We now begin using the theory established in the last two lectures to rigorously extract an asymptotic gradient shrinking soliton from the scaling limit of any given -solution. This will require a number of new tools, including the notion of a geometric limit of pointed Ricci flows
, which can be viewed as the analogue of the Gromov-Hausdorff limit in the category of smooth Riemannian flows. A key result here is Hamilton’s compactness theorem: a sequence of complete pointed non-collapsed Ricci flows with uniform bounds on curvature will have a subsequence which converges geometrically to another Ricci flow. This result, which one can view as an analogue of the Arzelá-Ascoli theorem for Ricci flows, relies on some parabolic regularity estimates for Ricci flow due to Shi.
Next, we use the estimates on reduced length from the Harnack inequality analysis in Lecture 13 to locate some good regions of spacetime of a -solution in which to do the asymptotic analysis. Rescaling these regions and applying Hamilton’s compactness theorem (relying heavily here on the
-noncollapsed nature of such solutions) we extract a limit. Formally, the reduced volume is now constant and so Lecture 14 suggests that this limit is a gradient soliton; however, some care is required to make this argument rigorous. In the next section we shall study such solitons, which will then reveal important information about the original
-solution.
Our treatment here is primarily based on Morgan-Tian’s book and the notes of Ye. Other treatments can be found in Perelman’s original paper, the notes of Kleiner-Lott, and the paper of Cao-Zhu. See also the foundational papers of Shi and Hamilton, as well as the book of Chow, Lu, and Ni.
— Geometric limits —
To develop the theory of geometric limits for pointed Ricci flows , we begin by studying such limits in the simpler context of pointed Riemannian manifolds
, i.e. a Riemannian manifold
together with a point
, which we shall call the origin or distinguished point of the manifold. To simplify the discussion, let us restrict attention to complete Riemannian manifolds (though for later analysis we will eventually have to deal with incomplete manifolds).
Definition 1. (Geometric limits) A sequence
of pointed d-dimensional connected complete Riemannian manifolds is said to converge geometrically to another pointed d-dimensional connected complete Riemannian manifold
if there exists a sequence
of connected neighbourhoods of
increasing to
(i.e.
) and a sequence of smooth embeddings
mapping
to
such that
- The closure of each
is compact and contained in
(note that this implies that every compact subset of
will be contained in
for sufficiently large n);
- The pullback metric
converges in the
topology to
(i.e. all derivatives of the metric converge uniformly on compact sets).
Example 1. The pointed round d-sphere of radius R converges geometrically to the pointed Euclidean space as
. Note how this example shows that the geometric limit of compact manifolds can be non-compact.
Example 2. If (M,g) is Hamilton’s cigar (Example 3 from Lecture 8), and is a sequence on M tending to infinity, then
converges geometrically to the pointed round 2-cylinder.
Example 3. The d-torus of length 1/n does not converge to a geometric limit as , despite being flat. More generally, the sequence needs to be locally uniformly non-collapsed in order to have a geometric limit.
Exercise 1. Show that the geometric limit of a sequence
, if it exists, is unique up to (pointed) isometry.
Geometric limits, as their name suggests, tend to preserve all (local) “geometric” or “intrinsic” information about the manifold, although global information of this type can be lost. Here is a typical example:
Exercise 2. Suppose that converges geometrically to
. Show that
for every
, and that we have the Fatou-type inequality
. Give an example to show that the latter inequality can be strict.
Here is the basic compactness theorem for such limits.
Theorem 1. (Compactness theorem) Let
be a sequence of connected complete Riemannian d-dimensional manifolds. Assume that
- (Uniform bounds on curvature and derivatives) For all
, one has the pointwise bound
on the ball
for all sufficiently large n and some constant
.
- (Uniform non-collapsing) For every
there exists
such that
for all
and
, and all sufficiently large n.
Then, after passing to a subsequence if necessary, the sequence
has a geometric limit.
Proof. (Sketch) Let be an arbitrary radius. From Cheeger’s lemma (Theorem 1 from Lecture 7) and hypothesis 2, we know that the injectivity radius on
is bounded from below by some small
for sufficiently large n. Also, from the curvature bounds and Bishop-Gromov comparison geometry (Lemma 1 from Lecture 9) we know that the volume of
is uniformly bounded from above for sufficiently large n.
Now find a maximal -net
of
, thus the balls
are disjoint and the balls
cover
. Volume counting shows that k is bounded for all sufficiently large n; by passing to a subsequence we may assume that it is constant. Similarly we may assume that all the distances
converge to a limit. Using the exponential map and some arbitrary identification of tangent spaces with
, we can identify each ball
with the standard Euclidean ball of radius
. Any pair
of separation less than
induces a smooth transition map from the Euclidean ball of radius
into some subset of
, which can be shown by comparison geometry to be uniformly bounded in
norms; applying the (
version of the) Arzelá-Ascoli theorem we may thus pass to a subsequence and assume that all these transition maps converge in
to a limit. It is then a routine matter to glue together all the limit transition maps to fashion an incomplete manifold to which the balls
converge geometrically (up to errors of
at the boundary). Furthermore, as one increases
, one can show (by a modification of Exercise 1) that these limits are compatible. Now letting
go to infinity (and using the usual diagonalisation trick on all the subsequences obtained), and then gluing together all the incomplete limits obtained, one can create the full geometric limit.
Remark 1. One could use ultrafilters here in place of subsequences, but this does not significantly affect any of the arguments.
Now we turn to geometric limits of pointed Ricci flows (Ricci flows with a specified origin
).
Definition 2. Let
be a sequence of pointed d-dimensional complete connected Ricci flows, each on its own time interval
. We say that a pointed d-dimensional complete connected Ricci flow
on a time interval I is a geometric limit of this sequence if
- Every compact subinterval of I is contained in
for all sufficiently large n.
- There exists neighbourhoods
of
as in Definition 1, compact time intervals
increasing to I, and smooth embeddings
preserving the origin such that the pullback of the flow
to
converges in spacetime
to
.
Exercise 3. Show that if a sequence of -noncollapsed Ricci flows (with a uniform value of
) converges geometrically to another Ricci flow, then the limit flow is also
-noncollapsed.
Now we present Hamilton’s compactness theorem for Ricci flows, which requires less regularity hypotheses than Theorem 1 due to the parabolic smoothing effects of Ricci flow (as captured by Shi’s estimates, see Theorem 3).
Theorem 2. (Hamilton compactness theorem) Let
and
be as in Definition 2, and let I be an open interval obeying hypothesis 1 of that definition. Let
be a time. Suppose that
- For every compact subinterval J of I containing
and every
, one has the curvature bound
on the cylinder
for some
and all sufficiently large n; and
- One has the non-collapsing bound
for some
and
, and all sufficiently large n.
- (Uniform lower bound on curvature) For any compact J, there is a K such that for any
, one has the curvature lower bound
on
for all sufficiently large n. (This is not quite implied by 1. because the curvature bound K there is allowed to depend on r, whereas here we require that K is independent of r.)
Then some subsequence of
converges geometrically to a limit
on I.
Condition 3 is technical (and was erroneously omitted in some of the literature), but it was recently observed by Topping that the claim fails without some hypothesis of this form. Fortunately, in the applications to the Poincare conjecture one always has a uniform lower bound on Ricci curvature, and so Condition 3 is not difficult to verify in practice.
Proof. By Shi’s estimates (Theorem 3) we can upgrade the bound on curvature in hypothesis 1 to bounds on derivatives of curvature. Indeed, these estimates imply that for any J, r as in that hypothesis, and any , we have
for some
and sufficiently large n.
Now we restrict to the time slice and apply Theorem 1. Passing to a subsequence, we can assume that
converges geometrically to a limit
.
For any radius r and any compact J in I containing , we can pull back the flow
to a (spatially incomplete) flow
on the cylinder
for sufficiently large n. By construction,
converges in
norm to
; in particular, it is uniformly bounded in each of the seminorms of this space. Also, each
is a Ricci flow with uniform bounds on any derivative of curvature for sufficiently large n.
Exercise 4. Using these facts, show that the sequence of flows is uniformly bounded in each of the seminorms of
for each fixed J, r, and for n sufficiently large.
By using the Arzelá-Ascoli theorem as before, we may thus pass to a further subsequence and assume that converges in
to a limiting flow
. Clearly this limit is a Ricci flow. Letting
and pasting together the resulting limits one obtains the desired geometric limit. (One has to verify that every geodesic in
starting from
can be extended to any desired length, thus establishing completeness by the Hopf-Rinow theorem, but this is easy to establish given all the uniform bounds on the metric and curvature, and their derivatives. It is here that hypothesis 3 is used to prevent the metric from shrinking too rapidly as one goes backwards in time. )
— Locating an asymptotic gradient shrinking soliton —
We now return to the study of -solutions
. We pick an arbitrary point
and consider the reduced length function
. Recall (see equation (18) from Lecture 11) that we had
(1)
for every . (This bound was obtained from the parabolic inequality
and the maximum principle.) Thus we can find a sequence
with
such that
. (2)
Now recall that as a consequence of Hamilton’s Harnack inequality, we have the pointwise estimates
(3)
and
(4)
(see equations (37), (38) from Lecture 13). From these bounds and Gronwall’s inequality, one easily sees that we can extend (2) to say that
(5)
for any in the cylinder
and any
and
. Applying (3) once more, together with the hypothesis of non-negative curvature more, we also obtain bounded normalised curvature on this cylinder:
(6).
If we thus introduce the rescaled flow by setting
,
, and
, we see that these flows obey hypothesis 1 of Theorem 2. Also, since the original
-solutions are
-noncollapsed, so are their rescalings, which (in conjunction with hypothesis 1) gives us hypothesis 2. We can thus invoke Theorem 2 and assume (after passing to a subsequence) that the rescaled flows converge geometrically to an ancient Ricci flow
on the time interval
. From Exercise 3 we see that this limit is also
-noncollapsed. Since the rescaled flows have non-negative curvature, the limit flow has non-negative curvature also. (Note however that we do not expect in general that
has bounded curvature (for instance, if the original
-solution was a round shrinking sphere terminating at the unit radius sphere, the limit object would be a round shrinking sphere terminating at a point). In particular we do not expect
to be a
-solution.)
Let be the rescaled length function, thus
. From (5) we see that
is uniformly bounded on compact subsets of
for n sufficiently large (where we identify compact subsets of
with subsets of
for n large enough). By the rescaled versions of (4) and (5) we also see that
is also uniformly bounded on such compact sets for sufficiently large n; thus the
are uniformly Lipschitz on each compact set. Applying the Arzelá-Ascoli theorem and passing to a subsequence, we may thus assume that the
converge uniformly on compact sets to some limit
, which is then locally Lipschitz.
Remark 2. We do not attempt to interpret as a reduced length function arising from some point at time t=0; indeed we expect the limiting flow to develop a singularity at this time.
We know that the reduced volume is non-increasing in
and ranges between 0 and
, and so converges to a limit
between 0 and
. This limit cannot equal
since this would mean that the
-solution is flat (by Theorem 1 from Lecture 14), which is absurd. The limit cannot be zero either, since the bounds (5) and the non-collapsing ensure a uniform lower bound on the reduced volume. By rescaling, we conclude that
(7)
for each fixed .
Let us now argue informally, and then return to make the argument rigorous later. Formally taking limits in (7), we conclude that
. (8)
On the other hand, from the proof of the monotonicity of reduced volume from Lecture 10 we have (formally, at least)
(9)
and hence by rescaling
. (10)
Formally taking limits, we obtain
. (11)
We can rewrite this as the assertion that is a subsolution of the backwards heat equation:
. (12)
This (formally) implies that the left-hand side of (8) is non-increasing in . On the other hand, this quantity is constant in
; and so (12) must be obeyed with equality, and thus
. (13)
Also, recall from Lecture 10 that
. (14)
Rescaling and taking limits, we formally conclude that the same is true for ;
. (15)
From (14) and (15) we obtain that the Perelman -functional
(16)
vanishes (cf. the last section of Lecture 11). In particular, it is constant. On the other hand, by (13) and the monotonicity formula for this functional (see Exercise 9 of Lecture 8) we have
. (17)
Combining this with the vanishing of (16) we thus conclude that
(18)
and thus is a gradient shrinking soliton as desired.
— Making the argument rigorous I. Spatial localisation —
Now we turn to the (surprisingly delicate) task of justifying the steps from (7) to (18).
The first task is to deduce (8) from (7). From the dominated convergence theorem it is not difficult to show that
(19)
for any fixed and r; the difficulty is to prevent the escape of mass of
to spatial infinity. (Fatou’s lemma will tell us that the left-hand side of (8) is less than or equal to the right, but this is not enough for our application.)
In order to prevent such an escape, one needs a lower bound on when
is large. (Note that estimates such as (3), (4) only provide upper bounds on
.) The problem is equivalent to that of upper bounding
in terms of
. To do this we need some control on quantities related to the distance function at extremely large distances. Remarkably, such bounds are possible. We begin with a lemma of Perelman (related to an earlier argument of Hamilton).
Lemma 1. Let
be a d-dimensional Riemannian manifold, let
, and let
. Suppose that
on the balls
and
. Then for any minimising geodesic
connecting x and y, we have
, where
is the velocity field.
Proof. We may assume that , since the claim is trivial otherwise. We recall the second variation formula
(20)
whenever one deforms a geodesic along a vector field Y (see equation (17) of Lecture 10). Since
is minimising, the left-hand side of (20) is non-negative when Y vanishes at the endpoints of
. Now let v be any unit vector at x, transported by parallel transport along
. Setting Y(t) to equal tv/r when
, equal to v when
, and equal to (d(x,y)-t)v/r when
, we conclude that
. (21)
Letting v vary over an orthonormal frame and summing, we soon obtain the claim.
The above lemma, combined with the Ricci flow equation, gives an upper bound as to how rapidly the distance function can grow as one goes backwards in time.
Corollary 1. Let
be a d-dimensional Ricci flow, let
, let t be a time, and let
. Suppose that
on
and on
. Then
(in the sense of forward difference quotients).
Using this estimate, we can now obtain a bound on distance in terms of reduced length.
Proposition 1. Let
be a d-dimensional
-solution, let
, and
. Then
. (22)
Proof. We use an argument of Ye. Write A for the expression inside the on the right-hand side, and let
be minimising
-geodesics from
to p, p’ respectively. By the fundamental theorem of calculus, we have
. (24)
Using (3) and the -Gauss lemma
we see that
move at speed
, and that all curvature tensors are
in a
-neighbourhood of either curve. Applying Corollary 1, the chain rule, and the Gauss lemma we conclude that
; (25)
inserting this into (24) we obtain the claim.
Combining this with (5) and rescaling we see that we have a bound of the form
(26)
for all x and some ; taking limits we also obtain
. (27)
On the other hand, from the Bishop-Gromov inequality we know that balls of radius r in either or
have volume
. These facts are enough to establish that the portion of (7) or (8) outside of the ball of radius r decays exponentially fast in r, uniformly in n, and this allows us to take limits in (19) as
to deduce (8) from (7).
— Making the argument rigorous II. Parabolic inequality for —
The next major task in making the previous arguments rigorous is to justify the passage from (10) to (11). First of all, because of the -cut locus, (10) is only valid in the sense of distributions. We would like to take limits and conclude that (11) holds in the sense of distributions as well. There is no difficulty taking limits with the linear terms
in (10), or in the zeroth order terms
; the only problem is in justifying the limit from
to
. We know that the
are uniformly locally Lipschitz, and converge locally uniformly to
; but this is unfortunately not enough to ensure that
converges in the sense of distributions to
, due to possible high frequency oscillations in
. To give a toy counterexample, the one-dimensional functions
are uniformly Lipschitz and converge uniformly to zero, but
converges in the distributional sense to
rather than zero.
Since is bounded and converges distributionally to zero, it will be locally asymptotically orthogonal
. From this and Pythagoras’ theorem we obtain
(28)
in the sense of distributions, where we pass to a subsequence in order to make the limits on both sides exist. (Note that converges locally uniformly to g and so there is no difficulty passing back and forth between those metrics.) The task is now to show that there is not enough oscillation to cause the second term on the right-hand side to be non-vanishing.
To do this, we observe that (10) provides an upper bound on ; indeed on any fixed compact set in
, we have
. This one-sided bound on the Laplacian is enough to rule out the oscillation problem. Indeed, as
converges locally uniformly to
, we see that
(29)
for any non-negative bump function and
chosen so that
on the support of
. Integrating by parts and disposing of a lower order term, we conclude that
. (30)
On the other hand, since is bounded converges weakly to zero, one has
. (31)
One can easily replace and
here by
and
. Combining (30) and (31) we conclude that the second term on the RHS of (28) is non-positive in the sense of distributions. But it is clearly also non-negative, and so it vanishes as required.
This gives (11); as a by-product of the argument we have also established the useful fact
(32)
in the sense of distributions. Combining this with the growth bounds (26), (27) on and $l_\infty$ from the previous section (which give exponential decay bounds on
,
and their first derivatives), it is not too difficult to then justify the remaining steps (12)-(18) of the argument rigorously; see Section 9.2 of Morgan-Tian for full details. (Note that once one reaches (13), one has a nonlinear heat equation for
, and it is not difficult to use the smoothing effects of the heat kernel to then show that the locally Lipschitz function
is in fact smooth.)
— The asymptotic gradient shrinking soliton is not flat —
Finally, we show that the asymptotic gradient shrinking soliton is non-trivial in the sense that its curvature is not identically zero at some time. For if the curvature did vanish everywhere at time t, then the equation (18) simplifies to
. On the other hand, being flat,
is the quotient of Euclidean space
by some discrete subgroup. Lifting
up to this space, we thus see that f is quadratic, and more precisely is equal to
plus an affine-linear function. Thus f has no periodicity whatsoever and so the above-mentioned discrete subgroup is trivial. If we now apply (8) we see that
. But on the other hand, as the original
-solution was not flat, its reduced volume was strictly less than
by Theorem 1 from Lecture 14, a contradiction. Thus the asymptotic gradient soliton is not flat.
— Appendix: Shi’s derivative estimates —
The purpose of this appendix is to prove the following estimate of Shi.
Theorem 3. Suppose that
is a complete d-dimensional Ricci flow on the time interval
, and that on the cylinder
one has the pointwise curvature bound
. Then on any slightly smaller cylinder
one has the curvature bounds
for any
, where the implied constant depend on
.
Proof. (Sketch) We induct on k. The case is trivial, so suppose that
and that the claim has already been proven for all smaller values of k. We allow all implied constants in the O() notation to depend on
. We refer to
and
as the “large cylinder” and “small cylinder” respectively.
We make some reductions. It is easy to see that we can take and T to be small.
Since on the cylinder, we see that the metric at later times of the large cylinder is comparable to the initial metric up to multiplicative constants. The curvature bound tells us that if
is small, then we are inside the conjugacy radius; pulling back under the exponential map, we may thus assume that the exponential map from
is injective on the large cylinder. Let
be the time-varying radial coordinate; observe that the annulus
will be contained between the large cylinder and small cylinder for T small enough.
Exercise 5. Show that if and T are small enough, then
on the large cylinder.
Let be a smooth non-negative radial cutoff to the large cylinder that equals 1 on the small cylinder. From the above exercise, the Gauss lemma, and the chain rule, we see that
.
Now we study the heat equation obeyed by the “energy densities” for various m.
Exercise 6. (Bochner-Weitzenböck type estimate) For any , show that
. (33)
(Hint: start with the equation and use the product rule and the definition of curvature repeatedly.)
From this exercise and the induction hypothesis we see that
(34)
for all , with the understanding that the second term on the right-hand side is absent when m=0. Telescoping this, we can thus find an expression
(35)
for some sufficiently large positive constant , which obeys the heat equation
. Also, by hypothesis we have E=O(1) at time zero. Applying the maximum principle, we obtain the claim.
Exercise 6. Suppose that in the hypotheses of Shi’s theorem that we also have for
on the large cylinder at time zero. Conclude that we have
on the small cylinder for all j.
Exercise 7. Let be a smooth compact manifold, and let
be a bounded solution to the heat equation
which obeys a pointwise bound
at time zero. Establish the bounds
on the spacetime
and all
, where the implied constant depends on (M,g), K, T, and k.
(Update, June 2: some corrections.)
(Update, Oct 18 2011: A recently discovered issue with the Hamilton compactness theorem in the case where the curvature bound is not uniform in r has been addressed.)
13 comments
Comments feed for this article
27 May, 2008 at 7:13 pm
Anonymous
I can’t make the slightest sense out of any of these lectures, but they must be incredibly exciting to those able to attend them and understand their content. I think it is not very often that there is such a deep new result, with such a level of exposition built around it so quickly. Wiles’ proof of FLT comes to mind but it really didn’t play out quite like this.
If you don’t mind my asking, how many people are showing up for the lectures? What fraction of them seem to actually follow what’s going on? What mix of grad students, faculty, undergrads, random onlookers? Are they writing out the exercises and turning them in? Just curious.
28 May, 2008 at 3:55 am
John Sidles
To extend Anonymous’ post, who else is tracking the on-line lectures, and why?
Our QSE Group’s interest in these lectures arises partly because the mathematics is beautiful and the results are historic and profound.
But we also have pragmatic motivations. In both the classical and quantum domains, we engineers expend a great deal of effort adapting state-spaces to suit the problems at-hand … typically this class of problems poses the biggest computational challenge in system engineering.
Sometimes the state-space adaptations we need are geometric — this leads us to study Ricci-flow methods. Sometimes the state-space adaptations are dimensional — this leads us to study topics like L1 optimization and compressive sampling.
Either way, Terence’s lectures are a tremendous source of mathematical information. The theorems are amazing and inspiring, and for us quantumm system engineers, the mathematical toolset deployed to prove the theorems is equally inspiring.
For which, thank you very much! Everybody understands that posting these lectures is a tremendous amount of work … please let us say that we hope you enjoy writing them as much as we enjoy studying them.
28 May, 2008 at 10:42 am
PS
Minor typo: In Definition 1 it looks like the freestanding “p” should be a “
“.
Another really small typo: The reference in Definition 2 should be to Definition 1 rather than 2.
In paragraph 3 of the proof of Hamilton’s compactness theorem it seems that the flow we want to pull back should have a “t” in it instead of a “t_0″.
28 May, 2008 at 2:12 pm
Terence Tao
PS: thanks for the corrections!
Dear anonymous: Actually, this class is more or less similar to other topics graduate classes I have taught in terms of audience, though I have had some visitors from the optimal transport program next door at IPAM drop in from time to time, and of course our resident Ricci flow experts, Petersen and Wylie, are regularly present. The exercises are intended for self-study only and I am not grading them.
28 May, 2008 at 11:19 pm
Andy Sanders
Just as a response to an above comment. I am a graduate student at the University of Maryland and have been following these notes on a “more than topical but less than thorough level.” I was motivated to read along since I study Geometry and Topology and am very interested in applications of PDE to topological and geometric problems, of which I will naively claim this proof is probably the most stupendous example of.
I guess while I’m at it, I’ll just let you know (Prof. Tao) that I am very appreciative of your exposition in these lectures. The wealth of material surrounding this topic is complemented nicely by your heuristic and intuitive explanations.
30 May, 2008 at 10:17 pm
285G, Lecture 16: Classification of asymptotic gradient shrinking solitons « What’s new
[…] and consider the rescaled manifolds . Using Hamilton’s compactness theorem (Theorem 2 from Lecture 15) we may assume that these manifolds converge geometrically to a limit of nonnegative Riemann […]
2 June, 2008 at 1:15 pm
285G, Lecture 17: The structure of κ-solutions « What’s new
[…] we saw in Lecture 15, every -solution has at least one asymptotic gradient shrinking soliton associated to it. Suppose […]
2 June, 2008 at 4:34 pm
Dan
I noticed that your proof of Shi’s local derivative estimates is a bit different from the “usual” one in the book by Chow et. al. (based on Shi’s original proof, and similar to Hamilton’s refined version). This “usual” proof involves finding a quantity that satisfies a nonlinear heat inequality of the form
and then observing that one can essentially “localize the maximum principle argument” when such an inequality holds.
On the other hand, your proof seems much simpler, being a straightforward modification of the proof of the global derivative estimates in Chow et. al. (which I believe is Bando’s proof). In your approach (for m=1), one controls the bad terms in the evolution of
by adding a
term whereas in the other approach one controls the bad terms by multiplying a
term. I was wondering whether you had any insight into the difference between the two approaches, or if there’s just not much to it beyond computations.
Also, regarding Exercise 5, since the Hessian depends on first derivatives of the metric, one naively expects that, under Ricci flow, time derivatives of the Hessian will depend on third derivatives of the metric, or first derivatives of the curvature. But these are not controlled by the hypotheses of Shi’s local estimates. This complication seems to be a source of technical annoyance in the proof of Shi’s local estimates in Chow et. al. Is there an easier way to get around this issue?
Finally, could you explain why your refer to the inequality (33) as a Bochner-Weitzenbock inequality?
Thanks.
2 June, 2008 at 5:06 pm
Terence Tao
Dear Dan,
I adapted the proof of Shi’s estimates from the one in Morgan-Tian’s book, which they credit to Peng Lu. He also uses a multiplicative weight to damp the bad term but after playing around with it for a while I found that an additive weight works just as well. The basic point is that the only way that the m^th derivative can increase rapidly is if the m-1^th derivative is falling rapidly, so just about any suitable combination of the two terms will balance out this problem.
You’re right that Exercise 5 is a little more subtle than I thought. One can break the circularity by using a continuity method. We’re already trying to prove a bound of the form
. Without loss of generality, it suffices to prove this under the bootstrap hypothesis
, since one can then use “continuous induction” to remove this hypothesis (the set of all T for which these bounds hold will be both open and closed, and contain 0). But once one has this bootstrap hypothesis in hand, first derivatives of curvature become integrable in time and one can control the variation of the Hessian. (As long as T is small compared with C, we can close the induction, and C is ultimately just going to be an absolute constant.)
As for the Bochner-Weitzenbock type inequality, I am reasoning by analogy with the usual such inequality, which says that if u is some harmonic function or form, then
looks like the square of a Hessian plus lower order curvature terms. The parabolic analogue of that fact is essentially what one has here, with the curvature tensor or its derivatives playing the role of the harmonic function or form.
2 June, 2008 at 10:18 pm
Terence Tao
Oops, I realised that my proposed fix using the continuity method does not work (it requires gradient control on the large cylinder, but one only deduces gradient control on the small cylinder, so the induction does not close). But there is a different way, which is to make the r coordinate (and the cutoff eta) time-dependent, using the metric at time t rather than at time 0. Then one gets the bound on
(and now also
) from comparison geometry.
4 June, 2008 at 8:38 am
285G, Lecture 18: The structure of high-curvature regions of Ricci flow « What’s new
[…] the cylinder increasingly resembles a cone. (For instance, one can use the bound from Lemma 1 of Lecture 15, where are geodesics emanating from the infinite curvature end, to establish this sort of thing.) […]
16 July, 2008 at 2:15 pm
Classification of Almost Quarter-Pinched Manifolds « What’s new
[…] effects of that flow allow one to take a limit in (this is the Hamilton compactness theorem, see Lecture 15 from my class) and extract a smooth limit manifold of non-negative isotropic curvature (after […]
29 March, 2009 at 4:12 am
Mikhail Gromov wins 2009 Abel prize « What’s new
[…] proof of the Poincaré conjecture), the concept of Gromov-Hausdorff convergence (a version of which is also key in the proof of the Poincaré conjecture), and Gromov’s celebrated theorem on groups of […]