Processing math: 68%
Research article Special Issues

Properties and applications of generalized 1-parameter 3-variable Hermite-based Appell polynomials

  • We present a novel framework for introducing generalized 3-variable 1-parameter Hermite-based Appell polynomials. These polynomials are characterized by generating function, series definition, and determinant definition, elucidating their fundamental properties. Moreover, utilizing a factorization method, we established recurrence relations, shift operators, and various differential equations, including differential, integrodifferential, and partial differential equations. Special attention is given to exploring the specific cases of 3-variable 1-parameter generalized Hermite-based Bernoulli, Euler, and Genocchi polynomials, offering insights into their unique features and applications.

    Citation: Mohra Zayed, Shahid Ahmad Wani. Properties and applications of generalized 1-parameter 3-variable Hermite-based Appell polynomials[J]. AIMS Mathematics, 2024, 9(9): 25145-25165. doi: 10.3934/math.20241226

    Related Papers:

    [1] Antonio Vitolo . Singular elliptic equations with directional diffusion. Mathematics in Engineering, 2021, 3(3): 1-16. doi: 10.3934/mine.2021027
    [2] Isabeau Birindelli, Giulio Galise . Allen-Cahn equation for the truncated Laplacian: Unusual phenomena. Mathematics in Engineering, 2020, 2(4): 722-733. doi: 10.3934/mine.2020034
    [3] Aleksandr Dzhugan, Fausto Ferrari . Domain variation solutions for degenerate two phase free boundary problems. Mathematics in Engineering, 2021, 3(6): 1-29. doi: 10.3934/mine.2021043
    [4] Daniela De Silva, Ovidiu Savin . On the boundary Harnack principle in Hölder domains. Mathematics in Engineering, 2022, 4(1): 1-12. doi: 10.3934/mine.2022004
    [5] David Cruz-Uribe, Michael Penrod, Scott Rodney . Poincaré inequalities and Neumann problems for the variable exponent setting. Mathematics in Engineering, 2022, 4(5): 1-22. doi: 10.3934/mine.2022036
    [6] Mikyoung Lee, Jihoon Ok . Local Calderón-Zygmund estimates for parabolic equations in weighted Lebesgue spaces. Mathematics in Engineering, 2023, 5(3): 1-20. doi: 10.3934/mine.2023062
    [7] 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
    [8] Alessia E. Kogoj, Ermanno Lanconelli, Enrico Priola . Harnack inequality and Liouville-type theorems for Ornstein-Uhlenbeck and Kolmogorov operators. Mathematics in Engineering, 2020, 2(4): 680-697. doi: 10.3934/mine.2020031
    [9] La-Su Mai, Suriguga . Local well-posedness of 1D degenerate drift diffusion equation. Mathematics in Engineering, 2024, 6(1): 155-172. doi: 10.3934/mine.2024007
    [10] 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
  • We present a novel framework for introducing generalized 3-variable 1-parameter Hermite-based Appell polynomials. These polynomials are characterized by generating function, series definition, and determinant definition, elucidating their fundamental properties. Moreover, utilizing a factorization method, we established recurrence relations, shift operators, and various differential equations, including differential, integrodifferential, and partial differential equations. Special attention is given to exploring the specific cases of 3-variable 1-parameter generalized Hermite-based Bernoulli, Euler, and Genocchi polynomials, offering insights into their unique features and applications.



    To Sandro, with friendship, admiration and much more.

    Let z=(x,y)Rn+1, with xRn and yR, n1, aR. Our aim is to study the boundary behaviour of solutions to a class of problems involving singular/degenerate operators in divergence form including

    Lau:=div(|y|aA(x,y)u),

    and their regularizations. The boundary here coincides with Σ:={y=0} the characteristic manifold, where the weight becomes degenerate or singular, and this happens respectively when a>0 and a<0. Accordingly, this class of operators is called degenerate elliptic.

    The first motivation for this work is to complete the study started in [21] on local regularity for solutions to degenerate/singular problems including the following

    div(|y|au)=|y|af+div(|y|aF)in B1. (1.1)

    In [21], we treated the regularity of even-in-y solutions (corresponding to Neumann boundary conditions), including the case of variable coefficients. We provided local C0,α and C1,α estimates, which are uniform as the parameter ε0+, for even solutions of regularized uniformly elliptic problems of the form

    div(ρaε(y)A(x,y)uε)=ρaε(y)fε+div(ρaε(y)Fε)in B1, (1.2)

    where the regularized family of weights ρaε is defined as:

    ρaε(y):={(ε2+y2)a/2min{εa,1}if a0,(ε2+y2)a/2max{εa,1}if a0. (1.3)

    A further motivation comes from a remarkable link between our operators and fractional powers of the Laplacian, from a Dirichlet-to-Neumann point of view, as highlighted in [4], when our weights belong to the A2-class; i.e., a(1,1).

    Goal of this paper is to deal with odd-in-y solutions to (1.1) (corresponding to Dirichlet boundary conditions), providing local regularity, when possible in the ε-stable sense, by proving uniform bounds for solutions to (1.2). Odd solutions make sense as energy solutions in the natural weighted Sobolev spaces whenever a(,1) (in the sense of ยง2). At first, we notice that can not expect, for the odd solutions, the same estimates as for the even ones, where the regularity results from the combined effect of the ellipticity and the boundary condition. In fact, the function y|y|a is La-harmonic with finite energy when a<1 (in case of A=I), and for a(0,1) is no more than Hölder continuous. We will refer to this special solution as the characteristic odd comparison solution. Similar, yet smoother, characteristic odd comparison solutions exist for the full regularized family of ε-problems (in a rather general setting). Nonetheless, one major obstruction in the study of regularity is the fact that our weights need not to be locally integrable when a1, preventing the application of classical regularity theory such as that developed for degenerate weights of the A2-Muckenhoupt class, starting from the seminal papers [8,9,10]. We point out that our singular/degenerate operators fall within the class of edge operators whose calculus was developed by Mazzeo and his collaborators (see in particular [7,15,16] and references therein). We shall adopt here a different perspective, exploiting suitably tailored Liouville type theorems as main tools (similarly to [20]). To this aim, a major hindrance is that the measure |y|adz is not absolutely continuous with respect to the Lebesgue measure. In order to overcome this difficulty, one can be guided by the following insight:

    Proposition 1.1. Let a(,1) and uH1,a(B1) be an odd energy solution to (1.1) in B1 (for simplicity with F=0). Then for any r<1 the ratio w=u/y|y|aH1,2a(Br) and it is an even energy solution to

    div(|y|2aw)=|y|2aˉf=|y|2afy|y|ain Br. (1.4)

    Proposition 1.1 allows the application of the results for even solutions already proved in [21], providing regularity up to the multiplicative factor y|y|a. Thanks to this observation it is natural to shift the study of regularity for odd solutions to that of even solutions of the auxiliary problem above. A similar perspective has been adopted in [19] for the obstacle problem in the same singular/degenerate setting.

    As an example, by the Schauder estimates in [21], when the forcing ˉf=:f/y|y|a in (1.4) is Ck,α, then the ratio w=u/y|y|a is locally Ck+2,α. Thus, we understand that the correct way to face the regularity of odd solutions consists in seeking C0,α and C1,α bounds for the ratio between the solution and the characteristic odd one, depending on the regularity of the same ratio of the right hand side. This point of view corresponds to (possibly higher order and/or non homogeneous) boundary Harnack principle at Σ in the sense of [3,6,10,13,14]. It is worthwhile noticing that, when a(,1), then the exponent 2a belongs to (1,+), placing Eq (1.4) in the so called super degenerate case, again outside the land of A2-Muckenhoupt weights theory, and which has been treated in [21] when associated with Neumann boundary conditions. Furthermore, looking at the right hand side of (1.4), we realize that the transition from the odd to the even case requires to pay a cost in terms of more stringent conditions on the forcing term f, in the sense that the ratio fy|y|a must possess some regularity (integrability at least); in other words, when a<0, it means that the forcing term is vanishing with a certain rate at Σ. In this regard, our results are connected with the recent paper [1], where a boundary Harnack principle with right hand side is established in the uniformly elliptic case.

    As already pointed out, our results are not limited to the A2-Muckenhoupt class of weights, which restricts a in the interval (1,1). Nonetheless, we wish to state the following corollary, which joins the results contained in this paper with the Schauder theory for even solutions developed in [21], concerning full regularity for energy solutions of degenerate or singular problems when the weight is A2-Muckenhoupt and A=I.

    Corollary 1.2. Let a(1,1), kN{0}, α(0,1) and consider uH1,a(B1) an energy solution to

    div(|y|au)=|y|afin B1.

    Let us consider the even and odd parts* (with respect to y) of the forcing term f. Let

    *Even and odd parts (in y) of a function are defined as usual as

    fe(x,y)=f(x,y)+f(x,y)2,fo(x,y)=f(x,y)f(x,y)2.
    f=fe+fo=fe+y|y|a˜fewith fe,˜feCk,α(B1).

    Then

    u=ue+uo=ue+y|y|a˜ue,with ue,˜ueCk+2,αloc(B1).

    As a next step, we aim at deepening the ε-stability of these estimates with respect to the family of regularized weights (1.3) (also including the variable coefficient case). In other words, we deal with odd-in-y solutions to the family of Eq (1.2). We will provide local uniform-in-ε regularity estimates, enlightening their delicate link with curvature issues related with the matrix A. As we shall see, also the notion of characteristic solution must be suitably adjusted in order to deal with the variable coefficient cases. Finally, we will apply our results to a family of degenerate/singular equations naturally associated with the euclidean Laplacian expressed in Fermi coordinates in the neighbourhood of an embedded hypersurface.

    Below we set the minimal assumptions on the matrix A that we need throughout the paper:

    Assumption 1.3 (HA). The matrix A=(aij) is (n+1,n+1)-dimensional and symmetric A=AT, has the following symmetry with respect to Σ: we have

    A(x,y)=JA(x,y)J,withJ=(In001).

    Moreover, A is continuous and satisfies the uniform ellipticity condition λ1|ξ|2A(x,y)ξξλ2|ξ|2, for all ξRn+1, for every (x,y) and some ellipticity constants 0<λ1λ2. Therefore, the characteristic manifold Σ is assumed to be invariant with respect to A when y=0; that is, there exists a suitable scalar function μ such that

    A(x,0)ey=μ(x,0)ey.

    Whenever the hypothesis on A are not specified, we always imply Assumption (HA). From now on, through out the paper, whenever not otherwise specified, in order to simplify the notations, we will work with A=I every time this condition is not playing a role in the proofs. In the perspective of Proposition 1.1, but considering odd solutions for the family of regularized problems in (1.2), it will be convenient to adopt the following notation on the matrix A.

    Notation 1.4 (HA+). The matrix A is written as:

    A(x,y)=μ(x,y)B(x,y),

    with

    1Cμ(x,y)C, (1.5)
    B(x,y)=(˜B(x,y)T(x,y)T(x,y)1),

    where ˜B is a (n,n)-dimensional matrix and T:Rn+1Rn (we denote by ˜A=μ˜B). We remark here that under our hypothesis on the symmetries of coefficients; one has, for y<0

    A(x,y)=μ(x,y)(˜B(x,y)T(x,y)T(x,y)1).

    The structural assumption on the matrix A is consistent with [5]. Moreover, it fits also with the metric induced by Fermi's coordinates, which allow to study phenomena of singularity or degeneration on a characteristic manifold Σ which is a generic (regular enough) n-dimensional hypersurface embedded in Rn+1. Hence, the objective will be to consider the ratio wε between odd solutions uε to (1.2) and functions of the form

    vaε(x,y)=(1a)y0ρaε(s)μ(x,s)1ds, (1.6)

    which play now the role of the characteristic odd solution for the regularized family of weights in the variable coefficients case. It is worthwhile stressing that the characteristic solutions vaε do not longer solve the homogenous problem, as a dependence on the curvature appears.

    As said, we wish to obtain uniform local regularity estimates for wε which will be even solutions to an auxiliary weighted problems having the following structure

    div(ρaε(vaε)2Awε)=ρaε(vaε)2fε+div(ρaε(vaε)2Fε)+ρaε(vaε)2bεwε. (1.7)

    The new weights appearing in the auxiliary equation are equivalent, though not equal, (using (1.5)) to

    ωaε(y)=ρaε(y)(1a)2(χaε(y))2 (1.8)

    where we have defined

    χaε(y):=y0ρaε(s)ds. (1.9)

    We remark that, as a(,1), such a class of weights is always super degenerate; indeed, at Σ, they behave like

    ωaε(y){y2if ε>0|y|2aif ε=0,

    with 2a(1,+).

    Our first main result concerns in fact the even solutions to the auxiliary family of Eq (1.7). It essentially consists in extending (in a non trivial way) the analogous result already obtained in [21] to the new family of weights ρaε(vaε)2.

    Theorem 1.5. Let a(,1) and, as ε0, let {wε} be a family of solutions in B+1 of (1.7) which are even-in-y; that is, satisfying the boundary condition

    ρaε(vaε)2ywε=0on 0B+1.

    1) Let r(0,1), β>1, p1>n+3+(a)+2, p2,p3>n+3+(a)+, and α(0,2n+3+(a)+p1](0,1n+3+(a)+p2](0,1n+3+(a)+p3]. Let's moreover take A with continuous coefficients and bεLp3(B+1,ωaε(y)dz)b. There is a positive constant c depending on a, b, n, β, p1, p2, p3, α and r only such that functions wε satisfy

    wεC0,α(B+r)c(wεLβ(B+1,ωaε(y)dz)+fεLp1(B+1,ωaε(y)dz)+FεLp2(B+1,ωaε(y)dz)).

    2) Let r(0,1), β>1, p1,p2>n+3+(a)+, and α(0,1n+3+(a)+p1](0,1n+3+(a)+p2]. Let Fε=(F1ε,...,Fn+1ε) with the y-component vanishing on Σ: Fn+1ε(x,0)=Fyε(x,0)=0 in 0B+1. Let's moreover take A with α-Hölder continuous coefficients and bεL2p2(B+1,ωaε(y)dz)b. There is a positive constant c depending on a, b, n, β, p1, p2, α and r only such that functions wε satisfy

    wεC1,α(B+r)c(wεLβ(B+1,ωaε(y)dz)+fεLp1(B+1,ωaε(y)dz)+FεC0,α(B+1)).

    We would like to remark here that local C2,α uniform estimates (up to Σ) with respect to the regularization can not be proven (for a counterexample we refer to [21, Remark 5.4]).

    When applying Theorem 1.5 to the quotient

    wε=uvaε (1.10)

    of a solution of (1.2) and the characteristic solution (1.6), we realise that the actual terms appearing in right hand side of (1.7) depend on the original forcings f,F jointly with the parameters μ,T,B of the matrix A written as in Notation (HA+). In particular, as shown in (2.8), we see the appearance of a drift term involving the x-derivatives of μ which, consequently, need to satisfy a C0,α condition. Our main result is Theorem 4.4. We give here below a simplified statement, suitable to be applied to the case of Laplacians in Fermi coordinates treated in subsection ยง1.1.

    Theorem 1.6. Let a(,1), the matrix A written as in Notation (HA+) with T0. As ε0 let {uε} be a family of solutions in B+1 of

    {div(ρaεAuε)=ρaεfε+div(ρaεFε)in B+1uε=0on 0B+1.

    Let also {vaε} be the family of solutions defined in (1.6) in B+1. Denote

    wε=uεvaε.

    1) Assume μ be Lipschitz continuous, r(0,1), β>1, p1>n+3+(a)+2, p2>n+3+(a)+, and α(0,2n+3+(a)+p1](0,1n+3+(a)+p2]. Let's moreover take A with continuous coefficients. There is a positive constant c depending on a, n, β, p1, p2, α and r only such that the wε satisfy

    wεC0,α(B+r)c(wεLβ(B+1,ωaε(y)dz)+fε/vaεLp1(B+1,ωaε(y)dz)+Fyε/(yvaε)Lp1(B+1,ωaε(y)dz)+Fε/vaεLp2(B+1,ωaε(y)dz)).

    2) Assume μC1,α(B+1), and let r(0,1), β>1, p1>n+3+(a)+, and α(0,1n+3+(a)+p1]. Let Fε=(F1ε,...,Fn+1ε) with the α-Hölder continuous ratio between the y-component and vaε vanishing on Σ: Fn+1ε(x,0)/vaε=Fyε(x,0)/vaε=0 in 0B+1. Let's moreover take A with α-Hölder continuous coefficients. There is a positive constant c depending on a, n, β, p1, α and r only such that

    wεC1,α(B+r)c(wεLβ(B+1,ωaε(y)dz)+fε/vaεLp1(B+1,ωaε(y)dz)+Fyε/(yvaε)Lp1(B+1,ωaε(y)dz)+Fε/vaεC0,α(B+1)).

    It is worthwhile noticing here that any energy odd solution to (1.2) for ε=0 (under suitable conditions on the matrix and the right hand side) can be approximated by a ε-sequence of solutions to (1.2) satisfying the hypothesis in our regularity results. The same happens for the auxiliary weighed problems solved by the even functions w=u/y|y|a. This is done in details in [21, Section 2 and 6].

    Remark 1.7. A special, yet fundamental, case is when we take A=I, so that μ1 and the family of fundamental comparison odd solutions vaε's are in fact the χaε's. Nevertheless, it has to be noticed that, in the presence of non trivial curvature, the ratio vaε/χaε may not be uniformly (in ε) bounded in C1,α(B+1). Furthermore, in the variable coefficient case, the χaε's are not in the kernel of the corresponding operators, as a (possibly weird) right hand side appears.

    Theorem 1.6 finds a natural application to the study of the boundary behaviour of solutions of operators degenerate/singular at embedded manifolds, as shown by the following result.

    Corollary 1.8. Let Σ be an n-dimensional hypersurface embedded in Rn+1, of class C3,α and let dΣ(X) denote the signed distance of X to Σ. Let a(,1), R>0 sufficiently small, and consider, as ε0, a family of solutions to

    {div(ρaεdΣuε)=ρaεdΣfε+div(ρaεdΣFε)in BR{dΣ(X)>0}uε=0on BRΣ.

    Let also {χaε} be the family of functions defined in (1.9) in BR. Denote

    wε=uεχaεdΣ,

    1) The same conclusion of point 1) of Theorem 1.6 holds with vaε replaced by χaε, y by dΣ(X) and en+1 by the normal ν at Σ.

    2) The same conclusion of point 2) of Theorem 1.6 holds in C1,α(Br{yε}) where c is independent of ε, and, again, vaε replaced by χaε, y by dΣ(X) and en+1 by the normal ν at Σ.

    Remark 1.9. In particular, letting ε0 we find C1,α(B+r) estimates in the degenerate/singular case, though not in the full ε-stable sense. The reason is the possible lack of uniform-in-ε smoothness of the ratio vaε/χaε.

    As already mentioned, the structural assumption on the matrix A done in Assumption (HA) with Notation (HA+) fits also with the metric induced by Fermi's coordinates around the characteristic manifold Σ (see [18]). Let Σ be an oriented regular enough hypersurface embedded in Rn+1. We are concerned with operators associated with Dirichlet energies of the form

    {dΣ(X)>0}(ρaεdΣ)(X)|u|2,

    with aR, XRn+1 and dΣ() the signed distance function to Σ. Let ge be the Euclidean metric on Rn+1 and denote by ν the unit normal vector field on Σ. We define Fermi coordinates in a tubular neighborhood of Σ as follows: let zΣ and yR, and define

    Z(z,y):=z+yν(z).

    Points z and yν(z) belong to Rn+1. Given yR, we define

    Σy:={Z(z,y)Rn+1 : zΣ}.

    Following Lemma 6.1 in [18], one has that the induced metric on Σy is given by

    gy=g02yh0+y2h0h0,

    where g0 is the induced metric on Σ, h0 is the second fundamental form on Σ and h0h0 its square; namely we have

    h0(t1,t2)=g0(t1ν,t2)

    for all t1,t2 on the tangent bundle of Σ. Notice in particular that, in local coordinates, the terms g0,h0,h0h0 depend only on z. Therefore, invoking Lemma 6.3 in [18], one finally has

    Zge=gy+dy2

    where gy is considered as a family of metrics on the tangent bundle of Σ, depending smoothly on y in a neighborhood of 0 in R.

    In other words, we are obtaining a quadratic form for v(z,y)=u(Z(z,y)) of the form

    y00ρaε(y)Σy(|gyv|2+|yv|2)detgy.

    Recall that the variation with respect to y of of the volume form of the parallel hypersurfaces Σy satisfy the equation:

    Hy=1detgyddydetgy. (1.11)

    Hence, by considering a parametrization of Σ of the form z=ψ(x) with xRn, then one obtains for w(x,y)=v(ψ(x),y)

    ρaε(y)Aww,

    where

    A(x,y)=(˜A(x,y)001)detgy.

    We remark that the matrix A satisfies Assumption (HA), and can be expressed as in Notation (HA+) with μ(x,y)=detgy. As ΣC3,α, we have μC1,α(B+r0) for r0 small enough. Hence we are in the position to apply Theorem 1.6. Next we have to compare the two families vaε and χaε. At first, in order to prove point 1) we remark that Proposition A.3 ensures uniform-in-ε C0,α estimates for the ratio vaε/χaε. Using (1.11), we infer that yμ(,0)C1,α(B+r0) and finally, by virtue of Proposition A.4, we obtain that also the ratio vaε/χaε satisfies the desired uniform bounds in C1,α(Br{yε}), for r<r0.

    Below is the list of symbols we shall use throughout this paper.

    Rn+1+=Rn×(0,+) z=(x,y) with xRn, y>0
    Σ={y=0} characteristic manifold
    B+r=Br{y>0} half ball
    +B+r=Sn+(r)=Br{y>0} upper boundary of the half ball
    0B+r=Br{y=0} flat boundary of the half ball
    ρaε(y)=(ε2+y2)a/2 regularized weight
    ωaε(y)=ρaε(y)πaε(y) regularized auxiliary weight
    Lρaεu=div(ρaε(y)A(x,y)u) regularized operator
    H1(Ω,ρaε(y)dz) weighted Sobolev space given by the completion of C(¯Ω)
    H10(Ω,ρaε(y)dz) weighted Sobolev space given by the completion of Cc(Ω)
    ˜H1(Ω,ρaε(y)dz) weighted Sobolev space given by the completion of Cc(¯ΩΣ)
    H1,a(Ω)=H1(Ω,|y|adz) weighted Sobolev space for ε=0
    ayu=|y|ayu "weighted" derivative
    y|y|a characteristic odd solution
    vaε characteristic odd solution in presence of A and ε>0
    a+=max{a,0}

    In this section we collect the natural notions of Sobolev spaces, and their main properties, needed to work in our degenerate or singular context (for further details see [21, Section 2]). Let ΩRn+1 be non empty, open and bounded. Denoting by C(¯Ω) the set of real functions u defined on ¯Ω such that the derivatives Dαu can be continuously extended to ¯Ω for all multiindices α, then for any aR, ε0 we define the weighted Sobolev space H1(Ω,ρaε(y)dz) as the closure of C(¯Ω) with respect to the norm

    uH1(Ω,ρaε(y)dz)=(Ωρaεu2+Ωρaε|u|2)1/2.

    To simplify the notation we will denote

    H1,a(Ω)=H1(Ω,|y|adz)=H1(Ω,ρa0(y)dz).

    In the same way, we define H10(Ω,ρaε(y)dz) as the closure of Cc(Ω) with respect to the norm

    uH10(Ω,ρaε(y)dz)=(Ωρaε|u|2)1/2.

    We will denote by ˜H1(Ω,ρaε(y)dz) the closure of Cc(¯ΩΣ) with respect to the norm H1(Ω,ρaε(y)dz). In particular, when a<1, there is a natural isometry (on balls B centered in a point on Σ of any radius)

    Taε:˜H1(B,ρaε(y)dz)˜H1(B):uv=ρaεu,

    where ˜H1(B) is endowed with the equivalent norm with squared expression

    Qε(v)=B|v|2+[(yρaε2ρaε)2+y(yρaε2ρaε)]v2Byρaε2ρaεyv2,

    (this is detailed in the appendix B.5). We remark that both in the super singular and super degenerate cases, that is when a(,1][1,+) and ε=0, when the weight is taken outside the A2-Muckenhoupt class, one has

    H1,a(Ω)=˜H1,a(Ω). (2.1)

    This happens for very opposite reasons: roughly speaking, when a1 then the singularity is so strong to force the function to annihiliate on Σ (we will call this case the super singular case). Instead, when a1, then the strong degeneracy leaves enough freedom to the function to allow it to be very irregular through Σ (we will call this case the super degenerate case). In the latter case, Σ has vanishing capacity with respect to the energy |y|a|u|2.

    The Sobolev embedding theorems are stated in details in [21] as inequalities which are uniform in ε. This point is fundamental in order to develop a local regularity theory which is stable with respect to the regularization parameter ε. Hence, following some results contained in [12], the critical Sobolev exponents do depend on how the weighted measures dμ=ρaε(y)dz scale on balls of small radius r>0: one can check that there exists b,d>0 independent from ε0 (in the locally integrable case a>1) such that for small radii

    μ(Br(z))brd.

    So, we can define the effective dimension

    d=n+1+a+=n(a),

    and the Sobolev optimal exponent is

    2(a)=2dd2=2(n+1+a+)n+a+1.

    For details one can refer to Theorems 2.4 and 2.5 in [21].

    In the very same way one can define weighted Sobolev spaces for the class of weights ωaε; that is, the spaces H1(Ω,ωaε(y)dz)=˜H1(Ω,ωaε(y)dz) (the equality is due to the fact that ωaε is always a super degenerate weight as a<1) and H10(Ω,ωaε(y)dz).

    In this case one can check that there exist two positive constants ¯b,¯d>0 independent on ε0 such that d¯μ=ωaε(y)dz has the following growth condition on small balls of radius r>0

    ¯μ(Br(z))¯br¯d,

    and the effective dimension is given by ¯d=n+1+2+(a)+=n+3+(a)+=¯n(a). Hence one can state the following

    Theorem 2.1. Let a(,1), n1, ε0 and uC1c(Ω). Then there exists a constant which does not depend on ε0 such that

    (Ωωaε|u|¯2(a))2/¯2(a)c(¯d,¯b,Ω)Ωωaε|u|2,

    where the optimal embedding exponent is

    ¯2(a)=2¯d¯d2=2(n+3+(a)+)n+(a)++1.

    Throughout the paper, we are going to consider different elliptic equations depending on different families of weights. Nevertheless, we will deal with right hand sides having forcing terms, terms expressed by the divergence of a given field and drift terms (we will see that any other possible term that will appear can be translated in one of these). In order to give an unified definition of energy solutions to weighted problems, we will consider a generic measurable weight function w, and define an energy solution u in B1 to

    div(wAu)=wf+div(wF)+wbuin B1. (2.2)

    We say that uH1(B1,wdz) is an energy solution to (2.2) if

    B1wA(x,y)uϕ=B1wfϕB1wFϕ+B1w(bu)ϕ,ϕCc(B1)H1(B1,wdz), (2.3)

    any time the terms in the right hand side give sense to the previous integrals. We remark that we are not assuming local integrability of the weight, and this is the reason why we must consider test functions in the suitable weighted Sobolev space.

    Now, we recall the consequent definition of energy solutions in case the weight term is given by ρaε(y), with aR and ε0 (the following definition is contained in [21]). Let us consider the following problem

    div(ρaεAu)=ρaεf+div(ρaεF)in B1. (2.4)

    We say that uH1(B1,ρaε(y)dz) is an energy solution to (2.4) if

    B1ρaεA(x,y)uϕ=B1ρaεfϕB1ρaεFϕ,ϕCc(B1)H1(B1,ρaε(y)dz). (2.5)

    We remark that the condition in (2.5) can be equivalently expressed testing with any ϕCc(B1Σ) if a(,1][1,+) and ε=0. In order to give a sense to energy solutions to (2.4) we need the following minimal hypothesis on the right hand side.

    Assumption 2.2 (Hfρaε). Let a(1,+). Then if n2 or n=1 and a+>0, the forcing term f in (2.4) belongs to Lp(B1,ρaε(y)dz) with p(2(a)) the conjugate exponent of 2(a); that is,

    (2(a))=2(n+1+a+)n+a++3.

    If n=1 and a+=0 then fLp(B1,ρaε(y)dz) with p>1.

    Let a(,1]. Then if n2, the condition on the forcing term is (ρaε)1/2fLp(B1) with p(2(a))=(2). If n=1, then any p>1 is allowed.

    Assumption 2.3 (HFρaε). Let a(1,+). The condition on the field F=(F1,...,Fn+1) in (2.4) is FLp(B1,ρaε(y)dz) with p2. Let a(,1]. Then the condition is (ρaε)1/2FLp(B1) with p2.

    We are concerned with local regularity of energy odd solutions to (2.4) with a(,1) and ε0. Our analysis relies in the validity of suitable Liouville type theorems which hold true whenever a>1; that is, when the weight |y|a is locally integrable. In order to ensure regularity results also in the super singular case a1, we will consider the ratio w between the odd solution u and the function vaε defined in (1.6) which is odd and satisfies

    div(ρaεAvaε)=divx(ρaεμ˜Bxvaε)+divx(T)in B1, (2.6)

    whenever the right hand side in the equation satisfies suitable integrability assumptions and the matrix A is written as in Notation (HA+). As we have already remarked in the introduction, such a function vaε plays the role of the characteristic odd solution y|y|a in presence of a matrix and of regularization.

    The following Lemma is a formal computation

    Lemma 2.4. Let aR, ε>0 and let u,v be solutions to

    div(ρaεAu)=ρaεf,div(ρaεAv)=ρaεgin B1,

    with v>0 and A satisfying Assumption (HA). Then the function w=u/v is solution to

    div(ρaεv2Aw)=ρaεvfρaεugin B1.

    Proof. Let recall ρ=ρaε. Then

    div(ρv2Aw)=div(ρv2A(uvuvv2))=div(ρvAuρuAv)=vdiv(ρAu)ρv(Au)+udiv(ρAv)+ρu(Av)=vdiv(ρAu)ρv(Au)+udiv(ρAv)+ρv(ATu)=ρvfρug.

    The new class of weights appearing in the auxiliary equation for the ratio w=u/vaε is given by ρaε(vaε)2 and it will be equivalent (using (1.5)) to

    ωaε(y)=ρaε(y)πaε(y)=ρaε(y)(1a)2(χaε(y))2=ρaε(y)((1a)y0ρaε(s)ds)2.

    We remark that, considering a(,1), such a class of weights is always super degenerate; that is, at Σ

    ωaε(y){y2if ε>0|y|2aif ε=0,

    with 2a(1,+).

    Formal computations show that the auxiliary equation for w (which corresponds to Eq (1.4) in Proposition 1.1 for ε=0 and A=I) in Br for any r<1 is given by

    div(ρaε(vaε)2Aw)=ρaε(vaε)2(¯f+Vw¯Fvaεvaε)+div(ρaε(vaε)2¯F)in Br, (2.7)

    with

    ¯f:=fvaε,¯F:=Fvaε

    and

    V:=divx(μ˜Bxvaε)vaε+divx(T)ρaεvaε.

    Actually we can rewrite the 0-order term, obtaining that the auxiliary equation for w in Br is given by

    div(ρaε(vaε)2Aw)=ρaε(vaε)2(¯f¯Fvaεvaε)+div(ρaε(vaε)2¯F)+divx(ρaε(vaε)2b˜Aw)ρaε(vaε)2(b˜AbIw+b˜Axw)+divx(ρaε(vaε)2¯Tw)ρaε(vaε)2(¯TbIw+¯Txw), (2.8)

    where for a (n,n)-dimensional matrix M

    bM=Mxvaεvaε,and¯T=Tρaεvaε.

    Thus we can write the equation the following form:

    div(ρaε(vaε)2Aw)=ρaε(vaε)2f+div(ρaε(vaε)2F1)+divx(ρaε(vaε)2F2w)+ρaε(vaε)2Vw+ρaε(vaε)2bxw. (2.9)

    We would like to prove that w is an even energy solution to (2.9) in Br in the sense that wH1(Br,ωaε(y)dz) and satisfies

    Brρaε(vaε)2Awϕ=Brρaε(vaε)2fϕBrρaε(vaε)2F1ϕBrρaε(vaε)2F2wxϕ+Brρaε(vaε)2Vwϕ+Brρaε(vaε)2(bxw)ϕ,

    for any ϕCc(BrΣ) (as we have already remarked, super degeneracy allows us to take test functions compactly supported away from Σ). In order to give a sense to energy solutions to (2.9) we need the following minimal hypothesis on the right hand side.

    Assumption 2.5 (Hfωaε). Let a(,1). Then the forcing term f in (2.9) belongs to Lp(B1,ωaε(y)dz) with p(¯2(a)) the conjugate exponent of ¯2(a); that is,

    (¯2(a))=2(n+3+(a)+)n+(a)++5.

    Assumption 2.6 (HF1ωaε). Let a(,1). Then the field term F1 in (2.9) belongs to Lp(B1,ωaε(y)dz) with p2.

    Assumption 2.7 (HF2ωaε). Let a(,1). Then the field term F2 in (2.9) belongs to Lp(B1,ωaε(y)dz) with p¯d=n+3+(a)+.

    Assumption 2.8 (HVωaε). Let a(,1). Then the 0-order term V in (2.9) belongs to Lp(B1,ωaε(y)dz) with

    p¯d2=n+3+(a)+2.

    Assumption 2.9 (Hbωaε).Let a(,1). Then the field b the drift term in (2.9) belongs to Lp(B1,ωaε(y)dz) with p¯d=n+3+(a)+.

    We will need the following important result, which contains also Proposition 1.1 when ε=0 and A=I.

    Proposition 2.10. Let a(,1), ε0 and let uεH1(B1,ρaε(y)dz) be an odd energy solution to (2.4) in B1. Then, fixed 0<r<1, the function wε=uε/vaε is an even energy solution in H1(Br,ωaε(y)dz) to (2.8), provided that the right hand side satisfies the suitable integrability assumptions stated above.

    Proof. First, we want to show that wεH1(Br,ρaε(y)(vaε)2(x,y)dz). We remark that since 1CμC and since the weight is super degenerate, we have that at Σ

    ρaε(vaε)2ωaε{|y|2aif ε=0|y|2if ε>0,

    with 2a(1,+), then the (H = W)-property does not necessarily hold (due to the lack of a Poincaré inequality, see [21]). Nevertheless, we can argue as follows: let ηCc(B1) be a radial decreasing cut off function such that 0η1 and η1 in Br. Let also for δ>0

    fδ(y)={0in B1{|y|δ}logyδin B1{δ|y|δe}1in B1{δe|y|}.

    Let φδ=ηfδ, then |φδ|1 and |φδ|c/y uniformly in δ>0. We remark that one can replace fδ with a function with the same properties which is C(B1). So,

    B1ρaε(vaε)2|φδwε|2B1ρaεu2εc. (2.10)

    Obviously in B1Σ Eq (2.8) holds. It is an easy consequence of Lemma 2.4, using that vaε is an odd energy solution to (2.6) in B1 and that vaε>0 in B1Σ. Then, testing the equation with φ2δwε, we obtain

    B1ρaε(vaε)2A(φδwε)(φδwε)=B1ρaε(vaε)2(φ2δAwεwε+2φδwεAwεφδ+w2εAφδφδ)=B1(RHS)φ2δwε+B1ρaε(vaε)2w2εAφδφδB1(RHS)φ2δwε+cB1ρaεy2u2εc, (2.11)

    and this is true by the weighted Hardy inequality in (B.12), weighted Sobolev embeddings (Theorem 2.1) and Assumptions (Hfωaε), (HF1ωaε), (HF2ωaε), (HVωaε) and (Hbωaε) which give a bound on the term with (RHS) of Eq (2.8). We remark that, fixed δ>0, the boundedness in norm H1(B1,ρaε(y)(vaε)2(x,y)dz) is enough to ensure that φδwε belongs to the same space. In fact, they have compact support away from Σ, and hence these norms are equivalent to the usual H1-norm. Since the bounds in (2.10) and (2.11) are uniform in δ>0, this is enough to have weak convergence for the sequence φδwε in H1(Br,ρaε(y)(vaε)2(x,y)dz) as δ0 and of course the limit is wε (it is almost everywhere pointwise limit).

    We have already remarked that in BrΣ Eq (2.8) holds. Then, one can conclude since the weighted Sobolev space H1(Br,ρaε(y)(vaε)2(x,y)dz) is super degenerate, and consequently test functions can be taken in Cc(BrΣ).

    In this section we present two important results which will be our main tool in order to prove regularity local estimates which are uniform with respect to ε0.

    Theorem 3.1. Let a(1,1), ε0 and w be a solution to

    {div(ρaε(y)w)=0in Rn+1+w(x,0)=0,

    and let us suppose that for some γ[0,1a), C>0 it holds

    |w(z)|C(1+|z|γ) (3.1)

    for every z. Then w is identically zero.

    Proof It is enough to prove the result only for ε{0,1}. In fact for any other value of ε>0 we can normalize the problem falling in the case ε=1.

    Case 1: ε=0.

    Let us consider wH1,aloc(Rn+1+) satisfying the conditions of the statement, that is, solution in the following sense

    Rn+1+yawϕ=0ϕCc(Rn+1+).

    Let us define

    E(r)=1rn+a1B+rya|w|2,H(r)=1rn+a+B+ryaw2.

    Note that, as the weight ya is locally integrable, (3.1) implies

    H(r)C(1+r2γ),r>0. (3.2)

    Now, defining wr(x)=w(rx), we have

    E(r)=B+1ya|wr|2andH(r)=Sn+ya(wr)2,

    and hence

    H(r)=2rE(r).

    We are looking for the best constant in the following trace Poincaré inequality

    B+1ya|u|2λ(a)Sn+yau2,u˜H1,a(B+1). (3.3)

    Actually we are able to provide the best constant λ(a) in (3.3), since it is given by the homogeneity of the unique (up to multiplicative constants) solution in ˜H1,a(B+1) to

    {Lau=0in B+1u>0in B+1u(x,0)=0uν=λ(a)uin Sn+,

    which is u(x,y)=y1a with λ(a)=1a. However λ(a) is the same as (B.9). Hence H(r)2λ(a)rH(r), and integrating, then we infer

    H(r)r2(1a)H(1).

    We obtain that if w is not trivial, the growth of H at infinity is at least r2(1a), in contradiction with (3.2) taking r large.

    Case 2: ε=1.

    Let us define

    E(r)=1rn+a1B+r(1+y2)a/2|w|2,H(r)=1rn+a+B+r(1+y2)a/2w2.

    Note that, as a>1, the ρaε's are uniformly locally integrable and thus (3.1) implies again

    H(r)C(1+r2γ),r>0,with γ<1a. (3.4)

    Hence,

    H(r)=2rE(r)arn+a+1+B+r(1+y2)a/21w2. (3.5)

    Moreover, defining wr(x)=w(rx) one has

    E(r)=B+1(1r2+y2)a/2|wr|2andH(r)=Sn+(1r2+y2)a/2(wr)2.

    By Lemma B.4 and Remark B.5, one can find for any radius r>0 the best constant λr(a) such that

    B+1(1r2+y2)a/2|u|2λr(a)Sn+(1r2+y2)a/2u2. (3.6)

    Defining ρk(y)=(1r2k+y2)a/2 with rk+ as k+, one can see

    λ(a)=minv˜H1(B+1){0}Qa(v)Sn+v2andλk(a)=minv˜H1(B+1){0}Qρk(v)Sn+v2.

    By Lemma B.4, λk(a)λ(a)=1a as k+.

    Now we want to prove that the correction term in (3.5) is of lower order as r+. By (B.13), we have that in ˜Cc(B+1)

    B+1ρr|u|2c0B+1ρryu2.

    Hence

    |arn+a+1+B+r(1+y2)a/21w2||a|rn+a+1+B+r(1+y2)a/21/2w2=|a|r2Sn+(1r2+y2)a/21/2(wr)2|a|r2Sn+(1r2+y2)a/2y1(wr)2|a|c0r2B+1(1r2+y2)a/2|wr|2=|a|c0r2E(r).

    Hence for r large enough

    H(r)2λr(a)rH(r),

    and since λr(a)λ(a)=1a, by integrating the above expression we deduce that, for all small δ, there exists r0>0 such that, for every r>r0

    H(r)r2(1aδ)H(r0),

    which says that if w is not trivial, the growth of H at infinity is at least r2(1aδ). Taking δ>0 so small that 1aδ>γ we find a contradiction with (3.4).

    Theorem 3.2. Let a(,1), ε0 and w be a solution to

    {div((ωaε(y))1w)=0in Rn+1+w=0in Rn×{0},

    and let us suppose that for some γ[0,1), C>0 it holds

    |w(z)|Cωaε(y)(1+|z|γ) (3.7)

    for every z=(x,y). Then w is identically zero.

    Proof. By a simple normalization argument, it is enough to prove the result only for ε{0,1}. We start with

    Case 1: ε=0.

    The case falls into the proof of Case 1 in [21, Theorem 3.4] replacing a(,1) with (a2)(,1).

    Case 2: ε=1.

    Let us now define

    E(r)=1rn+(a2)1B+r(ωa1(y))1|w|2,andH(r)=1rn+(a2)+B+r(ωa1(y))1w2.

    Note that, defining wr(x)=w(rx) one has

    E(r)=B+1(ωa1/r(y))1|wr|2,andH(r)=Sn+(ωa1/r(y))1(wr)2.

    First we remark that the growth condition (3.7) implies the following upper bound

    H(r)Cr2(a2)(1+r2γ),r>0, (3.8)

    (due to the local integrability of y2a) and heence

    Sn+ωa1/r(y)c

    uniformly in r>0. Therefore,

    H(r)=2rE(r)+Sn+ddr[(ωa1/r(y))1](wr)2. (3.9)

    By Lemma B.7 and Lemma B.8, one can find for any radius r>0 the best constant μr(a) such that

    B+1(ωa1/r(y))1|u|2μr(a)Sn+(ωa1/r(y))1|u|2. (3.10)

    Defining (ωak(y))1=(ωa1/rk(y))1 and μk=μrk with rk+ as k+, by Lemma B.8, μk(a)μ(a)=1(a2)=3a as k+.

    Now we want to prove that the correction term in (3.9) is of lower order as r+. By (B.27), we have that

    B+1(ωa1/r(y))1|u|2c0Sn+(ωa1/r(y))1yu2.

    Using

    Sn+ddr[(ωa1/r(y))1](wr)2=ar3Sn+1(1r2+y2)(ωa1/r(y))1(wr)22ar3Sn+y0(1r2+s2)a/21y0(1r2+s2)a/2(ωa1/r(y))1(wr)2,

    we can estimate the first term of the rest as follows

    |ar3Sn+1(1r2+y2)(ωa1/r(y))1(wr)2||a|r2Sn+1(1r2+y2)1/2(ωa1/r(y))1(wr)2|a|r2Sn+(ωa1/r(y))1y(wr)2cr2E(r).

    Moreover, when a0 the second term of the rest can be estimated as

    |2ar3Sn+y0(1r2+s2)a/21y0(1r2+s2)a/2(ωa1/r(y))1(wr)2|2|a|r2Sn+ryry0(1+s2)a/21ry0(1+s2)a/2(ωa1/r(y))1y(wr)22|a|r2Sn+(ωa1/r(y))1y(wr)2cr2E(r),

    and this is due to the fact that, calling z=ry[0,+), by the fact that

    f(z)=zz0(1+s2)a/21z0(1+s2)a/2

    is continuous and such that f(0)=0 and

    f(z)z+{czaif a(1,0]logzzif a=11zif a<1

    and hence f(z)c in [0,+). Instead, when a(0,1) the second term of the rest can be estimated as

    |2ar3Sn+y0(1r2+s2)a/21y0(1r2+s2)a/2(ωa1/r(y))1(wr)2|2|a|r2aSn+(ry)1ary0(1+s2)a/21ry0(1+s2)a/2(ωa1/r(y))1y1a(wr)22|a|r2aSn+(ωa1/r(y))1y1a(wr)22|a|r2aSn+(ωa1/r(y))1y(wr)2cr2aE(r),

    using the fact that 0y1 and by the fact that

    f(z)=z1az0(1+s2)a/21z0(1+s2)a/2

    is continuous and such that f(0)=0 and

    f(z)z+c

    and hence f(z)c in [0,+). Hence for r large enough

    H(r)2μr(a)rH(r),

    and since μr(a)μ(a)=1(a2), we can choose a small δ>0 such that 1(a2)δ>γ(a2). Hence, by integrating the above expression we deduce that there exists r0>0 such that, for every r>r0, we have μr(a)>1(a2)δ>γ(a2) and

    H(r)r2(1(a2)δ)H(r0), (3.11)

    which is in contradiction with (3.8) for r large if w is not trivial.

    Corollary 3.3. Let a(,1), ε0 and w be a solution to

    {div(ωaε(y)w)=0in Rn+1+ωaεyw=0in Rn×{0},

    and let us suppose that for some γ[0,1), C>0 it holds

    |w(z)|C(1+|z|γ) (3.12)

    for every z. Then w is constant.

    Proof. Again, it is enough to treat the cases ε{0,1}. Let us assume ε=1, the case ε=0 coincides with the case ε=0 in [21, Corollary 3.5], by replacing in the proof a(1,+) by (2a)(1,+). Then we have (by an even reflection across Σ) an even solution w to

    div(ωa1(y)w)=0in Rn+1.

    Such a solution is wH1,2loc(Rn+1)=H1loc(Rn+1,|y|2dz), with the growth condition (3.12). Now we observe that, as w is not constant with a sublinear growth at infinity, v=ωa1(y)yw can not be trivial, otherwise w would be globally harmonic and sublinear, in contradiction with the Liouville theorem in [17]. Hence, if w is not constant, v must be an odd and nontrivial solution to

    {div((ωa1(y))1v)=0in Rn+1+v=0in {y=0}.

    By (3.11), we know that the weighted average of v2 must satisfy a minimal growth rate as

    H(r)=1rn+(a2)+B+r(ωa1(y))1v2cr2(1(a2)δ),1δ>γ,

    for rr0 depending on δ>0 chosen. Therefore, by integrating, we obtain

    B+rωa1(y)(yw)2=r0dt+B+t(ωa1(y))1v2crn(a2)+22δ.

    On the other hand, we have, by (3.12)

    B+rωa1(y)(yw)2B+rωa1(y)|w|2cB+2rωa1(y)|w|2c(1+rn(a2)+2γ)

    in contradiction with the previous inequality when r is large, since 1δ>γ.

    As a first step in our regularity theory for odd solutions, we state some results on local uniform estimates for solutions to (1.7); that is,

    div(ρaε(vaε)2Auε)=ρaε(vaε)2fε+div(ρaε(vaε)2Fε)+ρaε(vaε)2bεuεin B1.

    Using a Moser iteration argument (see also [11, Section 8.4]), one can prove the following nowadays standard result.

    Proposition 4.1. Let a(,1) and ε0. Let uH1(B1,ωaε(y)dz) be an energy solution to (1.7). Let β>1,

    p1>¯d2=n+3+(a)+2,p2,p3>¯d.

    Let moreover bLp3(B1,ωaε(y)dz)b. Then, for any 0<r<1 there exists a positive constant independent of ε (depending on n, a, p1, p2, p3, β, b, r and α) such that

    uL(Br)c(uLβ(B1,ωaε(y)dz)+fLp1(B1,ωaε(y)dz)+FLp2(B1,ωaε(y)dz)).

    Proof. The proof follows the same steps as in [21, Proposition 2.17], but iterating the Sobolev embedding in Theorem 2.1.

    Now we show how the uniform local boundedness of solutions to (1.7) yields also local uniform bounds in Hölder spaces.

    Theorem 4.2. Let a(,1). As ε0, let {uε} be a family of solutions in B+1 of (1.7) satisfying boundary conditions (evenness)

    ρaε(vaε)2yuε=0on 0B+1.

    Let r(0,1), β>1, p1>¯d2=n+3+(a)+2, p2,p3>¯d, and α(0,1)(0,2n+3+(a)+p1](0,1n+3+(a)+p2](0,1n+3+(a)+p3]. Let bLp3(B1,ωaε(y)dz)b. Let moreover A satisfy assumption (HA) with continuous coefficients. Then, there is a positive constant depending on a, n, β, p1, p2, p3, b, α and r only such that

    uεC0,α(B+r)c(uεLβ(B+1,ωaε(y)dz)+fεLp1(B+1,ωaε(y)dz)+FεLp2(B1,ωaε(y)dz)).

    Proof. The proof follows the very same steps as in the proof of [21, Theorem 4.1]. First, one has to remark that the suitable Hölder continuity for ε0 fixed is given by the theory developed for even solutions to degenerate problems in [21]. Then, one can argue by contradiction with the usual blow up argument considering two blow up sequences

    vk(z)=(ηuk)(zk+rkz)(ηuk)(zk)Lkrαk,wk(z)=η(zk)(uk(zk+rkz)uk(zk))Lkrαk,

    (with the same asymptotic behaviour on compact sets) defined in the rescaled domains B(k)=Bzkrk (where B=B1+r2 and {zk} is one of the two sequences of points where Hölder seminorms blow up), the first possessing some uniform Hölder continuity, and the second one satisfying suitable problems on rescaled domains which blow up. In order to complete the proof we prove some steps.

    Step 1: blow-ups. The first thing to do is to characterize the possible asymptotic behaviours of the weights ρaε(vaε)2 in the rescaled points: that is,

    pk(z):=ρaε(yk+rky)(vaε(zk+rkz))2=(ε2k+(yk+rky)2)a/2(yk+rky0(ε2k+s2)a/2μ(xk+rkx,s)1ds)2.

    To this end, let us define by Γk=(εk,yk,rk) and denote νk=|Γk|. The latter is a bounded sequence and, up to subsequences, has finite limit ν=|(0,y,0)|0 (where we have assumed zkz=(x,y)). Taking possibly another subsequence, we may assume that the normalized sequence

    ˜Γk=Γkνk=(˜εk,˜yk,˜rk)=(εkνk,ykνk,rkνk)

    has a limit

    ˜Γk˜Γ=(˜ε,˜y,˜r)S2,

    and moreover that

    limk+˜yk˜rk=˜l[0,+].

    Thus we can consider ˜Σ=limΣk; that is,

    ˜Σ={{(x,y)Rn+1 : y=˜l}if˜l<+,if˜l=+.

    After rescaling the independent variables, we find new weights having the form:

    pk(z)=ν2ak(˜ε2k+(˜yk+˜rky)2)a/2(˜yk+˜rky0(˜ε2k+t2)a/2μ(xk+rkx,νkt)1dt)2,

    and, in order to study their asymptotics, we have to distinguish between different cases:

    Case 1. ν>0. Then, ˜r=˜ε=0 and ˜y=1. Moreover, it is easy to see, using that 1/μ is continuous, that pk(z)=c+o(1).

    Case 2. ν=0 and ˜ε=0 (˜y0˜r0). Using the continuity of 1/μ, up to a vertical translation of ˜l, one obtains

    pk(z)=ν2ak˜p(y)(1+o(1))

    where

    ˜p(y)={cif ˜r=0,c|y|2aif ˜r0.

    Case 3. ν=0 and ˜ε(0,1). Using again the continuity of 1/μ, , up to a vertical translation of ˜l, we obtain

    pk(z)=ν2ak˜p(y)(1+o(1))

    where

    ˜p(y)={cif ˜r=0,c ωa1(y)if ˜r0.

    in the second case up to a dilation of ˜ε˜r.

    Case 4. ν=0 and ˜ε=1 (˜y=0˜r=0). As usual, by the continuity of 1/μ, , up to a vertical translation of ˜l, one obtains, if ˜rk=o(˜yk)

    pk(z)=ν2ak˜y2kc(1+o(1)),

    and otherwise

    pk(z)=ν2ak˜r2kc|y|2(1+o(1)).

    Let us define

    hk={ν2akin Cases 1,2,3,ν2ak˜y2kin Case 4, and ˜rk=o(˜yk)ν2ak˜r2kin Case 4, otherwise,

    and ˜pk=pkhk. We have shown that, up to the suitable normalization, the rescaled weights ˜pk do converge uniformly to ˜p on compact sets of Rn+1˜Σ (or the whole Rn+1 whenever ˜Σ=). Note that this latter case is equivalent to the limiting weight ˜p be constant.

    Step 2: the limiting equation and uniform-in-k energy estimates. The equation for the rescaled variable wk becomes:

    div(˜pkA(zk+rk)wk)(z)=η(zk)Lkr2αk˜pk(z)fk(zk+rkz)+div(η(zk)Lkr1αk˜pk()Fk(zk+rk))(z)+rk˜pk(z)bk(zk+rkz)wk(z). (4.1)

    By a Caccioppoli type inequality, easily obtained by multiplying (4.1) by ¯η2wk, being ¯η a cut-off function, taking into account that the wk are uniformly bounded and that

    ● the first term in the right hand side of (4.1) is bounded in L1loc;

    ● the field η(zk)Lkr1αkFk(zk+rk) in the second term in the right hand side of (4.1) is bounded in L2loc(˜p(z)dz);

    rkbk(zk+rk) is bounded in L2loc(˜pk(z)dz);

    then we obtain uniform-in-k energy bounds holding on compact subsets of Rn+1:

    R>0,c>0,k,BR˜pkA(zk+rkz)wkwkc.

    The computations are very similar to the ones done in the following step.

    Step 3: the right hand side vanishes as k+. Next we wixh to check that the right hand sides in the rescaled equations vanish in L1loc(Rn+1˜Σ) (or L1loc(Rn+1) whenever ˜Σ=), and that consequently the limit w is an energy solution of

    div(˜pA(z)w)=0in Rn+1˜Σ, (4.2)

    even with respect to ˜Σ (when not empty). We use the continuity of the matrix A in order to obtain a constant coefficients matrix in the limit Eq (4.2) together with the fact that ˜Σ is invariant with respect to the limit matrix to have evenness across the characteristic hyperplane.

    Let us show that the right hand sides vanish in L1loc at least for one of the cases (the other cases are very similar), for instance when hk=ν2ak˜y2k; that is, Case 4, when ˜rk=o(˜yk). Indeed, let ϕCc(Rn+1): using the fact that for k large enough supp(ϕ)BRB(k), using Hölder inequality, we have

    |BR˜pk(z)fεk(zk+rkz)ϕ(z)dz|ϕL(BR)(1rn+1khkBrkR(zk)ρaεk(ζn+1)(vaεk(ζ))2|fεk(ζ)|p1dζ)1/p1(BR˜pk(z)dz)1/p1crn+1p1kν2ap1k˜y2p1k,

    and hence the first term in the right hand side converges to zero since α2n+3+(a)+p1, ˜rk=rkνk, the fact that 0rkνk and having

    η(zk)Lkr2αn+3+(a)+p1k(r(a)+kνak)1/p1(˜rk˜yk)2/p10.

    With analogous computations one can check that also the second term in the right hand side vanishes.

    Concerning the third term, one can estimate the integral as follows

    rk|BR˜pk(z)bk(zk+rkz)wk(z)ϕ(z)|rkϕL(BR)(BR˜pk|wk|2)12(BR˜pk)p322p3(1hkrn+1kBrkR(zk)ρaεk(ζn+1)(vaεk(ζ))2|bεk(ζ)|p3)1p3cb1p3r1n+3+(a)+p3k(r(a)+kνak)1/p3(˜rk˜yk)2/p3(BR˜pk|wk|2)1/2=tk(BR˜pk|wk|2)1/2. (4.3)

    The sequence tk converges to zero, having p3>n+3+(a)+. Moreover, the full term vanishes using the uniform energy bound obtained in the previous step.

    Step 4: the limit belongs to H1loc(Rn+1,˜pdz). At this point, always up to subsequences, we know that the (pointwise) convergence to w holds also in the weak H1loc(Rn+1˜Σ) topology. Now we wish to infer that the limit w belongs to the space H1loc(Rn+1,˜pdz) as the closure of C with respect to the weighted norm (as defined in ยง2). Let us start with the easiest case when ˜Σ= and the limiting weight ˜p is constant. Moreover, we know tha ˜pk converge uniformly to ˜p on compact sets. Thus the sequence wk converges weakly H1 to w on each compact subset. The convergence to the case when ˜Σ requires a more thorough analysis. In order to ensure that wH1loc(Rn+1,˜p(y)dz) also when ˜Σ, one can argue as follows: using the fact that μ is continuous with 1C<μ<C, then fixed a compact set of Rn+1, we can find positive constants ck,Ck (which are uniformly bounded from above and below by two constants respectively 0<c1<c2<+) such that

    ck˜ωεk(yk+rky)˜pk(x,y)Ck˜ωεk(yk+rky)andCkck1.

    where we have denoted

    ˜ωk=ωεkhk. (4.4)

    Now, reabsorbing the weights as in (B.30) and using the family of isometries given by

    ¯Tk(wk)=(˜ωk(yk+rky))1/2wk=Wk,

    one obtains uniform boundedness of the Wk's in H1loc(Rn+1), and hence they weakly converge in the same space to W. Coming back with the inverse isometry associated with the limit weight

    ¯T(w)=(˜p(y))1/2w=W,

    we obtain wH1loc(Rn+1,˜p(y)dz).

    Step 5: end of the proof. Next we wish to show that w solves the equation in (4.2) also across the limiting characteristic hyperplane ˜Σ. Indeed, using wH1loc(Rn+1,˜p(y)dz) jointly with Eq (4.2) holding in Rn+1Σ, using that Cc(¯BRΣ) is actually dense in H1(BR,˜p(y)dz) (all the weights here, including the limit one, are super degenerate), as we have already remarked in (2.1). Eventually, one can reach a contradiction by applying the suitable Liouville theorems in [17,21] and Corollary 3.3 for the case ˜p(y)=ωa1(y).

    As a further step we now provide C1,α uniform estimates.

    Theorem 4.3. Let a(,1). As ε0, let {uε} be a family of solutions in B+1 of (1.7) satisfying boundary conditions (evenness)

    ρaε(vaε)2yuε=0on 0B+1.

    Let r(0,1), β>1, p1,p2>¯d=n+3+(a)+. Let bL2p2(B1,ωaε(y)dz)b. Let Fε=(F1ε,...,Fn+1ε) with the y-component vanishing on Σ: Fn+1ε(x,0)=Fyε(x,0)=0 in 0B+1. Let moreover A satisfy assumption (HA) with α-Hölder continuous coefficients and α(0,1n+3+(a)+p1](0,1n+3+(a)+p2]. Then, there is a positive constant depending on a, n, β, p1, p2, b, α and r only such that

    uεC1,α(B+r)c(uεLβ(B+1,ωaε(y)dz)+fεLp1(B+1,ωaε(y)dz)+FεC0,α(B+1)).

    Proof. We wish to follow the same steps of proof of [21, Theorems 5.1 and 5.2]. Among others, we have to deal with an additional difficulty; that is, our weights here do depend on the full variable z=(x,y) and not on y only. For our purposes, we can take advantage of the fact that the ratio

    γaε(x,y)=:vaε(x,y)χaε(y)=y0ρaε(s)μ1(x,s)dsy0ρaε(s)ds (4.5)

    is uniformly bounded in C0,α with respect to ε (just apply Lemma A.3, using the fact that μ1C0,α since the matrix A possesses α-Hölder continuous coefficients). Hence, one can rewrite our operator as

    div(ρaε(y)(vaε(x,y))2A(x,y)uε)=div(ωaε(y)Aε(x,y)uε), (4.6)

    where, up to constants, the new family of matrices is defined as

    Aε(x,y)=(γaε(x,y))2A(x,y), (4.7)

    with coefficients which are uniformly bounded in C0,α with respect to ε.

    With these premises, we are now able to follow the construction made in [21, Theorems 5.1 and 5.2]. Just to give the idea, the contradiction argument uses two blow-up sequences

    vk(z)=η(ˆzk+rkz)Lkr1+αk(uk(ˆzk+rkz)uk(ˆzk)),wk(z)=η(ˆzk)Lkr1+αk(uk(ˆzk+rkz)uk(ˆzk)),

    for zB(k):=Bˆzkrk. Hence, one has to work with

    ¯vk(z)=vk(z)vk(0)z,¯wk(z)=wk(z)wk(0)z,

    or

    ¯vk(z)=vk(z)xvk(0)x,¯wk(z)=wk(z)xwk(0)x,

    respectively when d(zk,Σ)rk+ (in this case we choose ˆzk=zk), or d(zk,Σ)rkc uniformly in k (in this case we choose ˆzk=(xk,0) to be the projection on Σ of zk, where zk=(xk,yk)).

    Hence, reasoning as in the previous Theorem 4.2, one can characterize all possible rescalings of the weights (in facts the possible scalings of weights pk and ωk are the same), and prove that the limit w is an energy entire solution to the suitable limiting problem.

    We remark that in order to show that the limit equation has a constant coefficient matrix one has to reason as in [21, Remark 5.3], using the α-Hölder continuity of coefficients of the matrix (in this case we will invoke the uniform bounds with respect to ε in C0,α for the coefficients of Aεk).

    Nevertheless, we need also to deal with drift terms in the rescaled equations, and we wish to show that they vanish once testing with the suitable test function ϕ supported in BR. We assume here that hk=ν2ak (one of the possible cases). Hence, also in this case we use the fact that we know a priori that the sequence {uk} is uniformly locally bounded in C0,β spaces, for any choice of β(0,1) (follows from Theorem 4.2). Reasoning as in [21, Remark 5.3], this gives the following energy estimate

    Brpk(z)|uk|2cr2(1β)B2rpk(z). (4.8)

    Hence, we can estimate

    rk|BR˜pk(z)bk(ˆzk+rkz)wk(z)ϕ(z)|r1αkη(ˆzk)LkϕL(BR)BR˜pk(z)|bk(ˆzk+rkz)||uk(ˆzk+rkz)|cr1αkLk(BR˜pk(z)|bk(ˆzk+rkz)|2)1/2(BR˜pk(z)|uk(ˆzk+rkz)|2)1/2cr1αkLk(BR˜pk(z))1/2p2(1rn+1khkBrkR(ˆzk)ρaεk(ζn+1)(vaεk(ζ))2|bk(ζ)|2p2)1/2p2cr1βk(2BR˜pk(z))1/2cLk(rkνk)1/p2(r(a)+kνak)1/2p2r12(1αn+3+(a)+p2)krβ1+α2k0

    since Lk+, νkrk, p2>n+3+(a)+, α1n+3+(a)+p2 and choosing β>1+α2.

    The following is the main result of the paper (we have already stated it in a simplified version in Theorem 1.6 in the introduction). Let a(,1), the matrix A written as in Notation (HA+) and let uε be an odd energy solution to (1.2) in B+1; that is,

    {div(ρaεAuε)=ρaεfε+div(ρaεFε)in B+1uε=0on 0B+1. (4.9)

    Let also vaε be defined as in (1.6) in B+1. Then, we have already showed (in Proposition 2.10) that under suitable integrability assumptions on the terms in the right hand side and on coefficients of the matrix A, then functions

    wε=uεvaε

    are even energy solutions (for any R<1) to Eq (2.8); that is,

    div(ρaε(vaε)2Awε)=ρaε(vaε)2(¯fε¯Fεvaεvaε)+div(ρaε(vaε)2¯Fε)+divx(ρaε(vaε)2(b˜Aε+¯Tε)wε)ρaε(vaε)2((b˜Aε+¯Tε)bIεwε+(b˜Aε+¯Tε)xwε),

    with boundary condition

    ρaε(vaε)2ywε=0on 0B+R,

    and where we denote by

    ¯fε=fεvaε,¯Fε=Fεvaε,bMε=Mxvaεvaε,and¯Tε=Tρaεvaε,

    (M is a general (n,n)-dimensional matrix).

    Theorem 4.4. Let a(,1), the matrix A written as in Notation (HA+) and as ε0 let {uε} be a family of solutions in B+1 of (4.9).

    1) Let r(0,1), β>1, p1,p2>n+3+(a)+2, p3,p4>n+3+(a)+, and α(0,2n+3+(a)+p1](0,2n+3+(a)+p2](0,1n+3+(a)+p3](0,1n+3+(a)+p4]. Let also

    (b˜Aε+¯Tε)bIεLp2(B+1,ωaε(y)dz)b1,b˜Aε+¯TεLp4(B+1,ωaε(y)dz)b2.

    Let us moreover take A with continuous coefficients. There is a positive constant depending on a, n, β, p1, p2, p3, p4, b1, b2, α and r only such that functions

    wε=uεvaε

    satisfy

    wεC0,α(B+r)c(wεLβ(B+1,ωaε(y)dz)+¯fε¯FεvaεvaεLp1(B+1,ωaε(y)dz)+¯FεLp3(B+1,ωaε(y)dz)).

    2) Let r(0,1), β>1, p1,p2>n+3+(a)+, and α(0,1n+3+(a)+p1](0,1n+3+(a)+p2]. Let ¯Fε=(¯F1ε,...,¯Fn+1ε) with the y-component vanishing on Σ: ¯Fn+1ε(x,0)=¯Fyε(x,0)=0 in 0B+1. Let also

    (b˜Aε+¯Tε)bIεLp2(B+1,ωaε(y)dz)b1,b˜Aε+¯TεC0,α(B+1)b2.

    Let's moreover take A with α-Hölder continuous coefficients. There is a positive constant depending on a, n, β, p1, p2, b1, b2, α and r only such that functions

    wε=uεvaε

    satisfy

    wεC1,α(B+r)c(wεLβ(B+1,ωaε(y)dz)+¯fε¯FεvaεvaεLp1(B+1,ωaε(y)dz)+¯FεC0,α(B+1)).

    We remark that uniform estimates with respect to the regularization are optimal in C1,α-spaces (in [21, Remark 5.4] we provided a counterexample which show that C2,α estimates could not be uniform up to Σ as ε0).

    In order to prove our main result we have the following useful preliminary result on equations of the form

    div(ρaε(vaε)2Auε)=ρaε(vaε)2Vuε+div(ρaε(vaε)2Fuε)in B1. (4.10)

    Lemma 4.5. Let a(,1) and ε0. Let uH1(B1,ωaε(y)dz) be an energy solution to (4.10), where VLp1(B1,ωaε(y)dz) with p1>¯d2=n+3+(a)+2 and FLp2(B1,ωaε(y)dz) with p2>¯d=n+3+(a)+. Let

    VLp1(B1,ωaε(y)dz)b1,FLp2(B1,ωaε(y)dz)b2.

    1) Then, for any 0<r<1 and β>1 there exists a positive constant independent of ε (depending on n, a, r, β, p1, p2, b1, b2), m1>¯d2 and m2>¯d such that

    VuLm1(Br,ωaε(y)dz)+FuLm2(Br,ωaε(y)dz)cuLβ(B1,ωaε(y)dz).

    2) If moreover p1>¯d=n+3+(a)+ and FC0,α(B1) for some α(0,1),

    VLp1(B1,ωaε(y)dz)b1,FC0,α(B1)b2,

    then for any 0<r<1 and β>1 there exists a positive constant independent of ε (depending on n, a, r, β, p1, α, b1, b2), and m1>¯d such that

    VuLm1(Br,ωaε(y)dz)+FuC0,α(Br)cuLβ(B1,ωaε(y)dz).

    Proof. The proof is done applying Moser iterations on a finite number of small enough balls which cover Br. The radius of such balls is chosen in order to ensure coercivity of the quadratic forms. Hence, using the fact that the weighted integrability of V and F is suitably large, by a finite number of Moser iterations one can promote the integrability of u itself, up to guarantee that the products Vu and Fu have the desired integrability (this type of argument is classic, see for instance [11, Section 8.4]). Since the number of iterations is finite, one can control uniformly the constants in the iterative process, proving point 1). At to point 2), thanks to point 1) we can apply Theorem 4.2 in order to obtain that the solution is C0,α with a bound which is independent from ε. Hence, we obtain the second inequality taking into account the Hölder continuity of F.

    A relevant consequence of this result is that, under suitable conditions on the 0-order terms and divergence terms with the solution itself inside (the conditions stated in Theorem 4.4), we can treat Vw and div(ρv2Fw) respectively as a fixed forcing term and a divergence term with a given field. As a consequence, we obtain uniform local regularity estimates in Theorem 4.4 for solutions wε to (2.7) by simply applying Theorems 4.2 and 4.3.

    We would like to show an example of a set of hypothesis for which part 2) of our main Theorem 4.4 holds true; that is, local uniform C1,α estimates for the ratio of odd solutions and the fundamental ones.

    We remark that, as a, the decay of the data on Σ becomes stronger and stronger.

    Assumption 4.6 (C1,α). Let fε:=ymax{1,1a}gε with gε uniformly bounded in Lp(B+1,ωaε(y)dz) as ε0 and

    p>n+3+(a)+.

    Let Fε:=ymax{1,1a}Gε with Gε uniformly bounded in C0,α(B+1) as ε0 and

    α>11+min{2,2a}n+3+(a)+.

    Nevertheless, the matrix A, which satisfies Assumption (HA+), must also satisfy some regularity assumptions: AC0,α(B+1) with xμC0,α(B+1) and T=y˜T with ˜TC0,α(B+1).

    Moreover, when the weight is locally integrable; that is, a(1,1), we obtain local estimates for odd solutions working directly on the equation.

    Theorem 4.7. Let a(1,1) and as ε0 let {uε} be a family of solutions in B+1 of either

    div(ρaεAuε)=ρaεfε+div(ρaεFε) (4.11)

    satisfying the Dirichlet boundary condition

    uε=0on 0B+1.

    Let r(0,1), β>1, p1>n+1+a+2, p2>n+1+a+ and α(0,1)(0,1a)(0,2n+1+a+p1](0,1n+1+a+p2]. Let moreover A satisfy assumption (HA) with continuous coefficients. There are constants depending on a, n, β, p1, p2, α and r only such that

    uεC0,α(B+r)c(uεLβ(B+1,ρaε(y)dz)+fεLp1(B+1,ρaε(y)dz)+FεLp2(B+1,ρaε(y)dz)).

    Proof. The proof is obtained by contradiction following the very same passages of [21, Theorem 4.1], observing that in presence of the zero Dirichlet boundary condition at Σ we obtain a contradiction by applying the Liouville Theorem 3.1. The blow-up sequences invoked are centered in points ˆzkB+=B1+r2{y0}; that is,

    vk(z)=(ηuk)(ˆzk+rkz)(ηuk)(ˆzk)Lkrαk,wk(z)=η(ˆzk)(uk(ˆzk+rkz)uk(ˆzk))Lkrαk,

    with

    zB(k):=Bˆzkrk.

    Moreover, if yk/rk+ (where zk=(xk,yk)), then we choose ˆzk=zk, while if yk/rkc uniformly, then we choose ˆzk=(xk,0). In this second case, we remark that vk and wk are antisymmetric with respect to {y=0} so that the limit w will be odd in y .

    Work partially supported by the ERC Advanced Grant 2013 n. 339958 Complex Patterns for Strongly Interacting Dynamical Systems - COMPAT. Y. S. was partially supported by the Simons foundation.

    The authors declare no conflict of interest.

    In this appendix we are going to state and prove some technical results which will allow us to compare, from the regularity point of view, the variable coefficient case with the constant one.

    Remark A.1. Let a\in(-\infty, 1) , \varepsilon\geq0 . Then the family of functions

    \begin{equation} \psi_\varepsilon^a(y): = \frac{y\rho_\varepsilon^{-a}(y)}{\int_0^y\rho_\varepsilon^{-a}(s)\mathrm{d}s} \end{equation} (A.1)

    are monotone in y and uniformly bounded in L^{\infty}(B_1^+) by a constant which does not depend on \varepsilon . In fact, denoting t = y/\varepsilon , we have

    \begin{equation*} \psi_\varepsilon^a(y) = \psi_1^a\left(\frac{y}{\varepsilon}\right) = \psi_1^a\left(t\right) = \frac{t(1+t^2)^{-a/2}}{\int_0^t(1+s^2)^{-a/2}\mathrm{d}s}. \end{equation*}

    The latter function is continuous and monotone nondecreasing if a < 0 and nonincreasing if a\in(0, 1) . Since \psi_1^a has limit 1 as t\to0 and limit 1-a as t\to+\infty , then

    \sup\limits_{t \gt 0}\psi_1^a(t) = \max\{1, 1-a\}\qquad\mathrm{and}\qquad \inf\limits_{t \gt 0}\psi_1^a(t) = \min\{1, 1-a\}.

    Finally, note that the family \psi_\varepsilon^a can not be equicontinuous, nor uniformly bounded in C^{0, \alpha}(B_1^+) , while it enjoys the following property:

    \begin{equation} \exists c \gt 0\;:\; \forall \varepsilon\in[0, \varepsilon_0)\;, \Vert \psi_\varepsilon^a\Vert_{ {\rm Lip}(B_1\cap\{y \gt \sqrt{\varepsilon}\})} \lt c\;, \end{equation} (A.2)

    due to the fact that \psi_1^a is bounded, has a finite limit as t\to +\infty and its derivative vanishes as 1/t^2 .

    Lemma A.2. Let a\in(-\infty, 1) , \varepsilon\geq0 , \alpha\in(0, 1) and let g(x, y, s)\in C^{0, \alpha}_{x, y}(B_1^+) uniformly in s\in[0, y] , such that |g(x, y, s)|\leq c|y|^\alpha for (x, y)\in B_1^+ uniformly in s\in[0, y] . Then the family of functions

    \begin{equation*} \mathcal G_\varepsilon(x, y) = \frac{\int_0^y\rho_\varepsilon^{-a}(s)g(x, y, s)\mathrm{d}s}{\int_0^y\rho_\varepsilon^{-a}(s)\mathrm{d}s} \end{equation*}

    is uniformly bounded in C^{0, \alpha}(B_1^+) by a constant which does not depend on \varepsilon .

    Proof. We remark that the proof follows some ideas of the proof in [21, Lemma 7.5], where the case \varepsilon = 0 is done. The uniform Hölder continuity in the x -variable is trivial. Hence, fixed 0 < \delta < 1 , let us consider the following two sets

    I_1 = \{(y_1, y_2)\ : \ 0\leq y_1\leq y_2 \lt 1, \ y_2-y_1\geq\delta y_2\}

    and

    I_2 = \{(y_1, y_2)\ : \ 0\leq y_1\leq y_2 \lt 1, \ y_2-y_1 \lt \delta y_2\}.

    If we consider (y_1, y_2)\in I_1 , using that for i = 1, 2 , in the interval (0, y_i) it holds |g(x, y_i, s)|\leq cy_i^\alpha and thanks to the inequalities (y_2-y_1)^\alpha\geq\delta^\alpha y_2^\alpha\geq\delta^\alpha y_i^\alpha , then

    \begin{eqnarray*} \frac{|\mathcal G_\varepsilon(x, y_1)-\mathcal G_\varepsilon(x, y_2)|}{(y_2-y_1)^\alpha}&\leq&\frac{1}{(y_2-y_1)^\alpha}\sum\limits_{i = 1}^2|\mathcal G_\varepsilon(x, y_i)|\\ &\leq&\frac{c}{\delta^\alpha}\sum\limits_{i = 1}^2\frac{y_i^\alpha\int_0^{y_i}\rho_\varepsilon^{-a}(s)}{y_i^\alpha\int_0^{y_i}\rho_\varepsilon^{-a}(s)} = \frac{2c}{\delta^\alpha}. \end{eqnarray*}

    If we consider (y_1, y_2)\in I_2 , then

    \begin{eqnarray*} \frac{|\mathcal G_\varepsilon(x, y_1)-\mathcal G_\varepsilon(x, y_2)|}{(y_2-y_1)^\alpha}&\leq&\frac{1}{(y_2-y_1)^\alpha}\frac{\int_{0}^{y_2}\rho_\varepsilon^{-a}(s)|g(x, y_2, s)-g(x, y_1, s)|}{\int_0^{y_2}\rho_\varepsilon^{-a}(s)}\\ &&+\frac{1}{(y_2-y_1)^\alpha}\frac{\int_{y_1}^{y_2}\rho_\varepsilon^{-a}(s)|g(x, y_1, s)|}{\int_0^{y_2}\rho_\varepsilon^{-a}(s)}\\ &&+\frac{1}{(y_2-y_1)^\alpha}\frac{\left(\int_{0}^{y_1}\rho_\varepsilon^{-a}(s)|g(x, y_1, s)|\right)\left(\int_{y_1}^{y_2}\rho_\varepsilon^{-a}(s)\right)}{\left(\int_{0}^{y_1}\rho_\varepsilon^{-a}(s)\right)\left(\int_{0}^{y_2}\rho_\varepsilon^{-a}(s)\right)}\\ & = &J_1+J_2+J_3. \end{eqnarray*}

    Hence, J_1 can be bounded using the fact that |g(x, y_2, s)-g(x, y_1, s)|\leq c(y_2-y_1)^\alpha . Working on J_2 , there exists y_1\leq\xi\leq y_2 such that

    \begin{eqnarray*} J_2&\leq& c(y_2-y_1)^{1-\alpha}\frac{\rho_\varepsilon^{-a}(\xi)y_1^\alpha}{\int_0^{y_2}\rho_\varepsilon^{-a}(s)}\\ &\leq&c\left(\frac{y_2-y_1}{y_2}\right)^{1-\alpha}\frac{y_2\rho_\varepsilon^{-a}(\xi)}{\int_0^{y_2}\rho_\varepsilon^{-a}(s)}\\ &\leq&c\delta^{1-\alpha}\max\{1, (1-\delta)^{-a}\}\frac{y_2\rho_\varepsilon^{-a}(y_2)}{\int_0^{y_2}\rho_\varepsilon^{-a}(s)} \end{eqnarray*}

    using the fact that y_2-y_1 < \delta y_2 , the inequalities

    \begin{equation*} 1-\delta \lt \frac{y_1}{y_2}\leq\frac{\xi}{y_2}\leq 1, \end{equation*}

    and the fact that \rho_\varepsilon^{-a}(\xi)\leq \max\{1, (1-\delta)^{-a}\}\rho_\varepsilon^{-a}(y_2) (easy to check). Eventually, recalling y_2/\varepsilon = t\in[0, +\infty) , we have already remarked that the function defined in (1.1) is bounded uniformly in \varepsilon

    \frac{y_2\rho_\varepsilon^{-a}(y_2)}{\int_0^{y_2}\rho_\varepsilon^{-a}(s)} = \frac{t(1+t^2)^{-a/2}}{\int_0^{t}(1+s^2)^{-a/2}} = \psi(t)\leq\max\{1, 1-a\}.

    With analogous computations we can bound also J_3 .

    Proposition A.3. Let a\in(-\infty, 1) , \varepsilon\geq0 , \alpha\in(0, 1) and let \gamma\in C^{0, \alpha}(B_1^+) . Then the family of functions

    \begin{equation*} \mathcal G_\varepsilon(x, y) = \frac{\int_0^y\rho_\varepsilon^{-a}(s)\left(\gamma(x, s)-\gamma(x, 0)\right)\mathrm{d}s}{\int_0^y\rho_\varepsilon^{-a}(s)\mathrm{d}s} \end{equation*}

    is uniformly bounded in C^{0, \alpha}(B_1^+) by a constant which does not depend on \varepsilon .

    Proof Just notice that, since g(x, y, s): = \gamma(x, s)-\gamma(x, 0) for s\leq y , g satisfies conditions of the previous Lemma A.2. Indeed is \alpha -Hölder continuous in (x, y) uniformly in s\leq y and

    |g(x, y, s)| = |\gamma(x, s)-\gamma(x, 0)|\leq c|s|^\alpha\leq c|y|^\alpha.

    Proposition A.4. Let a\in(-\infty, 1) , \varepsilon\geq0 , \alpha\in(0, 1) and let \gamma\in C^{1, \alpha}(B_1^+) with \partial_y\gamma(x, 0)\in C^{1, \alpha}(B_1^+) . Consider the family of functions

    \begin{equation*} \mathcal G_\varepsilon(x, y) = \frac{\int_0^y\rho_\varepsilon^{-a}(s)\left(\gamma(x, s)-\gamma(x, 0)\right)\mathrm{d}s}{\int_0^y\rho_\varepsilon^{-a}(s)\mathrm{d}s}\;. \end{equation*}

    Then there exists c > 0 such that, for every \varepsilon\in[0, \varepsilon_0] , \mathcal G_\varepsilon is uniformly bounded in C^{1, \alpha}(B_1\cap\{y\geq \sqrt{\varepsilon}\}) by c .

    Proof. One can rewrite our function as

    \begin{eqnarray*} \mathcal G_\varepsilon(x, y)& = &\frac{\int_0^y\rho_\varepsilon^{-a}(s)\left(\gamma(x, s)-\gamma(x, 0)-\partial_y\gamma(x, 0)s\right)\mathrm{d}s}{\int_0^y\rho_\varepsilon^{-a}(s)\mathrm{d}s}+\partial_y\gamma(x, 0)\frac{\int_0^y\rho_\varepsilon^{-a}(s)s\, \mathrm{d}s}{\int_0^y\rho_\varepsilon^{-a}(s)\mathrm{d}s}\\ & = & \frac{\int_0^y\rho_\varepsilon^{-a}(s)\left(\int_0^s(\partial_y\gamma(x, \tau)-\partial_y\gamma(x, 0))\mathrm{d}\tau\right)\mathrm{d}s}{\int_0^y\rho_\varepsilon^{-a}(s)\mathrm{d}s}+\partial_y\gamma(x, 0)\frac{\int_0^y\rho_\varepsilon^{-a}(s)s\, \mathrm{d}s}{\int_0^y\rho_\varepsilon^{-a}(s)\mathrm{d}s}\\ & = & \mathcal{H}_\varepsilon(x, y)+\partial_y\gamma(x, 0)\frac{\int_0^y\rho_\varepsilon^{-a}(s)s\, \mathrm{d}s}{\int_0^y\rho_\varepsilon^{-a}(s)\mathrm{d}s}. \end{eqnarray*}

    First we show that the second term has the desired property uniformly in \varepsilon . At first we remark that \partial_y\gamma(x, 0)\in C^{1, \alpha}(B_1^+) . Now consider that the family of functions

    \begin{equation*} \xi_\varepsilon^a(y): = \frac{\int_0^y\rho_\varepsilon^{-a}(s)s\, \mathrm{d}s}{\int_0^y\rho_\varepsilon^{-a}(s)\mathrm{d}s} \end{equation*}

    is uniformly bounded in L^\infty(B_1^+) . In fact, denoting t = y/\varepsilon ,

    \begin{equation*} \xi_\varepsilon^a(y) = \varepsilon\xi_1^a(t) = \varepsilon\frac{\int_0^t(1+s^2)^{-a/2}s\, \mathrm{d}s}{\int_0^t(1+s^2)^{-a/2}\mathrm{d}s} = y \, \frac{\xi_1^a(t)}{t}, \end{equation*}

    is bounded in B_1^+ (uniformly with respect to \varepsilon\geq0 ). In fact, the first factor y is obviously bounded in [0, 1] and the second one is bounded for t\in[0, +\infty) . Now, let us consider the derivative in y ,

    \begin{equation*} \partial_y\xi_\varepsilon^a(y) = (\xi_1^a)'(t) = \psi_1^a(t)\left(1-\frac{\int_0^t(1+s^2)^{-a/2}s\, \mathrm{d}s}{t\int_0^t(1+s^2)^{-a/2}\mathrm{d}s}\right)\;. \end{equation*}

    We claim that \partial_y\xi_\varepsilon^a enjoys the property stated in (A.2), being the product of two functions, both bounded, having a finite limit as t\to +\infty and derivatives vanishing as 1/t^2 ..

    Eventually we consider \mathcal H_\varepsilon . Computing the gradient \nabla_x\mathcal H_\varepsilon , we obtain

    \begin{equation*} \nabla_x\mathcal H_\varepsilon(x, y) = \frac{\int_0^y\rho_\varepsilon^{-a}(s)\left(\tilde\gamma(x, s)-\tilde\gamma(x, 0)\right)\mathrm{d}s}{\int_0^y\rho_\varepsilon^{-a}(s)\mathrm{d}s} \end{equation*}

    where

    \tilde\gamma(x, s) = \nabla_x\gamma(x, s)-\nabla_x\partial_y\gamma(x, 0)s\in C^{0, \alpha}(B_1^+),

    and satisfies the assumptions in Proposition A.3.

    It remains to consider the partial derivative in y of \mathcal H_\varepsilon ; that is,

    \begin{eqnarray*} \partial_y\mathcal H_\varepsilon(x, y)& = &\frac{y\rho_\varepsilon^{-a}(y)}{\int_0^y\rho_\varepsilon^{-a}(s)\mathrm{d}s}\cdot\dfrac{\int_0^y\rho_\varepsilon^{-a}(s)\left(\frac{1}{y}\int_s^y(\partial_y\gamma(x, \tau)-\partial_y\gamma(x, 0))\mathrm{d}\tau\right)\mathrm{d}s}{\int_0^y\rho_\varepsilon^{-a}(s)\mathrm{d}s}\\ & = &\psi_\varepsilon^a(y)\cdot\mathcal I_\varepsilon(y). \end{eqnarray*}

    By Remark A.1, the family of functions \psi_\varepsilon^a enjoy the desired propery (A.2). Now we wish to conclude that \mathcal I_\varepsilon is uniformly bounded in C^{0, \alpha}(B_1^+) . To this aim, it is enough to prove that the function

    g(x, y, s) = \frac{1}{y}\int_s^y(\partial_y\gamma(x, \tau)-\partial_y\gamma(x, 0))\mathrm{d}\tau

    satisfies conditions in Lemma A.2. Using the Hölder continuity of \partial_y\gamma , obviously

    |g(x, y, s)|\leq\frac{1}{y}\int_s^y|\partial_y\gamma(x, \tau)-\partial_y\gamma(x, 0)|\mathrm{d}\tau\leq\frac{c|y|^\alpha(y-s)}{y}\leq c|y|^\alpha.

    The Hölder continuity of g in the x -variable is trivial. Nevertheless, following the reasonings in the proof of Lemma A.2, fixed 0 < \delta < 1 , let us consider the following two sets

    I_1 = \{(y_1, y_2)\ : \ 0\leq y_1\leq y_2 \lt 1, \ y_2-y_1\geq\delta y_2\}

    and

    I_2 = \{(y_1, y_2)\ : \ 0\leq y_1\leq y_2 \lt 1, \ y_2-y_1 \lt \delta y_2\}.

    If we consider (y_1, y_2)\in I_1 , using that for i = 1, 2 it holds |g(x, y_i, s)|\leq cy_i^\alpha and thanks to the inequalities (y_2-y_1)^\alpha\geq\delta^\alpha y_2^\alpha\geq\delta^\alpha y_i^\alpha , then

    \begin{eqnarray*} \frac{|g(x, y_1, s)-g(x, y_2, s)|}{(y_2-y_1)^\alpha}&\leq&\frac{1}{(y_2-y_1)^\alpha}\sum\limits_{i = 1}^2|g(x, y_i, s)|\\ &\leq&\frac{c}{\delta^\alpha}\sum\limits_{i = 1}^2\frac{y_i^\alpha}{y_i^\alpha} = \frac{2c}{\delta^\alpha}. \end{eqnarray*}

    If we consider (y_1, y_2)\in I_2 , then, using the fact that y_2-y_1 < \delta y_2

    \begin{eqnarray*} \frac{|g(x, y_1, s)-g(x, y_2, s)|}{(y_2-y_1)^\alpha}&\leq&\frac{1}{(y_2-y_1)^{\alpha}y_2}\int_{y_1}^{y_2}|\partial_y\gamma(x, \tau)-\partial_y\gamma(x, 0)|\mathrm{d}\tau\\ &&+\frac{1}{(y_2-y_1)^{\alpha}}\left|\frac{1}{y_2}-\frac{1}{y_1}\right|\int_{s}^{y_1}|\partial_y\gamma(x, \tau)-\partial_y\gamma(x, 0)|\mathrm{d}\tau\\ &\leq&c\delta^{1-\alpha}+c\frac{\delta^{1-\alpha}}{1-\delta}. \end{eqnarray*}

    In this appendix we are going to prove some useful inequalities, needed when working in weighted Sobolev spaces, specially whenever the weight does not belong to the A_2 class. These results will be the key of the validity of Liouville type theorems in Section 3.

    B.1. Hardy type inequalities

    At first, we deal with the validity of Hardy (trace) type inequalities and their spectral stability. These results will be the key tools in order to establish a class of Liouville theorems contained in this section. Let \mathbb{R}^{n+1}_+ = \mathbb{R}^{n+1}\cap\{y > 0\} , B_1^+ = B_1\cap\{y > 0\} and S^n_+ = S^{n}\cap\{y > 0\} . We define the space \tilde H^{1}(B_1^+) as the closure of C^\infty_c(\overline B_1^+\setminus\Sigma) with respect to the norm

    \left(\int_{B_1^+}|\nabla v|^2\right)^{1/2}.

    Then, we remark that the following trace Poincaré inequality holds

    \begin{equation} c\int_{S^{n}_+}v^2\leq\int_{B_1^+}|\nabla v|^2. \end{equation} (B.1)

    We first state the following Hardy inequality.

    Lemma B.1 (Hardy inequality). Let v\in \tilde H^{1}(B_1^+) . Then

    \begin{equation} \frac{1}{4}\int_{B_1^+}\frac{v^2}{y^2}\leq\int_{B_1^+}|\nabla v|^2. \end{equation} (B.2)

    Proof. The proof is an easy exercise based on the well known Hardy inequality on the half space

    \frac{1}{4}\int_{\mathbb{R}^{n+1}_+}\frac{v^2}{y^2}\leq\int_{\mathbb{R}^{n+1}_+}|\nabla v|^2,

    and using the Kelvin transform.

    Next, we will need a boundary version of the Hardy inequality

    Lemma B.2 (Boundary Hardy inequality). There exists c_0 > 0 such that, for every v\in\tilde H^1(B_1^+) , there holds

    \begin{equation} c_0\int_{S^n_+}\frac{v^2}{y}\leq\int_{B_1^+}|\nabla v|^2. \end{equation} (B.3)

    Proof. By taking the harmonic replacement of v on B_1^+ , we may assume without loss of generality that \(\Delta v = 0\) in B_1^+ . Now we consider the following inversion (stereographic projection) \Phi:B_1^+\subset\mathbb{R}^{n+1}\to\mathbb{R}^{n+1} such that

    \Phi:z = (x, y) = (x_1, ..., x_{n}, y)\mapsto \tilde z = (\tilde x, \tilde y) = (\tilde x_1, ..., \tilde x_{n}, \tilde y),

    with

    \Phi(z) = \frac{z+e_1}{|z+e_1|^2}-\frac{e_1}{2}\qquad\mathrm{and}\qquad \Phi^{-1}(\tilde z) = \frac{\tilde z+\frac{e_1}{2}}{|z+\frac{e_1}{2}|^2}-e_1.

    This map is conformal and such that \Phi(B_1^+) = \{\tilde x_1 > 0\}\cap\{\tilde y > 0\} and \Phi(S^n_+) = \{\tilde x_1 = 0\}\cap\{\tilde y > 0\} . Hence, the Kelvin transform

    \begin{equation*} w(\tilde z) = Kv(\tilde z): = \frac{1}{|\tilde z+\frac{e_1}{2}|^{n-1}}v(\Phi^{-1}(\tilde z)) \end{equation*}

    is harmonic in \{\tilde x_1 > 0\}\cap\{\tilde y > 0\} and such that

    \int_{B_1^+}|\nabla v|^2\mathrm{d}z = \int_{\{\tilde x_1 \gt 0\}\cap\{\tilde y \gt 0\}}|\nabla w|^2\mathrm{d}\tilde z.

    Using a fractional Hardy inequality (see [2]) on the n -dimensional half space \{\tilde x_1 = 0\}\cap\{\tilde y > 0\} , up to extending the function w = 0 in \{\tilde x_1 = 0\}\cap\{\tilde y < 0\} , we have

    \begin{eqnarray*} \int_{\{\tilde x_1 \gt 0\}\cap\{\tilde y \gt 0\}}|\nabla w|^2\mathrm{d}\tilde z &\geq& c\iint_{(\{\tilde x_1 = 0\}\cap\{\tilde y \gt 0\})^2}\frac{|w(\tilde \zeta_1)-w(\tilde \zeta_2)|^2}{|\tilde \zeta_1-\tilde \zeta_2|^{n+1}}\mathrm{d}\tilde \zeta_1\mathrm{d}\tilde \zeta_2\nonumber\\ &\geq& c\int_{\{\tilde x_1 = 0\}\cap\{\tilde y \gt 0\}}\frac{w^2(\tilde z)}{\tilde y}\mathrm{d}\tilde z. \end{eqnarray*}

    Finally we compute

    \begin{eqnarray*} &&\int_{S^n_+}\frac{v^2(z)}{y}\mathrm{d}\sigma(z) \\ & = &\int_{\{\tilde x_1 = 0\}\cap\{\tilde y \gt 0\}}\frac{w^2(\tilde z)}{\tilde y}\left|\tilde z+\frac{e_1}{2}\right|^{2(n-1)+2} \cdot|\Phi^{-1}_{\tilde x_2}(\tilde z)\wedge\Phi^{-1}_{\tilde x_3}(\tilde z)\wedge...\wedge\Phi^{-1}_{\tilde x_{n}}(\tilde z)\wedge\Phi^{-1}_{\tilde y}(\tilde z)|\mathrm{d}\tilde z\nonumber\\ &\leq&\int_{\{\tilde x_1 = 0\}\cap\{\tilde y \gt 0\}}\frac{w^2(\tilde z)}{\tilde y}\mathrm{d}\tilde z. \end{eqnarray*}

    B.2. A stability result

    The following Lemma is a stability result for quadratic forms which is usueful for our Liouville theorems.

    Lemma B.3. Let \{Q_k\}_{k\in\mathbb{N}} be a family of quadratic forms Q_k:\tilde H^1(B_1^+)\to[0, +\infty) defined by

    \begin{equation*} Q_k(v) = \int_{B_1^+}|\nabla v|^2+\int_{B_1^+}V_kv^2+\int_{S^n_+}W_kv^2. \end{equation*}

    Assume that the family \{Q_k\} satisfies the following conditions:

    i) |W_k|\leq c on S^n_+ and |V_k|\leq\frac{c}{y^2} in B_1^+ uniformly on k\in\mathbb{N} ;

    ii) there exists a constant C > 0 which does not depend on k\in\mathbb{N} such that for any v\in\tilde H^1(B_1^+)

    \begin{equation} \frac{1}{C}\|v\|_{\tilde H^1(B_1^+)}^2\leq Q_k(v)\leq C\|v\|_{\tilde H^1(B_1^+)}^2; \end{equation} (B.4)

    iii) V_k\to V in B_1^+ and W_k\to W on S^n_+ pointwisely as k\to+\infty , where

    Q(v) = \int_{B_1^+}|\nabla v|^2+\int_{B_1^+}Vv^2+\int_{S^n_+}Wv^2,

    with Q:\tilde H^1(B_1^+)\to[0, +\infty) satisfying |W|\leq c on S^n_+ , |V|\leq\frac{c}{y^2} in B_1^+ and

    \frac{1}{C}\|v\|_{\tilde H^1(B_1^+)}^2\leq Q(v)\leq C\|v\|_{\tilde H^1(B_1^+)}^2.

    Let

    \lambda_k = \min\limits_{v\in\tilde H^1(B_1^+)\setminus\{0\}}\frac{Q_k(v)}{\int_{S^n_+}v^2}, \qquad\lambda = \min\limits_{v\in\tilde H^1(B_1^+)\setminus\{0\}}\frac{Q(v)}{\int_{S^n_+}v^2}.

    Then, \lambda_k\to\lambda .

    Proof. Let \{v_k\}\subset \tilde H^1(B_1^+)\setminus\{0\} be a sequence of minimizers for \lambda_k ; that is, such that

    \lambda_k = Q_{k}(v_k) = \int_{B_1^+}|\nabla v_k|^2+\int_{B_1^+}V_kv_k^2+\int_{S^n_+}W_kv_k^2,

    and \int_{S^n_+}v^2_k = 1 . Since by compact embedding \tilde H^1(B_1^+)\hookrightarrow L^2(S^n_+) the minimum

    \min\limits_{v\in\tilde H^1(B_1^+)\setminus\{0\}}\frac{\|v\|^2_{\tilde H^1(B_1^+)}}{\int_{S^n_+}v^2} = \frac{\|u\|^2_{\tilde H^1(B_1^+)}}{\int_{S^n_+}u^2} = \nu \gt 0

    is achieved by u\in\tilde H^1(B_1^+)\setminus\{0\} and it is strictly positive by the trace Poincaré inequality, then there exists a positive constant C independent from k such that

    \frac{\nu}{C}\leq\lambda_k\leq C\nu.

    Moreover, we have that

    \frac{1}{C}\|v_k\|^2_{\tilde H^1(B_1^+)}\leq\lambda_k\leq C\nu

    and so there exists \overline v\in\tilde H^1(B_1^+) such that v_k\rightharpoonup\overline v in \tilde H^1(B_1^+) and, up to passing to a subsequence, v_k\to\overline v in L^2(S^n_+) . Moreover, the limit is non trivial by the condition \int_{S^n_+}\overline v^2 = 1 .

    We want to prove that the convergence is strong in \tilde H^1(B_1^+) . Testing the eigenvalue equation solved by v_k with v_k-\overline v , we have

    \int_{B_1^+}\nabla v_k\cdot\nabla(v_k-\overline v)+\int_{B_1^+}V_kv_k(v_k-\overline v)+\int_{S^n_+}W_kv_k(v_k-\overline v) = \lambda_k\int_{S^n_+}v_k(v_k-\overline v).

    Using the fact that |W_k|, |\lambda_k|\leq c uniformly in k , the strong convergence and the normalization in L^2(S^n_+) , by the Hölder inequality the terms over the half sphere S^n_+ go to 0 in the limit. So

    \begin{equation} \int_{B_1^+}\nabla v_k\cdot\nabla(v_k-\overline v)+\int_{B_1^+}V_kv_k(v_k-\overline v)\to0. \end{equation} (B.5)

    Hence,

    \begin{eqnarray} Q_{k}(v_k-\overline v)& = &\int_{B_1^+}|\nabla (v_k-\overline v)|^2+\int_{B_1^+}V_k(v_k-\overline v)^2+\int_{S^n_+}W_k(v_k-\overline v)^2\\ & = &\int_{B_1^+}\nabla v_k\cdot\nabla(v_k-\overline v)+\int_{B_1^+}V_kv_k(v_k-\overline v)-\int_{B_1^+}\nabla \overline v\cdot\nabla(v_k-\overline v)\\ &&-\int_{B_1^+}V\overline v(v_k-\overline v)+\int_{B_1^+}(V-V_k)\overline v(v_k-\overline v)+\int_{S^n_+}W_k(v_k-\overline v)^2\to0. \end{eqnarray} (B.6)

    In fact, the sum of the first two terms goes to 0 by (B.5), the sum of the second two by weak convergence in \tilde H^1(B_1^+) . The third term is such that

    \begin{eqnarray*} \int_{B_1^+}(V-V_k)\overline v(v_k-\overline v)&\leq&\left(\int_{B_1^+}(V-V_k)\overline v^2\right)^{1/2}\left(\int_{B_1^+}(V-V_k)(v_k-\overline v)^2\right)^{1/2}\\ &\leq& c\left(\int_{B_1^+}(V-V_k)\overline v^2\right)^{1/2}\to0. \end{eqnarray*}

    We used that V_k\to V , the fact that |V_k-V|\leq\frac{c}{y^2} and the Hardy inequality to ensure the dominated convergence theorem. Eventually the last term in (B.6) goes to 0 by the strong convergence in L^2(S^n_+) . Hence we obtain the strong convergence by (B.4).

    It is easy to see that Q_{k}(v_k)\to Q(\overline v) . This is enough to conclude because if we consider \tilde v the normalized in L^2(S^n_+) minimizer of \lambda , since it is competitor for the minimization of any Q_{k} , then

    \lambda_k = Q_{k}(v_k)\leq Q_{k}(\tilde v),

    and since Q_{k}(v_k)\to Q(\overline v) and Q_{k}(\tilde v)\to Q(\tilde v) , then by Q(\overline v)\leq Q(\tilde v) , and by the minimality of \overline v , we finally obtain that \overline v = \tilde v with \lambda_k\to\lambda .

    B.3. Quadratic forms for the odd case

    Let a\in(-\infty, 1) , \varepsilon\geq0 and consider a function u\in C^\infty_c(\overline B_1^+\setminus\Sigma) and define v = (\rho_\varepsilon^a)^{1/2}u\in C^\infty_c(\overline B_1^+\setminus\Sigma) . Let us define the quadratic form

    \begin{equation} \int_{B_1^+}\rho_\varepsilon^au^2 = Q_{\rho_\varepsilon^a}(v) = \int_{B_1^+}|\nabla v|^2+\int_{B_1^+}V_{\rho_\varepsilon^a}v^2+\int_{S^n_+}W_{\rho_\varepsilon^a}v^2, \end{equation} (B.7)

    where

    V_{\rho_\varepsilon^a}(y) = \frac{(\rho_\varepsilon^a)''}{2\rho_\varepsilon^a}-\left(\frac{(\rho_\varepsilon^a)'}{2\rho_\varepsilon^a}\right)^2 = \frac{a[(a-2)y^2+2\varepsilon^2]}{4(\varepsilon^2+y^2)^2}

    and

    W_{\rho_\varepsilon^a}(y) = -\frac{(\rho_\varepsilon^a)'y}{2\rho_\varepsilon^a} = -\frac{ay^2}{2(\varepsilon^2+y^2)}.

    Let

    \begin{equation*} Q_{a}(v) = \int_{B_1^+}|\nabla v|^2+\int_{B_1^+}V_{a}v^2+\int_{S^n_+}W_{a}v^2, \end{equation*}

    with V_a(y) = \frac{a(a-2)}{4y^2} = V_{\rho_0^a}(y) and W_{a}(y) = -\frac{a}{2} = W_{\rho_0^a}(y) . Eventually consider a sequence \varepsilon_k\to0 as k\to+\infty and define \rho_k = \rho_{\varepsilon_k}^a . Let us recall Q_k = Q_{\rho_k} and Q = Q_a .

    Lemma B.4. Under the previous hypothesis, the family \{Q_{k}\} = \{Q_{\rho_{\varepsilon_k}}\} defined in (B.7) and its limit Q satisfy the conditions in Lemma B.3.

    Proof. Condition i) is trivially satisfied. Moreover, combining i) , the trace Poincaré and the Hardy inequalities, we easily obtain the upper bound in ii) for any k\in\mathbb{N} with a constant independent on \varepsilon_k ; that is,

    Q_k(v)\leq c\|v\|_{\tilde H^1(B_1^+)}^2.

    Let us consider Q = Q_a and let us define u = y^{-a/2}v\in C^\infty_c(\overline B_1^+\setminus\Sigma) .

    \begin{eqnarray} Q_a(v)& = &\int_{B_1^+}|\nabla v|^2+\left(\frac{a^2}{4}-\frac{a}{2}\right)\int_{B_1^+}\frac{v^2}{y^2}-\frac{a}{2}\int_{S^n_+}v^2\\ & = &\int_{B_1^+}|\nabla v|^2+\left(\frac{a^2}{4}-\frac{a}{2}\right)\int_{B_1^+}\frac{v^2}{y^2}-\frac{a}{2}\int_{B_1^+}\mathrm{div}\left(\frac{v^2}{y}\vec{e_n}\right) = \int_{B_1^+}y^a|\nabla u|^2. \end{eqnarray} (B.8)

    First of all we notice that if a\leq 0 the lower bound follows trivially. So we can suppose that a\in(0, 1) . Since for a\neq1 , (\frac{a^2}{4}-\frac{a}{2}) > -\frac{1}{4} , hence by the Hardy inequality in (B.2), the quantity

    G_a(v) = \int_{B_1^+}|\nabla v|^2+\left(\frac{a^2}{4}-\frac{a}{2}\right)\int_{B_1^+}\frac{v^2}{y^2}

    defines an equivalent norm in \tilde H^1(B_1^+) . Hence by the compact embedding \tilde H^1(B_1^+)\hookrightarrow L^2(S^n_+) we have that the minimum in

    \xi(a) = \min\limits_{v\in\tilde H^1(B_1^+)\setminus\{0\}}\frac{G_a(v)}{\int_{S^n_+}v^2}

    is achieved. In fact, considering a minimizing sequence, we can take it such that \int_{S^n_+}v_k^2 = 1 and also such that v_k\in C^\infty_c(\overline B_1^+\setminus\Sigma) . So it is uniformly bounded in \tilde H^1(B_1^+) and v_k\rightharpoonup\overline v\in\tilde H^1(B_1^+) with G_a(v_k)\to\xi(a) . Moreover the convergence is strong in L^2(S^n_+) by compact embedding. Since \int_{S^n_+}v_k^2 = 1 , we also obtain convergence of the \tilde H^1_0 -norms of the v_k to that of the limit, yielding strong convergence in \tilde H^1(B_1^+) . In fact, by the lower semicontinuity of the norm

    \xi(a)\leq\frac{G_a(\overline v)}{\int_{S^n_+}\overline v^2}\leq\liminf\limits_{k\to+\infty}\frac{G_a(v_k)}{\int_{S^n_+}v_k^2} = \xi(a).

    Obviously by the condition \int_{S^n_+}\overline v^2 = 1 the limit \overline v is not trivial. This proves that \overline v achieves the minimum. Moreover, defining

    \begin{equation} \lambda(a) = \min\limits_{v\in\tilde H^1(B_1^+)\setminus\{0\}}\frac{Q_a(v)}{\int_{S^n_+}v^2} = \xi(a)-\frac{a}{2}\geq0, \end{equation} (B.9)

    we want to prove that actually \lambda(a) > 0 . First of all, such a minimum is nonnegative since the minimizing sequence can be taken in C^\infty_c(\overline B_1^+\setminus\Sigma) and so the equalities in (B.8) give this condition. By contradiction let \lambda(a) = 0 . Hence the minimizing sequence is such that Q_a(v_k)\to0 . Defining u_k = y^{-a/2}v_k , one has \int_{B_1^+}y^a|\nabla u_k|^2\to0 . Moreover, the strong convergence in \tilde H^1(B_1^+) gives the almost everywhere convergence of \nabla v_k\to\nabla\overline v which of course implies that \nabla u_k\to \nabla(y^{-a/2}\overline v) almost everywhere in B_1^+ . Hence, since \nabla(y^{-a/2}\overline v) = 0 almost everywhere, \overline v = cy^{a/2} , but \nabla\overline v does not belong to L^2(B_1^+) . This is a contradiction. So \lambda(a) > 0 . So we have the inequality

    Q_a(v)\geq\lambda(a)\int_{S^n_+}v^2,

    which says that

    Q_a(v)\geq\frac{\lambda(a)}{\frac{a}{2}+\lambda(a)}\left(\int_{B_1^+}|\nabla v|^2+\left(\frac{a^2}{4}-\frac{a}{2}\right)\int_{B_1^+}\frac{v^2}{y^2}\right),

    and by the equivalence of the norms we obtain the result for a constant which depends on a and \lambda(a) . Eventually, we have proved that also Q_a is an equivalent norm on \tilde H^1(B_1^+) .

    In order to find a lower bound for Q_k which is uniform in k , it is enough to remark that if a\geq0 , then Q_k\geq Q_a . If a < 0 , then one can check that

    Q_k(v)\geq \int_{B_1^+}|\nabla v|^2-\int_{B_1^+}\frac{a}{4(a-4)}\frac{v^2}{y^2},

    with \frac{a}{4(a-4)} < \frac{1}{4} and hence by the Hardy inequality in (B.2) we have also in this case an equivalent norm.

    Let us recall the definition of \tilde H^1(B_1^+, \rho_\varepsilon^a(y)\mathrm{d}z) as the closure of C^\infty_c(\overline B_1^+\setminus\Sigma) with respect to the norm

    \int_{B_1^+}\rho_\varepsilon^a|\nabla u|^2.

    Lemma B.5. Let a\in(-\infty, 1) , \varepsilon\geq0 and u\in\tilde H^1(B_1^+, \rho_\varepsilon^a(y)\mathrm{d}z) . Then the following inequalities hold true for a positive constant c independent of \varepsilon\in[0, 1]

    \begin{equation} c\int_{B_1^+}\rho_\varepsilon^au^2\leq\int_{B_1^+}\rho_\varepsilon^a|\nabla u|^2, \end{equation} (B.10)
    \begin{equation} c\int_{S^n_+}\rho^a_{\varepsilon}u^2\leq\int_{B_1^+}\rho_\varepsilon^a|\nabla u|^2, \end{equation} (B.11)
    \begin{equation} c\int_{B_1^+}\frac{\rho_\varepsilon^a}{y^2}u^2\leq\int_{B_1^+}\rho_\varepsilon^a|\nabla u|^2, \end{equation} (B.12)
    \begin{equation} c\int_{S^n_+}\frac{\rho_\varepsilon^a}{y}u^2\leq\int_{B_1^+}\rho_\varepsilon^a|\nabla u|^2, \end{equation} (B.13)
    \begin{equation} \left(\int_{B_1^+}(\rho_\varepsilon^a)^{2^*/2}|u|^{2^*}\right)^{2/2^*}\leq c\int_{B_1^+}\rho_\varepsilon^a|\nabla u|^2, \end{equation} (B.14)

    which are respectively the Poincaré inequality, the trace Poincaré inequality, the Hardy inequality, the trace Hardy inequality and a Sobolev type inequality.

    Proof The proof is performed for functions u\in C^\infty_c(\overline B_1^+\setminus\Sigma) and then extending the inqualities to u\in\tilde H^1(B_1^+, \rho_\varepsilon^a(y)\mathrm{d}z) by a density argument. By Lemma B.4 there exists a positive constant uniform in \varepsilon such that

    \begin{equation} \int_{B_1^+}\rho_\varepsilon^a|\nabla u|^2 = Q_{\rho_\varepsilon^a}((\rho_\varepsilon^a)^{1/2}u)\geq c\int_{B_1^+}|\nabla((\rho_\varepsilon^a)^{1/2}u)|^2, \end{equation} (B.15)

    then all the inequalities are obtained by the validity of them in \tilde H^1(B_1^+) .

    B.4. Quadratic forms for the auxiliary weights

    Consider now a\in(-\infty, 1) and define

    \pi_\varepsilon^a(y) = \left((1-a)\int_0^y\rho_\varepsilon^{-a}(s)\mathrm{d}s\right)^2,

    and

    \begin{equation*} \omega_\varepsilon^a(y) = \rho_\varepsilon^a(y)\pi_\varepsilon^a(y). \end{equation*}

    We observe that this weight is super degenerate; that is, at \Sigma

    \omega_\varepsilon^a(y)\sim\begin{cases} |y|^{2-a} & \mathrm{if \ }\varepsilon = 0\\ |y|^2 & \mathrm{if \ }\varepsilon \gt 0, \end{cases}

    with 2-a\in(1, +\infty).

    B.4.1. Super singular weights (\omega_\varepsilon^a)^{-1}

    Let us consider u\in C^\infty_c(\overline B_1^+\setminus\Sigma) and define v = (\omega_\varepsilon^a)^{-1/2}u\in C^\infty_c(\overline B_1^+\setminus\Sigma) . Then we consider the quadratic form

    \begin{equation} \int_{B_1^+}(\omega_\varepsilon^a)^{-1}|\nabla u|^2 = Q_{\omega_\varepsilon^a}(v) = \int_{B_1^+}|\nabla v|^2+\int_{B_1^+}V_{\omega_\varepsilon^a} v^2+\int_{S^n_+}W_{\omega_\varepsilon^a} v^2, \end{equation} (B.16)

    with

    V_{\omega_\varepsilon^a} = \frac{1}{4}[(\log\omega_\varepsilon^a)']^2-\frac{1}{2}(\log\omega_\varepsilon^a)'',

    and

    W_{\omega_\varepsilon^a} = \frac{1}{2}(\log\omega_\varepsilon^a)'y.

    Hence

    V_{\omega_0^a}(y) = \frac{(2-a)(4-a)}{4y^2}, \qquad\mathrm{and}\qquad W_{\omega_0^a}(y) = \frac{2-a}{2}.

    Eventually consider a sequence \varepsilon_k\to0 as k\to+\infty and define \omega_k = \omega_{\varepsilon_k}^a . Let us name Q_k = Q_{\omega_k} and Q = Q_{\omega_0^a} . In what follows it would be useful to consider for t > 0 , the continuous function defined in (A.1); that is,

    \begin{equation*} \psi(t) = \frac{t(1+t^2)^{-a/2}}{\int_0^{t}(1+s^2)^{-a/2}\mathrm{d}s}, \end{equation*}

    which is monotone nondecreasing if a < 0 and nonincreasing if a\in(0, 1) . Since \psi has limit 1 as t\to0 and limit 1-a as t\to+\infty , then

    \sup\limits_{t \gt 0}\psi(t) = \max\{1, 1-a\}\qquad\mathrm{and}\qquad \inf\limits_{t \gt 0}\psi(t) = \min\{1, 1-a\}.

    Let us finally define for any k\in\mathbb{N}

    \begin{equation} \tilde Q_k(v) = Q_{k}(v)+\left(-\frac{a}{2}\right)^+\int_{S^n_+}v^2. \end{equation} (B.17)

    First we need the following technical result.

    Lemma B.6. Let us define for a\in(-\infty, 1) and t\in[0, +\infty) the function

    \begin{equation} \Phi_a(t) = \left[\frac{\sqrt{2}t(1+t^2)^{-a/2}}{\int_0^t(1+s^2)^{-a/2}}+\frac{at^2}{\sqrt{2}(1+t^2)}\right]^2+\frac{at^2[(2-a)t^2-2]}{4(1+t^2)^2}. \end{equation} (B.18)

    Hence there exists a positive constant c_1(a) > -\frac{1}{4} such that

    \begin{equation} \inf\limits_{t \gt 0}\Phi_a(t) = c_1(a). \end{equation}

    Proof Step 1: a\in(-3, 1) .

    Whenever 0\leq a < 1 , there holds

    \min\limits_{t \gt 0}f_a(t) = \min\limits_{t \gt 0}\frac{at^2[(2-a)t^2-2]}{4(1+t^2)^2} = f_a\left(\frac{1}{\sqrt{3-a}}\right) = \frac{a}{4(a-4)} \gt -\frac{1}{4}.

    Moreover, if a < 0 ,

    \inf\limits_{t \gt 0}f_a(t) = \lim\limits_{t\to+\infty}f_a(t) = \frac{a(2-a)}{4}.

    Hence, whenever 1-\sqrt{2} < a < 0 , then, the infimum remains strictly larger that -1/4 .

    Moreover, for a < 0 , then f_a(t)\geq0 in \left[0, \sqrt{\frac{2}{2-a}} \ \right] . From now on we will consider a < 0 and t > \sqrt{\frac{2}{2-a}} . Now, let us compute the square in (B.18), and add 1/4 ; that is

    \begin{eqnarray*} \Phi_a(t)+\frac{1}{4}& = &\frac{2t^2(1+t^2)^{-a}}{\left(\int_0^t(1+s^2)^{-a/2}\right)^2}+\frac{2at^3(1+t^2)^{-1-a/2}}{\int_0^t(1+s^2)^{-a/2}}+\frac{a^2t^4}{2(1+t^2)^2}+f_a(t)+\frac{1}{4}\\ & = &\frac{2t^3(1+t^2)^{-1-a/2}}{\left(\int_0^t(1+s^2)^{-a/2}\right)^2} \ \cdot g_a(t)+\frac{t^4(a^2+2a+1)+t^2(-2a+2)+1}{4(1+t^2)^2}\\ & = &I_a(t)+J_a(t), \end{eqnarray*}

    with

    g_a(t) = \left(\frac{(1+t^2)^{1-a/2}}{t}+a\int_0^t(1+s^2)^{-a/2}\right).

    It is easy to see that

    \inf\limits_{t \gt 0}J_a(t)\begin{cases} \gt 0 & \mathrm{if \ } a\neq-1\\ = 0 & \mathrm{if \ }a = -1.\end{cases}

    Nevertheless, since

    g_a'(t) = \frac{(1+t^2)^{-a/2}}{t^2}(t^2-1),

    then g_a has its global minimum in t = 1 , and hence it is easy to see that

    g_a(1) = 2^{1-a/2}+a\int_0^1(1+s^2)^{-a/2}\geq 2^{1-a/2}+a\int_0^1(1+s)^{-a/2} = 2^{1-a/2}\frac{2+a}{2-a}-\frac{2a}{2-a} \gt 0,

    surely if a > -3 . Hence, when a\in(-3, -1)\cup(-1, 0) , we have the result since \inf_{t > 0}I_a(t)\geq0 and \inf_{t > 0}J_a(t) > 0 . In the case a = -1 one can see that

    \inf\limits_{t \gt 0}I_{-1}(t) = \min\limits_{t \gt 0}I_{-1}(t) \gt 0,

    using the explicit form

    I_{-1}(t) = \frac{2t^3(1+t^2)^{-1-a/2}}{\frac{1}{4}\left(t\sqrt{t^2+1}+\log(\sqrt{t^2+1}+t)\right)^2}\left(\frac{(1+t^2)^{1-a/2}}{t}-\frac{1}{2}\left(t\sqrt{t^2+1}+\log(\sqrt{t^2+1}+t)\right)\right).

    Step 2: a\leq-3 .

    We can express

    \Phi_a(t)+\frac{1}{4} = \frac{t^4}{(1+t^2)^2}\left(2\left(\frac{(1+t^2)^{-\frac{a}{2}+1}}{t\int_0^t(1+s^2)^{-\frac{a}{2}}}+\frac{a}{2}\right)^2+\frac{a(2-a)}{4}-\frac{a}{2t^2}+\frac{1}{4}\frac{(1+t^2)^2}{t^4}\right).

    Hence

    \tilde\Phi_a(t) = \frac{(1+t^2)^2}{t^4}\left(\Phi_a(t)+\frac{1}{4}\right),

    and \gamma_a(t) = \tilde\Phi_a(t/\sqrt{-a})-\frac{0.001}{4}\frac{(-a+t^2)^2}{t^4} ; that is,

    \begin{equation} \gamma_a(t) = 2a^2\left(\frac{(1+\frac{t^2}{-a})^{-\frac{a}{2}+1}}{t\int_0^t(1+\frac{s^2}{-a})^{-\frac{a}{2}}}-\frac{1}{2}\right)^2+\frac{a(2-a)}{4}+\frac{a^2}{2t^2}+\frac{0.999}{4}\frac{(-a+t^2)^2}{t^4}. \end{equation} (B.20)

    First need to highlight some fundamental properties of the functions

    \begin{equation*} w_a(t) = \frac{(1+\frac{t^2}{-a})^{-\frac{a}{2}+1}}{t\int_0^t(1+\frac{s^2}{-a})^{-\frac{a}{2}}}. \end{equation*}

    As a\to-\infty one has the pointwise convergence w_a(t)\to v(t) in (0, +\infty) (which is however uniform on compact subsets) with

    \begin{equation*} v(t) = \frac{e^{\frac{t^2}{2}}}{t\int_0^te^{\frac{s^2}{2}}}. \end{equation*}

    We wish to prove the following

    Claim: w_a/v\geq 1 in [0, +\infty) . At first, elementary computations show that, in a neighbourhood of t = 0 , the expansion

    w_a(t) = \frac{1}{t^2}+\frac{1}{2}+\frac{1}{-a}+o(1)\qquad\mathrm{and}\qquad v(t) = \frac{1}{t^2}+\frac{1}{2}+o(1),

    holds, while in a neighbourhood of t = +\infty we have

    w_a(t) = \frac{1-a}{-a}+o(1)\qquad\mathrm{and}\qquad v(t) = 1+o(1),

    implying that w_a/v > 1 near zero and at infinity. Thus, the claim is false if and only if there exists t_0 > 0 such that

    \begin{equation} \begin{cases} w_a(t_0) = v_a(t_0)\\ \left(\frac{w_a}{v}\right)'(t_0)\leq0. \end{cases} \end{equation} (B.21)

    Remark that, w_a and v solve respectively the following differential equations

    \begin{equation*} w_a'(t) = \frac{1}{t(1+\frac{t^2}{-a})}\left(\frac{1-a}{-a}t^2-1\right)w_a(t)-\frac{t}{1+\frac{t^2}{-a}}w_a^2(t) \end{equation*}

    and

    \begin{equation*} v'(t) = \frac{t^2-1}{t}v(t)-tv^2(t). \end{equation*}

    Using these equations we obtain

    \begin{equation*} \left(\frac{w_a}{v}\right)' = \frac{w_a}{v}\left(\frac{t}{-a+t^2}(2-t^2)-\frac{t}{1+\frac{t^2}{-a}}w_a+tv\right), \end{equation*}

    and (B.21) holds if and only if

    \begin{equation*} v(t_0)\leq 1-\frac{2}{t_0^2}. \end{equation*}

    Now we are going to show that, on the contrary,

    \begin{equation} v(t) \gt z(t): = 1-\frac{2}{t^2}. \end{equation} (B.22)

    In (0, \sqrt 2) we have v > 0 and z < 0 . Moreove the inequality (B.22) can be checked numerically (with error estimate) on [\sqrt 2, \sqrt 6] , and is also valid in a neighbourhood of t = +\infty , by the exapnsion

    v(t) = 1-\frac{1}{t^2}+o\left(\frac{1}{t^2}\right) \gt z(t).

    This was proved with the aid of the computing system Mathematica, by numerical computations with error estimates. So, the function v-z is positive near 0 and at +\infty , and hence denying (B.22) yields the existence of t_1\geq\sqrt 6 such that

    \begin{cases} v(t_1) = z(t_1)\\ (v-z)'(t_1)\leq0. \end{cases}

    It is easy to see that at such a point t_1 one has (v-z)'(t_1) > 0 if t_1\geq\sqrt 6 (using the fact that v(t_1) = z(t_1) ). Finally, we observe that the function v changes monotonicity only once on (0, +\infty) and its absolute minimum value 0, 77836\pm 10^{-5} is larger than 1/2 .

    Now we can turn back to (B.20), obtaining that

    \begin{equation} \gamma_a(t)\geq2a^2\left(v(t)-\frac{1}{2}\right)^2+\frac{a(2-a)}{4}+\frac{a^2}{2t^2}+\frac{0.999}{4}\frac{(-a+t^2)^2}{t^4}. \end{equation} (B.23)

    In order to complete the proof, we need to prove positivity of the right hand side. To this end we take advantage once more of numerical computations with error estimates. At first, as v(5.1) = 0.95774\pm10^{-5} and v'(5.1) = 0.001860\pm 10^{-5} > 0 , since v changes monotonicity only once, we infer positivity of the right hand side for all t\in [5.1, +\infty) , for all a\leq -2.96767 . The remaining values (a, t) lay in the compact rectangle [-43.3272, -2.96767]\times[1, 5.1] and can be easily dealt numerically with error controlled minimization.

    Lemma B.7. Let a\in(-\infty, 1) . The family \{\tilde Q_{k}\} = \{\tilde Q_{\omega_{\varepsilon_k}}\} defined in (B.17) and its limit \tilde Q satisfy the conditions in Lemma B.3.

    Proof. First, we want to prove property i) ; that is, there exists a positive constant c > 0 uniform in \varepsilon\to0 such that

    |V_{\omega_\varepsilon^a}(y)|\leq\frac{c}{y^2}\qquad\mathrm{and}\qquad |W_{\omega_\varepsilon^a}(y)|\leq c.

    We remark that there exists a positive constant c > 0 uniform in \varepsilon\to0 such that

    |(\log\rho_\varepsilon^a)'| = \left|\frac{(\rho_\varepsilon^a)'}{\rho_\varepsilon^a}\right|\leq |a|\frac{y}{\varepsilon^2+y^2}\leq\frac{c}{y}.

    Moreover

    |(\log\rho_\varepsilon^a)''|\leq \left|\frac{(\rho_\varepsilon^a)'}{\rho_\varepsilon^a}\right|\cdot\left|\frac{(\rho_\varepsilon^a)''}{(\rho_\varepsilon^a)'}\right|+\left|\frac{(\rho_\varepsilon^a)'}{\rho_\varepsilon^a}\right|^2\leq \frac{c}{y^2}.

    It remains to prove the following uniform bounds

    \begin{equation*} \left|\frac{(\pi_\varepsilon^a)'}{\pi_\varepsilon^a}\right|\leq \frac{c}{y}, \qquad\mathrm{and}\qquad\left|\frac{(\pi_\varepsilon^a)''}{(\pi_\varepsilon^a)'}\right|\leq \frac{c}{y}. \end{equation*}

    Then the result follows since we are considering the logarithm of a product by linearity of the derivative.

    \begin{eqnarray*} |(\log\pi_\varepsilon^a)'| = \left|\frac{(\pi_\varepsilon^a)'}{\pi_\varepsilon^a}\right|& = &2\frac{\rho_\varepsilon^{-a}(y)}{\int_0^y\rho_\varepsilon^{-a}(s)\mathrm{d}s}\\ & = &\frac{2}{y}\frac{\frac{y}{\varepsilon}(1+\left(\frac{y}{\varepsilon}\right)^2)^{-a/2}}{\int_0^{\frac{y}{\varepsilon}}(1+s^2)^{-a/2}\mathrm{d}s}\\ &\leq&\frac{2}{y}\sup\limits_{t \gt 0}\frac{t(1+t^2)^{-a/2}}{\int_0^{t}(1+s^2)^{-a/2}\mathrm{d}s}\leq \frac{2\max\{1, 1-a\}}{y}. \end{eqnarray*}

    Moreover,

    \left|\frac{(\pi_\varepsilon^a)''}{(\pi_\varepsilon^a)'}\right|\leq\frac{\rho_\varepsilon^{-a}(y)}{\int_0^y\rho_\varepsilon^{-a}(s)\mathrm{d}s}+|a|\frac{y}{\varepsilon^2+y^2}\leq\frac{\max\{1, 1-a\}+|a|}{y}.

    Eventually

    |(\log\pi_\varepsilon^a)''|\leq \left|\frac{(\pi_\varepsilon^a)'}{\pi_\varepsilon^a}\right|\cdot\left|\frac{(\pi_\varepsilon^a)''}{(\pi_\varepsilon^a)'}\right|+\left|\frac{(\pi_\varepsilon^a)'}{\pi_\varepsilon^a}\right|^2\leq \frac{c}{y^2}.

    Obviously, point i) implies the uniform upper bound in (B.4) by trace Poincaré and the Hardy inequalities. In order to prove the uniform lower bound and eventually proving ii) , we only have to prove that there exists a positive constant c_1 > -\frac{1}{4} uniform in \varepsilon\to0 such that

    V_{\omega_\varepsilon^a}(y)\geq\frac{c_1}{y^2}.

    In fact,

    W_{\omega_\varepsilon^a}(y)+\left(-\frac{a}{2}\right)^+\geq0.

    Let t = y/\varepsilon > 0 . Then

    V_{\omega_\varepsilon^a}(y) = \frac{\Phi_a(t)}{y^2},

    with \Phi_a as in definition (B.18). We can conclude by applying Lemma B.6.

    Eventually we remark that also condition iii) holds true.

    Let us define \tilde H^1(B_1^+, (\omega_\varepsilon^a(y))^{-1}\mathrm{d}z) as the closure of C^\infty_c(\overline B_1^+) with respect to the norm

    \int_{B_1^+}(\omega_\varepsilon^a)^{-1}|\nabla u|^2.

    Lemma B.8. Let a\in(-\infty, 1) and u\in\tilde H^1(B_1^+, (\omega_\varepsilon^a(y))^{-1}\mathrm{d}z) . Then the following inequalities hold true for a positive constant c independent of \varepsilon\in[0, 1]

    \begin{equation} c\int_{B_1^+}(\omega_\varepsilon^a)^{-1}u^2\leq\int_{B_1^+}(\omega_\varepsilon^a)^{-1}|\nabla u|^2, \end{equation} (B.24)
    \begin{equation} c\int_{S^n_+}(\omega_\varepsilon^a)^{-1}u^2\leq\int_{B_1^+}(\omega_\varepsilon^a)^{-1}|\nabla u|^2, \end{equation} (B.25)
    \begin{equation} c\int_{B_1^+}\frac{(\omega_\varepsilon^a)^{-1}}{y^2}u^2\leq\int_{B_1^+}(\omega_\varepsilon^a)^{-1}|\nabla u|^2, \end{equation} (B.26)
    \begin{equation} c\int_{S^n_+}\frac{(\omega_\varepsilon^a)^{-1}}{y}u^2\leq\int_{B_1^+}(\omega_\varepsilon^a)^{-1}|\nabla u|^2, \end{equation} (B.27)
    \begin{equation} c\left(\int_{B_1^+}((\omega_\varepsilon^a)^{-1})^{2^*/2}|u|^{2^*}\right)^{2/2^*}\leq \int_{B_1^+}(\omega_\varepsilon^a)^{-1}|\nabla u|^2, \end{equation} (B.28)

    which are respectively the Poincaré inequality, the trace Poincaré inequality, the Hardy inequality, the trace Hardy inequality and a Sobolev type inequality.

    Proof. First, we prove (B.25). Thanks to Lemma B.7 we can define for a sequence \varepsilon_k\to0

    \tilde\mu_k = \min\limits_{v\in\tilde H^1(B_1^+)\setminus\{0\}}\frac{\tilde Q_{k}(v)}{\int_{S^n_+}v^2} = \min\limits_{v\in\tilde H^1(B_1^+)\setminus\{0\}}\frac{Q_{k}(v)}{\int_{S^n_+}v^2}+\left(-\frac{a}{2}\right)^+ = \mu_k+\left(-\frac{a}{2}\right)^+,

    and

    \tilde\mu = \min\limits_{v\in\tilde H^1(B_1^+)\setminus\{0\}}\frac{\tilde Q(v)}{\int_{S^n_+}v^2} = \min\limits_{v\in\tilde H^1(B_1^+)\setminus\{0\}}\frac{Q(v)}{\int_{S^n_+}v^2}+\left(-\frac{a}{2}\right)^+ = \mu+\left(-\frac{a}{2}\right)^+.

    Actually, we are able to provide the value of \mu since u(x, y) = y^{1-(a-2)} is the unique function in \tilde H^{1, a-2}(B_1^+)\setminus\{0\} which solves

    \begin{equation*} \begin{cases} -L_{a-2}u = 0 &\mathrm{in} \ B_1^+\\ u \gt 0 &\mathrm{in} \ B_1^+\\ u(x, 0) = 0\\ \nabla u\cdot\nu = \mu u &\mathrm{in} \ S_+^{n}, \end{cases} \end{equation*}

    with \mu = 1-(a-2) = 3-a . Hence, by Lemma B.3, since \tilde\mu_k\to\tilde\mu , then \mu_k\to\mu = 3-a > 0 and one can find \varepsilon_0 > 0 such that for 0\leq\varepsilon_k\leq\varepsilon_{\overline k} = \varepsilon_0 one has \mu_k\geq\mu_{\overline k} > 0 . Hence one has (B.25) with a constant \mu_{\overline k} > 0 uniform in 0\leq\varepsilon\leq\varepsilon_0 . For the other inequalities, the proof is done taking functions u\in C^\infty_c(\overline B_1^+\setminus\Sigma) and then passing to functions u\in\tilde H^1(B_1^+, (\omega_\varepsilon^a(y))^{-1}\mathrm{d}z) by density. By Lemma B.7 there exists a positive constant uniform in \varepsilon such that

    \begin{equation} \int_{B_1^+}(\omega_\varepsilon^a)^{-1}|\nabla u|^2+\left(-\frac{a}{2}\right)^+\int_{S^n_+}(\omega_\varepsilon^a)^{-1}u^2 = \tilde Q_{\omega_\varepsilon^a}((\omega_\varepsilon^a)^{-1/2}u)\geq c\int_{B_1^+}|\nabla((\omega_\varepsilon^a)^{-1/2}u)|^2, \end{equation} (B.29)

    then all the inequalities are obtained by the validity of them in \tilde H^1(B_1^+) and using the trace Poincaré inequality (B.25).

    B.4.2. Super degenerate weights \omega_\varepsilon^a

    Let a\in(-\infty, 1) and let us consider u\in C^\infty(B_1^+) and define v = (\omega_\varepsilon^a)^{1/2}u\in C^\infty_c(\overline B_1^+\setminus\Sigma) . Then we consider the quadratic form

    \begin{equation} \int_{B_1^+}\omega_\varepsilon^a\left(|\nabla u|^2+u^2\right) = \overline Q_{\omega_\varepsilon^a}(v) = \int_{B_1^+}\left(|\nabla v|^2+v^2\right)+\int_{B_1^+}\overline V_{\omega_\varepsilon^a} v^2+\int_{S^n_+}\overline W_{\omega_\varepsilon^a} v^2, \end{equation} (B.30)

    with

    \overline V_{\omega_\varepsilon^a} = \frac{1}{4}[(\log\omega_\varepsilon^a)']^2+\frac{1}{2}(\log\omega_\varepsilon^a)'' = \frac{a}{4}\frac{(a-2)y^2+2\varepsilon^2}{(\varepsilon^2+y^2)^2},

    and

    \overline W_{\omega_\varepsilon^a} = -\frac{1}{2}(\log\omega_\varepsilon^a)'y.

    We remark that \overline V_{\omega_\varepsilon^a} = V_{\rho_\varepsilon^a} in (B.7). It is easy to check that the family of quadratic forms \overline Q_{\omega_\varepsilon^a} are equivalent norms in \tilde H^1(B_1^+) with constants which are uniform in \varepsilon ; i.e., the following holds

    Lemma B.9. Let a\in(-\infty, 1) . The family \{\overline Q_{k}\} = \{\overline Q_{\omega_{\varepsilon_k}}\} defined in (B.30) and its limit \overline Q satisfy the conditions in Lemma B.3.

    B.5. Isometries

    In this last section, we express a fundamental consequence of the previous estimate on uniform-in- \varepsilon equivalence of norms. Indeed, for all exponents a\neq 0 , the nature of the weighted Sobolev spaces changes drastically when switching between \varepsilon > 0 and \varepsilon = 0 . For this reason, we need to embed them isometrically in the common space \tilde H^1 uniformly as \varepsilon\to0 . To this aim, we can take advantage of some fundamental isometries between weighted spaces to \tilde H^1 , which allow, by reabsorbing the weight, to obtain uniform estimates in a common space to any element in the approximating sequence. Fixed a\in(-\infty, 1) and \varepsilon\geq0 , then the map

    T^a_\varepsilon:\tilde H^1(B_1^+, \rho_\varepsilon^a(y)\mathrm{d}z)\to\tilde H^1(B_1^+)\;: u\mapsto v = T^a_\varepsilon(u) = (\rho_\varepsilon^a)^{1/2}u

    is an isometry when we endow the space \tilde H^1(B_1^+) with the squared norm Q_{\rho_\varepsilon^a} . Indeed we have:

    \int_{B_1^+}\rho_\varepsilon^a|\nabla u|^2 = Q_{\rho_\varepsilon^a}(v).

    Is is worthwhile noticing that the family of quadratic forms Q_{\rho_\varepsilon^a} is uniformly bounded (above and below) with respect to \varepsilon\in[0, 1] .

    Eventually, we remark that, similarily, fixed a\in(-\infty, 1) and \varepsilon\geq0 , then the map

    \begin{equation} \overline T^a_\varepsilon: H^1(B_1^+, \omega_\varepsilon^a(y)\mathrm{d}z)\to\tilde H^1(B_1^+)\;: u\mapsto v = \overline T^a_\varepsilon(u) = (\omega_\varepsilon^a)^{1/2}u \end{equation} (B.31)

    is also an isometry when the latter space is endowed with the squared norm \overline Q_{\omega_\varepsilon^a}(v) as we have

    \int_{B_1^+}\omega_\varepsilon^a\left(|\nabla u|^2+u^2\right) = \overline Q_{\omega_\varepsilon^a}(v).

    Again, \overline Q_{\omega_\varepsilon^a} is uniformly bounded (above and below) with respect to \varepsilon\in[0, 1] . Once again, we remark that for these super degenerate weights Poincaré type inequalities do not hold true (see [21]) and hence we can not consider only the weighted L^2 -norm of the gradient in the equation above.



    [1] G. Dattoli, C. Chiccoli, S. Lorenzutta, G. Maino, A. Torre, Generalized Bessel functions and generalized Hermite polynomials, J. Math. Anal. Appl., 178 (1993), 509–516. https://doi.org/10.1006/jmaa.1993.1321 doi: 10.1006/jmaa.1993.1321
    [2] G. Dattoli, S. Lorenzutta, G. Maino, A. Torre, C. Cesarano, Generalized Hermite polynomials and super-Gaussian forms, J. Math. Anal. Appl., 203 (1996), 597–609. https://doi.org/10.1006/jmaa.1996.0399 doi: 10.1006/jmaa.1996.0399
    [3] P. Appell, J. K. de F{\acute{e}}riet, Fonctions hyperg{\rm\acute{e}}om{\rm\acute{e}}triques et hypersph{\rm\acute{e}}riques: polyn{\rm\hat{o}}mes d' Hermite, Gauthier-Villars, Paris, 1926.
    [4] B. Yılmaz, M. A. Özarslan, Differential equations for the extended 2D Bernoulli and Euler polynomials, Adv. Differ. Equ., 2013 (2013), 107. https://doi.org/10.1186/1687-1847-2013-107 doi: 10.1186/1687-1847-2013-107
    [5] S. Khan, G. Yasmin, R. Khan, N. A. M. Hassan, Hermite-based Appell polynomials: properties and applications, J. Math. Anal. Appl., 351 (2009), 756–764. https://doi.org/10.1016/j.jmaa.2008.11.002 doi: 10.1016/j.jmaa.2008.11.002
    [6] P. Appell, Sur une classe de polynômes, Ann. Sci. \acute{E}cole. Norm. Sup., 9 (1880), 119–144. https://doi.org/10.24033/asens.186
    [7] S. Roman, The umbral calculus, New York: Academic Press, 1984.
    [8] L. Bedratyuk, J. Flusser, T. Suk, J. Kostkova, J. Kautsky, Non-separable rotation moment invariants, Pattern Recogn., 127 (2022), 108607. https://doi.org/10.1016/j.patcog.2022.108607 doi: 10.1016/j.patcog.2022.108607
    [9] J. Flusser, T. Suk, L. Bedratyuk, T. Karella, 3D non-separable moment invariants, In: N. Tsapatsoulis, Computer analysis of images and patterns, CAIP 2023, Lecture Notes in Computer Science, Springer, 14184 (2023), 295–305. https://doi.org/10.1007/978-3-031-44237-7_28
    [10] M. X. He, P. E. Ricci, Differential equation of Appell polynomials via the factorization method, J. Comput. Appl. Math., 139 (2002), 231–237. https://doi.org/10.1016/S0377-0427(01)00423-X doi: 10.1016/S0377-0427(01)00423-X
    [11] L. Infeld, T. E. Hull, The factorization method, Rev. Mod. Phys., 23 (1951), 21–68. https://doi.org/10.1103/RevModPhys.23.21 doi: 10.1103/RevModPhys.23.21
    [12] S. A. Wani, S. Khan, Certain properties and applications of the 2D Sheffer and related polynomials, Bol. Soc. Mat. Mex., 26 (2020), 947–971. https://doi.org/10.1007/s40590-020-00280-5 doi: 10.1007/s40590-020-00280-5
    [13] S. A. Wani, S. Khan, Properties and applications of the Gould-Hopper-Frobenius-Euler polynomials, Tbilisi Math. J., 12 (2019), 93–104 https://doi.org/10.32513/tbilisi/1553565629 doi: 10.32513/tbilisi/1553565629
    [14] S. Araci, M. Riyasat, S. A. Wani, S. Khan, Differential and integral equations for the 3-variable Hermite-Frobenius-Euler and Frobenius-Genocchi polynomials, Appl. Math. Inf. Sci., 11 (2017), 1335–1346. https://doi.org/10.18576/amis/110510 doi: 10.18576/amis/110510
    [15] B. S. T. Alkahtani, I. Alazman, S. A. Wani, Some families of differential equations associated with multivariate Hermite polynomials, Fractal Fract., 7 (2023), 390. https://doi.org/10.3390/fractalfract7050390 doi: 10.3390/fractalfract7050390
    [16] Z. Ozat, M. A. Ozarslan, B. Cekim, On Bell based Appell polynomials, Turk. J. Math., 47 (2023), 5. https://doi.org/10.55730/1300-0098.3415 doi: 10.55730/1300-0098.3415
    [17] G. Baran, Z. Ozat, B. Cekim, M. A. Ozarslan, Some properties of degenerate Hermite Appell polynomials in three variables, Filomat, 37 (2023), 6537–6567. https://doi.org/10.2298/FIL2319537B doi: 10.2298/FIL2319537B
    [18] N. Biricik, B. Çekim, M. A. Özarslan, Sequences of twice-iterated \Deltaw-Gould–Hopper Appell polynomials, J. Taibah Univ. Sci., 18 (2023), 2286714. https://doi.org/10.1080/16583655.2023.2286714 doi: 10.1080/16583655.2023.2286714
    [19] S. Khan, N. Raza, General-Appell polynomials within the context of monomiality principle, Int. J. Anal., 2013 (2013), 328032. https://doi.org/10.1155/2013/328032 doi: 10.1155/2013/328032
    [20] S. Khan, G. Yasmin, R. Khan, N. A. M. Hassan, Hermite-based Appell polynomials: properties and applications, J. Math. Anal. Appl., 351 (2009), 756–764. https://doi.org/10.1016/j.jmaa.2008.11.002 doi: 10.1016/j.jmaa.2008.11.002
    [21] H. M. Srivastava, M. A. Özarslan, B. Yılmaz, Some families of differential equations associated with the Hermite-based Appell polynomials and other classes of Hermite-based polynomials, Filomat, 28 (2014), 695–708. https://doi.org/10.2298/FIL1404695S doi: 10.2298/FIL1404695S
    [22] H. Bateman, A. Erdélyi, Higher transcendental functions, McGraw-Hill Book Company, 1953.
    [23] J. Sándor, B. Crstici, Handbook of number theory II, Springer Science & Business Media, 2004.
  • This article has been cited by:

    1. Yannick Sire, Susanna Terracini, Stefano Vita, Liouville type theorems and regularity of solutions to degenerate or singular problems part I: even solutions, 2021, 46, 0360-5302, 310, 10.1080/03605302.2020.1840586
    2. Agnid Banerjee, Federico Buseghin, Nicola Garofalo, The thin obstacle problem for some variable coefficient degenerate elliptic operators, 2022, 223, 0362546X, 113052, 10.1016/j.na.2022.113052
    3. Alessandra De Luca, Veronica Felli, Stefano Vita, Strong unique continuation and local asymptotics at the boundary for fractional elliptic equations, 2022, 400, 00018708, 108279, 10.1016/j.aim.2022.108279
    4. Giorgio Tortone, The nodal set of solutions to some nonlocal sublinear problems, 2022, 61, 0944-2669, 10.1007/s00526-022-02197-5
    5. Hongjie Dong, Tuoc Phan, Hung Tran, Degenerate linear parabolic equations in divergence form on the upper half space, 2023, 0002-9947, 10.1090/tran/8892
    6. Hongjie Dong, Tuoc Phan, Weighted mixed-norm L estimates for equations in non-divergence form with singular coefficients: The Dirichlet problem, 2023, 285, 00221236, 109964, 10.1016/j.jfa.2023.109964
    7. Hongjie Dong, Seongmin Jeon, Stefano Vita, Schauder type estimates for degenerate or singular elliptic equations with DMO coefficients, 2024, 63, 0944-2669, 10.1007/s00526-024-02840-3
    8. Hongjie Dong, Tuoc Phan, Hung Vinh Tran, Nondivergence form degenerate linear parabolic equations on the upper half space, 2024, 286, 00221236, 110374, 10.1016/j.jfa.2024.110374
    9. David Jesus, Yannick Sire, Gradient regularity for fully nonlinear equations with degenerate coefficients, 2024, 0373-3114, 10.1007/s10231-024-01498-0
    10. Alessandro Audrito, Susanna Terracini, On the Nodal Set of Solutions to a Class of Nonlocal Parabolic Equations, 2024, 301, 0065-9266, 10.1090/memo/1512
    11. Hongjie Dong, Tuoc Phan, Yannick Sire, Sobolev Estimates for Singular-Degenerate Quasilinear Equations Beyond the A_2 Class, 2024, 34, 1050-6926, 10.1007/s12220-024-01729-z
    12. Alessandro Audrito, Gabriele Fioravanti, Stefano Vita, Schauder estimates for parabolic equations with degenerate or singular weights, 2024, 63, 0944-2669, 10.1007/s00526-024-02809-2
    13. Alessandro Audrito, On the existence and Hölder regularity of solutions to some nonlinear Cauchy–Neumann problems, 2023, 23, 1424-3199, 10.1007/s00028-023-00899-7
    14. Hongjie Dong, Tuoc Phan, Parabolic and elliptic equations with singular or degenerate coefficients: The Dirichlet problem, 2021, 374, 0002-9947, 6611, 10.1090/tran/8397
    15. Seongmin Jeon, Stefano Vita, Higher order boundary Harnack principles in Dini type domains, 2024, 412, 00220396, 808, 10.1016/j.jde.2024.08.059
    16. Susanna Terracini, Giorgio Tortone, Stefano Vita, Higher Order Boundary Harnack Principle via Degenerate Equations, 2024, 248, 0003-9527, 10.1007/s00205-024-01973-1
  • Reader Comments
  • © 2024 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(776) PDF downloads(33) Cited by(2)

Figures and Tables

Tables(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog