Skip to main content

ORIGINAL RESEARCH article

Front. Comput. Sci., 28 June 2023
Sec. Computer Vision
Volume 5 - 2023 | https://doi.org/10.3389/fcomp.2023.1131317

Difference of anisotropic and isotropic TV for segmentation under blur and Poisson noise

  • 1Department of Mathematics, University of California, Irvine, Irvine, CA, United States
  • 2Department of Mathematical Sciences, University of Texas at Dallas, Richardson, TX, United States
  • 3Department of Mathematics & Computer Science, Whittier College, Whittier, CA, United States

In this paper, we aim to segment an image degraded by blur and Poisson noise. We adopt a smoothing-and-thresholding (SaT) segmentation framework that finds a piecewise-smooth solution, followed by k-means clustering to segment the image. Specifically for the image smoothing step, we replace the least-squares fidelity for Gaussian noise in the Mumford-Shah model with a maximum posterior (MAP) term to deal with Poisson noise and we incorporate the weighted difference of anisotropic and isotropic total variation (AITV) as a regularization to promote the sparsity of image gradients. For such a nonconvex model, we develop a specific splitting scheme and utilize a proximal operator to apply the alternating direction method of multipliers (ADMM). Convergence analysis is provided to validate the efficacy of the ADMM scheme. Numerical experiments on various segmentation scenarios (grayscale/color and multiphase) showcase that our proposed method outperforms a number of segmentation methods, including the original SaT.

1. Introduction

Image segmentation partitions an image into multiple, coherent regions, where pixels of one region share similar characteristics such as colors, textures, and edges. It remains an important yet challenging problem in computer vision that has various applications, including magnetic resonance imaging (Duan et al., 2015; Tongbram et al., 2021; Li et al., 2022) and microscopy (Zosso et al., 2017; Bui et al., 2020). One of the most fundamental models for image segmentation is the Mumford-Shah model (Mumford and Shah, 1989) because of its robustness to noise. Given an input image f : Ω → ℝ defined on an open, bounded, and connected domain Ω ⊂ ℝ2, the Mumford-Shah model is formulated as

minu,ΓEMS(u,Γ):=λ2Ω(fu)2dx+μ2Ω\Γ|u|2dx+Length(Γ),    (1)

where u : Ω → ℝ is a piecewise-smooth approximation of the image f, Γ ⊂ Ω is a compact curve representing the region boundaries, and λ, μ > 0 are the weight parameters. The first term in (1) is the fidelity term that ensures that the solution u approximates the image f. The second term enforces u to be piecewise smooth on Ω\Γ. The last term measures the perimeter, or more mathematically the one-dimensional Haussdorf measure in ℝ2 (Bar et al., 2011), of the curve Γ. However, (1) is difficult to solve because the unknown set of boundaries needs to be discretized. One common approach involves approximating the objective function in (1) by a sequence of elliptic functionals (Ambrosio and Tortorelli, 1990).

Alternatively, Chan and Vese (2001) (CV) simplified (1) by assuming the solution u to be piecewise constant with two phases or regions, thereby making the model easier to solve via the level-set method (Osher and Sethian, 1988). Let the level-set function ϕ be Lipschitz continuous and be defined as follows:

{ϕ(x)>0if x is inside Γ,ϕ(x)=0if x is at Γ,ϕ(x)<0if x is outside Γ.

By the definition of ϕ, the curve Γ is represented by ϕ(x) = 0. The image region can be defined as either inside or outside the curve Γ. In short, the CV model is formulated as

minc1,c2,ϕECV(c1,c2,ϕ):=λ(Ω|fc1|2H(ϕ) dx                                               +  Ω|fc2|2(1H(ϕ)) dx)                                                +  νΩ|H(ϕ)| dx,    (2)

where λ, ν are weight parameters, the constants c1, c2 are the mean intensity values of the two regions, and H(ϕ) is the Heaviside function defined by H(ϕ) = 1 if ϕ ≥ 0 and H(ϕ) = 0 otherwise. A convex relaxation (Chan et al., 2006) of (2) was formulated as

minc1,c2,u[0,1]λ(Ω|fc1|2u dx+Ω|fc2|2(1u)dx)                                                                                       +νΩ|u|dx,

where an image segmentation ũ is obtained by thresholding u, that is

u˜(x)={1if u(x)>τ,0if u(x)τ,

for some value τ ∈ (0, 1). It can be solved efficiently by convex optimization algorithms, such as the alternating direction method of multipliers (ADMM) (Boyd et al., 2011) and primal-dual hybrid gradient (Chambolle and Pock, 2011). A multiphase extension of (2) was proposed in Vese and Chan (2002), but it requires that the number of regions to be segmented is a power of 2. For segmenting into an arbitrary number of regions, fuzzy membership functions were incorporated (Li et al., 2010).

Cai et al. (2013) proposed the smoothing-and-thresholding (SaT) framework that is related to the model (1). In the smoothing step of SaT, a convex variant of (1) is formulated as

u*=arg minuλ2Ω(fAu)2dx+μ2Ω|u|2dx+Ω|u|dx,    (3)

yielding a piecewise-smooth solution u*. The blurring operator A is included in the case when the image f is blurred. The total variation (TV) term Ω|u|dx is a convex approximation of the length term in (2) by the coarea formula (Chan et al., 2006). After the smoothing step, a thresholding step is applied to the smooth image u* to segment it into multiple regions. The two-stage framework has many advantages. First, the smoothing model (3) is strongly convex, so it can be solved by any convex optimization algorithm to obtain a unique solution u*. Second, the user can adjust the number of thresholds to segment u* and the threshold values to obtain a satisfactory segmentation result, thanks to the flexibility of the thresholding step. Furthermore, the SaT framework can be adapted to color images by incorporating an intermediate lifting step (Cai et al., 2017). Before performing the thresholding step, the lifting step converts the RGB space to Lab (perceived lightness, red- green and yellow-blue) color space and concatenates both RGB and Lab intensity values into a six-channel image. The multi-stage framework for color image segmentation is called smoothing, lifting, and thresholding (SLaT).

One limitation of (3) lies in the ℓ2 fidelity term that is statistically designed for images corrupted by additive Gaussian noise, and as a result, the smoothing step is not applicable to other types of noise distribution. In this paper, we aim at Poisson noise, which is commonly encountered when an image is taken by photon-capturing devices such as in positron emission tomography (Vardi et al., 1985) and astronomical imaging (Lantéri and Theys, 2005). By using the data fidelity term of AuflogAu (Le et al., 2007), we obtain a smoothing model that is appropriate for Poisson noise (Chan et al., 2014):

minuλΩ(AuflogAu)dx+μ2Ω|u|2dx+Ω|u|dx.    (4)

As a convex approximation of the length term in (1), the TV term in (4) can be further improved by nonconvex regularizations. The TV regularization is defined by the ℓ1 norm of the image gradient. Literature has shown that nonconvex regularizations often yield better performance than the convex ℓ1 norm in identifying sparse solutions. Examples of nonconvex regularization include ℓp, 0 < p < 1 (Chartrand, 2007; Xu Z. et al., 2012; Cao et al., 2013), ℓ1 − αℓ2, α ∈ [0, 1] (Lou et al., 2015a,b; Ding and Han, 2019; Li P. et al., 2020; Ge and Li, 2021), ℓ1/ℓ2 (Rahimi et al., 2019; Wang et al., 2020; Xu et al., 2021), and an error function (Guo et al., 2021). Lou et al. (2015c) designed a TV version of ℓ1 − αℓ2 called the weighted difference of anisotropic–isotropic total variation (AITV), which outperforms TV in various imaging applications, such as image denoising (Lou et al., 2015c), image reconstruction (Lou et al., 2015c; Li P. et al., 2020), and image segmentation (Bui et al., 2021, 2022; Wu et al., 2022b).

In this paper, we propose an AITV variant of (4) to improve the smoothing step of the SaT/SLaT framework for images degraded by Poisson noise and/or blur. Incorporating AITV regularization is motivated by our previous works (Park et al., 2016; Bui et al., 2021, 2022), where we demonstrated that AITV regularization is effective in preserving edges and details, especially under Gaussian and impulsive noise. To maintain similar computational efficiency as the original SaT/SLaT framework, we propose an ADMM algorithm that utilizes the ℓ1 − αℓ2 proximal operator (Lou and Yan, 2018). The main contributions of this paper are as follows:

• We propose an AITV-regularized variant of (4) and prove the existence of a minimizer for the model.

• We develop a computationally efficient ADMM algorithm and provide its convergence analysis under certain conditions.

• We conduct numerical experiments on various grayscale/color images to demonstrate the effectiveness of the proposed approach.

The rest of the paper is organized as follows. Section 2 describes the background information such as notations, Poisson noise, and the SaT/SLaT framework. In Section 3, we propose a simplified Mumford-Shah model with AITV and a MAP data fidelity term for Poisson noise. In the same section, we show that the model has a global minimizer and develop an ADMM algorithm with convergence analysis. In Section 4, we evaluate the performance of the AITV Poisson SaT/SLaT framework on various grayscale and color images. Lastly, we conclude the paper in Section 5.

2. Preliminaries

2.1. Notation

Throughout the rest of the paper, we represent images and mathematical models in discrete notations (i.e., vectors and matrices). An image is represented as an M × N matrix, and hence the image domain is denoted by Ω = {1, 2, …, M} × {1, 2, …, N}. We define two inner product spaces: X: = ℝM×N and Y: = X × X. Let uX. For shorthand notation, we define u ≥ 0 if ui,j ≥ 0 for all (i,j) ∈ Ω. The discrete gradient operator ∇:XY is defined by (∇u)i, j = [(xu)i, j, (yu)i, j], where

(xu)i,j={ui,jui,j1if 2jN,ui,1ui,Nifj=1,

and

(yu)i,j={ui,jui1,jif 2iM,u1,juM,jif i=1.

The space X is equipped with the standard inner product 〈·, ·〉X, and Euclidean norm ||·||2. The space Y has the following inner product and norms: for p = (p1, p2) ∈ Y and q = (q1, q2) ∈ Y,

p,qY=p1,q1X+p2,q2X,     ||p||1=i=1Mj=1N|(p1)i,j|+|(p2)i,j|,     ||p||2 =i=1Mj=1N|(p1)i,j|2+|(p2)i,j|2,   ||p||2,1 =i=1Mj=1N(p1)i,j2+(p2)i,j2.

For brevity, we omit the subscript X or Y in the inner product when its context is clear.

2.2. AITV regularization

There are two popular discretizations of total variation: the isotropic TV (Rudin et al., 1992) and the anisotropic TV (Choksi et al., 2011), which are defined by

||u||2,1=i=1Mj=1N|(xu)i,j|2+|(yu)i,j|2,||u||1=i=1Mj=1N|(xu)i,j|+|(yu)i,j|,

respectively. This work is based on the weighted difference between anisotropic and isotropic TV (AITV) regularization (Lou et al., 2015c), defined by

||u||1α||u||2,1=i=1Mj=1N(|(xu)i,j|+|(yu)i,j|                                                               α|(xu)i,j|2+|(yu)i,j|2),    (5)

for a weighting parameter α ∈ [0, 1]. The range of α ensures the non-negativity of the AITV regularization. Note that anisotropic TV is defined as the ℓ1 norm of the image gradient [(xu)i,j, (yu)i,j] at the pixel location (i, j) ∈ Ω, while isotropic TV is the ℓ2 norm on the gradient vector. As a result, AITV can be viewed as the ℓ1 − αℓ2 regularization on the gradient vector at every pixel, thereby enforcing sparsity individually at each gradient vector.

2.3. Poisson noise

Poisson noise follows the Poisson distribution with mean and variance η, whose probability mass function is given by

η(n)=e-ηηnn!,n0.    (6)

For a clean image gX, its intensity value at each pixel gi, j serves as the mean and variance for the corresponding noisy observation fX defined by

fi,j~Poisson(gi,j)(i,j)Ω.

To recover the image g from the noisy image f, we find its maximum a posteriori (MAP) estimation u, which maximizes the probability ℙ(u|f). By Bayes' theorem, we have

(u|f)=(f|u)(u)(f).

It further follows from the definition (6) that

(fi,j|ui,j)(ui,j)=ui,j(fi,j)(ui,j)=e-ui,jui,jfi,j(fi,j)!(ui,j).

Since Poisson noise is i.i.d. pixelwise, we have

(u|f)=(i,j)Ω(fi,j|ui,j)(ui,j)(fi,j)=(i,j)Ωe-ui,jui,jfi,j(fi,j)!(ui,j)(fi,j).

The MAP estimate of ℙ(u|f) is equivalent to its negative logarithm, thus leading to the following optimization problem:

minu0(i,j)Ωui,j-fi,jlogui,j-log(ui,j).    (7)

The last term −log ℙ(ui,j) can be regarded as an image prior or a regularization. For example, Le et al. (2007) considered the isotropic total variation as the image prior and proposed a Poisson denoising model

minu0u-flogu,11+||u||2,1,    (8)

where log is applied pixelwise and 11 is the matrix whose entries are all 1's. The first term in (8) is a concise notation that is commonly used as a fidelity term for Poisson denoising in various imaging applications (Le et al., 2007; Chan et al., 2014; Wen et al., 2016; Chang et al., 2018; Chowdhury et al., 2020a,b).

2.4. Review of Poisson SaT/SLaT

A Poisson SaT framework (Chan et al., 2014) consists of two steps. Given a noisy grayscale image fX corrupted by Poisson noise, the first step is the smoothing step that finds a piecewise-smooth solution u* from the optimization model:

u*=arg minu0λAu-flogAu,11+μ2||u||22+||u||2,1.    (9)

Then in the thresholding step, K − 1 threshold values τ1 ≤ τ2 ≤ … ≤ τK−1 are appropriately chosen to segment u* into K regions, where the kth region is given by

Ωk={(i,j)Ω:τk-1ui,j*<τk},

with τ0:=infxΩu*(x). The thresholding step is typically performed by k-means clustering.

The Poisson smoothing, lifting, and thresholding (SLaT) framework (Cai et al., 2017) extends the Poisson SaT framework to color images. For a color image f = (f1, f2, f3) ∈ X × X × X, the model (9) is applied to each color channel fi for i = 1, 2, 3, thus leading to a smoothed color image u*=(u1*,u2*,u3*). An additional lifting step (Luong, 1993) is performed to transform u* to (u1,u2,u3) in the Lab space (perceived lightness, red-green, and yellow-blue). The channels in Lab space are less correlated than in RGB space, so they may have useful information for segmentation. The RGB image and the Lab image are concatenated to form the multichannel image û: = (u1*,u2*,u3*,u1,u2,u3), followed by the thresholding stage. Generally, k-means clustering yields K centroids c1, …, cK as constant vectors, which are used to form the region

Ωk={(i,j)Ω:||ûi,j-ck||2=min1κK||ûi,j-cκ||2}

for k = 1, …, K such that Ωk's are disjoint and k=1KΩk=Ω.

After the thresholding step for both SaT/SLaT, we define a piecewise-constant approximation of the image f by

f~=(f~1,,f~d)such thatf~=k=1Kck,11Ωk=1,,d,    (10)

where ck, ℓ is the ℓth entry of the constant vector ck and

11Ωk={1if (i,j)Ωk,0if (i,j)Ωk.

Recall that d = 1 when f is grayscale, and d = 3 when f is color.

3. Proposed approach

To improve the Poisson SaT/SLaT framework, we propose to replace the isotropic TV in (9) with AITV regularization. In other words, in the smoothing step, we obtain the smoothed image u* from the optimization problem

u*=arg minuF(u):=λAuflogAu,11+μ2||u||22                                                       +||u||1α||u||2,1,    (11)

for α ∈ [0, 1]. We establish that this model admits a global solution. We then develop an ADMM algorithm to find a solution and provide the convergence analysis. The overall segmentation approach is described in Algorithm 1.

ALGORITHM 1
www.frontiersin.org

Algorithm 1. AITV Poisson SaT/SLaT.

3.1. Model analysis

To establish the solution's existence of the proposed model (11), we start with Lemma 1, a discrete version of Poincaré's inequality (Evans, 2010). In addition, we prove Lemma 2 and Proposition 3, leading to the global existence theorem (Theorem 4).

Lemma 1. There exists a constant C > 0 such that

||u-ū11||2C||u||2,1,    (12)

for every uX and ū:=1MNi=1Mj=1Nui,j.

Proof. We prove it by contradiction. Suppose there exists a sequence {uk}k=1 such that

||uk-ūk11||2>k||uk||2,1,    (13)

where ūk=1MNi=1Mj=1N(uk)i,j.. For every k, we normalize each element in the sequence by vk=uk-ūk11||uk-ūk11||2. It is straightforward that

v̄k=1MNi=1Mj=1N(vk)i,j=0,||vk||2=1k.    (14)

By (13), we have

||vk||2,1<1k.    (15)

As {vk}k=1 is bounded, there exists a convergent subsequence {vkj}j=1 such that vkjv* for v*X. It follows from (15) that ||v*||2,1=0. Since ker(∇) = {c11:c ∈ ℝ}, then v* is a constant vector. However, (14) implies that v̄*=0 and ||v*||2=1. This contradiction proves the lemma.               ⃣

Lemma 2. Suppose ||f|| < ∞ and mini,jfi,j>0. There exists a scalar u0 > 0 such that we have 2(xfi,j log x) ≥ x for any xu0 and (i, j) ∈ Ω.

Proof. For each (i,j) ∈ Ω, we want to show that there exists ui,j > 0 such that H(x): = x − 2fi,j log x ≥ 0 for xui,j. Since H(x) is strictly convex and it attains a global minimum at x = 2fi,j, it is increasing on the domain x > 2fi,j. Additionally as x dominates log (x) as x → +∞, there exists ui,j > 2fi,j > 0 such that ui,jlogui,j2fi,j, which implies that H(ui,j) = ui,j − 2fi,j log ui,j ≥ 0. As a result, for xui,j > 2fi,j, we obtain x − 2fi,j log x = H(x) ≥ H(ui,j) ≥ 0. Define u0:=maxi,jui,j, and hence we have 2(xfi,j log x) ≥ x for xu0ui,j, ∀(i, j) ∈ Ω.

Proposition 3. Suppose ker(A) ∩ ker(∇) = {0} and {uk}k=1X. If {(Auk,uk)}k=1 is bounded, then {uk}k=1 is bounded.

Proof. Since ker(A) ∩ ker(∇) = {0}, we have A11 ≠ 0. Simple calculations lead to

|u¯k|A1​​12 =  A(u¯k1​​1)2A(u¯k1​​1uk)2+Auk2                                Auku¯k1​​12+Auk2                                CAuk2,1+Auk2,    (16)

where the last inequality is due to Lemma 1. The boundedness of {Auk}k=1 and {uk}k=1 implies that {ūk}k=1 is also bounded by (16). We apply Lemma 1 to obtain

||uk||2||uku¯k11||2+||u¯k11||2<C||uk||2,1+||u¯k11||2<,

which thereby proves that {uk}k=1 is bounded.

Finally, we adapt the proof in Chan et al. (2014) to establish that F has a global minimizer.

Theorem 4. Suppose ||f|| < ∞ and mini,jfi,j>0. If λ > 0, μ ≥ 0, α ∈ [0, 1), and ker(A) ∩ ker(∇) = {0}, then F has a global minimizer.

Proof. It is straightforward that ||∇u||2, 1 ≤ ||∇u||1, thus ||∇u||1 − α||∇u||2, 1 ≥ 0 for α ∈ [0, 1). As a result, we have

F(u)λAu-flogAu,11=λi=1Mj=1N(Au)i,j-fi,jlog(Au)i,j.

Given a scalar f > 0, the function G(x) = xflog(x) attains its global minimum at x = f. Therefore, we have xfi,j log xfi,jfi,j log fi,j for all x > 0 and (i,j) ∈ Ω, which leads to a lower bound of F(u), i.e.,

F(u)λi=1Mj=1N(Au)i,jfi,jlog(Au)i,j           λi=1Mj=1Nfi,jfi,jlogfi,j=:F0.    (17)

As F(u) is lower bounded by F0, we can choose a minimizing sequence {uk}k=1 and hence F(uk) has a uniform upper bound, denoted by B1, i.e., F(uk) < B1 for all k ∈ ℕ. It further follows from (17) that

B1F(uk)λAuk-flogAuk,11F0,

which implies that {|Auk-flogAuk,11|}k=1 is uniformly bounded, i.e., there exists a constant B2 > 0 such that |〈Aukf log Auk, 11〉| < B2, ∀k. Using these uniform bounds, we derive that

(1α)||uk||1μ2||uk||22+||uk||1α||uk||2,1           =F(uk)λAukflogAuk,11B1+λB2.

As α < 1, the sequence {uk}k=1 is bounded.

To prove the boundedness of {Auk}k=1, we introduce the notations of x+ = max(x, 0) and x = −min(x, 0) for any x ∈ ℝ. Then x = x+x. By Lemma 2, there exists u0 > 0 such that 2(xfi,j log x) ≥ x, ∀xu0 and (i,j) ∈ Ω. We observe that

Auk1=i=1Mj=1N|(Auk)i,j|               i=1Mj=1Nmax{2((Auk)i,jfi,jlog(Auk)i,j),u0}              2i=1Mj=1N((Auk)i,jfi,jlog(Auk)i,j)++MNu0              =2i=1Mj=1N[((Auk)i,jfi,jlog(Auk)i,j)                 +((Auk)i,jfi,jlog(Auk)i,j)]+MNu0             =2AukflogAuk,1​​1               + 2i=1Mj=1N((Auk)i,jfi,jlog(Auk)i,j)+MNu0            2B2+2i=1Mj=1N|fi,jfi,jlogfi,j|+MNu0<.    (18)

This shows that {Auk}k=1 is bounded.

Since both {uk}k=1 and {Auk}k=1 are bounded, then {uk}k=1 is bounded due to Proposition 3. Therefore, there exists a subsequence {ukn}n=1 that converges to some u*X. As F is continuous and thus lower semicontinuous, we have

F(u*)lim infnF(ukn),

which means that u* minimizes F.

3.2. Numerical algorithm

To minimize (11), we introduce two auxiliary variables vX and w = (wx, wy) ∈ Y, leading to an equivalent constrained optimization problem:

minu,v,w  λvflogv,1​​1+μ2u22+w1αw2,1s.t.     Au=v,u=w.    (19)

The corresponding augmented Lagrangian is expressed as

β1,β2(u,v,w,y,z)  =λvflogv,1​​1+μ2u22+w1αw2,1  + y,Auv+β12Auv22+z,uw  +β22uw22,    (20)

where yX and z = (zx, zy) ∈ Y are Lagrange multipliers and β1, β2 are positive parameters. We then apply the alternating direction method of multipliers (ADMM) to minimize (19) that consists of the following steps per iteration k:

uk+1=arg minuLβ1,k,β2,k(u,vk,wk,yk,zk),    (21a)
vk+1=arg minvLβ1,k,β2,k(uk+1,v,wk,yk,zk),    (21b)
wk+1=arg minwLβ1,k,β2,k(uk+1,vk+1,w,yk,zk),    (21c)
yk+1=yk+β1,k(Auk+1-vk+1),    (21d)
zk+1=zk+β2,k(uk+1-wk+1),    (21e)
(β1,k+1,β2,k+1)=σ(β1,k,β2,k),    (21f)

where σ > 1.

Remark 1. The scheme presented in (21) slightly differs from the original ADMM (Boyd et al., 2011), the latter of which has σ = 1 in (21f). Having σ > 1 increases the weights of the penalty parameters β1, k, β2, k in each iteration k, thus accelerating the numerical convergence speed of the proposed ADMM algorithm. A similar technique has been used in Cascarano et al. (2021), Gu et al. (2017), Storath and Weinmann (2014), Storath et al. (2014), and You et al. (2019).

All the subproblems (21a)–(21c) have closed-form solutions. In particular, the first-order optimality condition for (21a) is

[β1,kAA-(μ+β2,k)Δ]uk+1=A(β1,kvk-yk)-(zk-β2,kwk),    (22)

where Δ = −∇∇ is the Laplacian operator. If ker(A) ∩ ker(∇) = {0}, then β1,kAA-(μ+β2,k)Δ is positive definite and thereby invertible, which implies that (22) has a unique solution uk+1. By assuming periodic boundary condition for u, the operators Δ and AA are block circulant (Wang et al., 2008), and hence (22) can be solved efficiently by the 2D discrete Fourier transform F. Specifically, we have the formula

uk+1    =1((A)*(β1,kvkyk)()*(zkβ2,kwk)β1,k(A)*(A)(μ+β2,k)(Δ)),    (23)

where F-1 is the inverse discrete Fourier transform, the superscript * denotes complex conjugate, the operation ° is componentwise multiplication, and division is componentwise. By differentiating the objective function of (21b) and setting it to zero, we can get a closed-form solution for vk+1 given by

vk+1=(β1,kAuk+1+ykλ11)+(β1,kAuk+1+ykλ11)2+4λβ1,kf2β1,k,    (24)

where the square root, squaring, and division are performed componentwise. Lastly, the w-subproblem (21c) can be decomposed componentwise as follows:

(wi,j)k+1= arg minwi,jwi,j1αwi,j2                     +β2,k2wi,j((uk+1)i,j+(zk)i,jβ2,k)22                =prox((uk+1)i,j+(zk)i,jβ2,k,α,1β2,k),    (25)

where the proximal operator for ℓ1 − αℓ2 on x ∈ ℝn is given by

prox(x,α,β)=arg miny||y||1-α||y||2+12β||x-y||22.    (26)

The proximal operator for ℓ1 − αℓ2 has a closed form solution summarized by Lemma 5.

Lemma 5. (Lou and Yan, 2018) Given x ∈ ℝn, β > 0, and α ∈ [0, 1], the optimal solution to (26) is given by one of the following cases:

1. When ||x|| > β, we have

x*=(||ξ||2+αβ)ξ||ξ||2,

where ξ = sign(x)° max (|x| − β, 0).

2. When (1 − α)β < ||x|| ≤ β, then x* is a 1-sparse vector such that one chooses i ∈ argmaxj(|xj|) and defines xi*=(|xi|+(α-1)β)sign(xi) and the remaining elements equal to 0.

3. When ||x|| ≤ (1 − α)β, then x* = 0.

In summary, we describe the ADMM scheme to solve (11) in Algorithm 2.

ALGORITHM 2
www.frontiersin.org

Algorithm 2. ADMM for the AITV-regularized smoothing model with Poisson fidelity (Equation 11).

3.3. Convergence analysis

We establish the subsequential convergence of ADMM described in Algorithm 2. The global convergence of ADMM (Wang et al., 2019) is inapplicable to our model as the gradient operator ∇ is non-surjective, which will be further investigated in future work. For the sake of brevity, we set β = β1 = β2 and denote

Lβ(u,v,w,y,z):=Lβ,β(u,v,w,y,z).

In addition, we introduce definitions of subdifferentials (Rockafellar and Wets, 2009), which defines a stationary point of a non-smooth objective function.

Definition 6. For a proper function h:ℝn → ℝ∪{+∞}, define dom(h): = {x ∈ ℝn:h(x) < +∞}.

(a) The regular subdifferential at x ∈ dom(h) is given by

^h(x):={w:lim infxx,xxh(x)-h(x)-w,x-x||x-x||0}.

(b) The (limiting) subdifferential at x ∈ dom(h) is given by

h(x):={w:xkx and wk^h(xk)with wkw and h(xk)h(x)}.

An important property of the limiting subdifferential is its closedness: for any (xk, vk) → (x, v) with vk ∈ ∂h(xk), if h(xk) → h(x), then v ∈ ∂h(x).

Lemma 7. Suppose that ker(A) ∩ ker(∇) = {0} and 0 ≤ α < 1. Let {(uk,vk,wk,yk,zk)}k=1 be a sequence generated by Algorithm 2. Then, we have

βk+1(uk+1,vk+1,wk+1,yk+1,zk+1)βk(uk,vk,wk,yk,zk)   ν2uk+1uk22β02vk+1vk22  +1σk1β0(yk+1yk22+zk+1zk22),    (28)

for some constant ν > 0.

Proof. If ker(A) ∩ ker(∇) = {0}, then β0AA+(β0+μ) is positive definite, and hence there exists ν > 0 such that

βk||Au||22+(βk+μ)||u||22β0||Au||22+(β0+μ)||u||22                                                         ν||u||22k,

which implies that Lβk(u,vk,wk,yk,zk) is strongly convex with respect to u with parameter ν. Additionally, Lβk(uk+1,v,wk,yk,zk) is strongly convex with respect to v with parameter β0 ≤ βk. It follows from Beck (2017), Theorem 5.25, that we have

βk(uk+1,vk,wk,yk,zk)βk(uk,vk,wk,yk,zk)ν2||uk+1uk||22,    (29)
βk(uk+1,vk+1,wk,yk,zk)βk(uk+1,vk,wk,yk,zk)β02||vk+1vk||22.    (30)

As wk+1 is the optimal solution to (21c), it is straightforward to have

Lβk(uk+1,vk+1,wk+1,yk,zk)-Lβk(uk+1,vk+1,wk,yk,zk)0.    (31)

Simple calculations by using (21d)–(21e) lead to

βk(uk+1,vk+1,wk+1,yk+1,zk+1)βk(uk+1,vk+1,wk+1,yk,zk)              =(βk(uk+1,vk+1,wk+1,yk+1,zk+1)                 βk(uk+1,vk+1,wk+1,yk+1,zk))               +(βk(uk+1,vk+1,wk+1,yk+1,zk)               βk(uk+1,vk+1,wk+1,yk,zk))       =zk+1zk,uk+1wk+1+yk+1yk,Auk+1vk+1      =1βk(yk+1yk22+zk+1zk22).    (32)

Lastly, we have

βk+1(uk+1,vk+1,wk+1,yk+1,zk+1)                      βk(uk+1,vk+1,wk+1,yk+1,zk+1)            =βk+1βk2(Auk+1vk+122+uk+1wk+122)           =βk+1βk2βk2(yk+1yk22+zk+1zk22).    (33)

Combining (29)–(33) together with the fact that βk=σkβ0 for σ > 1, we obtain

βk+1(uk+1,vk+1,wk+1,yk+1,zk+1)βk(uk,vk,wk,yk,zk)                ν2uk+1uk22β02vk+1vk22                   +βk+1+βk2βk2(yk+1yk22+zk+1zk22)               =ν2uk+1uk22β02vk+1vk22                  +σ+12σkβ0(yk+1yk22+zk+1zk22)             ν2uk+1uk22β02vk+1vk22                +1σk1β0(yk+1yk22+zk+1zk22).

This completes the proof.

Lemma 8. Suppose that ker(A) ∩ ker(∇) = {0} and 0 ≤ α < 1. Let {(uk,vk,wk,yk,zk)}k=1 be generated by Algorithm 2. If {yk}k=1 bounded, then the sequence {(uk,vk,wk,yk,zk)}k=1 is bounded, uk+1uk → 0, and vk+1vk → 0.

Proof. First we show that {zk}k=1 is bounded. Combining (21e) with the first-order optimality condition of (25), we have

(zk+1)i,j=(zk)i,j+βk((uk+1)i,j-(wk+1)i,j)(||(wk+1)i,j||1-α||(wk+1)i,j||2)(||(wk+1)i,j||1)-α(||(wk+1)i,j||2),    (34)

which implies that there exist ξ1 ∈ ∂||(wk+1)i,j||1 and ξ2 ∈ ∂||(wk+1)i,j||2 such that (zk+1)i,j = ξ1−αξ2 for each (i,j) ∈ Ω. Recall that for x ∈ ℝ2 the subgradients of the two norms are

||x||1={ξ2:ξi={sign(xi)if xi0ξi[1,1]if xi=0fori=1,2},    (35)
||x||2={ξ2:ξ={x||x||2                                                           if x0{ξ2:||ξ||21}      if x=0}    (36)

Therefore, we have ||ξ1|| ≤ 1, ||ξ2|| ≤ 1, and hence ||(zk+1)i,j|| ≤ 1+α (by the triangle inequality), i.e., {zk}k=1 is bounded.

By the assumption {(yk)}k=1 is bounded. There exist two constants C1, C2 > 0 such that ||yk+1-yk||22C1, ||zk+1-zk||22C1, ||yk||22C2, and |zk||22C2 for all k ∈ ℕ. Hence, we have from (28) that

βk+1(uk+1,vk+1,wk+1,yk+1,zk+1)βk(uk,vk,wk,yk,zk)ν2uk+1uk22    β02vk+1vk22+2C1σk1β0.    (37)

A telescoping summation of (37) leads to

βk+1(uk+1,vk+1,wk+1,yk+1,zk+1)β0(u0,v0,w0,y0,z0)+2C1β0i=0k1σi1  ν2i=0kui+1ui22β02i=0kvi+1vi22.    (38)

By completing two least-squares terms, we can rewrite Lβk+1 as

   βk+1(uk+1,vk+1,wk+1,yk+1,zk+1)=λvk+1flogvk+1,1​​1+μ2uk+122    +wk+11αwk+12,1    +βk+12Auk+1vk+1+yk+1βk+122yk+1222βk+1   +βk+12uk+1wk+1+zk+1βk+122zk+1222βk+1.    (39)

Combining (38) and (39), we have

λfflogf,1​​1+(1α)wk+11C2β0βk+1(uk+1,vk+1,wk+1,yk+1,zk+1)β0(u0,v0,w0,y0,z0)+2C1β0i=0k1σi1ν2i=0kui+1ui22β02i=0kvi+1vi22β0(u0,v0,w0,y0,z0)+2C1β0i=01σi1.    (40)

Since σ > 1, the infinite sum is finite, and hence we have ∀k ∈ ℕ,

                 ||wk+1||111α(β0(u0,v0,w0,y0,z0)λfflogf,11+2C1β0i=01σi1+C2β0)<,

which implies that {wk}k=1 is bounded. Also from (38) and (39), we have

λf-flogf,11-C2β0λvk+1-flogvk+1,11-C2β0λvk+1-flogvk+1,11-||yk+1||222βk+1-||zk+1||222βk+1Lβk+1(uk+1,vk+1,wk+1,yk+1,zk+1)Lβ0(u0,v0,w0,y0,z0)+2C1β0i=01σi-1<.

This shows that {vk-flogvk,11}k=1 is bounded. By emulating the computation in (18), it can be shown that {vk}k=1 is bounded.

It suffices to prove that {(Auk,uk)}k=1 is bounded in order to prove the boundedness of {uk}k=1 by Proposition 3. Using (21d), we have

||Auk+1||2||yk+1-yk||2βk+||vk+1||2C1β0+||vk+1||2.

As {vk}k=1 is proven to be bounded, then {Auk}k=1 is also bounded. We can prove {uk}k=1 is bounded similarly using (21e). Altogether, {(uk,vk,wk,yk,zk)}k=1 is bounded.

It follows from (40) that

ν2i=0k||ui+1-ui||22+β02i=0k||vi+1-vi||22Lβ0(u0,v0,w0,y0,z0)+2C1β0i=0k1σi-1-λf-flogf,11+C2β0.

As k → ∞, we see the right-hand side is finite, which forces the infinite summations on the left-hand side to converge, and hence we have uk+1uk → 0 and vk+1vk → 0.

Theorem 9. Suppose that ker(A) ∩ ker(∇) = {0} and 0 ≤ α < 1. Let {(uk,vk,wk,yk,zk)}k=1 be generated by Algorithm 2. If {yk}k=1 bounded, βk(vk+1vk) → 0, βk(wk+1wk) → 0, yk+1yk → 0, and zk+1zk → 0, then there exists a subsequence whose limit point (u*, v*, w*, y*, z*) is a stationary point of (19) that satisfies the following:

0=-μΔu*+Ay*+z*,    (41a)
0=λ(11-fv*)-y*,    (41b)
z*(||w*||1-α||w*||2,1),    (41c)
Au*=v*,    (41d)
u*=w*.    (41e)

Proof. By Lemma 8, the sequence {(uk,vk,wk,yk,zk)}k=1 is bounded, so there exists a subsequence {(ukn,vkn,wkn,ykn,zkn)}n=1 that converges to a point (u*, v*, w*, y*, z*). Additionally, we have uk+1uk → 0 and vk+1vk → 0. Since {(yk,zk)}k=1 is bounded, there exists a constant C > 0 such that ||yk+1yk||2 < C and ||zk+1zk||2 < C for each k ∈ ℕ. By (21e), we have

wk+1wk2 wk+1uk+12+uk+1uk2                          +ukwk2                          =zk+1zk2βk+uk+1uk2                         +zkzk12βk1                         2Cβk1+uk+1uk2.

As k → ∞, we have wk+1wk → 0. Altogether, we can derive the following results:

limn(ukn+1,vkn+1,wkn+1)=limn(ukn,vkn,wkn)=(u*,v*,w*).    (42)

Furthermore, the assumptions give us

limnβkn(vkn+1vkn)=0,limnβkn(wkn+1wkn)=0,            lim nykn+1ykn=0,            limnzkn+1zkn=0.

By (21d)–(21e), we have

||Au*v*||2 =limn||Aukn+1vkn+1||2                           =limn||ykn+1ykn||2βknlimnCβkn=0,||u*w*||2 =limn||ukn+1wkn+1||2                           =limn||zk+1zk||2βknlimnCβkn=0.

Hence, we have Au* = v* and ∇u* = w*.

The optimality conditions at iteration kn are the following:

μΔukn+1+Aykn+βknA(Aukn+1vkn)+zkn+βkn(ukn+1wkn)=0,    (43a)
λ(11-fvkn+1)-ykn-βkn(Aukn+1-vkn+1)=0,    (43b)
zkn+βkn(ukn+1-wkn+1)(||wkn+1||1-α||wkn+1||2,1).    (43c)

Expanding (43a) by substituting in (21d)–(21e) and taking the limit, we have

0=limnμΔukn+1+Aykn+βknA(Aukn+1vkn)       +zkn+βkn(ukn+1wkn)   =limnμΔukn+1+Aykn+βknA(Aukn+1vkn+1)      +βknA(vkn+1vkn)+zkn      +βkn(ukn+1wkn+1)      +βkn(wkn+1wkn)   =limnμΔukn+1+Aykn+A(ykn+1ykn)     +βknA(vkn+1vkn)+zkn     +(zkn+1zkn)+βkn(wkn+1wkn)  =μΔu*+Ay*+z*.

Substituting in (21d) into (43b) and taking the limit give us

0= limnλ(11fvkn+1)yknβkn(Aukn+1vkn+1)   = limnλ(11fvkn+1)ykn(ykn+1ykn)   =  λ(11fv*)y*.

Lastly, by substituting (21e) into (43c), we have

zkn+1(||wkn+1||1-α||wkn+1||2,1).

By continuity, we have ||wkn+1||1-α||wkn+1||2,1||w*||1-α||w*||2,1. Together with the fact that (wkn+1,zkn+1)(w*,z*), we have z*(||w*||1-α||w*||2,1) by closedness of the subdifferential.

Therefore, (u*, v*, w*, y*, z*) is a stationary point.

Remark 2. It is true that the assumptions in Theorem 9 are rather strong, but they are standard in the convergence analyses of other ADMM algorithms for nonconvex problems that fail to satisfy the conditions for global convergence in Wang et al. (2019). For example, Jung (2017), Jung et al. (2014), Li et al. (2016) and Li Y. et al. (2020) assumed convergence of the successive differences of the primal variables and Lagrange multipliers. Instead, we modify the convergence of the successive difference of the primal variables, i.e., βk(vk+1vk) → 0, βk(wk+1wk) → 0. Boundedness of the Lagrange multiplier (i.e., {yk}k=1) was also assumed in Liu et al. (2022) and Xu Y. et al. (2012), which required a stronger assumption than ours regarding the successive difference of the Lagrange multipliers.

4. Numerical experiments

In this section, we apply the proposed method of AITV Poisson SaT/SLaT on various grayscale and color images for image segmentation. For grayscale images, we compare our method with the original TV SaT (Chan et al., 2014), thresholded-Rudin-Osher-Fatemi (T-ROF) (Cai et al., 2019), and the Potts model (Potts, 1952) solved by either Pock's algorithm (Pock) (Pock et al., 2009) or Storath and Weinmann's algorithm (Storath) (Storath and Weinmann, 2014). For color images, we compare with TV SLaT (Cai et al., 2017), Pock's method (Pock et al., 2009), and Storath's method (Storath and Weinmann, 2014). We can solve (9) for TV SaT/SLaT via Algorithm 2 that utilizes the proximal operator corresponding to the ||·||2, 1 norm. The code for T-ROF is provided by the respective author1 and we can adapt it to handle blur by using a more general data fidelity term. Pock's method is implemented by the lab group2. Storath's method is provided by the original author3. Note that T-ROF, Pock's method, and Storath's method are designed for images corrupted with Gaussian noise. We apply the Anscombe transform (Anscombe, 1948) to the test images, after which the Poisson noise becomes approximately Gaussian noise. Since Storath's method is not for segmentation, we perform a post-processing step of k-means clustering to its piecewise-constant output. For the SLaT methods, we parallelize the smoothing step separately for each channel.

To quantitatively measure the segmentation performance, we use the DICE index (Dice, 1945) and peak signal-to-noise ratio (PSNR). Let S ⊂ Ω be the ground-truth region and S′ ⊂ Ω be a region obtained from the segmentation algorithm corresponding to the ground-truth region S. The DICE index is formulated by

DICE=2|SS||S|+|S|.

To compare the piecewise-constant reconstruction f~ according to (10) with the original test image f, we compute PSNR by

PSNR=20 log10(M×N)×Pij(fi,jf˜i,j)2,

where M × N is the image size and P=maxi,jfi,j.

Poisson noise is added to the test images by the MATLAB command poissrnd. To ease parameter tuning, we scale each test image to [0, 1] after its degradation with Poisson noise and/or blur. We set σ = 1.25 and β1, 0 = β2, 0 = 1.0, 2.0 in Algorithm 2 for grayscale and color images, respectively. The stopping criterion is either 300 iterations or when the relative error of uk is below ϵ = 10−4. We tune the fidelity parameter λ and the smoothing parameter μ for each image, which will be specified later. For T-ROF, Pock's method, and Storath's method, their parameters are manually tuned to give the best DICE indices for binary segmentation (Section 4.1) and the PSNR values for multiphase segmentation (Sections 4.2-4.3). All experiments are performed in MATLAB R2022b on a Dell laptop with a 1.80 GHz Intel Core i7-8565U processor and 16.0 GB RAM.

4.1. Grayscale, binary segmentation

We start with performing binary segmentation on the entire DRIVE dataset (Staal et al., 2004) that consists of 20 images shown in Figure 1. Each image has size 584 × 565 with modified pixel values of either 200 for the background or 255 for the vessels. Before adding Poisson noise, we set the peak value of the image to be P/2 or P/5, where P = 255. Note that a lower peak value indicates stronger noise in the image, thus more challenging for denoising. We examine three cases: (1) P/2 no blur, (2) P/5 no blur, and (3) P/2 with Gaussian blur specified by MATLAB command fspecial(“gaussian,” [10 10], 2). For the TV SaT method, we set λ = 14.5, μ = 0.5 for case (1), λ = 8.0, μ = 0.5 for case (2), and λ = 22.5, μ = 0.25 for case (3). For the AITV SaT method, the parameters λ and μ are set the same as TV SaT, and we have α = 0.3 for cases (1)–(2) and α = 0.8 for case (3).

FIGURE 1
www.frontiersin.org

Figure 1. The entire DRIVE dataset (Staal et al., 2004) for binary segmentation. The image size is 584×565 with background value of 200 and the pixel value for vessels to be 255.

Table 1 records the DICE indices and the computational time in seconds for the competing methods, averaged over 20 images. We observe that AITV SaT attains the best DICE indices for all three cases with comparable computational time to TV SaT and T-ROF, all of which are much faster than Pock and Storath. As visually illustrated in Figure 2, AITV SaT segments more of the thinner vessels compared to TV SaT and T-ROF in five images, thereby having the higher average DICE indices.

TABLE 1
www.frontiersin.org

Table 1. DICE and computational time in seconds of the binary segmentation methods averaged over 20 images in Figure 1 with standard deviations in parentheses.

FIGURE 2
www.frontiersin.org

Figure 2. Binary segmentation results of Figure 1 with peak P/2 under Gaussian blur and Poisson noise.

4.2. Grayscale, multiphase segmentation

We examine the multiphase segmentation on the entire BrainWeb dataset (Aubert-Broche et al., 2006) that consists of 20 grayscale images as shown in Figure 3. Each image is of size 104 × 87 and has four regions to segment: background, cerebrospinal fluid (CSF), gray matter (GM), and white matter (WM). The pixel values are 10 (background), 48 (CSF), 106 (GM), and 154 (WM). The maximum intensity P = 154. We consider two cases: (1) P/2 no blur and (2) P/2 with motion blur specified by fspecial(“motion,” 5, 225). For the SaT methods, we have μ = 1.0, α = 0.6, 0.7, and λ = 4.0, 5.0 for case (1) and case (2), respectively.

FIGURE 3
www.frontiersin.org

Figure 3. The entire BrainWeb dataset (Aubert-Broche et al., 2006) for grayscale, multiphase segmentation. Each image is of size 104 × 87. The pixel values are 10 (background), 48 (cerebrospinal fluid), 106 (gray matter), and 154 (white matter).

Across all 20 images of the BrainWeb dataset, Table 2 reports the average DICE indices for CSF, GM, and WM and average computational times in seconds of the segmentation methods. For both cases (1) and (2), AITV SaT attains the highest average DICE indices for segmenting CSF, GM, and WM. AITV SaT is comparable to TV SaT and T-ROF in terms of computational time.

TABLE 2
www.frontiersin.org

Table 2. DICE and computational time in seconds of the multiphase segmentation methods averaged over 20 images in Figure 3 with standard deviations in parentheses.

Figure 4 shows the segmentation results of the first image in Figure 3 for case (1). When segmenting CSF, the methods (TV SaT, AITV SaT, and Storath) yield similar visual results, while Pock fails to segment roughly half of the region. In addition, AITV SaT segments the most GM region with the least amount of noise artifacts than the other methods. Lastly, for WM segmentation, AITV SaT avoids the “holes” or “gaps” and segments fewer regions outside of the ground truth, thus outperforming TV SaT and Storath. For the three regions, T-ROF has the most noise artifacts in its segmentation results.

FIGURE 4
www.frontiersin.org

Figure 4. Segmentation result of the first image of Figure 3 with peak P/8 under Poisson noise with no blur. From top to bottom are segmentation results for CSF, GM, and WM.

4.3. Color segmentation

We perform color image segmentation on 10 images shown in Figure 5, which are selected from the PASCAL VOC 2010 dataset (Everingham et al., 2009). Each image has six different color regions. Figure 5A is of size 307 × 461; Figure 5B is of 500 × 367; Figures 5C, H, I are of 500 × 375; and Figures 5DG, J are of 375 × 500. Before adding Poisson noise to each channel of each image, we set the peak value P = 10. We choose the parameters of the SLaT methods to be λ = 1.5, μ = 0.05, and α = 0.6.

FIGURE 5
www.frontiersin.org

Figure 5. Test images from the PASCAL VOC 2010 dataset (Everingham et al., 2009) for color, multiphase segmentation. Each image has six regions. The image sizes are (A) 307 × 461, (B) 500 × 367, (C) 500 × 375, (D–G) 375 × 500, (H, I) 500 × 375, and (J) 375 × 500.

Figures 6, 7 present the piecewise-constant approximations via (10), showing similar segmentation results obtained by TV SLaT, AITV SLaT, and Storath. Quantitatively in Table 3, AITV SLaT has better PSNRs than TV SLaT and Pock for all the images and outperforms Storath for seven. Overall, the proposed AITV SLaT has the highest PSNR on average over 10 images with the lowest standard deviation and comparable speed as Storath.

FIGURE 6
www.frontiersin.org

Figure 6. Color image segmentation results of Figures 5AE.

FIGURE 7
www.frontiersin.org

Figure 7. Color image segmentation results of Figures 5FJ.

TABLE 3
www.frontiersin.org

Table 3. PSNR and computation time in seconds for color image segmentation of the images in Figure 5.

4.4. Parameter analysis

The proposed smoothing model (11) involves the following parameters:

The fidelity parameter λ weighs how close the approximation Au* is to the original image f. For a larger amount of noise, the value of λ should be chosen smaller.

The smoothing parameter μ determines how smooth the solution u* should be. A larger value of μ may improve denoising, but at a cost of smearing out the edges between adjacent regions, which may be segmented together if they have similar colors.

The sparsity parameter α ∈ [0, 1] determines how sparse the gradient vector at each pixel should be. More specifically, the closer the value α is to 1, the more ||∇u||1−α||∇u||2, 1 resembles ||∇u||0.

We perform sensitivity analysis on these parameters to understand how they affect the segmentation quality of AITV SaT/SLaT. We consider two types of tests in the case of P/8 with motion blur in Figure 3. In the first case (Figure 8), we fix μ = 1.0 and vary α, λ. In the second case (Figure 9), we fix λ = 5.0 and vary α, μ. Figure 8 reveals a concave relationship of the DICE index of each region with respect to the parameter λ, which implies there exists the optimal choice of λ. Additionally, when λ is small, a large value for α can improve the DICE indices. According to Figure 9, the DICE indices of the GM and WM regions decrease with respect to μ, while the DICE index of the CSF region is approximately constant. For α = 0.8, the DICE indices of the GM and WM regions are the largest when μ ≥ 1, but the large α is not optimal for CSF. Hence, an intermediate value of α, such as 0.6, is preferable to attain satisfactory segmentation quality for all three regions.

FIGURE 8
www.frontiersin.org

Figure 8. Sensitivity analysis on λ for the P/8 with motion blur case of Figure 3. The parameter μ = 1.0 is fixed. DICE indices averaged over 20 images for each brain region are plotted with respect to λ.

FIGURE 9
www.frontiersin.org

Figure 9. Sensitivity analysis on μ for the P/8 with motion blur case of Figure 3. The parameter λ = 5.0 is fixed. DICE indices averaged over 20 images for each brain region are plotted with respect to μ.

Lastly, in Figure 10, we conduct sensitive analysis for the case of P = 10 of Figure 5. We fix μ = 0.05, while varying λ, α in Figure 10A, which indicates that the optimal value for α is in the range of 0.5 ≤ α ≤ 0.7. Then we fix λ = 1.5 to examine μ and α in Figure 10B. For 0.2 ≤ α ≤ 0.7, PSNR decreases as μ increases. Again, α = 0.6 generally yields the best PSNR.

FIGURE 10
www.frontiersin.org

Figure 10. Sensitivity analysis of parameters for P = 10 case of Figure 5. (A) Sensitivity analysis on λ when μ = 0.05 fixed. (B) Sensitivity analysis on μ when λ = 1.5 fixed. Average PSNR is plotted.

5. Conclusion and future work

In this paper, we developed the AITV Poisson SaT/SLaT framework for image segmentation. In particular, we proposed a simplified Mumford-Shah model with the AITV regularization and Poisson fidelity for the smoothing step. The model was proven to have a global minimizer. Our numerical algorithm incorporated a specific splitting scheme for ADMM and the ℓ1−αℓ2 proximal operator for solving a subproblem. Convergence analysis established that the sequence generated by ADMM has a convergent subsequence to a stationary point of the nonconvex model. In our numerical experiments, the AITV Poisson SaT/SLaT yielded high-quality segmentation results within seconds for various grayscale and color images corrupted with Poisson noise and/or blur. For future directions, we are interested in other nonconvex regularization, such as ℓ1/ℓ2 on the gradient (Wang et al., 2021, 2022; Wu et al., 2022a), ℓp, 0 < p < 1, on the gradient (Hintermüller and Wu, 2013; Li Y. et al., 2020; Wu et al., 2021b), and transformed total variation (Huo et al., 2022), as alternatives to AITV. On the other hand, we can develop AITV variants of weighted TV (Li and Li, 2022) or adaptive TV (Wu et al., 2021a; Zhang et al., 2022). Moreover, we plan to determine how to make the sparsity parameter α in AITV adaptable to each image. In future work, we will adapt other segmentation algorithms (Li et al., 2010, 2016; Jung et al., 2014; Jung, 2017; Cai et al., 2019; Yang et al., 2022; Pang et al., 2023) designed for Gaussian noise or impulsive noise to Poisson noise.

Data availability statement

Publicly available datasets were analyzed in this study. This data can be found at: https://drive.grand-challenge.org/DRIVE/; https://github.com/casperdcl/brainweb; https://www2.eecs.berkeley.edu/Research/Projects/CS/vision/bsds/. Code for AITV Poisson SaT/SLaT is available at https://github.com/kbui1993/Official_Poisson_AITV_SaT_SLaT.

Author contributions

KB performed the experiments and analysis and drafted the manuscript. All authors contributed to the design, evaluation, discussions, and production of the manuscript. All authors contributed to the article and approved the submitted version.

Funding

This work was partially supported by NSF grants DMS-1846690, DMS-1854434, DMS-1952644, DMS-2151235, and a Qualcomm Faculty Award.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher's note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Footnotes

References

Ambrosio, L., and Tortorelli, V. M. (1990). Approximation of functional depending on jumps by elliptic functional via t-convergence. Commun. Pure Appl. Math. 43, 999–1036. doi: 10.1002/cpa.3160430805

CrossRef Full Text | Google Scholar

Anscombe, F. J. (1948). The transformation of Poisson, binomial and negative-binomial data. Biometrika 35, 246–254. doi: 10.1093/biomet/35.3-4.246

CrossRef Full Text | Google Scholar

Aubert-Broche, B., Griffin, M., Pike, G. B., Evans, A. C., and Collins, D. L. (2006). Twenty new digital brain phantoms for creation of validation image data bases. IEEE Trans. Med. Imaging 25, 1410–1416. doi: 10.1109/TMI.2006.883453

PubMed Abstract | CrossRef Full Text | Google Scholar

Bar, L., Chan, T. F., Chung, G., Jung, M., Kiryati, N., Mohieddine, R., et al. (2011). “Mumford and Shah model and its applications to image segmentation and image restoration,” in Handbook of Mathematical Methods in Imaging, ed O. Scherzer (New York, NY: Springer), 1095–1157. doi: 10.1007/978-0-387-92920-0_25

PubMed Abstract | CrossRef Full Text | Google Scholar

Beck, A. (2017). First-order Methods in Optimization. Philadelphia, PA: SIAM. doi: 10.1137/1.9781611974997

CrossRef Full Text | Google Scholar

Boyd, S., Parikh, N., Chu, E., Peleato, B., and Eckstein, J. (2011). Distributed optimization and statistical learning via the alternating direction method of multipliers. Found. Trends Mach. Learn. 3, 1–122. doi: 10.1561/2200000016

CrossRef Full Text | Google Scholar

Bui, K., Fauman, J., Kes, D., Torres Mandiola, L., Ciomaga, A., Salazar, R., et al. (2020). Segmentation of scanning tunneling microscopy images using variational methods and empirical wavelets. Pattern Anal. Appl. 23, 625–651. doi: 10.1007/s10044-019-00824-0

CrossRef Full Text | Google Scholar

Bui, K., Lou, Y., Park, F., and Xin, J. (2022). An efficient smoothing and thresholding image segmentation framework with weighted anisotropic-isotropic total variation. arXiv preprint arXiv:2202.10115. doi: 10.48550/arXiv.2202.10115

CrossRef Full Text | Google Scholar

Bui, K., Park, F., Lou, Y., and Xin, J. (2021). A weighted difference of anisotropic and isotropic total variation for relaxed Mumford-Shah color and multiphase image segmentation. SIAM J. Imaging Sci. 14, 1078–1113. doi: 10.1137/20M1337041

CrossRef Full Text | Google Scholar

Cai, X., Chan, R., Nikolova, M., and Zeng, T. (2017). A three-stage approach for segmenting degraded color images: smoothing, lifting and thresholding (SLaT). J. Sci. Comput. 72, 1313–1332. doi: 10.1007/s10915-017-0402-2

CrossRef Full Text | Google Scholar

Cai, X., Chan, R., Schonlieb, C.-B., Steidl, G., and Zeng, T. (2019). Linkage between piecewise constant Mumford-Shah model and Rudin-Osher-Fatemi model and its virtue in image segmentation. J. Sci. Comput. 41, B1310–B1340. doi: 10.1137/18M1202980

CrossRef Full Text | Google Scholar

Cai, X., Chan, R., and Zeng, T. (2013). A two-stage image segmentation method using a convex variant of the Mumford-Shah model and thresholding. SIAM J. Imaging Sci. 6, 368–390. doi: 10.1137/120867068

CrossRef Full Text | Google Scholar

Cao, W., Sun, J., and Xu, Z. (2013). Fast image deconvolution using closed-form thresholding formulas of Lq(q=12,23) regularization. J. Vis. Commun. Image Represent. 24, 31–41. doi: 10.1016/j.jvcir.2012.10.006

CrossRef Full Text | Google Scholar

Cascarano, P., Calatroni, L., and Piccolomini, E. L. (2021). Efficient ℓ0 gradient-based super-resolution for simplified image segmentation. IEEE Trans. Comput. Imaging 7, 399–408. doi: 10.1109/TCI.2021.3070720

CrossRef Full Text | Google Scholar

Chambolle, A., and Pock, T. (2011). A first-order primal-dual algorithm for convex problems with applications to imaging. J. Math. Imaging Vis. 40, 120–145. doi: 10.1007/s10851-010-0251-1

CrossRef Full Text | Google Scholar

Chan, R., Yang, H., and Zeng, T. (2014). A two-stage image segmentation method for blurry images with Poisson or multiplicative Gamma noise. SIAM J. Imaging Sci. 7, 98–127. doi: 10.1137/130920241

CrossRef Full Text | Google Scholar

Chan, T. F., Esedoglu, S., and Nikolova, M. (2006). Algorithms for finding global minimizers of image segmentation and denoising models. SIAM J. Appl. Math. 66, 1632–1648. doi: 10.1137/040615286

PubMed Abstract | CrossRef Full Text | Google Scholar

Chan, T. F., and Vese, L. A. (2001). Active contours without edges. IEEE Trans. Image Process. 10, 266–277. doi: 10.1109/83.902291

PubMed Abstract | CrossRef Full Text | Google Scholar

Chang, H., Lou, Y., Duan, Y., and Marchesini, S. (2018). Total variation-based phase retrieval for Poisson noise removal. SIAM J. Imaging Sci. 11, 24–55. doi: 10.1137/16M1103270

CrossRef Full Text | Google Scholar

Chartrand, R. (2007). Exact reconstruction of sparse signals via nonconvex minimization. IEEE Signal Process. Lett. 14, 707–710. doi: 10.1109/LSP.2007.898300

CrossRef Full Text | Google Scholar

Choksi, R., Gennip, Y. G., and Oberman, A. (2011). Anisotropic total variation regularized L1 approximation and denoising/deblurring of 2D bar codes. Inverse Probl. Imaging 5, 591–617. doi: 10.3934/ipi.2011.5.591

CrossRef Full Text | Google Scholar

Chowdhury, M. R., Qin, J., and Lou, Y. (2020a). Non-blind and blind deconvolution under Poisson noise using fractional-order total variation. J. Math. Imaging Vis. 62, 1238–1255. doi: 10.1007/s10851-020-00987-0

CrossRef Full Text | Google Scholar

Chowdhury, M. R., Zhang, J., Qin, J., and Lou, Y. (2020b). Poisson image denoising based on fractional-order total variation. Inverse Probl. Imaging 14, 77–96. doi: 10.3934/ipi.2019064

CrossRef Full Text | Google Scholar

Dice, L. R. (1945). Measures of the amount of ecologic association between species. Ecology 26, 297–302. doi: 10.2307/1932409

CrossRef Full Text | Google Scholar

Ding, L., and Han, W. (2019). αℓ1−βℓ2 regularization for sparse recovery. Inverse Probl. 35:125009. doi: 10.1088/1361-6420/ab34b5

PubMed Abstract | CrossRef Full Text | Google Scholar

Duan, Y., Chang, H., Huang, W., Zhou, J., Lu, Z., and Wu, C. (2015). The L0 regularized Mumford-Shah model for bias correction and segmentation of medical images. IEEE Trans. Image Process. 24, 3927–3938. doi: 10.1109/TIP.2015.2451957

PubMed Abstract | CrossRef Full Text | Google Scholar

Evans, L. C. (2010). Partial Differential Equations, Vol. 19. Providence: American Mathematical Society. doi: 10.1090/gsm/019

CrossRef Full Text | Google Scholar

Everingham, M., Van Gool, L., Williams, C. K., Winn, J., and Zisserman, A. (2009). The PASCAL visual object classes (VOC) challenge. Int. J. Comput. Vis. 88, 303–308. doi: 10.1007/s11263-009-0275-4

CrossRef Full Text | Google Scholar

Ge, H., and Li, P. (2021). The Dantzig selector: recovery of signal via ℓ1−αℓ2 minimization. Inverse Probl. 38:015006. doi: 10.1088/1361-6420/ac39f8

CrossRef Full Text | Google Scholar

Gu, S., Xie, Q., Meng, D., Zuo, W., Feng, X., and Zhang, L. (2017). Weighted nuclear norm minimization and its applications to low level vision. Int. J. Comput. Vis. 121, 183–208. doi: 10.1007/s11263-016-0930-5

CrossRef Full Text | Google Scholar

Guo, W., Lou, Y., Qin, J., and Yan, M. (2021). A novel regularization based on the error function for sparse recovery. J. Sci. Comput. 87, 1–22. doi: 10.1007/s10915-021-01443-w

CrossRef Full Text | Google Scholar

Hintermüller M. and Wu, T.. (2013). Nonconvex TVq-models in image restoration: analysis and a trust-region regularization-based superlinearly convergent solver. SIAM J. Imaging Sci. 6, 1385–1415. doi: 10.1137/110854746

CrossRef Full Text | Google Scholar

Huo, L., Chen, W., Ge, H., and Ng, M. K. (2022). Stable image reconstruction using transformed total variation minimization. SIAM J. Imaging Sci. 15, 1104–1139. doi: 10.1137/21M1438566

CrossRef Full Text | Google Scholar

Jung, M. (2017). Piecewise-smooth image segmentation models with l1 data-fidelity terms. J. Sci. Comput. 70, 1229–1261. doi: 10.1007/s10915-016-0280-z

CrossRef Full Text | Google Scholar

Jung, M., Kang, M., and Kang, M. (2014). Variational image segmentation models involving non-smooth data-fidelity terms. J. Sci. Comput. 59, 277–308. doi: 10.1007/s10915-013-9766-0

CrossRef Full Text | Google Scholar

Lantéri, H., and Theys, C. (2005). Restoration of astrophysical images-the case of Poisson data with additive Gaussian noise. EURASIP J. Adv. Signal Process. 2005, 1–14. doi: 10.1155/ASP.2005.2500

CrossRef Full Text | Google Scholar

Le, T., Chartrand, R., and Asaki, T. J. (2007). A variational approach to reconstructing images corrupted by Poisson noise. J. Math. Imaging Vis. 27, 257–263. doi: 10.1007/s10851-007-0652-y

CrossRef Full Text | Google Scholar

Li, F., Ng, M. K., Zeng, T. Y., and Shen, C. (2010). A multiphase image segmentation method based on fuzzy region competition. SIAM J. Imaging Sci. 3, 277–299. doi: 10.1137/080736752

CrossRef Full Text | Google Scholar

Li, F., Osher, S., Qin, J., and Yan, M. (2016). A multiphase image segmentation based on fuzzy membership functions and L1-norm fidelity. J. Sci. Comput. 69, 82–106. doi: 10.1007/s10915-016-0183-z

CrossRef Full Text | Google Scholar

Li, H., Guo, W., Liu, J., Cui, L., and Xie, D. (2022). Image segmentation with adaptive spatial priors from joint registration. SIAM J. Imaging Sci. 15, 1314–1344. doi: 10.1137/21M1444874

CrossRef Full Text | Google Scholar

Li, M.-M., and Li, B.-Z. (2022). A novel weighted anisotropic total variational model for image applications. Signal Image Video Process. 16, 211–218. doi: 10.1007/s11760-021-01977-4

CrossRef Full Text | Google Scholar

Li, P., Chen, W., Ge, H., and Ng, M. K. (2020). ℓ1−αℓ2 minimization methods for signal and image reconstruction with impulsive noise removal. Inverse Probl. 36:055009. doi: 10.1088/1361-6420/ab750c

CrossRef Full Text | Google Scholar

Li, Y., Wu, C., and Duan, Y. (2020). The TVp regularized Mumford-Shah model for image labeling and segmentation. IEEE Trans. Image Process. 29, 7061–7075. doi: 10.1109/TIP.2020.2997524

CrossRef Full Text | Google Scholar

Liu, H., Deng, K., Liu, H., and Wen, Z. (2022). An entropy-regularized ADMM for binary quadratic programming. J. Glob. Optim. 1–33. doi: 10.1007/s10898-022-01144-0

CrossRef Full Text | Google Scholar

Lou, Y., Osher, S., and Xin, J. (2015a). “Computational aspects of constrained L1L2 minimization for compressive sensing,” in Modelling, Computation and Optimization in Information Systems and Management Sciences, eds H. A. L. Thi, T. P. Dinh, and N. T. Nguyen (University of Lorraine, France: Springer), 169–180. doi: 10.1007/978-3-319-18161-5_15

CrossRef Full Text | Google Scholar

Lou, Y., and Yan, M. (2018). Fast L1-L2 minimization via a proximal operator. J. Sci. Comput. 74, 767–785. doi: 10.1007/s10915-017-0463-2

CrossRef Full Text | Google Scholar

Lou, Y., Yin, P., He, Q., and Xin, J. (2015b). Computing sparse representation in a highly coherent dictionary based on difference of L1 and L2. J. Sci. Comput. 64, 178–196. doi: 10.1007/s10915-014-9930-1

CrossRef Full Text | Google Scholar

Lou, Y., Zeng, T., Osher, S., and Xin, J. (2015c). A weighted difference of anisotropic and isotropic total variation model for image processing. SIAM J. Imaging Sci. 8, 1798–1823. doi: 10.1137/14098435X

CrossRef Full Text | Google Scholar

Luong, Q.-T. (1993). “Color in computer vision,” in Handbook of Pattern Recognition and Computer Vision, eds C. H. Chen, L. F. Pau, P. S. P. Wang (Singapore: World Scientific), 311–368. doi: 10.1142/9789814343138_0012

CrossRef Full Text | Google Scholar

Mumford, D. B., and Shah, J. (1989). Optimal approximations by piecewise smooth functions and associated variational problems. Commun. Pure Appl. Math. 42, 577–685. doi: 10.1002/cpa.3160420503

CrossRef Full Text | Google Scholar

Osher, S., and Sethian, J. A. (1988). Fronts propagating with curvature-dependent speed: algorithms based on Hamilton-Jacobi formulations. J. Comput. Phys. 79, 12–49. doi: 10.1016/0021-9991(88)90002-2

CrossRef Full Text | Google Scholar

Pang, Z.-F., Geng, M., Zhang, L., Zhou, Y., Zeng, T., Zheng, L., et al. (2023). Adaptive weighted curvature-based active contour for ultrasonic and 3T/5T MR image segmentation. Signal Process. 205:108881. doi: 10.1016/j.sigpro.2022.108881

CrossRef Full Text | Google Scholar

Park, F., Lou, Y., and Xin, J. (2016). “A weighted difference of anisotropic and isotropic total variation for relaxed Mumford-Shah image segmentation,” in 2016 IEEE International Conference on Image Processing (ICIP) (Phoenix, AZ), 4314–4318. doi: 10.1109/ICIP.2016.7533174

CrossRef Full Text | Google Scholar

Pock, T., Chambolle, A., Cremers, D., and Bischof, H. (2009). “A convex relaxation approach for computing minimal partitions,” in 2009 IEEE Conference on Computer Vision and Pattern Recognition (Miami, FL), 810–817. doi: 10.1109/CVPR.2009.5206604

CrossRef Full Text | Google Scholar

Potts, R. B. (1952). Some generalized order-disorder transformations. Math. Proc. Cambrid. Philos. Soc. 48, 106–109. doi: 10.1017/S0305004100027419

CrossRef Full Text | Google Scholar

Rahimi, Y., Wang, C., Dong, H., and Lou, Y. (2019). A scale-invariant approach for sparse signal recovery. SIAM J. Sci. Comput. 41, A3649–A3672. doi: 10.1137/18M123147X

CrossRef Full Text | Google Scholar

Rockafellar, R. T., and Wets, R. J.-B. (2009). Variational Analysis, Vol. 317. Berlin; Heidelberg: Springer Science & Business Media.

Google Scholar

Rudin, L. I., Osher, S., and Fatemi, E. (1992). Nonlinear total variation based noise removal algorithms. Phys. D 60, 259–268. doi: 10.1016/0167-2789(92)90242-F

CrossRef Full Text | Google Scholar

Staal, J., Abràmoff, M. D., Niemeijer, M., Viergever, M. A., and Van Ginneken, B. (2004). Ridge-based vessel segmentation in color images of the retina. IEEE Trans. Med. Imaging 23, 501–509. doi: 10.1109/TMI.2004.825627

PubMed Abstract | CrossRef Full Text | Google Scholar

Storath, M., and Weinmann, A. (2014). Fast partitioning of vector-valued images. SIAM J. Imaging Sci. 7, 1826–1852. doi: 10.1137/130950367

CrossRef Full Text | Google Scholar

Storath, M., Weinmann, A., and Demaret, L. (2014). Jump-sparse and sparse recovery using Potts functionals. IEEE Trans. Signal Process. 62, 3654–3666. doi: 10.1109/TSP.2014.2329263

CrossRef Full Text | Google Scholar

Tongbram, S., Shimray, B. A., Singh, L. S., and Dhanachandra, N. (2021). A novel image segmentation approach using FCM and whale optimization algorithm. J. Ambient. Intell. Human. Comput. 1–15. doi: 10.1007/s12652-020-02762-w

CrossRef Full Text | Google Scholar

Vardi, Y., Shepp, L. A., and Kaufman, L. (1985). A statistical model for positron emission tomography. J. Am. Stat. Assoc. 80, 8–20. doi: 10.1080/01621459.1985.10477119

PubMed Abstract | CrossRef Full Text | Google Scholar

Vese, L. A., and Chan, T. F. (2002). A multiphase level set framework for image segmentation using the Mumford and Shah model. Int. J. Comput. Vis. 50, 271–293. doi: 10.1023/A:1020874308076

CrossRef Full Text | Google Scholar

Wang, C., Tao, M., Chuah, C.-N., Nagy, J., and Lou, Y. (2022). Minimizing l1 over l2 norms on the gradient. Inverse Probl. 38:065011. doi: 10.1088/1361-6420/ac64fb

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, C., Tao, M., Nagy, J. G., and Lou, Y. (2021). Limited-angle CT reconstruction via the l1/l2 minimization. SIAM J. Imaging Sci. 14, 749–777. doi: 10.1137/20M1341490

CrossRef Full Text | Google Scholar

Wang, C., Yan, M., Rahimi, Y., and Lou, Y. (2020). Accelerated schemes for the l1/l2 minimization. IEEE Trans. Signal Process. 68, 2660–2669. doi: 10.1109/TSP.2020.2985298

CrossRef Full Text | Google Scholar

Wang, Y., Yang, J., Yin, W., and Zhang, Y. (2008). A new alternating minimization algorithm for total variation image reconstruction. SIAM J. Imaging Sci. 1, 248–272. doi: 10.1137/080724265

CrossRef Full Text | Google Scholar

Wang, Y., Yin, W., and Zeng, J. (2019). Global convergence of ADMM in nonconvex nonsmooth optimization. J. Sci. Comput. 78, 29–63. doi: 10.1007/s10915-018-0757-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Wen, Y., Chan, R. H., and Zeng, T. (2016). Primal-dual algorithms for total variation based image restoration under Poisson noise. Sci. China Math. 59, 141–160. doi: 10.1007/s11425-015-5079-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Wu, T., Gu, X., Wang, Y., and Zeng, T. (2021a). Adaptive total variation based image segmentation with semi-proximal alternating minimization. Signal Process. 183:108017. doi: 10.1016/j.sigpro.2021.108017

CrossRef Full Text | Google Scholar

Wu, T., Mao, Z., Li, Z., Zeng, Y., and Zeng, T. (2022a). Efficient color image segmentation via quaternion-based L1/L2 regularization. J. Sci. Comput. 93:9. doi: 10.1007/s10915-022-01970-0

CrossRef Full Text | Google Scholar

Wu, T., Shao, J., Gu, X., Ng, M. K., and Zeng, T. (2021b). Two-stage image segmentation based on nonconvex ℓ2−ℓp approximation and thresholding. Appl. Math. Comput. 403:126168. doi: 10.1016/j.amc.2021.126168

CrossRef Full Text | Google Scholar

Wu, T., Zhao, Y., Mao, Z., Shi, L., Li, Z., and Zeng, Y. (2022b). Image segmentation via Fischer-Burmeister total variation and thresholding. Adv. Appl. Math. Mech. 14, 960–988. doi: 10.4208/aamm.OA-2021-0126

CrossRef Full Text | Google Scholar

Xu, Y., Narayan, A., Tran, H., and Webster, C. G. (2021). Analysis of the ratio of l1 and l2 norms in compressed sensing. Appl. Comput. Harmon. Anal. 55, 486–511. doi: 10.1016/j.acha.2021.06.006

CrossRef Full Text | Google Scholar

Xu, Y., Yin, W., Wen, Z., and Zhang, Y. (2012). An alternating direction algorithm for matrix completion with nonnegative factors. Front. Math. China 7, 365–384. doi: 10.1007/s11464-012-0194-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Xu, Z., Chang, X., Xu, F., and Zhang, H. (2012). L1/2 regularization: a thresholding representation theory and a fast solver. IEEE Trans. Neural Netw. Learn. Syst. 23, 1013–1027. doi: 10.1109/TNNLS.2012.2197412

PubMed Abstract | CrossRef Full Text | Google Scholar

Yang, J., Guo, Z., Zhang, D., Wu, B., and Du, S. (2022). An anisotropic diffusion system with nonlinear time-delay structure tensor for image enhancement and segmentation. Comput. Math. Appl. 107, 29–44. doi: 10.1016/j.camwa.2021.12.005

CrossRef Full Text | Google Scholar

You, J., Jiao, Y., Lu, X., and Zeng, T. (2019). A nonconvex model with minimax concave penalty for image restoration. J. Sci. Comput. 78, 1063–1086. doi: 10.1007/s10915-018-0801-z

CrossRef Full Text | Google Scholar

Zhang, T., Chen, J., Wu, C., He, Z., Zeng, T., and Jin, Q. (2022). Edge adaptive hybrid regularization model for image deblurring. Inverse Probl. 38:065010. doi: 10.1088/1361-6420/ac60bf

CrossRef Full Text | Google Scholar

Zosso, D., An, J., Stevick, J., Takaki, N., Weiss, M., Slaughter, L. S., et al. (2017). Image segmentation with dynamic artifacts detection and bias correction. Inverse Probl. Imaging 11, 577–600. doi: 10.3934/ipi.2017027

CrossRef Full Text | Google Scholar

Keywords: image segmentation, total variation, nonconvex optimization, ADMM, Poisson noise

Citation: Bui K, Lou Y, Park F and Xin J (2023) Difference of anisotropic and isotropic TV for segmentation under blur and Poisson noise. Front. Comput. Sci. 5:1131317. doi: 10.3389/fcomp.2023.1131317

Received: 24 December 2022; Accepted: 03 May 2023;
Published: 28 June 2023.

Edited by:

You-wei Wen, Hunan Normal University, China

Reviewed by:

Saneera Hemantha Kulathilake, Rajarata University of Sri Lanka, Sri Lanka
Yide Zhang, California Institute of Technology, United States
Tingting Wu, Nanjing University of Posts and Telecommunications, China

Copyright © 2023 Bui, Lou, Park and Xin. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Kevin Bui, kevinb3@uci.edu

Download