
Half a century after the appearance of the celebrated paper by Serrin about overdetermined boundary value problems in potential theory and related symmetry properties, we reconsider semilinear polyharmonic equations under Dirichlet boundary conditions in the unit ball of Rn. We discuss radial properties (symmetry and monotonicity) of positive solutions of such equations and we show that, in conformal dimensions, the associated Green function satisfies elegant reflection and symmetry properties related to a suitable Kelvin transform (inversion about a sphere). This yields an alternative formula for computing the partial derivatives of solutions of polyharmonic problems. Moreover, it gives some hints on how to modify a counterexample by Sweers where radial monotonicity fails: we numerically recover strict radial monotonicity for the biharmonic equation in the unit ball of R4.
Citation: Filippo Gazzola, Gianmarco Sperone. Remarks on radial symmetry and monotonicity for solutions of semilinear higher order elliptic equations[J]. Mathematics in Engineering, 2022, 4(5): 1-24. doi: 10.3934/mine.2022040
[1] | Juan-Carlos Felipe-Navarro, Tomás Sanz-Perela . Semilinear integro-differential equations, Ⅱ: one-dimensional and saddle-shaped solutions to the Allen-Cahn equation. Mathematics in Engineering, 2021, 3(5): 1-36. doi: 10.3934/mine.2021037 |
[2] | Luigi Montoro, Berardino Sciunzi . Qualitative properties of solutions to the Dirichlet problem for a Laplace equation involving the Hardy potential with possibly boundary singularity. Mathematics in Engineering, 2023, 5(1): 1-16. doi: 10.3934/mine.2023017 |
[3] | Alberto Farina . Some results about semilinear elliptic problems on half-spaces. Mathematics in Engineering, 2020, 2(4): 709-721. doi: 10.3934/mine.2020033 |
[4] | Isabeau Birindelli, Giulio Galise . Allen-Cahn equation for the truncated Laplacian: Unusual phenomena. Mathematics in Engineering, 2020, 2(4): 722-733. doi: 10.3934/mine.2020034 |
[5] | Idriss Mazari . Some comparison results and a partial bang-bang property for two-phases problems in balls. Mathematics in Engineering, 2023, 5(1): 1-23. doi: 10.3934/mine.2023010 |
[6] | Giulia Bevilacqua . Symmetry break in the eight bubble compaction. Mathematics in Engineering, 2022, 4(2): 1-24. doi: 10.3934/mine.2022010 |
[7] | Yihong Du, Wenjie Ni . The Fisher-KPP nonlocal diffusion equation with free boundary and radial symmetry in R3. Mathematics in Engineering, 2023, 5(2): 1-26. doi: 10.3934/mine.2023041 |
[8] | Gerd Grubb . The principal transmission condition. Mathematics in Engineering, 2022, 4(4): 1-33. doi: 10.3934/mine.2022026 |
[9] | Francesca Colasuonno, Fausto Ferrari, Paola Gervasio, Alfio Quarteroni . Some evaluations of the fractional p-Laplace operator on radial functions. Mathematics in Engineering, 2023, 5(1): 1-23. doi: 10.3934/mine.2023015 |
[10] | Liliane Maia, Gabrielle Nornberg . Radial solutions for Hénon type fully nonlinear equations in annuli and exterior domains. Mathematics in Engineering, 2022, 4(6): 1-18. doi: 10.3934/mine.2022055 |
Half a century after the appearance of the celebrated paper by Serrin about overdetermined boundary value problems in potential theory and related symmetry properties, we reconsider semilinear polyharmonic equations under Dirichlet boundary conditions in the unit ball of Rn. We discuss radial properties (symmetry and monotonicity) of positive solutions of such equations and we show that, in conformal dimensions, the associated Green function satisfies elegant reflection and symmetry properties related to a suitable Kelvin transform (inversion about a sphere). This yields an alternative formula for computing the partial derivatives of solutions of polyharmonic problems. Moreover, it gives some hints on how to modify a counterexample by Sweers where radial monotonicity fails: we numerically recover strict radial monotonicity for the biharmonic equation in the unit ball of R4.
Let Ω⊂Rn (n≥2, possibly n=1971) be a bounded smooth domain and consider the problem
{−Δu=1inΩ,u=0,∂u∂ν=con∂Ω, | (1.1) |
where c∈R is a constant and ν denotes the outward unit normal to ∂Ω. Imposing both Dirichlet and Neumann conditions on ∂Ω makes the problem overdetermined so that, in general, (1.1) has no solution. In a celebrated paper published in 1971, Serrin [15] proved that if (1.1) admits a smooth solution, then Ω must necessarily be a ball. This paper has raised a great interest and, nowadays (half a century later), it has reached almost 600 citations in the Mathscinet. Serrin's original proof combines analytic arguments, such as the Maximum Principle and a refinement of Hopf's boundary Lemma, with geometric techniques such as the moving plane method inspired to the Alexandrov characterization of spheres [1,2]. Starting from [18], several different approaches have been devised as an alternative to Serrin's original proof, see [14] and the references therein for a fairly complete survey.
The moving plane method has been fruitfully used in symmetry results for semilinear elliptic equations, see [4,10] for second order equations and [3] for higher order problems. Let B⊂Rn be the unit ball and consider the semilinear polyharmonic problem under Dirichlet boundary conditions:
{(−Δ)mu=f(u)in B,u=∂u∂r=⋯=∂m−1u∂rm−1=0on ∂B. | (1.2) |
Here r=|x| denotes the radial variable and, hence, the outward normal direction to ∂B. The following result, valid for second order equations with m=1, is a restatement of [10,Theorem 1], combined with deep remarks by Spruck [10,Remark 1].
Theorem 1.1. ([4,10]) Let n≥2 and m=1. Assume that
eitherf∈Liploc(R+;R)orf∈C(R+;R)is nondecreasing. | (1.3) |
Then, every positive strong solution u∈W2,nloc(B)∩C(¯B) (u>0 in B) of (1.2) is radially symmetric (u=u(r)) and u′(r)<0 for all r∈(0,1). Moreover, if
f∈Liploc(R+;R)andf(0)≥0, | (1.4) |
then every nonnegative and non-trivial strong solution u∈W2,nloc(B)∩C(¯B) (u≥0, u≢0 in B) of (1.2) is radially symmetric (u=u(r)) and u′(r)<0 for all r∈(0,1).
This result was extended in [3,Theorem 1] to the case m≥2, in the following form:
Theorem 1.2. ([3]) Let n≥2 and m≥1. Assume that
f∈C(R+;R)is nondecreasing withf(0)≥0. | (1.5) |
Then, every nonnegative and non-trivial strong solution u∈Hm0(B)∩L∞(B) (u≥0, u≢0 in B) of (1.2) is radially symmetric (u=u(r)) and u′(r)<0 for all r∈(0,1).
In its original version, Theorem 1.2 was stated by requiring that the solution u was strictly positive, u>0 in B. But this is not necessary since [9,Theorem 5.1] ensures that if u≥0 is nontrivial, then u>0 (recall that f(u)≥0 in view of (1.5)). Assumption (1.5) is stronger than (1.3) and one may wonder whether Theorem 1.2 also holds under weaker assumptions. A nice example by Sweers [16] shows that, for a smooth and decreasing f, positive radial solutions may not be radially decreasing. Therefore, (1.3) is not enough to obtain the full statement of Theorem 1.1 (radial symmetry and monotonicity) in the higher order case m≥2. This discrepancy of assumptions between the cases m=1 and m≥2 is due to the lack of a maximum principle for polyharmonic operators; see [9,Section 1.2].
To find conditions, different from (1.5), ensuring both radial symmetry and radial monotonicity for (1.2) in the case m≥2 is a challenging problem. The first purpose of this paper is precisely to discuss some conditions on the source f which ensure the radial symmetry of the solutions of (1.2) and their radial monotonicity. In Section 2, particular attention is devoted to the regularity and monotonicity of f: for the biharmonic equation in B⊂Rn, in Theorem 2.3 we exhibit a Hölder-continuous function f for which the radial symmetry of solutions fails, while Proposition 2.4 displays a strictly decreasing and sign-changing f that still ensures the existence of a radially symmetric and strictly decreasing solution.
Contrary to the case of Navier boundary conditions, considered in [17], (1.2) cannot be reduced to a second order system when m≥2. This is why the moving plane procedure used in the proof of Theorem 1.2 was carried on by using fine estimates of the Green function G associated to the polyharmonic operator (−Δ)m, see [3,Section 2]. The second purpose of the present paper is to obtain new properties of the Green function in the so-called conformal dimensions n=2m. In Section 3 we prove that some estimates become explicit identities in conformal dimensions, see formula (3.6). Combined with the inversion in the sphere and the Kelvin transform, this enables us to obtain an elegant symmetry property of the Green function, see Theorem 3.2 and the sketchy representation in Figures 3 and 4. In turn, this result is used in Corollary 3.5 where we present an alternative formula for the computation of the partial derivatives of solutions of (1.2): since the sign of these partial derivatives is the fundamental feature to implement the moving plane procedure, this formula can become useful under suitable assumption on the source f. Indeed, based on this formula, Proposition 3.7 suggests that the radial symmetry and monotonicity of the solutions of (1.2) is ensured if ‖f′‖L∞(R+) is small.
Finally, as an application of Proposition 3.7, in Section 4 we revisit the counterexample by Sweers [16] where radial monotonicity of the solution fails: by appropriately modifying the source f (enforcing the conditions given in Proposition 3.7) we numerically obtain radially symmetric and strictly decreasing solutions of the biharmonic equation in the unit ball of R4 for a decreasing function f with sufficiently small derivative. Throughout this work, some open problems and questions are posed.
In this section we discuss radial properties (symmetry and monotonicity) of solutions of (1.2). Let us recall the weak formulation of (1.2) within the Sobolev space Hm0(B), which is a Hilbert space if endowed with the scalar product
(u,v)m={∫B(Δm/2u)(Δm/2v)if m is even∫B(∇Δ(m−1)/2u)⋅(∇Δ(m−1)/2v)if m is odd∀u,v∈Hm0(B). |
We denote the induced norm by ‖⋅‖m; in particular, ‖⋅‖0 is the L2(B)-norm. A function u∈Hm0(B) is called a weak solution of (1.2) if f(u)∈H−m(B) (the dual space of Hm0(B)) and (1.2) is satisfied in a weak sense, that is
(u,v)m=⟨f(u),v⟩for all v∈Hm0(B), |
where ⟨⋅,⋅⟩ stands for the duality product between H−m(B) and Hm0(B). In the present article, however, we will mostly deal with slightly more regular solutions: when f is continuous and u∈Hm0(B)∩L∞(B), we say that u is a strong solution of (1.2) if
(u,v)m=∫Bf(u)vfor all v∈Hm0(B); | (2.1) |
the integral exists since u∈L∞(B) and f is continuous. By elliptic regularity, any such strong solution u belongs to C2m−1,α(¯B), for some α∈(0,1), and all partial derivatives of order less than m vanish on ∂B. Moreover, if f is Hölder continuous, then u∈C2m,α(¯B) is a classical solution, see [9]. In the sequel, we always take (at least) f∈C(R;R).
Our first (elementary) result is a restatement of [3,Remark (iv)] and gives a different condition for radial symmetry (but not for radial monotonicity).
Theorem 2.1. Let n≥2 and m≥1. Let λ1 be the first Dirichlet eigenvalue for (−Δ)m in B. If
f∈C(R;R)satisfiesf(t)−f(s)t−s<λ1∀t>s, | (2.2) |
then there exists at most one strong solution u∈Hm0(B)∩L∞(B) of problem (1.2) which is, moreover, radially symmetric.
Proof. By contradiction, assume that (1.2) admits two strong solutions u1,u2∈Hm0(B)∩L∞(B). Then by (2.1) we have both that
(u1,v)m=∫Bf(u1)vand(u2,v)m=∫Bf(u2)vfor all v∈Hm0(B). |
Set w≐u1−u2, subtract these two equations and choose v=w as a test function to obtain
(w,w)m=∫B(f(u1)−f(u2))w. | (2.3) |
By (2.2) we know that
(f(u1(x))−f(u2(x)))w(x)≤λ1w(x)2∀x∈B, |
with strict inequality in a set of positive measure since w≢0 (recall u2≢u1). Therefore, (2.3) yields
‖w‖2m<λ1‖w‖20, |
which contradicts the Poincaré inequality and proves uniqueness.
Once uniqueness is established, it suffices to remark that if u is a nonradial solution of (1.2), then for any given nontrivial rotation A∈SO(n), also the function uA≐u∘A is a (different) strong solution of (1.2), which contradicts the just proved uniqueness statement.
We have so shown that, if (2.2) holds, then there exists at most a unique solution of (1.2) which is necessarily radially symmetric. This proves the theorem.
Theorem 2.1 states that one cannot expect radial monotonicity under the sole assumption (2.2). Note also that Theorem 2.1 does not require the positivity of the solution and its elementary proof is based on the invariance properties of polyharmonic operators under rotations, a fact that ensures uniqueness. Let us recall that symmetry is also ensured if f′(s)<λ2, for all s∈R (with λ2 being the second Dirichlet eigenvalue for (−Δ)m in B, see [6]), and uniqueness is guaranteed if f∈C(R+;R) is sublinear at 0 and +∞ (see [7]). All this raises a natural question:
Problem 2.2. Is radial symmetry related to uniqueness? Are there examples of multiple radial positive solutions? Except, of course, when f(u)=λ1u!
Some attention also deserves the regularity of f, as this feature is related to the existence of non-radial solutions. The counterexample in [10,Remark 1] shows that Theorem 1.1 (m=1) does not hold if the Lipschitz continuity assumption on f is relaxed to Hölder continuity. We extend this example to the biharmonic equation (m=2), showing that the monotonicity requirement in (1.5) cannot be dropped even in an arbitrarily small interval. We only deal with the biharmonic operator since the amount of computations grows very quickly as m increases.
Theorem 2.3. Let n≥2. For every ε>0 there exists a Hölder-continuous function gε:R+⟶R such that
gε(0)=0,gε(w)>−ε∀w∈R+,gε(ε)>0,gεis strictly increasing over[ε,∞), |
and such that the problem
{Δ2u=gε(u)inBu=∂u∂r=0on∂B, |
admits both
− a radial solution (u=u(r)) in C4(B), nonnegative (but not strictly positive) and radially decreasing (but not strictly radially decreasing): namely, u(r0)=0 and u′(r0)=0 for some r0∈(0,1);
− infinitely many nonnegative nontrivial solutions in C4(B) which are not radially symmetric.
Proof. For any p>4 define the function g:R+⟶R by
g(w)=4p(p−1)w1−4p[4(p−3)(p−2)−4(n(p−2)+2(p−2)2)w1/p+(n2+4(p−2)(p−1)+2n(2p−3))w2/p]∀w≥0. | (2.4) |
Then g∈C0,1−4p([0,∞);R)∩C∞((0,∞);R) and
g′(w)=4(p−2)(p−1)w−4/p[4(p−4)(p−3)+4(3(n−4)−2p2−p(n−3)+7p)w1/p+(n2−6n+8+4p2+4p(n−3))w2/p]∀w>0. |
The quantities between brackets in both the expressions of g(w) and g′(w) can be seen as second order polynomials in the variable w1/p. In particular, there must exist finite numbers M=M(n,p)>0, Γ1=Γ1(n,p)>0, Γ2=Γ2(n,p)∈(0,Γ1) such that
−M(n,p)≐min | (2.5) |
Two plots of g , for given values of p > 4 and n \geq 2 , are displayed in Figure 1.
Since m = 2 , the Eq (1.2) in radial coordinates becomes
\left\{\begin{array}{l} u^{(4)}(r)+\frac{2(n-1)}{r} u^{\prime \prime \prime}(r)+\frac{(n-1)(n-3)}{r^{2}} u^{\prime \prime}(r)\\-\frac{(n-1)(n-3)}{r^{3}} u^{\prime}(r)=g(u(r)) \quad \forall r \in(0,1) \\ u^{\prime}(0)=u^{\prime \prime \prime}(0)=u(1)=u^{\prime}(1)=0, \end{array}\right. | (2.6) |
where the conditions u'(0) = u'''(0) = 0 are needed to ensure the smoothness of u in {\bf B} , see [8]. By putting u(r) \doteq (1-r^2)^p , for r \in [0, 1] , one sees that such u satisfies u(1) = u'(1) = 0 (fulfilling the Dirichlet boundary conditions) and also the equation in (2.6), where g\big(u(r)\big) is as in (2.4). Hence,
u(x) = (1-\vert x\vert^2)^p\qquad \text{for $x \in {\bf B}$} |
is a \mathcal{C}^{4}(\overline{ {\bf B}}) -positive, radial and strictly radially decreasing solution ( u'(r) < 0 in (0, 1) ) of the problem
\left\{ \begin{aligned} & \Delta^2u = g(u)\quad\text{in} \; {\bf B} \\ & u = \frac{\partial u}{\partial r} = 0\quad\text{on} \;\partial {\bf B}. \end{aligned} \right. |
In order to obtain the statement, we need some rescaling. First we extend u to all \mathbb{R}^n by setting
\begin{equation} u(x) = \left \{ \begin{aligned} &(1-\vert x\vert^2)^p&&\qquad \text{if $x \in {\bf B}$}\\[3pt] &0 &&\qquad \text{if $x \not\in {\bf B}$}, \end{aligned} \right. \end{equation} | (2.7) |
so that u\in\mathcal{C}^{4}(\mathbb{R}^{n}) is nonnegative and nontrivial (but not strictly positive), radial and radially decreasing (but not strictly decreasing). We now argue as in [10,p. 220]: take any point x_{0} \in \mathbb{R}^{n} such that | x_{0} | = 3 and consider the function \hat{u} \in \mathcal{C}^{4}(\mathbb{R}^{n}) defined as
\begin{equation} \hat{u}(x) = u(x) + u(x-x_{0}) = \left \{ \begin{aligned} &(1-\vert x\vert^2)^p&&\qquad \text{if $| x | < 1$}\\[3pt] &(1-\vert x - x_{0} \vert^2)^p&&\qquad \text{if $ | x - x_{0} | < 1$}\\[3pt] &0 &&\qquad \text{otherwise,} \end{aligned} \right. \end{equation} | (2.8) |
which is a nonnegative solution of the problem
\left\{ \begin{aligned} & \Delta^2 \hat{u} = g(\hat{u})\quad\text{in} \; {\bf B}_5 \\ & \hat{u} = \frac{\partial \hat{u}}{\partial r} = 0\quad\text{on} \;\partial {\bf B}_5, \end{aligned} \right. |
where {\bf B}_5\subset \mathbb{R}^{n} is the (open) ball of radius 5 centered at the origin of \mathbb{R}^{n} , and g is given by (2.4). Nevertheless, \hat{u} is not radially symmetric!
Then we rescale the problem as follows. For u = u(r) as in (2.7) and for some \gamma > 0 , define v(r) \doteq u(5r)/\gamma for r \in [0, 1] , so that
v(r) = \left \{ \begin{aligned} &\frac{1}{\gamma}(1-25r^2)^p &&\qquad \text{if $ 0 \leq r < \dfrac{1}{5}$}\\[3pt] &0 &&\qquad \text{if $r\geq \dfrac{1}{5}$} \, , \end{aligned} \right. |
and v solves
\left\{ \begin{aligned} & \Delta^2v = g^\gamma(v)\quad\text{in} \; {\bf B} \\ & v = \frac{\partial v}{\partial r} = 0\quad\text{on} \;\partial {\bf B} \end{aligned} \right.\qquad\qquad\mbox{where}\quad g^\gamma(w) \doteq \dfrac{625}{\gamma} g(\gamma w) \qquad \forall w \geq 0. |
Finally, fix {\varepsilon} > 0 and let us construct the function g_{\varepsilon} in the statement. Choose \gamma_{0} > 0 such that
\gamma_{0} > \dfrac{1}{\varepsilon}\max\{\Gamma_1, \, 625M\} \quad \Longrightarrow \quad \frac{\Gamma_2}{\gamma_{0}} < \frac{\Gamma_1}{\gamma_{0}} < {\varepsilon} \ \ \mbox{ and } \ \ \frac{625M}{\gamma_{0}} < {\varepsilon}. |
Then g_{\varepsilon} \doteq g^{\gamma_{0}} satisfies all the assumptions of the statement and the function v is a radial solution, nonnegative but not strictly positive (since it vanishes in {\bf B}\setminus {\bf B}_{1/5} ), and radially decreasing but not strictly radially decreasing (since it is constant in {\bf B}\setminus {\bf B}_{1/5} ). Moreover, after rescaling the function \hat{u} in (2.8), we obtain infinitely many nonnegative nontrivial solutions which are not radially symmetric, one for each x_0 and, possibly with multiple bumps.
The next result shows that the monotonicity assumption in (1.5) is not necessary to obtain the statement of Theorem 1.2. For n \geq 2 , consider the function
\begin{equation} f_{n}(w) \doteq 1632 + 576(n-1) + 32(n-1)(n-3)\\-\left[ 1680 + 672(n-1) + 48(n-1)(n-3) \right]\sqrt{w} \quad \forall w \geq 0, \end{equation} | (2.9) |
which is strictly decreasing and sign-changing, see Figure 2 below.
However, we can prove
Proposition 2.4. Let f_n be as in (2.9). The unique strong solution of the problem
\begin{equation} \label{{prob}} \left\{ \begin{aligned} & \Delta^2u = f_{n}(u)\quad\mathit{\text{in}} \; {\bf B} \\[6pt] & u = \frac{\partial u}{\partial r} = 0\quad\mathit{\text{on}} \; \partial {\bf B}, \end{aligned} \right. \end{equation} | (2.10) |
is given by
u(x) = (1-|x|^4)^2 \qquad \forall x \in {\bf B} \, , |
and, hence, it is positive, radially symmetric and strictly decreasing in the radial variable.
Proof. The function f_{n} defined in (2.9) satisfies the assumption (2.2) so that, by arguing as in the proof of Theorem 2.1, we infer that (2.10) admits at most one strong solution, which is necessarily radially symmetric.
In radial coordinates, when m = 2 , the Eq (1.2) becomes (2.6). The fact that u(r) = (1-r^4)^2 , for r \in [0, 1] , is a (radial) solution follows by noticing that u(1) = u'(1) = 0 (fulfilling the Dirichlet boundary conditions), and then, by inserting the expressions of u(r) and f_{n}\big(u(r)\big) into (2.6).
We introduce here the Green function of the polyharmonic operator (-\Delta)^m under Dirichlet boundary conditions in {\bf B}\subset \mathbb{R}^n and we determine some of its properties. Let us firstly define
\theta(x,y)\doteq(1-|x|^2)(1-|y|^2)\qquad\forall x,y \in {\bf B}\, . |
Then for x, y \in {\bf B} , with x \not = y , Boggio [5,p.126] (see also [9,Section 2.6]) gave the following explicit representation of the Green function:
\begin{equation} \begin{aligned} G(x,y) & = k_n^m |x-y|^{2m-n} \int_1^{\big(\frac{\theta(x,y)}{|x-y|^2}+1\big)^{1/2}} \frac{(z^2-1)^{m-1}}{z^{n-1}}\:d z \\[6pt] & = \frac{k_n^m}{2} |x-y|^{2m-n} \int_0^{\text{ $\frac{\theta(x,y)}{|x-y|^2}$}}\frac{z^{m-1}}{(z+1)^{n/2}}\:dz = \frac{k_n^m}{2}H(|x-y|^2,\theta(x,y)), \end{aligned} \end{equation} | (3.1) |
where k_n^m is a positive constant defined by
\begin{equation} k_n^m = \dfrac{\Gamma \left( 1 + \dfrac{n}{2} \right)}{n \, \pi^{n/2} \, 4^{m-1} ((m-1)!)^{2}} \, , \end{equation} | (3.2) |
see [9,Lemma 2.27], and H:(0, \infty)\times[0, \infty)\rightarrow \mathbb{R} is defined as
\begin{equation} H(s,t) = s^{m-\frac{n}{2}}\int_0^{\frac{t}{s}}\frac{z^{m-1}}{(z+1)^{n/2}}\,dz \qquad \forall s > 0, \, t \geq 0. \end{equation} | (3.3) |
The following statement is a direct consequence of Boggio's work [5], elliptic regularity (see [9,Section 2.5]) and the estimates in [12].
Proposition 3.1. Let h \in L^{\infty}({\bf B}) , and let u \in H_0^m({\bf B}) be a weak solution of (-\Delta)^m u = h in {\bf B} under Dirichlet boundary conditions, that is
\begin{equation} \langle u, v \rangle_m = \int_ {\bf B} h v \qquad\mathit{\mbox{for all}}\;v\in H_0^m( {\bf B}). \end{equation} | (3.4) |
Then u \in \mathcal{C}^{2m-1, \alpha}(\overline\Omega) and it satisfies
\begin{equation} D^ku(x) = \int_{ {\bf B}}D_x^kG(x,y)h(y)\:dy\qquad \forall x \in \overline{ {\bf B}} \, , \end{equation} | (3.5) |
where D^k stands for any partial derivative of order |k| < 2m . In particular, u is a strong solution and D^k u \equiv 0 on \partial {\bf B} for |k|\le m -1 .
In conformal dimensions n = 2m , the Green function admits a simpler representation. Indeed, the change of variables \xi = \frac{1}{z+1} yields
\begin{aligned} H(s,t) & = \int_0^{\frac{t}{s}}\frac{z^{m-1}}{(z+1)^{m}}\,dz = \int_{0}^{\frac{t}{s}} \left( 1 - \frac{1}{z+1} \right)^{m-1} \frac{1}{z+1}\,dz = \int_{\frac{s}{s+t}}^{1} \dfrac{1}{\xi} (1-\xi)^{m-1} \, d\xi \\[6pt] & = \int_{\frac{s}{s+t}}^{1} \left[ \dfrac{1}{\xi} + \sum\limits_{k = 1}^{m-1} (-1)^{k} \binom{m-1}{k} \xi^{k-1} \right] d\xi \\[6pt] & = \log \left(1 + \dfrac{t}{s} \right) + \sum\limits_{k = 1}^{m-1} \dfrac{(-1)^{k}}{k} \binom{m-1}{k} \left( 1 - \left( \dfrac{s}{s+t} \right)^{k} \right) \qquad \forall s > 0, \ t \geq 0. \end{aligned} |
Therefore, by setting \kappa_{m} \doteq k_{2m}^{m} , for every x, y \in {\bf B} we have
\begin{equation} \label{{Greenconf}} G(x,y) = \dfrac{\kappa_{m}}{2} \left[ \log \left(1 + \dfrac{\theta(x,y)}{|x-y|^2} \right) + \sum\limits_{k = 1}^{m-1} \dfrac{(-1)^{k}}{k} \binom{m-1}{k} \left( 1 - \left( \dfrac{|x-y|^2}{|x-y|^2+\theta(x,y)} \right)^{k} \right) \right]. \end{equation} | (3.6) |
The main result of this section (Theorem 3.2) shows that, in conformal dimensions, the Green function satisfies an elegant reflection property. For this, given the point x = (x_{1}, x')\in {\bf B} with x_{1} > 0 , we define
\begin{equation} R_x\doteq\frac{1-\vert x\vert^2}{2x_{1}}\, ,\qquad y_{x}\doteq(x_{1}+R_x,x')\, ,\qquad \mathcal C_{x} \doteq\left\{y\in {\bf B}\; ; \, \vert y - y_{x} \vert > R_x\; \right\}, \end{equation} | (3.7) |
and the set
\begin{equation} \label{{esferaS}} \mathcal{S}_{x} \doteq \left\{y\in \mathbb{R}^{n}\; ; \; (1-\vert x\vert^2)(x_{1}-y_{1})+x_{1}\vert x-y\vert^2 = 0\right\} , \end{equation} | (3.8) |
which is a sphere with center at y_x and radius R_x (see the proof of Theorem 3.2 below and also Figure 4). We then denote by \mathcal{P}_{x} : \mathbb{R}^{n} \setminus \{y_{x}\} \longrightarrow \mathbb{R}^{n} the inversion in \mathcal{S}_{x} , given by the expression
\mathcal{P}_{x}(y) = y_{x} + R_x^2\frac{y-y_{x}}{\vert y-y_{x}\vert^2} \qquad \forall y \in \mathbb{R}^{n} \setminus \{y_{x}\} \, . |
Finally, the Kelvin transform of a function h : \mathcal C_{x} \longrightarrow \mathbb{R} with respect to \mathcal{S}_{x} is defined as
\mathcal{K}_{x}(h)(y) = h(\mathcal{P}_{x}(y)) \qquad \forall y \in \mathcal C_{x} \, . |
Then we prove
Theorem 3.2. For any x = (x_{1}, x')\in {\bf B}\subset\mathbb{R}^{2m} with x_1 > 0 and any u \in H^{2m}({\bf B}) \cap H_0^m({\bf B}) we have
\begin{equation} \begin{aligned} \dfrac{1 - \vert x \vert^{2}}{\kappa_{m} \, x_{1}} \frac{\partial u}{\partial x_{1}}(x) = \int_{\mathcal C_{x}} & \dfrac{\vert y-y_{x}\vert^2 - R_{x}^2}{|x-y|^2}\left(\frac{\theta(x,y)}{|x-y|^2+\theta(x,y)}\right)^m (-\Delta_{y})^{m} \left[ \mathcal{K}_{x}(u)(y) - u(y) \right] dy \, . \end{aligned} \end{equation} | (3.9) |
Proof. The following identities hold for the function H defined in (3.3): for all s, t > 0 we have
\begin{equation} \dfrac{\partial H}{\partial t}(s,t) = \frac{t^{m-1}}{(t+s)^m}\, ,\qquad \dfrac{\partial H}{\partial s} (s,t) = -\frac{t^{m}}{s(t+s)^m}\, , \qquad \dfrac{\partial^{2} H}{\partial s \partial t} (s,t) = -\frac{mt^{m-1}}{(t+s)^{m+1}}\, , \end{equation} | (3.10) |
see also [3]. Moreover, for every x, y \in {\bf B} we have
\begin{equation} \dfrac{\partial G}{\partial x_{1}} (x,y) = \kappa_{m} \left[ (x_1-y_1) \dfrac{\partial H}{\partial s} \left(|x-y|^2,\theta(x,y) \right)- x_{1}(1-|y|^2) \dfrac{\partial H}{\partial t} \left( |x-y|^2,\theta(x,y) \right) \right]. \end{equation} | (3.11) |
We then prove the announced symmetry property, namely that the zero level sets of \partial_{x_{1}} G(x, \cdot) are spherical caps. Let us denote the zero level set by
\mathcal{L}_{x} \doteq \left\{y\in {\bf B}\; ; \; \frac{\partial G}{\partial x_{1}}(x,y) = 0\right\}\, . |
By (3.10) and by (3.11) we have
\begin{equation} \begin{aligned} -\frac{1}{\kappa_{m}} \frac{\partial G}{\partial x_{1}}(x,y) & = x_{1}(1-|y|^2) \dfrac{\partial H}{\partial t}(|x-y|^2,\theta(x,y))-(x_1-y_1)\dfrac{\partial H}{\partial s}(|x-y|^2,\theta(x,y)) \\[6pt] & = x_{1}(1-|y|^2) \frac{\theta(x,y)^{m-1}}{(|x-y|^2+\theta(x,y))^m} + (x_{1}-y_{1})\frac{\theta(x,y)^m}{|x-y|^2(|x-y|^2 + \theta(x,y))^m}\\[6pt] & = \frac{\theta(x,y)^{m-1}}{|x-y|^2(|x-y|^2+\theta(x,y))^m}(1-\vert y\vert^2)\left[ x_{1}\vert x-y\vert^2 + (1-\vert x\vert^2)(x_{1}-y_{1}) \right]. \end{aligned} \end{equation} | (3.12) |
Therefore, \partial_{x_{1}} G(x, y) vanishes if and only if y\in {\bf B} belongs to the set \mathcal{S}_{x} defined in (3.8). Then we notice that y\in {\bf B}\cap\mathcal{S}_{x} if and only if
\begin{align} 0 = \frac1{x_{1}}\left[(1-\vert x\vert^2)(x_{1}-y_{1})+x_{1}\vert x-y\vert^2 \right]& = \frac{1-\vert x\vert^2}{x_{1}}(x_{1}-y_{1}) + (x_{1}-y_{1})^2 + \vert x'-y'\vert^2\\[6pt] & = 2R_x(x_{1}-y_{1}) + (x_{1}-y_{1})^2 + \vert x'-y'\vert^2\\[6pt] & = (R_x+x_{1}-y_{1})^2 + \vert x'-y'\vert^2 -R_x^2\\[6pt] & = \vert y_{x}-y\vert^2 - R_x^2\, . \end{align} | (3.13) |
Hence, \mathcal{S}_{x} is a sphere with center at y_x and radius R_x and the set of negativity of \partial_{x_{1}} G(x, y) is the cap \mathcal C_{x} defined in (3.7).
Next, we notice that, by definition of y_{x} and R_x , we have
\vert y_{x}\vert^2 = \left(x_{1}+\frac{1-\vert x\vert^2}{2x_{1}}\right)^2+\vert x'\vert^2 = x_{1}^2+(1-\vert x\vert^2)+R_x^2+\vert x'\vert^2 = 1+R_x^2\, . |
Then we compute
\begin{array}{l}\vert \mathcal{P}_{x}(y) \vert^2 = \vert \mathcal{P}_{x}(y) -y_{x}+y_{x}\vert^2 = \vert \mathcal{P}_{x}(y) -y_{x}\vert^2 + 2(\mathcal{P}_{x}(y)-y_{x})\cdot y_{x} + \vert y_{x}\vert^2\\[3pt] = \frac{R_x^4}{\vert y-y_{x}\vert^2} + 2R_x^2\frac{y-y_{x}}{\vert y-y_{x}\vert^2}\cdot y_{x} + \vert y_{x}\vert^2 \\= \frac{R_x^2}{\vert y-y_{x}\vert^2}\left(R_x^2 + 2(y-y_{x})\cdot y_{x} + \vert y_{x}\vert^2\frac{\vert y-y_{x}\vert^2}{R_x^2}\right)\\[3pt] = \frac{R_x^2}{\vert y-y_{x}\vert^2}((R_x^2 -\vert y-y_{x}\vert^2)+(\vert y-y_{x}\vert^2 + 2(y-y_{x})\cdot y_{x}+\vert y_{x}\vert^2)\\-\vert y_{x}\vert^2 + \vert y_{x}\vert^2\frac{\vert y-y_{x}\vert^2}{R_x^2})\\[3pt] = \frac{R_x^2}{\vert y-y_{x}\vert^2}\left((R_x^2 -\vert y-y_{x}\vert^2)+|y|^{2}-\vert y_{x}\vert^2 + \vert y_{x}\vert^2\frac{\vert y-y_{x}\vert^2}{R_x^2}\right)\\[3pt] = \frac{R_x^2}{\vert y-y_{x}\vert^2}\left((R_x^2 +|y|^{2}-\vert y_{x}\vert^2) + \frac{(\vert y_{x}\vert^2-R_x^2) \vert y-y_{x}\vert^2}{R_x^2} \right)\\[3pt] = \frac{R_x^2}{\vert y-y_{x}\vert^2}\left( (|y|^{2} -1) +\frac{\vert y-y_{x}\vert^2}{R_x^2}\right) = 1 - \frac{R_x^2}{\vert y-y_{x}\vert^2} (1 - |y|^{2}) \, . \end{array} |
This computation has two main consequences. First, that the unit sphere \partial {\bf B} is invariant under the inversion in \mathcal{S}_{x} , that is,
y\in\partial {\bf B}\ \Longleftrightarrow\ \mathcal{P}_x(y)\in\partial {\bf B}\, . |
Second, that
\begin{equation} \theta(x,\mathcal{P}_{x}(y)) = \frac{R_{x}^2}{\vert y_{x}-y\vert^2} \, \theta(x,y)\qquad\forall y\in {\bf B}\, . \end{equation} | (3.14) |
We next claim that if \mathcal C_{x} is the cap defined in (3.7), then {\bf B} = \mathcal C_{x} \cup \mathcal{P}_{x}(\mathcal C_{x}) \cup \mathcal{L}_{x} . To this end, it suffices to show that
\mathcal{P}_{x}( \mathcal C_{x}) = \left\{y\in {\bf B}\; ; \, \vert y - y_{x} \vert < R_x\right\}. |
Indeed, let y \in \mathcal C_{x} , so that \mathcal{P}_{x}(y) \in \mathcal{P}_{x}(\mathcal C_{x}) and \vert y - y_{x} \vert > R_x . By definition of \mathcal{P}_{x} we have
| \mathcal{P}_{x}(y) - y_{x} | = \dfrac{R_x^{2}}{| y - y_{x} |} < R_x, |
thus proving that \mathcal{P}_{x}(\mathcal C_{x}) \subset \left\{y\in {\bf B}\; ; \, \vert y - y_{x} \vert < R_x\right\} . On the other hand, let y \in {\bf B} be such that \vert y - y_{x} \vert < R_x . Since the function \mathcal{P}_{x} is bijective, there exists a point y_{0} \in \mathbb{R}^{n} such that \mathcal{P}_{x}(y_{0}) = y . By definition, we then have
| y_{0} - y_{x} | = \dfrac{R_x^{2}}{| y - y_{x} |} > R_x, |
implying y_{0} \in \mathcal C_{x} , and subsequently, that y \in \mathcal{P}_{x}(\mathcal C_{x}) and \left\{y\in {\bf B}\; ; \, \vert y-y_{x} \vert < R_x \right\} \subset \mathcal{P}_{x}(\mathcal C_{x}) .
Since {\bf B} = \mathcal C_{x} \cup \mathcal{P}_{x}(\mathcal C_{x}) \cup \mathcal{L}_{x} , by Proposition 3.1, we have
\begin{aligned} \frac{\partial u}{\partial x_{1}}(x) & = \int_{ \mathcal C_{x}} \frac{\partial G}{\partial x_{1}}(x,y) (-\Delta_{y})^{m}u(y) \;dy + \int_{\mathcal{P}_{x}( \mathcal C_{x})} \frac{\partial G}{\partial x_{1}}(x,y)(-\Delta_{y})^{m}u(y)\;dy\\[6pt] & = \int_{ \mathcal C_{x}} \frac{\partial G}{\partial x_{1}}(x,y)(-\Delta_{y})^{m}u(y)\;dy + \int_{ \mathcal C_{x}} \left(\frac{R_{x}}{\vert y-y_{x}\vert}\right)^{2n} \, \frac{\partial G}{\partial x_{1}}(x,\mathcal{P}_{x}(y)) ((-\Delta_{y})^{m} u)(\mathcal{P}_{x}(y)) \;dy\\[6pt] & = \int_{ \mathcal C_{x}} \frac{\partial G}{\partial x_{1}}(x,y)(-\Delta_{y})^{m}u(y)\;dy + \int_{ \mathcal C_{x}} \left(\frac{R_{x}}{\vert y-y_{x}\vert}\right)^{4m} \, \frac{\partial G}{\partial x_{1}}(x,\mathcal{P}_{x}(y)) ((-\Delta_{y})^{m} u)(\mathcal{P}_{x}(y))\;dy,\\[6pt] \end{aligned} |
where we used that the Jacobian matrix \mathcal{P}'_{x} satisfies
\left| \text{det}(\mathcal{P}'_{x}(y)) \right| = \left(\frac{R_{x}}{\vert y-y_{x}\vert}\right)^{2n}\qquad\forall y\in {\bf B}\, . |
In view of (3.12) and (3.13) we can write
\begin{equation} -\frac{1}{\kappa_{m}} \dfrac{1 - \vert x \vert^{2}}{x_{1}} \frac{\partial G}{\partial x_{1}}(x,y) = \dfrac{\vert y_{x}-y\vert^2 - R_x^2}{|x-y|^2} \left(\frac{\theta(x,y)}{\theta(x,y)+|x-y|^2}\right)^m \qquad \forall y \in {\bf B} \, , \end{equation} | (3.15) |
so that
\begin{equation} -\frac{1}{\kappa_{m}} \dfrac{1 - \vert x \vert^{2}}{x_{1}} \frac{\partial u}{\partial x_{1}}(x) = I_{1}(x) + I_{2}(x) \, , \end{equation} | (3.16) |
where
![]() |
To compute I_{2}(x) , we first notice that
\begin{equation} \vert y_{x}-\mathcal{P}_{x}(y)\vert^2-R_x^2 = \frac{R_x^4}{\vert y_{x}-y\vert^2}-R_x^2 = -\frac{R_x^2}{\vert y_{x}-y\vert^2}(\vert y_{x}-y\vert^2-R_x^2) < 0 \qquad \forall y \in \mathcal C_{x} \, , \end{equation} | (3.17) |
whereas, for every y \in {\bf B} we have
\begin{equation} \begin{aligned} |x-\mathcal{P}_{x}(y)|^{2} & = \vert x-y_{x}\vert^2 + 2(x-y_{x})\cdot(y_{x}-\mathcal{P}_{x}(y)) +\vert y_{x}-\mathcal{P}_{x}(y)\vert^2\\[3pt] & = \vert x-y_{x}\vert^2 + \frac{2R_x^2}{\vert y_{x}-y\vert^2}(x-y_{x})\cdot(y_{x}-y) +\frac{R_x^4}{\vert y-y_{x}\vert^2}\\[3pt] & = \frac{R_x^2}{\vert y_{x}-y\vert^2}\left( \frac{\vert y_{x}-y\vert^2}{R_x^2}\vert x-y_{x}\vert^2 + 2(x-y_{x})\cdot(y_{x}-y) + {R_x^2}\right)\\[3pt] & = \frac{R_x^2}{\vert y_{x}-y\vert^2}\left( {\vert y_{x}-y\vert^2} + 2(x-y_{x})\cdot(y_{x}-y) + \vert x-y_{x}\vert^2\right) = \frac{R_x^2}{\vert y_{x}-y\vert^2} \, |x-y|^2 \, . \end{aligned} \end{equation} | (3.18) |
After plugging (3.14), (3.17) and (3.18) into I_{2}(x) we get
I_{2}(x) = - \int_{\mathcal C_{x}} \dfrac{\vert y-y_{x}\vert^2 - R_{x}^2}{|x-y|^2}\left(\frac{\theta(x,y)}{|x-y|^2+\theta(x,y)}\right)^m \left(\frac{R_{x}}{| y - y_{x} |}\right)^{4m} ((-\Delta_{y})^{m} u)(\mathcal{P}_{x}(y)) dy |
and, after recalling (3.16), we deduce
\begin{array}{l} -\frac{1}{\kappa_{m}} \dfrac{1 - \vert x \vert^{2}}{x_{1}} \frac{\partial u}{\partial x_{1}}(x) \\[6pt] = \int_{\mathcal C_{x}} \dfrac{\vert y-y_{x}\vert^2 - R_{x}^2}{|x-y|^2}\left(\frac{\theta(x,y)}{|x-y|^2+\theta(x,y)}\right)^m \\ \left[ (-\Delta_{y})^{m} u(y) - \left(\frac{R_{x}}{| y - y_{x} |}\right)^{4m} ((-\Delta_{y})^{m} u)(\mathcal{P}_{x}(y)) \right] dy \, . \end{array} |
A direct computation (see [13,Proposition 7.3]) shows that
\begin{equation} (-\Delta_{y})^{m} \mathcal{K}_{x}(u)(y) = \left(\frac{R_{x}}{| y - y_{x} |}\right)^{4m} ((-\Delta_{y})^{m} u)(\mathcal{P}_{x}(y)) \qquad \forall y \in \mathcal{C}_{x} \, , \end{equation} | (3.19) |
which concludes the proof after changing the sign.
Problem 3.3. It is not clear how to derive an expression similar to (3.9) in non-conformal dimensions n \geq 2 , n \neq 2m . This can be seen after computing
\dfrac{\partial H}{\partial s}(s,t) = \left( n - \dfrac{m}{2} \right) s^{m-\frac{n}{2}-1}\int_0^{\frac{t}{s}}\frac{z^{m-1}}{(z+1)^{n/2}}\,dz - \dfrac{t^{m}}{s(t+s)^{n/2}} \qquad \forall s > 0, \ t \geq 0 \,, |
so that the determination of the level sets of \partial_{x_{1}} G(x, \cdot) as in (3.12) becomes a delicate task. Is it possible to obtain a representation formula similar to (3.9) in any dimension n \geq 2 ?
Clearly, a representation formula similar to (3.9) holds for any other directional derivative of u . Note that (3.7) may be rewritten as y_x = x+(R_x, 0) , which implies that y_x is an "horizontal translation" of x \in {\bf B} . Moreover, R_x\to0 if x\to\partial {\bf B} while R_x\to+\infty if x_1\to0 : the limit of the map x\mapsto R_x does not exist when x approaches the equator \partial {\bf B}\cap \{x_1 = 0\} . The level surfaces of x\mapsto R_x are spheres:
R_x = k\in(0,\infty)\ \Longrightarrow\ (x_1+k)^2+|x'|^2 = 1+k^2\, . |
The center of the level-surface sphere is then (x_1, x') = (-k, 0) and its radius is \sqrt{1+k^2} . For a fixed k > 0 , when x runs over the level surface R_x = k , the corresponding y_x , defined by (3.7), runs over a portion of the sphere centered in the origin with the same radius \sqrt{1+k^2} , see Figure 3.
Note also that \mathcal C_{x} , the domain of negativity of \partial_{x_{1}} G(x, y) , is the intersection between two balls. In Figure 4 we sketch the mutual position of the sphere \mathcal{S}_x and the ball {\bf B} . In particular, we emphasize that
|y_x|^2 = \left(x_1+\frac{1-|x|^2}{2x_1}\right)^2+|x'|^2 = 1+\frac{(1-|x|^2)^2}{4x_1^2} > 1\qquad\forall x\in {\bf B}\quad(x_1 > 0) |
which, again, tells us that y_x is exterior to {\bf B} and it approaches \partial {\bf B} as x approaches \partial {\bf B} , while it goes to infinity if x approaches the plane x_1 = 0 .
Theorem 3.2 has several relevant consequences. Firstly, we deduce
Corollary 3.4. Let m \geq 1 . For every x \in {\bf B} \subset \mathbb{R}^{2m} such that x_{1} > 0 , the following identities hold
\begin{equation} \begin{aligned} (1 - \vert x \vert^{2})^{m} = \dfrac{1}{\pi^{m}} \dfrac{(2m-1)!}{(m-1)!} \int_{\mathcal C_{x}} \dfrac{\vert y-y_{x}\vert^2 - R_{x}^2}{|x-y|^2} \\ \left(\frac{\theta(x,y)}{|x-y|^2+\theta(x,y)}\right)^m \left[ 1 - \left(\frac{R_{x}}{| y - y_{x} |}\right)^{4m} \right] dy \, , \end{aligned} \end{equation} | (3.20) |
\begin{equation} \begin{aligned} \mathbb{K}_m\doteq \pi^{m} \dfrac{(m-1)!}{(2m-1)!} = \int_{\mathcal C_{x}} \dfrac{\vert y-y_{x}\vert^2 - R_{x}^2}{|x-y|^2} \left(\frac{1 - \vert y \vert^{2}}{|x-y|^2+\theta(x,y)}\right)^m \left[ 1 - \left(\frac{R_{x}}{| y - y_{x} |}\right)^{4m} \right] dy \, . \end{aligned} \end{equation} | (3.21) |
Proof. We start by noticing that
\begin{aligned} & \Delta r^{2m} = 2m(2m + n -2) r^{2m-2} \, , \qquad \Delta^{2} r^{2m} = 2m(2m-2)(2m + n -2)(2m + n -4) r^{2m-4} \, , \\[6pt] & \quad \Delta^{3} r^{2m} = 2m(2m-2)(2m-4)(2m + n -2)(2m + n -4)(2m + n -6) r^{2m-6} \, , \end{aligned} |
where n = 2m . By induction we then obtain
\Delta^{m} r^{2m} = (2m)!! \, \dfrac{(2m + n -2)!!}{(n-2)!!} = (2m)!! \, \dfrac{(4m -2)!!}{(2m-2)!!} = 2^{2m-1} (2m)! \, . |
Consider the polyharmonic version of the original problem (1.1) proposed by Serrin:
\begin{equation} \left\{ \begin{array}{ll} (-\Delta)^{m}u = 1\qquad & \mbox{in } {\bf B}\\[3pt] u = \dfrac{\partial u}{\partial r} = \dots = \dfrac{\partial^{m-1}u}{\partial r^{m-1}} = 0\qquad & \mbox{on }\partial {\bf B}. \end{array}\right. \end{equation} | (3.22) |
A direct application of the binomial expansion yields
(-\Delta)^{m} (1 - r^{2})^{m} = (-\Delta)^{m} \left[ \sum\limits_{k = 0}^{m} \binom{m}{k} (-1)^{k} r^{2k} \right] = \Delta^{m}r^{2m} = 2^{2m-1}(2m)! \, , |
that is
\begin{equation} (-\Delta)^{m} U_{m}(r) = 1 \qquad \text{with} \qquad U_{m}(r) \doteq \dfrac{(1 - r^{2})^{m}}{2^{2m-1} (2m)!} \qquad \forall r \in [0,1]. \end{equation} | (3.23) |
Since U_{m} satisfies the Dirichlet boundary conditions in (3.22) _{2} , it is a strong \mathcal{C}^{\infty}(\overline{ {\bf B}}) -solution of (3.22). Applying identity (3.9) to the function U_{m} and recalling (3.2) (with n = 2m ), we derive equality (3.20). The identity (3.21) is a straightforward consequence of (3.20), after replacing \theta(x, y) .
It is remarkable that the right hand side of (3.21) does not depend on x\in {\bf B} . As a further consequence of Theorem 3.2, we give an alternative formula to compute the partial derivatives of the solutions of (1.2), which should be compared with Proposition 3.1.
Corollary 3.5. Let {\bf B}\subset \mathbb{R}^{2m} be the unit ball. Suppose that f \in \mathcal{C}(\mathbb{R}; \mathbb{R}) and that u\in H_0^m({\bf B}) \cap L^{\infty}({\bf B}) is a strong solution of (1.2). Then, for every x \in {\bf B} such that x_{1} > 0 , the following formula holds:
\begin{equation} \dfrac{\vert x \vert^{2}-1}{\kappa_{m} \, x_{1}} \frac{\partial u}{\partial x_{1}}(x) = \int_{\mathcal C_{x}}\dfrac{\vert y-y_{x}\vert^2 - R_{x}^2}{|x-y|^2}\left[\frac{\theta(x,y)}{|x-y|^2+\theta(x,y)}\right]^m \left[ f(u(y)) - \left(\frac{R_{x}}{| y - y_{x} |}\right)^{4m} f(u(\mathcal{P}_{x}(y))) \right]\!dy. \end{equation} | (3.24) |
Formula (3.24) follows directly from (3.9)–(3.19) and suggests the problem:
Problem 3.6. Since the moving plane procedure developed in [3,Section 3] requires that
\frac{\partial u}{\partial x_{1}}(x) < 0 \qquad \mathit{\text{for all}} \ x \in {\bf B} \ \ \mathit{\text{with}} \ \ x_{1} > 0 \ , |
is it possible to use (3.24) to prove this inequality and to relax assumption (1.5), but still ensuring the statement of Theorem 1.2?
In connection with Problem 3.6 we notice that, by looking at (3.24), it is clear that the overall sign of the integral over \mathcal{C}_{x} is determined by the behavior of the quantity
\Phi_{x}(y) \doteq f(u(y)) - \left(\frac{R_{x}}{| y - y_{x} |}\right)^{4m} f(u(\mathcal{P}_{x}(y))) \qquad \forall y \in \mathcal{C}_{x} \, . |
Fixing x \in {\bf B} such that x_{1} > 0 and u\in H_0^m({\bf B}) , notice that \Phi_{x}(y) = 0 for all y \in \partial \mathcal{C}_{x} \cap {\bf B} , while for y \in \partial \mathcal{C}_{x} \cap \partial {\bf B} (so that \mathcal{P}_{x}(y) \in \partial {\bf B} ) we have that \Phi_{x}(y) = 0 if f(0) = 0 , and \Phi_{x}(y) > 0 if f(0) > 0 .
Under suitable assumptions on the source f , in the next result we give an upper bound for the partial derivative of u (recall that |x| < 1 !), which shows that \partial_{x_{1}}u(x) < 0 for all x \in {\bf B} located far away from \partial {\bf B} , with x_1\geq\varepsilon for some \varepsilon > 0 , see Figure 5 for a schematic representation. Therefore, in order to solve Problem 3.6, one should mainly focus the attention on the complement of this region.
To this end, we define
\frac1\Gamma \doteq \inf\limits_{v\in\mathcal{C}^{2m}(\overline{ {\bf B}})\cap H^m_0( {\bf B})\setminus\{0\}}\frac{\|\Delta^mv\|_{L^{\infty}( {\bf B})}}{\|\nabla v\|_{L^{\infty}( {\bf B})}}\, , |
and we state:
Proposition 3.7. Assume that f \in W^{1, \infty}(\mathbb{R}_+; \mathbb{R}_+) satisfies
\begin{equation} f(s) \geq M \qquad \forall s \geq 0 \, , \ \ \mathit{\text{for some}} \ \ M > 0, \end{equation} | (3.25) |
and let u\in H_0^m({\bf B}) \cap L^{\infty}({\bf B}) be a strong solution of (1.2). For any point x = (x_{1}, x')\in {\bf B} with x_1 > 0 we have
\begin{aligned} \dfrac{\vert x \vert^{2}-1}{\kappa_{m} \, x_{1}} \frac{\partial u}{\partial x_{1}}(x) & \geq \left( M - 2\Gamma\|f\|_{L^{\infty}( \mathbb{R}_+)}\| f' \|_{L^{\infty}( \mathbb{R}_+)} \right) \mathbb{K}_m (1-|x|^2)^{m} \\[6pt] & \quad - 2\Gamma\|f\|_{L^{\infty}( \mathbb{R}_+)}\| f' \|_{L^{\infty}( \mathbb{R}_+)} \dfrac{1 - \vert x \vert^{2}}{\kappa_{m} \, x_{1}} \int_{\mathcal C_{x}} \left(\frac{R_{x}}{| y - y_{x} |}\right)^{4m} \frac{\partial G}{\partial x_{1}}(x,\mathcal{P}_x(y)) \, dy \, , \end{aligned} |
where \mathbb{K}_m > 0 is defined in (3.21).
Proof. In view of Corollary 3.4, by the Maximum Principle and assumption (3.25), any strong solution u\in H_0^m({\bf B}) \cap L^{\infty}({\bf B}) of (1.2) satisfies
u(x) \geq \dfrac{M}{2^{2m-1} (2m)!} (1 - |x|^{2})^{m} \qquad \forall x \in {\bf B}\, . |
By embedding theorems and elliptic regularity we also have that
\| \nabla u \|_{L^{\infty}( {\bf B})}\leq\Gamma\|f\|_{L^{\infty}( \mathbb{R}_+)} \, . |
Then, by the Mean Value Theorem, we obtain the inequality
|u(\mathcal{P}_x(y))-u(y)|\leq \Gamma\|f\|_{L^{\infty}( \mathbb{R}_+)}|\mathcal{P}_{x}(y) - y | \qquad \forall y \in \mathcal{C}_{x} \, . |
From this we deduce that
|f(u(\mathcal{P}_x(y)))-f(u(y))|\leq\Gamma\|f\|_{L^{\infty}( \mathbb{R}_+)}\|f'\|_{L^{\infty}( \mathbb{R}_+)}\,| \mathcal{P}_{x}(y) - y | \qquad \forall y \in \mathcal{C}_{x} \, , |
thus yielding
f(u(\mathcal{P}_{x}(y))) \leq f(u(y)) + \Gamma\|f\|_{L^{\infty}( \mathbb{R}_+)}\| f' \|_{L^{\infty}( \mathbb{R}_+)} \, | \mathcal{P}_{x}(y) - y | \qquad \forall y \in \mathcal{C}_{x} |
and, by using (3.25), we finally obtain
\begin{aligned} f(u(y))-\left(\frac{R_{x}}{|y-y_x|}\right)^{4m} f(u(\mathcal{P}_{x}(y))) & \geq \left[1-\left(\frac{R_{x}}{| y - y_{x} |}\right)^{4m}\right]M \\[6pt] & \quad -\left(\frac{R_{x}}{| y - y_{x} |}\right)^{4m}\Gamma\|f\|_{L^{\infty}( \mathbb{R}_+)}\| f' \|_{L^{\infty}( \mathbb{R}_+)} \, | \mathcal{P}_{x}(y)-y|\, . \end{aligned} |
By inserting this bound within (3.24), we get
\begin{equation} \dfrac{\vert x \vert^{2}-1}{\kappa_{m} \, x_{1}} \frac{\partial u}{\partial x_{1}}(x) \geq I_1(x)-I_2(x) \, , \end{equation} | (3.26) |
where
![]() |
From Corollary 3.4 we infer that
\begin{equation} I_1(x) = M \mathbb{K}_m (1-|x|^2)^m\, . \end{equation} | (3.27) |
In order to bound I_2(x) , we notice that
|\mathcal{P}_{x}(y)-y| < 2 \quad \text{and} \quad \frac{R_{x}}{| y - y_{x} |} < 1 \qquad \forall y\in\mathcal C_{x}\, . |
Hence,
\begin{equation} I_2(x) \leq 2\Gamma\|f\|_{L^{\infty}( \mathbb{R}_+)}\| f' \|_{L^{\infty}( \mathbb{R}_+)} \int_{\mathcal C_{x}}\dfrac{\vert y-y_{x}\vert^2 - R_{x}^2}{|x-y|^2}\left[\frac{\theta(x,y)}{|x-y|^2+\theta(x,y)}\right]^m dy \, . \end{equation} | (3.28) |
From (3.15) and Proposition 3.1, and by proceeding as in the proof of Corollary 3.4, we obtain
\begin{aligned} & \int_{\mathcal C_{x}}\dfrac{\vert y-y_{x}\vert^2 - R_{x}^2}{|x-y|^2}\left[\frac{\theta(x,y)}{|x-y|^2+\theta(x,y)}\right]^m dy = -\frac{1}{\kappa_{m}} \dfrac{1 - \vert x \vert^{2}}{x_{1}} \int_{\mathcal C_{x}} \frac{\partial G}{\partial x_{1}}(x,y) \, dy \\[6pt] & = -\frac{1}{\kappa_{m}} \dfrac{1 - \vert x \vert^{2}}{x_{1}} \left( \int_{ {\bf B}} \frac{\partial G}{\partial x_{1}}(x,y) \, dy - \int_{\mathcal{P}_{x}(\mathcal{C}_{x})} \frac{\partial G}{\partial x_{1}}(x,y) \, dy \right) \\[6pt] & = -\frac{1}{\kappa_{m}} \dfrac{1 - \vert x \vert^{2}}{x_{1}} \left[ \dfrac{1}{2^{2m-1} (2m)!} \dfrac{\partial}{\partial x_{1}} \left( (1 - |x|^{2})^{m}\right) - \int_{\mathcal C_{x}} \left(\frac{R_{x}}{| y - y_{x} |}\right)^{4m} \frac{\partial G}{\partial x_{1}}(x,\mathcal{P}_x(y)) \, dy \right] \\[6pt] & = \mathbb{K}_m (1-|x|^2)^m + \dfrac{1 - \vert x \vert^{2}}{\kappa_{m} \, x_{1}} \int_{\mathcal C_{x}} \left(\frac{R_{x}}{| y - y_{x} |}\right)^{4m} \frac{\partial G}{\partial x_{1}}(x,\mathcal{P}_x(y)) \, dy \, . \end{aligned} |
After plugging this into (3.28) we deduce that
\begin{equation} I_2(x) \leq 2\Gamma\|f\|_{L^{\infty}( \mathbb{R}_+)}\| f' \|_{L^{\infty}( \mathbb{R}_+)} \left[ \mathbb{K}_m (1-|x|^2)^m + \dfrac{1 - \vert x \vert^{2}}{\kappa_{m} \, x_{1}} \int_{\mathcal C_{x}} \left(\frac{R_{x}}{| y - y_{x} |}\right)^{4m} \frac{\partial G}{\partial x_{1}}(x,\mathcal{P}_x(y)) \, dy \right]. \end{equation} | (3.29) |
The conclusion is reached after inserting (3.27) and (3.29) into (3.26).
Proposition 3.7 suggests that the radial symmetry and monotonicity of the solutions of (1.2) can be ensured if \|f'\|_{L^{\infty}(\mathbb{R}_+)} is small. This assumption is not needed if m = 1 (see Theorem 1.1), while for m \geq 2 , the required condition is that f' be nonnegative (see Theorem 1.2). Therefore, one is led to analyze the cases where the negative part of f' is small, that is, \|(f')_{-}\|_{L^{\infty}(\mathbb{R}_+)} small. This issue is tackled numerically in the next section, for the biharmonic equation in {\bf B}\subset\mathbb{R}^{4} .
As a direct consequence of the Hopf-type lemma by Grunau-Sweers [11,Theorem 3.2] we obtain
Proposition 4.1. Assume that f \in W^{1, \infty}(\mathbb{R}_+; \mathbb{R}_+) satisfies (3.25), and let u\in H_0^m({\bf B}) \cap L^{\infty}({\bf B}) be a strong solution of (1.2). Then, there exists \gamma_{f} > 0 such that
x \cdot \nabla u(x) < 0 \qquad \mathit{\text{for all}} \ x \in {\bf B} \ \mathit{\text{such that}} \ \gamma_{f} < |x| < 1 \, . |
Guido Sweers [16] provided an explicit example of a non-decreasing radial solution of a linear biharmonic problem in {\bf B}\subset\mathbb{R}^{2} (equation (1.2) with m = 2 ). In this section we take advantage of his example and give numerical evidence to claim that Proposition 4.1 might be complemented with the statement that \gamma_{f} can be made arbitrarily small (possibly zero) provided that \|(f')_{-}\|_{L^{\infty}(\mathbb{R}_+)} is sufficiently small.
Consider the following radial function defined in \mathbb{R}^{4} (the conformal dimension for the biharmonic operator):
v(x) = \sum\limits_{k = 0}^{\infty} a_{k} \, r^{4k+2} \doteq \sum\limits_{k = 0}^{\infty} \dfrac{(-1)^{k}}{2^{4k+2}} \dfrac{r^{4k+2}}{(2k+2)! (2k+1)!} \qquad \forall x \in \mathbb{R}^{4} \, , |
whose plot in the interval [-10, 10] is displayed in Figure 6 below.
Since m = 2 and n = 4 , the biharmonic operator in radial coordinates becomes
\Delta^2 \psi(r) = \psi^{(4)}(r) + \dfrac{6}{r}\psi'''(r) + \dfrac{3}{r^{2}}\psi''(r) - \dfrac{3}{r^{3}}\psi'(r) \qquad \forall r \in (0,1). |
After noticing that
a_{k+1} = - \dfrac{1}{64 (k+2)(k+1)(2k+3)^{2}} \, a_{k} \qquad \forall k \in \mathbb{N} \, , |
one readily sees that
\Delta^{2} v = - v \quad \text{in} \quad \mathbb{R}^{4} \, . |
Let r_0\approx 9.2218 be the first nonzero local minimum of v ; numerically, we find that v(r_0)\approx-14.8388 . Then define the function
\begin{equation} u(r) \doteq v(r_{0}r) - v(r_{0}) \qquad \forall r \in [0,1] \, , \end{equation} | (4.1) |
which is radially symmetric and strictly positive in [0, 1) , but not decreasing, in the interval [0, 1] ; see the left picture in Figure 7 below.
By defining
f(w) = r_{0}^{4} (|v(r_{0})|-w) \qquad \forall w \geq 0 \, , |
we observe that f'(w)\equiv-r_{0}^{4} < 0 , so that f is decreasing and (2.2) is satisfied. By Theorem 2.1 we then know that there exists at most one strong solution of the following problem:
\left\{ \begin{aligned} & \Delta^2 u = f(u)\quad\text{in} \; {\bf B} \\[6pt] & u = \frac{\partial u}{\partial r} = 0\quad\text{on} \; \partial {\bf B}. \end{aligned} \right. |
The solution is radially symmetric and a simple computation shows that it coincides with u defined in (4.1). Note that u(0) = |v(r_{0})| and u(0.5) > 16.5 > |v(r_{0})| , so that f(u) is sign-changing in [0, 1] .
We now modify f in order to reduce \|(f')_{-}\|_{L^{\infty}(\mathbb{R}_+)} . We take
f_{\gamma}(w) \doteq r_{0}^{4}(|v(r_{0})|-\gamma w) \qquad \forall w \geq 0 \, , |
being \gamma > 0 a variable parameter. Then we consider the problem
\begin{equation} \left\{ \begin{aligned} & \Delta^2 U = f_{\gamma}(U)\quad\text{in}\; {\bf B} \\[6pt] & U = \frac{\partial U}{\partial r} = 0\quad\text{on} \; \partial {\bf B}, \end{aligned} \right. \end{equation} | (4.2) |
and prove the following result:
Proposition 4.2. For every \gamma > 0 , the unique strong solution U_{\gamma} \in H^{2}_{0}({\bf B}) \cap L^\infty({\bf B}) of (4.2) (which is radially symmetric) is analytic and can be written as
\begin{equation} U_{\gamma}(r) = \sum\limits_{k = 0}^{\infty} A_{k}(\gamma) r^{2k} = A_{0}(\gamma) + A_{1}(\gamma) r^{2} + A_{2}(\gamma) r^{4} + A_{3}(\gamma) r^{6} + ... \qquad \forall r \in [0,1], \end{equation} | (4.3) |
for some coefficients \{ A_{k}(\gamma) \}_{k \in \mathbb{N}} \subset \mathbb{R} that satisfy the following properties:
\left\{ \begin{aligned} & A_{0}(\gamma) = \sum\limits_{k = 2}^{\infty} (k-1) A_{k}(\gamma) \, , \qquad \quad A_{1}(\gamma) = - \sum\limits_{k = 2}^{\infty} k A_{k}(\gamma) \, , \\[6pt] & A_{2}(\gamma) = \dfrac{r_{0}^{4}}{192}(|v(r_{0})| - \gamma A_{0}(\gamma)) \, , \qquad A_{k}(\gamma) = - \dfrac{\gamma r_{0}^{4}}{16 k^{2}(k^{2}-1)} A_{k-2}(\gamma) \qquad \forall k \geq 3 \, . \end{aligned} \right. |
Moreover, there exists \gamma_{0} > 0 such that, for every \gamma \in (0, \gamma_{0}) , the function U_{\gamma} is positive in [0, 1) and strictly decreasing in the radial variable.
Proof. For every \gamma > 0 we notice that the condition (2.2) is satisfied. Then, Theorem 2.1 guarantees the existence of at most one strong solution U_{\gamma} \in H^{2}_{0}({\bf B}) \cap L^\infty({\bf B}) to problem (4.2) which is, moreover, radially symmetric and analytic in \overline{ {\bf B}} . Problem (4.2) in radial coordinates reads:
\begin{equation} \left\{ \begin{aligned} & U_{\gamma}^{(4)}(r) + \dfrac{6}{r} U_{\gamma}'''(r) + \dfrac{3}{r^{2}} U_{\gamma}''(r) - \dfrac{3}{r^{3}} U_{\gamma}'(r) = r_{0}^{4} \left(|v(r_{0})|-\gamma U_{\gamma}(r) \right) \qquad \forall r\in(0,1) \\[6pt] & U_{\gamma}(1) = U_{\gamma}'(1) = 0\, . \end{aligned} \right. \end{equation} | (4.4) |
Upon substitution of (4.3) into (4.4) _1 we obtain
192A_{2}(\gamma)+16\sum\limits_{k = 3}^{\infty}A_{k}(\gamma) k^{2}(k^{2}-1) r^{2k - 4} \\ = r_{0}^{4} (|v(r_{0})| - \gamma A_{0}(\gamma)) - \gamma r_{0}^{4} \sum\limits_{k = 3}^{\infty} A_{k-2}(\gamma) r^{2k - 4} \qquad \forall r \in (0,1), |
which yields the identities
\begin{equation} A_{2}(\gamma) = \dfrac{r_{0}^{4}}{192}(|v(r_{0})| - \gamma A_{0}(\gamma)) \, , \qquad A_{k}(\gamma) = - \dfrac{\gamma r_{0}^{4}}{16 k^{2}(k^{2}-1)} A_{k-2}(\gamma) \qquad \forall k \geq 3 \, . \end{equation} | (4.5) |
Moreover, the boundary conditions in (4.4) _2 imply that
\sum\limits_{k = 0}^{\infty} A_{k}(\gamma) = 0 \, , \ \sum\limits_{k = 1}^{\infty} k A_{k}(\gamma) = 0\ \Longrightarrow\ A_1(\gamma) = -\sum\limits_{k = 2}^{\infty}kA_k(\gamma)\, , \ A_0(\gamma) = \sum\limits_{k = 2}^{\infty}(k-1) A_k(\gamma)\, . |
In connection with (3.23) we define
U_{*}(r) \doteq \sigma_{0} (1 - r^{2})^{2} \doteq \dfrac{r_{0}^{4} |v(r_{0})|}{64} (1 - r^{2})^{2} \qquad \forall r \in [0,1]. |
By Corollary (3.4) and by continuous dependence, we have that
U_{\gamma} \to U_{*} \quad \text{in} \ \ \mathcal{C}^{3}(\overline{ {\bf B}}) \ \ \text{as} \ \ \gamma \to 0^{+} \, ; |
in fact, the convergence holds in a stronger norm. Notice that
U''_{*}(0) = -4 \sigma_{0} < 0 \, , \qquad U''_{*}(1) = 8 \sigma_{0} > 0 \, . |
Thus, by uniform convergence of U_\gamma'' and U_\gamma''' , there exist \delta_{0} \in (0, 1) and \gamma_{0} > 0 such that:
U''_{\gamma}(r) \leq - 2 \sigma_{0} \quad \text{if} \quad r \in [0, \delta_{0}], \qquad U''_{\gamma}(r) \geq 4 \sigma_{0} \quad \text{if} \quad r \in [1-\delta_{0},1], \qquad \forall \gamma \in (0, \gamma_{0}). |
Since U'_{\gamma}(0) = U'_{\gamma}(1) = 0 for every \gamma > 0 , this implies that
U'_{\gamma}(r) \leq - 2 \sigma_{0}r < 0 \quad \text{if} \quad r \in (0, \delta_{0}], \\ U'_{\gamma}(r) \leq 4 \sigma_{0}(r-1) < 0 \quad \text{if} \quad r \in [1-\delta_{0},1), \qquad \forall \gamma \in (0, \gamma_{0}). |
Moreover, by uniform convergence of U'_{\gamma} , there exists \gamma_{1} > 0 such that U'_{\gamma}(r) < 0 if r \in [\delta_{0}, 1-\delta_{0}] , for every \gamma \in (0, \gamma_{1}) . Therefore U'_{\gamma} < 0 in (0, 1) whenever \gamma < \min \{ \gamma_{0} \, , \gamma_{1} \} , which also implies that U_{\gamma} is strictly positive in [0, 1) , since U_{\gamma}(1) = 0 .
In fact, the function U_{\gamma} is also unique within the class of weak solutions. Proposition 4.2 does not come unexpected, since the solution of (4.4) is closely related to a Bessel function, see [16]. Therefore, the number of critical points of U_{\gamma} is increasing with respect to \gamma and becomes arbitrarily large as \gamma \to \infty , whereas there are no critical points for any sufficiently small \gamma . By taking \gamma = 1 , we numerically obtain the solution displayed in the right picture of Figure 7, which represents the function u in (4.1). Several numerical simulations were performed in order to yield a radially decreasing solution of problem (4.2). The obtained results may be summarized as follows:
● if \gamma \leq 0.317 , then the associated solution U_{\gamma} of (4.2) is radially symmetric and strictly decreasing in the radial variable in the interval [0, 1] (and therefore, positive), see Figure 8;
● if \gamma \geq 0.318 , then the associated solution U_{\gamma} of (4.2) is radially symmetric but not strictly decreasing in the interval [0, 1] , see Figure 9.
We numerically determined the position of the maximum point r_{\gamma} \in [0, 1] of U_{\gamma} as a function of \gamma \in [0, 1] , see the Table and Figure below. From these data we deduce that the map \gamma \in [0, 1] \mapsto r_{\gamma} is increasing (but not continuous!), giving thus consistency to the discussion following Proposition 4.1. Finally, in connection with the condition (3.25) given in Propositions 3.7 and 4.1, we observe that, for the unique radial solution U_{\gamma} \in H^{2}_{0}({\bf B}) of problem (4.2), the following inequality holds:
f_{\gamma}(U_{\gamma}(r)) \geq r_{0}^{4} \left[|v(r_{0})|-\gamma \max\limits_{t \in [0,1]} |U_{\gamma}(t)| \right] \qquad \forall r \in [0,1]. |
From the Table and Figure displayed below we further deduce that
\begin{equation} f_{\gamma}(U_{\gamma}(r)) \geq r_{0}^{4} \left( |v(r_{0})|-\gamma U_{\gamma}(0) \right) \qquad \forall r \in [0,1], \quad \forall \gamma \in [0,0.317]; \end{equation} | (4.6) |
recall that U_{\gamma} is nonnegative in [0, 1] for every \gamma \in [0, 0.317] . Numerical experiments yield the value of \gamma = 0.057 as a threshold for ensuring the positivity of the right-hand side of (4.6), in the sense that
f_{0.057}(U_{0.057}(r)) \geq 0.014 \qquad \text{and} \qquad f_{0.058}(U_{0.058}(r)) \geq -0.092 \qquad \forall r \in [0,1]. |
The first Author is supported by the PRIN project Direct and inverse problems for partial differential equations: theoretical aspects and applications and by INdAM. The second Author is supported by the Research Programme PRIMUS/19/SCI/01, by the program GJ19-11707Y of the Czech National Grant Agency GA\v{C}R, and by the University Centre UNCE/SCI/023 of the Charles University in Prague.
The authors declare no conflict of interest.
[1] | A. D. Alexandrov, Uniqueness theorem for surfaces in the large, Vestnik Leningradskogo Universiteta, 11 (1956), 5–17. |
[2] |
A. D. Alexandrov, A characteristic property of spheres, Annali di Matematica, 58 (1962), 303–315. doi: 10.1007/BF02413056
![]() |
[3] | E. Berchio, F. Gazzola, T. Weth, Radial symmetry of positive solutions to nonlinear polyharmonic Dirichlet problems, J. reine angew. Math., 620 (2008), 165–183. |
[4] |
H. Berestycki, L. Nirenberg, On the method of moving planes and the sliding method, Bol. Soc. Bras. Mat., 22 (1991), 1–37. doi: 10.1007/BF01244896
![]() |
[5] |
T. Boggio, Sulle funzioni di Green d'ordine m, Rend. Circ. Matem. Palermo, 20 (1905), 97–135. doi: 10.1007/BF03014033
![]() |
[6] |
R. Dalmasso, Symmetry properties in higher order semilinear elliptic equations, Nonlinear Anal. Theor., 24 (1995), 1–7. doi: 10.1016/0362-546X(94)E0030-K
![]() |
[7] |
R. Dalmasso, Existence and uniqueness results for polyharmonic equations, Nonlinear Anal. Theor., 36 (1999), 131–137. doi: 10.1016/S0362-546X(98)00049-2
![]() |
[8] |
F. Gazzola, H.-C. Grunau, Radial entire solutions for supercritical biharmonic equations, Math. Ann., 334 (2006), 905–936. doi: 10.1007/s00208-005-0748-x
![]() |
[9] | F. Gazzola, H.-C. Grunau, G. Sweers, Polyharmonic boundary value problems: positivity preserving and nonlinear higher order elliptic, Springer Science & Business Media, 2010. |
[10] |
B. Gidas, W.-M. Ni, L. Nirenberg, Symmetry and related properties via the maximum principle, Commun. Math. Phys., 68 (1979), 209–243. doi: 10.1007/BF01221125
![]() |
[11] | H.-C. Grunau, G. Sweers, Positivity for equations involving polyharmonic operators with Dirichlet boundary conditions, Math. Ann., 307 (1996), 589–626. |
[12] |
H.-C. Grunau, G. Sweers, Positivity properties of elliptic boundary value problems of higher order, Nonlinear Anal. Theor., 30 (1997), 5251–5258. doi: 10.1016/S0362-546X(96)00164-2
![]() |
[13] |
S. Mayboroda, V. Maz'ya, Regularity of solutions to the polyharmonic equation in general domains, Invent. Math., 196 (2014), 1–68. doi: 10.1007/s00222-013-0464-1
![]() |
[14] |
C. Nitsch, C. Trombetti, The classical overdetermined Serrin problem, Complex Var. Elliptic Equ., 63 (2018), 1107–1122. doi: 10.1080/17476933.2017.1410798
![]() |
[15] |
J. Serrin, A symmetry problem in potential theory, Arch. Rational Mech. Anal., 43 (1971), 304–318. doi: 10.1007/BF00250468
![]() |
[16] | G. Sweers, No Gidas-Ni-Nirenberg type result for semilinear biharmonic problems, Math. Nachr., 246 (2002), 202–206. |
[17] |
W. C. Troy, Symmetry properties in systems of semilinear elliptic equations, J. Differ. Equations, 42 (1981), 400–413. doi: 10.1016/0022-0396(81)90113-3
![]() |
[18] |
H. F. Weinberger, Remark on the preceding paper of Serrin, Arch. Rational Mech. Anal., 43 (1971), 319–320. doi: 10.1007/BF00250469
![]() |
1. | Enno Lenzmann, Tobias Weth, Symmetry breaking for ground states of biharmonic NLS via Fourier extension estimates, 2024, 152, 0021-7670, 777, 10.1007/s11854-023-0311-2 |