. would like to thank the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES) and the Fundação de Amparo à Ciência e Tecnologia de Pernambuco (FACEPE) for their support.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A. Limit Behavior of H G I 0 as α → −∞

To verify that H G I 0 ( μ , α , L ) converges to H Γ SAR ( L , μ ) as α , we show that the additional terms in H G I 0 cancel in the limit.
The Shannon entropy for the G I 0 distribution is given by
H G I 0 ( μ , α , L ) = H Γ SAR ( L , μ ) ln Γ ( L α ) + ( L α ) ψ ( 0 ) ( L α ) ( 1 α ) ψ ( 0 ) ( α ) + ln ( 1 α ) + ln Γ ( α ) L .
We aim to show that
lim α H G I 0 ( μ , α , L ) = H Γ SAR ( L , μ ) .
The additional terms in H G I 0 compared to H Γ SAR are
lim α ln Γ ( L α ) + ( L α ) ψ ( 0 ) ( L α ) ( 1 α ) ψ ( 0 ) ( α ) + ln ( 1 α ) + ln Γ ( α ) L .
For L = 1 , this becomes
lim α ln Γ ( 1 α ) + ln ( 1 α ) + ln Γ ( α ) A + ( 1 α ) ψ ( 0 ) ( 1 α ) ( 1 α ) ψ ( 0 ) ( α ) B 1 .
Simplifying A and B
A = ln ( 1 α ) Γ ( α ) Γ ( 1 α ) = ln ( 1 α ) Γ ( α ) α Γ ( α ) = ln 1 α α = ln 1 + 1 α .
B = ( 1 α ) ψ ( 0 ) ( 1 α ) ψ ( 0 ) ( α ) = ( 1 α ) ψ ( 0 ) ( 1 α 1 ) + 1 1 α 1 Because ψ ( 0 ) ( x + 1 ) = ψ ( 0 ) ( x ) + 1 x ψ ( 0 ) ( α ) = ( 1 α ) ψ ( 0 ) ( α ) 1 α ψ ( 0 ) ( α ) = 1 α + 1 .
Replacing A and B into (A2) and taking the limit, we obtain
lim α ln 1 + 1 α approaches 0 lim α 1 α approaches 0 + lim α 1 1 = 0 .
For the general case for L > 1 , we use Stirling’s approximation for large z:
Γ ( z ) 2 π z z e z ,
and
ψ ( 0 ) ( z ) ln ( z ) 1 2 z .
Therefore, the terms of Equation (A1) approximate to
ln Γ ( L α ) 1 2 ln 2 π ( L α ) ( L α ) ln ( L α ) + ( L α ) , ( L α ) ψ ( 0 ) ( L α ) ( L α ) ln ( L α ) 1 2 ( L α ) ( L α ) ln ( L α ) 1 2 , ( 1 α ) ψ ( 0 ) ( α ) ( 1 α ) ln ( α ) 1 α 2 α ,
and
ln Γ ( α ) 1 2 ln ( 2 π α ) α ln ( α ) + α .
Then, replacing these terms in (A1), we obtain
lim α [ 1 2 ln 2 π ( L α ) L ln ( L α ) + α ln ( L α ) + L α + L ln ( L α ) α ln ( L α ) 1 2 ln ( α ) + α ln ( α ) 1 α 2 α + ln ( 1 α ) + 1 2 ln ( 2 π α ) α ln ( α ) + α ] L .
We, then, simplify the expression to
lim α 1 2 ln 2 π ( L α ) + L 1 2 ln ( α ) 1 α 2 α + ln ( 1 + α ) + 1 2 ln ( 2 π α ) L .
We now group terms:
lim α 1 2 ln α L α + L 1 2 + α 1 2 α + ln 1 + α α L = 1 2 ln 1 + L 1 2 + 1 2 + ln 1 L = 0 .
Therefore
lim α H G I 0 ( μ , α , L ) = H Γ SAR ( L , μ ) ,
which concludes the proof.

Appendix B. Computational Information

We used a MacBook Pro, 14-inch 2021 computer with Apple M1 Max chip, 64 GB of memory, running macOS Sonoma 14.5 operating system, sourced in New Zealand. All the computational developments, including the text, were produced with RMarkdown v.2.27, RStudio v. 2024.04.1+748, and R v. 4.4.0.

References

  1. Moreira, A.; Prats-Iraola, P.; Younis, M.; Krieger, G.; Hajnsek, I.; Papathanassiou, K.P. A tutorial on synthetic aperture radar. IEEE Geosci. Remote Sens. Mag. 2013, 1, 6–43. [Google Scholar] [CrossRef]
  2. Mu, Y.; Huang, B.; Pan, Z.; Yang, H.; Hou, G.; Duan, J. An Enhanced High-Order Variational Model Based on Speckle Noise Removal With G0 Distribution. IEEE Access 2019, 7, 104365–104379. [Google Scholar] [CrossRef]
  3. Argenti, F.; Lapini, A.; Bianchi, T.; Alparone, L. A Tutorial on Speckle Reduction in Synthetic Aperture Radar Images. IEEE Geosci. Remote Sens. Mag. 2013, 1, 6–35. [Google Scholar] [CrossRef]
  4. Ferreira, J.d.A.; Nascimento, A.D.C. Shannon Entropy for the G I 0 Model: A New Segmentation Approach. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2020, 13, 2547–2553. [Google Scholar]
  5. Pressé, S.; Ghosh, K.; Lee, J.; Dill, K.A. Principles of maximum entropy and maximum caliber in statistical physics. Rev. Mod. Phys. 2013, 85, 1115–1141. [Google Scholar] [CrossRef]
  6. Mohammad-Djafari, A. Entropy, Information Theory, Information Geometry and Bayesian Inference in Data, Signal and Image Processing and Inverse Problems. Entropy 2015, 17, 3989–4027. [Google Scholar] [CrossRef]
  7. Avval, T.G.; Moeini, B.; Carver, V.; Fairley, N.; Smith, E.F.; Baltrusaitis, J.; Fernandez, V.; Tyler, B.J.; Gallagher, N.; Linford, M.R. The Often-Overlooked Power of Summary Statistics in Exploratory Data Analysis: Comparison of Pattern Recognition Entropy (PRE) to Other Summary Statistics and Introduction of Divided Spectrum-PRE (DS-PRE). J. Chem. Inf. Model. 2021, 61, 4173–4189. [Google Scholar] [CrossRef] [PubMed]
  8. Nascimento, A.D.C.; Horta, M.M.; Frery, A.C.; Cintra, R.J. Comparing Edge Detection Methods Based on Stochastic Entropies and Distances for PolSAR Imagery. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2014, 7, 648–663. [Google Scholar] [CrossRef]
  9. Nascimento, A.D.C.; Frery, A.C.; Cintra, R.J. Detecting Changes in Fully Polarimetric SAR Imagery With Statistical Information Theory. IEEE Trans. Geosci. Remote Sens. 2019, 57, 1380–1392. [Google Scholar] [CrossRef]
  10. Shannon, C.E. A Mathematical Theory of Communication. Bell Syst. Tech. J. 1948, 27, 379–423. [Google Scholar] [CrossRef]
  11. Wieczorkowski, R.; Grzegorzewski, P. Entropy estimators-improvements and comparisons. Comm. Statist. Simul. Comput. 1999, 28, 541–567. [Google Scholar] [CrossRef]
  12. Zamanzade, E.; Arghami, N.R. Testing normality based on new entropy estimators. J. Stat. Comput. Simul. 2012, 82, 1701–1713. [Google Scholar] [CrossRef]
  13. Alizadeh Noughabi, H. A new estimator of entropy and its application in testing normality. J. Stat. Comput. Simul. 2010, 80, 1151–1162. [Google Scholar] [CrossRef]
  14. Subhash, S.; Sunoj, S.M.; Sankaran, P.G.; Rajesh, G. Nonparametric estimation of quantile-based entropy function. Commun. Stat. Simul. Comput. 2021, 52, 1805–1821. [Google Scholar] [CrossRef]
  15. Pearson, K.; Henrici, O.M.F.E. VII. Mathematical contributions to the theory of evolution.—III. Regression, heredity, and panmixia. Philos. Trans. R. Soc. Lond. Ser. A Contain. Pap. Math. Phys. Character 1896, 187, 253–318. [Google Scholar]
  16. Walter, A. Hendricks, K.W.R. The Sampling Distribution of the Coefficient of Variation. Ann. Math. Stat. 1936, 7, 129–132. [Google Scholar]
  17. Tian, L. Inferences on the common coefficient of variation. Stat. Med. 2005, 24, 2213–2220. [Google Scholar] [CrossRef] [PubMed]
  18. Subrahmanya Nairy, K.; Aruna Rao, K. Tests of Coefficients of Variation of Normal Population. Commun. Stat.-Simul. Comput. 2003, 32, 641–661. [Google Scholar] [CrossRef]
  19. Chankham, W.; Niwitpong, S.A.; Niwitpong, S. The Simultaneous Confidence Interval for the Ratios of the Coefficients of Variation of Multiple Inverse Gaussian Distributions and Its Application to PM2.5 Data. Symmetry 2024, 16, 331. [Google Scholar] [CrossRef]
  20. Banik, S.; Kibria, B.M.G. Estimating the Population Coefficient of Variation by Confidence Intervals. Commun. Stat.-Simul. Comput. 2011, 40, 1236–1261. [Google Scholar] [CrossRef]
  21. Ulaby, F.; Kouyate, F.; Brisco, B.; Williams, T.H. Textural Information in SAR Images. IEEE Trans. Geosci. Remote Sens. 1986, GE-24, 235–245. [Google Scholar] [CrossRef]
  22. Touzi, R.; Lopes, A.; Bousquet, P. A statistical and geometrical edge detector for SAR images. IEEE Trans. Geosci. Remote Sens. 1988, 26, 764–773. [Google Scholar] [CrossRef]
  23. Frery, A.C.; Muller, H.J.; Yanasse, C.; Sant’Anna, S. A model for extremely heterogeneous clutter. IEEE Trans. Geosci. Remote Sens. 1997, 35, 648–659. [Google Scholar] [CrossRef]
  24. Cassetti, J.; Delgadino, D.; Rey, A.; Frery, A.C. Entropy Estimators in SAR Image Classification. Entropy 2022, 24, 509. [Google Scholar] [CrossRef] [PubMed]
  25. Nascimento, A.D.C.; Cintra, R.J.; Frery, A.C. Hypothesis Testing in Speckled Data With Stochastic Distances. IEEE Trans. Geosci. Remote Sens. 2010, 48, 373–385. [Google Scholar] [CrossRef]
  26. Vasicek, O. A test for normality based on sample entropy. J. R. Stat. Soc. Ser. B Methodol. 1976, 38, 54–59. [Google Scholar] [CrossRef]
  27. Correa, J.C. A new estimator of entropy. Commun. Stat.–Theory Methods 1995, 24, 2439–2449. [Google Scholar] [CrossRef]
  28. Al-Omari, A.I.; Haq, A. Novel entropy estimators of a continuous random variable. Int. J. Model. Simul. Sci. Comput. 2019, 10, 1950004. [Google Scholar] [CrossRef]
  29. Ebrahimi, N.; Pflughoeft, K.; Soofi, E.S. Two measures of sample entropy. Stat. Probab. 1994, 20, 225–234. [Google Scholar] [CrossRef]
  30. Al-Omari, A.I. Estimation of entropy using random sampling. J. Comput. Appl. Math. 2014, 261, 95–102. [Google Scholar] [CrossRef]
  31. Michelucci, U.; Venturini, F. Estimating Neural Network’s Performance with Bootstrap: A Tutorial. Mach. Learn. Knowl. Extr. 2021, 3, 357–373. [Google Scholar] [CrossRef]
  32. Ospina, R.; Marmolejo-Ramos, F. Performance of Some Estimators of Relative Variability. Front. Appl. Math. Stat. 2019, 5, 43. [Google Scholar] [CrossRef]
  33. Chaubey, Y.P.; Singh, M.; Sen, D. Symmetrizing and Variance Stabilizing Transformations of Sample Coefficient of Variation from Inverse Gaussian Distribution. Sankhya B 2017, 79, 217–246. [Google Scholar] [CrossRef]
  34. Takagi, K.; Kumagai, S.; Matsunaga, I.; Kusaka, Y. Application of inverse Gaussian distribution to occupational exposure data. Ann. Occup. Hyg. 1997, 41, 505–514. [Google Scholar] [CrossRef]
  35. Whitmore, G.; Yalovsky, M. A normalizing logarithmic transformation for inverse Gaussian random variables. Technometrics 1978, 20, 207–208. [Google Scholar] [CrossRef]
  36. Burnham, K.P.; Anderson, D.R. Multimodel Inference: Understanding AIC and BIC in Model Selection. Sociol. Methods Res. 2004, 33, 261–304. [Google Scholar] [CrossRef]
  37. Dziak, J.J.; Coffman, D.L.; Lanza, S.T.; Li, R.; Jermiin, L.S. Sensitivity and specificity of information criteria. Briefings Bioinform. 2019, 21, 553–565. [Google Scholar] [CrossRef]
  38. Gomez, L.; Alvarez, L.; Mazorra, L.; Frery, A.C. Fully PolSAR image classification using machine learning techniques and reaction-diffusion systems. Neurocomputing 2017, 255, 52–60. [Google Scholar] [CrossRef]
  39. Frery, A.C.; Gambini, J. Comparing samples from the 𝒢0 distribution using a geodesic distance. TEST 2019, 29, 359–378. [Google Scholar] [CrossRef]
  40. Neto, J.F.S.R.; Rodrigues, F.A.A. Improving Log-Cumulant-Based Estimation of Heterogeneity Information in SAR Imagery. IEEE Geosci. Remote Sens. Lett. 2023, 20, 1–5. [Google Scholar] [CrossRef]
Figure 1. H G I 0 converges to the H Γ SAR when α , with L = 8 .
Figure 1. H G I 0 converges to the H Γ SAR when α , with L = 8 .
Remotesensing 16 01973 g001
Figure 2. Bias and MSE of the entropy estimators for the Γ SAR with L = 5 .
Figure 2. Bias and MSE of the entropy estimators for the Γ SAR with L = 5 .
Remotesensing 16 01973 g002
Figure 3. Empirical densities obtained from S H ˜ AO ( Z ; L ) test under the null hypothesis.
Figure 3. Empirical densities obtained from S H ˜ AO ( Z ; L ) test under the null hypothesis.
Remotesensing 16 01973 g003
Figure 4. Normal Q–Q plots for n = 121 . The dashed green lines in the Q–Q plots represent theoretical expected quantiles for a normal distribution. Variations in the scales of the axes reflect differences in the range of values across datasets.
Figure 4. Normal Q–Q plots for n = 121 . The dashed green lines in the Q–Q plots represent theoretical expected quantiles for a normal distribution. Variations in the scales of the axes reflect differences in the range of values across datasets.
Remotesensing 16 01973 g004
Figure 5. Goodness-of-fit plots for evaluating the best distribution with CV data from Γ SAR (under the null hypothesis) with n = 49 , L = 5 , and μ = 1 .
Figure 5. Goodness-of-fit plots for evaluating the best distribution with CV data from Γ SAR (under the null hypothesis) with n = 49 , L = 5 , and μ = 1 .
Remotesensing 16 01973 g005
Figure 6. Goodness-of-fit plots for evaluating the best distribution with CV data from G I 0 (under the alternative hypothesis) with n = 49 , L = 5 , μ = 1 , and α = 3 .
Figure 6. Goodness-of-fit plots for evaluating the best distribution with CV data from G I 0 (under the alternative hypothesis) with n = 49 , L = 5 , μ = 1 , and α = 3 .
Remotesensing 16 01973 g006
Figure 7. Goodness-of-fit plots for evaluating the best distribution with CV MnAD data from Γ SAR (under the null hypothesis) with n = 49 , L = 5 , and μ = 1 .
Figure 7. Goodness-of-fit plots for evaluating the best distribution with CV MnAD data from Γ SAR (under the null hypothesis) with n = 49 , L = 5 , and μ = 1 .
Remotesensing 16 01973 g007
Figure 8. Goodness-of-fit plots for evaluating the best distribution with C V MnAD data from G I 0 (under the alternative hypothesis) with n = 49 , L = 5 , μ = 1 , and α = 3 .
Figure 8. Goodness-of-fit plots for evaluating the best distribution with C V MnAD data from G I 0 (under the alternative hypothesis) with n = 49 , L = 5 , μ = 1 , and α = 3 .
Remotesensing 16 01973 g008
Figure 9. Synthetic dataset: (a) Phantom. (b) Simulated image, varying α and μ , with L = 5 .
Figure 9. Synthetic dataset: (a) Phantom. (b) Simulated image, varying α and μ , with L = 5 .
Remotesensing 16 01973 g009
Figure 10. Results of applying the test statistics: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Figure 10. Results of applying the test statistics: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Remotesensing 16 01973 g010
Figure 11. Map of p-values: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Figure 11. Map of p-values: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Remotesensing 16 01973 g011
Figure 12. Results for a threshold of 0.05 of the p-value: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Figure 12. Results for a threshold of 0.05 of the p-value: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Remotesensing 16 01973 g012
Figure 13. SAR images: (a) Rotterdam, (b) Coast of Jalisco, (c) Illinois-Region 1, (d) Illinois-Region 2.
Figure 13. SAR images: (a) Rotterdam, (b) Coast of Jalisco, (c) Illinois-Region 1, (d) Illinois-Region 2.
Remotesensing 16 01973 g013
Figure 14. Optical images: (a) Rotterdam, (b) Coast Jalisco, (c) Illinois-Region 1, (d) Illinois-Region 2.
Figure 14. Optical images: (a) Rotterdam, (b) Coast Jalisco, (c) Illinois-Region 1, (d) Illinois-Region 2.
Remotesensing 16 01973 g014
Figure 15. Results of applying the test statistics, Rotterdam: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Figure 15. Results of applying the test statistics, Rotterdam: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Remotesensing 16 01973 g015
Figure 16. Results of applying the test statistics, Coast of Jalisco: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Figure 16. Results of applying the test statistics, Coast of Jalisco: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Remotesensing 16 01973 g016
Figure 17. Results of applying the test statistics, Illinois-Region 1: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Figure 17. Results of applying the test statistics, Illinois-Region 1: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Remotesensing 16 01973 g017
Figure 18. Results of applying the test statistics, Illinois-Region 2: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Figure 18. Results of applying the test statistics, Illinois-Region 2: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Remotesensing 16 01973 g018
Figure 19. Map of p-values, Rotterdam: (a) S H ˜ AO ( Z ; L ) . (b) T CV . (c) T CV MnAD .
Figure 19. Map of p-values, Rotterdam: (a) S H ˜ AO ( Z ; L ) . (b) T CV . (c) T CV MnAD .
Remotesensing 16 01973 g019
Figure 20. Map of p-values, Coast of Jalisco: (a) S H ˜ AO ( Z ; L ) . (b) T CV . (c) T CV MnAD .
Figure 20. Map of p-values, Coast of Jalisco: (a) S H ˜ AO ( Z ; L ) . (b) T CV . (c) T CV MnAD .
Remotesensing 16 01973 g020
Figure 21. Map of p-values, Illinois-Region 1: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Figure 21. Map of p-values, Illinois-Region 1: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Remotesensing 16 01973 g021
Figure 22. Map of p-values, Illinois-Region 2: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Figure 22. Map of p-values, Illinois-Region 2: (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Remotesensing 16 01973 g022
Figure 23. Results for a threshold of 0.05 of the p-value, Rotterdam. (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Figure 23. Results for a threshold of 0.05 of the p-value, Rotterdam. (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Remotesensing 16 01973 g023
Figure 24. Results for a threshold of 0.05 of the p-value, Coast of Jalisco. (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Figure 24. Results for a threshold of 0.05 of the p-value, Coast of Jalisco. (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Remotesensing 16 01973 g024
Figure 25. Results for a threshold of 0.05 of the p-value, Illinois-Region 1. (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Figure 25. Results for a threshold of 0.05 of the p-value, Illinois-Region 1. (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Remotesensing 16 01973 g025
Figure 26. Results for a threshold of 0.05 of the p-value, Illinois-Region 2. (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Figure 26. Results for a threshold of 0.05 of the p-value, Illinois-Region 2. (a) S H ˜ AO ( Z ; L ) , (b) T CV , and (c) T CV MnAD .
Remotesensing 16 01973 g026
Figure 27. Image of Illinois-Region 2 with reference samples.
Figure 27. Image of Illinois-Region 2 with reference samples.
Remotesensing 16 01973 g027
Table 1. Bias and MSE of the entropy estimators for the Γ SAR with L = 5 .
Table 1. Bias and MSE of the entropy estimators for the Γ SAR with L = 5 .
BiasMSE
μ n H ^ C H ^ E H ^ AO H ˜ C H ˜ E H ˜ AO H ^ C H ^ E H ^ AO H ˜ C H ˜ E H ˜ AO
19 0.210 0.280 0.377 0.106 0.054 0.156 0.140 0.173 0.236 0.121 0.116 0.133
25 0.071 0.112 0.149 0.033 0.004 0.035 0.032 0.040 0.050 0.032 0.033 0.033
49 0.036 0.060 0.081 0.020 0.015 0.006 0.016 0.019 0.021 0.016 0.017 0.017
81 0.009 0.026 0.039 0.002 0.031 0.018 0.008 0.009 0.010 0.009 0.010 0.009
121 0.001 0.013 0.022 0.004 0.030 0.023 0.005 0.006 0.006 0.006 0.007 0.007
109 0.227 0.297 0.394 0.119 0.060 0.156 0.146 0.180 0.247 0.126 0.117 0.136
25 0.088 0.129 0.166 0.052 0.013 0.050 0.040 0.048 0.059 0.040 0.035 0.039
49 0.017 0.043 0.064 0.001 0.035 0.012 0.016 0.017 0.019 0.018 0.019 0.017
81 0.008 0.025 0.038 0.001 0.031 0.019 0.009 0.009 0.010 0.009 0.011 0.010
121 0.004 0.007 0.017 0.008 0.037 0.027 0.006 0.006 0.006 0.006 0.008 0.007
Table 2. Tests accuracy and processing time for each bootstrap-improved estimator.
Table 2. Tests accuracy and processing time for each bootstrap-improved estimator.
EstimatorLn S ( Z ; L ) Time (s)
H ˜ C 225 0.00152 22.53
49 0.00515 40.35
81 0.00625 63.93
121 0.00751 97.06
825 0.04332 22.25
49 0.01659 33.42
81 0.00393 50.94
121 0.00261 97.35
H ˜ E 225 0.02204 4.66
49 0.03452 5.55
81 0.03195 6.89
121 0.03012 7.90
825 0.00801 4.81
49 0.01654 5.43
81 0.03036 6.38
121 0.03137 7.46
H ˜ AO 225 0.01935 4.61
49 0.00786 5.19
81 0.01995 6.70
121 0.01741 7.41
825 0.04020 4.74
49 0.00047 5.35
81 0.01176 6.21
121 0.02019 7.48
Table 3. Descriptive analysis of S H ˜ AO ( Z ; L ) , with L 3 , 5 , 8 , 11 and μ = 1 .
Table 3. Descriptive analysis of S H ˜ AO ( Z ; L ) , with L 3 , 5 , 8 , 11 and μ = 1 .
LnMeanSDVarSKEKp-Value
325 0.0280 0.1547 0.0239 0.0076 0.4734 0.0022
49 0.0003 0.1053 0.0111 0.0562 0.2610 0.0582
81 0.0124 0.0796 0.0063 0.0124 0.0536 0.5278
121 0.0187 0.0630 0.0040 0.0337 0.1826 0.5894
200 0.0215 0.0490 0.0024 0.0625 0.0473 0.0860
525 0.0379 0.1669 0.0278 0.0007 0.1420 0.3267
49 0.0015 0.1150 0.0132 0.1025 0.1998 0.2582
81 0.0145 0.0869 0.0075 0.1008 0.5297 0.1121
121 0.0198 0.0687 0.0047 0.0127 0.0222 0.2919
200 0.0236 0.0529 0.0028 0.0467 0.0977 0.3346
825 0.0464 0.1680 0.0282 0.0121 0.0980 0.6477
49 0.0031 0.1202 0.0144 0.1282 0.2000 0.0038
81 0.0137 0.0883 0.0078 0.0279 0.2554 0.5567
121 0.0200 0.0738 0.0055 0.0089 0.0686 0.7502
200 0.0260 0.0546 0.0030 0.0716 0.0349 0.3771
1125 0.0442 0.1735 0.0301 0.1422 0.2413 0.0981
49 0.0019 0.1201 0.0144 0.0503 0.1464 0.9576
81 0.0127 0.0917 0.0084 0.0172 0.0333 0.3179
121 0.0239 0.0729 0.0053 0.0127 0.2102 0.0596
200 0.0234 0.0572 0.0033 0.0233 0.1072 0.6740
Table 4. Size and power of the S H ˜ AO ( Z ) test statistic.
Table 4. Size and power of the S H ˜ AO ( Z ) test statistic.
SizePower
L n 1 % 5 % 10 % 1 % 5 % 10 %
325 0.0160 0.0620 0.1070 0.6900 0.8450 0.8340
49 0.0100 0.0480 0.0960 0.6890 0.8920 0.8480
81 0.0120 0.0490 0.1080 0.6260 0.8750 0.8540
121 0.0090 0.0690 0.1190 0.5680 0.8620 0.8230
525 0.0210 0.0660 0.1130 0.9120 0.9620 0.9880
49 0.0100 0.0460 0.1080 0.9470 0.9820 0.9960
81 0.0120 0.0560 0.1070 0.9580 0.9900 0.9960
121 0.0150 0.0640 0.1150 0.9420 0.9780 0.9950
825 0.0210 0.0650 0.1080 0.9930 0.9950 0.9970
49 0.0060 0.0470 0.0860 0.9980 1.0000 0.9970
81 0.0120 0.0490 0.1000 0.9930 0.9980 0.9990
121 0.0150 0.0650 0.1220 0.9970 0.9990 0.9980
1125 0.0130 0.0610 0.1000 0.9990 0.9990 0.9990
49 0.0100 0.0450 0.0920 0.9980 0.9990 0.9990
81 0.0170 0.0530 0.1050 1.0000 1.0000 1.0000
121 0.0160 0.0680 0.1180 0.9980 1.0000 0.9980
Table 5. AIC and BIC values for evaluating the best distribution with CV data from Γ SAR .
Table 5. AIC and BIC values for evaluating the best distribution with CV data from Γ SAR .
CriterionnNormalLognormalGammaWeibullInverse Gaussian
AIC25 38 , 031.9 38 , 266.7 38 , 311.8 36 , 413.2 38 , 261.6
49 47 , 698.2 47 , 913.7 47 , 905.6 45 , 554.0 47 , 911.6
81 55 , 382.1 55 , 494.7 55 , 494.9 53 , 220.4 55 , 493.8
121 61 , 344.9 61 , 470.8 61 , 453.8 58 , 876.0 61 , 470.5
BIC25 38 , 016.7 38 , 251.5 38 , 296.6 36 , 398.0 38 , 246.4
49 47 , 683.0 47 , 898.5 47 , 890.4 45 , 538.7 47 , 896.4
81 55 , 366.9 55 , 479.5 55 , 479.6 53 , 205.2 55 , 478.6
121 61 , 329.7 61 , 455.6 61 , 438.6 58 , 860.8 61 , 455.2
Table 6. AIC and BIC values for evaluating the best distribution with CV data from G I 0 .
Table 6. AIC and BIC values for evaluating the best distribution with CV data from G I 0 .
CriterionnNormalLognormalGammaWeibullInverse Gaussian
AIC25 8254.04 2186.31 3628.40 8383.58 2257.63
49 8821.79 1689.03 3483.10 9533.53 1835.29
81 8525.81 866.29 2853.31 9822.48 1057.91
121 8708.81 131.86 2341.06 10 , 506.49 398.53
BIC25 8269.27 2201.54 3643.63 8398.81 2272.86
49 8837.02 1704.26 3498.33 9548.76 1850.52
81 8541.04 881.52 2868.55 9837.72 1073.14
121 8724.04 147.09 2356.29 10 , 521.72 413.76
Table 7. AIC and BIC values for evaluating the best distribution with CV MnAD data from Γ SAR .
Table 7. AIC and BIC values for evaluating the best distribution with CV MnAD data from Γ SAR .
CriterionnNormalLognormalGammaWeibullInverse Gaussian
AIC25 38 , 375.56 39 , 147.85 39 , 066.29 36 , 652.49 39 , 143.61
49 48 , 386.11 48 , 795.32 48 , 745.48 46 , 240.91 48 , 793.83
81 56 , 072.87 56 , 322.32 56 , 290.02 53 , 836.04 56 , 322.12
121 62 , 217.14 62 , 394.77 62 , 369.32 59 , 861.80 62 , 394.57
BIC25 38 , 360.32 39 , 132.62 39 , 051.05 36 , 637.26 39 , 128.38
49 48 , 370.87 48 , 780.09 48 , 730.25 46 , 225.67 48 , 778.60
81 56 , 057.64 56 , 307.09 56 , 274.79 53 , 820.81 56 , 306.89
121 62 , 201.91 62 , 379.54 62 , 354.09 59 , 846.56 62 , 379.33
Table 8. AIC and BIC values for evaluating the best distribution with CV MnAD data from G I 0 .
Table 8. AIC and BIC values for evaluating the best distribution with CV MnAD data from G I 0 .
CriterionnNormalLognormalGammaWeibullInverse Gaussian
AIC25 13 , 302.39 15 , 575.23 15 , 158.42 11 , 933.35 15 , 565.03
49 23 , 265.27 24 , 529.72 24 , 284.47 20 , 986.94 24 , 522.28
81 29 , 908.19 30 , 960.38 30 , 747.90 25 , 233.41 30 , 946.54
121 36 , 496.78 37 , 128.07 36 , 991.41 32 , 366.52 37 , 123.20
BIC25 13 , 287.16 15 , 559.99 15 , 143.19 11 , 918.12 15 , 549.80
49 23 , 250.04 24 , 514.48 24 , 269.24 20 , 971.71 24 , 507.05
81 29 , 892.96 30 , 945.15 30 , 732.67 25 , 218.17 30 , 931.31
121 36 , 481.55 37 , 112.84 36 , 976.18 32 , 351.28 37 , 107.97
Table 9. Parameters of selected SAR images.
Table 9. Parameters of selected SAR images.
SiteMissionModeBandPolarizationSizeLResolution Rg/Az [m]Acquisition Date
RotterdamTerraSAR-XHSXHH 512 × 512 1 0.85 / 0.85 06-12-2018
Coast of JaliscoSentinel-1BGRD EWCVV 512 × 512 18 40 / 40 29-08-2021
Illinois-Region 1Sentinel-1BGRD SMCVV 512 × 512 36 10 / 10 30-06-2020
Illinois-Region 2Sentinel-1BGRD SMCVV 1024 × 1024 36 10 / 10 28-08-2021
Table 10. Analysis of test statistics for different areas of reference.
Table 10. Analysis of test statistics for different areas of reference.
Area of ReferenceTest StatisticMeanSD
Water S H ˜ AO ( Z ; L ) 0.050290.15415
T CV 0.158270.02584
T CV MnAD 0.130130.02183
Vegetation S H ˜ AO ( Z ; L ) 0.364820.28489
T CV 0.294210.10957
T CV MnAD 0.245010.09829
Urban S H ˜ AO ( Z ; L ) 0.829700.21910
T CV 0.568460.33938
T CV MnAD 0.471760.29486
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

Frery, A.C.; Alpala, J.; Nascimento, A.D.C. Identifying Heterogeneity in SAR Data with New Test Statistics. Remote Sens. 2024, 16, 1973. https://doi.org/10.3390/rs16111973

AMA Style

Frery AC, Alpala J, Nascimento ADC. Identifying Heterogeneity in SAR Data with New Test Statistics. Remote Sensing. 2024; 16(11):1973. https://doi.org/10.3390/rs16111973

Chicago/Turabian Style

Frery, Alejandro C., Janeth Alpala, and Abraão D. C. Nascimento. 2024. "Identifying Heterogeneity in SAR Data with New Test Statistics" Remote Sensing 16, no. 11: 1973. https://doi.org/10.3390/rs16111973

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