Research article

Applications of the lichnerowicz Laplacian to stress energy tensors

  • Received: 23 June 2017 Accepted: 30 August 2017 Published: 13 September 2017
  • A generalization of the Laplacian for p-forms to arbitrary tensors due to Lichnerowicz will be applied to a 2-tensor which has physical applications. It is natural to associate a divergencefree symmetric 2-tensor to a critical point of a specific variational problem and it is this 2-tensor that is studied. Numerous results are obtained for the stress-energy tensor, such as its divergence and Laplacian. A remarkable integral formula involving a symmetric 2-tensor and a conformal vector field is obtained as well.

    Citation: Paul Bracken. Applications of the lichnerowicz Laplacian to stress energy tensors[J]. AIMS Mathematics, 2017, 2(3): 545-556. doi: 10.3934/Math.2017.2.545

    Related Papers:

    [1] Rajesh Kumar, Sameh Shenawy, Lalnunenga Colney, Nasser Bin Turki . Certain results on tangent bundle endowed with generalized Tanaka Webster connection (GTWC) on Kenmotsu manifolds. AIMS Mathematics, 2024, 9(11): 30364-30383. doi: 10.3934/math.20241465
    [2] George Rosensteel . Differential geometry of collective models. AIMS Mathematics, 2019, 4(2): 215-230. doi: 10.3934/math.2019.2.215
    [3] J. Alberto Conejero, Antonio Falcó, María Mora–Jiménez . A pre-processing procedure for the implementation of the greedy rank-one algorithm to solve high-dimensional linear systems. AIMS Mathematics, 2023, 8(11): 25633-25653. doi: 10.3934/math.20231308
    [4] Yanlin Li, Aydin Gezer, Erkan Karakaş . Some notes on the tangent bundle with a Ricci quarter-symmetric metric connection. AIMS Mathematics, 2023, 8(8): 17335-17353. doi: 10.3934/math.2023886
    [5] Yinzhen Mei, Chengxiao Guo, Mengtian Liu . The bounds of the energy and Laplacian energy of chain graphs. AIMS Mathematics, 2021, 6(5): 4847-4859. doi: 10.3934/math.2021284
    [6] Xi Wang . Blow-up criterion for the 3-D inhomogeneous incompressible viscoelastic rate-type fluids with stress-diffusion. AIMS Mathematics, 2025, 10(1): 1826-1841. doi: 10.3934/math.2025084
    [7] Lito E. Bocanegra-Rodríguez, Yony R. Santaria Leuyacc, Paulo N. Seminario-Huertas . Exponential stability of a type III thermo-porous-elastic system from a new approach in its coupling. AIMS Mathematics, 2024, 9(8): 22271-22286. doi: 10.3934/math.20241084
    [8] Tong Wu, Yong Wang . Super warped products with a semi-symmetric non-metric connection. AIMS Mathematics, 2022, 7(6): 10534-10553. doi: 10.3934/math.2022587
    [9] Dijian Wang, Dongdong Gao . Laplacian integral signed graphs with few cycles. AIMS Mathematics, 2023, 8(3): 7021-7031. doi: 10.3934/math.2023354
    [10] Jahfar T K, Chithra A V . Central vertex join and central edge join of two graphs. AIMS Mathematics, 2020, 5(6): 7214-7233. doi: 10.3934/math.2020461
  • A generalization of the Laplacian for p-forms to arbitrary tensors due to Lichnerowicz will be applied to a 2-tensor which has physical applications. It is natural to associate a divergencefree symmetric 2-tensor to a critical point of a specific variational problem and it is this 2-tensor that is studied. Numerous results are obtained for the stress-energy tensor, such as its divergence and Laplacian. A remarkable integral formula involving a symmetric 2-tensor and a conformal vector field is obtained as well.


    1. Introduction

    Variational problems arise in various areas of mathematics and physics. Suppose (M,g) is a Riemannian manifold with volume form dvM, it is the case that functionals of the form

    I(φ,g)=Mσ(φ,g)dvM (1.1)

    very often occur [1,2]. Here φ could be a mapping between Riemannian manifolds of a vector bundle valued differential form. Given a variational problem starting from (1.1), the stress-energy tensor S can be derived by considering variations of the metric on M. If I has a critical point with respect to variations of φ, then the stress-energy tensor is divergence free, and there are conservation laws.

    To provide some motivation, let (M,g) and (N,h) be two smooth Riemannian manifolds which are connected, compact, orientable and without boundary, and φ:(M,g)(N,h) a smooth map. The differential of φ which is dφ can be thought of as a section of the bundle TMφ1TN, with norm |dφ|. If {xi} and {ua} constitute local coordinate systems around x and φ(x), respectively, then in terms of coordinates, we can write

    |dφ|2=gijhab(φ)φaxiφbxj, (1.2)

    where (φa/xi) is the local representation of dφ. Then the energy density of φ can be defined as e(φ)=1/2|dφ|2 and the energy density of the field is given by the positive functional E(φ)=Me(φ)dvM.

    A large class of maps which come up in physics, especially in gravity, are called harmonic. A mapping φ:MN is harmonic if and only if it is an extremal of the energy. Consequently, it is the case that a map φ is harmonic if and only if it satisifes the Euler-Lagrange equation

    τ(φ)=ddφ=trdφ.

    This defines the tension field of φ, and may be expressed in local coordinates on M and N as follows

    i(dφ)=(φajxi)dxjua+φaj(idxj)ua+φajdxjiua
    =φaijMΓkijφak+NΓbbγφγj. (1.3)

    The tension field is the trace of (1.3),

    τ(φ)a=gij(dφ)aij=Δφa+NΓabγφbiφγjgij. (1.4)

    Thus (1.4) is a semilinear, elliptic, second-order system. If N is the space R, a harmonic map is called a harmonic function [4,7,8].


    2. Energy Functional and Critical Point

    Now let us extend this idea to another object which may be defined on a manifold. Let M be a Riemannian manifold and E a Riemannian vector bundle over M, where each fiber carries a positive definite inner product denoted by ,E. Let Ωp(E) be the space of smooth p-forms which have values in E, where it is assumed throughout that p1. For ωΩp(E), define the energy functional

    I(ω,g)=Mω(ei1,,eip),ω(ei1,,eip)Edvm. (2.1)

    where {ei} is an orthonormal basis on M and repeated indices are summed for 1i1,,ipm and m=dimM. With respect to a local coordinate system {xi} on M and local frame {sa} of E, the norm of ω, which is the integrand of (2.1) can be written

    |ω|2=ω(ei1,,eip),ω(ei1,,eip)=gi1j1gipjpωai1ipωbj1jphab. (2.2)

    Suppose M is compact, then vary the integral (2.1) with respect to metric g. If g(u) is a smooth, one-parameter family of metrics such that g(0)=g, then the variation δg=g/u|u=0 is a smooth symmetric tensor on M.

    Theorem 1. For ωΩp(E) and p1,

    dIdu|u=0=MS(ω),gu|u=0dvM, (2.3)

    where S(ω) is the symmetric two-tensor defined by

    S(ω)=12|ω|2gpi2,,ipω(,ei2,,eip),ω(,ei2,,eip)E. (2.4)

    where {ei} is an orthonormal basis on M.

    Proof: Let {xi} be a local coordinate system on M, and {sa} a local frame for E,

    dI(ω)du|u=0=M|ω|2gijδgijdvM+M|ω|2(dvM)gijδgij. (2.5)

    The volume form on M is given by

    dvM=(detg)1/2dx1dxm,

    where detg is the determinant of the metric tensor gij and therefore,

    gijdvM=12(detg)1/2gij(detg)dx1dxm=12gijdvM.

    Differentiating the expression for the metric tensor gijgjk=δik, the following relation holds

    gisjsgij=gisigjsj.

    The first term in (2.5) is

    |ω|2gij=gij(gi1j1gipjpωai1ipωbj1jphab)
    =ps=1gi1j1gisigjsjgipjpωai1isipωbj1jsjphab.

    Therefore,

    |ω|2gijgikgjl=ps=1gi1j1δiskδjslgipjpωai1isipωbj1jsjphab
    =pgi2j2gipjpωaki2ipωblj2jphab.

    Definition 1. Let M be an arbitrary not necessarily compact Riemannian manifold, and let E be a Riemannian vector bundle over M. Let ωΩp(E) and define the stress-energy tensor of the form ω to be the following symmetric 2-tensor

    S(ω)=12|ω|2gpj1,,jpω(,ej2,,ejp),ω(,ej2,,ejp)E (2.6)

    at each point x where there is an orthonormal basis {ei}.

    Let the vector bundle E be endowed with a Riemannian connection denoted by E so

    Xs,tE=EXs,tE+s,EXtE, (2.7)

    Theorem 2. Let ωΩp(E) and S(ω) be the stress-energy tensor associated with ω, then for all xM and each XTxM,

    ÷S(ω)(X)=ω,dωX+pdω,ωX (2.8)

    where the contraction of a p-form with a vector field X is given by

    (ωX)(X1,,Xp1)=ω(X,X1,,Xp1).

    Proof: Let {ei} be an orthonormal basis at x and extend objects to a neighborhood of x. Suppose that eiej=0 at x for all i,j. The tensorial property allows one to evaluate at X=ek without loss of generality. Consequently,

    ÷S(X)=j(ejS)(ej,ek)=jejS(ej,ek)=jej{12|ω|g(ej,ek)pωej,ωek}
    =ekω,ωpjej(ωej),ωekpjωej,ej(ωek). (2.9)

    At x, it is the case that

    dω(ej1,,ejp1)=Eej(ω(ej,ej1,,ejp1)),

    and

    dω(ek,ei1,,eip)=Eek(ω(ei1,,eip))+pk=1(1)kEeik(ω(ek,ei1,,ˆeik,,eip)).

    Solving for Eek and substituting it into the divergence, we get

    ÷S(X)=dωek,ω
    i1,,ippk=1(1)kEeik(ω(ek,ei1,,ˆeik,,eip)),ω(ei1,,eip)E
    +pdω,ωXpjωej,ej(ωek).

    The double sum in this can be simplified to the form

    i1,,ippk=1(1)k+1Eeik(ω(ek,ei1,,ˆeik,,eip)),ω(e1,,eip)
    =pnj=1ej(ωek),ωej.

    The claim follows as a result of these

    ÷S(ω)=ω,dωek+pdω,ωek+pjej(ωek),ωejjωej,ej(ωek)
    =ω,dωek+pdω,,ωek.

    The form ωΩp(E) is called harmonic if and only if dω=dω=0. Substituting these derivatives into the right-hand side of (2.8), we prove the following result.

    Corollary 1. If ωΩp(E) is a harmonic form, then ÷S(ω)=0.

    Clearly, S is symmetric and so given a symmetric, divergence-free 2-tensor S, conservation laws can be formulated by contracting with a Killing vector field. If X is a Killing vector field, then SX is also divergence free. However, the converse of Corollary 1 is not true. If ÷S(ω)=0, it may not be concluded that ω is harmonic. However, when ω is a differential of a submersive mapping, there is equivalence.

    Corollary 2. Let φ:MN be a smooth mapping between Riemannian manifolds and let ω=dφ be the corresponding φ1TN-valued one-form on M. Then for each vector field X,

    ÷S(dφ)(X)=τ(φ),dφ(X). (2.10)

    In (2.10), τ(φ) is called the tension field of φ and is defined as

    τ(φ)=ddφ. (2.11)

    If φ is a submersive almost everywhere, then τ(φ)=0 if and only if ÷S(dφ)=0.

    Proof: Substitute ω=dφ into (2.8) and use the identity d(dφ)=0 to obtain,

    ÷S(dφ)(X)=dφ,ddφX+ddφ,dφX=τ(φ),dφX. (2.12)

    Clearly, if τ(φ)=0, then the right side of (2.12) vanishes, ÷S(dφ)(X)=0, and conversely.

    The form ω is a critical form with respect to variations of the metric, so the conditions under which the stress-energy tensor vanishes should be studied.

    Definition 2. The form ωΩp(E) is conformal if the map XωX is conformal for each xX, or equivalently, ω is conformal if and only if there is a real λ such that

    ωX,ωY=λ2X,Y,X,YTxM. (2.13)

    To obtain an expression for λ2, evaluate S(ω) in (2.13) on the vectors X=Y=eiTxM and carry out the trace on both sides

    mλ2=|ω|2. (2.14)

    If φ:MN is a smooth map between Riemannian manifolds, then dφΩ1(φ1TN) is a conformal form if and only if the map is conformal as well.

    Lemma 1. For ωΩp(E) which is not identically zero, the stress-energy tensor (2.6) vanishes identically if and only if m=2p and ω is conformal.

    Proof: Evaluating S(ω) on the pair (ei,ei) we obtain,

    S(ω)(ei,ei)=12|ω|2pωei,ωei.

    Tracing on both sides gives

    trS(ω)=(m2p)|ω|2.

    This vanishes for ω0 when m=2p. The definition of conformal form (2.13) gives equivalence.


    3. Weizenböck Formulas and Applications to the Stress-Energy Tensor

    Now consider the Laplacian of the tensor S(ω). The Laplacian on p-forms was extended to arbitrary tensors on a Riemannian manifold by Lichnerowicz [5,6]. For a 2-tensor Q, the Laplacian on Q is given by

    ΔQ(X,Y)=tr2Q(X,Y)+S(RicciM(X),Y)+S(X,RicciM(Y))
    2i,jRM(X,ei)Y,ejS(ei,ej). (3.1)

    where {ei} is an orthonormal basis. The curvature tensor on M is represented by RM with sign convention

    RM(X,Y)Z=XYZ+YXZ+[X,Y]Z. (3.2)

    In (3.2), RicciM is the Ricci tensor defined by RicciM(X,Y)=tr(ZRM(X,Z)Y).

    The following theorem given first by Lichnorowicz [5,6] is very useful and is presented without proof.

    Proposition 1. (Lichnerowicz) Let M be a Riemannian manifold with RicciM=0. Let Q be a 2-tensor on M, then the divergence commutes with the Laplacian

    ÷(ΔQ)=Δ(÷Q), (3.3)

    where the right-hand side is now the standard Laplacian on one-forms.

    Define the symmetric 2-tensor L(Q) to be

    L(Q)=Q(RicciM(X),Y)+Q(X,RicciM(Y))2i,jRM(X,ei)Y,ejQ(ei,ej). (3.4)

    Suppose EM is a vector bundle endowed with a metric and Riemannian connection and associated curvature RE. The Ricci operator on a p-form ω written as Riccip=Ricci is defined to be

    (Ricci(ω))(X1,,Xp)=i,k(Rp(ei,Xk)ω)(ei,X1,,ˆXk,,Xp). (3.5)

    In (3.5), Rp is the canonical curvature and is defined to be [4]

    (Rp(X,Y)ω)(X1,,Xp)=RE(X,Y)(ω(X1,,Xp))
    kω(X1,,RM(X,Y)Xk,,Xp). (3.6)

    The following theorem is due to Weizenböck [3].

    Proposition 2. (Weitzenböck) Let EM be a Riemannian vector bundle over a Riemannian manifold M. Then for any ωΩp(E),

    (i)Δω=tr2ω+Ricci(ω).(ii)12Δ|ω|2=Δω,ω|ω|2Ricci(ω),ω. (3.7)

    Proof: Let us prove (ii) by using (i) for tr2ω,

    12Δ|ω|2=12trd|ω|2=trω,ω=trω,ωω,ω
    =Δω,ωRicci(ω),ω|ω|2.

    It is worth noting a few important applications of this proposition. If ωΩ0(E), then (3.7) reduces to

    Δω=trdω,12Δ|ω|2=Δω,ω|ω|2. (3.8)

    If ωΩ1(E) and XΓ(TM), then

    Δω(X)=tr2ω(X)sRE(es,X)ω(es)+sω(RM(es,X)es), (3.9)

    where {es} is an orthonormal basis for TM and

    12Δ|ω|2=Δω,ω|ω|2+i,jRE(ei,ej)ω(ei),ω(ej)kω(Ricci(ek)),ω(ek). (3.10)

    Theorem 3. Let X,YΓ(TM), it holds that

    (tr2ω,ω)(X,Y)=(tr2ω)X,ωY+ωX,(tr2ω)Y
    +2i(eiω)X,(eiω)Y. (3.11)

    Proof: Differentiating once gives,

    ejω,ω(X,Y)=ejωX,ωY+ωX,ejωY.

    Differentiating a second time gives

    eiejω,ω(X,Y)=eiejωX,ωY+ejωX,ejωY+eiωX,ejωY
    +ωX,eiejωY.

    Tracing on both sides of this equation yields (3.11).

    Theorem 4. Let ωΩp(E) be a vector bundle valued p-form with stress-energy tensor S(ω). The Laplacian of S(ω) can be written in the following way,

    ΔS(ω)=2S(Δω,ω)2S(jejω,ejω)2S(Ricci(ω),ω)+L(S(ω)). (3.12)

    where RicciM is the Ricci tensor and the symmetric 2-tensor L(S) was introduced in (3.4). Substituting Weitzenböck formula (3.7) (i) into (3.11), it follows that

    tr2ω,ω(X,Y)=(ΔωRicci(ω))X,ωY+ωX,(ΔωRicci(ω))Y
    2ieiωX,eiωY
    =ΔωX,ωYRicci(ω)X,ωY+ωX,ΔωYωX,Ricci(ω)Y
    2ieiωX,eiωY. (3.13)

    Substituting (3) into (3.12), it follows by symmetry and linearity that,

    ΔS(ω)(X,Y)=2S(ΔωX,ωY)2S(ieiωX,eiωY)
    2S(Ricci(ω)X,Y)+L(S(ω))(X,Y).

    This is the required result.

    For any 2-tensor QΓ(2TM), define the p-th stress-energy tensor associated to Q to be the 2-tensor

    Sp(Q)=12(trgQ)gpsymQ, (3.14)

    where trQ=iQ(ei,ei), and {ei} is orthonormal with respect to the metric g. Moreover, define

    (symQ)(X,Y)=12(Q(X,Y)+Q(Y,X)). (3.15)

    Then symQ in (3.15) is called the symmetrization of Q. For a p-form ωΩp(E), this is simply

    S(ω)=Sp(ω,ω). (3.16)

    The previous result can be written in completely symmetric form in terms of Sp.

    Corollary 3. Let ωΩp(E) be a vector-bundle valued p-form with associated stress-energy tensor S(ω). The Laplacian of S(ω) is given by

    ΔS(ω)=2Sp(Δω,ω)2Sp(ieiωeiω)2Sp(Ricci(ω),ω)+L(S(ω)). (3.17)

    It is also worth writing this out for the special case of a 1-form.

    Corollary 4. Let ωΩ1(E) be a vector bundle valued 1-form, then (3.17) takes the form

    ΔS(ω)=2S(Δω,ω)2S(ieiω,eiω)+2S(iRE(ω),ω)
    2S(ω(RicciM()),ω)+L(S(ω)). (3.18)

    Moreover, for ωΩ1(E), a closed vector bundle valued 1-form,

    ΔS(ω)=2S(Δω,ω)2S(ω)+2S(iRE(ei,)ω(ei),ω())
    2S(ω(RicciM(),ω)+L(S(ω)). (3.19)

    where S(ω) denotes the stress-energy tensor of the TME valued one-form Xω.

    Proof: First (3.18) is a consequence of (3.17). For (3.19), since ω is closed, dω=0, and it follows that (eiω)(X)=(Xω)(ei), consequently ieiω,eiω=ω,ω.

    Theorem 5. Let φ:(M,g)(N,h) be a smooth map between Riemannian manifolds and let dφΩ1(φ1TN) be its exterior derivative.

    ΔS(dφ)=2S(τ(φ),dφ)2S(dφ)2S(dφ(RicciM()),dφ())
    2S(iRN(dφ(ei),dφ())dφ(ei),dφ())+L(S(dφ)), (3.20)

    where τ(φ)=ddφ denotes the tension field of the map φ.

    Proof: For the one-form ω, substitute ω=dφ into (3.19). Since it follows that Δdφ=(dd+dd)dφ=dτ(φ)=τ(φ), and RNdφ is the induced curvature in the bundle E=φ1TN.

    Theorem 6. Let φ:(M2,g)(N2,h) be a harmonic map between surfaces where M2 has Gaussian curvature cM, then the folowing holds,

    ΔS(dφ)=2S(dφ)+2cMS(dφ). (3.21)

    In particular, if M has constant curvature, then it follows that ÷S(dφ)=0.

    Proof: If φ is harmonic, then τ(φ)=0, so the first term in (3.20) vanishes. Suppose σ(Q)=12(trQ)gsymQ is a stress-energy tensor associated to the 2-tensor Q on M. Suppose z is an isothermal coordinate, then σ(Q) can be expressed in diagonal form σ(Q)=adzdz+ˉadˉzdˉz for some function a=a(z,ˉz). If σ(Q) is divergence free, then aˉz vanishes, so σ(Q)=adzdz defines a quadratic differential form on M2. When φ is a smooth map, then locally φh can be diagonalized to the form φh=λ1ω1ω1+λ2ω2ω2, where {ω1,ω2} is an orthonormal basis of 1-forms. If M and N have Gaussian curvatures cM and cN, then letting {e1,e2} be the basis dual to {ω1,ω2}, summing on i=1,2 we have RN(dφ(ei),dφ())dφ(ei),dφ()=cNλ1λ2g, and consequently σ(RN(dφ(ei),dφ())dφ(ei),dφ())=0. Moreover, L(S(φ))=4cMS(dφ) and σ(dφ(RicciM()),dφ())=cMS(dφ), so substituting these calculations into (3.20), equation (3.21) is the result.

    Recall that if ωΩp(E) is harmonic, then ÷S(ω)=0. If we use ω=dφΩ1(E), then ÷S(dφ)=0. Take the divergence of both sides of (3.21) and note that cM is constant, then the result follows since divergence and Laplacian commute be Proposition 1.

    Now τ(φ) defines an energy functional which is called the biharmonic energy functional

    E2(φ)=12M|τ(φ)|2dvM, (3.22)

    and has the associated stress-energy tensor S2(dφ),

    S2(dφ)=12|τ(φ)|2g+2S(τ(φ),dφ). (3.23)

    Theorem 7. Let φ:(M,g)(N,h) be a smooth map between Riemannian manifolds and let dφΩ1(φ1TN) be the derivative.

    (i) The Laplacian of S(dφ) is given as

    ΔS(dφ)=12|τ(φ)|2gS2(dφ)2S(dφ)2S(dφ(RicciM(),dφ())
    +2S(iRN(dφ(ei),dφ())dφ(ei),dφ()TN)+L(S(dφ)). (3.24)

    (ii) If φ:MN is an isometric immersion,

    12|τ(φ)|2gS2(dφ)2S(dφ)2S(RicciM)
    +2S(iRN(dφ(ei),dφ())dφ(ei),dφ())=0. (3.25)

    (iii) If N is a space form of constant curvature KN=1,0,+1,

    12|τ(φ)|2gS2(dφ)2S(dγ)2S(RicciM)+2(m1)KNS(dφ)=0. (3.26)

    Proof: (i) Let ω=dφ then solve for the second term in (3.23) and substitute it into the previous result (3.21) which gives (3.24). (ii) If φ:(M,g)(N,h) is an isometric immersion, then

    S(dφ)=12(m2)g, (3.27)

    hence the left-hand side of (3.24) as well as L(Sdφ)) both vanish. (iii) Finally, if we map into a space form, the second fundamental form dφ can be identified with the derivative of the associated Gauss map Γ and (3.26) follows.

    As an application of this theorem, suppose φ is a minimal immersion of a surface into Euclidean space, then τ(φ)=0, S2(dφ)=0 and σ(RicciM)=0. Then equation (3.26) implies that S(dγ)=0.

    If the divergence of both sides of (3.26) is worked out, since ÷S(RicciM)=0 as a consequence of the Bianchi identity, (3.26) implies that if φ:MN is an isometric immersion into a space form, it follows that

    12d|τ(φ)|2÷S2(dφ)2÷S(dγ)=0. (3.28)

    4. Further Results Including an Integral Theorem

    Let us examine the influence of conformal vector fields to finish [3]. A monotonicity formula describes the growth properties of extremals of the functional (2.1) and may be used to establish their regularity in appropriate situations.

    Let (M,g) be a Riemannian manifold of dimension m. A vector field X on M is called conformal if it satisfies

    LXg=ξg (4.1)

    for some function ξ:MR and L is the Lie derivative in the direction of X. Equivalently, X is conformal if and only if

    symg(i)=1m(÷X)g, (4.2)

    in which case, we have ξ=(2/m)÷X.

    Theorem 8. Let T be a symmetric 2-tensor, and let X be a conformal vector field, then

    ÷(TX)=(÷T)(X)+1m(÷X)trT. (4.3)

    Proof: Let {ei}m1 be a locally defined frame field. If X is a conformal vector field, the divergence can be calculated as follows,

    ÷(TX)=ei(TX)(ei)=(eiT)(X,ei)+T(eiX,ei)
    =(÷T)(X)+T(eiX,ejej,ei).

    Since T is a symmetric 2-tensor, the definition of conformal implies that

    ÷(TX)=(÷T)(X)+12(eiX,ej+ejX,ei)T(ei,ej)
    =(÷T)(X)+1m÷Xg(ei,ej)T(ei,ej)=(÷T)+1m÷XtrT,

    since we have trT=g(ei,ej)T(ei,ej).

    Integrate (4.3) over a compact region, U, to obtain the monotonicity formula. The divergence theorem permits us to write

    U÷(TX)dvM=UT(X,n)daM. (4.4)

    where daM is the volume form on U.

    Theorem 9. Let (M,g) be an oriented Riemannian manifold and let X be a conformal vector field on M. Suppose that UM is a compact region with a smooth boundary U. Then for any symmetric 2-tensor T on M,

    UT(X,n)daM=U(÷T)(X)dvM+1mM÷XtrTdvM, (4.5)

    where dvM is the volume element on M and daM is the volume element on U, n the outward pointing normal on U.

    This is a remarkable theorem as there are numerous applications of it, but only one will be presented here. Let φ:Bm(N,h) be a harmonic map from the Euclidean m-ball of radius R and S=e(φ)gφh the corresponding stress-energy tensor. Clearly, ÷S=0 and trS=etrgtrφh=em2e=(m2)e(φ). Take X to be the conformal vector field X=rr where r=|x|, xRm. By direct calculation, ÷X=m, hence substituting these facts into (4.5), the following result holds for r<R,

    BmS(rr,r)da=Bm(m2)e(φ)dvM. (4.6)

    Substituting for S yields

    (m2)Bme(φ)dvM=Bmr{e(φ)h(φr,φr)}daM. (4.7)

    The following conclusion can be drawn from (4.7). If m>2 and φ|Bmr=c for some r<R and constant c, then φ must be constant, that is, for e(φ)|Bm=12h(rφ,rφ) gives the upper bound

    Bme(φ)dvM0,

    which in turn implies that e(φ)0.


    Conflict of Interest

    The author declares no conflicts of interest in this paper.


    [1] P. Baird, Stress-energy tensors and the Lichnerowicz Laplacian, J. of Geometry and Physics, 58(2008), 1329-1342.
    [2] P. Blanchard and E. Brüning, Variational Methods in Mathematical Physics, Springer-Verlag, Berlin-Heidelberg, 1992.
    [3] S. S. Chern, Minimal surfaces in Euclidean Space of N dimensions: Symposium in Honor of Marston Morse, Princeton Univ. Press, 1965,187-198.
    [4] J Eells and L. Lemaine, Selected Topics in Harmonic Maps: CBMS Regional Conference Series in Mathematics, American Mathematical Society, Providence, RI, 50 (1983).
    [5] A. Lichnerowicz, Géometrie des Groups de Transformations, Dunod, Paris, 1958.
    [6] A. Lichnerowicz, Propagateurs et commutateurs en relativité générale, Publ. Math. Inst. Hautes Ëtudes Sci., 10 (1961), 293-344.
    [7] P. Peterson, Riemannian Geometry, Springer-Verlag, NY, 1998.
    [8] R. Schoen, The existence of weak solutions with prescribed singular behavior for a conformally invariant scalar equation, Comm. Pure and Applied Math. XLI, 317-392,1988.
  • Reader Comments
  • © 2017 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(4588) PDF downloads(1062) Cited by(0)

Article outline

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog