On the extended Bogomolny equations on with real symmetry breaking
Abstract
In this paper, we construct solutions to the extended Bogomolny equations on with certain boundary conditions and asymptotic conditions.
Let be the coordinate of . Roughly, both the boundary condition and the asymptotic condition say that a configuration (variables in the extended Bogomolny equations) approaches to certain model solutions when and resepctively. The boundary condition () is called “generalized Nahm pole boundary condition” and the asymptotic condition () is called “real symmetry breaking condition”.
For each triple of polynomials with complex coefficients with , and are co-prime, are monic, we construct a solution. This solution should be thought as an analog of the instanton solutions that Taubes created in [20] or the solutions that Dimakis created in [3], while their solutions satisfy a different asymptotic condition as .
The idea of the construction is almost identical with Dimakis’ construction in [3]. The main difference
is: We rely on the newly constructed model solution that satisfies the new asymptotic condition as the starting point.
Section 1 is a breaf introduction on background settings and terminologies. Section 2 constructs the model solution. Section 3 uses an analog of Dimakis’ argument to construct more solutions based on the model solution.
Acknowledgement
The author thanks Panagiotis Dimakis for explaining his work in [3] and suggesting the author to work on this topic. The author is also very grateful to Siqi He for carefully and patiently explaining many technical details in his paper (with Rafe Mazzeo) [11]. The author also thanks Clifford Taubes and Rafe Mazzeo for many helpful discussions and useful advice.
1 Introduction
In this section, we set up the basics for the extended Bogomolny equations. We need some preparation in linear algebra, which are summarized in appendix A. Readers are supposed to be familiar with them.
1.1 The extended Bogomolny equations
Suppose . Suppose and are coordinates of , where being the one for . We use to represent the partial derivatives. Sometimes we treat as and let be its coordinate.
One way to introduce the extended Bogomolny equations on is as follows: It is a set of equations on configurations. Each configuration is a collection of 6 -valued functions on . Here are the equations:
(1.1) |
Remark: If the background metric on is not the Euclidean metric, then there should be additional terms in the extended Bogomolny equations that come from the metric. But this is not what we focus on in this paper.
These equations are gauge invariant. Moreover, all bullets of (1.1) except the first one are also gauge invariant. Here an (or ) gauge transformation is represented by an -valued (or -valued) function that sends
to
An easy way to check that the gauge invariant properties of the equations is as follows: Define three operators acting on sections of a trivial bundle (that is to say, the bundle is just the trivial bundle but all the operators keep the structure):
Then all but the first bullet of the extended Bogomolny equations can be written as
which is clearly gauge invariant (as commutators of operators).
The first bullet of the extended Bogomolny equations is equivalent to
where represents the ad-joint. Then it is obviously gauge invariant. It is not gauge invariant in general because the operator in the equation relies on the structure.
It is proposed that the extended Bogomolny equations on are related with certain type of Kapustin-Witten equations, which are further conjectured to be related with the Jones polynomial of knots. The details of this proposal may be found in [5][21][22] [23] [24] . Some relevant studies include [2] [3][4][8][9] [10][11] [15] [16] [17] [18] [19] [20] . The author does not attempt to list all the relevant references here.
According to the proposal, an interesting solution should satisfy the so-called “generalized Nahm pole boundary condition” as and one of several certain asymptotic conditions as .
Full descriptions of the generalized Nahm pole condition can be found in [5] and [15] [16]. Several different but essentially equivalent definitions occurred in these literature. Here is what we choose to use in this paper: When , in a certain gauge, for some , where is a model solution which already satisfies the generalized Nahm pole boundary condition (to be described later) that goes like when . Note that we do not require that the inequality is uniform in .
We choose finitely many points with degrees (a positive integer assigned to each point) in . These points are called “knotted points”. If the choice of points are given, then according to proposition 6.1 in [10], there is a special solution to the extended Bogomolny equations that satisfies the generalized Nahm pole boundary condition and as . This special solution is the model solution that we use. Note that strictly speaking, different choices of knotted points give different versions of the “generalized Nahm pole boundary condition”.
On the other hand, there are three types of asymptotic conditions that we are interested in:
- •
-
•
The second type of asymptotic condition says that
as . This condition is called “real symmetry breaking condition” in Gaiotto and Witten’s paper [5]. And this is what we study in this paper. Sometimes we assume a stronger condition which says that:
where is a positive real number, is a constant element with norm .
-
•
The third type of asymptotic condition is to require , but and approach the same constant element with norm . This is called “complex symmetry breaking condition” in [5]. The author hopes to study it in the future.
1.2 The metric representation
Let be the trivial bundle over . It is convenient to use a pair to represent (at least locally) a configuration, where is an Hermitian metric on , is an matrix whose items are holomorphic functions.
Suppose is written as
where are holomorphic functions. Typically we assume they are polynomials. Then there is a configuration which trivially satisfies all but the first bullets of the extended Bogomolny equations, described by
Any gauge transformation sends to another configuration which also satisfies all but the first bullets of the extended Bogomolny equations. (Because they are gauge invariant.) Let be the Hermitian metric. (See appendix A.) Then we say corresponds to the pair . We call the pair a metric representation of . Note that this is not a - correspondence. In fact,
-
•
Each configuration can be locally represented by a pair if and only if it satisfies all but the first bullets of the extended Bogomolny equations.
-
•
It is possible that two different pairs and represent the same configuration . This happens if and only if there is a holomorphic gauge transformation (represented by a matrix whose items are holomorphic functions in and independent with ), such that
In fact, we may use different pairs of at different regions to represent the same configuration. They are connected by holomorphic gauge transformations in the overlaps. So technically speaking, a configuration corresponds to a Cech cocycle of pairs .
-
•
It is also possible that two different and correspond to the same pair . This happens if and only if and are gauge equivalent.
Here is the illustration of the above statements:
From to
Suppose is a configuration such that . Keep in mind that the structure of is preserved under the operators , whose definitions are in the last subsection.
Fix any . Then can be viewed as a d-bar operator on sections of . So it gives a holomorphic structure. Note that since , both and keeps this holomorphic structure.
On the slice, we may choose (at least locally) two holomorphic sections of , namely and . Since keeps the structure, we may assume that everywhere.
We may identify the choices of at different slices by requiring . Then and form a basis of . Under this basis, is just and is just . The operator is represented by an matrix whose items are holomorphic in and doesn’t depend on . We call this matrix . Note that is obvious in this basis.
Note that the definition of depends on a choice of the holomorphic basis . Different choices of the basis can be related by a holomorphic gauge transformation (an matrix whose items are holomorphic functions in ) which sends to .
A warning: The holomorphic gauge transformation is not an actual gauge transformation on the original configuarition . In fact, it doesn’t affect at all. It only changes by changing the pairs that we use to define it. Readers should not get confused by the (somehow misleading) term “gauge transformation” that we use here.
From to
Suppose we have a pair . Recall that the “trivial” configuration satisfies all but the first bullets of the extended Bogomolny equations.
We may choose an gange transformation such that . Then sends to some configuration . Clearly corresponds to .
Note that adding another gauge transformation to doesn’t change . In fact, if we replace by , we still have
So essentially, what we get is a configuration up to an gauge transformation.
For each pair , there is a preferred way to choose which we’ll frequently use in this paper: Each Hermitian metric can be written in the following format:
where is a positive real-valued function and is a complex-valued function. Then we have , where . Under the gauge transformation given by this specifically chosen , the trivial configuration becomes
We use to denote this particular configuration that corresponds to . Any other configuration that corresponds to is gauge equivalent to .
For each and , we write down the concrete formulas of in terms of and in two special situations:
Special case 1
Then , with
Proof.
In this special case, suppose . Then
The first bullet of the extended Bogomolny equations is:
We have
So indeed , with
∎
Special case 2
Then , with
Proof.
In this special case, suppose . Then
We have
And
So indeed
∎
1.3 The deformation of a configuration
In this paper, one general strategy to find a solution to the Bogomolny equations is: Construct a configuration that satisfies all but the first bullet of the equations first. (This can usually be done using the metric representation described in subsection 1.2.) And then seak an gauge transformation to make the first bullet vanish. So we need to study how the first bullet behaves under gauge transformations.
Suppose is a configration. We always assume it satisfies all but the first bullet of the extended Bogomolny equations. Recall that the first bullet can be written as
Suppose is an gauge transformation sending to . We may think as an -valued function on . Then . And the norm is invariant under this transformation.
On the other hand, suppose is an gauge transformation. If in addition, , then we call it a Hermitian gauge transformation.
The following fact is simple in linear algebra: For each , there is a unique such that is Hermitian. Because and are equivalent. If we are not sensitive to transformations on what we get, we may only consider Hermitian gauge transformation when we deform .
Suppose is a configuration. Given any section , there is a 1-parameter family of Hermitian gauge transformations parameterized by . Since , this deformation is perpendicular with gauge transformations initially.
On the other hand, any Hermitian valued function can be written as either or for some uniquely. Note that and act the same way on as gauge transformations. So effectively any Hermitian gauge transformation can be represented by for some .
Deformation of the equation
Suppose is a section in . We consider the Hermitian gauge transformation acting on . What we get is written as . Then the first bullet of the extended Bogomolny equations is written as
where are given by . Note that the last inequality used the fact that .
This formula seems awful. But actually we only need two facts:
The first fact
is
here the “remaining terms” include multi-linear terms and with any integer with convergent coefficients (comparible with the coefficients in the Taylor expansion of ). The operator behaves like the Laplacian, whose definition is
where , and are six components of .
Proof.
We have
Recall that is the connection defined using . And . We have . We have
And
Thus the first fact follows directly.
∎
There are two alternative ways to write the first fact that will be useful later:
where is the second order term and is the lower order terms; is the linear term and contains all the non-linear terms.
The second fact
is a Weitzenbock type of formula:
where the undefined linear algebra notations can be found in appendix A.
This is a very mysteries (at least to the author) and happy formula. It comes from [11] as proposition 5.1 there (with some adaptions). It may be also more or less originated from Donaldson, Uhlenbeck and Yau’s series of famous work on Kobayashi-Hitchin correspondence. Note that it is only true when , that is, Hermitian. It is generally false if only .
Proof.
The following identity can be checked directly:
Thus
Here we used the fact that is self-adjoint and . Moreover,
The last step used the fact that and are self-adjoint and . Similarly,
Finally,
So
∎
Adding Hermitian gauge transformations
We need to be very careful to add two Hermitian gauge transformations. If we have two different Hermitian gauge transformations and . If we apply first and then on a configuration , what we get is , which is generally not equivalent with .
Here is the correct way to add them: Let satisfy . Note that is Hermitian. So exists and is unique.
Moreover, . So is in . And . So is equivalent with .
Sometimes, we still use the sloppy notation to represent mentioned above if there is not a potential confusion. Note that if and are all bounded for , where is the th derivative and is a non-negative integer, then is also bounded accordingly. This is the key that guarantees the elliptic estimates later are not ruined by the non-commutative way of taking sums.
One should also be careful with the differences here: For example, . And the bound of may also depend on the lower order derivatives of and .
2 The model solution
This section studies a very special case in which the extended Bogomolny equations can be reduced to a single scalar equation. We assume the solution corresponds to a pair such that
In another word, can be diagonalized and can be made as a lower triangular matrix at the same time. This section studies solutions to the extended Bogomolny equations in this special case with generalized Nahm pole boundary condition at and real symmetry breaking condition as .
Based on the special case 1 in subsection 1.2, the extended Bogomolny equations are reduced to a scalar equation:
This section finds a solution to that gives us a solution to the extended Bogomolny equations with Nahm pole boundary condition as and real symmetry breaking condition as . This solution will serve as a model solution to construct more solutions later in this paper.
2.1 Boundary/assymptotic conditions
In order for a solution of to represent a solution to the extended Bogomolny equations that we are interested in, we need to translate the boundary/asymptotic conditions that we mentioned into conditions on .
The generalized Nahm pole boundary condition as
We define the generalized Nahm pole boundary condition in an indirect way: We quote Section 6 of [10] for a model solution that has the generalized Nahm pole boundary condition. The following theorem can be found as proposition 6.1 in [10].
Theorem 2.1.
For each non-zero polynomial , there is a unique solution to , denoted as , that has the following properties:
-
•
When , it satisfies the generalized Nahm pole boundary condition.
-
•
When , , uniformly with respect to .
The following feature of will be useful later: On the entire (either or ), we have
Note that the definition of here differs from the one in some literature, say [10], up to a sign. But it agrees with some other literature, say [3].
In this paper, taking the above theorem for granted, we can simply define a generilized Nahm pole boundary conditions for a solution as follows:
Let be the configuration that is represented by . Then to say that a general configuration satisfies the generalized Nahm pole boundary condition means: There is a positive number such that, possibly after an gauge transformation,
If we have a solution of , then the following condition on implies that it represents a solution with the generalized Nahm pole boundary condition:
Real symmetry breaking as
Definition 2.2.
We say that a configuration satisfies the real symmetry breaking condition if, possibly after an gauge transformation, for some , when
If we have a solution of , then the following condition on implies that it represents a solution with the real symmetry breaking: As ,
2.2 The construction of a solution
We use a version of the Perrons’ method. This method is somehow standard in many PDE books, say [6].
Super/sub-solutions
A function such that
is called a sub-solution. A function such that
is called a super-solution.
We always assume is continuous, but allow the possibility that is not differentiable. If this is the case, then a sub-solution/super-solution is defined in the weak sense: For example, to say
means for any non-negative smooth function on supported on a compact region, we have
If is both a sup-solution and a sub-solution, then it is a weak solution. The following argument is standard, showing that a weak solution is an actual smooth solution:
Suppose is a weak solution. Choose any ball whose closure is in . Since in the ball and , using the standard elliptic regularity argument, we see that , where is a slightly smaller ball in , is the Sobolev space. This further implies that using the Chain rule for derivatives, the fact that is continuous and bounded, and the Sobolev embedding/multiplication inequalities in a 3-dimensional space. This further implies that in a even smaller ball. We can do boot-strapping and conclude that for any positive in a smaller ball. So it is smooth in that small ball. Since the ball is arbitrary chose, is smooth everywhere in . In particular, it is an actual solution.
We construct a preferred super-solution and a preferred sub-solution such that: They both satisfy the generalized Nahm pole condition plus real symmetry breaking condition. And pointwise.
Here is the definition of :
Here is the definition of :
where
where is a large enough constant.
Clearly they all satisfy the desired boundary/assymptotic conditions. Moreover,
So is indeed a super-solution.
Note that although doesn’t exist, we do have
in the weak sense.
So replace by the above function on the right, we get when and we get
when .
Thus is indeed a sub-solution.
Perron’s argument
In general, if we have a collection of functions that are bounded above point-wise, then we define to be another function whose value at each point is:
For each point , let
We still use to denote the set .
Then clearly is a function that satisfies:
We prove that is an actual solution. We need two lemmas first:
Lemma 2.3.
If we have two sub-solutions , then is still a sub-solution.
Proof.
Clear is still continuous. Moreover,
in the weak sense. We write it as
So
So is also a sub-solution.
∎
Lemma 2.4.
Suppose is a ball whose closure is a compact subset of . Suppose . Then there is a unique solution in that is continuous up to the boundary, such that in and on . We call it a “lifting to an solution” of in . Moreover .
Proof.
We show the uniqueness first. Suppose two actual solutions in both equal on . Then
Since is continuous up to boundary in . If they are not the same in , without loss of generality, we may assume that is a positive maximal point of inside of with the property
Since we have
at , it contradicts with the fact that
To prove the existence, choose a large enough positive constant . Using a standard Dirichlet argument, we may construct a function such that
Using the fact that and maximal principle, we know . Then successively construct a sequence that have the same boundary values on and
in . Since and are fixed and since is bounded below by from the construction, we may assume is large enough such that, if is also bounded from the below by and if , then
Thus
implies
which further implies
So inductively we get is an increasing sequence and is a decreasing sequence.
On the other hand, we have . So
And we know that on the boundary. A maximal principle implies that in as well. And inductively we get for all . In particular, the increasing sequence has an upper bound . So it converges uniformly to a continuous function .
Clearly .
Finally, suppose is the Green’s function of for centered at any point and is the Poisson’s kernel on evaluated at . Since all are bounded uniformly in , using dominated convergent theorem,
This implies that
in . So is what we want.
∎
Note that in the above proof, suppose . Then its lifting to a solution in (while keeping the outside part of unchanged) is still in .
Given the above two lemmas, we prove that is an actual sotluion:
Proof.
Consider a ball whose closure is a compact subset of . Since elements in are continuous and bounded above uniformly by in the closure of , we may find an increasing sequence in that converges to on the closure of . Note that this implies that is also continuous in .
For each element in , replacing it with its “lifting to a solution” in makes it larger without violating being in . Without affecting the argument, we may assume each that we chose equal to its “lifting to a solution” in . In particular, each is an actual solution in .
Finally, suppose is the Green’s function of for centered at any point and is the Poisson’s kernel on evaluated at . Since all are bounded uniformly in and convergent uniformly to in , using dominated convergent theorem,
This implies that
in . Since is arbitrary, is a solution on the entire . ∎
The real symmetry breaking condition
Let have the same meaning as before. Remember that . We verify that satisfies the real symmetry breaking condition as . We always assume is a small enough constant, and (which means, all the inequalities only work when is large enough).
It suffices to verify that:
Since ,
Let . It only remains to verify that
The equation for is:
Then
So
Moreover, we have
We need a lemma:
Lemma 2.5.
Choose a point first. Suppose . Let be the ball of radius centered at . Then
Proof.
For each ball , we may choose a cut-off function that is supported in (the ball of radius with the same center) such that:
Then
where is a large enough constant.
Recall that
It implies
∎
Now we prove that .
Proof.
We assume and let be the Green’s function of the Laplacian centered at . Let be the ball of radius with the same center. Let be the same cut-off function as in the previous lemma. Then
where is evaluated at the point . Using integration by parts, one can verify that
We have , so
Moreover,
So
Remember that we are working on the region such that . We may choose , then we get
where .
∎
The generalized Nahm pole boundary condition
We verify the generalized Nahm pole boundary condition for . In this section, we always assume is a small enough constant and (which means, all inequalities only work when is small). Note that both and satisfy the generalized Nahm pole boundary condition.
Since , cleary we have
So it suffices to check that
Let . Then
So
The following lemma is similar with lemma 2.5, except that we are working in the region now and the definition of is also different.
Lemma 2.6.
Choose a point first. Suppose . Let be the ball of radius centered at . Then
Proof.
We choose the same cut-off function supported in and equal in as always. Then the same argument as in the last paragraph,
∎
Now we prove that . This is also similar with the last subsection (the real symmetry breaking).
Proof.
We assume and let be the Green’s function of the Laplacian centered at . Let be the ball of radius with the same center. Let be the same cut-off function as always. Then
where is evaluated at the point .
We have
Using the same method as the real symmetry breaking case,
We may choose . Then we get
∎
2.3 The uniqueness
Suppose all have the same meanings as in the last subsection. We prove that the solution of is unique under certain constraints:
Proposition 2.7.
Suppose is a solution of such that,
where is any fixed large constant and is another fixed real number. Then .
The author conjectures under weaker conditions it is still unique.
Conjecture 2.8.
Suppose is a solution of such that, for some ,
-
•
When , and .
-
•
When , and .
Then . (The inequalities do not necessarily uniform in in the conjecture.)
The remaining of this subsection proves proposition 2.7
Lemma 2.9.
Suppose is a smooth function on such that for some and , . Moreover, on the entire . Then .
Proof.
Let . Note that since is bounded on each fixed slice, is a well-defined function on . Moreover, . In particular, we have .
Note that since is continuous, is also a continuous function. We show that is a concave up function.
In fact, for any and . Let be a point such that
. Moreover, we may assume has a strict maximum over at . In particular, the acting on has a negative value at .
Note that . Since . This indicates that
Since is smooth, we actually have
in a neighbourhood of . So using the mean value theorem,
where is a number that can be arbitrarily close to .
So
Letting , we get
Since this is true at any point, we know that is a concave up function.
Fix any . Then since is concave up, for any other ,
Recall that , we have
But . We must have . This is true for any . Thus for any . And it implies that .
∎
Corollary 2.10.
If and are two solutions of on such that on the entire for some and . Then .
Proof.
Lemma 2.11.
Suppose is a large constant and . Then there is a second order differentiable function on with the following properties:
-
•
.
-
•
-
•
Proof.
In fact, we may choose a even larger and let
Then all three bullets are satisfied provided that is large enough. Here is the reason: The first two bullets are obvious. For the third bullet,
when is small, . So and hence the above expression is non-negative.
Otherwise, since is large, when , since has a lower bound, we may assume
So in any case, the third bullet is also true.
∎
Now here is the proof of proposition 2.7.
Proof.
Suppose is a solution of such that,
In the Perron’s argument, we may instead choose to be for a possibly larger constant and choose to be . Since is assumed to be large, by lemma 2.11, one can verify that they are indeed super/sub solutions.
Using these substituted and , we may run Perron’s argument again and get a solution, still call it . But recall that
And is a solution with . So we have .
On the other hand, we may modify the Perron’s argument to define
By the same reason, is also an actual solution. And . So in order to prove is unique, we only need to show that
Since we know that . And
Moreover,
Thus by lemma 2.9, we know that . ∎
A remark on poly-homogeneous expansion
Based on Section 6 of [10], has a poly-homogeneous expansion on , where is a preferred way to compactify as a manifold with boundaries and corners whose definition can be found in either appendix B or [10]. However, surprisingly, this is not the case for . In fact, may have a poly-homogeneous expansion on each boundary away from the corners of , but they do not seem to compatible at the corner. Even in the simplest case: Suppose there is no knot singularity as , but only Nahm pole sigular boundary condition with real symmetry condition. The solution is written explicitly as
This solution doesn’t seem to have a poly-homogeneous expansion at the corner of given by while at the same time, where .
3 The continuity method
In this section, we use the continuity method to construct more solutions to the extended Bogomolny equations with generalized Nahm pole boundary condition and the real symmetry condition. This construction is almost identical with Dimakis’ argument in [3]. By way of looking ahead, here is a brief sketch:
With the help of the solution constructed in the last section, for each triple , where are polynomials with , and are coprime, are monic, we construct an approximate solution that corresponds to it. Then we improve the approximate solution near . Finally, we use a continuity method to further improve it to get an actual solution.
3.1 The approximate solution
Suppose a triple is given. The approximate solution constructed in this section will be a pair with
We define separately in different regions. Recall that in general, we may write as
For an approximate solution, we only require that it behaves nicely near all boundaries/corners, but allow it to behave awfully in the middle area. We only need to define near each boundary and use any arbitrary smooth one to fill the inside.
Note that a preferred compactification of and preferred coordinates near boundaries/corners are used as always. See appendix B for the definitions of all types of boundaries/corners and local coordinates.
When is large (near type I boundary)
Recall that when is large, . In this region,
where , , and is the function constructed in section 2 which satisfies
and generalized Nahm pole bounary condition plus real symmetry breaking condition.
Since is large, we may assume that at all roots of and well-defined. Moreover, can be represented by (see special case 1 in subsection 1.2):
with
Note that and are only non-zero when . So except in the region . In this region, since we have assumed is large, , , and . So in fact, the first bullet of the extended Bogomolny equtauions
as , uniformly in .
When is small (near type III boundary)
We construction on a region such that is small and is bounded above. This region contains all the points with small .
We have to work in a different basis. (That is to say, a different choice of in the definition of , see subsection 1.2.)
Since and are coprime, we assume , where are also polynomials. Consider the holomorphic gauge transformation . (Recall, it is not an actual gauge transformation on the configuration, see subsection 1.2.) It sends to
In this basis, we choose
where is the version of but using instead of . That is to say, satisfies the following equation:
Using the new pair , its preferred configuration is gauge equivalent to . (Recall the definition of is in subsection 1.2.) This is in the special case 2 there. So the first bullet of the extended Bogomolny equations is
with
When is bounded, all the polynomials are bounded. Note that when is small and is bounded, and . Thus
uniformly in when is bounded.
Going back to the original basis, is written as
We still have since its norm doesn’t change under gauge transformations. This is what we want.
When is small (near type II boundary)
We assume is small. When is large at the same time, recall that is already defined by , . We call it
On the other hand, when is small at the same time, is also defined to be
When is neither too large nor too small and (or ) is small, we have
-
•
are all bounded.
-
•
(or ) and (or ).
-
•
As , and when is not small (to stay away from zeros of and ), , while . So
All these properties imply that in this region, as
So using a smooth cut-off function in to connect and , we get an defined on the entire region where is small such that when with is large and when is small. The error in the middle is
when is neither too small nor too big.
So uniformaly in and when is large.
To sum up
We have defined an near each boundary/corner with
We can define in the middle area smoothly and arbitrarily. The properties of the pair are:
-
•
is smooth on .
-
•
The configuration that corresponds to it satisfies the generalized Nahm pole boundary condition as (but of the version of , not ) and real symmetry breaking condition as .
-
•
When , uniformly in .
-
•
When , uniformly in .
-
•
When is large but is small, or when is large but is small, .
An estimate on
For later analysis, we will need an estimate on the point-wise norm . Here the norm is the sum of the norms on all its six components .
Recall that if is written in the format
then is written as
Examining the construction of , it is clear that when , all these terms are point-wise bounded above by a constant which doesn’t depend on or .
When , one also examines that uniformly in .
So overall, there exists a constant such that
on the entire .
A remark on geometrical meanings of and boundary conditions
We briefly explain the geometrical meanings of . There is no difference in this regard between our situation and the situation described by Dimakis in [3] without real symmetry breaking .
Note that zeros of are the “knotted points” for the generalized Nahm pole boundary condition with degrees. And zeros of are the zeros and vanishing orders of . Moreover, under a certain gauge, there is a “small section” described by . Readers may see [3] for more details. These data are irrelevant with the gauge transformations. So in particular, different choices of do not give equivalent solutions.
3.2 Improve the approximate solution when is small
Suppose is the approximate solution in the last subsection. We modify when is small and is bounded to make for any positive integer as . Without leading to an ambiguity, we may still use to denote the modified version after this subsection in this paper.
The idea of the modification is to inductively construct a smooth section of , . Suppose acts on and gets . (Note that the summation is in the sense described in subsection 1.3 written in a sloppy way.) Then has the property
when , with arbitrarily small. We assume all the derivatives of also have the correct vanishing order as . We further assume each is supported in a region where is small and is bounded. So it doesn’t ruin other good properties that has as listed in 3.1. In particular, we assume originally = 0 when and for some large constant . Then for each , we inductively assume when and .
Here is the construction:
Suppose is constructed. (When , we assume .) When is small, for any arbitrarily small ,
and when and at the same time.
We use to denote the operator for simplicity. We haven’t studied the mapping properties of yet, which is in fact awful. But luckily since we don’t care about the region when either or is large, we may modify to remain unchanged when both and are small, but equals when either or is large, where is a configuration with the same version of generalized Nahm pole boundary condition and without real symmetry breaking at . When is small and is bounded, as an elliptic operator with “edge” type (strictly speaking, is the elliptic operator of “edge” type locally), only non-leading terms are modified. This operator is already well studied in [16] and [11]. And it equals when is small and is bounded (whose bound is larger than , say when ).
Suppose is a smooth cut-off function in which is when is small and when is large. Suppose is a smooth cut-off function in which is when and when is large. By the same reason as the proof of proposition 7.1 in [11] or the arguments in paragraph 4.3 in [3], we have a smooth section such that when , in this region,
All the derivatives of also have the appropriate decay rates at . Note that may behave bad outside the region mentioned above.
In fact, the operator is locally invertible in the region is small and above the lower Fredhohm weight. To be more precise, if is supported in this region and when , then there is an such that in this region. The definition of can be found in [11], [3] or appendix C. The fact that the Fredholm weight is above can be found in [16], [11] or [3]. (Strictly speaking, near a corner both and goes to , we should use . However even there, is equivalent with .) Since we only require the equation holds in the local region mentioned above, this is a local property and there is no need to define the space in other parts of , say, when . This also means, is allowed to behave bad outside the region that we are interested in. Note that the small constant exists because of the fact that may have additional terms of which make it not reside exactly in .
We may take a third cut-off function in which is when and when . Recall the fact that when and is small. This implies that in the same region.
Thus let . We have is supported in the region and is small. And in particular, when and (We assume in this region),
So let be the configuration that we get after applying on . According to the first fact in subsection 1.3,
When and ,
When and ,
and
Finally, when and , we have thus . So the inductive construction is finished.
We take summation in a convergent way of (and in the sense described in subsection 1.3) and get a new configuration, still denoted as , with the additional property besides what are listed in subsection 3.1:
for any positive integer when uniformly in . Here “in a convergent way” means we may freely multiply a cut-off function in supported and equals to near (but becomes quickly when gets away from ) onto each (whose support depends on ) to make the infinite sum of converge. In the remaining of this paper, we will use this new instead of the old one.
3.3 The continuity argument
Let be the approximate solution from the last subsection. Let . Clearly . Consider acting on and call it . Note that vanishes up to any finite order on any boundaries. (We’ll need this property to make sure that still satisfies the same generalized Nahm pole condition and real symmetry condition as . ) Then acting on gets back to . We have
where the notation is defined in subsection 1.3.
On the other hand, by the first fact described in subsection 1.3
Thus whenever . And for any when just like . And it decays exponentially when just like . That is to say, for some small constant , as .
The following equation
has an obvious solution when . In general, suppose is a section of . Let be the set of values such that
has a solution with for some small and any non-negative integer , where the definition of can be found in appendix C. Then is non-empty. (Because .)
By way of looking ahead, we prove that by showing it is both relatively open and relatively closed. In particular, when , it gives an actual solution to the extended Bogomolny equations. In fact, we have the following five facts:
-
•
The initial is in for any non-negative integers . This is straightforward from the construction.
-
•
Suppose at some , there is a solution for all . Then for some small depending on , there is also a solution in when . This is proved in subsection 3.5.
-
•
For each , suppose when there is a solution . Then there is a constant which only depends on and such that
Note that the constant doesn’t depend on or . This is proved in subsection 3.4.
-
•
Suppose . If is a solution in for all and and some small . Then it is also in for all and . This is proved in subsection 3.5.
-
•
There is a constant which depends only on and , such that for any , if there is a solution . Then
Note that doesn’t depend on or . This is proved in subsection 3.4.
Proof that using the above facts
Let be the subset of such that additionally for all . Then .
The second bullet indicates that is relatively open in . If there is a sequence in converging to . Then by the third bullet and by choosing a sub-sequence, we may assume converges to an element in for all . We don’t need to add an on since it is true for all . Then by the fourth bullet the limit actually lies in . This limit solves the version of the equation. So is a relatively closed subset of . Thus . In particular, .
Finally, choose a sequence . According the the fifth bullet, by choosing a sub-sequence, we may assume converges to some . This limit solves the version of the equation. Thus .
3.4 A priori estimates
This subsection derives some a priori estimates, which prove the third and the fifth bullets of the facts listed in the end of subsection 3.3. These estimates are largely due to Dimakis’ estimates in [3]. A large portion of them are also nearly identical to Mazzeo and He’s work in [10] and Jacob and Walpuski’s work in [13]. Moreover, the local estimate may originate in Bando and Siu’s paper [1] and the local estimate may originate in Hildebrandt’s work in [12]. The author is not attempting to trace all the origins here, but wants to give the readers the impression that these estimates are all standard in analysis in some sense.
Local estimate
Suppose we have a solution
Then according to the second fact (WeitzenBock formula) described in subsection 1.3, we have
Lemma 3.1.
Proof.
The constants and may change from line to line. Keep in mind that only depends on and , while only depends on .
Note that originally we have
where depends on .
Point-wise we have
Consider the Green’s function for on . This Green’s function is
where is the reflection of about plane.
First we assume and prove the first bullet. Choose a ball of radius centered at . The maximal principle implies that
Note that has exponential decay as and has exponential decay with respect to the distance to outside of the ball . So
This implies
Repeating this finitely many times we get,
Next we assume and prove the second bullet.
We simply have
Originally we have
And the same maximal principle applies and we have
Note that is bounded above by since has an exponential decay in when .
On the other hand, note that
where means the coordinates of and respectively. On ,
So after all
Which means
Repeating this finitely many times we get (remember that may change from line to line)
∎
We also need a point-wise estimate of when . We use to represent when is small and when is not small.
Lemma 3.2.
(This is inspired by a similar Taubes’ estimate in [20])
There is a constant which only depends , such that,
Proof.
Since by lemma 3.1 is bounded above by on the entire . We only need to assume is small. Recall in the proof of lemma 3.1, the maximal principle that we have:
Suppose initially we suppose
where depends on .
Let be the ball of radius centered at . Then outside of this ball and
Note that is bounded. So centered at the point , we have
On the other hand, within the ball , since for any positive integer when is small, we have
So after all
Repeating this finitely many times we get
∎
The interior estimates for away from
(Bando and Siu’s estimates)
We continue to assume solves
We first work in a region away from the boundary.
Lemma 3.3.
Suppose is a ball of radius in which is far away from the boundary. We assume . Let be the Green’s function of centered at the center of the Ball (basically times over the distance to the center). Let
Then there exists a constant and which only depends on , such that
Proof.
Recall the fact that is bounded above by a constant that only depends on . So the point-wise norm (largest eigenvalue) of the operator is also bounded uniformly on the entire . Note that is a self-dual map. So all its eigen-values are real number. Thus
is bounded below uniformly (that is to say, the norm of all its eigenvalues are bounded below by a positive constant).
So for a constant which only depends on , but may change from line by line, we have
From now on, let be a standard smooth cut-off function which is on and outside of . We assume uniformly.
Step 1 We first show that is uniformly bounded above. We have
On a ball of radius centered at ,
Step 2 Let be the average of on the region , where and have the same center. That is to say,
Since is a constant, all its derivatives are . Moreover, since we are working in a region away from the boundary, we know that is uniformly bounded and hence
are all bounded uniformly. Thus
Suppose . (See subsection 1.3 for details on how to add Hermitian gauge transformations.) Then . We will verity that
In fact, for the first inequality, suppose the eigen-values of are and . Then . And the norm is equivalent with , which is also equivalent with when is bounded. On the other hand, we indeed have are all bounded and
Then the first inequality follows.
The second inequality is because
Again, are all bounded above and below away . So and can bound each other.
Applying the Weitzenbock formula in subsection 1.3 and the fact that are bounded above, we get
So we have
The last inequality above is the Poincare inequality
So
By some algebraic manipulations and the fact that is an increasing function, one sees from above that
for some . ∎
Note that the proof of lemma 3.1 implies that locally has a finite norm (the Morry-Campanato norm, see appendix D). By theorem D.2 in the same appendix, we know that
So away from the boundary (say, ),
Moreover, when is large, (Here means the value of the center of the ball, hence is a constant.)
which is bounded above uniformly. And the proof doesn’t need to change if we replace with , where is the -value of the center of the ball . In particular, we still have
(We need to use the first fact in subsection 1.3 and lemma 3.2 here.)
So in fact, we get when is not small (say ), we get
for some .
Moreover, by exactly the same reason, if and allow to depend on and an additional positive integer . Then
for some .
The estimates for near
(This is an “adapted to the edge boundary version” of Bando and Siu’s estimate, being analog with an estimate of He and Mazzeo introduced in [11].)
We cannot use the same argument to show that is still locally in near since we don’t have a uniform bound on near , which is required in the estimate of . In fact, we only have
when is small. On the other hand, the definition of is also adjusted near this boundary in a scale-invariant way. So in fact, we need to do the argument in a scale-invariant way near the boundary.
Consider a ball of radius centered at a point . We hope get an estimate for and some small .
Thanks to the fact that and the fact that for any positive integer when is small.
We still define
where is the ball of radius centered at . We assume .
In the ball, we have
Then by the same reason as in the step 1 of the proof of lemma 3.3,
On the other hand, in step 2, we have and . So
are all bounded by uniformly. So and , where has the same meaning as in the proof of lemma 3.3.
So what we get is
And like in step 2 of the proof of lemma 3.3,
So what we have is
In fact, what we get is
Doing algebraic manipulations and keep in mind that and we have for any , we get
for some .
This implies that
So
Transfer it to the edge version, we get
We may assume , so together with the fact that , we get
Even near the type III boundary or any corner, we have the same estimate. There should be written as . (We are somehow abuse of notations here. The letter here means the coordinate near the type III boundary as defined in appendix B, not the radius of a ball.) And here is what we get on the entire .
Proposition 3.4.
There is a constant which depends only on such that for all , uniformly we have
Moreover, if and allow to depend on and a positive integer , then
The estimates away from boundaries
(Hilderbrandt’s estimate)
In the following arguments, sometimes we need to shrink . So we allow to change from line to line, but always independent with . We first work in a region away from the boundary (say, ).
Recall that
So
Away from boundaries, this can be written as
where are all bounded.
Lemma 3.5.
Consider the region away from boundaries. Then for some ,
where depends only on .
Proof.
Let be a ball of radius whose center is away from all boundaries. We assume .
Then from the proof of lemma 3.3, we know
Let , where in and on , in and on .
Use to represent the average of over the ball . That is to say,
Similar definitions for and etc..
Note that we already have
On the other hand,
Since , we have on the entire ,
By maximal principle on , we have
So
Thus
We know for ,
If we fix , then this implies
So the Campanato norm of over is bounded above by .
We may use the smaller power between and as the new starting point and re-run the argument. After finitely many times of iterations, we get for some possibly smaller ,
Then by theorem D.2 in the appendix D, we know that
∎
The weighted estimates near boundaries
(an adapted version of Hilderbrandt’s estimate)
Near the boundary (but away from the boundary), we have
where only is bounded, and , . The fact that , are not bounded is because is not bounded there.
Thus on a ball of radius centered at , we want to re-run the proof of lemma 3.5.
To start, we have
Let . And the same argument as in the proof of lemma 3.5 leads to (we still have the same definition of )
As long as the exponent of is less or equal than we can run the same argument and gain an extra factor. After finitely many times of iterations, this will lead to
for some possibly smaller .
This implies the norm of is bounded above by on this ball. Note that doesn’t depend on the ball. So we conclude that, in a region near the boundary away from boundary and for possibly smaller , we have
Even when approaching the boundary, the argument doesn’t need to change and we get locally
On the other hand, near the boundary, we have
where are bounded, if and can be any integer if with the bounds of depends on and .
So examine the proof of lemma 3.5, we get:
where depends only on . When and when is allowed to depend on , the positive integer , then
All the higher weighted norms with follow by the standard elliptic Holder type interior estimates. So to conclude we have
Proposition 3.6.
There exists an such that
Here the constant depends only on and the positive integer . When and we assume depends only on , , positive integers and , we have stronger estimate:
Note that this proposition is the third and fifth bullets of the facts listed in subsection 3.3.
3.5 More regularities
This subsection proves the second and the fourth bullet of the facts listed in the end of subsection 3.3.
Suppose is a solution for some . We hope to study the operator . Unfortunately this operator seems to be hard to study. However, suppose . Consider alternatively the operator . Then it behaves much better and can be analysis-ed.
In fact, we have the following lemma.
Lemma 3.7.
Assume and in (for all positive integers ) is a solution of
Suppose for some small and all positive integers and some large enough . Then there exists a for all such that
Moreover, the norm of is bounded above by (a constant times) the norm of .
Proof.
We use two cut-off functions to divide into two parts: , . Here are all smooth functions with
We assume is supported in a region such that . And is supported in a region such that either . Clearly we only need to solve on each region.
We first consider . We apply Mazzeo’s theory of elliptic edge operators (and terminologies) here, which can be found in [14]. We quote theorem 5.8 in [11] (where the indicial weights are actually computed in [16]). In fact, when , the operator has the same normal operator as if and is the model solution in their construction. Here is the only fact that we need: The weight lies in the Fredholm range both near the boundary and the boundary (and is compatible with the corners). (Alternatively, readers may compute the Fredholm weight of directly. It is actually easier than Mazzeo and Witten’s computations in [16], which takes more situations into account.) In particular, near those boundaries,
has a solution in the support of with for all . Here we allow that is nonzero in a slightly larger region and doesn’t satisfy the equation outside of the support of . But we may add (which is supported in a relatively compact region) to there and throw it into the next step.
Here is the next step: We solve the equation for . Consider the following functional on , where for all :
We may first consider the Banach space defined by completion of smooth compact supported functions using the following norm:
Clearly, is bounded in and has a Dirichlet minimizor in . (Recall that for sufficiently large .) This minimizor is unique because of the convexity of . By a standard elliptic regularity argument, this minimizor, as a week solution of , is smooth in the interior of . Moreover, there is a Hardy type of inequality for :
In particular, the integral of both and over the region has at most a polynomial growth as . We have
But
So we get
Let be a cut-off function that equals when and equals when . Moreover, we assume And let . Then recall the Green’s function of is (defined in subsection 3.4) and the fact that has compact support, we have (centering at any point )
Note that when , has exponential decay while the integral of and on support have at most polynomial growth. So
And letting ,
We may divide into two parts: Let be a ball of radius , where is the value of . Then
When , the second integral above decays exponentially in . The first integral is bounded by (recall that )
So we get
Once we have this point-wise bound, then the bound on Holder norms of follows by a standard elliptic interior argument and are omitted. (It is actually much easier than the analysis in subsection 3.4 because we don’t have the quadratic term here.)
Finally, we add the two different that we get for and together and finish the proof.
∎
Suppose for some , we have a solution with
Consider a small variation on top of and a small variation on top of :
Note that is a bounded map from to . So when is small enough, by implicit function theorem, the equation
has a solution in . This proves the second bullet of the facts in subsection 3.3.
Suppose when , is a solution in . We have
where actually maps into for sufficiently small . In particular, is in . Then because and are all Fredholm weights for near those boundaries, and vanishes up to infinite order at all boundaries. So inductively lies in for all . This proves the fourth bullet of the facts in subsection 3.3.
Appendix A The linear algebra
This appendix summarizes the linear algebras that we use.
Suppose is the trivial bundle whose structure is fixed. Since we are working on the Euclidean space, we take the advantage that nearly everything can be represented by matrices. In the following list, all matrices have complex variable items. Note that we use to represent the usual complex ad-joint (conjugate of the transpose) of a matrix.
-
•
An Hermitian metric on is represented by an matrix such that , , and positive definite. Each such metric gives an structure. (The condition keeps the structure unchanged.) The inner product of two sections (represented by -d vectors with variable coefficients) is
Unless otherwise specified, we typically just use the inner product defined by , that is
-
•
Infinitesimal gauge transformations are represented by sections of . Infinitesimal gauge transformations are represented by sections of . And , where are the Hermitian elements.
-
•
One useful formula: Suppose is a differentiable parameter family of matrices. Then
where
Proof.
Let Then commutes. And . Let . Then
The other identity can be derived in the same way using . ∎
-
•
Another useful formula: Suppose are two matrices. Then
Proof.
This is straightforward:
∎
-
•
Suppose is Hermitian. Then the operator has a square root:
In fact, consider the function , where is any real number. Clearly is a real analytical function over . It has a convergent Taylor series expansion at any point. On the other hand, when is Hermitian, is also a self-adjoint operator. So it is diagonalizable and has real eigen-values. In particular, when , we use this expansion to define
Appendix B Compactification of
This appendix gives a preferred way to compactify as a manifold with boundaries and corners . In this paper, we have always assumed that is compactified in this way whenever we work near one of the boundaries/corners and a compactification is needed.
There are three types of boundaries of the compactification of . We call them “type I, II, III boundaries” respectively. Type I and type II boundaries intersect at a type A corner. Type II and type III boundaries interest at a type B corner. Type I and type III boundaries do not intersect.
Here are the definitions and local coordinates.
Type I boundary (, or equivalently )
Recall that . Here is the definition of : It is a smooth function from to that equals when is small and equals when is large.
So (or equivalently ) defines a boundary for the compatification . This is the type I boundary.
Type III boundaries ()
Here is the definition of : It is a smooth function from to .
Suppose is any root of (which corresponds to a “knotted point” at of the generalized Nahm pole boundary condition). When both and are small, we require . When either is large or is away from all roots of , we require that .
Then defines the type III boundaries of the compatification of . Note that we have blowed up at for each root of .
Type II boundary (, or when equivalently there)
Here is the definition of : It is a smooth function from to .
Away from the type I and type III boundaries (say, ), when is small, we require . When is large, we require .
Near a type I boundary, if is small, then we require . When is close to , we require that .
Near a type III boundary, we require .
Note that effectively, away from other boundaries, and define the same boundary. But we use instead of because it is also compactible with other boundaries.
Type A and type B corners
Type A corners are given by (or equivalently ). And type B corners are given by .
A remark on the coordinates:
When is large, since , we may freely choose to use either or there for the same meaning. But usually we use if we want to emphasize that it equals (not arbitrarily small) when .
When is not too small and is not too large and when is small, and can bound each other. So they are also interchangeable there in analysis.
Appendix C Weighted Holder spaces of (iterated) edge type
This appendix defines the Banach spaces , where , is a non-negative integer, are real numbers. These spaces are standard in the aspect of Mazzeo’s micro-local analysis theory (see [14]). They’ve also occurred in [16], [11] and [3]. (For the sake of convenience, the descriptions here may be modified compared to other literature in a non-essential way.)
C.1 The Holder spaces of (iterated) edge type
Suppose is a ball in far away from any boundary/corner (the distance to any boundary/corner is at least ). Then we define the Holder spaces over , where is a non-negative integer and . This space is given by the norm:
where
In a region far away from type II () and type III () boundaries, we take the supreme of all balls of radius of the above norm.
The operator that we want to study is which is introduced in subsection 3.5, where . This operator is of the “degenerate elliptic of (iterated) edge type” near a type II or a type III boundary as studied in [16], [11] and [3]. It is standard to modify the Holder spaces near those boundaries. The modified version will be denoted as .
We take the boundary as an example. One way to think about the modification is to re-define the distance between two points and near the boundary to make it re-scaling invariant under a re-scaling . This is done by modifying the metric near the boundary. In a direction that is perpendicular with the direction, there is nothing need to be changed. But in the direction, the metric should be instead of . Thus the distance between a point at and a point at (with all other perpendicular coordinates the same) is given by
which is clearly re-scaling invariant.
Another equivalent way to do the modification is to define it on each ball of radius whose center has an -value . On this ball, the of should be given by
Note that the weight here works equivalently as if the metric is re-scaled in this ball. (They bound each other in a way that doesn’t depend on near the boundary.)
When it comes to the boundary but away from the corner (or equivalently, when is not too small and is not too large), things are slightly different. The operator (strictly speaking, its ) has leading order terms which are made from combination of compositions of . So the re-scaling should be made in both and direction. And the metric should be addapted to be near the boundary.
Similarly, an equivalent way is to define it on each ball of radius centered at a point whose value is . On this ball , the of is given by
Near the corner, the metric is adjusted so it is dual re-scaling invariant in two directions that corresponds to the two boundaries.
Note that we do not need to modify the Holder norm when (type I boundary). Because the operator is not of the “degenerate elliptic of edge type” near this boundary.
For the higher Holder spaces, in the definition of , near boundary, we need to replace in by . And near boundary (or boundary but away from boundary), we need to replace by . This corresponds to the edge structure of the operator that we study. Nothing needs to be adjusted when .
C.2 The weighted Holder spaces
In order to be suitable for the operator (occurred in subsection 3.5) to map between, we need to add wights near boundaries/corners of the aforementioned Holder spaces . Here is the definition:
Here are several remarks:
-
•
If is an element in for some and all positive integers , then is also in for any other . If this is the case, then the concrete doesn’t matter and we may simply use to represent it.
-
•
Although all these norms and spaces are defined for functions, frequently we use them on sections of trivial bundles. The difference is only tautological so we don’t emphasize it.
-
•
If , then is actually the same space as , where we treat as a single boundary (without the blow-ups at each point like what we did in appendix B), and the subscript “0” means this is the ordinary edge type Holder norm as , not iterated edge type.
Appendix D Morrey-Camponato spaces and inequalities
The Morrey-Camponato spaces and their embedding inequalities are standard in analysis. We only state what we need. We always assume is a ball of radius whose closure is in the interior of .
Note that although the spaces and inequalities are stated for functions, they work the same tautologically for sections of trivial vector bundles. So we don’t emphasize the difference.
Definition D.1.
Suppose is a function in and is a real number. Let be the ball of radius centered at . Then the Morrey norm of is defined to be
The Camponato semi-norm is
where is the average of in the ball , that is to say
The Camponato norm is
We have some embedding theorems between different normed spaces. These are all standard in modern analysis so we omit the proofs. Here they are:
Theorem D.2.
Suppose is a real number. There is a constant which depends on . Suppose is a function in . Then
-
•
If , then , where is the Holder (semi-)norm.
-
•
.
-
•
If , then
Note that under a re-scaling, the two sides of all the inequalities scale in the same way. So the constant doesn’t depend on the radius of the ball.
There is another inequality which is standard in analysis
References
- [1] SHIGETOSHI BANDO and YUM-TONG SIU. STABLE SHEAVES AND EINSTEIN-HERMITIAN METRICS. Geometry and Analysis on Complex Manifolds, pages 39–50, 12 1994.
- [2] Panagiotis Dimakis. Model knot solutions for the twisted Bogomolny equations. 4 2022.
- [3] Panagiotis Dimakis. The extended Bogomolny equations on R^2 x R+ I: Nilpotent Higgs field. 9 2022.
- [4] Michael Gagliardo and Karen Uhlenbeck. Geometric aspects of the Kapustin-Witten equations. Journal of Fixed Point Theory and Applications, 11(2):185–198, 1 2014.
- [5] Davide Gaiotto and Edward Witten. Knot Invariants from Four-Dimensional Gauge Theory. Advances in Theoretical and Mathematical Physics, 16(3):935–1086, 6 2011.
- [6] David Gilbarg and Neil S. Trudinger. Elliptic Partial Differential Equations of Second Order. 224, 2001.
- [7] Qing Han. Elliptic partial differential equations / Qing Han, Fanghua Lin. 2011.
- [8] Siqi He and Rafe Mazzeo. Classification of Nahm pole solutions of the Kapustin-Witten equations on S^1 x x R+. 1 2019.
- [9] Siqi He and Rafe Mazzeo. Opers and the twisted Bogomolny equations. 2 2019.
- [10] Siqi He and Rafe Mazzeo. The extended bogomolny equations and generalized Nahm pole boundary condition. Geometry and Topology, 23(5):2475–2517, 2019.
- [11] Siqi He and Rafe Mazzeo. The extended Bogomolny equations with generalized Nahm pole boundary conditions, II. Duke Mathematical Journal, 169(12):2281–2335, 6 2020.
- [12] Stefan Hildebrandt. Harmonic mappings of Riemannian manifolds. pages 1–117, 1985.
- [13] Adam Jacob and Thomas Walpuski. Hermitian Yang–Mills metrics on reflexive sheaves over asymptotically cylindrical Kähler manifolds. https://doi.org/10.1080/03605302.2018.1517792, 43(11):1566–1598, 11 2019.
- [14] Rafe Mazzeo. Elliptic Theory of Differential Edge Operators I. Communications in Partial Differential Equations, 16(10), 1991.
- [15] Rafe Mazzeo and Edward Witten. The Nahm Pole Boundary Condition. arXiv e-prints, page arXiv:1311.3167, 11 2013.
- [16] Rafe Mazzeo and Edward Witten. The KW equations and the Nahm pole boundary condition with knots. arXiv preprint arXiv:1712.00835, 2017.
- [17] Clifford Henry Taubes. Sequences of Nahm pole solutions to the SU(2) Kapustin-Witten equations. 5 2018.
- [18] Clifford Henry Taubes. The R invariant solutions to the Kapustin Witten equations on (0,\infty) x R^2 x R with generalized Nahm pole asymptotics. arXiv preprint arXiv:1903.03539, 2019.
- [19] Clifford Henry Taubes. Lectures on the linearized Kapustin-Witten equations on (0,\infty) x Y. arXiv preprint arXiv:2008.09538, 2020.
- [20] Clifford Henry Taubes. The existence of instanton solutions to the R-invariant Kapustin-Witten equations on (0, \infty ) x R^2 x R. 2 2021.
- [21] Edward Witten. Fivebranes and Knots. Quantum Topology, pages 1–137, 1 2011.
- [22] Edward Witten. Khovanov homology and gauge theory. In Proceedings of the Freedman Fest, volume 18, pages 291–308. Mathematical Sciences Publishers, 10 2012.
- [23] Edward Witten. Two Lectures On The Jones Polynomial And Khovanov Homology. 1 2014.
- [24] Edward Witten. Two lectures on gauge theory and Khovanov homology. arXiv preprint arXiv:1603.03854, 2016.