Thesis
Thesis
TALHA ALI
1. Introduction
The Ricci flow is a geometric flow, which evolves the metric on a Riemannian
manifold in a manner similar to heat diffusion; this process results in the smoothing
of the irregularities in the metric. Given a Riemannian manifold M , with metric
tensor gij , me may compute its Ricci tensor Rij , which is an average of sectional
curvatures. If we further let the metric tensor be dependent upon an external
variable t, usually called time, then we may define the Ricci flow with the equation,
∂gij
= −2Rij .
∂t
Date: 5/8/10.
1
2 TALHA ALI
We will show in the paper, that the complicated formula for the Ricci tensor
actually reduces to an operator whose highest order component is a Laplacian. This
allows us to relate the Ricci flow to the heat flow. In essence, if given a general
manifold the Ricci flow distributes the curvature on the manifold, in a similar
manner to the heat flow. Areas of “concentrated” curvature are dispersed, and
areas of little or no curvature gain curvature. As the metric evolves the manifold
in turn becomes more “even” and smooth.
But the Ricci flow equation is not exactly a heat equation, among other tech-
nicalities it is not linear. Thus, there is no analytic theory which guarantees the
existence of solutions defined for all time. Unlike the heat flow, singularities can
form which cause the curvature to increase unboundedly. The Ricci flow equation
is in turn a reaction-diffusion equation for curvature on a manifold. As the metric
tensor evolves via the Ricci flow, it behaves like a heat equation plus some lower
order terms. In some cases, the lower order terms are negligible, however in other
cases the lower order are not negligible and the resulting metric has singularities.
In section 3, we will introduce the geodesic equation which allows the reader
to grasp the concept of the shortest distance between two points on a manifold. For
Euclidean flat space, this a fairly straightforward idea because the shortest distance
between two paths is a straight line. However the notion of a straight line does not
make sense in curved space, and thus the geodesic equation helps to describe the
minimizing path between two points.
We discuss the heat flow in section 6. The heat flow is a linear partial differ-
ential equation that describes the distribution of heat over time. The last section
of the paper will discuss the evolution of various geometric quantities such as the
Riemann tensor. The essential “game” in the Ricci flow is to control how these
quantities evolve. We will close the paper by discussing the Einstein-Hilbert func-
tional, which relates the Ricci flow to a gradient flow.
2. Riemannian Geometry
2.1. Tensor Algebra. The tensor product is the formal multiplication of vectors.
Let V and W be vector spaces over the field R, and let α ∈ R be a scalar and
v, v 0 ∈ V and w, w0 ∈ W be vectors. We can define the tensor product between v
and w as the formal symbol v ⊗ w, subject to the following linearity conditions:
v ⊗ (αw) = α(v ⊗ w) = (αv) ⊗ w
0
v ⊗ (w + w ) = v ⊗ w + v ⊗ w0
(v + v 0 ) ⊗ w = v ⊗ w + v 0 ⊗ w.
We define the tensor product between vector spaces V and W , denoted V ⊗ W , as
the linear span of tensors of the form v ⊗ w. Namely,
V ⊗ W = spanR {v ⊗ wv ∈ V, w ∈ W }.
We define the dual V ∗ of a vector space V to be the space of linear functionals
Ni
f : V → R. The definition of tensor product can be iterated. We can define V,
Ni,k
and V respectively by setting
Oi
V = V ⊗ V ⊗ ··· ⊗ V (i many copies of V )
Oi,k Oi Ok
V = V ⊗ V ∗ = V ⊗ V ⊗ · · · ⊗ V ⊗ V ∗ ⊗ V ∗ ⊗ · · · ⊗ V ∗.
s
Z f (b)
dγ dγ
L(γ(s)) = , ds
f (a) ds ds
Z f (b) s
dt dγ dγ
= , ds
f (a) ds dt dt
Z f −1 (f (b)) s
dγ dγ
= , dt
f −1 (f (a)) dt dt
Z b s
dγ dγ
= , dt
a dt dt
= L(γ(t))
We take it for granted that the shortest path between two points is a straight
line. However, this notion only makes sense in Euclidean space, which happens to
be flat and simply connected. If we consider other geometries perhaps with nonzero
sectional curvature, then the shortest path between two points must to be described
in a different way. The following equation describes how this may be done.
Definition 1 (Geodesic). Given a manifold M , a path γ is called a is geodesic if
E(γ) = inf E(η),
η
where the infimum is taken over all paths η : [a, b] → M with η(a) = γ(a) and
η(b) = γ(b).
There will be two vector fields involved in the calculation, if we fix t and vary s
d
we obtain the variation field ds ; likewise if fix s, and vary t we obtain the directional
d
field dt . Note that since the end points are fixed, there is no variation at the end
points. Theorem 2.1 tells us that E-minimizers and L-minimizers are the same up
to reparametrization, so we can minimize the E functional instead.
THE RICCI FLOW 7
d
If γs0 is a geodesic then we must have E(γs0 ) ≤ E(γs ), which implies ds E(γs ) s=s0
=
0. Computing this derivative, we get
Z b
d d d d
E(γs ) = h , idt
ds s=s0 ds a dt dt
Z b Z b
d d d d d
= h , idt = 2 h5 d , idt
a ds dt dt a
ds dt dt
Z b Z b
d d d d d
= 2 h , idt − 2 h , 5 d idt
a dt ds dt a ds dt dt
Z b
d d b d d
= 2 , −2 h , 5 d idt
ds dt a a ds dt dt
Z b
d d
= 0−2 h , 5 d idt
a ds
dt dt
The first term is zero since we have assumed there is no variation at the end points,
so we have
Z b
d d d
E(γs )|s=s0 = −2 h , 5 d idt
ds a ds dt dt
Since γs0 is the shortest path between the end points, we must have
Z b
d d d
0= E(γs )|s=s0 = −2 h , 5 d idt
ds a ds dt dt
d d
Since this must hold for any variational field ds , we must have 5 d dt ≡ 0. Hence
dt
d
we have, γs0 is a geodesic if and only if 5 d dt ≡ 0.
dt
Theorem 3.2. Let {xi } be a local coordinate chart on M and let γ(t) = (x1 (t), x2 (t), . . . , xn (t))
be a path on M . Then γ is a geodesic if and only if
d2 xk dxi dxj
+ Γkij =0
dt dt dt
dγ dxi ∂
=
dt dt ∂xi
8 TALHA ALI
If γ is a geodesic then
dγ
0 = ∇ dγ
dt dt
dxj ∂
= ∇ dxi ∂
dt ∂xi dt ∂xj
i
dx dxj ∂
= ∇ ∂i
dt ∂x dt ∂xj
dxi ∂ dxj ∂ dxi dxj ∂
= i j
+ ∇ ∂
dt ∂x dt ∂x dt dt ∂xi ∂xj
d dxk ∂ dxi dxj k ∂
= + Γ
dt dt ∂xk dt dt ij ∂xk
d2 xk dxi dxj k ∂
= ( 2 + Γ )
dt dt dt ij ∂xk
Hence we have
d2 xk dxi dxj
+ Γkij =0
dt dt dt
Proof. This follows from the existence theorems of ordinary differential equa-
tions.
which implies
C11 t + C12
C21 t + C22
x(t) =
..
.
Cn1 t + Cn2
Given initial conditions we may solve explicitly for C11 , C12 , C21 , C22 , ..., Cn1 , Cn2 .
We note that the solutions do indeed make sense because we would expect the
geodesics of Rn to be straight lines.
These are known as stereographic coordinates. Note that (x, y, z) are coordi-
nates on R3 , and (a, b) are stereographic coordinates on the sphere on R2 . It is clear
that we may take any point on the sphere except for the north pole and locate a
corresponding point on R2 via φ, the north pole is the point at infinity. φ is clearly
a bijection if we restrict its domain to all points except the north pole. It’s inverse
is,
2a 2b a2 + b2 − 1
φ−1 : R2 → S2 such that φ−1 (a, b) = ( , , ) = (x, y, z)
a2 + b2 + 1 a2 + b2 + 1 a2 + b2 + 1
and
N = lim φ−1 (a, b) = lim φ−1 (a, b) = lim φ−1 (a, b)
a→∞ b→∞ (a,b)→(∞,∞)
4. Exponential Map
The theorems for ODE’s guarantee the existence and uniqueness of solutions
to the geodesic equation, however the solutions may only exist for a very short
period of time.
Theorem 4.1. Given a Riemannian Manifold M, point p ∈ M and vector v ∈
Tp M , there is a unique geodesic γv : [0, ∞] → M with γ(0) = p, and γ̇(0) = v.
Proof. This follows from the existence and uniqueness theorems for ODE’s.
Definition 2 (Exponential Map). Given a Riemannian manifold M and point
p ∈ M , we define the exponential map centered at point p to be
expp : Tp M → M
such that expp (v) = γv (1) where γv the geodesic with γ(0) = p, and γ̇(0) = v.
We should note that exp is smooth, and defined globally. One should note that for
c > 0, expp (cv) = γcv (1) = γv (c).
Definition 3 (Exponential Normal Coordinates). Given a Riemannian Manifold
M and point p ∈ M , if we choose an orthonormal basis vi for Tp M , we can assign
the point expp (q i vi ) the coordinates (q1 , q2 , ..., qn ).
This construction only makes sense for a region S ⊆ Tp M , around the origin
of Tp M for which expp : S → M is diffeomorphism onto its image.
THE RICCI FLOW 11
The Gauss Lemma says that given a radial vector v in the tangent space, and
w orthogonal to v also in the tangent space, after applying the exp mapping, they
must still be orthogonal. We should note that this only holds true if v is a radial
vector, in general the exp map does not preserve orthogonality.
∂ ∂
g( i
, j )|p = δij
∂x ∂x
and
∂gij
|p = 0
∂xk
5. Harmonic Coordinates
5.1. N
The Laplacian.N If X, and Y are vector fields, and f is a function, that is
0,0 0,1
f∈ , then ∇f ∈ , in particular
∇f (X) = ∇X f = X(f )
N0,2
Likewise, ∇(∇f ) ∈ , in particular we have
In coordinates we have,
∂ ∂
f,ij = (∇∇f )( , )
∂xi ∂xj
∂ ∂
= (∇ ∂ j (∇f ( i )))) − ∇f ∇ ∂ j
∂x ∂x ∂x ∂xj
∂2f ∂
= i j
− Γkij k
∂x x ∂x
Definition 4 (The Laplacian). Given the metric g, we define the Laplacian ∆ of
function f to be
∂2f ∂
∆f , g ij f,ij = g ij i j − g ij Γkij k
∂x x ∂x
Theorem 5.2 (The existence of Harmonic Coordinates [4]). Given a Riemannian
manifold M with metric g, and let g be of class C k,α (for k ≥ 1) or respectively C ω
in a local coordinate chart about a neighborhood of some point p. Then there is a
neighborhood of p in which the harmonic coordinates exist, these new coordinates
being C k+1,α or respectively C ω functions of the original coordinates. Moreover, all
harmonic coordinate charts defined near p have this regularity.
THE RICCI FLOW 13
Proof. Existence and regularity theorems for second order partial differential
equations are beyond the scope of this paper. We refer the reader to Kazdan-Deturk
[4], lemma 1.2 for a formal proof.
The above two theorems give a sufficient condition for harmonic coordinates
and there existence.
1
=
r
6= 0
and
y
∆θ = ∆ arctan
x
∂ 2 (arctan xy ) ∂ 2 (arctan xy )
= +
∂x2 ∂y 2
−2xy 2xy
= + 2
(x2 + y 2 )2 (x + y 2 )2
= 0
However, the last two terms involve one derivative of the metric so we may
write
∂gij
Rij = ∂k Γkij − ∂i Γkkj + Q(gij , )
∂xk
Calculating further, and leaving out the first order terms we have
∂gij
Rij = ∂k Γkij − ∂i Γkkj + Q(gij , )
∂xk
1 1 ∂gij
= ∂k ( g kl (∂i gjl + ∂j gil − ∂l gij ) − ∂i ( g kl (∂j gkl + ∂k gjl − ∂l gkj ) + Q(gij , )
2 2 ∂xk
2 2 2 2 2 2
1 kl ∂ gjl ∂ gil ∂ gij ∂ gkl ∂ gjl ∂ gkj ∂gij
= g ( k i+ k j
− k l
− i j
− i k
+ i l
) + Q(gij , )
2 ∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂xk
1 ∂ 2 gij 1 ∂ 2 gil ∂ 2 gkl ∂ 2 gkj ∂gij
= − g kl k l + g kl ( k j − i j
+ ) + Q(gij , )
2 ∂x ∂x 2 ∂x ∂x ∂x ∂x ∂xi ∂xl ∂xk
We study the heat flow briefly because it helps us to understand the Ricci
flow. We shall start by describing the heat flow, and then derive the fundamental
solution.
∂u
= ∆u
∂t
Definition 7. Nonhomogenuous Heat Problem
∂u
= ∆u + f
∂t
6.1. Physical interpretation. This equation maybe derived from Fourier’s law.
If W ⊂ U is any smooth subregion, the rate of change of the total quantity within
W equals the negative of the net flux through ∂W .
Z Z
d
u dx = − F · N dS
dt W ∂W
From Gauss’s divergence theorem
Z Z
− F · N dS = − div F dV.
∂W W
Where F is the heat current density. Therefore,
∂u
(1) = −div F.
∂t
Since heat flows from higher concentration to lower concentration, F is proportional
to the gradient of u, however it points in the opposite direction. Hence, for a > 0
we have
F = −a∇u.
Taking a divergence and using 1, we have
∂u
= adiv(∇u) = a∆u.
∂t
Setting a = 1, we obtain the heat equation
∂u
= ∆u.
∂t
16 TALHA ALI
6.2. Fundamental Solution. The heat equation involves one derivative with re-
spect to time t and two derivates with respect to the spatial variable x ∈ Rn .
If u solves the heat equation then so does u(x, t) = u(λx, λ2 t) for some constant
λ ∈ R. This indicates that we are looking for a solution which is invariant under
the dilation scaling
(2) u(x, t) = λα u(λβ x, λt), ∀λ > 0, x ∈ Rn , t > 0
We want to look for a solution with a certain form; since t > 0, we can let λ = t−1
and v(y) = u(y, 1). Plugging into equation (2),
(3) u(x, t)λα u(λβ x, λt)
=
1 x
(4) = α u( β , 1)
t t
1 x
(5) = α v( β ).
t t
x
Inserting (5) into the homogeneous heat equation and letting y = tβ
, we have
∂u
(6) 0 = − ∆u
∂t
∂v
(7) = − ∆v
∂t
α βy ∂v(y) 1
(8) = − α+1 v(y) − α+1 − α+2β ∆v(y).
t t ∂t t
If we let β = 21 , then we can further simplify (8) to
1 1 ∂v(y)
0=− (αv(y) + y + ∆v(y))
tα+1 2 ∂t
which implies that
1 ∂v(y)
0 = αv(y) + y + ∆v(y).
2 ∂t
We simplify further by guessing that v is radial, meaning v only depends on the
distance from the origin |y|. So we let v(y) = w(|y|) = w(r), for some function
w : R → R. Therefore we have
r ∂w ∂ 2 w n − 1 ∂w
0 = αw(r) + + +
2 ∂r ∂r2 r ∂r
Note that since the laplacian is with respect to the spatial variables (x1 , x2 , . . . , xn ) ∈
∂2 n−1 ∂ n
Rn , we have ∆ = ∂r 2 + r ∂r . We reduce further by letting α = 2 . We have
n r ∂w ∂ 2 w n − 1 ∂w
w(r) +
0= + + .
2 2 ∂r ∂r2 r ∂r
We then multiply through by rn−1 ,
n n−1 rn ∂w ∂2w ∂w
(9) 0 = r w(r) + + rn−1 2 + (n − 1)rn−2
2 2 ∂r ∂r ∂r
∂ 1 n n−1 ∂w
(10) = ( r w(r) + r ).
∂r 2 ∂r
After we integrate (10), for some constant a, we have
1 n ∂w
a= r w(r) + rn−1 .
2 ∂r
THE RICCI FLOW 17
Assuming further that the limr→∞ w(r) = 0, and limr→∞ w0 (r) = 0, we can con-
clude that a = 0. Therefore we have
dw 1
= − rw,
dr 2
which is a linear ordinary differential equation. We may solve by separation of
variables, to get
r2
w(r) = Ce− 4
1 x
for some constant C. Applying our assumptions, u(x, t) = tα v( tβ ), β = 12 , α = n
2
and r = |y| = | txβ |, we can conclude that
|x|2
u(x, t) = Ce− 4t
2
|x|
1
(4πt)n/2
e− 4t x ∈ Rn , t > 0
Φ(x, t) =
0 x ∈ Rn , t < 0
This satisfies the heat equation and is known as the fundamental solution.
Z
Φ(x, t)dx = 1
Rn
18 TALHA ALI
Definition 9 (Dirac Delta Function). We define the Dirac delta function, denoted
δx , as
∞ x=0
δx =
0 x 6= 0
and constrained to Z ∞
δx dx = 1.
−∞
According to lemma 3 since the integral over the spacial variable x of the
fundamental solution Φ is 1, and if we take a limit as t → 0, then we may view the
fundamental solution as
2
|x|
1
(4πt)n/2
e− 4t t>0
Φ(x, t) =
δx t=0
∂Φ
Proof. Note that since Φ solves the heat equation, we have ∂t = ∆Φ. There-
fore,
∂ ∂Φ
(Φ ∗ g) = ( ∗ g)
∂t ∂t
= (∆Φ ∗ g) = ∆(Φ ∗ g)
THE RICCI FLOW 19
Theorem 6.1 provides a systematic way to obtain solutions to the heat problem,
if we are given the initial value problem
∂u
∂t = ∆u in Rn × (0, ∞)
u=g in Rn × {t = 0}
we may obtain a solution by taking the convolution Φ∗g. We verify this construction
is valid. At t = 0, Φ = δx , therefore we have u = δx ∗ g = g. For t > 0 we may
apply Theorem 6.1.
The proof of the following theorem illustrates a nice trick in computing the
evolution of various geometric quantities.
Theorem 7.1 (Variation of Christoffel symbols). If g(t) is a one parameter family
d
of metrics with dt gij = hij , then
d k 1
(11) Γij = g kl (hjl,i + hil,j − hij,l )
dt 2
We note that since both sides of (11) are components of tensors, equation (11)
is in fact a tensor equation, and therefore it holds for all coordinate systems, not
just normal coordinates.
Theorem 7.2 (Variation of Riemann curvature tensor). If g(t) is a one parameter
d
family of metrics with dt gij = hij , then we have
d l
R = Γ̇ljk,i − Γ̇lik,j .
dt ijk
Proof. By the ordinary formula for the Riemann curvature tensor we have
∂ l ∂ l
l
Rijk = Γ − Γ + Γpjk Γlip − Γpik Γljp .
∂xi jk ∂xj ik
20 TALHA ALI
Theorem 7.3 (Variation of Ricci curvature tensor). If g(t) is a one parameter
d
family of metrics with dt gij = hij , then
d 1 1
Rij = g lp (hjl,ip + hil,jp − hij,lp ) − H,ij ,
dt 2 2
where H = g ij hij .
Proof. By Theorem 7.2, and the ordinary definition of the Ricci tensor we
have,
d ∂ d p ∂ d p
(12) Rij = Γij − Γ .
dt p
∂x dt ∂xi dt pj
Substituting (11), we have
d ∂ d p ∂ d p
Rij = p
Γij − i
Γpj
dt ∂x dt ∂x dt
∂ 1 pl ∂hjl ∂hil ∂hij ∂ 1 pl ∂hpl ∂hjl ∂hpj
= g + − − g + −
∂xp 2 ∂xi ∂xj ∂xl ∂xi 2 ∂xj ∂xp ∂xl
2
∂ 2 hil ∂ 2 hij ∂ 2 hpl
1 pl ∂ hpj 1
= g l i
+ j p
− l p
− g pl
2 ∂x ∂x ∂x ∂x ∂x ∂x 2 ∂xi ∂xj
Computing in normal coordinates centered at an arbitrary point p ∈ M , we have
g ij |p = δ ij , therefore at p
1 ∂2H
d 1 ∂ ∂hjl ∂hil ∂hij
Rij = + − −
dt 2 ∂xl ∂xi ∂xj ∂xl 2 ∂xi ∂xj
1 1
= (hjl,il + ∂hil,jl − hij,ll ) − H,ij
2 2
Since this a tensor equation, it holds for all coordinate systems.
THE RICCI FLOW 21
Definition 11 (Adjugate). For a given matrix A, with entries Aij , the adjugate
matrix of A, denoted adj(A) = A,
e with entries A
eij is defined as
Lemma 4. From any invertible matrix A, such that the entries Aij (t) are functions
of t, we have
d dA −1
det(A) = det(A) tr( A )
dt dt
where the is the anti symmetric symbol. From this formula we can compute,
d d i1 i2 ...in
det(A) = A1i1 A2i2 . . . Anin
dt dt
dA1i1 dA2i2 dAnin
= i1 i2 ...in ( A2i2 . . . Anin + A1i1 . . . Anin + · · · + A1i1 A2i2 . . . )
dt dt dt
X X dAij dA e
= A
g jk = tr ( A)
j
dt dt
i=k
Where A
e is the adjugate matrix of A. From Cramer’s rule it can be shown,
1 e
A−1 = A.
det A
22 TALHA ALI
Thus we have,
d dA e
det(A) = tr ( A)
dt dt
dA
= tr ( det(A)A−1 )
dt
dA −1
= det(A) tr( A )
dt
Theorem 7.5 (Variation of the volume form). Let {xi } be a positively oriented
local coordinate chart on a Riemannian manifold M , and let gij (t), be the metric
on M at time t. Then
d 1
dV = HdV,
dt 2
ij d
where H = g hij , and hij = dt gij .
R
Lagrangian. We have already computed
R the Euler-Lagrange equation for (−R);
the Euler-Lagrange equation for L is Tij =R 0, where Tij is the stress-energy tensor
of field theory. The statement that L(g) = L − R is extremized is precisely that
1
Tij − Rgij + Ricij = 0.
2
This system of n2 equations is collectively known as Einstein’s Field Equations.
References
[1] B. Chow, P. Lu, L. Ni Hamilton’s Ricci Flow. American Mathematical Society Science Press,
Providence, Rhode Island, 2006
[2] J. Cheeger, D. Ebin Comparison Theorems in Riemannian Geometry.American Mathematical
Society Science Press, Providence, Rhode Island, 2008
[3] M. doCarmo Riemannian GeometryBirkhäuser, Boston, Massachusetts, 1992
[4] D. DeTurck and J. Kazdan, Some regularity theorems in Riemannian geometry, Annales sci-
entifiques de l École Norm. Sup.14 no. 3 (1981) 249–260