|
|
|
|
|
|
|
|
|
|
|
|
|
|
Citation: Nicola Abatangelo, Sven Jarohs, Alberto Saldaña. Fractional Laplacians on ellipsoids[J]. Mathematics in Engineering, 2021, 3(5): 1-34. doi: 10.3934/mine.2021038
[1] | María Ángeles García-Ferrero, Angkana Rüland . Strong unique continuation for the higher order fractional Laplacian. Mathematics in Engineering, 2019, 1(4): 715-774. doi: 10.3934/mine.2019.4.715 |
[2] | Lorenzo Brasco . Convex duality for principal frequencies. Mathematics in Engineering, 2022, 4(4): 1-28. doi: 10.3934/mine.2022032 |
[3] | Antonio Iannizzotto, Giovanni Porru . Optimization problems in rearrangement classes for fractional p-Laplacian equations. Mathematics in Engineering, 2025, 7(1): 13-34. doi: 10.3934/mine.2025002 |
[4] | 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 |
[5] | Alessandra De Luca, Veronica Felli . Unique continuation from the edge of a crack. Mathematics in Engineering, 2021, 3(3): 1-40. doi: 10.3934/mine.2021023 |
[6] | Luigi Provenzano, Alessandro Savo . Isoparametric foliations and the Pompeiu property. Mathematics in Engineering, 2023, 5(2): 1-27. doi: 10.3934/mine.2023031 |
[7] | Giovanni S. Alberti, Yves Capdeboscq, Yannick Privat . On the randomised stability constant for inverse problems. Mathematics in Engineering, 2020, 2(2): 264-286. doi: 10.3934/mine.2020013 |
[8] | Alessandro Michelangeli, Raffaele Scandone . On real resonances for three-dimensional Schrödinger operators with point interactions. Mathematics in Engineering, 2021, 3(2): 1-14. doi: 10.3934/mine.2021017 |
[9] | Catharine W. K. Lo, José Francisco Rodrigues . On the obstacle problem in fractional generalised Orlicz spaces. Mathematics in Engineering, 2024, 6(5): 676-704. doi: 10.3934/mine.2024026 |
[10] | Peter Bella, Mathias Schäffner . Local boundedness for p-Laplacian with degenerate coefficients. Mathematics in Engineering, 2023, 5(5): 1-20. doi: 10.3934/mine.2023081 |
The fractional Laplacian (−Δ)s, s>0, is a pseudodifferential operator with Fourier symbol |⋅|2s which can be evaluated pointwisely via a hypersingular integral (see (2.1) below). This operator has many applications in mathematical modeling and the set of solutions of boundary value problems involving the fractional Laplacian has a rich and complex mathematical structure, see [6,9,17].
One of the main obstacles in the study of this operator is the difficulty of evaluating explicitly (−Δ)s, even on simple functions, see for example [1,3,14,15] and the references therein for some of the few exceptions that are available in the literature. For the same reason, explicit solutions of boundary value problems are rare.
In this paper, we show some explicit formulas for the evaluation of the fractional Laplacian of polynomial-like functions supported in ellipsoids. Our first result concerns the explicit expression of the torsion function of an ellipsoid. Let
Hs0(Ω):={u∈Hs(Rn):u=0 in Rn∖Ω} for any s>0 |
and Hs(Rn) denotes the usual fractional Sobolev space of order s>0 (see, for example, [4], for standard existence and uniqueness results in this setting). If s=m∈N, then Hs0(Ω) is the usual Sobolev space Hm0(Ω).
Theorem 1.1. Let n≥2, s>0, A∈Rn×n be a symmetric positive definite matrix, and let
E:={x∈Rn:Ax⋅x<1}. |
Then, there is κ=κ(n,s,A)>0 such that us:Rn→R given by us(x):=(1−Ax⋅x)s+ solves pointwisely
(−Δ)sus=κin E, | (1.1) |
and us is the unique (weak) solution of (1.1) in Hs0(E).
Here f+ denotes the positive part of f. The explicit value of κ(n,s,a) can be computed in terms of hypergeometric functions 2F1 (see (2.8), (2.4), and (3.11)). In particular, for (two-dimensional) ellipses with axes of length 1√a1 and 1√a2 we have that
κ=4sΓ(1+s)2as+121a−1/222F1(s+1,12;1;1−a1a2) for a1,a2>0, |
see Remark 3.4. The name torsion function comes from elasticity theory, where u1 denotes the Prandtl torsion stress function describing the deformation of an elastic body subject to surface forces. The function u1 also has applications in fluid mechanics (modelling the pressure gradient of a flow in a viscous fluid), see [24] and the references therein. A solution of (1.1) in general domains for any s>0 is usually also called torsion function, and its explicit expression is often useful for checking inequalities and to formulate or disprove general conjectures (see, for example, [23,24,28]).
Theorem 1.1 relies on the following more general result, which is an extension of [15,Corollary 4] to ellipsoidal domains.
Theorem 1.2. Under the assumptions of Theorem 1.1, let j∈Z, j≥−⌊s⌋−1, and us+j(x)=(1−Ax⋅x)s+j+, x∈Rn. Then us+j solves pointwisely
(−Δ)sus+j=fjin E,us+j=0in Rn∖E, |
where fj is the polynomial of degree (2j)+ given by
fj(x)={Cn,s,jj∑k=0(−1)kΓ(s+12+k)Γ(12+k)(jk)∫∂E(u1(x)+(Ax⋅θ)2)j−k(Ax⋅θ)2k|θ|n+2s|Aθ|dθ,if j≥0,0,if j≤−1, | (1.2) |
and, under the convention Γ(t)−1=0 for t∈Z∖N,
Cn,s,j=22s−1Γ(n2+s)Γ(1+s+j)π(n−1)/2Γ(12+s)Γ(1+j). |
Theorem 1.2 can be in turn deduced as a particular case of Theorem 3.2 (see also Corollaries 3.3 and 3.5). The proof of Theorem 3.2 relies on direct computations mainly inspired by [14,15].
Using this approach, we can also calculate the evaluation of (−Δ)s of functions such as
xiusandx2ius | (1.3) |
for i=1,…,n, see Lemmas 3.6 and 3.7. With a similar strategy one may compute the fractional Laplacian (−Δ)s of xkius for any k∈N (although the length of the expressions increases considerably with k).
These formulas are of independent interest since, as mentioned earlier, there are very few examples of explicit computations regarding fractional Laplacians. However, one of our main motivations in studying these expressions is related to the problem of the positivity preserving property (p.p.p., from now on) for higher-order elliptic operators, which we describe next.
We say that the operator (−Δ)s satisfies a p.p.p. (in Ω) if
u≥0a.e.inΩ,wheneveru∈Hs0(Ω) and (−Δ)su≥0 pointwisely inΩ. | (1.4) |
Property (1.4) is sometimes called weak maximum principle and it holds for general domains if s∈(0,1]. The p.p.p. is one of the cornerstones in the analysis of linear and nonlinear second-order elliptic problems, and it is involved in results regarding existence of solutions, uniqueness, regularity, symmetry, monotonicity, geometry of level sets, etc.
Whenever s>1, the verification of (1.4) is a delicate issue; it can be shown that (1.4) holds for any s>0 whenever Ω is a ball [2,12] or a halfspace [1]; however, (1.4) does not hold in general. For s>1, the validity of (1.4) depends strongly on the geometry of Ω, but hitherto there is no way of knowing which domains satisfy (1.4) and which ones do not. The classification of domains satisfying (1.4) is a long-standing open problem in the theory of higher-order elliptic equations, see [18,Section 1.2].
One way of approaching this problem is to find first some examples of domains where (1.4) does not hold, and to try to identify a common nature. In particular, the ellipse is known to be incompatible with the p.p.p. whenever it is eccentric enough. This striking example shows that convexity, smoothness, and symmetry are not properties that guarantee the validity of (1.4). Next we include a list of references concerned with ellipses and the absence of a p.p.p.:
i) The first available result dates back to [16] for the bilaplacian s=2 in dimension n=2, where it is shown that an ellipse with axes ratio 5/3 does not satisfy (1.4). Later, in [22], it is mentioned a ratio of about 1.17 is enough.
ii) In [26] a machinery is designed to extend the two-dimensional examples to higher dimensions. We remark that this approach strongly relies on a separation of variables that is not available for the fractional Laplacian (2.1).
iii) For s=n=2, [31] builds an explicit and elementary example: an ellipse with axes ratio equal to 5; the explicit sign-changing solution is a polynomial of degree 7.
iv) A thorough analysis for s=n=2 is performed in [27], finding a counterexample in terms of a polynomial of degree 6 in an ellipse with axes ratio equal to √19≈4.359. The authors also show that it is not possible to construct a counterexample in an ellipse with polynomials of degree less than 6; moreover, it is also shown that counterexamples with degree 6 polynomials are only possible if the axes ratio is larger than ≈4.352 (this threshold also appears in our analysis, see Section 4.1).
v) The first example for s=3 and n=2 was given in [32]: in this case, the ellipse has an axes ratio equal to 12 and the explicit sign-changing solution is a polynomial of degree 8.
vi) Finally, [33] suggests that, for s=4 and the same ellipse as in [32], it is possible to find an explicit nodal solution which is a polynomial of degree 12.
Other domains where a general p.p.p. fails are some domains with corners [10] (in particular squares), cones [25], domains with holes [19], elongated rectangles [13], some large cylindrical domains [21], and some limaçons and cardioids [11]. For a survey on this subject for the bilaplacian in the context of the "Boggio-Hadamard conjecture", we refer to [18,Section 1.2] and the references therein.
All the techniques mentioned above are either incompatible or very hard to extend to the fractional setting s∈(0,∞)∖N, this case requires new ideas. Nevertheless, we believe that the study of p.p.p. in the fractional regime is relevant, since it offers a novel perspective on the subject using the continuity of the solution mapping, see [23].
For fractional powers there is only one known counterexample to (1.4), given in [4] (see also [2,Theorem 1.11]), where it is shown that, for s∈(k,k+1) with k a positive odd integer, two disjoint balls and dumbbell shaped domains do not satisfy p.p.p.
In the following, we show that, using our explicit computations in ellipsoids, we can construct counterexamples to (1.4) in any dimension n≥2 and for s∈(1,√3+3/2), where √3+3/2≈3.232. We follow the ideas from the above mentioned paper [31], where a counterexample in ellipses is built in terms of an explicit polynomial. For n≥2, let
Ea:={{x=(x1,…,xn)∈Rn:∑ni=1aix2i<1}, if a∈Rn with ai>0,{x=(x1,…,xn)∈Rn:∑n−1i=1x2i+ax2n<1}, if a>1. | (1.5) |
For functions in Hs0(Ea)∩C2s+γ(Ea) with γ>0 and s>1, the fractional Laplacian can be evaluated via the hypersingular integral (2.1), but it can also be evaluated as a composition of operators (see [5,Corollary 1.4]), namely,
(−Δ)su={(−Δ)(−Δ)s−1u for s∈(1,2),(−Δ)2(−Δ)s−2u for s∈(2,3). |
We emphasize that the order of the differential operators cannot be interchanged freely in the context of boundary value problems. For more details, see [5,29].
Theorem 1.3. Let n≥2 and s∈(1,2). There are a0=a0(s,n)>1 and 0ε=0ε(s,n)∈(0,1) such that, for every a>a0 and ε∈(0,0ε), the function Uε:Rn→R given by
Uϵ(x):=((1−x1)2−ϵ)(1−n−1∑i=1x2i−ax2n)s+,x∈Rn, |
belongs to Hs0(Ea), it changes sign in Ea, and (−Δ)sUϵ>0 in Ea.
For larger values of s one can still construct a counterexample, but the shape of Uε is slightly more involved.
Theorem 1.4. Let n≥2 and s∈(1,√3+3/2). There are constants a0>1, 0ε∈(0,1), γ≥0, and δ≥0, depending only on s and n, such that, for every a>a0 and ε∈(0,0ε), the function Uε:Rn→R given by
Uϵ(x):=(p(x)−ε)(1−n−1∑i=1x2i−ax2n)s+,where p(x):=(1−x1)2+γ(1−x1)−δ(n−1∑k=2x2k+ax2n), x∈Rn, | (1.6) |
belongs to Hs0(Ea), Uϵ changes sign in Ea, and (−Δ)sUϵ>0 in Ea.
We emphasize that Theorem 1.4 is the first counterexample to (1.4) in the range s∈(2,3). In contrast to the results in [31] and [32] which rely on explicit computations of polynomials that can be verified quickly with a computer, the fractional case is much more complex, even with the explicit form of the fractional Laplacian (−Δ)sUε, since these formulas are given in terms of hypergeometric functions which are in general difficult to manipulate. To overcome this difficulty, we use an asymptotic analysis as the length of one of the axis in the ellipsoid goes to zero; it turns out that a suitable normalization of the hypergeometric functions simplifies in the limit and its asymptotic behaviour can be determined with precision (see Lemma). This is enough to guarantee the positivity of (−Δ)sUε for thin enough ellipsoids.
As to the upper bound √3+3/2 for s in Theorem 1.4, it is a technical limitation of our asymptotic approach involving polynomials of the form (1.6). Surprisingly, for some (relatively) small values of a one can obtain counterexamples for slightly larger s (up to around 3.8), and we explore this fact in Section 4.1, where we do a computer-assisted analysis in two dimensions. We also remark that, as expected, a0↑∞ as s↓1, as can be seen in Figure 2.
We believe that counterexamples for any s>3 can be found in suitable ellipses, but this requires a more involved analysis with polynomials p of degree strictly higher than two, and we do not pursue this here. See the discussion in Section 4.1 and see [33] for a counterexample to the p.p.p. for s=4 in terms of a polynomial of degree 12.
Via a point inversion transformation, one can use Theorem 1.4 to show that a wide variety of shapes do not satisfy (1.4) either. To be more precise, in [1] (see also [12]) the following result is shown.
Proposition 1.5 (Proposition 1.6 in [1]). Let v∈Rn, c,s>0, u∈C∞c(Rn∖{−ν}), and x∈Rn∖{−ν}. Then
(−Δ)s(u∘σ(x)|x+ν|n−2s)=c2s(−Δ)su(σ(x))|x+ν|n+2s,whereσ(x):=cx+ν|x+ν|2−ν. | (1.7) |
To understand the geometrical meaning of the point inversion transformation σ, see Figure 4. We have the following consequences of Theorems 3.2, 1.4, and Proposition 1.5. Let n≥1, c>0, ν∈Rn∖∂Ea, and
Ω(a,c,ν):={{x∈Rn:n∑i=1ai(cxi+νi|x+ν|2−νi)2<1}, if a∈Rn with ai>0,{x∈Rn:n−1∑i=1(cxi+νi|x+ν|2−νi)2+a(cxn+νn|x+ν|2−νn)2<1}, if a>1. | (1.8) |
Corollary 1.6. Let n≥1, c>0, a∈Rn with ai>0, and ν∈Rn∖∂Ea. Then −ν∉¯Ω(a,c,ν) and, for s>0, the function
ws(x):=1|x+ν|n−2s(1−n∑i=1ai(cxi+νi|x+ν|2−νi)2)s+,x∈Rn, | (1.9) |
is a pointwise solution of
(−Δ)sws(x)=k|x+ν|n+2sin Ω(a,c,ν),ws=0in Rn∖Ω(a,c,ν) | (1.10) |
for some constant k=k(n,s,c,a)>0.
Corollary 1.7. Let n≥2, a,c>0, and ν∈Rn∖∂Ea such that Ω(a,c,ν) is a bounded domain. Then −ν∉¯Ω(a,c,ν) and, for every s∈(1,√3+3/2), there is a0=a0(s,n)>1 such that Ω(a,c,ν) does not satisfy (1.4) for every a>a0. For the case Ω(a,c,ν) unbounded, the claim still holds under the assumption n>4s.
To see some of the different (bounded and unbounded) domains represented by Ω(a,c,ν) for n=2 and n=3, see Figures 5 and 6 in Section 5.
The paper is organized as follows. In Section 2 we introduce some of the most relevant notation and important definitions. In Section 3 we show Theorems 1.1 and 1.2 and deduce the explicit formulas regarding functions of the type (1.3) in ellipsoids. Section 4 is devoted to the construction of counterexamples, and contains the proofs of Theorems 1.3 and 1.4, as well as those of Corollaries 1.6 and 1.7.
Any positive power s>0 of the (minus) Laplacian, i.e., (−Δ)s, has the same Fourier symbol (see [30,Chapter 5] or [5,Theorem 1.8]) as the following hypersingular integral,
Lm,su(x):=cn,m,s2∫Rnδmu(x,y)|y|n+2s dy,x∈Rn, | (2.1) |
where n∈N is the dimension, m∈N, s∈(0,m),
δmu(x,y):=m∑k=−m(−1)k(2mm−k)u(x+ky) for x,y∈Rn |
is a finite difference of order 2m, and cn,m,s is the positive constant given by
cn,m,s:={4sΓ(n2+s)πn/2Γ(−s)(m∑k=1(−1)k(2mm−k)k2s)−1,s∈(0,m)∖N,4sΓ(n2+s)s!2πn/2(m∑k=2(−1)k−s+1(2mm−k)k2sln(k))−1,s∈{1,…,m−1}. | (2.2) |
In particular, if ⌊s⌋ denotes the floor of s, then
(−Δ)su(x)=(−Δ)⌊s⌋(−Δ)s−⌊s⌋u(x)=Lm,su(x) |
for x∈Ω and for any u∈C2s+β(Ω)∩Hs0(Ω), with β>0, see [5,Corollary 1.4].
Let n≥1, a∈Rn, ai>0, and A=diag(ak)nk=1 a diagonal matrix. Then, for x,y∈Rn,
⟨x,y⟩a:=Ax⋅y and |x|a:=√⟨x,x⟩a |
define an equivalent scalar product and norm in Rn (note that the converse is also true for any symmetric positive definite matrix A, after a suitable rotation of the axes). Let Ea⊂Rn denote the open unitary ball with respect to the a-norm, i.e.,
Ea:={x∈Rn:|x|a<1}. |
In Section 4 we use a to denote a positive real number, in this case we use the convention given in (1.5).
For β>−1, let the function uβ:Rn→R be given by
uβ(x):=(1−|x|2a)β+,x∈Rn. |
We also let
μ(dθ)=dθ|θ|n+2s|Aθ|, | (2.3) |
where dθ denotes the surface measure of ∂Ea, and
J0:=∫∂Eaμ(dθ), | (2.4) |
J(k)i:=aik∫∂Eaθ2ikμ(dθ),k∈{1,…,n}, i∈N. | (2.5) |
These integrals appear frequently in our explicit evaluations. In the particular case a1=…=an−1=1, the integrals J0 and J(k)i can be computed explicitly as well as their asymptotic profile as an↑∞, see Lemma A.1.
We use the gamma, beta, and hypergeometric functions in our analysis, see [7,Chapter 6 and Chapter 15] for general properties of these functions. We collect here the definitions and some integral representations.
1). (Gamma function) For z>0 we denote by
Γ(z)=∫∞0tz−1e−tdt |
the gamma function. If z∈(−∞,0)∖Z, we let Γ(z) be given by the iterative definition Γ(z+1)=zΓ(z).
2). (Beta function) For a,b>0 we denote by
B(a,b)=Γ(a)Γ(b)Γ(a+b) |
the beta function. Note that in this case
B(a,b)=∫10(1−t)a−1tb−1dt=∫∞0ta−1(1+t)a+bdt. |
3). (Hypergeometric function) For a,b,c,z∈R with |z|<1, 2F1(a,b;c;z) denotes the hypergeometric function
2F1(a,b;c;z):=∞∑k=0(a)k(b)k(c)kzkk!, | (2.6) |
where (q)k is the Pochhammer symbol given by (q)0=1 and (q)k=∏k−1i=0(q+i). Note that if q∉Z∩(−∞,0], then (q)k=Γ(q+k)Γ(q) for k∈N0 and hence in particular, if a,b,c∉Z∩(−∞,0], then
2F1(a,b;c;z)=Γ(c)Γ(a)Γ(b)∞∑k=0Γ(a+k)Γ(b+k)Γ(c+k)zkk!. |
If instead q∈Z∩(−∞,0], then
(q)k=0 fork+q≥1. | (2.7) |
Moreover, if c>b>0, then by using the meromorphic extension of the hypergeometric function we have for z<1
2F1(a,b;c;z)=1B(b,c−b)∫10tb−1(1−t)c−b−1(1−zt)−adt. | (2.8) |
Lemma 3.1. Let s>0 and β>0. Then, for i,j,k∈{1,…,n} and x∈Ea,
(−Δ)s(xiuβ(x))=−12(β+1)ai∂i(−Δ)suβ+1(x), | (3.1) |
(−Δ)s(xixjuβ(x))=12(β+1)ai(δi,j(−Δ)suβ+1(x)+12(β+2)aj∂ij(−Δ)suβ+2(x)), | (3.2) |
where δi,j is the Kronecker delta. In particular,
(−Δ)s(x1uβ(x))=−12(β+1)a1∂1(−Δ)suβ+1(x), | (3.3) |
(−Δ)s(x21uβ(x))=12(β+1)a1((−Δ)suβ+1(x)+12(β+2)a1∂21(−Δ)suβ+2(x)). | (3.4) |
Proof. Let us first notice that, for any β>0 and x∈Ea,
∂iuβ+1(x)=−2(β+1)(1−|x|2a)β+(Ax)i=−2(β+1)aixiuβ(x), | (3.5) |
∂i(xjuβ+1(x))=δi,juβ+1(x)−2(β+1)aixixjuβ(x). | (3.6) |
Identity (3.5) directly gives (3.1). Iterating the same idea, from (3.6) one deduces
(−Δ)s(xixjuβ(x))=12(β+1)ai(δi,j(−Δ)suβ+1(x)−∂i(−Δ)s(xjuβ+1(x)))== 12(β+1)ai(δi,j(−Δ)suβ+1(x)+12(β+2)aj∂ij(−Δ)suβ+2(x)). |
Note that the interchange between derivative ∂i and fractional Laplacian (−Δ)s is allowed in this case by the Lebesgue dominated convergence theorem, see for example [2,Proposition B.2].
Theorem 3.2. Let s>0 and β>−1. Then
(−Δ)suβ(x)= 22s−1Γ(12+s)Γ(1+β)cn,m,sΓ(1+β−s)Γ(12)c1,m,s ×× ∫∂Ea(u1(x)+⟨x,θ⟩2a)β−s2F1(s+12,−β+s;12;⟨x,θ⟩2au1(x)+⟨x,θ⟩2a)μ(dθ)for x∈Ea, | (3.7) |
where cn,m,s is given in (2.2). Here, Γ(t)−1=0 if t∈Z∖N.
Proof. We consider spherical coordinates with respect to the a-norm by writing any z∈Rn as z=tθ with t>0 and θ∈∂Ea. This transformation has the Jacobian tn−1/|Aθ|, since, by the coarea formula (notice that ∇|x|a=Ax/|x|a),
∫Rnf(x)dx=∫Rnf(x)|∇|x|a||∇|x|a|dx=∫∞0∫t∂Eaf(x)|x|a|Ax|dxdt=∫∞0∫∂Eaf(tθ)|Aθ|dθtn−1dt. |
We recall notation (2.3) and write
(−Δ)suβ(x)=cn,m,s2∫Rnδmuβ(x,y)|y|n+2sdy=cn,m,s2∫∂Ea∫∞0δmuβ(x,tθ)t1+2sdtμ(dθ) ==cn,m,s4∫∂Ea∫Rδmuβ(x,tθ)|t|1+2sdtμ(dθ). |
We now focus on the inner integral: recall that
δmuβ(x,tθ)=m∑k=−m(−1)k(2mm−k)uβ(x+ktθ)=m∑k=−m(−1)k(2mm−k)(1−|x+ktθ|2a)β+. |
Apply the change of variables
t=−⟨x,θ⟩a+τ√1−|x|2a+⟨x,θ⟩2a, |
rearrange
1−|x+ktθ|2a=1−|x−k⟨x,θ⟩aθ+kτθ√1−|x|2a+⟨x,θ⟩2a|2a== 1−|x|2a−k2⟨x,θ⟩2a−k2τ2(1−|x|2a+⟨x,θ⟩2a)+2k⟨x,θ⟩2a−2k(1−k)⟨x,θ⟩2aτ√1−|x|2a+⟨x,θ⟩2a= (1−|x|2a1−|x|2a+⟨x,θ⟩2a+(2k−k2)⟨x,θ⟩2a1−|x|2a+⟨x,θ⟩2a−k2τ2−2k(1−k)τ⟨x,θ⟩a√1−|x|2a+⟨x,θ⟩2a) ××(1−|x|2a+⟨x,θ⟩2a)= (1−(1−k)2⟨x,θ⟩2a1−|x|2a+⟨x,θ⟩2a−k2τ2−2k(1−k)τ⟨x,θ⟩a√1−|x|2a+⟨x,θ⟩2a)(1−|x|2a+⟨x,θ⟩2a)= (1−((1−k)⟨x,θ⟩a√1−|x|2a+⟨x,θ⟩2a+kτ)2)(1−|x|2a+⟨x,θ⟩2a), |
and deduce
∫Rδmuβ(x,tθ)|t|1+2sdt=(1−|x|2a+⟨x,θ⟩2a)β−s ×× ∫Rm∑k=−m(−1)k(2mm−k)(1−((1−k)⟨x,θ⟩a√1−|x|2a+⟨x,θ⟩2a+kτ)2)β+|τ−⟨x,θ⟩a√1−|x|2a+⟨x,θ⟩2a|1+2sdτ |
which amounts to (after a translation in the τ variable)
∫Rδmuβ(x,tθ)|t|1+2sdt= (1−|x|2a+⟨x,θ⟩2a)β−s∫Rm∑k=−m(−1)k(2mm−k)(1−(˜xθ+kτ)2)β+|τ|1+2sdτ, | (3.8) |
where ˜xθ:=⟨x,θ⟩a(1−|x|2a+⟨x,θ⟩2a)−1/2. Now, using a particular case of*[15,Corollary 4], we know that
* In the notations of [15,Corollary 4], we fix V(x)≡1, l=0, δ=n=1, σ=β and ρ=s.
(−Δ)s(1−z2)β+=c1,m,s2∫Rm∑k=−m(−1)k(2mm−k)(1−(z+kτ)2)β+|τ|1+2sdτ=22sΓ(12+s)Γ(1+β)Γ(1+β−s)Γ(12)2F1(s+12,−β+s;12;z2)for z∈(−1,1). | (3.9) |
Therefore, by (3.8) and (3.9),
(−Δ)suβ(x)=cn,m,s4∫∂Ea(1−|x|2a+⟨x,θ⟩2a)β−s∫Rm∑k=−m(−1)k(2mm−k)(1−(˜xθ+kτ)2)β+|τ|1+2sdτμ(dθ)=cn,m,s2c1,m,s∫∂Ea(1−|x|2a+⟨x,θ⟩2a)β−s(−Δ)s(1−˜x2θ)β+μ(dθ),=22s−1Γ(12+s)Γ(1+β)cn,m,sΓ(1+β−s)Γ(12)c1,m,s∫∂Ea(1−|x|2a+⟨x,θ⟩2a)β−s2F1(s+12,−β+s;12;˜x2θ)μ(dθ). |
In the next corollaries we collect some consequences of Theorem 3.2. For this let
kn,s:=22s−1Γ(n/2+s)πn/2. | (3.10) |
Corollary 3.3. Let s>0. Then it holds
(−Δ)sus(x)=Γ(1+s)kn,sJ0for x∈Ea. | (3.11) |
Moreover, for any ℓ∈N such that s−ℓ>−1, it also holds
(−Δ)sus−ℓ(x)=0for x∈Ea, | (3.12) |
with J0 as in (2.4).
Proof. Both statements follow by just considering respectively β=s and β=s−ℓ in (3.7). Note that for (3.11) we are using that 2F1(s+12,0;12;t)=1 for t∈(−1,1) and, moreover, since
cn,m,s=4sΓ(n/2+s)πn/2Γ(−s)(m∑k=−m(−1)k(2mm−k))−1,s∈(0,m)∖N, |
we have
cn,m,sc1,m,s=Γ(n/2+s)√ππn/2Γ(12+s). |
Note that the same holds for s∈N and hence
22s−1Γ(12+s)cn,m,sΓ(12)c1,m,s=22s−1Γ(n/2+s)πn/2=kn,s. |
Proof of Theorem 1.1. Using the rotation invariance of the fractional Laplacian, we may assume that A is a diagonal matrix. By (3.11), we have that
τ(x):=1Γ(1+s)kn,sJ0(1−|x|2a)s+,x∈Rn, | (3.13) |
satisfies pointwisely that
(−Δ)sτ(x)=1for x∈Ea. | (3.14) |
Moreover, τ∈Hs0(Ea). For s∈N this is clear, so let s∉N and m∈N such that s∈(m,m+1). We argue with the Gagliardo-Nirenberg interpolation inequality (see, e.g., [8,Theorem 1]),
‖f‖Ws,p(Ea)≤C‖f‖θWs1,p1(Ea)‖f‖θWs2,p2(Ea)for all\; f∈Ws1,p1(Ea)∩Ws2,p2(Ea), | (3.15) |
which for some C independent of f is satisfied for 1<p,p1,p2≤∞, 0<s1<s<s2 satisfying for some θ∈(0,1) the relation
s=θs1+(1−θ)s2and1p=θp1+1−θp2. |
Next note that for any β,β′∈Nn0 with |β|=m and |β′|=m+1 there is a constant ˜C>0 such that
|∂βτ(x)|≤˜C(1−|x|a)s−mand|∂β′τ(x)|≤˜C(1−|x|a)s−m−1forx∈Ea |
so that τ∈Wm,∞(Ea) and also τ∈Wm+1,p2(Ea) for 1<p2<11+m−s. By (3.15) with θ=1+m−s, s1=m, s2=m+1, and p1=∞, we then have τ∈Ws,p(Ea) for all p=p21−θ<1(1+m−s)(s−m). Since (1+m−s)(s−m)≤14, we have in particular τ∈Hs(Ea)=Ws,2(Ea). Since also τ/dist(⋅,∂Ea)s∈L∞(Ea), it follows that τ∈Hs0(Ea) also for s∉N (see, for example [34,Section 4.3.2,equation 7]). But then, by uniqueness of weak solutions, τ is the unique weak solution of (3.14) in Hs0(Ea).
Remark 3.4 (Torsion function in an ellipse). In two dimensions, the constant of the torsion function τ can be computed explicitly with a direct computation. Let α1,α2>0 and E={x∈R2:α1x21+α2x22<1}. For a=α2α1 let τ be given by (3.13). Finally, let ˜τ:E→R be given by
˜τ(x):=α−s1τ(α1/21x)=1αs1Γ(1+s)kn,sJ0(1−α1x21−aα1x22)s+=122sΓ(1+s)2αs+1/21α−1/222F1(s+1,12;1;1−α1α2)(1−α1x21−α2x22)s+, |
since, by Lemma A.1,
J0=a−1/2B(12,12)2F1(s+1,12;1;1−1a)=π(α2α1)−1/22F1(s+1,12;1;1−α1α2). |
Then, for x∈E, (−Δ)s˜τ(x)=(−Δ)sτ(ax)=1.
The case β=s+j with j∈N in Theorem 3.2 is particularly useful, and therefore we state it as a corollary.
Corollary 3.5. Let j∈N. Then, for x∈Ea,
(−Δ)sus+j(x)= Γ(1+s+j)Γ(1+j)kn,s∫∂Ea(u1(x)+⟨x,θ⟩2a)j2F1(s+12,−j;12;⟨x,θ⟩2au1(x)+⟨x,θ⟩2a)μ(dθ). | (3.16) |
In the particular cases j=1,2, Table 1 hold.
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Proof. Identity (3.16) simply follows by considering β=s+j in (3.7). In order to deduce the particular cases listed in Table 1, we need to remark that, as one of the arguments in the hypergeometric function is a negative integer, then the hypergeometric function reduces to a polynomial, see (2.7). Such polynomials for j=1,2 can be found in Table 2. The calculation of kn,s follows as in the proof of Corollary 3.3.
j | (v+w)j2F1(s+12,−j;12;wv+w) for t∈(−1,1) |
1 | v−2sw |
2 | v2−4svw+4s(s−1)3w2 |
Proof of Theorem 1.2. Using the rotation invariance of the fractional Laplacian, we may assume that A is a diagonal matrix. As mentioned above, this Theorem follows immediately from Corollary 3.3 for j∈Z∖N0 and from Corollary 3.5, since for j∈N0 and v,w≥0 we have
(v+w)j2F1(s+12,−j;12;wv+w)=(v+w)jj∑k=0(s+12)k(−j)k(12)kk!wk(v+w)k==j∑k=0Γ(s+12+k)Γ(12)Γ(s+12)Γ(12+k)(jk)(−1)kwk(v+w)j−k. |
Recall the definitions of J0, J(k)i, and μ respectively given in (2.4), (2.5), and (2.3).
Lemma 3.6. Let U(x):=(1−a1/2kxk)us(x) for x∈Rn, k∈N. Then, for any x∈Ea,
(−Δ)sU(x)kn,sΓ(1+s)=J0−(J0+2sJ(k)1)a1/2kxk. | (3.17) |
Proof. From Lemma 3.1 and Corollary 3.5 it follows that
(−Δ)sU(x)kn,sΓ(1+s)=J0+a−1/2k2∂k∫∂Ea(u1(x)−2s⟨x,θ⟩2a)μ(dθ) ==J0−a1/2kJ0xk−2sa1/2k∫∂Ea⟨x,θ⟩aθkμ(dθ)=J0−(J0+2sJ(k)1)a1/2kxk, |
since, by symmetry, ∫∂Eaθjθkμ(dθ)=0 for j∈{1,…,n}∖{k}.
Lemma 3.7. Let U(x):=(1−a1/2kxk)2us(x) for x∈Rn. Then for any x∈Ea we have
(−Δ)sU(x)kn,sΓ(1+s)=[J0+5sJ(k)1+2s(s−1)J(k)2]akx2k−2[J0+2sJ(k)1]a1/2kxk+J0−sJ(k)1++sn∑i=1i≠k[J(k)1+2(s−1)aiak∫∂Eaθ2iθ2kμ(dθ)]aix2i. |
Proof. Using (3.3) and (3.4) of Lemma 3.1, we have
(−Δ)sU(x)= (−Δ)sus(x)+a−1/2ks+1∂k(−Δ)sus+1(x)+12(s+1)(−Δ)sus+1(x)++14(s+1)(s+2)ak∂2k(−Δ)sus+2(x), x∈Ea. | (3.18) |
Using the identities in Table 1, we have
(−Δ)sU(x)Γ(1+s)kn,s=J0+1a1/2k∂k∫∂Ea(u1(x)−2s⟨x,θ⟩2a)μ(dθ) ++12∫∂Ea(u1(x)−2s⟨x,θ⟩2a)μ(dθ)+18ak∂2k∫∂Ea(u1(x)2−4su1(x)⟨x,θ⟩2a+4s(s−1)3⟨x,θ⟩4a)μ(dθ). | (3.19) |
In order to compute (3.19), we consider the following differential identities
∂ku1(x)=−2akxk, ∂2ku1(x)=−2ak, | (3.20) |
∂k⟨x,θ⟩2a=2ak⟨x,θ⟩aθk,∂2k⟨x,θ⟩2a=2a2kθ2k,∂2ku1(x)2=∂k(2u1(x)∂ku1(x))=2(∂ku1(x))2+2u1(x)∂2ku1(x)=8a2kx2k−4aku1(x),∂2k(u1(x)⟨x,θ⟩2a)=⟨x,θ⟩2a∂2ku1(x)+2∂ku1(x)∂k⟨x,θ⟩2a+u1(x)∂2k⟨x,θ⟩2a=−2ak⟨x,θ⟩2a−8a2k⟨x,θ⟩aθkxk+2a2ku1(x)θ2k, | (3.21) |
∂2k⟨x,θ⟩4a=∂k(2⟨x,θ⟩2a∂k⟨x,θ⟩2a)=2(∂k⟨x,θ⟩2a)2+2⟨x,θ⟩2a∂2k⟨x,θ⟩2a=12a2k⟨x,θ⟩2aθ2k. | (3.22) |
In view of (3.20) and (3.22), equation (3.19) can be rewritten
(−Δ)sU(x)Γ(1+s)kn,s == J0−2a1/2kJ0xk−4sa1/2kJ(k)1xk+12J0u1(x)−sn∑i=1aiJ(i)1x2i+18ak(8a2kJ0x2k−4J0aku1(x)+8sakn∑i=1aiJ(i)1x2i+32sa2kJ(k)1x2k−8sakJ(k)1u1(x)+16s(s−1)a2kn∑i=1a2ix2i∫∂Eaθ2iθ2kμ(dθ))= J0−2a1/2kJ0xk−4sa1/2kJ(k)1xk+akJ0x2k+5sakJ(k)1x2k−sJ(k)1+sJ(k)1n∑i=1i≠kaix2i+2s(s−1)akx2kJ(k)2+2s(s−1)akn∑i=1i≠ka2ix2i∫∂Eaθ2iθ2kμ(dθ)= [J0+5sJ(k)1+2s(s−1)J(k)2]akx2k−2[J0+2sJ(k)1]a1/2kxk+J0−sJ(k)1+sJ(k)1n∑i=1i≠kaix2i+2s(s−1)akn∑i=1i≠ka2ix2i∫∂Eaθ2iθ2kμ(dθ). |
Lemma 3.8. Let U(x)=us(x)∑ni=1i≠kaix2i for x∈Rn. Then, for any x∈Ea,
(−Δ)sU(x)kn,sΓ(1+s)=s[J0−J(k)1+2(s−1)(J(k)1−J(k)2)]akx2k−s(J0−J(k)1)++n∑i=1i≠k[(s+1)J0+4sJ(i)1−sJ(k)1+2s(s−1)J(i)2+2s(s−1)n∑h=1h≠k,iaiah∫∂Eaθ2hθ2iμ(dθ)]aix2i. |
Proof. Using Lemma 3.1,
(−Δ)sU(x)=n∑i=1i≠kai(−Δ)s(x2ius(x))=12(s+1)((n−1)(−Δ)sus+1(x)+12(s+2)n∑i=1i≠k1ai∂2i(−Δ)sus+2(x)). |
By Table 1 and by suitably adjusting (3.20)–(3.22) to the current situation, we deduce
(−Δ)sU(x)Γ(1+s)kn,s=n−12∫∂Ea(u1(x)−2s⟨x,θ⟩2a)μ(dθ)+n∑i=1i≠k18ai∂2i∫∂Ea(u1(x)2−4su1(x)⟨x,θ⟩2a+4s(s−1)3⟨x,θ⟩4a)μ(dθ)=n−12J0u1(x)−s(n−1)n∑j=1ajJ(j)1x2j+n∑i=1i≠k18ai(8a2iJ0x2i−4J0aiu1(x)+8sain∑h=1ahJ(h)1x2h+32sa2iJ(i)1x2i−8saiJ(i)1u1(x)+16s(s−1)a2in∑h=1a2hx2h∫∂Eaθ2hθ2iμ(dθ))=n∑i=1i≠k(aiJ0x2i+4saiJ(i)1x2i−sJ(i)1u1(x)+2s(s−1)ain∑h=1a2hx2h∫∂Eaθ2hθ2iμ(dθ)). |
Observe that
n∑i=1i≠kJ(i)1=∫∂Ea(1−akθ2k)μ(dθ)=J0−J(k)1n∑i=1i≠kaiak∫∂Eaθ2kθ2iμ(dθ)=∫∂Eaakθ2k(1−akθ2k)μ(dθ)=J(k)1−J(k)2, |
then
(−Δ)sU(x)Γ(1+s)kn,s ==n∑i=1i≠k[J0+4sJ(i)1+2s(s−1)J(i)2]aix2i−s(J0−J(k)1)u1(x)+2s(s−1)n∑i=1i≠kain∑h=1h≠ia2hx2h∫∂Eaθ2hθ2iμ(dθ)=s[J0−J(k)1+2(s−1)(J(k)1−J(k)2)]akx2k−s(J0−J(k)1)++n∑i=1i≠k[(s+1)J0+4sJ(i)1−sJ(k)1+2s(s−1)J(i)2+2s(s−1)n∑h=1h≠k,iaiah∫∂Eaθ2hθ2iμ(dθ)]aix2i. |
Remark 3.9. Consider a1=…=an−1=1 and an=a. In this particular case one has
n∑i=2[ai∫∂Eaθ2iθ21μ(dθ)]aix2i=n−1∑i=2[∫∂Eaθ2iθ21μ(dθ)]x2i+[a∫∂Eaθ2nθ21μ(dθ)]ax2n= 1n−2[n−1∑i=2∫∂Eaθ2iθ21μ(dθ)]n−1∑i=2x2i+[a∫∂Eaθ2nθ21μ(dθ)]ax2n= 1n−2[J(1)1−J(1)2−a∫∂Eaθ2nθ21μ(dθ)]n−1∑i=2x2i+[a∫∂Eaθ2nθ21μ(dθ)]ax2n |
and therefore
(−Δ)s((1−x1)us(x))kn,sΓ(1+s)=J0−[J0+2sJ(1)1]x1,(−Δ)s((1−x1)2us(x))kn,sΓ(1+s)=[J0+5sJ(1)1+2s(s−1)J(1)2]x21−2[J0+2sJ(1)1]x1+J0−sJ(1)1+sn−1∑i=2[J(1)1+2(s−1)∫∂Eaθ2iθ21μ(dθ)]x2i+s[J(1)1+2(s−1)a∫∂Eaθ2nθ21μ(dθ)]ax2n,(−Δ)s(us(x)(n−1∑i=2x2i+ax2n))kn,sΓ(1+s)=s[J0−J(1)1+2(s−1)(J(1)1−J(1)2)]x21−s(J0−J(1)1)++n−1∑i=2[(s+1)J0+4sJ(i)1−sJ(1)1+2s(s−1)J(i)2+2s(s−1)n−1∑h=2h≠i∫∂Eaθ2hθ2iμ(dθ)+2s(s−1)a∫∂Eaθ2nθ21μ(dθ)]x2i+[(s+1)J0+4sJ(n)1−sJ(1)1+2s(s−1)J(n)1−2s(s−1)a∫∂Eaθ2nθ21μ(dθ)]ax2n. |
Note also that, in this case,
(−Δ)s((1−a1/2xn)us(x))kn,sΓ(1+s)=J0−(J0+2sJ(n)1)a1/2xn(−Δ)s((1−a1/2xn)2us(x))kn,sΓ(1+s)=[J0+5sJ(n)1+2s(s−1)J(n)2]ax2n−2[J0+2sJ(n)1]a1/2xn+J0−sJ(n)1+s[J(n)1+2(s−1)n−1(J(n)1−J(n)2)]n−1∑i=1x2i. |
For the sake of clarity we summarize the above in Table 3 for the particular case n=2.
p(x) | Γ(1+s)−1k−1n,s(−Δ)s(pus)(x) for x∈Ea |
1−x1 | J0−[J0+2sJ1]x1 |
(1−x1)2 | [J0+5sJ1+2s(s−1)J2]x21−2[J0+2sJ1]x1+J0−sJ1 |
+s[J1+2(s−1)(J1−J2)]ax22 | |
ax22 | s[J0−J1+2(s−1)(J1−J2)]x21−s(J0−J1) |
+[(2s+1)(s+1)J0−s(4s+1)J1+2s(s−1)J2]ax22 |
In the following, we give a counterexample to the positivity preserving property (see (1.4)) of (−Δ)s, s>1, in an ellipsoid Ea, where we choose a1=…=an−1=1, and an=a>1 sufficiently large. To this end, we consider
U(x):=p(x)us(x),x∈Rn, | (4.1) |
where p is a polynomial of degree two such that p−ϵ is sign-changing for every ϵ>0. Note that once we have shown that there is a constant k>0 such that
(−Δ)sU≥kinEa, | (4.2) |
it follows, by linearity, that for a suitable ϵ>0 small the function Uϵ:=(p−ϵ)us has a nonnegative fractional Laplacian while the function itself is sign-changing in Ea.
We begin with a heuristic explanation of the strategy. We choose p(x)=p2(x1)+γp1(x1)−δq(x) for constants γ,δ≥0 to be fixed later and where
p2(x1)=(1−x1)2,p1(x1)=1−x1,andq(x)=n−1∑k=2x2k+ax2n,x∈Rn. |
From Lemmas 3.6, 3.7, and 3.8 it follows that
(−Δ)s(p2us)=P2(x1)+R2(x2,…,xn)forsomedegree2polynomialsP2andR2,(−Δ)s(p1us)=P1(x1),forsomedegree1polynomialP1,(−Δ)s(qus)=Q(x1)+R0(x2,…,xn)forsomedegree2polynomialsQandR0. |
To achieve (4.2) we then need, in particular, that δ satisfies
R2−δR0≥0,in Ea. | (4.3) |
The choice of γ is far more delicate, but from a geometric point of view it can be made intuitively optimal: indeed, in the worst case scenario, the polynomial P2(x1) has two real roots P2,−<P2,+<1, while P1(x1) always has one P1,+. In this case, it holds that P2,+ and P1,+ are both of the order
1−P2,+=O(1a)=1−P1,+,as a↑∞. |
But then, if we aim at having P2(x1)+γP1(x1)>0 in (−1,1), it is enough to verify (see Figure 1)
P2,+<P1,+ | (4.4) |
and consequently choose
γ=−P′2(P2,+)P′1. | (4.5) |
(noticing that the derivative of P1 is a negative constant): with this choice of γ, we will have P2(x1)≥−γP1(x1) in (−1,1) by convexity.
By taking δ>0 such that (4.3) is satisfied, and replacing P2(x1) with P2(x1)+δQ(x1), the range of possible choices of s so that (4.4) is satisfied can even be enlarged.
The conditions that need to be verified in this argument and their compatibility (on top of an asymptotic analysis as a↑∞) is basically the technical reason why the strategy stops working at finite s: nevertheless we expect that increasing the degrees of the involved polynomials could give some more flexibility in the computations, resulting in a wider range for s.
Theorem 1.4 follows directly from the next result.
Theorem 4.1. Let
p(x):=(1−x1)2+γ(1−x1)−δ(n−1∑k=2x2k+ax2n). | (4.6) |
Then, for every s∈(1,√3+3/2), there are γ,δ≥0, and a0>1 such that the following holds: for every a≥a0 there is K>0 such that
(−Δ)s(pus)(x)≥KJ0a2for allx∈Ea. |
In particular, for every a≥a0 there is ϵ>0 such that the function Uϵ=(p−ϵ)us∈Hs0(Ea) satisfies
(−Δ)sUϵ(x)>0for allx∈Ea. |
Proof of Theorem 4.1. In the following, we perform an asymptotic analysis letting a↑∞. To this end, let us first recall (2.4) and (2.5). By Lemma A.1, we have
j1:=lima↑∞aJ(1)1J0=12s−1andj2:=lima↑∞a2J(1)2J0={+∞,if s∈(1,32];3(2s−1)(2s−3)=3j12s−3,if s>32. | (4.7) |
Moreover, lima↑∞aJ(1)2J0=lima↑∞J(1)2J(1)1=0 for all s>1. Let
A:=(1−sδ)J0+s(5−δ(2s−3))J(1)1+2s(s−1)(1+δ)J(1)2,B:=J0+2sJ(1)1,andC:=(1+δs)J0−s(1+δ)J(1)1. | (4.8) |
We denote by
P1(x1)=J0−Bx1,P2,δ(x1)=Ax21−2Bx1+C,Qδ(x2,…,xn)=sn−1∑i=2[J(1)1+2(s−1)∫∂Eaθ2iθ21μ(dθ)]x2i+s[J(1)1+2(s−1)a∫∂Eaθ2nθ21μ(dθ)]ax2n−δn−1∑i=2[(s+1)J0+4sJ(i)1−sJ(1)1+2s(s−1)J(i)2+2s(s−1)n−1∑h=2h≠i∫∂Eaθ2hθ2iμ(dθ)+2s(s−1)a∫∂Eaθ2nθ21μ(dθ)]x2i−δ[(s+1)J0+4sJ(n)1−sJ(1)1+2s(s−1)J(n)1−2s(s−1)a∫∂Eaθ2nθ21μ(dθ)]ax2n, | (4.9) |
so that, for x∈Ea, we have
(−Δ)s(pU)(x)Γ(1+s)kn,s=P2,δ(x1)+γP1(x1)+Qδ(x2,…,xn),x∈Ea. |
We first note that the discriminant of P2,δ is given by
B2−AC=J20+4sJ0J(1)1+4s2(J(1)1)2−((1−sδ)J0+s(5−δ(2s−3))J(1)1+2s(s−1)(1+δ)J(1)2)(J0(1+sδ)−s(1+δ)J(1)1)=J20+4sJ0J(1)1+4s2(J(1)1)2−(1−s2δ2)J20−s(1+sδ)(5−δ(2s−3))J0J(1)1−2s(s−1)(1+δ)(1+sδ)J0J(1)2+s(1+δ)(1−sδ)J0J(1)1+s2(5−δ(2s−3))(1+δ)(J(1)1)2+2s2(s−1)(1+δ)2J(1)1J(1)2=s2δ2J20+s(4−(1+sδ)(5−δ(2s−3))+(1+δ)(1−sδ))J0J(1)1+s2(4+(5−δ(2s−3))(1+δ))(J(1)1)2−2s(s−1)(1+δ)(1+sδ)J0J(1)2+2s2(s−1)(1+δ)2J(1)1J(1)2=s2δ2J20−2sδ(2s+1+s(2−s)δ)J0J(1)1+s2(9+(8−2s)δ−(2s−3)δ2)(J(1)1)2−2s(s−1)(1+δ)(1+sδ)J0J(1)2+2s2(s−1)(1+δ)2J(1)1J(1)2. |
If s∈(1,3/2] and δ=0, then
a2(B2−AC)J20=9s2(aJ(1)1J0)2−2s(s−1)a2J(1)2J0+2s2(s−1)aJ(1)1J0aJ(1)2J0↓−∞,as a↑∞, |
so that there is a0>0 such that P2,0 is positive for all a≥a0. On the other hand, if s∈(3/2,2) and δ=0, then, using (4.7),
a2(B2−AC)J20=9s2(aJ(1)1J0)2−2s(s−1)a2J(1)2J0+2s2(s−1)aJ(1)1J0aJ(1)2J0⟶9s2j21−6s(s−1)2s−3j1=3sj1(3s2s−1−2(s−1)2s−3)=3sj1(2s−1)(2s−3)(s−2)(2s+1)<0as a↑∞. |
The claim in the case s∈(1,2) hence follows by choosing δ=γ=0, noting that Q0≥0 since it is the sum of nonnegative terms.
In the following we assume s≥2. Moreover, we assume that δ is such that
A=(1−sδ)J0+s(5−δ(2s−3))J(1)1+2s(s−1)(1+δ)J(1)2>0: | (4.10) |
this is asymptotically satisfied as a↑∞ if sδ<1.
For the positivity of Qδ first note that, by symmetry, J(k)i=J(1)i for k∈{1,…,n−1} and i∈N; furthermore,
J(n)1=J0−(n−1)J(1)1,∫∂Eaθ21θ2kμ(dθ)=13J(1)2,anda∫∂Eaθ2nθ21μ(dθ)=J(1)1−n+13J(1)2, | (4.11) |
where the last two identities follow from Lemma and the first identity is a consequence of the definition of J(n)1 and of Ea. Hence, again by symmetry, the fact that θ22+…+θ2n−1+aθ2n=1−θ21 for θ∈∂Ea, and (4.11),
Qδ(x2,…,xn)=n−1∑i=2x2i[sJ(1)1+2s(s−1)∫∂Eaθ22θ21μ(dθ)−δ[(s+1)J0+3sJ(1)1+2s(s−1)J(1)2+2s(s−1)∫∂Ea(θ22+…+θ2n−1+aθ2n)θ2iμ(dθ)−2s(s−1)J(i)2]]+ax2n[sJ(1)1+2s(s−1)a∫∂Eaθ2nθ21μ(dθ)−δ((s+1)J0+4sJ(n)1−sJ(1)1+2s(s−1)J(n)1−2s(s−1)a∫∂Eaθ2nθ21μ(dθ))]=n−1∑i=2x2i[sJ(1)1+23s(s−1)J(1)2−δ((s+1)J0+3sJ(1)1+2s(s−1)∫∂Ea(1−θ21)θ22μ(dθ))]+ax2n[sJ(1)1+2s(s−1)(J(1)1−n+13J(1)2)−δ((s+1)J0+2s(s+1)J(n)1−sJ(1)1−2s(s−1)(J(1)1−n+13J(1)2))]=n−1∑i=2x2i[sJ(1)1+23s(s−1)J(1)2−δ((s+1)J0+s(2s+1)J(1)1−23s(s−1)J(1)2)]+ax2n[s(2s−1)J(1)1−2s(s−1)n+13J(1)2−δ((2s+1)(s+1)J0−s[2(s+1)(n−1)+2s−1]J(1)1+2s(s−1)n+13J(1)2)]. |
This combined with the asymptotic estimates in Lemma gives Qδ≥0 for a sufficiently large, if
sJ(1)1−δ(s+1)J0>0ands(2s−1)J(1)1−δ(s+1)(2s+1)J0>0. |
Note that the second inequality implies the first and in view of the last inequality, we choose
δ=O(1a)andδ<1alima↑∞s(2s−1)aJ(1)1(s+1)(2s+1)J0=s(s+1)(2s+1)1a; | (4.12) |
remark how this choice for δ also fulfills (4.10) for a large.
Note that, in view of (4.10), the largest root of P2,δ is given by
P2,+:=B+√B2−ACA, | (4.13) |
provided† B2≥AC. We remark that‡
† If this is not the case, then P2,δ is positive and it is sufficient to take γ=0.
‡ We use the asymptotic behaviours stated in (A.3), on top of identities (A.9) and (A.10): mind that all this relies on the restriction s>3/2.
B2−AC=s2δ2J20−2sδ(2s+1)J0J(1)1+9s2(J(1)1)2−2s(s−1)J0J(1)2=J20(s2δ2−2s(2s+1)2s−1δa+9s2(2s−1)21a2−6s(s−1)(2s−1)(2s−3)1a2)+o(J20a2),as a↑∞. |
The root of P1 is given by
P1,+:=J0B. | (4.14) |
As explained above, with γ as in (4.5) we have P2,δ+γP1>0 in [−1,1], if (and only if) we can find δ such that
P2,+<P1,+, | (4.15) |
where the strict inequality is needed due to the asymptotic analysis. This inequality is moreover equivalent to
B2+B√B2−AC<J0A. |
Asymptotically, this is satisfied if and only if
1+4s2s−11a+√s2δ2−2s(2s+1)2s−1δa+9s2(2s−1)21a2−6s(s−1)(2s−1)(2s−3)1a2<1−sδ+5s2s−11a, |
which is equivalent to
s2δ2−2s(2s+1)2s−1δa+9s2(2s−1)21a2−6s(s−1)(2s−1)(2s−3)1a2<(−sδ+s2s−11a)2for δ<12s−11a, |
i.e.,
δ>−1s+1(3(s−1)2s−3−4s2s−1)1afor δ<12s−11a. | (4.16) |
As the condition aδ<1/(2s−1) is already implied by (4.12), we are left to verify what values of s allow for a non-empty range of δ as resulting from (4.12) and (4.16): these are those values that satisfy
−3(s−1)2s−3+4s2s−1<s2s+1, |
which in particular holds for s∈[2,√3+3/2).
Proof of Theorem 1.3. This follows directly from the first part of the proof of Theorem 4.1.
Theorem 4.1 shows that the fractional Laplacian (−Δ)s does not satisfy a positivity preserving property in the ellipse Ea for a large enough. Its proof uses an asymptotic analysis as a↑∞ and constructs an explicit counterexample for any a sufficiently large (a>a0 for some a0>1) and for s∈(1,s0) with s0:=√3+3/2≈3.232. In this section we fix n=2 and address the following questions:
i) How large is a0?
ii) What can be said for s≥s0?
The answer to these questions depends on the explicit calculations developed in Section 3, which involve several hypergeometric functions. These functions can be expressed as a series (2.6) or as an integral (2.8). However, direct calculations using these representations are usually hard to perform; nevertheless, computers are very efficient and precise manipulating and approximating the values of hypergeometric functions, and we use this to answer questions i) and ii).
Let
p(x):=(1−x1)2,x∈R2 | (4.17) |
then the value of (−Δ)s(pus) in Ea can be computed explicitly in terms of hypergeometric functions, see Table 3. In particular,
(−Δ)s(pus)>0 inEa if B2−AC<0, | (4.18) |
where A, B, and C are given in (4.8). In Figure 2 we present a plot of the nodal regions of D(a,s):=B2−AC (note that A, B, and C are all explicit functions of a and s).
In particular, Figure 2 shows that (4.18) holds for all s∈(1,2) and a>a0 for some a0>0, as stated in Theorem 4.1, however a0↑∞ as s↓1, whereas for s=3/2 we have a0<115. Note that, if s↑2, then we also have that a0↑∞ whenever p has the simple form (4.17); but, by using a more general polynomial p as in (4.6) for suitable δ and γ, one can obtain a counterexample for s larger.
If s≥s0, then the asymptotic analysis in the proof of Theorem 4.1 cannot be successfully implemented. However, one can show that a counterexample can be obtained for some s≥s0 if a is not very large.
To be more precise, let γ be as in (4.5) and let
δ=s(J(1)1+2(s−1)(J(1)1−J(1)2))(s+1)J0−sJ(1)1−2s(s−1)(J(1)1−J(1)2)+2s(s+1)J(2)1=s(a−1)(2F1(12,s+1;1;a−1a)−2F1(12,s+1;2;a−1a))2F1(12,s+1;1;a−1a)+((a−1)s+a−2)2F1(32,s+1;1;a−1a). |
This choice of δ is such that Qδ≡0 (see (4.9) and use (A.2) and (A.10)).
Let P1,+ and P2,+ be as in (4.13) and (4.14). Then a counterexample can be successfully constructed if P1,+>P2,+, see (4.15). Let
h(a,s):=P1,+−P2,+. |
Then we can compute numerically that h(11,s)>0 for s∈[3,3.8456), see Figure 3. Observe also that h(20,3.8)<0; in particular, this implies that large values of a are not always optimal to construct a counterexample.
To argue the optimality and the consistency of our approach, we remark that the root of the mapping a↦h(a,2) can be computed numerically, and it is given by b0≈18.94281916344395 (see Figure 3), which is the same threshold found in [27,Theorem 5.2], obtained with different arguments than ours in the study of the bilaplacian in two-dimensional ellipses.
For a,c>0 and ν∈Rn, let σ and Ω=Ω(a,c,ν) be defined as in Corollary 1.6, namely, for x∈Rn∖{−ν},
σ(x):=cx+ν|x+ν|2−ν,Ω=Ω(a,c,ν):={x∈Rn:n−1∑i=1σi(x)2+aσn(x)2<1}. |
The geometrical meaning of the point inversion transformation σ is that of an inversion with respect to the boundary of a sphere of radius √c centered in −ν, see Figure 4. Note that if c=1 and ν=0, then σ is the usual Kelvin transform.
Varying ν and c gives rise to a wide variety of shapes, as illustrated in Figures 5 and 6 below. See also [20], where a point inversion transformation is used to show the existence of domains for which the bilaplacian's torsion function is sign-changing. We thank G. Sweers for sharing references [20,21,11] with us.
Proof of Corollary 1.6. We argue as in [1,Proposition 1.6]. Fix c>0, a∈Rn with ai>0, ν∈Rn∖∂Ea, σ as in (1.7), Ω:=Ω(a,c,ν)=σ(Ea), and let Ksz(x):=|x+v|2s−nz(σ(x)) for x∈Rn∖{−v} and z∈C(Rn). Note that −ν∉¯Ω. Then, if us(x):=(1−∑ni=1aix2i)s+ we have that ws=Ksus. By [1,Lemma 3.3], one can compute (−Δ)sws pointwisely in Ω. Then, for every ϕ∈C∞c(Ω),
∫Ωws(−Δ)sϕ = c2n−4s∫ΩKs(Ksws)(x)(−Δ)sKs(Ksϕ)(x) dx=c2n−2s∫ΩKsws(σ(x))|x+v|n−2s(−Δ)sKsϕ(σ(x))|x+ν|n+2s dx = cn−2s∫EaKsws(y)(−Δ)sKsϕ(y) dy=∫Eaus(−Δ)sKsϕ, |
by a change of variables (y=σ(x)) and by Proposition 1.5, where we used that Ks(Ksz)=c2s−nz and that the Jacobian for x↦σ(x) is cn|x+v|−2n. Integrating by parts (see, for example, [5,Lemma 1.5]),
∫Ω(−Δ)sws(x)ϕ(x) dx =∫Ea(−Δ)sus(y)ϕ(σ(y))|x+ν|n−2s dy=κ∫Ωcn|x+ν|−2n|σ(x)+ν|n−2sϕ(x) dy=∫Ωkϕ(x)|x+ν|n+2s dy |
for some constant k>0. Since this holds for any ϕ∈C∞c(Ω), we have that (−Δ)sws(x)=k|x+ν|−n−2s pointwisely in Ω, as claimed.
Proof of Corollary 1.7. We use the notation from the proof of Corollary 1.6. Assume that Ω is bounded or that n>4s and let Uε be given by Theorem 1.4. Then, a direct calculation shows that W:=KsUε∈L2(Rn). Moreover, W is sign changing and, by Proposition 1.5 and Plancherel's Theorem,
∫Rn|ξ|2s|ˆW|2=∫ΩW(−Δ)sW=c2s∫ΩUε(σ(x))|x+ν|n−2s(−Δ)sUε(σ(x))|x+ν|2s+n=c2s∫ΩUε(σ(x))|x+ν|n−2sP(σ(x))|x+ν|2s+n=cn+2s∫EaUε(x)|σ(x)+ν|n−2sP(x)|σ(x)+ν|2s+n|x+ν|−2n=cn+2s∫EaUε(x)P(x)<∞, |
where ˆW is the Fourier transform of W and P is a polynomial of degree two given by Lemmas 3.6 and 3.7. In particular W∈Hs0(Rn). Arguing as in Corollary 1.6, we obtain that (−Δ)sW>0 pointwisely in Ω.
NA is supported by the Alexander von Humboldt Foundation. We thank the anonymous referee for the careful reading of the manuscript and for several helpful comments and suggestions.
The authors declare no conflict of interest.
Recall that μ is defined in (2.3) with a diagonal matrix A with entries a1=…=an−1=1 and an=a.
Lemma A.1. Let n≥2, k∈{1,…,n}, and
J(k)i=aik∫∂Eaθ2ikμ(dθ),i∈N0 | (A.1) |
as in (2.5), where J0:=J(1)0=…=J(n)0. Then
J(n)i=a−1/2ωn−2B(i+12,n−12)2F1(s+n2,i+12;i+n2;1−1a)andJ(k)i=J(1)i=a−1/2ωn−2B(i+12,n−12)2F1(s+n2,12;i+n2;1−1a)fork=1,…,n−1, | (A.2) |
where ωd=2π(d+1)/2Γ((d+1)/2)=|Sd| for d∈N0. Moreover, lima↑∞J(n)iJ0=1 and,
1). if s>i−12, then
lima↑∞aiJ(1)iJ0=Γ(i+12)Γ(12+s−i)Γ(12)Γ(12+s)=i−1∏k=01+2k2s−2k−1; | (A.3) |
2). If s≤i−12, then
lima↑∞a12J(1)i=ωn−2B(i+12,n−12)B(i−s−12,12)B(12,n−12+i) |
and in particular lima↑∞aiJ(1)iJ0=∞ and lima↑∞ai−jJ(1)iJ0=0 for j∈{1,…,i} with s>i−j−12.
Proof. Let θ=(sin(ϕn−1)Pn−2(ϕ′),a−1/2cos(ϕn−1)), with ϕn−1∈(−π,π) and Pn−2(ϕ′) is the parametrization of ∂Bn−11(0)∩{xn>0}, that is P0≡1 and for n>2,
Pn−2=(Pn−3(ϕ1,…,ϕn−3)sin(ϕn−2),cos(ϕn−2)),ϕk∈(0,π)for k=1,…,n−2 |
Then
detJTθJθ=cos2(ϕn−1)+a−1sin2(ϕn−1) |
for n=2 and for n>2 we have
detJTθJθ=det(cos(ϕn−1)PTn−2(ϕ′)−a−1/2sin(ϕn−1)sin(ϕn−1)JTPn−2(ϕ′)0)(cos(ϕn−1)Pn−2(ϕ′)sin(ϕn−1)JPn−2(ϕ′)−a−1/2sin(ϕn−1)0)=det(cos2(ϕn−1)+a−1sin2(ϕn−1)00sin2(ϕn−1)JTPn−2(ϕ′)JPn−2(ϕ′))=(cos2(ϕn−1)+a−1sin2(ϕn−1))(sin2(ϕn−1))n−2detJTPn−2(ϕ′)JPn−2(ϕ′)=(cos2(ϕn−1)+a−1sin2(ϕn−1))(sin2(ϕn−1))n−2n−2∏k=1sin2(k−1)ϕk. |
We begin with k=n, where the above parametrization gives
J(n)i=ωn−22∫π−πcos2i(ϕn−1)(cos2(ϕn−1)+1asin2(ϕn−1))1/2(sin2(ϕn−1))n/2−1(sin2(ϕn−1)+1acos2(ϕn−1)))s+n/2(sin2(ϕn−1)+acos2(ϕn−1))1/2dϕn−1=ωn−22a1/2∫π−πcos2i(ϕn−1)(1−cos2(ϕn−1))n/2−1(1−(1−1a)cos2(ϕn−1)))s+n/2dϕn−1=2ωn−2a1/2∫π/20cos2i(ϕn−1)(1−cos2(ϕn−1))n/2−1(1−(1−1a)cos2(ϕn−1)))s+n/2dϕn−1 |
by symmetry. With the change of variable ϕn−1=arccos(t), dϕn−1dt=−1√1−t2 (and afterwards t2=τ) it follows that
2∫π/20cos2i(ϕn−1)(1−cos2(ϕn−1))n/2−1(1−(1−1a)cos2(ϕn−1))s+n/2dϕn−1=2∫10t2i(1−t2)(n−3)/2(1−(1−1a)t2)s+n/2dt=∫10τi−1/2(1−τ)(n−3)/2(1−(1−1a)τ)−s−n/2dτ=B(i+12,n−12)2F1(s+n2,i+12;i+n2;1−1a), |
where we have used the integral representation of the hypergeometric function 2F1. This proves (A.2) for k=n.
In the following, given two functions f and g, we use notation f∼g as a↑∞, if lima↑∞f(a)g(a)=1. With the change of variable t=(a−1)τ we have as a↑∞
B(i+12,n−12)2F1(s+n2,i+12;i+n2;1−1a)∼∫11/2τi−1/2(1−τ)(n−3)/2(1−(1−1a)τ)−s−n/2dτ=as+n/2(a−1)i+1/2∫a−1(a−1)/2ti−1/2(1−ta−1)(n−3)/2(a−t)−s−n/2dt=as+n/2(a−1)i+n/2−1∫a−1(a−1)/2ti−1/2(a−1−t)(n−3)/2(a−t)−s−n/2dt=as+n/2(a−1)i+n/2−1∫(a−1)/20(a−1−t)i−1/2t(n−3)/2(t+1)−s−n/2dt=as+n/2(a−1)(n−1)/2∫(a−1)/20(1−ta−1)i−1/2t(n−3)/2(t+1)−s−n/2dt∼as+1/2∫∞0t(n−3)/2(t+1)−s−n/2dt=as+1/2B(n−12,s+12). | (A.4) |
Note now that the asymptotic behaviour of Ji follows from (A.4) and it reads
J(n)i∼ωn−2B(n−12,s+12)asas a↑∞. | (A.5) |
so that lima↑∞J(n)iJ0=1 as claimed.
For k=1,…,n−1, by symmetry, it follows that J(k)i=J(1)i. Moreover, with the above parametrization we have
θ1=n−1∏k=1sin(ϕk),withϕk∈(0,π), |
so that with a similar calculation as for k=n we have
J(1)i=(n−2∏k=1∫π0sink−1+2i(ϕk)dϕk)××∫π−π(cos2(ϕn−1)+1asin2(ϕn−1))1/2(sin2(ϕn−1))n/2−1+i(sin2(ϕn−1)+1acos2(ϕn−1)))s+n/2(sin2(ϕn−1)+acos2(ϕn−1))1/2dϕn−1=4a1/2(n−2∏k=1Γ(12)Γ(i+k2)Γ(i+k+12))∫π/20(1−cos2(ϕn−1))n/2−1+i(1−(1−1a)cos2(ϕn−1)))s+n/2dϕn−1=2π(n−2)/2a1/2(n−2∏k=1Γ(i+k2)Γ(i+k+12))∫10τ−1/2(1−τ)(n−3)/2+i(1−(1−1a)τ)−s−n/2dτ=2π(n−2)/2a1/2B(12,n−12+i)Γ(i+12)Γ(i+n−12)2F1(s+n2,12;i+n2;1−1a), | (A.6) |
from which (A.2) follows for k=1,…,n−1. Note that if s<i−12, then, using again the integral representation of the hypergeometric function and the dominated convergence theorem, we have
lima↑∞2F1(s+n2,12;i+n2;1−1a) == B(12,n−12+i)−1lima↑∞∫10τ−1/2(1−τ)(n−3)/2+i(1−(1−1a)τ)−s−n/2 dτ= B(12,n−12+i)−1∫10τ−1/2(1−τ)i−s−3/2 dτ= B(12,n−12+i)−1B(i−s−12,12) |
by the integral representation of the beta function. Hence in this case
lima↑∞a12J(1)i=ωn−2B(i+12,n−12)B(i−s−12,12)B(12,n−12+i), | (A.7) |
which shows the first part in 2. If s>i−12 then with the change of variable t=(a−1)τ we have from (A.6) as a↑∞
aiJ(1)i=2π(n−2)/2a−i+1/2Γ(i+12)Γ(i+n−12)B(12,n−12+i)2F1(s+n2,12;i+n2;1−1a)∼2π(n−2)/2a−i+1/2Γ(i+12)Γ(i+n−12)∫11/2τ−1/2(1−τ)(n−3)/2+i(1−(1−1a)τ)−s−n/2dτ=2π(n−2)/2as+(n−1)/2+i(a−1)1/2Γ(i+12)Γ(i+n−12)∫(a−1)(a−1)/2t−1/2(1−ta−1)(n−3)/2+i(a−t)−s−n/2dt=2π(n−2)/2as+(n−1)/2+i(a−1)n/2+i−1Γ(i+12)Γ(i+n−12)∫(a−1)(a−1)/2t−1/2(a−1−t)(n−3)/2+i(a−t)−s−n/2dt=2π(n−2)/2as+(n−1)/2+i(a−1)n/2+i−1Γ(i+12)Γ(i+n−12)∫(a−1)/20(a−1−t)−1/2t(n−3)/2+i(t+1)−s−n/2dt=2π(n−2)/2as+(n−1)/2+i(a−1)(n−1)/2+iΓ(i+12)Γ(i+n−12)∫(a−1)/20(1−ta−1)−1/2t(n−3)/2+i(t+1)−s−n/2dt∼2π(n−2)/2asΓ(i+12)Γ(i+n−12)∫∞0t(n−3)/2+i(t+1)−s−n/2dt=2π(n−2)/2asΓ(i+12)Γ(i+n−12)B(i+n−12,s−i+12). | (A.8) |
Finally, we have with (A.4)
\begin{align*} \lim\limits_{a\uparrow\infty}\frac{a^iJ_i^{(1)}}{J_0} & = \frac{2\pi^{(n-2)/2} \Gamma\big(i+\frac12\big) B\big(i+\frac{n-1}2, s-i+\frac12\big)}{\omega_{n-2}\Gamma\big(i+\frac{n-1}2\big)\, B\big(\frac{n-1}{2}, s+\frac{1}2\big)}\\ & = \frac{ \Gamma\big(i+\frac12\big) \Gamma\big(s-i+\frac12\big)}{\pi^{1/2} \Gamma\big(s+\frac12\big)} = \prod\limits_{k = 0}^{i-1}\frac{1+2k}{2s-2k-1}, \qquad \text{ if } s \gt i-\frac{1}{2}. \end{align*} |
as claimed in 1. If instead s < i-\frac{1}2 , then by (A.7) we have
\begin{align*} \lim\limits_{a\uparrow\infty }\frac{a^{i-j}J_i^{(1)}}{J_0} = \left\{\begin{aligned} & +\infty & & \text{if }i-j \gt s+\frac12, \\ & 0 & &\text{if }i-j \lt s+\frac12. \end{aligned}\right. \end{align*} |
The case of s = i-\frac{1}{2} now follows similarly, noting that in this case a^{1/2}J_i = O(\ln(a)) for a\uparrow\infty .
Lemma A.2. In the notations of Lemma A.1, we have
\begin{align} \int_{\partial E_a}\theta_i^2\, \theta_k^2 \; \mu(d\theta) & = \frac13 J_2^{(1)} & for ~i, k\in\{1, \ldots, n-1\}, \ i\neq k, \end{align} | (A.9) |
\begin{align} a\int_{\partial E_a}\theta_i^2\, \theta_n^2 \; \mu(d\theta) & = J_1^{(1)}-\frac{n+1}{3}J_2^{(1)} & for ~i\in\{1, \ldots, n-1\}. \end{align} | (A.10) |
Proof. The proof of (A.9) closely follows the computation in the proof of (A.2). Indeed, by symmetry,
\begin{align*} \int_{\partial E_a}\theta_i^2\, \theta_k^2 \; \mu(d\theta) = \int_{\partial E_a}\theta_1^2\, \theta_2^2 \; \mu(d\theta) \qquad\text{for }i, k\in\{1, \ldots, n-1\}, \ i\neq k, \end{align*} |
and with
J: = \int_{-\pi}^{\pi}\frac{\big(\cos^2(\phi_{n-1})+\frac{1}{a}\sin^2(\phi_{n-1})\big)^{1/2}\big(\sin^2(\phi_{n-1})\big)^{n/2+1}}{\big(\sin^2(\phi_{n-1})+\frac{1}{a}\cos^2(\phi_{n-1}))\big)^{s+n/2}\big(\sin^2(\phi_{n-1})+a\cos^2(\phi_{n-1})\big)^{1/2}} \;d\phi_{n-1}\Bigg(\prod\limits_{k = 2}^{n-2}\int_{0}^{\pi}\sin^{k+3}(\phi_k)\;d\phi_k\Bigg), |
observing that \theta_2 = \cos(\phi_1)\prod_{k = 2}^{n-1}\sin(\phi_k) , we have
\begin{multline*} \int_{\partial E_a}\theta_1^2\, \theta_2^2 \; \mu(d\theta) = J\int_{0}^{\pi}\sin^{2}(\phi_1)\, \cos^2(\phi_1)\;d\phi_1 = J\Big(\int_{0}^{\pi}\sin^{2}(\phi_1)\;d\phi_1-\int_{0}^{\pi}\sin^{4}(\phi_1)\;d\phi_1\Big) = \\ = J_2^{(1)}\Big(\frac{\frac\pi2}{\frac38\pi}-1\Big) = \frac13 J_2^{(1)}. \end{multline*} |
For the proof of (A.10) we proceed as follows using again the symmetry:
\begin{align*} a\int_{\partial E_a}\theta_i^2\, \theta_n^2 \; \mu(d\theta) = a\int_{\partial E_a}\theta_1^2\, \theta_n^2 \; \mu(d\theta) = \int_{\partial E_a}\theta_1^2\, \Big(1-\sum\limits_{i = 1}^{n-1}\theta_i^2\Big) \; \mu(d\theta) = J_1^{(1)}-J_2^{(1)}-\frac{n-2}{3}J_2^{(1)}, \end{align*} |
where we have used (A.9) in the last identity.
[1] | Abatangelo N, Dipierro S, Fall MM, et al. (2019) Positive powers of the Laplacian in the half-space under Dirichlet boundary conditions. Discrete Contin Dyn Syst 39: 1205-1235. |
[2] | Abatangelo N, Jarohs S, Saldaña A (2018) Green function and Martin kernel for higher-order fractional Laplacians in balls. Nonlinear Anal 175: 173-190. |
[3] | Abatangelo N, Jarohs S, Saldaña A (2018) Integral representation of solutions to higher-order fractional Dirichlet problems on balls. Commun Contemp Math 20: 1850002. |
[4] | Abatangelo N, Jarohs S, Saldaña A (2018) On the loss of maximum principles for higher-order fractional Laplacians. P Am Math Soc 146: 4823-4835. |
[5] | Abatangelo N, Jarohs S, Saldaña A (2018) Positive powers of the Laplacian: from hypersingular integrals to boundary value problems. Commun Pure Appl Anal 17: 899-922. |
[6] | Abatangelo N, Valdinoci E (2019) Getting acquainted with the fractional Laplacian, In: Contemporary Research in Elliptic PDEs and Related Topics, Cham: Springer, 1-105. |
[7] | Abramowitz M, Stegun IA (1964) Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, U.S. Government Printing Office, Washington, DC. |
[8] | Brezis H, Mironescu P (2018) Gagliardo-Nirenberg inequalities and non-inequalities: The full story. Ann I H Poincaré Anal Non Linéaire 35: 1355-1376. |
[9] | Bucur C, Valdinoci E (2016) Nonlocal Diffusion and Applications, Cham: Springer. |
[10] | Coffman CV, Duffin RJ (1980) On the structure of biharmonic functions satisfying the clamped plate conditions on a right angle. Adv Appl Math 1: 373-389. |
[11] | Dall'Acqua A, Sweers G (2005) The clamped-plate equation for the limaçon. Ann Mat Pura Appl 184: 361-374. |
[12] | Dipierro S, Grunau HC (2017) Boggio's formula for fractional polyharmonic Dirichlet problems. Ann Mat Pura Appl 196: 1327-1344. |
[13] | Duffin RJ (1949) On a question of Hadamard concerning super-biharmonic functions. J Math Phys 27: 253-258. |
[14] | Dyda B (2012) Fractional calculus for power functions and eigenvalues of the fractional Laplacian. Fract Calc Appl Anal 15: 536-555. |
[15] | Dyda B, Kuznetsov A, Kwaśnicki M (2017) Fractional Laplace operator and Meijer G-function. Constr Approx 45: 427-448. |
[16] | Garabedian PR (1951) A partial differential equation arising in conformal mapping. Pacific J Math 1: 485-524. |
[17] | Garofalo M (2019) Fractional thoughts, In: New Developments in the Analysis of Nonlocal Operators, Providence, RI: Amer. Math. Soc., 1-135. |
[18] | Gazzola F, Grunau HC, Sweers G (2010) Polyharmonic Boundary Value Problems, Berlin: Springer-Verlag. |
[19] | Grunau HC, Robert F (2013) Uniform estimates for polyharmonic Green functions in domains with small holes, In: Recent Trends in Nonlinear Partial Differential Equations. Ⅱ. Stationary Problems, Providence, RI: Amer. Math. Soc., 263-272. |
[20] | Grunau HC, Sweers G (2014) A clamped plate with a uniform weight may change sign. Discrete Contin Dyn Syst Ser S 7: 761-766. |
[21] | Grunau HC, Sweers G (2014) In any dimension a "clamped plate" with a uniform weight may change sign. Nonlinear Anal 97: 119-124. |
[22] | Hedenmalm H, Jakobsson S, Shimorin S (2002) A biharmonic maximum principle for hyperbolic surfaces. J Reine Angew Math 550: 25-75. |
[23] | Jarohs S, Saldaña A, Weth T (2020) A new look at the fractional poisson problem via the logarithmic Laplacian. J Funct Anal 279: 108732. |
[24] | Keady G, McNabb A (1993) The elastic torsion problem: solutions in convex domains. New Zealand J Math 22: 43-64. |
[25] | Kozlov VA, Kondrat'ev VA, Maz'ya VG (1989) On sign variability and the absence of "strong" zeros of solutions of elliptic equations. Izv Akad Nauk SSSR Ser Mat 53: 328-344. |
[26] | Nakai M, Sario L (1977) On Hadamard's problem for higher dimensions. J Reine Angew Math 291: 145-148. |
[27] | Render H, Ghergu M (2012) Positivity properties for the clamped plate boundary problem on the ellipse and strip. Math Nachr 285: 1052-1062. |
[28] | Ros-Oton X, Serra J (2015) Local integration by parts and Pohozaev identities for higher order fractional Laplacians. Discrete Contin Dyn Syst 35: 2131-2150. |
[29] | Saldaña A (2020) On fractional higher-order Dirichlet boundary value problems: between the Laplacian and the bilaplacian. arXiv: 1810.08435. |
[30] | Samko SG, Kilbas AA, Marichev OI (1993) Fractional Integrals and Derivatives, Yverdon: Gordon and Breach Science Publishers. |
[31] | Shapiro HS, Tegmark M (1994) An elementary proof that the biharmonic Green function of an eccentric ellipse changes sign. SIAM Rev 36: 99-101. |
[32] | Sweers G (2016) An elementary proof that the triharmonic Green function of an eccentric ellipse changes sign. Arch Math 107: 59-62. |
[33] | Sweers G (2019) Correction to: An elementary proof that the triharmonic Green function of an eccentric ellipse changes sign. Arch Math 112: 223-224. |
[34] | Triebel H (1978) Interpolation Theory, Function Spaces, Differential Operators, Amsterdam-New York: North-Holland Publishing Co. |
1. | Alberto Saldaña, 2021, 775, 978-1-4704-6728-9, 255, 10.1090/conm/775/15595 | |
2. | Serena Dipierro, Luca Lombardini, Partial differential equations from theory to applications: Dedicated to Alberto Farina, on the occasion of his 50th birthday, 2023, 5, 2640-3501, 1, 10.3934/mine.2023050 | |
3. | Serena Dipierro, Giorgio Poggesi, Jack Thompson, Enrico Valdinoci, Quantitative stability for overdetermined nonlocal problems with parallel surfaces and investigation of the stability exponents, 2024, 188, 00217824, 273, 10.1016/j.matpur.2024.06.011 |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
\ j\ | \big(v+w \big)^{j} {\, {}_2F_1}\Big(s+\frac12, -j;\frac12;\frac{w}{v+w}\Big) for t\in(-1, 1) |
1 | v-2sw |
2 | v^2-4svw+\frac{4s(s-1)}3w^2 |
\ p(x)\ | \Gamma(1+s)^{-1}k_{n, s}^{-1} (-\Delta)^s(pu_{s})(x)\ for x\in E_a |
1-x_1 | J_0-\big[J_0+2sJ_1\big]x_1 |
(1-x_1)^2 | \big[J_0+5sJ_1+2s(s-1)J_2\big]x_1^2 -2\big[J_0+2sJ_1\big]x_1 +J_0-sJ_1 |
\quad +s\big[J_1+2(s-1)\big(J_1-J_2\big)\big]ax_2^2 | |
ax_2^2 | s\Big[J_0-J_1+2(s-1)\big(J_1-J_2\big)\Big]x_1^2-s\big(J_0-J_1\big) |
\quad +\Big[(2s+1)(s+1)J_0-s(4s+1)J_1 +2s(s-1)J_2\Big]ax_2^2\; |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
\ j\ | \big(v+w \big)^{j} {\, {}_2F_1}\Big(s+\frac12, -j;\frac12;\frac{w}{v+w}\Big) for t\in(-1, 1) |
1 | v-2sw |
2 | v^2-4svw+\frac{4s(s-1)}3w^2 |
\ p(x)\ | \Gamma(1+s)^{-1}k_{n, s}^{-1} (-\Delta)^s(pu_{s})(x)\ for x\in E_a |
1-x_1 | J_0-\big[J_0+2sJ_1\big]x_1 |
(1-x_1)^2 | \big[J_0+5sJ_1+2s(s-1)J_2\big]x_1^2 -2\big[J_0+2sJ_1\big]x_1 +J_0-sJ_1 |
\quad +s\big[J_1+2(s-1)\big(J_1-J_2\big)\big]ax_2^2 | |
ax_2^2 | s\Big[J_0-J_1+2(s-1)\big(J_1-J_2\big)\Big]x_1^2-s\big(J_0-J_1\big) |
\quad +\Big[(2s+1)(s+1)J_0-s(4s+1)J_1 +2s(s-1)J_2\Big]ax_2^2\; |