Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Symmetry results for Serrin-type problems in doubly connected domains

  • In this work, we employ the technique developed in [2] to prove rotational symmetry for a class of Serrin-type problems for the standard Laplacian. We also discuss in some length how our strategy compares with the classical moving plane method.

    Citation: Stefano Borghini. Symmetry results for Serrin-type problems in doubly connected domains[J]. Mathematics in Engineering, 2023, 5(2): 1-16. doi: 10.3934/mine.2023027

    Related Papers:

    [1] Rolando Magnanini, Giorgio Poggesi . Interpolating estimates with applications to some quantitative symmetry results. Mathematics in Engineering, 2023, 5(1): 1-21. doi: 10.3934/mine.2023002
    [2] Filippo Gazzola, Gianmarco Sperone . Remarks on radial symmetry and monotonicity for solutions of semilinear higher order elliptic equations. Mathematics in Engineering, 2022, 4(5): 1-24. doi: 10.3934/mine.2022040
    [3] Chao Xia, Jiabin Yin . Two overdetermined problems for anisotropic p-Laplacian. Mathematics in Engineering, 2022, 4(2): 1-18. doi: 10.3934/mine.2022015
    [4] Pier Domenico Lamberti, Michele Zaccaron . Spectral stability of the curlcurl operator via uniform Gaffney inequalities on perturbed electromagnetic cavities. Mathematics in Engineering, 2023, 5(1): 1-31. doi: 10.3934/mine.2023018
    [5] Michiaki Onodera . Linear stability analysis of overdetermined problems with non-constant data. Mathematics in Engineering, 2023, 5(3): 1-18. doi: 10.3934/mine.2023048
    [6] E. Artioli, G. Elefante, A. Sommariva, M. Vianello . Homogenization of composite materials reinforced with unidirectional fibres with complex curvilinear cross section: a virtual element approach. Mathematics in Engineering, 2024, 6(4): 510-535. doi: 10.3934/mine.2024021
    [7] Masashi Misawa, Kenta Nakamura, Yoshihiko Yamaura . A volume constraint problem for the nonlocal doubly nonlinear parabolic equation. Mathematics in Engineering, 2023, 5(6): 1-26. doi: 10.3934/mine.2023098
    [8] 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
    [9] Antonio Greco, Francesco Pisanu . Improvements on overdetermined problems associated to the p-Laplacian. Mathematics in Engineering, 2022, 4(3): 1-14. doi: 10.3934/mine.2022017
    [10] Bruno Bianchini, Giulio Colombo, Marco Magliaro, Luciano Mari, Patrizia Pucci, Marco Rigoli . Recent rigidity results for graphs with prescribed mean curvature. Mathematics in Engineering, 2021, 3(5): 1-48. doi: 10.3934/mine.2021039
  • In this work, we employ the technique developed in [2] to prove rotational symmetry for a class of Serrin-type problems for the standard Laplacian. We also discuss in some length how our strategy compares with the classical moving plane method.



    In his pioneering paper [35], Serrin refined Alexandrov's method of moving planes [6] and employed it to prove that a function defined in a bounded domain having constant nonzero Laplacian and satisfying constant Dirichlet and Neumann boundary condition is necessarily rotationally symmetric. In fact, in [35] it was shown that similar results hold more generally for a vast class of elliptic PDEs. Serrin's result and the method of moving planes have given rise to a prolific field of research on overdetermined boundary value problems. Nowadays Serrin's technique has been employed to study analogous problems in a number of works, among which we mention [8,14,23,27]. An alternative proof of Serrin's result was given by Weinberger [38] using a completely different method, essentially based on a comparison argument with the model solutions. Weinberger's method and in particular its use of a P-function has also experienced quite some success [16,18,19,22,34]. For further insights on these methods and their applications, see also [28] and references therein.

    The above mentioned papers all gave rise to interesting characterizations of solutions supported in a ball. In this work, we are instead interested in characterizations of solutions supported in an annulus. Let us explain our setting in more detail. Let EiEoR2 be two nonempty simply connected bounded domains, whose boundaries Eo=Γo and Ei=Γi are simple closed smooth curves, and let Ω=Eo¯Ei. By construction, Ω is a bounded domain that is not simply connected, with boundary

    Ω=ΓiΓo.

    We are interested in studying pairs (Ω,u), where Ω has the above structure and u:ΩR is a solution to the following problem

    {Δu=2,in Ω,u=a,  uν=α,on Γi,u=b,  uν=β,on Γo, (1.1)

    where a,b,α,β are constants and ν is the unit-normal pointing outside Ω. This overdetermined boundary value problem has been well studied, especially in the case where a>b and α0. Under this assumption, the method of moving planes shows its strength and allows to prove some powerful rigidity results. A particularly nice one is the following, due to Reichel.

    Theorem 1.1. ([31,Theorem 2]). Let (Ω,u) be a solution to problem (1.1) such that a>b and b<u<a in Ω. Then Ω is an annulus and u is rotationally symmetric with u/|x|<0.

    It is worth mentioning that Reichel's result is slightly stronger, as it also works for more general PDEs and it allows Ei (and thus Γi) to be disconnected. In [36], Sirakov generalized previous results [1,5,40] and proved that Reichel's thesis remains true if one allows different values of a and α on different connected components of Γi. He was also able to replace the hypothesis u<a with the weaker α0. We also remark that similar symmetry results have been proven to hold for more general families of elliptic operators. In [36] it is shown that the same result works for quasi-linear regular strongly elliptic operator. Similar results also hold for quasi-linear possibly degenerate elliptic operators [5], fully nonlinear operators [32] and for the fractional Laplacian [37]. For further symmetry results on this and related problems, see also [7,16,17,29,33].

    On the other hand, it seems that less attention has been brought to the case a<b. One possible explaination is that the moving plane method works well when the model solution are monotonically decreasing along the radial coordinate, but seems to be harder to employ when the model solutions are monotonically increasing. In this work, our aim is to discuss this case and to prove the analogue of Theorem 1.1, namely:

    Theorem 1.2. Let (Ω,u) be a solution to problem (1.1) such that a<b and a<u<b in Ω. Then Ω is an annulus and u is rotationally symmetric with u/|x|>0.

    The strategy of the proof is based on the method developed in [2]. There, problem (1.1) has been studied in the case a=b=0, using a new comparison strategy. Of course, under that hypothesis, monotonicity is lost and there are some complications, especially concerning the behaviour of the solution near the set of the maxima of the function u. However, we will show that the core of the method employed in [2] is perfectly suited to study the case discussed in this paper. In fact, we can adapt most of the argument in [2] to prove Theorem 1.2.

    Perhaps surprisingly, our method proves to be less effective in the more studied case b<a. We will try to explain along the work why this is the case. Nevertheless, we will be able to employ our method to prove a slightly weaker version of Theorem 1.1 (namely, we will need to assume the additional condition 2a+α22b+β2, see Theorem 4.1). As we will see, the proof is also less elementary than the one for Theorem 1.2. It does seem then that the moving plane method is still the one best suited to characterize rotationally symmetric solutions that are radially decreasing. Nevertheless, our alternative proof may be of interest: it is novel and it expands on our method in ways that may find further applications in the future.

    Let us conclude with some final comments on the above results and on future applications. First of all, one natural question would be whether similar characterizations can be achieved for model solutions that are not monotonic along the radial coordinate. In this case, the situation is less clear as there are negative results: in [25] it is proven that there exist solutions to (1.1) with α=β that are not rotationally symmetric. In the case a=b=0, there are both positive ([2,Theorem B and Theorem C]) and negative results ([2,Theorem A]). Another natural question concerns whether our method can be applied to more general elliptic equations (in the same way as the Weinberger's comparison method [38] was then developed in [18,19,22]) or to other problems in Riemannian geometry. In fact, the most part of our arguments (with the notable exception of the Pohozaev identity) do not rely on the structure of the Euclidean space at all. As a couple of examples of further applications, this method has proven to be quite successful to characterize static spacetimes in General Relativity [10,11,12,13] and somewhat similar techniques have been employed for other problems in General Relativity [4,21] and for p-harmonic functions in manifolds with nonnegative Ricci curvature in [3,20].

    The paper is structured as follows. In Section 2 we show that for any acceptable choice of a,b,α,β there is a rotationally symmetric solution solving problem (1.1). This is necessary in order to start our method, as we need to select a model solution to compare with. Section 3 is the heart of the paper: a crucial gradient estimate (Theorem 3.1) is introduced and exploited to obtain some important length bounds for the boundary components (Proposition 3.2). In Subsection 3.3 these two results are then used, in combination with a divergence theorem argument, to prove Theorem 1.2. Unfortunately, the method described in Subsection 3.3 does not work to prove Theorem 1.1. In Section 4 we then develop a new alternative argument, based on a combination of the Pohozaev identity and the isoperimetric inequality, leading to the proof of Theorem 4.1, which is a weaker version of Theorem 1.1. Finally, in Section 5 we further comment on the complications that one has to deal with when studying the case a>b, trying to analyze where they come from and how one may try to overcome them. Ultimately, the purpose of this final section is that of understanding what are the limits of our method and how far they can be pushed.

    Our first aim is that of understanding whether there are some relations that are always in place between the constants a,α,b,β. The main tool that will help us in this regard is the well known Pohozaev identity [30]. In our framework, this formula has the following form

    Ω4udA=12Ω(4u+|u|2)X|νd=12(4b+β2)ΓoX|νd+12(4a+α2)ΓiX|νd,

    where X=(x1,x2) is the position vector and ν is the exterior normal to Ω. We use d, dA to denote the line element and area element respectively, whereas | is the standard scalar product between vectors of R2. Recalling that divX=2 and using the divergence theorem, we deduce

    Ω4udA=(4b+β2)|Eo|(4a+α2)|Ei|. (2.1)

    From this formula we deduce some relations between a,b,α,β, summarized in the following result.

    Proposition 2.1. Let (Ω,u) be a solution to problem (1.1).

    If a>b and b<u<a in Ω, then 4a+α2>4b+β2, α0, β<0,

    If a<b and a<u<b in Ω, then 4a+α2>4b+β2, α<0, β0.

    Proof. We start by recalling that Ω=Eo¯Ei, so that in particular |Eo|=|Ω|+|Ei|. As a consequence, we can rewrite (2.1) in the following two forms

    Ω(4u4aα2)dA=(4b+β24aα2)|Eo|,Ω(4u4bβ2)dA=(4b+β24aα2)|Ei|.

    If b<u<a in Ω, then the left hand side of the first equation is negative, whereas if a<u<b in Ω, then the left hand side of the second equation is negative. In both cases, we deduce 4a+α2>4b+β2.

    The signs for α and β follow immediately from the hypothesis that a is the maximum (resp. minimum) of u and b is the minimum (resp. maximum) of u. The fact that β0 when b<u<a and that α0 when a<u<b is a consequence of the Hopf Lemma (recall Δu=2<0).

    In order to start our comparison argument, we first need to take a close look at the model solutions. In particular, we first have to make sure that we always have a model solution to compare with, for any acceptable value of a,b,α,β.

    It can be easily checked that the rotationally symmetric solutions to Δu=2 have the form

    u=L12|x|2+Mlog|x|, (2.2)

    with L,MR. When M0, these solutions are monotonically decreasing with respect to |x|. When M>0, these solutions are monotonically increasing for |x|<M and they are monotonically decreasing for |x|>M.

    Remark 1. As in [2], one can fix the value of L by means of a rescaling of the function and of the domain, ending up with the one-parameter family

    u=1|x|22+Mlog|x|.

    However, for our purposes in this paper this is not necessary, as it is actually easier to work directly with (2.2).

    We expect that for any acceptable choice (according to Proposition 2.1) of a,b,α,β, there should be a solution of the form (2.2) solving our problem (1.1). This is true in the case a<b, as guaranteed by the following lemma:

    Lemma 2.2. If a<b, α<0, β0, 4a+α2>4b+β2, then there exist constants L and M>0 and two radii 0<ri<roM so that the function u=L|x|2/2+Mlog|x| satisfies problem (1.1) in the annulus Ω={ri<|x|<ro}.

    Proof. The proof of the lemma is more of a long exercise. We start by imposing the conditions we want on L,M,ro,ri, namely

    {Lr2i/2+Mlogri=aLr2o/2+Mlogro=bMr2i=αriMr2o=βro (2.3)

    and we employ the last two equations to obtain

    ri=α±α2+4M2,ro=β±β2+4M2. (2.4)

    Since we are assuming α<0 and β0, in order for ro,ri to be positive, we need to choose the plus sign in (2.4) and we need M>0.

    Substituting in the first two equations, we then deduce that we have a solution if and only if we can find a zero for the function

    F(M)=4a+α24bβ2+αα2+4M+ββ2+4M+4Mlog(β+β2+4Mα+α2+4M). (2.5)

    Taking the limit at zero and infinity, we compute

    limM0F(M)=4a4b<0,limM+F(M)=4a+α24bβ2>0.

    It follows by continuity of F that there is a value 0<M< such that F(M)=0, as wished.

    The case a>b appears to be more complicated. In the next lemma we will show that we can still find a model to compare with as long as we assume the additional condition 2a+α22b+β2. This is the first instance in which we notice complications in the case a>b. More specifically, it seems that negative values of M are harder to deal with. In this respect, the hypothesis 2a+α22b+β2 is helpful as it forces M0.

    Lemma 2.3. Let a>b, α0, β<0, 4a+α2>4b+β2. If we further have 2a+α22b+β2, then there exist constants L and M0 and two radii Mri<ro so that the function u=L|x|2/2+Mlog|x| satisfies problem (1.1) in the annulus Ω={ri<|x|<ro}.

    Proof. Proceeding as in the proof of Lemma 2.2, we find again the expressions (2.4) for ro and ri. Notice that, in contrast with the previous case, both signs for ± in (2.4) are possible. However, since we are ultimately interested in having M0, this again essentially forces us to choose the plus sign in the formulæ (2.4) for ri and ro. Therefore, as in the previous lemma, we reduce ourselves to finding a zero for the function (2.5). We compute

    F(0)=4a+2α24b2β2,limM+F(M)=4a+α24bβ2>0.

    Since we are assuming 4a+2α24b2β20, we conclude again that there is a value 0M< such that F(M)=0.

    In the previous section we have developed a way to compare a general solution of (1.1) with a model solution (2.2). Building on that, in this section we develop our comparison technique. In Subsection 3.1 we prove a crucial gradient estimate, that is then exploited in Subsection 3.2 to prove length bounds for the boundary components Γi and Γo. The results in both these subsection work in both the case a<b and b<a. In Subsection 3.3 we will then specialize to the case a<b and we will prove Theorem 1.2.

    Let u be a solution to problem (1.1) for some fixed values of a,α,b,β. In the case a>b, assume further that 2a+α22b+β2. Then, Lemmata 2.2 and 2.3 tell us that there exist constants LR, M0, ro>0, ri>0 such that the model solution

    u=L12|x|2+Mlog|x|,

    in the annulus {ri<|x|<ro}, solves problem (1.1) for the same values of a,α,b,β.

    From now on, the constants LR, M0, ro>0, ri>0 will always be the ones prescribed by Lemmata 2.2 and 2.3. We are now ready to set up our comparison argument, in the spirit of [2]. We start by defining the pseudo-radial function Ψ:ΩR implicitly via the following identity

    u=L12Ψ2+MlogΨ. (3.1)

    It is easily seen that the function G(Ψ)=LΨ2/2+MlogΨ is monotone in (0,M] and [M,+), hence the pseudo-radial function Ψ=G1(u) is well defined as long as Ψ2M inside Ω. From now on we will always assume that a<u<b or b<u<a, and from this it easily follows that the function Ψ actually takes values between ri (attained at Γi) and ro (attained at Γo). On the other hand, Lemmata 2.2 and 2.3 tell us that either ri<roM or Mri<ro. Therefore, for all our purposes, the function Ψ:Ω[ri,ro] will always be well defined.

    We also introduce the two functions

    W=|u|2andW0=(MΨ2Ψ)2. (3.2)

    The function W0 corresponds to the value that |u|2 would have on the model solution. In fact, it is easily checked that, for the model solutions (2.2), it holds W0W pointwise. Our strategy is then that of comparing W and W0 in Ω, ultimately trying to prove that they have to coincide.

    First of all, notice that, by construction, W and W0 coincide on the boundary of Ω, namely when Ψ=ro and when Ψ=ri. We will now show that, under our hypotheses, W is actually controlled by W0 on the whole Ω.

    Theorem 3.1 (Gradient Estimates). Let (Ω,u) be a solution to problem (1.1) with respect to some choice of a,α,b,β. Suppose that a<u<b or b<u<a. In the case b<u<a, assume further that 2a+α22b+β2. Let W, W0 be defined as above. Then it holds

    WW0inΩ, (3.3)

    Moreover, if W=W0 at some point in the interior of Ω, then (Ω,u) is rotationally symmetric.

    Remark 2. This theorem is clearly inspired by [2,Theorem 3.1], however we mention that similar gradient estimates have already found applications in a number of problems. One of the most notable ones is the paper [9] and the subsequent developments in [15,26], where a gradient comparison argument very much resembling Theorem 3.1 has been obtained and exploited for static spacetimes in General Relativity. The introduction of the pseudo-radial function is instead more recent (it was exploited in the series of papers [11,12,13] for static spacetimes). This function is really helpful as it allows to have an explicit formula for W0. This is crucial in the computations that follows: we will see that Ψ will appear in all the important computations in this section.

    Proof. The proof is essentially the same as the one in [2], but let us give some comments. Let LR, M0, ro>0, ri>0 be the constants prescribed by Lemmata 2.2 and 2.3.

    The crucial estimate is deduced starting from the following quantity

    2u+2M(MΨ2)2dudu+[1M(MΨ2)2|u|2]gR2,

    where dudu is the 2×2 matrix with entries uxiuxj and gR2 is the 2×2 identity matrix. The heuristic behind the choice of the quantity above is essentially the fact that it can be computed to be zero on the model solution. Computing explicitly its square norm and using the fact that it is nonnegative, we obtain the following estimate for the hessian of u:

    |2u|22M(MΨ2)2(WW0)|u+2[1+2M2Ψ2(MΨ2)2|u|2M2(MΨ2)4|u|4].

    Starting now from the equality Δ|u|2=2|2u|2 and plugging in this estimate, after some computations we obtain an elliptic inequality for the quantity WW0

    Δ(WW0)4M(MΨ2)2(WW0)|u+4M(MΨ2)2[1M(MΨ2)2|u|2](WW0).

    Considering the function Fγ=γ(WW0), where γ=γ(Ψ)>0, one gets

    Fγ2ΨMΨ2[γγ2MΨ(MΨ2)]Fγ|u2ΨMΨ2[γγ2MΨ(MΨ2)]Fγ+WW0[(γγ)(γγ)2+Ψ2+5MΨ(MΨ2)γγ4M2Ψ2(MΨ2)2]Fγ, (3.4)

    where we have used to denote the differentiation with respect to Ψ. We now need to find a function γ such that the coefficients of the zero order terms have the right sign. A good choice is to set

    γγ=2MΨ(MΨ2),

    which corresponds to choosing

    γ=ΨW0=Ψ2|MΨ2|.

    We know that Ψ2M inside Ω, hence γ is well defined in the interior of Ω. However, Ψ2 may attain the value M on the boundary Ω: this happens when α=0 (which implies from (2.4) that ri=M) or when β=0 (which implies from (2.4) that ro=M). Let us for the moment assume that this does not happen, so that Fγ is well defined on the whole Ω. With this choice of γ, it is easily seen that Fγ satisfies

    ΔFγ8MΨ2(MΨ2)4|u|2Fγ0. (3.5)

    Since M0, it follows that Fγ satisfies the Maximum Principle in Ω. Furthermore, W=W0 on Ω by construction, hence the Maximum Principle implies that Fγ0 (equivalently, WW0) on the whole Ω. Furthermore, if the equality W=W0 holds at one point p in the interior of N, then, applying the Strong Maximum Principle in a neighborhood of p, we deduce that W=W0 on the whole Ω. It is then easy to conclude the desired rigidity statement with standard arguments, see for instance [10,Theorem 4.2].

    It remains to discuss the case where Ψ2=M on one of the boundary components, say Γi (for Γo, the same argument applies). In this case, we consider ε>0 and a small neighborhood Uε={xΩ:d(x,Γi)<ε} of Γi, and we work on the domain Ωε=ΩUε. Clearly Fγ is well defined and satisfies the Maximum Principle in Ωε. We will then obtain the desired result by taking ε0, provided we can prove that Fγ0 as we approach Γi. It is easily seen that Fγ goes to zero if and only if W/W0 goes to zero, which in turn is equivalent to |u|2/au0, see [2,Lemma A.1]. The latter is granted by the Reverse Łojasiewicz Inequality [11,Theorem 2.2] (to be more precise, [11,Theorem 2.2] cannot be applied directly, as it is written for interior points only; however the proof extends without modifications to the case at hand).

    Following again [2], we can now exploit the gradient estimate proven above to obtain length bounds for both Γi and Γo.

    Proposition 3.2. Let (Ω,u) be a solution to problem (1.1) with respect to some choice of a,α,b,β. Suppose that a<u<b or b<u<a. In the case b<u<a, assume further that 2a+α22b+β2. Then

    |Γi|2πriand|Γo|2πro. (3.6)

    Proof. We will prove the length bound for Γi only, as the one for Γo is proven in the same way (up to some sign differences coming from the fact that Ψ is pointing outside Ω on Γo). Let us also suppose for the moment that α0, so that the gradient of u does not vanish on Γi. Recalling that Ψ=G1(u), where G(x)=Lx2/2+Mlogx, we easily compute

    Ψ=ΨMΨ2u=uW0.

    In particular the norm |Ψ|=W/W0 is equal to 1 on Γi by construction and is less than 1 inside Ω thanks to (3.3). In particular, the inequality |Ψ|2|Ψ0 holds on Γi and the curvature κ of Γi at any point of Γi satisfies

    κ=|Ψ|2ΔΨ12|Ψ|2|Ψ|Ψ|3=ΔΨ12|Ψ|2|ΨΔΨ=1/Ψ=1/ri. (3.7)

    Integrating κri on Γi, recalling that the total curvature of a simple curve is equal to 2π, we get

    2π|Γi|/ri.

    This concludes the proof of the length bound for Γi when α0.

    If instead α=0, we have W=W0=0 on Γi, so |Ψ|=W/W0 is not well defined and we need to find a workaround. Let pΓi and let us extend u to a smooth function in a neighborhood U of p. We can do it by means of Whitney Extension Theorem [39], that clearly applies to our case since u is smooth up to the boundary (see for instance [24,Theorem 6.19]). With a slight abuse of notation, let us still call u the smooth function defined on U that coincides with the original u in ¯ΩU. In the spirit of [11,Theorem 3.1], we consider the signed distance function r to Γi. Namely, |r(x)|=dist(x,Γi) and r>0 inside Ω, r<0 outside Ω. If U is taken to be small enough, the function r is known to be smooth with |r|1 in U. We can then proceed exactly as in [11,Theorem 3.1] to show that u=a+(Δu/2)r2+r3ω in U, where ω is a smooth function. In particular, in ¯ΩU we have

    u=ar2+r3ω. (3.8)

    Recalling Δu=2, from this expansion we easily compute that Δr=3ω on Γi. Furthermore, recalling that u=G(Ψ)=LΨ2/2+r2ilogΨ (notice that, if α=0, from (2.3) we have M=r2i), starting from the Taylor series for G at ri, with some computations one also gets an expansion for Ψ in ¯ΩU:

    Ψ=ri+r+(16riω2)r2+o(r3).

    Computing |Ψ| from this expansion using |r|=1 in U and recalling that |Ψ|1 thanks to the gradient estimates, we deduce 3ω1/ri. We now use the fact that Γi={r=0} is a level set of r to compute its curvature as follows

    κ=|r|2Δr12|r|2|r|r|3=Δr=3ω1/ri.

    Integrating on Γi we then obtain the desired length bound.

    In this subsection, we focus on the case a<b and we prove Theorem 1.2. For simplicity, let us start by discussing the case where u does not vanish on the boundary of Ω, namely for the moment let us assume β0 (notice that we know that α is strictly positive in the case a<b, so α cannot vanish). Using the divergence theorem and recalling that Δu=2, we compute

    Ω2MΨ2dA=ΩΔuMΨ2dA=Ω(1MΨ2)|udAΩu|νMΨ2d=Ω2Ψ2(MΨ2)3|u|2dAΩu|νMΨ2d, (3.9)

    where ν is the outward unit normal to Ω=ΓiΓo. Since we are in the case a<b, then ν=u/|u| on Γi and ν=u/|u| on Γo. Setting W=|u|2 as usual, and recalling the expression (3.2) of W0 in terms of Ψ, the identity above can be written as

    Ω2Ψ2(MΨ2)3(W0W)dA=Γi|u|MΨ2dΓo|u|MΨ2d. (3.10)

    Notice that the left hand side is nonnegative because of (3.3). Furthermore, since we are focusing on the case a<b, we have Ψ<ro<M hence

    |u|MΨ2=1ΨWW0.

    Since by construction W=W0 on Ω, from (3.10) we immediately deduce

    0|Γi|ri|Γo|ro. (3.11)

    On the other hand, the length bounds in Proposition 3.2 give us the opposite inequality. Therefore the equality must hold in (3.11), and from (3.10) in particular we obtain that WW0 in Ω. The result now follows from the rigidity statement in Theorem 3.1.

    It remains to discuss the case where β is equal to zero. In this case, the boundary term in (3.9) is ill-defined, as one has Ψ2=r2o=M on Γo. To avoid this problem, we just need to apply the divergence theorem in Ω{u>bε}, and then take the limit as ε0. From (3.8) it follows easily that the quantity u|ν/(MΨ2) goes to 1/Ψ=1/ro as we approach Γo. The proof then proceeds exactly as in the case β0.

    In this section we focus on the case a>b and we prove Theorem 1.1 under the additional hypothesis that 2a+α22b+β2. Let us give the precise statement for the convenience of the reader.

    Theorem 4.1. Let (Ω,u) be a solution to problem (1.1) such that a>b, b<u<a in Ω and 2a+α22b+β2. Then Ω is an annulus and u is rotationally symmetric with u/|x|<0.

    As discussed, all the arguments in Subsections 3.1 and 3.2 still work in this case. Unfortunately however, the strategy employed in Subsection 3.3 to complete the proof for a<b, does not work when a>b. In fact, one can check that in this case the inequality that one gets at the end has the opposite sign of (3.11), hence it does not combine with the length bounds from Proposition 3.2 anymore. We will then need to exploit a more delicate argument based on sharp estimates coming from the Pohozaev identity and the isoperimetric inequality.

    Let us start by writing

    4u=2φ+(4u2φ)=φΔu+(4u2φ),

    where

    φ=2uΨ44MΨ2+4M2logΨ+k2(MΨ2), (4.1)

    for kR constant to be chosen later. The function φ is constant on the level sets of u, hence we can consider the function ˙φ representing the derivative of φ with respect to u. More precisely, ˙φ:ΩR is the function satisfying φ=˙φu. We then use the divergence theorem to write the following sequence of identities

    Ω4udA=Ω[φΔu+(4u2φ)]dA=Ω[˙φ|u|2+4u2φ]dAΩφuνd=Ω˙φ(W2φ4u˙φ)dAαφi|Γi|βφo|Γo|, (4.2)

    where we have denoted by φi and φo the (constant) value of φ on Γi, Γo, respectively.

    The choice of φ in (4.1) has been done so that it holds (2φ4u)/˙φ=W0. Combining (4.2) with the Pohozaev identity (2.1), we obtain

    Ω˙φ(WW0)dA=(4b+β2)|Eo|(4a+α2)|Ei|+αφi|Γi|+βφo|Γo|.

    Since |Eo|=|Ei|+|Ω|, we can write the above equation in the following way:

    Ω˙φ(WW0)dA=(4b+β24aα2)|Ei|+(4b+β2)|Ω|+αφi|Γi|+βφo|Γo|.

    Recall from Proposition 2.1 that 4a+α2>4b+β2. Furthermore, combining the isoperimetric inequality with the length bound in Proposition 3.2 we get

    |Ei||Γi|24π12ri|Γi|.

    Finally, we have the identity

    2|Ω|=ΩΔudA=Ωuνd=α|Γi|+β|Γo|.

    Combining these insights together, we obtain the following inequality:

    Ω˙φ(WW0)dA[4b+β24aα22ri+αφi2bαβ22α]|Γi|+[βφo2bββ32]|Γo|.

    Recall that the constants appearing in the above formula are related as follows:

    a=Lr2i2+Mlogri,b=Lr2o2+Mlogro,α=riMri,β=Mroro.

    Using these identities and formula (4.1) for φ, with some computations we can rewrite the above inequality as

    Ω˙φ(WW0)dAM(4b+β24LM)+k2(|Γi|ri|Γo|ro). (4.3)

    Let us now analyze ˙φ. Differentiating (4.1), we get

    ˙φ=Ψ2(MΨ2)3[4MΨ2Ψ44M2logΨk].

    One can check that the quantity in square brakets is monotonically decreasing in Ψ. Since we know from Lemma 2.3 that Ψ2r2iM, in order for ˙φ to be positive in Ω it is then sufficient to choose

    k4Mr2ir4i4M2logri=4LM+M24aMα2r2i

    Choosing k exactly equal to the above value, recalling from (3.3) that WW0, we then get from (4.3):

    0M(4b+β24aα2r2i/M)2(|Γi|ri|Γo|ro)

    Since r2iM>0 and 4b+β24aα2<0, this implies

    |Γi|ri|Γo|ro0

    But we have the opposite inequality from Proposition 3.2, therefore everything must be an equality. As a consequence, W=W0 on the whole Ω and we conclude using the rigidity statement in Theorem 3.1.

    In the previous sections we have shown that we are able to deal with the case a>b and prove Theorem 1.1 if we assume the additional hypothesis 2a+α22b+β2. While this hypothesis is somewhat restrictive, it still allows to get some nontrivial applications. As a particularly relevant example, we now argue that our weaker version of Theorem 1.1, namely Theorem 4.1, is still strong enough to deal with the cases of interest in [2]. There, Theorem 1.1 was invoked in the proof of [2,Theorem B] on a domain Ωo (the outer domain, in the terminology of [2]) for a function u satisfying problem (1.1) with α=0, b=0 and β2/2a1 (the latter inequality followed from [2,Theorem 2.1]). Under those hypotheses, the inequality 2a+α22b+β2 is trivially satisfied. In other words, Theorem 4.1 is enough for the intended applications in [2].

    The rest of this section is devoted to further comment on the hypothesis 2a+α22b+β2. Specifically, we discuss what can go wrong when this hypothesis is not in place and what are the complications that one would need to overcome in order to prove Theorem 1.1 in full generality using our approach.

    The first basic step that one needs in order to start our comparison argument is of course to have a model to compare with. In other words, given a solution (Ω,u) of problem (1.1) for some a,b,α,β with a>b, we need to have a model solution (2.2) solving the same problem. This is granted by Lemma 2.3 when 2a+α22b+β2. Unfortunately, an analogous result seems harder to prove when 2a+α2>2b+β2.

    In case we are able to show that there is a model solution to compare with, there is still another crucial complication, namely the proof of Theorem 3.1. In fact, in order to prove the gradient estimate WW0, we relied on the Maximum Principle applied to the elliptic inequality (3.5). Unfortunately, when M<0 (this can happen when 2a+α2>2b+β2), the zeroth order term of (3.5) has the wrong sign, hence we cannot apply the Maximum Principle. It is possible with some work to find a workaround, at least under the additional hypothesis that r2iM.

    Theorem 5.1 (Gradient Estimates, M<0). Let (Ω,u) be a solution to problem (1.1) and suppose that b<u<a. Suppose that there exist L,M,ro,ri such that the corresponding model solution (2.2) solves the problem for the same a,b,α,β. Let W, W0 be defined as in (3.2). If r2iM>0, then it holds

    WW0inΩ, (5.1)

    Moreover, if W=W0 at some point in the interior of Ω, then (Ω,u) is rotationally symmetric.

    Proof. We start by observing that we can write (3.4) as

    Fγ2ΨΨ2M[γγ+2MΨ(Ψ2M)]Fγ|u2ΨΨ2M[γγ+2MΨ(Ψ2M)]F2γγW0+WW0[(γγ)(γγ)2+Ψ25MΨ(Ψ2M)γγ+4MΨ2(Ψ2M)]Fγ. (5.2)

    If we ask the quantity in the last square bracket to be equal to zero, we obtain the following formula for γ

    γ=Ψ2|2Ψ2(Ψ2+M)(logΨk)|.

    where kR is a constant. In particular we get

    γγ+2MΨ(Ψ2M)=Ψ4+4MΨ2M24MΨ2(logΨk)Ψ(Ψ2M)[2Ψ2(Ψ2+M)(logΨk)]. (5.3)

    Since Ψ2r2iM>0, choosing k to be big enough we have that the right hand side of (5.3) is negative at any point in Ω. As a consequence, for such k, from (5.2) we get

    ΔFγ2Ψ4+4MΨ2M24MΨ2(logΨk)(Ψ2M)2[2Ψ2(Ψ2+M)(logΨk)]Fγ|u0.

    We can then apply the Maximum Principle to find out that Fγ is nonpositive on the whole Ω. This concludes the proof.

    The question still remains open of whether it is possible to remove the hypothesis r2iM in the above theorem. Looking at the proof, it is clear that the freedom in the choice of k should allow to weaken that hypothesis. However, it does seem hard to remove it entirely.

    The rest of our arguments does not seem to depend heavily on the hypothesis 2a+α22b+β2. In other words, even in the case 2a+α2>2b+β2, our method should allow to conclude the desired rigidity, provided we have a model solution to compare with and the gradient estimate WW0 is in force. In order to prove Theorem 1.1 in full generality using our approach, these seem to be the two issues that are left to address.

    The author is a member of the Gruppo Nazionale per l'Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM).

    The author declares no conflict of interest.



    [1] A. Aftalion, J. Busca, Radial symmetry of overdetermined boundary-value problems in exterior domains, Arch. Rational Mech. Anal., 143 (1998), 195–206. http://dx.doi.org/10.1007/s002050050103 doi: 10.1007/s002050050103
    [2] V. Agostiniani, S. Borghini, L. Mazzieri, On the Serrin problem for ring-shaped domains, 2021, arXiv: 2109.11255.
    [3] V. Agostiniani, M. Fogagnolo, L. Mazzieri, Sharp geometric inequalities for closed hypersurfaces in manifolds with nonnegative Ricci curvature, Invent. Math., 222 (2020), 1033–1101. http://dx.doi.org/10.1007/s00222-020-00985-4 doi: 10.1007/s00222-020-00985-4
    [4] V. Agostiniani, L. Mazzieri, F. Oronzio, A Green's function proof of the positive mass theorem, 2021, arXiv: 2108.08402.
    [5] G. Alessandrini, A symmetry theorem for condensers, Math. Method. Appl. Sci., 15 (1992), 315–320. http://dx.doi.org/10.1002/mma.1670150503 doi: 10.1002/mma.1670150503
    [6] A. D. Alexandrov, A characteristic property of spheres, Annali di Matematica, 58 (1962), 303–315. http://dx.doi.org/10.1007/BF02413056 doi: 10.1007/BF02413056
    [7] L. Barbu, C. Enache, A free boundary problem with multiple boundaries for a general class of anisotropic equations, Appl. Math. Comput., 362 (2019), 124551. http://dx.doi.org/10.1016/j.amc.2019.06.065 doi: 10.1016/j.amc.2019.06.065
    [8] B. Barrios, L. Montoro, B. Sciunzi, On the moving plane method for nonlocal problems in bounded domains, JAMA, 135 (2018), 37–57. http://dx.doi.org/10.1007/s11854-018-0031-1 doi: 10.1007/s11854-018-0031-1
    [9] R. Beig, W. Simon, On the uniqueness of static perfect-fluid solutions in general relativity, Commun. Math. Phys., 144 (1992), 373–390. https://dx.doi.org/10.1007/BF02101098 doi: 10.1007/BF02101098
    [10] S. Borghini, Static Black Hole uniqueness for nonpositive masses, Nonlinear Anal., 220 (2022), 112843. http://dx.doi.org/10.1016/j.na.2022.112843 doi: 10.1016/j.na.2022.112843
    [11] S. Borghini, P. T. Chruściel, L. Mazzieri, On the uniqueness of Schwarzschild-de Sitter spacetime, 2019, arXiv: 1909.05941.
    [12] S. Borghini, L. Mazzieri, On the mass of static metrics with positive cosmological constant: Ⅰ, Class. Quantum Grav., 35 (2018), 125001. http://dx.doi.org/10.1088/1361-6382/aac081 doi: 10.1088/1361-6382/aac081
    [13] S. Borghini, L. Mazzieri, On the mass of static metrics with positive cosmological constant: Ⅱ, Commun. Math. Phys., 377 (2020), 2079–2158. http://dx.doi.org/10.1007/s00220-020-03739-8 doi: 10.1007/s00220-020-03739-8
    [14] B. Brandolini, C. Nitsch, P. Salani, C. Trombetti, On the stability of the Serrin problem, J. Differ. Equations, 245 (2008), 1566–1583. http://dx.doi.org/10.1016/j.jde.2008.06.010 doi: 10.1016/j.jde.2008.06.010
    [15] P. T. Chruściel, W. Simon, Towards the classification of static vacuum spacetimes with negative cosmological constant, J. Math. Phys., 42 (2001), 1779–1817. http://dx.doi.org/10.1063/1.1340869 doi: 10.1063/1.1340869
    [16] G. Ciraolo, L. Vezzoni, On Serrin's overdetermined problem in space forms, Manuscripta Math., 159 (2019), 445–452. http://dx.doi.org/10.1007/s00229-018-1079-z doi: 10.1007/s00229-018-1079-z
    [17] A. Enciso, D. Peralta-Salas, Symmetry for an overdetermined boundary problem in a punctured domain, Nonlinear Anal. Theor., 70 (2009), 1080–1086. http://dx.doi.org/10.1016/j.na.2008.01.034 doi: 10.1016/j.na.2008.01.034
    [18] A. Farina, B. Kawohl, Remarks on an overdetermined boundary value problem, Calc. Var., 31 (2008), 351–357. http://dx.doi.org/10.1007/s00526-007-0115-8 doi: 10.1007/s00526-007-0115-8
    [19] A. Farina, E. Valdinoci, A pointwise gradient estimate in possibly unbounded domains with nonnegative mean curvature, Adv. Math., 225 (2010), 2808–2827. http://dx.doi.org/10.1016/j.aim.2010.05.008 doi: 10.1016/j.aim.2010.05.008
    [20] M. Fogagnolo, L. Mazzieri, A. Pinamonti, Geometric aspects of p-capacitary potentials, Ann. Inst. H. Poincaré Anal. Non Linéaire, 36 (2019), 1151–1179. http://dx.doi.org/10.1016/j.anihpc.2018.11.005 doi: 10.1016/j.anihpc.2018.11.005
    [21] M. Fogagnolo, A. Pinamonti, New integral estimates in substatic Riemannian manifolds and the Alexandrov theorem, 2021, arXiv: 2105.04672.
    [22] N. Garofalo, J. L. Lewis, A symmetry result related to some overdetermined boundary value problems, Amer. J. Math., 111 (1989), 9–33. http://dx.doi.org/10.2307/2374477 doi: 10.2307/2374477
    [23] B. Gidas, W. M. Ni, L. Nirenberg, Symmetry and related properties via the maximum principle, Commun. Math. Phys., 68 (1979), 209–243. https://dx.doi.org/10.1007/BF01221125 doi: 10.1007/BF01221125
    [24] D. Gilbarg, N. S. Trudinger, Elliptic partial differential equations of second order, 2 Eds., Berlin, Heidelberg: Springer, 2001. https://dx.doi.org/10.1007/978-3-642-61798-0
    [25] N. Kamburov, L. Sciaraffia, Nontrivial solutions to Serrin's problem in annular domains, Ann. Inst. H. Poincaré Anal. Non Linéaire, 38 (2021), 1–22. http://dx.doi.org/10.1016/j.anihpc.2020.05.001 doi: 10.1016/j.anihpc.2020.05.001
    [26] D. A. Lee, A. Neves, The Penrose inequality for asymptotically locally hyperbolic spaces with nonpositive mass, Commun. Math. Phys., 339 (2015), 327–352. http://dx.doi.org/10.1007/s00220-015-2421-x doi: 10.1007/s00220-015-2421-x
    [27] L. Ma, B. Liu, Symmetry results for decay solutions of elliptic systems in the whole space, Adv. Math., 225 (2010), 3052–3063. http://dx.doi.org/10.1016/j.aim.2010.05.022 doi: 10.1016/j.aim.2010.05.022
    [28] R. Magnanini, Alexandrov, Serrin, Weinberger, Reilly: simmetry and stability by integral identities, Bruno Pini Mathematical Analysis Seminar, Italy: University of Bologna, 2017,121–141. http://dx.doi.org/10.6092/issn.2240-2829/7800
    [29] L. E. Payne, G. A. Philippin, On two free boundary problems in potential theory, J. Math. Anal. Appl., 161 (1991), 332–342. http://dx.doi.org/10.1016/0022-247X(91)90333-U doi: 10.1016/0022-247X(91)90333-U
    [30] S. I. Pohožaev, On the eigenfunctions of the equation Δu+λf(u)=0, (Russian), Dokl. Akad. Nauk SSSR, 165 (1965), 36–39.
    [31] W. Reichel, Radial symmetry by moving planes for semilinear elliptic BVPs on annuli and other non-convex domains, In: Progress in partial differential equations: elliptic and parabolic problems, Math. Inst. I, 1995,164–182.
    [32] W. Reichel, Radial symmetry for an electrostatic, a capillarity and some fully nonlinear overdetermined problems on exterior domains, Z. Anal. Anwend., 15 (1996), 619–635. http://dx.doi.org/10.4171/ZAA/719 doi: 10.4171/ZAA/719
    [33] W. Reichel, Radial symmetry for elliptic boundary-value problems on exterior domains, Arch. Rational Mech. Anal., 137 (1997), 381–394. http://dx.doi.org/10.1007/s002050050034 doi: 10.1007/s002050050034
    [34] A. Roncoroni, A Serrin-type symmetry result on model manifolds: an extension of the Weinberger argument, C. R. Math., 356 (2018), 648–656. http://dx.doi.org/10.1016/j.crma.2018.04.012 doi: 10.1016/j.crma.2018.04.012
    [35] J. Serrin, A symmetry problem in potential theory, Arch. Rational Mech. Anal., 43 (1971), 304–318. http://dx.doi.org/10.1007/BF00250468 doi: 10.1007/BF00250468
    [36] B. Sirakov, Symmetry for exterior elliptic problems and two conjectures in potential theory, Ann. Inst. H. Poincaré Anal. Non Linéaire, 18 (2001), 135–156. http://dx.doi.org/10.1016/S0294-1449(00)00052-4 doi: 10.1016/S0294-1449(00)00052-4
    [37] N. Soave, E. Valdinoci, Overdetermined problems for the fractional Laplacian in exterior and annular sets, J. Anal. Math., 137 (2019), 101–134. http://dx.doi.org/10.1007/s11854-018-0067-2 doi: 10.1007/s11854-018-0067-2
    [38] H. F. Weinberger, Remark on the preceding paper of Serrin, Arch. Rational Mech. Anal., 43 (1971), 319–320. http://dx.doi.org/10.1007/BF00250469 doi: 10.1007/BF00250469
    [39] H. Whitney, Analytic extensions of differentiable functions defined in closed sets, Trans. Amer. Math. Soc., 36 (1934), 63–89. http://dx.doi.org/10.2307/1989708 doi: 10.2307/1989708
    [40] N. B. Willms, G. M. L. Gladwell, D. Siegel, Symmetry theorems for some overdetermined boundary value problems on ring domains, Z. Angew. Math. Phys., 45 (1994), 556–579. http://dx.doi.org/10.1007/BF00991897 doi: 10.1007/BF00991897
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1915) PDF downloads(205) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog