Riemannian manifold should not be confused with Riemann surface.
In differential geometry, a Riemannian manifold is a geometric space on which many geometric notions such as distance, angles, length, volume, and curvature are defined. Euclidean space, the n
Formally, a Riemannian metric (or just a metric) on a smooth manifold is a choice of inner product for each tangent space of the manifold. A Riemannian manifold is a smooth manifold together with a Riemannian metric. The techniques of differential and integral calculus are used to pull geometric data out of the Riemannian metric. For example, integration leads to the Riemannian distance function, whereas differentiation is used to define curvature and parallel transport.
Any smooth surface in three-dimensional Euclidean space is a Riemannian manifold with a Riemannian metric coming from the way it sits inside the ambient space. The same is true for any submanifold of Euclidean space of any dimension. Although John Nash proved that every Riemannian manifold arises as a submanifold of Euclidean space, and although some Riemannian manifolds are naturally exhibited or defined in that way, the idea of a Riemannian manifold emphasizes the intrinsic point of view, which defines geometric notions directly on the abstract space itself without referencing an ambient space. In many instances, such as for hyperbolic space and projective space, Riemannian metrics are more naturally defined or constructed using the intrinsic point of view. Additionally, many metrics on Lie groups and homogeneous spaces are defined intrinsically by using group actions to transport an inner product on a single tangent space to the entire manifold, and many special metrics such as constant scalar curvature metrics and Kähler–Einstein metrics are constructed intrinsically using tools from partial differential equations.
Riemannian geometry, the study of Riemannian manifolds, has deep connections to other areas of math, including geometric topology, complex geometry, and algebraic geometry. Applications include physics (especially general relativity and gauge theory), computer graphics, machine learning, and cartography. Generalizations of Riemannian manifolds include pseudo-Riemannian manifolds, Finsler manifolds, and sub-Riemannian manifolds.
In 1827, Carl Friedrich Gauss discovered that the Gaussian curvature of a surface embedded in 3-dimensional space only depends on local measurements made within the surface (the first fundamental form). This result is known as the Theorema Egregium ("remarkable theorem" in Latin).
A map that preserves the local measurements of a surface is called a local isometry. Call a property of a surface an intrinsic property if it is preserved by local isometries and call it an extrinsic property if it is not. In this language, the Theorema Egregium says that the Gaussian curvature is an intrinsic property of surfaces.
Riemannian manifolds and their curvature were first introduced non-rigorously by Bernhard Riemann in 1854. However, they would not be formalized until much later. In fact, the more primitive concept of a smooth manifold was first explicitly defined only in 1913 in a book by Hermann Weyl.
Élie Cartan introduced the Cartan connection, one of the first concepts of a 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. Specifically, the Einstein field equations are constraints on the curvature of spacetime, which is a 4-dimensional pseudo-Riemannian manifold.
Let
M
p\inM
TpM
M
p
TpM
M
p
However,
TpM
A Riemannian metric
g
M
p
gp:TpM x TpM\toR
\| ⋅ \|p:TpM\toR
\|v\|p=\sqrt{gp(v,v)}
M
g
(M,g)
A Riemannian metric is not to be confused with the distance function of a metric space, which is also called a metric.
If
(x1,\ldots,xn):U\toRn
M
\left\{ | \partial |
\partialx1 |
|p,...c,
\partial | |
\partialxn |
|p\right\}
TpM
p\inU
p
gij|p:=g
|
\right| | ||||
|
\right|p\right)
n2
gij:U\toR
n x n
U
gp
p
\{dx1,\ldots,dxn\}
g=\sumi,jgijdxi ⊗ dxj.
The Riemannian metric
g
gij:U\toR
(U,x).
g
gij
There are situations in geometric analysis in which one wants to consider non-smooth Riemannian metrics. See for instance (Gromov 1999) and (Shi and Tam 2002). However, in this article,
g
See main article: Musical isomorphism.
In analogy to how an inner product on a vector space induces an isomorphism between a vector space and its dual given by
v\mapsto\langlev, ⋅ \rangle
g
(p,v)\mapstogp(v, ⋅ )
TM
T*M
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
(M,g)
(N,h)
f:M\toN
g=f\asth
gp(u,v)=hf(p)(dfp(u),dfp(v))
p\inM
u,v\inTpM.
One says that a smooth map
f:M\toN,
p\inM
U
f:U\tof(U)
An oriented
n
(M,g)
n
dVg
M
M
M
\intMdVg
Let
x1,\ldots,xn
Rn.
gcan
can\left(\sum | |
g | |
i |
ai
\partial | |
\partialxi |
,\sumjbj
\partial | |
\partialxj |
\right)=\sumiaibi
gcan=(dx1)2+ … +(dxn)2
can | |
g | |
ij |
=\deltaij
\deltaij
can) | |
(g | |
ij |
=\begin{pmatrix} 1&0& … &0\\ 0&1& … &0\\ \vdots&\vdots&\ddots&\vdots\\ 0&0& … &1 \end{pmatrix}.
(Rn,gcan)
See main article: Riemannian submanifold.
Let
(M,g)
i:N\toM
M
i*g
g
N
(N,i*g)
(M,g)
In the case where
N\subseteqM
i:N\toM
i(x)=x
i*g
g
N
i*g
*g | |
i | |
p(v,w) |
=gi(p)(dip(v),dip(w)),
dip(v)
v
i.
Examples:
n
Sn=\{x\inRn+1:(x1)2+ … +(xn+1)2=1\}
is a smooth embedded submanifold of Euclidean space
Rn+1
Sn
a,b,c
\left\{x\inR3:
x2 | |
a2 |
+
y2 | |
b2 |
+
z2 | |
c2 |
=1\right\}
is a smooth embedded submanifold of Euclidean space
R3
f:Rn\toR
Rn+1
(M,g)
\widetilde{M}\toM
\widetildeM
M
\widetildeM
On the other hand, if
N
\tildeg
i:N\toM
\tildeg=i*g
Let
(M,g)
(N,h)
M x N
g
h
\widetilde{g}
M x N,
T(p,q)(M x N)\congTpM ⊕ TqN,
\widetilde{g}p,q((u1,u2),(v1,v2))=gp(u1,v1)+hq(u2,v2).
(U,x)
M
(V,y)
N
(U x V,(x,y))
M x N.
gU
g
(U,x)
hV
h
(V,y)
\widetilde{g}
(U x V,(x,y))
\widetilde{g}=\sumij\widetilde{g}ijdxidxj
(\widetilde{g}ij)=\begin{pmatrix}gU&0\ 0&hV\end{pmatrix}.
Tn=S1 x … x S1
S1
Tn
R x … x R
R
Rn
Let
g1,\ldots,gk
M.
f1,\ldots,fk
M
f1g1+\ldots+fkgk
M.
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 using only that a smooth manifold is a locally Euclidean topological space, for this result it is necessary to use that smooth manifolds are Hausdorff and paracompact. The reason is that the proof makes use of a partition of unity.
Let
M
\{(U\alpha,\varphi\alpha)\}\alpha
U\alpha\subseteqM
\varphi\alpha\colonU\alpha\to\varphi\alpha(U
n | |
\alpha)\subseteqR |
Let
\{\tau\alpha\}\alpha
\operatorname{supp}(\tau\alpha)\subseteqU\alpha
\alpha\inA
Define a Riemannian metric
g
M
g=\sum\alpha\tau\alpha ⋅ \tilde{g}\alpha
\tilde{g}\alpha=
* | |
\varphi | |
\alpha |
gcan.
gcan
Rn
*g | |
\varphi | |
\alpha |
can
\varphi\alpha
\tilde{g}\alpha
U\alpha
\tau\alpha ⋅ \tilde{g}\alpha
M
\operatorname{supp}(\tau\alpha)\subseteqU\alpha
U\alpha
g
An alternative proof uses the Whitney embedding theorem to embed
M
M
(M,g),
F:M\toRN
N
F
RN
g.
An admissible curve is a piecewise smooth curve
\gamma:[0,1]\toM
\gamma'(t)\inT\gamma(t)M
t\mapsto\|\gamma'(t)\|\gamma(t)
[0,1]
L(\gamma)
\gamma:[0,1]\toM
1 | |
L(\gamma)=\int | |
0 |
\|\gamma'(t)\|\gamma(t)dt.
(M,g)
dg:M x M\to[0,infty)
dg(p,q)=inf\{L(\gamma):\gammaanadmissiblecurvewith\gamma(0)=p,\gamma(1)=q\}.
(M,dg)
(M,dg)
M
In verifying that
(M,dg)
p ≠ q
dg(p,q)>0
There must be some precompact open set around p which every curve from p to q must escape. By selecting this open set to be contained in a coordinate chart, one can reduce the claim to the well-known fact that, in Euclidean geometry, the shortest curve between two points is a line. In particular, as seen by the Euclidean geometry of a coordinate chart around p, any curve from p to q must first pass though a certain "inner radius." The assumed continuity of the Riemannian metric g only allows this "coordinate chart geometry" to distort the "true geometry" by some bounded factor.
To be precise, let
(U,x)
x(p)=0
q\notinU.
V\nix
U
\overline{V}\subsetU.
g
\overline{V},
λ
g(X,X)\geqλ\|X\|2
r\inV
X\inTrM,
\| ⋅ \|
\sup\{r>0:Br(0)\subsetx(V)\}
\gamma:[0,1]\toM
\delta>0
\gamma(\delta)\notinV;
\gamma(\delta)\in\partialV.
The length of
\gamma
\gamma
[0,\delta].
\delta\|\gamma'(t)\|dt. | |
L(\gamma)\geq\sqrt{λ}\int | |
0 |
x(\partialV)\subsetRn
L(\gamma)\geq\sqrt{λ}R.
The observation about comparison between lengths measured by g and Euclidean lengths measured in a smooth coordinate chart, also verifies that the metric space topology of
(M,dg)
M
Although the length of a curve is given by an explicit formula, it is generally impossible to write out the distance function
dg
M
dg:M x M\toR
(M,g)
If one works with Riemannian metrics that are merely continuous but possibly not smooth, the length of an admissible curve and the Riemannian distance function are defined exactly the same, and, as before,
(M,dg)
(M,dg)
M
The diameter of the metric space
(M,dg)
\operatorname{diam}(M,dg)=\sup\{dg(p,q):p,q\inM\}.
(M,dg)
(M,dg)
dg:M x M\toR
If
(M,dg)
See main article: Affine connection.
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
akX(M)
M
\nabla:akX(M) x akX(M)\toakX(M)
M
(X,Y)\mapsto\nablaXY
f\inCinfty(M)
\nabla | |
f1X1+f2X2 |
Y=f1
\nabla | |
X1 |
Y+f2
\nabla | |
X2 |
Y,
\nablaXfY=X(f)Y+f\nablaXY
\nablaXY
Y
X
See main article: Levi-Civita connection.
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.
A connection
\nabla
Xl(g(Y,Z)r)=g(\nablaXY,Z)+g(Y,\nablaXZ)
\nabla
\nablaXY-\nablaYX=[X,Y],
[ ⋅ , ⋅ ]
A Levi-Civita connection is a torsion-free connection that preserves the metric. Once a Riemannian metric is fixed, there exists a unique Levi-Civita connection. Note that the definition of preserving the metric uses the regularity of
g
If
\gamma:[0,1]\toM
\gamma
X:[0,1]\toTM
X(t)\inT\gamma(t)M
t\in[0,1]
akX(\gamma)
\gamma
\gamma
f:[0,1]\toR
(fX)(t)=f(t)X(t)
X\inakX(\gamma).
Let
X
\gamma
\tildeX
\gamma
X(t)=\tildeX\gamma(t)
\tildeX
X
Given a fixed connection
\nabla
M
\gamma:[0,1]\toM
Dt:akX(\gamma)\toakX(\gamma)
\gamma
Dt(aX+bY)=aDtX+bDtY,
Dt(fX)=f'X+fDtX,
\tildeX
X
DtX(t)=\nabla\gamma'(t)\tildeX
See main article: Geodesic.
Geodesics are curves with no intrinsic acceleration. Equivalently, geodesics are curves that locally take the shortest path between two points. They are the generalization of straight lines in Euclidean space to arbitrary Riemannian manifolds. An ant living in a Riemannian manifold walking straight ahead without making any effort to accelerate or turn would trace out a geodesic.
Fix a connection
\nabla
M
\gamma:[0,1]\toM
\gamma
Dt\gamma'
\gamma
Dt\gamma'=0
t
\gamma
For every
p\inM
v\inTpM
\gamma:I\toM
I
\gamma(0)=p
\gamma'(0)=v
I
\gamma(0)=p
\gamma'(0)=v
\gamma(0)=p
\gamma'(0)=v
Every curve
\gamma:[0,1]\toM
\gamma
R2
S2
See main article: Hopf–Rinow theorem.
The Riemannian manifold
M
(-infty,infty)
R2
R2\smallsetminus\{(0,0)\}
R2
p=(1,1)
v=(1,1)
R
The Hopf–Rinow theorem characterizes geodesically complete manifolds.
Theorem: Let
(M,g)
(M,dg)
dg
M
M
See main article: Parallel transport.
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 along a curve in the setting of a general Riemannian manifold. Given a fixed connection, there is a unique way to do parallel transport.
Specifically, call a smooth vector field
V
\gamma
\gamma
DtV=0
\gamma:[0,1]\toM
\gamma(0)=p
\gamma(1)=q
v\inTpM
TqM
\gamma
v
\gamma
q
R2\smallsetminus\{0,0\}
dx2+dy2=dr2+r2d\theta2
dr2+d\theta2
Warning: This is parallel transport on the punctured plane along the unit circle, not parallel transport on the unit circle. Indeed, in the first image, the vectors fall outside of the tangent space to the unit circle.
See main article: Riemann curvature tensor.
The Riemann curvature tensor measures precisely the extent to which parallel transporting vectors around a small rectangle is not the identity map. The Riemann curvature tensor is 0 at every point if and only if the manifold is locally isometric to Euclidean space.
Fix a connection
\nabla
M
R:akX(M) x akX(M) x akX(M)\toakX(M)
R(X,Y)Z=\nablaX\nablaYZ-\nablaY\nablaXZ-\nabla[X,Z
[X,Y]
(1,3)
See main article: Ricci curvature.
Fix a connection
\nabla
M
Ric(X,Y)=\operatorname{tr}(Z\mapstoR(Z,X)Y)
\operatorname{tr}
See main article: Einstein manifold.
The Ricci curvature tensor
Ric
g
Ric=λg
λ
n
See main article: Scalar curvature.
A Riemannian manifold is said to have constant curvature if every sectional curvature equals the number . This is equivalent to the condition that, relative to any coordinate chart, the Riemann curvature tensor can be expressed in terms of the metric tensor as
Rijkl=\kappa(gilgjk-gikgjl).
| ||||||||||
|
A Riemannian space form is a Riemannian manifold with constant curvature which is additionally connected and geodesically complete. A Riemannian space form is said to be a spherical space form if the curvature is positive, a Euclidean space form if the curvature is zero, and a hyperbolic space form or hyperbolic manifold if the curvature is negative. In any dimension, the sphere with its standard Riemannian metric, Euclidean space, and hyperbolic space are Riemannian space forms of constant curvature,, and respectively. Furthermore, the Killing–Hopf theorem says that any simply-connected spherical space form is homothetic to the sphere, any simply-connected Euclidean space form is homothetic to Euclidean space, and any simply-connected hyperbolic space form is homothetic to hyperbolic space.
Using the covering manifold construction, any Riemannian space form is isometric to the quotient manifold of a simply-connected Riemannian space form, modulo a certain group action of isometries. For example, the isometry group of the -sphere is the orthogonal group . Given any finite subgroup thereof in which only the identity matrix possesses as an eigenvalue, the natural group action of the orthogonal group on the -sphere restricts to a group action of, with the quotient manifold inheriting a geodesically complete Riemannian metric of constant curvature . Up to homothety, every spherical space form arises in this way; this largely reduces the study of spherical space forms to problems in group theory. For instance, this can be used to show directly that every even-dimensional spherical space form is homothetic to the standard metric on either the sphere or real projective space. There are many more odd-dimensional spherical space forms, although there are known algorithms for their classification. The list of three-dimensional spherical space forms is infinite but explicitly known, and includes the lens spaces and the Poincaré dodecahedral space.
The case of Euclidean and hyperbolic space forms can likewise be reduced to group theory, based on study of the isometry group of Euclidean space and hyperbolic space. For example, the class of two-dimensional Euclidean space forms includes Riemannian metrics on the Klein bottle, the Möbius strip, the torus, the cylinder, along with the Euclidean plane. Unlike the case of two-dimensional spherical space forms, in some cases two space form structures on the same manifold are not homothetic. The case of two-dimensional hyperbolic space forms is even more complicated, having to do with Teichmüller space. In three dimensions, the Euclidean space forms are known, while the geometry of hyperbolic space forms in three and higher dimensions remains an area of active research known as hyperbolic geometry.
Let be a Lie group, such as the group of rotations in three-dimensional space. Using the group structure, any inner product on the tangent space at the identity (or any other particular tangent space) can be transported to all other tangent spaces to define a Riemannian metric. Formally, given an inner product on the tangent space at the identity, the inner product on the tangent space at an arbitrary point is defined by
gp(u,v)=ge(dL
p-1 |
(u),dL | |
p-1 |
(v)),
The Levi-Civita connection and curvature of a general left-invariant Riemannian metric can be computed explicitly in terms of, the adjoint representation of, and the Lie algebra associated to . These formulas simplify considerably in the special case of a Riemannian metric which is bi-invariant (that is, simultaneously left- and right-invariant). All left-invariant metrics have constant scalar curvature.
Left- and bi-invariant metrics on Lie groups are an important source of examples of Riemannian manifolds. Berger spheres, constructed as left-invariant metrics on the special unitary group SU(2), are among the simplest examples of the collapsing phenomena, in which a simply-connected Riemannian manifold can have small volume without having large curvature. They also give an example of a Riemannian metric which has constant scalar curvature but which is not Einstein, or even of parallel Ricci curvature. Hyperbolic space can be given a Lie group structure relative to which the metric is left-invariant.[1] Any bi-invariant Riemannian metric on a Lie group has nonnegative sectional curvature, giving a variety of such metrics: a Lie group can be given a bi-invariant Riemannian metric if and only if it is the product of a compact Lie group with an abelian Lie group.
A Riemannian manifold is said to be homogeneous if for every pair of points and in, there is some isometry of the Riemannian manifold sending to . This can be rephrased in the language of group actions as the requirement that the natural action of the isometry group is transitive. Every homogeneous Riemannian manifold is geodesically complete and has constant scalar curvature.
Up to isometry, all homogeneous Riemannian manifolds arise by the following construction. Given a Lie group with compact subgroup which does not contain any nontrivial normal subgroup of, fix any complemented subspace of the Lie algebra of within the Lie algebra of . If this subspace is invariant under the linear map for any element of, then -invariant Riemannian metrics on the coset space are in one-to-one correspondence with those inner products on which are invariant under for every element of . Each such Riemannian metric is homogeneous, with naturally viewed as a subgroup of the full isometry group.
The above example of Lie groups with left-invariant Riemannian metrics arises as a very special case of this construction, namely when is the trivial subgroup containing only the identity element. The calculations of the Levi-Civita connection and the curvature referenced there can be generalized to this context, where now the computations are formulated in terms of the inner product on, the Lie algebra of, and the direct sum decomposition of the Lie algebra of into the Lie algebra of and . This reduces the study of the curvature of homogeneous Riemannian manifolds largely to algebraic problems. This reduction, together with the flexibility of the above construction, makes the class of homogeneous Riemannian manifolds very useful for constructing examples.
See main article: Symmetric space.
A connected Riemannian manifold is said to be symmetric if for every point of there exists some isometry of the manifold with as a fixed point and for which the negation of the differential at is the identity map. Every Riemannian symmetric space is homogeneous, and consequently is geodesically complete and has constant scalar curvature. However, Riemannian symmetric spaces also have a much stronger curvature property not possessed by most homogeneous Riemannian manifolds, namely that the Riemann curvature tensor and Ricci curvature are parallel. Riemannian manifolds with this curvature property, which could loosely be phrased as "constant Riemann curvature tensor" (not to be confused with constant curvature), are said to be locally symmetric. This property nearly characterizes symmetric spaces; Élie Cartan proved in the 1920s that a locally symmetric Riemannian manifold which is geodesically complete and simply-connected must in fact be symmetric.
Many of the fundamental examples of Riemannian manifolds are symmetric. The most basic include the sphere and real projective spaces with their standard metrics, along with hyperbolic space. The complex projective space, quaternionic projective space, and Cayley plane are analogues of the real projective space which are also symmetric, as are complex hyperbolic space, quaternionic hyperbolic space, and Cayley hyperbolic space, which are instead analogues of hyperbolic space. Grassmannian manifolds also carry natural Riemannian metrics making them into symmetric spaces. Among the Lie groups with left-invariant Riemannian metrics, those which are bi-invariant are symmetric.
Based on their algebraic formulation as special kinds of homogeneous spaces, Cartan achieved an explicit classification of symmetric spaces which are irreducible, referring to those which cannot be locally decomposed as product spaces. Every such space is an example of an Einstein manifold; among them only the one-dimensional manifolds have zero scalar curvature. These spaces are important from the perspective of Riemannian holonomy. As found in the 1950s by Marcel Berger, any Riemannian manifold which is simply-connected and irreducible is either a symmetric space or has Riemannian holonomy belonging to a list of only seven possibilities. Six of the seven exceptions to symmetric spaces in Berger's classification fall into the fields of Kähler geometry, quaternion-Kähler geometry, G2 geometry, and Spin(7) geometry, each of which study Riemannian manifolds equipped with certain extra structures and symmetries. The seventh exception is the study of 'generic' Riemannian manifolds with no particular symmetry, as reflected by the maximal possible holonomy group.
The statements and theorems above are for finite-dimensional manifolds—manifolds whose charts map to open subsets of
\Rn.
Riemannian metrics are defined in a way similar to the finite-dimensional case. However, there is a distinction between two types of Riemannian metrics:
M
g:TM x TM\to\R,
x\inM
gx:TxM x TxM\to\R
TxM.
M
gx
TxM
g
M
(H,\langle ⋅ , ⋅ \rangle)
x\inH,
H
TxH.
gx(u,v)=\langleu,v\rangle
x,u,v\inH
(M,g)
\operatorname{Diff}(M)
M.
\mu
M.
L2
\operatorname{Diff}(M)
G
f\in\operatorname{Diff}(M),
u,v\inTf\operatorname{Diff}(M).
x\inM,u(x)\inTf(x)M
Gf(u,v)=\intMgf(x)(u(x),v(x))d\mu(x)
Length of curves and the Riemannian distance function
dg:M x M\to[0,infty)
dg
g
M
dg
g
dg
(M,g)
L2
\operatorname{Diff}(M)
In the case of strong Riemannian metrics, one part of the finite-dimensional Hopf–Rinow still holds.
Theorem: Let
(M,g)
dg
However, a geodesically complete strong Riemannian manifold might not be metrically complete and it might have closed and bounded subsets that are not compact. Further, a strong Riemannian manifold for which all closed and bounded subsets are compact might not be geodesically complete.
If
g
. do Carmo . Manfredo Perdigão . Manfredo do Carmo . Riemannian geometry . . Boston, MA . Mathematics: Theory & Applications . 978-0-8176-3490-2 . 1138207 . 1992. Translated from the second Portuguese edition of 1979 original. 0752.53001.