Movatterモバイル変換


[0]ホーム

URL:


Jump to content
WikipediaThe Free Encyclopedia
Search

Hodge star operator

From Wikipedia, the free encyclopedia
Exterior algebraic map taking tensors from p forms to n-p forms

Inmathematics, theHodge star operator orHodge star is alinear map defined on theexterior algebra of afinite-dimensionalorientedvector space endowed with anondegeneratesymmetric bilinear form. Applying the operator to an element of the algebra produces theHodge dual of the element. This map was introduced byW. V. D. Hodge.

For example, in an oriented 3-dimensional Euclidean space, an oriented plane can be represented by theexterior product of two basis vectors, and its Hodge dual is thenormal vector given by theircross product; conversely, any vector is dual to the oriented plane perpendicular to it, endowed with a suitable bivector. Generalizing this to ann-dimensional vector space, the Hodge star is a one-to-one mapping ofk-vectors to(n – k)-vectors; the dimensions of these spaces are thebinomial coefficients(nk)=(nnk){\displaystyle {\tbinom {n}{k}}={\tbinom {n}{n-k}}}.

Thenaturalness of the star operator means it can play a role in differential geometry when applied to the cotangentbundle of apseudo-Riemannian manifold, and hence todifferentialk-forms. This allows the definition of the codifferential as the Hodge adjoint of theexterior derivative, leading to theLaplace–de Rham operator. This generalizes the case of 3-dimensional Euclidean space, in whichdivergence of a vector field may be realized as the codifferential opposite to thegradient operator, and theLaplace operator on a function is the divergence of its gradient. An important application is theHodge decomposition of differential forms on aclosed Riemannian manifold.

Formal definition fork-vectors

[edit]

LetV be ann-dimensionalorientedvector space with a nondegenerate symmetric bilinear form,{\displaystyle \langle \cdot ,\cdot \rangle }, referred to here as a scalar product. (In more general contexts such as pseudo-Riemannian manifolds andMinkowski space, the bilinear form may not be positive-definite.) This induces ascalar product onk-vectorsα,βkV{\textstyle \alpha ,\beta \in \bigwedge ^{\!k}V}, for0kn{\displaystyle 0\leq k\leq n}, by defining it on simplek-vectorsα=α1αk{\displaystyle \alpha =\alpha _{1}\wedge \cdots \wedge \alpha _{k}} andβ=β1βk{\displaystyle \beta =\beta _{1}\wedge \cdots \wedge \beta _{k}} to equal theGram determinant[1]: 14 

α,β=det(αi,βji,j=1k){\displaystyle \langle \alpha ,\beta \rangle =\det \left(\left\langle \alpha _{i},\beta _{j}\right\rangle _{i,j=1}^{k}\right)}

extended tokV{\textstyle \bigwedge ^{\!k}V} through linearity.

The unitn-vectorωnV{\displaystyle \omega \in {\textstyle \bigwedge }^{\!n}V} is defined in terms of an orientedorthonormal basis{e1,,en}{\displaystyle \{e_{1},\ldots ,e_{n}\}} ofV as:

ω:=e1en.{\displaystyle \omega :=e_{1}\wedge \cdots \wedge e_{n}.}

(Note: In the general pseudo-Riemannian case, orthonormality meansei,ej{δij,δij}{\displaystyle \langle e_{i},e_{j}\rangle \in \{\delta _{ij},-\delta _{ij}\}} for all pairs of basis vectors.)TheHodge star operator is a linear operator on theexterior algebra ofV, mappingk-vectors to (nk)-vectors, for0kn{\displaystyle 0\leq k\leq n}. It has the following property, which defines it completely:[1]: 15 

α(β)=α,βω{\displaystyle \alpha \wedge ({\star }\beta )=\langle \alpha ,\beta \rangle \,\omega } for allk-vectorsα,βkV.{\displaystyle \alpha ,\beta \in {\textstyle \bigwedge }^{\!k}V.}

Dually, in the spacenV{\displaystyle {\textstyle \bigwedge }^{\!n}V^{*}} ofn-forms (alternatingn-multilinear functions onVn{\displaystyle V^{n}}), the dual toω{\displaystyle \omega } is thevolume formdet{\displaystyle \det }, the function whose value onv1vn{\displaystyle v_{1}\wedge \cdots \wedge v_{n}} is thedeterminant of then×n{\displaystyle n\times n} matrix assembled from the column vectors ofvj{\displaystyle v_{j}} inei{\displaystyle e_{i}}-coordinates. Applyingdet{\displaystyle \det } to the above equation, we obtain the dual definition:

det(αβ)=α,β{\displaystyle \det(\alpha \wedge {\star }\beta )=\langle \alpha ,\beta \rangle } for allk-vectorsα,βkV.{\displaystyle \alpha ,\beta \in {\textstyle \bigwedge }^{\!k}V.}

Equivalently, takingα=α1αk{\displaystyle \alpha =\alpha _{1}\wedge \cdots \wedge \alpha _{k}},β=β1βk{\displaystyle \beta =\beta _{1}\wedge \cdots \wedge \beta _{k}}, andβ=β1βnk{\displaystyle {\star }\beta =\beta _{1}^{\star }\wedge \cdots \wedge \beta _{n-k}^{\star }}:

det(α1αkβ1βnk) = det(αi,βj).{\displaystyle \det \left(\alpha _{1}\wedge \cdots \wedge \alpha _{k}\wedge \beta _{1}^{\star }\wedge \cdots \wedge \beta _{n-k}^{\star }\right)\ =\ \det \left(\langle \alpha _{i},\beta _{j}\rangle \right).}

This means that, writing an orthonormal basis ofk-vectors aseI = ei1eik{\displaystyle e_{I}\ =\ e_{i_{1}}\wedge \cdots \wedge e_{i_{k}}} over all subsetsI={i1<<ik}{\displaystyle I=\{i_{1}<\cdots <i_{k}\}} of[n]={1,,n}{\displaystyle [n]=\{1,\ldots ,n\}}, the Hodge dual is the (n – k)-vector corresponding to the complementary setI¯=[n]I={i¯1<<i¯nk}{\displaystyle {\bar {I}}=[n]\smallsetminus I=\left\{{\bar {i}}_{1}<\cdots <{\bar {i}}_{n-k}\right\}}:

eI=steI¯,{\displaystyle {\star }e_{I}=s\cdot t\cdot e_{\bar {I}},}

wheres{1,1}{\displaystyle s\in \{1,-1\}} is thesign of the permutationi1iki¯1i¯nk{\displaystyle i_{1}\cdots i_{k}{\bar {i}}_{1}\cdots {\bar {i}}_{n-k}}andt{1,1}{\displaystyle t\in \{1,-1\}} is the productei1,ei1eik,eik{\displaystyle \langle e_{i_{1}},e_{i_{1}}\rangle \cdots \langle e_{i_{k}},e_{i_{k}}\rangle }. In the Riemannian case,t=1{\displaystyle t=1}.

Since Hodge star takes an orthonormal basis to an orthonormal basis, it is anisometry on the exterior algebraV{\textstyle \bigwedge V}.

Geometric explanation

[edit]

The Hodge star is motivated by the correspondence between a subspaceW ofV and its orthogonal subspace (with respect to the scalar product), where each space is endowed with anorientation and a numerical scaling factor. Specifically, a non-zero decomposablek-vectorw1wkkV{\displaystyle w_{1}\wedge \cdots \wedge w_{k}\in \textstyle \bigwedge ^{\!k}V} corresponds by thePlücker embedding to the subspaceW{\displaystyle W} with oriented basisw1,,wk{\displaystyle w_{1},\ldots ,w_{k}}, endowed with a scaling factor equal to thek-dimensional volume of the parallelepiped spanned by this basis (equal to theGramian, the determinant of the matrix of scalar productswi,wj{\displaystyle \langle w_{i},w_{j}\rangle }). The Hodge star acting on a decomposable vector can be written as a decomposable (nk)-vector:

(w1wk)=u1unk,{\displaystyle {\star }(w_{1}\wedge \cdots \wedge w_{k})\,=\,u_{1}\wedge \cdots \wedge u_{n-k},}

whereu1,,unk{\displaystyle u_{1},\ldots ,u_{n-k}} form an oriented basis of theorthogonal spaceU=W{\displaystyle U=W^{\perp }\!}. Furthermore, the (nk)-volume of theui{\displaystyle u_{i}}-parallelepiped must equal thek-volume of thewi{\displaystyle w_{i}}-parallelepiped, andw1,,wk,u1,,unk{\displaystyle w_{1},\ldots ,w_{k},u_{1},\ldots ,u_{n-k}} must form an oriented basis ofV{\displaystyle V}.

A generalk-vector is a linear combination of decomposablek-vectors, and the definition of Hodge star is extended to generalk-vectors by defining it as being linear.

Examples

[edit]

Two dimensions

[edit]

In two dimensions with the normalized Euclidean metric and orientation given by the ordering(x,y), the Hodge star onk-forms is given by1=dxdydx=dydy=dx(dxdy)=1.{\displaystyle {\begin{aligned}{\star }\,1&=dx\wedge dy\\{\star }\,dx&=dy\\{\star }\,dy&=-dx\\{\star }(dx\wedge dy)&=1.\end{aligned}}}

Three dimensions

[edit]

A common example of the Hodge star operator is the casen = 3, when it can be taken as the correspondence between vectors and bivectors. Specifically, forEuclideanR3 with the basisdx,dy,dz{\displaystyle dx,dy,dz} ofone-forms often used invector calculus, one finds thatdx=dydzdy=dzdxdz=dxdy.{\displaystyle {\begin{aligned}{\star }\,dx&=dy\wedge dz\\{\star }\,dy&=dz\wedge dx\\{\star }\,dz&=dx\wedge dy.\end{aligned}}}

The Hodge star relates the exterior and cross product in three dimensions:[2](uv)=u×v(u×v)=uv.{\displaystyle {\star }(\mathbf {u} \wedge \mathbf {v} )=\mathbf {u} \times \mathbf {v} \qquad {\star }(\mathbf {u} \times \mathbf {v} )=\mathbf {u} \wedge \mathbf {v} .} Applied to three dimensions, the Hodge star provides anisomorphism betweenaxial vectors andbivectors, so each axial vectora is associated with a bivectorA and vice versa, that is:[2]A=a,  a=A{\displaystyle \mathbf {A} ={\star }\mathbf {a} ,\ \ \mathbf {a} ={\star }\mathbf {A} }.

The Hodge star can also be interpreted as a form of the geometric correspondence between anaxis of rotation and aninfinitesimal rotation (see also:3D rotation group#Lie algebra) around the axis, with speed equal to the length of the axis of rotation. A scalar product on a vector spaceV{\displaystyle V} gives anisomorphismVV{\displaystyle V\cong V^{*}\!} identifyingV{\displaystyle V} with itsdual space, and the vector spaceL(V,V){\displaystyle L(V,V)} is naturally isomorphic to thetensor productVVVV{\displaystyle V^{*}\!\!\otimes V\cong V\otimes V}. Thus forV=R3{\displaystyle V=\mathbb {R} ^{3}}, the star mapping:V2VVV{\textstyle \textstyle {\star }:V\to \bigwedge ^{\!2}\!V\subset V\otimes V} takes each vectorv{\displaystyle \mathbf {v} } to a bivectorvVV{\displaystyle {\star }\mathbf {v} \in V\otimes V}, which corresponds to a linear operatorLv:VV{\displaystyle L_{\mathbf {v} }:V\to V}. Specifically,Lv{\displaystyle L_{\mathbf {v} }} is askew-symmetric operator, which corresponds to an infinitesimal rotation: that is, the macroscopic rotations around the axisv{\displaystyle \mathbb {v} } are given by thematrix exponentialexp(tLv){\displaystyle \exp(tL_{\mathbf {v} })}. With respect to the basisdx,dy,dz{\displaystyle dx,dy,dz} ofR3{\displaystyle \mathbb {R} ^{3}}, the tensordxdy{\displaystyle dx\otimes dy} corresponds to a coordinate matrix with 1 in thedx{\displaystyle dx} row anddy{\displaystyle dy} column, etc., and the wedgedxdy=dxdydydx{\displaystyle dx\wedge dy\,=\,dx\otimes dy-dy\otimes dx} is the skew-symmetric matrix[010100000]{\displaystyle \scriptscriptstyle \left[{\begin{array}{rrr}\,0\!\!&\!\!1&\!\!\!\!0\!\!\!\!\!\!\\[-.5em]\,\!-1\!\!&\!\!0\!\!&\!\!\!\!0\!\!\!\!\!\!\\[-.5em]\,0\!\!&\!\!0\!\!&\!\!\!\!0\!\!\!\!\!\!\end{array}}\!\!\!\right]}, etc. That is, we may interpret the star operator as:v=adx+bdy+cdzv  Lv =[0cbc0aba0].{\displaystyle \mathbf {v} =a\,dx+b\,dy+c\,dz\quad \longrightarrow \quad {\star }{\mathbf {v} }\ \cong \ L_{\mathbf {v} }\ =\left[{\begin{array}{rrr}0&c&-b\\-c&0&a\\b&-a&0\end{array}}\right].} Under this correspondence, cross product of vectors corresponds to the commutatorLie bracket of linear operators:Lu×v=LvLuLuLv=[Lu,Lv]{\displaystyle L_{\mathbf {u} \times \mathbf {v} }=L_{\mathbf {v} }L_{\mathbf {u} }-L_{\mathbf {u} }L_{\mathbf {v} }=-\left[L_{\mathbf {u} },L_{\mathbf {v} }\right]}.

Four dimensions

[edit]

In casen=4{\displaystyle n=4}, the Hodge star acts as anendomorphism of the second exterior power (i.e. it maps 2-forms to 2-forms, since4 − 2 = 2). If the signature of themetric tensor is all positive, i.e. on aRiemannian manifold, then the Hodge star is aninvolution. If the signature is mixed, i.e.,pseudo-Riemannian, then applying the operator twice will return the argument up to a sign – see§ Duality below. This particular endomorphism property of 2-forms in four dimensions makesself-dual and anti-self-dual two-forms natural geometric objects to study. That is, one can describe the space of 2-forms in four dimensions with a basis that "diagonalizes" the Hodge star operator with eigenvalues±1{\displaystyle \pm 1} (or±i{\displaystyle \pm i}, depending on the signature).

For concreteness, we discuss the Hodge star operator in Minkowski spacetime wheren=4{\displaystyle n=4} with metric signature(− + + +) and coordinates(t,x,y,z){\displaystyle (t,x,y,z)}. Thevolume form is oriented asε0123=1{\displaystyle \varepsilon _{0123}=1}. Forone-forms,dt=dxdydz,dx=dtdydz,dy=dtdzdx,dz=dtdxdy,{\displaystyle {\begin{aligned}{\star }dt&=-dx\wedge dy\wedge dz\,,\\{\star }dx&=-dt\wedge dy\wedge dz\,,\\{\star }dy&=-dt\wedge dz\wedge dx\,,\\{\star }dz&=-dt\wedge dx\wedge dy\,,\end{aligned}}}while for2-forms,(dtdx)=dydz,(dtdy)=dzdx,(dtdz)=dxdy,(dxdy)=dtdz,(dzdx)=dtdy,(dydz)=dtdx.{\displaystyle {\begin{aligned}{\star }(dt\wedge dx)&=-dy\wedge dz\,,\\{\star }(dt\wedge dy)&=-dz\wedge dx\,,\\{\star }(dt\wedge dz)&=-dx\wedge dy\,,\\{\star }(dx\wedge dy)&=dt\wedge dz\,,\\{\star }(dz\wedge dx)&=dt\wedge dy\,,\\{\star }(dy\wedge dz)&=dt\wedge dx\,.\end{aligned}}}

These are summarized in the index notation as(dxμ)=ημλελνρσ13!dxνdxρdxσ,(dxμdxν)=ημκηνλεκλρσ12!dxρdxσ.{\displaystyle {\begin{aligned}{\star }(dx^{\mu })&=\eta ^{\mu \lambda }\varepsilon _{\lambda \nu \rho \sigma }{\frac {1}{3!}}dx^{\nu }\wedge dx^{\rho }\wedge dx^{\sigma }\,,\\{\star }(dx^{\mu }\wedge dx^{\nu })&=\eta ^{\mu \kappa }\eta ^{\nu \lambda }\varepsilon _{\kappa \lambda \rho \sigma }{\frac {1}{2!}}dx^{\rho }\wedge dx^{\sigma }\,.\end{aligned}}}

Hodge dual of three- and four-forms can be easily deduced from the fact that, in the Lorentzian signature,2=1{\displaystyle {\star }^{2}=1} for odd-rank forms and2=1{\displaystyle {\star }^{2}=-1} for even-rank forms. An easy rule to remember for these Hodge operations is that given a formα{\displaystyle \alpha }, its Hodge dualα{\displaystyle {\star }\alpha } may be obtained by writing the components not involved inα{\displaystyle \alpha } in an order such thatα(α)=dtdxdydz{\displaystyle \alpha \wedge ({\star }\alpha )=dt\wedge dx\wedge dy\wedge dz}.[verification needed] An extra minus sign will enter only ifα{\displaystyle \alpha } containsdt{\displaystyle dt}. (For(+ − − −), one puts in a minus sign only ifα{\displaystyle \alpha } involves an odd number of the space-associated formsdx{\displaystyle dx},dy{\displaystyle dy} anddz{\displaystyle dz}.)

Note that the combinations(dxμdxν)±:=12(dxμdxνi(dxμdxν)){\displaystyle (dx^{\mu }\wedge dx^{\nu })^{\pm }:={\frac {1}{2}}{\big (}dx^{\mu }\wedge dx^{\nu }\mp i{\star }(dx^{\mu }\wedge dx^{\nu }){\big )}}take±i{\displaystyle \pm i} as the eigenvalue for Hodge star operator, i.e.,(dxμdxν)±=±i(dxμdxν)±,{\displaystyle {\star }(dx^{\mu }\wedge dx^{\nu })^{\pm }=\pm i(dx^{\mu }\wedge dx^{\nu })^{\pm },}and hence deserve the name self-dual and anti-self-dual two-forms. Understanding the geometry, or kinematics, of Minkowski spacetime in self-dual and anti-self-dual sectors turns out to be insightful in bothmathematical andphysical perspectives, making contacts to the use of thetwo-spinor language in modern physics such asspinor-helicity formalism ortwistor theory.

Conformal invariance

[edit]

The Hodge star is conformally invariant onn-forms on a2n-dimensional vector spaceV{\displaystyle V}, i.e. ifg{\displaystyle g} is a metric onV{\displaystyle V} andλ>0{\displaystyle \lambda >0}, then the induced Hodge starsg,λg:ΛnVΛnV{\displaystyle {\star }_{g},{\star }_{\lambda g}:\Lambda ^{n}V\to \Lambda ^{n}V}are the same.

Example: Derivatives in three dimensions

[edit]

The combination of the{\displaystyle {\star }} operator and theexterior derivatived generates the classical operatorsgrad,curl, anddiv onvector fields in three-dimensional Euclidean space. This works out as follows:d takes a 0-form (a function) to a 1-form, a 1-form to a 2-form, and a 2-form to a 3-form (and takes a 3-form to zero). For a 0-formf=f(x,y,z){\displaystyle f=f(x,y,z)}, the first case written out in components gives:df=fxdx+fydy+fzdz.{\displaystyle df={\frac {\partial f}{\partial x}}\,dx+{\frac {\partial f}{\partial y}}\,dy+{\frac {\partial f}{\partial z}}\,dz.}

The scalar productidentifies 1-forms with vector fields asdx(1,0,0){\displaystyle dx\mapsto (1,0,0)}, etc., so thatdf{\displaystyle df} becomesgradf=(fx,fy,fz){\textstyle \operatorname {grad} f=\left({\frac {\partial f}{\partial x}},{\frac {\partial f}{\partial y}},{\frac {\partial f}{\partial z}}\right)}.

In the second case, a vector fieldF=(A,B,C){\displaystyle \mathbf {F} =(A,B,C)} corresponds to the 1-formφ=Adx+Bdy+Cdz{\displaystyle \varphi =A\,dx+B\,dy+C\,dz}, which has exterior derivative:dφ=(CyBz)dydz+(CxAz)dxdz+(BxAy)dxdy.{\displaystyle d\varphi =\left({\frac {\partial C}{\partial y}}-{\frac {\partial B}{\partial z}}\right)dy\wedge dz+\left({\frac {\partial C}{\partial x}}-{\frac {\partial A}{\partial z}}\right)dx\wedge dz+\left({\partial B \over \partial x}-{\frac {\partial A}{\partial y}}\right)dx\wedge dy.}

Applying the Hodge star gives the 1-form:dφ=(CyBz)dx(CxAz)dy+(BxAy)dz,{\displaystyle {\star }d\varphi =\left({\partial C \over \partial y}-{\partial B \over \partial z}\right)\,dx-\left({\partial C \over \partial x}-{\partial A \over \partial z}\right)\,dy+\left({\partial B \over \partial x}-{\partial A \over \partial y}\right)\,dz,}which becomes the vector fieldcurlF=(CyBz,Cx+Az,BxAy){\textstyle \operatorname {curl} \mathbf {F} =\left({\frac {\partial C}{\partial y}}-{\frac {\partial B}{\partial z}},\,-{\frac {\partial C}{\partial x}}+{\frac {\partial A}{\partial z}},\,{\frac {\partial B}{\partial x}}-{\frac {\partial A}{\partial y}}\right)}.

In the third case,F=(A,B,C){\displaystyle \mathbf {F} =(A,B,C)} again corresponds toφ=Adx+Bdy+Cdz{\displaystyle \varphi =A\,dx+B\,dy+C\,dz}. Applying Hodge star, exterior derivative, and Hodge star again:φ=AdydzBdxdz+Cdxdy,dφ=(Ax+By+Cz)dxdydz,dφ=Ax+By+Cz=divF.{\displaystyle {\begin{aligned}{\star }\varphi &=A\,dy\wedge dz-B\,dx\wedge dz+C\,dx\wedge dy,\\d{\star \varphi }&=\left({\frac {\partial A}{\partial x}}+{\frac {\partial B}{\partial y}}+{\frac {\partial C}{\partial z}}\right)dx\wedge dy\wedge dz,\\{\star }d{\star }\varphi &={\frac {\partial A}{\partial x}}+{\frac {\partial B}{\partial y}}+{\frac {\partial C}{\partial z}}=\operatorname {div} \mathbf {F} .\end{aligned}}}

One advantage of this expression is that the identityd2 = 0, which is true in all cases, has as special cases two other identities: (1)curl gradf = 0, and (2)div curlF = 0. In particular,Maxwell's equations take on a particularly simple and elegant form, when expressed in terms of the exterior derivative and the Hodge star. The expressiond{\displaystyle {\star }d{\star }} (multiplied by an appropriate power of −1) is called thecodifferential; it is defined in full generality, for any dimension, further in the article below.

One can also obtain theLaplacianΔf = div grad f in terms of the above operations:Δf=ddf=2fx2+2fy2+2fz2.{\displaystyle \Delta f={\star }d{\star }df={\frac {\partial ^{2}f}{\partial x^{2}}}+{\frac {\partial ^{2}f}{\partial y^{2}}}+{\frac {\partial ^{2}f}{\partial z^{2}}}.}

The Laplacian can also be seen as a special case of the more generalLaplace–deRham operatorΔ=dδ+δd{\displaystyle \Delta =d\delta +\delta d} where in three dimensions,δ=(1)kd{\displaystyle \delta =(-1)^{k}{\star }d{\star }} is the codifferential fork{\displaystyle k}-forms. Any functionf{\displaystyle f} is a 0-form, andδf=0{\displaystyle \delta f=0} and so this reduces to the ordinary Laplacian. For the 1-formφ{\displaystyle \varphi } above, the codifferential isδ=d{\displaystyle \delta =-{\star }d{\star }} and after some straightforward calculations one obtains the Laplacian acting onφ{\displaystyle \varphi }.

Duality

[edit]

Applying the Hodge star twice leaves ak-vector unchangedup to a sign: forηkV{\displaystyle \eta \in {\textstyle \bigwedge }^{k}V} in ann-dimensional spaceV, one has

η=(1)k(nk)sη,{\displaystyle {\star }{\star }\eta =(-1)^{k(n-k)}s\,\eta ,}

wheres is the parity of thesignature of the scalar product onV, that is, the sign of thedeterminant of the matrix of the scalar product with respect to any basis. For example, ifn = 4 and the signature of the scalar product is either(+ − − −) or(− + + +) thens = −1. For Riemannian manifolds (including Euclidean spaces), we always haves = 1.

The above identity implies that the inverse of{\displaystyle {\star }} can be given as

1: kVnkVη(1)k(nk)sη{\displaystyle {\begin{aligned}{\star }^{-1}:~{\textstyle \bigwedge }^{\!k}V&\to {\textstyle \bigwedge }^{\!n-k}V\\\eta &\mapsto (-1)^{k(n-k)}\!s\,{\star }\eta \end{aligned}}}

Ifn is odd thenk(nk) is even for anyk, whereas ifn is even thenk(nk) has the parity ofk. Therefore:

1={sn is odd(1)ksn is even{\displaystyle {\star }^{-1}={\begin{cases}s\,{\star }&n{\text{ is odd}}\\(-1)^{k}s\,{\star }&n{\text{ is even}}\end{cases}}}

wherek is the degree of the element operated on.

On manifolds

[edit]

For ann-dimensional orientedpseudo-Riemannian manifoldM, we apply the construction above to eachcotangent spaceTpM{\displaystyle {\text{T}}_{p}^{*}M} and its exterior powerskTpM{\textstyle \bigwedge ^{k}{\text{T}}_{p}^{*}M}, and hence to the differentialk-formsζΩk(M)=Γ(kTM){\textstyle \zeta \in \Omega ^{k}(M)=\Gamma \left(\bigwedge ^{k}{\text{T}}^{*}\!M\right)}, theglobal sections of thebundlekTMM{\textstyle \bigwedge ^{k}\mathrm {T} ^{*}\!M\to M}. The Riemannian metric induces a scalar product onkTpM{\textstyle \bigwedge ^{k}{\text{T}}_{p}^{*}M} at each pointpM{\displaystyle p\in M}. We define theHodge dual of ak-formζ{\displaystyle \zeta }, definingζ{\displaystyle {\star }\zeta } as the unique (nk)-form satisfyingηζ = η,ζω{\displaystyle \eta \wedge {\star }\zeta \ =\ \langle \eta ,\zeta \rangle \,\omega }for everyk-formη{\displaystyle \eta }, whereη,ζ{\displaystyle \langle \eta ,\zeta \rangle } is a real-valued function onM{\displaystyle M}, and thevolume formω{\displaystyle \omega } is induced by the pseudo-Riemannian metric. Integrating this equation overM{\displaystyle M}, the right side becomes theL2{\displaystyle L^{2}} (square-integrable)scalar product onk-forms, and we obtain:Mηζ = Mη,ζ ω.{\displaystyle \int _{M}\eta \wedge {\star }\zeta \ =\ \int _{M}\langle \eta ,\zeta \rangle \ \omega .}

More generally, ifM{\displaystyle M} is non-orientable, one can define the Hodge star of ak-form as a (nk)-pseudo differential form; that is, a differential form with values in thecanonical line bundle.

Computation in index notation

[edit]

We compute in terms oftensor index notation with respect to a (not necessarily orthonormal) basis{x1,,xn}{\textstyle \left\{{\frac {\partial }{\partial x_{1}}},\ldots ,{\frac {\partial }{\partial x_{n}}}\right\}} in a tangent spaceV=TpM{\displaystyle V=T_{p}M} and its dual basis{dx1,,dxn}{\displaystyle \{dx_{1},\ldots ,dx_{n}\}} inV=TpM{\displaystyle V^{*}=T_{p}^{*}M}, having the metric matrix(gij)=(xi,xj){\textstyle (g_{ij})=\left(\left\langle {\frac {\partial }{\partial x_{i}}},{\frac {\partial }{\partial x_{j}}}\right\rangle \right)} and its inverse matrix(gij)=(dxi,dxj){\displaystyle (g^{ij})=(\langle dx^{i},dx^{j}\rangle )}. The Hodge dual of a decomposablek-form is:(dxi1dxik) = |det[gij]|(nk)!gi1j1gikjkεj1jndxjk+1dxjn.{\displaystyle {\star }\left(dx^{i_{1}}\wedge \dots \wedge dx^{i_{k}}\right)\ =\ {\frac {\sqrt {\left|\det[g_{ij}]\right|}}{(n-k)!}}g^{i_{1}j_{1}}\cdots g^{i_{k}j_{k}}\varepsilon _{j_{1}\dots j_{n}}dx^{j_{k+1}}\wedge \dots \wedge dx^{j_{n}}.}

Hereεj1jn{\displaystyle \varepsilon _{j_{1}\dots j_{n}}} is theLevi-Civita symbol withε1n=1{\displaystyle \varepsilon _{1\dots n}=1}, and weimplicitly take the sum over all values of the repeated indicesj1,,jn{\displaystyle j_{1},\ldots ,j_{n}}. The factorial(nk)!{\displaystyle (n-k)!} accounts for double counting, and is not present if the summation indices are restricted so thatjk+1<<jn{\displaystyle j_{k+1}<\dots <j_{n}}. The absolute value of the determinant is necessary since it may be negative, as for tangent spaces toLorentzian manifolds.

An arbitrary differential form can be written as follows:α = 1k!αi1,,ikdxi1dxik = i1<<ikαi1,,ikdxi1dxik.{\displaystyle \alpha \ =\ {\frac {1}{k!}}\alpha _{i_{1},\dots ,i_{k}}dx^{i_{1}}\wedge \dots \wedge dx^{i_{k}}\ =\ \sum _{i_{1}<\dots <i_{k}}\alpha _{i_{1},\dots ,i_{k}}dx^{i_{1}}\wedge \dots \wedge dx^{i_{k}}.}

The factorialk!{\displaystyle k!} is again included to account for double counting when we allow non-increasing indices. We would like to define the dual of the componentαi1,,ik{\displaystyle \alpha _{i_{1},\dots ,i_{k}}} so that the Hodge dual of the form is given byα=1(nk)!(α)ik+1,,indxik+1dxin.{\displaystyle {\star }\alpha ={\frac {1}{(n-k)!}}({\star }\alpha )_{i_{k+1},\dots ,i_{n}}dx^{i_{k+1}}\wedge \dots \wedge dx^{i_{n}}.}

Using the above expression for the Hodge dual ofdxi1dxik{\displaystyle dx^{i_{1}}\wedge \dots \wedge dx^{i_{k}}}, we find:[3](α)jk+1,,jn=|det[gab]|k!αi1,,ikgi1j1gikjkεj1,,jn.{\displaystyle ({\star }\alpha )_{j_{k+1},\dots ,j_{n}}={\frac {\sqrt {\left|\det[g_{ab}]\right|}}{k!}}\alpha _{i_{1},\dots ,i_{k}}\,g^{i_{1}j_{1}}\cdots g^{i_{k}j_{k}}\,\varepsilon _{j_{1},\dots ,j_{n}}\,.}

Although one can apply this expression to any tensorα{\displaystyle \alpha }, the result is antisymmetric, since contraction with the completely anti-symmetric Levi-Civita symbol cancels all but the totally antisymmetric part of the tensor. It is thus equivalent to antisymmetrization followed by applying the Hodge star.

The unit volume formω=1nV{\textstyle \omega ={\star }1\in \bigwedge ^{n}V^{*}} is given by:ω=|det[gij]|dx1dxn.{\displaystyle \omega ={\sqrt {\left|\det[g_{ij}]\right|}}\;dx^{1}\wedge \cdots \wedge dx^{n}.}

Codifferential

[edit]

The most important application of the Hodge star on manifolds is to define thecodifferentialδ{\displaystyle \delta } onk{\displaystyle k}-forms. Letδ=(1)n(k+1)+1s d=(1)k1d{\displaystyle \delta =(-1)^{n(k+1)+1}s\ {\star }d{\star }=(-1)^{k}\,{\star }^{-1}d{\star }}whered{\displaystyle d} is theexterior derivative or differential, ands=1{\displaystyle s=1} for Riemannian manifolds. Thend:Ωk(M)Ωk+1(M){\displaystyle d:\Omega ^{k}(M)\to \Omega ^{k+1}(M)}whileδ:Ωk(M)Ωk1(M).{\displaystyle \delta :\Omega ^{k}(M)\to \Omega ^{k-1}(M).}

The codifferential is not anantiderivation on the exterior algebra, in contrast to the exterior derivative.

The codifferential is theadjoint of the exterior derivative with respect to the square-integrable scalar product:η,δζ = dη,ζ,{\displaystyle \langle \!\langle \eta ,\delta \zeta \rangle \!\rangle \ =\ \langle \!\langle d\eta ,\zeta \rangle \!\rangle ,}whereζ{\displaystyle \zeta } is ak{\displaystyle k}-form andη{\displaystyle \eta } a(k1){\displaystyle (k\!-\!1)}-form. This property is useful as it can be used to define the codifferential even when the manifold is non-orientable (and the Hodge star operator not defined). The identity can be proved from Stokes' theorem for smooth forms:0 = Md(ηζ) = M(dηζ+(1)k1η1dζ) = dη,ζη,δζ,{\displaystyle 0\ =\ \int _{M}d(\eta \wedge {\star }\zeta )\ =\ \int _{M}\left(d\eta \wedge {\star }\zeta +(-1)^{k-1}\eta \wedge {\star }\,{\star }^{-1}d\,{\star }\zeta \right)\ =\ \langle \!\langle d\eta ,\zeta \rangle \!\rangle -\langle \!\langle \eta ,\delta \zeta \rangle \!\rangle ,}providedM{\displaystyle M} has empty boundary, orη{\displaystyle \eta } orζ{\displaystyle {\star }\zeta } has zero boundary values. (The proper definition of the above requires specifying atopological vector space that is closed and complete on the space of smooth forms. TheSobolev space is conventionally used; it allows the convergent sequence of formsζiζ{\displaystyle \zeta _{i}\to \zeta } (asi{\displaystyle i\to \infty }) to be interchanged with the combined differential and integral operations, so thatη,δζiη,δζ{\displaystyle \langle \!\langle \eta ,\delta \zeta _{i}\rangle \!\rangle \to \langle \!\langle \eta ,\delta \zeta \rangle \!\rangle } and likewise for sequences converging toη{\displaystyle \eta }.)

Since the differential satisfiesd2=0{\displaystyle d^{2}=0}, the codifferential has the corresponding propertyδ2=(1)ns2dd=(1)nk+k+1s3d2=0.{\displaystyle \delta ^{2}=(-1)^{n}s^{2}{\star }d{\star }{\star }d{\star }=(-1)^{nk+k+1}s^{3}{\star }d^{2}{\star }=0.}

TheLaplace–deRham operator is given byΔ=(δ+d)2=δd+dδ{\displaystyle \Delta =(\delta +d)^{2}=\delta d+d\delta }and lies at the heart ofHodge theory. It is symmetric:Δζ,η=ζ,Δη{\displaystyle \langle \!\langle \Delta \zeta ,\eta \rangle \!\rangle =\langle \!\langle \zeta ,\Delta \eta \rangle \!\rangle }and non-negative:Δη,η0.{\displaystyle \langle \!\langle \Delta \eta ,\eta \rangle \!\rangle \geq 0.}

The Hodge star sendsharmonic forms to harmonic forms. As a consequence ofHodge theory, thede Rham cohomology is naturally isomorphic to the space of harmonick-forms, and so the Hodge star induces an isomorphism of cohomology groups:HΔk(M)HΔnk(M),{\displaystyle {\star }:H_{\Delta }^{k}(M)\to H_{\Delta }^{n-k}(M),}which in turn gives canonical identifications viaPoincaré duality ofH k(M) with itsdual space.

In coordinates, with notation as above, the codifferential of the formα{\displaystyle \alpha } may be written asδα= 1k!gml(xlαm,i1,,ik1Γmljαj,i1,,ik1)dxi1dxik1,{\displaystyle \delta \alpha =\ -{\frac {1}{k!}}g^{ml}\left({\frac {\partial }{\partial x_{l}}}\alpha _{m,i_{1},\dots ,i_{k-1}}-\Gamma _{ml}^{j}\alpha _{j,i_{1},\dots ,i_{k-1}}\right)dx^{i_{1}}\wedge \dots \wedge dx^{i_{k-1}},}where hereΓmlj{\displaystyle \Gamma _{ml}^{j}} denotes theChristoffel symbols of{x1,,xn}{\textstyle \left\{{\frac {\partial }{\partial x_{1}}},\ldots ,{\frac {\partial }{\partial x_{n}}}\right\}}.

Poincare lemma for codifferential

[edit]

In analogy to thePoincare lemma forexterior derivative, one can define its version for codifferential, which reads[4]

Ifδω=0{\displaystyle \delta \omega =0}forωΛk(U){\displaystyle \omega \in \Lambda ^{k}(U)}, whereU{\displaystyle U}is astar domain on a manifold, then there isαΛk+1(U){\displaystyle \alpha \in \Lambda ^{k+1}(U)}such thatω=δα{\displaystyle \omega =\delta \alpha }.

A practical way of findingα{\displaystyle \alpha } is to use cohomotopy operatorh{\displaystyle h}, that is a local inverse ofδ{\displaystyle \delta }. One has to define ahomotopy operator[4]

Hβ=01Kβ|F(t,x)tkdt,{\displaystyle H\beta =\int _{0}^{1}{\mathcal {K}}\lrcorner \beta |_{F(t,x)}t^{k}dt,}

whereF(t,x)=x0+t(xx0){\displaystyle F(t,x)=x_{0}+t(x-x_{0})} is the linear homotopy between its centerx0U{\displaystyle x_{0}\in U} and a pointxU{\displaystyle x\in U}, and the (Euler) vectorK=i=1n(xx0)ixi{\displaystyle {\mathcal {K}}=\sum _{i=1}^{n}(x-x_{0})^{i}\partial _{x^{i}}} forn=dim(U){\displaystyle n=\dim(U)} is inserted into the formβΛ(U){\displaystyle \beta \in \Lambda ^{*}(U)}. We can then define cohomotopy operator as[4]

h:Λ(U)Λ(U),h:=η1H{\displaystyle h:\Lambda (U)\rightarrow \Lambda (U),\quad h:=\eta {\star }^{-1}H\star },

whereηβ=(1)kβ{\displaystyle \eta \beta =(-1)^{k}\beta } forβΛk(U){\displaystyle \beta \in \Lambda ^{k}(U)}.

The cohomotopy operator fulfills (co)homotopy invariance formula[4]

δh+hδ=ISx0,{\displaystyle \delta h+h\delta =I-S_{x_{0}},}

whereSx0=1sx0{\displaystyle S_{x_{0}}={\star }^{-1}s_{x_{0}}^{*}{\star }}, andsx0{\displaystyle s_{x_{0}}^{*}} is thepullback along the constant mapsx0:xx0{\displaystyle s_{x_{0}}:x\rightarrow x_{0}}.

Therefore, if we want to solve the equationδω=0{\displaystyle \delta \omega =0}, applying cohomotopy invariance formula we get

ω=δhω+Sx0ω,{\displaystyle \omega =\delta h\omega +S_{x_{0}}\omega ,} wherehωΛk+1(U){\displaystyle h\omega \in \Lambda ^{k+1}(U)} is a differential form we are looking for, and "constant of integration"Sx0ω{\displaystyle S_{x_{0}}\omega } vanishes unlessω{\displaystyle \omega } is a top form.

Cohomotopy operator fulfills the following properties:[4]h2=0,δhδ=δ,hδh=h{\displaystyle h^{2}=0,\quad \delta h\delta =\delta ,\quad h\delta h=h}. They make it possible to use it to define[4]anticoexact forms onU{\displaystyle U} byY(U)={ωΛ(U)|ω=hδω}{\displaystyle {\mathcal {Y}}(U)=\{\omega \in \Lambda (U)|\omega =h\delta \omega \}}, which together withexact formsC(U)={ωΛ(U)|ω=δhω}{\displaystyle {\mathcal {C}}(U)=\{\omega \in \Lambda (U)|\omega =\delta h\omega \}} make adirect sum decomposition[4]

Λ(U)=C(U)Y(U){\displaystyle \Lambda (U)={\mathcal {C}}(U)\oplus {\mathcal {Y}}(U)}.

This direct sum is another way of saying that the cohomotopy invariance formula is a decomposition of unity, and theprojector operators on the summands fulfillsidempotence formulas:[4](hδ)2=hδ,(δh)2=δh{\displaystyle (h\delta )^{2}=h\delta ,\quad (\delta h)^{2}=\delta h}.

These results are extension of similar results for exterior derivative.[5]

Citations

[edit]
  1. ^abHarley Flanders (1963)Differential Forms with Applications to the Physical Sciences,Academic Press
  2. ^abPertti Lounesto (2001)."§3.6 The Hodge dual".Clifford Algebras and Spinors,Volume 286 of London Mathematical Society Lecture Note Series (2nd ed.). Cambridge University Press. p. 39.ISBN 0-521-00551-5.
  3. ^Frankel, T. (2012).The Geometry of Physics (3rd ed.). Cambridge University Press.ISBN 978-1-107-60260-1.
  4. ^abcdefghKycia, Radosław Antoni (2022-07-29)."The Poincare Lemma for Codifferential, Anticoexact Forms, and Applications to Physics".Results in Mathematics.77 (5) 182.arXiv:2009.08542.doi:10.1007/s00025-022-01646-z.ISSN 1420-9012.S2CID 221802588.
  5. ^Edelen, Dominic G. B. (2005).Applied exterior calculus (Revised ed.). Mineola, N.Y.ISBN 978-0-486-43871-9.OCLC 56347718.{{cite book}}: CS1 maint: location missing publisher (link)

References

[edit]
Scope
Mathematics
Notation
Tensor
definitions
Operations
Related
abstractions
Notable tensors
Mathematics
Physics
Mathematicians
Retrieved from "https://en.wikipedia.org/w/index.php?title=Hodge_star_operator&oldid=1300981725"
Categories:
Hidden categories:

[8]ページ先頭

©2009-2025 Movatter.jp