Loading [MathJax]/jax/element/mml/optable/GeneralPunctuation.js
Review

A Review of the interrelations of terrestrial carbon sequestration and urban forests

Running title: Terrestrial carbon sequestration and urban forests
  • Received: 25 September 2020 Accepted: 10 November 2020 Published: 13 November 2020
  • Rapid urbanization poses major challenges to the mankind with huge impact on the society and the economy. This paper aims to review relevant literature, mainly recent studies, focused on significant aspects of the interrelations of urban forests and terrestrial carbon sequestration, discussing their implications with reference to urban forests and their roles in climate mitigation, given that the real challenge lies in understanding the integration of carbon sequestration having huge pollution mitigating potentials with other mitigation options. Findings suggest that despite indications of studies that urban forests play significant mitigation roles; they have not been accorded adequate importance vis-à-vis management of ecological disturbances. Findings also suggest that urban forests significantly contribute to terrestrial carbon sequestration and these contributions are in addition to their multiple social-economic-cultural-aesthetic benefits. This paper underscoring the significant contributions of urban forests in maintaining ecological equilibrium, and managing balance between emissions and sequestration to ensure sustainability, offers usefulness to the future researchers, academics, urban-planners and policymakers.

    Citation: Kumari Anjali, YSC Khuman, Jaswant Sokhi. A Review of the interrelations of terrestrial carbon sequestration and urban forests[J]. AIMS Environmental Science, 2020, 7(6): 464-485. doi: 10.3934/environsci.2020030

    Related Papers:

    [1] Bruno Bianchini, Giulio Colombo, Marco Magliaro, Luciano Mari, Patrizia Pucci, Marco Rigoli . Recent rigidity results for graphs with prescribed mean curvature. Mathematics in Engineering, 2021, 3(5): 1-48. doi: 10.3934/mine.2021039
    [2] Mattia Fogagnolo, Andrea Pinamonti . Strict starshapedness of solutions to the horizontal p-Laplacian in the Heisenberg group. Mathematics in Engineering, 2021, 3(6): 1-15. doi: 10.3934/mine.2021046
    [3] Rolando Magnanini, Giorgio Poggesi . Interpolating estimates with applications to some quantitative symmetry results. Mathematics in Engineering, 2023, 5(1): 1-21. doi: 10.3934/mine.2023002
    [4] Michiaki Onodera . Linear stability analysis of overdetermined problems with non-constant data. Mathematics in Engineering, 2023, 5(3): 1-18. doi: 10.3934/mine.2023048
    [5] 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
    [6] Luigi Provenzano, Alessandro Savo . Isoparametric foliations and the Pompeiu property. Mathematics in Engineering, 2023, 5(2): 1-27. doi: 10.3934/mine.2023031
    [7] Serena Dipierro, Giovanni Giacomin, Enrico Valdinoci . The fractional Malmheden theorem. Mathematics in Engineering, 2023, 5(2): 1-28. doi: 10.3934/mine.2023024
    [8] La-Su Mai, Suriguga . Local well-posedness of 1D degenerate drift diffusion equation. Mathematics in Engineering, 2024, 6(1): 155-172. doi: 10.3934/mine.2024007
    [9] Antonio Greco, Francesco Pisanu . Improvements on overdetermined problems associated to the p-Laplacian. Mathematics in Engineering, 2022, 4(3): 1-14. doi: 10.3934/mine.2022017
    [10] Giacomo Ascione, Daniele Castorina, Giovanni Catino, Carlo Mantegazza . A matrix Harnack inequality for semilinear heat equations. Mathematics in Engineering, 2023, 5(1): 1-15. doi: 10.3934/mine.2023003
  • Rapid urbanization poses major challenges to the mankind with huge impact on the society and the economy. This paper aims to review relevant literature, mainly recent studies, focused on significant aspects of the interrelations of urban forests and terrestrial carbon sequestration, discussing their implications with reference to urban forests and their roles in climate mitigation, given that the real challenge lies in understanding the integration of carbon sequestration having huge pollution mitigating potentials with other mitigation options. Findings suggest that despite indications of studies that urban forests play significant mitigation roles; they have not been accorded adequate importance vis-à-vis management of ecological disturbances. Findings also suggest that urban forests significantly contribute to terrestrial carbon sequestration and these contributions are in addition to their multiple social-economic-cultural-aesthetic benefits. This paper underscoring the significant contributions of urban forests in maintaining ecological equilibrium, and managing balance between emissions and sequestration to ensure sustainability, offers usefulness to the future researchers, academics, urban-planners and policymakers.


    In this paper, the object under investigation is a triple (M,g0,u) satisfying the following two conditions:

    (a) (M,g0) is a smooth, connected, noncompact, complete, asymptotically flat, n-dimensional Riemannian manifold, with n3, with one end, and with nonempty smooth compact boundary M, which is a priori allowed to have several connected components.

    (b) uC(M) satisfies the system

    {uRicg0D2g0u0in M,Δg0u=0 in M,u=0on M,u1at , (1.1)

    where Ricg0, Dg0 and Δg0 are the Ricci tensor, the Levi–Civita connection, and the Laplace operator of the metric g0, respectively.

    If the equality holds in the first equation of (1.1), the triple (M,g0,u) is said static. For clarity, we recall the definition to which we refer for asymptotically flat manifolds.

    Definition 1.1. A smooth, connected, noncompact, n-dimensional Riemannian manifold (with or without compact boundary) (N,h), with n3, is said to be asymptotically flat if there exists a compact subset KN such that NK is a finite disjoint union of ends Nk with the following properties. Every Nk is diffeomorphic to Rn minus a closed ball by a coordinate chart ψk and, if ˜h:=(ψk)h=˜hijdxidxj, we have

    ˜hij=δij+O(|x|p), (1.2)
    r˜hij=O(|x|(p+1)), (1.3)
    rs˜hij=O(|x|(p+2)), (1.4)
    R˜hL1(μ˜h), (1.5)

    for some p>(n2)/2. Here, δ is the Kronecker delta, and the coordinate charts ψk are called charts at infinity.

    Throughout the paper, we will refer to a triple (M,g0,u) that satisfies conditions (a) and (b) as to a sub-static harmonic triple. A fundamental sub-static harmonic triple is the so called Schwarzschild solution, which is given by

    M=[(2m)1n2,+)×Sn1,g0=drdr12mr2n+r2gSn1,u=12mr2n. (1.6)

    It is well–known that both the metric g0 and the potential u, which a priori are well defined only in ˚M, extend smoothly up to the boundary and (M,g0) is called (spatial) Schwarzschild manifold. The parameter m>0 is the ADM mass mADM of the Schwarzschild manifold. We refer the reader to Section 5 for the definition of the mADM associated with a general asymptotically flat manifold. Here, we limit ourselves to recall that the decay conditions (1.2)–(1.5) guarantee that mADM is a geometric invariant [4,8].

    Associated with a sub-static harmonic triple, specifically with the potential u ranging in [0,1), let us consider the following family of functions depending on the parameter β0:

    [0,1) tVβ(t):=(1t2)β(n1n2){u=t}|Du|β+1dσ.

    In [2] it was proven that if (M,g0,u) is a static triple, then, for every β2, the function Vβ is strictly nonincreasig unless (M,g0,u) is the Schwarzschild solution. The main purpose of this paper is to extend this result to the sub-static case and to the optimal threshold β=n2n1. This is the content of Theorem 3.1, where the monotonicity of the above family - equipped with a corresponding rigidity statement - is expressed in terms of the functions Fβ(τ), where τ=1+t21t21, to be consistent with [3] and in light of the more advanced analysis contained therein. This generalisation suggests that our approach is robust enough and likely to be exported to other contexts. In a similar way, S. Brendle shows in [7] how some structure conditions for the metric are sufficient to prove an Alexandrov-type theorem and how such structure generalises to the sub-static case.

    Let us now be slightly more detailed on how our Theorem 3.1 is proved. We adopt the main strategy proposed in [2], which essentially consists in obtaining the monotonicity as a consequence of a fundamental integral identity derived in a suitable conformally-related setting (see Proposition 4.3). A delicate point is justifying such identity in a region where critical points of the potential are present. One of the main differences with [2] is that, whereas in the static case the analyticity of the potential guarantees the local finiteness of the singular values, which made the argument simpler in many occurrences, in the present sub-static setting the metric and in turn the potential are not a priori analytic. Nevertheless, standard measure properties of the critical set of harmonic functions (summarised in Theorem 2.3) are enough to obtain the fundamental integral identity, which in turn implies the monotonicity of Fβ and, coupled with Sard's Theorem, also its differentiability.

    Observe that the difficulty in treating the critical points under the threshold β=1 can be read off directly from formulæ (3.2) and (4.5), the first one displaying the derivative of Fβ and the second one expressing the mean curvature on a equipotential set in terms of the Hessian of the potential itself. In fact, calling Φβ the conformal version of Fβ and looking at formula (4.26) containing the equivalent characterisation of Φβ derived from the integral identity (4.18), one realises that problems arise already when β<2.

    Let us stress that the monotonicity is obtained from the nonnegativity of the right-hand side of our fundamental integral identity. It is above the threshold β=n2n1 that this is guaranteed, thanks to the Refined Kato Inequality for harmonic functions. The optimality of such inequality reflects a corresponding optimality of β=n2n1 in our result. Moreover, let us remark that the (nonnegative) right-hand side of (4.18) is obtained as the divergence of a suitable modification of a specific vector filed with nonnegative divergence (see (4.19)), in the limit of a vanishing neighbourhood of the critical set. The crucial point in the construction is to maintain the divergence of the modified vector field nonnegative. It would be interesting to see whether a similar construction can be performed for other families of metrics, including special solutions as rigid case.

    A straightforward application of the monotonicity of Fβ is comparing Fβ(1) with Fβ(+), in turn yielding a "capacitary version" of the Riemannian-Penrose inequality (Theorem 1.1 below). The capacity comes naturally into play when computing Fβ(1) and Fβ(+), the latter value via the asymptotic expansions of the metric and of the potential. We recall that the capacity Cap(M,g0) of M is defined as

    Cap(M,g0):=1(n2)|Sn1|inf{M|Dg0v|2g0dμg0:vLiploc(M),v0,v=0onM,v1at}.

    Throughout the paper, we will use the short–hand notation C for the capacity. Comparing (1.6) with either (2.1) or (2.2), it is straightforward, in the case of the Schwarzschild solution, that mADM=C. For a general sub-static harmonic triple, the following inequality holds.

    Theorem 1.1 (Capacitary Riemannian Penrose Inequality). Let (M,g0,u) be a sub-static harmonic triple with associated capacity C and suppose that M is connected. Then

    C12(|M||Sn1|)n2n1. (1.7)

    Moreover, the equality in (1.7) holds if and only if (M,g0) is isometric to the Schwarzschild manifold with mADM=C.

    Whereas the above inequality has been obtained as a consequence of the monotonicity of Fβ, at every fixed βn2n1, we remark that one could possibly push the above described analysis one step forward, at the same time exploiting the full power of the optimality threshold. Indeed, we believe that considering p-harmonic functions defined at the exterior of a bounded domain Ω lying in M, it may be possible to derive, as done in [1] for the Euclidean case and in the simultaneous limit as βn2n1 and p1, a Minkowski-like inequality for Ω (see [21] for a Minkowski-like inequality in the static, asymptotically flat case and [5] for the nonnegative Ricci case).

    Concerning the treatment of general sub-static metrics and the derivation of related geometric inequalities, besides the already cited [7] we also would like to mention [19], where an integral formula is obtained and applied to prove Hentze-Karcher-type inequalities. For the case of asymptotically hyperbolic sub-static manifolds (specifically, for adS-Reissner-Nordström manifolds), we refer the interested reader to [13] and [27].

    We remark that our results are not based on the Positive Mass Theorem. By contrast, we observe that using this celebrated result, more precisely a consequence of it contained in [17,Theorem 1.5], one can prove the following uniqueness statement. We refer the reader to Definition 1.1 for the notation and terminology.

    Theorem 1.2 (Uniqueness Theorem for sub-static harmonic triples). Let (M,g0,u) be a sub-static harmonic triple with associated capacity C. Suppose that there is a chart at infinity such that

    R˜g0=O(|x|q), (1.8)

    for some q>n. Then (M,g0) is the Schwarzschild manifold with associated ADM mass given by C.

    It remains an open question to see whether it is possible to remove the assumption on the decay of R˜g0 and get the same conclusion.

    The paper is organised as follows. In Section 2, we recall and discuss some preparatory material, namely the asymptotic expansions of the metric and of the potential, and classical measure properties of the critical set of the potential, with a close look on related integral quantities. In Section 3, we prove the Monotonicity and Outer Rigidity Theorem 3.1, and the consequent Capacitary Riemannian Penrose Inequality contained in Theorem 1.1. To do this, we use from Section 4 some corresponding results obtained in a suitable conformally-related setting. The biggest technical effort is contained in such section. In the Appendix we also provide an alternative proof of the monotonicity of our monotone quantities. Finally, Section 5 is devoted to the proof of Theorem 1.2.

    Let (M,g0,u) be a sub-static harmonic triple. We observe, as a first consequence of system (1.1), that the scalar curvature Rg0 is nonnegative. Since u satisfies the last three conditions of system (1.1), by the Maximum Principle we have

    ˚M=MM={0<u<1}.

    Also, by the forth condition in (1.1), each level set of u is compact. Moreover, from the Hopf Lemma, it follows that |Dg0u|g0>0 on M. In particular, zero is a regular value of u. Furthermore, from the first two conditions in (1.1) restricted to M it is easy to deduce that D2g0u0 on M. In turn, the function |Dg0u|g0 attains a positive constant value on each connected component of M, and the boundary M is a totally geodesic hypersurface in M.

    We now deal with the asymptotic behaviour of the potential u at . By Theorem 2.2 below, this is given by:

    u=1C|x|n2+o2(|x|2n)  as  |x|+, (2.1)

    being

    C=1(n2)|Sn1|M|Dg0u|g0dσg0. (2.2)

    Here, σg0 is the canonical measure on the boundary M seen as a Riemannian submanifold of (M,g0), and we have used the standard notation o2, which means that, in any chart at infinity ψ, denoted by ˜u the function uψ1, the following conditions hold true.

    ˜u=1C|x|2n+o(|x|2n), (2.3)
    i˜u=(n2)C|x|nxi+o(|x|1n), (2.4)
    ij˜u=(n2)C|x|n2(nxixj|x|2δij)+o(|x|n). (2.5)

    Let us remark that we can always suppose, without loss of generality, that the considered chart at infinity admits a diffeomorphic extension to the closure of the coordinate domain. We will make this implicit assumption throughout the paper, so that K (see Definition 1.1) is a connected hypersurface of M and the quantities related to the metric can be pushed–forward in Rn outside an open ball and be smooth here. We also observe that formula (2.2) is nothing but an equivalent characterisation of the capacity of M.

    Let (N,h) be a smooth, connected, noncompact, complete, asymptotically flat, n-dimensional Riemannian manifold, with n3, with one end and with nonempty smooth compact boundary N. We adopt the following notation.

    B and BR a generic open ball and the open ball of radius R>0 centred in the origin of (Rn,de), respectively;

    || the euclidean norm of Rn;

    |Sn1| the hypersurface area of the unit sphere inside Rn with the canonical metric;

    De and Δe the Levi–Civita connection and the Laplace operator of (Rn,gRn), respectively;

    Dh and Δh the Levi–Civita connection and the Laplace operator of (N,h), respectively;

    σe the canonical measure on a Riemannian submanifold of (Rn,gRn);

    σh the canonical measure on a Riemannian submanifold of (N,h);

    ||e the norm induced by gRn on the tangent spaces to the manifold Rn;

    ||h the norm induced by h on the tangent spaces to the manifold N.

    ● If ψ is a chart at infinity of (N,h) according to Definition 1.1, we denote by ˜h the push–forward metric ψh of h by ψ, having coordinate expression ˜hij(x)dxidxj. In this context, D˜h and Δ˜h denote the Levi–Civita connection and the Laplace operator of ˜h, respectively, while σ˜h is the canonical measure on a Riemannian submanifold of (RnB,˜h) and ||˜h is the norm induced by ˜h on the tangent spaces. Moreover, Ric˜h and R˜h are the Ricci tensor and the scalar curvature of ˜h, respectively.

    Proposition 2.1. Let ψ be a chart at infinity of N (according to Definition 1.1). The decays

    ˜hijδij=O2(|x|p), (2.6)
    (Rm˜h)lijk=O(|x|(p+2)), (2.7)
    (Ric˜h)ij=O(|x|(p+2)), (2.8)
    R˜h=O(|x|(p+2)), (2.9)

    hold true for some p>n22. Moreover,

    |νi˜hνie|=O(|x|p), (2.10)
    dσ˜h=(1+O(|x|p)dσe, (2.11)

    where νe is the –pointing unit normal with respect to the Euclidean metric and σe the associated canonical measure on BR, while ν˜h is the –pointing unit normal with respect to ˜h and σ˜h the associated canonical measure on BR.

    Proof. From ˜hik˜hkj=δij it is easy to get

    i˜hkl=˜hkr˜hlsi˜hrsij˜hkl=˜hka˜hrb˜hls(i˜hab)(j˜hrs)+˜hla˜hsb˜hkr(i˜hab)(j˜hrs)˜hkr˜hlsij˜hrs.

    These formulae coupled with (1.2), (1.3) and (1.4) give (2.6). Decay (2.7) is another direct consequence of Definition 1.1, keeping in mind that

    (Rm˜h)lijk=iΓljkjΓlik+ΓsjkΓlisΓsikΓljs,Γkij=˜hkl2[i˜hlj+j˜hlil˜hij]. (2.12)

    Decays (2.8)–(2.9) are obtained by contractions of the Riemannian tensor. Now, observe that

    ν˜h=˜hijxixj˜hlkxlxk,

    and that

    |νieνi˜h|=|xi|x|˜hijxj˜hlkxlxk|=|(δij˜hij)xj˜hlkxlxk+xi(1|x|1˜hlkxlxk)|Cj|δij˜hij|+|(˜hlkδlk)xlxk|˜hlkxlxk(˜hlkxlxk+|x|). (2.13)

    Observe also that

    ˜hij(x)vivjC1vivjδij, (2.14)

    for some C>0, for any xRnB. Since trivially |xkxl||x|2, from (2.13) and (2.14), coupled with (2.6), we get decay (2.10). Concerning decay (2.11), recall first that, using a coordinate chart (y1,,yn1) on BR, we have that dσ˜h=det˜hBRdy1dyn1 with ˜hBR=˜hBRαβdyαdyβ, where ˜hBRαβ=˜h(yα,yβ). Now, using the specific local parametrization x=x(y1,,yn1) of BR, given by the inverse of stereographic projection from its north pole with the diffeomorphism pSn1RpBR, we have that

    ˜h(yα,yβ)(x(y))=˜hij(x(y))xiyα(y)xjyβ(y)=(˜hij(x(y))±δij)xiyα(y)xjyβ(y)=4R2(|y|2+1)2(δαβ+O(Rp)),

    because

    |(˜hij(x(y))δij)xiyα(y)xjyβ(y)|i,j|˜hij(x(y))δij||yα|e|yβ|e=4R2(|y|2+1)2O(Rp).

    Hence, on BR,

    dσ˜h=(2R|y|2+1)n1det(δαβ+O(Rp))dy1dyn1=(1+O(Rp))dσe,

    where in the last identity we have used the Leibniz formula for the determinant and Taylor–expanded the square root.

    The following result is well–known. For completeness, we provide the statement, along with its proof, which is an extension of [20,Lemma A.2.] to every n3.

    Theorem 2.2. Let (N,h) be a smooth, connected, noncompact, complete, asymptotically flat, n-dimensional Riemannian manifold, with n3, with one end, and with nonempty smooth compact boundary N. If vC(N) is the solution to

    {Δhv=0 in N,v=1on N,v0at . (2.15)

    then

    v=C|x|n2+o2(|x|2n)  as  |x|,withC=1(n2)|Sn1|N|Dhv|hdσh. (2.16)

    We remark that the asymptotic behaviour of the potential u at , given by formula (2.1), is a simply consequence of the above theorem observing that u=1v when (N,h)=(M,g0).

    Proof. Step 1: Construction of a barrier function. Let ψ be a chart at infinity for N. From now on by C we will denote some positive constant, which may change from line to line. By Definition 1.1, there exist p>(n2)/2 and R11 such that

    RnBR1Rn¯B
    |˜hijδij|C|x|p|k˜hij| C|x|(p+1)|kl˜hij| C|x|(p+2) (2.17)

    for every xRnBR1. By (2.6), the same conditions as in (2.17) are satisfied by ˜hij(x) for all xRnBR1. Then, for every fC(RnBR1), writing

    Δ˜hf=δijijf+σijijf+bjjf, (2.18)

    where

    σij:=˜hijδij,bj:=˜hklΓjkl=12˜hkl˜hiji˜hlk˜hki˜hlji˜hkl,

    we have that

    |σij|C|x|p,|bj|C|x|(p+1),|lbj|C|x|(p+2) (2.19)

    in RnBR1. For a fixed 0<ε<p and for a>0 to be chosen later, consider the function

    ϕa=a(1|x|n21|x|n2+ε).

    By direct computation one can check that

    jϕa=a(n2|x|nn2+ε|x|n+ε)xjijϕa=a[n(n2)|x|n+2(n+ε)(n2+ε)|x|n+2+ε]xixja[n2|x|nn2+ε|x|n+ε]δij,

    and in turn that

    |iϕa|aC|x|1nand|ijϕa|aC|x|n.

    Therefore, by (2.18) and (2.19), we obtain that

    Δ˜hϕa=a[(n2+ε)ε|x|(n+ε)+O(|x|(n+p))],

    and hence there exists R2>R1 independent of a such that Δ˜hϕa<0 in RnBR2, for every a>0. We now choose a>0 so that ϕa=1 on BR2, that is a=[1Rn221Rn2+ε2]1. Since ϕa is ˜h–superharmonic in RnBR2 and since ˜v:=vψ1<1 on BR2, by the Maximum Principle

    ˜vϕainRnBR2. (2.20)

    Step 2: Asymptotic expansion of v. Note that from (2.20) one gets in particular that ˜vC|x|2n. We now apply Shauder's Interior estimates ([12,Lemma 6.20]) to Δ˜h˜v=0 in Rn¯BR2, where the operator Δ˜h is defined as in (2.18) and its coefficients satisfy the estimates in (2.19). Recalling that the Hölder norms are weighted by the (Euclidean) distance de(,BR2) from BR2 and since de(x,BR2)|x| when |x|>>1, from such estimates we get

    |i˜v(x)|C|x|1n|ij˜v(x)|C|x|n (2.21)

    in RnBR2 (up to a bigger R2). Combining (2.19) and (2.21), the equation Δ˜h˜v=0 can be equivalent written as Δe˜v=f where

    |f(x)|C|x|(n+p). (2.22)

    We consider a smooth extension of ˜v on Rn, still denoted by ˜v, which is zero in a ball centred in the origin, and the smooth extension of f given by Δe˜v, still denoted by f. By a classical representation formula and due to (2.22), the function

    w(x)=1n(n2)ωnRnf(y)|xy|n2dy,

    is well–defined and fulfils Δew=f on Rn. Now, one can rewrite w in Rn{O} as

    w(x)=1n(n2)ωn1|x|n2Rnf(y)dy+1n(n2)ωn1|x|n2RnB|x|2(O)f(y)dy1n(n2)ωnB|x|2(x)f(y)|xy|n2dy1n(n2)ωnRn(B|x|2(x)B|x|2(O))f(y)|xy|n2dy1n(n2)ωnB|x|2(O)[1|xy|n21|x|n2]f(y)dy,

    and show that each summand can be bounded by C|x|(n2+γ), where γ=min{1,p} if p1 and γ(1/2,1) if p=1, except the first one. Therefore, we have that

    w(x)=1n(n2)ωn1|x|n2Rnf(y)dy+z(x),|z(x)|C|x|(n2+γ),

    in Rn{O}. Since the function ˜vw is harmonic and bounded on Rn, then it is constant and this constant is zero, using the fact that ˜vw0 for |x|. Hence

    ˜v=C|x|n2+z(x) (2.23)

    in RnBR2. We observe that

    Δ˜hz=Δ˜h(˜vC|x|n2)=C(σijij1|x|n2+bkk1|x|n2)=:Cˆz,

    and that

    |ˆz(x)|C|x|(n+γ),|kˆz(x)|C|x|(n+γ+1).

    Therefore, applying Shauder's Interior estimates to Δ˜hz=Cˆz in Rn¯BR2, we get

    |z(x)|C|x|(n2+γ)|iz(x)|C|x|(n1+γ)|ijz(x)|C|x|(n+γ) (2.24)

    in RnBR2 (up to a bigger R2). From (2.23) and (2.24) we obtain in particular (2.16).

    Step 3: Characterization of C. First of all we remark that 0<v<1 on ˚N, v:N(0,1] is proper, and, from the Hopf Lemma, |Dhv|h>0 on N. In particular, 1 is a regular value of v. Let K be the compact set on the complement of which the chart ψ is defined. For every R>R2, applying the Divergence Theorem to the function v on K{|ψ|<R} we obtain that

    0=K{|ψ|<R}Δhvdμh=Nh(Dhv,νh)dσh+{|ψ|=R}h(Dhv,νh)dσh,

    where νh is the outward unit normal vector field with respect to h along N and {|ψ|=R}. Then, it follows that

    N|Dhv|hdσh=Nh(Dhv,νh)dσh={|ψ|=R}h(Dhv,νh)dσh=BR˜h(D˜h˜v,ν˜h)dσ˜h,

    where ˜v=vψ1. Now, thanks to (1.2), (2.6), (2.10) and (2.11), which are true for γ too, and also by identity (2.23) and the second in (2.24), and keeping in mind that |i˜v|C|x|1n, we have that

    BR˜h(D˜h˜v,ν˜h)dσ˜h=BRgRn(De˜v,νe)dσe+O(Rγ)=BRgRn(De(C|x|2n+z),x|x|)dσe+O(Rγ)=BRgRn(De(C|x|2n),x|x|)dσe+O(Rγ)=C(n2)|Sn1|+O(Rγ).

    Hence

    N|Dhv|hdσh=limRBR˜h(D˜h˜v,ν˜h)dσ˜h=C(n2)|Sn1|.

    Let (N,h) and ι:SN be respectively a m–dimensional Riemannian manifold and a s–dimensional Riemannian submanifold of N. Let k be a positive real number. We set

    B(S) the smallest σ–algebra containing all open sets of S;

    (S,Λ(S),μιh) the canonical space of measure on the Riemannian manifold (S,ιh) (see [14,Section 3.4]);

    HkS the k-dimensional Hausdorff measure on (S,dS), being dS the distance function of S;

    HkS;N the k-dimensional Hausdorff measure on (S,dS;N) where dS;N is the distance function of N restricted to S×S;

    HkNS the k-dimensional Hausdorff measure of N restricted to S.

    By definition of the Hausdorff measure and by [26,Proposition 12.7], HkS;N, HkNS and HkS coincide on B(S), and by [26,Proposition 12.6] and by [25,Proposition 2.17], HsS and μιh coincide on Λ(S). The same results still hold when N is a manifold with boundary.

    For the ease of the reader, we collect in the next theorem some results about the measure of the level sets of the potential u and

    Crit(u):={|Dg0u|g0=0},

    which are well–known in the Euclidean setting (see, e.g., [15,16]).

    Theorem 2.3. Let (M,g0,u) be a sub-static harmonic triple. Then the following statements hold true.

    (i) For every t[0,1), the level set {u=t} is compact and has finite (n1)–Hausdorff measure in M;

    (ii) Crit(u) is a compact subset of M and its Hausdorff dimension in M is less than or equal to (n2);

    (iii) The set of the critical values of u has zero Lebesgue measure, and for every t[0,1) regular value of u there exists ϵt>0 such that (tϵt,t+ϵt)[0,1) does not contain any critical value of u.

    Proof. Each level set of u is compact, due to the forth condition in (1.1), while the compactness of Crit(u) follows by (2.4). Now, consider the nontrivial case where Crit(u) and let p be a point of a critical level set {u=t}. Take a chart (Up,ψp) centred at the point p with ψp(Up)=B1. Setting ˜g0=(ψp)g0=˜g0;ijdxidxj, there exists C>0 such that

    C1vivjδij˜g0;ij(x)vivjCvivjδij, (2.25)

    for each (v1,,vn)Rn and for each x¯B12. The same condition is satisfied by the coefficients ˜gij0. In particular, setting ˜u=uψ1p, we have that

    {|D˜g0˜u|˜g0=0}B12={|De˜u|e=0}B12.

    We observe that Δ˜g0v=aijijv+biiv=0 is an elliptic partial differential equation with coefficient C in ¯B12. We recall that if v is a C–solution of the above equation and if v vanishes to infinite order at a point x0B12, i.e., for every k>0

    limr01rkBr(x0)v2dx=0,

    then v is identically zero in B12 (see [11,Theorem 1.2]). Applying this fact to ˜ut, one can argue that ˜ut has finite order of vanishing at O. Then, by using [16,Theorem 1.7], there exists 0<ρ<12 such that

    Hn1(Bρ{˜u=t})<.

    Since ˜u is nonconstant in B12 and by the structure and regularity of Δ˜g0, [15,Theorem 1.1] yields

    Hn2(Bρ{|De˜u|e=0})<.

    Hence, since the restriction of ψp to ψ1p(B1/2) is bilipschitz due to (2.25) and since the measures Hkψ1p(B1/2) and HkMψ1p(B1/2) coincide on borel sets, statements (i) and (ii) are true locally. In turn, by compactness of {u=t} and Crit(u), they are true globally. To prove (iii), observe that by Sard's Theorem, the set of the critical values of u has zero Lebesgue measure. Now, suppose by contradiction that there exists ¯t[0,1) regular value such that, for all m¯m with 1¯m<1¯t, the interval (¯t1m,¯t+1m)[0,1) contains critical values. Hence there is a sequence {tm}m¯m of critical values such that tm¯t. In particular, there exists a sequence {pm}m¯m of critical points contained in the set {0u¯t+1m} and such that u(pm)=tm. Then, by compactness and up to a subsequence, pmp. In turn, 0=|Dg0u|g0(pm)|Dg0u|g0(p) and tm=u(pm)u(p). Hence |Dg0u|g0(p)=0 and u(p)=¯t, which is absurd. This concludes the proof of (iii).

    Remark 2.1. It is useful to observe that:

    (i) for every t(0,1), the set {ut} is connected;

    (ii) for every t1, the level set {u=t} is regular and diffeomorphic to Sn1;

    (iii) For every t(0,1), {ut}=¯{u>t} and {0ut}=¯{0<u<t}.

    We check (ii) first. We start by observing that due to (2.1) |Dg0u|g00 in {ut0}, for some 0<t0<1. This fact establishes a diffemorphism between {ut0} and {u=t0}×[t0,1) and tells us at the same time that the level sets {u=t} are pairwise diffeomorphic, for every tt0. It is thus sufficient to show that {u=t0} is connected. Suppose by contradiction that this is not the case. Without loss of generality we can assume that {u=t0} can be decomposed into the disjoint union of two connected sets C1 and C2, indeed the same argument works a fortiori if the connected components are more than two. Now, note that by definition of asymptotically flat manifold, there exists a compact set KM such that M˚K is diffeomorphic to Rn˚B by a chart at infinity ψ, where B is a suitable ball, and we can suppose, up to a bigger t0, that {ut0}M˚K. Now, in view of the asymptotic expansion of u, there exist two positive constants A<B such that

    A|x|n21uB|x|n2.

    In particular, setting R0=[B/(1t0)]1/(n2), we have that

    {|x|>R0}{ut0}{C1×[t0,1)}{C2×[t0,1)}.

    At the same time, we have that {|x|>R0} is connected and each Ci×[t0,1) is a closed set of M, so that indeed {|x|>R0}Ci×[t0,1), for some i{1,2}.Therefore, we have that

    {C1×[t0,1)}{C2×[t0,1)}={ut0}M˚K=[(M˚K){|x|R0}][(M˚K){|x|>R0}][(M˚K){|x|R0}]{Ci×[t0,1)},

    which gives the contradiction that the noncompact set {Cj×[t0,1)}, where j{1,2}{i}, is contained into the compact one (M˚K){|x|R0}. Therefore, {u=t0} is connected. Now, setting ˜u:=uψ1, we have that, up to a bigger t0 and due to (2.5), the set {˜u=t0} is a compact and connected hypersurface of Rn having strictly positive sectional curvature, as a Riemannian submanifold of (Rn,gRn). Hence, {˜u=t0} is diffeomorphic to Sn1 by the Gauss map (see [10,Section 5.B] for more details). Statement (ii) thus follows, being {u=t0} and {˜u=t0} diffeomorphic.

    To see (i), observe first that if t is a regular value of u, then Et:={ut} is a n–dimensional submanifold with boundary {u=t}. By Theorem 2.3 and by the Maximum Principle, every connected component C of Et is unbounded. Since u1 at , we have that u(C)=[t,1), and hence C{u=t0}, for every t0(t,1). Then, Et is connected by (ii). If t is a critical value of u, we let ¯t>t be a regular value of u such that {u=¯t} is connected and let {tm} be a nondecreasing sequence of regular value of u such that tm<t and tmt. Hence, {{tmu¯t}}mN is a nonincreasing family of connected and compact sets in M, which is Hausdorff, and in turn the intersection {tu¯t} is still connected. In particular, we deduce that Et={tu¯t}{u¯t} is connected.

    To check (iii), note first that for every t(0,1) regular value of u, the equalities are always true. If t(0,1) is a critical value of u, by Theorem 2.3 and by the Maximum Principle the interior of {0ut} and the interior of {ut} are both disjoint from {u=t}, so that (iii) is still true.

    Let (M,g0,u) be a sub-static harmonic triple, and let t[0,1) be a real number. We consider the spaces of measure

    ({u=t},B({u=t}),Hn1M{u=t})and(S:={u=t}Crit(u),Λ(S),σg0:=μιg0).

    Let f:SR be a continuous and consider the zero–extension ˚f of f defined on {u=t} as

    ˚f(p)={f(p)ifpS,0ifp{u=t}Crit(u).

    By Theorem 2.3 and by definition of the Lebesgue integral, we have that ˚fL1(Hn1M{u=t}) iff fL1(σg0) and

    {u=t}˚fd(Hn1M{u=t})=Sfdσg0. (2.26)

    Similarly, if f:{u=t}R is a continuous function, then fL1(Hn1M{u=t}) iff f|SL1(σg0) and

    {u=t}fd(Hn1M{u=t})=Sf|Sdσg0. (2.27)

    In the rest of this paper, we will confuse the integrals of cases (2.26) and (2.27), denoting both by {u=t}fdσg0.

    In this section, we state and prove our Monotonicity and Outer Rigidity Theorem, which is then used to prove the Capacitary Riemannian Penrose Inequality (1.7). From now on and unless otherwise stated, (M,g0,u) will always be a sub-static harmonic triple, and, when referring to such triple, the subscript g0 will be dropped. The only exception is |Sn1|, which always stands for the Euclidean volume of Sn1.

    Theorem 3.1 (Monotonicity and Outer Rigidity Theorem). Let (M,g0,u) be a sub-static harmonic triple, and let Fβ:τ[1,+)[0,+) be the function defined by

    Fβ(τ):=(1+τ)βn1n2{u=τ1τ+1}|Du|β+1dσ, (3.1)

    for every β0. Then, the following properties hold true.

    (i) Differentiability, Monotonicity and Outer Rigidity: for every β>n2n1, the function Fβ is continuously differentiable with nonpositive derivative in (1,+). Moreover, if there exists τ0(1,+) such that Fβ(τ0)=0 for some β>n2n1, then, setting t0=τ01τ0+1, the Riemannian submanifold {ut0}

    is isometric to

    ([r0,+)×Sn1,drdr12Cr2n+r2gSn1),r0=[C(1+τ0)]1n2.

    (ii) Convexity: for every β>n2n1, the function Fβ is convex on [1,).

    We remark that the functions Fβ are well–defined, in view of Theorem 2.3 and since the integrand function in (3.1) is bounded on every level set of u. Note that, from Theorem 3.1 and by a simple argument based on the Dominated Convergence Theorem, the monotonicity and the convexity of Fβ extend to the case β=n2n1. Moreover, on the values τ such that {u=τ1τ+1} is regular and thus on a.e. τ>1 due to Theorem 2.3 (iii), the function Fβ is twice differentiable for each β>n2n1, with first and second derivative given by

    Fβ(τ)=β(τ+1)βn1n232τ1{u=τ1τ+1}|Du|β[Hn1n22u1u2|Du|]dσ, (3.2)
    Fβ (3.3)

    In the computation, we have used the first normal variation of the volume and the mean curvature of \{u = t\} , and the Divergence Theorem. The symbols \mathrm{H} and \mathrm{h} stand respectively for the mean curvature and the second fundamental form of the smooth (n-1) –dimensional submanifold \big\{u = \sqrt{\frac{\tau-1}{\tau+1}}\, \big\} , with respect to the \infty –pointing unit normal vector field \nu = \frac{\mathrm{D}u}{\vert \mathrm{D}u \vert} . Also, \mathrm{D}^{\mathrm{T}} denotes the tangential part of the gradient, that is

    \mathrm{D}^{\mathrm{T}}f = \mathrm{D}f-g_{0}(\mathrm{D}f,\nu)\nu,\,

    for every f\in C^{1}(M) .

    To prove Theorem 3.1, we use the results of Section 4, which are obtained in the conformal setting defined by

    \begin{align} g = (1-u^{2})^{\frac{2}{n-2}}g_{0}\,,\,\,\quad\quad\quad\,\, \varphi = \log\Big(\frac{1+u}{1-u}\Big)\,. \end{align} (3.4)

    Denoting by \nabla and \Delta_{g} the Levi–Civita connection and the Laplace–Beltrami operator of g , the triple (M, g, \varphi) satisfies the following system.

    \begin{equation} \left\{ \begin{aligned} \mathrm{Ric}_{g}-\coth(\varphi)\,\nabla ^{2}\varphi+\frac{1}{n-2}\,d\varphi \otimes d\varphi-\frac{1}{n-2}\,\vert \nabla \varphi \vert^{2}_{g}\,\ g \,&\geq0 &&\mathrm{in}\, \mathring{M},\\ \Delta_{g}\varphi & = 0 &&\mathrm{in} \, M,\\ \varphi & = 0 &&\mathrm{on} \, \, \partial M,\\ \varphi &\to +\infty\,&&\mathrm{at}\,\infty. \end{aligned} \right . \end{equation} (3.5)

    Moreover, we have that

    \begin{align} \vert \nabla \varphi \vert_{g}^{2} = 4\,\vert \mathrm{D}u \vert^{2}(1-u^{2})^{-2\frac{n-1}{n-2}}\longrightarrow\big( 2\,\mathcal{C}\big)^{-\frac{2}{n-2}}(n-2)^{2}\,\,{\rm{at}}\, \,\infty\,,\, \end{align} (\star)

    as we will see in the proof of Lemma 4.1.

    Remark 3.1. Since \mathrm{Crit}(\varphi) = \mathrm{Crit}(u) = \{\vert \nabla \varphi \vert_{g} = 0\} by the equality in ( \star ) and since \{\varphi = s\} = \{u = \tanh\big(\frac{s}{2}\big)\} by (3.4), using Theorem 2.3 and Remark 2.1 we deduce that: \mathrm{Crit}(\varphi) has zero \mu_{g} –measure and zero (n-1) –Hausdorff measure in (M, g) ; the level sets of \varphi have finite (n-1) –Hausdorff measure in (M, g) and in particular the smooth (n-1) –dimensional submanifolds \{\varphi = s\}\setminus\mathrm{Crit}(\varphi) have finite g –area, i.e., finite \sigma_{g} –measure. Moreover, \{\varphi\geq s\} is connected for every s\geq 0 and there exists s_0\geq0 such that \{\varphi = s\} is regular and diffeomorphic to \mathbb{S}^{n-1} , for every s\geq s_0 . Similar comments as those at the end of Subsection 2.2 can be made, regarding the relation between integration and \mathrm{Crit}(\varphi) .

    Let \Phi_{\beta}:[0, \infty)\to \mathbb{R} be the function defined by formula

    \begin{align*} \Phi_{\beta}(s): = \int\limits_{\{\varphi = s\}}\vert \nabla \varphi \vert^{\beta+1}_{g}\,d\sigma_{g}\,, \end{align*}

    for every \beta\geq 0 . For the convenience of the reader, we anticipate from Section 4 the properties of \Phi_{\beta} that we are going to use.

    ( \circ ) For every \beta\geq0 , the function \Phi_{\beta}(s) is continuous in [0, +\infty) .

    ( \diamond ) For every \beta > \frac{n-2}{n-1} , the function \Phi_{\beta} is continuously differentiable in (0, +\infty) . The derivative \Phi'_{\beta} is nonpositive, satisfies for every S > s > 0

    \begin{align} \frac{\Phi_{\beta}'(S)}{\sinh(S)}-\frac{\Phi_{\beta}'(s)}{\sinh(s)}\geq0\,, \end{align} (3.6)

    and admits for every s > 0 the integral representation

    \begin{align*} \Phi'_{\beta}(s) = -\beta\, \sinh( s) \,\int\limits_{\{\varphi > s\}}\frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,(\beta-2)\,\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}+\vert \nabla^{2} \varphi \vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\, \Big]}{\sinh\varphi}\,d\mu_{g}\leq 0\,, \end{align*}

    where Q is defined in (4.10).

    ( \diamond\, \diamond ) If there exists s_{0} > 0 such that \Phi'_{\beta}(s_{0}) = 0 for some \beta > \frac{n-2}{n-1} , then \{\varphi = s_{0}\} is connected and (\{\varphi\geq s_{0}\}, g) is isometric to \big([0, +\infty)\times \{\varphi = s_{0}\}, d\rho\otimes d\rho+g_{\{\varphi = s_{0}\}}) , where \rho is the g –distance function to \{\varphi = s_{0}\} and \varphi is an affine function of \rho in \{\varphi\geq s_{0}\} . If \Phi_{\beta} is constant for some \beta > \frac{n-2}{n-1} , then \partial M is connected and (M, g) is isometric to \big([0, +\infty)\times \partial M, d\rho\otimes d\rho+g_{\partial M}) , where \rho is the g –distance function to \partial M and \varphi is an affine function of \rho .

    In the above list we have gathered and summarised the results contained in Lemma 4.5, Proposition 4.6, and Corollary 4.7.

    Proof of Theorem 3.1. Step 1 : Differentiability, Monotonicity and Convexity. For every \beta\geq0 and for all \tau \in[1, +\infty) , we note that

    \begin{align} F_{\beta}(\tau) = 2^{\frac{\beta}{n-2}-1}\,\, \Phi_{\beta}\Bigg(\log\Bigg(\,\frac{\sqrt{\tau+1}+\sqrt{\tau-1}}{\sqrt{\tau+1}-\sqrt{\tau-1}}\,\Bigg) \Bigg) \,. \end{align} (3.7)

    Consequently, by (\circ) , we deduce that for every \beta\geq0 the function F_{\beta} is continuous in [1, +\infty) . By (\diamond) and by (3.7), we obtain immediately that for every \beta > \frac{n-2}{n-1} the function F_{\beta} is continuously differentiable in (1, +\infty) , with

    \begin{align} F'_{\beta}(\tau) = 2^{\frac{\beta}{n-2}-1}\,\frac{1}{\sqrt{\tau^{2}-1}}\,\Phi'_{\beta}\Bigg(\log\Bigg(\,\frac{\sqrt{\tau+1}+\sqrt{\tau-1}}{\sqrt{\tau+1}-\sqrt{\tau-1}}\,\Bigg) \Bigg)\,. \end{align} (3.8)

    In particular, from (\diamond) we get F'_{\beta}\leq 0 . As for the convexity, noticing that

    \sqrt{\tau^{2}-1} = \sinh \Bigg(\log\Bigg(\,\frac{\sqrt{\tau+1}+\sqrt{\tau-1}}{\sqrt{\tau+1}-\sqrt{\tau-1}}\,\Bigg) \Bigg)

    and that the function \log\big(\, \frac{\sqrt{\tau+1}+\sqrt{\tau-1}}{\sqrt{\tau+1}-\sqrt{\tau-1}}\, \big) is nondecreasing, from (3.6) we obtain that F'_{\beta}\, is nondecreasing in (1, +\infty) . Therefore also by the continuity of F_{\beta} at 1 , F_{\beta} is convex in [1, +\infty) .

    Step 2 : Outer Rigidity. Let us assume that there exists \tau_{0} \in (1, \infty) such that F'_{\beta}(\tau_{0}) = 0 for some \beta > \frac{n-2}{n-1} . Then, by equality (3.8), \Phi'_{\beta}(s_{0}) = 0 for s_{0} = \log\big(\, \frac{\sqrt{\tau_{0}+1}+\sqrt{\tau_{0}-1}}{\sqrt{\tau_{0}+1}-\sqrt{\tau_{0}-1}}\, \big) , and hence, by (\diamond\, \diamond) , (\{\varphi\geq s_{0}\}, g) is isometric to \big([0, +\infty)\times \{\varphi = s_{0}\}, d\rho\otimes d\rho+g_{\{\varphi = s_{0}\}}) , where \rho is the g –distance function to \{\varphi = s_{0}\} and \varphi = (n-2)\, (2\mathcal{C})^{-\frac{1}{n-2}}\rho+s_{0} , because \vert \nabla \rho\vert_{g}\equiv1 and in view of the limit in (\star) . Setting t_{0} = \tanh \frac{s_{0}}{2} , consider N the submanifold with boundary \{\varphi\geq s_{0}\} = \{u\geq t_{0}\} . Writing

    \begin{align*} &N &\,& &&[0,+\infty)\times \partial N &\,& &&[s_{0},+\infty)\times \partial N &\,& &&[t_{0},1)\times \partial N \\ &g &\,& &&d\rho\otimes d\rho+g_{\partial N} &\,& & &\frac{d\varphi\otimes d\varphi}{(n-2)^{2}\, (2\mathcal{C})^{-\frac{2}{n-2}}}+g_{\partial N} &\,& &&\begin{array}{l} \frac{{{2^{2{\kern 1pt} \frac{{n - 1}}{{n - 2}}}}{\mkern 1mu} {\mkern 1mu} {{\cal C}^{\frac{2}{{n - 2}}}}}}{{{{(n - 2)}^2}{\mkern 1mu} {{(1 - {u^2})}^2}}}{\mkern 1mu} \\ {\mkern 1mu} du \otimes du + {g_{\partial N}} \end{array}\\ &p &\mapsto& &&\big(\rho,q\big) &\mapsto& &&\big (\varphi = (n-2)\, (2\mathcal{C})^{-\frac{1}{n-2}} \,\rho, q\big) &\mapsto& && \Big(u = \tanh\, \frac{\varphi}{2} , q\Big)\,, \end{align*}

    the Riemannian manifolds in the first row, whose metrics are indicated in the second row, are pairwise isometric through the applications written in the third row. We recall that the application p\to (\rho, q) in the third row is the inverse of the diffeomorphism given by the normal exponential map, i.e., the application which associates to every point p of N the couple having as first coordinate the g –distance of p from \partial N and as second coordinate the point q of \partial N that realizes such distance. Then, in view of (3.4) and with the same notation as above, the following Riemannian manifolds are isometric.

    \begin{align} &N \;\;\;\; [t_{0},1)\times \partial N &\,& \;\;\;\; [r_{0},+\infty)\times \partial N\\ &g_{0} \;\; \;\; \frac{2^{2\,\frac{n-1}{n-2}}\,\, \mathcal{C}^{\frac{2}{n-2}} }{(n-2)^{2}\,(1-u^{2})^{2\,\frac{n-1}{n-2}} }\,\, du\otimes du +(1-u^{2})^{-\frac{2}{n-2}} g_{\partial N} &\,& \;\;\;\; \frac{dr \otimes dr}{1-2\,\mathcal{C} \,r^{2-n}}+ (2\,\mathcal{C})^{-\frac{2}{n-2}}\, r^{2}g_{\partial N} \\ &p \;\;\;\; \mapsto\;\; \big(u, q\big) &\mapsto& \;\;\;\;\Big (r = \Big(\,\frac{2\mathcal{C}}{1-u^{2}}\,\Big)^{\frac{1}{n-2}},q\Big)\,,\, \end{align} (3.9)

    where r_{0} = \big(\frac{2\mathcal{C}}{1-t_{0}^{2}}\big)^{\frac{1}{n-2}} . Doing some computations, we obtain that

    (3.10)

    where the convection followed for the Riemannian curvature tensor is that given in [23], c is a suitable positive constant and q is the point of \partial N that realizes the g –distance of p from \partial N . Denoting by \Theta the diffeomorphism from N to [r_{0}, +\infty)\times \partial N introduced in (3.9), for every q_{0}\in \partial N we consider the curve

    \gamma:r\in [r_{0},+\infty)\to \Theta^{-1}(r,q_{0})\in M

    and observe from (3.10) that

    (3.11)

    At the same time, we have that

    \begin{align} r^{4}\,\vert \mathrm{Rm}\vert^{2}\,(\gamma(r)) \xrightarrow{r\rightarrow +\infty} 0\,. \end{align} (3.12)

    This is because g_{0} is asymptotically flat according to Definition 1.1 and by (2.3), which yields in particular

    \begin{align*} \vert \mathrm{Rm}\vert = O(\vert x\vert^{-(p+2)})\quad\quad{\rm{and}}\quad\quad \frac{r}{\vert x\vert}&\xrightarrow{\vert x\vert\rightarrow +\infty}1\,, \end{align*}

    for some p > \frac{n-2}{2} . Combining (3.11) and (3.12), the arbitrariness of the point q_{0} in \partial N gives that

    Hence the sectional curvature of the Riemannian manifold (\partial N, g_{\partial N}) is constant and identically equal to (2\mathcal{C}\, )^{-\frac{2}{n-2}} . Then, being all the level sets \{u = t\} with t\approx1 regular and diffeomorphic to \mathbb{S}^{n-1} as observed in Remark 2.1, for [10,Section 3.F] (\partial N, g_{\partial N}) and (\mathbb{S}^{n-1}, (2\mathcal{C}\, )^{\frac{2}{n-2}}g_{ \mathbb{S}^{n-1}}) are isometric. Then, (\{u\geq t_{0}\}, g_{0}) is isometric to the submanifold \big(\, [r_{0}, +\infty)\times \mathbb{S}^{n-1}\, , \frac{dr \otimes dr}{1-2\mathcal{C}r^{2-n}}+r^{2}g_{ \mathbb{S}^{n-1}} \big) of the Schwarzschild manifold with associated ADM mass given by \mathcal{C} .

    Proof of Theorem 1.1. Spep 1 : Inequality. By Theorem 3.1, we have that F_{\beta}(\tau_{0})\geq \lim_{\tau\to+\infty}\, F_{\beta}(\tau) , for every \tau_{0} > 1 . In particular, since F_{\beta} is continuous in [1, +\infty) due to the step 1 of Theorem 3.1, we have that

    \begin{align} F_{\beta}(1)\geq \lim\limits_{\tau\to+\infty}\,F_{\beta}(\tau)\,, \end{align} (3.13)

    for every \beta > \frac{n-2}{n-1} . Since \mathrm{D}^{2}u \equiv 0 on \partial M and since \partial{M} of M is connected, \vert \mathrm{D} u \vert is constantly equal to \frac{(n-2)\, \mathcal{C}\, \vert \mathbb{S}^{n-1}\vert\, }{\vert \partial M\vert}\, , by formula (2.2). In particular, we have that

    \begin{align} F_{\beta}(1) = \frac{2^{\beta\,\frac{n-1}{n-2}} \,\,(n-2)^{\beta+1}\,\,\mathcal{C}^{\beta+1}\,\,\vert \mathbb{S}^{n-1}\vert^{\beta+1}}{\vert \partial M\vert^{\beta}} \end{align} (3.14)

    By (\star) , we know that

    \begin{align*} \frac{\vert \mathrm{D}u \vert\,\,\,\,\,\,}{\,\,(1-u^{2})^{\frac{n-1}{n-2}}}\, \longrightarrow \, 2^{-\frac{n-1}{n-2}}\, (n-2)\,\mathcal{C}^{-\frac{1}{n-2}}\quad \,\, {\rm{at}}\,\,\infty \end{align*}

    Therefore, fixed \varepsilon > 0 , there exists 1 < \tau_{0} < +\infty such that

    \begin{align*} \vert \mathrm{D}u \vert\geq \,(1-u^{2})^{\frac{n-1}{n-2}}\Big( 2^{-\frac{n-1}{n-2}}\, (n-2)\,\mathcal{C}^{-\frac{1}{n-2}}-\varepsilon\Big) \end{align*}

    in \big\{u\geq\sqrt{\frac{\tau_{0}-1}{\tau_{0}+1}}\, \big\} and the level sets \big\{u = \sqrt{\frac{\tau-1}{\tau+1}}\, \big\} are regular for all \tau\geq \tau_{0} . Therefore, for every \tau\geq \tau_{0} we have that

    \begin{align*} F_{\beta}(\tau)& = (1+\tau)^{\beta\, \frac{n-1}{n-2}}\,\int\limits_{\big\{u = \sqrt{\frac{\tau-1}{\tau+1}}\,\big\}}\,\,\vert \mathrm{D}u \vert^{\beta+1}\,d\sigma\\ &\geq (1+\tau)^{\beta\, \frac{n-1}{n-2}}\,\int\limits_{\big\{u = \sqrt{\frac{\tau-1}{\tau+1}}\,\big\}}\,\, (1-u^{2})^{\beta\,\frac{n-1}{n-2}}\,\Big( 2^{-\frac{n-1}{n-2}}\, (n-2)\,\mathcal{C}^{-\frac{1}{n-2}}-\varepsilon\Big)^{\beta} \,\,\vert \mathrm{D}u \vert\,d\sigma\\ & = 2^{\beta\,\frac{n-1}{n-2}}\,\Big( 2^{-\frac{n-1}{n-2}}\, (n-2)\,\mathcal{C}^{-\frac{1}{n-2}}-\varepsilon\Big)^{\beta} \,\,\int\limits_{\partial M }\,\vert \mathrm{D}u \vert\,d\sigma\\ & = 2^{\beta\,\frac{n-1}{n-2}}\,(n-2)\,\mathcal{C} \,\Big( 2^{-\frac{n-1}{n-2}}\, (n-2)\,\mathcal{C}^{-\frac{1}{n-2}}-\varepsilon\Big)^{\beta}\,\vert \mathbb{S}^{n-1}\vert\,, \end{align*}

    where in the second equality we have used the Divergence Theorem couple with the fact that u is harmonic, and in the third equality we have used formula (2.2). Since \varepsilon is arbitrary, we get

    \lim\limits_{\tau\to+\infty}\,F_{\beta}(\tau)\geq\,(n-2)^{\beta+1}\,\mathcal{C}^{1-\frac{\beta}{n-2}} \,\vert \mathbb{S}^{n-1}\vert \,.

    In a similar way we can obtain the reverse inequality, so that

    \begin{align} \lim\limits_{\tau\to+\infty}\,F_{\beta}(\tau) = \,(n-2)^{\beta+1}\,\mathcal{C}^{1-\frac{\beta}{n-2}} \,\vert \mathbb{S}^{n-1}\vert\,. \end{align} (3.15)

    Joining the formulas in (3.13), (3.14) and (3.15), we obtain the desired inequality (1.7).

    Step 2 : Rigidity. If (M, g_{0}) is isometric to the Schwarzschild manifold with ADM mass m > 0 , then the right–hand side and the left–hand side of (1.7) are both equal to m , by direct computation.

    Suppose now that the equality holds in (1.7). Then, by Step 1 and for every \beta > \frac{n-2}{n-1} , the function F_{\beta} is constant. In turn, \Phi_{\beta} is constant, being

    \Phi_{\beta}(s) = 2^{1-\frac{\beta}{n-2}}\, F_{\beta}\Bigg(\,\frac{1+\tanh^{2}\big( \frac{s}{2}\big)}{1-\tanh^{2}\big( \frac{s}{2}\big)}\,\Bigg)\,.

    Finally, (\diamond\, \diamond) and the very same argument of the proof of the Outer Rigidity in Theorem 3.1 imply first that (M, g) is isometric to

    \big([0,+\infty)\times \partial M,d\rho\otimes d\rho+g_{\partial M}),

    where \rho is the g –distance to \partial M and \varphi is an affine function of \rho , and secondly that (M, g_{0}) is isometric to the Schwarzschild manifold with ADM mass \mathcal{C} .

    Let us consider the conformal change g of the metric g_0 introduced in (3.4) which is well–defined being 0\leq u < 1 in M . The metric g is complete, since any g –geodesic \gamma parametrized by g –arc length defined on a bounded interval [0, a) can be extended to a continuous path on [0, a] . Indeed, if \gamma has infinity length with respect to g_{0} , there exists a sequence \{t_{m}\}_{m\in \mathbb{N}} such that \gamma(t_{m})\to \infty (being \gamma not contained in any compact set) and using, in the computation of g –length of \gamma , the passage from g to g_{0} , the asymptotic flatness of (M, g_{0}) and the asymptotic expansion of u in (2.3) we obtain that \gamma has infinity length with respect to g . Hence \gamma has finite length with respect to g_{0} and, being g_{0} complete, it follows that g is complete (see [24,Section 1.1] and [9]). We also recall that the metric g is asymptotically cylindrical (see [2,Section 3.1]). The other main element of the conformal setting is the C^{\infty} –function \varphi , defined in (3.4). Now, the reverse changes are

    g_{0} = \Big(\cosh\frac{\varphi}{2}\Big)^{\frac{4}{n-2}}g\,, \quad\quad\quad u = \tanh \frac{\varphi}{2}\,.

    Recalling that we denote by the symbols \nabla and \Delta_{g} the Levi–Civita connection and the Laplace–Beltrami operator of g , by the formulas in [6,Theorem 1.159], we obtain

    \begin{align} \mathrm{D}u& = \frac{1}{2}\Big(\cosh \frac{\varphi}{2}\Big)^{-\frac{2n}{n-2}}\,\,\nabla \varphi\,, \end{align} (4.1)
    \begin{align} \mathrm{D}^{2}u& = \frac{1}{2}\, \frac{1}{\cosh^{2}\frac{\varphi}{2}}\,\nabla^{2}\varphi- \frac{n}{2(n-2)}\,\frac{\sinh\frac{\varphi}{2}}{\cosh^{3}\frac{\varphi}{2}}\, d\varphi \otimes d\varphi+\frac{1}{2(n-2)}\,\frac{\sinh\frac{\varphi}{2}}{\cosh^{3}\frac{\varphi}{2}}\,\vert \nabla \varphi \vert^{2}_{g}\,g\,, \end{align} (4.2)
    \begin{align} \Delta u& = \frac{1}{2}\Big(\cosh \frac{\varphi}{2}\Big)^{-\frac{2n}{n-2}}\Delta_{g}\varphi\,, \end{align} (4.3)
    \begin{align} \mathrm{Ric}& = \mathrm{Ric}_{g}-\tanh\Big(\frac{\varphi}{2}\Big)\,\nabla ^{2}\varphi+\Big[\frac{1}{n-2}\tanh^{2}\Big( \frac{\varphi}{2}\Big)-\frac{1}{2}\frac{1}{\cosh^{2}\frac{\varphi}{2}}\Big]d\varphi \otimes d\varphi \\ &-\frac{1}{(n-2)}\,\Big[\,\frac{1}{2}\,\frac{1}{\cosh^{2}\frac{\varphi}{2}}+\tanh^{2}\Big( \frac{\varphi}{2}\Big)\Big]\vert \nabla \varphi \vert^{2}_{g}\,g\,.\nonumber \end{align}

    Translating system (1.1) in terms of g and \varphi , we get system (3.5). Moreover, on \{\varphi = s\}\setminus\mathrm{Crit}(\varphi) we consider the \infty –pointing normal unit vector fields

    \begin{align*} \nu = \frac{ \mathrm{D}u }{\vert \mathrm{D}u \vert}\,, \quad \quad \quad \nu_{g} = \frac{\nabla \varphi}{\vert \nabla \varphi \vert_{g}}\,, \end{align*}

    the mean curvatures

    \begin{align} \mathrm{H} = -\, \frac{\mathrm{D}^{2}u(\mathrm{D}u,\mathrm{D}u)}{\vert \mathrm{D}u \vert^{3}}\,,\quad \quad \quad \mathrm{H}_{g} = -\,\frac{\nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)}{\vert \nabla \varphi \vert_{g}^{3}}\,,\, \end{align} (4.4)

    and the second fundamental forms

    \begin{align*} \mathrm{h}(X,Y) = \frac{\mathrm{D}^{2}u\,(X,Y)}{\vert \mathrm{D}u \vert}\,,\quad\quad\quad \mathrm{h}_{g}(X,Y) = \frac{\nabla ^{2}\varphi\,(X,Y)}{\vert \nabla \varphi \vert_{g}}\,, \end{align*}

    for any X, Y tangent vector fields to the considered submanifold.

    Reversing formulas (4.1), (4.2) and (4.3), we get

    \begin{array}{l} \nabla \varphi = \frac{2}{(1-u^{2})^{\frac{n}{n-2}}}\, \mathrm{D}u\,,\\ \nabla ^{2}\varphi = \frac{2}{1-u^{2}}\,\Big[\,\mathrm{D}^{2}u\,+\, \frac{n}{n-2}\,\frac{2u}{1-u^{2}}\, du\otimes du\,-\,\frac{1}{n-2}\,\frac{2u}{1-u^{2}}\,\vert \mathrm{D}u \vert^{2}\,g_{0}\,\Big]\,,\\ \vert \nabla^{2} \varphi \vert_{g}^{2} = \frac{4}{(1-u^{2})^{\frac{2n}{n-2}}}\,\vert \mathrm{D}^{2}u \vert^{2}+\frac{16n}{n-2}\,\frac{u}{(1-u^{2})^{\frac{3n-2}{n-2}}}\,\mathrm{D}^{2}u(\mathrm{D}u,\mathrm{D}u)+ \\ \frac{16n(n-1)}{(n-2)^{2}}\, \frac{u^{2}}{(1-u^{2})^{4\big(\frac{n-1}{n-2}\big)}}\,\vert \mathrm{D}u \vert^{4}\,. \end{array}

    These equalities, jointly with the asymptotic flatness of (M, g_{0}) and the asymptotic expansion of u given in Section 2, allow us to obtain an upper bound for the functions \vert \nabla \varphi \vert_{g} and \vert \nabla^{2} \varphi \vert_{g} , and for the g –areas of the level sets of \varphi sufficiently "close" to infinity. This is the content of the following lemma.

    Lemma 4.1. There exists 0\leq s_{0} < +\infty such that

    \begin{align} \sup\limits_{M}\vert \nabla \varphi \vert_{g}\,+\,\sup\limits_{M}\vert \nabla^{2} \varphi \vert_{g}\,+\,\sup\limits_{s\geq s_{0}}\int\limits_{\{\varphi = s\}}d\sigma_{g} < \,+\infty\,. \end{align} (4.5)

    Proof. Let \psi be a chart at infinity. Considering \widetilde{g}_{0} = \psi_{*}g_{0} = \widetilde{g}_{0;ij}dx^{i}\otimes dx^{j} , by formulas (2.6) and (2.4), the coordinate expression of

    \vert \nabla \varphi \vert_{g}^{2} = \frac{4\,\vert \mathrm{D}u \vert^{2}}{(1-u^{2})^{2\frac{n-1}{n-2}}}

    is

    \begin{align*} \psi_{*}\vert \nabla \varphi \vert_{g}^{2}& = \frac{4\,\psi_{*}\vert \mathrm{D}u \vert^{2}}{(1-\widetilde{u}^{2})^{2\frac{n-1}{n-2}}} = \frac{4\,\widetilde{g}_{0}^{ij}\,\partial_{i}\widetilde{u}\,\partial_{j}\widetilde{u} }{\bigl\{1-[1-\mathcal{C}\vert x\vert^{2-n}+o(\vert x\vert^{2-n})]^{2}\bigl\}^{2\,(\frac{n-1}{n-2})}}\\ & = \frac{4\,\big[ \delta^{ij}+O\big(\vert x\vert^{-p}\big)\big]\,\big[(n-2)^{2}\,\mathcal{C}^{2}\,\vert x\vert^{-2n}\,x^{i}\,x^{j}+o(\vert x\vert^{2-2n})\big]}{\big[ 2\,\mathcal{C}\,\vert x\vert^{2-n}+o(\vert x\vert^{2-n})\big]^{2\,(\frac{n-1}{n-2})}}\\ & = \frac{4\,(n-2)^{2}\,\mathcal{C}^{2}\,\vert x\vert^{2-2n}+o(\vert x\vert^{2-2n})}{\big(2\,\mathcal{C}\big)^{2\,(\frac{n-1}{n-2})}\, \vert x\vert^{2-2n}\,\big(1+o(1)\big)} = \frac{4\,(n-2)^{2}\,\mathcal{C}^{2}+o(1)}{\big(2\,\mathcal{C}\big)^{2\,(\frac{n-1}{n-2})}\,\big(1+o(1)\big)}\,. \end{align*}

    Hence

    \begin{equation} \vert \nabla \varphi \vert_{g}^{2}\longrightarrow \frac{4\,(n-2)^{2}\,\mathcal{C}^{2}}{\big(2\,\mathcal{C}\big)^{2\,(\frac{n-1}{n-2})}} = \big( 2\,\mathcal{C}\big)^{-\frac{2}{n-2}}(n-2)^{2}\,\,{\rm{at}}\, \,\infty\,. \end{equation} (4.6)

    Moreover, by limit (4.6) there exist a constant L > 0 and a value s_{0} > 0 of \varphi such that every s\geq s_{0} is a regular value of \varphi and (1-u^{2})^{\frac{n-1}{n-2}}\leq L\, \, \vert \mathrm{D}u \vert on \{\varphi\geq s_{0}\} . Then

    \begin{align*} \int\limits_{\{\varphi = s\}}d\sigma_{g} = \int\limits_{\{u = \tanh\frac{s}{2}\}}(1-u^{2})^{\frac{n-1}{n-2}}\,d\sigma\leq L\int\limits_{\{u = \tanh\frac{s}{2}\}}\vert \mathrm{D}u \vert\,d\sigma = L\int\limits_{\partial M}\vert \mathrm{D}u \vert\,d\sigma\,, \end{align*}

    where in the last equality we have applied the Divergence Theorem. Consequently, we have that \sup\limits_{s\geq s_{0}}\int\limits_{\{\varphi = s\}}d\sigma_{g} < +\infty . Similarly, we have that

    \begin{align*} \psi_{*}\vert \mathrm{D}^{2}u \vert^{2}\!& = \!\widetilde{g}_{0}^{\,i_{1}i_{2}}\widetilde{g}_{0}^{\,j_{1}j_{2}}(\mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u})_{i_{1}j_{1}}\!(\mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u})_{i_{2}j_{2}}\! = \!\widetilde{g}_{0}^{\,i_{1}i_{2}}\widetilde{g}_{0}^{\,j_{1}j_{2}}\big[\partial _{i_{1}}\partial_{j_{1}}\widetilde{u}-\Gamma_{\widetilde{g}_{0};i_{1}j_{1}}^{k_{1}}\partial _{k_{1}}\widetilde{u}\big] \big[\partial _{i_{2}}\partial_{j_{2}}\widetilde{u}-\Gamma_{\widetilde{g}_{0};i_{2}j_{2}}^{k_{2}}\partial _{k_{2}}\widetilde{u}\big]\\ & = [ \delta^{\,i_{1}i_{2}}\delta^{\,j_{1}j_{2}}+O\big(\vert x\vert^{-p}\big)\big]\,\big[\partial _{i_{1}}\partial_{j_{1}}\widetilde{u}-O(\vert x\vert^{-(p+n)})\big]\, \big[\partial _{i_{2}}\partial_{j_{2}}\widetilde{u}-O(\vert x\vert^{-(p+n)})\big]\\ & = \big[(n-1)(n-2)\mathcal{C}\big]^{2}\vert x\vert^{-2n}+o(\vert x\vert^{-2n}) \end{align*}

    due to formulas (2.6), (2.5) and (5.3). Moreover,

    \begin{align*} \psi_{*}\mathrm{D}^{2}u(\mathrm{D}u,\mathrm{D}u)& = \mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u}(\mathrm{D}_{\widetilde{g}_{0}}\widetilde{u},\mathrm{D}_{\widetilde{g}_{0}}\widetilde{u}) = \widetilde{g}_{0}^{\,ir}\,\widetilde{g}_{0}^{\,js}\,\partial_{r}\widetilde{u}\,\partial_{s}\widetilde{u}\,\big(\partial _{i}\partial_{j}\widetilde{u}-\Gamma_{\widetilde{g}_{0};ij}^{k}\partial _{k}\widetilde{u})\,\\ & = -(n-1)\,(n-2)^{3}\,\mathcal{C}^{3}\vert x\vert^{2-3n}+o(\vert x\vert^{2-3n})\,. \end{align*}

    All in all,

    \begin{align*} \psi_{*}\vert \nabla^{2} \varphi \vert_{g}^{2}& = \frac{4}{(1-\widetilde{u}^{2})^{\frac{2n}{n-2}}}\,\psi_{*}\vert \mathrm{D}^{2}\widetilde{u} \vert^{2}+\frac{16n}{n-2}\,\frac{\widetilde{u}}{(1-\widetilde{u}^{2})^{\frac{3n-2}{n-2}}}\,\psi_{*}\mathrm{D}^{2}\widetilde{u}(\mathrm{D}\widetilde{u},\mathrm{D}\widetilde{u})\\ &+\frac{16n(n-1)}{(n-2)^{2}}\,\frac{\widetilde{u}^{2}}{(1-\widetilde{u}^{2})^{4\big(\frac{n-1}{n-2}\big)}}\,\psi_{*}\vert \mathrm{D}\widetilde{u} \vert^{4}\\ & = (n-1)\,(n-2)^{2}\,(2\,\mathcal{C})^{-\frac{4}{n-2}}\,\Bigr\{\frac{n-1+o(1)}{1+o(1)}-\frac{2\,n\,\widetilde{u}+o(1)}{1+o(1)} +\frac{16\,n\,\widetilde{u}^{2}+o(1)}{1+o(1)}\Bigr\}\,, \end{align*}

    which gives

    \begin{align*} \vert \nabla^{2} \varphi \vert_{g}^{2}\longrightarrow (n-1)\,(15n-1)\,(n-2)^{2}\,(2\,\mathcal{C})^{-\frac{4}{n-2}}\quad\,\,{\rm{at}}\, \,\infty\,. \end{align*}

    In particular, since \varphi is smooth, we have that

    \sup\limits_{M}\vert \nabla \varphi \vert_{g}\,+\,\sup\limits_{M}\vert \nabla^{2} \varphi \vert_{g}\, < \,+\infty\,.

    Remark 4.1. Note that \sup\limits_{s\geq 0}\int\limits_{\{\varphi = s\}}d\sigma_{g}\in(0, +\infty] , since we cannot a priori exclude that there exist a critical value \overline{s} > 0 and a sequence \{s_{m}\}\subset (0, +\infty) such that s_{m}\to\overline{s} and

    \int\limits_{\{\varphi = s_{m}\}}d\sigma_{g}\to +\infty\,.

    As it will be clear in the proof of the integral identity (4.17), which is at the core of the conformal–monotonicity result (Proposition 4.6), it is useful to introduce a suitable vector field with nonnegative divergence. To do this, let us focus on the set \mathring{M}\setminus\mathrm{Crit}(\varphi) and notice first that the classical Bochner formula, applied to the g –harmonic function \varphi , becomes

    \begin{align} \frac{1}{2}\,\Delta_{g}\vert \nabla \varphi \vert^{2}_{g}\, = \,\vert \nabla^{2} \varphi \vert_{g}^{2}\,+\,\mathrm{Ric}_{g}(\nabla \varphi ,\nabla \varphi )+\,g(\nabla\Delta_{g}\varphi,\nabla \varphi )\, = \,\vert \nabla^{2} \varphi \vert_{g}^{2}\,+\,\mathrm{Ric}_{g}(\nabla \varphi ,\nabla \varphi )\,. \end{align} (4.7)

    Then, we obtain

    \begin{align} \Delta_{g}\vert \nabla \varphi \vert^{\beta}_{g}& = \mathrm{div}_{g}\Big(\nabla \vert \nabla \varphi \vert^{\beta}_{g}\Big) = \mathrm{div}_{g}\Big(\frac{\beta}{2}\vert \nabla \varphi \vert^{\beta-2}_{g}\nabla\vert \nabla \varphi \vert^{2}_{g}\Big)\\ & = \frac{\beta}{2}\Big[g(\nabla \vert \nabla \varphi \vert^{\beta-2}_{g},\nabla\vert \nabla \varphi \vert^{2}_{g})+ \vert \nabla \varphi \vert^{\beta-2}_{g}\Delta_{g}\vert \nabla \varphi \vert^{2}_{g} \Big]\\ & = \beta\,\vert\nabla \varphi \vert^{\beta-2}_{g}\Big[\,(\beta-2)\,\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}+\vert \nabla^{2} \varphi \vert_{g}^{2}+\mathrm{Ric}_{g}(\nabla \varphi ,\nabla \varphi )\Big]\,, \end{align} (4.8)

    where in the third equality we have used (4.7). Now, observe from the nonnegativity of the tensor

    \begin{equation} Q\,: = \,\mathrm{Ric}_{g}-\coth(\varphi)\,\nabla ^{2}\varphi+\frac{1}{n-2}\,d\varphi \otimes d\varphi-\frac{1}{n-2}\,\vert \nabla \varphi \vert^{2}_{g}\,\ g \end{equation} (4.9)

    (see (3.5)) that

    \begin{align} Q( \nabla \varphi, \nabla \varphi) = \mathrm{Ric}_{g}( \nabla \varphi, \nabla \varphi)-\coth(\varphi)\,\nabla ^{2}\varphi( \nabla \varphi, \nabla \varphi)\geq 0\,. \end{align} (4.10)

    Therefore, by adding and subtracting the term \beta\, \vert\nabla \varphi \vert^{\beta-2}_{g}\, \coth(\varphi)\, \nabla ^{2}\varphi(\nabla \varphi, \nabla \varphi) on the right–hand side of (4.8), we get

    \begin{align} \Delta_{g}\vert \nabla \varphi \vert^{\beta}_{g}-\beta\,\vert\nabla \varphi \vert^{\beta-2}_{g}&\,\coth(\varphi)\,\nabla ^{2}\varphi( \nabla \varphi, \nabla \varphi)\\ & = \beta\vert \nabla \varphi \vert^{\beta-2}_{g}\Big[\,(\beta-2)\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}+\vert \nabla^{2} \varphi \vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\,\Big]\,. \end{align} (4.11)

    Since

    \beta\,\vert\nabla \varphi \vert^{\beta-2}_{g}\,\coth(\varphi)\,\nabla ^{2}\varphi( \nabla \varphi, \nabla \varphi) = \coth(\varphi)\, g(\nabla\vert \nabla \varphi \vert_{g}^{\beta}, \nabla \varphi)

    and since, setting

    \begin{align} Y_{\beta}: = \frac{\nabla\, \vert \nabla \varphi \vert^{\beta}_{g}}{\sinh \varphi}\,, \end{align} (4.12)

    there holds

    \mathrm{div}_{g}\,Y_{\beta} = \frac{\Delta_{g}\vert \nabla \varphi \vert^{\beta}_{g}}{\sinh \varphi}-\frac{\cosh \varphi}{\sinh^2 \varphi}\, \,g(\nabla\vert \nabla \varphi \vert_{g}^{\beta}, \nabla \varphi)\,,

    from (4.11) we get

    \begin{align*} \sinh (\varphi)\, \mathrm{div}_{g}\,Y_{\beta} = \beta\vert \nabla \varphi \vert^{\beta-2}_{g}\Big[\,(\beta-2)\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}+\vert \nabla^{2} \varphi \vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\,\Big]\,. \end{align*}

    Note that, by the refined Kato inequality for harmonic function

    \begin{align} \frac{n}{n-1}\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}\leq \vert \nabla^{2} \varphi \vert_{g}^{2}\,, \end{align} (4.13)

    we have that

    \begin{align} (\beta-2)\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}&+\vert \nabla^{2} \varphi \vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\\ & = \Big(\beta-\frac{n-2}{n-1}\Big)\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}+\Bigg[\vert \nabla^{2} \varphi \vert_{g}^{2}-\frac{n}{n-1}\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}\Bigg]+\\& Q(\nabla \varphi ,\nabla \varphi )\geq0\,,\, \end{align} (4.14)

    whenever \beta\geq\frac{n-2}{n-1} . Hence, \mathrm{div}_{g}\, Y_{\beta}\geq 0 for every \beta\geq\frac{n-2}{n-1} . This fact will be heavily used in the proof of the forthcoming results. It will also be useful to have a precise estimate of \int\limits_{\{\vert \nabla \varphi \vert_{g} = \delta\}}\big\vert\, \nabla\vert \nabla \varphi \vert_{g}\, \, \big\vert_{g}\, d\sigma_{g} in terms of a suitable power of \delta , close to \mathrm{Crit}(\varphi) , that is when \delta\to 0^{+} . This is the content of the following lemma.

    Lemma 4.2. There exists \delta_{0} > 0 such that

    \begin{align} \sup\Bigg\{\delta^{-\frac{1}{n-1}}\int\limits_{\{\vert \nabla \varphi \vert_{g} = \delta\}}\big\vert\,\nabla\vert \nabla \varphi \vert_{g}\,\,\big\vert_{g}\,d\sigma_{g}\,:\, 0 < \delta < \delta_{0}\,\,\mathit{{regular \;value \;of}}\,\,\vert \nabla \varphi \vert_{g}\,\Bigg\} < +\infty\,. \end{align} (4.15)

    We recall that the set of the critical values of \vert \nabla \varphi \vert^{2}_{g} has zero Lebesgue measure by Sard's Theorem, whereas we have no information regarding the local \mathcal{H} –dimension of \mathrm{Crit}(\vert \nabla \varphi \vert^{2}_{g}) .

    Proof. Applying Sard's Theorem to the smooth function \vert \nabla \varphi \vert^{2}_{g} there exists \varepsilon_{0} > 0 such that \varepsilon_{0} is a regular value of \vert \nabla \varphi \vert^{2}_{g} and

    \varepsilon_{0} < \min\Big\{\min\limits_{\{\varphi = 0\}}\vert \nabla \varphi \vert^{2}_{g},{\rm{ the \;limit\; of}}\; \vert \nabla \varphi \vert^{2}_{g} \;{\rm{at}} \; \infty \Big\}\,,

    where the limit in the previous expression is the (finite and positive) value computed in (4.6). In particular, \{\vert \nabla \varphi \vert^{2}_{g}\leq \varepsilon_{0}\} is compactly contained in \mathring{M} , and for every 0 < \delta < \delta_{0} regular value of \vert \nabla \varphi \vert_{g} we have that

    \begin{align*} \delta^{-\frac{1}{n-1}}\int\limits_{\{\vert \nabla \varphi \vert_{g} = \delta\}}\big\vert\,\nabla\vert \nabla \varphi \vert_{g}\,\,\big\vert_{g}\,d\sigma_{g}& = \int\limits_{\{\vert \nabla \varphi \vert_{g} = \delta\}}\sinh(\varphi)\, \frac{\vert \nabla \varphi \vert_{g}^{-\frac{1}{n-1}}\big\vert\,\nabla\vert \nabla \varphi \vert_{g}\,\,\big\vert_{g}}{\sinh(\varphi)}\,d\sigma_{g}\\ &\leq c\,\int\limits_{\{\vert \nabla \varphi \vert_{g} = \delta\}}\frac{\vert \nabla \varphi \vert_{g}^{-\frac{1}{n-1}}\big\vert\,\nabla\vert \nabla \varphi \vert_{g}\,\,\big\vert_{g}}{\sinh(\varphi)}\,d\sigma_{g}\,. \end{align*}

    Now, consider the smooth vector field

    \begin{align*} Z: = 2\,\frac{n-1}{n-2} \,Y_{\frac{n-2}{n-1}} = \frac{1}{\sinh \varphi}\,\, \frac{\nabla\vert \nabla \varphi \vert^{2}_{g}}{\quad\vert \nabla \varphi \vert^{\frac{n}{n-1}}_{g}}\,, \end{align*}

    with \mathrm{div}_{g}Z\geq0 . Set

    \begin{align} U_{\mu}: = \{\vert \nabla \varphi \vert^{2}_{g} < \mu\}\quad {\rm{for \;every}}\; \mu > 0 \,. \end{align} (4.16)

    Then, for every 0 < \varepsilon < \varepsilon_{0} regular value of the function \vert \nabla \varphi \vert^{2}_{g} , we apply the Divergence Theorem to the smooth vector field Z on U_{\varepsilon_{0}} \setminus \overline{U_{\varepsilon}} , and we get

    \begin{align*} \int\limits_{\{\vert \nabla \varphi \vert^{2}_{g} = \varepsilon_{0}\}}\frac{1}{\sinh \varphi}\,\,\frac{\quad\big\vert\,\nabla\vert \nabla \varphi \vert^{2}_{g}\,\,\big\vert_{g}}{\quad\vert \nabla \varphi \vert^{\frac{n}{n-1}}_{g}}\,d\sigma_{g}&-\int\limits_{\{\vert \nabla \varphi \vert^{2}_{g} = \varepsilon\}}\frac{1}{\sinh \varphi}\,\,\frac{\quad\big\vert\,\nabla\vert \nabla \varphi \vert^{2}_{g}\,\,\big\vert_{g}}{\quad\vert \nabla \varphi \vert^{\frac{n}{n-1}}_{g}}\,d\sigma_{g}\\ & = \int\limits_{U_{\varepsilon_{0}} \setminus\overline{U_{\varepsilon}}}\mathrm{div}_{g}Z\,d\mu_{g}\geq 0\,. \end{align*}

    Then, it follows

    \begin{align*} \int\limits_{\{\vert \nabla \varphi \vert^{2}_{g} = \varepsilon_{0}\}}\frac{1}{\sinh \varphi}\,\,\frac{\quad\big\vert\,\nabla\vert \nabla \varphi \vert^{2}_{g}\,\,\big\vert_{g}}{\quad\vert \nabla \varphi \vert^{\frac{n}{n-1}}_{g}}\,d\sigma_{g}\geq\int\limits_{\{\vert \nabla \varphi \vert^{2}_{g} = \varepsilon\}}\frac{1}{\sinh \varphi}\,\,\frac{\quad\big\vert\,\nabla\vert \nabla \varphi \vert^{2}_{g}\,\,\big\vert_{g}}{\quad\vert \nabla \varphi \vert^{\frac{n}{n-1}}_{g}}\,d\sigma_{g}\,. \end{align*}

    Therefore, setting

    c_{1}: = \int\limits_{\{\vert \nabla \varphi \vert^{2}_{g} = \varepsilon_{0}\}}\frac{1}{\sinh \varphi}\,\,\frac{\quad\big\vert\,\nabla\vert \nabla \varphi \vert^{2}_{g}\,\,\big\vert_{g}}{\quad\vert \nabla \varphi \vert^{\frac{n}{n-1}}_{g}}\,d\sigma_{g} > 0\,,

    we obtain

    \begin{align*} \frac{1}{\varepsilon^{\frac{1}{2}\frac{n}{n-1}}}\int\limits_{\{\vert \nabla \varphi \vert^{2}_{g} = \varepsilon\}}\frac{\quad\big\vert\,\nabla\vert \nabla \varphi \vert^{2}_{g}\,\,\big\vert_{g}}{\sinh \varphi}\,d\sigma_{g}\leq c_{1}\,. \end{align*}

    Consequently, the desired statement follows keeping in mind that: if \delta is a regular value of \vert \nabla \varphi \vert_{g} , then \delta^{2} is a regular value of \vert \nabla \varphi \vert_{g}^{2} ; in M\setminus\mathrm{Crit}(\varphi) we have \nabla\vert \nabla \varphi \vert^{2}_{g} = 2\vert \nabla \varphi \vert_{g}\nabla\vert \nabla \varphi \vert_{g} .

    We underline that from now on we will use Remark 3.1 widely.

    The following proposition contains the integral identity which is the main tool of our analysis.

    Proposition 4.3. Let (M, g_{0}, u) be a sub-static harmonic triple, and let g and \varphi be the metric and the function defined in (3.4). Then, for every \beta > \frac{n-2}{n-1} and for every S > s > 0 regular values of \varphi , it holds

    \begin{array}{l} \int\limits_{\{\varphi=s\}} \frac{|\nabla \varphi|_{g}^{\beta} \mathrm{H}_{g}}{\sinh \varphi} d \sigma_{g}-\int\limits_{|\varphi=S|} \frac{|\nabla \varphi|_{g}^{\beta} \mathrm{H}_{g}}{\sinh \varphi} d \sigma_{g} \\\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; =\int\limits_{\mid s < \varphi < S)} \frac{|\nabla \varphi|_{g}^{\beta-2}\left[\left.\left.(\beta-2)|\nabla| \nabla \varphi\right|_{g}\right|_{g} ^{2}+\left|\nabla^{2} \varphi\right|_{g}^{2}+Q(\nabla \varphi, \nabla \varphi)\right]}{\sinh \varphi} d \mu_{g}, \end{array} (4.17)

    where the tensor Q is defined as in (4.9).

    Proof. The case \beta\geq 2 is an easy adaptation of the argument used in [2] but it is anyway a consequence of the following argument. We focus on the unknown case \frac{n-2}{n-1} < \beta < 2 . In \mathring{M}\setminus\mathrm{Crit}(\varphi) we consider the smooth vector field Y_{\beta} , defined in (4.12) and satisfying

    \begin{align} 0\leq\mathrm{div}_{g}\,Y_{\beta}& = \frac{\beta\,\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,(\beta-2)\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}+\vert \nabla^{2} \varphi \vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\,\Big]}{\sinh \varphi}\,, \end{align} (4.18)

    as already explained. Set

    \begin{align*} E_{s}^{S}: = \{s < \varphi < S\},\quad {\rm{for \;every}}\; S > s > 0 \,.\, \end{align*}

    When E_{s}^{S}\cap \mathrm{Crit}(\varphi) = \emptyset , then the statement is a straightforward application of the Divergence Theorem. Now, suppose that E_{s}^{S}\cap \mathrm{Crit}(\varphi)\neq\emptyset . In this case we consider, for every \varepsilon > 0 sufficiently small, a smooth nondecreasing cut–off function \chi_{\varepsilon}:[0, +\infty)\to [0, 1] satisfying the following conditions

    \begin{align*} \chi_{\varepsilon}(\tau) = 0\,{{\rm{in}}}\; \Big[\,0,\frac{1}{2} \varepsilon \,\Big] \,,\quad 0\leq\chi_{\varepsilon}'(\tau)\leq c\varepsilon^{-1}\,\rm{in \Big[\,\frac{1}{2} \varepsilon ,\frac{3}{2} \varepsilon \,\Big] }\,,\quad \chi_{\varepsilon}(\tau) = 1\,{{\rm{in}}} \; \Big[\,\frac{3}{2} \varepsilon ,+\infty\,\Big) \,,\, \end{align*}

    where c is a positive real constant independent of \varepsilon . We then define the smooth function \Xi_{\varepsilon}:M\to [0, 1] as

    \begin{align*} \Xi_{\varepsilon} = \chi_{\varepsilon}\circ \vert \nabla \varphi \vert^{2}_{g}\,, \end{align*}

    and apply the Divergence Theorem to the smooth vector field \Xi_{\varepsilon}\, Y_{\beta} in E_{s}^{S} . In this way, we get

    \begin{align*} \int\limits_{\{\varphi = s\}} \frac{\vert \nabla \varphi \vert^{\beta}_{g}\,\,\mathrm{H}_{g}}{\sinh\varphi}\,d\sigma_{g}&-\int\limits_{\{\varphi = S\}} \frac{\vert \nabla \varphi \vert^{\beta}_{g}\,\,\mathrm{H}_{g}}{\sinh\varphi}\,d\sigma_{g} = \beta^{-1}\Bigg[\int\limits_{E_{s}^{S}}\Xi_{\varepsilon}\,\mathrm{div}_{g}\,Y_{\beta}\,d\mu_{g}+\int\limits_{E_{s}^{S}}g(\nabla\Xi_{\varepsilon},Y_{\beta})\,d\mu_{g}\Bigg]\\ & = \int\limits_{E_{s}^{S}}\frac{\,\Xi_{\varepsilon}\,\,\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,(\beta-2)\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}+\vert \nabla^{2} \varphi \vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\,\Big]}{\sinh \varphi}\,d\mu_{g}\\ &+\int\limits_{ \big(U_{\frac{3}{2}\varepsilon} \setminus\overline{U_{\frac{1}{2}\varepsilon}}\big)\cap \, E_{s}^{S}}\,\frac{\chi'_{\varepsilon}\big(\,\vert \nabla \varphi \vert^{2}_{g}\,\big)\,\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big\vert\nabla\vert \nabla \varphi \vert^{2}_{g}\,\Big\vert_{g}^{2}}{2\sinh \varphi}\,d\mu_{g}\,, \end{align*}

    where U_{\mu} is defined in (4.16). Note that \{\chi_{\varepsilon}\} can always be chosen to be nondecreasing in \varepsilon so that, in turn, \{\Xi_{\varepsilon}\} is nondecreasing. Therefore, applying the Monotone Convergence Theorem, when \varepsilon\to 0^+ , the first term on the right of the second equality tends to

    \begin{align*} \int\limits_{E_{s}^{S}}\frac{\,\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,(\beta-2)\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}+\vert \nabla^{2} \varphi \vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\,\Big]}{\sinh \varphi}\,d\mu_{g}\,. \end{align*}

    For obtaining the desired statement, we show

    \begin{align} \lim\limits_{\varepsilon\to 0^{+}}\int\limits_{ \big(U_{\frac{3}{2}\varepsilon} \setminus\overline{U_{\frac{1}{2}\varepsilon}}\big)\cap \, E_{s}^{S}}\,\frac{\chi'_{\varepsilon}\big(\,\vert \nabla \varphi \vert^{2}_{g}\,\big)\,\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big\vert\nabla\vert \nabla \varphi \vert^{2}_{g}\,\Big\vert_{g}^{2}}{2\sinh \varphi}\,d\mu_{g} = 0\,. \end{align} (4.19)

    First we observe that

    \begin{align*} \int\limits_{ \big(U_{\frac{3}{2}\varepsilon} \setminus\overline{U_{\frac{1}{2}\varepsilon}}\big)\cap \, E_{s}^{S}}\,\frac{\chi'_{\varepsilon}\big(\,\vert \nabla \varphi \vert^{2}_{g}\,\big)\,\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big\vert\nabla\vert \nabla \varphi \vert^{2}_{g}\,\Big\vert_{g}^{2}}{2\sinh \varphi}\,d\mu_{g} &\leq\,\,\int\limits_{ U_{\frac{3}{2}\varepsilon} \setminus\overline{U_{\frac{1}{2}\varepsilon}}}\,\frac{\chi'_{\varepsilon}\big(\,\vert \nabla \varphi \vert^{2}_{g}\,\big)\,\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big\vert\nabla\vert \nabla \varphi \vert^{2}_{g}\,\Big\vert_{g}^{2}}{2\sinh \varphi}\,d\mu_{g}\\ &\\ &\leq\frac{c}{2\varepsilon}\int\limits_{\frac{1}{2}\varepsilon}^{\frac{3}{2}\varepsilon}\,s^{\frac{\beta-2}{2}}ds\int\limits_{\{\vert \nabla \varphi \vert^{2}_{g} = s\}}\frac{\big\vert\,\nabla\vert \nabla \varphi \vert^{2}_{g}\,\,\big\vert_{g}}{\sinh \varphi}\,d\sigma_{g} \end{align*}

    where, keeping in mind the properties satisfied by \chi_{\varepsilon} , in the first inequality we have used the nonnegativity of the integrand function and in the last one the Coarea Formula. Note that there exist \varepsilon_{0}, \, c_{1} > 0 such that the inequality

    \begin{align*} \frac{1}{s^{\frac{1}{2}\frac{n}{n-1}}}\int\limits_{\{\vert \nabla \varphi \vert^{2}_{g} = s\}}\frac{\quad\big\vert\,\nabla\vert \nabla \varphi \vert^{2}_{g}\,\,\big\vert_{g}}{\sinh \varphi}\,d\sigma_{g}\leq c_{1} \end{align*}

    is true a.e. s\in[\frac{1}{2}\varepsilon, \frac{3}{2}\varepsilon] for every 0 < \varepsilon < \frac{2}{3}\varepsilon_{0} , by both Sard's Theorem applied to the smooth function \vert \nabla \varphi \vert^{2}_{g}\, and by Lemma 4.2. Then, we get

    \begin{align*} \int\limits_{ \big(U_{\frac{3}{2}\varepsilon} \setminus\overline{U_{\frac{1}{2}\varepsilon}}\big)\cap \, E_{s}^{S}}\,\frac{\chi'_{\varepsilon}\big(\,\vert \nabla \varphi \vert^{2}_{g}\,\big)\,\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big\vert\nabla\vert \nabla \varphi \vert^{2}_{g}\,\Big\vert_{g}^{2}}{2\sinh \varphi}\,d\mu_{g} &\leq\frac{c}{2\varepsilon}\int\limits_{\frac{1}{2}\varepsilon}^{\frac{3}{2}\varepsilon}\,s^{\frac{\beta-2}{2}}ds\int\limits_{\{\vert \nabla \varphi \vert^{2}_{g} = s\}}\frac{\big\vert\,\nabla\vert \nabla \varphi \vert^{2}_{g}\,\,\big\vert_{g}}{\sinh \varphi}\,d\sigma_{g}\\ &\leq\frac{c\,c_{1}}{2\varepsilon}\,\int\limits_{\frac{1}{2}\varepsilon}^{\frac{3}{2}\varepsilon}\,s^{\frac{\beta-2}{2}+\frac{1}{2}\,\frac{n}{n-1}}\,ds\\ & = \frac{c\,c_{1}}{2\varepsilon}\,\int\limits_{\frac{1}{2}\varepsilon}^{\frac{3}{2}\varepsilon}\,s^{\frac{1}{2}(\beta-\frac{n-2}{n-1})}\,ds\\ &\leq c_{2}\,\varepsilon^{\frac{1}{2}(\beta-\frac{n-2}{n-1})}\,,\, \end{align*}

    where c_{2} > 0 is sufficiently big constant. This implies the limit in (4.19), because \beta > \frac{n-2}{n-1} .

    Corollary 4.4. For every S > s > 0 regular values of \varphi there exists r_{s, S}\geq0 such that

    \begin{align*} \int\limits_{\{\varphi = s\}} \frac{\vert \nabla \varphi \vert^{\frac{n-2}{n-1}}_{g}\,\,\mathrm{H}_{g}}{\sinh\varphi}\,d\sigma_{g}&-\int\limits_{\{\varphi = S\}} \frac{\vert \nabla \varphi \vert^{\frac{n-2}{n-1}}_{g}\,\,\mathrm{H}_{g}}{\sinh\varphi}\,d\sigma_{g}\\ & = \int\limits_{\{s < \varphi < S\}}\frac{\vert \nabla \varphi \vert^{-\frac{n}{n-1}}_{g}\,\Big[\,\vert \nabla^{2} \varphi \vert_{g}^{2}-\frac{n}{n-1}\,\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\, \Big]}{\sinh\varphi}\,d\mu_{g}+r_{s,S}\,. \end{align*}

    Proof. Let \{\beta_{m}\}_{m\in \mathbb{N}} be a sequence such that \beta_{m} > \frac{n-2}{n-1} and \beta_{m}\to\frac{n-2}{n-1} . Due to Proposition 4.3, we have

    \begin{align*} \int\limits_{\{\varphi = s\}} \frac{\vert \nabla \varphi \vert^{\frac{n-2}{n-1}}_{g}\,\,\mathrm{H}_{g}}{\sinh\varphi}\,d\sigma_{g}&-\int\limits_{\{\varphi = S\}} \frac{\vert \nabla \varphi \vert^{\frac{n-2}{n-1}}_{g}\,\,\mathrm{H}_{g}}{\sinh\varphi}\,d\sigma_{g}\\ & = \lim\limits_{m\to+\infty} \Bigg[\int\limits_{\{\varphi = s\}} \frac{\vert \nabla \varphi \vert^{\beta_{m}}_{g}\,\,\mathrm{H}_{g}}{\sinh\varphi}\,d\sigma_{g}-\int\limits_{\{\varphi = S\}} \frac{\vert \nabla \varphi \vert^{\beta_{m}}_{g}\,\,\mathrm{H}_{g}}{\sinh\varphi}\,d\sigma_{g}\Bigg]\\ & = \lim\limits_{m\to+\infty}\int\limits_{\{s < \varphi < S\}}\frac{\vert \nabla \varphi \vert^{\beta_{m}-2}_{g}\,\Big[\,(\beta_{m}-2)\,\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}+\vert \nabla^{2} \varphi \vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\, \Big]}{\sinh\varphi}\,d\mu_{g}\\ &\geq \int\limits_{\{s < \varphi < S\}}\frac{\vert \nabla \varphi \vert^{-\frac{n}{n-1}}_{g}\,\Big[\,\vert \nabla^{2} \varphi \vert_{g}^{2}-\frac{n}{n-1}\,\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\, \Big]}{\sinh\varphi}\,d\mu_{g}\,, \end{align*}

    where the first equality is consequence of the Dominate Converge Theorem keeping in mind that s and S are regular values of \varphi while the inequality follows from Fatou's Lemma. Since \{\beta_{m}\}_{m\in N} is arbitrary, the quantity

    \begin{align*} r_{s,S}: = &\lim\limits_{\beta\to\frac{n-2}{n-1}^+} \Bigg\{\,\int\limits_{\{s < \varphi < S\}}\frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,(\beta-2)\,\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}+\vert \nabla^{2} \varphi \vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\, \Big]}{\sinh\varphi}\,d\mu_{g}\Bigg\}\\ &\quad- \int\limits_{\{s < \varphi < S\}}\frac{\vert \nabla \varphi \vert^{-\frac{n}{n-1}}_{g}\,\Big[\,\vert \nabla^{2} \varphi \vert_{g}^{2}-\frac{n}{n-1}\,\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\, \Big]}{\sinh\varphi}\,d\mu_{g} \end{align*}

    is well–defined. Moreover, it is nonnegative as above and therefore we get the statement.

    Remark 4.2. For every \beta > \frac{n-2}{n-1} and for every s > 0 regular value of the function \varphi :

    \begin{align} \int\limits_{\{\varphi = s\}} \frac{\vert \nabla \varphi \vert^{\beta}_{g}\,\,\mathrm{H}_{g}}{\sinh\varphi}\,d\sigma_{g} = \int\limits_{\{\varphi > s\}}\frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,(\beta-2)\,\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}+\vert \nabla^{2} \varphi \vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\, \Big]}{\sinh\varphi}\,d\mu_{g}\,.\, \end{align} (4.20)

    For every S big enough, which is a regular value of \varphi , by Lemma 4.5 with (4.6) we have

    \Bigg\vert \,\int\limits_{\{\varphi = S\}}\vert \nabla \varphi \vert^{\beta}_{g}\,\,\mathrm{H}_{g}\,d\sigma_{g}\,\Bigg\vert\leq \int\limits_{\{\varphi = S\}}\vert \nabla \varphi \vert^{\beta-1}_{g}\,\vert \nabla^{2} \varphi \vert_{g}\,d\sigma_{g}\,\leq \widetilde{c}\,.

    In particular,

    \begin{align*} \lim\limits_{S\to +\infty}\,\frac{1}{\sinh(S)}\,\int\limits_{\{\varphi = S\}}\vert \nabla \varphi \vert^{\beta}_{g}\,\,\mathrm{H}_{g}\,d\sigma_{g} = \,0\,. \end{align*}

    Therefore, the desired identity can be obtained by the Monotone Convergence Theorem, by passing to the limit as S\to+\infty in (4.17).

    Remark 4.3. For every \beta > \frac{n-2}{n-1} , as consequence of integral identity (4.17), we have

    \begin{align} \vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,(\beta-2)\,\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}+\vert \nabla^{2} \varphi \vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\, \Big]\,\in L^{1}_{\rm{loc}}\big(\mathring{M},\mu_{g}\big)\,. \end{align} (4.21)

    Since

    \begin{align*} \int\limits_{K}\vert \nabla \varphi \vert^{\beta-3}_{g}\,\,\vert \nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)\vert\,d\mu_{g}\leq \int\limits_{K}\vert \nabla \varphi \vert^{\beta-1}_{g}\,\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g} \,d\mu_{g} = \int\limits_{K}\vert \nabla \varphi \vert^{\frac{\beta}{2}}_{g}\,\vert \nabla \varphi \vert^{\frac{\beta-2}{2}}_{g}\,\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g} \,d\mu_{g} \end{align*}

    for every K\subset \mathring{M} compact, by Hölder's Inequality from (4.21) with (4.14) we get that

    \begin{align} \vert \nabla \varphi \vert^{\beta-3}_{g}\,\,\nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)\,\in L^{1}_{\rm{loc}}\big(\mathring{M},\mu_{g}\big)\,. \end{align} (4.22)

    We need a final lemma before stating the (last and) most important result of this section.

    Lemma 4.5. Let (M, g_{0}, u) be a sub-static harmonic triple, and let g and \varphi be the metric and the function defined in (3.4). Then, the following statements hold true.

    (i) For every \beta\geq0 and for every S > s > 0 :

    \begin{align} \int\limits_{\{\varphi = S\}}\frac{\vert \nabla \varphi \vert^{\beta+1}_{g}}{\sinh\varphi}\,d\sigma_{g}&-\int\limits_{\{\varphi = s\}}\frac{\vert \nabla \varphi \vert^{\beta+1}_{g}}{\sinh\varphi}\,d\sigma_{g} = \, \\ & = \int\limits_{\{s < \varphi < S\}}\frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,\beta\,\nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)\,-\coth(\varphi)\vert \nabla \varphi \vert^{4}_{g} \Big]}{\sinh\varphi}\,d\mu_{g}\,. \end{align} (4.23)

    (ii) For every \beta\geq0 and for every s > 0 :

    \begin{align*} \int\limits_{\{\varphi = s\}}\frac{\vert \nabla \varphi \vert^{\beta+1}_{g}}{\sinh\varphi}\,d\sigma_{g} = \int\limits_{\{\varphi > s\}}\frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,\coth(\varphi)\vert \nabla \varphi \vert^{4}_{g}-\beta\,\nabla ^{2}\varphi(\nabla \varphi, \nabla \varphi)\, \Big]}{\sinh\varphi}\,d\mu_{g}\,. \end{align*}

    (iii) The function \Phi_{\beta}:[0, \infty)\to \mathbb{R} , defined by formula

    \begin{align} \Phi_{\beta}(s): = \int\limits_{\{\varphi = s\}}\vert \nabla \varphi \vert^{\beta+1}_{g}\,d\sigma_{g} \end{align} (4.24)

    for every \beta\geq0 , is continuous and admits for every s > 0 the integral representation

    \begin{align*} \Phi_{\beta}(s): = \sinh (s)\,\int\limits_{\{\varphi > s\}}\frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,\coth(\varphi)\vert \nabla \varphi \vert^{4}_{g}-\beta\,\nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)\, \Big]}{\sinh\varphi}\,d\mu_{g}\,. \end{align*}

    This lemma can be proved as [2,Proposition 4.1]. In the Appendix we provide an alternative proof which is self contained and does not make use of any fine property of the measure of \mathrm{Crit}(\varphi) : we just need to know very classical properties of it (see Remark 3.1).

    Proposition 4.6. Let (M, g_{0}, u) be a sub-static harmonic triple, let g and \varphi be the metric and the function defined in (3.4), and let \Phi_{\beta}:[0, \infty)\to \mathbb{R} be the function defined by formula (4.24) for every \beta\geq0 . Then for every \beta > \frac{n-2}{n-1} , the function \Phi_{\beta} is continuously differentiable. The derivative \Phi'_{\beta} is nonpositive and admits for every s > 0 the integral representation

    \begin{equation} \Phi'_{\beta}(s) = -\beta\, \sinh( s) \,\int\limits_{\{\varphi > s\}}\frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,(\beta-2)\,\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}+\vert \nabla^{2} \varphi \vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\, \Big]}{\sinh\varphi}\,d\mu_{g}\leq 0\,. \end{equation} (4.25)

    Moreover, if there exists s_{0} > 0 such that \Phi'_{\beta}(s_{0}) = 0 for some \beta > \frac{n-2}{n-1} , then (\{\varphi\geq s_{0}\}, g) is isometric to \big([0, +\infty)\times \{\varphi = s_{0}\}, d\rho\otimes d\rho+g_{\{\varphi = s_{0}\}}) , where \rho is the g –distance function to \{\varphi = s_{0}\} and \varphi is an affine function of \rho in \{\varphi\geq s_{0}\} .

    The following proof is essentially the same as in [2]. For completeness, we include it here, in a slightly refined version.

    Proof. Step 1 : Continuous Differentiability and Monotonicity. Let \beta > \frac{n-2}{n-1} . Note that the boundary \partial M is a regular level set of \varphi and then, by Theorem 2.3 and the relationship between \mathrm{Crit}(u) and \mathrm{Crit}(\varphi) , there exists \epsilon_{0} such that the interval [0, \epsilon_{0}] doesn't contain critical values of the function \varphi . Therefore, for every 0 < \epsilon\leq\epsilon_{0} , applying first the Divergence Theorem to the smooth vector field \vert \nabla \varphi \vert^{\beta}_{g}\, \nabla \varphi in \{0 < \varphi < \epsilon\} and later the Coarea Formula, we get

    \Phi_{\beta}(\epsilon) = \Phi_{\beta}(0)-\beta\,\int\limits_{0}^{\epsilon}\,ds\,\int\limits_{\{\varphi = s\}}\vert \nabla \varphi \vert^{\beta}_{g}\,\mathrm{H}_{g}\,d\sigma_{g}\,.

    Being

    \int\limits_{\{\varphi = s_{1}\}}\vert \nabla \varphi \vert^{\beta}_{g}\,\mathrm{H}_{g}\,d\sigma_{g}-\int\limits_{\{\varphi = s_{2}\}}\vert \nabla \varphi \vert^{\beta}_{g}\,\mathrm{H}_{g}\,d\sigma_{g} = \beta^{-1}\int\limits_{\{s_{1} < \varphi < s_{2}\}}\mathrm{div}_{g}\big(\nabla \vert \nabla \varphi \vert^{\beta}_{g}\big)\,d\mu_{g}

    for every 0\leq s_{1} < s_{2}\leq \epsilon_{0} , by the Dominated Convergence Theorem the function

    s\in[0,\epsilon_{0}]\to\int\limits_{\{\varphi = s\}}\vert \nabla \varphi \vert^{\beta}_{g}\,\mathrm{H}_{g}\,d\sigma_{g}\in \mathbb{R}

    is continuous and therefore, by the Fundamental Theorem of Calculus \Phi_{\beta} is continuously differentiable on the closed interval [0, \epsilon_{0}] .

    Let s_{0} be a regular value of the function \varphi . By Remark 4.3, we can define the function \Psi_{\beta}:(0, +\infty)\to \mathbb{R} by

    \begin{align*} \Psi_{\beta}(s) = \begin{cases}\int\limits_{\{\varphi = s_{0}\}} \frac{\vert \nabla \varphi \vert^{\beta}_{g}\,\,\mathrm{H}_{g}}{\sinh\varphi}\,d\sigma_{g}+ \int\limits_{\{s < \varphi < s_{0}\}}\frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,(\beta-2)\,\big\vert\nabla\vert \nabla \varphi \vert_{g}\,\big\vert_{g}^{2}+\vert \nabla^{2} \varphi \vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\, \Big]}{\sinh\varphi}\,d\mu_{g} &{\rm{if}}\,s\leq s_{0}\\ \\ \int\limits_{\{\varphi = s_{0}\}} \frac{\vert \nabla \varphi \vert^{\beta}_{g}\,\,\mathrm{H}_{g}}{\sinh\varphi}\,d\sigma_{g}- \int\limits_{\{s_{0} < \varphi < s\}}\frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,(\beta-2)\,\big\vert\nabla\vert \nabla \varphi \vert_{g}\,\big\vert_{g}^{2}+\vert \nabla^{2} \varphi \vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\, \Big]}{\sinh\varphi}\,d\mu_{g} &{\rm{if}}\,s > s_{0} \,, \end{cases} \end{align*}

    which satisfies the following properties

    (i) for every s > 0 regular value of the function \varphi , we have \Psi_{\beta}(s) = \int\limits_{\{\varphi = s\}} \frac{\vert \nabla \varphi \vert^{\beta}_{g}\, \, \mathrm{H}_{g}}{\sinh\varphi}\, d\sigma_{g} ;

    (ii) the function \Psi_{\beta} is continuous on its definition interval (0, +\infty) .

    The first statement follows immediately from Proposition 4.3. As for the second statement, we first observe that

    \begin{align} \Psi_{\beta}(s)-\Psi_{\beta}(\overline{s}) = \int\limits_{\{s < \varphi < \overline{s}\}}\frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,(\beta-2)\,\big\vert\nabla\vert \nabla \varphi \vert_{g}\,\big\vert_{g}^{2}+\vert \nabla^{2} \varphi \vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\, \Big]}{\sinh\varphi}\,d\mu_{g} \end{align} (4.26)

    for every couple 0 < s < \overline{s} < +\infty . Always by Remark 4.3 and by the Dominated Convergence Theorem, we can deduce the right and the left continuity of \Psi_{\beta} on the interval (0, +\infty) .

    We consider \Upsilon_{\beta}:s\in(0, +\infty)\to\frac{\Phi_{\beta}(s)}{\sinh s}\in \mathbb{R}\, . For every (s, \overline{s}) couple of real number such that 0 < s < \overline{s} < +\infty , we have

    \begin{align*} \frac{\Upsilon_{\beta}(\overline{s})-\Upsilon_{\beta}(s)}{\overline{s}-s}& = \frac{1}{\overline{s}-s}\int\limits_{\{s < \varphi < \overline{s}\}}\frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,\beta\,\nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)\,-\coth(\varphi)\vert \nabla \varphi \vert^{4}_{g} \Big]}{\sinh\varphi}\,d\mu_{g}\\ & = \frac{1}{\overline{s}-s}\int\limits_{s}^{\overline{s}}d\tau\int\limits_{\{\varphi = \tau\}}\frac{\vert \nabla \varphi \vert^{\beta-3}_{g}\,\Big[\,\beta\,\nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)\,-\coth(\varphi)\vert \nabla \varphi \vert^{4}_{g} \Big]}{\sinh\varphi}\,d\sigma_{g}\\ &\overset{(\star)}{ = } -\frac{\beta}{\overline{s}-s}\,\int\limits_{s}^{\overline{s}}\Psi_{\beta}(\tau)\,d\tau-\frac{1}{\overline{s}-s}\,\int\limits_{s}^{\overline{s}}\coth(\tau)\Upsilon_{\beta}(\tau)\, d\tau\,, \end{align*}

    where the first equality follows from Lemma 4.5 (i) , the second equality from the Coarea Formula keeping in mind (4.22). Moreover, the last equality follows from (i) and from Sard's Theorem. Using the continuity of both the functions \Upsilon_{\beta} and \Psi_{\beta} , passing to the limit in (\star) for either s\to \overline{s} or \overline{s}\to s yields that the function \Upsilon_{\beta} is C^{1} , and

    \Upsilon_{\beta}'(\,\cdot\, ) = -\beta\, \Psi_{\beta}(\,\cdot\,)-\coth(\,\cdot\,)\,\Upsilon_{\beta}(\,\cdot\,)\,.

    Since \Phi_{\beta}(s) = \sinh(s)\Upsilon_{\beta}(s) for every s > 0 , then \Phi_{\beta}\in C^{1}(0, +\infty) and \Phi'_{\beta}(s) = -\beta\, \sinh(s)\Psi_{\beta}(s) . Moreover, by (4.26), we can see

    \begin{align*} \frac{\Phi_{\beta}'(S)}{\sinh(S)}-\frac{\Phi_{\beta}'(s)}{\sinh(s)}& = -\beta\,\Psi_{\beta}(S)+\beta\,\Psi_{\beta}(s)\\ & = \beta\,\int\limits_{\{s < \varphi < S\}}\frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,(\beta-2)\,\big\vert\nabla\vert \nabla \varphi \vert_{g}\,\big\vert_{g}^{2}+\vert \nabla^{2} \varphi \vert_{g}^{2}+Q(\nabla \varphi ,\nabla \varphi )\, \Big]}{\sinh\varphi}\,d\mu_{g} \end{align*}

    for every 0 < s < S < +\infty .

    Finally the integral representation (4.25) follows in the limit as S\to +\infty of the above identity, by using the Monotone Convergence Theorem, and by the fact that

    \lim\limits_{S\to+\infty}\,\frac{\Phi_{\beta}'(S)}{\sinh(S)}\, = -\beta \lim\limits_{S\to+\infty}\,\Psi_{\beta}(S) = -\beta \lim\limits_{S\to+\infty}\,\int\limits_{\{\varphi = S\}} \frac{\vert \nabla \varphi \vert^{\beta}_{g}\,\,\mathrm{H}_{g}}{\sinh\varphi}\,d\sigma_{g} = 0\,.

    Step 2 : Outer Rigidity. Let \beta > \frac{n-2}{n-1} and suppose \Phi'_{\beta}(s_{0}) = 0 for some s_{0} > 0 . By (4.25) with (4.14) we deduce that

    \begin{equation*} \Big(\beta-\frac{n-2}{n-1}\Big)\,\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}\equiv0 \quad{\rm{and}} \quad \vert \nabla^{2} \varphi \vert_{g}^{2}-\frac{n}{n-1}\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}\equiv0\quad {\rm{in}}\,\{\varphi\geq s_{0}\}\setminus\mathrm{Crit}(\varphi)\,. \end{equation*}

    Consequently \nabla^{2} \varphi\equiv 0 in \{\varphi\geq s_{0}\} being \mu_{g}\big(\mathrm{Crit}(\varphi)\, \big) = 0 , and hence \vert \nabla \varphi \vert_{g}^{2}\equiv a^{2} with a > 0 since \{\varphi\geq s_{0}\} is connected, due to Remark 3.1. Then, \{\varphi\geq s_{0}\} , with the induced Riemanninan metric, is a noncompact, connected and complete Riemannian manifold (being properly embedded in M ), with smooth, compact and totally geodesic boundary, and with \mathrm{Ric}_{g}\geq0 (from the inequality in (3.5)). Applying [18,Theorem C], we can thus deduce that the level set \{\varphi \, = \, s_{0}\} is connected (this is true in general and not only in the rigid case, if s_0\gg0 , as observed in Remark 3.1), and that \{\varphi\geq s_{0}\} is isometric to the product [0, + \infty)\times\{\varphi \, = \, s_{0}\} . Moreover, the isometry from the product [0, + \infty)\times\{\varphi \, = \, s_{0}\} to \{\varphi\geq s_{0}\} is given by the normal exponential map.

    Now we want to prove that \varphi is an affine function of \rho on \{\varphi\geq s_{0}\} . First, we remark that every integral curve \gamma_{p} of \nabla \varphi outgoing from a point p of \{\varphi \, = \, s_{0}\} is defined on the interval [0, +\infty) , and it is contained in \{\varphi\geq s_{0}\} , by the completeness and since \vert\nabla \varphi\vert_{g} > 0 . Furthermore, \varphi \circ \gamma_{p}(t) = a^{2}t+s_{0} for every t\in[0, +\infty) , and all the curves \gamma_{p} realize the distance between the hypersurfaces \{\varphi = s_{0}\} and \{\varphi = s_{1}\} with s_{1} > s_{0} . Indeed, for any curve \sigma:[0, l]\to \{\varphi\geq s_{0}\} parametrized by arc–length joining a point of \{\varphi = s_{0}\} to a point of \{\varphi = s_{1}\} we have

    \begin{align*} L_{g}(\, \sigma\,)& = \int\limits_{0}^{l} \vert \,\dot{\sigma}(\tau)\,\vert_{g}\,\,d\tau\geq \Bigg\vert \int\limits_{0}^{l} g\Big(\,\dot{\sigma}(\tau),\frac{1}{a}\nabla \varphi \,\big(\sigma(t)\,\big)\,\Big)\,\,d\tau \Bigg\vert = \frac{1}{a}\vert \varphi \circ \sigma\,(l)-\varphi \circ \sigma(0)\vert \\ & = at = L_{g}(\, \gamma_{\sigma(0)}\,|_{[0,t]}\,) = L_{g}(\, \gamma_{\cdot}\,|_{[0,t]}\,)\,, \end{align*}

    where s_{1}, s_{0} and t satisfy s_{1} = a^{2}t+s_{0} . Since \xi = \frac{\, \nabla \varphi }{a} is the unit inner normal vector field of the boundary \{\varphi \, = \, s_{0}\} and we just know that the normal exponential map is a diffeomorphism, \exp^{\bot}(t\xi_{p}) is a point having distance from \{\varphi = s_{0}\} equal to t\, , and therefore

    \varphi\big(\exp^{\bot}(t\xi_{p})\,\big) = \varphi \circ \gamma_{p}\Big(\,\frac{t}{a}\,\Big) = at+s_{0} = a\, \rho\big(\exp^{\bot}(t\xi_{p})\big)+s_{0}\,.

    This tell us that \varphi is an affine function of \rho on \{\varphi\geq s_{0}\} .

    While the previous proposition contains an outer rigidity result, with the following corollary we provide a global rigidity result.

    Corollary 4.7. Let (M, g_{0}, u) be a sub-static harmonic triple, let g and \varphi be the metric and the function defined in (3.4), and let \Phi_{\beta}:[0, \infty)\to \mathbb{R} be the function defined by formula (4.24) for every \beta\geq0 . If \Phi_{\beta} is constant for some \beta > \frac{n-2}{n-1} , then \partial M is connected and (M, g) is isometric to \big([0, +\infty)\times \partial M, d\rho\otimes d\rho+g_{\partial M}) , where \rho is the g –distance function to \partial M and \varphi is an affine function of \rho .

    Proof. We obtain immediately that \Phi_{\beta}'(s) = 0 for every s > 0 . Thus, by formula (4.25) with (4.14) we have that

    \int\limits_{\{\varphi > s\}}\Bigg\{\Big(\beta-\frac{n-2}{n-1}\Big)\,\big\vert\nabla\vert \nabla \varphi \vert_{g}\,\big\vert_{g}^{2}+\Bigg[\,\vert \nabla^{2} \varphi \vert_{g}^{2}-\frac{n}{n-1}\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}\,\Bigg]+Q(\nabla \varphi ,\nabla \varphi )\Bigg\}\,d\mu_{g} = 0

    for every s > 0 . In turn, by the Monotone Convergence Theorem, we get

    \int\limits_{M}\Bigg\{\Big(\beta-\frac{n-2}{n-1}\Big)\,\big\vert\nabla\vert \nabla \varphi \vert_{g}\,\big\vert_{g}^{2}+\Bigg[\,\vert \nabla^{2} \varphi \vert_{g}^{2}-\frac{n}{n-1}\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}\,\Bigg]+Q(\nabla \varphi ,\nabla \varphi )\Bigg\}\,d\mu_{g} = 0\,.

    Then, we deduce that

    \Big(\beta-\frac{n-2}{n-1}\Big)\,\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}\equiv0 \quad{\rm{and}} \quad \vert \nabla^{2} \varphi \vert_{g}^{2}-\frac{n}{n-1}\Big\vert\nabla\vert \nabla \varphi \vert_{g}\,\Big\vert_{g}^{2}\equiv0\quad {\rm{in}}\, M\setminus\mathrm{Crit}(\varphi)\,,

    due to Kato Inequality for harmonic functions (4.13) and by (4.10). Consequently \nabla^{2} \varphi\equiv 0 in M . The very same argument of the proof of Outer Rigidity in Proposition 4.6 implies that \partial M is connected and (M, g) is isometric to

    \big([0,+\infty)\times \partial M,d\rho\otimes d\rho+g_{\partial M})

    where \rho is the g –distance to \partial M and \varphi is an affine function of \rho .

    This section is devoted to the proof of the Black–Hole uniqueness result for a sub–static harmonic triple, Theorem 1.2. We first recall the classical definition of ADM mass, together with an alternative characterization of it.

    Let (N, h) and \psi be an asymptotically flat manifold with one end and a chart at infinity of N , respectively. We consider \widetilde{h}: = \psi_{*}h = \widetilde{h}_{ij}dx^{i}\otimes dx^{j} and we set

    \begin{align*} m(r)& = \frac{1}{2(n-1)\vert \mathbb{S}^{n-1}\vert}\int\limits_{\partial B_{r}}(\partial_{j}\widetilde{h}_{ij}-\partial_{i}\widetilde{h}_{jj})\nu_{e}^{i}d\sigma_{e}\,,\\ m_{I}(r)& = -\frac{1}{(n-2)(n-1)\vert \mathbb{S}^{n-1}\vert}\int\limits_{\partial B_{r}}\Big(\mathrm{Ric}_{\widetilde{h}}-\frac{1}{2}\mathrm{R}_{\widetilde{h}}\,\widetilde{h}\Big)(X,\nu_{\widetilde{h}})\,d\sigma_{\widetilde{h}}\,, \end{align*}

    where \nu_{e} and \sigma_{e} are the \infty –pointing unit normal and the canonical measure on \partial B_{r} as Riemannian submanifold of (\mathbb{R}^{n}\setminus \overline{B}, g_{ \mathbb{R}^{n}}) , respectively, and \nu_{\widetilde{h}} and \sigma_{\widetilde{h}} are the \infty –pointing unit normal and the canonical measure on \partial B_{r} as Riemannian submanifold of (\mathbb{R}^{n}\setminus \overline{B}, \widetilde{h}) , respectively. Also, \mathrm{Ric}_{\widetilde{h}} and \mathrm{R}_{\widetilde{h}} are the Ricci tensor and the scalar curvature of \widetilde{h} respectively, and X is the Euclidean conformal Killing vector field x^{i}\, \frac{\partial }{\partial x^{i}} . The ADM mass is well defined as

    \begin{equation*} m_{\mathrm{ADM}}: = \lim\limits_{r\to +\infty}\, m(r)\,, \end{equation*}

    and independent of the chosen chart at infinity. Moreover (see [22]), it can be equivalently expressed as

    \begin{equation} m_{\mathrm{ADM}} = \lim\limits_{r\to +\infty}\, m_{I}(r)\,. \end{equation} (5.1)

    From the alternative definition of ADM mass, given by (5.1), and using the Positive Mass Theorem, more precisely a consequence of it contained in [17,Theorem 1.5], one can prove the following uniqueness statement. For the notation and terminology, we refer the reader to Definition 1.1 and Section 2.

    Proof of Theorem 1.2. By condition (1.8) and by the fact that \mathrm{D}_{g_{0}}^{2}u\equiv 0 on \partial M , which in turn implies \mathrm{H}_{\partial M}^{g_{0}}\equiv 0 , we have that the hypothesis of [17,Theorem 1.5] are fulfilled, so that

    \begin{equation*} m_{\mathrm{ADM}}\geq \mathcal{C}. \end{equation*}

    Now, we want to show that the reverse inequality holds. Let \psi be a chart at infinity of M (according to Definition 1.1) and consider \widetilde{g}_{0} = \psi_{*}{g_{0}} . Recalling that \widetilde{u} stands for u\circ\psi^{-1} , we rewrite characterization (5.1) as

    \begin{align*} m_{\mathrm{ADM}}& = \lim\limits_{r\to +\infty}\Bigg\{-\frac{1}{(n-2)(n-1)\vert \mathbb{S}^{n-1}\vert}\int\limits_{\partial B_{r}}\Bigg(\mathrm{Ric}_{\widetilde{g}_{0}}-\frac{\mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u}}{\widetilde{u}} \,\Bigg)(X,\nu_{e})\,d\sigma_{\widetilde{g}_{0}}\\ &-\frac{1}{(n-2)(n-1)\vert \mathbb{S}^{n-1}\vert}\int\limits_{\partial B_{r}}\Bigg(\mathrm{Ric}_{\widetilde{g}_{0}}-\frac{\mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u}}{\widetilde{u}}\, \Bigg)(X,\nu_{\widetilde{g}_{0}}-\nu_{e})\,d\sigma_{\widetilde{g}_{0}}\\ &- \frac{1}{(n-2)(n-1)\vert \mathbb{S}^{n-1}\vert}\int\limits_{\partial B_{r}} \frac{\mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u}}{\widetilde{u}}(X,\nu_{\widetilde{g}_{0}})\,d\sigma_{\widetilde{g}_{0}}+\\ & \frac{1}{2(n-2)(n-1)\vert \mathbb{S}^{n-1}\vert}\int\limits_{\partial B_{r}}\mathrm{R}_{\widetilde{g}_{0}}\, \widetilde g_0(X,\nu_{\widetilde{g}_{0}})\,d\sigma_{\widetilde{g}_{0}}\Bigg\}\,. \end{align*}

    We note first that since \nu_{e} = \frac{x^{i}}{\vert x\vert}\, \frac{\partial }{\partial x^{i}} = \frac{1}{\vert x\vert}\, X and \widetilde{u}\, \mathrm{Ric}_{\widetilde{g}_{0}}-\mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u} \geq 0 from the first equation in (1.1), we have

    \begin{align} \int\limits_{\partial B_{r}}\Bigg(\mathrm{Ric}_{\widetilde{g}_{0}}-\frac{\mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u}}{\widetilde{u}}\, \Bigg)(X,\nu_{e})\,d\sigma_{\widetilde{g}_{0}} = \frac{1}{r}\int\limits_{\partial B_{r}}\Bigg(\mathrm{Ric}_{\widetilde{g}_{0}}-\frac{\mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u}}{\widetilde{u}}\, \Bigg)(X,X)\,d\sigma_{\widetilde{g}_{0}}\geq0\,. \end{align} (5.2)

    Secondly, recalling that (\mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u})_{ij} = \partial_{i}\partial_{j}\widetilde{u}-\Gamma_{ij}^{k}\partial_{k}\widetilde{u} , where \Gamma_{ij}^{k} are the Christoffel symbols related to \widetilde{g}_{0} , and using (1.3), (2.12), and the asymptotic expansions of \widetilde{u} , we get

    \begin{align} \vert (\mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u})_{ij}-(\mathrm{D}_{e}^{2}\widetilde{u})_{ij}\vert& = \vert \Gamma_{ij}^{k}\partial_{k}\widetilde{u}\vert = O\big(\vert x\vert^{-(n+p)}\big) \end{align} (5.3)
    \begin{align} (\mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u})_{ij}& = O(\vert x\vert^{-n}) \,. \end{align} (5.4)

    Decay (5.4) coupled with (2.8) (2.10) and (2.11) yields

    \begin{align} \Big\vert \int\limits_{\partial B_{r}}\Bigg(\mathrm{Ric}_{\widetilde{g}_{0}}-\frac{\mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u}}{\widetilde{u}}\,\Bigg)(X,\nu_{\widetilde{g}_{0}}-\nu_{e})\,d\sigma_{\widetilde{g}_{0}}\Big\vert \leq C \int\limits_{\partial B_{r}}\frac{1}{\vert x\vert ^{p+\min\{p+2,n\}-1}}\,d\sigma_{e} = \frac{C}{r^{p+\min\{p+2,n\}-n}}\xrightarrow {}0 \,, \end{align} (5.5)

    being p > \frac{n-2}{2} . Thirdly, we observe that

    \begin{align} \int\limits_{\partial B_{r}} \frac{\mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u}}{\widetilde{u}}(X,\nu_{\widetilde{g}_{0}})\,d\sigma_{\widetilde{g}_{0}}\xrightarrow[r\to +\infty]{}-(n-1)(n-2)\,\mathcal{C}\vert \mathbb{S}^{n-1}\vert\,. \end{align} (5.6)

    Indeed

    \begin{align*} \int\limits_{\partial B_{r}} \frac{\mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u}}{\widetilde{u}}(X,\nu_{\widetilde{g}_{0}})\,d\sigma_{\widetilde{g}_{0}}& = \int\limits_{\partial B_{r}} \frac{\mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u}}{\widetilde{u}}(X,\nu_{\widetilde{g}_{0}}-\nu_{e})\,d\sigma_{\widetilde{g}_{0}}+\int\limits_{\partial B_{r}} \frac{\mathrm{D}_{\widetilde{g}_{0}}^{2}\widetilde{u}-\mathrm{D}_{e}^{2}\widetilde{u}}{\widetilde{u}}(X,\nu_{e})\,d\sigma_{\widetilde{g}_{0}}\\ &+\int\limits_{\partial B_{r}} \frac{\mathrm{D}_{e}^{2}\widetilde{u}}{\widetilde{u}}(X,\nu_{e})\,d\sigma_{\widetilde{g}_{0}}\,, \end{align*}

    and one can show, with similar estimates as before, that the first two terms of this sum tend to 0 for r\to+\infty . It is also easy to see, using (2.5) and (2.11), that

    \int\limits_{\partial B_{r}} \frac{\mathrm{D}_{e}^{2}\widetilde{u}}{\widetilde{u}}(X,\nu_{e})\,d\sigma_{\widetilde{g}_{0}}\xrightarrow[r\to +\infty]{}-(n-1)(n-2)\,\mathcal{C}\vert \mathbb{S}^{n-1}\vert\,.

    Hence, (5.6) is proven. Gathering (5.2), (5.5), and (5.6), we have finally obtained

    \begin{align} m_{\mathrm{ADM}}&\leq\mathcal{C}+\limsup\limits_{r\to +\infty}\frac{1}{2(n-2)(n-1)\vert \mathbb{S}^{n-1}\vert}\int\limits_{\partial B_{r}}\mathrm{R}_{\widetilde{g}_{0}}\,g(X,\nu_{\widetilde{g}_{0}})\,d\sigma_{\widetilde{g}_{0}}\,. \end{align} (5.7)

    We remark that the above inequality is true for any \psi chart at infinity of M . From now on we assume that \psi satisfies condition (1.8) regarding the decay rate of \mathrm{R}_{\widetilde{g}_{0}} at \infty . Since

    \begin{align*} \widetilde{g}_{0}(X,\nu_{\widetilde{g}_{0}})& = \widetilde{g}_{0;ij}X^{i}\nu_{\widetilde{g}_{0}}^{j} = \big(\,\delta_{ij}+O(\vert x\vert^{-p})\,\big)X^{i} (\nu_{\widetilde{g}_{0}}^{j}\pm\nu_{e}^{j})\\ & = g_{ \mathbb{R}^{n}}(X,\nu_{e})+O(\vert x\vert^{-p+1}) = \vert x\vert+O(\vert x\vert^{-p+1})\,, \end{align*}

    also using (2.11) we obtain

    \begin{align} \Big\vert\int\limits_{\partial B_{r}}\mathrm{R}_{\widetilde{g}_{0}}\,\widetilde{g}_{0}(X,\nu_{\widetilde{g}_{0}})\,d\sigma_{\widetilde{g}_{0}}\Big \vert\leq C \int\limits_{\partial B_{r}}r^{-q}\big(\,r+O(r^{-p+1})\,\big) \,d\sigma_{e}\leq C r^{-q+n}\xrightarrow[r\to +\infty]{} 0 \,. \end{align} (5.8)

    The fact that m_{\mathrm{ADM}}\leq \mathcal{C} thus follows from (5.7). All in all, the rigidity case m_{\mathrm{ADM}} = \mathcal{C} of [17,Theorem 1.5] holds, which implies that (M, g_{0}) is the Schwarzschild manifold.

    The authors are members of the Gruppo Nazionale per l'Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA), which is part of the Istituto Nazionale di Alta Matematica (INdAM), and they are partially funded by the GNAMPA project "Aspetti geometrici in teoria del potenziale lineare e nonlineare". F. O. thanks L. Benatti and C. Mantegazza for useful discussions during the preparation of the manuscript.

    The authors declare no conflict of interest.

    In this appendix we provide a proof of Lemma 4.5 which is alternative and more self contained than the corresponding in [2]. We underline that we will use Remark 3.1 widely.

    Proof of Lemma 4.5 (i). In \mathring{M}\setminus\mathrm{Crit}(\varphi) and for every \beta\geq 0 , we consider the smooth vector field

    \begin{align*} X_{\beta}: = \frac{\vert \nabla \varphi \vert^{\beta}_{g}\, \nabla \varphi}{\sinh \varphi}\,, \end{align*}

    which is such that

    \begin{align*} \mathrm{div}_{g}\,X_{\beta}& = \frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,\beta\,\nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)\,-\coth(\varphi)\vert \nabla \varphi \vert^{4}_{g} \Big]}{\sinh\varphi}\, \end{align*}

    If \{s\leq \varphi\leq S\}\cap \mathrm{Crit}(\varphi) = \emptyset , then the statement is a straightforward application of the Divergence Theorem. Now, suppose that \{s\leq \varphi\leq S\}\cap \mathrm{Crit}(\varphi) \neq\emptyset . Since there always exists \overline{s}\in (s, S) regular value of \varphi , up to splitting the right–hand side of (4.23) into two subintegrals, we can suppose without loss of generality that one among s and S is a regular value of \varphi . To fix the ideas, suppose that S is the regular value. We are going to change the function \varphi in a neighbourhood of the set \mathrm{Crit}(\varphi) . To do this, for every \varepsilon > 0 sufficiently small, applying Sard's Theorem to the smooth function \varphi , we can fix a positive real number \delta(\varepsilon) such that s+\delta(\varepsilon) < S is a regular value of \varphi and \delta(\varepsilon) < d\, \varepsilon , where d > 0 will be specified later. Then, considering a smooth nonincreasing cut–off function \zeta_{\varepsilon}:[0, +\infty)\to [0, 1] satisfying the conditions

    \begin{align} \zeta_{\varepsilon}(\tau) = 1\quad {\rm{in}} \; \Big[\,0,\frac{1}{2} \varepsilon \,\Big] \,,\quad \vert \zeta_{\varepsilon}'(\tau)\vert \leq \frac c\varepsilon\rm{\quad in \Big[\,\frac{1}{2} \varepsilon ,\frac{3}{2} \varepsilon \,\Big] }\,,\quad \zeta_{\varepsilon}(\tau) = 0 \quad {\rm{in}}\; \Big[\,\frac{3}{2} \varepsilon ,+\infty\,\Big) \,,\, \end{align} (A.1)

    where c is a positive real constant independent of \varepsilon , we define

    \varphi_{\varepsilon}: = \varphi-\zeta_{\varepsilon}(\vert \nabla \varphi\vert^{2}_{g})\,\delta(\varepsilon)\,.

    Clearly,

    \begin{align} \nabla \varphi_{\varepsilon} = \nabla \varphi-\delta(\varepsilon)\, \zeta'_{\varepsilon}\big( \,\vert \nabla \varphi \vert^{2}_{g}\,\big)\,\nabla\vert \nabla \varphi \vert^{2}_{g}, \end{align} (A.2)

    and

    \begin{align} \varphi = \varphi_{\varepsilon}\,\,{\rm{in}}\,\, \Big\{\vert \nabla \varphi \vert^{2}_{g}\geq\frac{3}{2}\varepsilon\,\Big\}\,. \end{align} (A.3)

    Note that s is a regular value for the function \varphi_{\varepsilon} . To see this, let p be a point of \{\varphi_{\varepsilon} = s\} and distinguish the two cases

    \vert \nabla \varphi \vert^{2}_{g}\,(p)\leq\frac{1}{2}\varepsilon\,;\quad \quad\quad \vert \nabla \varphi \vert^{2}_{g}\,(p)\overset{(\bigstar)}{ > }\frac{1}{2}\varepsilon.

    In the first case, \zeta_{\varepsilon}(\vert \nabla \varphi\vert^{2}_{g})\equiv 1 so that s = \varphi_{\varepsilon}(p) = \varphi(p)-\delta(\varepsilon) and \nabla \varphi_{\varepsilon}(p) = \nabla \varphi(p) . Since s+\delta(\varepsilon) is a regular value for \varphi , \nabla \varphi_{\varepsilon}(p)\neq 0 . In the second case, observing that s\leq\varphi(p)\leq s+\delta(\varepsilon) and therefore p\in\{s\leq \varphi\leq S\} , we have from (A.2) that in p

    \begin{align*} \vert \nabla \varphi_{\varepsilon}\vert_{g}&\geq \vert \nabla \varphi \vert_{g}-\delta(\varepsilon)\,\vert\zeta'_{\varepsilon}\vert\big( \,\vert \nabla \varphi \vert^{2}_{g}\,\big)\big\vert \nabla\vert \nabla \varphi \vert^{2}_{g}\big\vert_{g} = \vert \nabla \varphi \vert_{g}\Big(1-2\delta(\varepsilon)\,\vert\zeta'_{\varepsilon}\vert\big( \,\vert \nabla \varphi \vert^{2}_{g}\,\big)\big\vert \nabla\vert \nabla \varphi \vert_{g}\big\vert_{g}\Big)\\ &\geq\vert \nabla \varphi \vert_{g}\Big(1-2d\,\varepsilon\, \frac{c}{\varepsilon}\, \max\limits_{\{s\leq \varphi\leq S\}}\big\vert\nabla\vert \nabla \varphi \vert_{g}\,\big\vert_{g}\,\Big)\,, \end{align*}

    where c is the constant appearing in (A.1). Now, observe that \max\limits_{\{s\leq \varphi\leq S\}}\big\vert\nabla\vert \nabla \varphi \vert_{g}\, \big\vert_{g}\, > 0 , since otherwise, due to the presence of critical points in \{s\leq \varphi\leq S\} , there should be a connected component of \{s\leq \varphi\leq S\} where \nabla \varphi\equiv0 . But this is impossible because \{s\leq \varphi\leq S\} = \overline{\{s < \varphi < S\}} (by Remark 2.1) and by the size of \mathrm{Crit}(\varphi) . Hence, choosing

    d\leq \frac{1}{4\,c\,\max\limits_{\{s\leq \varphi\leq S\}}\big\vert\nabla\vert \nabla \varphi \vert_{g}\,\big\vert_{g }}\,,

    from above we obtain \vert\nabla \varphi_{\varepsilon}\vert_{g}(p)\geq \frac{\vert\nabla \varphi\vert_{g}}{2}(p) . In particular, from (\bigstar) we get that \vert\nabla \varphi_{\varepsilon}\vert_{g}(p) > \frac{\varepsilon}{4} .

    Now, we apply the Divergence Theorem to the smooth vector field \Xi_{4\varepsilon} X_{\beta} on \{s < \varphi_{\varepsilon} < S\} , where

    \Xi_{\varepsilon}\,: = \,1- \zeta_{\varepsilon}(\vert \nabla \varphi\vert^{2}_{g}).

    Recalling that U_{\mu} is defined as in (4.16), we obtain

    \begin{align*} \int\limits_{\{\varphi_{\varepsilon} = S\}}\,g\Big(\,\Xi_{4\varepsilon} X_{\beta},\frac{\nabla \varphi_{\varepsilon}}{\vert \nabla \varphi_{\varepsilon}\vert_{g}}\,\Big)d\sigma_{g}&- \int\limits_{\{\varphi_{\varepsilon} = s\}}\,g\Big(\,\Xi_{4\varepsilon} X_{\beta},\frac{\nabla \varphi_{\varepsilon}}{\vert \nabla \varphi_{\varepsilon}\vert_{g}}\,\Big)\,d\sigma_{g}\\ & = \int\limits_{\{s < \varphi_{\varepsilon} < S\}}\Xi_{4\varepsilon}\frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,\beta\,\nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)\,-\coth(\varphi)\vert \nabla \varphi \vert^{4}_{g} \Big]}{\sinh\varphi}\,d\mu_{g}\\ &-2\int\limits_{ \big(U_{6\varepsilon} \setminus\overline{U_{2\varepsilon}}\big)\cap \{s < \varphi_{\varepsilon} < S\}}\,\frac{\,\zeta'_{4\varepsilon}\big(\,\vert \nabla \varphi \vert^{2}_{g}\,\big)\,\vert \nabla \varphi \vert^{\beta}_{g}\,\nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)\,}{\sinh \varphi}\,d\mu_{g}\,. \end{align*}

    Note that \{\varphi = S\} is compactly contained in \{\vert \nabla \varphi\vert^{2}_{g} > \frac{3}{2}\, \varepsilon\} for every \varepsilon sufficiently small, and \Xi_{4\varepsilon}\equiv0 in \{\vert \nabla \varphi\vert^{2}_{g}\leq2\, \varepsilon\}\supset \{\vert \nabla \varphi\vert^{2}_{g}\leq\frac{3}{2}\, \varepsilon\} . Then, by (A.3) we get

    \begin{align} \int\limits_{\{\varphi = S\}}\,\frac{\vert \nabla \varphi \vert^{\beta+1}_{g}}{\sinh\varphi}\,d\sigma_{g}&- \int\limits_{\big\{\,\varphi = s,\vert \nabla \varphi \vert^{2}_{g}\geq\frac{3}{2}\varepsilon\,\big\}}\,\Xi_{4\varepsilon}\,\frac{\vert \nabla \varphi \vert^{\beta+1}_{g}}{\sinh\varphi}\,d\sigma_{g}\\ & = \int\limits_{\{s < \varphi < S\}}\Xi_{4\varepsilon}\,\frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,\beta\,\nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)\,-\coth(\varphi)\vert \nabla \varphi \vert^{4}_{g} \Big]}{\sinh\varphi}\,d\mu_{g}\\ &-2\int\limits_{ \big(U_{6\varepsilon} \setminus\overline{U_{2\varepsilon}}\big)\cap \{s < \varphi < S\}}\,\frac{\,\zeta'_{4\varepsilon}\big(\,\vert \nabla \varphi \vert^{2}_{g}\,\big)\,\vert \nabla \varphi \vert^{\beta}_{g}\,\nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)\,}{\sinh \varphi}\,d\mu_{g}\,.\, \end{align} (A.4)

    Looking at the left–hand side of (A.4), note that

    \begin{align*} \Bigg\vert\int\limits_{ \big(U_{6\varepsilon}\setminus\overline{U_{2\varepsilon}}\big)\cap \{s < \varphi < S\}}\zeta'_{4\varepsilon}\big(\,\vert \nabla \varphi \vert^{2}_{g}\,\big) \vert \nabla \varphi \vert^{\beta}_{g} \nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)\,d\mu_{g}\Bigg\vert &\leq\frac{c}{4\,\varepsilon}\int\limits_{ U_{6\varepsilon}} \vert \nabla \varphi \vert^{\beta+2}_{g}\vert \nabla ^{2}\varphi\vert_{g}\,d\mu_{g}\\ &\leq C\frac{\varepsilon^{\frac{\beta}{2}+1}}{\varepsilon}\mu_{g}(U_{6\varepsilon} )\to 0\,, \end{align*}

    where in the second inequality we have used Lemma 4.1 and the fact that U_{\varepsilon} is contained in a compact set for every \varepsilon < < 1 (which is a consequence of (4.6)). Moreover, by the Dominated Convergence Theorem, we have that

    \begin{align*} \lim\limits_{\varepsilon\to 0^+}\int\limits_{\{s < \varphi < S\}}\Xi_{4\varepsilon}\,&\frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,\beta\,\nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)\,-\coth(\varphi)\vert \nabla \varphi \vert^{4}_{g} \Big]}{\sinh\varphi}\,d\mu_{g}\\ & = \int\limits_{\{s < \varphi < S\}}\,\frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,\beta\,\nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)\,-\coth(\varphi)\vert \nabla \varphi \vert^{4}_{g} \Big]}{\sinh\varphi}\,d\mu_{g}\,. \end{align*}

    Finally, note that \{\zeta_{\varepsilon}\} can always be chosen to be nonincreasing in \varepsilon so that, in turn, \{\Xi_{\varepsilon}\} in nondecreasing. Therefore, looking at the left–hand side of (A.4), we have that

    \begin{align*} \lim\limits_{\varepsilon\to 0^+}\int\limits_{\big\{\,\varphi = s,\vert \nabla \varphi \vert^{2}_{g}\geq\frac{3}{2}\varepsilon\,\big\}}\,\Xi_{4\varepsilon}\,\frac{\vert \nabla \varphi \vert^{\beta+1}_{g}}{\sinh\varphi}\,d\sigma_{g}& = \lim\limits_{\varepsilon\to 0^+}\int\limits_{\{\,\varphi = s\}}\,\Xi_{4\varepsilon}\,\frac{\vert \nabla \varphi \vert^{\beta+1}_{g}}{\sinh\varphi}\,d\sigma_{g}\\ & = \int\limits_{\{\,\varphi = s \}}\,\frac{\vert \nabla \varphi \vert^{\beta+1}_{g}}{\sinh\varphi}\,d\sigma_{g}\,, \end{align*}

    by the Monotone Convergence Theorem. All in all, passing to the limit as \varepsilon\to 0^+ in (A.4), yields the desired identity.

    Proof of Lemma 4.5 (ii). Lemma 4.1 implies

    \lim\limits_{S\to+\infty}\int\limits_{\{\varphi = S\}}\frac{\vert \nabla \varphi \vert^{\beta+1}_{g}}{\sinh\varphi}\,d\sigma_{g} = 0\,.

    Note that

    \begin{align*} \frac{\vert \nabla \varphi \vert^{\beta-2}_{g}\,\Big[\,\coth(\varphi)\vert \nabla \varphi \vert^{4}_{g}-\beta\,\nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)\, \Big]}{\sinh\varphi}\in L^{1}\Big(\{\varphi\geq s\};\mu_{g}\Big) \end{align*}

    because its absolute value belongs to L^{1}_{\rm{loc}}\big(\{\varphi\geq s\}, \mu_{g}\big) immediately and to L^{1}\big(\{\varphi\geq S\}, \mu_{g}\big) for S sufficiently big, applying the Coarea Formula coupled with (4.6) and Lemma 4.5. Therefore, passing to the limit as S\to +\infty in (4.23) and using the Dominated Convergence Theorem gives the desired identity.

    Proof of Lemma 4.5 (iii). Let \beta\geq0 . We are assuming that the boundary \partial M is a regular level set of \varphi so that there exists \epsilon > 0 such that [0, \epsilon]\cap\mathrm{Crit}(\varphi) = \emptyset . Therefore, applying the Divergence Theorem to the smooth vector field \vert \nabla \varphi \vert^{\beta}_{g}\, \nabla \varphi in \{0 < \varphi < \epsilon\} yields

    \Phi_{\beta}(\epsilon)-\Phi_{\beta}(0) = \int\limits_{\{0 < \varphi < \epsilon\}}\,\beta\,\vert \nabla \varphi \vert^{\beta-2}_{g}\,\nabla ^{2}\varphi(\nabla \varphi,\nabla \varphi)\,d\mu_{g}\,.

    In turn, the absolute continuity of the integral implies the continuity of \Phi_{\beta} at 0 . By point (i) and again by the absolute continuity of the integral, we obtain the right and the left continuity of the function

    \begin{equation*} \Upsilon_{\beta}:s\in(0,+\infty)\to\frac{\Phi_{\beta}(s)}{\sinh s}\in \mathbb{R}\,. \end{equation*}

    Hence, \Phi_{\beta} is continuous also in (0, +\infty) . The integral representation of \Phi_{\beta} follows directly from point (ii) .



    [1] Marzluff JM (2008) Urban ecology: an international perspective on the interaction between humans and nature, Springer.
    [2] Kant N, Anjali K (2020) Climate Strategy Proactivity (CSP): A Stakeholders-Centric Concept, In: Filho WL, Azul AM, Brandli L, et al. (Eds.), Partnerships for the Goals, Encyclopedia of the UN Sustainable Development Goals (Living Reference), Switzerland AG, Springer Nature, 1-16.
    [3] Gruber N, Friedlingstein P, Field C, et al. (2004) The vulnerability of the carbon cycle in the 21st century: An assessment of carbon-climate-human interactions, In: Gruber N, Friedlingstein P, Field CB, et al. (Eds.), The Global Carbon Cycle: Integrating Humans, Climate, and the Natural World, Washington, D.C.; London: Island Press, 45-76.
    [4] Lal M, Singh R (2000) Carbon Sequestration Potential of Indian Forests. Environ Monit Assess 60: 315-327.
    [5] Ravindranath N, Ostwald M (2008) Methods for estimating above ground biomass, In: Ravindranath N, Ostwald M (Eds.), Carbon inventory methods: handbook for greenhouse gas inventory, carbon mitigation and round wood production projects, Netherlands, Springer Science + Business Media B.V.
    [6] IPCC (2000) A special report of the IPCC: Land use, Land-use change and Forestry, UK, Cambridge University Press.
    [7] Cox HM (2012) A Sustainability Initiative to Quantify Carbon Sequestration by Campus Trees. J Geog 111: 173-183.
    [8] United Nations (2016) Urbanization and Development: World Cities Report 2016. Nairobi, United Nations Human Settlements Programme, 2016.
    [9] Nowak DJ (2016) Urban Forests, Assessing the Sustainability of Agricultural and Urban Forests in the United States, 37-52.
    [10] Fang J, Yu G, Liu L, et al. (2018) Climate change, human impacts, and carbon sequestration in China. Proc Natl Acad Sci 115: 4015-4020.
    [11] Yeshaneh E, Wagner W, Exner-Kittridge M, et al. (2013) Identifying land use/cover dynamics in the Koga catchment, Ethiopia, from multi-scale data, and implications for environmental change. ISPRS Int J Geo-Inf 302-323.
    [12] Schwela D (2000) Air Pollution and health in urban areas. Rev Environ Health 15: 13-42.
    [13] Mcpherson EG, Rowntree RA (1993) Energy conservation potential of urban tree. J Aboriculture 321-331.
    [14] McPherson EG (1998) Atmospheric carbon dioxide reduction by Sacramento's urban forest. J Aboriculture 215-223.
    [15] Nowak DJ, Crane DE (2002) Carbon storage and sequestration by urban trees in the USA. Environ Pollut 116: 381-389.
    [16] Yang J, McBride J, Zhou J, et al. (2005) The urban forest in Beijing and its role in air pollution reduction. Urban For Urban Green 3: 65-78.
    [17] Nowak DJ, Crane DE, Stevens JC (2006) Air pollution removal by urban trees and shrubs in the United States. Urban For Urban Green 4: 115-123.
    [18] Zhao X, Lynch JG, Chen Q (2010) Reconsidering Baron and Kenny: Myths and Truths about Mediation Analysis. J Consum Res 37: 197-206.
    [19] Nowak DJ, Hirabayashi S, Bodine A, et al. (2014) Tree and Forest Effects on Air Quality and Human Health in the United States. Environ Pollut 119-129.
    [20] Skofronick-Jackson G, Petersen WA, Berg W, et al. (2016) The Global Precipitation Measurement (GPM) Mission for Science and Society. Bull Am Meteorol Soc 98: 1679-1695.
    [21] Livesley SJ, McPherson EG, Calfapietra C (2016) The Urban Forest and Ecosystem Services: Impacts on Urban Water, Heat, and Pollution Cycles at the Tree, Street, and City Scale. J Environ Qual 45: 119-124.
    [22] McCarthy MP, Best M, Betts R (2010) Climate change in cities due to global warming and urban effects. Geophys Res Lett 37.
    [23] Borelli S, Conigliaro M, Pineda F (2018) Urban forests in the global context. Unasylva 250 69: 3-10.
    [24] Konijnendijk CC, Randrup TB (2004) Urban Forestry, LANDSCAPE AND PLANNING, 471-478.
    [25] BATJES NH, SOMBROEK WG (1997) Possibilities for carbon sequestration in tropical and subtropical soils. Glob Chang Biol 3: 161-173.
    [26] Canadell JG, Kirschbaum MUF, Kurz WA, et al. (2007) Factoring out natural and indirect human effects on terrestrial carbon sources and sinks. Environ Sci Policy 10: 370-384.
    [27] Lal R (2008) Carbon sequestration. Philos Trans R Soc Biol Sci 363: 815-830.
    [28] IPCC (2014) Climate change 2014: Synthesis report summary for policymakers. Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change, Geneva, Switzerland.
    [29] Salunkhe O, Khare PK, Kumari R, et al. (2018) A systematic review on the aboveground biomass and carbon stocks of Indian forest ecosystems. Ecol Process 7-17.
    [30] Afzal M, Akhtar AM (2013) Factors affecting carbon sequestration in trees. J Agric Res 51: 61-69.
    [31] Sahu C, Sahu KS (2015) Air Pollution Tolerance Index (APTI), Anticipated Performance Index (API), Carbon Sequestration and Dust Collection Potential of Indian Tree Species: A Review. Emerg Res Manag & Technology 4-11.
    [32] Stern NS (2007) The Economics of Climate Change: The Stern Review, Cambridge, UK, Cambridge University Press.
    [33] Stephson NL, Das AJ, Condit R, et al. (2014) Rate of tree carbon accumulation increases continuously with tree size. Nature 507: 90-93.
    [34] WMO (2017) Statement on the state of the global climate change.
    [35] Raupach MR, Canadell JG (2010) Carbon and the Anthropocene. Curr Opin Environ Sustain 210-218.
    [36] Akbari H (2002) Shade trees reduce building energy use and CO2 emissions from power plants. Environ Pollut 116: 119-126.
    [37] McHale MR, McPherson EG, Burke IC (2007) The potential of urban tree plantings to be cost effective in carbon credit markets. Urban For Urban Green 49-60.
    [38] Hartmann F, Perego P, Young A (2013) Carbon Accounting: Challenges for Research in Management Control and Performance Measurement. Abacus 49: 539-563.
    [39] Cairns RD, Lasserre P (2006) Implementing carbon credits for forests based on green accounting. Ecol Econ 56: 610-621.
    [40] Gazioğlu C, Okutan V (2016) Underwater Noise Pollution at the Strait of Istanbul (Bosphorus). Int J Environ Geoinformatics 3: 26-39.
    [41] Sasaki N, Kim S (2009) Biomass carbon sinks in Japanese forests: 1966-2012. Forestry 82: 105-115.
    [42] Suryawanshi M, Patel A, Kale T, et al. (2014) Carbon sequestration potential of tree species in the environment of North Maharashtra University Campus, Jalgaon (MS) India. Biosci Discov 5: 175-179.
    [43] Jha KK (2015) Carbon storage and sequestration rate assessment and allometric model development in young teak plantations of tropical moist deciduous forest, India. J For Res 26: 589-604.
    [44] Das M, Mukherjee A (2015) Carbon sequestration potential with height and girth of selected trees in the Golapbag Campus, Burdwan, West Bengal(India). Indian J Sci Res 10: 53-57.
    [45] Guarna B (2012) An Analysis of Carbon Sequestration on Clarkson University 's Campus. B.S. ES & P '12. 2011-2012 Capstone.
    [46] IPCC (2007) Synthesis Report. Contribution of Working Groups I, II and III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change.
    [47] Paustian K, Lehmann J, Ogle S, et al. (2016) Climate-Smart Soils. Nature 532: 49-57.
    [48] Sedjo R, Sohngen B, Jagger P (1998) Carbon Sinks in the Post-Kyoto World. RRF Clim Issue 895-908.
    [49] Khurana P (2012) A study on carbon sequestration in natural forests of India. J Appl Nat Sci 4: 132-136.
    [50] Melkania N (2009) Carbon Sequestration in Indian Natural and Planted Forests. Indian For 380-387.
    [51] Brown S (1996) Tropical forests and the global carbon cycle: estimating state and change in biomass density BT-Forest Ecosystems, Forest Management and the Global Carbon Cycle, In: Apps MJ, Price DT (Eds.), Berlin, Heidelberg, Springer Berlin Heidelberg, 135-144.
    [52] Kiran GS, Kinnary S (2011) Carbon Sequatration by urban trees on roadside of Vadodara city. Int J Eng Sci Technol 3: 3066-3070.
    [53] Chavan BL, Rasal GB (2010) Sequestered standing carbon stock in selective tree species grown in University campus at Aurangabad. Int J Eng Sci Technol 12: 3003-3007.
    [54] Yunusa A, Linatoc A (2018) Inventory of Vegetation and Assessment of Carbon Storage Capacity towards a Low Carbon Campus: a Case Study of Universiti Tun Husein. Traektoria Nauk = Path Sci 4: 3001-3006.
    [55] Ramachandran A, Jayakumar S, Haroon R, et al. (2007) Carbon sequestration: estimation of carbon stock in natural Forestsusing Geospatial technology in the Eastern Ghats of TN, India. Curr Sci 323-331.
    [56] Singh R, Pal D, Tripathi N (2018) Land use/Land cover change detection analysis using remote sensing and GIS of Dhanbad District. India. Eurasian J For Sci 6: 1-12.
    [57] Lal R (2004) Soil carbon sequestration impacts on global climate change and food security. Science (80-) 304: 1623-1627.
    [58] Kumar BM (2006) Carbon sequestration potential of tropical garden, In: Kumar BM, Nair P (Eds.), Tropical Home garden: a time tested example of sustainable agroforstry, Netherlands, Springer, 185-204.
    [59] Sheikh M, Kumar M (2010) Carbon sequestration potential of trees on two aspects in sub tropical forest. Int J Conserv Sci 1143-1148.
    [60] Marak T, Khare N (2017) Carbon Sequestration Potential of Selected Tree Species in the Campus of SHUATS. Int J Sci Res Dev 5: 63-66.
    [61] Keenan TF, Williams CA (2018) The Terrestrial Carbon Sink. Annu Rev Environ Resour 43: 219-243.
    [62] Chavan BL, Rasal GB (2009) Carbon Storage in Selective Tree Species in University campus at Aurangabad, Maharashtra, India, International Conference and Exhibition on RAEP, Agra, india., 119-130.
    [63] Yirdaw M (2018) Carbon Stock Sequestered by Selected Tree Species Plantation in Wondo Genet College, Ethiopia. J Earth Sci Clim Chang 9: 1-5.
    [64] Kongsager R, Napier J, Mertz O (2013) The carbon sequestration potential of tree crop plantations. Mitig Adapt Strateg Glob Chang 18: 1197-1213.
    [65] Verchot L, Noordwijk M, Kandii S, et al. (2007) Climate change: Linking adaptation and mitigation through agroforestry. Mitig Adapt Strateg Glob Chang 12: 901-918.
    [66] Soto-Pinto L, Anzueto M, Mendoza J, et al. (2010) Carbon sequestration through agroforestry in indigineous community of chiapas, Mexico. Agrofor Syst 78: 1-39.
    [67] Nowak DJ (1993) Atmospheric carbon reduction by urban trees. J EnvironManagement 37: 207-217.
    [68] Montagnini F, Nair PKR (2004) Carbon sequestration: An un- derexploited environmental benefit of agroforestry systems. Agro- for Syst 61: 281-295.
    [69] Brack CL (2002) Pollution mitigation and carbon sequestration by an urban forest. Environ Pollut 116: 195-200.
    [70] Churkina G (2012) Carbon cycle in urban ecosystem, In: Lal R, Augustin B (Eds.), Carbon Sequestration in urban ecosystems, Netherlands, Springer.
    [71] Joshi RK, Dhyani S (2019) Biomass, carbon density and diversity of tree species in tropical dry deciduous forests in Central India. Acta Ecol Sin 39: 289-299.
    [72] Flora G, Athista GM, Derisha L, et al. (2018) Estimation of Carbon Storage in the Tree Growth of St. Mary's College (Autonomous) Campus, Thoothukudi, Tamilnadu, India. Int J Emerg Technol Innov Res 5: 260-266.
    [73] Ketterings QM, Coe R, van Noordwijk M, et al. (2001) Reducing uncertainty in the use of allometric biomass equations for predicting above-ground tree biomass in mixed secondary forests. For Ecol Manage 146: 199-209.
    [74] ANAMIKA A, PRADEEP C (2016) Urban Vegetation and Air Pollution Mitigation: Some Issues from India. Chinese J Urban Environ Stud 04: 1-10.
    [75] Salazar S, Sanchez L, Galiendo P, et al. (2010) AG tree biomass equation and nutrient pool for a para climax chestnut stand and for a climax oak stand. Sci Res Essays 1294-1301.
    [76] Parresol B (1999) Assessing tree and stand biomass: A review with examples and critical comparisions. For Sci 45: 573-593.
    [77] Pragasan L (2014) Carbon Stock Assessment in the Vegetation of the Chitteri Reserve Forest of the Eastern Ghats in India Based on Non-Destructive Method Using Tree Inventory Data. J Earth Sci Clim Chang 11-21.
    [78] Brown S, Lugo AE (1982) The Storage and Production of Organic Matter in Tropical Forests and Their Role in the Global Carbon Cycle. Biotropica 14: 161-187.
    [79] Haripriya GS (2000) Estimates of biomass in Indian forests. Biomass and Bioenergy 19: 245-258.
    [80] Haghparast H, Delbari AS, Kulkarni D (2013) Carbon sequestration in Pune University Campus with special reference to Geographical information system. J Ann Biol Res 2013, 4 4: 169-175.
    [81] Vashum KT, Jayakumar S (2012) Methods to Estimate Above-Ground Biomass and Carbon Stock in Natural Forests: A Review. J Ecosyst Ecogr 2-4.
    [82] Nowak DJ (2012) Contrasting natural regeneration and tree planting in 14 North American cities. Urban For Urban Green 11: 374-382.
    [83] Nowak DJ, Dwyer JF (2007) Understanding the benefits and costs of urban forest ecosystems, In: Kuser J (Ed.), Urban and community forestry in the Northeast, New York: Springer, 25-46.
    [84] Patel J, Vasava A, Patel N (2017) Occurence of the Forest Owlet Heteroglaux blewitti in Navsari and Valsad Districts of Gujrat, India. Indian Birds 13: 78-79.
    [85] Yang J, Yu Q, Gong P (2008) Quantifying air pollution removal by green roofs in Chicago. Atmos Environ 42: 7266-7273.
    [86] Chiesura A (2004) The role of urban parks for the sustainable city. Landsc Urban Plan 68: 129-138.
    [87] Vailshery LS, Jaganmohan M, Nagendra H (2013) Effect of street trees on microclimate and air pollution. Urban For Urban Green 12: 408-415.
    [88] Dhadse S, Gajghate DG, Chaudhari PR, et al. (2011) Interaction of urban vegetation cover to sequester air pollution from ambient air environment, In: Moldovennu AM (Ed.), Air pollution: new developments, Rijeka, Croatia, InTech.
    [89] Dhyani SK, Ram A, Dev I (2016) Potential of agroforestry systems in carbon sequestration in India. Indian J Agric Sci 86: 1103-1112.
    [90] Chaudhry P (2016) Urban greening: a review of some Indian studies. Brazilian J Biol Sci 3: 425-432.
    [91] Jim CY, Chen WY (2006) Recreation amenity use and contingent valuation of urban green spaces in Guangzhou, China. Landsc Urban Plan 52: 117-133.
    [92] Stoner T, Rapp C (2008) Open Spaces Sacred Places.
    [93] Tripathi M, Joshi H (2015) Carbon flow in Delhi urban forest ecosystems. Ann Biol Res 6: 13-17.
    [94] Nowak DJ, Crane D, Stevens J, et al. (2008) A ground-based method of assessing urban forest structure and ecosystem services. Arboric Urban For 34: 347-358.
    [95] Peper PJ, McPherson EG (2003) Evaluation of four methods for estimating leaf area of isolated trees. Urban For Urban Green 2: 19-29.
    [96] Dobbs C, Martinez-Harms M, Kendal D (2017) Ecosystem services, In: Ferrini F, Bosch CK van den, A. Fini (Eds.), Routledge handbook of urban forestry, Abingdon, UK, Routledge.
    [97] Ramaiah M, Avtar R (2019) Urban Green Spaces and Their Need in Cities of Rapidly Urbanizing India: A Review. Urban Sci 3.
    [98] Arnold A V, Edward D (2014) A model for Low Carbon Campus in Kerala. Int J Sci Eng Res 5: 95-99.
    [99] Jim CY (2018) Protecting heritage trees in urban and peri-urban environments. Unasylva 250 69: 1.
    [100] Arya A, Shalini Negi S, Kathota JC, et al. (2017) Carbon Sequestration Analysis of dominant tree species using Geo-informatics Technology in Gujarat State (INDIA). Int J Environ Geoinformatics 4: 79-93.
    [101] Narayana J, Savinaya MS, Manjunath S, et al. (2017) Distribution and diversity of flora and fauna in and around Kuvempu University campus, Bhadra Wildlife Sanctuary range, Karnataka. Int J Plant Anim Env Sci 7: 89-99.
    [102] Khera N, Mehta V, Sabata BC (2009) Interrelationship of birds and habitat features in urban green spaces in Delhi, India. Urban For Urban Green 8: 187-196.
    [103] Nath A jyothi, Das A kumar (2011) Carbon storage and sequestration in bamboo-based smallholder homegardens of Barak Valley, Assam. Curr Sci 100: 229-233.
    [104] Roy S, Byrne J, Pickering C (2012) A systematic quantitative review of urban tree benefits, costs, and assessment methods across cities in different climatic zones. Urban For Urban Green 11: 351-363.
    [105] Song XP, Tan PY, Edwards P, et al. (2017) The economic benefits and costs of trees in urban forest stewardship: A systematic review. Urban For Urban Greening2.
    [106] Nowak DJ (2010) Urban Biodiversity and Climate Change, In: Muller N, Warner P, Kelcey J (Eds.), Urban Biodiversity and Design, Hoboken, NJ, Wiley-Blackwell Publishing, 101-117.
    [107] IGNOU (2020) IGNOU Profile 2020, Indira Gandhi National Open University, New Delhi, India.
    [108] Silver WL, Ostertag R, Lugo AE (2000) The potential for carbon sequestration through reforestation of abandoned tropical agricultural and pasture lands. Restor Ecol 394-407.
    [109] Nero BF, Conche DC, Anning A, et al. (2017) Urban Green Spaces Enhance Climate Change Mitigation in Cities of the Global South: the case of kumasi Ghana. Procedia Eng 68-83.
    [110] Abdollahi KK, Ning ZH, Appeaning A (2000) Global climate change & the urban forest, GCRCC/Franklin Press, Baton Rouge, LA.
    [111] Wilby R, Perry G (2006) Climate change, biodiversity and the Urban Environment: A critical review based on London, UK. Prog Phys Geogr 30: 73-98.
    [112] Gill S., Handley J., Ennos A., et al. (2007) Adapting Cities for Climate Change: The Role of the Green Infrastructure. Built Environ 33: 115-133.
    [113] Uddin S, Al Ghadban A, Dousari AA, et al. (2010) A remote sensing classification for Land cover changes and micro climate in Kuwait. Int J Sustain Dev Plan 5: 367-377.
    [114] Chaudhry P, Tewari VP (2011) Urban forestry in India: development and research scenario. Interdiscip Environ Rev 12: 80.
    [115] Bolund P, Hunhammar S (1999) Ecosystem services in urban areas. Ecol Econ 29: 293-301.
    [116] Ferrini F, Fini A (2011) Sustainable management techniques for trees in the urban areas. J Biodivers Ecol Sci 1: 1-20.
    [117] Tripathi M (2016) Carbon Management with Urban Trees. Indian J Fundam Appl Life Sci 6: 40-46.
    [118] Chaudhry P, Dollo M, Bagra K, et al. (2011) Traditional biodiversity conservation and natural resource management system of some tribes of Arunachal Pradesh, India. Interdiscip Environ Rev 12: 338.
    [119] Ugle P, Rao S, Ramachandra TV (2010) Carbon sequestration potential of Urban trees, Wetlands, Biodiversity and Climate Change Conference at IISc, Bengaluru, Bengaluru, India, 1-12.
    [120] Ren Z, Zhal C, Shen G, et al. (2015) Effects of forest type and urbanization on carbon storage of urban forest in changchun, NE. China, Springer.
    [121] Potadar VR, Patil SS (2017) Potential of carbon sequestration and storage by trees in and around B. A. M. University campus of Aurangabad city in Maharashtra, India. Int J Sci Dev Res 2: 28-33.
    [122] Velasco E, Roth M, Norford L, et al. (2016) Does urban vegetation enhance carbon sequestration? Landsc Urban Plan 148: 99-107.
  • This article has been cited by:

    1. Luca Benatti, Mattia Fogagnolo, Lorenzo Mazzieri, The asymptotic behaviour of p-capacitary potentials in asymptotically conical manifolds, 2022, 0025-5831, 10.1007/s00208-022-02515-4
    2. Mattia Fogagnolo, Andrea Pinamonti, New integral estimates in substatic Riemannian manifolds and the Alexandrov Theorem, 2022, 163, 00217824, 299, 10.1016/j.matpur.2022.05.007
    3. Luca Benatti, Mattia Fogagnolo, Lorenzo Mazzieri, Minkowski inequality on complete Riemannian manifolds with nonnegative Ricci curvature, 2024, 17, 1948-206X, 3039, 10.2140/apde.2024.17.3039
    4. Brian Harvie, Ye-Kai Wang, A rigidity theorem for asymptotically flat static manifolds and its applications, 2024, 0002-9947, 10.1090/tran/9134
  • 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(6343) PDF downloads(527) Cited by(6)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog