Citation: Tú Nguyen-Dumont, Jenna Stewart, Ingrid Winship, Melissa C. Southey. Rare genetic variants: making the connection with breast cancer susceptibility[J]. AIMS Genetics, 2015, 2(4): 281-292. doi: 10.3934/genet.2015.4.281
[1] | Habib ur Rehman, Poom Kumam, Kanokwan Sitthithakerngkiet . Viscosity-type method for solving pseudomonotone equilibrium problems in a real Hilbert space with applications. AIMS Mathematics, 2021, 6(2): 1538-1560. doi: 10.3934/math.2021093 |
[2] | Francis Akutsah, Akindele Adebayo Mebawondu, Austine Efut Ofem, Reny George, Hossam A. Nabwey, Ojen Kumar Narain . Modified mildly inertial subgradient extragradient method for solving pseudomonotone equilibrium problems and nonexpansive fixed point problems. AIMS Mathematics, 2024, 9(7): 17276-17290. doi: 10.3934/math.2024839 |
[3] | Habib ur Rehman, Wiyada Kumam, Poom Kumam, Meshal Shutaywi . A new weak convergence non-monotonic self-adaptive iterative scheme for solving equilibrium problems. AIMS Mathematics, 2021, 6(6): 5612-5638. doi: 10.3934/math.2021332 |
[4] | Yasir Arfat, Muhammad Aqeel Ahmad Khan, Poom Kumam, Wiyada Kumam, Kanokwan Sitthithakerngkiet . Iterative solutions via some variants of extragradient approximants in Hilbert spaces. AIMS Mathematics, 2022, 7(8): 13910-13926. doi: 10.3934/math.2022768 |
[5] | Saud Fahad Aldosary, Mohammad Farid . A viscosity-based iterative method for solving split generalized equilibrium and fixed point problems of strict pseudo-contractions. AIMS Mathematics, 2025, 10(4): 8753-8776. doi: 10.3934/math.2025401 |
[6] | Lu-Chuan Ceng, Shih-Hsin Chen, Yeong-Cheng Liou, Tzu-Chien Yin . Modified inertial subgradient extragradient algorithms for generalized equilibria systems with constraints of variational inequalities and fixed points. AIMS Mathematics, 2024, 9(6): 13819-13842. doi: 10.3934/math.2024672 |
[7] | Yali Zhao, Qixin Dong, Xiaoqing Huang . A self-adaptive viscosity-type inertial algorithm for common solutions of generalized split variational inclusion and paramonotone equilibrium problem. AIMS Mathematics, 2025, 10(2): 4504-4523. doi: 10.3934/math.2025208 |
[8] | Austine Efut Ofem, Jacob Ashiwere Abuchu, Godwin Chidi Ugwunnadi, Hossam A. Nabwey, Abubakar Adamu, Ojen Kumar Narain . Double inertial steps extragadient-type methods for solving optimal control and image restoration problems. AIMS Mathematics, 2024, 9(5): 12870-12905. doi: 10.3934/math.2024629 |
[9] | Rose Maluleka, Godwin Chidi Ugwunnadi, Maggie Aphane . Inertial subgradient extragradient with projection method for solving variational inequality and fixed point problems. AIMS Mathematics, 2023, 8(12): 30102-30119. doi: 10.3934/math.20231539 |
[10] | Lu-Chuan Ceng, Li-Jun Zhu, Tzu-Chien Yin . Modified subgradient extragradient algorithms for systems of generalized equilibria with constraints. AIMS Mathematics, 2023, 8(2): 2961-2994. doi: 10.3934/math.2023154 |
We are interested in the symmetry and monotonicity of solutions to the problem
$ {−Δ1u=f(u),inΩ,u>0inΩ,u=0,on∂Ω, $
|
(1.1) |
where $ \Delta_1u = \text{div}(\frac{D u}{|D u|}) $, $ \Omega $ is a smooth bounded domain in $ \mathbb{R}^N $, $ N\geq 2 $, and strictly convex. The purpose of the paper is to investigate a priori estimates and symmetric properties of the solutions when the domain is assumed to have symmetric properties and $ f $ is supposed to satisfy the following conditions $ (H_1) $, $ (H_2) $ and $ (H_4) $. We also assume that $ f $ satisfies the following conditions $ (H_3) $ and $ (H_5) $ to use mountain pass lemma to get a nontrivial solution.
$ (H_1) $: $ f:[0, +\infty) $ is a locally Lipschitz continuous function and $ f(s)\geq 0 $ for $ \forall $ $ s\in [0, +\infty) $.
$ (H_2) $: $ f(s)\leq C_1(1+s^{1^\ast-1}) $, for $ \forall $ $ s\in [0, +\infty) $, with $ 1^\ast = \frac{N}{N-1} $ and a constant $ C_1 > 0 $.
$ (H_3) $: There exists $ \theta > 1 $, and $ k_0 > 0 $ such that
$ 0<θF(s)≤sf(s),s≥k0. $
|
$ (H_4) $: There exists a constant $ C_2 > 0 $ such that
$ lim infs→+∞1∗F(s)−sf(s)sf(s)≥C2, $
|
where $ F(s) = \int_0^sf(t)dt $.
$ (H_5) $: There exists a constant $ \alpha\in (0, \frac{1}{N-1}) $ such that
$ lims→0|f(s)|sα<∞. $
|
We point out that the similar $ p $-Laplace problems $ (p > 1) $ have many applications and have been studied for a long time, more precisely, Dirichlet problems for the $ p $-Laplace operator,
$ {−Δpu=f(u),inΩ,u>0inΩ,u=0,on∂Ω. $
|
(1.2) |
In the case $ p = 2 $, the problem (1.2) $ -\Delta_p u = f(u) $ has been widely studied. Gidas and Spruck [27] prove a priori bounds for nonlinearities $ f $ for $ N\geq3 $ behave as a subcritical power at infinity, introducing the blow up method together with Liouville type theorems for solutions in $ \mathbb{R}^N $. Figneiredo, Lions and Nussbaum [19] consider the existence and a priori estimates of positive solutions of the problem (1.2) when $ f $ satisfies the superlinear grow at infinity. They prove a priori bound for positive solutions of the problem (1.2) under the hypothesis $ \lim_{s\rightarrow \infty}\frac{f(s)}{s^{\frac{N+2}{N-2}}} = 0 $, together with the monotonic results by Gidas, Ni and Nirenberg [28] obtained by the Alexandrov-Serrin moving plane method [37]. The moving plane method has been improved and simplified by Beresticky and Nirenberg [7] with the aid of the maximum principle in small domain. With the help of the blow up procedure, Azizieh and Clément [5] prove a priori estimates for the problem (1.2) in the case of $ \Omega $ being a strictly convex domain and $ f $ satisfying some suitable assumption. Damascelli and Pacella [14,15] apply the moving plane method to prove some monotonic and symmetric results for the $ p $-Laplace equation in the singular case $ 1 < p < 2 $, also see [6,13]. The results are later extended to the case $ p > 2 $ in the papers [12,17,18]. Damascelli and Pardo [16] used the technique introduced in [19] that allowed to give the a priori estimates for solutions in case $ 1 < p < N $, case $ p = N $, and case $ p > N $. Esposito, Montoro and Sciunzi [24] study symmetric and monotonic properties of singular positive solutions to the problem (1.2) via moving plane method under suitable assumptions on $ f $. However, all the above mentioned papers can not deal with the case $ p = 1 $. In this paper, we can extend the case $ p > 1 $ to the case $ p = 1 $.
Obviously, the problem of $ \Delta_1 $ is different from $ \Delta_p $ ($ p > 1 $). When $ p = 1 $, it is necessary to replace $ W^{1, 1} $ by $ BV $, the space of functions of bounded variation. A function $ u\in L^1(\Omega) $ is called a function of bounded variation, whose partial derivatives in the sense of distribution are Radon measures. We point out that the space $ W^{1, p}(\Omega) $ is reflexive, however, the space $ BV(\Omega) $ is not reflexive, so that we can not follow the arguments on $ \Delta_p $. The 1-Laplace operator $ \Delta_1 $ introduces some extra difficulties and special features. The first difficulty occurs by defining the quotient $ \frac{Du}{|Du|} $, $ Du $ being just a Radon measure. To deal with the 1-Laplacian operator, we need the theory of pairing of $ L^\infty $ divergence measure vector fields (see the pioneering works [3,4,8]).
Demengel [21] is concerned with existence of solution in $ BV(\Omega) $ to the problem $ -\text{div}z+z\text{sign}u = f|u|^{1^\ast-2}u $ with $ z\cdot\nabla u = \nabla u $ in $ \Omega $ and $ -z\cdot \gamma = u $ on $ \partial\Omega $. Demengel [22] is devoted to the elliptic equations with 1-Laplacian operator
$ {−Δ1u=f(x,u),inΩ,u=0,on∂Ω, $
|
(1.3) |
and introduces the concept of locally almost 1-harmonic functions in $ \Omega $. The comparison principle, the first eigenvalue and related eigenfunctions for the 1-Laplacian operator are established in [22]. Kawohl and Schuricht [30] consider a number of problems that are associated with the 1-Laplace operator $ \Delta_1 $, the formal limit of the $ p $-Laplace operator as $ p\rightarrow1 $, by investigating the underlying variational problem. Since the corresponding solution typically belongs to $ BV $ and not to $ W^{1, 1} $, they have to study the minimizers of the functionals containing the total variation. In particular, they look for constrained minimizers subject to a prescribed $ L^1 $ norm which can be considered as an eigenvalue problem for the 1-Laplace operator. Degiovanni and Magrone [20] are concerned with the problem (1.3) with $ f(x, u) = \lambda\frac{u}{|u|}+|u|^{1^\ast-2}u $. It is proved that for every $ \lambda\geq \lambda_1 $, the problem (1.3) admits a nontrivial solution by the non-standard linking methods. Salas and Segura de León [35] study the problem (1.3) with $ f(x, u) $ satisfying subcritical growth; i.e., $ |f(x, u)|\leq C(1+|u|^{q}) $ with $ 0 < q < 1^{\ast}-1 $. They prove that for the problem (1.3) there exists at least two nontrivial solutions, one nonnegative and one nonpositive, by using known existence results for the $ p $-Laplacian $ (p > 1) $ and considering the limit as $ p\rightarrow1^+ $. De Cicco, Giachetti, Oliva and Petitta [9] study the existence and regularity of special distributional nonnegative solutions to the boundary value singular problem (1.3) with $ f(x, u) = h(u)g(x) $. They show existence of nonnegative solutions to (1.3) with $ u^{\max\{1, \gamma\}}\in BV(\Omega) $. These solutions are obtained as a limit as $ p\rightarrow1^+ $ of nonnegative solutions of the $ p $-Laplacian problems $ -\Delta_p u_p = h(u_p)g $ with $ u_p = 0 $ on $ \partial\Omega $. We also refer to [33,34,35,36,38] for the a priori estimates and gradient estimates of solutions. In this paper we can study the monotonicity and symmetry of positive solution to the 1-Laplace problem and show the a priori estimates for the solution.
By the theory of pairing of $ L^\infty $ divergence measure vector fields, we introduce the following definition of solutions to the problem (1.1).
Definition 1.1. We say that $ u\in BV_{loc}(\Omega) $, $ u > 0 $, is a solution to problem (1.1) if there exists a vector field $ z\in \mathcal{DM}^\infty(\Omega) $ with $ \|z\|_{L^\infty}\leq 1 $ such that
$ −divz=f(u),inD′(Ω), $
|
(1.4) |
$ (z,Du)=|Du|as measures inΩ, $
|
(1.5) |
$ [z,γ]∈sign(−u) on ∂Ω, $
|
(1.6) |
where $ \gamma $ is the unit exterior normal on $ \partial\Omega $, and the spaces $ BV_{loc}(\Omega) $ and $ \mathcal{DM}^\infty(\Omega) $ are given in Section 2.
To state more precisely some known result about the monotonicity and symmetry of solutions of the problem (1.1), we need some notations. Let $ \nu $ be a direction in $ \mathbb{R}^N $. For a real number $ \mu $ we define
$ Tνμ={x∈RN∣x⋅ν=μ} $
|
$ Ωνμ={x∈Ω∣x⋅ν<μ} $
|
$ xνμ=Rνμ(x)=x+2(μ−x⋅ν)ν,x∈RN $
|
and
$ a(ν)=infx∈Ωx⋅ν. $
|
(1.7) |
If $ \mu > a(\nu) $ then $ \Omega_\mu^\nu $ is nonempty, thus we set
$ (Ωνμ)′=Rνμ(Ωνμ). $
|
Following [6] and [12,13,14,15,16,17,18], we observe that $ \mu-a(\nu) $ small then $ (\Omega_\mu^\nu)' $ is contained in $ \Omega $ and will remain in it, at least until one of the following occurs:
$ (A) $ $ (\Omega_\mu^\nu)' $ becomes internally tangent to $ \partial\Omega $.
$ (B) $ $ T_\mu^\nu $ is orthogonal to $ \partial\Omega $.
Let $ \Pi_1(\nu) $ be the set of those $ \mu > a(\nu) $ such that for each $ \eta < \mu $ none of the conditions $ (A) $ and $ (B) $ holds and define
$ μ1(ν)=supΠ1(ν). $
|
(1.8) |
Moreover, let
$ Π2(ν)={μ>a(ν)∣(Ωνη)′⊂Ω,∀η∈(a(ν),μ]} $
|
and
$ μ2(ν)=supΠ2(ν). $
|
(1.9) |
Since $ \Omega $ is supposed to be smooth, note that neither $ \Pi_1(\nu) $ nor $ \Pi_2(\nu) $ are empty and $ \Pi_1(\nu)\subset \Pi_2(\nu) $, so that $ \mu_1(\nu)\leq\mu_2(\nu) $.
We deal with solutions to the problem (1.1) in the sense of Definition 1.1. Our main result is stated as follows.
Theorem 1.2. Let $ \Omega $ be a smooth bounded domain in $ \mathbb{R}^N $, $ N\geq2 $, which is strictly convex. Assume the nonlinearity $ f $ satisfies the conditions $ (H_1)-(H_5) $. Then there exists a nontrivial positive solution $ u $ to the problem (1.1) in the sense Definition 1.1, bounded in $ L^\infty(\Omega) $ (i.e., $ u\in L^\infty(\Omega) $), and for any direction $ \nu $ and for $ \mu $ in the interval $ (a(\nu), \mu_1(\nu)] $,
$ u(x)≤u(xνμ),a.e.x∈Ωνμ, $
|
(1.10) |
where $ a(\nu) $ and $ \mu_1(\nu) $ are given by (1.7) and (1.8) respectively.
If $ f $ is locally Lipschitz continuous in the closed interval $ [0, +\infty) $, the condition (1.10) holds for any $ \mu $ in the interval $ (a(\nu), \mu_2(\nu)] $.
Corollary 1.3. Let the smooth bounded domain $ \Omega\subset\mathbb{R}^N $, $ N\geq 2 $, be strictly convex with respect to a direction $ \nu $ and symmetric with respect to the hyperplane $ T_0^\nu = \{ x\in \mathbb{R}^N\mid x\cdot \nu = 0\} $. Assume that the nonlinearity $ f $ satisfies the conditions $ (H_1)-(H_5) $, which is locally Lipschitz continuous in the closed interval $ [0, +\infty) $ and strictly positive in $ (0, +\infty) $. Then there exists a nontrivial positive solution $ u $ to the problem (1.1) in the sense Definition 1.1, bounded in $ L^\infty(\Omega) $, almost everywhere symmetric, i.e., $ u(x) = u(x_0^\nu) $ and nondecreasing in the $ \nu $-direction a.e. in $ \Omega_0^\nu $.
Remark 1.4. Since the moving plane procedure can be performed in the same way but in the opposite direction, then it is obvious that Corollary 1.3 is obtained by Theorem 1.2 (see Corollary 2.4 of [16]).
Throughout this paper, $ \Omega $ denotes an bounded subset of $ \mathbb{R}^N $ with Lipschitz boundary. The symbol $ |\Omega| $ stands for its $ N $ dimensional Lebesgue measure and $ H^{N-1}(E) $ for the $ N-1 $ dimensional Hausdorff measure of a set $ E\subset \mathbb{R}^N $. An outward normal with vector $ \gamma = \gamma(x) $ is defined for $ H^{N-1} $ a.e. $ x\in \partial\Omega $. We will denote by $ W_0^{1, p}(\Omega) $ the usual Sobolev space, of measureable functions having weak gradient in $ L^p(\Omega; \mathbb{R}^N) $ and zero trace on $ \partial\Omega $. If $ 1 < p < N $, denote by $ p^\ast = \frac{Np}{N-p} $ its critical Sobolev exponent. $ BV(\Omega) $ will denote the space of functions of bounded variation
$ BV(Ω)={v∈L1(Ω)∣Dvis a bounded Radon measure} $
|
where $ Dv:\Omega\rightarrow \mathbb{R}^N $ is the distributional gradient of $ u $. It is endowed with the norm by
$ ‖v‖BV=∫Ω|Dv|+∫Ω|v|dx, $
|
where
$ ∫Ω|Dv|=sup{∫Ωvdivφdx∣φ∈C10(Ω;RN),|φ(x)|≤1,x∈Ω}. $
|
$ BV(\Omega) $ is a Banach space which is non-reflexive and non-separable. The notion of a trace on the boundary can be extended to functions $ v\in BV(\Omega) $ and this fact allows us to write $ v|_{\partial \Omega} $. Moreover, the trace defines a linear bounded operator $ i:BV(\Omega)\hookrightarrow L^1(\partial \Omega) $ which is onto. By the trace, we have an equivalent norm on $ BV(\Omega) $
$ ‖v‖BV=∫Ω|Dv|+∫∂Ω|v|dHN−1, $
|
where $ H^{N-1} $ denotes the $ N-1 $ dimensional Hausdorff measure. We will often use this norm in what follows. In addition, the following continuous embeddings hold
$ BV(Ω)↪Lm(Ω),1≤m≤NN−1, $
|
which are compact for $ 1\leq m < \frac{N}{N-1} $ (see for instance [25,41]). We denote by $ \mathcal{M}(\Omega) $ the space of Radon measures with finite total variation over $ \Omega $, by
$ DM∞(Ω)={z∈L∞(Ω;RN)∣divz∈M(Ω)} $
|
and by
$ DM∞loc(Ω)={z∈L∞(Ω;RN)∣divz∈M(Ω′),Ω′⊂⊂Ω}. $
|
The theory of $ L^\infty $ divergence measure vector fields is due to Anzellotti [4] and Chen and Frid [8]. We define the following distribution $ (z, Dv) $
$ ⟨(z,Dv),φ⟩=−∫Ωvφdivzdx−∫Ωvz⋅∇φdx $
|
(2.1) |
for $ \forall $ $ \varphi\in C_c^1(\Omega) $. In Anzellotti's theory we need some compatibility conditions, such as $ \text{div} z\in L^1(\Omega) $ and $ v\in BV(\Omega)\cap L^\infty (\Omega) $ or $ \text{div} z $ a Radon measure with finite total variation and $ v\in BV(\Omega)\cap L^\infty (\Omega)\cap C(\Omega) $.
Lemma 2.1 ([34,35]). Let $ v\in BV_{loc}(\Omega)\cap L^1(\Omega, \mu) $ and $ z\in \mathcal{DM}_{loc}^\infty (\Omega) $. Then the distribution $ (z, Dv) $ defined in (2.1) previously satisfies
$ |⟨(z,Dv),φ⟩|≤‖φ‖L∞‖z‖L∞(U)∫U|Dv|, $
|
for all open set $ U\subset\subset \Omega $ and all $ \varphi \in C_c^1(U) $.
Lemma 2.2 ([34,35]). The distribution $ (z, Dv) $ is a Radon measure. It and its total variation $ |(z, Dv)| $ are absolutely continuous with respect to the measure $ |Dv| $ and
$ |∫B(z,Dv)|≤∫B|(z,Dv)|≤‖z‖L∞(U)∫B|Dv|, $
|
holds for all Borel sets $ B $ and for all open sets $ U $ such that $ B\subset U\subset \Omega $.
Lemma 2.3 ([10,11,34]). Let $ z\in \mathcal{DM}_{loc}^\infty (\Omega) $ and let $ v\in BV(\Omega)\cap L^\infty (\Omega) $. Then $ zv\in \mathcal{DM}_{loc}^\infty (\Omega) $. Moreover, the following formula holds in the sense of measures
$ div(z,v)=(divz)v+(z,Dv). $
|
It follows from Anzellotti's theory that every $ z\in \mathcal{DM}^\infty (\Omega) $ has a weak trace on $ \partial \Omega $ of the normal component of $ z $ which is denoted by $ [z, \gamma] $ with $ \gamma $ the unit exterior normal on $ \partial\Omega $, which satisfies
$ ‖[z,γ]‖L∞(∂Ω)≤‖z‖L∞, $
|
and
$ v[z,γ]=[vz,γ] $
|
for all $ z\in \mathcal{DM}^\infty (\Omega) $ and $ v\in BV(\Omega)\cap L^\infty(\Omega) $.
Lemma 2.4 (Green formula [10,11,34]). Let $ z\in \mathcal{DM}_{loc}^\infty (\Omega) $, $ \varpi = \text{div} z $ and $ v\in BV(\Omega) $ and assume $ v \in L^1(\Omega, \mu) $. Then $ vz\in \mathcal{DM}^\infty (\Omega) $ and the following holds
$ ∫Ωvdϖ+∫Ω(z,Dv)=∫∂Ω[vz,γ]dHN−1. $
|
Lemma 2.5 ([34,35]). Let $ z\in \mathcal{DM}_{loc}^\infty (\Omega) $ and $ v\in BV(\Omega)\cap L^\infty(\Omega) $. If $ vz\in \mathcal{DM}^\infty (\Omega) $, then
$ |[vz,γ]|≤|v|∂Ω‖z‖L∞(Ω),HN−1a.e.on∂Ω. $
|
Let $ p_0: = \min\{\theta, \frac{N}{N-1}\} $, with $ \theta > 1 $ given by $ (H_3) $. For each $ 1 < p < p_0 $, let us consider the following problem
$ {−Δpw=f(w),inΩ,w>0inΩ,w=0,on∂Ω, $
|
(3.1) |
where $ \Omega $ is a bounded smooth domain in $ \mathbb{R}^N $, $ N\geq 2 $, $ 1 < p < p_0 $ and $ f:[0, +\infty)\rightarrow \mathbb{R} $ satisfies the conditions $ (H_1)-(H_5) $. We need the following propositions and a priori estimates of $ p $-Laplace equation to prove Theorem 1.2.
Definition 3.1. We say $ u_p\in W_0^{1, p}(\Omega) $, $ u_p\geq0 $, is a weak solution to the problem (3.1) in the sense that
$ ∫Ω|∇up|p−2∇up⋅∇φdx=∫Ωf(up)φdx, $
|
(3.2) |
for $ \forall $ $ \varphi\in W_0^{1, p}(\Omega) $.
If $ u_p\in W^{1, p}(\Omega) $ is a weak solution of the problem (3.1) with $ f $ satisfying the critical growth, then $ u_p\in C^{1, \alpha}(\Omega) $ with $ \alpha\in (0, 1) $ (see [23,31,40]), so that we suppose from the beginning a $ C^1 $ regularity for the solution. Next, we recall some results on the monotonicity and estimates of solutions for the $ p $-Laplace equation. One can refer to [1,16,19,29,32] for the proof of the following Proposition 3.2-3.7.
Proposition 3.2 ([16]). Let $ \Omega $ be a smooth bounded domain in $ \mathbb{R}^N $, $ N\geq2 $, $ 1 < p < \infty $, $ f:[0, \infty)\rightarrow \mathbb{R} $ a continuous function which is locally Lipschitz continuous in $ (0, \infty) $ and strictly positive in $ (0, \infty) $ if $ p > 2 $. Let $ w\in C^1(\overline{\Omega}) $ be a weak solution of (3.1). Then for any direction $ \nu $ and for $ \mu $ in the interval $ (a(\nu), \mu_1(\nu)] $, we have
$ w(x)≤w(xνμ),a.e.x∈Ωνμ. $
|
(3.3) |
If $ f $ is locally Lipschitz continuous in the closed interval $ [0, +\infty) $, then (3.3) holds for any $ \mu $ in the interval $ (a(\nu), \mu_2(\nu)] $, where $ a(\nu) $, $ \mu_1(\nu) $ and $ \mu_2(\nu) $ are given by (1.7), (1.8) and (1.9).
Proposition 3.3 ([16,19]). Let $ \Omega $ be a strictly convex bounded smooth domain, and define $ \Omega_\delta = \{x\in \Omega\mid \text{dist}(x, \partial \Omega) > \delta\} $, for $ \delta > 0 $. Then the following result holds for a weak solution $ w\in C^1(\Omega) $ of the problem (3.1) with $ f $ satisfying the condition $ (H_1) $
$ {∃σ,ε>0depending only onΩ, such that ∀x∈Ω∖Ωεthereis a part of a cone Ixwith(i)w(ξ)≥w(x),∀ξ∈Ix,(ii)Ix⊂Ωε2,(iii)|Ix|≥σ. $
|
$ I_x $ is a part of a cone $ K_x $ with vertex in $ x $, where all the $ K_x $ are congruent to a fixed cone $ K $, and if $ x\in \Omega\setminus \Omega_\frac{\varepsilon}{2} $, then $ I_x = K_x\cap \Omega_\frac{\varepsilon}{2} $.
Proposition 3.4 ([32]). Let us define
$ λ1=infw∈W1,p00(Ω){∫Ω|∇w|p0dx∣∫Ω|w|p0dx=1}, with p0=min{θ,NN−1}>1, $
|
where $ \theta $ is given by $ (H_3) $. Then, $ \lambda_1 $ is the first eigenvalue of the operator $ -\Delta_{p_0} $ ($ \lambda_1\leq \lambda $ for any eigenvalue $ \lambda $), it is simple, i.e., there is only an eigenfunction up to multiplication by a constant, and it is isolated. Moreover a first eigenfunction does not change sign in $ \Omega $ and by the strong maximum principle it is in fact either strictly positive or strictly negative in $ \Omega $. So we can select a unique eigenfunction $ \phi_1 $ such that
$ ∫Ωϕp01dx=1, and ϕ1>0 in Ω. $
|
The following extension of the Picone's identity for the $ p $-Laplacian has been proved in [1].
Proposition 3.5 (Picone's identity [1]). Let $ v_1, v_2\geq0 $ be differentiable functions in an open set $ \Omega $, with $ v_2 > 0 $ and $ p > 1 $. Set
$ L(v1,v2)=|∇v1|p+(p−1)vp1vp2|∇v2|p−pvp−11vp−12|∇v2|p−2∇v1⋅∇v2 $
|
and
$ R(v1,v2)=|∇v1|p−|∇v2|p−2∇(vp1vp−12)⋅∇v2. $
|
Then $ R(v_1, v_2) = L(v_1, v_2)\geq 0 $.
As a consequence we have
$ |∇v2|p−2∇(vp1vp−12)⋅∇v2≤|∇v1|p. $
|
The following extension of the Pohozaev's identity for the $ p $-Laplacian has been given by [29].
Proposition 3.6 (Pohozaev's identity for $ p $-Laplace [29]). Let $ w\in W_0^{1, p}(\Omega)\cap L^\infty(\Omega) $, $ p > 1 $, be a weak solution of the problem
$ {−Δpw=f(w),inΩ,w=0,on∂Ω, $
|
where $ \Omega $ is a bounded smooth domain in $ \mathbb{R}^N $, $ N\geq2 $ and $ f:[0, +\infty)\rightarrow \mathbb{R} $ is a continuous function. Denote $ F(w) = \int_0^w f(s)ds $. Then
$ N∫ΩF(w)dx−N−pp∫Ωf(w)wdx=p−1p∫∂Ω|∂w∂γ|p(x⋅γ)dHN−1, $
|
where $ \gamma $ is the unit exterior normal on $ \partial \Omega $.
We need also local $ W^{1, \infty}(\Omega) $ result at the boundary. This result follows from the global estimates by Lieberman [31] extending the local interior estimates by Dibenedetto [23].
Proposition 3.7 ([16]). Let $ \Omega $ be a smooth bounded domain in $ \mathbb{R}^N $, $ N\geq2 $, and $ w\in C^1(\overline{\Omega}) $ be a solution of the problem
$ {−Δpw=h,inΩ,w>0inΩ,w=0,on∂Ω, $
|
with $ h\in L^{(p^\ast)'}(\Omega) $. For $ \delta > 0 $, let $ \Omega_\delta = \{x\in \Omega \mid \text{dist}(x, \partial \Omega) > \delta\} $ and suppose that $ w, h\in L^\infty (\Omega\setminus \Omega_\delta) $ with
$ ‖h‖L∞(Ω∖Ωδ)≤M and ‖w‖L∞(Ω∖Ωδ)≤M. $
|
Then there exists a constant $ C > 0 $ only depending on $ M $ and $ \delta $ such that
$ ‖∇w‖L∞(∂Ω)≤C. $
|
Next, we will give the estimate of the solution for the problem (3.1).
Theorem 3.8. If $ u_p $ is a weak solution to the problem (3.1) and $ f $ satisfies the conditions $ (H_2)-(H_4) $, then $ u_p $ satisfies
$ ‖up‖W1,p0(Ω)≤C′, $
|
(3.4) |
where the constant $ C' > 0 $ is not dependent on $ p $.
Proof. By $ 1 < p < p_0 $, Proposition 3.4, Proposition 3.5 with $ v_2 = u_p $, $ v_1 = \phi_1 $ and Young's inequality, we have
$ ∫Ωf(up)up−1pϕp1dx=∫Ω−div(|∇up|p−2∇up)ϕp1up−1pdx=∫Ω|∇up|p−2∇up⋅∇(ϕp1up−1p)dx≤∫Ω|∇ϕ1|pdx≤pp0∫Ω|∇ϕ1|p0dx+p0−pp0|Ω|≤∫Ω|∇ϕ1|p0dx+|Ω|≤λ1+|Ω|. $
|
(3.5) |
By the condition $ (H_3) $, there exists a constant $ C_3 > 0 $ such that
$ sθ−1≤C3f(s), for s≥k1, $
|
that is
$ sθ−p≤C3f(s)sp−1, for s≥k1, $
|
(3.6) |
where $ k_1 = \max\{k_0, 1\} $ and $ k_0 $ is given by $ (H_3) $.
Indeed, from $ (H_3) $, it holds
$ θt≤f(t)F(t), for t≥k0. $
|
(3.7) |
Setting $ k_1 = \max\{k_0, 1\} $ and integrating the above inequality (3.7) with respect to $ t $ on the interval $ [k_1, s] $, one has
$ θlnsk1≤lnF(s)F(k1), for s≥k1. $
|
That is
$ F(s)≥F(k1)(sk1)θ, for s≥k1. $
|
(3.8) |
Setting $ C_3: = \frac{k_1^\theta}{\theta F(k_1)} $ in (3.8), we get
$ F(s)≥sθθC3, for s≥k1. $
|
(3.9) |
Considering (3.9) and $ sf(s)\geq \theta F(s) $, for $ s\geq k_1 $, one gets the inequality (3.6).
Now, taking into account (3.5), (3.6) with $ s = u_p $ and Young's inequality, we get
$ ∫Ωuθ−ppϕp1dx=∫{0≤up≤k1}uθ−ppϕp1dx+∫{up>k1}uθ−ppϕp1dx≤kθ−p1∫{0≤up≤k1}ϕp1dx+C3∫{up>k1}f(up)up−1pϕp1dx≤kθ−p1∫Ωϕp1dx+C3∫{up>k1}f(up)up−1pϕp1dx=kθ−p1∫Ωϕp1dx+C3∫Ωf(up)up−1pϕp1dx−C3∫{0<up≤k1}f(up)up−1pϕp1dx≤kθ−p1∫Ωϕp1dx+(λ1+|Ω|)C3∫Ωϕp1dx≤(kθ−p1+(λ1+|Ω|)C3)∫Ωϕp1dx≤(kθ−p1+(λ1+|Ω|)C3)(pp0∫Ωϕp01dx+p0−pp0|Ω|)≤(kτ1+(λ1+|Ω|)C3)(|Ω|+1):=C4, $
|
(3.10) |
where $ \lambda_1+|\Omega| $ is given by (3.5) and $ -C_3\int_{\{ 0 < u_p\leq k_0\}}\frac{f(u_p)}{u_p^{p-1}}\phi_1^p dx\leq 0 $ is given by the condition $ (H_1) $ ($ f(s)\geq0 $, for all $ s\geq0 $) respectively, and the last inequality is given by Proposition 3.4 with $ \int_\Omega \phi_1^{p_0}dx = 1 $ and $ k_1 = \max\{k_0, 1\}\geq 1 $. By Proposition 3.3 and (3.10), for any $ x\in \Omega\setminus\Omega_\delta $, we have that
$ σ(infx∈Ωδ2ϕp1)[up(x)]θ−p≤∫Ix[up(y)]θ−pϕp1(y)dy≤∫Ω[up(y)]θ−pϕp1(y)dy≤C4, $
|
i.e.,
$ up(x)≤(C4σinfx∈Ωδ2ϕp1)1θ−p=(C4σ(infx∈Ωδ2ϕ1)p)1θ−p=(C4σ)1θ−p(infx∈Ωδ2ϕ1)−pθ−p≤(C4σ+1)1θ−p0[(infx∈Ωδ2ϕ1)−1θ−1+(infx∈Ωδ2ϕ1)−p0θ−p0]:=C5, $
|
(3.11) |
where the constant $ C_5 $ may be depend on $ C_4 $, $ \sigma $, $ \theta $, $ p_0 $ and $ \phi_1 $ by (3.11), but are independent of $ p $. Estimate (3.11) gives the uniform $ L^\infty $ bounds near the boundary: $ \exists $ $ \delta > 0 $ and $ C_5 > 0 $ such that
$ ‖up‖L∞(Ω∖Ωδ)≤C5, $
|
(3.12) |
for $ \forall $ $ u_p\in W_0^{1, p}(\Omega) $ satisfying the problem (3.1). On the other hand, from the condition $ (H_2) $ and (3.12), we have
$ ‖f(up)‖L∞(Ω∖Ωδ)≤C1(1+‖up‖1N−1L∞(Ω∖Ωδ))≤C6. $
|
(3.13) |
It is clear that $ f(u_p(\cdot))\in L^{(p^\ast)'} $ is given by the condition $ (H_2) $ and Sobolev embedding. By Proposition 3.7, (3.12) and (3.13), we get
$ ‖∂up∂γ‖L∞(∂Ω)≤C7, $
|
(3.14) |
where the constant $ C_7 > 0 $ is only depending on $ C_5 $, $ C_6 $ and $ \delta $. By Proposition 3.6 (Pohozaev's identity)
$ p∗∫ΩF(s)dx−∫Ωf(up)updx=p−1N−p∫∂Ω|∂up∂γ|p(x⋅γ)dHN−1, $
|
$ p^\ast = \frac{Np}{N-p} > \frac{N}{N-1} = 1^\ast $ and $ (H_4) $, there exists a large enough constant $ k_2 > 0 $ such that
$ f(s)s≤C2(1∗F(s)−f(s)s)≤C2(p∗F(s)−f(s)s) $
|
(3.15) |
as $ s\geq k_2 $, so that by the condition $ (H_2) $ and taking $ s = u_p $ in (3.15)
$ ∫Ω|∇up|pdx=∫Ωf(up)updx=∫{0<up≤k2}f(up)updx+∫{up>k2}f(up)updx≤C1k2(1+k1N−12)|Ω|+C2p−1N−pC7|∂Ω|≤C1k2(1+k1N−12)|Ω|+C2p0−1N−p0C7|∂Ω|:=C8. $
|
(3.16) |
That is
$ ‖up‖W1,p0(Ω)≤C1p8≤C8+1:=C′. $
|
From the definitions of $ C_5 $, $ C_6 $, $ C_7 $ and $ C_8 $, i.e., (3.11)-(3.14) and (3.16), we obtain that the constant $ C' $ is not dependent on $ p $. The proof of Theorem 3.8 is completed.
The following existence result holds.
Theorem 3.9. Let $ f $ satisfy the conditions $ (H_1) $, $ (H_2) $, $ (H_3) $ and $ (H_5) $. Then there exists a nontrivial positive solution $ u_p $ to the problem (3.1).
Proof. By the conditions $ (H_1) $, $ (H_2) $, $ (H_3) $ and $ (H_5) $, it is well known that there exists a nontrivial solution $ u_p\geq0 $ to the problem (3.1). The positive solution $ u_p $ is obtained using the mountain pass lemma by Ambrosetti and Rabinowitz [2] for the following truncated functional $ J_p^+: W_0^{1, p}(\Omega)\rightarrow \mathbb{R} $ given by
$ J+p(w)=1p∫Ω|∇w|pdx−∫ΩF+(w)dx, $
|
(3.17) |
where $ F_+(s) = \int_0^sf_+(t)dt $ and
$ f+(s)={f(s),s≥0,0,s<0. $
|
(3.18) |
We claim that $ J_p^+ $ satisfies the structure of mountain pass lemma and the $ (P-S) $ condition.
Indeed, by the condition $ (H_5) $, $ 0 $ is a local minimum of $ J_p^+ $. From the condition $ (H_3) $, there exist two constants $ \widetilde{C} $, $ \widehat{C} > 0 $, such that
$ F+(s)≥˜Csθ−ˆC, $
|
for all $ s\in [0, +\infty) $ with $ \theta > 1 $. This implies that
$ J+p(w)≤1p‖w‖pW1,p0−˜C‖w‖θLθ+ˆC|Ω|, $
|
(3.19) |
for $ \forall $ $ w\in W_0^{1, p}(\Omega) $. We can choose a $ w_0\in W_0^{1, p}(\Omega) $ and $ \|w_0\|_{W_0^{1, p}} = 1 $ such that
$ J+p(tw0)≤tpp−˜Ctθ‖w0‖θLθ+ˆC|Ω|→−∞, $
|
as $ t\rightarrow +\infty $, with $ 1 < p < p_0: = \min\{\theta, \frac{N}{N-1}\} $. Whence there exists a large number $ t_0 > 0 $ such that
$ J+p(t0w0)<0. $
|
(3.20) |
We set $ e: = t_0w_0\in W_0^{1, p}(\Omega) $. Since $ (H_2) $ and the embedding $ W_0^{1, p}(\Omega)\hookrightarrow L^{1^\ast}(\Omega) $, $ 1^\ast = \frac{N}{N-1} < \frac{Np}{N-p} = p^\ast $, is compact, we obtain that $ f_+ $ satisfies the subcritical grow, i.e.,
$ |f+(s)|≤C1(1+s1∗−1), with 1∗<p∗. $
|
(3.21) |
Considering (3.21) and $ (H_3) $, $ J_p^+ $ satisfies the $ (P-S) $ condition.
In this section we prove our main results concerning the case $ p = 1 $, namely Theorem 1.2. Under the same assumption of Theorem 1.2, we divide the proof into few steps.
Step 1. Existence of a solution $ u $ and a field $ z $.
Step 2. $ (z, Du) = |Du| $ as measures in $ \Omega $.
Step 3. $ [z, \gamma]\in $ sign$ (-u) $ on $ \partial\Omega $.
Step 4. The monotonicity of solution $ u $.
Step 5. $ u\in L^\infty (\Omega) $.
Step 6. $ u $ is nontrivial.
Step 1. Existence of a solution $ u $ for the problem (1.1) and existence of a field $ z\in \mathcal{DM}^\infty(\Omega) $ satisfying (1.4) and $ \|z\|_{L^\infty}\leq1 $.
Proof of Step 1: From Theorem 3.8, we obtain that $ u_p $ is bounded in $ W_0^{1, p}(\Omega)\hookrightarrow L^m(\Omega) $, with $ 1\leq m\leq \frac{N}{N-1} < p^\ast = \frac{Np}{N-p} $, $ 1 < p < p_0 < 2\leq N $.
$ up→u strongly in Lm(Ω), $
|
(4.1) |
$ up(x)→u(x) a.e. x∈Ω, $
|
(4.2) |
$ ∃g∈Lm(Ω), such that |up(x)|≤g(x), $
|
(4.3) |
as $ p\rightarrow 1^+ $.
Next, we will show that there exists a vector field $ z $ satisfying (1.4). Recalling Theorem 3.8, we obtain that $ \{u_p\} $ is bounded in $ W_0^{1, p}(\Omega)\subset BV(\Omega) $. So that for $ 1\leq r < p' = \frac{p}{p-1} $, we have
$ ∫Ω||∇up|p−2∇up|rdx=∫Ω|∇up|r(p−1)dx≤(∫Ω|∇up|pdx)rp′|Ω|1−rp′, $
|
and thus
$ ‖|∇up|p−2∇up‖Lr(Ω)≤C1p′8|Ω|1r−1p′, $
|
(4.4) |
where the constant $ C_8 $ is given by (3.16). This implies that $ |\nabla u_p|^{p-2}\nabla u_p $ is bounded in $ L^r(\Omega; \mathbb{R}^N) $ with respect to $ p $. Then there exists $ z_r\in L^r(\Omega; \mathbb{R}^N) $ such that
$ |∇up|p−2∇up⇀zr,weaklyinLr(Ω;RN), $
|
(4.5) |
as $ p\rightarrow 1^+ $. A standard diagonal argument shows that there exists a unique vector field $ z $ which is defined on $ \Omega $ independently of $ r $, such that
$ |∇up|p−2∇up⇀z,weaklyinLr(Ω;RN), $
|
(4.6) |
as $ p\rightarrow 1^+ $. By applying the semicontinuity of the $ L^r $ norm the previous inequality (4.4) implies
$ ‖z‖Lr(Ω)≤lim infp→1+‖|∇up|p−2∇up‖Lr≤|Ω|1r,∀r<∞, $
|
so that, letting $ r\rightarrow \infty $ we have $ z\in L^\infty(\Omega; \mathbb{R}^N) $ and
$ ‖z‖L∞(Ω;RN)≤1. $
|
Using $ \varphi \in C_c^1(\Omega) $ with $ \varphi\geq0 $ as a test function in (3.1), we have
$ ∫Ω|∇up|p−1∇up∇φdx=∫Ωf(up)φdx. $
|
(4.7) |
Taking $ p\rightarrow 1^+ $ in the left hand side of (4.7) and by (4.6), we get
$ limp→1+∫Ω|∇up|p−1∇up∇φdx=∫Ωz⋅∇φdx, $
|
(4.8) |
for $ \forall $ $ \varphi \in C_c^1(\Omega) $. On the other hand, thanks to (4.2) and $ f(s) $ a locally Lipschitz continuous function, we have
$ f(up(x))→f(u(x)), a.e. x∈Ω. $
|
Moreover, we deduce from $ (H_2) $ and (4.3) that
$ |f(up(⋅))|≤C1(1+|up(⋅)|1N−1)≤C1(1+|g(⋅)|1N−1)∈LN(Ω). $
|
Consequently, by the Dominated Convergence Theorem, we get
$ limp→1+∫Ωf(up(x))φ(x)dx=∫Ωf(u(x))φ(x)dx, $
|
(4.9) |
for $ \forall $ $ \varphi\in C_c^1(\Omega) $. Therefore, (4.7), (4.8) and (4.9) imply that
$ −divz=f(u) in D′(Ω). $
|
(4.10) |
Step 2. $ (z, Du) = |Du| $ as measures in $ \Omega $.
Before proving $ (z, Du) = |Du| $, we need the following lemma for which one can refer to [9].
Lemma 4.1 ([9]). Under the same assumptions of Theorem 1.2, the following identity holds
$ −∫Ωuφdivzdx=∫Ωf(u)uφdx, $
|
(4.11) |
for $ \forall $ $ \varphi \in C_c^1(\Omega) $.
Proof of Step 2: We take $ u_p\varphi \in W_0^{1, p}(\Omega) $ as a test function in (3.1) with $ 0\leq \varphi \in C_c^1(\Omega) $, $ \max_{x\in\Omega}|\varphi(x)| = M_0 $ and get
$ ∫Ω|∇up|pφdx+∫Ωup|∇up|p−2∇up⋅∇φdx=∫Ωf(up)upφdx. $
|
(4.12) |
By Young's inequality and Fatou's Lemma, we estimate the first integral term in (4.12)
$ ∫Ω|Du|φdx≤lim infp→1+∫Ω|∇up|φdx≤lim infp→1+[1p∫Ω|∇up|pφdx+p−1p∫Ωφdx]=lim infp→1+∫Ω|∇up|pφdx $
|
(4.13) |
On the other hand, by (4.6) we have
$ limp→1+∫Ωup|∇up|p−2∇up⋅∇φdx=∫Ωuz⋅∇φdx. $
|
(4.14) |
From
$ |f(up)upφ|≤M0C1|up|(1+|up|1N−1)≤M0C1|g(⋅)|(1+|g(⋅)|1N−1)∈L1(Ω), $
|
and the Dominated Convergence Theorem, we obtain the right hand side of (4.12) is as follows
$ limp→1+∫Ωf(up)upφdx=∫Ωf(u)uφdx. $
|
(4.15) |
From (4.12)-(4.15), we have
$ ∫Ω|Du|φdx+∫Ωuz⋅∇φdx≤∫Ωf(u)uφdx. $
|
(4.16) |
By (4.16) and Lemma 4.1, we also have
$ ∫Ω|Du|φdx+∫Ωuz⋅∇φdx≤−∫Ωuφdivzdx. $
|
Therefore, by (2.1), we get
$ ∫Ω|Du|φdx≤−∫Ωuz⋅∇φdx−∫Ωuφdivzdx=∫Ω(z,Du)φdx. $
|
The arbitrariness of $ \varphi $ implies that
$ |Du|≤(z,Du) $
|
as measures in $ \Omega $. On the other hand, since $ \|z\|_{L^\infty}\leq1 $, and
$ (z,Du)≤‖z‖L∞|Du|≤|Du| $
|
as measures in $ \Omega $, we have
$ |Du|=(z,Du). $
|
Step 3. The boundary condition $ [z, \gamma]\in $ sign$ (-u) $ on $ \partial\Omega $.
Proof of Step 3: It is easy to check that this fact is equivalent to show
$ ∫∂Ω(|u|+u[z,γ])dHN−1=0. $
|
(4.17) |
Choosing $ u_p $ as a test function in (3.1), we have
$ ∫Ω|∇up|pdx=∫Ωf(up)updx. $
|
Since $ u_p\in W_0^{1, p}(\Omega) $ is bounded, by the fact that $ u_p = 0 $ on $ \partial \Omega $ and Young's inequality, we get
$ ∫Ω|∇up|dx+∫∂Ω|up|dHN−1≤1p∫Ω|∇up|pdx+p−1p|Ω|=1p∫Ωf(up)updx+p−1p|Ω|. $
|
(4.18) |
We use the lower semicontinuity (4.18) to pass to the limit as $ p\rightarrow 1^+ $ and obtain
$ ∫Ω|Du|dx+∫∂Ω|u|dHN−1≤lim infp→1+(∫Ω|∇up|dx+∫∂Ω|up|dHN−1)≤lim infp→1+[1p∫Ωf(up)updx+p−1p|Ω|]=∫Ωf(u)udx, $
|
(4.19) |
where the last equality is given by the Dominated Convergence Theorem and
$ |f(up)up|≤C1|up|(1+|up|1N−1)≤C1|g(⋅)|(1+|g(⋅)|1N−1)∈L1(Ω). $
|
Furthermore, by Lemma 2.3 and Lemma 2.4, we have
$ ∫Ωf(u)udx=−∫Ωudivzdx=∫Ω(z,Du)dx−∫∂Ωu[z,γ]dHN−1. $
|
(4.20) |
From $ (z, Du) = |Du| $, (4.19) and (4.20), we have
$ ∫∂Ω(|u|+u[z,γ])dHN−1≤0. $
|
(4.21) |
The inequality (4.21) and $ |u|\geq |u|\|z\|_{L^\infty}\geq |u[z, \gamma]|\geq -u[z, \gamma] $ give the desired equality (4.17) and we conclude that
$ [z,γ]∈ sign(−u) on ∂Ω. $
|
Step 4. The monotonicity of the solution $ u $ of problem (1.1).
Proof of Step 4: By Proposition 3.2, we obtain $ u_p $ satisfies the following result. For any direction $ \nu $ and $ \mu $ in the interval $ (a(\nu), \mu_1(\nu)] $, then
$ up(x)≤up(xνμ),∀x∈Ωνμ, $
|
(4.22) |
where $ a(\nu) $ and $ \mu_1(\nu) $ are given by (1.7) and (1.8). Considering this fact and $ u_p(x)\rightarrow u(x) $ a.e. in $ \Omega $, taking $ p\rightarrow 1^+ $ in (4.22), we have
$ u(x)≤u(xνμ), a.e. x∈Ωνμ. $
|
(4.23) |
We get the result of monotonicity for the solution $ u $. Inequality (4.23) also holds for any $ \mu\in (a(\nu), \mu_2(\nu)] $ by Proposition 3.2, if $ f $ is locally Lipschitz continuous, and $ a(\nu) $ and $ \mu_2(\nu) $ are given by (1.7) and (1.9).
Step 5. The boundedness of the solution $ u $, i.e., $ u\in L^\infty(\Omega) $.
Before proving $ u\in L^\infty(\Omega) $, we need to prove the following lemma.
Lemma 4.2. For every $ \varepsilon > 0 $ there exists $ k_3 > 0 $ which does not depend on $ p $, such that
$ ∫Ak(1+u1N−1p)Ndx<ε $
|
(4.24) |
for every $ k\geq k_3 $ and $ \forall $ $ p\in (1, p_0) $, with $ A_k = \{ x\in\Omega \mid u_p(x) > k\} $.
Proof of Lemma 4.2: Using Sobolev embedding $ W_0^{1, p}(\Omega)\subset BV(\Omega)\hookrightarrow L^\frac{N}{N-1}(\Omega) $, Theorem 3.8 and Holder's inequality, we obtain that
$ |Ak|N−1N≤1k(∫AkuNN−1pdx)N−1N≤1kS1∫Ak|∇up|dx≤S1k|Ak|p−1p(∫Ak|∇up|pdx)1p≤S1k|Ω|p−1pC1p8≤S1k(1+|Ω|)(C8+1), $
|
(4.25) |
where $ S_1 $ is given by the best Sobolev constant
$ S1={Γ(1+N2)}1N√πN, $
|
see [26,39], and $ |A_k| $ stands for its $ N $ dimensional Lebesgue measure. Inequality (4.25) implies that $ \lim_{k\rightarrow \infty}|A_k| = 0 $. It holds that for $ \forall $ $ \varepsilon > 0 $, there exists a large number $ k_4 > 0 $ such that
$ |Ak|<ε2N, for all k≥k4. $
|
(4.26) |
On the other hand, by Sobolev embedding $ u_p\in W_0^{1, p}(\Omega)\subset BV(\Omega)\hookrightarrow L^\frac{N}{N-1}(\Omega) $, Theorem 3.8 and (4.3), we get
$ up∈LNN−1(Ω) $
|
and
$ 0≤∫Ak|up(x)|NN−1dx≤∫Ak|g(x)|NN−1dx, $
|
(4.27) |
which implies that $ u_p(x) < \infty $ a.e. in $ \Omega $. Considering (4.27), $ \lim_{k\rightarrow \infty}|A_k| = 0 $ and by absolute continuity of integrable function, we have
$ limk→∞∫Ak|up(x)|NN−1dx≤limk→∞∫Ak|g(x)|NN−1dx=0. $
|
(4.28) |
From (4.28), for $ \forall $ $ \varepsilon > 0 $, $ \exists $ $ k_5 > 0 $ large enough (not depend on $ p $) and $ \delta > 0 $ small enough such that as $ k\geq k_5 $, we have $ |A_k| < \delta $ and
$ ∫Ak|up(x)|NN−1dx≤∫Ak|g(x)|NN−1dx<ε2N, $
|
(4.29) |
From (4.26) and (4.29), we obtain
$ ∫Ak(1+u1N−1p)Ndx≤2N−1(|Ak|+∫Ak|up(x)|NN−1dx)≤2N−1(ε2N+ε2N)=ε, $
|
for all $ k\geq k_3: = \max\{ k_4, k_5\} $. The proof of Lemma 4.2 is completed.
Proof of Step 5: Next, we would like to use Stampacchia truncation [38] to prove the boundedness of the positive solution $ u $. For every $ k > 0 $, we define the auxiliary function $ G_k:[0, \infty)\rightarrow \mathbb{R} $ as
$ Gk(s)={s−k,s>k,0,0<s≤k. $
|
(4.30) |
Then, choosing $ G_k(u_p) $ as a test function in (3.1), we get
$ ∫Ω|∇Gk(up)|pdx=∫Ωf(up)Gk(up)dx. $
|
(4.31) |
By (4.31), $ (H_2) $, Sobolev embedding, Young's inequality and Holder's inequality, we have
$ (∫ΩGk(up)NN−1dx)N−1N≤S1∫Ω|∇Gk(up)|dx≤S1p∫Ω|∇Gk(up)|pdx+S1(p−1)p|Ω|=S1p∫Ωf(up)Gk(up)dx+S1(p−1)p|Ω|≤S1pC1∫Ω(1+u1N−1p)Gk(up)dx+S1(p−1)p|Ω|≤S1C1[∫Ak(1+u1N−1p)Ndx]1N(∫AkGk(up)NN−1dx)N−1N+S1(p−1)p|Ω|. $
|
(4.32) |
By Lemma 4.2 and taking $ \varepsilon = \frac{1}{(2C_1S_1)^N} $, there exists $ k_3 > 0 $ which does not depend on $ p $, such that
$ ∫Ak(1+u1N−1p)Ndx<1(2C1S1)N, $
|
(4.33) |
for all $ k\geq k_3 $ and $ p\in (1, p_0) $. Consequently, from (4.32) and (4.33) we obtain
$ 0≤∫ΩGk(up)NN−1dx≤[2S1(p−1)|Ω|p]NN−1. $
|
(4.34) |
Since $ u_p(x)\rightarrow u(x) $ a.e. $ x\in \Omega $ and Fatou's Lemma, we can pass to the limit on $ p\rightarrow 1^+ $ in (4.34), to conclude that
$ ∫Ω(u(x)−k)NN−1dx=0, $
|
for $ \forall $ $ k\geq k_3 > 0 $. Thus $ u\in L^\infty(\Omega) $.
Step 6. $ u $ is nontrivial.
Proof of Step 6: For $ \forall $ $ v\in BV(\Omega) $, we define the functional $ J^+:BV(\Omega)\rightarrow \mathbb{R} $ as
$ J+(v)=∫Ω|Dv|+∫∂Ω|v|dHN−1−∫ΩF+(v)dx, $
|
where $ F_+(s) = \int_0^s f_+(t)dt $ and $ f_+ $ is given by (3.18).
We will say that $ v_0\in BV(\Omega) $ is a critical point of $ J^+ $ if there exists $ z\in \mathcal{DM}^\infty(\Omega) $ with $ \|z\|_{L^\infty}\leq 1 $ such that
$ −∫Ωφdivzdx=∫Ωf(v0)φdx, for all φ∈C1c(Ω), $
|
$ (z,Dv0)=|Dv0|as measures inΩ, $
|
$ [z,γ]∈sign(−v0) on ∂Ω, $
|
where $ \gamma $ is the unit exterior normal on $ \partial\Omega $. The critical points of $ J^+ $ coincide with solutions to the problem (1.1) in the sense Definition 1.1, for which one can refer to [9] or [35].
We shall show that $ 0 $ is a local minimum of $ J^+ $.
Indeed, by the condition $ (H_5) $, there exists small enough $ \delta > 0 $ such that
$ |f(s)|≤C9|s|α, $
|
for $ \forall $ $ |s|\in (0, \delta) $ and for some constant $ C_9 > 0 $ with $ \alpha\in (0, \frac{1}{N-1}) $. Moreover, by the definition of $ F_+(s) $, we have
$ F+(s)=∫s0f+(t)dt≤∫s0|f(t)|dt≤C91+α|s|1+α $
|
(4.35) |
for $ \forall $ $ |s|\in (0, \delta) $. By (4.35) and the norm $ \|v\|_{BV} = \int_\Omega |Dv|+ \int_{\partial\Omega} |v|dH^{N-1} $, $ v\in BV(\Omega) $, it holds
$ J+(v)=‖v‖BV−∫ΩF+(v)dx≥‖v‖BV−C91+α∫Ω|v|1+αdx≥‖v‖BV−C10‖v‖1+αBV, $
|
where the last inequality is given by the embedding $ BV(\Omega)\hookrightarrow L^{1+\alpha}(\Omega) $, $ \alpha\in (0, \frac{1}{N-1}) $. Choosing a positive constant $ \rho < \min\{\delta, (\frac{1}{2C_{10}})^\frac{1}{\alpha}\} $, we obtain
$ J+(v)≥12‖v‖BV>0, $
|
(4.36) |
for $ \forall $ $ v\in BV(\Omega) $ and $ \|v\|_{BV}\leq \rho $. This implies that $ 0 $ is a local minimum of $ J^+ $.
Now, we introduce the auxiliary functional
$ Ip(w)=J+p(w)+p−1p|Ω|, $
|
(4.37) |
where $ J_p^+ $ is given by (3.17). By Young's inequality and (4.18), we can fix $ p\in (1, p_0) $ and obtain
$ Ip(w)=J+p(w)+p−1p|Ω|=1p∫Ω|∇w|pdx−∫ΩF+(w)dx+p−1p|Ω|≥∫Ω|∇w|dx+∫∂Ω|w|dHN−1−∫ΩF+(w)dx=J+(w), $
|
(4.38) |
for $ \forall $ $ w\in W_0^{1, p}(\Omega) \subset BV(\Omega) $, with $ p_0 = \min\{\theta, \frac{N}{N-1}\} $. From (4.38) and (3.19), one gets
$ J+(w)≤J+p(w)+p−1p|Ω|≤1p‖w‖pW1,p0−˜C‖w‖θLθ+(p−1p+ˆC)|Ω|, $
|
(4.39) |
for all $ w\in W_0^{1, p}(\Omega) $ with $ p < p_0 < \theta $. Recalling the structure of mountain pass lemma in Theorem 3.9, we can deduce that there exists $ e = t_0 w_0\in W_0^{1, p}(\Omega)\subset BV(\Omega) $ and $ \|e\|_{BV} > \rho $ such that $ J(e) < 0 $ by (3.20).
Obviously, the critical points of $ I_p $ are identical with the critical points of $ J_p^+ $. Then $ u_p $ given by Theorem 3.9 is a critical point of $ J_p^+ $, and also a critical point of $ I_p $, which implies that the critical point $ u_p $ satisfies
$ Ip(up)=infη∈Γpmaxt∈[0,1]Ip(η(t)), $
|
(4.40) |
where $ \Gamma_p = \{\eta\in C([0, 1], W_0^{1, p}(\Omega)) \mid \eta(0) = 0, \eta(1) = e\} $. Considering any path $ \eta \in \Gamma_p $ and the continuity of the map $ t\rightarrow I_p(\eta(t)) $, there exists $ t_0 > 0 $ such that $ \|\eta (t_0)\|_{BV} = \rho $. From (4.36), (4.38), (4.40) and $ \|\eta (t_0)\|_{BV} = \rho $, we obtain that
$ Ip(up)=infη∈Γpmaxt∈[0,1]Ip(η(t))≥ρ2. $
|
(4.41) |
On the other hand, choosing $ u_p $ as a test function in (3.1), by the Dominated Convergence Theorem, (4.2) and (4.20), we have
$ limp→1+1p∫Ω|∇up|pdx=limp→1+1p∫Ωf(up)updx=∫Ωf(u)udx=∫Ω(z,Du)−∫∂Ωu[z,γ]dHN−1=∫Ω|Du|+∫∂Ω|u|dHN−1, $
|
(4.42) |
where the last equality is given by Step 2 and Step 3. From $ (H_2) $, (4.2) and (4.3), we can apply the Dominated Convergence Theorem to obtain
$ limp→1+∫ΩF+(up)dx=∫ΩF+(u)dx. $
|
(4.43) |
By (4.37), (4.42) and (4.43), we can get
$ limp→1+Ip(up)=limp→1+[J+p(up)+p−1p|Ω|]=limp→1+J+p(up)=J+(u). $
|
(4.44) |
Summarizing (4.41) and (4.44) we obtain that
$ J+(u)≥ρ2>0, $
|
with $ 0 < \rho < \min\{ \delta, (\frac{1}{2C_{10}})^\frac{1}{\alpha}\} $, and then $ u $ is nontrivial, because $ J^+(0) = 0 $.
The proof of Theorem 1.2 is completed.
The author sincerely thanks the editors and reviewers for their valuable suggestions and useful comments. This work was supported by the Natural Science Foundation of Jiangsu Province of China (BK20180638).
For the publication of this article, no conflict of interest among the authors is disclosed.
[1] |
Easton DF, Pharoah PD, Antoniou AC, et al. (2015) Gene-panel sequencing and the prediction of breast-cancer risk. N Engl J Med 372: 2243-2257. doi: 10.1056/NEJMsr1501341
![]() |
[2] |
Desmond A, Kurian AW, Gabree M, et al. (2015) Clinical Actionability of Multigene Panel Testing for Hereditary Breast and Ovarian Cancer Risk Assessment. JAMA Oncol 1: 943-951. doi: 10.1001/jamaoncol.2015.2690
![]() |
[3] |
Campuzano O, Sarquella-Brugada G, Mademont-Soler I, et al. (2014) Identification of Genetic Alterations, as Causative Genetic Defects in Long QT Syndrome, Using Next Generation Sequencing Technology. PLoS One 9: e114894. doi: 10.1371/journal.pone.0114894
![]() |
[4] |
Kapoor NS, Curcio LD, Blakemore CA, et al. (2015) Multigene Panel Testing Detects Equal Rates of Pathogenic BRCA1/2 Mutations and has a Higher Diagnostic Yield Compared to Limited BRCA1/2 Analysis Alone in Patients at Risk for Hereditary Breast Cancer. Ann Surg Oncol 22: 3282-3288. doi: 10.1245/s10434-015-4754-2
![]() |
[5] |
Antoniou A, Pharoah PDP, Narod S, et al. (2003) Average risks of breast and ovarian cancer associated with BRCA1 or BRCA2 mutations detected in case Series unselected for family history: a combined analysis of 22 studies. Am J Hum Genet 72: 1117-1130. doi: 10.1086/375033
![]() |
[6] |
Lovelock PK, Spurdle AB, Mok MT, et al. (2007) Identification of BRCA1 missense substitutions that confer partial functional activity: potential moderate risk variants? Breast Cancer Res 9: R82. doi: 10.1186/bcr1826
![]() |
[7] |
Spurdle AB, Whiley PJ, Thompson B, et al. (2012) BRCA1 R1699Q variant displaying ambiguous functional abrogation confers intermediate breast and ovarian cancer risk. J Med Genet 49: 525-532. doi: 10.1136/jmedgenet-2012-101037
![]() |
[8] |
Michailidou K, Hall P, Gonzalez-Neira A, et al. (2013) Large-scale genotyping identifies 41 new loci associated with breast cancer risk. Nat Genet 45: 353-361, 361e351-352. doi: 10.1038/ng.2563
![]() |
[9] | Meeks HD, Song H, Michailidou K, et al. (2016) BRCA2 Polymorphic Stop Codon K3326X and the Risk of Breast, Prostate, and Ovarian Cancers. J Natl Cancer Inst 108. |
[10] |
Antoniou AC, Casadei S, Heikkinen T, et al. (2014) Breast-cancer risk in families with mutations in PALB2. N Engl J Med 371: 497-506. doi: 10.1056/NEJMoa1400382
![]() |
[11] |
Southey MC, Teo ZL, Dowty JG, et al. (2010) A PALB2 mutation associated with high risk of breast cancer. Breast Cancer Res 12: R109. doi: 10.1186/bcr2796
![]() |
[12] | Southey MC, Teo ZL, Winship I (2013) PALB2 and breast cancer: ready for clinical translation! Appl Clin Genet 6: 43-52. |
[13] |
Teo ZL, Park DJ, Provenzano E, et al. (2013) Prevalence of PALB2 mutations in Australasian multiple-case breast cancer families. Breast Cancer Res 15: R17. doi: 10.1186/bcr3392
![]() |
[14] | Teo ZL, Sawyer SD, James PA, et al. (2013) The incidence of PALB2 c.3113G>A in women with a strong family history of breast and ovarian cancer attending familial cancer centres in Australia. Fam Cancer 12: 587-595. |
[15] |
Wong MW, Nordfors C, Mossman D, et al. (2011) BRIP1, PALB2, and RAD51C mutation analysis reveals their relative importance as genetic susceptibility factors for breast cancer. Breast Cancer Res Treat 127: 853-859. doi: 10.1007/s10549-011-1443-0
![]() |
[16] | Dansonka-Mieszkowska A, Kluska A, Moes J, et al. (2010) A novel germline PALB2 deletion in Polish breast and ovarian cancer patients. BMC Med Genet 11: 20. |
[17] |
Foulkes WD, Ghadirian P, Akbari MR, et al. (2007) Identification of a novel truncating PALB2 mutation and analysis of its contribution to early-onset breast cancer in French-Canadian women. Breast Cancer Res 9: R83. doi: 10.1186/bcr1828
![]() |
[18] |
Erkko H, Xia B, Nikkila J, et al. (2007) A recurrent mutation in PALB2 in Finnish cancer families. Nature 446: 316-319. doi: 10.1038/nature05609
![]() |
[19] |
Tischkowitz M, Capanu M, Sabbaghian N, et al. (2012) Rare germline mutations in PALB2 and breast cancer risk: a population-based study. Hum Mutat 33: 674-680. doi: 10.1002/humu.22022
![]() |
[20] |
Plon SE, Eccles DM, Easton D, et al. (2008) Sequence variant classification and reporting: recommendations for improving the interpretation of cancer susceptibility genetic test results. Hum Mutat 29: 1282-1291. doi: 10.1002/humu.20880
![]() |
[21] | Scott CL, Jenkins MA, Southey MC, et al. (2003) Average age-specific cumulative risk of breast cancer according to type and site of germline mutations in BRCA1 and BRCA2 estimated from multiple-case breast cancer families attending Australian family cancer clinics. Hum Genet 112: 542-551. |
[22] |
Tavtigian SV, Oefner PJ, Babikyan D, et al. (2009) Rare, evolutionarily unlikely missense substitutions in ATM confer increased risk of breast cancer. Am J Hum Genet 85: 427-446. doi: 10.1016/j.ajhg.2009.08.018
![]() |
[23] |
Le Calvez-Kelm F, Lesueur F, Damiola F, et al. (2011) Rare, evolutionarily unlikely missense substitutions in CHEK2 contribute to breast cancer susceptibility: results from a breast cancer family registry case-control mutation-screening study. Breast cancer research : BCR 13: R6. doi: 10.1186/bcr2810
![]() |
[24] | Thusberg J, Vihinen M (2009) Pathogenic or not? And if so, then how? Studying the effects of missense mutations using bioinformatics methods. Hum Mutat 30: 703-714. |
[25] | Zuckerkandl E (1965) [Remarks on the evolution of polynucleotides compared to that of polypeptides]. Bull Soc Chim Biol (Paris) 47: 1729-1730. |
[26] |
Jukes TH, King JL (1971) Deleterious mutations and neutral substitutions. Nature 231: 114-115. doi: 10.1038/231114a0
![]() |
[27] |
Jordan DM, Ramensky VE, Sunyaev SR (2010) Human allelic variation: perspective from protein function, structure, and evolution. Curr Opin Struct Biol 20: 342-350. doi: 10.1016/j.sbi.2010.03.006
![]() |
[28] |
Ng PC, Henikoff S (2001) Predicting deleterious amino acid substitutions. Genome Res 11: 863-874. doi: 10.1101/gr.176601
![]() |
[29] |
Ferrer-Costa C, Gelpi JL, Zamakola L, et al. (2005) PMUT: a web-based tool for the annotation of pathological mutations on proteins. Bioinformatics 21: 3176-3178. doi: 10.1093/bioinformatics/bti486
![]() |
[30] |
Chun S, Fay JC (2009) Identification of deleterious mutations within three human genomes. Genome Res 19: 1553-1561. doi: 10.1101/gr.092619.109
![]() |
[31] |
Adzhubei IA, Schmidt S, Peshkin L, et al. (2010) A method and server for predicting damaging missense mutations. Nat Methods 7: 248-249. doi: 10.1038/nmeth0410-248
![]() |
[32] | Tavtigian SV, Deffenbaugh AM, Yin L, et al. (2006) Comprehensive statistical study of 452 BRCA1 missense substitutions with classification of eight recurrent substitutions as neutral. J Med Genet 43: 295-305. |
[33] |
Bromberg Y, Rost B (2007) SNAP: predict effect of non-synonymous polymorphisms on function. Nucleic Acids Res 35: 3823-3835. doi: 10.1093/nar/gkm238
![]() |
[34] |
Goldgar DE, Easton DF, Deffenbaugh AM, et al. (2004) Integrated evaluation of DNA sequence variants of unknown clinical significance: application to BRCA1 and BRCA2. Am J Hum Genet 75: 535-544. doi: 10.1086/424388
![]() |
[35] |
Easton DF, Deffenbaugh AM, Pruss D, et al. (2007) A systematic genetic assessment of 1,433 sequence variants of unknown clinical significance in the BRCA1 and BRCA2 breast cancer-predisposition genes. Am J Hum Genet 81: 873-883. doi: 10.1086/521032
![]() |
[36] |
Goldgar DE, Easton DF, Byrnes GB, et al. (2008) Genetic evidence and integration of various data sources for classifying uncertain variants into a single model. Hum Mutat 29: 1265-1272. doi: 10.1002/humu.20897
![]() |
[37] |
Spurdle AB, Lakhani SR, Healey S, et al. (2008) Clinical classification of BRCA1 and BRCA2 DNA sequence variants: the value of cytokeratin profiles and evolutionary analysis--a report from the kConFab Investigators. J Clin Oncol 26: 1657-1663. doi: 10.1200/JCO.2007.13.2779
![]() |
[38] |
Spurdle AB, Healey S, Devereau A, et al. (2012) ENIGMA--evidence-based network for the interpretation of germline mutant alleles: an international initiative to evaluate risk and clinical significance associated with sequence variation in BRCA1 and BRCA2 genes. Hum Mutat 33: 2-7. doi: 10.1002/humu.21628
![]() |
[39] |
Vallee MP, Francy TC, Judkins MK, et al. (2012) Classification of missense substitutions in the BRCA genes: a database dedicated to Ex-UVs. Hum Mutat 33: 22-28. doi: 10.1002/humu.21629
![]() |
[40] |
Couch FJ, Rasmussen LJ, Hofstra R, et al. (2008) Assessment of functional effects of unclassified genetic variants. Hum Mutat 29: 1314-1326. doi: 10.1002/humu.20899
![]() |
[41] | Wu K, Hinson SR, Ohashi A, et al. (2005) Functional evaluation and cancer risk assessment of BRCA2 unclassified variants. Cancer Res 65: 417-426. |
[42] |
Mitui M, Nahas SA, Du LT, et al. (2009) Functional and computational assessment of missense variants in the ataxia-telangiectasia mutated (ATM) gene: mutations with increased cancer risk. Hum Mutat 30: 12-21. doi: 10.1002/humu.20805
![]() |
[43] |
Roeb W, Higgins J, King MC (2012) Response to DNA damage of CHEK2 missense mutations in familial breast cancer. Hum Mol Genet 21: 2738-2744. doi: 10.1093/hmg/dds101
![]() |
[44] |
Kato S, Han SY, Liu W, et al. (2003) Understanding the function-structure and function-mutation relationships of p53 tumor suppressor protein by high-resolution missense mutation analysis. Proc Natl Acad Sci U S A 100: 8424-8429. doi: 10.1073/pnas.1431692100
![]() |
[45] |
Iversen ES, Jr., Couch FJ, Goldgar DE, et al. (2011) A computational method to classify variants of uncertain significance using functional assay data with application to BRCA1. Cancer Epidemiol Biomarkers Prev 20: 1078-1088. doi: 10.1158/1055-9965.EPI-10-1214
![]() |
[46] | Guidugli L, Pankratz VS, Singh N, et al. (2013) A classification model for BRCA2 DNA binding domain missense variants based on homology-directed repair activity. Cancer Res 73: 265-275. |
[47] |
Rahman N (2014) Mainstreaming genetic testing of cancer predisposition genes. Clin Med 14: 436-439. doi: 10.7861/clinmedicine.14-4-436
![]() |
[48] | http://apps.ccge.medschl.cam.ac.uk/consortia/bcac/index.html. |
[49] |
French JD, Ghoussaini M, Edwards SL, et al. (2013) Functional variants at the 11q13 risk locus for breast cancer regulate cyclin D1 expression through long-range enhancers. Am J Hum Genet 92: 489-503. doi: 10.1016/j.ajhg.2013.01.002
![]() |
[50] |
Goldgar DE, Healey S, Dowty JG, et al. (2011) Rare variants in the ATM gene and risk of breast cancer. Breast Cancer Res 13: R73. doi: 10.1186/bcr2919
![]() |
[51] |
Gorringe KL, Choong DY, Visvader JE, et al. (2008) BARD1 variants are not associated with breast cancer risk in Australian familial breast cancer. Breast Cancer Res Treat 111: 505-509. doi: 10.1007/s10549-007-9799-x
![]() |
[52] |
Seal S, Thompson D, Renwick A, et al. (2006) Truncating mutations in the Fanconi anemia J gene BRIP1 are low-penetrance breast cancer susceptibility alleles. Nat Genet 38: 1239-1241. doi: 10.1038/ng1902
![]() |
[53] |
Schrader KA, Masciari S, Boyd N, et al. (2011) Germline mutations in CDH1 are infrequent in women with early-onset or familial lobular breast cancers. J Med Genet 48: 64-68. doi: 10.1136/jmg.2010.079814
![]() |
[54] |
Bogdanova N, Schurmann P, Waltes R, et al. (2008) NBS1 variant I171V and breast cancer risk. Breast Cancer Res Treat 112: 75-79. doi: 10.1007/s10549-007-9820-4
![]() |
[55] |
Tommiska J, Seal S, Renwick A, et al. (2006) Evaluation of RAD50 in familial breast cancer predisposition. Int J Cancer 118: 2911-2916. doi: 10.1002/ijc.21738
![]() |
[56] |
Rennert G, Lejbkowicz F, Cohen I, et al. (2012) MutYH mutation carriers have increased breast cancer risk. Cancer 118: 1989-1993. doi: 10.1002/cncr.26506
![]() |
[57] |
Sharif S, Moran A, Huson SM, et al. (2007) Women with neurofibromatosis 1 are at a moderately increased risk of developing breast cancer and should be considered for early screening. J Med Genet 44: 481-484. doi: 10.1136/jmg.2007.049346
![]() |
[58] |
Pradella LM, Evangelisti C, Ligorio C, et al. (2014) A novel deleterious PTEN mutation in a patient with early-onset bilateral breast cancer. BMC Cancer 14: 70. doi: 10.1186/1471-2407-14-70
![]() |
[59] |
Figer A, Kaplan A, Frydman M, et al. (2002) Germline mutations in the PTEN gene in Israeli patients with Bannayan-Riley-Ruvalcaba syndrome and women with familial breast cancer. Clin Genet 62: 298-302. doi: 10.1034/j.1399-0004.2002.620407.x
![]() |
[60] |
Meindl A, Hellebrand H, Wiek C, et al. (2010) Germline mutations in breast and ovarian cancer pedigrees establish RAD51C as a human cancer susceptibility gene. Nat Genet 42: 410-414. doi: 10.1038/ng.569
![]() |
[61] |
Loveday C, Turnbull C, Ruark E, et al. (2012) Germline RAD51C mutations confer susceptibility to ovarian cancer. Nat Genet 44: 475-476; author reply 476. doi: 10.1038/ng.2224
![]() |
[62] | Boardman LA, Couch FJ, Burgart LJ, et al. (2000) Genetic heterogeneity in Peutz-Jeghers syndrome. Hum Mutat 16: 23-30. |
[63] |
Evans DG, Birch JM, Thorneycroft M, et al. (2002) Low rate of TP53 germline mutations in breast cancer/sarcoma families not fulfilling classical criteria for Li-Fraumeni syndrome. J Med Genet 39: 941-944. doi: 10.1136/jmg.39.12.941
![]() |
[64] |
Mouchawar J, Korch C, Byers T, et al. (2010) Population-based estimate of the contribution of TP53 mutations to subgroups of early-onset breast cancer: Australian Breast Cancer Family Study. Cancer Res 70: 4795-4800. doi: 10.1158/0008-5472.CAN-09-0851
![]() |
[65] |
Park DJ, Lesueur F, Nguyen-Dumont T, et al. (2012) Rare mutations in XRCC2 increase the risk of breast cancer. Am J Hum Genet 90: 734-739. doi: 10.1016/j.ajhg.2012.02.027
![]() |
[66] |
Park DJ, Tao K, Le Calvez-Kelm F, et al. (2014) Rare mutations in RINT1 predispose carriers to breast and Lynch syndrome-spectrum cancers. Cancer Discov 4: 804-815. doi: 10.1158/2159-8290.CD-14-0212
![]() |
[67] |
Kiiski JI, Pelttari LM, Khan S, et al. (2014) Exome sequencing identifies FANCM as a susceptibility gene for triple-negative breast cancer. Proc Natl Acad Sci U S A 111: 15172-15177. doi: 10.1073/pnas.1407909111
![]() |
1. | E. Azroul, A. Benkirane, M. Shimi, M. Srati, On a class of nonlocal problems in new fractional Musielak-Sobolev spaces, 2022, 101, 0003-6811, 1933, 10.1080/00036811.2020.1789601 | |
2. | Yu Cheng, Zhanbing Bai, Existence results of non-local integro-differential problem with singularity under a new fractional Musielak–Sobolev space* , 2025, 58, 1751-8113, 045205, 10.1088/1751-8121/adaa3e |