This paper deals with the existence of weak solutions for semilinear elliptic equation with nonlinearity on the boundary. We establish the existence of a maximal and a minimal weak solution between an ordered pair of sub- and supersolution for both monotone and nonmonotone nonlinearities. We use iteration argument when the nonlinearity is monotone. For the nonmonotone case, we utilize the surjectivity of a pseudomonotone and coercive operator, Zorn's lemma and a version of Kato's inequality.
Citation: S. Bandyopadhyay, M. Chhetri, B. B. Delgado, N. Mavinga, R. Pardo. Maximal and minimal weak solutions for elliptic problems with nonlinearity on the boundary[J]. Electronic Research Archive, 2022, 30(6): 2121-2137. doi: 10.3934/era.2022107
[1] | Abdeljabbar Ghanmi, Hadeel Z. Alzumi, Noureddine Zeddini . A sub-super solution method to continuous weak solutions for a semilinear elliptic boundary value problems on bounded and unbounded domains. Electronic Research Archive, 2024, 32(6): 3742-3757. doi: 10.3934/era.2024170 |
[2] | Yaojun Ye, Qianqian Zhu . Existence and nonexistence of global solutions for logarithmic hyperbolic equation. Electronic Research Archive, 2022, 30(3): 1035-1051. doi: 10.3934/era.2022054 |
[3] | Yijun Chen, Yaning Xie . A kernel-free boundary integral method for reaction-diffusion equations. Electronic Research Archive, 2025, 33(2): 556-581. doi: 10.3934/era.2025026 |
[4] | Yan Dong . Local Hölder continuity of nonnegative weak solutions of inverse variation-inequality problems of non-divergence type. Electronic Research Archive, 2024, 32(1): 473-485. doi: 10.3934/era.2024023 |
[5] | Limin Guo, Weihua Wang, Cheng Li, Jingbo Zhao, Dandan Min . Existence results for a class of nonlinear singular $ p $-Laplacian Hadamard fractional differential equations. Electronic Research Archive, 2024, 32(2): 928-944. doi: 10.3934/era.2024045 |
[6] | Youngmok Jeon, Dongwook Shin . Immersed hybrid difference methods for elliptic boundary value problems by artificial interface conditions. Electronic Research Archive, 2021, 29(5): 3361-3382. doi: 10.3934/era.2021043 |
[7] | Guanrong Li, Yanping Chen, Yunqing Huang . A hybridized weak Galerkin finite element scheme for general second-order elliptic problems. Electronic Research Archive, 2020, 28(2): 821-836. doi: 10.3934/era.2020042 |
[8] | Lin Shen, Shu Wang, Yongxin Wang . The well-posedness and regularity of a rotating blades equation. Electronic Research Archive, 2020, 28(2): 691-719. doi: 10.3934/era.2020036 |
[9] | Li-ming Xiao, Cao Luo, Jie Liu . Global existence of weak solutions to a class of higher-order nonlinear evolution equations. Electronic Research Archive, 2024, 32(9): 5357-5376. doi: 10.3934/era.2024248 |
[10] | Hui Jian, Min Gong, Meixia Cai . Global existence, blow-up and mass concentration for the inhomogeneous nonlinear Schrödinger equation with inverse-square potential. Electronic Research Archive, 2023, 31(12): 7427-7451. doi: 10.3934/era.2023375 |
This paper deals with the existence of weak solutions for semilinear elliptic equation with nonlinearity on the boundary. We establish the existence of a maximal and a minimal weak solution between an ordered pair of sub- and supersolution for both monotone and nonmonotone nonlinearities. We use iteration argument when the nonlinearity is monotone. For the nonmonotone case, we utilize the surjectivity of a pseudomonotone and coercive operator, Zorn's lemma and a version of Kato's inequality.
We consider an elliptic equation with nonlinear boundary condition of the form
{−Δu+u=0inΩ;∂u∂η=f(x,u)on∂Ω, | (1.1) |
where Ω⊂RN(N≥2) is a bounded domain with C2,α (0<α<1) boundary ∂Ω, and ∂/∂η:=η(x)⋅∇ denotes the outer normal derivative on the boundary ∂Ω. Here f:∂Ω×R→R is a Carathéodory function, that is, f(⋅,u) is measurable for each u and f(x,⋅) is continuous for a.e. x∈∂Ω.
In this paper, we investigate the existence of maximal and minimal weak solutions (to be clarified later) between an ordered pair of sub- and supersolution of (1.1) for both monotone and nonmonotone nonlinearities. We use monotone iteration procedure when the nonlinearity is monotone. The nonmonotone case required a careful use of the surjectivity of a bounded, pseudomonotone and coercive operator, Zorn's lemma and a version of Kato's inequality up to the boundary. This proof, for the nonmonotone case, is motivated by the works in [1] and [2].
Elliptic equations with nonlinear boundary conditions have attracted a lot of attention over the last decades, see for instance [3,4,5,6,7,8,9] and references therein. Motivation to study equations with nonlinear boundary conditions stems from the fact that, when the reaction near the boundary depends on the density itself, linear boundary conditions (Dirichlet, Neumann, or Robin) are often inadequate to study chemical, biological, or ecological processes, see [10,11,12,13] and references therein, for specific applications.
The existence of a solution between an ordered pair of sub- and supersolution of elliptic boundary value problems has been studied extensively. For the linear boundary conditions, the sub–supersolution method for classical solutions were developed in [14,15,16] to study the solvability of quasi-linear and semi-linear equations using monotone iteration method. This method also yields the existence of a maximal and a minimal solution. These iterative methods can be thought as a generalization of the Perron arguments on sub- and superharmonic functions for existence of solutions of the boundary value problem. For relatively recent results on the existence of maximal and minimal solutions, for the linear boundary conditions, we refer readers to [2,17,18] for the Laplacian case, and [1,19] for the p-Laplacian case.
For the nonlinear boundary case, see [20] and [13,Ch. 4] where the existence of maximal and minimal classical solutions was established for the monotone case. To the best of our knowledge, our results concerning the existence of maximal and minimal weak solutions are new for both monotone and nonmonotone cases.
We begin with the definitions of weak solution and weak sub- and supersolution. For this, we make use of the real Lebesgue space Lr(∂Ω) and the Sobolev space H1(Ω).
Definition 1.1. We say that a function u∈H1(Ω) is a weak solution to (1.1) whenever:
(i) f(.,u(.))∈L2(N−1)N(∂Ω) if N>2, and f(.,u(.))∈Lr(∂Ω) for r>1 if N=2, and
(ii) ∫Ω(∇u∇ψ+uψ)=∫∂Ωf(x,u)ψforall ψ∈H1(Ω).
Definition 1.2. We say that a function ¯u∈H1(Ω) is a weak supersolution to (1.1) whenever:
(i) f(.,¯u(.))∈L2(N−1)N(∂Ω) if N>2, and f(.,¯u(.))∈Lr(∂Ω) for r>1 if N=2, and
(ii) ∫Ω(∇¯u∇ψ+¯uψ)≥∫∂Ωf(x,¯u)ψ for all 0≤ψ∈H1(Ω).
A weak subsolution u_ is defined by reversing the inequality in (ii) above.
Remark 1.3. Let Γ:H1(Ω)→Lr(∂Ω) be the trace operator given by Γu=u|∂Ω. It is known that, see e.g. [21], [22,Thm 2.79], and [23,Chapter 6], Γ is continuous (compact) if
{1≤r≤2(N−1)N−2(1≤r<2(N−1)N−2) if N>2r≥1(r≥1) if N=2. | (1.2) |
Therefore, the integrals on the right hand side of (ii) of Definition 1.1 and Definition 1.2 make sense since (i) holds, and 2(N−1)N is the conjugate of 2(N−1)N−2 when N>2.
We state and prove our results for the case N>2, since the case N=2 follows clearly using (1.2). We state our first result concerning maximal and minimal solutions for the monotone case.
Theorem 1.4. Suppose there exists a pair of weak sub- and supersolution u_ and ¯u, respectively, satisfying u_≤¯u in ¯Ω. Assume that
(H1) there exists k≥0 such that the map s↦f(x,s)+ks is nondecreasing for all u_≤s≤¯u, and for all x∈∂Ω.
Then, there exist a minimal weak solution u∗ and a maximal weak solution u∗ to (1.1), in the sense that if u is any weak solution to (1.1) such that u_≤u≤¯u, then u∗≤u≤u∗.
Next, we note that if f is locally Lipschitz with respect to the second variable u, and the interval [u_,¯u] is bounded, then f satisfies the hypothesis (H1). For functions f that do not satisfy the monotonicity condition given in (H1), we have the following existence result.
Theorem 1.5. Suppose there exists a pair of weak sub- and supersolution u_ and ¯u, respectively, satisfying u_≤¯u in ¯Ω. Assume that
(H2) there exists a K∈Lr(∂Ω), r>2(N−1)N, such that |f(x,s)|≤K(x) a.e. x∈∂Ω, for all s satisfying u_(x)≤s≤¯u(x).
Then (1.1) has at least one weak solution u such that u_≤u≤¯u.
Finally, we state a result that guarantees the existence of a maximal and a minimal weak solution without assuming monotonicity condition (H1) on the nonlinearity f.
Theorem 1.6. Assume hypotheses of Theorem 1.5 hold. Then, there exist a minimal weak solution u∗ and a maximal weak solution u∗ to (1.1), in the sense that if u is any weak solution to (1.1) such that u_≤u≤¯u, then u∗≤u≤u∗.
In [24], Hess proved the existence of a solution, assuming that
∫∂Ω supu_(x)≤s≤¯u(x)|f(x,s)|q<∞, | (1.3) |
for q=2. Our result, Theorem 1.4, is sharper, needing only that condition (1.3) hold for q=2(N−1)N=2−2N<2.
In Section 2, we collect some known results that will be helpful in the sequel. We also state and prove a version of Kato's inequality for our setting, see Theorem 2.4 and Corollary 2.5. In Section 3, we prove Theorem 1.4 using monotone iteration method. In Section 4, we prove Theorem 1.5 by showing that an appropriately defined operator is surjective. We also prove Theorem 1.6 in Section 4 by utilizing Theorem 1.5, Zorn's Lemma and Theorem 2.4. In Section 5, we discuss applications of our results.
Here we collect some results that we use in the sequel. First, we recall an existence and uniqueness result for a linear problem.
Proposition 2.1. ([4,8]) Let h∈Lq(∂Ω) for q≥1. Then, the linear problem
{−Δv+v=0inΩ;∂v∂η=hon∂Ω |
has a unique solution v∈W1,m(Ω) and
‖v‖W1,m(Ω)≤C‖h‖Lq(∂Ω),where1≤m≤Nq/(N−1). |
In particular, if q=2(N−1)N, then u∈H1(Ω).
Next, let X be a reflexive Banach space and A:X→X∗. We say that the operator A is coercive if
⟨A(ψ),ψ⟩‖ψ‖X→∞ as ‖ψ‖X→∞. |
We say that A is pseudomonotone, whenever
vn⇀vinXandlim supn→∞⟨A(vn),vn−v⟩≤0imply |
lim infn→∞⟨A(vn),vn−ψ⟩≥⟨A(v),v−ψ⟩for anyψ∈X. | (2.1) |
We will utilize the following surjectivity result in the proof of Theorem 1.5.
Proposition 2.2. ([25,Thm. II. 2.8], [22,Thm. 2.99]) Let X be a reflexive Banach space. If A:X→X∗ is a bounded, pseudomonotone and coercive operator, then for each b∈X∗, Au=b has a solution.
Finally, we say that a subset Y of a partially ordered set (X,≤) is a chain if x≤y or y≤x for every x,y∈Y. Then, to prove Theorem 1.6, we use the following version of Zorn's lemma (see [22]):
Proposition 2.3 (Zorn's lemma). If in a partially ordered set (X,≤), every chain Y has an upper bound, then X possesses a maximal element.
In [26], authors established Kato's inequality up to the boundary for a function u∈W1,1(Ω). Here, we state and prove a version of Kato's inequality up to the boundary, that is necessary in the proof of Theorem 1.6. This result can be rephrased as the maximum of two weak subsolutions is also a weak subsolution. In particular, the maximum of two weak solutions is a weak subsolution.
Theorem 2.4. Let u1 and u2 be functions in H1(Ω) such that there exist f1 and f2 in Lr(∂Ω), for r≥2(N−1)N, satisfying
∫Ω(∇ui∇ψ+uiψ)≤∫∂Ωfiψforall 0≤ψ∈H1(Ω), | (2.2) |
for i=1,2. Then, u:=max{u1,u2} satisfies
∫Ω(∇u∇ψ+uψ)≤∫∂Ωfψforall 0≤ψ∈H1(Ω), |
where f(x):={f1(x)if u1(x)>u2(x)f2(x)if u1(x)≤u2(x), a.e. x∈∂Ω.
Proof. Define
Ω1:={x∈Ω:u1(x)>u2(x)}and Ω2:=Ω∖Ω1 |
and
Γ1:={x∈∂Ω:u1(x)>u2(x)}andΓ2:=∂Ω∖Γ1. |
Fix 0≤ψ∈H1(Ω). Then,
I=∫Ω∇u∇ψ+∫Ωuψ=∫Ω1(∇u1∇ψ+u1ψ)⏟I1+∫Ω2(∇u2∇ψ+u2ψ)⏟I2. |
Consider a sequence ξn∈C1(R) such that
ξn(t):={1if t≥1/n0if t≤0, |
and ξ′n>0 on (0,1/n). Then, define the sequence of functions
rn(x):=ξn((u1−u2)(x))forx∈¯Ω. |
Observe that rn∈H1(Ω) and rn converges pointwise to χΩ1∪Γ1, where the characteristic function is defined as χΩ1∪Γ1(x):={1if x∈Ω1∪Γ10if otherwise. Moreover, ||rn||L∞(Ω)∩L∞(∂Ω)≤1 and supp(∇rn)⊂¯Dn, where Dn:={x∈Ω:0<u1(x)−u2(x)<1n}. Then, using Lebesgue Dominated Convergence Theorem, we have that
I1=limn→∞[∫Ωrn∇u1∇ψ+∫Ωrnu1ψ]. |
Since rn∈H1(Ω)∩L∞(Ω)∩L∞(∂Ω), it follows that rnψ∈H1(Ω) for any test function ψ∈H1(Ω)∩L∞(Ω). Recalling that ∇rn=0 on Ω∖Dn, and that u1 satisfies (2.2), we can write
∫Ωrn∇u1∇ψ+rnu1ψ=∫Ω∇u1∇(rnψ)+u1(rnψ)−∫Dnψ∇u1∇rn≤∫∂Ωf1rnψ−∫Dnψ∇u1∇rn. | (2.3) |
Taking the limit as n→∞ in the first term of the right-hand side of (2.3), using the Lebesgue Dominated Convergence Theorem, we get
limn→∞∫∂Ωf1rnψ=∫Γ1f1ψ. |
Likewise, for I2 we have
I2=limn→∞[∫Ω(1−rn)∇u2∇ψ+∫Ω(1−rn)u2ψ], |
and
∫Ω(1−rn)∇u2∇ψ+∫Ω(1−rn)u2ψ=∫Ω∇u2∇[(1−rn)ψ]+u2(1−rn)ψ+∫Dnψ∇u2∇rn≤∫∂Ωf2(1−rn)ψ+∫Dnψ∇u2∇rn. | (2.4) |
Taking the limit as n→∞ in the first term of the right-hand side of (2.4) and using the Lebesgue Dominated Convergence Theorem, we get
limn→∞∫∂Ωf2(1−rn)ψ=∫Γ2f2ψ. |
Using the fact that ∇rn=ξ′n(u1−u2)∇(u1−u2), the sum of the second terms of the right-hand side of (2.3) and (2.4) yields
−∫Dnψ∇u1∇rn+∫Dnψ∇u2∇rn=−∫Dnψ∇(u1−u2)∇rn=−∫Dnψξ′n(u1−u2)|∇(u1−u2)|2≤0, | (2.5) |
since ψ≥0andξ′n≥0. Adding (2.3) and (2.4), taking the limit, and using (2.5), we get
I=I1+I2≤∫Γ1f1ψ+∫Γ2f2ψ=∫∂Ωfψ. |
Thus, u:=max{u1,u2} satisfies
∫Ω(∇u∇ψ+uψ)≤∫∂Ωfψfor all 0≤ψ∈H1(Ω), |
completing the proof of Theorem 2.4.
Likewise, we have a result for the minimum of two supersolutions.
Corollary 2.5. Let u1 and u2 be functions in H1(Ω) such that there exist f1 and f2 in Lr(∂Ω), for r≥2(N−1)N, satisfying
∫Ω(∇ui∇ψ+uiψ)≥∫∂Ωfiψfor all 0≤ψ∈H1(Ω), |
for i=1,2. Then, u:=min{u1,u2} satisfies
∫Ω(∇u∇ψ+uψ)≥∫∂Ωfψ,forall 0≤ψ∈H1(Ω), |
where
f(x):={f1(x)ifu1(x)<u2(x)f2(x)ifu1(x)≥u2(x),a.e.x∈∂Ω. |
Proof. Using the fact that min{u1,u2}=max{−u1,−u2}, the proof follows from Theorem 2.4.
We will construct a monotone operator, and show that the iterative scheme starting with a weak subsolution (supersolution) will converge to a minimal (maximal) weak solution.
Let J:={u∈H1(Ω):u_≤u≤¯u}. Define the linear map T:J→H1(Ω) by T(u)=v, where v satisfies
{−Δv+v=0inΩ;∂v∂η+kv=f(x,u)+kuon∂Ω. |
Step 1. T is well-defined and maps J into itself.
For every u∈J, we have u_≤u≤¯u. Then using (H1) and the fact that u_ and ¯u are sub and supersolutions, we get
f(x,u_)+ku_≤f(x,u)+ku≤f(x,¯u)+k¯u, |
and
0≤|u|≤max{|u_|,|¯u|}≤|u_|+|¯u|. |
Taking into account the definitions of u_ and ¯u, we have that f(.,u_(.)),f(.,¯u(.)) are in L2(N−1)N(∂Ω). Since u_,¯u∈H1(Ω), then by the continuity of the trace operator (1.2) and the embedding of L2(N−1)N−2(∂Ω) into L2(N−1)N(∂Ω), for every u∈J, we have
‖f(x,u)+ku‖L2(N−1)N(∂Ω)≤‖f(x,u_)+ku_‖L2(N−1)N(∂Ω)+‖f(x,¯u)+k¯u‖L2(N−1)N(∂Ω)≤C. | (3.1) |
Therefore, f(.,u(.))+ku(.)∈L2(N−1)N(∂Ω). Then, Proposition 2.1 implies that v=T(u)∈H1(Ω) is unique. Thus, the map T is well-defined.
Further, if u,w∈J with u≤w, then by the weak maximum principle and the fact that f satisfies (H1), T(u)≤T(w), that is, the map T is nondecreasing. Moreover, repeating the argument and using Definition 1.2(ii), it follows that
u_≤T(u_)≤T(¯u)≤¯u. | (3.2) |
Hence, T maps J to J.
Step 2. There exist weakly convergent monotone sequences in H1(Ω).
Let's construct monotone sequences {un} and {wn} successively from the (linear) iteration process
un=T(un−1) with u0=u_ and wn=T(wn−1) with w0=¯u. |
Using (3.2) and the monotonicity of T, we get
u_=u0≤⋯≤un≤⋯≤wn≤⋯≤w0=¯u. | (3.3) |
We show that {un} is convergent. The proof for {wn} is analogous. We see that un=T(un−1) satisfies
∫Ω(∇un∇ψ+unψ)+k∫∂Ωunψ=∫∂Ω(f(x,un−1)+kun−1)ψ, |
for all ψ∈H1(Ω). Letting un=T(un−1) as a test function, we get
∫Ω(|∇un|2+u2n)+k∫∂Ωu2n=∫∂Ω(f(x,un−1)+kun−1)un. | (3.4) |
Since un−1,un∈J, using Hölder's inequality in (3.4), and the bound (3.1), we have
‖un‖2H1(Ω)≤‖un‖2H1(Ω)+k‖un‖2L2(∂Ω)≤‖f(x,un−1)+kun−1‖L2(N−1)N(∂Ω)‖un‖L2(N−1)N−2(∂Ω)≤C(‖¯u‖L2(N−1)N−2(∂Ω)+‖u_‖L2(N−1)N−2(∂Ω)). |
Hence, there exists a uniform constant C′>0, depending on Ω, f, k, u_ and ¯u, such that
‖un‖H1(Ω)≤C′. | (3.5) |
By the reflexivity of H1(Ω), (3.5), there is a subsequence (relabeled) un which converges weakly to u∗ in H1(Ω).
Step 3. f(x,un)+kun converges weakly to f(x,u∗)+ku∗ in L2(N−1)N(∂Ω).
Since the sequence un in Step 2 is nondecreasing and bounded (see (3.3)), it converges pointwise to u∗, that is,
u∗(x)=limn→∞un(x)∈J. | (3.6) |
Using the fact that f is continuous in the second variable u for a.e x∈∂Ω and (3.6), we have that
f(x,u∗(x))+ku∗=limn→∞f(x,un(x))+kun(x). |
By (3.1), f(x,un)+kun is bounded in L2(N−1)N(∂Ω). Then, Lebesgue Dominated Convergence Theorem yields
‖(f(x,un)+kun)−(f(x,u∗)+ku∗)‖L2(N−1)N(∂Ω)→0 as n→∞. |
Therefore, f(x,un)+kun converges strongly (hence weakly) to f(x,u∗)+ku∗ in L2(N−1)N(∂Ω). Thus, for all ψ∈H1(Ω)⊂L2(N−1)N−2(∂Ω), we have
limn→∞∫∂Ω(f(x,un)+kun)ψ=∫∂Ω(f(x,u∗)+ku∗)ψ. | (3.7) |
Step 4. u∗ is a weak solution to (1.1).
First, since u∗∈H1(Ω), the continuity of the trace (1.2) and L2(N−1)N−2(∂Ω)↪L2(N−1)N(∂Ω), imply that u∗∈L2(N−1)N(∂Ω). Therefore, for some positive constant C″, we have
‖f(x,u∗)‖L2(N−1)N(∂Ω),=‖f(x,u∗)+ku∗−ku∗‖L2(N−1)N(∂Ω),≤‖f(x,u∗)+ku∗‖L2(N−1)N(∂Ω),+‖ku∗‖L2(N−1)N(∂Ω),≤C″. |
Second, from the monotone iteration, we know that un=T(un−1) satisfies
∫Ω(∇un∇ψ+unψ)+k∫∂Ωunψ=∫∂Ω(f(x,un−1)+kun−1)ψ. |
Observe that un converges weakly to u∗ in H1(Ω), strongly in L2(∂Ω) (see Step 2) and f(x,un)+kunconverges weakly to f(x,u∗)+ku∗in L2(N−1)N(∂Ω) (see Step 3). Then taking the limit as n→∞ and using (3.7), we get for any ψ∈H1(Ω)
∫Ω(∇u∗∇ψ+u∗ψ)+∫∂Ωku∗ψ=limn→∞(∫Ω(∇un∇ψ+unψ)+∫∂Ωkunψ)=limn→∞(∫∂Ω(f(x,un−1)+kun−1)ψ)=∫∂Ω(f(x,u∗)+ku∗)ψ. |
Hence,
∫Ω(∇u∗∇ψ+u∗ψ)=∫∂Ωf(x,u∗)ψfor allψ∈H1(Ω). |
Moreover, we also have f(x,u∗)∈L2(N−1)N(∂Ω). Thus u∗ is a weak solution to (1.1).
Step 5. u∗ is the minimal weak solution in the interval [u_,¯u].
Let v be a weak solution to (1.1) with u_≤v≤¯u. Then v is a weak supersolution, and u_≤v. Repeating the above iteration procedure with u0=u_, we get u_≤u∗≤v. Thus u∗ is a weak minimal solution.
Similarly, we can construct the maximal weak solution u∗ from the sequence {wn} with w0=¯u. This completes the proof of Theorem 1.4.
We prove Theorem 1.5 by applying Proposition 2.2 to an appropriate operator related to our problem (1.1). Then, Theorem 1.6 is proved by using Zorn's lemma and Theorem 2.4. Theorem 1.5 guarantees that the set defined for the Zorn's lemma in the proof of Theorem 1.6 is nonempty.
Let us consider a modified problem
{−Δu+u=0inΩ;∂u∂η=g(x,u)on∂Ω, | (4.1) |
where
g(x,s):={f(x,u_(x)),s<u_(x),f(x,s),u_(x)≤s≤¯u(x),f(x,¯u(x)),s>¯u(x) | (4.2) |
is the truncated function. We observe that g is a Carathéodory function, since f is a Carathéodory function. We note that a weak solution u of (4.1) is a weak solution of (1.1) whenever u_≤u≤¯u.
Our plan is to establish the existence of a weak solution u of (4.1), and verify that u_≤u≤¯u. For the existence part, we use Proposition 2.2. For this, we define the map B:H1(Ω)→(H1(Ω))∗ given by
⟨B(v),ψ⟩:=∫Ω(∇v∇ψ+vψ)−∫∂Ωg(x,v)ψ, | (4.3) |
for all ψ∈H1(Ω).
First, we show that B is well-defined and bounded. The first integral of (4.3) is well-defined since v,ψ∈H1(Ω). By the Hölder's inequality combined with the continuity of trace operator (1.2) and hypothesis (H2), we get
∫{u_≤v≤¯u}|f(x,v)ψ|≤‖K‖Lr(∂Ω)‖ψ‖Lr′(∂Ω), | (4.4) |
where r′<2(N−1)N−2 is the conjugate of r. Then, the definition of g given in (4.2), Definition 1.2(i), and (4.4) yield
|∫∂Ωg(x,v)ψ|≤∫{v<u_}|f(x,u_)ψ|+∫{u_≤v≤¯u}|f(x,v)ψ|+∫{v>¯u}|f(x,¯u)ψ|≤C2‖ψ‖H1(Ω), | (4.5) |
where the last inequalities of (4.5) follow by (4.4) and (1.2), and the constant C2 depends only on K and Ω.
Second, we show that B is pseudomonotone, see definition (2.1). For this, we set B=L−G, where L,G:H1(Ω)→(H1(Ω))∗ are defined by
⟨L(v),ψ⟩:=∫Ω(∇v∇ψ+vψ) and ⟨G(v),ψ⟩:=∫∂Ωg(x,v)ψ, |
for all ψ∈H1(Ω). Then we show that B is pseudomonotone in the following steps. Let vn⇀v in H1(Ω).
Step 1: Lvn→Lv in (H1(Ω))∗.
Since vn⇀v, ⟨L(vn)−L(v),ψ⟩→0 as n→∞ for all ψ∈H1(Ω). Hence,
‖L(vn)−L(v)‖(H1(Ω))∗=sup‖ψ‖H1(Ω)≤1|⟨L(vn)−L(v),ψ⟩|→0 as n→∞, |
as desired.
Step 2: G(vn)→G(v) in (H1(Ω))∗.
Suppose that vn⇀v in H1(Ω) but G(vn)↛G(v) in (H1(Ω))∗. Then there exists ε0>0 and a subsequence {vnj} such that
‖G(vnj)−G(v)‖(H1(Ω))∗≥ε0. | (4.6) |
Using the fact that {vnj} is bounded in H1(Ω) and the compactness of the trace operator (1.2), there exists a subsequence {v′nj} such that v′nj→v in Lr′(∂Ω), where r′<2(N−1)N−2. By [27,Theorem 4.9], there exists a subsequence {v″nj} such that
v″nj(x)→v(x) a.e. x∈∂Ω. |
Since g(x,.) is continuous for a.e. x∈∂Ω, then g(x,v″nj(x))→g(x,v(x)) a.e. x∈∂Ω and g(x,v″nj(x)) is bounded in Lr(∂Ω) by (H2). Using the Lebesgue Dominated Convergence Theorem, we get
‖g(.,v″nj(.))−g(.,v(.))‖Lr(∂Ω)→0 as n→∞. |
By the Hölder's inequality, for all ψ∈H1(Ω), we get
⟨G(v″nj)−G(v),ψ⟩→0 as j→∞. |
Therefore, ‖G(v″nj)−G(v)‖(H1(Ω))∗=sup‖ψ‖H1(Ω)≤1|⟨G(v″nj)−G(v),ψ⟩|→0 as j→∞. Hence, G(v″nj)→G(v) in (H1(Ω))∗ as j→∞, a contradiction to (4.6).
Step 3: B is pseudomonotone.
Let vn⇀v in H1(Ω). Using Step 1-Step 2, we get that
B(vn)→B(v)in(H1(Ω))∗. |
Therefore, ⟨B(vn),ψ⟩→⟨B(v),ψ⟩ as n→∞ for all ψ∈H1(Ω). Furthermore, by [27,Proposition 3.5 (iv)], ⟨B(vn),vn⟩→⟨B(v),v⟩ as n→∞. Hence,
⟨B(vn),vn−ψ⟩→⟨B(v),v−ψ⟩ as n→∞, |
establishing that B is pseudomonotone.
Finally, we show that B is coercive, i.e., ⟨B(ψ),ψ⟩/‖ψ‖H1(Ω)→∞ as ‖ψ‖H1(Ω)→∞. For any ψ∈H1(Ω), using (4.5) in the definition of the operator B, we have
⟨B(ψ),ψ⟩≥‖ψ‖2H1(Ω)−C2‖ψ‖H1(Ω)≥12‖ψ‖2H1(Ω)−C3. |
Hence B is coercive. Thus B satisfies the hypotheses of Proposition 2.2 with X=H1(Ω). Therefore, for b=0∈(H1(Ω))∗, there exists u∈H1(Ω) such that
⟨B(u),ψ⟩=0∀ψ∈H1(Ω). |
Moreover, g(x,.) is bounded in L2(N−1)N−2(∂Ω) by (H2), and therefore in L2(N−1)N(∂Ω) by continuous embedding of L2(N−1)N−2(∂Ω) into L2(N−1)N(∂Ω). Hence u is a weak solution of (4.1). It remains to prove that u is a weak solution of (1.1). For this, we will show that u_≤u≤¯u in ¯Ω, so that g=f in (4.1).
Clearly, (u−¯u)+:=max{0,u−¯u}∈H1(Ω) and (u_−u)+:=max{0,u_−u}∈H1(Ω). Then, using the weak formulation of (4.1) with the test function ψ:=(u−¯u)+≥0 in H1(Ω), and the facts that ¯u is a supersolution of (1.1) and (u−¯u)+=0 in {u≤¯u}, we have
∫Ω(∇u∇(u−¯u)++u(u−¯u)+)=∫∂Ωg(x,u)(u−¯u)+=∫{u>¯u}f(x,¯u)(u−¯u)+=∫∂Ωf(x,¯u)(u−¯u)+≤∫Ω∇¯u∇(u−¯u)++∫Ω¯u(u−¯u)+. | (4.7) |
Then, (4.7) yields
0≤∫Ω|∇(u−¯u)+|2+∫Ω|(u−¯u)+|2=∫Ω∇(u−¯u)∇(u−¯u)++∫Ω(u−¯u)(u−¯u)+≤0, |
which implies that ‖(u−¯u)+‖H1(Ω)=0. That is, u≤¯u a.e. in Ω. Using the the continuity of the trace operator (1.2), we get that ‖(u−¯u)+‖L2(N−1)N−2(∂Ω)=0. Hence, u≤¯u a.e. in ¯Ω.
Analogously, taking the test function ψ:=(u_−u)+≥0 and using the fact that u_ is a subsolution of (1.1), we obtain that
0≤∫Ω|∇(u_−u)+|2+∫Ω|(u_−u)+|2=∫Ω∇(u_−u)∇(u_−u)++∫Ω(u_−u)(u_−u)+≤0, |
Therefore, u_≤u a.e. in ¯Ω, and hence u_≤u≤¯u a.e. in ¯Ω. Thus, u is a weak solution of (1.1), completing the proof of Theorem 1.5.
We will use Zorn's Lemma and Proposition 2.3, to prove our result. Consider the set
A:={u∈H1(Ω):u_(x)≤u(x)≤¯u(x) a.e. in ¯Ω and u is a weak solution of (1.1)}, |
and we note that A is nonempty by Theorem 1.5. Let {ui}i∈I⊆A be a family of chain. Since ui is a weak solution of (1.1), taking ui as the test function and using (4.4), we get
‖ui‖H1(Ω)=∫Ω(|∇ui|2+u2i)=∫∂Ωf(x,ui)ui≤C, |
where C depends on u_,¯u,K,∂Ω but independent of i∈I. By the separability and reflexivity of H1(Ω), there exists an increasing sequence un such that
un⇀u:=supi∈Iui in H1(Ω). |
Clearly, u is an upper bound of the chain {ui}i∈I. It suffices to show that u∈A. Since {un} is nondecreasing and u_≤un≤¯u, we have that un(x)→u(x), and un(x)≤u(x) for all n, and u_(x)≤u(x)≤¯u(x) pointwise a.e. in ¯Ω. Furthermore, since f is Carathéodory, we have that
f(x,un(x))→f(x,u(x)) as n→∞. |
This, in conjunction with (H2), and the Lebesgue Dominated Convergence Theorem yields ‖f(x,un)−f(x,u)‖Lr(∂Ω)→0 as n→∞. Therefore, using Hölder's inequality, we deduce that
|∫∂Ωf(x,un)ψ−∫∂Ωf(x,u)ψ|≤∫∂Ω|f(x,un)−f(x,u)||ψ|≤‖f(x,un)−f(x,u)‖Lr(∂Ω)‖ψ‖Lr′(∂Ω)→0, |
which yields
limn→∞∫∂Ωf(x,un)ψ=∫∂Ωf(x,u)ψ for all ψ∈H1(Ω). |
Taking the limit as n→∞, we get for any ψ∈H1(Ω)
∫Ω(∇u∇ψ+uψ)=limn→∞∫Ω(∇un∇ψ+unψ)=limn→∞∫∂Ωf(x,un)ψ=∫∂Ωf(x,u)ψ. |
Hence, u is a weak solution of (1.1), thus concluding that u∈A.
By Zorn's Lemma, there exists a maximal element u∗∈A. It remains to show that u∗ is maximal in the sense that if ˆu is any other weak solution of (1.1) between u_ and ¯u, then ˆu≤u∗. So, let ˆu be a weak solution of (1.1) between u_ and ¯u, and u∗ is the maximal element of A. By Proposition 2.4, u=max{ˆu,u∗} is a subsolution of (1.1). Then, by Theorem 1.5, there exists a weak solution u0 of (1.1) satisfying
u_≤u≤u0≤¯u. |
Thus, u0∈A. On the other hand, u∗≤max{ˆu,u∗}=u≤u0. But u∗ is maximal element of A, so necessarily u∗=u0. Therefore, we readily see that ˆu≤u≤u0=u∗, and hence
u_≤ˆu≤u∗≤¯u, |
as desired. The existence of a minimal element u∗ of A is proved analogously. This completes the proof of Theorem 1.6.
In this section, we apply our existence results, Theorem 1.4 and Theorem 1.5, to problems involving sublinear nonlinearities. In particular, in each case we construct an ordered pair of weak sub- and supersolution. We apply Theorem 1.4 to establish Theorem 5.1 and Theorem 1.5 in Remark 5.2 below.
Theorem 5.1. Consider
{−Δu+u=0inΩ;∂u∂η=λf(u)on∂Ω, | (5.1) |
where λ>0 parameter and f:[0,∞)→[0,∞) is locally Lipschitz continuous function satisfying
(i) f(0)=0 with f′(0)>0, and
(ii) lims→∞f(s)s=0.
Then (5.1) has a positive weak solution for λ>μ1f′(0), where μ1>0 is the first eigenvalue of the Steklov eigenvalue problem
{−Δφ1+φ1=0inΩ;∂φ1∂η=μ1φ1on∂Ω, | (5.2) |
and 0<φ1∈H1(Ω) is the corresponding eigenfunction.
Proof. Let λ>μ1f′(0) be fixed. Using hypothesis (i), we verify that u_:=ϵφ1 is a subsolution of (5.1) for ϵ≈0. Indeed, we observe that since λ>μ1f′(0) is fixed, ξ(s):=μ1s−λf(s) satisfies ξ(0)=0 and ξ′(0)<0, then ξ(s)<0 for s≈0. Therefore, for all 0≤ψ∈H1(Ω), the following holds for ϵ≈0
∫Ω∇u_∇ψ+∫Ωu_ψ=μ1∫∂Ω(ϵφ1)ψ≤λ∫∂Ωf(ϵφ1)ψ=λ∫∂Ωf(u_)ψ. |
Next, using hypothesis (ii), we show that there exists Mλ>0 such that ¯u:=Me is a weak supersolution of (5.1) for all M≥Mλ, where e is the unique positive solution of
{−Δe+e=0in Ω;∂e∂η=1on ∂Ω. |
We observe that while f is not assumed to be nondecreasing, ¯f(t):=maxs∈[0,t]f(s) is nondecreasing, and f(t)≤¯f(t) for all t≥0. Moreover, due to hypothesis (ii), ¯f satisfies the sublinear condition at infinity
limt→+∞¯f(t)t=0. |
Therefore, there exists Mλ>0 such that for all M≥Mλ
¯f(M‖e‖L∞(∂Ω))M‖e‖L∞(∂Ω)≤1λ‖e‖L∞(∂Ω) or equivalently λ¯f(M‖e‖L∞(∂Ω))≤M. |
Then ¯u=Me∈H1(Ω) satisfies
∫Ω∇¯u∇ψ+∫Ω¯uψ=M∫∂Ωψ≥λ∫∂Ω¯f(M‖e‖L∞(∂Ω))ψ≥λ∫∂Ω¯f(Me)ψ≥λ∫∂Ωf(Me)ψ=λ∫∂Ωf(¯u)ψ |
for all 0≤ϕ∈H1(Ω). Therefore, ¯u is a weak supersolution of (1.1) for each λ>μ1f′(0). Clearly ¯u=Me≥ϵ(λ)φ1=u_ a.e. in ¯Ω. We remark that since f is locally Lipschitz and [u_,¯u] is bounded, f satisfies hypothesis (H1) of Theorem 1.4. Hence, there exists a positive weak solution u of (5.1) such that ϵφ1≤u≤Me a.e. in ¯Ω for any λ>μ1f′(0). This completes the proof.
Remark 5.2. On the other hand, if f is continuous (not necessarily Lipschitz), satisfies hypothesis (ii) of Theorem 5.1 and f(s)>0 for s≥0, the problem (5.1) has a positive weak solution for each λ>0. Indeed, it is easy to see that u_≡0 is a strict weak subsolution and for each λ>0, there exists Mλ>0 such that ¯u=Me is a weak supersolution for all M≥Mλ, as in the proof of Theorem 5.1. Then, the result follows by Theorem 1.5.
We would like to thank the referee for carefully reading the manuscript and providing valuable suggestions. We would like to thank Professor Jesús Jaramillo for helpful discussions about Step 2 in the proof of Theorem 1.5. The fifth author is supported by grants PID2019-103860GB-I00, MICINN, Spain, and by UCM-BSCH, Spain, GR58/08, Grupo 920894. All authors acknowledge MSRI for bringing this group together for collaboration.
The authors declare there is no conflicts of interest.
[1] | M. Cuesta Leon, Existence results for quasilinear problems via ordered sub- and supersolutions, Ann. Fac. Sci. Toulouse Math., 6 (1997), 591–608. http://www.numdam.org/item?id=AFST_1997_6_6_4_591_0 |
[2] |
J. Schoenenberger-Deuel, P. Hess, A criterion for the existence of solutions of non-linear elliptic boundary value problems, Proc. Roy. Soc. Edinburgh Sect. A, 74 (1976), 49–54. https://doi-org.proxy.swarthmore.edu/10.1017/s030821050001653x doi: 10.1017/s030821050001653x
![]() |
[3] |
J. M. Arrieta, R. Pardo, A. Rodríguez-Bernal, Infinite resonant solutions and turning points in a problem with unbounded bifurcation, Internat. J. Bifur. Chaos Appl. Sci. Engrg., 20 (2010), 2885–2896. https://doi.org/10.1142/S021812741002743X doi: 10.1142/S021812741002743X
![]() |
[4] | J. M. Arrieta, R. Pardo, A. Rodríguez-Bernal, Bifurcation and stability of equilibria with asymptotically linear boundary conditions at infinity, Proc. Roy. Soc. Edinburgh Sect. A, 137 (2007), 225–252. |
[5] |
J. M. Arrieta, R. Pardo, A. Rodríguez-Bernal, Equilibria and global dynamics of a problem with bifurcation from infinity, J. Differ. Equ., 246 (2009), 2055–2080. https://doi.org/10.1016/j.jde.2008.09.002 doi: 10.1016/j.jde.2008.09.002
![]() |
[6] |
P. Liu, J. Shi, Bifurcation of positive solutions to scalar reaction-diffusion equations with nonlinear boundary condition, J. Differ. Equ., 264 (2018), 425–454. https://doi.org/10.1016/j.jde.2017.09.014 doi: 10.1016/j.jde.2017.09.014
![]() |
[7] |
N. Mavinga, Generalized eigenproblem and nonlinear elliptic equations with nonlinear boundary conditions, Proc. Roy. Soc. Edinburgh Sect. A, 142 (2012), 137–153. https://doi.org/10.1017/S0308210510000065 doi: 10.1017/S0308210510000065
![]() |
[8] |
N. Mavinga, R. Pardo, Bifurcation from infinity for reaction-diffusion equations under nonlinear boundary conditions, Proc. Roy. Soc. Edinburgh Sect. A, 147 (2017), 649–671. https://doi.org/10.1017/S0308210516000251 doi: 10.1017/S0308210516000251
![]() |
[9] | C. Morales-Rodrigo, A. Suárez, Some elliptic problems with nonlinear boundary conditions, in Spectral theory and nonlinear analysis with applications to spatial ecology, World Sci. Publ., Hackensack, NJ, 2005,175–199. https://doi.org/10.1142/9789812701589_0009 |
[10] |
J. M. Arrieta, A. N. Carvalho, A. Rodríguez-Bernal, Parabolic problems with nonlinear boundary conditions and critical nonlinearities, J. Differ. Equ., 156 (1999), 376–406. https://doi.org/10.1006/jdeq.1998.3612 doi: 10.1006/jdeq.1998.3612
![]() |
[11] |
R. S. Cantrell, C. Cosner, On the effects of nonlinear boundary conditions in diffusive logistic equations on bounded domains, J. Differ. Equ., 231 (2006), 768–804. https://doi.org/10.1016/j.jde.2006.08.018 doi: 10.1016/j.jde.2006.08.018
![]() |
[12] |
A. A. Lacey, J. R. Ockendon, J. Sabina, Multidimensional reaction diffusion equations with nonlinear boundary conditions, SIAM J. Appl. Math., 58 (1998), 1622–1647. https://doi.org/10.1137/S0036139996308121 doi: 10.1137/S0036139996308121
![]() |
[13] | C. V. Pao, Nonlinear parabolic and elliptic equations, Plenum Press, New York, 1992. |
[14] |
K. Akô, On the Dirichlet problem for quasi-linear elliptic differential equations of the second order, J. Math. Soc. Japan, 13 (1961), 45–62, https://doi.org/10.2969/jmsj/01310045 doi: 10.2969/jmsj/01310045
![]() |
[15] |
H. Amann, On the existence of positive solutions of nonlinear elliptic boundary value problems, Indiana Univ. Math. J., 21 (1971), 125–146, https://doi.org/10.1512/iumj.1971.21.21012 doi: 10.1512/iumj.1971.21.21012
![]() |
[16] |
D. H. Sattinger, Monotone methods in nonlinear elliptic and parabolic boundary value problems, Indiana Univ. Math. J., 21 (1971), 979–1000. https://doi.org/10.1512/iumj.1972.21.21079 doi: 10.1512/iumj.1972.21.21079
![]() |
[17] |
H. Amann, M. G. Crandall, On some existence theorems for semi-linear elliptic equations, Indiana Univ. Math. J., 27 (1978), 779–790. https://doi.org/10.1512/iumj.1978.27.27050 doi: 10.1512/iumj.1978.27.27050
![]() |
[18] | E. N. Dancer, G. Sweers, On the existence of a maximal weak solution for a semilinear elliptic equation, Differ. Integral Equ., 2 (1989), 533–540. |
[19] |
D. Motreanu, A. Sciammetta, E. Tornatore, A sub-supersolution approach for {N}eumann boundary value problems with gradient dependence, Nonlinear Anal. Real World Appl., 54 (2020), 103096. https://doi.org/10.1016/j.nonrwa.2020.103096 doi: 10.1016/j.nonrwa.2020.103096
![]() |
[20] | H. Amann, Nonlinear elliptic equations with nonlinear boundary conditions, in New developments in differential equations (Proc. 2nd Scheveningen Conf., Scheveningen, 1975), 1976, 43–63. North–Holland Math. Studies, Vol. 21. |
[21] | R. A. Adams, J. J. F. Fournier, Sobolev spaces, vol. 140 of Pure and Applied Mathematics (Amsterdam), 2nd edition, Elsevier/Academic Press, Amsterdam, 2003. |
[22] | S. Carl, V. K. Le, D. Motreanu, Nonsmooth variational problems and their inequalities, Springer Monographs in Mathematics, Springer, New York, 2007. https://doi.org/10.1007/978-0-387-46252-3 Comparison principles and applications. |
[23] | A. Kufner, O. John, S. Fučík, Function spaces, Noordhoff International Publishing, Leyden; Academia, Prague, 1977, Monographs and Textbooks on Mechanics of Solids and Fluids; Mechanics: Analysis. |
[24] |
P. Hess, On the solvability of nonlinear elliptic boundary value problems, Indiana Univ. Math. J., 25 (1976), 461–466. https://doi.org/10.1512/iumj.1976.25.25036 doi: 10.1512/iumj.1976.25.25036
![]() |
[25] | J.-L. Lions, Quelques méthodes de résolution des problèmes aux limites non linéaires, Dunod; Gauthier-Villars, Paris, 1969. |
[26] |
H. Brezis, A. C. Ponce, Kato's inequality up to the boundary, Commun. Contemp. Math., 10 (2008), 1217–1241. https://doi.org/10.1142/S0219199708003241 doi: 10.1142/S0219199708003241
![]() |
[27] | H. Brezis, Functional analysis, Sobolev spaces and partial differential equations, Universitext, Springer, New York, 2011. |
1. | Shalmali Bandyopadhyay, Maya Chhetri, Briceyda B. Delgado, Nsoki Mavinga, Rosa Pardo, Bifurcation and multiplicity results for elliptic problems with subcritical nonlinearity on the boundary, 2024, 411, 00220396, 28, 10.1016/j.jde.2024.07.041 |