Processing math: 0%
Home Symmetric and Asymmetric Solutions of p-Laplace Elliptic Equations in Hollow Domains
Article Open Access

Symmetric and Asymmetric Solutions of p-Laplace Elliptic Equations in Hollow Domains

  • Ryuji Kajikiya EMAIL logo
Published/Copyright: June 20, 2017

Abstract

In the present paper, we study the p-Laplace equation in a hollow symmetric bounded domain. Let H and G be closed subgroups of the orthogonal group such that H. Then we prove the existence of a positive solution which is H-invariant and G-non-invariant. Furthermore, we give several examples of H, G and Ω, and find symmetric and asymmetric solutions.

MSC 2010: 35J20; 35J25

1 Introduction and Problems

We study the existence of symmetric and asymmetric positive solutions for the p-Laplace elliptic equation

(1.1) - Δ p u = f ( x , u ) , u > 0 in  Ω , u = 0 on  Ω ,

where Δpu:=div(|u|p-2u) is the p-Laplacian with p2, Ω is a bounded domain in N with piecewise smooth boundary Ω, and f(x,u) is a continuous function on Ω¯×[0,). In the present paper, we consider a domain Ω which has a symmetry and a hole like an annulus. We denote the orthogonal group by O(N), which is the set of all N×N orthogonal matrices. Let G be a closed subgroup of O(N). We call Ω a G-invariant domain if g(Ω)=Ω for all gG. We call f(x,u) a G-invariant function if f(gx,u)=f(x,u) for all gG, xΩ and u[0,). We call a solution u(x) of (1.1) a G-invariant solution if u(gx)=u(x) for gG and xΩ. We define the Lagrangian functional I(u) for (1.1) by

I ( u ) := Ω ( 1 p | u | p - F ( x , u ) ) 𝑑 x , F ( x , u ) := 0 u f ( x , s ) 𝑑 s .

We denote the Fréchet derivative of I(u) by I(u), which is computed as follows:

I ( u ) v = Ω ( | u | p - 2 u v - f ( x , u ) v ) 𝑑 x for  u , v W 0 1 , p ( Ω ) .

Here W01,p(Ω) is the usual Sobolev space. We put

J ( u ) := I ( u ) u = Ω ( | u | p - f ( x , u ) u ) 𝑑 x .

We define the Nehari manifold𝒩 and the minimum energyI0 by

𝒩 := { u W 0 1 , p ( Ω ) { 0 } : J ( u ) = 0 } ,
(1.2) I 0 := inf { I ( u ) : u 𝒩 } .

We call u(x) a minimum energy solution if u𝒩 and I(u)=I0. A minimum energy solution is not necessarily unique. We shall prove in Section 4 that I0>0 and that there exists a minimum energy solution which becomes a positive solution of (1.1) after replacing u by -u, if necessary.

For a closed subgroup G of O(N), we put

W 0 1 , p ( Ω , G ) := { u W 0 1 , p ( Ω ) : u ( g x ) = u ( x ) , g G , x Ω } ,
(1.3) 𝒩 ( G ) := 𝒩 W 0 1 , p ( Ω , G ) ,
(1.4) I G := inf { I ( u ) : u 𝒩 ( G ) } .

We call u a G-minimum solution if u𝒩(G) and I(u)=IG. Such a minimizer exists and becomes a G-invariant positive solution of (1.1); this will be proved in Section 4. To avoid confusion, we call a usual minimum energy solution a global minimum solution. Since a G-minimum solution exists, we are interested in a G-non-invariant positive solution. The purpose of the present paper is to solve the two problems below.

Problem A.

Find G and Ω such that G is a closed subgroup of O(N), Ω is a G-invariant bounded domain, and no global minimum solution is G-invariant.

Problem B.

Find H, G and Ω such that H and G are closed subgroups of O(N) satisfying HGO(N), Ω is a G-invariant bounded domain, and no H-minimum solution is G-invariant.

Comparing the two problems above, we see that Problem A is included in Problem B. Indeed, we put H:={e}, with e being the unit matrix. Then a global minimum solution is equal to an H-minimum solution. Therefore, an answer to Problem A follows from that to Problem B with H={e}.

The goal of the present paper is to solve Problem B. We shall explain the reason why we have an interest in this problem. We start with G=O(N) and introduce the result by Coffman [6]. Let AN(a,ε) be an annulus defined by

A N ( a , ε ) := { x N : a < | x | < a + ε } .

We consider the Emden–Fowler equation

(1.5) - Δ u = u p , u > 0 in  A N ( a , ε ) , u = 0 on  A N ( a , ε ) .

Two solutions u(x) and v(x) are called equivalent if u(gx)=v(x) for a certain gO(N). Coffman [6], Li [15] and Byeon [3] proved the next result.

Proposition 1.1.

The number of non-equivalent positive solutions for (1.5) diverges to infinity as ε+0. Furthermore, no global minimum solution is radially symmetric when ε is small enough.

For more results on the annulus, we refer to [1, 5, 9, 10, 16, 17, 18, 19, 21, 22, 26, 29, 30]. When Ω is not an annulus, the existence of multiple positive solutions is proved in [4].

To explain the reason why we study Problems A and B, we shall propose some problems. First, we consider a hollow regular polygon, which is a domain enclosed by a small and a large regular n-gon with a common center and with their sides parallel to each other. To give a strict definition, let Pn be an interior of a regular n-gon centered at the origin. For ε>0, we define

( 1 + ε ) P n := { ( 1 + ε ) x : x P n } .

Remove P¯n from (1+ε)Pn and define

(1.6) HP n ( ε ) := ( 1 + ε ) P n P ¯ n .

This is a hollow regular n-gon. Define

(1.7) G n := { ρ ( 2 j π n ) : j = 0 , 1 , , n - 1 } , ρ ( θ ) := ( cos θ - sin θ sin θ cos θ ) .

Then HPn(ε) is Gn-invariant. Suppose that f(x,u) is also Gn-invariant. Observing Proposition 1.1, we consider the problem below.

Problem 1.2.

Let Ω=HPn(ε) be the hollow regular n-gon, with ε>0 small enough. Does (1.1) have a positive solution without Gn-invariance?

For the homogeneous nonlinear term f(x,u)=uq, we solved the above problem in the affirmative for p=2 (see [14]) and for p2 (see [13]). In the corresponding papers, we used the functional R(u) defined by

R ( u ) := ( Ω | u | 2 d x ) ( Ω | u | q + 1 d x ) - 2 / ( q + 1 ) .

Then a global minimum solution is defined by a minimizer of R(u) in 𝒩. In equation (1.1), the nonlinear term f(x,u) is not assumed to have a homogeneity like f(u)=uq. Therefore, we cannot use R(u) in the present paper, and hence we employ the Lagrangian functional I(u) and develop a new method. Another aim of the present paper is to extend the results in [13, 14] to a more general equation (1.1).

We shall propose a new problem. Let L be one of the axes of reflection symmetry for a regular n-gon Pn. Denote the reflection group with respect to L by HL. We call u an L-solution if u is an HL-minimum solution for (1.1) with Ω=HPn(ε). Recall that two solutions are called equivalent if they are transformed by an orthogonal matrix. We consider the next problem for Ω=HPn(ε).

Problem 1.3.

Find HPn(ε), L1 and L2 such that L1 and L2 are axes of symmetry for Pn and no L1-solution is equivalent to any L2-solution.

For the annulus, we consider an H-invariant non-radial solution for a closed subgroup H of O(N).

Problem 1.4.

Let Ω be the annulus AN(a,ε). Find all closed subgroups H of O(N) such that no H-minimum solution is radially symmetric.

An answer to the problem above follows from that to Problem B by putting G=O(N). To state a new problem, we consider HP3(ε) given by (1.6) with n=3, that is, it is a hollow equilateral triangle. Then HP3(ε) has three axes of reflection symmetry. Let L be one of these axes.

Problem 1.5.

Let Ω=HP3(ε) be the hollow equilateral triangle. Does (1.1) have a positive solution which is reflectionally symmetric with respect to L and not invariant under the rotation of 120?

We shall consider a cube and a cross polytope in N. Let CN be the following N-dimensional cube:

C N := { ( x 1 , , x N ) N : | x i | < 1  for  1 i N } .

We define a hollow cubeHCN(ε) by

(1.8) HC N ( ε ) := ( 1 + ε ) C N C ¯ N .

Let E be the union of all edges of CN. Let SCN(ε) be an ε-neighborhood of E. This is a skeleton cube. We denote a regular cross polytope in N by CPN. Then CPN is a convex hull of all points Pi+ and Pi- with i=1,2,,N, where

P i ± := ( 0 , , 0 , ± 1 , 0 , , 0 )

is a point in N whose i-th element is ±1 and the others are 0. Therefore, CPN is defined by

CP N := { i = 1 N t i + P i + + i = 1 N t i - P i - : t i ± 0 , i = 1 N t i + + i = 1 N t i - < 1 } = { ( x 1 , , x N ) N : | x 1 | + + | x N | < 1 } .

Define

(1.9) HCP N ( ε ) := ( 1 + ε ) CP N CP ¯ N ,

which is a hollow cross polytope. Let SCPN(ε) be an ε-neighborhood of the union of all edges of CPN. This is a skeleton cross polytope.

Let H be the set of all orthogonal matrices gO(N) whose each element is equal to 0 or 1, that is, H is the set of all N×N permutation matrices. Let G be the set of gO(N) whose each element is equal to 0, 1 or -1. We define Ω by HCN(ε), SCN(ε), HCPN(ε), SCPN(ε) or AN(a,ε) for ε>0 small. It is easy to verify that Ω is G-invariant because CN and CPN are G-invariant. The group H is isomorphic to the symmetric group SN, which is the set of all permutations

σ = ( 1 N σ ( 1 ) σ ( N ) ) .

A function u(x) is H-invariant if and only if

(1.10) u ( x 1 , , x N ) = u ( x σ ( 1 ) , , x σ ( N ) ) for  σ S N , x Ω .

A function u(x) is G-invariant if and only if

(1.11) u ( x 1 , , x N ) = u ( τ 1 x σ ( 1 ) , , τ N x σ ( N ) )

for σSN, τ1,,τN{1,-1} and xΩ.

Problem 1.6.

Let Ω be one of the following: HCN(ε), SCN(ε), HCPN(ε), SCPN(ε) or AN(a,ε), with ε>0 being small. Does there exist a positive solution u(x) which satisfies (1.10) and does not satisfy (1.11)?

All the problems in this section are reduced to Problem B. Therefore, we shall solve Problem B. This paper is organized in five sections. In Section 2, we state the main results. In Section 3, by using the main theorems, we solve all the problems that appear in Section 1. In Section 4, we prove the existence of a global minimum solution and a G-minimum solution. In Section 5, we prove the main theorems.

2 Main Results

In this section, we state the main results. For a closed subgroup G of O(N), we define the orbit of G through xN by

(2.1) G ( x ) := { g x : g G } .

We suppose the two assumptions below.

Assumption 2.1.

Let G, H and D be such that G and H are closed subgroups of O(N), D is a G-invariant bounded domain in N, 0D¯, HG and H(x)G(x) for xD¯. Here H(x) and G(x) are orbits defined by (2.1).

Assumption 2.2.

Let G be a closed subgroup of O(N) and let D be a G-invariant bounded domain in N whose closure does not contain the origin. Suppose that p2 and assume the following conditions:

  1. f ( x , u ) is a G-invariant continuous function on D¯×[0,) such that f(x,0)=0 and f(x,u)>0 for u>0 and xD¯.

  2. f ( x , u ) has a continuous partial derivative fu(x,u), and there exists a constant q(p,) (hence q>p2) such that

    (2.2) u ( f ( x , u ) u q - 1 ) = f u ( x , u ) u - ( q - 1 ) f ( x , u ) u q > 0

    for u>0 and xD¯.

  3. There exist constants r,C>0 such that |fu(x,u)|C(ur-2+1) for u0 and xD¯, where r satisfies p<r<Np/(N-p) when p<N, and p<r< when pN.

As a consequence of the assumption above, we have fu(x,0)=0.

Example 2.3.

Some examples of f(x,u) satisfying Assumption 2.2 are

a ( x ) u r - 1 , a ( x ) u r - 1 + b ( x ) u s - 1 , a ( x ) u r - 1 1 + u t ,
a ( x ) u r - 1 log ( 1 + u ) , a ( x ) u r - 1 tanh u , a ( x ) u r - 1 arctan u ,

where we assume that r,s(p,Np/(N-p)) for p<N and r,s(p,) for pN. We assume also 0<t<r-p, and that a(x), b(x) are G-invariant positive continuous functions on D¯. The sum of these functions also satisfies Assumption 2.2.

We shall give a domain Ω as a G-invariant subdomain of D. We extend f(x,u) as an odd function of u, i.e., f(x,-u):=-f(x,u). Observe the definition of IG in (1.4). To give an affirmative answer to Problem B, it is enough to show that IH<IG. Indeed, this inequality ensures that no H-minimum solution is G-invariant. We denote the first eigenvalue of the Dirichlet p-Laplacian in the domain Ω by λ1(Ω), i.e.,

- Δ p u = λ 1 ( Ω ) u p - 1 u > 0 in  Ω , u = 0 on  Ω .

We state the first main result, which is an answer to Problem B.

Theorem 2.4.

Let Assumptions 2.1 and 2.2 hold. Then there exists a constant Λ>0 such that if Ω is a G-invariant subdomain of D which satisfies λ1(Ω)>Λ, then we have that IH<IG. Therefore, no H-minimum solution is G-invariant.

Observe the definitions of 𝒩(G) and IG given in (1.3) and (1.4), respectively. The inclusion HG implies 𝒩(G)𝒩(H), which ensures that IHIG. Consequently, the stronger symmetry has higher energy. The essential point of Theorem 2.4 is the strict inequality IH<IG. The theorem above guarantees the existence of at least two positive solutions: one is an H-invariant G-non-invariant positive solution and the other is a G-invariant positive solution.

It is well known that the first eigenvalue λ1(Ω) diverges to infinity as the volume of Ω tends to 0. Therefore, we have the corollary below.

Corollary 2.5.

Under Assumptions 2.1 and 2.2, there exists a constant ε>0 such that if Ω is a G-invariant subdomain of D satisfying |Ω|<ε, then IH<IG. Here |Ω| denotes the volume of Ω.

Put H={e}, with e being the unit matrix. Then an H-minimum solution coincides with a global minimum solution. Furthermore, the assumption H(x)={x}G(x) is equivalent to a condition that G(x) has at least two points. In other words, there do not exist any fixed points of G in D¯, where x is called a fixed point of G if gx=x for all gG. Therefore, Theorem 2.4 is reduced to the result below, which is an answer to Problem A.

Theorem 2.6.

Let Assumption 2.2 hold and assume that for each xD¯, G(x) includes at least two points. Then there exists a constant Λ>0 such that if Ω is a G-invariant subdomain of D which satisfies λ1(Ω)>Λ, then we have that I0<IG. Therefore, no global minimum solution is G-invariant.

In the theorems above, we first fix D and then define the subset Ω of D. This is complicated. We shall explain the reason why we need this procedure. To this end, we consider the equation,

(2.3) - Δ u = u q , u > 0 in  Ω ( a , b ) , u = 0 on  Ω ( a , b ) ,

where Ω(a,b) is the annulus a<|x|<b in N and q is assumed to satisfy 1<q<(N+2)/(N-2) for N3 and 1<q< for N=2. Then Dancer [8] proved the following result.

Lemma 2.7 ([8]).

For any b>0, there exists an ε>0, depending on b, such that if 0<a<ε, then (2.3) has a unique positive solution. Moreover, it becomes radially symmetric. In particular, a global minimum solution is radially symmetric.

Let B(b) denote a ball centered at the origin with radius b>0. Since Ω(a,b)B(b), we have that

λ 1 ( Ω ( a , b ) ) λ 1 ( B ( b ) ) .

As b0, the right-hand side diverges to infinity and so does λ1(Ω(a,b)). Putting H:={e}, with e being the unit matrix, and G:=O(N), we compare Lemma 2.7 and Theorem 2.4. Let Λ>0 be any number. If we do not set D, then we can choose b>0 sufficient small so that λ1(Ω(a,b))>Λ. Lemma 2.7 says that a global minimum solution is radially symmetric when a>0 is small enough. Hence, the conclusion of Theorem 2.4 does not hold. This is caused by the assumption that the radius a>0 of the inner hole of Ω(a,b) is very small. We first fix D and then define the subset Ω of D in Theorem 2.4. This procedure prevents the inner hole of Ω(a,b) from shrinking.

Let H be a closed subgroup of O(N). Then it is a linear isometric transformation group on N. Hence, it becomes a transformation group on the unit sphere

S N - 1 := { x N : | x | = 1 } .

H is said to be transitive on SN-1 if H(x)=SN-1 for xSN-1, where H(x) is the orbit defined by (2.1). All the transitive Lie groups have been listed by Montgomery and Samelson [23], and Borel [2] (see also [11, p. 186, Theorem 2.6] or [25, p. 267, Theorem 3]).

Theorem 2.8 ([23, 2]).

Let N2 and H be a connected closed subgroup of SO(N). Here SO(N) is the special orthogonal group (rotation group). Then H is transitive on SN-1 if and only if H is locally isomorphic to one of the following groups:

  1. SO ( N ) ,

  2. SU ( m ) , U(m) if N=2m,

  3. Sp ( m ) , Sp(m)Sp(1), Sp(m)U(1) if N=4m,

  4. Spin ( 9 ) if N = 16 ,

  5. Spin ( 7 ) if N = 8 ,

  6. G 2 if N = 7 .

Let H be a closed subgroup of O(N) which is not necessarily connected. Then it is transitive on SN-1 if and only if the connected component of H, including the unit matrix, is locally isomorphic to one of the Lie groups listed in Theorem 2.8.

Let Ω=AN(a,ε) be the annulus for ε small enough. Let H be a closed subgroup of O(N) and put G:=O(N). Then G(x)=SN-1 for xSN-1. If H is not transitive on SN-1, then H(x)G(x) for xSN-1, and so H(x)G(x) for x0. Theorem 2.4 says that if H is not transitive, then no H-minimum solution is G-invariant (i.e., radially symmetric). On the other hand, if H is transitive, any H-invariant function is radially symmetric. Consequently, we have the next result.

Theorem 2.9.

Let Assumption 2.2 hold with G=O(N) and D=AN(a,1). Let Ω=AN(a,ε) be the annulus with ε>0 small enough. Then the following are equivalent:

  1. No H -minimum solution is radially symmetric.

  2. H is not transitive on S N - 1 .

The theorem above is a complete answer to Problem 1.4. We give an application of this theorem to a non-transitive group.

Example 2.10.

Let

G := SO ( m ) × SO ( N - m ) = { ( g 0 0 h ) : g SO ( m ) , h SO ( N - m ) } .

Clearly, G is not transitive. Therefore, for small ε>0, a G-minimum solution u(x) of (1.1) on AN(a,ε) is not radially symmetric, that is, u=u(|x|,|x′′|) for xm, x′′N-m, with x=(x,x′′)AN(a,ε) but uu(|x|).

Since any finite subgroup is not transitive, we have the next result.

Corollary 2.11.

Let the assumptions of Theorem 2.9 hold. If H is a finite subgroup of O(N), then for small ε>0, no H-minimum solution of (1.1) on AN(a,ε) is radially symmetric.

3 Answers to the Problems

By applying the main theorems stated in Section 2, we give an affirmative answer to all the problems given in Section 1. Throughout this section, we assume that f(x,u) satisfies Assumption 2.2. In order to use Theorem 2.4, we have only to verify that G and H satisfy Assumption 2.1.

We recall the definition of equivalence given in Section 1. Solutions u and v are called equivalent if u(gx)=v(x) for xΩ, with a certain gO(N). Define HPn(ε) and Gn by (1.6) and (1.7), respectively. Then Ω:=HPn(ε) is Gn-invariant. If m is a divisor of n, then Gm is a subgroup of Gn and hence Ω is Gm-invariant. The next result gives an answer to Problem 1.2.

Theorem 3.1.

Put Ω=HPn(ε) with ε>0 small enough. Let k and m be any two divisors of n satisfying 1k<mn. Let uk and um be a Gk- and Gm-minimum solution, respectively. Then uk is not equivalent to um. Moreover, if k is a divisor of m, then I(uk)<I(um). Therefore, (1.1) has at least d non-equivalent solutions, where d is a number of all divisors of n.

In the theorem above, for k=1 and m=n, u1 is a global minimum solution and un is a Gn-minimum solution. By Theorem 3.1, I0=I(u1)<I(un)=IGn, and hence a global minimum solution is not Gn-invariant. Problem 1.2 is solved, in the affirmative.

Proof of Theorem 3.1.

We use a method we developed in [14]. Let k and m be any two divisors of n such that k<m. Let uk and um be a Gk- and Gm-minimum solution, respectively. We divide the proof into two cases.

(i)  Assume that k is a divisor of m. Then GkGm and Gk(x)Gm(x) in D¯, where D:=HPn(1). Hence, Ω=HPn(ε)D for 0<ε<1. By Theorem 2.4, IGk<IGm, i.e., I(uk)<I(um). Since any orthogonal transformation leaves I() invariant, uk is not equivalent to um.

(ii)  Assume that k is not a divisor of m. Suppose on the contrary that uk is equivalent to um. Then uk(x)=um(h0x) for a certain h0H, where

H := { h O ( 2 ) : h ( Ω ) = Ω } .

Then uk is h0-1Gmh0-invariant because for gGm,

u k ( h 0 - 1 g h 0 x ) = u m ( g h 0 x ) = u m ( h 0 x ) = u k ( x ) .

Therefore, uk is both Gk- and h0-1Gmh0-invariant, and hence it is invariant under the group Gkh0-1Gmh0, where A denotes a group generated by A. By [14, Lemma 4.2], we have that h-1Gmh=Gm for any hO(2). Accordingly, uk is GkGm-invariant. Let r be the least common multiple of k and m. Then there exist μ,ν such that μ/k+ν/m=1/r. This shows that GkGm=Gr. Therefore, uk is Gr-invariant. However, k is a divisor of r, which contradicts (i). Therefore, uk is not equivalent to um. The proof is complete. ∎

Theorem 3.1 remains valid for the two-dimensional annulus Ω=A2(a,ε). Hence, we have the next result.

Corollary 3.2.

Let Ω=A2(a,ε) be the two-dimensional annulus. Then the number of non-equivalent positive solutions for (1.1) diverges to infinity as ε+0.

Let us consider Problem 1.3. Let Pn be a regular n-gon. We call Pn an even (resp. odd) polygon if n is even (resp. odd). Instead of a middle point of a side, we say a midpoint for simplicity. For an odd polygon, an axis of reflection symmetry goes through one vertex and one midpoint. For an even polygon, an axis of symmetry passes through either two vertices or two midpoints. Recall that for an axis of symmetry L, we call u(x) an L-solution if it is an HL-minimum solution, where HL is a reflection group with respect to L. Let n be odd and L1 and L2 be two axes of symmetry for Pn. Choose a rotation matrix gSO(2) satisfying g(L2)=L1. Then g leaves Pn invariant. Let u(x) be an L1-solution. Then v(x):=u(gx) is an L2-solution. Hence, the L1-solution u is equivalent to the L2-solution v. Consequently, for an odd polygon, we cannot find L1 and L2 satisfying the condition of Problem 1.3. Consider an even polygon. Choose the center of Pn as the coordinate origin 0. For a point x0, we denote the line through x and the origin by Lx. For a vertex v and a midpoint m, Lv and Lm are axes of symmetry. For any two vertices v and v, an Lv-solution is equivalent to a certain Lv-solution for the same reason given in the odd polygon case. Similarly, an Lm-solution is equivalent to an Lm-solution for the midpoints m and m. Thus, our target is Lv and Lm in an even polygon. Our answer to Problem 1.3 is as follows.

Theorem 3.3.

Let Ω:=HP4n(ε) be a hollow regular 4n-gon, where ε>0 is small enough and n is a positive integer. Then no Lv-solution is equivalent to any Lm-solution for any vertex v and any midpoint m.

The theorem above comes from Theorem 2.4 with the help of elementary Euclidean geometry. To prove the theorem, we first prove the following lemma.

Lemma 3.4.

Let Pn be an even polygon. Then the following assertions hold:

  1. If n is even and not a multiple of 4 , then there exist a vertex v and a midpoint m such that Lv is perpendicular to Lm.

  2. If n is a multiple of 4 , then for any vertex v and any midpoint m, Lv is not perpendicular to Lm.

Proof.

We give a rigorous proof by using the complex plane. Put Pn in the complex plane so that all its vertices lie on the unit circle |z|=1 and the point z=1 is one of them. We consider P4n with a positive integer n. Then all the vertices are written as Vk=e2kπi/4n, with k=0,1,2,,4n-1. Note that V0 is the point z=1. Denote the origin by O. If V is a point on the unit circle such that OV¯OV0¯, i.e., the line segment OV¯ is perpendicular to the x-axis, then V is equal to either eπi/2 or e3πi/2. Such points are achieved at Vk=e2kπi/4n for k=n and k=3n. These are vertices of P4n. Therefore, for any midpoint M of P4n, the segment OM¯ is not perpendicular to OV0¯. Thus, assertion (ii) holds.

Consider P2n with an odd integer n3. Then all the vertices are represented as Vk=e2πki/2n, with k=0,1,2,,2n-1. Denote all the midpoints of P2n by Mk, with k=0,1,,2n-1. Since a midpoint does not lie on the unit circle, we put Nk:=Mk/|Mk| so that Nk=e2π(2k+1)i/4n, with k=0,1,,2n-1. Recall that n is odd. The points Nk, with 2k+1=n,3n lie on the y-axis. This proves assertion (i). The proof is complete. ∎

For two straight lines in the plane 2, we have the next lemma.

Lemma 3.5.

Let L1 and L2 be two different lines passing through the origin, which are not perpendicular to each other. Let G be a group generated by reflections with respect to L1 and L2. Then G(x) includes at least three points for x0, where G(x) is the orbit defined by (2.1).

Proof.

The lemma follows from elementary geometry, nevertheless we give a proof. Let x be a point that does not lie on L1 and L2. The reflection with respect to L1 (resp. L2) moves x to another point x1 (resp. x2). Then xxi, i=1,2, and x1x2. Thus, we have at least three points.

Let xL1 and x0. Then it is not on L2. A reflection with respect to L2 moves x to another point x. Then x does not lie on the line L1 because L1 and L2 are not perpendicular to each other. The reflection by L1 moves x to another point x′′. Since xL1, x′′ is not equal to x. Hence, we have at least three points, x, x and x′′, which are different from each other. In the case where xL2, the argument above remains valid. The proof is complete. ∎

Throughout this section, |A| stands for the cardinal number of a set A. In the lemma above, the conclusion |G(x)|3 follows from the assumption that two lines are not perpendicular. Indeed, if L1L2, then |G(x)|=2 for x(L1L2){0}. In Lemma 3.5, the estimate |G(x)|3 is optimal. Indeed, there exists an example of L1, L2 and x where we have |G(x)|=3. We choose the x-axis as L1, and take L2 so that the angle between L1 and L2 is 60. Choose a point (complex number) z=1 on L1. Move it by all compositions of reflections by L1 and L2. Then G(1)={1,e2πi/3,e4πi/3} and |G(1)|=3. Thus, 3 is optimal.

Some readers may have an interest in |G(x)|. Let us investigate it, which is of independent interest. Assume that L1 and L2 pass through the origin. Let α be the angle between L1 and L2. Then the following assertions hold:

  1. | G ( x ) | = for any x0 if and only if α/π is an irrational number.

  2. | G ( x ) | < for any x0 if and only if α/π is a rational number.

  3. | G ( x ) | = 3 if and only if α=π/3 (or 2π/3) and x(L1L2L3){0}. Here L3 is a line through the origin whose angle between Li is π/3 for i=1,2.

To show the claims above, we choose L1 as the x-axis. Let fi be a function in the complex plane which is defined by the reflection with respect to Li for i=1,2. Then f1(z)=z¯=re-iθ and f2(z)=rei(2α-θ) for z=reiθ. Since G is all multiple compositions of f1 and f2, we have

G ( z ) = { r e i ( 2 n α + θ ) : n } { r e i ( 2 n α - θ ) : n }

for z=reiθ. This expression proves (i), (ii) and (iii).

We consider a reflection group with respect to a straight line. It is easy to verify that

{ g O ( 2 ) : det g = - 1 } = { ρ ~ ( θ ) : θ } , ρ ~ ( θ ) := ( cos θ sin θ sin θ - cos θ ) .

From a direct computation, it follows that ρ~(θ)2=e, with e being the unit matrix. Hence, ρ~(θ) is a reflection matrix. In fact, it represents a reflection with respect to a line whose angle between the x-axis is θ/2. Therefore, a subgroup H of O(2) is a reflection group with respect to a line if and only if H={e,g}, with a matrix gO(2) satisfying detg=-1. We shall show Theorem 3.3.

Proof of Theorem 3.3.

Let Ω:=HP4n(ε). Recall that the center of P4n is the coordinate origin 0 and for a point x0, Lx denotes a line through x and 0. Let v be any vertex and m be any midpoint of P4n. Let uv and um be an Lv-solution and an Lm-solution, respectively. That is, uv is a minimum energy solution in all positive solutions having reflection symmetry with respect to Lv. Denote the reflection group with respect to Lv or Lm by Hv or Hm, respectively. We shall show that uv is not equivalent to um. Suppose on the contrary that they are equivalent. Then uv(x)=um(kx) for a certain kK, where

K := { k O ( 2 ) : k ( Ω ) = Ω } .

By using the same method as in the proof of Theorem 3.1, we see that uv is k-1Hmk-invariant. Since Hm is a reflection group, it takes the form Hm={e,h} for a certain matrix hO(2) satisfying deth=-1. Then k-1Hmk={e,k-1hk} and det(k-1hk)=-1. Hence, k-1Hmk is also a reflection group. Since k-1 leaves P4n invariant, it moves the line Lm to another line passing through a midpoint. Denote it by Lm, that is, k-1(Lm)=Lm. We shall show that each point in Lm is a fixed point of k-1Hmk. Note that h(x)=x for xLm because Hm={e,h} is a reflection group with respect to Lm. For xLm, we can write x=k-1y with yLm, and so k-1hk(x)=k-1h(y)=k-1(y)=x. Thus, k-1Hmk does not move any point in Lm, and therefore it is a reflection with respect to Lm, that is, k-1Hmk=Hm. Consequently, uv is invariant under the action of both Hv and Hm. Let G be a group generated by the union of Hv and Hm. Then u is G-invariant. Moreover, HvG and Hv(x)G(x). Since Lv and Lm are not perpendicular to each other by Lemma 3.4, we have that |G(x)|3 for any x0 by Lemma 3.5. On the other hand, |Hv(x)|2 for any x0 because Hv is a reflection. Thus, Hv(x)G(x) for x0. Theorem 2.4 proves that IHv<IG. However, uv is an Hv-minimum solution and it is G-invariant. A contradiction occurs. Therefore, no Lv-solution is equivalent to any Lm-solution. The proof is complete. ∎

In Theorem 3.1, we considered the hollow polygon. Let us generalize it to regular polytopes in N. It is known that all the regular polytopes are the tetrahedron, the cube, the octahedron, the dodecahedron and the icosahedron in 3; the 5-cell, the 8-cell, the 16-cell, the 24-cell, the 120-cell and the 600-cell in 4; and the (N+1)-cell, the 2N-cell and the 2N-cell in N for N5 (see [7]). Let P be an interior of any regular polytope in N. Put

HP ( ε ) := ( 1 + ε ) P P ¯ .

This is a hollow regular polytope. Define

SO ( P ) := { g SO ( N ) : g ( P ) = P } ,

which is called a regular polytope group. Clearly, the orbit SO(P)(x) has at least two points for any x0. Hence, Theorem 2.6 ensures the next result.

Corollary 3.6.

For ε>0 small enough, no global minimum solution for HP(ε) is SO(P)-invariant.

Define

(3.1) O ( P ) := { g O ( N ) : g ( P ) = P } .

In the literature, the definition above is often referred to a regular polytope group. Since SO(P)O(P), the SO(P)-non-invariance implies the O(P)-non-invariance. So, Corollary 3.6 is valid for O(P) as well.

Corollary 3.6 remains valid for semiregular polytopes, with SO(P) replaced by O(P). A polytope P is called semiregular if it is a convex polytope whose faces are (N-1)-dimensional regular polytopes and its symmetric group is transitive on the set of vertices. More precisely, for any vertices x and y of P, there exists an isometric transformation which leaves P invariant and maps x to y (see [20, p. 15]). Therefore, all vertices of a semiregular polytope lie on a certain sphere. Choose the center of this sphere as the coordinate origin. Then the symmetric group of P is defined by (3.1). Accordingly, for any vertices x, y, there exists a gO(P) such that gx=y.

Let E be the union of all edges of a semiregular polytope P. Let SP(ε) be an ε-neighborhood of E. This is a skeleton semiregular polytope.

Theorem 3.7.

Let P be an interior of a semiregular polytope centered at the origin in RN. Let Ω be either the hollow semiregular polytope HP(ε) or the skeleton semiregular polytope SP(ε). For ε>0 small enough, no global minimum solution is O(P)-invariant.

For example, consider a soccer ball (a truncated icosahedron) in 3 . Let P be the interior of a soccer ball centered at the origin, and define Ω:=(1+ε)PP¯. This is a hollow soccer ball. For ε>0 small enough, no global minimum solution is invariant under the soccer ball group. The soccer ball is constructed by truncating vertices from the icosahedron. Therefore, the soccer ball group is equal to the icosahedral group.

Proof of Theorem 3.7.

Put G:=O(P). If we could prove that G(x) has at least two points for x0, then Theorem 2.6 proves Theorem 3.7. Although this claim seems clear, we give a proof to make the paper rigorous. Suppose on the contrary that G(x0)={x0} at some x00. Let M be the orthogonal complement of x0, which is defined by

M := { y N : ( x 0 , y ) = 0 } ,

where (x0,y) is a standard inner product in N. Since gx0=x0, we have that g-1x0=x0. Then

( g y , x 0 ) = ( y , g t x 0 ) = ( y , g - 1 x 0 ) = ( y , x 0 ) = 0

for gG and yM, where gt is the transpose of g. Therefore, g(M)M for gG. This shows that g(M)=M for gG. Let x1 be a vertex of P. Choose λ so that x1λx0+M, where λx0+M is a hyperplane defined by

λ x 0 + M := { λ x 0 + y : y M } .

Since g(M)=M and gx0=x0 for gG, it follows that g(λx0+M)=λx0+M. Denote the set of all vertices by V. Since G is transitive on V (this is a definition of a semiregular polytope), it is written as

V = { g x 1 : g G } .

Since x1λx0+M and g(λx0+M)=λx0+M, we have that Vλx0+M. Taking the convex hull of both sides, we have Pλx0+M. This is impossible because dimM=N-1. The proof is complete. ∎

Example 3.8.

We give an easy example of G and Ω. Let G:={e,-e}, with e being the unit matrix. In this case, a G-invariant function is even, i.e., u(-x)=u(x). Let D be a bounded domain whose closure does not contain the origin and which is point symmetric with respect to the origin, that is, xD implies -xD. Let Ω be a point symmetric subdomain of D for which λ1(Ω) is large enough. Then no global minimum solution is even. This assertion follows from Theorem 2.6 because G(x)={x,-x} for any xD¯.

We shall give two examples of H, G and Ω and in these examples, we shall solve Problems 1.5 and 1.6.

Example 3.9.

Let us consider Problem 1.5. We generalize it to a hollow regular n-gon. Let Ω:=HPn(ε) be a hollow regular n-gon with ε>0 small enough and let L be an axis of reflection symmetry of Ω. We choose the center of the n-gon as the coordinate origin. Let H be the reflection group with respect to L. Let G be a group generated by the union of H and the rotations of 2πj/n, with j=0,1,,n-1, that is, G is a dihedral group. It is clear that HG and H(x)G(x) for xD¯, where D:=HPn(1). We shall show that H(x)G(x) for xD¯. For x0, |H(x)|2 because H is a reflection. However, |G(x)|n3 because G includes rotations of 2πj/n, with j=0,1,,n-1. Therefore, H(x)G(x) for xD¯. By Theorem 2.4, no H-minimum solution is G-invariant. Therefore, Problem 1.5 is solved in the affirmative.

Example 3.10.

We consider Problem 1.6. Define HCN(ε) and HCPN(ε) by (1.8) and (1.9), respectively, and SCN(ε) and SCPN(ε) by the ε-neighborhood of the union of all the edges of CN and CPN, respectively. Let H and G be as defined after (1.8), i.e., H is the set of gO(N), each element of which is equal to 0 or 1, and G is the set of gO(N) each element of which is equal to 0, 1 or -1. We shall show that H(x)G(x) for x0. Since HG, we have that H(x)G(x). Let x=(x1,,xN)0 be any point. Since x0, at least one element xi is not 0. Suppose that it is positive. For any yH(x), at least one element yj is positive because (y1,,yN) is a permutation of (x1,,xN). However, there exists a point in G(x), all elements of which are non-positive because we can define it by (τ1x1,,τNxN), where τi:=-1 if xi>0 and τi=1 if xi0. Thus, H(x)G(x). By Theorem 2.4, no H-minimum solution is G-invariant. Choose an H-minimum solution. Then it satisfies (1.10) but does not satisfy (1.11). Consequently, Problem 1.6 is solved in the affirmative. Moreover, in the annulus AN(a,ε), an H-minimum solution and a G-minimum solution are not radially symmetric. Indeed, since they are finite groups, Corollary 2.11 ensures the assertion above.

4 Existence of G-Minimum Solutions

In this section, we shall prove the existence of G-minimum solutions. We denote the Lq(Ω) norm of u by uq. We employ the norm up in the space W01,p(Ω) because of the Poincaré inequality. Throughout this section, we suppose Assumption 2.2. Recall that f(x,u) is extended as an odd function of u. We shall investigate the sign of J(tu) as t varies in (0,).

Lemma 4.1.

For each uW01,p(Ω){0}, there exists a unique μ>0 such that J(tu)>0 for t(0,μ), J(μu)=0, and J(tu)<0 for t(μ,). Therefore, μu belongs to the Nehari manifold N.

Proof.

Since f(x,u)/uq-1 is increasing with respect to u by (2.2), we have

0 < f ( x , u ) u q - 1 f ( x , 1 ) C for  0 < u < 1 , x Ω ¯ ,

for a constant C>0. Accordingly, |f(x,u)|C|u|q-1 for |u|1 because f(x,u) is odd with respect to u. This inequality with Assumption 2.2 (iii) shows that

(4.1) | f ( x , u ) | C ( | u | q - 1 + | u | r - 1 ) for  u , x Ω ,

for some C>0. Recall that f(x,u) is odd with respect to u, and uf(x,u)>0 for u0. Let uW01,p(Ω){0}. Then J(tu), for t>0, is written as

J ( t u ) = t p Ω ( | u | p - f ( x , t | u | ) | t u | p - 1 | u | p ) 𝑑 x .

By (4.1), we have

Ω f ( x , t | u | ) | t u | p - 1 | u | p 𝑑 x C Ω ( t q - p | u | q + t r - p | u | r ) 𝑑 x 0 as  t + 0 .

Accordingly, J(tu)>0 for small t>0.

We shall show that J(tu)<0 for t>0 large enough. By Assumption 2.2 (i) and (ii), we have a constant c0>0 such that

(4.2) f ( x , u ) u q - 1 f ( x , 1 ) c 0 for  u 1 .

This shows that

min x Ω ¯ f ( x , u ) u - ( p - 1 ) as  u .

Let uW01,p(Ω){0}. For ε>0, we put

E := { x Ω : | u ( x ) | > ε } .

Since u(x)0, we can choose ε>0 so small that the Lebesgue measure of E (denoted by |E|) is positive. Then

Ω f ( x , t | u | ) | t u | p - 1 | u | p d x E f ( x , t ε ) ( t ε ) p - 1 ε p d x ε p | E | min x Ω ¯ f ( x , t ε ) ( t ε ) - ( p - 1 ) as  t .

Thus, J(tu)<0 for t>0 large. Since f(x,u)/up-1 is strictly increasing with respect to u, J(tu) has a unique zero μ>0. Accordingly, J(tu)>0 in (0,μ), and J(μu)=0 and J(tu)<0 in (μ,). The proof is complete. ∎

In view of Lemma 4.1, for uW01,p(Ω){0}, we define μ(u) by μ>0 satisfying J(μu)=0.

Lemma 4.2.

μ ( ) is continuous.

Proof.

Let un converge to u0W01,p(Ω){0}. Define μ¯:=lim infnμ(un). If μ¯>0, for μ(0,μ¯), we have that J(μun)>0 for all large n, by Lemma 4.1. As n, we have J(μu0)0. If μ>μ¯, then J(μun)<0 along a subsequence. Hence, J(μu0)0. Consequently, J(μu0)0 when μ<μ¯ and J(μu0)0 when μ>μ¯. Thus, μ¯=μ(u0). The same argument shows that lim supnμ(un)=μ(u0). Therefore, μ() is continuous. ∎

Let Lp(Ω,G) and C0(Ω,G) denote the set of G-invariant functions in Lp(Ω) and C0(Ω), respectively. The space W01,p(Ω,G) has already been defined in Section 1. Then the following lemma holds.

Lemma 4.3.

Let G be a closed subgroup of O(N) and let Ω be a G-invariant open set in RN. Then C0(Ω,G) is dense in Lp(Ω,G) and W01,p(Ω,G).

Proof.

In [12, Lemma 3.1], it is proved that C0(Ω,G) is dense in Lp(Ω,G). The same method, as in that lemma, is still valid for proving the density in W01,p(Ω,G). ∎

By Lemmas 4.1 and 4.3, 𝒩 and 𝒩(G) are not empty. Therefore, I0 and IG are well defined. We shall show that 𝒩 is bounded away from the origin.

Lemma 4.4.

There exists a constant c>0 such that upc for uN. Moreover, I0, given by (1.2), is positive.

Proof.

Let q and r be the exponents given in Assumption 2.2. From (4.2) and Assumption 2.2 (iii), it follows that qr, and hence p<qr. By this inequality and (4.1), for any ε>0, there exists a constant C>0 such that |f(x,u)|ε|u|p-1+C|u|r-1 for u and xΩ¯. Using the Sobolev embedding, we estimate J(u):

J ( u ) = u p p - Ω f ( x , u ) u 𝑑 x u p p - ε u p p - C u r r u p p - ε C u p p - C u p r

for some C>0. Let u𝒩. Since J(u)=0, we have

C u p r - p 1 - ε C .

Choose ε>0 so small that the right-hand side is positive. Hence, we have the first assertion of the lemma.

It follows from (2.2) that fu(x,u)u>(q-1)f(x,u). Integrating it over [0,u], we have F(x,u)(1/q)uf(x,u) for u0. This inequality is still valid for u because both sides are even with respect to u. Let u𝒩. Since J(u)=0 and upc, we get

I ( u ) 1 p u p p - 1 q Ω u f ( x , u ) 𝑑 x = 1 p u p p - 1 q u p p ( q - p p q ) c p .

Therefore, I0((q-p)/pq)cp. The proof is complete. ∎

In the next lemma, we shall show the existence of a global minimum solution only. However, the same method is applicable to a G-minimum solution with the help of the principle of symmetric criticality by Palais [27].

Lemma 4.5.

There exists a global minimum solution. Any global minimum solution u is a critical point of I, and it becomes a positive solution of (1.1) after replacing u by -u if necessary.

Proof.

The lemma can be proved in a standard method (we refer the readers to [31, Theorems 4.2 and 4.3]). By the strong maximum principle, a global minimum solution u must be positive after replacing u by -u if necessary. ∎

5 Proof of the Main Theorem

In this section, we shall prove Theorem 2.4 only, which readily yields all the other theorems in Section 2. We always suppose Assumptions 2.1 and 2.2. For a G-invariant bounded domain Ω, we have already defined Lp(Ω,G) and H01(Ω,G), which are G-invariant function spaces in Lp(Ω) and H01(Ω), respectively. Here we denote the orthogonal complement of L2(Ω,G) in L2(Ω) by L2(Ω,G), and the orthogonal complement of H01(Ω,G) in H01(Ω) by H01(Ω,G), that is,

L 2 ( Ω , G ) := { u L 2 ( Ω ) : ( u , v ) L 2 = 0  for  v L 2 ( Ω , G ) } ,
H 0 1 ( Ω , G ) := { u H 0 1 ( Ω ) : ( u , v ) H 0 1 = 0  for  v H 0 1 ( Ω , G ) } .

Here the inner products are defined by

( u , v ) L 2 := Ω u ( x ) v ( x ) 𝑑 x , ( u , v ) H 0 1 := Ω u ( x ) v ( x ) 𝑑 x .

We present four lemmas below, which have been proved in [14].

Lemma 5.1 ([14, Lemma 5.2]).

Let Ω be a G-invariant bounded domain. Then the following assertions hold:

  1. If u L p ( Ω , G ) and v L q ( Ω ) L 2 ( Ω , G ) , with 1 / p + 1 / q = 1 and 1 p , then Ω u ( x ) v ( x ) 𝑑 x = 0 .

  2. H 0 1 ( Ω , G ) L 2 ( Ω , G ) .

We here introduce a Haar measure. Since a closed subgroup G of O(N) is a compact Lie group, it has a unique Haar measure dg which satisfies

G f ( g ) 𝑑 g = G f ( g g ) 𝑑 g = G f ( g g ) 𝑑 g = G f ( g - 1 ) 𝑑 g , G f ( g ) 𝑑 g > 0 if  f 0 , f 0 , G 1 𝑑 g = 1 ,

for any gG and any real valued integrable function f on G. For the Haar measure, we refer to [28] or [24, Chapter 2].

Let M(N) be the set of all N×N real matrices. We define the norm by

g := max | x | 1 | g x | for  g M ( N ) ,

where |x| is the standard Euclidean norm. Let G be a closed subgroup of O(N). For hG and r>0, we define a ball B(h,r;G) in G by

B ( h , r ; G ) := { g G : g - h < r } .

We define the volume of B(h,r;G) (denoted by |B(h,r;G)|) by

| B ( h , r ; G ) | := B ( h , r ; G ) 1 𝑑 g ,

where dg is the Haar measure of G.

Lemma 5.2 ([14, Lemma 5.6]).

The volume |B(h,r;G)| is independent of h.

Let H and G be closed subgroups of O(N) satisfying HG. Since H and G are compact Lie groups, we can define

Q ( x , g ) := min h H | g x - h x | , P ( x ) := max g G Q ( x , g ) .

Lemma 5.3 ([14, Lemma 5.5]).

We have that

| P ( x ) - P ( y ) | 2 | x - y | for  x , y N .

By Assumption 2.1, for any xD¯, there exists a gG such that hxgx for any hH. This assertion is equivalent to a condition that P(x)>0 for xD¯. By Lemma 5.3, P(x) is continuous. Hence, the minimum of P(x) on D¯ is positive and we define

(5.1) ε := 1 4 min D ¯ P ( x ) > 0 .

This definition implies that for any xD¯, there exists a gG such that

(5.2) | g x - h x | 4 ε for all  h H .

Denote a ball in N centered at x with radius r>0 by B(x,r). Let ε>0 be defined by (5.1). Since D¯ is compact, we can take a finite covering B(xi,ε/4) with some x1,,xkN such that

(5.3) D ¯ i = 1 k B ( x i , ε / 4 ) ,

where k is the smallest integer satisfying the inclusion above, and hence it depends only on D and ε. Let Ω be a G-invariant subdomain of D satisfying λ1(Ω)>Λ, where Λ>0 will be determined later on. Let u be a G-minimum solution. We put u(x)=0 outside Ω. Choose a point y0N satisfying

sup y N B ( y , ε / 4 ) f ( x , u ) u 𝑑 x = B ( y 0 , ε / 4 ) f ( x , u ) u 𝑑 x .

If B(y0,ε/4)Ω=, then the right-hand side is zero and a contradiction occurs. Thus, we can choose a point x0B(y0,ε/4)Ω, which ensures that B(y0,ε/4)B(x0,ε/2). Then

(5.4) sup y N B ( y , ε / 4 ) f ( x , u ) u 𝑑 x B ( x 0 , ε / 2 ) f ( x , u ) u 𝑑 x .

Combining (5.3) and (5.4), we have

(5.5) Ω f ( x , u ) u 𝑑 x = D f ( x , u ) u 𝑑 x k B ( x 0 , ε / 2 ) f ( x , u ) u 𝑑 x .

Note that k is independent of Ω and u(x). We define a radially symmetric function Φ(x)C0(N) which satisfies 0Φ(x)1 in N,

(5.6) Φ ( x ) = 1 for  | x | ε , Φ ( x ) = 0 for  | x | 2 ε ,
(5.7) | Φ ( x ) | 2 / ε for  ε < | x | < 2 ε ,

where ε has been defined by (5.1). Let dg and dh be the Haar measures of G and H, respectively. We define

(5.8) ϕ ( x ) := G Φ ( x - g x 0 ) 𝑑 g - H Φ ( x - h x 0 ) 𝑑 h ,

where x0 has been defined before (5.4). Then ϕC0(N). Define

Ω := { x N : dist ( x , Ω ) < 2 ε } , dist ( x , Ω ) := inf { | x - y | : y Ω } .

Then ΩΩ. Since Ω is G-invariant, so is Ω. Since x0Ω, we have that hx0,gx0Ω for hH and gG. Therefore, the support of ϕ(x) is in Ω.

Lemma 5.4.

We have ϕC0(Ω)H01(Ω,G)H01(Ω,H).

Proof.

This lemma is proved in [14, Theorem 2.2]. ∎

Let u be a G-minimum solution and let x0Ω satisfy (5.4). Let ϕ be given by (5.8). We define

(5.9) g ( t , s ) := I ( ( 1 + t ) ( 1 + s ϕ ) u ) for  t , s .

Since uW01,p(Ω) and ϕC0(Ω), the function (1+sϕ)u belongs to W01,p(Ω), and therefore g(t,s) is well defined.

We shall explain our idea to prove Theorem 2.4. For a small s>0, we shall show that there exists a number t such that

( 1 + t ) ( 1 + s ϕ ) u 𝒩 ( H ) , I ( ( 1 + t ) ( 1 + s ϕ ) u ) < I ( u ) .

Then it follows that

I H I ( ( 1 + t ) ( 1 + s ϕ ) u ) < I ( u ) = I G .

Thus, we get Theorem 2.4. To accomplish the method above, we investigate the behavior of g(t,s) near (t,s)=(0,0).

Lemma 5.5.

Let u be a G-minimum solution and let ϕ be defined by (5.8). Define g(t,s) by (5.9). Then we have that

(5.10) g ( t , s ) = g ( 0 , 0 ) + t 2 2 g t t ( 0 , 0 ) + s 2 2 g s s ( 0 , 0 ) + o ( t 2 + s 2 )

as t,s0, where o(t2+s2)/(t2+s2)0 as t,s0, and gtt(0,0) denotes the second partial derivatives with respect to t. Moreover, we have the representations

(5.11) g t t ( 0 , 0 ) = Ω ( ( p - 1 ) f ( x , u ) u - f u ( x , u ) u 2 ) 𝑑 x ,
(5.12) g s s ( 0 , 0 ) = ( p - 1 ) Ω ( f ( x , u ) u ϕ 2 + | u | p - 2 | ϕ | 2 u 2 ) 𝑑 x - Ω f u ( x , u ) u 2 ϕ 2 𝑑 x .

Proof.

Since I(u)=0, we have

g t ( 0 , 0 ) = I ( u ) u = 0 , g s ( 0 , 0 ) = I ( u ) ϕ u = 0 .

The Taylor theorem shows that

(5.13) g ( t , s ) = g ( 0 , 0 ) + t 2 2 g t t ( 0 , 0 ) + t s g t s ( 0 , 0 ) + s 2 2 g s s ( 0 , 0 ) + o ( t 2 + s 2 )

as t,s0. By Assumption 2.2 (iii) and since p2, I(u) has a second derivative which is a bilinear form, i.e.,

I ′′ ( u ) [ v , w ] = Ω ( ( p - 1 ) | u | p - 2 v w - f u ( x , u ) v w ) 𝑑 x

for u,v,wW01,p(Ω). Noting that I(u)=0, we compute the second derivatives of g(t,s):

(5.14) g t t ( 0 , 0 ) = I ′′ ( u ) [ u , u ] = Ω ( ( p - 1 ) | u | p - f u ( x , u ) u 2 ) 𝑑 x ,
(5.15) g t s ( 0 , 0 ) = I ′′ ( u ) [ u , ϕ u ] + I ( u ) ϕ u = ( p - 1 ) Ω ( | u | p ϕ + | u | p - 2 ( u ϕ ) u ) 𝑑 x - Ω f u ( x , u ) u 2 ϕ 𝑑 x ,

where uϕ=i=1N(u/xi)(ϕ/xi), and

g s s ( 0 , 0 ) = I ′′ ( u ) [ ϕ u , ϕ u ]
= ( p - 1 ) Ω ( | u | p ϕ 2 + 2 | u | p - 2 ( u ϕ ) u ϕ ) 𝑑 x
(5.16) + Ω ( ( p - 1 ) | u | p - 2 | ϕ | 2 u 2 - f u ( x , u ) u 2 ϕ 2 ) 𝑑 x .

We shall show that gts(0,0)=0. Put u(x)=0 outside Ω. Then u belongs to W01,p(Ω,G). Accordingly, |u|pL1(Ω,G). Since ϕH01(Ω,G) by Lemma 5.4, it belongs to L2(Ω,G) by Lemma 5.1. We use Lemma 5.1 (i) to obtain

Ω | u | p ϕ d x = Ω | u | p ϕ d x = 0 ,

which is the first term on the right-hand side of (5.15). In the same way, we have

Ω f u ( x , u ) u 2 ϕ 𝑑 x = 0 .

Therefore, the third term of (5.15) vanishes. If uC0(Ω,G), then we have

Ω | u | p - 2 ( u ϕ ) u d x = - Ω div ( | u | p - 2 u u ) ϕ d x = 0

because div(|u|p-2uu) is G-invariant. The space C0(Ω,G) is dense in W01,p(Ω,G) by Lemma 4.3. For uW01,p(Ω,G), we choose a sequence un in C0(Ω,G) converging to u in W01,p(Ω,G). Substituting un in the equation above and letting n, we obtain

Ω | u | p - 2 ( u ϕ ) u d x = 0 ,

which is the second term on the right-hand side of (5.15). Consequently, gts(0,0)=0. Hence, (5.13) is reduced to (5.10).

We shall prove the expressions (5.11) and (5.12). Multiplying (1.1) by u and integrating it over Ω, we get

(5.17) Ω | u | p d x = Ω f ( x , u ) u d x .

Substituting this relation into (5.14), we obtain (5.11). Multiplying (1.1) by uϕ2 and integrating it, we have

Ω ( | u | p ϕ 2 + 2 | u | p - 2 ( u ϕ ) u ϕ ) 𝑑 x = Ω f ( x , u ) u ϕ 2 𝑑 x .

Substituting this relation into (5.16), we get (5.12). The proof is complete. ∎

We are now in a position to prove Theorem 2.4.

Proof of Theorem 2.4.

Let u be a G-minimum solution and let x0Ω satisfy (5.4). For this x0, by (5.2), there exists a g0G such that

| g 0 x 0 - h x 0 | 4 ε for  h H .

Hence, |x-hx0|2ε for xB(g0x0,2ε) and hH. This inequality and (5.6) show that

Φ ( x - h x 0 ) = 0 for  x B ( g 0 x 0 , 2 ε ) , h H ,

which, by (5.8), implies that

(5.18) ϕ ( x ) = G Φ ( x - g x 0 ) 𝑑 g for  x B ( g 0 x 0 , 2 ε ) .

Since D is bounded, there exists a constant R>0 such that |x|R for xD. We define ν:=ε/(2R) and d:=|B(g,ν;G)|. By Lemma 5.2, d is independent of gG. For xB(g0x0,ε/2) and gB(g0,ν;G), we have that

| x - g x 0 | | x - g 0 x 0 | + | g 0 x 0 - g x 0 | < ε / 2 + g 0 - g | x 0 | < ε .

This inequality and (5.6) prove that

Φ ( x - g x 0 ) = 1 for  x B ( g 0 x 0 , ε / 2 ) , g B ( g 0 , ν ; G ) ,

which, by (5.18), implies that

(5.19) ϕ ( x ) B ( g 0 , ν ; G ) Φ ( x - g x 0 ) 𝑑 g | B ( g 0 , ν ; G ) | = d > 0

for xB(g0x0,ε/2). Thus, ϕ0 in Ω. We define g(t,s) by (5.9). We shall show that gtt(0,0) and gss(0,0) are negative. We use (5.11) and (2.2) to obtain

g t t ( 0 , 0 ) = Ω ( ( p - 1 ) f ( x , u ) u - f u ( x , u ) u 2 ) 𝑑 x Ω ( ( p - 1 ) f ( x , u ) u - ( q - 1 ) f ( x , u ) u ) 𝑑 x < 0 ,

because u is a positive solution. We again write (5.12) as

(5.20) g s s ( 0 , 0 ) = ( p - 1 ) Ω ( f ( x , u ) u ϕ 2 + | u | p - 2 | ϕ | 2 u 2 ) 𝑑 x - Ω f u ( x , u ) u 2 ϕ 2 𝑑 x .

Since |Φ(x)|2/ε in N, by (5.6) and (5.7), we have

| ϕ ( x ) | G | Φ ( x - g x 0 ) | d g + H | Φ ( x - h x 0 ) | d h 4 / ε .

Using the inequality above and employing the Hölder inequality, we estimate the second term in (5.20):

Ω | u | p - 2 | ϕ | 2 u 2 d x 16 ε - 2 Ω | u | p - 2 u 2 d x 16 ε - 2 u p p - 2 u p 2 .

Since λ1(Ω) is the first eigenvalue, we have that λ1(Ω)uppupp. Using this inequality with the assumption λ1(Ω)>Λ and employing (5.17) and (5.5), we get

Ω | u | p - 2 | ϕ | 2 u 2 d x 16 ε - 2 Λ - 2 / p u p p = 16 ε - 2 Λ - 2 / p Ω f ( x , u ) u d x 16 k ε - 2 Λ - 2 / p B ( x 0 , ε / 2 ) f ( x , u ) u d x .

Since u and f are G-invariant, the right-hand side is equal to

16 k ε - 2 Λ - 2 / p B ( g 0 x 0 , ε / 2 ) f ( x , u ) u 𝑑 x .

Since ϕ(x)d in B(g0x0,ε/2) by (5.19), we have

Ω | u | p - 2 | ϕ | 2 u 2 d x 16 k d - 2 ε - 2 Λ - 2 / p B ( g 0 x 0 , ε / 2 ) f ( x , u ) u ϕ 2 d x 16 k d - 2 ε - 2 Λ - 2 / p Ω f ( x , u ) u ϕ 2 d x .

Denoting the coefficient of the last integral by a, we get

Ω | u | p - 2 | ϕ | 2 u 2 d x a Ω f ( x , u ) u ϕ 2 d x .

Using this inequality, we estimate (5.20) as

g s s ( 0 , 0 ) Ω ( ( p - 1 ) ( a + 1 ) f ( x , u ) u ϕ 2 - f u ( x , u ) u 2 ϕ 2 ) 𝑑 x
(5.21) - [ q - 1 - ( p - 1 ) ( a + 1 ) ] Ω f ( x , u ) u ϕ 2 𝑑 x ,

where we have used (2.2). Recall that a=16kd-2ε-2Λ-2/p. Here we choose Λ>0 so large that

(5.22) ( p - 1 ) [ 16 k d - 2 ε - 2 Λ - 2 / p + 1 ] < q - 1 .

Then the right-hand side of (5.21) is negative. We note that the constants k, d and ε in (5.22) depend only on G, H and D and are independent of Ω and u(x).

We have proved that gtt(0,0) and gss(0,0) are negative. It then follows from (5.10) that

g ( t , s ) = I ( ( 1 + t ) ( 1 + s ϕ ) u ) < g ( 0 , 0 ) = I ( u ) ,

when |t|,|s| are small enough and (t,s)(0,0). For vW01,p(Ω){0}, we recall that μ(v) is a unique positive number satisfying μ(v)v𝒩. Since u is a G-minimum solution, it belongs to 𝒩. Therefore, μ(u)=1. Since μ() is continuous by Lemma 4.2, μ((1+sϕ)u) converges to μ(u)=1 as s0. Define t(s):=μ((1+sϕ)u)-1. Then t(s)0 as s0. Since u is G-invariant and ϕ is H-invariant, (1+sϕ)u is H-invariant. Thus, we find that (1+t(s))(1+sϕ)u belongs to 𝒩(H). When s>0 is small enough, so is |t(s)|. Therefore, for s>0 small, we have

I H I ( ( 1 + t ( s ) ) ( 1 + s ϕ ) u ) < I ( u ) = I G .

The proof is complete. ∎


Communicated by Paul Rabinowitz


Award Identifier / Grant number: 16K05236

Funding statement: This work was supported by JSPS KAKENHI Grant Number 16K05236.

References

[1] D. Arcoya, Positive solutions for semilinear Dirichlet problems in an annulus, J. Differential Equations 94 (1991), 217–227. 10.1016/0022-0396(91)90090-VSearch in Google Scholar

[2] A. Borel, Le plan projectif des octaves et les sphères comme espaces homogènes, C. R. Acad. Sci. Paris 230 (1950), 1378–1380. Search in Google Scholar

[3] J. Byeon, Existence of many nonequivalent nonradial positive solutions of semilinear elliptic equations on three-dimensional annuli, J. Differential Equations 136 (1997), 136–165. 10.1006/jdeq.1996.3241Search in Google Scholar

[4] J. Byeon and K. Tanaka, Multi-bump positive solutions for a nonlinear elliptic problem in expanding tubular domains, Calc. Var. Partial Differential Equations 50 (2014), 365–397. 10.1007/s00526-013-0639-zSearch in Google Scholar

[5] D. Castorina and F. Pacella, Symmetry of positive solutions of an almost-critical problem in an annulus, Calc. Var. Partial Differential Equations 23 (2005), 125–138. 10.1007/s00526-004-0292-7Search in Google Scholar

[6] C. V. Coffman, A nonlinear boundary value problem with many positive solutions, J. Differential Equations 54 (1984), 429–437. 10.1016/0022-0396(84)90153-0Search in Google Scholar

[7] H. S. M. Coxeter, Regular Polytopes, 3rd ed., Dover, New York, 1973. Search in Google Scholar

[8] E. N. Dancer, On the number of positive solutions of some weakly nonlinear equations on annular regions, Math. Z. 206 (1991), 551–562. 10.1007/BF02571362Search in Google Scholar

[9] E. N. Dancer, Global breaking of symmetry of positive solutions on two-dimensional annuli, Differential Integral Equations 5 (1992), 903–913. 10.57262/die/1370955427Search in Google Scholar

[10] E. N. Dancer, Some singularly perturbed problems on annuli and a counterexample to a problem of Gidas, Ni and Nirenberg, Bull. Lond. Math. Soc. 29 (1997), 322–326. 10.1112/S0024609396002391Search in Google Scholar

[11] V. V. Gorbatsevich, A. L. Onishchik and E. B. Vinberg, Foundations of Lie Theory and Lie Transformation Groups, Springer, Berlin, 1997. Search in Google Scholar

[12] R. Kajikiya, Least energy solutions of the Emden–Fowler equation in hollow thin symmetric domains, J. Math. Anal. Appl. 406 (2013), no. 1, 277–286. 10.1016/j.jmaa.2013.04.068Search in Google Scholar

[13] R. Kajikiya, Positive solutions of the p-Laplace Emden–Fowler equation in hollow thin symmetric domains, Math. Bohem. 139 (2014), 145–154. 10.21136/MB.2014.143845Search in Google Scholar

[14] R. Kajikiya, Multiple positive solutions of the Emden–Fowler equation in hollow thin symmetric domains, Calc. Var. Partial Differential Equations 52 (2015), 681–704. 10.1007/s00526-014-0729-6Search in Google Scholar

[15] Y. Y. Li, Existence of many positive solutions of semilinear elliptic equations in annulus, J. Differential Equations 83 (1990), 348–367. 10.1016/0022-0396(90)90062-TSearch in Google Scholar

[16] S.-S. Lin, Positive radial solutions and non-radial bifurcation for semilinear elliptic equations in annular domains, J. Differential Equations 86 (1990), 367–391. 10.1016/0022-0396(90)90035-NSearch in Google Scholar

[17] S.-S. Lin, Existence of positive nonradial solutions for nonlinear elliptic equations in annular domains, Trans. Amer. Math. Soc. 332 (1992), 775–791. 10.1090/S0002-9947-1992-1055571-1Search in Google Scholar

[18] S.-S. Lin, Existence of many positive nonradial solutions for nonlinear elliptic equations on an annulus, J. Differential Equations 103 (1993), 338–349. 10.1006/jdeq.1993.1053Search in Google Scholar

[19] S.-S. Lin, Asymptotic behavior of positive solutions to semilinear elliptic equations on expanding annuli, J. Differential Equations 120 (1995), 255–288. 10.1006/jdeq.1995.1112Search in Google Scholar

[20] P. McMullen and E. Schulte, Abstract Regular Polytopes, Cambridge University Press, Cambridge, 2002. 10.1017/CBO9780511546686Search in Google Scholar

[21] N. Mizoguchi, Generation of infinitely many solutions of semilinear elliptic equation in two-dimensional annulus, Comm. Partial Differential Equations 21 (1996), 221–227. 10.1080/03605309608821181Search in Google Scholar

[22] N. Mizoguchi and T. Suzuki, Semilinear elliptic equations on annuli in three and higher dimensions, Houston J. Math. 22 (1996), 199–215. Search in Google Scholar

[23] D. Montgomery and H. Samelson, Transformation groups of spheres, Ann. of Math. (2) 44 (1943), 454–470. 10.2307/1968975Search in Google Scholar

[24] D. Montgomery and L. Zippin, Topological Transformation Groups, Robert E. Krieger, New York, 1974. Search in Google Scholar

[25] A. L. Onishchik, Topology of Transitive Transformation Groups, Johann Ambrosius Barth, Leipzig, 1994. Search in Google Scholar

[26] F. Pacard, Radial and nonradial solutions of -Δu=λf(u), on an annulus of n, n3, J. Differential Equations 101 (1993), 103–138. 10.1006/jdeq.1993.1007Search in Google Scholar

[27] R. S. Palais, The principle of symmetric criticality, Comm. Math. Phys. 69 (1979), 19–30. 10.1007/BF01941322Search in Google Scholar

[28] L. S. Pontrjagin, Topological Groups, 3rd ed., Gordon and Breach Science, New York, 1986. Search in Google Scholar

[29] T. Suzuki, Positive solutions for semilinear elliptic equations on expanding annuli: Mountain pass approach, Funkcial. Ekvac. 39 (1996), 143–164. Search in Google Scholar

[30] Z.-Q. Wang and M. Willem, Existence of many positive solutions of semilinear elliptic equations on an annulus, Proc. Amer. Math. Soc. 127 (1999), 1711–1714. 10.1090/S0002-9939-99-04708-5Search in Google Scholar

[31] M. Willem, Minimax Theorems, Birkhäuser, Berlin, 1996. 10.1007/978-1-4612-4146-1Search in Google Scholar

Received: 2017-02-09
Revised: 2017-05-18
Accepted: 2017-05-19
Published Online: 2017-06-20
Published in Print: 2018-04-01

© 2018 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 8.5.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2017-6023/html
Scroll to top button