Research article Special Issues

A new glance to the Alt-Caffarelli-Friedman monotonicity formula

  • In this paper we revisit the proof of the Alt-Caffarelli-Friedman monotonicity formula. Then, in the framework of the Heisenberg group, we discuss the existence of an analogous monotonicity formula introducing a necessary condition for its existence, recently proved in [18].

    Citation: Fausto Ferrari, Nicolò Forcillo. A new glance to the Alt-Caffarelli-Friedman monotonicity formula[J]. Mathematics in Engineering, 2020, 2(4): 657-679. doi: 10.3934/mine.2020030

    Related Papers:

    [1] Morteza Fotouhi, Andreas Minne, Henrik Shahgholian, Georg S. Weiss . Remarks on the decay/growth rate of solutions to elliptic free boundary problems of obstacle type. Mathematics in Engineering, 2020, 2(4): 698-708. doi: 10.3934/mine.2020032
    [2] Daniela De Silva, Ovidiu Savin . Uniform density estimates and Γ-convergence for the Alt-Phillips functional of negative powers. Mathematics in Engineering, 2023, 5(5): 1-27. doi: 10.3934/mine.2023086
    [3] Catharine W. K. Lo, José Francisco Rodrigues . On an anisotropic fractional Stefan-type problem with Dirichlet boundary conditions. Mathematics in Engineering, 2023, 5(3): 1-38. doi: 10.3934/mine.2023047
    [4] Filippo Gazzola, Gianmarco Sperone . Remarks on radial symmetry and monotonicity for solutions of semilinear higher order elliptic equations. Mathematics in Engineering, 2022, 4(5): 1-24. doi: 10.3934/mine.2022040
    [5] Donatella Danielli, Rohit Jain . Regularity results for a penalized boundary obstacle problem. Mathematics in Engineering, 2021, 3(1): 1-23. doi: 10.3934/mine.2021007
    [6] 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
    [7] Hugo Tavares, Alessandro Zilio . Regularity of all minimizers of a class of spectral partition problems. Mathematics in Engineering, 2021, 3(1): 1-31. doi: 10.3934/mine.2021002
    [8] Tatsuya Miura . Polar tangential angles and free elasticae. Mathematics in Engineering, 2021, 3(4): 1-12. doi: 10.3934/mine.2021034
    [9] Matteo Novaga, Marco Pozzetta . Connected surfaces with boundary minimizing the Willmore energy. Mathematics in Engineering, 2020, 2(3): 527-556. doi: 10.3934/mine.2020024
    [10] Daniela De Silva, Giorgio Tortone . Improvement of flatness for vector valued free boundary problems. Mathematics in Engineering, 2020, 2(4): 598-613. doi: 10.3934/mine.2020027
  • In this paper we revisit the proof of the Alt-Caffarelli-Friedman monotonicity formula. Then, in the framework of the Heisenberg group, we discuss the existence of an analogous monotonicity formula introducing a necessary condition for its existence, recently proved in [18].


    The Alt-Caffarelli-Friedman monotonicity formula was introduced in [1] as a fundamental tool for studying the main properties of the solutions of two-phase free boundary problems. Roughly saying, following [1], the result says that there exists r0>0 such that for every non-negative u1,u2C(B1(0))H1(B1(0)), Δui0, i=1,2, u1(0)=u2(0)=0 and u1u2=0 in B1(0), where B1(0) is the Euclidean ball centered at 0 of radius 1 in Rn, then

    Φ(r):=r4Br(0)|u1(x)|2|x|n2dxBr(0)|u2(x)|2|x|n2dx (1.1)

    is well defined, bounded and monotone increasing in [0,r0).

    Alt, Caffarelli and Friedman used this result for proving the Lipschitz continuity of critical points of a functional like the following one

    E(v):=Ω(|v|2+χ{v>0})dx (1.2)

    defined on a set KH1(Ω), where ΩRn is a given bounded open set and K is determined by some known conditions on v given on Ω, where χ{v>0} denotes, as usual, the characteristic function of the set {v>0}.

    The critical points of the previous functional E satisfy the following two-phase free boundary problem

    {Δu=0in Ω+(u):={xΩ:u(x)>0},Δu=0in Ω(u):=Int({xΩ:u(x)0}),|u+|2|u|2=1on F(u):=Ω+(u)Ω, (1.3)

    see [1]. Thus, solutions of (1.3) satisfy, at least in a "weak" sense, the following property: for every PF(u)

    (u+ν(P))2(uν(P))2=limr0+Φ(r)C,

    where u+:=sup{u,0}, u:=sup{u,0}, ν is the unit vector, pointing inside Ω+(u) at PF(u) and inside Ω(u) at PF(u) when this makes sense in the smooth case. See [9] for a more general viscosity meaning.

    Hence, if one of the two phases, let us say u, is sufficiently regular at PF(u), see [27], then by the Hopf maximum principle it results uν(P)>0 so that, as a by-product, u+ν(P) has to be bounded. In this way, the solutions of the free boundary problem are globally Lipschitz.

    After [1] many other important papers on this topic appeared. We remind some of them, without pretending to cite all the literature about this topic. In [7] it was proved that the monotonicity formula holds for linear uniformly elliptic operators in divergence form with Hölder continuous coefficients, in [8] a formula for non-homogeneous free boundary problems was discovered, in [39] the Riemannian case was treated, while in [33] the non-divergence form case has been faced. Some very partial results have been obtained also in the nonlinear case in lower dimension: See [15] for the pLaplace case.

    Moreover, this formula became popular for other applications as well. Among them, there are further two-phase problems, see [6] for the elliptic homogeneous case, [2] and [19] for the parabolic homogeneous setting, and [14] for the elliptic linear non-homogeneous problems. In addition we also recall some segregation problems, see for instance: [34,35,37,38]. In this way, during the last decade, the Alt-Caffarelli-Friedman monotonicity formula has quickly increased its importance in literature.

    The existence of such a tool for elliptic degenerate operators, for instance sublaplacians on groups, as far as we know, has not yet been understood. Anyhow, concerning other similar formulas about sublaplacians we find in the literature some important contributions, see [26] and in particular [28], where the authors deal with the frequency function of Almgren in Carnot groups. Moreover, see [12,13] for further papers in non-commutative setting dealing with other free boundary problems, namely the obstacle problem.

    We also gently warn the reader about the existence of results about two-phase problems in the Heisenberg group, like [16,20] in particular, where the following parallel version of the Euler equations (1.3), of a two-phase problem in this non-commutative framework, has been achieved:

    {ΔHnu=0in Ω+(u):={xΩ:u(x)>0}ΔHnu=0in Ω(u):=Int({xΩ:u(x)0})|Hnu+|2|Hnu|2=1on F(u):=Ω+(u)Ω. (1.4)

    In the Section 3 of this paper, we shall introduce the main notation that we need for working on this subject. Nevertheless, we ask to the reader that is not customary with this language to continue to follow this colloquial presentation having in mind that, in the Heisenberg group, there exists a natural translation of the classical Euclidean tools in terms of parallel intrinsic notions in the non-commutative structure Hn. Hence, what we are going to discuss in a while in this introduction, it should be easily interpreted by all. We remark that, in this particular non-commutative context, the gradient jump |u+|2|u|2=1 now is governed by the jump of the horizontal gradient Hn of the solutions u of (1.4) in the Heisenberg group Hn. As a first consequence, in this degenerate case associated with the sublaplacian ΔHn, a new geometric problem, that in the Euclidean two-phase problem did not exist, now appears. In fact, since classical smooth free boundaries of (1.4), in principle, might have characteristic points, then the jump of the horizontal gradient of u on F(u) could not be satisfied pointwise, because the horizontal gradient vanishes on characteristic points, see Section 3. It is also worthwhile to recall that it has been already proved that a minimum u of the functional

    EHn(v):=Ω(|Hnv|2+χ{v>0})dx,

    ΩHn, see Section 3 in [20], is endowed with a locally bounded horizontal gradient Hnu and moreover that every minimum u satisfies ΔHnu=0in Ω+(u), as well as ΔHnu=0in Ω(u), even if no word has been spent about the behavior of the free boundary of these local minima. Indeed, an alternative way of proving that a local minimum of the functional EHn is intrinsically Lipschitz, instead of using the monotonicity formula, has been shown in [20]. The proof of the monotonicity formula in the Euclidean framework is quite long and based on many highly non-trivial results. Thus, we like to revisit it in Section 2, by commenting the key points of the proof and then focusing our attention to the parallel steps that we would need to prove in the Heisenberg group, including the statement of our following main result proved in [18] as well.

    In order to reach our goal, let us introduce the following family of functionals depending on a real number β>0:

    Jβ,H1(r)=rβBH1r(0)H1u12|ζ|2H1dζBH1r(0)H1u22|ζ|2H1dζ. (1.5)

    Following the main steps of the Euclidean proof, in [18] we proved the following result as a corollary of an estimate of the first eigenvalue of an operator defined on the boundary of the Koranyi ball of radius one. In fact, as people who usually work in this sub-Riemannian field well know, this set takes the place of the boundary of the classical Euclidean ball of radius one, when we need to work with the fundamental solution of the sublaplacian ΔH1, see [17].

    Theorem 1.1. If there exists a positive number β for which Jβ,H1 is monotone for every non-negative u1,u2H1H1(BH11(0)), such that ΔH1ui0, ui(0)=0, i=1,2 and u1u2=0, then β4.

    We stated this result in the first Heisenberg group only, because we did not prove a monotonicity formula for all the Heisenberg groups, but simply we have proved that if this formula holds in the non-commutative framework given by H1, then the right exponent β has to be smaller or equal than 4. The proof in higher Heisenberg groups requires more computations, but it may be obtained with some further efforts, that we do not discuss here, following the same ideas. On the other hand, the breakthrough that we would need for concluding that, at least in H1, the sharp exponent β is exactly 4, depends on a long standing open question. In fact the best profile of the set that realizes the equality in the isoperimetric inequality in the Heisenberg group (and as a byproduct the descendant Polya-Szëgo inequality on the surface of the Koranyi ball of radius one) is still open, see [11] for an introduction to this problem. So that, considering previous arguments, we have decided to state our result only in H1. We shall discuss this part in Section 5. In the remaining Section 4, we describe the main tools we need for obtaining the key estimate on the Rayleigh quotient in H1, see [18] for the details.

    In this section, following the original paper [1] and [9], we try to focus on the main steps we need to achieve for proving the Alt-Caffarelli-Friedman monotonicity formula in the Euclidean setting.

    After a straightforward differentiation, it results

    Φ(r)=I1(r)I2(r)r5(4+r(I1I1+I2I2)), (2.1)

    where for i=1,2:

    Ii(r)=Br(0)|ui(x)|2|x|n2dx.

    By a rescaling argument, we can write

    Φ(r)=I1(r)I2(r)r5(4+B1(0)|u1(x)|2dσB1(0)|u1(x)|2|x|n2dx+B1(0)|u2(x)|2dσB1(0)|u2(x)|2|x|n2dx). (2.2)

    Precisely, we have

    Ii(r)=Br(0)|ui(x)|2|x|n2dx=B1(0)|ui(ry)|2|ry|n2rndy=r2B1(0)|ui(ry)|2|y|n2dy,

    and

    Ii(r)=Br(0)|ui(x)|2|x|n2dx=r0(Bρ(0)|ui(x)|2|x|n2dσ(x))dρ=r0(B1(0)|ui(ρy)|2ρn2ρn1dσ(y))dρ=r0ρ(B1(0)|ui(ρy)|2dσ(y))dρ,

    where here y denotes the coordinates on B1(0). Thus, we get

    IiIi=ddrr0ρ(B1(0)|ui(ρy)|2dσ(y))dρr2B1(0)|ui(ry)|2|y|n2dy=rB1(0)|ui(ry)|2dσ(y)r2B1(0)|ui(ry)|2|y|n2dy=1rB1(0)|ui(ry)|2dσ(y)B1(0)|ui(ry)|2|y|n2dy,

    which implies, if we define

    (ui)r(x)=ui(rx)r,xB1(0),

    that

    IiIi=1rB1(0)|(ui)r(y)|2dσ(y)B1(0)|(ui)r(y)|2|y|n2dy,

    where (ui)r is defined in B1(0). As a consequence, if we write y=x and (ui)r=ui the last equality gives

    rIiIi=B1(0)|ui(x)|2dσ(x)B1(0)|ui(x)|2|x|n2dx,

    and so (2.1) becomes (2.2).

    Now, if

    4+B1(0)|u1(x)|2dσB1(0)|u1(x)|2|x|n2dx+B1(0)|u2(x)|2dσB1(0)|u2(x)|2|x|n2dx0

    then, from (2.2), Φ(r)0. Hence, in order to prove that the previous inequality holds, the following ratios

    Ji(r):=B1(0)|ui(x)|2dσB1(0)|ui(x)|2|x|n2dx,

    for i=1,2, have to be estimated.

    Since the gradient may split in two orthogonal parts involving the radial part and the tangential part, respectively denoted by ρui and θui, it results

    |ui(x)|2=|ρui(x)|2+|θui(x)|2.

    Then, we can rewrite Ji as

    Ji(r)=B1(0)(|ρui(x)|2+|θui(x)|2)dσB1(0)|ui(x)|2|x|n2dx. (2.3)

    At this point, we estimate the numerator and denominator of (2.3) separately.

    As regards the numerator, we define first

    λ(Γi):=infvH10(Γi)Γi|θv(x)|2dσΓiv(x)2dσ,

    where

    Γi:={xB1(0):ui(x)>0}

    and λ(Γi), i=1,2, is the Rayleigh quotient. By the definition of λ(Γi), we thus obtain, for every βi(0,1),

    B1(0)|θui(x)|2dσ=Γi|θui(x)|2dσλ(Γi)Γiui(x)2dσ=(1βi+βi)λ(Γi)Γiui(x)2dσ=βiλ(Γi)Γiui(x)2dσ+(1βi)λ(Γi)Γiui(x)2dσ,

    hence, by Cauchy inequality, we have

    B1(0)(|ρui(x)|2+|θui(x)|2)dσΓi|ρui(x)|2dσ+βiλ(Γi)Γiui(x)2dσ+(1βi)λ(Γi)Γiui(x)2dσ2(Γi|ρui(x)|2dσ)1/2(βiλ(Γi)Γiui(x)2dσ)1/2+(1βi)λ(Γi)Γiui(x)2dσ. (2.4)

    Concerning the denominator, instead, we compute

    Δ(u2i)=nj=12x2j(u2i)=nj=1xj(2uiuixj)=2(|ui|2+uiΔui)2|ui|2,

    since uiΔui0 by the assumptions on ui.

    Consequently, we achieve the following estimate:

    B1(0)|ui(x)|2|x|n2dx(Γi|ρui(x)|2dσ)12(Γiu2i(x)dσ)12+n22Γiu2i(x)dσ. (2.5)

    In fact, the previous inequality follows after an integration by parts, using the facts that |x|2n is, up to a multiplicative constant, the fundamental solution of Δ and 0F(ui), i=1,2, and by the Hölder inequality because:

    B1(0)|ui(x)|2|x|n2dx12B1(0)Δ(u2i)(x)|x|2ndx=12(B1(0)div(|x|2n(u2i)(x))dxB1(0)(|x|2n)(u2i)(x)dx)=12(B1(0)|x|2n(u2i)(x)x|x|dσB1(0)div(u2i(x)(|x|2n))dx+B1(0)u2i(x)Δ(|x|2n)dx)=12(Γi2ui(x)ρui(x)dσ+(n2)Γiu2i(x)|x|1ndσ)=Γiui(x)ρui(x)dσ+n22Γiu2i(x)dσ(Γi|ρui(x)|2dσ)12(Γiu2i(x)dσ)12+n22Γiu2i(x)dσ.

    Now, putting together (2.4) and (2.5), we get, in view of (2.3),

    Ji(r)2(Γi|ρui(x)|2dσ)12(Γiβiλ(Γi)u2i(x)dσ)12+(1βi)λ(Γi)Γiu2i(x)dσ(Γi|ρui(x)|2dσ)12(Γiu2i(x)dσ)12+n22Γiu2i(x)dσ, (2.6)

    and setting ξi=(Γi|ρui(x)|2dσ)12 and ηi=(Γiu2i(x)dσ)12, it holds

    Ji(r)2(βiλ(Γi))12ξiηi+(1βi)λ(Γi)η2iξiηi+n22η2i=2(βiλ(Γi))12+(1βi)λ(Γi)ηiξi1+n22ηiξiinfz02(βiλ(Γi))12+(1βi)λ(Γi)z1+n22z=2min{λ(Γi)n2(1βi),(βiλ(Γi))12}.

    The last equality easily follows by elementary arguments. Now, if it were possible to choose βi(0,1) in such a way that

    λ(Γi)n2(1βi)=(βiλ(Γi))12

    we would realize, by denoting αi:=(βiλ(Γi))12, that the previous equation is satisfied if and only if

    α2i+(n2)αiλ(Γi)=0.

    On the other hand, since a function u=ραg(θ), θ:=(θ1,,θn1), is harmonic in a cone determined by a domain Γ whenever

    ρα2((α(α1)+α(n1))g(θ)+Δθg)=0,

    we deduce that there exists αi such that

    αi(αi1)+αi(n1)=λ(Γi),

    namely

    α2i+(n2)αiλ(Γi)=0.

    By the structure of the equation, it immediately comes out that there always exists a strictly positive solution αi=αi(Γi), which is called the characteristic constant of Γi.

    Therefore, we have to prove the existence of βi(0,1) such that

    (n2)+(n2)2+4λ(Γi)2=(βiλ(Γi))12. (2.7)

    Specifically, (2.7) is equivalent to solving

    4λ(Γi)(n2)+(n2)2+4λ(Γi)=2(βiλ(Γi))12,

    that is

    2λ(Γi)12(n2)+(n2)2+4λ(Γi)=β12i.

    Since the continuous positive function defined in [0,+) as

    zz(n2)+(n2)2+z2

    is strictly increasing, (z(n2)+(n2)2+z2)(0)=0 and sup[0,+)z(n2)+(n2)2+z2=1, we conclude that for every λ(Γi)>0, there exists βi such that (2.7) holds. In particular, we get

    βi=(2λ(Γi)12(n2)+(n2)2+4λ(Γi))2.

    Hence, with previous choice of βi, if we denote

    αi:=min{λ(Γi)n2(1βi),(βiλ(Γi))12},

    which is also the exponent corresponding to the eigenvalue given by the Rayleigh quotient λ(Γi), we conclude that, whenever

    α1+α22, (2.8)

    then Φ0.

    So, for completing this proof, we would need to know that (2.8) holds.

    To this end, by [36] we know that αi(Γi)αi(Γi), where ΓiB1(0) is a spherical cap, namely a set of the form

    Γi=B1(0){xn>s},1<s<1,

    such that Hn1(Γi)=Hn1(Γi). Here Hn1 denotes the (n1)-dimensional Hausdorff measure on B1(0).

    Precisely, [36] shows that if uC(B1(0),R), then

    {B1(0)updHn1B1(0)updHn11p<,uL(B1(0))uL(B1(0)), (2.9)

    where u is the symmetrized function of u, depending only on the latitude of the argument. Moreover, we also have that u#(Hn1)|B(R)=u#(Hn1)|B(R), that is the pushforward measures of u and u coincide in the Borel sets of R, which entails

    B1(0)ϕudHn1=B1(0)ϕudHn1, (2.10)

    for any function ϕ:RR μ-measurable, where μ is the outer measure defined on the power set P(R) of R as

    μ(F)=inf{i=1μ(Ai):AiB(R),Fi=1Ai},

    with μ=u#(Hn1)|B(R)=u#(Hn1)|B(R) and FP(R). Hence, choosing ϕ=x2 in (2.10), we obtain

    B1(0)u2dHn1=B1(0)(u)2dHn1,

    which gives, together with (2.9), λ(Γi)λ(Γi), and thus, using the expression of αi(Γi), αi(Γi)αi(Γi), since u is defined on Γi, if u is defined on Γi. The fact that Hn1(Γi)=Hn1(Γi) derives from a property of u which says that

    Hn1(u1[ρ,))=Hn1((u)1[ρ,)),ρR.

    On the other hand, from [24] we achieve that αi(Γi)ψ(si), where si=Hn1(Γi)Hn1(B1(0)) and ψ(s), s(0,1), is convex and decreasing. In particular, ψ(s) is defined as

    ψ(s):={12log14s+32,0<s14,2(1s),14s<1. (2.11)

    Precisely, the proof of αi(Γi)ψ(si) is organized in some steps.

    First of all, we denote α(E)=α(s,n), where α(E) is the characteristic constant of the spherical cap EB1(0), s=Hn1(E)Hn1(B1(0)), and n is the dimension. At this point, Theorem 2 in [24] tells us that α(s,n) is a monotone decreasing function of n for fixed s, so

    α(s,)=limnα(s,n) (2.12)

    is well defined and satisfies α(s,)α(s,n) for every n. It is thus sufficient to show that α(s,)ψ(s) defined in (2.11). To this end, Theorem F in [24], which is taken by [30], says that α(s)ψ(s), where

    s:=he(1/2)t2dt,

    with h=h(α) the largest real zero of

    F(x)=e(1/4)x2Hα(x2)

    satisfying

    d2Fdx2+(α+1214x2)F=0 (2.13)

    and

    F(0)F(0)=21/2Γ(1α2)Γ(α2),

    where Γ is the Euler gamma function. In particular, Hα(x) is the Hermite's function of order α and α(s) here is the α that appears in the equation (2.13).

    Now, Theorem 3 in [24] shows that α(s,) defined in (2.12) is equal to α(s) of Theorem F, since s of α(s,n) converges to s of α(s) as n goes to i.e.,

    Hn1(E)Hn1(B1(0))nhe(1/2)t2dt.

    Hence, being α(s,n)α(s,) for all n, we finally have that α(s,n)ψ(s) for every n and for all s(0,1).

    As a consequence, recalling that si=Hn1(Γi)Hn1(B1(0)), i{1,2}, s1+s2212, because Γ1Γ2=, hence, since ψ(s) defined in (2.11) is convex and decreasing, we get

    α1+α2ψ(s1)+ψ(s2)2(12ψ(s1)+12ψ(s2))2ψ(s1+s22)2ψ(12)=2,

    which finally gives (2.8).

    An alternative proof of this result is given in [9], where, following an unpublished paper by W. Beckner, C. Kenig, J. Pipher and [5], the two authors directly show that α(s1)+α(s2)2, exploiting the properties of α(s) of Theorem F in [24], which is the first Dirichlet eigenvalue on [h,) associated to the Hermite operator

    d2dx2+(14x212).

    We denote by Hn the set R2n+1, nN, n1, endowed with the non-commutative inner law in such a way that for every P(x1,y1,t1)R2n+1, M(x2,y2,t2)R2n+1, xiRn, yiRn, i=1,2:

    PM:=(x1+x2,y1+y2,t1+t2+2(x2,y1x1,y2)),

    where , denotes the usual inner product in Rn. Let Xi=(ei,0,2yi) and Yi=(0,ei,2xi), i=1,,n, where {ei}1in is the canonical basis for Rn.

    We use the same symbol to denote the vector fields associated with the previous vectors, so that for i=1,,n,

    Xi=xi+2yit,Yi=yi2xit.

    The commutator between the vector fields is

    [Xi,Yi]=4t,i=1,,n,

    otherwise is 0. The intrinsic gradient of a smooth function u in a point P is

    Hnu(P)=ni=1(Xiu(P)Xi(P)+Yiu(P)Yi(P)).

    Now, there exists a unique metric on HHnP=span{X1(P),,Xn(P),Y1(P),,Yn(P)} which makes orthonormal the set of vectors {X1,,Xn,Y1,,Yn}. Thus, for every PHn and for every U,WHHnP, U=nj=1(α1,jXj(P)+β1,jYj(P)), V=nj=1(α2,jXj(P)+β2,jYj(P)), we have

    U,V=nj=1(α1,jα2,j+β1,jβ2,j).

    In particular, we get a norm associated with the metric on the space span{X1,,Xn,Y1,,Yn}, which is

    U∣=nj=1(α21,j+β21,j).

    For example, the norm of the intrinsic gradient of a smooth function u in P is

    Hnu(P)∣=ni=1((Xiu(P))2+(Yiu(P))2).

    Moreover, if Hnu(P)0, then

    |Hnu(P)Hnu(P)|=1.

    If Hnu(P)=0, instead, we say that the point P is characteristic for the smooth surface {u=u(P)}. Hence, for every point M{u=u(P)}, which is not characteristic, it is well defined the intrinsic normal to the surface {u=u(P)} as follows:

    ν(M)=Hnu(M)Hnu(M).

    At this point, we introduce in the Heisenberg group Hn the following gauge norm:

    |(x,y,t)|Hn:=4(x2+y2)2+t2.

    In particular, for every positive number r, the gauge ball of radius r centered in 0 is

    BHnr(0):={PHn:|P|Hn<r}.

    In the Heisenberg group, a dilation semigroup is defined as follows: for every r>0 and for every P=(x,y,t)Hn, let

    δr(P):=(rx,ry,r2t).

    Let P:=(ξ,η,σ)Hn and O=(0,0,0), then we define

    dK(P,O):=|P|Hn.

    For every P,THn is well defined

    dK(P,T)=|P1T|Hn,

    that is a distance dK on the Heisenberg group Hn, known as the Koranyi distance. This distance is left invariant, that is for every P,T,RHn

    dK(RP,RT)=dK(P,T).

    As a consequence, we may perform our computation supposing to deal with dK(P,O)=|P|Hn, where O=(0,0,0). This may be obtained simply considering a left translation by T1. In particular, for every i=1,,n we obtain:

    Xi|P|Hn=|P|3Hn((ξ2+η2)ξi+σηi)

    and

    Yi|P|Hn=|P|3Hn((ξ2+η2)ηiσξi).

    Moreover, for every i=1,,n:

    X2i|P|Hn=3|P|7Hn((ξ2+η2)ξi+σηi)2+|P|3Hn(2ξ2i+(ξ2+η2)+2η2i)

    and

    Y2i|P|Hn=3|P|7Hn((ξ2+η2)ηiσξi)2+|P|3Hn(2η2i+(ξ2+η2)+2ξ2i).

    As a consequence,

    Hn|P|Hn2=ni=1((Xi|P|Hn)2+(Yi|P|Hn)2)=(ξ2+η2)|P|2Hn, (3.1)

    and

    ΔHn|P|Hn=(2n+1)(ξ2+η2)|P|3Hn. (3.2)

    Thus, for every i=1,,n, denoting by Q:=2n+2 homogeneous dimension we get:

    Xi|P|2QHn=(2Q)|P|1QHn|P|3Hn((ξ2+η2)ξi+σηi),
    Yi|P|2QHn=(2Q)|P|1QHn|P|3Hn((ξ2+η2)ηiσξi),

    and

    ΔHn|P|2QHn=(2Q)|P|2QHn(ξ2+η2)(1Q+2n+1)=0.

    In conclusion, |P|2QHn is, up to a constant, the fundamental solution of the sublaplacian ΔHn in the Heisenberg group, with the pole in the origin, and Γ(P,R)=c|P1R|2QHn is the fundamental solution of the sublaplacian ΔHn. The definition of Hnsubharmonic function, as well as the one of Hnsuperharmonic function in a set ΩHn, can be stated, as usual, in the classical way, requiring respectively that ΔHnu(P)0 for every PΩ, for the Hnsubharmonicity, and that ΔHnu(P)0 for every PΩ for having Hnsuperharmonicity. We refer to [3] for further details. Concerning the natural Sobolev spaces to consider in the Heisenberg group Hn, we refer to the literature, see for instance [25]. Here, we simply recall that:

    L1,2(Ω):={fL2(Ω):Xif,YifL2(Ω),i=1,,n}

    is a Hilbert space with respect to the norm

    |f|L1,2(Ω)=(Ωni((Xif)2+(Yif)2)+|f|2dx)12.

    Moreover

    H1Hn(Ω)=¯C(Ω)L1,2(Ω)||L1,2(Ω).

    Now, if EHn is a measurable set, a notion of Hn-perimeter measure |E|Hn has been introduced in [25] in a more general setting, even if here we recall some results in the framework of the Heisenberg group, the simplest non-trivial example of Carnot group. We refer to [21,22,23,25] for a detailed presentation. For our applications, we restrict ourselves to remind that, if E has locally finite Hn-perimeter (is a Hn-Caccioppoli set), then |E|Hn is a Radon measure in Hn, invariant under group translations and Hn-homogeneous of degree Q1. Moreover, the following representation theorem holds (see [10]).

    Proposition 3.1. If E is a Hn:=R2n+1-Caccioppoli set with Euclidean C1 boundary, then there is an explicit representation of the Hn-perimeter in terms of the Euclidean 2n-dimensional Hausdorff measure H2n

    PΩ,EHn(E)=EΩ(nj=1(Xj,nE2R2n+1+Yj,nE2R2n+1))1/2dH2n,

    where nE=nE(x) is the Euclidean unit outward normal to E.

    We also have:

    Proposition 3.2. If E is a regular bounded open set with Euclidean C1 boundary and ϕ is a horizontal vector field, continuously differentiable on ¯Ω, then

    EdivHn ϕdx=Eϕ,νHndPEHn,

    where νHn(x) is the intrinsic horizontal unit outward normal to E, given by the (normalized) projection of nE(x) on the fiber HHnx of the horizontal fiber bundle HHn.

    Remark 3.3. The definition of νHn is correctly stated, since HHnx is transversal to the tangent space of E at x, for PEHn(E)-a.e. xE (see [32]).

    Now, adapting the approach described in [1] and recalled in Section 2 to the Heisenberg case, we conclude, by applying the definition of solution in the sense of the domain variation to the functional

    EHn(v):=Ω(|Hnv|2+χ{v>0})dx,

    ΩHn, that the parallel two-phase problem to (1.3) is, see [16]:

    {ΔHnu=0in Ω+(u):={xΩ:u(x)>0},ΔHnu=0in Ω(u):=Int({xΩ:u(x)0}),|Hnu+|2|Hnu|2=1on F(u):=Ω+(u)Ω. (3.3)

    Thus, it seems natural to consider, as a candidate for an Alt-Caffarelli-Friedman monotonicity formula in the Heisenberg group, the following function:

    Jβ,Hn(r)=rβBHnr(0)Hnu+2|ζ|Q2HndζBHnr(0)Hnu2|ζ|Q2Hndζ, (3.4)

    where β>0 is a suitable fixed exponent and u+:=sup{u,0} and u:=sup{u,0}, being 0F(u).

    In this section we mainly discuss some results proved in [18].

    Lemma 4.1. There exists a positive constant c=c(Q) such that for every nonnegative Hnsubharmonic function in C(BHn1(0)), if u(0)=0, then there exists r0 such that for every 0<ρ<r0:

    BHnρ(0)Hnu(ζ)2|ζ|Q2HndζcρQBHn2ρ(0)BHnρ(0)u2(ζ)dζ.

    Lemma 4.2. For every nonnegative Hnsubharmonic functions uiC(BHn1(0)), i=1,2, such that u1u2=0 and u1(0)=u2(0)=0, we have

    Jβ,Hn(1)Jβ,Hn(1)=BHn1(0)Hnu1(κ)2x2+y2dPBHn1(0)Hn(κ)BHn1(0)Hnu1(κ)2|κ|Q2Hndκ+BHn1(0)Hnu2(κ)2x2+y2dPBHn1(0)Hn(κ)BHn1(0)Hnu2(κ)2|κ|Q2Hndκβ.

    Moreover, Jβ,Hn will be monotone increasing in the interval [0,r0), for some r0>0, if and only if Jβ,Hn(1)Jβ,Hn(1)0 for every u1, u2 satisfying the hypotheses of this lemma.

    In order to obtain some estimates of Jβ,Hn(1)Jβ,Hn(1), we need to read the Kohn-Laplace operator ΔHn in terms of radial coordinates. The problem has been faced in [31], by using an abstract and elegant approach, see also [4,29]. In [18] we describe the H1 case in details, with an explicit computation.

    Precisely, we consider the following coordinates in H1:

    T(ρ,φ,θ):={x=ρsinφcosθy=ρsinφsinθt=ρ2cosφ. (4.1)

    From (4.1), we obtain the values of ρ, φ and θ with respect to the cartesian coordinates x, y and t, that is:

    {ρ=((x2+y2)2+t2)1/4θ=arctan(yx)φ=arccos(tρ2). (4.2)

    Recalling the vector fields

    {X=x+2ytY=y2xt, (4.3)

    and the operators:

    H1(X,Y),ΔH1=X2+Y2, (4.4)

    we determine the following: H1ρ, H1θ, H1φ, by using (4.3), (4.2) and (4.4).

    Lemma 4.3. Let ρ,φ,θ defined as in (4.1). Then:

    H1ρ=ρ3((x2+y2)x+ty,(x2+y2)ytx),H1φ=2ρ(x2+y2)(tH1ρ+ρ(y,x))

    and

    H1θ=1x2+y2(y,x).

    In addition, we obtain the properties described in the following lemma.

    Lemma 4.4. Let ρ,φ,θ defined as in (4.1). Then:

    |H1φ|2=4(x2+y2)ρ4,|H1ρ|2=x2+y2ρ2,|H1θ|2=1x2+y2.

    Moreover, it results:

    H1φ,H1ρ=0,H1ρ,H1θ=cosφρ,H1φ,H1θ=2(x2+y2)ρ4

    and

    ΔH1θ=0,ΔH1ρ=3(x2+y2)ρ3,ΔH1φ=4cosφρ2.

    Let now H1u(P)HH1P and define

    eρ:=H1ρ|H1ρ|,andeφ:=H1φ|H1φ|.

    We observe that eρ,eφR2=0, see Lemma 4.4. Then, whenever eρ,eφ exist we have:

    span{eρ(P),eφ(P)}=HH1P.

    As a consequence,

    H1u(P)=H1u(P),eρ(P)eρ(P)+H1u(P),eφ(P)eφ(P)

    and denoting ρH1u(P)=H1u(P),eρ(P)eρ(P) and φH1u(P)=H1u(P),eφ(P)eφ(P), we have

    |H1u(P)|2=H1u(P),eρ(P)2+H1u(P),eφ(P)2

    and

    |H1u(P)|2=|ρH1u(P)|2+|φH1u(P)|2. (4.5)

    We may summarize this fact as follows.

    Lemma 4.5. The couple (H1ρ)(P), (H1φ)(P) determines a basis of HH1P, for every P=(x,y,t), such that x2+y20.

    At this point, assuming that u=ραf(θ,φ), we compute ΔH1u obtaining the following result.

    Lemma 4.6. Let u=ραf(θ,φ), then:

    ΔH1u=ΔH1(ραf(θ,φ))=ρα2(α(α+2)(sinφ)f(θ,φ)2α(cosφ)fθ+1sinφ2fθ2+4sinφ2fφθ+4sinφ2fφ2+4cosφfφ).

    In particular, if u=ραf(φ), then

    ΔH1u=ρα2(α(α+2)(sinφ)f(φ)+4φ(sinφfφ)).

    Thus, whenever f satisfies

    α(α+2)(sinφ)f(θ,φ)2α(cosφ)fθ+1sinφ2fθ2+4sinφ2fφθ+4sinφ2fφ2+4cosφfφ=0

    on ΓBH11(0) or

    α(α+2)(sinφ)f(φ)+4φ(sinφfφ)=0

    on ΓBH11(0) for f depending only on φ, then u=ραf(θ,φ) is H1-harmonic in the set

    PΓ:={(x,y,t)H1:(x,y,t)=δλ(ξ,η,τ),λ>0,(ξ,η,τ)ΓBH11(0)}.

    In fact, if uλ(x,y,t)=u(δλ(x,y,t)), then whenever u=ραf(θ,ϕ), uλ=λαu(x,y,t) and if u is H1harmonic on ΓBH11(0), we obtain:

    ΔH1uλ(x,y,t)=λαΔH1u(x,y,t)=0. (4.6)

    For instance, if Γ={(x,y,t)BH11(0):x2+y2<Mt}, where M>0 is a constant, then

    PΓ={(x,y,t)H1:x2+y2<Mt}.

    Moreover, if we add a boundary condition to the equation

    α(α+2)(sinφ)f(θ,φ)2α(cosφ)fθ+1sinφ2fθ2+4sinφ2fφθ+4sinφ2fφ2+4cosφfφ=0,

    by requiring that f=0 on PΓ, then u=ραf satisfies

    {ΔH1u=0,(x,y,t)PΓ,u=0,(x,y,t)PΓ.

    Of course, if we fix Γ and e assume that f=0 on PΓ as well, then the equation

    α(α+2)(sinφ)f(θ,φ)2α(cosφ)fθ+1sinφ2fθ2+4sinφ2fφθ+4sinφ2fφ2+4cosφfφ=0

    has a solution only for some particular values of α. Indeed, focusing our attention to the case in which f depends only on φ, we clearly obtain the following eigenvalues problem:

    {4(sinφf)=λφ0,φ1sinφf,0φ0<φ<φ1π,f(φ0)=0=f(φ1),

    where Γ:={(x,y,t)BH11(0):cosφ1<t(x2+y2)2+t2<cosφ0}. Hence, the exponent α is related to the first eigenvalue λφ0,φ1 via the relationship:

    λφ0,φ1=α(α+2).

    On the other hand, the first eigenvalue λφ0,φ1 is determined by the Rayleigh quotient given, in this case, by

    λφ0,φ1:=inffH10(φ0,φ1)4φ1φ0sin(φ)f(φ)2dφφ1φ0sin(φ)f(φ)2dφ.

    Thus, it is fundamental to know if the result by [24], that is the cap on B1(0) having the same Hn1 measure of some sets Σ on B1(0) has the smallest Rayleigh quotient, is true even in the Heisenberg case.

    Let say that we would like to know if there exists a set ΓBH11(0) such that for every ΓBH11(0),

    PBH11(0)H1(Γ)=PBH11(0)H1(Γ),

    it results

    αH1(Γ)αH1(Γ),

    where αH1(Γ) denotes the unique positive solution to the equation

    α(α+2)=λ(Γ),

    λ(Γ) is the first eigenvalue of the problem

    {Lθ,φf=λ(Γ)fin ΩR2,f=0on Ω,

    with

    Lθ,φ=divθ,φ(A(θ,φ)θ,φ)

    where T(Ω)=Γ and A(θ,φ) is the matrix-valued function

    A(θ,φ)=[1sinφ(4+2α)sinφ2αsinφ4sinφ]. (4.7)

    In particular, it holds:

    λ(Γ)=inffH10(Ωθ,φ)Ωθ,φ(1sinφ(fθ)2+4sinφfθfφ+4sinφ(fφ)2)dθdφΩθ,φ(sinφ)f2dθdφ. (4.8)

    The existence in the Heisenberg group of the properties of the characteristic number associated with the set Γ, as far as we know, is still unknown. This part corresponds to the topic discussed in [36] in the Euclidean setting. In fact, just for having an idea about the difficulty in solving the problem, we remark that

    PBH11(0)H1(Γ)=Ωsin(φ)dθdφ,

    where Γ=T({1}×Ω). At this point, we may decide to symmetrize the set Ω in many ways. For instance, for every φ, we might define Ωφ in such a way that

    H1(Ωφ)=2θφ=H1(Ωφ),

    and consider Ω:=φΠ2(Ω)Ωφ, being Π2(Ω):={φ:Ωφ}. Unfortunately, the lack of an isoperimetric result does not permit to conclude anything.

    What we can do is to give an estimate. In fact, let

    λφ(Σ):=infvH10(Σ)Σ|φH1v(ξ)|2x2+y2dPBH11(0)H1(ξ)Σv2(ξ)x2+y2dPBH11(0)H1(ξ), (4.9)

    be the Rayleigh quotient, where ΣBH11(0) is a rectifiable set, then the following result holds, see [18].

    Theorem 4.7. Let u1,u2C(BH11(0))H1H1(BH11(0)) be nonnegative, such that u1u2=0 in BH11(0) and ui(0)=0, ΔH1ui0, i=1,2. Then

    2i=1BH11(0)|H1ui(ξ)|2x2+y2dPBH11(0)H1(ξ)BH11(0)|H1ui(ξ)|2|ξ|2H1dξ22i=1(1+λφ(Σi)1), (4.10)

    where Σi=BH11(0){ui>0}.

    In fact, recalling the particular structure described in (4.5), we have:

    |H1u|2=|ρH1u|2+|φH1u|2.

    Thus, as well as in the Euclidean setting, we obtain the following lower bound for each function u:=ui, i=1,2:

    BH11(0)|H1u(ξ)|2x2+y2dPH1(ξ)BH11(0)|H1u(ξ)|2|ξ|2H1dξAρ+AφAu+A1/2uA1/2ρ,

    where

    Aρ:=BH11(0)|ρH1u(ξ)|2x2+y2dPH1(ξ),Aφ:=BH11(0)|φH1u(ξ)|2x2+y2dPH1(ξ),Au:=BH11(0)u2(ξ)x2+y2dPH1(ξ), (4.11)

    having denoted the perimeter measure PBH11(0)H1 simply by PH1 and being ξ=(x,y,t)H1. Hence, recalling the definition (4.9) we conclude. In addition, it results that 1+λφ(Σi)1 is the positive solution of α2+2αλφ(Σi)=0, see [18] for the details.

    In [24], it is proved that for every ΣB1(0) such that Hm1(Σ)=Hm1(Σφ0), where Σφ0 is a cap with width φ0, if fH10(B1(0)) and B1(0)Rm is the Euclidean ball of radius one centered at 0, then

    Σf2dσΣf2dσH(f)=Σφ0f2dσΣφ0(f)2dσ=φ00(f)(φ)2sinm2(φ)dφφ00f(φ)2sinm2(φ)dφH(F)=λE(Σφ0)

    with F the solution of the eigenvalues problem

    {F+(m2)cot(φ)F+λE(Σφ0)F=0,φ]0,φ0[,F(φ0)=0,F(0)=0,

    where λE(Σφ0) is the first eigenvalue and φ0[0,π], see Lemma 1 in [24]. Moreover, see Lemma 2 in [24], the function w=ρα(Σφ0)F is harmonic in the Euclidean cone having as a cap Σφ0 on B1(0), with opening φ0, and where α(Σφ0) is the characteristic number associated with Σφ0, always in the Euclidean framework. In case m=3, we exactly obtain

    λ0,φ0=4λE(Σφ0),

    so that the relationship with αH1 becomes

    α(α+2)=4λE(Σφ0)

    and

    αH1(ΣBH11(0)φ0)=1+4λE(Σφ0)1.

    Thus, we can deduce that the minimum is realized when φ0=π2, and since λE(Σπ2)=2, we conclude that αH1(ΣBH11(0)π2)=2. Furthermore,

    1+4λE(Σφ0)11+4λE(Σ)1,

    nevertheless we cannot conclude, in general, that 1+4λE(Σ)11+λH1(Σ)1, except when the cap on BH11(0) depends only on φ0. In this particular case, for functions having these caps depending only on φ, by choosing β=8 in (3.4), we conclude that Jβ,H1(1)Jβ,H1(1)0 and Lemma 4.2 applies. Indeed, by straightforward computation, by considering the function u=αt+βt, for some positive numbers α,β, and denoting

    Jαt+βt8,H1(r)=r8BH1r(0)α2Hnt+2|ζ|2H1dζBH1r(0)β2H1t2|ζ|2H1dζ,

    it results that

    ddrJαt+βt8,H1(r)=0.

    This is perfectly coherent with the fact that (0,0,0) is a characteristic point of the surface {(x,y,t)BH11(0):t=0}, so that in (0,0,0) it holds H1t(0,0,0)(0,0), as well as we know that

    ddrJαt+βt8,H1(r)=0=∣H1αt+(0,0,0)∣∣H1βt(0,0,0).

    We recall that in this case:

    {(x,y,t)BH11(0):αt+βt=0}={(x,y,t)BH11(0):t=0}.

    We observe here, as a by-product, that αt+βt cannot be a classical solution of (3.3), since the free boundary condition is not fulfilled in (0,0,0). Moreover, we remark that this situation corresponds to the case in which the Koranyi ball is split in two parts separated by the plane {(x,y,t)H1,t=0}.

    On the other hand, whenever we fix α,β0 such that α2β2=1, the function u=αx+βx is a solution of the two-phase problem (3.3), as well as in the Euclidean case, but in this case, testing J8,H1 on this function, we would get that the differential J8,H1 is negative, losing in this way the desired monotonicity property of our formula, see [18].

    In fact, the following result holds.

    Lemma 5.1. For every a,bR, such that a0 or b0, let u=(ax+by)+, defined in BH11(0). Then

    BH11(0)|H1u|2x2+y2dPH1(ξ)BH11(0)|H1u|2|ξ|2H1dξ=2.

    As a consequence of previous Lemma 5.1 and Lemma 4.2, we obtain the proof of Theorem 1.1.

    Proof of Theorem 1.1. From Lemma 4.2 we know that Jβ,H1(1)Jβ,H1(1)0 if and only if

    BH11(0)H1u1(κ)2x2+y2dPBH11(0)H1(κ)BH11(0)H1u1(κ)2|κ|2H1dκ+BH11(0)H1u2(κ)2x2+y2dPBH11(0)H1(κ)BH11(0)H1u2(κ)2|κ|2H1dκβ0. (5.1)

    Let now u1=(ax+by)+ and u2=(ax+by), defined for every a,bR, such that a0 or b0. Then, we invoke Lemma 5.1, concluding that

    BH11(0)H1u1(κ)2x2+y2dPBH11(0)H1(κ)BH11(0)H1u1(κ)2|κ|2H1dκ+BH11(0)H1u2(κ)2x2+y2dPBH11(0)H1(κ)BH11(0)H1u2(κ)2|κ|2H1dκ=4. (5.2)

    Thus, if β>4, then u1=(ax+by)+ and u2=(ax+by) satisfy the hypotheses of Lemma 4.2, but Jβ,H1(1)Jβ,H1(1)<0 when Jβ,H1 is tested on u1=(ax+by)+ and u2=(ax+by). Hence, in order to preserve the increasing monotonicity of Jβ,H1, from (5.2) we are forced to suppose that β4.

    The authors are partially supported by INDAM-GNAMPA-2019 project: Proprietà di regolarità delle soluzioni viscose con applicazioni a problemi di frontiera libera.

    The authors declare no conflict of interest.



    [1] Alt W, Caffarelli L, Friedman A (1984) Variational problems with two phases and their free boundaries. T Am Math Soc 282: 431–461. doi: 10.1090/S0002-9947-1984-0732100-6
    [2] Athanasopoulos I, Caffarelli L, Salsa S (1996) Caloric functions in Lipschitz domains and the regularity of solutions to phase transition problems. Ann Math 143: 413–434. doi: 10.2307/2118531
    [3] Bonfiglioli A, Lanconelli E, Uguzzoni F (2007) Stratified Lie Groups and Potential Theory for Their Sub-Laplacians, Berlin: Springer.
    [4] Birindelli I (2003) Superharmonic functions in the Heisenberg group: estimates and Liouville theorems. NODEA–Nonlinear Diff 10: 171–185.
    [5] Brascamp HJ, Lieb EH (1976) On extensions of the Brunn-Minkowski and Prèkopa-Leindler Theorems, including inequalities for log concave functions, and with an application to the diffusion equation. J Funct Anal 22: 366–389. doi: 10.1016/0022-1236(76)90004-5
    [6] Caffarelli LA (1987) A Harnack inequality approach to the regularity of free boundaries. I. Lipschitz free boundaries are C1,α. Rev Mat Iberoamericana 3: 139–162.
    [7] Caffarelli LA (1988) A Harnack inequality approach to the regularity of free boundaries. III. Existence theory, compactness, and dependence on X. Ann Scuola Norm Sci 15: 583–602.
    [8] Caffarelli LA, Jerison D, Kenig CE (2002) Some new monotonicity theorems with applications to free boundary problems. Ann Math 155: 369–404. doi: 10.2307/3062121
    [9] Caffarelli L, Salsa S (2005) A Geometric Approach to Free Boundary Problems, Providence RI: American Mathematical Society.
    [10] Capogna L, Danielli D, Garofalo N (1994) The geometric Sobolev embedding for vector fields and the isoperimetric inequality. Commun Anal Geom 22: 203–215.
    [11] Capogna L, Danielli D, Pauls S, et al. (2007) An Introduction to the Heisenberg Group and the Sub-Riemannian Isoperimetric Problem, Basel: Birkhauser Verlag.
    [12] Danielli D, Garofalo N, Petrosyan A (2007)The sub-elliptic obstacle problem: C1,a regularity of the free boundary in Carnot groups of step two. Adv Math 211: 485–516.
    [13] Danielli D, Garofalo N, Salsa S (2003) Variational inequalities with lack of ellipticity. I. Optimal interior regularity and non-degeneracy of the free boundary. Indiana U Math J 52: 361–398.
    [14] De Silva D, Ferrari F, Salsa S (2014) Two-phase problems with distributed sources: regularity of the free boundary. Anal PDE 7: 267–310. doi: 10.2140/apde.2014.7.267
    [15] Dipierro S, Karakhanyan AL (2018) A new discrete monotonicity formula with application to a two-phase free boundary problem in dimension two. Commun Part Diff Eq 43: 1073–1101. doi: 10.1080/03605302.2018.1499776
    [16] Dzhugan A, Ferrari F (2020) Domain variation solutions for degenerate elliptic operators. arXiv:2001.07174.
    [17] Folland GB (1975) Subelliptic estimates and function spaces on nilpotent Lie groups. Ark Mat 13: 161–207. doi: 10.1007/BF02386204
    [18] Ferrari F, Forcillo N (2020) Some remarks about the existence of an Alt-Caffarelli-Friedman monotonicity formula in the Heisenberg group. arXiv:2001.04393.
    [19] Ferrari F, Salsa S (2010) Regularity of the solutions for parabolic two-phase free boundary problems. Commun Part Diff Eq 35: 1095–1129. doi: 10.1080/03605301003717126
    [20] Ferrari F, Valdinoci E (2011) Density estimates for a fluid jet model in the Heisenberg group. J Math Anal Appl 382: 448–468. doi: 10.1016/j.jmaa.2011.04.057
    [21] Franchi B, Serapioni R, Cassano FS (1996) Meyers-Serrin type theorems and relaxation of variational integrals depending on vector fields. Houston J Math 22: 859–890.
    [22] Franchi B, Serapioni R, Cassano FS (2003) On the structure of finite perimeter sets in step 2 Carnot groups. J Geom Anal 13: 421–466. doi: 10.1007/BF02922053
    [23] Franchi B, Serapioni R, Cassano FS (2003) Regular hypersurfaces, intrinsic perimeter and implicit function theorem in Carnot groups. Commun Anal Geom 11: 909–944. doi: 10.4310/CAG.2003.v11.n5.a4
    [24] Friedland S, Hayman WK (1976) Eigenvalue inequalities for the Dirichlet problem on spheres and the growth of subharmonic functions. Comment Math Helv 51: 133–161. doi: 10.1007/BF02568147
    [25] Garofalo N, Nhieu DM (1996) Isoperimetric and Sobolev inequalities for Carnot-Carathéodory spaces and the existence of minimal surfaces. Commun Pure Appl Math 49: 1081–1144. doi: 10.1002/(SICI)1097-0312(199610)49:10<1081::AID-CPA3>3.0.CO;2-A
    [26] Garofalo N, Lanconelli E (1990) Frequency functions on the Heisenberg group, the uncertainty principle and unique continuation. Ann I Fourier 40: 313–356. doi: 10.5802/aif.1215
    [27] Gilbarg D, Trudinger NS (2001) Elliptic Partial Differential Equations of Second Order Classics in Mathematics, Berlin: Springer-Verlag.
    [28] Garofalo N, Rotz K (2015) Properties of a frequency of Almgren type for harmonic functions in Carnot groups. Calc Var Partial Dif 54: 2197–2238. doi: 10.1007/s00526-015-0862-x
    [29] Greiner PC (1980) Spherical harmonics on the Heisenberg group. Can Math Bull 23: 383–396. doi: 10.4153/CMB-1980-057-9
    [30] Hayman WK, Ortiz EL (1976) An upper bound for the largest zero of Hermite's function with applications to subharmonic functions. P Roy Soc Edinb A 75: 182–197.
    [31] Jerison DS (1981) The Dirichlet problem for the Kohn Laplacian on the Heisenberg group, II. J Funct Anal 43: 224–257. doi: 10.1016/0022-1236(81)90031-8
    [32] Magnani V (2001) Differentiability and area formula on stratified Lie groups. Houston J Math 27: 297–323.
    [33] Matevosyan N, Petrosyan A (2011) Almost monotonicity formulas for elliptic and parabolic operators with variable coefficients. Commun Pure Appl Math 64: 271–311. doi: 10.1002/cpa.20349
    [34] Noris B, Tavares H, Terracini S, et al. (2010) Uniform Hölder bounds for nonlinear Schrödinger systems with strong competition. Commun Pure Appl Math 63: 267–302. doi: 10.1002/cpa.20309
    [35] Quitalo V (2013) A free boundary problem arising from segregation of populations with high competition. Arch Ration Mech Anal 210: 857–908. doi: 10.1007/s00205-013-0661-5
    [36] Sperner jr E (1973) Zur symmetrisierung von funktionen auf sphären. Math Z 134: 317–327. doi: 10.1007/BF01214695
    [37] Terracini S, Tortone G, Vita S (2018) On s-harmonic functions on cones. Anal PDE 11: 1653– 1691. doi: 10.2140/apde.2018.11.1653
    [38] Terracini S, Verzini G, Zilio A (2016) Uniform Hölder bounds for strongly competing systems involving the square root of the laplacian. J Eur Math Soc 18: 2865–2924. doi: 10.4171/JEMS/656
    [39] Teixeira EV, Zhang L (2011) Monotonicity theorems for Laplace Beltrami operator on Riemannian manifolds. Adv Math 226: 1259–1284. doi: 10.1016/j.aim.2010.08.006
  • This article has been cited by:

    1. Nicola Garofalo, A note on monotonicity and Bochner formulas in Carnot groups, 2022, 0308-2105, 1, 10.1017/prm.2022.58
    2. Fausto Ferrari, Nicolò Forcillo, Alt–Caffarelli–Friedman monotonicity formula and mean value properties in Carnot groups with applications, 2024, 17, 1972-6724, 333, 10.1007/s40574-023-00393-5
    3. Federico Buseghin, Nicolò Forcillo, Nicola Garofalo, A sub-Riemannian maximum modulus theorem, 2024, 1864-8258, 10.1515/acv-2023-0066
    4. Fausto Ferrari, Davide Giovagnoli, Some counterexamples to Alt–Caffarelli–Friedman monotonicity formulas in Carnot groups, 2024, 0373-3114, 10.1007/s10231-024-01490-8
    5. Maria Esteban, The importance of Luis Caffarelli’s work in the study of fluids, 2025, 62, 0273-0979, 301, 10.1090/bull/1861
    6. Sabri Bensid, Analysis of a free boundary problem on stratified Lie group, 2025, 549, 0022247X, 129438, 10.1016/j.jmaa.2025.129438
  • Reader Comments
  • © 2020 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(3589) PDF downloads(309) Cited by(6)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog