Levi-Civita connection

From Wikipedia, the free encyclopedia

In Riemannian or pseudo-Riemannian geometry (in particular the Lorentzian geometry of general relativity), the Levi-Civita connection is the unique affine connection on the tangent bundle of a manifold (i.e. affine connection) that preserves the (pseudo-)Riemannian metric and is torsion-free.

The fundamental theorem of Riemannian geometry states that there is a unique connection which satisfies these properties.

In the theory of Riemannian and pseudo-Riemannian manifolds the term covariant derivative is often used for the Levi-Civita connection. The components (structure coefficients) of this connection with respect to a system of local coordinates are called Christoffel symbols.

History[edit]

The Levi-Civita connection is named after Tullio Levi-Civita, although originally "discovered" by Elwin Bruno Christoffel. Levi-Civita,[1] along with Gregorio Ricci-Curbastro, used Christoffel's symbols[2] to define the notion of parallel transport and explore the relationship of parallel transport with the curvature, thus developing the modern notion of holonomy.[3]

In 1869, Christoffel discovered that the components of the intrinsic derivative of a vector field, upon changing the coordinate system, transform as the components of a contravariant vector. This discovery was the real beginning of tensor analysis.

In 1906, L. E. J. Brouwer was the first mathematician to consider the parallel transport of a vector for the case of a space of constant curvature.[4][5]

In 1917, Levi-Civita pointed out its importance for the case of a hypersurface immersed in a Euclidean space, i.e., for the case of a Riemannian manifold embedded in a "larger" ambient space.[1] He interpreted the intrinsic derivative in the case of an embedded surface as the tangential component of the usual derivative in the ambient affine space. The Levi-Civita notions of intrinsic derivative and parallel displacement of a vector along a curve make sense on an abstract Riemannian manifold, even though the original motivation relied on a specific embedding

In 1918, independently of Levi-Civita, Jan Arnoldus Schouten obtained analogous results.[6] In the same year, Hermann Weyl generalized Levi-Civita's results.[7][8]

Notation[edit]

The metric g can take up to two vectors or vector fields X, Y as arguments. In the former case the output is a number, the (pseudo-)inner product of X and Y. In the latter case, the inner product of Xp, Yp is taken at all points p on the manifold so that g(X, Y) defines a smooth function on M. Vector fields act (by definition) as differential operators on smooth functions. In local coordinates , the action reads

where Einstein's summation convention is used.

Formal definition[edit]

An affine connection is called a Levi-Civita connection if

  1. it preserves the metric, i.e., .
  2. it is torsion-free, i.e., for any vector fields and we have , where is the Lie bracket of the vector fields and .

Condition 1 above is sometimes referred to as compatibility with the metric, and condition 2 is sometimes called symmetry, cf. Do Carmo's text.[9]

Fundamental theorem of (pseudo) Riemannian Geometry[edit]

Theorem Every pseudo Riemannian manifold has a unique Levi Civita connection .

proof: If a Levi-Civita connection exists, it must be unique. To see this, unravel the definition of the action of a connection on tensors to find

Hence we can write condition 1 as

By the symmetry of the metric tensor we then find:

By condition 2, the right hand side is therefore equal to

and we find the Koszul formula

Hence, if a Levi-Civita connection exists, it must be unique, because is arbitrary, is non degenerate, and the right hand side does not depend on .

To prove existence, note that for given vector field and , the right hand side of the Koszul expression is function-linear in the vector field , not just real linear. Hence by the non degeneracy of , the right hand side uniquely defines some new vector field which we suggestively denote as in the left hand side. By substituting the Koszul formula, one now checks that for all vector fields , and all functions

Hence the Koszul expression does, in fact, define a connection, and this connection is compatible with the metric and is torsion free, i.e. is a (hence the) Levi-Civita connection.

Note that with minor variations the same proof shows that there is a unique connection that is compatible with the metric and has prescribed torsion.

Christoffel symbols[edit]

Let be an affine connection on the tangent bundle. Choose local coordinates with coordinate basis vector fields and write for . The Christoffel symbols of with respect to these coordinates are defined as

The Christoffel symbols conversely define the connection on the coordinate neighbourhood because

that is,

An affine connection is compatible with a metric iff

i.e., if and only if

An affine connection is torsion free iff

i.e., if and only if

is symmetric in its lower two indices.

As one checks by taking for , coordinate vector fields (or computes directly), the Koszul expression of the Levi-Civita connection derived above is equivalent to a definition of the Christoffel symbols in terms of the metric as

where as usual are the coefficients of the dual metric tensor, i.e. the entries of the inverse of the matrix .

Derivative along curve[edit]

The Levi-Civita connection (like any affine connection) also defines a derivative along curves, sometimes denoted by D.

Given a smooth curve γがんま on (M, g) and a vector field V along γがんま its derivative is defined by

Formally, D is the pullback connection γがんま*∇ on the pullback bundle γがんま*TM.

In particular, is a vector field along the curve γがんま itself. If vanishes, the curve is called a geodesic of the covariant derivative. Formally, the condition can be restated as the vanishing of the pullback connection applied to :

If the covariant derivative is the Levi-Civita connection of a certain metric, then the geodesics for the connection are precisely those geodesics of the metric that are parametrised proportionally to their arc length.

Parallel transport[edit]

In general, parallel transport along a curve with respect to a connection defines isomorphisms between the tangent spaces at the points of the curve. If the connection is a Levi-Civita connection, then these isomorphisms are orthogonal – that is, they preserve the inner products on the various tangent spaces.

The images below show parallel transport of the Levi-Civita connection associated to two different Riemannian metrics on the plane, expressed in polar coordinates. The metric of left image corresponds to the standard Euclidean metric , while the metric on the right has standard form in polar coordinates (when ), and thus preserves the vector tangent to the circle. This second metric has a singularity at the origin, as can be seen by expressing it in Cartesian coordinates:

Parallel transports under Levi-Civita connections
Cartesian transport
This transport is given by the metric .
Polar transport
This transport is given by the metric .

Example: the unit sphere in R3[edit]

Let ⟨ , ⟩ be the usual scalar product on R3. Let S2 be the unit sphere in R3. The tangent space to S2 at a point m is naturally identified with the vector subspace of R3 consisting of all vectors orthogonal to m. It follows that a vector field Y on S2 can be seen as a map Y : S2R3, which satisfies

Denote as dmY the differential of the map Y at the point m. Then we have:

Lemma — The formula

defines an affine connection on S2 with vanishing torsion.

Proof

It is straightforward to prove that satisfies the Leibniz identity and is C(S2) linear in the first variable. It is also a straightforward computation to show that this connection is torsion free. So all that needs to be proved here is that the formula above produces a vector field tangent to S2. That is, we need to prove that for all m in S2

Consider the map f that sends every m in S2 to Y(m), m, which is always 0. The map f is constant, hence its differential vanishes. In particular
The equation (1) above follows. Q.E.D.

In fact, this connection is the Levi-Civita connection for the metric on S2 inherited from R3. Indeed, one can check that this connection preserves the metric.

Behavior under conformal rescaling[edit]

If the metric in a conformal class is replaced by the conformally rescaled metric of the same class , then the Levi-Civita connection transforms according to the rule[10]

where is the gradient vector field of i.e. the vector field -dual to , in local coordinates given by . Indeed, it is trivial to verify that is torsion-free. To verify metricity, assume that is constant. In that case,

As an application, consider again the unit sphere, but this time under stereographic projection, so that the metric (in complex Fubini–Study coordinates ) is:

This exhibits the metric of the sphere as conformally flat, with the Euclidean metric , with . We have , and so
With the Euclidean gradient , we have
These relations, together with their complex conjugates, define the Christoffel symbols for the two-sphere.

See also[edit]

Notes[edit]

  1. ^ a b Levi-Civita, Tullio (1917). "Nozione di parallelismo in una varietà qualunque" [The notion of parallelism on any manifold]. Rendiconti del Circolo Matematico di Palermo (in Italian). 42: 173–205. doi:10.1007/BF03014898. JFM 46.1125.02. S2CID 122088291.
  2. ^ Christoffel, Elwin B. (1869). "Ueber die Transformation der homogenen Differentialausdrücke zweiten Grades". Journal für die reine und angewandte Mathematik. 1869 (70): 46–70. doi:10.1515/crll.1869.70.46. S2CID 122999847.
  3. ^ See Spivak, Michael (1999). A Comprehensive introduction to differential geometry (Volume II). Publish or Perish Press. p. 238. ISBN 0-914098-71-3.
  4. ^ Brouwer, L. E. J. (1906). "Het krachtveld der niet-Euclidische, negatief gekromde ruimten". Koninklijke Akademie van Wetenschappen. Verslagen. 15: 75–94.
  5. ^ Brouwer, L. E. J. (1906). "The force field of the non-Euclidean spaces with negative curvature". Koninklijke Akademie van Wetenschappen. Proceedings. 9: 116–133. Bibcode:1906KNAB....9..116B.
  6. ^ Schouten, Jan Arnoldus (1918). "Die direkte Analysis zur neueren Relativiteitstheorie". Verhandelingen der Koninklijke Akademie van Wetenschappen te Amsterdam. 12 (6): 95.
  7. ^ Weyl, Hermann (1918). "Gravitation und Elektrizitat". Sitzungsberichte Berliner Akademie: 465–480.
  8. ^ Weyl, Hermann (1918). "Reine Infinitesimal geometrie". Mathematische Zeitschrift. 2 (3–4): 384–411. Bibcode:1918MatZ....2..384W. doi:10.1007/bf01199420. S2CID 186232500.
  9. ^ Carmo, Manfredo Perdigão do (1992). Riemannian geometry. Francis J. Flaherty. Boston: Birkhäuser. ISBN 0-8176-3490-8. OCLC 24667701.
  10. ^ Arthur Besse (1987). Einstein manifolds. Springer. p. 58.

References[edit]

External links[edit]