Gluing Harmonic Maps
Abstract.
In this paper, we consider harmonic maps from closed, two-dimensional Riemannian manifolds into a closed, Riemannian target manifold of dimension two or higher. We develop a gluing theory for such harmonic maps. In addtion, we develop the properties of this gluing map and apply them to the phenomenon of energy bubbling.
1. Introduction
There is extensive work on gluing in various contexts. Taubes [26] [25] [24] discussed gluing for anti-self-dual (ASD) connections and Yang-Mills connections. Gluing for ASD connections has also been explored by Donaldson [6], Mrowka [19], Feehan and Leness [8]. Gluing for Seiberg-Witten monopoles is discussed by Frøyshov [9] and Parker [20], and for Non-Abelian monopoles by Feehan and Leness [7]. Brendle [3] discussed gluing for yang-Mills connections. Brendle and Kapouleas [4] studied the gluing method for Eguchi-Hanson metrics. Additionally, gluing in the context of pseudoholomorphic curves is covered in Fukaya [10], Abouzaid [1], McDuff and Salamon [15], as well as McDuff and Wehrheim [17] [18]. Malchiodi, Rupflin, and Sharp [14] and Rupflin [22] considered gluing for almost-harmonic maps.
In this paper we consider two harmonic maps, and , where and are Riemannian manifolds of dimension 2 and is a Riemannian manifold. We consider under what conditions can we glue these two maps and, when the gluing map exists, what properties does the gluing map possess. In particular, we show surjectivity of the gluing map, and thus in the bubbling sequence will lie in the image of the gluing map.
1.1. Gluing of Manifolds and Maps
Loosely speaking, we are given two compact Riemannian manifolds of dimension 2, and , a Riemannian manifold , and two harmonic maps and . Suppose , then we connect and by punching holes at and , then gluing them by a neck. Denote this by , where and are parameters of neck. We glue the two maps using cutoff functions.
Here are the details: Consider gluing and whose domains are both . We denote
Since and are harmonic maps, we know they are smooth. Since is compact, we know there exists such that and (Here the norms are in the sense of the round metric).
Let be less than the injective radius of . By the upper bound of the differential of the harmonic maps, we can compute that when , we have and . Thus there exists such that and .
Consider some fixed nondecreasing smooth function satisfying:
We define our pre-gluing:
(1) |
We will need to specify which weighted norm we are using. For the reason mentioned in section 10.3 in [16], we use the same weight as in chapter 10 in [16]. Namely,
and we let the metric be
(2) |
Note that similar setup can be done for general closed Riemannian manifolds of dimension 2.
1.2. Existence of the Gluing Map
Given Riemannian manifolds and , where is closed and of dimension 2. Consider a smooth map . For , we can consider the perturbation of by under the exponential map, which we write as . We define to be the space consisting all that is in each coordinate chart, with the usual Sobolev space structure. We can define spaces similarly.
We define the harmonic map operator from the space to as
where denotes the metric on , and is the parallel transport from to along the geodesic . See Section 2.1 for details. We will deal with a second order nonlinear system of PDEs.
First, we define the space we will be working with, which is a space of pairs of harmonic maps satisfying certain conditions. These conditions will allow us to carry out a construction similar to that in chapter 10 of [16].
Definition 1.1 (Space of Harmonic Map Pairs).
Fix domain manifolds (closed, of dimension 2, with Riemannian metric) and and target Riemannian manifold of dimension no less than 2. Fix a constant . Fix points and . Let denote all pairs of harmonic maps such that
-
(1)
-
(2)
and for .
-
(3)
for the harmonic map operator , and are both surjective.
-
(4)
Furthermore, let . is surjective.
In this paper, we will need to specify the parameters of the neck in connected sum. To avoid repeating too much, we define the following symbol.
Definition 1.2 (Set of Parameter Pairs).
For any , we define to be the set of all pairs of such that , .
We will show that for pairs of harmonic maps in definition 1.1, we can glue the pair and then perturb it to get another harmonic map. This is stated as the following theoerm, which is proved in Section 6:
Theorem 1.1 (Existence of Gluing Map).
Given any precompact subset of . There exists , such that for each pair of , there exists a gluing map such that each element in is a smooth harmonic map, where is the glued manifold as defined in 2. Furthermore, for any , we can choose such that, for any , there exists satisfying
where denotes the pregluing of defined in (1).
1.3. Surjectivity of the Gluing Map and Bubbling
Consider bubbling described in [21]. By derivative estimates and facts about the bubbling phenomenon, we will eventually show the following using the result in Section 8:
Theorem 1.2 (Bubble Lies in Image of Gluing Map).
Let and be closed Riemannian manifolds and the dimension of be , let be a sequence of harmonic maps such that each bubble point only contains a single bubble. Denote the bubble points by and the corresponding bubble map by . Denote the limit modulo bubble points by . Then there exists such that, suppose for each , for some and . Then there exists such that the following holds:
Let denote the gluing map of and the maps of the bubbles with . Then there exists a subsequence of that lies in the image of the gluing map .
1.4. Outline
In Section 2, we introduce the setup of the problem. We define the harmonic map operator and compute the linearization. Then we explore the structure of the space of harmonic maps, which behaves locally like a smooth manifold. In Section 3, we introduce the implicit function theorem needed in the gluing construction. In Section 4, we explain how to choose the appropriate metric and the cutoff functions to glue the domain manifolds. In Section 5, we construct an approximate right inverse, and then get the exact right inverse using Newton iteration. In Section 6, we apply the implicit function theorem to find a way to perturb the preglued map into a harmonic map. In Section 7, we cite a way to ’suspend’ the harmonic map into a conformal harmonic map, which will be used in the proof of surjectivity. In Section 8, by derivative estimates and properties of conformal harmonic maps, we show that the gluing map constructed in Section 6 is locally surjective under certain conditions, and finally we will be able to apply this to the situation of bubbling to show that, under certain conditions, the bubble lies in the image of the gluing map. The appendix contains technical details including proof of Lemma 2.1, norm and derivative estimates, as well as related facts in elliptic PDE theory and functional analysis.
1.5. Acknowledgement
The author expresses deep gratitude to Professor Paul Feehan for his invaluable guidance and unwavering support, and to Professor Dan Ketover for his insightful comments and suggestions. Special thanks are also extended to Professor Jason Lotay, Zilu Ma, Gregory Parker, Junsheng Zhang, Xiao Ma, Liuwei Gong, and Jiakai Li for their helpful discussions. This work is also based in part on research supported by the National Science Foundation under Grant No. 1440140, while the author was in residence at the Simons Laufer Mathematical Sciences Institute in Berkeley, California, during Fall 2022 as an associate of the program Analytic and Geometric Aspects of Gauge Theory.
2. Basic Setup
2.1. The operator and its Linearization
Let’s follow chapter 10 in [16] and consider gluing harmonic maps on (The Riemann sphere) or any Riemann manifolds of dimension 2. For simplicity, we consider the case when . However, we will see that the gluing can be used for general Riemann manifolds of dimension 2.
First, let’s consider the general definition of harmonic maps. Suppose , are Riemannian manifolds and is a smooth map from the domain manifold to the target manifold . We say is a harmonic map if it satisfies the harmonic map equation. There are multiple ways of writing the harmonic map equation. One would be
(3) |
where we consider an isometric embedding and is the second fundamental from of the embedding.
However, the above will depend on the ambient Euclidean space. In particular, since we are using the implicit function theorem later, the above will make it difficult to utilize the surjectivity onto the tangent space. An alternate form would be
In coordinates, this is
where we are denoting by . We know that the above is a well defined vector, independent of choice of coordinates on and . We can also verify this by directly computing the transformation law in different coordinates.
In this paper we use another form obtained from considering coordinate charts in and computing the Euler-Lagrange equation of the energy. See section 1.1 from [13] for details. The equation is
or, equivalently,
We define the operator as
It can be verified directly by coordinate change that the above is a tensor.
Next, we consider the linearization of the above differential operator. The idea is that, for a perturbation of and , the operator will spit out a vector that belongs to a different fiber in the tangent bundle. Thus we have to pull back the vector before comparing them.
Here are the details: Given , we know that for , is embedded into where . We define the following:
(4) | ||||
(5) | ||||
(6) |
We would like to compute . Consider the path:
We use normal coordinates for to get an explicit expression for :
Consider a basis of in normal coordinates. Using we can view it as a basis of . is locally given by
where is the Christoffel symbol on .
In normal coordinates, is just . We have
We denote by and by . Since we are using normal coordinates, we have
Thus we get
(7) |
Note that in the above we are only able to compute the exact form of the operators when considering the normal coordinates at exactly that point. For certain estimates, we will need to consider the operator’s coordinate form in a single coordinate chart. When we change the coordinates, it will be different for every point in the neighborhood, so we will need some uniform bound on the change of coordinates.
Lemma 2.1 (Uniform Boundedness of Coordinate Change).
Given a Riemannian manifold , for any , consider the normal coordinates centered at , which we denote by , there exists a neighborhood of and a corresponding family of normal coordinates such that each coordinate is a normal coordinate at . Furthermore, any function of the form and has bounded norm on only depending on and .
The proof is in the appendix.
2.2. Structure of the Moduli Space
This part follows from section 4.2.4 of [5] or section A.4 in [16]. Suppose is Fredholm. It follows that the kernel and image of are closed and admit topological complements. For simplicity, we write and . So we can write , , where and are finite-dimensional and is a linear isomorphism from to .
Let be a connected open neighburhood of in . By definition we know is Fredholm. If is surjective, then by the implicit function theorem of Banach spaces we know there is a diffeomorphism from one neighborhood of in to another, such that . Furthermore, we know that
Now consider the general case when is not necessarily surjective. Consider projection of onto , we will have the derivative be surjective. Then we obtain
Proposition 2.1.
The Fredholm map from a neighborhood of is locally right equivalent to a map of the form
where is a linear isomorphism from to , and are finite-dimensional, and the derivative of vanishes at .
As an immediate corollary we obtain a finite-dimensional model for a neighborhood of in the zero set . Note that elements in are smooth by elliptic regularity (use partition of unity to reduce the case on manifolds to that on Euclidean space).
Explicitly, there exists a diffeomorphism from some open set containing , such that
and
From the above we can get a coordinate chart for by the isomorphism between and , where is the dimension of the kernel. Thus we can view the vectors in the kernel as tangent vectors of the moduli space. Furthermore, a smooth path in the open neighborhood can be represented as . Similarly, for a sufficiently small open neighborhood, we can find a smooth frame.
In the setting for theorem 1.1, since the linearizations are assumed to be surjective, together with proposition C.2 in the appendix, we know that the above can be applied.
Consider , note that for linear operators between Banach spaces, surjective operators form open sets in the sense of operator norm. Thus locally we can still find charts as above.
3. The implicit function theorem
We first state a version of implicit function theorem for general Banach spaces:
Proposition 3.1 (Implicit Function Theorem for General Banach Spaces).
(See proposition A.3.4 in [16]) Let and be Banach spaces, be an open set, and be a continuously differentiable map. Let be such that is surjective and has a (bounded linear) right inverse . Choose positive constants and such that , , and
(9) |
Suppose that satisfies
(10) |
Then there exists a unique such that
(11) |
Moreover, .
Then we can get a version of implicit function theorem that can be used in our setting:
Theorem 3.1 (Implicit Function Theorem).
Let be a Riemann surface and a Riemannian manifold. Let be a map from to . defined in (5) is a continuously differentiable map from to . Let be such that is surjective and has a (bounded linear) right inverse . Choose positive constants and such that , and
(12) |
Suppose that satisfies
(13) |
Then there exists a unique such that
(14) |
Moreover, .
4. Pregluing
From now on we consider gluing two harmonic maps from to some closed manifold of dimension no less than 2. However, we will see that the same procedure can be applied in the general case.
More specifically, consider defined in definition 1.1. Let be respectively (here we are using the projective plane as the coordinate chart for ). Furthermore, suppose is less than the injective radius of . Consider the setup in Section 1.1.
Let . Note that this definition is only for the brevity of notations. In the process of estimates, will denote the radius for polar coordinates. We will need the -small perturbations of :
(15) |
Note that for all and for all .
We set up some simplified notation. We define
for . Given such that , denote
We define
Let be the orthogonal complement of the kernel of , and define
Similarly, we have .Since the operators are small perturbations of (See proposition B.2), we know these right inverses still exist.
We will need to show that for any harmonic map is a Fredholm operator. This will be shown in section C.
and from that get an estimate: There are positive constants only depending on such that
(16) |
for all , , and . This will be shown in lemma C.3.
5. Approximate right inverse
The idea of our proof is that, first find an approximate right inverse, then we use this to find the real right inverse and apply the implicit function theorem. In this section, we will define the approximate right inverse and prove certain estimates that will be useful later.
In the definition below, note that for any pair of positive numbers such that , we have . Thus we can always shrink if necessary.
Definition 5.1.
Let be any precompact subset of . Let be the same as in C.3. For any and any , we define
along the preglued map defined by (1) as follows:
Given we first define the pair
by cutting off along the circle :
(17) |
Second, define
and note that the vector fields have the same central value :
Third, let denote a cutoff function defined as follows: for , for , and
where : is a cut-off function such that if and if .
Fourth, define by
(18) |
Let us consider why should be defined as such. We would like to have . Hence . Also, we want . Note that holds for .
For , we use coordinates for the domain, and pick normal coordinates at the point . Due to the special structure of the coordinates in the domain, after calculation, we find that the linearization is simpler than expected because certain terms in the Laplace operator cancel out. We have
where is in normal coordinates at .
Now consider . We have
Since we want , for , let , we have .
In the rest of this section, we will get estimates that will be useful in our construction of the right inverse. Let us denote the -norm (resp. -norm) on a certain region by (resp. ).
Lemma 5.1.
For any precompact subset of , we can choose small enough only depending on , such that for any and any , the approximate right inverse defined in 5.1 satisfies:
for every , where .
Proof.
In this region,
Therefore over this annulus the vector field takes values in the fixed vector space . By (7), we see that the operators are all equal to the usual Laplace-Beltrami operator. In particular, they vanish on the constant function . Furthermore, the definition of implies that in the region .
Let’s consider the polar coordinates. Given the weighted metric, we have
We also recall that, by our construction of the cutoff function, we have
where the constants are universal.
Hence, when , note that , we find
We would like to estimate the norm of under the weighted norm. For the second part, we have:
We can consider using the Sobolev embedding on the manifold, which, in our case, is the 2 dimensional sphere. First, we use the Hölder’s inequality to get:
where .
For the norm part, we want to use the Sobolev embedding on the sphere. From chapter 2 in [2] we know that for compact manifold the Sobolev embedding holds. Note that here we can treat in the same way as real-valued functions on the sphere. , where is univeral. Thus we have
if we apply lemma C.3. Here we are considering with the round metric.
We also know that
For the first part (), note that we have control over the norm of by the norm. We consider the Sobolev embedding of into where and . For every ball and every , we have
(19) |
Note that the constant comes from the extension. Thus for balls of different radius, the constant remains the same.
Thus we have the estimate:
Furthermore, note that,
Thus by choosing small enough only depending on , we have the desired result. ∎
Lemma 5.2.
For any precompact subset of , we can choose small enough only depending on , such that for any and any , the approximate right inverse defined in 5.1 satisfies:
for every , where .
Proof.
We still use the polar coordinates. Since the weight is different, here we have
and
We know that . We also know that . Again, we use coordinates for . We use coordinates for . We know that in the annulus , . We choose the normal coordinates at and we can write as , where is for the normal coordinates at . Note that in the annulus , . We have
Thus
Note that for , we have , thus . Furthermore, we have . Using the same coordinates as above ( for the domain, and normal coordinates at ), we have
Thus we know that .
Now we have
We want to estimate the weighted norm of the above formula.
We can compute that
where is a universal constant. Thus,
Let’s first estimate
where denotes the annulus . The last inequality comes from Hölder’s inequality.
For the part of the cutoff function, we have
where is a universal constant.
Before we look at the part involving , here are some thoughts about measuring the norm of the derivative of . Let us note that with the coordinate chart given by stereographic projection, there is a weight for the derivative of , since in the coordinate chart, at the points away from the origin, say , the vector is dilated with a ratio of . We have
Recall that using the coordinates from the stereographic projection of the Riemann sphere , the norm should be defined as
Note that here we cannot directly consider the range of to be . By the definition of the norm for maps mapping to vector bundles, we need to consider the local coordinates of that vector bundle. However, for , since , these two definitions are equivalent.
Now let’s come back to the part involving . Denote by . Since , we know . Thus it is easy to see that, in the annulus ,
We know that
where is universal.
Then we use the Sobolev embedding theorem for closed Riemann manifolds, as well as C.3:
From the definition of , we have
Thus we have
On the other hand,
We can consider changing coordinates . Since , we have . Also, in the new coordinate system, . We have
(20) |
where is a universal constant.
Thus we have
Similar as in the proof of 5.1, we know
Thus we get
Hence we can choose small enough only depending on to get the desired inequality. ∎
Proposition 5.1.
For any precompact subset of , we can choose and only depending on , such that for any and any , the approximate right inverse defined in 5.1 satisfies:
(21) |
for every .
Proof.
We have for each , and must prove that
(22) |
Since and , the term on the left hand side vanishes for and for . For , we can apply the previous two lemmas. The first equality is proved.
Before we prove the second inequality, we take a closer look of how the norm is defined on the weighted sphere.
We still consider the stereographic projection of the weighted (which is the connected sum of the original two s). We consider .
For , we have
Since is precompact, we know that the norm of and are uniformly bounded. Thus, there exists only depending on such that
Furthermore, since , we have , , and . Thus for we have
We know that the image of under is the same as the image of under , which is the same as the image of under . Since is compact and is precompact, we know there is a uniform injective radius only depending on .
We can choose coordinate charts as follows: We pick a set of points on and a radius such that the geodesic balls of radius and centered at those points yield a covering of . Furthermore, since is bounded, we can choose small enough that for any of these geodesic balls, say , let the radius of the image be less than the injective radius of , then considering the normal coordinates at , we have a coordinate chart for the image of . We know that this coordinate chart will only depend on .
Now on each of these coordinate charts, we can consider the vectors in coordinates, and thus talk about the derivatives in coordinates. That is how we define the space for . Since we choose , we know that in all these coordinate charts, if we consider the Riemann metric matrix of , we have where is a universal constant. Thus we know that this norm is equivalent to the norm, which is formed by using the stereographic projection coordinate chart with the weight.
Next, we consider :
By definition,
For , we can consider the coordinate change and do the same as above for . We want to show that this norm is equivalent to the norm for . Namely, for defined on (in , not ), we want to show that is equivalent to .
From the construction of the norm (details written in the case ), we know the norm is equivalent to
We can use change of variables in the integration to directly verify that this is equivalent to the norm. For example, consider and the change of variables we have
Now we have shown that norm is equivalent to norm, where is the weighted sphere. In particular, norm is equivalent to norm.
Now let’s come back to showing
From (18) and what we proved above, we know that we are only left to consider for and for .
For , when we consider the norm, for parts where there is no derivative on , the second order derivative part can be controlled directly. The other parts can be estiamted in the same way as when there’s derivative on . Thus we only need to consider the parts (note that we will have to power by , which is not written out in the formula)
For the first two parts, we use estimate for derivatives of as well as (19). For the last part, the estimate is similar to the estimate immediately above (19).
For , for parts where there is no derivative of , we can just estimate
Similar as before, the second derivative part can be controlled directly. The other parts can be controlled in the same way as when there’s derivative on .
Now we are only left with
The estimate for the first three part is the same: we use (20) and change of coordinates. The estimate for the last part is the same as estimating
The proof is complete. ∎
From the above proposition 5.1 we can construct the ’real’ right inverse:
Definition 5.2.
For any precompact subset of , we can choose only depending on , such that for any and any , we define
and we have
where only depends on .
6. Construction of the gluing map
Let us further tailor the implicit function theorem for our setting:
Theorem 6.1.
For any precompact subset of and as in definition 5.1, consider . Let denote with the weighted metric defined in the pregluing. Let , consider Banach spaces and . Let be an open subset of , be a continuously differentiable map. Suppose we have the following:
-
(1)
Consider , is surjective and has a linear right inverse such that for some constant .
-
(2)
There exists a positive constant such that , and
for all .
-
(3)
There exists some that satisfies
Then there exists a unique such that
Moreover,
We can choose to be , and in this section we will show that the conditions in the above theorem are satisfied.
First, let us estimate the norm of so that we know what should be in the third condition of the theorem.
From (1) we know that for and .
Recall that for normal coordinates, we have the Taylor expansion of the metric:
and can be written as:
For , we can choose normal coordinates at and have that all lie in the image of the normal coordinate chart by making small enough.
For , we know that is constant, so we have
For , we have:
We have . For small enough, the exponential map is a smooth isometry for . We have
Note that we may further assume is decreasing with respect to while and are increasing.
Since , and the exponential map is a smooth isometry for , we have , and
Similar to the above paragraph, we may assume is increasing with respect to .
We now start to divide into parts and estimate:
(Note that here is a function defined on instead of , which is different from the we started with. Actually it is (that we started with) composed with absolute value function)
where is increasing with respect to .
Things are the same for derivatives with respect to instead of .
where is increasing with respect to .
We can estimate from the formula of that
where here is increasing with respect to .
For , we have:
Similar as we did for , for small enough, we have:
where is decreasing with respect to and are increasing.
Since for , similar as the case when , we have:
We now divide into parts and estimate, eventually we will get the same estimate as in the case .
Now we can take a look at what we should choose to be and in the condition of the implicit function theorem. First, from the previous section, we have
So we should choose .
Since we want
while we have proved
where is increasing with respect to .
We first choose and from proposition 5.1. We know that the results of the proposition still hold if we make smaller. Furthermore, we have seen that is increasing with respect to .
For , we first require . This upper bound for only depends on .
Then we can consider the -norm of being smaller than . By making small enough only depending on , we can make as small as we like. Now we want to have
Consider . Let’s recall how we computed the linearization. In coordinates, we can write the harmonic map equation as
= . We choose normal coordinates at ,
To compute the linearization, note that the coordinate vectors are parallel along radial geodesics under normal coordinates
Also recall that
Let’s estimate :
Note that at every point, we are considering the normal coordinates at that point, so we will need to consider a coordinate change to get the real norm. By lemma 2.1, we can directly use the above formula to estimate :
note that we are using coordinate chart centered at , so we have:
For , we have , so everything does not depend on or . So for sufficiently small only depending on , we can assume .
For , note that the norm is controlled by the norm of , where the constant only depends on . Thus we can also estimate that for sufficiently small, we have .
For , we have , and we also have for sufficiently small. The rest is similar.
Thus we can apply the implicit function theorem. In particular, we know there exists a unique such that is a harmonic map.
Although we are considering s, everything can be done the same way for general closed Riemannian manifolds of dimension 2. Combining the above, we arrive at the following theorem:
Theorem 1.1 (Existence of Gluing Map).
Given any precompact subset of . There exists , such that for each pair of , there exists a gluing map such that each element in satisfies the harmonic map equation, where is the glued manifold as defined in 2. Furthermore, for any , we can choose such that, for any , there exists satisfying
where denotes the pregluing of defined in (1).
7. A trick for estimate
From [21] we know there is a trick to ’suspend’ a harmonic map to get a conformal harmonic map. Indeed, consider a harmonic map from a disk to a closed Riemannian manifold and the Hopf differential
Note that is holomorphic, and vanishes is conformal. The suspension of is defined by finding the unique solution to
and let
Then is harmonic and conformal. In [21], it’s instead of . This change is due to our desire to make target manifold of compact, so that we can apply estimates from [23]. We will also use estimates given in [21]:
For any subdomain ,
(23) |
For the circle ,
(24) |
Lemma 7.1 (Isopermetric Inequality).
(See lemma 2.4 in [21]) Let be a closed Riemannian manifold and let be a minimal surface in . Then there are constants such that if , then .
The only difference is that here we are considering instead of , but that doesn’t matter. The lemma essentially follows from the isoperimetric inequality. The next two lemmas are analogous to lemma 4.7.3 and 4.7.5 in [16]. One major difference between our case and the case of symplectic geometry is that, in the latter, there is the existence of Darboux coordinate charts, which gives better constant for the isoperimetric inequality. We will see that this will affect our choice of parameters later.
Proposition 7.1 (Propositions of Harmonic Maps).
(See proposition 1.1 in [21]) There are positive constants and , depending only on and such that
-
(1)
(Sup Estimate) If is harmonic and is a geodesic disk of radius with energy , then
-
(2)
(Uniform Convergence) If is a sequence of harmonic maps from a disk with for all , then there is a subsequence that converges in .
-
(3)
(Energy Gap) Any non-trivial harmonic map has energy .
-
(4)
(Removable Singularities) Any smooth finite-energy harmonic map from a punctured disk to extends to a smooth harmonic map on .
Lemma 7.2 (Compactness of Harmonic Maps).
(See lemma 1.2 in [21]) Let be a sequence of metrics on converging in to , and a sequence of -harmonic maps with . Then there is a subsequence of , a finite set of ”bubble points” , and an -harmonic map such that
-
(1)
in uniformly on compact sets in ,
-
(2)
the energy densities converge as measures to plus a sum of point measures with mass :
Now we are ready to prove the following lemmas. These are analogues to lemma 4.7.3 and 4.7.5 in [16].
Lemma 7.3.
Let be a compact Riemannian manifold of dimension 2. Then there exist constants such that the following holds:
If and is a conformal harmonic map satisfying
then
-
(1)
-
(2)
for .
Proof.
For every conformal harmonic map on , we have
As long as where is as in the previous lemma, and is small enough as in lemma 7.1. In other words, we can choose such that and .
Without loss of generality, we consider , where is a conformal harmonic map and . Let . Define such that
This is a conformal map composite with , so it is still a conformal harmonic map.
For define by
Let , we have
Thus
Now consider the function
Abbreviate . Then
On the other hand,
Thus
for .
To prove the seond part, we will need the following estimate:
First assume that
Then we have
Next assume
and denote . Then . Hence
With the above estimate, we can prove the second part of the lemma. For the above estimate implies
Fix a point such that and , we have
We have a similar estimate for . Thus
for and and . Thus the proof is complete. ∎
Lemma 7.4.
Let , , , and as . Let be a sequence of conformal harmonic maps. Suppose
-
(1)
, where is a conformal harmonic map, and the convergence is uniform on compact subsets of
-
(2)
, where is a conformal harmonic map, and the convergence is uniform on compact subsets of
-
(3)
Then,
-
(1)
-
(2)
Proof.
Abbreviate
By choosing small enough (and large enough), we can let the energy be arbitrarily small and thus apply the previous lemma.
For such that , we know . Let in the previous lemma, there exists and such that
Let ,
Take the limit , we get .
To prove the second assertion, fix an arbitrary . We only need to prove that, when is small enough, there exists an integer ( can depend on ) such that,
First choose and as in the first part of this proof. Then choose small enough such that and
Then choose large enough such that, for , we have
Now suppose and , then
The proof is complete. ∎
8. Surjectivity of the gluing map
The theorems in this section are analogues to section 10.7 in [16]. From now on, we always assume our target manifold to be closed.
Theorem 8.1.
Fix two constants and . Consider any precompact open subset . Then, depending only on , we can choose positive constants , , and such that the following holds:
Given , , and is a harmonic map whose domain is the glued sphere, satisfying , where . Then belongs to the image of the gluing map.
Proof.
Fix a pair and an element as in the statement of the theorem (We can let be smaller later if necessary). Let us denote by
the set of all pairs that satisfy
(25) |
for some vector field . (Note that by lemma D.1 we know , and hence by choosing small enough only depending on , there is at most one such vector field .) Since is fixed, we know is an open set.
We must find a constant depending only on with the following significance: if there is an element such that then there is an element such that the corresponding vector field belongs to the image of the operator . To prove this, note that is locally a manifold of dimension , where is some finite number. From lemma D.3 we know that is also the dimension of the kernel of for such that and small enough only depending on .
Next we choose a smooth framing
over . We can make a small enough open neighborhood (how small only depends on ) of and choose framing using the coordinate chart. Define
where is the pregluing (). We show that this is a framing of for for sufficiently small only depending on . In fact, by lemma D.3 we know that, in this case, the linear map
is an isomorphism from the kernel of to the kernel of for and , and that there is a uniform bound on the inverse of this isomorphism only depending on . Hence there are positive constants and only depending on such that the following holds whenever : If and , then
Hence is a framing. Increasing if necessary, we also have that every smooth path in satisfies the inequality
where denotes the Levi-Civita connection on . To see this, assume first that is any smooth vector field along . Then by lemma D.4 the -norm of can be estimated by a constant times . Now apply this inequality to and use lemma D.3 (), the result of lemma D.2, as well as the proof of lemma D.2 ().
Define the smooth map by
where and is the unique vector field along that satisfies (25). Thus is the coordinate vector of the projection of onto the kernel of along the image of . We would like to find where vanishes. We have
Hence our assumption implies that there is such that , for some small we can choose. To find an actual zero of we aim to apply the implicit function theorem. So we will need to estimate the inverse of .
Let us denote by
the isomorphism given by the framing , that is
Then, by definition of , we have
for . We shall prove that there is a constant such that
for every . Note that the above is the operator norm. Since is an isomorphism, we can apply proposition 3.1 for after proving the above inequality.
Along a smooth path . Denote and let be the unique smooth family of vector fields along satisfying
Then we have
where and . Differentiate to obtain
Thus
Combining the above, we have
Since , we have
On the other hand, we have the pointwise estimate
Similarly, we have an inequality for the first and second derivatives of in terms of the first and second derivatives of for a fixed . Hence we have
For the remaining term, we claim that
We proved a similar result for instead of in the appendix. Here we can also use D.2 in that proof. Combining these we get an estimate for the commutator and hence for . Note that , thus we get an estimate for the commutator of the operator with .
Thus for we obtain
Hence we have proved
Now we apply the implicit function theorem. ∎
With the previous result, now we can consider when there is a single bubble for a sequence of harmonic maps on .
Theorem 8.2.
There exists such that if we fix and , consider any precompact , then there exists an and only depending on such that the following holds:
Given , , and if is a harmonic map whose domain is the glued sphere, satisfying
where . Then belongs to the image of the gluing map.
Proof.
From the previous theorem, we only need to prove that there exists and such that for every , with the given conditions, we have where .
Suppose, by contradiction, that the assertion is wrong for some constant . Then there are sequences
such that diverges to infinity, is a harmonic map on the glued sphere with parameters , and
and, for every ,
By passing to a subsequence, we can assume converges in to a pair . Since the norm of first order derivatives are bounded, no bubbles form. Furthermore, away from the neck, converges uniformly on compact subsets to on the complement of a finite set of bubble points.
By assumption, , . By unique continuation of harmonic maps, this implies that on . Similarly, converges to on the complement of a finite set of bubble points.
First, we would like to prove that converges to uniformly on compact subsets of , and converges to , uniformly on compact subsets of . Let’s take for example. To avoid writing too many superscripts, we write as below. Since and are harmonic maps, they satisfy the harmonic map equations:
where
First pull back the vector on the left hand side of the second equation and then subtracting the two equations, we will get a quasilinear system of second order elliptic equations for . Here we cite the result in [12], section 8.4:
Theorem 8.3.
Let and that satisfies the system:
Suppose the following inequalities hold for , and the arbitrary :
(26) | |||
(27) | |||
(28) | |||
(29) |
where is a sufficiently small quantity determined only by , and as . Also suppose the above constant satisfies
Then, for arbitrary , the quantity is bounded in terms of distance from to , and mes (measure of ).
Now let’s show that the elliptic system for actually satisfies the conditions in the theorem. We construct the system as follows: first, we write out the equations for at each point using the normal coordinates at that point. Then, we consider disks of small enough radius and do coordinate change to get the equations under the same coordinates for the small disk.
Consider normal coordinates at , the satisfies:
where
Then, consider the coordinate change. Suppose this point is in the disk centered at , denote the normal coordinates at as . We can use our construction in lemma 2.1. We have .
For the Laplacian part,
In the above, for the parts where does not take derivatives, we can make the norm of arbitrarily small by taking large enough . Thus is satisfied. For the parts where the change of coordinates does not take any derivatives, note that
Thus the part with second order derivative remains the same after the coordinate change, and the rest belongs to in the above theorem, which has the required bound.
The only part that is slightly less obvious is the part:
Note that for , we can make the disk small enough such that is arbitrarily small. On the other hand,
Thus we also have control when . When , the term belongs to part in the theorem. Other terms belong to part. Furthermore, all conditions in the theorem are met.
For the parts involving Christoffle symbols, it is probably the easiest if we replace the by . We can do this since originally this is the normal coordinates. Then do the coordinate change. We can make the norm of these terms arbitrarily small since norm of can be arbitrarily small. Thus the condition in the theorem is still satisfied.
From the theorem, we get a bound for the maximum of . Thus there can be no bubble anywhere other than the neck. Furthermore, by the main estimate 3.2 in [23], we get a uniform bound for norm of away from the neck for any .
Next, we consider the neck. Fix , Consider restricted to . We know this is a harmonic map, thus we can apply the trick of making it into a conformal harmonic map, which we denote by (See section 7).
We would like to apply lemma 7.4, so we have to verify that the conditions are met. First, we show that converges to some conformal harmonic map , uniformly on compact subsets of . Let’s revisit how the conformal harmonic maps are obtained. We are given and they converge to on every compact subset. Then we consider the weighted sphere which represents two spheres glued together. The sphere for and is represented in the weighted sphere by . Then we define the by setting
on the disk of a fixed radius and centered at , where
Note that here we are simply considering the canonical norm on the disk. The process is similar for . We would like to show that converge to on compact subsets. We may not be able to do this directly because we have converge on the (standard) sphere while is defined on the disk. A direct esitmate will encounter multiplication by to some exponent. One possible way around this is to use compactness for the conformal harmonic map sequence and use energy estimate by the energy of the original harmonic maps to show that there exists a converging subsequence.
Here is the argument: Since the energy of the conformal harmonic maps are controlled by the energy of the original harmonic maps, we know the energy is bounded. By Uhlenbeck compactness, there exists a subsequence, which we still denote by , that converges in to some limiting harmonic map on every compact subset of . Since the energy is uniformly bounded, we know that is a removable singularity, so is a harmonic map on . We have to show that there is no bubble point. We know that the maps restricted to , which are the ’s, have no bubble point on . By the energy estimate in [21], we know
for any compact .
We also need to consider on . Consider any compact subset of , The corresponding shrinks as . The energy is conformal invariant. Thus, since ’s have no bubble point, we know ’s have no bubble point. Denote the limit as . We have defined on . By the property of removable singularities, we know is a harmonic map on .
Furthermore, by looking at , since the convergence is in the sense of , we know the limits and will also be conformal.
To apply lemma 7.4, we still need to show that . First, given a small positive number that only depends on the target manifold, we will need to show that the energy will be below the for small enough and large enough . We can get this from the conditions.
Then we know from the lemma . We know from the construction of the pregluing that, for the pregluing , we have
Thus for large enough , there exists a unique small vector field along such that
We claim that
We can focus on the ’neck’ part, which is . Since is conformal, we know is harmonic and conformal. The neck part is transformed to , where . This sequence of harmonic maps converge to in all derivatives, uniformly in all compact subsets.
We know from lemma 7.4 that there exists such that for any ,
for sufficiently large. Since
thus, if we first focus on , we have
For any such that , we have . Thus
This shows that
We know from the construction of the pregluing that
Thus we have the estimate for when . The estimate for when is similar. Note that here can be slightly larger than 2.
Next, we try to show
First, we have to show
Same as before, consider . We have . Note that although logarithm is multivalued, it doesn’t matter in our case since we only need to consider the local values of the derivatives. To write out the calculations explicitly, we adopt another expression:
Let . By direct calculation
To estimate the second order derivatives of and , it is easier to look at and , since derivatives with respect to and can be obtained by linear combination (both first and second order derivatives). From this, we know that
We would like to show that
First, consider the term
From previous estimates, we have
Furthermore,
Thus
We will get the desired result if we let .
Next, consider the term
We have
We divide the annulus into parts as follows: We are considering , here we can change to for any fixed constant and it won’t affect the argument, since we only care about the limit as . We choose depending on and , such that . By Hölder’s inequality, we have
where we abbreviate as . For the inequality to hold, we will have to choose carefully. We must satisfy the following:
For the part , we have
Here we have used the main estimate 3.2 in [23] (This only allows us to estimate a real subset, but this is not essential. We can consider a stripe wider on both sides by .), as well as the conformal invariance of energy.
Again, using Hölder’s inequality,
From previous part of the proof,
if we choose small enough such that .
Furthermore, we know
On the other hand,
as long as . Note that this holds even when due to the way we cut the rings. The constant only depends on .
Combining the above, we have
We consider
From previous argument, we need
Let
We can write , where . We will need to show that there exists such that
That is
We only need to show
This is obviously true.
Thus we have
where . Since
We have shown
We still have to consider the pregluing. This estimate has already been done in the section ’Construction of the gluing map’.
For the part , we can use symmetry.
Thus the proof is finished. ∎
9. In the case of bubbling
We know that bubbling can occur, for example, from [21] and [23]. Let’s now consider the case when a bubble occurs and try to apply the gluing theorem. For the bubbling, we can use the same construction as in [21]. We can take the sequence that ”bubbles off” and glue the limit. Denote the bubble in the limit as and the rest of the limit as . We can glue and . Note that in the construction, we have control of the energy on the neck. We can follow the same procedure as in [21] to show that the conditions of lemma 7.4 is met. Then we can use a proof that is similar to the proof of sujectivity to show that we can choose a subsequence of the bubbling and that sequence, except for finitly many maps, will lie in the image of the gluing map. Thus we arrive at theorem 1.2.
Appendix A Uniform Boundedness of Coordinate Change
Consider a smooth Riemannian manifold and a point . By the uniformly normal neighborhood lemma, there exists a neighborhood containing and such that
-
(1)
For all , there exists a unique geodesic of length less than joining to . Moreover, is minimizing.
-
(2)
For any , , and is a diffeomorphism on .
Consider such a uniformly normal neighborhood of and any point , there exists a unique smooth geodesic such that and .
Consider a fixed normal coordinate chart centered at . Denote the corresponding coordinates of by . Note that is equal to the coordinates of . Consider a vector and let for be the parallel transport along . We have the following system of ODEs:
Consider , we can transform the above system into a system of first order ODEs. Thus we know will be a smooth function of , , and . Here, . Denote the parallel transport by
(30) |
In coordinates (normal coordinates at we fixed above) we have:
Equivalently, we can write as
Note that for the matrix, the -th column is simply written in coordinates, so the matrix is a smooth function of . Also note that is the identity matrix. We can write where is some matrix whose norm goes to uniformly as goes to . Similarly, we can control .
Now we are ready to prove the following lemma:
Lemma 2.1 (Uniform Boundedness of Coordinate Change).
Given a Riemannian manifold , for any , consider the normal coordinates centered at , which we denote by , there exists a neighborhood of and a corresponding family of normal coordinates such that each coordinate is a normal coordinate at . Furthermore, any function of the form and has bounded norm on only depending on and .
Proof.
Consider the uniformly normal neighborhood of as above. Pick an orthonormal frame at . For any point , consider the parallel transport along the unique minimizing geodesic from to , then we get an orthonormal frame for all points in . Then we consider the normal coordinates determined by these frames. Now we have constructed the family of coordinates. Next we prove the properties.
Consider a point and two normal coordinates: One centered at , which we denote by , and the other centered at , which we denote by . We know that from to there exists a unique minimizing geodesic , and the vector written in coordinates at is exactly the coordinate . Similarly, we can find from to and written in coordinates at is the coordinate .
defined previously gives us relation between the frame at and the frame at . We can make smaller such that the closure of is a subset of a uniformly normal neighborhood. Thus we know and are smooth functions of , and the norms of and are bounded on for any nonnegative integer .
Suppose . Note that
Thus . We know that is the coordinates at of .
Consider in coordinates at . We know and . . Also, , as the solution of the geodesic equations, is a smooth function of the initial values and . Furthermore, denote the geodesic from to by , we have , where each entry in is a smooth function of . Thus is a smooth function of and . That is, is a smooth function of and . Shrinking if necessary, we know and have bounded norm. We also know the higher order derivatives of with respect to have bounded norms. The derivatives of with respect to can be obtained iteratively in terms of lower order derivatives and the derivatives of with respect to . The proof is complete. ∎
Appendix B Perturbations in the Pregluing
First, let’s identify the Sobolev spaces regarding the perturbed map with the space regarding the original map. Namely, we identify (resp. ) with (resp. ). For simplicity, we only write out the case for . The case for is completely the same.
Proposition B.1 (Equivalence of Sobolev Spaces under Perturbations).
For any precompact subset of and , there exists such that for any , we have:
(\romannum1) .
(\romannum2) .
where the constants in the equivalence relation only depend on .
Proof.
Recall . For , we have . For , whereas . For any , for , we can use parallel transport from to to get (For we can just take ). In other words:
where is parallel transport along , is a geodesic from to . Since is bounded by , for small enough , there always exists such a unique geodesic , whose image is a subset of .
Now we have defined the map, we have to show that the map will induce the isomorphisms. First we have to show that this map maps (resp. ) to (resp. ) functions, then we have to show that after identifying the two spaces, the two norms are equivalent. We know the map is identity if we only consider . Recall that for the norms, we consider a finite cover of such that on each set of the cover there is a coordinate chart for and the image of that set has a coordinate chart on . We can choose a cover such that there is a coordinate chart in the cover that contains . For small enough , this chart will contain all . We can choose the cover such that all other charts in this cover will only contain . Then we only have to consider the chart containing . We also make the small enough so that, for this chart, we can choose the coordinates on to be the normal coordinates at .
Now we can only consider the chart centered at . First, we identify the and spaces (both for and ) with and . We construct the map by parallel transport along the geodesic ending at . Recall the parallel transport map defined in 30. Since we can write where the norm of goes to as , we have
for small enough.
Next, we consider the case for . We can consider a fixed cutoff function on this chart such that the function is equal to for all for small enough. The section multiplied by this cutoff function together with the parts on other charts will be an equivalent norm. This way, we can use approximation by compactly supported smooth sections for . For a smooth section with compact support ,
Note that , where can be or . we want to show that the norm of is bounded by some constant that does not depend on or . This is obviously true for . For , we want to show the norm of is bounded:
By definition of , we have
Since the norm of is bounded by some constant times , we know we can control the norm of .
For the second order derivatives, we have
We only need to show that , where only depends on . Again, this is obviously true for , so we only need to consider . We write out :
The only term that matters is . We can get by direct computation:
Then we have
by the Sobolev embedding.
So far we have shown that . The proof for is the same. For the other direction, consider . For small enough so that the entries of are sufficiently small, exists and is just a function of the entries of . Thus we have finished the proof. ∎
Proposition B.2.
For any precompact subset of and any , there exists such that for any and any , under the identification in proposition B.1,
Proof.
Still, we only have to consider the chart centered at and . As in the proof of the proposition B.1, let and be sections of and respectively such that they are the same after the identification. We want to prove that for small enough, for any such pair of sections. We have
Consider the right hand side. For the first part, by (7) and lemma 2.1, we only need to control . In the proof of proposition B.1, we have , where the norm of converge to 0 as converge to 0. So the second order derivative part of can be controlled. For the first order derivative part, we can use Sobolev embedding to control the norm (of first order derivative), where . Then consider its multiplication with the characteristic function of the neck, and use Holder’s inequality. Since the measure of the neck converge to 0, we can control the norm. The second part can be controlled similarly. ∎
Appendix C Apriori Estimates for the Differential Operators
We try to get an estimate for the operators .
Proposition C.1.
Let be a smooth map from to . Then defined in (6) is locally a second order strongly elliptic system.
Proof.
Lemma C.1 ( estimate).
Consider any precompact subset of . For every and , we have
where only depends on .
Proof.
Essentially we only need to show that .
Lemma C.2.
Let and be Banach spaces and be a bounded linear operator. Suppose is compactly embedded into and
for all . Then is semi-Fredholm. That is, the kernel of is finite dimensional and the range of is closed.
Proof.
First, if we choose any sequence in the closed unit ball of , by compact embedding of in , we can choose a converging subsequence in . Furthermore, in , the norm in is bounded by norm in , thus we get a converging subsequence in . Thus the closed unit ball of is compact, which means that is finite dimensional.
Next, we prove that the range of is closed. Consider any sequence such that . We would like to show there exists such that . Since is finite dimensional, there exists such that . Furthermore, we know that is closed. Without loss of generality, we can assume that for all .
First, we assume that the sequence is bounded. Since is compactly embedded in , we can choose a subsequence, still denoted , that converges in . Then
for all positive integers . Thus converges in , thus we can find as the limit and we will have . In fact, we know that .
we show that is bounded. If not, then by choosing a subsequence, we can assume without loss of generality that . Thus
Consider the sequence and apply the above argument and get such that and . However, we have , a contradiction. ∎
Proposition C.2.
Let be a closed Riemannian manifold of dimension 2 and another Riemannian manifold. Let be a smooth map from to . Then defined in 6 is semi-Fredholm.
Proof.
We know that is compactly embedded into . Furthermore, same as in C.1, we have the estimate
for any . Thus we can apply the above lemma to and the proof is complete. ∎
The following is an analog of lemma 10.6.1 in [16]:
Lemma C.3.
Consider any precompact subset of . Then there are positive constants and only depending on such that, for all and , the following holds for :
(\romannum1) For every , we have
(\romannum2) For every , we have
(\romannum3) For every we have
Proof.
Note that the part for and the part for are symmetric, so we only need to consider one, and the other can be proved in the same way. Let’s consider . For any , note that only differ from in . In , we know that and are both close to , so we can do a parallel transport of to get . From lemma C.1 we know that
From the previous section we know that the first estimate is proved.
To prove the remaining estimates we consider the following abstract functional analytic setting. Suppose we have a surjective Fredholm operator
of index between two Banach spaces. (Think of the case , , and ) We assume is equipped with an inner product and denote the corresponding norm by
(Think of the inner product or ) We assume further that there are positive constants and such that
(31) |
for every (This is true because the or norm is controlled by the norm) and
(32) |
This holds in our setting because the kernel of is finite dimensional. Denote by the right inverse of whose image is the orthogonal complement of the kernel with respect to the above inner product. (Note that here is just a subspace of the Hilbert space . We can consider , which is finite dimensional, and its orthogonal space in , which we denote by . It is easy to show that is closed (in ), , and . Thus we have . Then we can construct the the same way as for Hilbert spaces)
Now we prove the norm of is bounded: Since we already have , we can consider on and we know the inverse will be bounded by the open mapping theorem.
Now suppose that is another bounded linear operator ( for example) such that
Since we have and so is surjective with right inverse . However, we wish to understand the right inverse whose image is the orthogonal complement of the kernel of .
As a first step we observe that, if , then and . Hence
(33) |
Since , we find . The last inequality holds because is the orthogonal projection of onto the kernel of (Note that the image of is ). But we saw above that . Hence
This proves (\romannum2).
Now assume that is orthogonal to the kernel of . Let be an orthonormal basis of and consider the basis of defined by
(The map is an isomorphism between the two kernels. We can see that since in the above argument, we have, if , then and . Note that the index of Fredholm operators stays the same if we have for fixed and small enough . Since the index is the same for and , we know the map is surjective.)
Since we have
Moreover, and hence, by (33),
Combining this with the previous identity we find
Hence
If and we deduce that for every that is orthogonal to the kernel of .
Now recall that and that is surjective, so that runs over all elements in . It follows that there is a constant such that
How small must be chosen depends only on the operator norm of and the constants . This finishes the proof. ∎
Appendix D Derivative Estimates
First, we consider Sobolev embeddings on the weighted sphere. We show that the constant in the inequality does not depend on or , which will be needed in certain proofs. The following lemma is an analog of lemma 10.3.1 in [16].
Lemma D.1.
Proof.
We let be small enough as in section 4.
For , the metric is the Fubini-Study metric on . Each point is contained in a disc of radius . Then we get the result from the usual Sobolev embedding.
For , we can consider the coordinate change and the result follows in the same way as above. ∎
The following two lemmas are analogs of lemma 10.6.2 and lemma 10.6.3 in [16]. In this subsection, we always assume that the target manifold is closed.
Lemma D.2.
Fix two constants and . Consider any precompact . Then, depending only on , we can choose positive constants and as well as an open set containing the closure of such that the following holds:
Fix a pair , let be a smooth path and be the corresponding path of preglued maps. Let be a smooth family of sections of the pull-back bundle along , that is, for all . Then
for .
Proof.
First, we simplify by not considering the pregluing map. Suppose we have a path . Thus, let
be a smooth map, denote by the corresponding family of linearized harmonic map operators, and let be a smooth family of sections of pull-back bundles. Denote by the right inverse whose image is the -orthogonal complement of the kernel of . Define by , .
We first attempt to get the estimate
(34) |
We will need the following:
(35) |
To prove the above inequality, we would like to make sure that the constant does not depend on the choice of the path. In fact, we can revisit lemma C.1, and get . Since , we have .
From the proof of boundedness of in lemma C.3), we know there is a uniform bound since is precompact.
We can calculate
Thus
From the above inequalities,
Now suppose that is a path in the kernel of (the existence is guaranteed by ODE theory) such that . Then, since ,
Moreover, , hence
Now choose an orthonormal (in norm) frame of (First get linear independence by ODE theory and then use Gram-Schmidt process) where denotes the Fredholm index of . Then
Next, instead of considering , we must consider the pair and we are interested in the operators on the preglued maps . Let be a family of sections of the pull-back bundle, denote by
the section of pull-back bundle along and corresponding to as in (17). Abbreviate , , and let , .
We verify that in this case we still have the above inequalities:
(36) |
Under proposition B.1, we know
and we have (Use Hölder’s inequality and note that the integration of characteristic function of the domain, which is the area, is bounded uniformly).
The uniform bound can be found as in the proof of lemma C.3.
For the norm, it is easy to show the equivalence under perturbations as in proposition B.1. For , the equivalence is already proved in that proposition. Thus we get the bound independent of or .
The proof is the same as before.
Furthermore, we have:
Hence, by what we have proved,
The pregluing map defined as in (18) is a bounded linear operator (boundedness can be shown using the estimates in the proof of proposition 5.1, and the bound only depends on .) that commutes with the covariant derivative along every path (because the cutoff function depends only on the points in the domain and so is independent of ). Hence
This proves the first inequality of the lemma. For the second inequality, we go back to (D). The critical part is and .
For , we have to estimate
where .
The rest is same as before. ∎
Denote the pregluing of two harmonic maps by .
Lemma D.3.
Fix any and any , for every precompact subset of , we can choose positive constants only depending on such that the following holds:
For any , let denote the pregluing , and , then
(37) |
for every . Thus the operator is an isomorphism from the kernel of to the kernel of .
Proof.
Similar as in (36), we have
We shall prove the estimate
for and .
We have
for . To prove this, first note that the kernel is finite dimensional and thus the norm of and on an open set of the ball control the norm on all of (Note that they cannot be identically zero on any open set). Then we must find a uniform bound for all . We can show that the bound is continuous on by considering of for any smooth path on as in the proof of D.2.
Furthermore, we have from the proof of proposition 5.1. Combining the above inequalities, we get
Now we prove the estimate
Abbreviate . Note that by construction, vanishes outside and .
To estimate for , recall
for . Then is given by
Differentiating with respect to , we get
Solve the second equation for and insert into the first equation, then set , we get
(38) |
where , and
for . This shows that the norm of in is bounded by times the norm of . The area of the annuli (in the weighted metric) are bounded by . Thus
For , the same can be proved by direct computation, or we can consider the coordinate change and use symmetry.
To show that the operator is an isomorphism as stated in the lemma, we also show
Again, we only have to consider and , and we can proceed in the same way as above.
Then, together with previous estimates regarding and , we know
∎
Lemma D.4.
Fix any and any , for every precompact subset of , we can choose positive constants only depending on such that the following holds:
For any , let denote the pregluing , and , then
for every .
Proof.
can be calculated as in the proof of lemma D.3. Note that for , , thus . For , , thus is also the identity map. For , , thus . For these parts, the -norm can be estimated by .
We are now left with and .
For , with the same calculations in the proof of lemma 10.5,
where , , , and
We would like to control
where denote all possible combinations of and where .
Now consider . From the proof of lemma D.3, we know
Furthermore, from the formula for , we know
Thus the estimate holds.
For , we can change the coordinates and then the proof is the same (Note that we can calculate that the two coordinates are equivalent in the sense of Riemannian metric, and the norm for is equivalent to the norm of on ). ∎
References
- [1] Mohammed Abouzaid “Framed bordism and Lagrangian embeddings of exotic spheres” In Ann. of Math. (2) 175.1, 2012, pp. 71–185 DOI: 10.4007/annals.2012.175.1.4
- [2] Thierry Aubin “Nonlinear analysis on manifolds. Monge-Ampère equations” 252, Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences] Springer-Verlag, New York, 1982, pp. xii+204 DOI: 10.1007/978-1-4612-5734-9
- [3] Simon Brendle “On the construction of solutions to the Yang-Mills equations in higher dimensions”, 2003 arXiv: https://arxiv.org/abs/math/0302093
- [4] Simon Brendle and Nikolaos Kapouleas “Gluing Eguchi-Hanson metrics and a question of Page” In Comm. Pure Appl. Math. 70.7, 2017, pp. 1366–1401 DOI: 10.1002/cpa.21678
- [5] S.. Donaldson and P.. Kronheimer “The geometry of four-manifolds” Oxford Science Publications, Oxford Mathematical Monographs The Clarendon Press, Oxford University Press, New York, 1990, pp. x+440
- [6] Simon K. Donaldson “Connections, cohomology and the intersection forms of -manifolds” In J. Differential Geom. 24.3, 1986, pp. 275–341
- [7] Paul M.. Feehan and Thomas G. Leness “ monopoles. III: Existence of gluing and obstruction maps” 91 pages, arXiv:math/9907107
- [8] Paul M.. Feehan and Thomas G. Leness “Donaldson invariants and wall-crossing formulas. I: Continuity of gluing and obstruction maps” 86 pages, arXiv:math/9812060
- [9] Kim A. Frøyshov “Compactness and gluing theory for monopoles” msp.warwick.ac.uk/gtm/2008/15/ 15, Geometry & Topology Monographs Geometry & Topology Publications, Coventry, 2008
- [10] Kenji Fukaya, Yong-Geun Oh, Hiroshi Ohta and Kaoru Ono “Kuranishi structures and virtual fundamental chains”, Springer Monographs in Mathematics Springer, Singapore, 2020, pp. xv+638 DOI: 10.1007/978-981-15-5562-6
- [11] David Gilbarg and Neil S. Trudinger “Elliptic partial differential equations of second order” 224, Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences] Springer-Verlag, Berlin, 1983, pp. xiii+513 DOI: 10.1007/978-3-642-61798-0
- [12] Olga A. Ladyzhenskaya and Nina N. Ural’tseva “Linear and quasilinear elliptic equations” Translated from the Russian by Scripta Technica, Inc, Translation editor: Leon Ehrenpreis Academic Press, New York-London, 1968, pp. xviii+495
- [13] Fanghua Lin and Changyou Wang “The analysis of harmonic maps and their heat flows” World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2008, pp. xii+267 DOI: 10.1142/9789812779533
- [14] Andrea Malchiodi, Melanie Rupflin and Ben Sharp “Łojasiewicz inequalities near simple bubble trees”, 2023 arXiv: https://arxiv.org/abs/2007.07017
- [15] Dusa McDuff and Dietmar Salamon “-holomorphic curves and symplectic topology” 52, American Mathematical Society Colloquium Publications American Mathematical Society, Providence, RI, 2012
- [16] Dusa McDuff and Dietmar Salamon “-holomorphic curves and symplectic topology” 52, American Mathematical Society Colloquium Publications American Mathematical Society, Providence, RI, 2012, pp. xiv+726
- [17] Dusa McDuff and Katrin Wehrheim “Smooth Kuranishi atlases with isotropy” arXiv:1508.01556 In Geom. Topol. 21.5, 2017, pp. 2725–2809 DOI: 10.2140/gt.2017.21.2725
- [18] Dusa McDuff and Katrin Wehrheim “The topology of Kuranishi atlases” In Proc. Lond. Math. Soc. (3) 115.2, 2017, pp. 221–292 DOI: 10.1112/plms.12032
- [19] Tomasz S. Mrowka “A local Mayer-Vietoris principle for Yang–Mills moduli spaces”, 1988 URL: https://search.proquest.com/docview/303683572
- [20] Gregory J. Parker “Gluing -Harmonic Spinors and Seiberg-Witten Monopoles on 3-Manifolds”, 2024 arXiv: https://arxiv.org/abs/2402.03682
- [21] Thomas H. Parker “Bubble tree convergence for harmonic maps” In J. Differential Geom. 44.3, 1996, pp. 595–633 URL: http://projecteuclid.org/euclid.jdg/1214459224
- [22] Melanie Rupflin “Lojasiewicz inequalities for almost harmonic maps near simple bubble trees”, 2022 arXiv: https://arxiv.org/abs/2101.05527
- [23] J. Sacks and K. Uhlenbeck “The existence of minimal immersions of -spheres” In Ann. of Math. (2) 113.1, 1981, pp. 1–24 DOI: 10.2307/1971131
- [24] Clifford Henry Taubes “A framework for Morse theory for the Yang–Mills functional” In Invent. Math. 94, 1988, pp. 327–402 DOI: 10.1007/BF01394329
- [25] Clifford Henry Taubes “Self-dual connections on -manifolds with indefinite intersection matrix” In J. Differential Geom. 19, 1984, pp. 517–560
- [26] Clifford Henry Taubes “Self-dual Yang–Mills connections on non-self-dual -manifolds” In J. Differential Geom. 17, 1982, pp. 139–170