Hodge star operator
{{Short description|Exterior algebraic map taking tensors from p forms to n-p forms}}
In mathematics, the Hodge star operator or Hodge star is a linear map defined on the exterior algebra of a finite-dimensional oriented vector space endowed with a nondegenerate symmetric bilinear form. Applying the operator to an element of the algebra produces the Hodge dual of the element. This map was introduced by W. V. D. Hodge.
For example, in an oriented 3-dimensional Euclidean space, an oriented plane can be represented by the exterior product of two basis vectors, and its Hodge dual is the normal vector given by their cross product; conversely, any vector is dual to the oriented plane perpendicular to it, endowed with a suitable bivector. Generalizing this to an {{math|n}}-dimensional vector space, the Hodge star is a one-to-one mapping of {{math|k}}-vectors to {{math|(n – k)}}-vectors; the dimensions of these spaces are the binomial coefficients .
The naturalness of the star operator means it can play a role in differential geometry, when applied to the cotangent bundle of a pseudo-Riemannian manifold, and hence to Differential form. This allows the definition of the codifferential as the Hodge adjoint of the exterior derivative, leading to the Laplace–de Rham operator. This generalizes the case of 3-dimensional Euclidean space, in which divergence of a vector field may be realized as the codifferential opposite to the gradient operator, and the Laplace operator on a function is the divergence of its gradient. An important application is the Hodge decomposition of differential forms on a closed Riemannian manifold.
Formal definition for ''k''-vectors
Let {{math|V}} be an Dimension (vector space) oriented vector space with a nondegenerate symmetric bilinear form , referred to here as a scalar product. (In more general contexts such as pseudo-Riemannian manifolds and Minkowski space, the bilinear form may not be positive-definite.) This induces a scalar product on Multivector {{nowrap|,}} for , by defining it on simple {{math|k}}-vectors and to equal the Gram determinantHarley Flanders (1963) Differential Forms with Applications to the Physical Sciences, Academic Press{{rp|14}}
:
extended to through linearity.
The unit {{math|n}}-vector is defined in terms of an oriented orthonormal basis of {{math|V}} as:
:
(Note: In the general pseudo-Riemannian case, orthonormality means
\langle e_i,e_j\rangle
\in\{\delta_{ij},-\delta_{ij}\} for all pairs of basis vectors.)
The Hodge star operator is a linear operator on the exterior algebra of {{math|V}}, mapping {{math|k}}-vectors to ({{math|n – k}})-vectors, for . It has the following property, which defines it completely:{{rp|15}}
: for all {{math|k}}-vectors
Dually, in the space of {{math|n}}-forms (alternating {{math|n}}-multilinear functions on ), the dual to is the volume form , the function whose value on is the determinant of the matrix assembled from the column vectors of in -coordinates. Applying to the above equation, we obtain the dual definition:
: for all {{math|k}}-vectors
Equivalently, taking , , and :
:
\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 of {{math|k}}-vectors as over all subsets n – k}})-vector corresponding to the complementary set
:
where
and
Since Hodge star takes an orthonormal basis to an orthonormal basis, it is an isometry on the exterior algebra
Geometric explanation
The Hodge star is motivated by the correspondence between a subspace {{math|W}} of {{math|V}} and its orthogonal subspace (with respect to the scalar product), where each space is endowed with an orientation and a numerical scaling factor. Specifically, a non-zero decomposable {{math|k}}-vector
:
where
A general {{math|k}}-vector is a linear combination of decomposable {{math|k}}-vectors, and the definition of Hodge star is extended to general {{math|k}}-vectors by defining it as being linear.
Examples
= Two dimensions =
In two dimensions with the normalized Euclidean metric and orientation given by the ordering {{math|(x, y)}}, the Hodge star on {{math|k}}-forms is given by
{\star} \, 1 &= dx \wedge dy \\
{\star} \, dx &= dy \\
{\star} \, dy &= -dx \\
{\star} ( dx \wedge dy ) &= 1 .
\end{align}
= Three dimensions =
A common example of the Hodge star operator is the case {{math|1=n = 3}}, when it can be taken as the correspondence between vectors and bivectors. Specifically, for Euclidean R3 with the basis
{\star} \,dx &= dy \wedge dz \\
{\star} \,dy &= dz \wedge dx \\
{\star} \,dz &= dx \wedge dy.
\end{align}
The Hodge star relates the exterior and cross product in three dimensions:
The Hodge star can also be interpreted as a form of the geometric correspondence between an axis of rotation and an infinitesimal 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 space
\,0\!\! & \!\!1 & \!\!\!\!0\!\!\!\!\!\!
\\[-.5em]
\,\!-1\!\!&\!\!0\!\!&\!\!\!\!0\!\!\!\!\!\!
\\[-.5em]
\,0\!\! & \!\!0\!\! & \!\!\!\!0\!\!\!\!\!\!
\end{array}\!\!\!\right], etc. That is, we may interpret the star operator as:
\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 commutator Lie bracket of linear operators:
= Four dimensions =
In case
For concreteness, we discuss the Hodge star operator in Minkowski spacetime where
{\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{align}
while for 2-forms,
{\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{align}
These are summarized in the index notation as
{\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{align}
Hodge dual of three- and four-forms can be easily deduced from the fact that, in the Lorentzian signature,
Note that the combinations
take
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 both mathematical and physical perspectives, making contacts to the use of the two-spinor language in modern physics such as spinor-helicity formalism or twistor theory.
= Conformal invariance =
The Hodge star is conformally invariant on {{math|n}}-forms on a {{math|2n}}-dimensional vector space
are the same.
= Example: Derivatives in three dimensions=
The combination of the
The scalar product identifies 1-forms with vector fields as
In the second case, a vector field
\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:
which becomes the vector field
\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,
{\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{align}
One advantage of this expression is that the identity {{math|1=d{{i sup|2}} = 0}}, which is true in all cases, has as special cases two other identities: (1) {{math|1=curl grad f = 0}}, and (2) {{math|1=div curl F = 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 expression
One can also obtain the Laplacian {{math|1=Δf = div grad f}} in terms of the above operations:
The Laplacian can also be seen as a special case of the more general Laplace–deRham operator
Duality
Applying the Hodge star twice leaves a {{math|k}}-vector unchanged up to a sign: for
:
where {{mvar|s}} is the parity of the signature of the scalar product on {{math|V}}, that is, the sign of the determinant of the matrix of the scalar product with respect to any basis. For example, if {{math|n {{=}} 4}} and the signature of the scalar product is either {{math|(+ − − −)}} or {{math|(− + + +)}} then {{math|s {{=}} −1}}. For Riemannian manifolds (including Euclidean spaces), we always have {{math|s {{=}} 1}}.
The above identity implies that the inverse of
:
{\star}^{-1}: ~ {\textstyle\bigwedge}^{\!k} V &\to {\textstyle\bigwedge}^{\!n-k} V \\
\eta &\mapsto (-1)^{k(n-k)} \!s\, {\star} \eta
\end{align}
If {{mvar|n}} is odd then {{math|k(n − k)}} is even for any {{mvar|k}}, whereas if {{mvar|n}} is even then {{math|k(n − k)}} has the parity of {{mvar|k}}. Therefore:
:
where {{mvar|k}} is the degree of the element operated on.
On manifolds
For an n-dimensional oriented pseudo-Riemannian manifold M, we apply the construction above to each cotangent space
for every k-form
\ =\ \int_M \langle\eta,\zeta\rangle\ \omega.
More generally, if
= Computation in index notation =
We compute in terms of tensor index notation with respect to a (not necessarily orthonormal) basis
{\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}.
Here
An arbitrary differential form can be written as follows:
\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 factorial
{\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 of
({\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 tensor
The unit volume form
= Codifferential =
The most important application of the Hodge star on manifolds is to define the codifferential
where
while
The codifferential is not an antiderivation on the exterior algebra, in contrast to the exterior derivative.
The codifferential is the adjoint of the exterior derivative with respect to the square-integrable scalar product:
where
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,
provided
Since the differential satisfies
The Laplace–deRham operator is given by
and lies at the heart of Hodge theory. It is symmetric:
and non-negative:
The Hodge star sends harmonic forms to harmonic forms. As a consequence of Hodge theory, the de Rham cohomology is naturally isomorphic to the space of harmonic {{mvar|k}}-forms, and so the Hodge star induces an isomorphism of cohomology groups
which in turn gives canonical identifications via Poincaré duality of {{math|H k(M)}} with its dual space.
In coordinates, with notation as above, the codifferential of the form
where here
== Poincare lemma for codifferential ==
In analogy to the Poincare lemma for exterior derivative, one can define its version for codifferential, which reads{{Cite journal |last=Kycia |first=Radosław Antoni |date=2022-07-29 |title=The Poincare Lemma for Codifferential, Anticoexact Forms, and Applications to Physics |url=https://doi.org/10.1007/s00025-022-01646-z |journal=Results in Mathematics |language=en |volume=77 |issue=5 |pages=182 |doi=10.1007/s00025-022-01646-z |issn=1420-9012|arxiv=2009.08542 |s2cid=221802588 }}
: If
A practical way of finding
:
where
:
where
The cohomotopy operator fulfills (co)homotopy invariance formula
:
where
Therefore, if we want to solve the equation
:
Cohomotopy operator fulfills the following properties:
:
This direct sum is another way of saying that the cohomotopy invariance formula is a decomposition of unity, and the projector operators on the summands fulfills idempotence formulas:
These results are extension of similar results for exterior derivative.{{Cite book |last=Edelen |first=Dominic G. B. |url=https://www.worldcat.org/oclc/56347718 |title=Applied exterior calculus |date=2005 |isbn=978-0-486-43871-9 |edition=Revised |location=Mineola, N.Y. |oclc=56347718}}
Citations
{{reflist}}
References
{{refbegin}}
- David Bleecker (1981) Gauge Theory and Variational Principles. Addison-Wesley Publishing. {{isbn|0-201-10096-7}}. Chpt. 0 contains a condensed review of non-Riemannian differential geometry.
- {{cite book|author-last=Jost|author-first=Jürgen|author-link=Jürgen Jost|title=Riemannian Geometry and Geometric Analysis|date=2002|publisher=Springer-Verlag|isbn=3-540-42627-2}}
- Charles W. Misner, Kip S. Thorne, John Archibald Wheeler (1970) Gravitation. W.H. Freeman. {{isbn|0-7167-0344-0}}. A basic review of differential geometry in the special case of four-dimensional spacetime.
- Steven Rosenberg (1997) The Laplacian on a Riemannian manifold. Cambridge University Press. {{isbn|0-521-46831-0}}. An introduction to the heat equation and the Atiyah–Singer theorem.
- [https://sites.science.oregonstate.edu/~tevian/onid/Courses/MTH434/2009/dual.pdf Tevian Dray (1999) The Hodge Dual Operator]. A thorough overview of the definition and properties of the Hodge star operator.
{{refend}}
{{Tensors}}
{{DEFAULTSORT:Hodge Dual}}