Ricci Tensor

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Ricci curvature

In differential geometry, the Ricci curvature tensor, named after Gregorio Ricci-Curbastro, represents
the amount by which the volume of a narrow conical piece of a small geodesic ball in a curved
Riemannian manifold deviates from that of the standard ball in Euclidean space. As such, it provides one
way of measuring the degree to which the geometry determined by a given Riemannian metric might
differ from that of ordinary Euclidean n-space. The Ricci tensor is defined on any pseudo-Riemannian
manifold, as a trace of the Riemann curvature tensor. Like the metric itself, the Ricci tensor is a
symmetric bilinear form on the tangent space of the manifold (Besse 1987, p. 43).[1]

In relativity theory, the Ricci tensor is the part of the curvature of spacetime that determines the degree to
which matter will tend to converge or diverge in time (via the Raychaudhuri equation). It is related to the
matter content of the universe by means of the Einstein field equation. In differential geometry, lower
bounds on the Ricci tensor on a Riemannian manifold allow one to extract global geometric and
topological information by comparison (cf. comparison theorem) with the geometry of a constant
curvature space form. If the Ricci tensor satisfies the vacuum Einstein equation, then the manifold is an
Einstein manifold, which have been extensively studied (cf. Besse 1987). In this connection, the Ricci
flow equation governs the evolution of a given metric to an Einstein metric; the precise manner in which
this occurs ultimately leads to the solution of the Poincaré conjecture.

Contents
Definition
Properties
Direct geometric meaning
Applications
Global geometry and topology
Behavior under conformal rescaling
Trace-free Ricci tensor
Kähler manifolds
Generalization to affine connections
Discrete Ricci curvature
See also
Footnotes
References
External links

Definition
Suppose that (M, g) is an n-dimensional Riemannian manifold, equipped with its Levi-Civita
connection ∇. The Riemannian curvature tensor of M is the (1, 3)-tensor defined by

on vector fields X, Y, Z. Let TpM denote the tangent space of M at a point p. For any pair of tangent
vectors ξ and η in TpM, the Ricci tensor Ric evaluated at (ξ, η) is defined to be the trace of the linear
map TpM → TpM given by

In local coordinates (using the Einstein summation convention), one has

where

In terms of the Riemann curvature tensor and the Christoffel symbols, one has

Due to the symmetries of the Riemann curvature tensor, it is possible for there to be a disagreement on
the sign convention, since

Properties
As a consequence of the Bianchi identities, the Ricci tensor of a Riemannian manifold is symmetric, in
the sense that

It thus follows that the Ricci tensor is completely determined by knowing the quantity Ric(ξ,ξ) for all
vectors ξ of unit length. This function on the set of unit tangent vectors is often simply called the Ricci
curvature, since knowing it is equivalent to knowing the Ricci curvature tensor.

The Ricci curvature is determined by the sectional curvatures of a Riemannian manifold, but generally
contains less information. Indeed, if ξ is a vector of unit length on a Riemannian n-manifold, then
Ric(ξ,ξ) is precisely (n − 1) times the average value of the sectional curvature, taken over all the 2-
planes containing ξ. There is an (n − 2)-dimensional family of such 2-planes, and so only in dimensions
2 and 3 does the Ricci tensor determine the full curvature tensor. A notable exception is when the
manifold is given a priori as a hypersurface of Euclidean space. The second fundamental form, which
determines the full curvature via the Gauss–Codazzi equation, is itself determined by the Ricci tensor and
the principal directions of the hypersurface are also the eigendirections of the Ricci tensor. The tensor
was introduced by Ricci for this reason.
If the Ricci curvature function Ric(ξ,ξ) is constant on the set of unit tangent vectors ξ, the Riemannian
manifold is said to have constant Ricci curvature, or to be an Einstein manifold. This happens if and only
if the Ricci tensor Ric is a constant multiple of the metric tensor g.

The Ricci curvature is usefully thought of as a multiple of the Laplacian of the metric tensor (Chow &
Knopf 2004, Lemma 3.32). Specifically, in harmonic local coordinates the components satisfy

where is the Laplace–Beltrami operator, here regarded as acting on the functions gij. This
fact motivates, for instance, the introduction of the Ricci flow equation as a natural extension of the heat
equation for the metric. Alternatively, in a normal coordinate system based at p, at the point p

Direct geometric meaning


Near any point p in a Riemannian manifold (M,g), one can define preferred local coordinates, called
geodesic normal coordinates. These are adapted to the metric so that geodesics through p correspond to
straight lines through the origin, in such a manner that the geodesic distance from p corresponds to the
Euclidean distance from the origin. In these coordinates, the metric tensor is well-approximated by the
Euclidean metric, in the precise sense that

In fact, by taking the Taylor expansion of the metric applied to a Jacobi field along a radial geodesic in
the normal coordinate system, one has

In these coordinates, the metric volume element then has the following expansion at p:

which follows by expanding the square root of the determinant of the metric.

Thus, if the Ricci curvature Ric(ξ,ξ) is positive in the direction of a vector ξ, the conical region in M
swept out by a tightly focused family of geodesic segments of length emanating from p, with initial
velocity inside a small cone about ξ, will have smaller volume than the corresponding conical region in
Euclidean space, at least provided that is sufficiently small. Similarly, if the Ricci curvature is negative
in the direction of a given vector ξ, such a conical region in the manifold will instead have larger volume
than it would in Euclidean space.

The Ricci curvature is essentially an average of curvatures in the planes including ξ. Thus if a cone
emitted with an initially circular (or spherical) cross-section becomes distorted into an ellipse (ellipsoid),
it is possible for the volume distortion to vanish if the distortions along the principal axes counteract one
another. The Ricci curvature would then vanish along ξ. In physical applications, the presence of a
nonvanishing sectional curvature does not necessarily indicate the presence of any mass locally; if an
initially circular cross-section of a cone of worldlines later becomes elliptical, without changing its
volume, then this is due to tidal effects from a mass at some other location.

Applications
Ricci curvature plays an important role in general relativity, where it is the key term in the Einstein field
equations.

Ricci curvature also appears in the Ricci flow equation, where a time-dependent Riemannian metric is
deformed in the direction of minus its Ricci curvature. This system of partial differential equations is a
non-linear analog of the heat equation, and was first introduced by Richard S. Hamilton in the early
1980s. Since heat tends to spread through a solid until the body reaches an equilibrium state of constant
temperature, Ricci flow may be hoped to produce an equilibrium geometry for a manifold for which the
Ricci curvature is constant. Recent contributions to the subject due to Grigori Perelman now show that
this program works well enough in dimension three to lead to a complete classification of compact 3-
manifolds, along lines first conjectured by William Thurston in the 1970s.

On a Kähler manifold, the Ricci curvature determines the first Chern class of the manifold (mod torsion).
However, the Ricci curvature has no analogous topological interpretation on a generic Riemannian
manifold.

Global geometry and topology


Here is a short list of global results concerning manifolds with positive Ricci curvature; see also classical
theorems of Riemannian geometry. Briefly, positive Ricci curvature of a Riemannian manifold has strong
topological consequences, while (for dimension at least 3), negative Ricci curvature has no topological
implications. (The Ricci curvature is said to be positive if the Ricci curvature function Ric(ξ,ξ) is
positive on the set of non-zero tangent vectors ξ.) Some results are also known for pseudo-Riemannian
manifolds.

1. Myers' theorem states that if the Ricci curvature is bounded from below on a complete
π
Riemannian manifold by (n − 1)k > 0, then the manifold has diameter ≤ √k , with equality
only if the manifold is isometric to a sphere of a constant curvature k. By a covering-space
argument, it follows that any compact manifold of positive Ricci curvature must have finite
fundamental group.
2. The Bishop–Gromov inequality states that if a complete m-dimensional Riemannian
manifold has non-negative Ricci curvature, then the volume of a ball is less than or equal to
the volume of a ball of the same radius in Euclidean m-space. Moreover, if vp(R) denotes
the volume of the ball with center p and radius R in the manifold and V(R) = cmRm denotes
vp(R)
the volume of the ball of radius R in Euclidean m-space then the function V(R) is
nonincreasing. (The last inequality can be generalized to arbitrary curvature bound and is
the key point in the proof of Gromov's compactness theorem.)
3. The Cheeger–Gromoll splitting theorem states that if a complete Riemannian manifold with
Ric ≥ 0 contains a line, meaning a geodesic γ such that d(γ(u),γ(v)) = |u − v| for all
u, v ∈ ℝ, then it is isometric to a product space ℝ × L. Consequently, a complete manifold
of positive Ricci curvature can have at most one topological end. The theorem is also true
under some additional hypotheses for complete Lorentzian manifolds (of metric signature
(+ − − ...)) with non-negative Ricci tensor (Galloway 2000).
These results show that positive Ricci curvature has strong topological consequences. By contrast,
excluding the case of surfaces, negative Ricci curvature is now known to have no topological
implications; Lohkamp (1994) has shown that any manifold of dimension greater than two admits a
Riemannian metric of negative Ricci curvature. (For surfaces, negative Ricci curvature implies negative
sectional curvature; but the point is that this fails rather dramatically in all higher dimensions.)

Behavior under conformal rescaling


If the metric g is changed by multiplying it by a conformal factor e2f, the Ricci tensor of the new,
conformally-related metric g̃ = e2fg is given (Besse 1987, p. 59) by

where Δ = d*d is the (positive spectrum) Hodge Laplacian, i.e., the opposite of the usual trace of the
Hessian.

In particular, given a point p in a Riemannian manifold, it is always possible to find metrics conformal to
the given metric g for which the Ricci tensor vanishes at p. Note, however, that this is only pointwise
assertion; it is usually impossible to make the Ricci curvature vanish identically on the entire manifold by
a conformal rescaling.

For two dimensional manifolds, the above formula shows that if f is a harmonic function, then the
conformal scaling g ↦ e2fg does not change the Ricci tensor (although it still changes its trace with
respect to the metric unless f=0).

Trace-free Ricci tensor


In Riemannian geometry and general relativity, the trace-free Ricci tensor of a pseudo-Riemannian
manifold (M,g) is the tensor defined by

where Ric is the Ricci tensor, S is the scalar curvature, g is the metric tensor, and n is the dimension of
M. The name of this object reflects the fact that its trace automatically vanishes:

If n ≥ 3, the trace-free Ricci tensor vanishes identically if and only if

for some constant λ.

In mathematics, this is the condition for (M,g) to be an Einstein manifold. In physics, this equation
states that (M,g) is a solution of Einstein's vacuum field equations with cosmological constant.
Kähler manifolds
On a Kähler manifold X, the Ricci curvature determines the curvature form of the canonical line bundle
(Moroianu 2007, Chapter 12). The canonical line bundle is the top exterior power of the bundle of
holomorphic Kähler differentials:

The Levi-Civita connection corresponding to the metric on X gives rise to a connection on κ. The
curvature of this connection is the two form defined by

where J is the complex structure map on the tangent bundle determined by the structure of the Kähler
manifold. The Ricci form is a closed 2-form. Its cohomology class is, up to a real constant factor, the first
Chern class of the canonical bundle, and is therefore a topological invariant of X (for compact X) in the
sense that it depends only on the topology of X and the homotopy class of the complex structure.

Conversely, the Ricci form determines the Ricci tensor by

In local holomorphic coordinates zα, the Ricci form is given by

where ∂ is the Dolbeault operator and

If the Ricci tensor vanishes, then the canonical bundle is flat, so the structure group can be locally
reduced to a subgroup of the special linear group SL(n,C). However, Kähler manifolds already possess
holonomy in U(n), and so the (restricted) holonomy of a Ricci-flat Kähler manifold is contained in
SU(n). Conversely, if the (restricted) holonomy of a 2n-dimensional Riemannian manifold is contained
in SU(n), then the manifold is a Ricci-flat Kähler manifold (Kobayashi & Nomizu 1996, IX, §4).

Generalization to affine connections


The Ricci tensor can also be generalized to arbitrary affine connections, where it is an invariant that plays
an especially important role in the study of projective geometry (geometry associated to unparameterized
geodesics) (Nomizu & Sasaki 1994). If ∇ denotes an affine connection, then the curvature tensor R is the
(1,3)-tensor defined by

for any vector fields X, Y, Z. The Ricci tensor is defined to be the trace:
In this more general situation, the Ricci tensor is symmetric if and only if there exist locally a parallel
volume form for the connection.

Discrete Ricci curvature


Discrete notions of Ricci curvature have been defined on graphs and networks, where they quantify local
divergence properties of edges. Olliver's Ricci curvature is defined using optimal transport theory. A
second notion, Forman's Ricci curvature, is based on topological arguments.

See also
Curvature of Riemannian manifolds
Scalar curvature
Ricci calculus
Ricci decomposition
Ricci-flat manifold
Christoffel symbols
Basic introduction to the mathematics of curved spacetime

Footnotes
1. It is assumed that the manifold carries its unique Levi-Civita connection. For a general affine
connection, the Ricci tensor need not be symmetric.

References
Besse, A.L. (1987), Einstein manifolds, Springer, ISBN 978-3-540-15279-8.
Chow, Bennet & Knopf, Dan (2004), The Ricci Flow: an introduction, American
Mathematical Society, ISBN 0-8218-3515-7.
Eisenhart, L.P. (1949), Riemannian geometry, Princeton Univ. Press.
Galloway, Gregory (2000), "Maximum Principles for Null Hypersurfaces and Null Splitting
Theorems", Annales de l'Institut Henri Poincaré A, 1: 543–567, arXiv:math/9909158 (https://
arxiv.org/abs/math/9909158), Bibcode:2000AnHP....1..543G (https://ui.adsabs.harvard.edu/
abs/2000AnHP....1..543G), doi:10.1007/s000230050006 (https://doi.org/10.1007%2Fs0002
30050006).
Kobayashi, S.; Nomizu, K. (1963), Foundations of Differential Geometry, Volume 1,
Interscience.
Kobayashi, Shoshichi; Nomizu, Katsumi (1996), Foundations of Differential Geometry, Vol.
2, Wiley-Interscience, ISBN 978-0-471-15732-8.
Lohkamp, Joachim (1994), "Metrics of negative Ricci curvature", Annals of Mathematics,
Second Series, Annals of Mathematics, 140 (3): 655–683, doi:10.2307/2118620 (https://doi.
org/10.2307%2F2118620), ISSN 0003-486X (https://www.worldcat.org/issn/0003-486X),
JSTOR 2118620 (https://www.jstor.org/stable/2118620), MR 1307899 (https://www.ams.org/
mathscinet-getitem?mr=1307899).
Moroianu, Andrei (2007), Lectures on Kähler geometry, London Mathematical Society
Student Texts, 69, Cambridge University Press, arXiv:math/0402223 (https://arxiv.org/abs/m
ath/0402223), doi:10.1017/CBO9780511618666 (https://doi.org/10.1017%2FCBO97805116
18666), ISBN 978-0-521-68897-0, MR 2325093 (https://www.ams.org/mathscinet-getitem?
mr=2325093)
Nomizu, Katsumi; Sasaki, Takeshi (1994), Affine differential geometry (https://archive.org/de
tails/affinedifferenti0000nomi), Cambridge University Press, ISBN 978-0-521-44177-3.
Ricci, G. (1903–1904), "Direzioni e invarianti principali in una varietà qualunque", Atti R.
Inst. Veneto, 63 (2): 1233–1239.
L.A. Sidorov (2001) [1994], "Ricci tensor" (https://www.encyclopediaofmath.org/index.php?ti
tle=r/r081800), in Hazewinkel, Michiel (ed.), Encyclopedia of Mathematics, Springer
Science+Business Media B.V. / Kluwer Academic Publishers, ISBN 978-1-55608-010-4
L.A. Sidorov (2001) [1994], "Ricci curvature" (https://www.encyclopediaofmath.org/index.ph
p?title=R/r081780), in Hazewinkel, Michiel (ed.), Encyclopedia of Mathematics, Springer
Science+Business Media B.V. / Kluwer Academic Publishers, ISBN 978-1-55608-010-4
Najman, Laurent and Romon, Pascal (2017): Modern approaches to discrete curvature,
Springer (Cham), Lecture notes in mathematics

External links
Z. Shen, C. Sormani "The Topology of Open Manifolds with Nonnegative Ricci Curvature" (h
ttps://arxiv.org/abs/math/0606774) (a survey)
G. Wei, "Manifolds with A Lower Ricci Curvature Bound" (https://arxiv.org/abs/math/061210
7) (a survey)

Retrieved from "https://en.wikipedia.org/w/index.php?title=Ricci_curvature&oldid=931203030"

This page was last edited on 17 December 2019, at 16:16 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By using
this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia
Foundation, Inc., a non-profit organization.

You might also like