Next Article in Journal
Protection of NOx Sensors from Sulfur Poisoning in Glass Furnaces by the Optimization of a “SO2 Trap”
Previous Article in Journal
Sensitive and Compact Evanescent-Waveguide Optical Detector for Sugar Sensing in Commercial Beverages
Previous Article in Special Issue
The Effect of Fractionation during the Vacuum Deposition of Stabilized Amorphous Selenium Alloy Photoconductors on the Overall Charge Collection Efficiency
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Development of Adaptive Point-Spread Function Estimation Method in Various Scintillation Detector Thickness for X-ray Imaging

1
Precision Medical Device Research Center, Korea Electrotechnology Research Institute (KERI), 111, Hanggaul-ro, Sangnok-gu, Ansan-si 15588, Gyeonggi-do, Republic of Korea
2
Department of Radiological Science, College of Health Science, Gachon University, 191, Hambangmoe-ro, Yeonsu-gu, Incheon 21936, Republic of Korea
3
Department of Biomedical Engineering, Eulji University, 553, Sanseong-daero, Sujeong-gu, Seongnam-si 13135, Gyeonggi-do, Republic of Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Sensors 2023, 23(19), 8185; https://doi.org/10.3390/s23198185
Submission received: 29 August 2023 / Revised: 21 September 2023 / Accepted: 27 September 2023 / Published: 30 September 2023
(This article belongs to the Special Issue Advanced Materials and Technologies for Radiation Detectors)

Abstract

:
An indirect conversion X-ray detector uses a scintillator that utilizes the proportionality of the intensity of incident radiation to the amount of visible light emitted. A thicker scintillator reduces the patient’s dose while decreasing the sharpness. A thin scintillator has an advantage in terms of sharpness; however, its noise component increases. Thus, the proposed method converts the spatial resolution of radiographic images acquired from a normal-thickness scintillation detector into a thin-thickness scintillation detector. Note that noise amplification and artifacts were minimized as much as possible after non-blind deconvolution. To accomplish this, the proposed algorithm estimates the optimal point-spread function (PSF) when the structural similarity index (SSIM) and feature similarity index (FSIM) are the most similar between thick and thin scintillator images. Simulation and experimental results demonstrate the viability of the proposed method. Moreover, the deconvolution images obtained using the proposed scheme show an effective image restoration method in terms of the human visible system compared to that of the traditional PSF measurement technique. Consequently, the proposed method is useful for restoring degraded images using the adaptive PSF while preventing noise amplification and artifacts and is effective in improving the image quality in the present X-ray imaging system.

1. Introduction

A flat-panel detector (FPD) is an important device in digital radiographic imaging systems. The FPD is located behind the object to be imaged in the direction of radiation from the X-ray and gamma-ray sources. As radiation penetrates an object, it converts the attenuated information into electrical signals and then converts them into digital signals for a subsequent real-time display [1]. The inner structures of the objects in the radiation imaging scanner are represented by the relative intensity of the signals captured by the pixel array owing to the difference in the X-ray attenuation coefficients of the inner structures. These electrical signals are converted into digital signal outputs using a readout integrated circuit (ROIC) and provided to a backend computer system [2].
Digital radiography (DR) using FPD has the advantages of fast image-acquisition speed and wide dynamic range compared to conventional film and computed tomography (CR) techniques [3]. The FPD has fast image acquisition, allowing real-time diagnostics. When the radiation reaches the FPD, it is immediately converted into an electrical signal and digital image. The FPD consists of a pixel array of sensing elements called pixels, and an electronic circuit that includes the ROIC and a gate driver. This reduces the image distortion caused by blurring owing to long exposure times and ensures an accurate diagnosis [4]. Moreover, FPD has a wide dynamic range under various illumination conditions. Combining a phosphor or scintillator and a backplane (pixel array) in an FPD results in high light sensitivity, which allows for the sensitive detection of radiation [5].
Depending on the method used to convert radiation into electrical signals, methods can be broadly classified as direct or indirect [6,7]. A direct conversion detector converts the energy deposited into electric charges using photoconductors or semiconductor materials, including amorphous selenium (a-Se), thallium bromide, and gadolinium compounds, using materials with low atomic numbers [8]. In contrast, an indirect conversion detector improves the photon-to-electric signal conversion ratio by inserting a scintillator (e.g., thallium-doped cesium iodide [CsI:Tl], gadolinium oxysulfide [Gadox], barium fluoride, or cadmium tungstate [CdWO4]) in the process of converting photons to electrical signals in the direct conversion detector [9,10]. The spatial resolution of the indirect conversion detector is restricted by the blurring of the scattered light within the scintillators compared to that of the direct conversion detector [11]. Swank [12] investigated the noise factor when the resultant signal was formed by integrating scintillation pulses. This depends on the shape of the pulse-height distribution in the X-ray phosphors. Therefore, direct conversion detectors have a relative advantage for high-resolution imaging, such as mammography [13,14], and it is helpful to improve the image quality by optimizing the scintillator thickness for optimal detection quantum efficiency (DQE) with noise reduction, although the resolution and sharpness are reduced [15,16].
In general, the finite focal spot size, detector pixel size, and magnification affect the acquisition of information by distorting the original signal [17,18], which can be expressed by Equation (1), assuming a linear shift-invariant system [19].
g m , n = P S F m , n   f m , n + N ( m , n ) ,
where m and n describe the orthogonal coordinate indices, g is the degraded image, and f is the clean image. f is recovered from g by deconvolving the point-spread function (PSF), which is referred to as image deconvolution. Here, the PSF is considered to be the result of the influence of the finite focal spot size and detector pixel size. The magnitude within the object is assumed to be the same when the thickness of the object is much smaller than the SDD. The represents the 2D convolution operator, and N is the noise component. Notably, the PSF is most important to restore the blurred image, and there are three main techniques in which PSF can be measured in radiographic images: the pinhole, slit, and edge phantom [20,21,22] methods. Here, the area in which radiation passes is required to be ϕ   ~   10   μ m in pinhole and slit phantom. The radiographic images acquired by these devices contain a blurred component owing to finite system conditions. A profile that includes the pinhole and slit areas can be represented as a PSF, as shown in Equation (2):
P S F m , n = 1 σ 2 π exp m 2 + n 2 2 σ 2 ,
where σ denotes the standard deviation of the Gaussian distribution function. In particular, the slit method was tilted by approximately 1.5°–3° to improve the measurement accuracy through oversampling [23,24]. The edge phantom was constructed from a 1 mm Tungsten (W) foil plate fixed on a 3 mm thick lead plate in accordance with the IEC 62220-1-1:2015 protocol [25,26]. The line spread function (LSF) of the edge phantom image was calculated by differentiating the profile between the penetrated radiation regions and vice versa.
Once the PSF is obtained through the above process, the restored image can be obtained by deconvolution of the degraded image, g . An approach that uses a pre-measured PSF is called a non-blind method. The advantage of this method is that it is more likely to succeed in deconvolution because it uses the exact PSF of an acquisition system [27,28]. Traditionally, Richardson–Lucy and Wiener filters are classic deconvolution methods [29,30], and fast deconvolution is performed to obtain restored images with improved sharpness. However, these images suffer from ringing artifacts [27]. Image prior-based deconvolution methods, including non-local-based approaches [31,32], fields of experts [33,34], and patch-based priors [35,36], can effectively prevent artifacts. However, these methods rely on empirical information and suffer from highly non-convex optimization problems and high computational costs [37]. Recently, machine learning-based deconvolution methods [38,39] have shown outstanding performance in image restoration; moreover, when combined with existing non-blind methods, they overcome the limitations of real-world applications to limited training data [37,40].
Despite the existing effective non-blind deconvolution methods, the noise in Equation (1) is amplified when high-pass filtering is performed by deconvolution, and the problem of image quality degradation remains [41]. To overcome these problems, a deconvolution approach based on regularization terms can be used to perform high-pass filtering with PSF while minimizing noise amplification [42,43,44]. In particular, penalty terms based on total variation (TV) and total generalized variation (TGV) are very effective at improving the sharpness of noisy and blurry images [45,46]. However, TV- and TGV-based regulation-term methods often tend to produce staircase artifacts [47]. Another approach is the deconvolution method after the noise reduction process, where deconvolution is performed considering the total PSF, composed of blurring in the imaging system and smoothing of the noise reduction algorithm [48]. This method can minimize noise amplification during deconvolution by removing the noise. However, restoring the lost detailed information using noise reduction algorithms is difficult.
Therefore, this study aims to convert the resolution of radiographic images acquired with a normal-thickness scintillator to that of a thin scintillator. The proposed method extracts the PSF with the most similar image quality between a high-resolution image convolving various PSF and a low-resolution image. Thus, a high-resolution image with minimal noise amplification was obtained by deconvolution using the PSF obtained from a low-exposure image with a normal-thickness scintillator. Simulations and experiments are conducted to verify the viability of the proposed framework. The following sections briefly describe the proposed scheme and discuss the results.

2. Materials and Methods

2.1. Imaging System Performance

Figure 1 shows the exposure conditions for measuring the modulation transfer function (MTF) of the detector for each scintillator thickness. All the detectors were configured using a Gd2O2S:Tb scintillator and a CMOS image sensor (Rad-icon, Teledyne DALSA, California, USA). Here, Figure 2 shows the scanning electron microscope images (SEM) of the three detectors, and the detectors 1, 2, and 3 have scintillator thicknesses of 84 μm, 96 μm, and 140 μm, respectively. The scintillation thicknesses, pixel sizes, matrix sizes, and analog-to-digital conversion (ADC) resolutions of the detectors used in the experiments are summarized in Table 1. The experimental exposure conditions were documented following the guidelines of the IEC 62220-1-1:2015 protocol. According to the Radiation Quality (RQA)-5, the SDD was set to 150 cm, tube voltage to 70 kV, and additional filter of 21.0 mm Al. An X-ray tube (L10321, Hamamatsu Photonics K.K., Shizuoka-ken, Japan; focal spot size: 5 μm) was used to investigate the imaging system performance and conduct an experimental study. Figure 3a shows the MTF plots of all the detectors, where the x-axis is the spatial frequency. Radiographic images were acquired according to IEC 62220-1-1:2015 guidelines, while tilting a slit camera (07-24-1, Nuclear Associate Corp., Washington, USA) with a slit width of 10 μm to avoid aliasing [49]. MTF plots were calculated using Equation (3):
M T F f = F { L S F } ,
where f is a 1D coordinate in the frequency domain, and F is an operator of the Fourier transformation. Figure 3b shows the LSF predicted by the inverse calculation of the measured MTF using Equations (2) and (3). The σ of the LSFs in Figure 3b were approximately 1.79 (pixels), 2.61 (pixels), and 5.13 (pixels) for detectors 1, 2, and 3, respectively.
Normalized noise power spectrum (NNPS) [50] representation is standard and useful for understanding the noise characteristics of an imaging system. This demonstrates the variation influenced by spatial frequency, indicating a limitation in utilizing the simulation for the design of degraded images based on Equation (1). Consequently, the analysis of the noise component was carried out using the Poisson–Gaussian mixture model as Equation (4), which has been recognized as the most suitable approach within the radiation degradation model due to its emphasis on noise characteristics [51,52].
g x = P S F x f x + η ( x ) δ ,   η 2 x = α x + β 2 ,
where η ( x ) represents the standard deviation of the noise distribution at the pixel position x , which can be decomposed into Poisson noise, α , and the variance of Gaussian noise, β 2 . Additionally, δ denotes zero-mean independent random noise with a standard deviation of one. Earlier research grounded in this model has substantiated its effectiveness [53,54]. Notably, Sutour et al. [55] anticipated noise parameters through non-parametric detection within uniform regions based on Kendall’s τ -coefficient [56,57]. This approach employs a robust polynomial noise level function (NLF) estimation method that relies on the correlation between the mean and variance observed in uniform regions.
As illustrated in Figure 4, the NLF of the three detectors is depicted using the average of the white images, denoted as S ¯ . The Poisson parameters, α , for detectors 1, 2, and 3 were approximately 0.37, 0.29, and 0.08, respectively. Moreover, their respective Gaussian parameters, β , were approximately 9.12, 7.10, and 3.31. These detector-specific PSF and noise parameter outcomes, derived from MTF and NLF measurements, were then employed in the simulation process to generate an image that authentically represents the underlying physical phenomena.

2.2. Data Acquisition

2.2.1. Simulations

Figure 5a shows a 3D numerical dice phantom used in the simulation. The geometrical acquisition conditions were the same as those of Section 2.1. The dice phantom has 300 × 300 × 300 voxels and the size of voxels is 0.1 × 0.1 × 0.1 mm3. In the simulation, a degradation image is created according to Equations (1) and (4) based on the previously calculated σ of PSF and α and β of noise parameters. To generate a noisy image, we used the imnoise (∙) and poissrnd (∙) functions in the MATLABTM (version 8.3) toolbox. The dice phantom composed of the methyl methacrylate (PMMA) and attenuation images were obtained based on the ray-tracing method. The SDD was set to 150 cm, the source-to-object (SOD) was set to 100 cm, and the tube voltage was set to 70 kV at mono-energy spectrum. Note that, the effect of scatter radiation is not considered for the proof of principle for Equation (4).

2.2.2. Experiments

Figure 5b also shows a high-resolution line chart phantom (Type 38, CN 69761, Active Radsys, Italy) and Figure 5c an electronic device used in the experiment. Here, we used high-resolution patterns in the line chart, which had 20 groups from 0.6 to 5.0 lp/mm. This pattern is enclosed in plastic, which can be assumed to be a tissue-equivalent material. The line chart phantom and electronic device have thicknesses of 0.1 mm and 2 mm or less. The acquired signal was subjected to preprocessing, including dark analysis, image lag analysis, and a uniformity test to evaluate the image quality [22]. The specifications of the normal workstation were as follows: OS, Windows 10; CPU, Intel Core i7 10700; and RAM, 64 GB.

2.3. Proposed Restoration Framework Based on the Adaptive PSF Estimation

As mentioned, the resolution and amount of noise are tradeoffs according to the scintillator thickness when the exposure conditions and pixel size of the TFT panel are the same. In general, thick scintillation results in less noise and a higher resolution than thin scintillation. The goal of the proposed method is to restore the resolution from that of a thick scintillation detector to that of a thin scintillation detector while minimizing noise amplification. Algorithm 1 shows a simplified illustration of the proposed restoration framework based on suitable PSF estimation using the factors for image quality assessment.
Algorithm 1 Structure of proposed scheme to estimate adaptive PSF
1:  Input: Initial matrix IMG1, IMG2
2:  Output: Complete matrix FSFσ
3:  FunctionInitialize ():
4:    Sigmaval = 0.01 to end (empirically);
5:    Preallocation (SSIMval, FSIMval);
6:  END
7:  Function Main ():
8:    For val = Sigmaval (start): Sigmaval (end) do
9:         PSFval  Input sigma in Equation (2) (val);
10:       IMG1_blur = IMG1  PSFval;
11:       SSIMval (val) Calculate the Equation (5) (IMG1_blur, IMG2);
12:       FSIMval (val) Calculate the Equation (6) (IMG1_blur, IMG2);
13:    END For
14:    SSIMσ = find the index (max (SSIMval));
15:    FSIMσ = find the index (max (FSIMval));
16:    FSFσ = average (SSIMσ, FSIMσ);
17: Return FSFσ
18: END
In brief, radiographic images were obtained using the scintillator thickness (here, we denote the thin scintillator thickness image as IMG1 and the thick scintillator thickness image as IMG2). Then, the PSFval was generated using one of a set of the sigma value (Sigmaval) and blurred image (IMG1_blur) designed when performing the convolution between IMG1 and PSFval. The structural similarity index (SSIM) and feature similarity index (FSIM) of IMG1_blur and IMG2 were computed. The SSIM factor is considered to be correlated with image quality perception from the perspectives of loss of correlation, luminance distortion, and contrast distortion [58]. This factor can be defined by Equation (5):
S S I M = l x , y α c x , y β s x , y γ = 2 μ x μ y + C 1 μ x 2 + μ y 2 + C 1 × 2 σ x y + C 2 σ x 2 + σ y 2 + C 2 × σ x y + C 3 σ x y + C 3 ,
C 1 = k 1 L 2 , C 2 = k 2 L 2 , C 3 = C 2 2 ,
where l x , y , c x , y , and s x , y are the luminance comparison function, contrast comparison function, and structure comparison function, respectively. u x , y , σ x , y , and σ x y are the local means, the standard deviations, and cross-covariance, respectively, of images x and y . C is a small positive constant (e.g., k 1 and k 2 are 0.01 and 0.03 as a default value, respectively), and L is the dynamic range. α , β , and γ are weight constants, and we set those constants to one in this study. In addition, by applying the default as C 3 = C 2 2 , the SSIM can be organized as 2 μ x μ y + C 1 2 σ x y + C 2 μ x 2 + μ y 2 + C 1 σ x 2 + σ y 2 + C 2 . The closer to the SSIM value, the better the adaptive σ -value of the PSF. In addition, FSIM is also a full-reference image quality assessment, and it uses important low-dimensional properties such as edges and zero crossings to determine image quality [59]. FSIM uses phase congruency (PC) and gradient of magnitude (GM) maps. FSIM was defined as follows:
F S I M = x Ω [ S P C x ] α · [ S G x ] β · P C ( x ) x Ω P C ( x ) , S P C x = 2 P C 1 x · P C 2 x + T 1 P C 1 x + P C 2 x + T 1 , S G x = 2 G M 1 x · G M 2 x + T 2 G M 1 2 x + G M 2 2 x + T 2 ,
where P C 1,2 and G M 1,2 represent the PC and GM maps of the input images, respectively. T 1 is a positive constant that controls the stability of S P C and T 2 is a positive constant according to the dynamic range of G M value. Ω denotes the whole image spatial domain, and α and β are balance parameters between PC and GM features. In this study, we followed the reference paper and used α = β = 1 . We found the index as the maximum value in the obtained SSIM (SSIMval) and FSIM (FSIMval) results. The optimal P S F σ can be used to calculate the average derived sigma index values. Finally, non-blind deconvolution was performed using the P S F σ as Equation (7):
f * = argmin IMG 2 Q P S F σ I M G 2 I M G 2 2 2 + λ I M G 2 1 ,
where Q is the set of feasible I M G 2 , P S F σ I M G 2 I M G 2 is the fidelity term, I M G 2 is the regularization term, and λ is the balancing factor due to control of the signal-to-noise ratio (i.e., λ = 0.01 was used in this study). Here, the regularization term was set to the total variation-based l1-norm calculation to avoid artifacts [44,48]. The solving method to find the f * used the augmented Lagrangian of the problem technique [60], and the tolerance used 1 × 10 4 with an iteration loop stop condition.

2.4. Quantitative Evaluation of Image Performance

The profile, contrast-to-noise ratio (CNR) [61], and GM [59] were selected for quantitative image quality assessment. The CNR is commonly used to assess the distinguishability of an object or feature from background noise. GM is a metric used to formulate the sharpness of an image. The CNR and GM were calculated as follows:
C N R   v a l u e = μ T μ B σ T 2 + σ B 2 ,
G M   v a l u e = I f h 2 + I f v 2 ,
where μ T , B and σ T , B denote the mean and standard deviation of the target or background region-of-interest (ROI), respectively. Generally, a higher CNR value indicates that it is easier to identify objects in an image by considering noise characteristics.
I denotes the input image and f h and f v are gradient operators according to the horizontal and vertical directions, such as the Sobel, Canny, Prewitt, Scharr, Laplacian, and hybrid edge operators [62]. The closer the GM value of the initially calculated thin scintillator image, the more suitable the image restoration.

3. Results and Discussion

3.1. Simulations

Figure 6 shows the plots of the calculated SSIM and FSIM values between the thick scintillator-based simulated radiographic image, IMG2, and the thin scintillator-based blurred image by convolving the PSF IMG1_blur. Here, the simulation image was designed using the numerical phantom shown in Figure 5a, and artifacts were added based on Equations (1) and (4) with the precalculated system performance evaluation parameters measured on the three detectors. The adaptive sigma for the PSF, determined by the highest SSIM value between detectors 1 and 2, was computed as 0.49 (pixels), while the corresponding value for the FSIM factor was approximately 0.41 (pixels). The average sigma across all values was calculated to be 0.45 (pixels). Similarly, the suitable sigma for detectors 3 and 1, determined through the same process, was approximately 0.63.
Figure 7 illustrates various image scenarios: (1) the reference image; (2) a degraded image assuming a thin scintillator-based detector 1; (3) a degraded image assuming a thick scintillator-based detector 2; (4) a restored image from detector 2 using the proposed PSF (σ = 0.45 pixels); (5) deconvolution images from detector 2 employing alternative PSF (σ = 2.61 pixels for Detector 1). Notably, the non-blind deconvolution based on the proposed PSF demonstrates effective resolution restoration while minimizing noise amplification, emphasizing its utility for thin scintillator-based simulated image restoration.
Figure 8 presents CNR and GM results for degraded and restored images using degradation parameters from three detectors: (a) results employing simulation images of detectors 1 and 2, and (b) results using detectors 1 and 3. Here, the CNR value is calculated and averaged using the values within the three ROIs depicted in the thin scintillator image (2) of Figure 7. In Figure 8a, average CNR values are approximately 23.43, 24.21, 22.04, and 11.21 for degraded images with detectors 1 (2), detector 2 (3), proposed image from detector 2 (4), and deconvolution images using σ = 2.61 pixels (5), respectively. Similarly, in Figure 8b, the average CNR values are approximately 23.43, 25.01, 24.28, and 11.74 for degraded images with detectors 1 (2) and 3 (3), proposed image from detector 3 (4), and deconvolution images using σ = 5.13 pixels (5), respectively. The GM values in Figure 8a,b were approximately 3.68, 4.57, 4.88, and 3.35, and 1.88, 4.57, 2.73, and 1.54, respectively. These results indicate that the resolution of the proposed image was improved compared with those of images (2) and (5). A quantitative assessment of the high-resolution image derived from the degraded image, which assumes the thick scintillator-based detector 2 (3), further validates a satisfactory resolution enhancement. Note that, utilizing the suitable PSF during non-blind deconvolution achieves resolution improvement and effectively curbs noise amplification in the resulting image (5).

3.2. Experiments

Figure 9 shows the plots of the SSIM and FSIM values used to calculate the adaptive sigma value of the PSF. Here, (a) compares the experimental images (IMG2 and IMG1_blur) of detectors 1 and 2, and (b) compares those of detectors 1 and 3. The maximum values of SSIM and FSIM were predicted for sigma values of (a) 1.03 and 1.14, and (b) 1.26 and 1.43, respectively. The average sigma value was calculated as 1.09 between detector 1 and 2, and 1.35 between detector 1 and 3. Here, we can see that the first peak occurs around 0.4 pixels and then the SSIM and FSIM values decrease and then increase again. This is a different tendency compared to that of the simulation result in Figure 6, and this is expected to be caused by asymmetric PSF, various noises, and physical distortions such as scatter radiation.
Figure 10 presents illustrative instances, including a radiographic image captured with a thin scintillator in detector 1, a degraded image attained with a thick scintillator, the resultant image using the proposed approach from a thick scintillator image, and deconvolution images employing the measured PSF for assessing system performance. Specifically, (a) showcases outcomes between detectors 1 and 2, while (b) delineates outcomes between detectors 1 and 3. The application of the proposed framework for image restoration enhances overall image quality concerning the resolution and noise characteristics in contrast to radiographic images captured with a thin scintillator (detector 1) and thick scintillators (detectors 2 and 3). These findings underscore the significance of deriving the optimal PSF by considering comprehensive quantitative evaluation criteria. Moreover, the deconvolution results for a PSF corresponding to system performance (σ = 2.61 pixels and 5.13 pixels in detectors 2 and 3, respectively) considerably improved the resolution compared with the proposed method. However, it has limitations in that noise is significantly amplified, and artifacts may occur, as indicated by the yellow arrows, owing to excessive high-frequency filtering. These results indicate that the optimal PSF is derived by considering the quantitative evaluation factors.
Figure 11 presents the 1D normalized intensity profiles measured along (a) the line AB and (b) line CD, as indicated in Figure 10a and Figure 10b, respectively. The intensity profile of the restored image, acquired using the proposed method, exhibits differentiation between 2.0 lp/mm and 4.6 lp/mm. Noticeable improvement in resolution is apparent when compared to the thick scintillator images ((2) and (5)). Clear separation in the line bars is also evident when contrasted with the thin scintillator image (1), except for the 3.7 lp/mm case in Figure 11b. At 3.7 lp/mm, the profile result of the proposed restored image does not distinctly differentiate from the profiles of (1) and (7). However, this can be improved by increasing the sigma size of the PSF, although it may result in noise amplification.
Figure 12 shows the experimental results of an electronic device radiographic image, a restored image using the proposed approach, and a deconvoluted image using the PSF of the imaging system performance. Here, (a) shows the results obtained using detectors 1 and 2, and (b) shows the results obtained using detectors 1 and 3. The magnified image shows a zoomed-in view of Box B in Figure 12.
Figure 13 shows the results of the CNR and GM with the ROIs in Figure 12. The average CNR values in Figure 13a,b were calculated approximately as 29.04, 32.51, 26.59, and 22.37 for images (1) to (4), and 34.75, 31.41, and 6.13 for images (5) to (7). Moreover, the average GM values in Figure 13a,b were determined to be approximately 5.70, 3.59, 5.33, and 5.64 for images (1) to (4), and 1.30, 4.53, and 4.84 for images (5) to (7).
The purpose of the proposed method is to find the optimal PSF considering the human visible system, and then the best results can be achieved by applying a contextualized image restoration algorithm. Figure 14 shows the experimental results of the iterative restoration method using a regularization term, a restoration method based on the Wiener filter, and the Lucy–Richardson–Rosen method [29,30,63]. Here, the results above use detector 2 and another result below uses detector 3. Wiener filtering, also known as minimum mean-square-error filtering, takes into account the stochastic characteristic of blur and noise. It is reconstructed by Equation (10):
F ^ = H * H 2 + K G ,
where F ^ is restored by the Wiener filter using the degraded image, G , in frequency domain, H is the PSF component in frequency domain, and H * is complex conjugate of H . K is the Wiener constant, which is related to the signal-to-noise ratio. The Wiener filter approximates inverse filtering by directly dividing the PSF when there are few noise components in the image. In contrast, the Wiener filter approaches zero and becomes similar to a frequency-rejection filter in a noisy image which has very large noise compared to the original image. The Lucy–Richardson–Rosen method replaces the correlation in the Lucy–Richardson algorithm by the non-linear reconstruction method [63]. The iterative Lucy–Richardson algorithm is calculated by Equation (11):
g ^ ( n + 1 ) = g ^ ( n ) f ( n ) g ^ ( n ) P S F P S F ,
where P S F denotes the complex conjugate of PSF, and f ( n ) is the initial guess. f ( n + 1 ) is generated by convolving between P S F and f ( n ) . The ratio between f ( n + 1 ) and f ( n ) is correlated with the PSF. This correlation is replaced by the non-linear reconstruction method, and the resulting residual is multiplied by the previous guess until the mismatch between the two metrics is within tolerance. This algorithm is also based on inverse filtering, and noise estimation is important. Therefore, it is essential that the noise characteristics are well derived and optimized for image restoration. These characteristics are demonstrated in Figure 14. The image from detector 2 has relatively more noise components, compared to that of the image from detector 3 at the same radiation dose. Therefore, the higher resolution image of the regularization method can be seen while optimizing the relative noise component, compared to that of the wiener filter and the Lucy–Richardson algorithm. However, when reconstructing the image of detector 3, which is less noisy, the result is not an optimal image owing to the obstacle of the regularization term of obtaining the high resolution. Comparatively, the Lucy–Richardson–Rosen method, which optimizes for noise characteristics, results in a much higher resolution. Therefore, the PSF obtained through the proposed method can achieve the best restoration result when deconvolution is performed by selecting the optimal reconstruction method according to the characteristics of the image.
In summary, the proposed method is considerably more effective in restoring the image by considering the overall image quality and predicting the appropriate PSF. However, this study had several limitations. First, some noise components are emphasized in high-pass filtering. Although the optimal PSF was predicted by accounting for noise amplification, high-frequency filtering inevitably introduces noise amplification. Several noise reduction approaches offer solutions to this problem. Chen et al. introduced an object function to solve for an optimal noise reduction image using an approximate TV regularization term and the weight of the fidelity term [64]. This model demonstrates that noise can be removed while preserving details and edges. Another approach for optimizing the parameters in the noise reduction algorithm is presented. Seo et al. optimized the similarity and weight parameters for a fast non-local means algorithm in low-dose computed tomography [65]. The optimal parameters of the noise reduction algorithm considering the X-ray exposure conditions yielded improved image quality in both quantitative and qualitative measurements. We believe that this method can be applied to reduce the noise component while maintaining sharpness and that it can significantly compensate for the problems of the proposed method.
Another limitation is the radiation–scatter component. The proposed method performs image restoration using the degraded image assumed in Equations (1) and (4). However, there is no contribution from the scatter component images in the assumed model, which have a significant impact on low-contrast radiographic image [66]. When X-rays penetrate an object, the transmitted radiation consists of primary and scattered components. The scattered component depends on the radiation energy, the field of view, and the thickness of the object [67,68]. Many studies on degradation models based on radiation–scatter characteristics have been published. The key idea behind these models is that the scatter component is modeled by adding or convolving primary radiation, and the scatter signal is mainly distributed at low frequencies [69,70,71]. Therefore, the low-frequency region contains various types of degradation information, including the effect on the scintillation thickness, and quantitative image analysis is required for an accurate image performance evaluation of the proposed algorithm. In continuous R&D, we plan to investigate the impact of the proposed method on scatter radiation using Monte Carlo simulations to describe a real radiographic imaging system.

4. Conclusions

The proposed framework predicts an adaptive PSF to convert an image from a normal-thickness scintillator to a thin scintillator. The main objective is to suppress noise amplification and artifacts as much as possible after deconvolution using adaptive PSF. The experimental results show that the average CNR value for the proposed image was approximately 31.41, which is approximately 1.08 times larger than that of the thin scintillator-based detector 1 image. The GM value of the proposed image was also approximately 4.53, which was approximately 3.48 times higher than that of the thick scintillator-based detector 3 images. These results indicate that the proposed method effectively solves the tradeoff problem between improving sharpness and preventing noise amplification. In conclusion, the proposed software is expected to be applicable to radiographic systems.

Author Contributions

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

Funding

This work was supported by the Korea Medical Device Development Fund grant funded by the Korean government (the Ministry of Science and ICT, the Ministry of Trade, Industry and Energy, the Ministry of Health & Welfare, the Ministry of Food and Drug Safety) (Project Number: 1711194312, RS-2020-KD000014).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Seibert, J.A. Flat-panel detectors: How much better are they? Pediatr. Radiol. 2006, 36, 173–181. [Google Scholar] [CrossRef] [PubMed]
  2. Beuville, E.J.; Belding, M.; Costello, A.N.; Hansen, R.; Petronio, S.M. High performance, low-noise, 128-channel readout integrated circuit for flat panel X-ray detector systems. In Proceedings of the SPIE 5368, Medical Imaging 2004: Physics of Medical Imaging, San Diego, CA, USA, 6 May 2004. [Google Scholar] [CrossRef]
  3. Aldrich, J.E.; Duran, E.; Dunlop, P.; Mayo, J.R. Optimization of dose and image quality for computed radiography and digital radiography. J. Digit. Imaging 2006, 19, 126–131. [Google Scholar] [CrossRef] [PubMed]
  4. Bosmans, H.; Hauwere, A.D.; Lemmens, K.; Zanca, F.; Thierens, H.; Ongeval, C.V.; Herck, K.V.; Steen, A.V.; Martens, P.; Bleyen, L.; et al. Technical and clinical breast cancer screening performance indicators for computed radiography versus direct digital radiography. Eur. Radiol. 2013, 23, 2891–2898. [Google Scholar] [CrossRef] [PubMed]
  5. James, J.; Davies, A.; Cowen, A.; O’Connor, P. Developments in digital radiography: An equipment update. Eur. Radiol. 2001, 11, 2616–2626. [Google Scholar] [CrossRef] [PubMed]
  6. Yaffe, M.J.; Rowlands, J.A. X-ray detectors for digital radiography. Phys. Med. Biol. 1997, 42, 1. [Google Scholar] [CrossRef] [PubMed]
  7. Lanca, L.; Silva, A. Digital Imaging Systems for Plain Radiography; Springer: New York, NY, USA, 2013; pp. 9–19. [Google Scholar] [CrossRef]
  8. Kasap, S.O.; Kabir, M.Z.; Rowlands, J.A. Recent advances in X-ray photoconductors for direct conversion X-ray image detectors. Curr. Appl. Phys. 2006, 6, 288–292. [Google Scholar] [CrossRef]
  9. Howansky, A.; Mishchenko, A.; Lubinsky, A.R.; Zhao, W. Comparison of CsI:Tl and Gd2O2S:Tb indirect flat panel detector X-ray imaging performance in front-and back irradiation geometries. Med. Phys. 2019, 46, 4857–4868. [Google Scholar] [CrossRef]
  10. Liu, X.; Xu, T.; Li, Y.; Zang, Z.; Peng, X.; Wei, H.; Zha, W.; Wang, F. Enhanced X-ray photon response in solution-synthesized CsPbBr3 nanoparticles wrapped by reduced graphene oxide. Sol. Energy Mater. Sol. Cells 2018, 187, 249–254. [Google Scholar] [CrossRef]
  11. Samei, E.; Flynn, M.J. An experimental comparison of detector performance for direct and indirect digital radiography systems. Med. Phys. 2003, 30, 608–622. [Google Scholar] [CrossRef]
  12. Swank, R.K. Absorption and noise in X-ray phosphors. J. Appl. Phys. 1973, 44, 4199–4203. [Google Scholar] [CrossRef]
  13. Kasap, S.; Frey, J.B.; Belev, G.; Tousignant, O.; Mani, H.; Greenspan, J.; Laperriere, L.; Bubon, O.; Reznik, A.; DeCrescenzo, G.; et al. Amorphous and polycrystalline photoconductors for direct conversion flat panel X-ray image sensors. Sensors 2011, 11, 5112–5157. [Google Scholar] [CrossRef]
  14. Grynko, O.; Thibault, T.; Pineau, E.; Reznik, A. Engineering of a blocking layer structure for low-lag operation of the a-PbO-based X-ray detector. IEEE Trans. Electron Devices 2021, 68, 2335–2341. [Google Scholar] [CrossRef]
  15. Chae, K.J.; Goo, J.M.; Ahn, S.Y.; Yoo, J.Y.; Yoon, S.H. Application of deconvolution algorithm of point spread function in improving image quality: An observer preference study on chest radiography. Korean J. Radiol. 2018, 19, 147–152. [Google Scholar] [CrossRef]
  16. Han, B.; Park, M.; Kim, K.; Lee, Y. Characterization of flexible amorphous silicon thin-film transistor-based detectors with positive-intrinsic-negative diode in radiography. Diagnostics 2022, 12, 2103. [Google Scholar] [CrossRef]
  17. Huda, W.; Abrahams, R.B. X-ray-based medical imaging and resolution. Am. J. Roentgenol. 2015, 204, W393–W397. [Google Scholar] [CrossRef] [PubMed]
  18. Nagesh, S.V.S.; Rana, R.; Russ, M.; Ionita, C.N.; Bednarek, D.R.; Rudin, S. Focal spot deblurring for high resolution direct conversion X-ray detectors. Proc. SPIE-Int. Soc. Opt. Eng. 2016, 9783, 97833R. [Google Scholar] [CrossRef] [PubMed]
  19. Gonzalez, R.C.; Woods, R.E. Digital Image Processing, Chapter 5: Image Restoration and Reconstruction, 3rd ed.; Pearson Education: Sydney, Australia, 2009; pp. 311–350. [Google Scholar]
  20. Samei, E.; Flynn, M.J.; Reimann, D.A. A method for measuring the presampled MTF of digital radiographic systems using an edge test device. Med. Phys. 1998, 25, 102–113. [Google Scholar] [CrossRef] [PubMed]
  21. Moy, J.P. Signal-to-noise ratio and spatial resolution in X-ray electronic imagers: Is the MTF a relevant parameters? Med. Phys. 2000, 27, 86–93. [Google Scholar] [CrossRef]
  22. Donini, B.; Rivetti, S.; Lanconelli, N.; Bertolini, M. Free software for performing physical analysis of systems for digital radiography and mammography. Med. Phys. 2014, 41, 051903. [Google Scholar] [CrossRef]
  23. Konstantinidis, A.; Martini, N.; Koukou, V.; Fountos, G.; Kalyvas, N.; Valais, I.; Michail, C. RAD_IQ: A free software for characterization of digital X-ray imaging devices based on the novel IEC 62220-1-1:2015 International Standard. J. Phys. Conf. Ser. 2021, 2090, 012107. [Google Scholar] [CrossRef]
  24. IEC 62220-1-1:2015; Medical Electrical Equipment—Characteristics of Digital X-ray Imaging Devices—Part 1-1: Determination of the Detective Quantum Efficiency—Detectors Used in Radiographic Imaging. International Electrotechnical Commission: Geneva, Switzerland, 2015.
  25. Michail, C.; Valais, I.; Martini, N.; Koukou, V.; Kalyvas, N.; Bakas, A.; Kandarakis, I.; Fountos, G. Determination of the detective quantum efficiency (DQE) of CMOS/CsI imaging detectors following the novel IEC 62220-1-1:2015 International standard. Radiat. Meas. 2016, 94, 8–17. [Google Scholar] [CrossRef]
  26. Linardatos, D.; Koukou, V.; Martini, N.; Konstantinidis, A.; Bakas, A.; Fountos, G.; Valais, I.; Michail, C. On the response of a micro non-destructive testing X-ray detector. Materials 2021, 14, 888. [Google Scholar] [CrossRef] [PubMed]
  27. Banham, M.R.; Katsaggelos, A.K. Digital image restoration. IEEE Signal Process. Mag. 1997, 14, 24–41. [Google Scholar] [CrossRef]
  28. Samarasinghe, P.D.; Kennedy, R.A.; Li, H. On non-blind image restoration. In Proceedings of the 2009 3rd International Conference on Signal Processing and Communication Systems, Omaha, NE, USA, 28–30 September 2009. [Google Scholar] [CrossRef]
  29. Biggs, D.S.C.; Andrews, M. Acceleration of iterative image restoration algorithms. Appl. Opt. 1997, 36, 1766–1775. [Google Scholar] [CrossRef]
  30. Chen, H.; Cen, Z.; Wang, C.; Lan, S.; Li, X. Image restoration via improved wiener filter applied to optical sparse aperture systems. Optik 2017, 147, 350–359. [Google Scholar] [CrossRef]
  31. Buades, A.; Coll, B.; Morel, J.-M. A non-local algorithm for image denoising. In Proceedings of the 2005 IEEE Computer Society Conference on Computer Vision and Pattern Recognition (CVPR’05), San Diego, CA, USA, 20–26 June 2005. [Google Scholar] [CrossRef]
  32. Zhao, M.; Zhang, W.; Wang, Z.; Hou, Q. Satellite image deconvolution based on nonlocal means. Appl. Opt. 2010, 49, 6286–6294. [Google Scholar] [CrossRef]
  33. Schmidt, U.; Schelten, K.; Roth, S. Bayesian deblurring with integrated noise estimation. In Proceedings of the CVPR 2011, Colorado Springs, CO, USA, 20–25 June 2011. [Google Scholar] [CrossRef]
  34. Schmidt, U.; Rother, C.; Nowozin, S.; Jancsary, J.; Roth, S. Discriminative non-blind deblurring. In Proceedings of the 2013 IEEE Conference on Computer Vision and Pattern Recognition, Portland, OR, USA, 23–28 June 2013. [Google Scholar] [CrossRef]
  35. Daniel, Z.; Weiss, Y. From learning models of natural image patches to whole image restoration. In Proceedings of the 2011 IEEE Conference on Computer Vision, Barcelona, Spain, 6–13 November 2011. [Google Scholar] [CrossRef]
  36. Sun, L.; Cho, S.; Wang, J.; Hays, J. Good image priors for non-blind deconvolution. In Proceedings of the European Conference on Computer Vision (ECCV), Zurich, Switzerland, 6–12 September 2014; pp. 231–246. [Google Scholar] [CrossRef]
  37. Zhang, J.; Pan, J.; Lai, W.-S.; Lau, R.W.H.; Yang, M.-H. Learning fully convolutional networks for iterative non-blind deconvolution. In Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition (CVPR), Honolulu, HI, USA, 21–26 July 2017. [Google Scholar] [CrossRef]
  38. Schuler, C.; Burger, H.C.; Harmeling, S.; Scholkopf, B. A machine learning approach for non-blind image deconvolution. In Proceedings of the 2013 IEEE Conference on Computer Vision and Pattern Recognition, Portland, OR, USA, 23–28 June 2013. [Google Scholar] [CrossRef]
  39. Quan, Y.; Chen, Z.; Zheng, H.; Ji, H. Learning deep non-blind image deconvolution without ground truths. In Proceedings of the European Conference on Computer Vision, Tel Aviv, Israel, 23–27 October 2022; pp. 642–659. [Google Scholar] [CrossRef]
  40. Son, H.; Lee, S. Fast non-blind deconvolution via regularized residual networks with long/short skip-connections. In Proceedings of the 2017 IEEE International Conference on Computational Photography (ICCP), Stanford, CA, USA, 12–14 May 2017. [Google Scholar] [CrossRef]
  41. Tao, S.; Dong, W.; Tang, Z.; Wang, Q. Fast non-blind deconvolution method for blurred image corrupted by Poisson noise. In Proceedings of the 2017 International Conference on Progress in Informatics and Computing (PIC), Nanjing, China, 15–17 December 2017. [Google Scholar] [CrossRef]
  42. Figueiredo, M.A.; Bioucas-Dias, J.M. Restoration of Poissonian images using althernating direction optimization. IEEE Trans. Image Process. 2010, 19, 3133–3145. [Google Scholar] [CrossRef]
  43. Tao, S.; Dong, W.; Xu, Z.; Tang, Z. Fast total variation deconvolution for blurred image contaminated by Poisson noise. J. Vis. Commun. Image Represent. 2016, 38, 582–594. [Google Scholar] [CrossRef]
  44. Ren, D.; Zuo, W.; Zhang, D.; Xu, J.; Zhang, L. Partial deconvolution with inaccurate blur kernel. IEEE Trans. Image Process. 2018, 27, 511–524. [Google Scholar] [CrossRef]
  45. Rodriguez, P. Total variation regularization algorithms for images corrupted with different noise models: A review. J. Electr. Comput. Eng. 2013, 2013, 217021. [Google Scholar] [CrossRef]
  46. Lu, H.; Wei, J.; Liu, Q.; Wang, Y.; Deng, X. A dictionary learning method with total generalized variation for MRI reconstruction. Int. J. Biomed. Imaging 2016, 2016, 7512471. [Google Scholar] [CrossRef] [PubMed]
  47. Knoll, F.; Bredies, K.; Pock, T.; Stollberger, R. Second order total generalized variation (TGV) for MRI. Magn. Reson. Med. 2011, 65, 480–491. [Google Scholar] [CrossRef]
  48. Lee, M.-H.; Yun, C.-S.; Kim, K.; Lee, Y.; The Alzheimer Disease Neuroimaging Initative. Image restoration algorithm incorporating methods to remove noise and blurring from positron emission tomography imaging for Alzheimer’s disease diagnosis. Phys. Medica 2022, 103, 181–189. [Google Scholar] [CrossRef] [PubMed]
  49. Buhr, E.; Gunther-Kohfahl, S.; Neitzel, U. Accuracy of a simple method for deriving the presampled modulation transfer function of a digital radiographic system from an edge image. Med. Phys. 2003, 30, 2323–2331. [Google Scholar] [CrossRef]
  50. Dobbins, J.T., III; Samei, E.; Ranger, N.T.; Chen, Y. Intercomparison of methods for image quality characterization. II. Noise power spectrum. Med. Phys. 2006, 33, 1466–1475. [Google Scholar] [CrossRef]
  51. Foi, A.; Trimeche, M.; Katkovnik, V.; Egiazarian, K. Practical Poissonian-Gaussian noise modeling and fitting for single-image raw-data. IEEE Trans. Image Process. 2008, 17, 1737–1754. [Google Scholar] [CrossRef] [PubMed]
  52. Luisier, F.; Blu, T.; Unser, M. Image denoising in mixed Poisson-Gaussian noise. IEEE Trans. Image Process. 2011, 20, 696–708. [Google Scholar] [CrossRef]
  53. Ding, Q.; Long, Y.; Zhang, X.; Fessler, J.A. Statistical image reconstruction using mixed Poisson-Gaussian noise model for X-ray CT. arXiv 2018. [Google Scholar] [CrossRef]
  54. Lee, S.; Lee, M.S.; Kang, M.G. Poisson-Gaussian noise analysis and estimation for low-dose X-ray images in the NSCT domain. Sensors 2018, 18, 1019. [Google Scholar] [CrossRef]
  55. Sutour, C.; Deledalle, C.-A.; Aujol, J.-F. Estimation of the noise level function based on a nonparametric detection of homogeneous image regions. SIAM J. Imaging Sci. 2015, 8, 2622–2661. [Google Scholar] [CrossRef]
  56. Kendall, M.G. A new measure of rank correlation. Biometrika 1938, 30, 81–93. [Google Scholar] [CrossRef]
  57. Kendall, M.G. The treatment of ties in ranking problems. Biometrika 1945, 33, 239–251. [Google Scholar] [CrossRef] [PubMed]
  58. Wang, Z.; Bovik, A.C.; Sheikh, H.R.; Simoncelli, E.P. Image quality assessment: From error visibility to structural similarity. IEEE Trans. Image Process. 2004, 13, 600–612. [Google Scholar] [CrossRef] [PubMed]
  59. Zhang, L.; Zhang, L.; Mou, X.; Zhang, D. FSIM: A feature similarity index for image quality assessment. IEEE Trans. Image Process. 2011, 20, 2378–2386. [Google Scholar] [CrossRef]
  60. Chan, S.H.; Khoshabeh, R.; Gibson, K.B.; Gill, P.E.; Nguyen, T.Q. An augmented lagrangian method for video restortaion. IEEE Trans. Image Process. 2011, 20, 3097–3111. [Google Scholar] [CrossRef] [PubMed]
  61. Koutalonis, M.; Delis, H.; Spyrou, G.; Costaridou, L.; Tzanakos, G.; Panayiotakis, G. Contrast-to-noise ratio in magnification mammography: A monte carlo study. Phys. Med. Biol. 2007, 52, 3185. [Google Scholar] [CrossRef]
  62. Kesarwani, A.; Purohit, K.; Dalui, M.; Kisku, D.R. Measuring the degree of suitability of edge detection operators prior to an application. In Proceedings of the 2020 IEEE Applied Signal Processing Conference (ASPCON), Kolkata, India, 7–9 October 2020. [Google Scholar] [CrossRef]
  63. Jayavel, A.; Gopinath, S.; Angamuthu, P.P.; Arockiaraj, F.G.; Bleahu, A.; Xavier, A.P.; Smith, D.; Han, M.; Slobozhan, I.; Ng, S.H.; et al. Improved classification of blurred images with deep-learning networks using lucy-richardson-rosen algorithm. Photonics 2023, 10, 396. [Google Scholar] [CrossRef]
  64. Chen, Q.; Montesinos, P.; Sun, Q.S.; Heng, P.A.; Xia, D.S. Adaptive total variation denoising based on difference curvature. Image Vis. Comput. 2010, 28, 298–306. [Google Scholar] [CrossRef]
  65. Seo, K.-H.; Kang, S.-H.; Shim, J.; Lee, Y. Optimization of smoothing factor for fast non-local means algorithm in high pitch based low-dose computed tomography images with tin-filter. Radiat. Phys. Chem. 2023, 206, 110762. [Google Scholar] [CrossRef]
  66. Johns, P.C.; Yaffe, M. Scattered radiation in fan beam imaging systems. Med. Phys. 1982, 9, 231–239. [Google Scholar] [CrossRef]
  67. Smith, S.W.; Kruger, R.A. A signal processing model of diagnostic X-ray scatter. Med. Phys. 1986, 13, 831–835. [Google Scholar] [CrossRef] [PubMed]
  68. Boone, J.M.; Lindfors, K.K.; Cooper III, V.N.; Seibert, J.A. Scatter/primary in mammography: Comprehensive results. Med. Phys. 2000, 27, 2408–2416. [Google Scholar] [CrossRef] [PubMed]
  69. Boone, J.M.; Seibert, J.A. An analytical model of the scattered radiation distribution in diagnostic radiology. Med. Phys. 1988, 15, 721–725. [Google Scholar] [CrossRef] [PubMed]
  70. Li, H.; Mohan, R.; Zhu, X.R. Scatter kernel estimation with an edge-spread function method for cone-beam computed tomography imaging. Phys. Med. Biol. 2008, 53, 6729–6748. [Google Scholar] [CrossRef] [PubMed]
  71. Monnin, P.; Verdun, F.R.; Bosmans, H.; Perez, S.R.; Marshall, N.W. A comprehensive model for X-ray projection imaging system efficiency and image quality characterization in the presence of scattered radiation. Phys. Med. Biol. 2017, 62, 5691–5722. [Google Scholar] [CrossRef]
Figure 1. Simplified illustration of the exposure condition to measure the modulation transfer function (MTF) according to the scintillation thickness. Here, measurements were implemented in compliance with IEC 62220-1-1:2015 RQA-5 protocol.
Figure 1. Simplified illustration of the exposure condition to measure the modulation transfer function (MTF) according to the scintillation thickness. Here, measurements were implemented in compliance with IEC 62220-1-1:2015 RQA-5 protocol.
Sensors 23 08185 g001
Figure 2. Scanning electron microscopy (SEM) images to measure each scintillator thickness.
Figure 2. Scanning electron microscopy (SEM) images to measure each scintillator thickness.
Sensors 23 08185 g002
Figure 3. (a) MTF graphs of three detectors and (b) the calculated line spread functions (LSFs) based on Equations (2) and (3).
Figure 3. (a) MTF graphs of three detectors and (b) the calculated line spread functions (LSFs) based on Equations (2) and (3).
Sensors 23 08185 g003
Figure 4. Predicted noise level function (NLF) of three detectors through the relationship between the mean and variance in homogeneous regions.
Figure 4. Predicted noise level function (NLF) of three detectors through the relationship between the mean and variance in homogeneous regions.
Sensors 23 08185 g004
Figure 5. (a) A 3D numerical phantom in simulation, (b) a high-resolution line chart phantom, and (c) an electronic device in experiment.
Figure 5. (a) A 3D numerical phantom in simulation, (b) a high-resolution line chart phantom, and (c) an electronic device in experiment.
Sensors 23 08185 g005
Figure 6. Graphs of SSIM and FSIM values between thick scintillator-based image and blurred thin scintillator-based image by convolving the PSF in the simulation.
Figure 6. Graphs of SSIM and FSIM values between thick scintillator-based image and blurred thin scintillator-based image by convolving the PSF in the simulation.
Sensors 23 08185 g006
Figure 7. Simulation results illustrating: (1) the reference image; (2) the degraded image assuming a thin scintillator-based detector 1; (3) the degraded image assuming a thick scintillator-based detector 2; (4) the restored image obtained through simulation using the proposed PSF for detector 2; (5) deconvolution image derived from the simulated image of detector 2 using an alternative PSF.
Figure 7. Simulation results illustrating: (1) the reference image; (2) the degraded image assuming a thin scintillator-based detector 1; (3) the degraded image assuming a thick scintillator-based detector 2; (4) the restored image obtained through simulation using the proposed PSF for detector 2; (5) deconvolution image derived from the simulated image of detector 2 using an alternative PSF.
Sensors 23 08185 g007
Figure 8. Contrast-to-noise ratio (CNR) and gradient of magnitude (GM) bar graphs; (a) the results using the simulation images of detector 1 and detector 2, and (b) those of detector 1 and detector 3.
Figure 8. Contrast-to-noise ratio (CNR) and gradient of magnitude (GM) bar graphs; (a) the results using the simulation images of detector 1 and detector 2, and (b) those of detector 1 and detector 3.
Sensors 23 08185 g008
Figure 9. Plots of the SSIM and FSIM values; (a) detectors 1 and 2, and (b) detectors 1 and 3.
Figure 9. Plots of the SSIM and FSIM values; (a) detectors 1 and 2, and (b) detectors 1 and 3.
Sensors 23 08185 g009
Figure 10. Experimental results of the radiographic image with a thin scintillator for detector 1, the degraded image with a thick scintillator, the proposed image derived from the thick scintillator image, and the deconvolution images utilizing the measured PSF to evaluate system performance. Here, (a) represents the outcomes between detectors 1 and 2, and (b) similarly depicts the outcomes between detectors 1 and 3.
Figure 10. Experimental results of the radiographic image with a thin scintillator for detector 1, the degraded image with a thick scintillator, the proposed image derived from the thick scintillator image, and the deconvolution images utilizing the measured PSF to evaluate system performance. Here, (a) represents the outcomes between detectors 1 and 2, and (b) similarly depicts the outcomes between detectors 1 and 3.
Sensors 23 08185 g010
Figure 11. Normalized intensity profiles as measured along (a) line AB and (b) line CD are indicated in Figure 10a and Figure 10b, respectively.
Figure 11. Normalized intensity profiles as measured along (a) line AB and (b) line CD are indicated in Figure 10a and Figure 10b, respectively.
Sensors 23 08185 g011
Figure 12. Examples of electronic device images, a restored image using the proposed method, and a deconvolution image using the PSF obtained by measuring the system performance. Here, the images of (1) to (4) a show the results using detectors 1 and 2, and the images of (1) and (5) to (7) show results using detectors 1 and 3.
Figure 12. Examples of electronic device images, a restored image using the proposed method, and a deconvolution image using the PSF obtained by measuring the system performance. Here, the images of (1) to (4) a show the results using detectors 1 and 2, and the images of (1) and (5) to (7) show results using detectors 1 and 3.
Sensors 23 08185 g012
Figure 13. Bar graphs of CNR and GM results with ROIs in Figure 12. Here, (a) compares detector 1 and 2, and (b) compares detector 1 and 3.
Figure 13. Bar graphs of CNR and GM results with ROIs in Figure 12. Here, (a) compares detector 1 and 2, and (b) compares detector 1 and 3.
Sensors 23 08185 g013
Figure 14. Experimental results of the iterative restoration method using regularization term-based minimization, a restoration method based on the Wiener filter, and the Lucy–Richardson–Rosen method. Here, the results above use detector 2 and another result below uses detector 3.
Figure 14. Experimental results of the iterative restoration method using regularization term-based minimization, a restoration method based on the Wiener filter, and the Lucy–Richardson–Rosen method. Here, the results above use detector 2 and another result below uses detector 3.
Sensors 23 08185 g014
Table 1. Detector specifications.
Table 1. Detector specifications.
NameScintillator Thickness (μm)Pixel Size
(μm)
Pixel Matrix (Pixels)ADC σ of LSF
(Pixels)
Detector 18448512 × 102412-bit1.79
Detector 29648512 × 102412-bit2.61
Detector 314048512 × 102412-bit5.13
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

Cha, B.K.; Lee, Y.; Kim, K. Development of Adaptive Point-Spread Function Estimation Method in Various Scintillation Detector Thickness for X-ray Imaging. Sensors 2023, 23, 8185. https://doi.org/10.3390/s23198185

AMA Style

Cha BK, Lee Y, Kim K. Development of Adaptive Point-Spread Function Estimation Method in Various Scintillation Detector Thickness for X-ray Imaging. Sensors. 2023; 23(19):8185. https://doi.org/10.3390/s23198185

Chicago/Turabian Style

Cha, Bo Kyung, Youngjin Lee, and Kyuseok Kim. 2023. "Development of Adaptive Point-Spread Function Estimation Method in Various Scintillation Detector Thickness for X-ray Imaging" Sensors 23, no. 19: 8185. https://doi.org/10.3390/s23198185

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