Processing math: 100%
Research article Special Issues

Convex duality for principal frequencies

  • We consider the sharp Sobolev-Poincaré constant for the embedding of W1,20(Ω) into Lq(Ω). We show that such a constant exhibits an unexpected dual variational formulation, in the range 1<q<2. Namely, this can be written as a convex minimization problem, under a divergence–type constraint. This is particularly useful in order to prove lower bounds. The result generalizes what happens for the torsional rigidity (corresponding to q=1) and extends up to the case of the first eigenvalue of the Dirichlet-Laplacian (i.e., to q=2).

    Citation: Lorenzo Brasco. Convex duality for principal frequencies[J]. Mathematics in Engineering, 2022, 4(4): 1-28. doi: 10.3934/mine.2022032

    Related Papers:

    [1] Isabeau Birindelli, Kevin R. Payne . Principal eigenvalues for k-Hessian operators by maximum principle methods. Mathematics in Engineering, 2021, 3(3): 1-37. doi: 10.3934/mine.2021021
    [2] Hitoshi Ishii . The vanishing discount problem for monotone systems of Hamilton-Jacobi equations. Part 1: linear coupling. Mathematics in Engineering, 2021, 3(4): 1-21. doi: 10.3934/mine.2021032
    [3] Takeyuki Nagasawa, Kohei Nakamura . Asymptotic analysis for non-local curvature flows for plane curves with a general rotation number. Mathematics in Engineering, 2021, 3(6): 1-26. doi: 10.3934/mine.2021047
    [4] L. Dieci, Fabio V. Difonzo, N. Sukumar . Nonnegative moment coordinates on finite element geometries. Mathematics in Engineering, 2024, 6(1): 81-99. doi: 10.3934/mine.2024004
    [5] Karl K. Brustad, Erik Lindgren, Peter Lindqvist . The infinity-Laplacian in smooth convex domains and in a square. Mathematics in Engineering, 2023, 5(4): 1-16. doi: 10.3934/mine.2023080
    [6] Plamen Stefanov . Conditionally stable unique continuation and applications to thermoacoustic tomography. Mathematics in Engineering, 2019, 1(4): 789-799. doi: 10.3934/mine.2019.4.789
    [7] Pengfei Guan . A weighted gradient estimate for solutions of $ L^p $ Christoffel-Minkowski problem. Mathematics in Engineering, 2023, 5(3): 1-14. doi: 10.3934/mine.2023067
    [8] Antonio Iannizzotto, Giovanni Porru . Optimization problems in rearrangement classes for fractional $ p $-Laplacian equations. Mathematics in Engineering, 2025, 7(1): 13-34. doi: 10.3934/mine.2025002
    [9] Neil S. Trudinger . On the local theory of prescribed Jacobian equations revisited. Mathematics in Engineering, 2021, 3(6): 1-17. doi: 10.3934/mine.2021048
    [10] Lina Zhao, Eun-Jae Park . A locally conservative staggered least squares method on polygonal meshes. Mathematics in Engineering, 2024, 6(2): 339-362. doi: 10.3934/mine.2024014
  • We consider the sharp Sobolev-Poincaré constant for the embedding of W1,20(Ω) into Lq(Ω). We show that such a constant exhibits an unexpected dual variational formulation, in the range 1<q<2. Namely, this can be written as a convex minimization problem, under a divergence–type constraint. This is particularly useful in order to prove lower bounds. The result generalizes what happens for the torsional rigidity (corresponding to q=1) and extends up to the case of the first eigenvalue of the Dirichlet-Laplacian (i.e., to q=2).



    Let ΩRN be an open set, we denote by W1,20(Ω) the completion of C0(Ω) with respect to the Sobolev norm

    φW1,2(Ω)=φL2(Ω)+φL2(Ω;RN), for every   φC0(Ω).

    In what follows, we will always consider for simplicity sets with finite measure. This guarantees that we have at our disposal the Poincaré inequality

    CΩΩ|φ|2dxΩ|φ|2dx, for every   φW1,20(Ω).

    Consequently, the space W1,20(Ω) can be equivalently endowed with the norm

    φW1,20(Ω):=φL2(Ω;RN).

    For 1q2, we consider the generalized principal frequency

    λ1(Ω;q):=minφW1,20(Ω){0}Ω|φ|2dx(Ω|φ|qdx)2q,

    already considered in [26] and more recently in [4,13,35], among others. The fact that the minimum above is attained in W1,20(Ω) follows from the compactness of the embedding W1,20(Ω)Lq(Ω). The latter holds true since we are assuming that Ω has finite measure, see [5,Theorem 2.8].

    Two important cases deserve to be singled-out from the very beginning: q=1 and q=2. In the first case, this quantity actually coincides with the reciprocal of the so-called torsional rigidity of Ω

    1λ1(Ω;1)=T(Ω):=maxφW1,20(Ω){0}(Ω|φ|dx)2Ω|φ|2dx.

    For q=2, on the other hand, the quantity λ1(Ω;2) is nothing but the first eigenvalue of the Dirichlet-Laplacian or principal frequency of Ω. This is the smallest real number λ such that the Helmholtz equation

    Δu=λu, in   Ω,

    admits a nontrivial weak solution in W1,20(Ω). For simplicity, we will simply denote this quantity by λ1(Ω).

    In general, the quantity λ1(Ω;q) can not be explicitly computed. It is then quite useful to seek for (possibly sharp) estimates in terms of geometric quantities of the set Ω. Some particular instances of results in this direction are given by:

    ● the Faber-Krahn inequality (see for example [26,Theorem 2])

    λ1(Ω;q)(λ1(B;q)|B|12N2q)|Ω|12N2q,

    which is valid for every open set ΩRN with finite measure. Here B is any Ndimensional open ball;

    ● the Hersch-Makai–type inequality (see [8,Theorem 1.1])

    λ1(Ω;q)(π2,q2)2|Ω|q2qR2Ω,

    which is valid for every open bounded convex set ΩRN. Here RΩ is the inradius, i.e., the radius of a largest ball contained in Ω and π2,q is the one-dimensional constant defined by

    π2,q=infφW1,20((0,1)){0}φL2((0,1))φLq((0,1)).

    This inequality is the extension to the range 1q2 of [28,equation (3')] by Makai (case q=1) and of [22,Théorème 8.1] by Hersch (case q=2);

    ● the Pólya–type inequality (see [6,Main Theorem])

    λ1(Ω;q)(π2,q2)2(P(Ω)|Ω|12+1q)2,

    which is again valid for open bounded convex sets. Here P(Ω) stands for the perimeter of Ω. This inequality generalizes the original result by Pólya [31] for the cases q=1 and q=2.

    We point out that all the previous estimates are sharp. All the exponents appearing above are of course dictated by scale invariance.

    As a general rule, we can assert that lower bounds on λ1(Ω;q) are harder to obtain with respect to upper bounds, since every generalized principal frequency is defined as an infimum. It would then be interesting to investigate whether λ1(Ω;q) admits a sort of "dual'' equivalent formulation, in terms of a supremum. This is the main goal of the present paper.

    At this aim, it is interesting to have a closer look at the case q=1. It is well-known that the torsional rigidity can be equivalently rewritten as an unconstrained concave maximization problem, i.e.,

    maxφW1,20(Ω){2ΩφdxΩ|φ|2dx}=T(Ω). (1.1)

    As such, it admits in a natural way a dual convex minimization problem

    minϕL2(Ω;RN){Ω|ϕ|2dx:divϕ=1 in   Ω}=maxφW1,20(Ω){2ΩφdxΩ|φ|2dx}=T(Ω), (1.2)

    which gives yet another equivalent definition of torsional rigidity. Here the divergence constraint has to be intended in distributional sense, i.e.,

    Ωϕ,φdx=Ωφdx, for every   φC10(Ω). (1.3)

    By means of a standard density argument, it is easily seen that we can enlarge the class of competitors in (1.3) to the whole W1,20(Ω), since ϕL2(Ω;RN).

    For a better understanding of the contents of this paper, it may be useful to recall the proof of (1.2). At first, one observes that for every admissible vector field and every φW1,20(Ω), we have

    2ΩφdxΩ|φ|2dx=2Ωϕ,φdxΩ|φ|2dx,

    by virtue of (1.3). We can now use Young's inequality

    2ϕ,φ|φ|2|ϕ|2.

    By integrating this inequality, from the identity above we get

    2ΩφdxΩ|φ|2dxΩ|ϕ|2dx.

    The arbitrariness of both φ and ϕ automatically gives

    maxφW1,20(Ω){2ΩφdxΩ|φ|2dx}minϕL2(Ω;RN){Ω|ϕ|2dx:divϕ=1 in   Ω}.

    On the other hand, by taking φ=w to be the optimal function for the problem on the left-hand side, this satisfies the relevant Euler-Lagrange equation. The latter is given by

    Δw=1, in   Ω.

    Thus ϕ0=w is an admissible vector field and we have

    Ω|ϕ0|2dx=2Ωw,ϕ0dxΩ|w|2=2ΩwdxΩ|w|2dx.

    This proves that

    minϕL2(Ω;RN){Ω|ϕ|2dx:divϕ=1 in   Ω}maxφW1,20(Ω){2ΩφdxΩ|φ|2dx},

    as well. Thus (1.2) holds true and we also have obtained that the unique (by strict convexity) minimal vector field ϕ0L2(Ω;RN) must be of the form

    ϕ0=w,

    with w being the unique W1,20(Ω) solution of Δw=1. In conclusion, getting back to the notation λ1(Ω;1), we obtain the following dual characterization of the relevant generalized principal frequency

    1λ1(Ω;1)=minϕL2(Ω;RN){Ω|ϕ|2dx:divϕ=1 in   Ω}. (1.4)

    The main result of the present paper asserts that the dual characterization (1.4) is not an isolated exception. Actually, it is possible to prove that

    1λ1(Ω;q),

    coincides with the minimum of a constrained convex minimization problem, for the whole range 1q2. The deep reason behind this result is a hidden convex structure of the problem which defines λ1(Ω;q), see Remark 4.2 below. Such a convex structure, which apparently is still not very popular, fails for q>2 and this explains why our result has 1q2 as the natural range of validity.

    In order to precisely state the result, we need at first to introduce the following convex lower semicontinuous function Gq:R×RN[0,+], defined for 1<q2 by:

    Gq(s,ξ)={|ξ|q|s|q1, if ξRN,s<0,0, if s=0,ξ=0,+, otherwise, (1.5)

    see [30,Lemma 5.17]. We then distinguish between the cases q<2 and q=2.

    Theorem 1.1 (Sub-homogeneous case). Let 1<q<2 and let ΩRN be an open set, with finite measure. If we set

    A(Ω)={(f,ϕ)L1loc(Ω)×L2loc(Ω;RN):divϕ+f1in  Ω},

    then we have

    1λ1(Ω;q)=(q1)(q1)2qinf(f,ϕ)A(Ω)Gq(f,ϕ)2qL22q(Ω), (1.6)

    where Gq is defined in (1.5). Moreover, if wW1,20(Ω) denotes the unique positive solution of

    maxφW1,20(Ω){2qΩ|φ|qdxΩ|φ|2dx},

    we get that the pair (f0,ϕ0) defined by

    ϕ0=wwq1and  f0=(q1)|w|2wq,

    is a minimizer for the problem in (1.6).

    Remark 1.2. Observe that the previous result is perfectly consistent with the case q=1. Indeed, by formally taking the limit as q goes to 1 in the statement above, the role of the dual variable f becomes immaterial and we get back (1.2), together with the optimality condition ϕ0=w.

    For the limit case q=2, corresponding to the first eigenvalue of the Dirichlet-Laplacian, we have the following dual characterization.

    Theorem 1.3 (Homogeneous case). Let ΩRN be an open set, with finite measure. If we set

    A(Ω)={(f,ϕ)L1loc(Ω)×L2loc(Ω;RN):divϕ+f1in  Ω},

    then we have

    1λ1(Ω)=inf(f,ϕ)A(Ω)G2(f,ϕ)L(Ω), (1.7)

    where G2 is defined in (1.5). Moreover, if we denote by UW1,20(Ω) any positive first eigenfunction of Ω, we get that the pair (f0,ϕ0) defined by

    ϕ0=1λ1(Ω)UUand  f0=1λ1(Ω)|U|2U2,

    is a minimizer for the problem in (1.7).

    Remark 1.4. It may be worth recalling that the existence of a (sort of) dual formulation for λ1 is not a complete novelty. A related result can be traced back in the literature and attributed to the fundamental contributions of Protter and Hersch. This is called maximum principle for λ1 and reads as follows

    λ1(Ω)=maxϕinfxΩ[divϕ(x)|ϕ(x)|2],

    under suitable regularity assumptions on Ω and on the admissible vector fields. It is not difficult to see that

    ϕ0(x)=UU,

    is a maximizer for the previous problem, at least formally. Here U is again any positive first eigenfunction of Ω. We refer to the paper [21] by Hersch for a presentation of this result and for a detailed discussion about its physical interpretation. \end{oss}

    Remark 1.5. As a last observation, we wish to point out the interesting papers [18] and [19], where yet another equivalent characterization for the torsional rigidity T(Ω) is obtained, when ΩR2 is a simply connected open set. Such a characterization is in terms of a minimization problem among holomorphic functions (see [18,Theorem 1.2]) and thus it is suitable for giving upper bounds on T(Ω) (see [19]).

    We start by exposing some preliminary facts in Section 2. In Section 3 we consider a certain convex function and show that its Legendre-Fenchel transform is related to the function Gq above. The core of the paper is Section 4, where Proposition 4.1 permits to rewrite the value λ1(Ω;q) as an unconstrained concave maximization problem, exactly as in the case of the torsional rigidity. We can then prove our main results in Section 5. Finally, in the last section we briefly show some applications of our results to geometric estimates for principal frequencies.

    We first recall that it is possible to rewrite the minimization problem which defines λ1(Ω;q) as an unconstrained optimization problem, in the regime 1q<2. This generalizes formula (1.1). The proof is standard, we include it for completeness.

    Proposition 2.1. Let 1q<2 and let ΩRN be an open set, with finite measure. Then we have

    maxφW1,20(Ω){2qΩ|φ|qdxΩ|φ|2dx}=2qq(1λ1(Ω;q))q2q.

    Moreover, the maximization problem on the left-hand side admits a unique non-negative solution w, which has the following properties

    wL(Ω)and  1wLloc(Ω).

    Proof. Existence of a maximizer follows by a standard application of the Direct Method in the Calculus of Variations. The fact that a non-negative maximizer exists is a consequence of the fact that the functional is even, thus we can always replace φ by |φ| without decreasing the energy.

    We also observe that for φW1,20(Ω){0} and t>0, the quantity

    2qtqΩ|φ|qdxt2Ω|φ|2dx,

    is strictly positive for t sufficiently small. This shows that

    maxφW1,20(Ω){2qΩ|φ|qdxΩ|φ|2dx}>0,

    and thus φ0 can not be a maximizer. Observe that the same argument, together with the locality of the functional, imply that for any maximizer w we must have w0 on every connected component of Ω. By coupling this information with the optimality condition, we get that any non-negative maximizer w must be a nontrivial weak solution of the Euler-Lagrange equation

    Δw=wq1, in   Ω.

    In particular, w is a weakly superharmonic function and by the strong minimum principle, we get that 1/wLloc(Ω). The fact that wL(Ω) follows from standard Ellipic Regularity.

    Finally, uniqueness of the positive maximizer can be found in [7,Lemma 2.2], where the uniqueness result of [9,Theorem 1] is extended to the case of open sets, not necessarily smooth.

    In order to prove the claimed equality between the extremum values, it is sufficient to exploit the different homogeneities of the two integrals and the fact that the maximum problem is equivalently settled on W1,20(Ω){0}. We then have

    maxφW1,20(Ω){2qΩ|φ|qdxΩ|φ|2dx}=maxφW1,20(Ω){0},t>0{2qtqΩ|φ|qdxt2Ω|φ|2dx}.

    It is easily seen that, for every φW1,20(Ω){0} the function

    t2qtqΩ|φ|qdxt2Ω|φ|2dx,

    is maximal for

    t0=(Ω|φ|qdxΩ|φ|2dx)12q.

    With such a choice of t, we get

    2qtq0Ω|φ|qdxt20Ω|φ|2dx=((Ω|φ|qdx)2qΩ|φ|2dx)q2q2qq.

    By recalling the definition of λ1(Ω;q), we get the desired conclusion.

    We also record the following technical result: this will be useful somewhere in the paper. More sophisticated results about the dependence on q of the quantity λ1(Ω;q) can be found in [1,Theorem 1] and [17].

    Lemma 2.2. Let ΩRN be an open set, with finite measure. Then we have

    limq1+λ1(Ω;q)=λ1(Ω;1)and  limq2λ1(Ω;q)=λ1(Ω).

    Proof. For 1<q<2 and for every φW1,20(Ω){0}, we have by Hölder's inequality

    Ω|φ|2dx(Ω|φ|qdx)2q|Ω|12qΩ|φ|2dxΩ|φ|2dx|Ω|12qλ1(Ω).

    By taking the infimum over φ, this leads to

    λ1(Ω;q)|Ω|12qλ1(Ω).

    On the other hand, if UW1,20(Ω) is any minimizer for λ1(Ω), then we have

    λ1(Ω;q)Ω|U|2dx(Ω|U|qdx)2q=Ω|U|2dx(Ω|U|qdx)2qλ1(Ω).

    The last two displays eventually prove the desired result for q converging to 2. The other result can be proved in exactly the same way.

    Remark 2.3. The assumption on the finiteness of the measure is sufficient, but in general not necessary, for the previous result to hold. However, as observed in [6,Remark 2.2], for a general open set ΩRN it may happen that

    lim supq2λ1(Ω;q)<λ1(Ω).

    In order to prove the main result of this paper, we will need to study a particular convex function Fq:R×RN[0,+] and its Legendre-Fenchel transform

    Fq(s,ξ)=sup(t,x)R×RN[st+ξ,xFq(t,x)].

    We refer to the classical monograph [32] for the basic properties of this transform.

    Lemma 3.1. Let 1<q<2, we consider the convex lower semicontinuous function F:R×RNR{+} defined by

    Fq(t,x)={|x|2t2q2,if  xRN,t>0,0,if  t=0,x=0,+,otherwise.

    Then its Legendre-Fenchel transform is given by the convex lower semicontinuous function

    Fq(s,ξ)={αq|ξ|2q2q|s|2(1q)2q,if  ξRN,s<0,0,if  s=0,ξ=0,+,otherwise,

    where the constant αq is given by

    αq=2q2q(q1q)2(q1)2q(12)q2q.

    Proof. We divide the proof in various parts, according to the claim that we are going to prove.

    Lower semicontinuity. In order to verify the semicontinuity of Fq, we need to prove that the epigraph

    epi(Fq)={((t,x);)RN+1×R:Fq(t,x)},

    is a closed set. We take {((tn,xn);n)}nNepi(Fq) such that

    limntn=t,limnxn=x,limnn=.

    By using the definition of Fq and that of epigraph, the fact that

    Fq(tn,xn)n, for every   nN, (3.1)

    automatically entails that

    {(tn,xn)}nN((0,+)×RN){(0,0)}.

    This in particular implies that the limit point t is such that t0. The same can be said for , since Fq always takes positive values.

    We now observe that if t>0, we would have tn>0 for n large enough. In this case, we can simply pass to the limit in (3.1) and get

    Fq(t,x)=|x|2t2q2=limn|xn|2t2q2nlimnn=,

    thus proving that ((t,x);)epi(Fq).

    Let us now suppose that t=0 and assume by contradiction that ((0,x);)epi(Fq). This means that

    Fq(0,x)>.

    By recalling that 0 and that Fq(0,0)=0, this would automatically gives that x0. On the other hand, by (3.1), we get that

     either tn=0  and   xn=0 or tn>0   and   |xn|2nt22qn.

    This entails that

    x=limnxn=0,

    which gives a contradiction. This finally proves that the epigraph is closed.

    Convexity. We need to prove that for every t0,t1R, x0,x1RN and λ[0,1], we have

    Fq(λt0+(1λ)t1,λx0+(1λ)x1)λFq(t0,x0)+(1λ)Fq(t1,x1). (3.2)

    We observe that for t0,t10, every x0,x1RN{0} and every λ[0,1], we trivially have (3.2), since both terms on the right-hand side are equal to +. We are thus confined to prove (3.2) for

    (t0,x0),(t1,x1)((0,+)×RN){(0,0)}.

    Moreover, if at least one between (t0,x0) and (t1,x1) coincides with (0,0), then again the desired inequality follows by a straighforward computation. Finally, we can assume that

    (t0,x0),(t1,x1)(0,+)×RN.

    We introduce the function

    F2(t,x)=|x|2t1, if xRN,t>0,

    and we observe that for every 1<q<2 we have

    Fq(t,x)=F2(t22q,x), for (t,x)(0,+)×RN.

    By using that tF2(t,x) is decreasing, that tt22q is concave (since 1<q<2) and that (t,x)F2(t,x) is convex (see for example [30,Lemma 5.17]), we get

    Fq(λt0+(1λ)t1,λx0+(1λ)x1)=F2((λt0+(1λ)t1)22q,λx0+(1λ)x1)F2(λt22q0+(1λ)t22q1,λx0+(1λ)x1)λF2(t22q0,x0)+(1λ)F2(t22q1,x1)=λFq(t0,x0)+(1λ)Fq(t1,x1),

    as desired.

    Computation of Fq. We now come to the computation of the Legendre-Fenchel transform. This is lengthy but elementary. We first observe that Fq is positively 2/qhomogeneous, that is for every τ>0 we have

    Fq(τt,τx)=τ2qFq(t,x), for every   (t,x)R×RN.

    Correspondingly, Fq will be positively 2/(2q)homogeneous, by standard properties of the Legendre-Fenchel transform. Thanks to this remark, it is sufficient to compute for ξRN

    Fq(1,ξ),Fq(0,ξ) and Fq(1,ξ).

    By definition, we have

    Fq(s,ξ)=sup(t,x)R×RN[ts+x,ξFq(t,x)]=supt0,xRN[ts+|x||ξ|Fq(t,x)]=sup(t,m)E[ts+m|ξ|m2t2q2],

    where we set*

    * For notational simplicity, we use the convention that m2t2q2=0 when both t=0 and m=0.

    E={(t,m)R×R:t>0,m0}{(0,0)}.

    We thus easily get for ξRN

    Fq(1,ξ)=sup(t,m)E[t+m|ξ|m2t2q2]=+,

    and

    Fq(0,0)=sup(t,m)E[m2t2q2]=0.

    Moreover, for ξ0 we have

    Fq(0,ξ)=sup(t,m)E[m|ξ|m2t2q2]=+,

    as can be seen by taking

    t=nN and m=12|ξ|n22q,

    and letting n go to +. We are left with computing for ξRN

    Fq(1,ξ)=sup(t,m)E[t+m|ξ|m2t2q2].

    We observe at first that we easily have

    Fq(1,0)=sup(t,m)E[tm2t2q2]=0.

    We thus take ξRN{0} and we make a preliminary observation: if (t,m)E are such that

    m>0 and t=|ξ|2m,

    we get

    Fq(1,ξ)|ξ|2m(|ξ|2)2q2m2q,

    and the last quantity can be made strictly positive, for m>0 small enough, thanks to the fact that 2/q>1. On the contrary, every point (t,0)E can not be a maximizer for the problem which defines Fq(1,ξ), since on these points

    t+m|ξ|m2t2q2=t0.

    This simple observation implies that we can rewrite the maximization problem for Fq(1,ξ) as

    Fq(1,ξ)=supt>0,m>0[t+m|ξ|m2t2q2].

    Moreover, this quantity is strictly positive. In order to explicitly compute it, we will exploit the homogeneity of the function (t,m)m2t2q2. Indeed, we first observe that by taking λ>0 and replacing (t,m) by (λt,λm) we get

    Fq(1,ξ)=supt>0,m>0,λ>0[λt+λm|ξ|λ2qm2t2q2].

    Now we observe that the derivative of the function

    h(λ)=λt+λm|ξ|λ2qm2t2q2,

    is given by

    h(λ)=t+m|ξ|2qλ2q1m2t2q2.

    We now distinguish two cases: if m|ξ|t0, the previous computation shows that h is decreasing on (0,+) and thus

    h(λ)limλ0+h(λ)=0.

    On the other hand, if m|ξ|t>0, then we get that h has a unique maximum point at

    λ0=(q2m|ξ|tm2t2q2)q2q,

    thus

    h(λ)(q2m|ξ|tm2t2q2)q2q(t+m|ξ|)(q2m|ξ|tm2t2q2)22qm2t2q2=(m|ξ|tmqt1q)22q(q2)q2q2q2.

    This discussion entails that

    Fq(1,ξ)=(q2)q2q2q2supt>0,m>0{(m|ξ|tmqt1q)22q:m|ξ|>t}.

    We are left with computing such a supremum. We may notice that the objective function only depends on the ratio t/m, indeed we have

    m|ξ|tmqt1q=(tm)q(mt|ξ|1).

    Thus, if we set τ=t/m, we finally arrive at the problem

    Fq(1,ξ)=(q2)q2q2q2supτ>0{(τq(|ξ|τ1))22q:|ξ|>τ}.

    It is easily seen that the function

    f(τ)=τq(|ξ|τ1),

    is maximal in the interval (0,|ξ|) for

    τ0=q1q|ξ|.

    Thus we obtain

    supτ>0{τq(|ξ|τ1):|ξ|>τ}=1q(q1q)q1|ξ|q,

    which eventually leads to

    Fq(1,ξ)=(q2)q2q2q2(1q(q1q)q1)22q|ξ|2q2q.

    Thanks to the positive homogeneity of Fq already discussed, we get the desired conclusion.

    Remark 3.2 (Relation between Fq and Gq). From the previous result, we get more generally that for every C>0, we have

    (CFq)(s,ξ)=CFq(sC,ξC)=Cq2qFq(s,ξ).

    This easily follows from the properties of the Legendre-Fenchel transform, together with the fact that Fq is 2/(2q)positively homogeneous. In particular, by taking C=1/(2q), we have

    (12qFq)(s,ξ)=(2q)q2qFq(s,ξ).

    By recalling the definition (1.5) of Gq, we easily see that

    (Gq(s,ξ))22q=1αqFq(s,ξ),

    and thus

    (12qFq)(s,ξ)=(2q)q2qFq(s,ξ)=(2q)q2qαq(Gq(s,ξ))22q.

    Finally, by using that

    αq=2q2q(q1q)2(q1)2q(12)q2q,

    we get the relation

    (12qFq)(s,ξ)=2q2(q1)(q1)22q(Gq(s,ξ))22q. (3.3)

    We are going to use this identity in the proof of the main result.

    By combining Proposition 2.1 and the convexity of the function Fq above, we can rewrite the variational problem which defines λ1(Ω;q) as a concave optimization problem. This property is crucial for the proof of Theorem 1.1.

    Proposition 4.1. Let 1<q<2 and let ΩRN be an open set, with finite measure. We define the following subset of W1,20(Ω)

    Xq(Ω)={ψW1,20(Ω)L(Ω):ΩFq(ψ,ψ)dx<+}.

    Then Xq(Ω) is convex and we have

    2qq(1λ1(Ω;q))q2q=maxφW1,20(Ω){2qΩ|φ|qdxΩ|φ|2dx}=supψXq(Ω){2qΩψdx1q2ΩFq(ψ,ψ)dx}. (4.1)

    Finally, the last supremum is attained by a function vXq(Ω) of the form

    v=wq,

    where w is the same as in Proposition 2.1.

    Proof. Convexity of Xq(Ω) immediately follows from the convexity of the function Fq. We now come to the proof of (4.1). The first identity is already contained in Proposition 2.1. Let us take wW1,20(Ω)L(Ω) to be the positive maximizer of

    maxφW1,20(Ω){2qΩ|φ|qdxΩ|φ|2dx}.

    We now set v=wq and observe that vW1,20(Ω)L(Ω), since v is the composition of a function in W1,20(Ω)L(Ω) with a locally Lipschitz function, vanishing at the origin. From the chain rule in Sobolev spaces, we get

    v=qwq1w=qvq1qw,

    where we also used the relation between w and v, to replace wq1. Since w>0 in Ω, we have the same property for v, as well. Thus we can infer

    w=1qv1q1v, a.\, e. in   Ω.

    By raising to the power 2 and integrating, we get

    Ω|w|2dx=1q2Ω|v|2v2q2dx=1q2ΩFq(v,v)dx,

    which shows that vXq(Ω). By recalling that w is optimal, this also shows that

    maxφW1,20(Ω){2qΩ|φ|qdxΩ|φ|2dx}supψXq(Ω){2qΩψdx1q2ΩFq(ψ,ψ)dx}.

    On the other hand, let ψXq(Ω). Thanks to the form of the function Fq, this in particular implies that

    ψ(x)0, for a. e.   xΩ.

    For every ε>0, we introduce the C1 function

    gε(τ)=(εq+τ)1qε, for every   τ0.

    Then we set φε=gεψ and observe that φεW1,20(Ω), thanks to the fact that gε is C1 with bounded derivative and gε(0)=0. Again by the chain rule, we have

    φε=gε(ψ)ψ=1q(εq+ψ)1q1ψ.

    By raising to the power 2 and integrating, we get

    Ω|φε|2dx=1q2Ω(εq+ψ)2q2|ψ|2dx1q2ΩFq(ψ,ψ)dx.

    In the last inequality, we used the well-known fact that ψ vanishes almost everywhere on the zero set of the Sobolev function ψ (see for example [27,Theorem 6.19]). This in turn implies that

    maxφW1,20(Ω){2qΩ|φ|qdxΩ|φ|2dx}2qΩ|φε|qdxΩ|φε|2dx2qΩgε(ψ)qdx1q2ΩFq(ψ,ψ)dx.

    It is only left to pass to the limit as ε goes to 0 in the integral containing gε(ψ). This can be done by a standard application of the Lebesgue Dominated Convergence Theorem. This finally leads to

    maxφW1,20(Ω){2qΩ|φ|qdxΩ|φ|2dx}2qΩψdx1q2ΩFq(ψ,ψ)dx.

    By arbitrariness of ψXq(Ω), we eventually get the desired conclusion (4.1).

    The above discussion also prove the last statement, about a maximizer of the problem settled over Xq(Ω).

    Remark 4.2. The previous result is crucially based on the fact that the Dirichlet integral, apart from being convex in the usual sense, enjoys a suitable form of "hidden'' convexity. In other words, for φ positive we have that

    φΩ|φ|2dx,

    remains convex also with respect to the new variable ψ=φq, for the whole range 1q2. In the limit case q=2, this remarkable fact has been proved by Benguria, see [2,Theorem 4.3] and [3,Lemma 4]. For 1<q<2 this property seems to have been first detected in [25,Proposition 4], see also [29,Proposition 1.1] and [34,Example 5.2].

    Actually, we can restrict the maximization to smooth compactly supported functions, without affecting the value of the supremum. This is the content of the following result.

    Lemma 4.3. With the notation of Proposition 4.1, we have

    supψXq(Ω){2qΩψdx1q2ΩFq(ψ,ψ)dx}=supψXq(Ω)C10(Ω){2qΩψdx1q2ΩFq(ψ,ψ)dx}.

    Proof. We just need to prove that

    supψXq(Ω){2qΩψdx1q2ΩFq(ψ,ψ)dx}supψXq(Ω)C10(Ω){2qΩψdx1q2ΩFq(ψ,ψ)dx}.

    We take φC0(Ω) not identically zero and set ψ=|φ|qC10(Ω). As above, we have

    |ψ|=q|φ|q1|φ|=qψq1q|φ|,

    which holds everywhere on Ω. This in particular implies that ψ vanishes on every point where ψ vanishes. Thus we have

    Fq(ψ,ψ)={|ψ|2ψ2q2, if   ψ0,0, if   ψ=0.

    By integrating and recalling the relation above between ψ, ψ and φ, we then obtain

    ΩFq(ψ,ψ)dxq2Ω|φ|2dx.

    This in turn implies

    2qΩψdx1q2ΩFq(ψ,ψ)dx2qΩ|φ|qdxΩ|φ|2dx.

    By arbitrariness of φC0(Ω), we get

    supψXq(Ω)C10(Ω){2qΩψdx1q2ΩFq(ψ,ψ)dx}supφC0(Ω){2qΩ|φ|qdxΩ|φ|2dx}. (4.2)

    On the other hand, by density of C0(Ω) in W1,20(Ω), it is easily seen that

    supφC0(Ω){2qΩ|φ|qdxΩ|φ|2dx}=maxφW1,20(Ω){2qΩ|φ|qdxΩ|φ|2dx}.

    The desired conclusion now follows by combining Proposition 4.1 and (4.2).

    We recall the definition

    A(Ω)={(f,ϕ)L1loc(Ω)×L2loc(Ω;RN):divϕ+f1 in   Ω},

    where the condition divϕ+f1 has to be intended in distributional sense, i.e.,

    Ω[ϕ,ψ+fψ]dxΩψdx, for every   ψC10(Ω) such that ψ0.

    In particular, for every (f,ϕ)A(Ω) and every ψXq(Ω)C10(Ω), we can write

    2qΩψdx1q2ΩFq(ψ,ψ)dx2qΩ[fψ+ϕ,ψ]dx1q2ΩFq(ψ,ψ)dx=2qΩ[fψ+ϕ,ψ12qFq(ψ,ψ)]dx.

    By Lemma 3.1 and Eq (3.3) from Remark 3.2, the following inequality holds almost everywhere

    fψ+ϕ,ψ12qFq(ψ,ψ)2q2(q1)(q1)22q(Gq(f,ϕ))22q.

    This simply follows from the definition of Legendre-Fenchel transform. By integrating this inequality and taking the supremum over ψ, we obtain

    supψXq(Ω)C10(Ω){2qΩψdx1q2ΩFq(ψ,ψ)dx}2qq(q1)(q1)22qΩ(Gq(f,ϕ))22qdx.

    By combining Proposition 4.1 and Lemma 4.3 and taking the infimum over admissible pairs (f,ϕ), we get

    1λ1(Ω;q)(q1)(q1)2qinf(f,ϕ)A(Ω)(Ω(Gq(f,ϕ))22qdx)2qq.

    In order to prove the reverse inequality and identify a minimizing pair, we take

    ϕ0=wwq1 and f0=(q1)|w|2wq,

    where w is the function of Proposition 2.1. In light of the properties of w, both ϕ0 and f0 have the required integrability properties. Moreover, it is not difficult to see that these are admissible, since it holds

    divϕ0+f0=1, in   Ω,

    in distributional sense. It is sufficient to use the equation solved by w. By recalling the definition (1.5) of Gq, when w0 we have

    Gq(f0,ϕ0)=|ϕ0|q|f0|1q=(q1)1q|w|2q.

    On the other hand, since both f0 and ϕ0 vanish when w=0, in this case we have that Gq(f0,ϕ0) would vanish, as well. In conclusion, we get

    Ω(Gq(f0,ϕ0))22qdx=(q1)2(1q)2qΩ|w|2dx.

    Thus we obtain

    (q1)(q1)2qinf(f,ϕ)A(Ω)(Ω(Gq(f,ϕ))22qdx)2qq(Ω|w|2dx)2qq.

    We can now use that by Proposition 2.1

    Ω|w|2dx=q2q(2qΩ|w|qdxΩ|w|2dx)=q2qmaxφW1,20(Ω){2qΩ|φ|qdxΩ|φ|2dx}=(1λ1(Ω;q))q2q.

    The first identity above follows by testing the Euler-Lagrange equation for w, i.e.,

    Ωw,φdx=Ωwq1φdx, for every   φW1,20(Ω),

    with φ=w itself. This finally proves that the reverse inequality holds

    (q1)(q1)2qinf(f,ϕ)A(Ω)(Ω(Gq(f,ϕ))22qdx)2qq1λ1(Ω;q),

    as well. The second part of the proof also proves that (f0,ϕ0) is an optimal pair.

    In order to prove the inequality

    1λ1(Ω)inf(f,ϕ)A(Ω)G2(f,ϕ)L(Ω), (5.1)

    we will go through an approximation argument, for simplicity. Let (f,ϕ)A(Ω) be an admissible pair, we can suppose that

    M:=G2(f,ϕ)L(Ω)<+.

    Thanks to the definition of G2, this implies in particular that the pair (f,ϕ) has the following properties:

    |f(x)|=0 implies that |ϕ(x)|=0, for a. e.   xΩ,

    and

    |ϕ(x)|2M|f(x)|, for a. e.   xΩ.

    We thus obtain for every 1<q<2

    |ϕ|q|f|q1Mq2|f|2q2L22qloc(Ω),

    and thus

    (Gq(f,ϕ))22qMq2q|f|L1loc(Ω).

    This estimate guarantees that for every open set Ω compactly contained in Ω, we have

    (Ω(Gq(f,ϕ))22qdx)2qqM(Ω|f|dx)2qq=G2(f,ϕ)L(Ω)(Ω|f|dx)2qq. (5.2)

    By Theorem 1.1 applied to Ω, we have for every 1<q<2

    1λ1(Ω;q)(q1)(q1)2qGq(f,ϕ)2qL22q(Ω).

    By using (5.2) on the right-hand side and taking the limit as q goes to 2, we get

    1λ1(Ω)G2(f,ϕ)L(Ω), (5.3)

    by virtue of Lemma 2.2. We can now take an increasing sequence of open sets {Ωn}nN compactly contained in Ω and invading it, i.e., such that

    Ω=nNΩn.

    By using that

    This simply follows from the properties of the sequence {Ωn}nN and the fact that

    λ1(Ω)=minφW1,20(Ω){0}Ω|φ|2dxΩ|φ|2dx=infφC0(Ω){0}Ω|φ|2dxΩ|φ|2dx.
    limnλ1(Ωn)=λ1(Ω),

    and applying (5.3) to each Ωn, we get

    1λ1(Ω)G2(f,ϕ)L(Ω).

    By arbitrariness of (f,ϕ)A(Ω), we get (5.1) as desired.

    In order to prove the reverse inequality, we take

    ϕ0=1λ1(Ω)UU and f0=1λ1(Ω)|U|2U2,

    where U is any positive first eigenfunction of Ω. It is easily seen that this pair is admissible for the variational problem

    inf(f,ϕ)A(Ω)G2(f,ϕ)L(Ω).

    Moreover, we have

    G2(f0,ϕ0)1λ1(Ω), a. e. in Ω,

    where we also used that f0 vanishes if and only if ϕ0 vanishes and in this case G2(f0,ϕ0)=0. This gives

    inf(f,ϕ)A(Ω)G2(f,ϕ)L(Ω)G2(f0,ϕ0)L(Ω)=1λ1(Ω),

    thus concluding the proof.

    In this section, we briefly sketch some geometric estimates for the generalized principal frequencies, that can be inferred from our main result.

    We start by recalling the Diaz-Weinstein inequality for the torsional rigidity. This is given by the following estimate

    In [16] the case N=2 is considered and a slightly different proof is given. The definition of torsional rigidity in [16] coincides with ours for simply connected sets, up to a multiplicative factor 4.

    T(Ω)1N2I2(Ω), where I2(Ω)=minx0RNΩ|xx0|2dx, (6.1)

    see [16,formula (11)], which is valid for open sets ΩRN such that

    Ω|x|2dx<+.

    The quantity I2(Ω) is sometimes called polar moment of inertia of Ω. It is easily seen that the minimum in its definition is uniquely attained at the centroid of Ω, i.e., at the point

    xΩ=1|Ω|Ωxdx.

    The Diaz-Weinstein inequality can be proved by appealing to the dual formulation for the torsional rigidity. Indeed, for every x0Ω, it is sufficient to use the admissible vector field

    ϕ0=x0xN,

    in the dual problem (1.2). This automatically gives

    T(Ω)1N2Ω|xx0|2dx,

    and thus (6.1) follows by arbitrariness of x0RN. We observe that this estimate is sharp, as equality is attained for a ball. Indeed, recall that the unique W1,20(Ω) solution of Δu=1 in a ball of radius R and center x0 is given by

    w(x)=R2|xx0|22N, for every  xRN   such that |xx0|<R.

    Thus, by observing that ϕ0=w and recalling the discussion in Subsection 1.2, we get the claimed optimality.

    We now show how the previous argument can be extended to the case 1<q<2. We fix again a point x0RN and take a constant α>1/N, then we choose the pair

    ϕ0(x)=α(x0x) and f0(x)=1αN.

    Observe that this solves

    divϕ0+f0=1, in   RN.

    Thus the pair (f0,ϕ0) is admissible for the dual problem (1.6), for every 1<q<2. By Theorem 1.1 we immediately get

    1λ1(Ω;q)(q1)(q1)2qGq(f0,ϕ0)2qL22q(Ω)=(q1)(q1)2qα2(αN1)(q1)2q(Ω|xx0|2q2qdx)2qq.

    We now observe that the quantity

    α2(αN1)(q1)2q,

    is minimal for α=q/N. By making such a choice for α and using the arbitrariness of x0, we then get the following

    Corollary 6.1 (Diaz-Weinstein–type estimate). Let 1<q<2 and let ΩRN be an open set such that

    Ω|x|2q2qdx<+.

    Then we have

    λ1(Ω;q)(qN)2(I2q2q(Ω))2qq,whereI2q2q(Ω)=minx0RNΩ|xx0|2q2qdx. (6.2)

    Remark 6.2. We notice that for 1<q<2 the estimate (6.2) does not appear to be sharp. In order to get the sharp constant, it seems unavoidable the use of more sophisticated arguments, based on radially symmetric decreasing rearrangements. These permit to show that both quantities

    λ1(Ω;q) and I2q2q(Ω),

    are minimal for a ball, among sets with given measure. By combining these two facts, then we get that the sharp constant in (6.2) is given by

    λ1(B1;q)(I2q2q(B1))2qq,

    where B1={xRN:|x|<1}.

    Nevertheless, we believe that the duality-based proof exposed above is interesting anyway: this gives a cheap way to get a scale invariant geometric estimate with a simple explicit constant, by means of an elementary argument.

    We recall that the Cheeger constant for an open set ΩRN is given by

    h1(Ω)=inf{P(E)|E|:EΩ bounded with |E|>0}.

    Here P(E) is the distributional perimeter of a set E. The Cheeger constant has the following dual characterization

    1h1(Ω)=minϕL(Ω;RN){ϕL(Ω):divϕ=1 in   Ω}. (6.3)

    This characterization seems to have first appeared in [33,Section 4]. The fact that the minimum in (6.3) is attained easily follows from the Direct Method in the Calculus of Variations.

    We take an optimal vector field ϕΩ in (6.3) and then for every ε>0 we make the choice

    ϕ0=(1+ε)ϕΩ and f0=ε.

    By observing that (f0,ϕ0)A(Ω), from Theorem 1.1 we get

    1λ1(Ω;q)(q1)(q1)2qGq(f0,ϕ0)2qL22q(Ω)=(q1)(q1)2q(1+ε)2ε(q1)2q(Ω|ϕΩ|2q2qdx)2qq(q1)(q1)2q(1+ε)2ε(q1)2qϕΩ2L(Ω)|Ω|2qq.

    We now notice that the quantity

    (1+ε)2ε(q1)2q,

    is minimal with the choice ε=(q1). Then such a choice leads to the estimate

    1λ1(Ω;q)q21h1(Ω)2|Ω|2qq.

    Thus, we proved the following

    Corollary 6.3. Let 1<q<2 and let ΩRN be an open set, with finite measure. Then we have the Cheeger-type inequality

    (h1(Ω)q)2|Ω|2qqλ1(Ω;q). (6.4)

    Remark 6.4. By taking the limits as q goes to 1 and as q goes to 2 in (6.4), we recover

    h1(Ω)2|Ω|T(Ω) and (h1(Ω)2)2λ1(Ω),

    respectively. The first estimate has been proved in [11,Theorem 2], while the second one is the classical Cheeger inequality for the Laplacian, see [14]. Both inequalities are sharp in the following sense: by taking the Ndimensional unit ball B1(0), one may prove that

    limN|B1(0)|T(B1(0))1h1(B1(0))2=1 and limNλ1(B1(0))h1(B1(0))2=14.

    The first fact can be easily seen, by recalling that

    T(B1(0))=|B1(0)|N(N+2) and h1(B1(0))=N.

    The second fact has been recently observed in [20,Theorem 1.3] and is based on asymptotics for zeros of Bessel functions.

    This somehow suggests that the general estimate (6.4) should be sharp, as well, by using a similar argument. However, the task of computing the exact asymptotics for λ1(B1(0);q), as the dimension N goes to , does not seem easy.

    We now take ΩRN to be an open bounded convex set. We will employ in a dual way a trick by Kajikiya (see [23,24] and also [8]), in conjunction with our duality result. This will give us a sharp lower bound on λ1(Ω;q) in terms on the inradius and the perimeter of the set.

    We indicate by dΩ:¯ΩR the distance function from the boundary Ω, while RΩ will be the inradius of Ω. We recall that this coincides with the supremum of the distance function, that is

    RΩ=supxΩdΩ(x).

    We take g to be the unique positive solution of

    maxφW1,20((1,1)){2q11|φ|qdt11|φ|2dt}.

    This satisfies the equation

    g=gq1, in   (1,1), with g(1)=g(1)=0.

    This is a concave even function, which is increasing on (1,0) and decreasing on (0,1). We then "transplant'' this function to Ω, by setting

    u(x)=R22qΩg(dΩ(x)RΩ1), for xΩ.

    By using the equation solved by g, the fact that |dΩ|=1 almost everywhere and the weak superharmonicity§ of dΩ, we get that

    § This follows from the convexity of Ω.

    Δuuq1, in   Ω,

    in weak sense. This entails that the pair

    ϕ0=uuq1 and f0=(q1)|u|2uq,

    is admissible for the dual problem (1.6). By Theorem 1.1, we then get

    1λ1(Ω;q)(q1)(q1)2qGq(f0,ϕ0)2qL22q(Ω)=(Ω|u|2dx)2qq.

    By using the explicit form of u and again the fact that |dΩ|=1 almost everywhere in Ω, the previous estimate can be rewritten as

    1λ1(Ω;q)R2Ω(Ω|g(dΩRΩ1)|2dx)2qq. (6.5)

    We observe that this is already a geometric estimate in nuce, since the right-hand side only depends on elementary geometric quantities of Ω (i.e., the distance function and the inradius) and on the universal one-dimensional function g. Moreover, it is not difficult to see that (6.5) is sharp (see Remark 6.6 below).

    Let us try to derive from (6.5) a more explicit estimate. At this aim, we can use the Coarea Formula with respect to the distance function, so to get

    Ω|g(dΩRΩ1)|2dx=RΩ0|g(tRΩ1)|2P(Ωt)dt,

    where Ωt={xΩ:dΩ(x)>t}. We now recall that tP(Ωt) is monotone decreasing, in a convex set (see [10,Lemma 2.2.2]). Thus we automatically get

    Ω|g(dΩRΩ1)|2dxP(Ω)RΩ0|g(tRΩ1)|2dt.

    A simple change of variable then leads to

    Ω|g(dΩRΩ1)|2dxP(Ω)RΩ01|g(τ)|2dτ.

    We can insert this estimate in (6.5) to obtain

    1λ1(Ω;q)Rq+2qΩP(Ω)2qq(01|g(τ)|2dτ)2qq.

    By using that g is even and the identity

    11|g(τ)|2dτ=11|g(τ)|qdτ,

    we have

    01|g|2dτ=1211|g|2dτ=12q2q[2q11|g|qdτ11|g|2dτ]=12(1λ1((1,1);q))q2q

    In the last equality we used the optimality of g and Proposition 2.1, for the one-dimensional set Ω=(1,1). This gives

    1λ1(Ω;q)Rq+2qΩP(Ω)2qq(12)2qq1λ1((1,1);q).

    If we now recall the definition

    π2,q=infφW1,20((0,1)){0}φL2((0,1))φLq((0,1)),

    and use the scaling properties of Sobolev-Poincaré constants, we get

    λ1((1,1);q)=(π2,q)222+qq.

    Thus, we finally obtain the following

    Corollary 6.5. Let 1<q<2 and let ΩRN be an open bounded convex set. Then we have

    λ1(Ω;q)(π2,q2)2P(Ω)q2qRq+2qΩ. (6.6)

    Remark 6.6. When compared with the Hersch-Makai–type inequality

    λ1(Ω;q)(π2,q2)2|Ω|q2qR2Ω, (6.7)

    already recalled in the Introduction, we see that the estimate (6.6) is slightly weaker. Indeed, the former implies the latter, by recalling that for a open bounded convex set we have

    |Ω|RΩP(Ω).

    Nevertheless, inequality (6.6) is still sharp: it is sufficient to take the "slab–type'' sequence

    ΩL=(L2,L2)N1×(0,1),

    with L diverging to +. For this family of sets we have (see [8,Lemma A.2])

    λ1(ΩL;q)(π2,q)2L(N1)2qq and P(ΩL)2LN1, as L+,

    and RΩL=1/2, for L>1.

    We also observe that this slight discrepancy between (6.6) and (6.7) is lost in the limit as q converges to 2: in both cases the estimates boil down to

    λ1(Ω)(π2)21R2Ω,

    which is the original Hersch sharp inequality from [22,Théorème 8.1]. We also refer to [12,Theorem 5.5], [15,Theorem 5.1] and [23,Theorem 2.1] for other proofs and extensions of this result.

    We wish to thank Francesco Maggi for first introducing us to the hidden convexity of the Dirichlet integral. We also thank Rafael Benguria, Guillaume Carlier and Eleonora Cinti for some comments on a preliminary version of the paper.

    The author declares no conflict of interest.



    [1] G. Anello, F. Faraci, A. Iannizzotto, On a problem of Huang concerning best constants in Sobolev embeddings, Annali di Matematica, 194 (2015), 767–779. doi: 10.1007/s10231-013-0397-8
    [2] R. Benguria, The von Weizsäcker and exchange corrections in the Thomas-Fermi theory, Ph. D. thesis of Princeton University, 1979.
    [3] R. Benguria, H. Brézis, E. H. Lieb, The Thomas-Fermi-von Weizsäcker theory of atoms and molecules, Commun. Math. Phys., 79 (1981), 167–180. doi: 10.1007/BF01942059
    [4] M. van den Berg, Estimates for the torsion function and Sobolev constants, Potential Anal., 36 (2012), 607–616. doi: 10.1007/s11118-011-9246-9
    [5] M. S. Berger, M. Schechter, Embedding theorems and quasi-linear elliptic boundary value problems for unbounded domains, T. Am. Math. Soc., 172 (1972), 261–278. doi: 10.1090/S0002-9947-1972-0312241-X
    [6] L. Brasco, On principal frequencies and isoperimetric ratios in convex sets, Ann. Fac. Sci. Toulouse Math., 29 (2020), 977–1005. doi: 10.5802/afst.1653
    [7] L. Brasco, G. Franzina, B. Ruffini, Schrödinger operators with negative potentials and Lane-Emden densities, J. Funct. Anal., 274 (2018), 1825–1863. doi: 10.1016/j.jfa.2017.10.005
    [8] L. Brasco, D. Mazzoleni, On principal frequencies, volume and inradius in convex sets, Nonlinear Differ. Equ. Appl., 27 (2020), 12. doi: 10.1007/s00030-019-0614-2
    [9] H. Brezis, L. Oswald, Remarks on sublinear elliptic equations, Nonlinear Anal., 10 (1986), 55–64. doi: 10.1016/0362-546X(86)90011-8
    [10] D. Bucur, G. Buttazzo, Variational Methods in Shape Optimization Problems, Boston, MA: Birkhäuser Boston, Inc., 2005.
    [11] H. Bueno, G. Ercole, Solutions of the Cheeger problem via torsion functions, J. Math. Anal. Appl., 381 (2011), 263–279. doi: 10.1016/j.jmaa.2011.03.002
    [12] G. Buttazzo, S. Guarino Lo Bianco, M. Marini, Sharp estimates for the anisotropic torsional rigidity and the principal frequency, J. Math. Anal. Appl., 457 (2018), 1153–1172. doi: 10.1016/j.jmaa.2017.03.055
    [13] T. Carroll, J. Ratzkin, Interpolating between torsional rigidity and principal frequency, J. Math. Anal. Appl., 379 (2011), 818–826. doi: 10.1016/j.jmaa.2011.02.004
    [14] J. Cheeger, A lower bound for the smallest eigenvalue of the Laplacian, In: Proceedings of the Princeton conference in honor of Professor S. Bochner, Princeton University Press, 1970,195–199.
    [15] F. Della Pietra, G. di Blasio, N. Gavitone, Sharp estimates on the first Dirichlet eigenvalue of nonlinear elliptic operators via maximum principle, Adv. Nonlinear Anal., 9 (2020), 278–291.
    [16] J. B. Diaz, A. Weinstein, The torsional rigidity and variational methods, Am. J. Math., 70 (1948), 107–116. doi: 10.2307/2371935
    [17] G. Ercole, Absolute continuity of the best Sobolev constant of a bounded domain, J. Math. Anal. Appl., 404 (2013), 420–428. doi: 10.1016/j.jmaa.2013.03.044
    [18] M. Fleeman, E. Lundberg, The Bergman analytic content of planar domains, Comput. Meth. Funct. Th., 17 (2017), 369–379. doi: 10.1007/s40315-016-0189-4
    [19] M. Fleeman, B. Simanek, Torsional rigidity and Bergman analytic content of simply connected regions, Comput. Meth. Funct. Th., 19 (2019), 37–63. doi: 10.1007/s40315-018-0252-4
    [20] I. Ftouhi, On the Cheeger inequality for convex sets, J. Math. Anal. Appl., 504 (2021), 125443. doi: 10.1016/j.jmaa.2021.125443
    [21] J. Hersch, Physical interpretation and strengthing of M. Protter's method for vibrating nonhomogeneous membranes; its analogue for Schrödinger's equation, Pacific J. Math., 11 (1961), 971–980. doi: 10.2140/pjm.1961.11.971
    [22] J. Hersch, Sur la fréquence fondamentale d'une membrane vibrante: évaluations par défaut et principe de maximum, Z. Angew. Math. Phys., 11 (1960), 387–413. doi: 10.1007/BF01604498
    [23] R. Kajikiya, A priori estimate for the first eigenvalue of the pLaplacian, Differ. Integral Equ., 28 (2015), 1011–1028.
    [24] R. Kajikiya, A priori estimates of positive solutions for sublinear elliptic equations, T. Am. Math. Soc., 361 (2009), 3793–3815. doi: 10.1090/S0002-9947-09-04875-2
    [25] B. Kawohl, Symmetry results for functions yielding best constants in Sobolev-type inequalities, DCDS, 6 (2000), 683–690. doi: 10.3934/dcds.2000.6.683
    [26] M. T. Kohler-Jobin, Symmetrization with equal Dirichlet integrals, SIAM J. Math. Anal., 13 (1982), 153–161. doi: 10.1137/0513011
    [27] E. H. Lieb, M. Loss, Analysis, 2 Eds., Providence, RI: American Mathematical Society, 2001.
    [28] E. Makai, On the principal frequency of a membrane and the torsional rigidity of a beam, In: Studies in math. analysis and related topics, Stanford: Stanford Univ. Press, 1962,227–231.
    [29] A. I. Nazarov, On the symmetry of extremals in the weight embedding theorem. Function theory and mathematical analysis, J. Math. Sci., 107 (2001), 3841–3859. doi: 10.1023/A:1012336127123
    [30] F. Santambrogio, Optimal transport for applied mathematicians, Basel: Birkhäuser, 2015.
    [31] G. Pólya, Two more inequalities between physical and geometrical quantities, J. Indian Math. Soc., 24 (1960), 413–419.
    [32] R. T. Rockafellar, Convex analysis, Princeton, N.J.: Princeton University Press, 1970.
    [33] G. Strang, L1 and L approximation of vector fields in the plane, In: North-Holland Mathematics Studies, 81 (1983), 273–288.
    [34] P. Takáč, L. Tello, M. Ulm, Variational problems with a phomogeneous energy, Positivity, 6 (2002), 75–94. doi: 10.1023/A:1012088127719
    [35] D. Zucco, Dirichlet conditions in Poincaré-Sobolev inequalities: the sub-homogeneous case, Calc. Var., 58 (2019), 89. doi: 10.1007/s00526-019-1547-7
  • This article has been cited by:

    1. Dario Mazzoleni, Benedetta Pellacci, Calculus of variations and nonlinear analysis: advances and applications, 2023, 5, 2640-3501, 1, 10.3934/mine.2023059
    2. Delio Mugnolo, Marvin Plümer, On torsional rigidity and ground-state energy of compact quantum graphs, 2023, 62, 0944-2669, 10.1007/s00526-022-02363-9
  • Reader Comments
  • © 2022 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(1861) PDF downloads(69) Cited by(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog