Jump to content

Riemannian manifold

From Wikipedia, the free encyclopedia
(Redirected from Riemannian metric)

In differential geometry, a Riemannian manifold (or Riemannian space) (M, g), so called after the German mathematician Bernhard Riemann, is a real, smooth manifold M equipped with a smoothly-varying positive-definite inner product gp on the tangent space TpM at each point p.

The family gp of inner products is called a Riemannian metric (or a Riemannian metric tensor, or just a metric). It is a special case of a metric tensor. Riemannian geometry is the study of Riemannian manifolds.

A Riemannian metric makes it possible to define several geometric notions on a Riemannian manifold, such as angle at an intersection, length of a curve, area of a surface and higher-dimensional analogues (volume, etc.), extrinsic curvature of submanifolds, and intrinsic curvature of the manifold itself.

The requirement that gp is smoothly-varying amounts to that for any smooth coordinate chart (U, x) on M, the n2 functions

are smooth functions, i.e., they are infinitely differentiable. The section Riemannian manifolds with continuous metrics handles the case where the are merely continuous.

History[edit]

Riemannian manifolds were first conceptualized by Bernhard Riemann.

In 1828, Carl Friedrich Gauss proved the Theorema Egregium ("remarkable theorem" in Latin), which says that the Gaussian curvature of a surface can be determined entirely by measuring distances along paths on the surface. That is, the Gaussian curvature of a surface is an intrinsic property that does not depend on how the surface might be embedded in 3-dimensional space. See Differential geometry of surfaces.

Bernhard Riemann extended Gauss's theory to higher-dimensional spaces called manifolds in a way that also allows distances and angles to be measured and the notion of curvature to be defined, again in a way that is intrinsic to the manifold and not dependent upon its embedding in higher-dimensional spaces.

Elie Cartan introduced the Cartan connection, one of the first concepts of connection. Levi-Civita defined the Levi-Civita connection, a special connection on a Riemannian manifold.

Albert Einstein used the theory of pseudo-Riemannian manifolds (a generalization of Riemannian manifolds) to develop general relativity. In particular, his equations for gravitation are constraints on the curvature of spacetime. Other applications of Riemannian geometry include computer graphics and artificial intelligence.

Definition[edit]

Riemannian metrics and Riemannian manifolds[edit]

Let be a smooth manifold. For each point , there is an associated vector space called the tangent space of at . Vectors in are thought of as the vectors tangent to at .

However, does not come equipped with an inner product, which would give tangent vectors a concept of length and angle. This is an important deficiency because calculus teaches that to calculate the length of a curve, the length of vectors tangent to the curve must be defined.

A Riemannian metric on assigns to each a positive-definite inner product in a smooth way (see the section on regularity below). This induces a norm defined by . A smooth manifold endowed with a Riemannian metric is a Riemannian manifold, denoted . A Riemannian metric is a special case of a metric tensor.

The Riemannian metric in coordinates[edit]

If are smooth local coordinates on , the vectors

form a basis of the vector space for any . Relative to this basis, one can define the Riemannian metric's components at each point by

One could consider these as individual functions or as a single matrix-valued function on . The requirement that is a positive-definite inner product says exactly that this matrix-valued function is a symmetric positive-definite matrix at .

In terms of the tensor algebra, the Riemannian metric can be written in terms of the dual basis of the cotangent bundle as

Regularity of the Riemannian metric[edit]

The Riemannian metric is continuous if are continuous in any smooth coordinate chart The Riemannian metric is smooth if are smooth in any smooth coordinate chart. One can consider many other types of Riemannian metrics in this spirit, such as Lipschitz Riemannian metrics or measurable Riemannian metrics.

There are situations in which the metrics are not smooth. Riemannian metrics produced by methods of geometric analysis, in particular, can be less than smooth. See for instance (Gromov 1999) and (Shi and Tam 2002). The section Riemannian manifolds with continuous metrics handles the case where the are merely continuous, but is smooth in this article unless stated otherwise.

Isometries[edit]

An isometry is a function between Riemannian manifolds which preserves all of the structure of Riemannian manifolds. If two Riemannian manifolds have an isometry between them, they are called isometric, and they are considered to be the same manifold for the purpose of Riemannian geometry.

Specifically, if and are two Riemannian manifolds, a diffeomorphism is called an isometry if , that is, if

for all and For example, translations and rotations are both isometries from to itself.

One says that a smooth map not assumed to be a diffeomorphism, is a local isometry if every has an open neighborhood such that is an isometry (and thus a diffeomorphism).

Examples[edit]

Euclidean space[edit]

Let denote the standard coordinates on Then define by

Phrased differently: relative to the standard coordinates, the local representation is given by the constant value

This is clearly a Riemannian metric, and is called the standard Riemannian structure on It is also referred to as Euclidean space of dimension n and gijcan is also called the (canonical) Euclidean metric.

Embedded submanifolds[edit]

The sphere with the round metric is an embedded Riemannian submanifold of .

Let be a Riemannian manifold and let be an embedded submanifold of . The restriction of g to vectors tangent along N defines a Riemannian metric on N.

  • The n-sphere is a smooth embedded submanifold of with its standard metric. The Riemannian metric this induces on is called the round metric.
  • An ellipsoid in is an embedded submanifold, hence it has a Riemannian metric.
  • The graph of a smooth function is an embedded submanifold, so it has a Riemannian metric.

Immersions[edit]

Let be a Riemannian manifold and let be a differentiable map. Then one may consider the pullback of via , which is a symmetric 2-tensor on defined by

where is the pushforward of by

In this setting, generally will not be a Riemannian metric on since it is not positive-definite. For instance, if is constant, then is zero. In fact, is a Riemannian metric if and only if is an immersion, meaning that the linear map is injective for each

  • An important example occurs when is not simply connected, so that there is a covering map This is an immersion, and so the universal cover of any Riemannian manifold automatically inherits a Riemannian metric. More generally, but by the same principle, any covering space of a Riemannian manifold inherits a Riemannian metric.
  • Also, an immersed submanifold of a Riemannian manifold inherits a Riemannian metric.

Products[edit]

Let and be two Riemannian manifolds, and consider the Cartesian product with the usual product smooth structure. The Riemannian metrics and naturally put a Riemannian metric on which can be described in a few ways.

  • Considering the decomposition one may define
  • Let be a smooth coordinate chart on and let be a smooth coordinate chart on Then is a smooth coordinate chart on For convenience let denote the collection of positive-definite symmetric real matrices. Denote the coordinate representation of relative to by and denote the coordinate representation of relative to by Then the local coordinate representation of relative to is given by

For example, the n-torus is defined as the n-fold product If one gives each copy of its standard Riemannian metric, considering as an embedded submanifold, then one can consider the product Riemannian metric on It is called a flat torus.

Positive combinations of metrics[edit]

Let be Riemannian metrics on If are any positive numbers, then is another Riemannian metric on

Every smooth manifold admits a Riemannian metric[edit]

Theorem: Every smooth manifold admits a (non-canonical) Riemannian metric.

This is a fundamental result. Although much of the basic theory of Riemannian metrics can be developed by only using that a smooth manifold is locally Euclidean, for this result it is necessary to include in the definition of "smooth manifold" that it is Hausdorff and paracompact. The reason is that the proof makes use of a partition of unity.

Proof

Let be a differentiable manifold and a locally finite atlas so that are open subsets and are diffeomorphisms. Such an atlas exists because the manifold is paracompact.

Let be a differentiable partition of unity subordinate to the given atlas, i.e. such that for all .

Then define the metric on by

where is the Euclidean metric on and is its pullback along .

This is readily seen to be a metric on .

An alternative proof uses the Whitney embedding theorem to embed into Euclidean space and then pulls back the metric from Euclidean space to . On the other hand, the Nash embedding theorem states that, given any smooth Riemannian manifold there is an embedding for some such that the pullback by of the standard Riemannian metric on is That is, the entire structure of a smooth Riemannian manifold can be encoded by a diffeomorphism to a certain embedded submanifold of some Euclidean space. Therefore, one could argue that nothing can be gained from the consideration of abstract smooth manifolds and their Riemannian metrics. However, there are many natural smooth Riemannian manifolds, such as the set of rotations of three-dimensional space and the hyperbolic space, of which any representation as a submanifold of Euclidean space will fail to represent their remarkable symmetries and properties as clearly as their abstract presentations do.

Metric space structure[edit]

An admissible curve is a piecewise smooth curve whose velocity is nonzero everywhere it is defined. The nonnegative function is defined on the interval except for at finitely many points. The length of an admissible curve is defined as

.

The integrand is bounded and continuous except at finitely many points, so it is integrable. For a connected Riemannian manifold, define by

Theorem: is a metric space, and the metric topology on coincides with the topology on .[1]

In verifying that satisfies all of the axioms of a metric space, the most difficult part is checking that implies .

Although the length of a curve is given by an explicit formula, it is generally impossible to write out the distance function by any explicit means. In fact, if is compact, there always exist points where is non-differentiable, and it can be remarkably difficult to even determine the location or nature of these points, even in seemingly simple cases such as when is an ellipsoid.

Diameter[edit]

The diameter of the metric space is

The Hopf–Rinow theorem shows that if is complete and has finite diameter, it is compact. Conversely, if is compact, then the function has a maximum, since it is a continuous function on a compact metric space. This proves the following.

If is complete, then it is compact if and only if it has finite diameter.

This is not the case without the completeness assumption; for counterexamples one could consider any open bounded subset of a Euclidean space with the standard Riemannian metric.

More generally, and with the same one-line proof, every compact metric space has finite diameter. However, it is not true that a complete metric space of finite diameter must be compact. For an example of a complete and non-compact metric space of finite diameter, consider

with the uniform metric

So, although all of the terms in the above corollary of the Hopf–Rinow theorem involve only the metric space structure of it is important that the metric is induced from a Riemannian manifold.

Connections, geodesics, and curvature[edit]

Connections[edit]

An (affine) connection is an additional structure on a Riemannian manifold that defines differentiation of one vector field with respect to another. Connections contain geometric data, and two Riemannian manifolds with different connections have different geometry.

Let denote the space of vector fields on . A connection

on is a bilinear map such that

  1. For any function , ,
  2. The product rule holds.

The expression is called the covariant derivative of with respect to .[1]

Levi-Civita connection[edit]

Two Riemannian manifolds with different connections have different geometry. Thankfully, there is a natural connection associated to a Riemannian manifold called the Levi-Civita connection.

There are two extra conditions a connection could satisfy:

  1. is parallel with respect to if ,
  2. is torsion-free if , where is the Lie bracket.

A torsion-free connection for which is parallel with respect to is called a Levi-Civita connection. Once a Riemannian metric is fixed, there exists a unique Levi-Civita connection.[1]

Covariant derivative along a curve[edit]

If is a smooth curve, a smooth vector field along is a smooth map such that for all . The set of smooth vector fields along is a vector space under pointwise vector addition and scalar multiplication. One can also pointwise multiply a smooth vector field along by a smooth function :

for .

Let be a smooth vector field along . If is a smooth vector field on a neighborhood of the image of such that , then is called an extension of .

Given a fixed connection on and a smooth curve , there is a unique operator , called the covariant derivative along , such that:

  1. ,
  2. ,
  3. If is an extension of , then .[1]

Geodesics[edit]

In Euclidean space , the maximal geodesics are straight lines.
In the round sphere , the maximal geodesics are great circles.

Geodesics are curves with no intrinsic acceleration. They are the generalization of straight lines in Euclidean space to arbitrary Riemannian manifolds.

Fix a connection on . Let be a smooth curve. The acceleration of is the vector field along . If for all , is called a geodesic.

For every and , there exists a geodesic defined on some open interval containing 0 such that and . Any two such geodesics agree on their common domain.[1] Taking the union over all open intervals containing 0 on which a geodesic satisfying and exists, one obtains a geodesic called a maximal geodesic of which every geodesic satisfying and is a restriction.

Examples[edit]

  • The maximal geodesics of with its standard Riemannian metric are exactly the straight lines.
  • The maximal geodesics of with the round metric are exactly the great circles.
Geodesics of a metric space[edit]

There is also a notion of a geodesic of a metric space. Relative to the metric space , a path is a unit-speed geodesic if for every there is an interval containing it such that

Informally, one may say that one is asking for to locally 'stretch itself out' as much as it can, subject to the unit-speed constraint. The idea is that if is admissible and for all where the derivative exists, then one automatically has by applying the triangle inequality to a Riemann sum approximation of the integral defining the length of So the unit-speed geodesic condition as given above is requiring and to be as far from one another as possible. The fact that we are only looking for curves to locally stretch themselves out is reflected by the first two examples given below; the global shape of may force even the most innocuous geodesics to bend back and intersect themselves.

Unit-speed geodesics, as defined here, are necessarily continuous, and in fact Lipschitz, but they are not necessarily differentiable or piecewise differentiable.

The Hopf–Rinow theorem[edit]

The punctured plane is not geodesically complete because it does not contain the straight line between and .

The Riemannian manifold with its Levi-Civita connection is geodesically complete if the domain of every maximal geodesic is . The plane is geodesically complete. On the other hand, the punctured plane with the restriction of the Riemannian metric from is not geodesically complete as there is no geodesic from to .

If is geodesically complete, then it is "non-extendable" in the sense that it is not isometric to an open proper submanifold of any other Riemannian manifold. The converse is not true, however: there exist non-extendable manifolds that are not complete.[citation needed]

The Hopf–Rinow theorem characterizes geodesically complete manifolds.

Theorem: Let be a connected Riemannian manifold with a smooth metric. The following are equivalent:

  • The metric space is complete (every -Cauchy sequence converges),
  • A subset of is compact if and only if it is closed and -bounded,
  • is geodesically complete.

Parallel transport[edit]

In Euclidean space, all tangent spaces are canonically identified with each other via translation, so it is easy to move vectors from one tangent space to another. Parallel transport is a way of moving vectors from one tangent space to another in the setting of a general manifold. Given a fixed connection, there is a way to do parallel transport.[1]

Riemann curvature tensor[edit]

Fix a connection on . The Riemann curvature tensor is the map defined by

where is the Lie bracket of vector fields.[1] The Riemann curvature tensor is a -tensor field.

Ricci curvature tensor[edit]

Fix a connection on . The Ricci curvature tensor is

where is the trace.[1]

Scalar curvature[edit]

Riemannian manifolds with continuous metrics[edit]

Throughout this section, Riemannian metrics will be assumed to be continuous but not necessarily smooth.

  • Isometries between Riemannian manifolds with continuous metrics are defined the same as in the smooth case.
  • One can consider Riemannian submanifolds of Riemannian manifolds with continuous metrics. The pullback metric of a continuous metric through a smooth function is still a continuous metric.
  • The product of Riemannian manifolds with continuous metrics is defined the same as in the smooth case and yields a Riemannian manifold with a continuous metric.
  • The positive combination of continuous Riemannian metrics is a continuous Riemannian metric.
  • The length of an admissible curve is defined exactly the same as in the case when the metric is smooth.
  • The metric is defined exactly the same as in the case when the metric is smooth. As before, is a metric space, and the metric topology on coincides with the topology on .
  • The metric space geodesics of a Riemannian manifold can be considered just as in the case when the metric is smooth.

Infinite-dimensional manifolds[edit]

The statements and theorems above are for finite-dimensional manifolds—manifolds whose charts map to open subsets of These can be extended, to a certain degree, to infinite-dimensional manifolds; that is, manifolds that are modeled after a topological vector space; for example, Fréchet, Banach and Hilbert manifolds.

Definitions[edit]

Riemannian metrics are defined in a way similar to the finite-dimensional case. However there is a distinction between two types of Riemannian metrics:

  • A weak Riemannian metric on is a smooth function such that for any the restriction is an inner product on
  • A strong Riemannian metric on is a weak Riemannian metric, such that induces the topology on Note that if is not a Hilbert manifold then cannot be a strong metric.

Examples[edit]

  • If is a Hilbert space, then for any one can identify with By setting for all one obtains a strong Riemannian metric.
  • Let be a compact Riemannian manifold and denote by its diffeomorphism group. The latter is a smooth manifold (see here) and in fact, a Lie group. Its tangent bundle at the identity is the set of smooth vector fields on Let be a volume form on Then one can define the weak Riemannian metric, on Let Then for and define The weak Riemannian metric on induces vanishing geodesic distance, see Michor and Mumford (2005).

Metric space structure[edit]

Length of curves is defined in a way similar to the finite-dimensional case. The function is defined in the same manner and is called the geodesic distance. In the finite-dimensional case, the proof that this function is a metric uses the existence of a pre-compact open set around any point. In the infinite case, open sets are no longer pre-compact and so this statement may fail.

  • If is a strong Riemannian metric on , then separates points (hence is a metric) and induces the original topology.
  • If is a weak Riemannian metric but not strong, may fail to separate points or even be degenerate.

For an example of the latter, see Valentino and Daniele (2019).

Hopf–Rinow theorem[edit]

In the case of strong Riemannian metrics, part of the finite-dimensional Hopf–Rinow still works.

Theorem: Let be a strong Riemannian manifold. Then metric completeness (in the metric ) implies geodesic completeness.

A proof can be found in (Lang 1999, Chapter VII, Section 6). The other statements of the finite-dimensional case may fail. An example can be found here.

If is a weak Riemannian metric, then no notion of completeness implies the other in general.

See also[edit]

Footnotes[edit]

  1. ^ a b c d e f g h Lee, John M. (2018). Introduction to Riemannian Manifolds. Springer-Verlag. ISBN 978-3-319-91754-2.

References[edit]

External links[edit]