This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On the extended Bogomolny equations on 2×+\mathbb{R}^{2}\times\mathbb{R}^{+} with real symmetry breaking

Weifeng Sun

Abstract

In this paper, we construct solutions to the extended Bogomolny equations on X=2×+X=\mathbb{R}^{2}\times\mathbb{R}^{+} with certain boundary conditions and asymptotic conditions.

Let yy be the coordinate of +\mathbb{R}^{+}. 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 y0y\rightarrow 0 and y+y\rightarrow+\infty resepctively. The boundary condition (y0y\rightarrow 0) is called “generalized Nahm pole boundary condition” and the asymptotic condition (y+y\rightarrow+\infty) is called “real symmetry breaking condition”.

For each triple of polynomials with complex coefficients (P(z),Q(z),R(z))(P(z),Q(z),R(z)) with degR<degQ\mathrm{deg}R<\mathrm{deg}Q, and Q,RQ,R are co-prime, P,QP,Q 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 y+y\rightarrow+\infty.

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 X=2×(0,+)X=\mathbb{R}^{2}\times(0,+\infty). Suppose x1,x2x_{1},x_{2} and yy are coordinates of XX, where yy being the one for +\mathbb{R}^{+}. We use 1,2,y\partial_{1},\partial_{2},\partial_{y} to represent the partial derivatives. Sometimes we treat 2\mathbb{R}^{2} as \mathbb{C} and let z=x1+ix2z=x_{1}+ix_{2} be its coordinate.

One way to introduce the extended Bogomolny equations on XX is as follows: It is a set of equations on configurations. Each configuration is a collection of 6 su(2)su(2)-valued functions Ψ=(A1,A2,Ay,Φ1,Φ2,Φ3)\Psi=(A_{1},A_{2},A_{y},\Phi_{1},\Phi_{2},\Phi_{3}) on XX. Here are the equations:

{1A22A1+[A1,A2][Φ1,Φ2]yΦ3[Ay,Φ3]=0,yΦ1+[Ay,Φ1]+[Φ2,Φ3]=0,yΦ2+[Ay,Φ2]+[Φ3,Φ1]=0,1AyyA1+[A1,Ay]+2Φ3+[A2,Φ3]=0,2AyyA2+[A2,Ay]1Φ3[A1,Φ3]=0,2Φ1[A2,Φ1]+1Φ2+[A1,Φ2]=0,1Φ1+[A1,Φ1]+2Φ2+[A2,Φ2]=0.\left\{\begin{array}[]{lr}\partial_{1}A_{2}-\partial_{2}A_{1}+[A_{1},A_{2}]-[\Phi_{1},\Phi_{2}]-\partial_{y}\Phi_{3}-[A_{y},\Phi_{3}]=0,&\\ \partial_{y}\Phi_{1}+[A_{y},\Phi_{1}]+[\Phi_{2},\Phi_{3}]=0,&\\ \partial_{y}\Phi_{2}+[A_{y},\Phi_{2}]+[\Phi_{3},\Phi_{1}]=0,&\\ \partial_{1}A_{y}-\partial_{y}A_{1}+[A_{1},A_{y}]+\partial_{2}\Phi_{3}+[A_{2},\Phi_{3}]=0,&\\ \partial_{2}A_{y}-\partial_{y}A_{2}+[A_{2},A_{y}]-\partial_{1}\Phi_{3}-[A_{1},\Phi_{3}]=0,&\\ -\partial_{2}\Phi_{1}-[A_{2},\Phi_{1}]+\partial_{1}\Phi_{2}+[A_{1},\Phi_{2}]=0,&\\ \partial_{1}\Phi_{1}+[A_{1},\Phi_{1}]+\partial_{2}\Phi_{2}+[A_{2},\Phi_{2}]=0.&\end{array}\right. (1.1)

Remark: If the background metric on XX 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 SU(2)SU(2) gauge invariant. Moreover, all bullets of (1.1) except the first one are also SL(2,)SL(2,\mathbb{C}) gauge invariant. Here an SU(2)SU(2) (or SL(2,)SL(2,\mathbb{C})) gauge transformation is represented by an SU(2)SU(2)-valued (or SL(2,)SL(2,\mathbb{C})-valued) function gg that sends

(A1+iA2,AyiΦ3,Φ1iΦ2)(A_{1}+iA_{2},A_{y}-i\Phi_{3},\Phi_{1}-i\Phi_{2})

to

(g(A1+iA2)g1((1+i2)g)g1,g(AyiΦ3)g1(yg)g1,g(Φ1iΦ2)g1).(g(A_{1}+iA_{2})g^{-1}-((\partial_{1}+i\partial_{2})g)g^{-1},g(A_{y}-i\Phi_{3})g^{-1}-(\partial_{y}g)g^{-1},g(\Phi_{1}-i\Phi_{2})g^{-1}).

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 SL(2,)SL(2,\mathbb{C}) bundle (that is to say, the bundle is just the trivial 2\mathbb{C}^{2} bundle but all the operators keep the SL(2,)SL(2,\mathbb{C}) structure):

D1=1+i2+(A1+iA2),D2=Φ1iΦ2,D3=y+AyiΦ3.D_{1}=\partial_{1}+i\partial_{2}+(A_{1}+iA_{2}),~{}D_{2}=\Phi_{1}-i\Phi_{2},~{}~{}D_{3}=\partial_{y}+A_{y}-i\Phi_{3}.

Then all but the first bullet of the extended Bogomolny equations can be written as

[D1,D2]=[D2,D3]=[D3,D1]=0,[D_{1},D_{2}]=[D_{2},D_{3}]=[D_{3},D_{1}]=0,

which is clearly SL(2,)SL(2,\mathbb{C}) gauge invariant (as commutators of operators).

The first bullet of the extended Bogomolny equations is equivalent to

[D1,D1]+[D2,D2]+[D3,D3]=0,[D_{1},D_{1}^{*}]+[D_{2},D_{2}^{*}]+[D_{3},D_{3}^{*}]=0,

where * represents the ad-joint. Then it is obviously SU(2)SU(2) gauge invariant. It is not SL(2,)SL(2,\mathbb{C}) gauge invariant in general because the * operator in the equation relies on the SU(2)SU(2) structure.

It is proposed that the extended Bogomolny equations on XX 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 Ψ\Psi should satisfy the so-called “generalized Nahm pole boundary condition” as y0y\rightarrow 0 and one of several certain asymptotic conditions as y+y\rightarrow+\infty.

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 y0y\rightarrow 0, in a certain SU(2)SU(2) gauge, |ΨΨmod|=O(y1+ϵ)|\Psi-\Psi_{\text{mod}}|=O(y^{-1+\epsilon}) for some ϵ>0\epsilon>0, where Ψmod\Psi_{\text{mod}} is a model solution which already satisfies the generalized Nahm pole boundary condition (to be described later) that goes like O(y1)O(y^{-1}) when y0y\rightarrow 0. Note that we do not require that the inequality is uniform in zz.

We choose finitely many points with degrees (a positive integer assigned to each point) in 2\mathbb{R}^{2}. 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 |Ψ|0|\Psi|\rightarrow 0 as y+y\rightarrow+\infty. 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 first type says that, under a certain SU(2)SU(2) gauge, |Ψ|0|\Psi|\rightarrow 0 as y+y\rightarrow+\infty. Solutions with this type of asymptotic condition are well studied in Section 6 of [10], [20], [19] and [3].

  • The second type of asymptotic condition says that

    |A1|,|A2|,|Ay|,|Φ1|,|Φ2|0,|Φ3|1|A_{1}|,|A_{2}|,|A_{y}|,|\Phi_{1}|,|\Phi_{2}|\rightarrow 0,~{}~{}~{}|\Phi_{3}|\rightarrow 1

    as y+y\rightarrow+\infty. 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:

    A1,A2,Ay,Φ1,Φ2=O(yϵ),Φ3=σ+O(yϵ),A_{1},A_{2},A_{y},\Phi_{1},\Phi_{2}=O(y^{\epsilon}),\Phi_{3}=\sigma+O(y^{\epsilon}),

    where ϵ\epsilon is a positive real number, σ\sigma is a constant su(2)su(2) element with norm 11.

  • The third type of asymptotic condition is to require |A1|,|A2|,|Ay|,|Φ3|0|A_{1}|,|A_{2}|,|A_{y}|,|\Phi_{3}|\rightarrow 0, but Φ1\Phi_{1} and Φ2\Phi_{2} approach the same constant su(2)su(2) element with norm 11. This is called “complex symmetry breaking condition” in [5]. The author hopes to study it in the future.

1.2 The metric representation

Let EE be the trivial 2\mathbb{C}^{2} bundle over XX. It is convenient to use a pair (H,φ)(H,\varphi) to represent (at least locally) a configuration, where HH is an SU(2)SU(2) Hermitian metric on EE, φ\varphi is an sl(2,)sl(2,\mathbb{C}) matrix whose items are holomorphic functions.

Suppose φ\varphi is written as

φ=(A(z)B(z)P(z)A(z)),\varphi=\begin{pmatrix}A(z)&B(z)\\ P(z)&-A(z)\end{pmatrix},

where A,B,PA,B,P are holomorphic functions. Typically we assume they are polynomials. Then there is a configuration Ψφ\Psi_{\varphi} which trivially satisfies all but the first bullets of the extended Bogomolny equations, described by

A1=A2=Ay=Φ3=0,Φ1iΦ2=φ=(A(z)B(z)P(z)A(z)).A_{1}=A_{2}=A_{y}=\Phi_{3}=0,~{}\Phi_{1}-i\Phi_{2}=\varphi=\begin{pmatrix}A(z)&B(z)\\ P(z)&-A(z)\end{pmatrix}.

Any SL(2,)SL(2,\mathbb{C}) gauge transformation gg sends Ψφ\Psi_{\varphi} to another configuration Ψ\Psi which also satisfies all but the first bullets of the extended Bogomolny equations. (Because they are SL(2,)SL(2,\mathbb{C}) gauge invariant.) Let H=ggH=g^{*}g be the Hermitian metric. (See appendix A.) Then we say Ψ\Psi corresponds to the pair (H,φ)(H,\varphi). We call the pair (H,φ)(H,\varphi) a metric representation of Ψ\Psi. Note that this is not a 11-11 correspondence. In fact,

  • Each configuration Ψ\Psi can be locally represented by a pair (H,φ)(H,\varphi) if and only if it satisfies all but the first bullets of the extended Bogomolny equations.

  • It is possible that two different pairs (H1,φ1)(H_{1},\varphi_{1}) and (H2,φ2)(H_{2},\varphi_{2}) represent the same configuration Ψ\Psi. This happens if and only if there is a holomorphic SL(2,)SL(2,\mathbb{C}) gauge transformation (represented by a matrix gg whose items are holomorphic functions in zz and independent with yy), such that

    H1=gH2g,φ1=g1φ2g.H_{1}=g^{*}H_{2}g,~{}~{}\varphi_{1}=g^{-1}\varphi_{2}g.

    In fact, we may use different pairs of (H,φ)(H,\varphi) at different regions to represent the same configuration. They are connected by SL(2,)SL(2,\mathbb{C}) holomorphic gauge transformations in the overlaps. So technically speaking, a configuration Ψ\Psi corresponds to a Cech cocycle of pairs (H,φ)(H,\varphi).

  • It is also possible that two different Ψ1\Psi_{1} and Ψ2\Psi_{2} correspond to the same pair (H,φ)(H,\varphi). This happens if and only if Ψ1\Psi_{1} and Ψ2\Psi_{2} are SU(2)SU(2) gauge equivalent.

Here is the illustration of the above statements:

From Ψ\Psi to (H,φ)(H,\varphi)

Suppose Ψ\Psi is a configuration such that [D1,D2]=[D2,D3]=[D3.D1]=0[D_{1},D_{2}]=[D_{2},D_{3}]=[D_{3}.D_{1}]=0. Keep in mind that the SL(2,)SL(2,\mathbb{C}) structure of EE is preserved under the operators D1,D2,D3D_{1},D_{2},D_{3}, whose definitions are in the last subsection.

Fix any y=y0>0y=y_{0}>0. Then D1D_{1} can be viewed as a d-bar operator on sections of E|y=y0E|_{y=y_{0}}. So it gives E|y=y0E|_{y=y_{0}} a holomorphic structure. Note that since [D1,D2]=[D1,D3]=0[D_{1},D_{2}]=[D_{1},D_{3}]=0, both D2D_{2} and D3D_{3} keeps this holomorphic structure.

On the y=y0y=y_{0} slice, we may choose (at least locally) two holomorphic sections of EE, namely s1s_{1} and s2s_{2}. Since D1D_{1} keeps the SL(2,)SL(2,\mathbb{C}) structure, we may assume that s1s2=1s_{1}\wedge s_{2}=1 everywhere.

We may identify the choices of s1,s2s_{1},s_{2} at different yy slices by requiring D3s1=D3s2=0D_{3}s_{1}=D_{3}s_{2}=0. Then s1s_{1} and s2s_{2} form a basis of EE. Under this basis, D1D_{1} is just 2¯=1+i22\bar{\partial}=\partial_{1}+i\partial_{2} and D3D_{3} is just y\partial_{y}. The operator D2D_{2} is represented by an sl(2,)sl(2,\mathbb{C}) matrix whose items are holomorphic in zz and doesn’t depend on yy. We call this matrix φ\varphi. Note that [D1,D2]=[D2,D3]=[D3,D1]=0[D_{1},D_{2}]=[D_{2},D_{3}]=[D_{3},D_{1}]=0 is obvious in this basis.

Note that the definition of (H,φ)(H,\varphi) depends on a choice of the holomorphic basis s1,s2s_{1},s_{2}. Different choices of the basis can be related by a holomorphic SL(2,)SL(2,\mathbb{C}) gauge transformation gg (an SL(2,)SL(2,\mathbb{C}) matrix whose items are holomorphic functions in zz) which sends (H,φ)(H,\varphi) to (gHg,g1φg)(g^{*}Hg,g^{-1}\varphi g).

A warning: The holomorphic SL(2,)SL(2,\mathbb{C}) gauge transformation gg is not an actual gauge transformation on the original configuarition Ψ\Psi. In fact, it doesn’t affect Ψ\Psi at all. It only changes (H,φ)(H,\varphi) by changing the pairs (s1,s2)(s_{1},s_{2}) that we use to define it. Readers should not get confused by the (somehow misleading) term “gauge transformation” that we use here.

From (H,φ)(H,\varphi) to Ψ\Psi

Suppose we have a pair (H,φ)(H,\varphi). Recall that the “trivial” configuration Ψφ\Psi_{\varphi} satisfies all but the first bullets of the extended Bogomolny equations.

We may choose an SL(2,)SL(2,\mathbb{C}) gange transformation gg such that gg=Hg^{*}g=H. Then gg sends Ψφ\Psi_{\varphi} to some configuration Ψ\Psi. Clearly Ψ\Psi corresponds to (H,φ)(H,\varphi).

Note that adding another SU(2)SU(2) gauge transformation uu to gg doesn’t change H=ggH=g^{*}g. In fact, if we replace gg by ugug, we still have

(ug)(ug)=gg=H.(ug)^{*}(ug)=g^{*}g=H.

So essentially, what we get is a configuration Ψ\Psi up to an SU(2)SU(2) gauge transformation.

For each pair (H,φ)(H,\varphi), there is a preferred way to choose gg which we’ll frequently use in this paper: Each SU(2)SU(2) Hermitian metric HH can be written in the following format:

H=(h+h1|w|2h1w¯h1wh1),H=\begin{pmatrix}h+h^{-1}|w|^{2}&h^{-1}\bar{w}\\ h^{-1}w&h^{-1}\end{pmatrix},

where hh is a positive real-valued function and ww is a complex-valued function. Then we have H=ggH=g^{*}g, where g=(h120h12wh12)g=\begin{pmatrix}h^{\frac{1}{2}}&0\\ h^{-\frac{1}{2}}w&h^{-\frac{1}{2}}\end{pmatrix}. Under the SL(2,)SL(2,\mathbb{C}) gauge transformation given by this specifically chosen gg, the trivial configuration Ψφ\Psi_{\varphi} becomes

{Φ=Φ1iΦ2=gφg1=(AwBhBh1(wA+Pw2B+wA)wBA),Φ3=i2h(yhyw¯ywyh),Ay=12h(0yw¯yw0),A1=12h(i2h2w¯2¯wi2h),A2=i2h(1h2w¯2¯w1h).\left\{\begin{array}[]{lr}\Phi=\Phi_{1}-i\Phi_{2}=g\varphi g^{-1}=\begin{pmatrix}A-wB&hB\\ h^{-1}(wA+P-w^{2}B+wA)&wB-A\end{pmatrix},&\\ \Phi_{3}=\dfrac{i}{2h}\begin{pmatrix}-\partial_{y}h&-\partial_{y}\bar{w}\\ -\partial_{y}w&\partial_{y}h\end{pmatrix},&\\ A_{y}=\dfrac{1}{2h}\begin{pmatrix}0&\partial_{y}\bar{w}\\ -\partial_{y}w&0\end{pmatrix},&\\ A_{1}=\dfrac{1}{2h}\begin{pmatrix}-i\partial_{2}h&2\partial\bar{w}\\ -2\bar{\partial}w&i\partial_{2}h\end{pmatrix},&\\ A_{2}=\dfrac{i}{2h}\begin{pmatrix}\partial_{1}h&2\partial\bar{w}\\ 2\bar{\partial}w&-\partial_{1}h\end{pmatrix}.\end{array}\right.

We use ΨH,φ\Psi_{H,\varphi} to denote this particular configuration that corresponds to (H,φ)(H,\varphi). Any other configuration that corresponds to (H,φ)(H,\varphi) is SU(2)SU(2) gauge equivalent to ΨH,φ\Psi_{H,\varphi}.

For each H=(eu+eu|w|2euw¯euweu)H=\begin{pmatrix}e^{u}+e^{-u}|w|^{2}&e^{-u}\bar{w}\\ e^{-u}w&e^{-u}\end{pmatrix} and φ\varphi, we write down the concrete formulas of V(H,φ):=V(ΨH,φ)V(H,\varphi):=V(\Psi_{H,\varphi}) in terms of HH and φ\varphi in two special situations:

Special case 1

H=(eu+eu|w|2euw¯euweu),φ=(00P(z)0).H=\begin{pmatrix}e^{u}+e^{-u}|w|^{2}&e^{-u}\bar{w}\\ e^{-u}w&e^{-u}\end{pmatrix},~{}~{}\varphi=\begin{pmatrix}0&0\\ P(z)&0\end{pmatrix}.

Then V(H,φ)=12(EF¯FE)V(H,\varphi)=\dfrac{1}{2}\begin{pmatrix}E&\bar{F}\\ F&-E\end{pmatrix}, with

E=Δu+e2u(4|¯w|2+|yw|2+|P|2),F=eu(Δw2(yu)(yw)8(¯w)(u)).E=\Delta u+e^{-2u}(4|\bar{\partial}w|^{2}+|\partial_{y}w|^{2}+|P|^{2}),~{}~{}F=e^{-u}(\Delta w-2(\partial_{y}u)(\partial_{y}w)-8(\bar{\partial}w)(\partial u)).
Proof.

In this special case, suppose h=euh=e^{u}. Then

{Φ=Φ1iΦ2=(00h1P0),Φ3=i2h(yhyw¯ywyh),Ay=12h(0yw¯yw0),A1=12h(i2h2w¯2¯wi2h),A2=i2h(1h2w¯2¯w1h).\left\{\begin{array}[]{lr}\Phi=\Phi_{1}-i\Phi_{2}=\begin{pmatrix}0&0\\ h^{-1}P&0\end{pmatrix},&\\ \Phi_{3}=\dfrac{i}{2h}\begin{pmatrix}-\partial_{y}h&-\partial_{y}\bar{w}\\ -\partial_{y}w&\partial_{y}h\end{pmatrix},&\\ A_{y}=\dfrac{1}{2h}\begin{pmatrix}0&\partial_{y}\bar{w}\\ -\partial_{y}w&0\end{pmatrix},&\\ A_{1}=\dfrac{1}{2h}\begin{pmatrix}-i\partial_{2}h&2\partial\bar{w}\\ -2\bar{\partial}w&i\partial_{2}h\end{pmatrix},&\\ A_{2}=\dfrac{i}{2h}\begin{pmatrix}\partial_{1}h&2\partial\bar{w}\\ 2\bar{\partial}w&-\partial_{1}h\end{pmatrix}.\end{array}\right.

The first bullet of the extended Bogomolny equations is:

V(H,φ)=(1A22A1+[A1,A2])(yΦ3+[Ay,Φ3]+[Φ1,Φ2]).V(H,\varphi)=(\partial_{1}A_{2}-\partial_{2}A_{1}+[A_{1},A_{2}])-(\partial_{y}\Phi_{3}+[A_{y},\Phi_{3}]+[\Phi_{1},\Phi_{2}]).

We have

{1A22A1+[A1,A2]=(2i)h2(h(¯h)|¯h|2+|¯w|2h(¯w¯)2(w¯)(¯h)h(¯w)2(¯w)(h)h(¯h)+|¯h|2|¯w|2),yΦ3+[Ay,Φ3][Φ1,Φ2]=i2h2(h(y2h)+(yh)2|yw|2|P|2h(y2w¯)+2(yh)(yw¯)h(y2w)+2(yh)(yw)h(y2h)(yh)2+|yw|2+|P|2).\left\{\begin{array}[]{lr}\partial_{1}A_{2}-\partial_{2}A_{1}+[A_{1},A_{2}]&\\ ~{}~{}~{}~{}=\dfrac{(2i)}{h^{2}}\begin{pmatrix}h(\partial\bar{\partial}h)-|\bar{\partial}h|^{2}+|\bar{\partial}w|^{2}&h(\bar{\partial}\partial\bar{w})-2(\partial\bar{w})(\bar{\partial}h)\\ h(\partial\bar{\partial}w)-2(\bar{\partial}w)(\partial h)&-h(\partial\bar{\partial}h)+|\bar{\partial}h|^{2}-|\bar{\partial}w|^{2}\end{pmatrix},&\\ \partial_{y}\Phi_{3}+[A_{y},\Phi_{3}]-[\Phi_{1},\Phi_{2}]&\\ ~{}~{}~{}~{}=\dfrac{i}{2h^{2}}\begin{pmatrix}-h(\partial_{y}^{2}h)+(\partial_{y}h)^{2}-|\partial_{y}w|^{2}-|P|^{2}&-h(\partial_{y}^{2}\bar{w})+2(\partial_{y}h)(\partial_{y}\bar{w})\\ -h(\partial_{y}^{2}w)+2(\partial_{y}h)(\partial_{y}w)&h(\partial_{y}^{2}h)-(\partial_{y}h)^{2}+|\partial_{y}w|^{2}+|P|^{2}\end{pmatrix}.\end{array}\right.

So indeed V(H,φ)=12(EF¯FE)V(H,\varphi)=\dfrac{1}{2}\begin{pmatrix}E&\bar{F}\\ F&-E\end{pmatrix}, with

E=h2(hΔh|h|2+4|¯w|2+|yw|2+|P|2)=Δu+e2u(4|¯w|2+|yw|2+|P|2),E=h^{-2}(h\Delta h-|\nabla h|^{2}+4|\bar{\partial}w|^{2}+|\partial_{y}w|^{2}+|P|^{2})=\Delta u+e^{-2u}(4|\bar{\partial}w|^{2}+|\partial_{y}w|^{2}+|P|^{2}),
F=h2(h(Δw)2(yh)(yw)8(¯w)(h))=eu(Δw2(yu)(yw)8(¯w)(u)).F=h^{-2}(h(\Delta w)-2(\partial_{y}h)(\partial_{y}w)-8(\bar{\partial}w)(\partial h))=e^{-u}(\Delta w-2(\partial_{y}u)(\partial_{y}w)-8(\bar{\partial}w)(\partial u)).

Special case 2

H=(eu00eu),φ=(A(z)B(z)P(z)A(z)).H=\begin{pmatrix}e^{u}&0\\ 0&e^{-u}\end{pmatrix},~{}~{}\varphi=\begin{pmatrix}A(z)&B(z)\\ P(z)&-A(z)\end{pmatrix}.

Then V(H,φ)=12(EF¯FE)V(H,\varphi)=\dfrac{1}{2}\begin{pmatrix}E&\bar{F}\\ F&-E\end{pmatrix}, with

E=Δu+e2u|P|2e2u|B|2,F=2eu(BA¯)eu(AP¯+PA¯).E=\Delta u+e^{-2u}|P|^{2}-e^{2u}|B|^{2},~{}~{}~{}F=2e^{u}(B\bar{A})-e^{-u}(A\bar{P}+P\bar{A}).
Proof.

In this special case, suppose h=euh=e^{u}. Then

{Φ=Φ1iΦ2=(AhBh1PA),Φ3=i2h(yh00yh),Ay=0,A1=12h(i2h00i2h),A2=i2h(1h001h).\left\{\begin{array}[]{lr}\Phi=\Phi_{1}-i\Phi_{2}=\begin{pmatrix}A&hB\\ h^{-1}P&-A\end{pmatrix},&\\ \Phi_{3}=\dfrac{i}{2h}\begin{pmatrix}-\partial_{y}h&0\\ 0&\partial_{y}h\end{pmatrix},&\\ A_{y}=0,&\\ A_{1}=\dfrac{1}{2h}\begin{pmatrix}-i\partial_{2}h&0\\ 0&i\partial_{2}h\end{pmatrix},&\\ A_{2}=\dfrac{i}{2h}\begin{pmatrix}\partial_{1}h&0\\ 0&-\partial_{1}h\end{pmatrix}.\end{array}\right.

We have

{1A22A1+[A1,A2]=(2i)h2(h(¯h)|¯h|200h(¯h)+|¯h|2),yΦ3+[Ay,Φ3]=i2h2(h(y2h)+(yh)200h(y2h)(yh)2).\left\{\begin{array}[]{lr}\partial_{1}A_{2}-\partial_{2}A_{1}+[A_{1},A_{2}]&\\ ~{}~{}~{}~{}=\dfrac{(2i)}{h^{2}}\begin{pmatrix}h(\partial\bar{\partial}h)-|\bar{\partial}h|^{2}&0\\ 0&-h(\partial\bar{\partial}h)+|\bar{\partial}h|^{2}\end{pmatrix},&\\ \partial_{y}\Phi_{3}+[A_{y},\Phi_{3}]&\\ ~{}~{}~{}~{}=\dfrac{i}{2h^{2}}\begin{pmatrix}-h(\partial_{y}^{2}h)+(\partial_{y}h)^{2}&0\\ 0&h(\partial_{y}^{2}h)-(\partial_{y}h)^{2}\end{pmatrix}.\end{array}\right.

And

[Φ1,Φ2]=12i[Φ,Φ]=12i((AhBh1PA)(A¯h1P¯hB¯A¯)(A¯h1P¯hB¯A¯)(AhBh1PA))[\Phi_{1},\Phi_{2}]=\dfrac{1}{2i}[\Phi,\Phi^{*}]=\dfrac{1}{2i}(\begin{pmatrix}A&hB\\ h^{-1}P&-A\end{pmatrix}\begin{pmatrix}\bar{A}&h^{-1}\bar{P}\\ h\bar{B}&-\bar{A}\end{pmatrix}-\begin{pmatrix}\bar{A}&h^{-1}\bar{P}\\ h\bar{B}&-\bar{A}\end{pmatrix}\begin{pmatrix}A&hB\\ h^{-1}P&-A\end{pmatrix})
=12i(h2|B|2h2|P|2h1(AP¯+P¯A)2h(BA¯)h1(A¯P+PA¯)2h(B¯A)h2|P|2h2|B|2).=\dfrac{1}{2i}\begin{pmatrix}h^{2}|B|^{2}-h^{-2}|P|^{2}&h^{-1}(A\bar{P}+\bar{P}A)-2h(B\bar{A})\\ h^{-1}(\bar{A}P+P\bar{A})-2h(\bar{B}A)&h^{-2}|P|^{2}-h^{2}|B|^{2}\end{pmatrix}.

So indeed

E=h2(hΔh|h|2h4|B|2+|P|2)=Δu+e2u|P|2e2u|B|2,E=h^{-2}(h\Delta h-|\nabla h|^{2}-h^{4}|B|^{2}+|P|^{2})=\Delta u+e^{-2u}|P|^{2}-e^{2u}|B|^{2},
F=2h(BA¯)h1(AP¯+PA¯)=2eu(BA¯)eu(AP¯+PA¯).F=2h(B\bar{A})-h^{-1}(A\bar{P}+P\bar{A})=2e^{u}(B\bar{A})-e^{-u}(A\bar{P}+P\bar{A}).

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 Ψ0\Psi_{0} 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 SL(2,)SL(2,\mathbb{C}) gauge transformation to make the first bullet vanish. So we need to study how the first bullet behaves under SL(2,)SL(2,\mathbb{C}) gauge transformations.

Suppose Ψ\Psi 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

V(Ψ):=[D1,D1]+[D2,D2]+[D3,D3]=0.V(\Psi):=[D_{1},D_{1}^{*}]+[D_{2},D_{2}^{*}]+[D_{3},D_{3}^{*}]=0.

Suppose uu is an SU(2)SU(2) gauge transformation sending Ψ\Psi to u(Ψ)u(\Psi). We may think uu as an SU(2)SU(2)-valued function on XX. Then V(u(Ψ))=uV(Ψ)u1V(u(\Psi))=uV(\Psi)u^{-1}. And the norm |V(Ψ)||V(\Psi)| is invariant under this transformation.

On the other hand, suppose gg is an SL(2,)SL(2,\mathbb{C}) gauge transformation. If in addition, g=gg^{*}=g, then we call it a Hermitian SL(2,)SL(2,\mathbb{C}) gauge transformation.

The following fact is simple in linear algebra: For each gSL(2,)g\in SL(2,\mathbb{C}), there is a unique uSU(2)u\in SU(2) such that ugug is Hermitian. Because ug(Ψ)ug(\Psi) and g(Ψ)g(\Psi) are SU(2)SU(2) equivalent. If we are not sensitive to SU(2)SU(2) transformations on what we get, we may only consider Hermitian SL(2,)SL(2,\mathbb{C}) gauge transformation when we deform Ψ\Psi.

Suppose Ψ\Psi is a configuration. Given any section sisu(2)s\in isu(2), there is a 1-parameter family of Hermitian SL(2,)SL(2,\mathbb{C}) gauge transformations etse^{ts} parameterized by tt\in\mathbb{R}. Since sl(2,)=su(2)isu(2)sl(2,\mathbb{C})=su(2)\oplus isu(2), this deformation is perpendicular with SU(2)SU(2) gauge transformations initially.

On the other hand, any Hermitian SL(2,)SL(2,\mathbb{C}) valued function gg can be written as either ese^{s} or es-e^{s} for some sisu(2)s\in isu(2) uniquely. Note that es-e^{s} and ese^{s} act the same way on Ψ\Psi as SL(2,)SL(2,\mathbb{C}) gauge transformations. So effectively any Hermitian SL(2,)SL(2,\mathbb{C}) gauge transformation can be represented by ese^{s} for some sisu(2)s\in isu(2).

Deformation of the equation

Suppose ss is a section in isu(2)isu(2). We consider the Hermitian SL(2,)SL(2,\mathbb{C}) gauge transformation ese^{s} acting on Ψ\Psi. What we get is written as Ψs\Psi_{s}. Then the first bullet of the extended Bogomolny equations is written as

V(Ψ,s):=V(Ψs)=i=13[Di(Di(es))es,Di((Di(es))es)]V(\Psi,s):=V(\Psi_{s})=\sum\limits_{i=1}^{3}[D_{i}-(D_{i}(e^{s}))e^{-s},D_{i}^{*}-((D_{i}(e^{s}))e^{-s})^{*}]
=i=13[Di(Di(es))es,Di+es(Di(es))],=\sum\limits_{i=1}^{3}[D_{i}-(D_{i}(e^{s}))e^{-s},D_{i}^{*}+e^{-s}(D^{*}_{i}(e^{s}))],

where D1,D2,D3D_{1},D_{2},D_{3} are given by Ψ\Psi. Note that the last inequality used the fact that s=ss^{*}=s.

This formula seems awful. But actually we only need two facts:

The first fact

is

V(Ψ,s)=V(Ψ)(γ(s)+γ(s))(ΔΨs)+12(γ(s)γ(s))([V(Ψ),s])+``remaining terms",V(\Psi,s)=V(\Psi)~{}~{}-~{}~{}(\gamma(s)+\gamma(-s))(\Delta_{\Psi}s)~{}~{}+~{}~{}\dfrac{1}{2}(\gamma(s)-\gamma(-s))([V(\Psi),s])~{}~{}+~{}~{}``\text{remaining terms}",

here the “remaining terms” include multi-linear terms R(sk(Dis)2)R(s^{\otimes k}\otimes(D_{i}s)^{\otimes 2}) and R(sk(Dis)2)R(s^{\otimes k}\otimes(D_{i}^{*}s)^{\otimes 2}) with any integer k0k\geq 0 with convergent coefficients (comparible with the coefficients in the Taylor expansion of ese^{s}). The operator ΔΨ\Delta_{\Psi} behaves like the Laplacian, whose definition is

ΔΨs:=i=1,2,y(i2s)+i=1,2,3([Φi,[Φi,s]]),\Delta_{\Psi}s:=\sum\limits_{i=1,2,y}(\nabla_{i}^{2}s)+\sum\limits_{i=1,2,3}([\Phi_{i},[\Phi_{i},s]]),

where is=is+[Ai,s]\nabla_{i}s=\partial_{i}s+[A_{i},s], and Ai,ΦiA_{i},\Phi_{i} are six components of Ψ\Psi.

Proof.

We have

V(Ψs)=i=13[Di(Di(es))es,Di+es(Di(es))]=i=13([Di,Di])V(\Psi_{s})=\sum\limits_{i=1}^{3}[D_{i}-(D_{i}(e^{s}))e^{-s},D_{i}^{*}+e^{-s}(D_{i}^{*}(e^{s}))]=\sum\limits_{i=1}^{3}([D_{i},D_{i}^{*}])
+i=13(es(DiDi(es))+(DiDi(es))es)+``remaning terms"+\sum\limits_{i=1}^{3}(e^{-s}(D_{i}D_{i}^{*}(e^{s}))+(D_{i}^{*}D_{i}(e^{s}))e^{-s})+``\text{remaning terms}"
=V(Ψ)+i=13(γ(s)(DiDis)+γ(s)(DiDis))+``remaining terms".=V(\Psi)+\sum\limits_{i=1}^{3}(\gamma(-s)(D_{i}D_{i}^{*}s)+\gamma(s)(D_{i}^{*}D_{i}s))+``\text{remaining terms}".

Recall that 1,2,y\nabla_{1},\nabla_{2},\nabla_{y} is the connection defined using A1,A2,AyA_{1},A_{2},A_{y}. And Φ=Φ1iΦ2\Phi=\Phi_{1}-i\Phi_{2}. We have i=i+Ai=i\nabla_{i}^{*}=\partial_{i}^{*}+A_{i}^{*}=-\nabla_{i}. We have

D1D1s+D2D2s+D3D3sD_{1}D_{1}^{*}s+D_{2}D_{2}^{*}s+D_{3}D_{3}^{*}s
=(1+i2)(1i2)s+[Φ1iΦ2,[Φ1+iΦ2,s]]+(yiΦ3)(y+iΦ3)s=(\nabla_{1}+i\nabla_{2})(\nabla_{1}^{*}-i\nabla_{2}^{*})s+[\Phi_{1}-i\Phi_{2},[\Phi_{1}^{*}+i\Phi_{2}^{*},s]]+(\nabla_{y}-i\Phi_{3})(\nabla_{y}^{*}+i\Phi_{3}^{*})s
=i=1,2,y(i2s)i=1,2,3([Φi,[Φi,s]])12[V(Ψ),s]=ΔΨ(s)12[V(Ψ),s].=-\sum\limits_{i=1,2,y}(\nabla_{i}^{2}s)-\sum\limits_{i=1,2,3}([\Phi_{i},[\Phi_{i},s]])-\dfrac{1}{2}[V(\Psi),s]=-\Delta_{\Psi}(s)-\dfrac{1}{2}[V(\Psi),s].

And

i=13DiDis=i=13(DiDis+[Di,Di]s)=ΔΨs+12[V(Ψ),s].\sum\limits_{i=1}^{3}D_{i}^{*}D_{i}s=\sum\limits_{i=1}^{3}(D_{i}^{*}D_{i}s+[D_{i},D_{i}^{*}]s)=-\Delta_{\Psi}s+\dfrac{1}{2}[V(\Psi),s].

Thus the first fact follows directly.

There are two alternative ways to write the first fact that will be useful later:

V(Ψ,s)=V(Ψ)+L~Ψ(s)+Q~(s)=V(Ψ)+L(s)+Q(s),V(\Psi,s)=V(\Psi)+\tilde{L}_{\Psi}(s)+\tilde{Q}(s)=V(\Psi)+L(s)+Q(s),

where L~Ψ=(γ(s)+γ(s))ΔΨ\tilde{L}_{\Psi}=-(\gamma(s)+\gamma(-s))\Delta_{\Psi} is the second order term and Q~(s)\tilde{Q}(s) is the lower order terms; LΨ=ΔΨL_{\Psi}=-\Delta_{\Psi} is the linear term and Q(s)Q(s) contains all the non-linear terms.

The second fact

is a Weitzenbock type of formula:

<V(Ψs)V(Ψ),s>=Δ(|s|2)+i=132|v(2s)Di(s)|2,<V(\Psi_{s})-V(\Psi),s>=-\Delta(|s|^{2})+\sum\limits_{i=1}^{3}2|v(-2s)D_{i}^{*}(s)|^{2},

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 sisu(2)s\in isu(2), that is, Hermitian. It is generally false if only ssl(2,)s\in sl(2,\mathbb{C}).

Proof.
<V(Ψs)V(Ψ),s>=<i=13([Di(Di(es))es,Di+es(Di(es))][Di,Di]),s><V(\Psi_{s})-V(\Psi),s>=<\sum\limits_{i=1}^{3}([D_{i}-(D_{i}(e^{s}))e^{-s},D_{i}^{*}+e^{-s}(D_{i}^{*}(e^{s}))]-[D_{i},D_{i}^{*}]),s>
=i=13<(Di(esDi(es))+Di(Di(es)es)[(Di(es)es,esDi(es))]),s>.=\sum\limits_{i=1}^{3}<(D_{i}(e^{-s}D_{i}^{*}(e^{s}))+D_{i}^{*}(D_{i}(e^{s})e^{-s})-[(D_{i}(e^{s})e^{-s},e^{-s}D_{i}^{*}(e^{s}))]),s>.

The following identity can be checked directly:

Di(esDi(es))+Di(Di(es)es)[(Di(es)es,esDi(es))]D_{i}(e^{-s}D_{i}^{*}(e^{s}))+D_{i}^{*}(D_{i}(e^{s})e^{-s})-[(D_{i}(e^{s})e^{-s},e^{-s}D_{i}^{*}(e^{s}))]
=es(Di(e2sDi(e2s)))es(DiDi(es)DiDi(es))es.=e^{s}(D_{i}(e^{-2s}D_{i}^{*}(e^{2s})))e^{-s}-(D_{i}D_{i}^{*}(e^{s})-D_{i}^{*}D_{i}(e^{s}))e^{-s}.

Thus

<V(Ψs)V(Ψ),s>=i=13<es(Di(e2sDi(e2s)))es(DiDi(es)DiDi(es))es,s><V(\Psi_{s})-V(\Psi),s>=\sum\limits_{i=1}^{3}<e^{s}(D_{i}(e^{-2s}D_{i}^{*}(e^{2s})))e^{-s}-(D_{i}D_{i}^{*}(e^{s})-D_{i}^{*}D_{i}(e^{s}))e^{-s},s>
=i=13(<Di(e2sDi(e2s)),s>)<[V(Ψ),es]es,s>.=\sum\limits_{i=1}^{3}(<D_{i}(e^{-2s}D_{i}^{*}(e^{2s})),s>)-<[V(\Psi),e^{s}]e^{-s},s>.

Here we used the fact that AdesAd_{e^{s}} is self-adjoint and Ades(s)=sAd_{e^{s}}(s)=s. Moreover,

<[V(Ψ),es]es,s>=<γ(s)([V(Ψ),s]),s>=<[V(Ψ),s],s>=0.<[V(\Psi),e^{s}]e^{-s},s>=<\gamma(s)([V(\Psi),s]),s>=<[V(\Psi),s],s>=0.
<D1(e2sD1(e2s)),s>=<(1+i2)(e2sD1(e2s))+[A1+iA2,(e2sD1(e2s))],s><D_{1}(e^{-2s}D_{1}^{*}(e^{2s})),s>=<(\partial_{1}+i\partial_{2})(e^{-2s}D_{1}^{*}(e^{2s}))+[A_{1}+iA_{2},(e^{-2s}D_{1}^{*}(e^{2s}))],s>
=1(<γ(2s)(D1(2s)),s>)+2(<iγ(2s)(D1(2s)),s>)+<γ(2s)D1(2s),D1s>=\partial_{1}(<\gamma(-2s)(D_{1}^{*}(2s)),s>)+\partial_{2}(<i\gamma(-2s)(D_{1}^{*}(2s)),s>)+<\gamma(-2s)D_{1}^{*}(2s),D_{1}^{*}s>
=21(<D1(s),s>)+22(<iD1s,s>)+2|v(2s)D1(s)|2.=2\partial_{1}(<D_{1}^{*}(s),s>)+2\partial_{2}(<iD_{1}^{*}s,s>)+2|v(-2s)D_{1}^{*}(s)|^{2}.

The last step used the fact that v(2s)v(-2s) and γ(2s)\gamma(-2s) are self-adjoint and γ(2s)(s)=s\gamma(-2s)(s)=s. Similarly,

<D2(e2sD2(e2s)),s>=2|v(2s)D2(s)|2.<D_{2}(e^{-2s}D_{2}^{*}(e^{2s})),s>=2|v(-2s)D_{2}^{*}(s)|^{2}.
<D3(e2sD3(e2s)),s>=<y(e2sD3(e2s))+[AyiΦ3,(e2sD3(e2s))],s><D_{3}(e^{-2s}D_{3}^{*}(e^{2s})),s>=<\partial_{y}(e^{-2s}D_{3}^{*}(e^{2s}))+[A_{y}-i\Phi_{3},(e^{-2s}D_{3}^{*}(e^{2s}))],s>
=2y(<D3(s),s>)+2|v(2s)D3(s)|2.=2\partial_{y}(<D_{3}^{*}(s),s>)+2|v(-2s)D_{3}^{*}(s)|^{2}.

Finally,

21(<D1(s),s>)+22(<iD1s,s>)+2y(<D3(s),s>)2\partial_{1}(<D_{1}^{*}(s),s>)+2\partial_{2}(<iD_{1}^{*}s,s>)+2\partial_{y}(<D_{3}^{*}(s),s>)
=21<1s+i2s,s>+22<i1s2s,s>+2y<ys,s>=Δ(|s|2).=2\partial_{1}<-\partial_{1}s+i\partial_{2}s,s>+2\partial_{2}<-i\partial_{1}s-\partial_{2}s,s>+2\partial_{y}<-\partial_{y}s,s>=-\Delta(|s|^{2}).

So

<V(Ψs)V(Ψ),s>=Δ(|s|2)+i=132|v(2s)Di(s)|2.<V(\Psi_{s})-V(\Psi),s>=-\Delta(|s|^{2})+\sum\limits_{i=1}^{3}2|v(-2s)D_{i}^{*}(s)|^{2}.

Adding Hermitian SL(2,)SL(2,\mathbb{C}) gauge transformations

We need to be very careful to add two Hermitian SL(2,)SL(2,\mathbb{C}) gauge transformations. If we have two different Hermitian SL(2,)SL(2,\mathbb{C}) gauge transformations es1e^{s_{1}} and es2e^{s_{2}}. If we apply es2e^{s_{2}} first and then es1e^{s_{1}} on a configuration Ψ\Psi, what we get is es1es2(Ψ)e^{s_{1}}e^{s_{2}}(\Psi), which is generally not equivalent with es1+s2(Ψ)e^{s_{1}+s_{2}}(\Psi).

Here is the correct way to add them: Let σ\sigma satisfy e2σ=es2e2s1es2e^{2\sigma}=e^{s_{2}}e^{2s_{1}}e^{s_{2}}. Note that es2e2s1es2e^{s_{2}}e^{2s_{1}}e^{s_{2}} is Hermitian. So σisu(2)\sigma\in isu(2) exists and is unique.

Moreover, (es1es2eσ)(es1es2eσ)=I(e^{s_{1}}e^{s_{2}}e^{-\sigma})^{*}(e^{s_{1}}e^{s_{2}}e^{-\sigma})=I. So u=(es1es2eσ)u=(e^{s_{1}}e^{s_{2}}e^{-\sigma}) is in SU(2)SU(2). And es1es2=ueσe^{s_{1}}e^{s_{2}}=ue^{\sigma}. So es1es2(Ψ)e^{s_{1}}e^{s_{2}}(\Psi) is SU(2)SU(2) equivalent with eσ(Ψ)e^{\sigma}(\Psi).

Sometimes, we still use the sloppy notation s1+s2s_{1}+s_{2} to represent σ\sigma mentioned above if there is not a potential confusion. Note that if |ks1||\nabla^{k}s_{1}| and |ks2||\nabla^{k}s_{2}| are all bounded for kKk\leq K, where k\nabla^{k} is the kkth derivative and KK is a non-negative integer, then |Kσ||\nabla^{K}\sigma| 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, s1+s2s2+s1s_{1}+s_{2}\neq s_{2}+s_{1}. And the bound of |Kσ||\nabla^{K}\sigma| may also depend on the lower order derivatives of s1s_{1} and s2s_{2}.

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 Ψ\Psi corresponds to a pair (H,φ)(H,\varphi) such that

H=(eu00eu),φ=(00φ(z)0).H=\begin{pmatrix}e^{u}&0\\ 0&e^{-u}\end{pmatrix},~{}~{}~{}\varphi=\begin{pmatrix}0&0\\ \varphi(z)&0\end{pmatrix}.

In another word, HH can be diagonalized and φ\varphi 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 y0y\rightarrow 0 and real symmetry breaking condition as y+y\rightarrow+\infty.

Based on the special case 1 in subsection 1.2, the extended Bogomolny equations are reduced to a scalar equation:

Δu+e2u|P(z)|2=0.()\Delta u+e^{-2u}|P(z)|^{2}=0.~{}~{}~{}~{}~{}(*)

This section finds a solution to ()(*) that gives us a solution to the extended Bogomolny equations with Nahm pole boundary condition as y0y\rightarrow 0 and real symmetry breaking condition as y+y\rightarrow+\infty. 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 uu.

The generalized Nahm pole boundary condition as y0y\rightarrow 0

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 P(z)P(z), there is a unique solution to ()(*), denoted as u=u0u=u_{0}, that has the following properties:

  • When y0y\rightarrow 0, it satisfies the generalized Nahm pole boundary condition.

  • When R+R\rightarrow+\infty, u0=NlnR+lny+O(1)u_{0}=N\ln R+\ln y+O(1), uniformly with respect to yR\dfrac{y}{R}.

The following feature of u0u_{0} will be useful later: On the entire XX (either y0y\rightarrow 0 or y+y\rightarrow+\infty), we have

eu0|P(z)|=O(1y).e^{-u_{0}}|P(z)|=O(\dfrac{1}{y}).

Note that the definition of u0u_{0} 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 uu as follows:

Let Ψ0\Psi_{0} be the configuration that is represented by u0u_{0}. Then to say that a general configuration Ψ\Psi satisfies the generalized Nahm pole boundary condition means: There is a positive number ϵ\epsilon such that, possibly after an SU(2)SU(2) gauge transformation,

Ψ=Ψ0+O(y1+ϵ),asy0.\Psi=\Psi_{0}+O(y^{-1+\epsilon}),~{}~{}\text{as}~{}~{}y\rightarrow 0.

If we have a solution of ()(*), then the following condition on uu implies that it represents a solution with the generalized Nahm pole boundary condition:

|(uu0)|=O(y1+ϵ),u=u0+O(yϵ),asy0.|\nabla(u-u_{0})|=O(y^{-1+\epsilon}),~{}~{}~{}u=u_{0}+O(y^{\epsilon}),~{}~{}~{}\text{as}~{}~{}y\rightarrow 0.

Real symmetry breaking as y+y\rightarrow+\infty

Definition 2.2.

We say that a configuration (A1,A2,Ay,Φ1,Φ2,Φ3)(A_{1},A_{2},A_{y},\Phi_{1},\Phi_{2},\Phi_{3}) satisfies the real symmetry breaking condition if, possibly after an SU(2)SU(2) gauge transformation, for some ϵ>0\epsilon>0, when y+y\rightarrow+\infty

A1,A2,Ay,Φ1,Φ2=O(yϵ),Φ3=12(i00i)+O(yϵ).A_{1},A_{2},A_{y},\Phi_{1},\Phi_{2}=O(y^{-\epsilon}),~{}~{}\Phi_{3}=\dfrac{1}{2}\begin{pmatrix}-i&0\\ 0&i\end{pmatrix}+O(y^{-\epsilon}).

If we have a solution of ()(*), then the following condition on uu implies that it represents a solution with the real symmetry breaking: As y+y\rightarrow+\infty,

1u,2u,eu|P(z)|=O(yϵ),yu=1+O(yϵ).\partial_{1}u,\partial_{2}u,e^{-u}|P(z)|=O(y^{-\epsilon}),~{}~{}\partial_{y}u=1+O(y^{-\epsilon}).

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 uu such that

Δue2u|P(z)|20-\Delta u-e^{-2u}|P(z)|^{2}\leq 0

is called a sub-solution. A function uu such that

Δue2u|P(z)|20-\Delta u-e^{-2u}|P(z)|^{2}\geq 0

is called a super-solution.

We always assume uu is continuous, but allow the possibility that uu is not differentiable. If this is the case, then a sub-solution/super-solution is defined in the weak sense: For example, to say

Δue2u|P(z)|20-\Delta u-e^{-2u}|P(z)|^{2}\geq 0

means for any non-negative smooth function vv on XX supported on a compact region, we have

X(uΔve2u|P(z)|2v)0.\int_{X}(-u\Delta v-e^{-2u}|P(z)|^{2}v)\geq 0.

If uu 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 uu is a weak solution. Choose any ball whose closure is in XX. Since Δu+e2u|P(z)|2=0\Delta u+e^{-2u}|P(z)|^{2}=0 in the ball and e2u|P(z)|2L2(B)e^{-2u}|P(z)|^{2}\in L^{2}(B), using the standard elliptic regularity argument, we see that uW2,2(B)u\in W^{2,2}(B^{\prime}), where BB^{\prime} is a slightly smaller ball in BB, W2,2W^{2,2} is the Sobolev space. This further implies that e2u|P(z)|2W2,2(B)e^{-2u}|P(z)|^{2}\in W^{2,2}(B^{\prime}) using the Chain rule for derivatives, the fact that uu is continuous and bounded, and the Sobolev embedding/multiplication inequalities in a 3-dimensional space. This further implies that uW2,4u\in W^{2,4} in a even smaller ball. We can do boot-strapping and conclude that uW2,nu\in W^{2,n} for any positive nn in a smaller ball. So it is smooth in that small ball. Since the ball is arbitrary chose, uu is smooth everywhere in XX. In particular, it is an actual solution.

We construct a preferred super-solution u1u_{1} and a preferred sub-solution u2u_{2} such that: They both satisfy the generalized Nahm pole condition plus real symmetry breaking condition. And u1u2u_{1}\geq u_{2} pointwise.

Here is the definition of u1u_{1}:

u1=u0+y.u_{1}=u_{0}+y.

Here is the definition of u2u_{2}:

u2=u0+f(y),u_{2}=u_{0}+f(y),

where f(y)={0,yC;yClnyCClnC,yC,f(y)=\begin{cases}0,&y\leq C;\\ y-C\ln y-C-C\ln C,&y\geq C,\end{cases}

where CC is a large enough constant.

Clearly they all satisfy the desired boundary/assymptotic conditions. Moreover,

Δu1+e2u1|P(z)|2=Δu0+e2ye2u0|P(z)|2Δu0+e2u0|P(z)|2=0.\Delta u_{1}+e^{-2u_{1}}|P(z)|^{2}=\Delta u_{0}+e^{-2y}e^{-2u_{0}}|P(z)|^{2}\leq\Delta u_{0}+e^{-2u_{0}}|P(z)|^{2}=0.

So u1u_{1} is indeed a super-solution.

Δu2+e2u2|P(z)|2=Δu0+f′′(y)+e2f(y)e2u0|P(z)|2=(e2f(y)1)e2u0|P(z)|2+f′′(y).\Delta u_{2}+e^{-2u_{2}}|P(z)|^{2}=\Delta u_{0}+f^{\prime\prime}(y)+e^{-2f(y)}e^{-2u_{0}}|P(z)|^{2}=(e^{-2f(y)}-1)e^{-2u_{0}}|P(z)|^{2}+f^{\prime\prime}(y).

Note that although f′′(y)f^{\prime\prime}(y) doesn’t exist, we do have

f′′(y){0,yCCy2,y>Cf^{\prime\prime}(y)\geq\begin{cases}0,&y\leq C\\ \dfrac{C}{y^{2}},&y>C\end{cases}

in the weak sense.

So replace f′′(y)f^{\prime\prime}(y) by the above function on the right, we get 0 when yCy\leq C and we get

(e2f(y)1)e2u0|P(z)|2+f′′(y)(e2f(y)1)e2u0|P(z)|2+Cy2e2u0|P(z)|2+Cy20,(e^{-2f(y)}-1)e^{-2u_{0}}|P(z)|^{2}+f^{\prime\prime}(y)\geq(e^{-2f(y)}-1)e^{-2u_{0}}|P(z)|^{2}+\dfrac{C}{y^{2}}\geq-e^{-2u_{0}}|P(z)|^{2}+\dfrac{C}{y^{2}}\geq 0,

when y>Cy>C.

Thus u2u_{2} is indeed a sub-solution.

Perron’s argument

In general, if we have a collection of functions u𝒰u\in\mathcal{U} that are bounded above point-wise, then we define sup𝒰\sup\mathcal{U} to be another function whose value at each point pp is:

sup{u(p)|u𝒰}.\sup\{u(p)~{}|~{}u\in\mathcal{U}\}.

For each point pZp\in Z, let

u3(p)=sup{u(p)|uis a sub-solution withuu2on the entireX}.u_{3}(p)=\sup\{u(p)~{}|~{}u~{}\text{is a sub-solution with}~{}u\leq u_{2}~{}\text{on the entire}~{}X\}.

We still use 𝒰\mathcal{U} to denote the set {u(p)|uis a sub-solution withuu2on the entireX}\{u(p)~{}|~{}u~{}\text{is a sub-solution with}~{}u\leq u_{2}~{}\text{on the entire}~{}X\}.

Then clearly u3u_{3} is a function that satisfies:

u1u3u2.u_{1}\leq u_{3}\leq u_{2}.

We prove that u3u_{3} is an actual solution. We need two lemmas first:

Lemma 2.3.

If we have two sub-solutions v1,v2v_{1},v_{2}, then max{v1,v2}\max\{v_{1},v_{2}\} is still a sub-solution.

Proof.

Clear max{v1,v2}\max\{v_{1},v_{2}\} is still continuous. Moreover,

Δ(max{v1,v2})Δv1,Δ(max{v1,v2})Δv2\Delta(\max\{v_{1},v_{2}\})\geq\Delta v_{1},~{}~{}~{}\Delta(\max\{v_{1},v_{2}\})\geq\Delta v_{2}

in the weak sense. We write it as

Δ(max{v1,v2})max{Δv1,Δv2}.\Delta(\max\{v_{1},v_{2}\})\geq\max\{\Delta v_{1},\Delta v_{2}\}.

So

Δ(max{v1,v2})+e2max{v1,v2}|P(z)|2max{Δv1,Δv2}+e2max{v1,v2}|P(z)|2\Delta(\max\{v_{1},v_{2}\})+e^{-2\max\{v_{1},v_{2}\}}|P(z)|^{2}\geq\max\{\Delta v_{1},\Delta v_{2}\}+e^{-2\max\{v_{1},v_{2}\}}|P(z)|^{2}
min{Δv1+e2v1|P(z)|2,Δv2+e2v2|P(z)|2}0.\geq\min\{\Delta v_{1}+e^{-2v_{1}}|P(z)|^{2},\Delta v_{2}+e^{-2v_{2}}|P(z)|^{2}\}\geq 0.

So {v1,v2}\{v_{1},v_{2}\} is also a sub-solution.

Lemma 2.4.

Suppose BB is a ball whose closure is a compact subset of XX. Suppose v𝒰v\in\mathcal{U}. Then there is a unique solution v0v_{0} in BB that is continuous up to the boundary, such that v0vv_{0}\geq v in BB and v0=vv_{0}=v on B\partial B. We call it a “lifting to an solution” of vv in BB. Moreover vv2v\leq v_{2}.

Proof.

We show the uniqueness first. Suppose two actual solutions v0,v0v_{0},v_{0}^{\prime} in BB both equal vv on B\partial B. Then

Δ(v0v0)+(e2v0e2v0)|P(z)|2=0.\Delta(v_{0}-v_{0}^{\prime})+(e^{-2v_{0}}-e^{-2v_{0}^{\prime}})|P(z)|^{2}=0.

Since v0v0v_{0}-v_{0}^{\prime} is continuous up to boundary in BB. If they are not the same in BB, without loss of generality, we may assume that pp is a positive maximal point of v0v0v_{0}-v_{0}^{\prime} inside of BB with the property

Δ(v0v0)(p)<0.\Delta(v_{0}-v_{0}^{\prime})(p)<0.

Since we have

(e2v0e2v0)|P(z)|20(e^{-2v_{0}}-e^{-2v_{0}^{\prime}})|P(z)|^{2}\leq 0

at pp, it contradicts with the fact that

Δ(v0v0)+(e2v0e2v0)|P(z)|2=0.\Delta(v_{0}-v_{0}^{\prime})+(e^{-2v_{0}}-e^{-2v_{0}^{\prime}})|P(z)|^{2}=0.

To prove the existence, choose a large enough positive constant CC. Using a standard Dirichlet argument, we may construct a function v1v_{1} such that

Δv1Cv1=e2v|P(z)|2Cvin B andv1=von B.\Delta v_{1}-Cv_{1}=-e^{-2v}|P(z)|^{2}-Cv~{}~{}\text{in $B$ and}~{}~{}v_{1}=v~{}~{}\text{on $\partial B$}.

Using the fact that Δv1Cv1ΔvCv\Delta v_{1}-Cv_{1}\leq\Delta v-Cv and maximal principle, we know v1vv_{1}\geq v. Then successively construct a sequence {vn}\{v_{n}\} that have the same boundary values on B\partial B and

Δvn+1Cvn+1=e2vn|P(z)|2Cvn\Delta v_{n+1}-Cv_{n+1}=-e^{-2v_{n}}|P(z)|^{2}-Cv_{n}

in BB. Since BB and P(z)P(z) are fixed and since vv is bounded below by u1u_{1} from the construction, we may assume CC is large enough such that, if vnv_{n} is also bounded from the below by u1u_{1} and if vn+1vnv_{n+1}\geq v_{n}, then

e2vn+1|P(z)|2+Cvn+1e2vn|P(z)|2+Cvn.e^{-2v_{n+1}}|P(z)|^{2}+Cv_{n+1}\geq e^{-2v_{n}}|P(z)|^{2}+Cv_{n}.

Thus

vn+1vnv_{n+1}\geq v_{n}

implies

Δvn+2Cvn+2Δvn+1Cvn+1,\Delta v_{n+2}-Cv_{n+2}\leq\Delta v_{n+1}-Cv_{n+1},

which further implies

vn+2vn+1.v_{n+2}\geq v_{n+1}.

So inductively we get vnv_{n} is an increasing sequence and Δvn\Delta v_{n} is a decreasing sequence.

On the other hand, we have vu2v\leq u_{2}. So

e2u2|P(z)|2+Cu2e2v|P(z)|2+Cv.e^{-2u_{2}}|P(z)|^{2}+Cu_{2}\geq e^{-2v}|P(z)|^{2}+Cv.
Δu2Cu2e2u2Cu2e2v|P(z)|2Cv=Δv1Cv1.\Delta u_{2}-Cu_{2}\leq-e^{-2u_{2}}-Cu_{2}\leq-e^{-2v}|P(z)|^{2}-Cv=\Delta v_{1}-Cv_{1}.

And we know that v1u2v_{1}\leq u_{2} on the boundary. A maximal principle implies that v1u2v_{1}\leq u_{2} in BB as well. And inductively we get vnu2v_{n}\leq u_{2} for all nn. In particular, the increasing sequence vnv_{n} has an upper bound u2u_{2}. So it converges uniformly to a continuous function v0v_{0}.

Clearly vv0u2v\leq v_{0}\leq u_{2}.

Finally, suppose GG is the Green’s function of Δ\Delta for BB centered at any point pBp\in B and PP is the Poisson’s kernel on B\partial B evaluated at pp. Since all vnv_{n} are bounded uniformly in B¯\bar{B}, using dominated convergent theorem,

v0(p)=limn+vn+1(p)=(limn+BG(Cvn+1e2vn|P(z)|2Cvn))+BPv)v_{0}(p)=\lim\limits_{n\rightarrow+\infty}v_{n+1}(p)=(\lim\limits_{n\rightarrow+\infty}\int_{B}G(Cv_{n+1}-e^{-2v_{n}}|P(z)|^{2}-Cv_{n}))+\int_{\partial B}Pv)
=BGe2v0|P(z)|2+BPv0.=-\int_{B}Ge^{-2v_{0}}|P(z)|^{2}+\int_{\partial B}Pv_{0}.

This implies that

Δv0+e2v0|P(z)|2=0\Delta v_{0}+e^{-2v_{0}}|P(z)|^{2}=0

in BB. So v0v_{0} is what we want.

Note that in the above proof, suppose v𝒰v\in\mathcal{U}. Then its lifting to a solution in BB (while keeping the outside part of BB unchanged) is still in 𝒰\mathcal{U}.

Given the above two lemmas, we prove that u3u_{3} is an actual sotluion:

Proof.

Consider a ball BB whose closure is a compact subset of XX. Since elements in 𝒰\mathcal{U} are continuous and bounded above uniformly by u2u_{2} in the closure of BB, we may find an increasing sequence vnv_{n} in 𝒰\mathcal{U} that converges to u3u_{3} on the closure of BB. Note that this implies that u3u_{3} is also continuous in BB.

For each element in 𝒰\mathcal{U}, replacing it with its “lifting to a solution” in BB makes it larger without violating being in 𝒰\mathcal{U}. Without affecting the argument, we may assume each vnv_{n} that we chose equal to its “lifting to a solution” in BB. In particular, each vnv_{n} is an actual solution in BB.

Finally, suppose GG is the Green’s function of Δ\Delta for BB centered at any point pBp\in B and PP is the Poisson’s kernel on B\partial B evaluated at pp. Since all vnv_{n} are bounded uniformly in B¯\bar{B} and convergent uniformly to u3u_{3} in B¯\bar{B}, using dominated convergent theorem,

u3(p)=limn+vn(p)=limn+(BG(e2vn|P(z)|2)+BPvn)u_{3}(p)=\lim\limits_{n\rightarrow+\infty}v_{n}(p)=\lim\limits_{n\rightarrow+\infty}(\int_{B}G(-e^{-2v_{n}}|P(z)|^{2})+\int_{\partial B}Pv_{n})
=BGe2u3|P(z)|2+BPu3.=-\int_{B}Ge^{-2u_{3}}|P(z)|^{2}+\int_{\partial B}Pu_{3}.

This implies that

Δv3+e2v3|P(z)|2=0\Delta v_{3}+e^{-2v_{3}}|P(z)|^{2}=0

in BB. Since BB is arbitrary, u3u_{3} is a solution on the entire XX. ∎

The real symmetry breaking condition

Let u1,u2,u3u_{1},u_{2},u_{3} have the same meaning as before. Remember that u1u3u2u_{1}\leq u_{3}\leq u_{2}. We verify that u3u_{3} satisfies the real symmetry breaking condition as y+y\rightarrow+\infty. We always assume ϵ\epsilon is a small enough constant, and y+y\rightarrow+\infty (which means, all the inequalities only work when yy is large enough).

It suffices to verify that:

1u3,2u3,eu3|P(z)|=O(yϵ),yu3=1+O(yϵ).\partial_{1}u_{3},\partial_{2}u_{3},e^{-u_{3}}|P(z)|=O(y^{-\epsilon}),~{}~{}\partial_{y}u_{3}=1+O(y^{-\epsilon}).

Since u3u1=u0+yu_{3}\geq u_{1}=u_{0}+y,

eu3|P(z)|eyeu0|P(z)|=O(ey).e^{-u_{3}}|P(z)|\leq e^{-y}e^{-u_{0}}|P(z)|=O(e^{-y}).

Let v=u3yNlnRv=u_{3}-y-N\ln R. It only remains to verify that

|v|=O(yϵ).|\nabla v|=O(y^{-\epsilon}).

The equation for vv is:

Δv+NR2+e2u3|P(z)|2=0.\Delta v+NR^{-2}+e^{-2u_{3}}|P(z)|^{2}=0.

Then

e2u3|P(z)|2=O(e2y).e^{-2u_{3}}|P(z)|^{2}=O(e^{-2y}).

So

Δv=O(y2).\Delta v=O(y^{-2}).

Moreover, we have

|v|max{|u1yNlnR,u2yNlnR|}=O(lny).|v|\leq\max\{|u_{1}-y-N\ln R,u_{2}-y-N\ln R|\}=O(\ln y).

We need a lemma:

Lemma 2.5.

Choose a point (z,y)(z,y) first. Suppose ry4r\leq\dfrac{y}{4}. Let BrB_{r} be the ball of radius rr centered at (z,y)(z,y). Then

Br|v|2=O(r|lny|2).\int_{B_{r}}|\nabla v|^{2}=O(r|\ln y|^{2}).
Proof.

For each ball BrB_{r}, we may choose a cut-off function χ\chi that is supported in B2rB_{2r} (the ball of radius 2r2r with the same center) such that:

χ=1inBr,and|kχ|=O(rk),for any non-negative integer k.\chi=1~{}~{}\text{in}~{}~{}B_{r},~{}~{}~{}\text{and}~{}~{}|\nabla^{k}\chi|=O(r^{-k}),~{}~{}\text{for any non-negative integer $k$}.

Then

B2r|(χv)|2=B2rΔ(χv)(χv)B2r(|Δv||v|+2|(χv)||χ||v|+(|Δχ|+2|χ|2)|v|2)\int_{B_{2r}}|\nabla(\chi v)|^{2}=-\int_{B_{2r}}\Delta(\chi v)\cdot(\chi v)\leq\int_{B_{2r}}(|\Delta v||v|+2|\nabla(\chi v)||\nabla\chi||v|+(|\Delta\chi|+2|\nabla\chi|^{2})|v|^{2})
12B2r|(χv)|2+B2r(|Δv||v|+Cr2|v|2),\leq\dfrac{1}{2}\int_{B_{2r}}|\nabla(\chi v)|^{2}+\int_{B_{2r}}(|\Delta v||v|+\dfrac{C}{r^{2}}|v|^{2}),

where CC is a large enough constant.

Recall that

Δv=e2v|P(z)|2=O(y2),v=O(lny).\Delta v=e^{-2v}|P(z)|^{2}=O(y^{-2}),~{}~{}v=O(\ln y).

It implies

Br|v|2B2r|(χv)|22B2r(|Δv||v|+Cr2|v|2)=O(r3y2|lny|)+O(r(lny)2)=O(r(lny)2).\int_{B_{r}}|\nabla v|^{2}\leq\int_{B_{2r}}|\nabla(\chi v)|^{2}\leq 2\int_{B_{2r}}(|\Delta v||v|+\dfrac{C}{r^{2}}|v|^{2})=O(r^{3}y^{-2}|\ln y|)+O(r(\ln y)^{2})=O(r(\ln y)^{2}).

Now we prove that |v|=O(yϵ)|\nabla v|=O(y^{-\epsilon}).

Proof.

We assume ry8r\leq\dfrac{y}{8} and let GG be the Green’s function of the Laplacian centered at (z,y)(z,y). Let BrB_{r} be the ball of radius rr with the same center. Let χ\chi be the same cut-off function as in the previous lemma. Then

(v)(z,y)=B2rGΔ(χv),(\nabla v)(z,y)=\int_{B_{2r}}G\Delta(\chi\nabla v),

where (v)(z,y)(\nabla v)(z,y) is v\nabla v evaluated at the point (z,y)(z,y). Using integration by parts, one can verify that

|(v)(z,y)|B2rG|Δ(χv)|B2r(8(|G||(χ)|+|G||2χ|)|v|+|(Gχ)||Δv|).|(\nabla v)(z,y)|\leq\int_{B_{2r}}G|\Delta(\chi\nabla v)|\leq\int_{B_{2r}}(8(|\nabla G||(\nabla\chi)|+|G||\nabla^{2}\chi|)|\nabla v|+|\nabla(G\chi)||\Delta v|).

We have |G||(χ)|+|G||2χ|=O(r3)|\nabla G||(\nabla\chi)|+|G||\nabla^{2}\chi|=O(r^{-3}), so

B2r(8(|G||(χ)|+|G||2χ|)|v|)B2r(8(|G||(χ)|+|G||2χ|)2)12(B2r|v|2)12\int_{B_{2r}}(8(|\nabla G||(\nabla\chi)|+|G||\nabla^{2}\chi|)|\nabla v|)\leq\int_{B_{2r}}(8(|\nabla G||(\nabla\chi)|+|G||\nabla^{2}\chi|)^{2})^{\frac{1}{2}}(\int_{B_{2r}}|\nabla v|^{2})^{\frac{1}{2}}
=O(r32r12|lny|)=O(r1lny).=O(r^{-\frac{3}{2}}\cdot r^{\frac{1}{2}}|\ln y|)=O(r^{-1}\ln y).

Moreover,

B2r|(Gχ)|=O(r2).\int_{B_{2r}}|\nabla(G\chi)|=O(r^{2}).

So

B2r|(Gχ)||Δv|=O(r2y2).\int_{B_{2r}}|\nabla(G\chi)||\Delta v|=O(r^{2}y^{-2}).

Remember that we are working on the region such that y+y\rightarrow+\infty. We may choose r=yr=\sqrt{y}, then we get

|v|=O(r1|lny|)+O(r2y2)=O(y12|lny|)=O(yϵ),|\nabla v|=O(r^{-1}|\ln y|)+O(r^{2}y^{-2})=O(y^{-\frac{1}{2}}|\ln y|)=O(y^{-\epsilon}),

where 0<ϵ<120<\epsilon<\dfrac{1}{2}.

The generalized Nahm pole boundary condition

We verify the generalized Nahm pole boundary condition for u3u_{3}. In this section, we always assume ϵ\epsilon is a small enough constant and y0y\rightarrow 0 (which means, all inequalities only work when yy is small). Note that both u1u_{1} and u2u_{2} satisfy the generalized Nahm pole boundary condition.

Since u1u3u2u_{1}\leq u_{3}\leq u_{2}, cleary we have

u3=u0+O(yϵ).u_{3}=u_{0}+O(y^{\epsilon}).

So it suffices to check that

|(u3u0)|=O(y1+ϵ).|\nabla(u_{3}-u_{0})|=O(y^{-1+\epsilon}).

Let v=u3u0=O(y)v=u_{3}-u_{0}=O(y). Then

Δv+(e2v1)e2u0|P(z)|2=0.\Delta v+(e^{-2v}-1)e^{-2u_{0}}|P(z)|^{2}=0.

So

Δv=(e2v1)e2u0|P(z)|2=O(y1).\Delta v=-(e^{-2v}-1)e^{-2u_{0}}|P(z)|^{2}=O(y^{-1}).

The following lemma is similar with lemma 2.5, except that we are working in the region y0y\rightarrow 0 now and the definition of vv is also different.

Lemma 2.6.

Choose a point (z,y)(z,y) first. Suppose ry4r\leq\dfrac{y}{4}. Let BrB_{r} be the ball of radius rr centered at (z,y)(z,y). Then

Br|v|2=O(ry2).\int_{B_{r}}|\nabla v|^{2}=O(ry^{2}).
Proof.

We choose the same cut-off function supported in B2rB_{2r} and equal 11 in BrB_{r} as always. Then the same argument as in the last paragraph,

Br|v|2=O(B2r(|Δv||v|+1r2|v|2))=O(r3)+O(ry2)=O(ry2).\int_{B_{r}}|\nabla v|^{2}=O(\int_{B_{2r}}(|\Delta v||v|+\dfrac{1}{r^{2}}|v|^{2}))=O(r^{3})+O(ry^{2})=O(ry^{2}).

Now we prove that |v|=O(y1+ϵ)|\nabla v|=O(y^{-1+\epsilon}). This is also similar with the last subsection (the real symmetry breaking).

Proof.

We assume ry8r\leq\dfrac{y}{8} and let GG be the Green’s function of the Laplacian centered at (z,y)(z,y). Let BrB_{r} be the ball of radius rr with the same center. Let χ\chi be the same cut-off function as always. Then

(v)(z,y)=B2rGΔ(χv),(\nabla v)(z,y)=\int_{B_{2r}}G\Delta(\chi\nabla v),

where (v)(z,y)(\nabla v)(z,y) is v\nabla v evaluated at the point (z,y)(z,y).

We have

|G||(χ)|+|G||2χ|=O(r3),B2r|(Gχ)|=O(r2).|\nabla G||(\nabla\chi)|+|G||\nabla^{2}\chi|=O(r^{-3}),~{}~{}~{}\int_{B_{2r}}|\nabla(G\chi)|=O(r^{2}).

Using the same method as the real symmetry breaking case,

|(v)(z,y)|B2r(8(|G||(χ)|+|G||2χ|)2)12(B2r|v|2)12+B2r|(Gχ)||Δv||(\nabla v)(z,y)|\leq\int_{B_{2r}}(8(|\nabla G||(\nabla\chi)|+|G||\nabla^{2}\chi|)^{2})^{\frac{1}{2}}(\int_{B_{2r}}|\nabla v|^{2})^{\frac{1}{2}}+\int_{B_{2r}}|\nabla(G\chi)||\Delta v|
=O(r32r12y)+O(r2y1).=O(r^{-\frac{3}{2}}\cdot r^{\frac{1}{2}}y)+O(r^{2}y^{-1}).

We may choose r=y16r=\dfrac{y}{16}. Then we get

|v|=O(r32r12y)+O(r2y1)=O(1).|\nabla v|=O(r^{-\frac{3}{2}}\cdot r^{\frac{1}{2}}y)+O(r^{2}y^{-1})=O(1).

2.3 The uniqueness

Suppose u0,u1,u2,u3u_{0},u_{1},u_{2},u_{3} 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 uu is a solution of ()(*) such that,

|uu1|C(yϵ+y1ϵ),|u-u_{1}|\leq C(y^{\epsilon}+y^{1-\epsilon}),

where CC is any fixed large constant and ϵ(0,1)\epsilon\in(0,1) is another fixed real number. Then u=u3u=u_{3}.

The author conjectures under weaker conditions it is still unique.

Conjecture 2.8.

Suppose uu is a solution of ()(*) such that, for some ϵ>0\epsilon>0,

  • When y+y\rightarrow+\infty, eu|P(z)|=O(yϵ)e^{-u}|P(z)|=O(y^{-\epsilon}) and |(uy)|=O(yϵ)|\nabla(u-y)|=O(y^{-\epsilon}).

  • When y0y\rightarrow 0, u=u0+O(yϵ)u=u_{0}+O(y^{\epsilon}) and |(uu0)|=O(y1+ϵ)|\nabla(u-u_{0})|=O(y^{-1+\epsilon}).

Then u=u3u=u_{3}. (The inequalities do not necessarily uniform in zz in the conjecture.)

The remaining of this subsection proves proposition 2.7

Lemma 2.9.

Suppose vv is a smooth function on XX such that for some ϵ>0\epsilon>0 and C>0C>0, |v|C(yϵ+y1ϵ)|v|\leq C(y^{\epsilon}+y^{1-\epsilon}). Moreover, Δv0\Delta v\geq 0 on the entire XX. Then v0v\leq 0.

Proof.

Let f(y)=supzv(z,y)f(y)=\sup\limits_{z\in\mathbb{C}}v(z,y). Note that since vv is bounded on each fixed yy slice, f(y)f(y) is a well-defined function on y[0,+)y\in[0,+\infty). Moreover, f(y)C(yϵ+y1ϵ)f(y)\leq C(y^{\epsilon}+y^{1-\epsilon}). In particular, we have f(0)=0f(0)=0.

Note that since vv is continuous, f(y)f(y) is also a continuous function. We show that f(y)f(y) is a concave up function.

In fact, for any ϵ>0\epsilon>0 and y=y0>0y=y_{0}>0. Let z0z_{0} be a point such that f(y0)v(y0,z0)<ϵf(y_{0})-v(y_{0},z_{0})<\epsilon. Moreover, we may assume v(z,y0)ϵ|zz0|2v(z,y_{0})-\epsilon|z-z_{0}|^{2} has a strict maximum over zz\in\mathbb{C} at z=z0z=z_{0}. In particular, the 12+22\partial_{1}^{2}+\partial_{2}^{2} acting on v(z,y0)ϵ|zz0|2v(z,y_{0})-\epsilon|z-z_{0}|^{2} has a negative value at z=z0z=z_{0}.

Note that Δ(v(z,y0)ϵ|zz0|2)2ϵ\Delta(v(z,y_{0})-\epsilon|z-z_{0}|^{2})\geq-2\epsilon. Since Δ=12+22+y2\Delta=\partial_{1}^{2}+\partial_{2}^{2}+\partial_{y}^{2}. This indicates that

(y2v)(z0,y0)2ϵ.(\partial_{y}^{2}v)(z_{0},y_{0})\geq-2\epsilon.

Since y2v\partial_{y}^{2}v is smooth, we actually have

(y2v)>3ϵ(\partial_{y}^{2}v)>-3\epsilon

in a neighbourhood of (z0,y0)(z_{0},y_{0}). So using the mean value theorem,

lim infh0v(z0,y0h)+v(z0,y0+h)2v(z0,y0)h2=y2v(z0,ξ)>3ϵ,\liminf\limits_{h\rightarrow 0}\dfrac{v(z_{0},y_{0}-h)+v(z_{0},y_{0}+h)-2v(z_{0},y_{0})}{h^{2}}=\partial_{y}^{2}v(z_{0},\xi)>-3\epsilon,

where ξ\xi is a number that can be arbitrarily close to y0y_{0}.

So

lim infh0f(y0h)+f(y0+h)2f(y0)+2ϵh2>3ϵ.\liminf\limits_{h\rightarrow 0}\dfrac{f(y_{0}-h)+f(y_{0}+h)-2f(y_{0})+2\epsilon}{h^{2}}>-3\epsilon.

Letting ϵ0\epsilon\rightarrow 0, we get

lim infh0f(y0h)+f(y0+h)2f(y0)h2=0.\liminf\limits_{h\rightarrow 0}\dfrac{f(y_{0}-h)+f(y_{0}+h)-2f(y_{0})}{h^{2}}=0.

Since this is true at any point, we know that f(y)f(y) is a concave up function.

Fix any y0>0y_{0}>0. Then since f(y)f(y) is concave up, for any other y>0y>0,

f(y)y0+f(0)(yy0)f(y0)y.f(y)y_{0}+f(0)(y-y_{0})\geq f(y_{0})y.

Recall that f(0)=0f(0)=0, we have

f(y)f(y0)y0y.f(y)\geq\dfrac{f(y_{0})}{y_{0}}y.

But f(y)C(yϵ+y1ϵ)f(y)\leq C(y^{\epsilon}+y^{1-\epsilon}). We must have f(y0)0f(y_{0})\leq 0. This is true for any y0y_{0}. Thus f(y)0f(y)\leq 0 for any y>0y>0. And it implies that v0v\leq 0.

Corollary 2.10.

If v1v_{1} and v2v_{2} are two solutions of ()(*) on XX such that 0v2v1C(yϵ+y1ϵ)0\leq v_{2}-v_{1}\leq C(y^{\epsilon}+y^{1-\epsilon}) on the entire XX for some C>0C>0 and ϵ(0,1)\epsilon\in(0,1). Then v1=v2v_{1}=v_{2}.

Proof.

We have

Δ(v2v1)=(e2v1e2v2)|P(z)|20.\Delta(v_{2}-v_{1})=(e^{-2v_{1}}-e^{-2v_{2}})|P(z)|^{2}\geq 0.

So from the lemma 2.9, v2v10v_{2}-v_{1}\leq 0. Hence v1=v2v_{1}=v_{2}. ∎

Lemma 2.11.

Suppose CC is a large constant and ϵ(0,1)\epsilon\in(0,1). Then there is a second order differentiable function f(y)f(y) on y[0,+)y\in[0,+\infty) with the following properties:

  • f(0)=0f(0)=0.

  • f(y)yC(yϵ+y1ϵ).f(y)\leq y-C(y^{\epsilon}+y^{1-\epsilon}).

  • f′′(y)+Cy2(e2f(y)1)0.f^{\prime\prime}(y)+\dfrac{C}{y^{2}}(e^{-2f(y)}-1)\geq 0.

Proof.

In fact, we may choose a even larger CC^{\prime} and let

f(y)=yC(yϵ+y1ϵ).f(y)=y-C^{\prime}(y^{\epsilon}+y^{1-\epsilon}).

Then all three bullets are satisfied provided that CC^{\prime} is large enough. Here is the reason: The first two bullets are obvious. For the third bullet,

f′′(y)+Cy2(ef(y)1)=1y2(Cϵ(1ϵ)(yϵ+y1ϵ)+Ce2f(y)C).f^{\prime\prime}(y)+\dfrac{C}{y^{2}}(e^{-f(y)}-1)=\dfrac{1}{y^{2}}(C^{\prime}\epsilon(1-\epsilon)(y^{\epsilon}+y^{1-\epsilon})+Ce^{-2f(y)}-C).

when yy is small, yC(yϵ+y1ϵ)y\leq C^{\prime}(y^{\epsilon}+y^{1-\epsilon}). So e2f(y)C0e^{-2f(y)}-C\geq 0 and hence the above expression is non-negative.

Otherwise, since CC^{\prime} is large, when y>C(yϵ+y1ϵ)y>C^{\prime}(y^{\epsilon}+y^{1-\epsilon}), since yy has a lower bound, we may assume

Cϵ(1ϵ)(yϵ+y1ϵ)C0.C^{\prime}\epsilon(1-\epsilon)(y^{\epsilon}+y^{1-\epsilon})-C\geq 0.

So in any case, the third bullet is also true.

Now here is the proof of proposition 2.7.

Proof.

Suppose uu is a solution of ()(*) such that,

|uu1|C(yϵ+y1ϵ).|u-u_{1}|\leq C(y^{\epsilon}+y^{1-\epsilon}).

In the Perron’s argument, we may instead choose u1u_{1} to be u0+y+C(yϵ+y1ϵ)u_{0}+y+C(y^{\epsilon}+y^{1-\epsilon}) for a possibly larger constant CC and choose u2u_{2} to be u0+yC(yϵ+y1ϵ)u_{0}+y-C(y^{\epsilon}+y^{1-\epsilon}). Since CC is assumed to be large, by lemma 2.11, one can verify that they are indeed super/sub solutions.

Using these substituted u1u_{1} and u2u_{2}, we may run Perron’s argument again and get a solution, still call it u3u_{3}. But recall that

u3=sup{u|uis a sub-solution withuu2on the entireX}.u_{3}=\sup\{u~{}|~{}u~{}\text{is a sub-solution with}~{}u\leq u_{2}~{}\text{on the entire}~{}X\}.

And uu is a solution with uu2u\leq u_{2}. So we have u3uu_{3}\geq u.

On the other hand, we may modify the Perron’s argument to define

u3=inf{u(p)|uis a super-solution withuu1on the entireX}.u_{3}^{\prime}=\inf\{u(p)~{}|~{}u~{}\text{is a super-solution with}~{}u\geq u_{1}~{}\text{on the entire}~{}X\}.

By the same reason, u3u_{3}^{\prime} is also an actual solution. And u3uu_{3}^{\prime}\leq u. So in order to prove uu is unique, we only need to show that

u3=u3.u_{3}=u_{3}^{\prime}.

Since we know that u3u3u_{3}^{\prime}\leq u_{3}. And

Δ(u3u3)=(e2u3e2u3)|P(z)|20.\Delta(u_{3}-u_{3}^{\prime})=(e^{-2u_{3}}-e^{-2u_{3}^{\prime}})|P(z)|^{2}\leq 0.

Moreover,

u3u3|u1u2|2C(yϵ+y1ϵ).u_{3}-u_{3}^{\prime}\leq|u_{1}-u_{2}|\leq 2C(y^{\epsilon}+y^{1-\epsilon}).

Thus by lemma 2.9, we know that u3=u3u_{3}=u_{3}^{\prime}. ∎

A remark on poly-homogeneous expansion

Based on Section 6 of [10], u0u_{0} has a poly-homogeneous expansion on X^\hat{X}, where X^\hat{X} is a preferred way to compactify XX 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 u3u_{3}. In fact, u3u_{3} may have a poly-homogeneous expansion on each boundary away from the corners of X^\hat{X}, but they do not seem to compatible at the corner. Even in the simplest case: Suppose there is no knot singularity as y0y\rightarrow 0, but only Nahm pole sigular boundary condition with real symmetry condition. The solution is written explicitly as

u=ln(eyey2).u=\ln(\dfrac{e^{y}-e^{-y}}{2}).

This solution doesn’t seem to have a poly-homogeneous expansion at the corner of X^\hat{X} given by R+R\rightarrow+\infty while ψ0\psi\rightarrow 0 at the same time, where y=Rsinψy=R\sin\psi.

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 u3u_{3} constructed in the last section, for each triple (P(z),Q(z),R(z))(P(z),Q(z),R(z)), where P,Q,RP,Q,R are polynomials with degR<degQ\mathrm{deg}R<\mathrm{deg}Q, QQ and RR are coprime, P,QP,Q are monic, we construct an approximate solution (H,φ)(H_{*},\varphi) that corresponds to it. Then we improve the approximate solution near y0y\rightarrow 0. Finally, we use a continuity method to further improve it to get an actual solution.

3.1 The approximate solution

Suppose a triple (P,Q,R)(P,Q,R) is given. The approximate solution constructed in this section will be a pair (H,φ)(H_{*},\varphi) with

φ=(00P(z)0).\varphi=\begin{pmatrix}0&0\\ P(z)&0\end{pmatrix}.

We define HH_{*} separately in different regions. Recall that in general, we may write HH_{*} as

H=(h+h1|w|2h1w¯h1wh1).H_{*}=\begin{pmatrix}h+h^{-1}|w|^{2}&h^{-1}\bar{w}\\ h^{-1}w&h^{-1}\end{pmatrix}.

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 HH_{*} near each boundary and use any arbitrary smooth one to fill the inside.

Note that a preferred compactification of XX 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 ρ\rho is large (near type I boundary)

Recall that when ρ\rho is large, ρ2=|z|2+y2\rho^{2}=|z|^{2}+y^{2}. In this region,

H=(h+h1|w|2h1w¯h1wh1),H_{*}=\begin{pmatrix}h+h^{-1}|w|^{2}&h^{-1}\bar{w}\\ h^{-1}w&h^{-1}\end{pmatrix},

where h=eu3h=e^{u_{3}}, w=χ(|z|y)R(z)Q(z)w=\chi(\dfrac{|z|}{y})\cdot\dfrac{R(z)}{Q(z)}, and u3u_{3} is the function constructed in section 2 which satisfies

Δu3+e2u3|P(z)|2=0\Delta u_{3}+e^{-2u_{3}}|P(z)|^{2}=0

and generalized Nahm pole bounary condition plus real symmetry breaking condition.

Since ρ\rho is large, we may assume that χ(|z|y)=0\chi(\dfrac{|z|}{y})=0 at all roots of Q(z)Q(z) and ww well-defined. Moreover, V(H,φ)V(H,\varphi) can be represented by (see special case 1 in subsection 1.2):

V(H,φ)=12(EF¯FE),V(H,\varphi)=\dfrac{1}{2}\begin{pmatrix}E&\bar{F}\\ F&-E\end{pmatrix},

with

E=Δu3+e2u3(4|¯w|2+|yw|2+|P|2)=e2u3(4|¯w|2+|yw|2),E=\Delta u_{3}+e^{-2u_{3}}(4|\bar{\partial}w|^{2}+|\partial_{y}w|^{2}+|P|^{2})=e^{-2u_{3}}(4|\bar{\partial}w|^{2}+|\partial_{y}w|^{2}),
F=eu3(Δw2(yu3)(yw)8(¯w)(u3)).F=e^{-u_{3}}(\Delta w-2(\partial_{y}u_{3})(\partial_{y}w)-8(\bar{\partial}w)(\partial u_{3})).

Note that Δw,yw\Delta w,\partial_{y}w and ¯w\bar{\partial}w are only non-zero when 1<|z|y<21<\dfrac{|z|}{y}<2. So E=F=0E=F=0 except in the region 1<|z|y<21<\dfrac{|z|}{y}<2. In this region, since we have assumed ρ\rho is large, eu3=O(ey)=O(eρ),e^{-u_{3}}=O(e^{-y})=O(e^{-\rho}), |w|=O(ρdegQ+degR)|w|=O(\rho^{-\mathrm{deg}Q+\mathrm{deg}R}), |w|=O(ρdegQ+degR1)|\nabla w|=O(\rho^{-\mathrm{deg}Q+\mathrm{deg}R}-1), |u3|=O(1)|\nabla u_{3}|=O(1) and |Δw|=O(ρ2)|\Delta w|=O(\rho^{-2}). So in fact, the first bullet of the extended Bogomolny equtauions

|V(H,φ)|=O(eρ),|V(H,\varphi)|=O(e^{-\rho}),

as ρ+\rho\rightarrow+\infty, uniformly in |z|y\dfrac{|z|}{y}.

When rr is small (near type III boundary)

We construction HH_{*} on a region such that yy is small and |z||z| is bounded above. This region contains all the points with small rr.

We have to work in a different basis. (That is to say, a different choice of s1,s2s_{1},s_{2} in the definition of (H,φ)(H,\varphi), see subsection 1.2.)

Since QQ and RR are coprime, we assume QS+TR=1QS+TR=1, where S,TS,T are also polynomials. Consider the holomorphic SL(2,)SL(2,\mathbb{C}) gauge transformation u=(QTRS)u=\begin{pmatrix}Q&T\\ -R&S\end{pmatrix}. (Recall, it is not an actual gauge transformation on the configuration, see subsection 1.2.) It sends φ\varphi to

u1φu=(STRQ)(00P0)(QTRS)=(PQTPT2PQ2PQT).u^{-1}\varphi u=\begin{pmatrix}S&-T\\ R&Q\end{pmatrix}\begin{pmatrix}0&0\\ P&0\end{pmatrix}\begin{pmatrix}Q&T\\ -R&S\end{pmatrix}=\begin{pmatrix}-PQT&-PT^{2}\\ PQ^{2}&PQT\end{pmatrix}.

In this basis, we choose

uHu=(eu300eu3),u^{*}Hu=\begin{pmatrix}e^{u_{3}^{\prime}}&0\\ 0&e^{-u_{3}^{\prime}}\end{pmatrix},

where u3u_{3}^{\prime} is the version of u3u_{3} but using PQ2PQ^{2} instead of PP. That is to say, u3u_{3}^{\prime} satisfies the following equation:

Δu3+e2u3|P(z)Q(z)2|2=0.\Delta u_{3}^{\prime}+e^{-2u_{3}^{\prime}}|P(z)Q(z)^{2}|^{2}=0.

Using the new pair (Hu,φu)=(uHu,u1φu)(H_{u},\varphi_{u})=(u^{*}Hu,u^{-1}\varphi u), its preferred configuration ΨHu,φu\Psi_{H_{u},\varphi_{u}} is SU(2)SU(2) gauge equivalent to ΨH,φ\Psi_{H,\varphi}. (Recall the definition of ΨH,φ\Psi_{H,\varphi} is in subsection 1.2.) This is in the special case 2 there. So the first bullet of the extended Bogomolny equations is

V(Hu,φu)=12(EF¯FE)V(H_{u},\varphi_{u})=\dfrac{1}{2}\begin{pmatrix}E&\bar{F}\\ F&-E\end{pmatrix}

with

E=Δu3+e2u3|PQ2|2+e2u3|PT2|2,F=2eu3(PT2P¯Q¯T¯)+eu3|PQ|2(TQ¯+T¯Q).E=\Delta u_{3}^{\prime}+e^{-2u_{3}^{\prime}}|PQ^{2}|^{2}+e^{2u_{3}^{\prime}}|PT^{2}|^{2},~{}~{}F=2e^{u_{3}^{\prime}}(PT^{2}\bar{P}\bar{Q}\bar{T})+e^{-u_{3}^{\prime}}|PQ|^{2}(T\bar{Q}+\bar{T}Q).

When |z||z| is bounded, all the polynomials P,Q,R,T,SP,Q,R,T,S are bounded. Note that when yy is small and |z||z| is bounded, eu3=O(1)e^{u_{3}^{\prime}}=O(1) and eu3|PQ|2=O(1y)e^{-u_{3}^{\prime}}|PQ|^{2}=O(\dfrac{1}{y}). Thus

E=e2u3|PT2|2=O(1),F=O(1y),E=e^{2u_{3}^{\prime}}|PT^{2}|^{2}=O(1),~{}~{}F=O(\dfrac{1}{y}),

uniformly in zz when |z||z| is bounded.

Going back to the original basis, HH is written as

H=(g1)(eu300eu3)g1=(S¯R¯T¯Q¯)(eu300eu3)(STRQ)H_{*}=(g^{-1})^{*}\begin{pmatrix}e^{u_{3}^{\prime}}&0\\ 0&e^{-u_{3}^{\prime}}\end{pmatrix}g^{-1}=\begin{pmatrix}\bar{S}&\bar{R}\\ -\bar{T}&\bar{Q}\end{pmatrix}\begin{pmatrix}e^{u_{3}^{\prime}}&0\\ 0&e^{-u_{3}^{\prime}}\end{pmatrix}\begin{pmatrix}S&-T\\ R&Q\end{pmatrix}
=(|S|2eu3+eu3|R|2TS¯eu3+R¯Qeu3T¯Seu3+RQ¯eu3|T|2eu3+|Q|2eu3).=\begin{pmatrix}|S|^{2}e^{u_{3}^{\prime}}+e^{-u_{3}^{\prime}}|R|^{2}&-T\bar{S}e^{u_{3}^{\prime}}+\bar{R}Qe^{-u_{3}^{\prime}}\\ -\bar{T}Se^{u_{3}^{\prime}}+R\bar{Q}e^{-u_{3}^{\prime}}&|T|^{2}e^{u_{3}^{\prime}}+|Q|^{2}e^{-u_{3}^{\prime}}\end{pmatrix}.

We still have |V(H,φ)|=O(1y)|V(H_{*},\varphi)|=O(\dfrac{1}{y}) since its norm doesn’t change under SU(2)SU(2) gauge transformations. This is what we want.

When ψ\psi is small (near type II boundary)

We assume ψ\psi is small. When |z||z| is large at the same time, recall that HH_{*} is already defined by h=eu3h=e^{u_{3}}, w=R(z)Q(z)w=\dfrac{R(z)}{Q(z)}. We call it

H1=(eu3+eu3|R(z)|2|Q(z)|2eu3R¯(z)Q¯(z)1eu3R(z)Q(z)1eu3).H_{1}=\begin{pmatrix}e^{u_{3}}+e^{-u_{3}}|R(z)|^{2}|Q(z)|^{-2}&e^{-u_{3}}\bar{R}(z)\bar{Q}(z)^{-1}\\ e^{-u_{3}}R(z)Q(z)^{-1}&e^{-u_{3}}\end{pmatrix}.

On the other hand, when |z||z| is small at the same time, HH_{*} is also defined to be

H2=(|S|2eu3+eu3|R|2TS¯eu3+R¯Qeu3T¯Seu3+RQ¯eu3|T|2eu3+|Q|2eu3)=(eu~+eu~|w~|2eu~w~¯eu~w~eu~).H_{2}=\begin{pmatrix}|S|^{2}e^{u_{3}^{\prime}}+e^{-u_{3}^{\prime}}|R|^{2}&-T\bar{S}e^{u_{3}^{\prime}}+\bar{R}Qe^{-u_{3}^{\prime}}\\ -\bar{T}Se^{u_{3}^{\prime}}+R\bar{Q}e^{-u_{3}^{\prime}}&|T|^{2}e^{u_{3}^{\prime}}+|Q|^{2}e^{-u_{3}^{\prime}}\end{pmatrix}=\begin{pmatrix}e^{\tilde{u}}+e^{-\tilde{u}}|\tilde{w}|^{2}&e^{-\tilde{u}}\bar{\tilde{w}}\\ e^{-\tilde{u}}\tilde{w}&e^{-\tilde{u}}\end{pmatrix}.

When |z||z| is neither too large nor too small and ψ\psi (or yy) is small, we have

  • |S|,|T||S|,|T| are all bounded.

  • eu3=O(y)e^{u_{3}^{\prime}}=O(y) (or O(ψ)O(\psi)) and eu3=O(y)e^{u_{3}}=O(y) (or O(ψ)O(\psi)).

  • As y0y\rightarrow 0, and when |z||z| is not small (to stay away from zeros of PP and QQ), u3=lnyln|P|+O(y)u_{3}=-\ln y-\ln|P|+O(y), while u3=lnyln(|PQ2|)+O(y)u_{3}^{\prime}=-\ln y-\ln(|PQ^{2}|)+O(y). So

    u3u3=2ln|Q|+O(y).u_{3}^{\prime}-u_{3}=-2\ln|Q|+O(y).

All these properties imply that in this region, as y0y\rightarrow 0

u~=u3+O(y),w~=R(z)Q(z)+O(y).\tilde{u}=u_{3}+O(y),~{}~{}\tilde{w}=\dfrac{R(z)}{Q(z)}+O(y).

So using a smooth cut-off function in zz to connect u3u_{3} and u~\tilde{u}, we get an HH_{*} defined on the entire region where ψ\psi is small such that H=H1H_{*}=H_{1} when with |z||z| is large and H=H2H_{*}=H_{2} when |z||z| is small. The error in the middle is

V(H,φ)V(H1,φ)=O(y1),V(H,φ)V(H2,φ)=O(y1).V(H_{*},\varphi)-V(H_{1},\varphi)=O(y^{-1}),~{}~{}~{}V(H_{*},\varphi)-V(H_{2},\varphi)=O(y^{-1}).

when |z||z| is neither too small nor too big.

So V(H,φ)=O(y1)V(H_{*},\varphi)=O(y^{-1}) uniformaly in zz and V(H1,φ)=0V(H_{1},\varphi)=0 when |z||z| is large.

To sum up

We have defined an HH_{*} near each boundary/corner with

φ=(00P(z)0).\varphi=\begin{pmatrix}0&0\\ P(z)&0\end{pmatrix}.

We can define HH_{*} in the middle area smoothly and arbitrarily. The properties of the pair (H,φ)(H_{*},\varphi) are:

  • HH_{*} is smooth on XX.

  • The configuration that corresponds to it satisfies the generalized Nahm pole boundary condition as y0y\rightarrow 0 (but of the version of PQ2PQ^{2}, not PP) and real symmetry breaking condition as y+y\rightarrow+\infty.

  • When y0y\rightarrow 0, V(H,φ)=O(1y)V(H_{*},\varphi)=O(\dfrac{1}{y}) uniformly in zz.

  • When R+R\rightarrow+\infty, V(H,φ)=O(eR)V(H_{*},\varphi)=O(e^{-R}) uniformly in t|z|\dfrac{t}{|z|}.

  • When |z||z| is large but yy is small, or when yy is large but |z||z| is small, V(H,φ)=0V(H_{*},\varphi)=0.

An estimate on ΨH,φ\Psi_{H_{*},\varphi}

For later analysis, we will need an estimate on the point-wise norm |ΨH,φ||\Psi_{H_{*},\varphi}|. Here the norm is the sum of the norms on all its six components A1,A2,Ay,Φ1,Φ2,Φ3A_{1},A_{2},A_{y},\Phi_{1},\Phi_{2},\Phi_{3}.

Recall that if (H,φ)(H_{*},\varphi) is written in the format

H=(h+h1|w|2h1w¯h1wh1),φ=(00P(z)0),H=\begin{pmatrix}h+h^{-1}|w|^{2}&h^{-1}\bar{w}\\ h^{-1}w&h^{-1}\end{pmatrix},~{}~{}\varphi=\begin{pmatrix}0&0\\ P(z)&0\end{pmatrix},

then ΨH,φ\Psi_{H_{*},\varphi} is written as

{Φ=Φ1iΦ2=(00h1P0),Φ3=i2h(yhyw¯ywyh),Ay=12h(0yw¯yw0),A1=12h(i2h2w¯2¯wi2h),A2=i2h(1h2w¯2¯w1h).\left\{\begin{array}[]{lr}\Phi=\Phi_{1}-i\Phi_{2}=\begin{pmatrix}0&0\\ h^{-1}P&0\end{pmatrix},&\\ \Phi_{3}=\dfrac{i}{2h}\begin{pmatrix}-\partial_{y}h&-\partial_{y}\bar{w}\\ -\partial_{y}w&\partial_{y}h\end{pmatrix},&\\ A_{y}=\dfrac{1}{2h}\begin{pmatrix}0&\partial_{y}\bar{w}\\ -\partial_{y}w&0\end{pmatrix},&\\ A_{1}=\dfrac{1}{2h}\begin{pmatrix}-i\partial_{2}h&2\partial\bar{w}\\ -2\bar{\partial}w&i\partial_{2}h\end{pmatrix},&\\ A_{2}=\dfrac{i}{2h}\begin{pmatrix}\partial_{1}h&2\partial\bar{w}\\ 2\bar{\partial}w&-\partial_{1}h\end{pmatrix}.\end{array}\right.

Examining the construction of HH, it is clear that when y1y\geq 1, all these terms are point-wise bounded above by a constant which doesn’t depend on yy or zz.

When y0y\rightarrow 0, one also examines that |ΨH,φ|=O(1y)|\Psi_{H_{*},\varphi}|=O(\dfrac{1}{y}) uniformly in zz.

So overall, there exists a constant C>0C>0 such that

|ΨH,φ|C(1y+1)|\Psi_{H_{*},\varphi}|\leq C(\dfrac{1}{y}+1)

on the entire XX.

A remark on geometrical meanings of (P,Q,R)(P,Q,R) and boundary conditions

We briefly explain the geometrical meanings of (P,Q,R)(P,Q,R). 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 PQ2PQ^{2} are the “knotted points” for the generalized Nahm pole boundary condition with degrees. And zeros of PP are the zeros and vanishing orders of φ\varphi. Moreover, under a certain SL(2,)SL(2,\mathbb{C}) gauge, there is a “small section” described by (R(z)Q(z))\begin{pmatrix}R(z)\\ Q(z)\end{pmatrix}. Readers may see [3] for more details. These data are irrelevant with the SU(2)SU(2) gauge transformations. So in particular, different choices of (P,Q,R)(P,Q,R) do not give SU(2)SU(2) equivalent solutions.

3.2 Improve the approximate solution when yy is small

Suppose Ψ=ΨH,φ\Psi_{*}=\Psi_{H_{*},\varphi} is the approximate solution in the last subsection. We modify Ψ{\Psi}_{*} when yy is small and |z||z| is bounded to make V(Ψ)=O(yn)V(\Psi_{*})=O(y^{n}) for any positive integer nn as y0y\rightarrow 0. Without leading to an ambiguity, we may still use Ψ{\Psi}_{*} to denote the modified version after this subsection in this paper.

The idea of the modification is to inductively construct a smooth section sks_{k} of isu(2)isu(2), k0k\geq 0. Suppose es0+s1++ske^{s_{0}+s_{1}+\cdots+s_{k}} acts on Ψ\Psi_{*} and gets Ψk\Psi_{k}. (Note that the summation is in the sense described in subsection 1.3 written in a sloppy way.) Then Ψk{\Psi}_{k} has the property

V(Ψk)=O(ykϵ),sk=O(yk+1ϵ),V(\Psi_{k})=O(y^{k-\epsilon}),~{}~{}~{}s_{k}=O(y^{k+1-\epsilon}),

when y0y\rightarrow 0, with ϵ\epsilon arbitrarily small. We assume all the derivatives of sks_{k} also have the correct vanishing order as y0y\rightarrow 0. We further assume each sks_{k} is supported in a region where yy is small and |z||z| is bounded. So it doesn’t ruin other good properties that Ψ\Psi_{*} has as listed in 3.1. In particular, we assume originally V(Ψ)V(\Psi_{*}) = 0 when y1My\leq\dfrac{1}{M} and |z|M|z|\geq M for some large constant M>0M>0. Then for each Ψk\Psi_{k}, we inductively assume V(Ψk)=0V(\Psi_{k})=0 when y1My\leq\dfrac{1}{M} and |z|(22(k+1))M|z|\geq(2-2^{-(k+1)})M.

Here is the construction:

Suppose Ψk1\Psi_{k-1} is constructed. (When k=0k=0, we assume Ψk1=Ψ\Psi_{k-1}=\Psi_{*}.) When yy is small, for any arbitrarily small ϵ>0\epsilon>0,

V(Ψk1)=O(yk1ϵ),V(\Psi_{k-1})=O(y^{k-1-\epsilon}),

and V(Ψk1)=0V(\Psi_{k-1})=0 when y1My\leq\dfrac{1}{M} and |z|(22k)M|z|\geq(2-2^{-k})M at the same time.

We use Lk1L_{k-1} to denote the operator LΨk1L_{\Psi_{k-1}} for simplicity. We haven’t studied the mapping properties of Lk1L_{k-1} yet, which is in fact awful. But luckily since we don’t care about the region when either yy or |z||z| is large, we may modify Lk1L_{k-1} to remain unchanged when both yy and |z||z| are small, but equals LΨk1L_{\Psi_{k-1}^{\prime}} when either yy or |z||z| is large, where Ψk1\Psi_{k-1}^{\prime} is a configuration with the same version of generalized Nahm pole boundary condition and without real symmetry breaking at y+y\rightarrow+\infty. When yy is small and |z||z| is bounded, as an elliptic operator with “edge” type (strictly speaking, y2LΨk1y^{2}L_{\Psi_{k-1}^{\prime}} is the elliptic operator of “edge” type locally), only non-leading terms are modified. This operator LΨk1L_{\Psi_{k-1}^{\prime}} is already well studied in [16] and [11]. And it equals Lk1L_{k-1} when yy is small and |z||z| is bounded (whose bound is larger than 2M2M, say when |z|100M|z|\leq 100M).

Suppose χ1(y)\chi_{1}(y) is a smooth cut-off function in yy which is 11 when yy is small and 0 when yy is large. Suppose χ2(z)\chi_{2}(z) is a smooth cut-off function in zz which is 11 when |z|100M|z|\leq 100M and 0 when |z||z| 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 s~k\tilde{s}_{k} such that when y0y\rightarrow 0, in this region,

LΨk1(s~k)=χ1(y)χ2(z)V(Ψk1),s~k=O(yk+1ϵ).L_{\Psi_{k-1}^{\prime}}(\tilde{s}_{k})=-\chi_{1}(y)\chi_{2}(z)V(\Psi_{k-1}),~{}~{}~{}\tilde{s}_{k}=O(y^{k+1-\epsilon}).

All the derivatives of s~k\tilde{s}_{k} also have the appropriate decay rates at y0y\rightarrow 0. Note that s~k\tilde{s}_{k} may behave bad outside the region mentioned above.

In fact, the operator LΨk1L_{\Psi_{k-1}^{\prime}} is locally invertible in the region yy is small and |z|100M|z|\leq 100M above the lower Fredhohm weight. To be more precise, if VV is supported in this region and VyδCiek,αV\in y^{\delta}C^{k,\alpha}_{\text{ie}} when δ>2\delta>-2, then there is an s~kyδ+2ϵCiek,α\tilde{s}_{k}\in y^{\delta+2-\epsilon}C^{k,\alpha}_{\text{ie}} such that LΨk1s~k=VL_{\Psi_{k-1}^{\prime}}\tilde{s}_{k}=V in this region. The definition of Ciek,αC^{k,\alpha}_{\text{ie}} can be found in [11], [3] or appendix C. The fact that the Fredholm weight is above 2-2 can be found in [16], [11] or [3]. (Strictly speaking, near a corner both ψ\psi and rr goes to 0, we should use ψδrδCiek,α\psi^{\delta}r^{\delta}C^{k,\alpha}_{\text{ie}}. However even there, ψr\psi r is equivalent with yy.) 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 XX, say, when R+R\rightarrow+\infty. This also means, s~k\tilde{s}_{k} is allowed to behave bad outside the region that we are interested in. Note that the small constant ϵ\epsilon exists because of the fact that s~k\tilde{s}_{k} may have additional terms of O(yδ+2lny)O(y^{\delta+2}\ln y) which make it not reside exactly in yδ+2Ciek,αy^{\delta+2}C^{k,\alpha}_{\text{ie}}.

We may take a third cut-off function χ3(z)\chi_{3}(z) in zz which is 11 when |z|(22k)M|z|\leq(2-2^{-k})M and 0 when |z|(22(k+1))M|z|\geq(2-2^{-(k+1)})M. Recall the fact that V(Ψk1)=0V(\Psi_{k-1})=0 when |z|(22k)M|z|\geq(2-2^{-k})M and yy is small. This implies that LΨk1(s~k)=0L_{\Psi_{k-1}^{\prime}}(\tilde{s}_{k})=0 in the same region.

Thus let sk=χ3(z)χ1(y)s~ks_{k}=\chi_{3}(z)\chi_{1}(y)\tilde{s}_{k}. We have sks_{k} is supported in the region |z|(22k+1)M|z|\leq(2-2^{k+1})M and yy is small. And in particular, when |z|(22k+1)M|z|\leq(2-2^{k+1})M and y1My\leq\dfrac{1}{M} (We assume χ1(y)=1\chi_{1}(y)=1 in this region),

Lk1(sk)=LΨk1(sk)=LΨk1(χ3(z))s~k.L_{k-1}(s_{k})=L_{\Psi_{k-1}^{\prime}}(s_{k})=L_{\Psi_{k-1}^{\prime}}(\chi_{3}(z))\tilde{s}_{k}.

So let Ψk\Psi_{k} be the configuration that we get after applying eeke^{e_{k}} on Ψk1\Psi_{k-1}. According to the first fact in subsection 1.3,

V(Ψk)=V(Ψk1)+Lk1(sk)+Q(sk).V(\Psi_{k})=V(\Psi_{k-1})+L_{k-1}(s_{k})+Q(s_{k}).

When |z|(22k)M|z|\leq(2-2^{k})M and y1My\leq\dfrac{1}{M},

V(Ψk)=V(Ψk1)+Lk1(sk)+Q(sk)=Q(sk)=O(ykϵ).V(\Psi_{k})=V(\Psi_{k-1})+L_{k-1}(s_{k})+Q(s_{k})=Q(s_{k})=O(y^{k-\epsilon}).

When (22k)M|z|(22k+1)M(2-2^{k})M\leq|z|\leq(2-2^{k+1})M and y1My\leq\dfrac{1}{M},

V(Ψk1)=Lk1(s~k)=0V(\Psi_{k-1})=L_{k-1}(\tilde{s}_{k})=0

and

Lk1(χ3(z)s~k)|2χ3||s~k|+|χ3||s~k|=O(ykϵ).L_{k-1}(\chi_{3}(z)\tilde{s}_{k})\lesssim|\nabla^{2}\chi_{3}||\tilde{s}_{k}|+|\nabla\chi_{3}||\nabla\tilde{s}_{k}|=O(y^{k-\epsilon}).
V(Ψk)=V(Ψk1)+Lk1(sk)+Q(sk)=Lk1(χ3(z)s~k)+Q(sk)=O(ykϵ).V(\Psi_{k})=V(\Psi_{k-1})+L_{k-1}(s_{k})+Q(s_{k})=L_{k-1}(\chi_{3}(z)\tilde{s}_{k})+Q(s_{k})=O(y^{k-\epsilon}).

Finally, when |z|(22k+1)M|z|\geq(2-2^{k+1})M and y1My\leq\dfrac{1}{M}, we have sk=0s_{k}=0 thus Ψk=Ψk1\Psi_{k}=\Psi_{k-1}. So the inductive construction is finished.

We take summation in a convergent way of {sk}\{s_{k}\} (and in the sense described in subsection 1.3) and get a new configuration, still denoted as Ψ\Psi_{*}, with the additional property besides what are listed in subsection 3.1:

V(Ψ)=O(yn),V(\Psi_{*})=O(y^{n}),

for any positive integer nn when y0y\rightarrow 0 uniformly in zz. Here “in a convergent way” means we may freely multiply a cut-off function in yy supported and equals to 11 near y=0y=0 (but becomes 0 quickly when yy gets away from 0) onto each sks_{k} (whose support depends on sks_{k}) to make the infinite sum of sks_{k} converge. In the remaining of this paper, we will use this new Ψ\Psi_{*} instead of the old one.

3.3 The continuity argument

Let Ψ{\Psi}_{*} be the approximate solution from the last subsection. Let s=V(Ψ)s_{*}=V(\Psi_{*}). Clearly sisu(2)s_{*}\in isu(2). Consider es2e^{\frac{s_{*}}{2}} acting on Ψ\Psi_{*} and call it Ψ0\Psi_{0}. Note that ss_{*} vanishes up to any finite order on any boundaries. (We’ll need this property to make sure that Ψ0\Psi_{0} still satisfies the same generalized Nahm pole condition and real symmetry condition as Ψ\Psi_{*}. ) Then es2e^{-\frac{s_{*}}{2}} acting on Ψ0\Psi_{0} gets back to Ψ\Psi_{*}. We have

V(Ψ0,s)=s,V(\Psi_{0},-s_{*})=s_{*},

where the notation V(Ψ0,s)V(\Psi_{0},-s_{*}) is defined in subsection 1.3.

On the other hand, by the first fact described in subsection 1.3

s=V(Ψ0)LΨ0(s)+Q(s).s_{*}=V(\Psi_{0})-L_{\Psi_{0}}(s_{*})+Q(-s_{*}).

Thus V(Ψ0)=0V(\Psi_{0})=0 whenever s=0s_{*}=0. And V(Ψ0)=O(yn)V(\Psi_{0})=O(y^{n}) for any nn when y0y\rightarrow 0 just like ss_{*}. And it decays exponentially when R0R\rightarrow 0 just like ss_{*}. That is to say, for some small constant ϵ>0\epsilon>0 ,V(Ψ0)=O(eϵR)V(\Psi_{0})=O(e^{-\epsilon R}) as R+R\rightarrow+\infty.

The following equation

V(Ψ0,s)+ts=0V(\Psi_{0},s)+ts=0

has an obvious solution s=ss=-s_{*} when t=1t=1. In general, suppose ss is a section of isu(2)isu(2). Let S[0,1]S\subseteq[0,1] be the set of tt values such that

V(Ψ0,s)+ts=0V(\Psi_{0},s)+ts=0

has a solution with s𝒳1/2ϵ,1/2ϵ,1+ϵks\in\mathcal{X}^{k}_{1/2-\epsilon,1/2-\epsilon,-1+\epsilon} for some small ϵ>0\epsilon>0 and any non-negative integer kk, where the definition of 𝒳1/2ϵ,1/2ϵ,1+ϵk\mathcal{X}^{k}_{1/2-\epsilon,1/2-\epsilon,-1+\epsilon} can be found in appendix C. Then SS is non-empty. (Because 1S1\in S.)

By way of looking ahead, we prove that S=[0,1]S=[0,1] by showing it is both relatively open and relatively closed. In particular, when t=0t=0, it gives an actual solution to the extended Bogomolny equations. In fact, we have the following five facts:

  • The initial s1=ss_{1}=-s_{*} is in 𝒳12,12,lk\mathcal{X}^{k}_{\frac{1}{2},\frac{1}{2},-l} for any non-negative integers k,lk,l. This is straightforward from the construction.

  • Suppose at some t=t>0t=t_{\diamond}>0, there is a solution st𝒳12,12,lks_{t_{\diamond}}\in\mathcal{X}^{k}_{\frac{1}{2},\frac{1}{2},-l} for all k,lk,l. Then for some small ϵ>0\epsilon>0 depending on tt_{\diamond}, there is also a solution in 𝒳12,12,lk\mathcal{X}^{k}_{\frac{1}{2},\frac{1}{2},-l} when tϵ<t<t+ϵt_{\diamond}-\epsilon<t<t_{\diamond}+\epsilon. This is proved in subsection 3.5.

  • For each 0<t010<t_{0}\leq 1, suppose when t0t1t_{0}\leq t\leq 1 there is a solution st𝒳12,12,lks_{t}\in\mathcal{X}^{k}_{\frac{1}{2},\frac{1}{2},-l}. Then there is a constant CC which only depends on k,l,t0k,l,t_{0} and Ψ0\Psi_{0} such that

    st𝒳12,12,lkC.||s_{t}||_{\mathcal{X}^{k}_{\frac{1}{2},\frac{1}{2},-l}}\leq C.

    Note that the constant CC doesn’t depend on sts_{t} or tt. This is proved in subsection 3.4.

  • Suppose t>0t>0. If sts_{t} is a solution in 𝒳12ϵ,12ϵ,lk\mathcal{X}^{k}_{\frac{1}{2}-\epsilon,\frac{1}{2}-\epsilon,-l} for all kk and ll and some small ϵ>0\epsilon>0. Then it is also in 𝒳12,12,lk\mathcal{X}^{k}_{\frac{1}{2},\frac{1}{2},-l} for all kk and ll. This is proved in subsection 3.5.

  • There is a constant CC which depends only on kk and Ψ0\Psi_{0}, such that for any 0t10\leq t\leq 1, if there is a solution st𝒳12,12,1ks_{t}\in\mathcal{X}^{k}_{\frac{1}{2},\frac{1}{2},-1}. Then

    st𝒳12,12,1kC.||s_{t}||_{\mathcal{X}^{k}_{\frac{1}{2},\frac{1}{2},-1}}\leq C.

    Note that CC doesn’t depend on sts_{t} or tt. This is proved in subsection 3.4.

Proof that S=[0,1]S=[0,1] using the above facts

Let SS^{\prime} be the subset of SS such that additionally st𝒳12,12,lks_{t}\in\mathcal{X}^{k}_{\frac{1}{2},\frac{1}{2},-l} for all k,lk,l. Then 1S1\in S^{\prime}.

The second bullet indicates that SS^{\prime} is relatively open in [0,1][0,1]. If there is a sequence {tn}\{t_{n}\} in SS^{\prime} converging to t0>0t_{0}>0. Then by the third bullet and by choosing a sub-sequence, we may assume stns_{t_{n}} converges to an element in 𝒳12ϵ,12ϵ,lk\mathcal{X}^{k}_{\frac{1}{2}-\epsilon,\frac{1}{2}-\epsilon,-l} for all k,lk,l. We don’t need to add an ϵ\epsilon on l-l since it is true for all ll. Then by the fourth bullet the limit st0s_{t_{0}} actually lies in 𝒳12,12,lk\mathcal{X}^{k}_{\frac{1}{2},\frac{1}{2},-l}. This limit solves the t0t_{0} version of the equation. So S(0,1]S^{\prime}\cap(0,1] is a relatively closed subset of (0,1](0,1]. Thus (0,1]S(0,1]\subset S^{\prime}. In particular, (0,1]S(0,1]\subset S.

Finally, choose a sequence tn0t_{n}\rightarrow 0. According the the fifth bullet, by choosing a sub-sequence, we may assume stns_{t_{n}} converges to some s0𝒳12ϵ,12ϵ,1+ϵks_{0}\in\mathcal{X}^{k}_{\frac{1}{2}-\epsilon,\frac{1}{2}-\epsilon,-1+\epsilon}. This limit s0s_{0} solves the t=0t=0 version of the equation. Thus 0S0\in S.

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 C0,αC^{0,\alpha} estimate may originate in Bando and Siu’s paper [1] and the local Ck,αC^{k,\alpha} 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 LL^{\infty} estimate

Suppose we have a solution

V(Ψ0,st)+tst=0.V(\Psi_{0},s_{t})+ts_{t}=0.

Then according to the second fact (WeitzenBock formula) described in subsection 1.3, we have

t|st|2=<V(Ψ0,st),st>=<V(Ψ0),st>12Δ(|st|2)+i=13|v(s)Dis|2.-t|s_{t}|^{2}=<V(\Psi_{0},s_{t}),s_{t}>=<V(\Psi_{0}),s_{t}>-\dfrac{1}{2}\Delta(|s_{t}|^{2})+\sum\limits_{i=1}^{3}|v(-s)D_{i}^{*}s|^{2}.
Lemma 3.1.

(Dimakis’ estimate in [3]) For each k1k\geq 1, there exists a constant C1C_{1} which only depends on V(Ψ)V(\Psi), kk and t0>0t_{0}>0, such that for each t(t0,1]t\in(t_{0},1], if st𝒳12,12,lks_{t}\in\mathcal{X}^{k}_{\frac{1}{2},\frac{1}{2},-l} for all k,lk,l. Then

sup|ρkst|C1.\sup|\rho^{k}s_{t}|\leq C_{1}.

Moreover, when k=1k=1, if we only assume t[0,1]t\in[0,1] and st𝒳12,12,1ks_{t}\in\mathcal{X}^{k}_{\frac{1}{2},\frac{1}{2},-1} for all kk, then there is a constant CC which doesn’t depend on tt such that

sup|ρst|C.\sup|\rho s_{t}|\leq C.

Note that the definition of ρ\rho can be found in appendix B.

Proof.

The constants C1C_{1} and CC may change from line to line. Keep in mind that C1C_{1} only depends on Ψ0,k\Psi_{0},k and t0t_{0}, while CC only depends on Ψ0\Psi_{0}.

Note that originally we have

|st|ρkCk(st),|s_{t}|\leq\rho^{-k}C_{k}(s_{t}),

where Ck(st)C_{k}(s_{t}) depends on sts_{t}.

Point-wise we have

Δ(|st|2)+2t|st|22|V(Ψ0)||st|-\Delta(|s_{t}|^{2})+2t|s_{t}|^{2}~{}\leq~{}2|V(\Psi_{0})|\cdot|s_{t}|

Consider the Green’s function for Δ+2t-\Delta+2t on XX. This Green’s function is

G2t,q(p)=e2t|pq|4π|pq|e2t|pq¯|4π|pq¯|,G_{2t,q}(p)=\dfrac{e^{-\sqrt{2t}|p-q|}}{4\pi|p-q|}-\dfrac{e^{-\sqrt{2t}|p-\bar{q}|}}{4\pi|p-\bar{q}|},

where q¯\bar{q} is the reflection of qq about y=0y=0 plane.

First we assume tt0>0t\geq t_{0}>0 and prove the first bullet. Choose a ball BB of radius ρ2\dfrac{\rho}{2} centered at qq. The maximal principle implies that

|st|2X2G2t,q|V(Ψ0)||st|=BX2G2t,q|V(Ψ0)||st|+X\B2G2t,q|V(Ψ0)||st||s_{t}|^{2}\leq\int_{X}2G_{2t,q}|V(\Psi_{0})||s_{t}|=\int_{B\cap X}2G_{2t,q}|V(\Psi_{0})||s_{t}|+\int_{X\backslash B}2G_{2t,q}|V(\Psi_{0})||s_{t}|
2k+1Ck(st)(BXρkG2t,q|V(Ψ0)|+X\BG2t,q|V(Ψ0)|).\leq 2^{k+1}C_{k}(s_{t})(\int_{B\cap X}{\rho}^{-k}G_{2t,q}|V(\Psi_{0})|+\int_{X\backslash B}G_{2t,q}|V(\Psi_{0})|).

Note that |V(Ψ0)||V(\Psi_{0})| has exponential decay as ρ+\rho\rightarrow+\infty and G2t,qG_{2t,q} has exponential decay with respect to the distance to qq outside of the ball BB. So

(BXρkG2t,q|V(Ψ0)|+X\BG2t,q|V(Ψ0)|)C1yk.(\int_{B\cap X}{\rho}^{-k}G_{2t,q}|V(\Psi_{0})|+\int_{X\backslash B}G_{2t,q}|V(\Psi_{0})|)\leq C_{1}y^{-k}.

This implies

|st|C1Ck(st)ρk.|s_{t}|\leq\sqrt{C_{1}C_{k}(s_{t})}{\rho}^{-k}.

Repeating this finitely many times we get,

|st|C1ρk|s_{t}|\leq C_{1}{\rho}^{-k}

Next we assume t[0,1]t\in[0,1] and prove the second bullet.

We simply have

Δ(|st|2)2|V(Ψ0)||st|.-\Delta(|s_{t}|^{2})~{}\leq~{}2|V(\Psi_{0})|\cdot|s_{t}|.

Originally we have

|st|C(st)ρ1C(st).|s_{t}|\leq C(s_{t})\rho^{-1}\leq C(s_{t}).

And the same maximal principle applies and we have

|st|2X2G0,q|V(Ψ0)||st|2C(st)X|G0,q||V(Ψ0)||s_{t}|^{2}\leq\int_{X}2G_{0,q}|V(\Psi_{0})||s_{t}|\leq 2C(s_{t})\int_{X}|G_{0,q}||V(\Psi_{0})|
=2C(st)(XB|G0,q||V(Ψ0)|+X\B|G0,q||V(Ψ0)|).=2C(s_{t})(\int_{X\cap B}|G_{0,q}||V(\Psi_{0})|+\int_{X\backslash B}|G_{0,q}||V(\Psi_{0})|).

Note that X\BG0,q|V(Ψ0)|\displaystyle\int_{X\backslash B}G_{0,q}|V(\Psi_{0})| is bounded above by Cρ2C\rho^{-2} since |V(Ψ0)||V(\Psi_{0})| has an exponential decay in ρ\rho when ρ+\rho\rightarrow+\infty.

On the other hand, note that

|G0,q(p)|=14π4y(p)y(q)|pq||pq¯|(|pq|+|pq¯|),|G_{0,q}(p)|=\dfrac{1}{4\pi}\dfrac{4y(p)y(q)}{|p-q||p-\bar{q}|(|p-q|+|p-\bar{q}|)},

where y(p),y(q)y(p),y(q) means the yy coordinates of pp and qq respectively. On X\BX\backslash B,

X\BG0,q|V(Ψ0)|CX\By(p)y(q)|pq|3|V(Ψ0)|𝑑pCρ2Xy|V(Ψ0)|Cρ2.\int_{X\backslash B}G_{0,q}|V(\Psi_{0})|\leq C\int_{X\backslash B}\dfrac{y(p)y(q)}{|p-q|^{3}}|V(\Psi_{0})|dp\leq C\rho^{-2}\int_{X}y|V(\Psi_{0})|\leq C\rho^{-2}.

So after all

|st|2CC(st)ρ2.|s_{t}|^{2}\leq CC(s_{t})\rho^{-2}.

Which means

|st|CC(st)ρ1.|s_{t}|\leq\sqrt{CC(s_{t})}\rho^{-1}.

Repeating this finitely many times we get (remember that CC may change from line to line)

|st|Cρ1.|s_{t}|\leq C\rho^{-1}.

We also need a point-wise estimate of |st||s_{t}| when y0y\rightarrow 0. We use y^\hat{y} to represent yy when yy is small and 11 when yy is not small.

Lemma 3.2.

(This is inspired by a similar Taubes’ estimate in [20])

There is a constant CC which only depends V(Ψ0)V(\Psi_{0}), such that,

|st|Cy12.|s_{t}|\leq Cy^{\frac{1}{2}}.
Proof.

Since by lemma 3.1 |st||s_{t}| is bounded above by CC on the entire XX. We only need to assume yy is small. Recall in the proof of lemma 3.1, the maximal principle that we have:

|st|2X2G0,q|V(Ψ0)||st|.|s_{t}|^{2}\leq\int_{X}2G_{0,q}|V(\Psi_{0})||s_{t}|.

Suppose initially we suppose

|st|min{C(st)y12,C},|s_{t}|\leq\min\{C(s_{t})y^{\frac{1}{2}},C\},

where C(st)C(s_{t}) depends on sts_{t}.

Let BB be the ball of radius y(q)2\dfrac{y(q)}{2} centered at qq. Then |G0,q|Cy2y(q)|G_{0,q}|\leq Cy^{-2}y(q) outside of this ball and

X\B2G0,q|V(Ψ0)||st|CC(st)y(q)X\By2|V(Ψ0)|.\int_{X\backslash B}2G_{0,q}|V(\Psi_{0})||s_{t}|\leq CC(s_{t})y(q)\int_{X\backslash B}y^{-2}|V(\Psi_{0})|.

Note that Xy2|V(Ψ0)|\displaystyle\int_{X}y^{-2}|V(\Psi_{0})| is bounded. So centered at the point qq, we have

X\B2G0,q|V(Ψ0)||st|CC(st)y(q).\int_{X\backslash B}2G_{0,q}|V(\Psi_{0})||s_{t}|\leq CC(s_{t})y(q).

On the other hand, within the ball BB, since |V(Ψ0)|=O(yn)|V(\Psi_{0})|=O(y^{n}) for any positive integer nn when yy is small, we have

B2G0,q|V(Ψ0)||st|CC(st)y.\int_{B}2G_{0,q}|V(\Psi_{0})||s_{t}|\leq CC(s_{t})y.

So after all

|st|2CC(st)y.|s_{t}|^{2}\leq CC(s_{t})y.

Repeating this finitely many times we get

|st|Cy.|s_{t}|\leq C\sqrt{y}.

The interior ρ1C0,α\rho^{-1}C^{0,\alpha} estimates for sts_{t} away from y=0y=0

(Bando and Siu’s estimates) 

We continue to assume sk𝒳12,12,lks_{k}\in\mathcal{X}^{k}_{\frac{1}{2},\frac{1}{2},-l} solves

V(Ψ0,st)+tst=0.V(\Psi_{0},s_{t})+ts_{t}=0.

We first work in a region away from the y=0y=0 boundary.

Lemma 3.3.

Suppose BrB_{r} is a ball of radius rr in XX which is far away from the y=0y=0 boundary. We assume r1r\leq 1. Let GG be the Green’s function of Δ\Delta centered at the center of the Ball (basically 4π4\pi times 11 over the distance to the center). Let

f(r):=BrG|st|2.f(r):=\int_{B_{r}}G|\nabla s_{t}|^{2}.

Then there exists a constant CC and α(0,1)\alpha\in(0,1) which only depends on Ψ0\Psi_{0}, such that

f(r)Cr2α.f(r)\leq Cr^{2\alpha}.
Proof.

Recall the fact that |st||s_{t}| is bounded above by a constant that only depends on V(Ψ0)V(\Psi_{0}). So the point-wise norm (largest eigenvalue) of the operator adstad_{-s_{t}} is also bounded uniformly on the entire XX. Note that adstad_{s_{t}} is a self-dual map. So all its eigen-values are real number. Thus

|v(s)|=|eads1ads||v(-s)|=|\sqrt{\dfrac{e^{ad_{-s}}-1}{ad_{-s}}}|

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 CC which only depends on V(Ψ0)V(\Psi_{0}), but may change from line by line, we have

Δ(|st|2)+1Ci=13|Dist|2+t|st|2|V(Ψ0)||st|.-\Delta(|s_{t}|^{2})+\dfrac{1}{C}\sum\limits_{i=1}^{3}|D_{i}^{*}s_{t}|^{2}+t|s_{t}|^{2}\leq|V(\Psi_{0})||s_{t}|.

From now on, let χr\chi_{r} be a standard smooth cut-off function which is 11 on BrB_{r} and 0 outside of B2rB_{2r}. We assume |kχr|Crk|\nabla^{k}\chi_{r}|\leq\dfrac{C}{r^{k}} uniformly.

Step 1 We first show that f(r)f(r) is uniformly bounded above. We have

Δ(|st|2)+1C|st|2|V(Ψ0)||st|C.-\Delta(|s_{t}|^{2})+\dfrac{1}{C}|\nabla s_{t}|^{2}\leq|V(\Psi_{0})||s_{t}|\leq C.

On a ball BrB_{r} of radius rr centered at pp,

BrG|st|2B2rχrG|st|2CB2rχrG(Δ(|st|2)+C)\int_{B_{r}}G|\nabla s_{t}|^{2}\leq\int_{B_{2r}}\chi_{r}G|\nabla s_{t}|^{2}\leq C\int_{B_{2r}}\chi_{r}G(\Delta(|s_{t}|^{2})+C)
C|st|2(p)+CB2r|st|2((Δχr)G+(χrG))+CC.\leq C|s_{t}|^{2}(p)+C\int_{B_{2r}}|s_{t}|^{2}((\Delta\chi_{r})G+(\nabla\chi_{r}\nabla G))+C\leq C.

Step 2 Let s¯t\bar{s}_{t} be the average of sts_{t} on the region B2r\BrB_{2r}\backslash B_{r}, where B2rB_{2r} and BrB_{r} have the same center. That is to say,

s¯t=1Vol(B2r\Br)B2r\Brst.\bar{s}_{t}=\dfrac{1}{\text{Vol}(B_{2r}\backslash B_{r})}\int_{B_{2r}\backslash B_{r}}s_{t}.

Since s¯t\bar{s}_{t} is a constant, all its derivatives are 0. Moreover, since we are working in a region away from the y=0y=0 boundary, we know that |Ψ0||\Psi_{0}| is uniformly bounded and hence

|D1s¯t|,|D2st¯|,|D3s¯t|,|D1(γ(s¯t))|,|D2(γ(s¯t))|,|D3(γ(s¯t))||D_{1}\bar{s}_{t}|,~{}|D_{2}\bar{s_{t}}|,~{}|D_{3}\bar{s}_{t}|,~{}|D_{1}(\gamma(-\bar{s}_{t}))|,~{}|D_{2}(\gamma(-\bar{s}_{t}))|,~{}|D_{3}(\gamma(-\bar{s}_{t}))|

are all bounded uniformly. Thus

|V(Ψ0,s¯t)||V(Ψ0)|+|LΨ0s¯t|+|Q(s¯t)|C.|V(\Psi_{0},\bar{s}_{t})|\leq|V(\Psi_{0})|+|L_{\Psi_{0}}\bar{s}_{t}|+|Q(\bar{s}_{t})|\leq C.

Suppose e2st=es¯te2σes¯te^{2s_{t}}=e^{\bar{s}_{t}}e^{2\sigma}e^{\bar{s}_{t}}. (See subsection 1.3 for details on how to add Hermitian gauge transformations.) Then σisu(2)\sigma\in isu(2). We will verity that

|σ|C|sts¯t|,|s|C|σ|.|\sigma|\leq C|s_{t}-\bar{s}_{t}|,~{}~{}~{}|\nabla s|\leq C|\nabla\sigma|.

In fact, for the first inequality, suppose the eigen-values of σ\sigma are λ1\lambda_{1} and λ2\lambda_{2}. Then tr(eσ)=eλ1+eλ2\mathrm{tr}(e^{\sigma})=e^{\lambda_{1}}+e^{\lambda_{2}}. And the norm |σ||\sigma| is equivalent with |λ1|+|λ2||\lambda_{1}|+|\lambda_{2}|, which is also equivalent with tr(eσ)2\mathrm{tr}(e^{\sigma})-2 when |σ||\sigma| is bounded. On the other hand, we indeed have |st|,|s¯t|,|σ||s_{t}|,|\bar{s}_{t}|,|\sigma| are all bounded and

tr(e2σ)=tr(es¯te2stes¯t)=tr(e2ste2s¯t)=tr(e2(sts¯t)).\mathrm{tr}(e^{2\sigma})=\mathrm{tr}(e^{-\bar{s}_{t}}e^{2s_{t}}e^{-\bar{s}_{t}})=\mathrm{tr}(e^{2s_{t}}e^{-2\bar{s}_{t}})=\mathrm{tr}(e^{2(s_{t}-\bar{s}_{t})}).

Then the first inequality follows.

The second inequality is because

γ(2st)(2(st))e2st=es¯tγ(2σ)(2σ)e2σes¯t.\gamma(2s_{t})(2\nabla(s_{t}))e^{2s_{t}}=e^{\bar{s}_{t}}\gamma(2\sigma)(2\nabla\sigma)e^{2\sigma}e^{\bar{s}_{t}}.

Again, |γ(2st)|,|e2st|,|e2σ|,|γ(2σ)|,|es¯t||\gamma(2s_{t})|,|e^{2s_{t}}|,|e^{2\sigma}|,|\gamma(2\sigma)|,|e^{\bar{s}_{t}}| are all bounded above and below away 0. So |st||\nabla s_{t}| and |σ||\nabla\sigma| can bound each other.

Applying the Weitzenbock formula in subsection 1.3 and the fact that st,s¯t,σs_{t},\bar{s}_{t},\sigma are bounded above, we get

12Δ(|σ|2)+1C|σ|2(|V(Ψ0,st))|+|V(Ψ0,s¯t)|)|σ|C.-\dfrac{1}{2}\Delta(|\sigma|^{2})+\dfrac{1}{C}|\nabla\sigma|^{2}\leq~{}~{}(|V(\Psi_{0},s_{t}))|+|V(\Psi_{0},\bar{s}_{t})|)|\sigma|\leq~{}C.

So we have

BrG|s|2CBrG|σ|2CB2rχrG|σ|2CB2rχrG(1+Δ(|σ|2))\int_{B_{r}}G|\nabla s|^{2}\leq C\int_{B_{r}}G|\nabla\sigma|^{2}\leq C\int_{B_{2r}}\chi_{r}G|\nabla\sigma|^{2}\leq C\int_{B_{2r}}\chi_{r}G(1+\Delta(|\sigma|^{2}))
Cr2+CB2r(|Δ(χr)G|+|χr||G|)|σ|2Cr2+C1r3B2r\Br|σ|2Cr2+CB2r\BrG|st|2.\leq Cr^{2}+C\int_{B_{2r}}(|\Delta(\chi_{r})G|+|\nabla\chi_{r}||\nabla G|)|\sigma|^{2}\leq Cr^{2}+C\dfrac{1}{r^{3}}\int_{B_{2r}\backslash B_{r}}|\sigma|^{2}\leq Cr^{2}+C\int_{B_{2r}\backslash B_{r}}G|\nabla s_{t}|^{2}.

The last inequality above is the Poincare inequality

B2r\Br|σ|2CB2r\Br|sts¯t|2Cr2B2r\Br|st|2.\int_{B_{2r}\backslash B_{r}}|\sigma|^{2}\leq C\int_{B_{2r}\backslash B_{r}}|s_{t}-\bar{s}_{t}|^{2}\leq Cr^{2}\int_{B_{2r}\backslash B_{r}}|\nabla s_{t}|^{2}.

So f(r)Cr2+C(f(2r)f(r)).f(r)\leq Cr^{2}+C(f(2r)-f(r)).

By some algebraic manipulations and the fact that f(r)f(r) is an increasing function, one sees from above that

f(r)Cr2α,f(r)\leq Cr^{2\alpha},

for some α(0,1)\alpha\in(0,1). ∎

Note that the proof of lemma 3.1 implies that locally |st||\nabla s_{t}| has a finite L2,1+2αL^{2,1+2\alpha} norm (the Morry-Campanato norm, see appendix D). By theorem D.2 in the same appendix, we know that

[st]C0,α(B)C.[s_{t}]_{C^{0,\alpha}(B)}\leq C.

So away from the y=0y=0 boundary (say, y1y\geq 1),

|st|C0,αC.|s_{t}|_{C^{0,\alpha}}\leq C.

Moreover, when ρ\rho is large, (Here ρ\rho means the ρ\rho value of the center of the ball, hence is a constant.)

Δ(|ρst|2)+1Ci=13|ρDist|2+t|ρkst|2|V(Ψ0)|ρ2|st|,-\Delta(|\rho s_{t}|^{2})+\dfrac{1}{C}\sum\limits_{i=1}^{3}|\rho D_{i}^{*}s_{t}|^{2}+t|\rho^{k}s_{t}|^{2}\leq|V(\Psi_{0})|\rho^{2}|s_{t}|,

which is bounded above uniformly. And the proof doesn’t need to change if we replace sts_{t} with ρst\rho s_{t}, where ρ\rho is the ρ\rho-value of the center of the ball BB. In particular, we still have

|V(Ψ0,ρs¯t)|C.|V(\Psi_{0},\rho\bar{s}_{t})|\leq C.

(We need to use the first fact in subsection 1.3 and lemma 3.2 here.)

So in fact, we get when yy is not small (say y1y\geq 1), we get

stρ1C0,αC||s_{t}||_{\rho^{-1}C^{0,\alpha}}\leq C

for some α(0,1)\alpha\in(0,1).

Moreover, by exactly the same reason, if tt0>0t\geq t_{0}>0 and allow CC to depend on t0t_{0} and an additional positive integer ll. Then

stρlC0,αC||s_{t}||_{\rho^{-l}C^{0,\alpha}}\leq C

for some α(0,1)\alpha\in(0,1).

The y12Cie0,αy^{\frac{1}{2}}C_{\text{\text{ie}}}^{0,\alpha} estimates for sts_{t} near y=0y=0

(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 sts_{t} is still locally in C0,αC^{0,\alpha} near y=0y=0 since we don’t have a uniform bound on |Ψ0||\Psi_{0}| near y=0y=0, which is required in the estimate of |V(Ψ0,s¯t)||V(\Psi_{0},\bar{s}_{t})|. In fact, we only have |Ψ0|Cy|\Psi_{0}|\leq\dfrac{C}{y} when yy is small. On the other hand, the definition of Cie0,αC^{0,\alpha}_{\text{ie}} 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 BB of radius y(q)4\dfrac{y(q)}{4} centered at a point qq. We hope get an y12Cie0,αy^{\frac{1}{2}}C^{0,\alpha}_{\text{ie}} estimate for sts_{t} and some small ϵ>0\epsilon>0.

Thanks to the fact that |st|=O(y)|s_{t}|=O(\sqrt{y}) and the fact that |V(Ψ0)|=O(yn)|V(\Psi_{0})|=O(y^{n}) for any positive integer nn when yy is small.

We still define

f(r)=BrG|st|2,f(r)=\int_{B_{r}}G|\nabla s_{t}|^{2},

where BrB_{r} is the ball of radius rr centered at qq. We assume 0<ry(q)40<r\leq\dfrac{y(q)}{4}.

In the ball, we have

Δ(|st|2)+1Ci=13|Dist|2+t|st|2|V(Ψ0)||st|Cy(q).-\Delta(|s_{t}|^{2})+\dfrac{1}{C}\sum\limits_{i=1}^{3}|D_{i}^{*}s_{t}|^{2}+t|s_{t}|^{2}\leq|V(\Psi_{0})||s_{t}|\leq Cy(q).

Then by the same reason as in the step 1 of the proof of lemma 3.3,

f(r)C|st|2(q)+CB2r|st|2(|Δχr||G|+|χr||G|)+Cy(q)Cy(q).f(r)\leq C|s_{t}|^{2}(q)+C\int_{B_{2r}}|s_{t}|^{2}(|\Delta\chi_{r}||G|+|\nabla\chi_{r}||\nabla G|)+Cy(q)\leq Cy(q).

On the other hand, in step 2, we have |Ψ0|=O(y1)|\Psi_{0}|=O(y^{-1}) and |st|Cy|s_{t}|\leq C\sqrt{y}. So

|D1s¯t|,|D2st¯|,|D3s¯t|,|D1(γ(s¯t))|,|D2(γ(s¯t))|,|D3(γ(s¯t))||D_{1}\bar{s}_{t}|,~{}|D_{2}\bar{s_{t}}|,~{}|D_{3}\bar{s}_{t}|,~{}|D_{1}(\gamma(-\bar{s}_{t}))|,~{}|D_{2}(\gamma(-\bar{s}_{t}))|,~{}|D_{3}(\gamma(-\bar{s}_{t}))|

are all bounded by Cy12Cy^{-\frac{1}{2}} uniformly. So |LΨ0s¯t|Cy12|L_{\Psi_{0}}\bar{s}_{t}|\leq Cy^{-\frac{1}{2}} and |Q(s¯t)|Cy1|Q(\bar{s}_{t})|\leq Cy^{-1}, where s¯t\bar{s}_{t} has the same meaning as in the proof of lemma 3.3.

So what we get is

|V(Ψ0,s¯t)|Cy(q)1.|V(\Psi_{0},\bar{s}_{t})|\leq Cy(q)^{-1}.

And like in step 2 of the proof of lemma 3.3,

12Δ(|σ|2)+1C|σ|2Cy(q)1y(q)12=Cy(q)12.-\dfrac{1}{2}\Delta(|\sigma|^{2})+\dfrac{1}{C}|\nabla\sigma|^{2}\leq Cy(q)^{-1}\cdot y(q)^{\frac{1}{2}}=Cy(q)^{-\frac{1}{2}}.

So what we have is

BrG|s|2CB2rχrG(y(q)12+Δ(|σ|2))\int_{B_{r}}G|\nabla s|^{2}\leq C\int_{B_{2r}}\chi_{r}G(y(q)^{-\frac{1}{2}}+\Delta(|\sigma|^{2}))
Cy(q)12r2+C1r3B2r\Br|σ|2Cy(q)12r2+CB2r\BrG|st|2.\leq Cy(q)^{-\frac{1}{2}}r^{2}+C\dfrac{1}{r^{3}}\int_{B_{2r}\backslash B_{r}}|\sigma|^{2}\leq Cy(q)^{-\frac{1}{2}}r^{2}+C\int_{B_{2r}\backslash B_{r}}G|\nabla s_{t}|^{2}.

In fact, what we get is

f(r)Cy(q)12r2+C(f(2r)f(r)).f(r)\leq Cy(q)^{-\frac{1}{2}}r^{2}+C(f(2r)-f(r)).

Doing algebraic manipulations and keep in mind that ry(q)4r\leq\dfrac{y(q)}{4} and we have f(r)Cy(q)f(r)\leq Cy(q) for any 0<ry(q)40<r\leq\dfrac{y(q)}{4}, we get

f(r)Cy(q)12αr2αf(r)\leq Cy(q)^{1-2\alpha}r^{2\alpha}

for some α(0,1)\alpha\in(0,1).

This implies that

Br|st|2Cy(q)12αr2α+1.\int_{B_{r}}|\nabla s_{t}|^{2}\leq Cy(q)^{1-2\alpha}r^{2\alpha+1}.

So

[st]C0,αCy(q)12α.[s_{t}]_{C^{0,\alpha}}\leq Cy(q)^{1-2\alpha}.

Transfer it to the edge version, we get

[st]Cie0,αCy(q)1α.[s_{t}]_{C^{0,\alpha}_{\text{ie}}}\leq Cy(q)^{1-\alpha}.

We may assume α<12\alpha<\dfrac{1}{2}, so together with the fact that |st|Cy12|s_{t}|\leq Cy^{\frac{1}{2}}, we get

sty12Cie0,αC.||s_{t}||_{y^{\frac{1}{2}}C^{0,\alpha}_{\text{ie}}}\leq C.

Even near the type III boundary or any corner, we have the same estimate. There yy should be written as rψr\psi. (We are somehow abuse of notations here. The letter rr 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 XX.

Proposition 3.4.

There is a constant CC which depends only on Ψ0\Psi_{0} such that for all tt, uniformly we have

stρ1ψ12r12Cie0,α=st𝒳1/2,1/2,10,αC.||s_{t}||_{\rho^{-1}\psi^{\frac{1}{2}}r^{\frac{1}{2}}C^{0,\alpha}_{\text{ie}}}=||s_{t}||_{\mathcal{X}^{0,\alpha}_{1/2,1/2,-1}}\leq C.

Moreover, if tt0>0t\geq t_{0}>0 and allow CC to depend on t0t_{0} and a positive integer ll, then

stρlψ12r12Cie0,α=st𝒳1/2,1/2,l0,αC.||s_{t}||_{\rho^{-l}\psi^{\frac{1}{2}}r^{\frac{1}{2}}C^{0,\alpha}_{\text{ie}}}=||s_{t}||_{\mathcal{X}^{0,\alpha}_{1/2,1/2,-l}}\leq C.

The C1,αC^{1,\alpha} estimates away from boundaries

(Hilderbrandt’s estimate) 

In the following arguments, sometimes we need to shrink α\alpha. So we allow α\alpha to change from line to line, but always independent with sts_{t}. We first work in a region away from the y=0y=0 boundary (say, y1y\geq 1).

Recall that

V(Ψ0,st)+tst=0.V(\Psi_{0},s_{t})+ts_{t}=0.

So

V(Ψ0)+LΨ0st+Q(st)+tst=0.V(\Psi_{0})+L_{\Psi_{0}}s_{t}+Q(s_{t})+ts_{t}=0.

Away from boundaries, this can be written as

Δst=A+B((st))+C(stst),\Delta s_{t}=A+B(\nabla(s_{t}))+C(\nabla s_{t}\otimes\nabla s_{t}),

where A,B,CA,B,C are all bounded.

Lemma 3.5.

Consider the region away from boundaries. Then for some α(0,1)\alpha\in(0,1),

stC0,αC,||\nabla s_{t}||_{C^{0,\alpha}}\leq C,

where CC depends only on Ψ0\Psi_{0}.

Proof.

Let BrB_{r} be a ball of radius rr whose center is away from all boundaries. We assume r1r\leq 1.

Then from the proof of lemma 3.3, we know

Br|st|2Cr1+2α.\int_{B_{r}}|\nabla s_{t}|^{2}\leq Cr^{1+2\alpha}.

Let st=g+hs_{t}=g+h, where Δh=0\Delta h=0 in BrB_{r} and h=sth=s_{t} on Br\partial B_{r}, Δg=Δst\Delta g=\Delta s_{t} in BrB_{r} and g=0g=0 on Br\partial B_{r}.

Use (st¯)r(\overline{s_{t}})_{r} to represent the average of sts_{t} over the ball BrB_{r}. That is to say,

(st¯)r:=1Vol(Br)Brst.(\overline{s_{t}})_{r}:=\dfrac{1}{\text{Vol}(B_{r})}\int_{B_{r}}s_{t}.

Similar definitions for g¯r\bar{g}_{r} and h¯r\bar{h}_{r} etc..

Let f(r):=Br|st(st¯)r|2f(r):=\displaystyle\int_{B_{r}}|\nabla s_{t}-(\overline{\nabla s_{t}})_{r}|^{2}. Using theorem D.2 in the appendix D, it suffices to prove that f(r)Cr3+2αf(r)\leq Cr^{3+2\alpha^{\prime}} for some possibly smaller 0<αα0<\alpha^{\prime}\leq\alpha.

Note that we already have

f(r)Cr1+2α.f(r)\leq Cr^{1+2\alpha}.

By lemma D.3 in the appendix D, for any 0<r1<r0<r_{1}<r,

1r13Br1|hh¯r1|2C(r1r)21r3Br|hh¯r|2.\dfrac{1}{r_{1}^{3}}\int_{B_{r_{1}}}|\nabla h-\overline{\nabla h}_{r_{1}}|^{2}\leq C(\dfrac{r_{1}}{r})^{2}\cdot\dfrac{1}{r^{3}}\int_{B_{r}}|\nabla h-\overline{\nabla h}_{r}|^{2}.

On the other hand,

Br|g|2=Br<Δg,g>C(Br|g|(1+|st|2).\int_{B_{r}}|\nabla g|^{2}=\int_{B_{r}}<\Delta g,g>~{}\leq~{}C(\int_{B_{r}}|g|(1+|\nabla s_{t}|^{2}).

Since stC0,αC||s_{t}||_{C^{0,\alpha}}\leq C, we have on the entire BrB_{r},

|st(st¯)r|Crα.|s_{t}-(\overline{s_{t}})_{r}|\leq Cr^{\alpha}.

By maximal principle on BrB_{r}, we have

|h(st¯)r|Crα.|h-(\overline{s_{t}})_{r}|\leq Cr^{\alpha}.

So

|g|=|sth|Crα.|g|=|s_{t}-h|\leq Cr^{\alpha}.

Thus

Br|g|2CrαBr|st|2Cr1+3α.\int_{B_{r}}|\nabla g|^{2}\leq Cr^{\alpha}\int_{B_{r}}|\nabla s_{t}|^{2}\leq Cr^{1+3\alpha}.

We know for rRr\leq R,

f(r)=Br|st(st¯)r|2Br|hh¯r|2+Br|g|2f(r)=\displaystyle\int_{B_{r}}|\nabla s_{t}-(\overline{\nabla s_{t}})_{r}|^{2}\leq\int_{B_{r}}|\nabla h-\overline{\nabla h}_{r}|^{2}+\int_{B_{r}}|\nabla g|^{2}
C(rR)5f(R)+Cr1+3α.\leq C(\dfrac{r}{R})^{5}f(R)+Cr^{1+3\alpha}.

If we fix R=1R=1, then this implies

f(r)Cr1+3α.f(r)\leq Cr^{1+3\alpha}.

So the Campanato norm 2,1+3α\mathcal{L}^{2,1+3\alpha} of st\nabla s_{t} over B1B_{1} is bounded above by CC.

By theorem D.2 in the appendix D, this implies that

Br|st|2Cmax{r1+3α,r3}.\int_{B_{r}}|\nabla s_{t}|^{2}\leq C\max\{r^{1+3\alpha},r^{3}\}.

We may use the smaller power between r1+3αr^{1+3\alpha} and r3r^{3} as the new starting point and re-run the argument. After finitely many times of iterations, we get for some possibly smaller α\alpha^{\prime},

f(r)Cr3+2α.f(r)\leq Cr^{3+2\alpha^{\prime}}.

Then by theorem D.2 in the appendix D, we know that

stC0,αC.||\nabla s_{t}||_{C^{0,\alpha^{\prime}}}\leq C.

The weighted Cie1,αC^{1,\alpha}_{\text{ie}} estimates near boundaries

(an adapted version of Hilderbrandt’s estimate) 

Near the y=0y=0 boundary (but away from the r=0r=0 boundary), we have

Δst=A+B(st)+C(stst),\Delta s_{t}=A+B(\nabla s_{t})+C(\nabla s_{t}\otimes s_{t}),

where only CC is bounded, and A=O(1y2)A=O(\dfrac{1}{y^{2}}), B=O(1y)B=O(\dfrac{1}{y}). The fact that AA, BB are not bounded is because |Ψ0||\Psi_{0}| is not bounded there.

Thus on a ball BB of radius y(q)4\dfrac{y(q)}{4} centered at y(q)y(q), we want to re-run the proof of lemma 3.5.

To start, we have

Br|st|2Cy(q)12r1+2α.\int_{B_{r}}|\nabla s_{t}|^{2}\leq Cy(q)^{\frac{1}{2}}r^{1+2\alpha}.

Let R=y(q)4R=\dfrac{y(q)}{4}. And the same argument as in the proof of lemma 3.5 leads to (we still have the same definition of f(r)f(r))

f(r)C(rR)5f(R)+C(y(q))12r1+3αCy(q)12αr1+3α.f(r)\leq C(\dfrac{r}{R})^{5}f(R)+C(y(q))^{\frac{1}{2}}r^{1+3\alpha}\leq Cy(q)^{\frac{1}{2}-\alpha}r^{1+3\alpha}.

As long as the exponent of rr is less or equal than 33 we can run the same argument and gain an extra y(q)αrαy(q)^{-\alpha}r^{\alpha} factor. After finitely many times of iterations, this will lead to

f(r)Cy(q)32r3+2α,f(r)\leq Cy(q)^{-\frac{3}{2}}r^{3+2\alpha^{\prime}},

for some possibly smaller α\alpha^{\prime}.

This implies the y(q)12C0,αy(q)^{\frac{1}{2}}C^{0,\alpha^{\prime}} norm of y(q)sty(q)\nabla s_{t} is bounded above by CC on this ball. Note that CC doesn’t depend on the ball. So we conclude that, in a region near the y=0y=0 boundary away from r=0r=0 boundary and for possibly smaller α\alpha, we have

sty12Cie1,αC.||s_{t}||_{y^{\frac{1}{2}}C^{1,\alpha}_{ie}}\leq C.

Even when approaching the r=0r=0 boundary, the argument doesn’t need to change and we get locally

str12ψ12Cie1,αC.||s_{t}||_{r^{\frac{1}{2}}\psi^{\frac{1}{2}}C^{1,\alpha}_{ie}}\leq C.

On the other hand, near the ρ+\rho\rightarrow+\infty boundary, we have

Δρlst=A+B(lρst)+C(lρstρlst),\Delta\rho^{l}s_{t}=A+B(\nabla^{l}\rho s_{t})+C(\nabla^{l}\rho s_{t}\otimes\rho^{l}s_{t}),

where A,B,CA,B,C are bounded, l=1l=1 if t[0,1]t\in[0,1] and ll can be any integer if tt0>0t\geq t_{0}>0 with the bounds of A,B,CA,B,C depends on t0t_{0} and ll.

So examine the proof of lemma 3.5, we get:

st𝒳1/2,1/2,11,αC,||s_{t}||_{\mathcal{X}^{1,\alpha}_{1/2,1/2,-1}}\leq C,

where CC depends only on Ψ0\Psi_{0}. When tt0>0t\geq t_{0}>0 and when CC is allowed to depend on t0t_{0}, the positive integer ll, then

st𝒳1/2,1/2,l1,αC.||s_{t}||_{\mathcal{X}^{1,\alpha}_{1/2,1/2,-l}}\leq C.

All the higher weighted Ciek,αC^{k,\alpha}_{\text{ie}} norms with k2k\geq 2 follow by the standard elliptic Holder type interior estimates. So to conclude we have

Proposition 3.6.

There exists an α(0,1)\alpha\in(0,1) such that

st𝒳1/2,1/2,1k,αC.||s_{t}||_{\mathcal{X}^{k,\alpha}_{1/2,1/2,-1}}\leq C.

Here the constant CC depends only on Ψ0\Psi_{0} and the positive integer kk. When tt0>0t\geq t_{0}>0 and we assume CC depends only on t0t_{0}, Ψ0\Psi_{0}, positive integers kk and ll, we have stronger estimate:

st𝒳1/2,1/2,lk,αC.||s_{t}||_{\mathcal{X}^{k,\alpha}_{1/2,1/2,-l}}\leq C.

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 ss is a solution for some tt. We hope to study the operator Ls:=LΨsL_{s}:=L_{\Psi_{s}}. Unfortunately this operator seems to be hard to study. However, suppose t>0t>0. Consider alternatively the operator Ls,t:=Ls+tL_{s,t}:=L_{s}+t. Then it behaves much better and can be analysis-ed.

In fact, we have the following lemma.

Lemma 3.7.

Assume t>0t>0 and ss in 𝒳12,12,lk\mathcal{X}^{k}_{\frac{1}{2},\frac{1}{2},-l} (for all positive integers k,lk,l) is a solution of

V(Ψ0,s)+ts=0.V(\Psi_{0},s)+ts=0.

Suppose F𝒳32+ϵ,32+ϵ,lkF\in\mathcal{X}^{k}_{-\frac{3}{2}+\epsilon,-\frac{3}{2}+\epsilon,-l} for some small ϵ>0\epsilon>0 and all positive integers kk and some large enough l>0l>0. Then there exists a u𝒳12,12,lku\in\mathcal{X}^{k}_{\frac{1}{2},\frac{1}{2},-l} for all kk such that

Ls,t(u)=F.L_{s,t}(u)=F.

Moreover, the 𝒳12,12,lk+2\mathcal{X}^{k+2}_{\frac{1}{2},\frac{1}{2},-l} norm of uu is bounded above by (a constant times) the 𝒳32+ϵ,32+ϵ,lk\mathcal{X}^{k}_{-\frac{3}{2}+\epsilon,-\frac{3}{2}+\epsilon,-l} norm of FF.

Proof.

We use two cut-off functions χ1,χ2\chi_{1},\chi_{2} to divide FF into two parts: Fi=χiFF_{i}=\chi_{i}F, i=1,2i=1,2. Here χi\chi_{i} are all smooth functions with

χ1+χ2=1.\chi_{1}+\chi_{2}=1.

We assume χ1\chi_{1} is supported in a region such that y<min{2,2Rl}y<\min\{2,2R^{-l}\}. And χ2\chi_{2} is supported in a region such that either y>min{1,Rl}y>\min\{1,R^{-l}\}. Clearly we only need to solve Ls,t(ui)=FiL_{s,t}(u_{i})=F_{i} on each region.

We first consider F1F_{1}. 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 y0y\rightarrow 0, the operator Ls,tL_{s,t} has the same normal operator as if t=0t=0 and Ψ\Psi is the model solution in their construction. Here is the only fact that we need: The weight r12ψ12r^{\frac{1}{2}}\psi^{\frac{1}{2}} lies in the Fredholm range both near the r0r\rightarrow 0 boundary and the ψ0\psi\rightarrow 0 boundary (and is compatible with the corners). (Alternatively, readers may compute the Fredholm weight of Ls,tL_{s,t} directly. It is actually easier than Mazzeo and Witten’s computations in [16], which takes more situations into account.) In particular, near those boundaries,

Ls,t(u)=F1L_{s,t}(u)=F_{1}

has a solution in the support of χ1\chi_{1} with ur12ψ12Cieku\in r^{\frac{1}{2}}\psi^{\frac{1}{2}}C^{k}_{\text{ie}} for all kk. Here we allow that uu is nonzero in a slightly larger region and doesn’t satisfy the equation outside of the support of χ1\chi_{1}. But we may add Ls,t(u)F1L_{s,t}(u)-F_{1} (which is supported in a relatively compact region) to F2F_{2} there and throw it into the next step.

Here is the next step: We solve the equation for F2F_{2}. Consider the following functional on uu, where u𝒳12,12,lku\in\mathcal{X}^{k}_{\frac{1}{2},\frac{1}{2},-l} for all k,lk,l:

A(u):=X(i=1,2,y|iu|2+i=1,2,3|[Φi,u]|2+<F2,u>).A(u):=\int_{X}(\sum\limits_{i=1,2,y}|\nabla_{i}u|^{2}+\sum\limits_{i=1,2,3}|[\Phi_{i},u]|^{2}+<F_{2},u>).

We may first consider the Banach space \mathbb{H} defined by completion of smooth compact supported functions using the following norm:

u2=X(i=1,2,y|iu|2+i=1,2,3|[Φi,u]|2).||u||_{\mathbb{H}}^{2}=\int_{X}(\sum\limits_{i=1,2,y}|\nabla_{i}u|^{2}+\sum\limits_{i=1,2,3}|[\Phi_{i},u]|^{2}).

Clearly, A(u)A(u) is bounded in \mathbb{H} and has a Dirichlet minimizor in \mathbb{H}. (Recall that |F2|CRl|F_{2}|\leq CR^{-l} for sufficiently large ll.) This minimizor is unique because of the convexity of A(u)A(u). By a standard elliptic regularity argument, this minimizor, as a week solution of Ls,t(u)=F2L_{s,t}(u)=F_{2}, is smooth in the interior of XX. Moreover, there is a Hardy type of inequality for uu\in\mathbb{H}:

Xy2|u|2CXi=1,2,y|iu|2Xi=1,2,y|iu|2u2<+.\int_{X}y^{-2}|u|^{2}\leq C\int_{X}\sum\limits_{i=1,2,y}|\partial_{i}u|^{2}\leq\int_{X}\sum\limits_{i=1,2,y}|\nabla_{i}u|^{2}\leq||u||_{\mathbb{H}}^{2}<+\infty.

In particular, the integral of both |u||u| and |u||\nabla u| over the region R<λR<\lambda has at most a polynomial growth as λ+\lambda\rightarrow+\infty. We have

|<Lt,su,u>|=|<F2,u>||F2||u|.|<L_{t,s}u,u>|=|<F_{2},u>|\leq|F_{2}||u|.

But

<Lt,su,u>=<ΔΨu+tu,u>=12Δ(|u|2)+i=1,2,y|Aiu|2+i=1,2,3|[Φi,u]|2+t|u|2<L_{t,s}u,u>=<-\Delta_{\Psi}u+tu,u>=-\dfrac{1}{2}\Delta(|u|^{2})+\sum\limits_{i=1,2,y}|\nabla_{A_{i}}u|^{2}+\sum\limits_{i=1,2,3}|[\Phi_{i},u]|^{2}+t|u|^{2}
12Δ(|u|2)+|u|2+t|u|2=<Δu,u>(Δ|u|+t|u|)|u|.\geq-\dfrac{1}{2}\Delta(|u|^{2})+|\nabla u|^{2}+t|u|^{2}=-<\Delta u,u>~{}~{}~{}\geq~{}~{}~{}(-\Delta|u|+t|u|)|u|.

So we get

Δ|u|+t|u||F2|.-\Delta|u|+t|u|\leq|F_{2}|.

Let χλ\chi_{\lambda} be a cut-off function that equals 11 when RλR\leq\lambda and equals 0 when R2λR\geq 2\lambda. Moreover, we assume |χλ|=O(R1),|2χλ|=O(R2).|\nabla\chi_{\lambda}|=O(R^{-1}),~{}|\nabla^{2}\chi_{\lambda}|=O(R^{-2}). And let uλ:=uχλu_{\lambda}:=u\chi_{\lambda}. Then recall the Green’s function of Δ+t-\Delta+t is Gt,qG_{t,q} (defined in subsection 3.4) and the fact that uλu_{\lambda} has compact support, we have (centering at any point qXq\in X)

|uλ(q)|XGt,qΔ|uλ|Csupport(χλ)Gt,q(R2|u|+R1|u|)+XGt,q|F2|).|u_{\lambda}(q)|\leq\int_{X}G_{t,q}\Delta|u_{\lambda}|\leq C\int_{\text{support}(\nabla\chi_{\lambda})}G_{t,q}(R^{-2}|u|+R^{-1}|\nabla u|)+\int_{X}G_{t,q}|F_{2}|).

Note that when λ+\lambda\rightarrow+\infty, Gt,qG_{t,q} has exponential decay while the integral of |u||u| and |u||\nabla u| on support(χλ)(\nabla\chi_{\lambda}) have at most polynomial growth. So

limλ+support(χλ)Gt,q(R2|u|+R1|u|)=0.\lim\limits_{\lambda\rightarrow+\infty}\int_{\text{support}(\nabla\chi_{\lambda})}G_{t,q}(R^{-2}|u|+R^{-1}|\nabla u|)=0.

And letting λ+\lambda\rightarrow+\infty,

|u(q)|XGt,q|F2|.|u(q)|\leq\int_{X}G_{t,q}|F_{2}|.

We may divide XX into two parts: Let BB be a ball of radius R(q)2\dfrac{R(q)}{2}, where R(q)R(q) is the RR value of qq. Then

XGt,q|F2|=XBGt,q|F2|+X\BGt,q|F2|.\int_{X}G_{t,q}|F_{2}|=\int_{X\cap B}G_{t,q}|F_{2}|+\int_{X\backslash B}G_{t,q}|F_{2}|.

When R(q)+R(q)\rightarrow+\infty, the second integral above decays exponentially in R(q)R(q). The first integral is bounded by (recall that |F2|CRl|F_{2}|\leq CR^{-l})

CRlBXGt,qCRlXGt,qCRl.CR^{-l}\int_{B\cap X}G_{t,q}\leq CR^{-l}\int_{X}G_{t,q}\leq CR^{-l}.

So we get

|u|CRl.|u|\leq CR^{-l}.

Once we have this point-wise bound, then the bound on Holder norms ρlCk,α\rho^{-l}C^{k,\alpha} of uu 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 uu that we get for F1F_{1} and F2F_{2} together and finish the proof.

Suppose for some t>0t>0, we have a solution ss with

V(Ψ0,s)+ts=0.V(\Psi_{0},s)+ts=0.

Consider a small variation ϵα\epsilon\alpha on top of ss and a small variation ϵ-\epsilon on top of tt:

V(Ψ0,s+ϵα)+(tϵ)(s+ϵα)=ϵ(Ls,t(α)s)+ϵ2(Q(α)α).V(\Psi_{0},s+\epsilon\alpha)+(t-\epsilon)(s+\epsilon\alpha)=\epsilon(L_{s,t}(\alpha)-s)+\epsilon^{2}(Q(\alpha)-\alpha).

Note that Q(α)αQ(\alpha)-\alpha is a bounded map from 𝒳12,12,lk+2,α\mathcal{X}_{\frac{1}{2},\frac{1}{2},-l}^{k+2,\alpha} to 𝒳32+ϵ,32+ϵ,lk,α\mathcal{X}^{k,\alpha}_{-\frac{3}{2}+\epsilon,-\frac{3}{2}+\epsilon,-l} . So when ϵ\epsilon is small enough, by implicit function theorem, the equation

Ls,t(α)=s+ϵ(αQ(α))L_{s,t}(\alpha)=s+\epsilon(\alpha-Q(\alpha))

has a solution in 𝒳μ,v,lk\mathcal{X}^{k}_{\mu,v,-l}. This proves the second bullet of the facts in subsection 3.3.

Suppose when t>0t>0, sts_{t} is a solution in 𝒳12ϵ,12ϵ,lk,α\mathcal{X}^{k,\alpha}_{\frac{1}{2}-\epsilon,\frac{1}{2}-\epsilon,-l}. We have

V(Ψ0)+L0,t(st)+Q(st)=0,V(\Psi_{0})+L_{0,t}(s_{t})+Q(s_{t})=0,

where Q(st)Q(s_{t}) actually maps 𝒳12ϵ,12ϵ,lk+2,α\mathcal{X}_{\frac{1}{2}-\epsilon,\frac{1}{2}-\epsilon,-l}^{k+2,\alpha} into 𝒳12ϵ,12ϵ,lk,α\mathcal{X}^{k,\alpha}_{-1-2\epsilon,-1-2\epsilon,-l} for sufficiently small ϵ>0\epsilon>0. In particular, Q(st)Q(s_{t}) is in 𝒳32,32,lk,α\mathcal{X}^{k,\alpha}_{-\frac{3}{2},-\frac{3}{2},-l}. Then because μ[12ϵ,12]\mu\in[\dfrac{1}{2}-\epsilon,\dfrac{1}{2}] and v[12ϵ,12]v\in[\dfrac{1}{2}-\epsilon,\dfrac{1}{2}] are all Fredholm weights for L0,tL_{0,t} near those boundaries, and V(Ψ0)V(\Psi_{0}) vanishes up to infinite order at all boundaries. So inductively sts_{t} lies in 𝒳12,12,lk,α\mathcal{X}^{k,\alpha}_{\frac{1}{2},\frac{1}{2},-l} for all k,lk,l. 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 EE is the trivial 2\mathbb{C}^{2} bundle whose SL(2,)SL(2,\mathbb{C}) 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 SU(2)SU(2) Hermitian metric on EE is represented by an 2×22\times 2 matrix HH such that H=HH^{*}=H, detH=1\det H=1, and positive definite. Each such metric gives EE an SU(2)SU(2) structure. (The condition detH=1\det H=1 keeps the SL(2,)SL(2,\mathbb{C}) structure unchanged.) The inner product of two sections s1,s2Γ(E)s_{1},s_{2}\in\Gamma(E) (represented by 22-d vectors with variable coefficients) is

    <s1,s2>H=12tr(H1s1Hs2+H1s2Hs1).<s_{1},s_{2}>_{H}=\dfrac{1}{2}\mathrm{tr}(H^{-1}s_{1}^{*}Hs_{2}+H^{-1}s_{2}^{*}Hs_{1}).

    Unless otherwise specified, we typically just use the inner product defined by H=IH=I, that is

    <s1,s2>=12tr(s1s2+s2s1)<s_{1},s_{2}>=\dfrac{1}{2}\mathrm{tr}(s_{1}^{*}s_{2}+s_{2}^{*}s_{1})
  • Infinitesimal SL(2,)SL(2,\mathbb{C}) gauge transformations are represented by sections of sl(2,)sl(2,\mathbb{C}). Infinitesimal SU(2)SU(2) gauge transformations are represented by sections of su(2)su(2). And sl(2,)=su(2)isu(2)sl(2,\mathbb{C})=su(2)\oplus isu(2), where isu(2)isu(2) are the Hermitian sl(2,)sl(2,\mathbb{C}) elements.

  • One useful formula: Suppose s(t)s(t) is a differentiable 11-parameter family of matrices. Then

    ddtes=esγ(s)(dsdt)=γ(s)(dsdt)es,\dfrac{d}{dt}e^{s}=e^{s}~{}\gamma(-s)(\dfrac{ds}{dt})=\gamma(s)(\dfrac{ds}{dt})~{}e^{s},

    where

    γ(s)(M)=(eads1ads)(M)=k=0+(ads)k(k+1)!(M),\gamma(s)(M)=(\dfrac{e^{ad_{s}}-1}{ad_{s}})(M)=\sum\limits_{k=0}^{+\infty}\dfrac{(ad_{s})^{k}}{(k+1)!}(M),
    ads(M)=[s,M]=sMMs.ad_{s}(M)=[s,M]=sM-Ms.
    Proof.

    Let Ls(M)=sM,Rs(M)=Ms.L_{s}(M)=sM,R_{s}(M)=Ms. Then Ls,Rs,adsL_{s},R_{s},ad_{s} commutes. And Ls=Rs+adsL_{s}=R_{s}+ad_{s}. Let Cnm=n!m!(nm)!C_{n}^{m}=\dfrac{n!}{m!(n-m)!}. Then

    ddt(es)=ddt(n=0+snn!)=n=0+1n!(k=0n1sn1k(dsdt)sk)=n=0+1n!(k=0n1RskLsn1k(dsdt))\dfrac{d}{dt}(e^{s})=\dfrac{d}{dt}(\sum\limits_{n=0}^{+\infty}\dfrac{s^{n}}{n!})=\sum\limits_{n=0}^{+\infty}\dfrac{1}{n!}(\sum\limits_{k=0}^{n-1}s^{n-1-k}(\dfrac{ds}{dt})s^{k})=\sum\limits_{n=0}^{+\infty}\dfrac{1}{n!}(\sum\limits_{k=0}^{n-1}R_{s}^{k}L_{s}^{n-1-k}(\dfrac{ds}{dt}))
    =n=0+k=0n11n!l=0n1kCn1kl(Rsl+k(ads)n1kl)(dsdt)~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}=\sum\limits_{n=0}^{+\infty}\sum\limits_{k=0}^{n-1}\dfrac{1}{n!}\sum\limits_{l=0}^{n-1-k}C_{n-1-k}^{l}(R_{s}^{l+k}(ad_{s})^{n-1-k-l})(\dfrac{ds}{dt})
    (letv=l+k)=n=0+1n!v=0n1(l=0vCn1v+ll)Rsv(ads)n1v(dsdt)(\text{let}~{}~{}v=l+k)~{}~{}=\sum\limits_{n=0}^{+\infty}\dfrac{1}{n!}\sum\limits_{v=0}^{n-1}(\sum\limits_{l=0}^{v}C_{n-1-v+l}^{l})R_{s}^{v}(ad_{s})^{n-1-v}(\dfrac{ds}{dt})
    =n=0+1n!v=0n1CnvRsv(ads)n1v(dsdt)=\sum\limits_{n=0}^{+\infty}\dfrac{1}{n!}\sum\limits_{v=0}^{n-1}C_{n}^{v}R_{s}^{v}(ad_{s})^{n-1-v}(\dfrac{ds}{dt})
    (letj=n1v)=(v=0+1v!Rsv)(j=0+1(j+1)!(ads)j)(dsdt)=γ(s)(dsdt)es.(\text{let}~{}~{}j=n-1-v)~{}~{}=(\sum\limits_{v=0}^{+\infty}\dfrac{1}{v!}R_{s}^{v})(\sum\limits_{j=0}^{+\infty}\dfrac{1}{(j+1)!}(ad_{s})^{j})(\dfrac{ds}{dt})=\gamma(s)(\dfrac{ds}{dt})~{}~{}e^{s}.

    The other identity can be derived in the same way using Rs=Ls+adsR_{s}=L_{s}+ad_{-s}. ∎

  • Another useful formula: Suppose s,Ms,M are two 2×22\times 2 matrices. Then

    esMes=eads(M)=M+(eads1ads)(adsM)=M+γ(s)([M,s]).e^{-s}Me^{s}=e^{ad_{-s}}(M)=M+(\dfrac{e^{ad_{-s}}-1}{ad_{-s}})(ad_{-s}M)=M+\gamma(-s)([M,s]).
    Proof.

    This is straightforward:

    esMes=k,l=1+1k!l!(1)k(Ls)k(Rs)lM=v=0+k=0v(1)k(Ls)k(Rs)vkMe^{-s}Me^{s}=\sum\limits_{k,l=1}^{+\infty}\dfrac{1}{k!l!}(-1)^{k}(L_{s})^{k}(R_{s})^{l}M=\sum\limits_{v=0}^{+\infty}\sum\limits_{k=0}^{v}(-1)^{k}(L_{s})^{k}(R_{s})^{v-k}M
    =v=0+1v!Cvk(1)k(Ls)k(Rs)vkM=v=0+1v!(RsLs)vM=eadsM.=\sum\limits_{v=0}^{+\infty}\dfrac{1}{v!}C_{v}^{k}(-1)^{k}(L_{s})^{k}(R_{s})^{v-k}M=\sum\limits_{v=0}^{+\infty}\dfrac{1}{v!}(R_{s}-L_{s})^{v}M=e^{ad_{-s}}M.

  • Suppose ss is Hermitian. Then the operator γ(s)=(eads1ads)\gamma(s)=(\dfrac{e^{ad_{s}}-1}{ad_{s}}) has a square root:

    v(s)=γ(s)=eads1ads.v(s)=\sqrt{\gamma(s)}=\sqrt{\dfrac{e^{ad_{s}}-1}{ad_{s}}}.

    In fact, consider the function fp(x)=(ex1x)p=exp(pln(ex1x))f_{p}(x)=(\dfrac{e^{x}-1}{x})^{p}=\exp(p\cdot\ln(\dfrac{e^{x}-1}{x})), where pp is any real number. Clearly fp(x)f_{p}(x) is a real analytical function over xx\in\mathbb{R}. It has a convergent Taylor series expansion at any point. On the other hand, when ss is Hermitian, adsad_{s} is also a self-adjoint operator. So it is diagonalizable and has real eigen-values. In particular, when p=12p=\dfrac{1}{2}, we use this expansion to define

    v(s)=f1/2(ads).v(s)=f_{1/2}(ad_{s}).

Appendix B Compactification of XX

This appendix gives a preferred way to compactify XX as a manifold with boundaries and corners X^\hat{X}. In this paper, we have always assumed that XX 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 XX. 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 (ρ+\rho\rightarrow+\infty, or equivalently R+R\rightarrow+\infty)

Recall that R2=|z|2+y2R^{2}=|z|^{2}+y^{2}. Here is the definition of ρ\rho: It is a smooth function from XX to [1,+)[1,+\infty) that equals 11 when RR is small and equals RR when RR is large.

So ρ=+\rho=+\infty (or equivalently 1ρ=0\dfrac{1}{\rho}=0) defines a boundary for the compatification XX. This is the type I boundary.

Type III boundaries (r=0r=0)

Here is the definition of rr: It is a smooth function from XX to (0,1](0,1].

Suppose z0z_{0} is any root of P(z)P(z) (which corresponds to a “knotted point” at z=z0,y=0z=z_{0},y=0 of the generalized Nahm pole boundary condition). When both yy and |zz0||z-z_{0}| are small, we require r2=|zz0|2+y2r^{2}=|z-z_{0}|^{2}+y^{2}. When either yy is large or zz is away from all roots of P(z)P(z), we require that r=1r=1.

Then r=0r=0 defines the type III boundaries of the compatification of XX. Note that we have blowed up at (z0,0)(z_{0},0) for each root z0z_{0} of P(z)P(z).

Type II boundary (ψ=0\psi=0, or when ρ=r=1\rho=r=1 equivalently y=0y=0 there)

Here is the definition of ψ\psi: It is a smooth function from XX to (0,1](0,1].

Away from the type I and type III boundaries (say, r=ρ=1r=\rho=1), when yy is small, we require ψ=y\psi=y. When yy is large, we require ψ=1\psi=1.

Near a type I boundary, if yR\dfrac{y}{R} is small, then we require ψ=yR\psi=\dfrac{y}{R}. When yR\dfrac{y}{R} is close to 11, we require that ψ=1\psi=1.

Near a type III boundary, we require ψ=yr\psi=\dfrac{y}{r}.

Note that effectively, away from other boundaries, ψ=0\psi=0 and y=0y=0 define the same boundary. But we use ψ\psi instead of yy because it is also compactible with other boundaries.

Type A and type B corners

Type A corners are given by ψ=0,ρ=+\psi=0,\rho=+\infty (or equivalently ψ=1ρ=0\psi=\dfrac{1}{\rho}=0). And type B corners are given by ψ=r=0\psi=r=0.

A remark on the coordinates:

When ρ\rho is large, since ρ=R\rho=R, we may freely choose to use either RR or ρ\rho there for the same meaning. But usually we use ρ\rho if we want to emphasize that it equals 11 (not arbitrarily small) when R0R\rightarrow 0.

When rr is not too small and ρ\rho is not too large and when yy is small, yy and ψ\psi 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 𝒳μ,v,lk,α\mathcal{X}^{k,\alpha}_{\mu,v,l}, where α(0,1)\alpha\in(0,1), kk is a non-negative integer, μ,v,l\mu,v,l 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 BB is a ball in XX far away from any boundary/corner (the distance to any boundary/corner is at least 11). Then we define the Holder spaces Ck,α(B)C^{k,\alpha}(B) over BB, where kk is a non-negative integer and α(0,1)\alpha\in(0,1). This space is given by the norm:

uCk,α(B):=j=0kjuL(B)+[ku]C0,α(B),||u||_{C^{k,\alpha}(B)}:=\sum\limits_{j=0}^{k}||\nabla^{j}u||_{L^{\infty}(B)}+[\nabla^{k}u]_{C^{0,\alpha}(B)},

where

[u]C0,α(B):=supp,qB,pq|u(p)u(q)||pq|α.[u]_{C^{0,\alpha}(B)}:=\sup\limits_{p,q\in{B},p\neq q}\dfrac{|u(p)-u(q)|}{|p-q|^{\alpha}}.

In a region far away from type II (ψ=0\psi=0) and type III (r=0r=0) boundaries, we take the supreme of all balls of radius 11 of the above norm.

The operator that we want to study is Ls,tL_{s,t} which is introduced in subsection 3.5, where 0<t10<t\leq 1. 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 []Cie0,α[\cdot]_{C^{0,\alpha}_{\text{ie}}}.

We take the boundary r=0r=0 as an example. One way to think about the modification is to re-define the distance between two points pp and qq near the boundary to make it re-scaling invariant under a re-scaling rλrr\rightarrow\lambda r. This is done by modifying the metric near the boundary. In a direction that is perpendicular with the ψ\partial_{\psi} direction, there is nothing need to be changed. But in the r\partial_{r} direction, the metric should be 1r2dr2\dfrac{1}{r^{2}}dr^{2} instead of dr2dr^{2}. Thus the distance between a point at r1r_{1} and a point at r2r_{2} (with all other perpendicular coordinates the same) is given by

|r1r21r𝑑r|=|ln(r1)ln(r2)|,|\int_{r_{1}}^{r_{2}}\dfrac{1}{r}dr|=|\ln(r_{1})-\ln(r_{2})|,

which is clearly re-scaling invariant.

Another equivalent way to do the modification is to define it on each ball BB of radius r02\dfrac{r_{0}}{2} whose center has an rr-value r0r_{0}. On this ball, the []Cie0,α[\cdot]_{C_{\text{ie}}^{0,\alpha}} of uu should be given by

supp,qB,pqr0α|u(p)u(q)||pq|α.\sup\limits_{p,q\in B,p\neq q}\dfrac{r_{0}^{\alpha}|u(p)-u(q)|}{|p-q|^{\alpha}}.

Note that the weight r0αr_{0}^{\alpha} 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 r0r_{0} near the boundary.)

When it comes to the ψ=0\psi=0 boundary but away from the corner (or equivalently, y=0y=0 when rr is not too small and |z||z| is not too large), things are slightly different. The operator LΨ+tL_{\Psi}+t (strictly speaking, its y2(LΨ+t)y^{2}(L_{\Psi}+t)) has leading order terms which are made from combination of compositions of yy,y1,y2y\partial_{y},y\partial_{1},y\partial_{2}. So the re-scaling should be made in both yy and zz direction. And the metric should be addapted to be 1y2(dy2+dx12+dx22)\dfrac{1}{y^{2}}(dy^{2}+dx_{1}^{2}+dx_{2}^{2}) near the boundary.

Similarly, an equivalent way is to define it on each ball BB of radius y02\dfrac{y_{0}}{2} centered at a point whose yy value is y0y_{0}. On this ball BB, the []Cie0,α[\cdot]_{C^{0,\alpha}_{\text{ie}}} of uu is given by

supp,qB,pqy0α|u(p)u(q)||pq|α.\sup\limits_{p,q\in B,p\neq q}\dfrac{y_{0}^{\alpha}|u(p)-u(q)|}{|p-q|^{\alpha}}.

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 ρ+\rho\rightarrow+\infty (type I boundary). Because the operator LΨ+tL_{\Psi}+t is not of the “degenerate elliptic of edge type” near this boundary.

For the higher Holder spaces, in the definition of Ciek,αC^{k,\alpha}_{\text{ie}}, near r=0r=0 boundary, we need to replace ru\partial_{r}u in u\nabla u by rrur\partial_{r}u. And near ψ=0\psi=0 boundary (or y=0y=0 boundary but away from r=0r=0 boundary), we need to replace u\nabla u by yuy\nabla u. This corresponds to the edge structure of the operator LΨ+tL_{\Psi}+t that we study. Nothing needs to be adjusted when ρ+\rho\rightarrow+\infty.

C.2 The weighted Holder spaces

In order to be suitable for the operator Ls,tL_{s,t} (occurred in subsection 3.5) to map between, we need to add wights near boundaries/corners of the aforementioned Holder spaces Ciek,αC^{k,\alpha}_{\text{ie}}. Here is the definition:

𝒳μ,v,δk,α:=ψμrvρδCiek,α={ψμrvρδu|uCiek,α}.\mathcal{X}^{k,\alpha}_{\mu,v,\delta}:=\psi^{\mu}r^{v}\rho^{\delta}C^{k,\alpha}_{\text{ie}}=\{\psi^{\mu}r^{v}\rho^{\delta}u~{}|u\in C^{k,\alpha}_{\text{ie}}\}.

Here are several remarks:

  • If uu is an element in 𝒳μ,v,δk,α\mathcal{X}^{k,\alpha}_{\mu,v,\delta} for some α(0,1)\alpha\in(0,1) and all positive integers kk, then uu is also in 𝒳μ,v,δk,α\mathcal{X}^{k,\alpha^{\prime}}_{\mu,v,\delta} for any other α(0,1)\alpha\in(0,1). If this is the case, then the concrete α\alpha doesn’t matter and we may simply use 𝐗μ,v,δk\mathbf{X}^{k}_{\mu,v,\delta} 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 v=μv=\mu, then ψμrμCiek,α\psi^{\mu}r^{\mu}C^{k,\alpha}_{\text{ie}} is actually the same space as yμC0k,αy^{\mu}C^{k,\alpha}_{0}, where we treat y=0y=0 as a single boundary (without the blow-ups at each r=0r=0 point like what we did in appendix B), and the subscript “0” means this is the ordinary edge type Holder norm as y0y\rightarrow 0, 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 BB is a ball of radius RR whose closure is in the interior of XX.

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 uu is a function in XX and λ\lambda is a real number. Let Br(x)B_{r}(x) be the ball of radius rRr\leq R centered at xBx\in B. Then the Morrey norm of uu is defined to be

uL2,λ(B)=(supBr(x)rλ(Br(x)B|u|2))12.||u||_{L^{2,\lambda}(B)}=(\sup\limits_{B_{r}(x)}r^{-\lambda}(\int_{B_{r}(x)\cap B}|u|^{2}))^{\frac{1}{2}}.

The Camponato semi-norm is

[u]2,λ(B)=(supBr(x)rλ(Br(x)B|uu¯r,x|2))12,[u]_{\mathcal{L}^{2,\lambda}(B)}=(\sup\limits_{B_{r}(x)}r^{-\lambda}(\int_{B_{r}(x)\cap B}|u-\bar{u}_{r,x}|^{2}))^{\frac{1}{2}},

where u¯r,x\bar{u}_{r,x} is the average of uu in the ball Br(x)B_{r}(x), that is to say

u¯r,x:=1Vol(Br(x)B)Br(x)Bu.\bar{u}_{r,x}:=\dfrac{1}{\text{Vol}(B_{r}(x)\cap B)}\int_{B_{r}(x)\cap B}u.

The Camponato norm is

u2,λ(B)=(1Vol(B)B|u|2)12+[u]2,λ(B).||u||_{\mathcal{L}^{2,\lambda}(B)}=(\dfrac{1}{\text{Vol}(B)}\int_{B}|u|^{2})^{\frac{1}{2}}+[u]_{\mathcal{L}^{2,\lambda}(B)}.

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 λ>0\lambda>0 is a real number. There is a constant C>0C>0 which depends on α\alpha. Suppose uu is a function in XX. Then

  • If 0<λ<10<\lambda<1, then [u]C0,λ(B)CuL2,1+2λ(B)[u]_{C^{0,\lambda}(B)}\leq C||\nabla u||_{L^{2,1+2\lambda}(B)}, where C0,λC^{0,\lambda} is the Holder (semi-)norm.

  • [u]2,λ+2(B)CuL2,λ(B)[u]_{\mathcal{L}^{2,\lambda+2}(B)}\leq C||\nabla u||_{L^{2,\lambda}(B)}.

  • If λ3\lambda\leq 3, then uL2,λ(B)Cu2,λ(B)||u||_{L^{2,\lambda}(B)}\leq C||u||_{\mathcal{L}^{2,\lambda}(B)}

Note that under a re-scaling, the two sides of all the inequalities scale in the same way. So the constant CC doesn’t depend on the radius of the ball.

There is another inequality which is standard in analysis

Lemma D.3.

(Lemma 10.3.1 in [7]) We use the same notations as in the definition D.1. Suppose uu is a smooth function on BB such that

Δu=0.\Delta u=0.

Suppose xx is the center of the ball BB. Then there is a constant CC such that for any 0<r1<r2<R0<r_{1}<r_{2}<R

Br1|uu¯x,r1|2C(r1r2)5Br2|uu¯x,r2|2.\int_{B_{r_{1}}}|u-\bar{u}_{x,r_{1}}|^{2}\leq C(\dfrac{r_{1}}{r_{2}})^{5}\int_{B_{r_{2}}}|u-\bar{u}_{x,r_{2}}|^{2}.

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 Σ\Sigma 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.