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

Generalized neck analysis of harmonic maps from surfaces

Hao Yin Hao Yin, School of Mathematical Sciences, University of Science and Technology of China, Hefei, China [email protected]
Abstract.

In this paper, we study the behavior of a sequence of harmonic maps from surfaces with uniformly bounded energy on the generalized neck domain. The generalized neck domain is a union of ghost bubbles and annular neck domains, which connects non-trivial bubbles. An upper bound of the energy density is proved and we use it to study the limit of the nullity and index of the sequence.

1. Introduction

Let (N,h)(N,h) be a closed Riemannian manifold isometrically embedded in p\mathbb{R}^{p} and BB be the unit ball of 2\mathbb{R}^{2}. We study a sequence of harmonic maps uiu_{i} from BB to NN satisfying

(U1) the energy B|ui|2𝑑x\int_{B}\left|\nabla u_{i}\right|^{2}dx is uniformly bounded;

(U2) for any r>0r>0, uiu_{i} converges smoothly to uu_{\infty} on BBrB\setminus B_{r} (where BrB_{r} is the ball of radius rr centered at the origin) and

(1) limr0limiBr|ui|2𝑑x>0.\lim_{r\to 0}\lim_{i\to\infty}\int_{B_{r}}\left|\nabla u_{i}\right|^{2}dx>0.

The energy concentration (as in (1)) leads to the existence of a sequence of pairs (xi,λi)(x_{i},\lambda_{i}) with xi0x_{i}\to 0 in BB and λi0\lambda_{i}\to 0 such that

vi(y)=ui(xi+λiy)v_{i}(y)=u_{i}(x_{i}+\lambda_{i}y)

converges to a (nontrivial) harmonic map ω\omega from 2\mathbb{R}^{2} to NN, which is known as a bubble. It is well known that there may be more than one bubbles developing at one concentration point.

These bubbles, according to their positions xix_{i} and scales λi\lambda_{i}, are organized in the form of a tree. There are several expositions about the construction of the bubble tree in the literature (see [DT95, Par96]). For technical reasons, a special type of bubbles, known as ghost bubbles (in the sense that they carry no energy in the limit), is introduced as connectors in the bubble tree. We refer to Section 2 for the exact formulation.

An edge in the tree represents an annular domain which is known as the neck. The ratio between the outer radius and the inner radius of the neck goes to infinity. While we know the limit of the sequence of scaled maps, the study of uiu_{i} in the neck domain is less obvious and known as the neck analysis. The energy identity theorem and the no neck theorem imply that the energy and the oscillation of uiu_{i} vanish in the neck domain. Indeed, a decay of the gradient of uiu_{i} (regarded as a map on cylinder due to the conformal invariance of the problem) was proved. Recently, the author [Yin19] proved some higher order estimate for uiu_{i} in the neck domain which allows us to obtain a normal form in the center piece of the neck.

It is well possible that the neck above is connected to a ghost bubble. It is worth emphasizing that a ghost bubble is not a real one and it serves the same purpose of connecting real bubbles (or the weak limit map uu_{\infty}) as the necks. It is natural to pursue a deeper understanding of uiu_{i} on the ghost bubble than the mere vanishing of energy (by its definition). This is the main topic of this paper. As a first step, we obtain an upper bound of |ui|\left|\nabla u_{i}\right|.

For a precise formulation of our main result, we need to introduce the generalized neck domain, on which our upper bound of |ui|\left|\nabla u_{i}\right| holds. It is helpful to keep the following simple case in mind, which we illustrate in the following figure. It involves a ghost bubble on top of which two real ones sit.

Refer to caption
Figure 1. A simple bubble tree with ghost bubble

The picture on the right shows a disk with two smaller ones removed. It is the simplest example of a so called generalized neck domain. Its image (as shown on the left) consists of a ghost bubble and three necks that are connected to it.

In general, let’s fix the sequence uiu_{i}. For some parameter δ>0\delta>0 and ll (real) bubbles given by (yi(j),λi(j))(y_{i}^{(j)},\lambda_{i}^{(j)}) with j=1,,lj=1,\cdots,l, the generalized neck domain is (up to scaling and translation)

Ωi=B(0,δ)j=1,,lB(yi(j),δ1λi(j)).\Omega_{i}=B(0,\delta)\setminus\bigcup_{j=1,\cdots,l}B(y_{i}^{(j)},\delta^{-1}\lambda_{i}^{(j)}).

Moreover, we assume that

(O1) the barycenter of the bubbles is the origin in the sense that

0=1lj=1lyi(j);0=\frac{1}{l}\sum_{j=1}^{l}y_{i}^{(j)};

(O2) the bubbles (yi(j),λi(j))(y_{i}^{(j)},\lambda_{i}^{(j)}) for j=1,,lj=1,\cdots,l are disjoint in the sense that

B(yi(j1),Rλi(j1))B(yi(j2),Rλi(j2))=B(y_{i}^{(j_{1})},R\lambda_{i}^{(j_{1})})\cap B(y_{i}^{(j_{2})},R\lambda_{i}^{(j_{2})})=\emptyset

for any R>0R>0, any j1j2j_{1}\neq j_{2} and sufficiently large ii (depending on RR, j1j_{1} and j2j_{2});

(O3) there is some ε0\varepsilon_{0} depending only on NN such that for sufficiently large ii,

Ωi|ui|2𝑑x<ε0.\int_{\Omega_{i}}\left|\nabla u_{i}\right|^{2}dx<\varepsilon_{0}.
Remark 1.1.

(1) These assumptions arise naturally in the construction of bubble tree. This is going to be clear in Section 2.

(2) The assumption (O3) above is a consequence of the energy identity theorem and the definition of the ghost bubble.

(3) In the construction in Section 2, the generalized neck domain is going to be a translation and a scaling of the Ωi\Omega_{i} defined above. By the nature of the problem, this does not matter.

The main result of this paper is

Theorem 1.2.

Suppose that uiu_{i} is a sequence of harmonic maps satisfying (U1) and (U2) and that Ωi\Omega_{i} is a generalized neck domain defined above, then there is some constant CC such that

|ui|g~iConΩi\left|\nabla u_{i}\right|_{\tilde{g}_{i}}\leq C\qquad\text{on}\quad\Omega_{i}

where g~i\tilde{g}_{i} is a conformal metric defined in terms of complex coordinate zz by

(2) g~i:=(1+j=1l(λi(j))2|zxi(j)|4)dzdz¯onΩi.\tilde{g}_{i}:=\left(1+\sum_{j=1}^{l}\frac{(\lambda_{i}^{(j)})^{2}}{\left|z-x_{i}^{(j)}\right|^{4}}\right)dz\wedge d\bar{z}\qquad\text{on}\quad\Omega_{i}.
Remark 1.3.

The constant CC in the above theorem depends not only on NN, but also on the particular sequence. This shall be clear in the proofs. We only remark that even if CC is universal, this upper bound depends on the sequence in the sense that the geometry of Ωi\Omega_{i} (hence g~i\tilde{g}_{i}) depends on the relative position and size of the bubbles. It is also in this sense that CC depends on the sequence and it is inevitable, if one considers a family of such sequences which brings the tree structure to a sudden change.

The upper bound of u\nabla u is measured with respect to the metric g~i\tilde{g}_{i} (see (2)) defined on the multi-connected domain Ωi\Omega_{i}. This metric is related to a specific sequence of uiu_{i} presented in Section 2.4. On one hand, we feel obliged to give explicit examples to assure the readers that complicated patterns of ghost bubble domain do occur. On the other hand, the maps uiu_{i} in this example are holomorphic curves (hence minimal surfaces). As if by coincidence, g~i\tilde{g}_{i} is the pullback metric of this family of uiu_{i}. Moreover, this feature of being induced metric of minimal surfaces will be useful in an application, which will be explained in a minute.

There is another way of understanding the upper bound. For that purpose, we need a different conformal metric g¯i\bar{g}_{i}. With this metric, the necks are long cylinders of radius 11 and the ghost bubbles are multi-way connectors of uniformly bounded geometry. For any point xΩix\in\Omega_{i}, let d(x)d(x) be the distance from xx to the boundary to Ωi\Omega_{i} w.r.t g¯i\bar{g}_{i}. Then the upper bound is equivalently formulated by

|ui|g¯iCed(x).\left|\nabla u_{i}\right|_{\bar{g}_{i}}\leq Ce^{-d(x)}.

Indeed, the proof of Theorem 1.2 relies on this equivalent formulation (see Theorem 4.2). For the proof, we combine known techniques with some new estimate. The known ones include the three circle lemma, the ordinary differential inequality and the sharp decay estimate (see Rade [Rad93]). The new estimate generalizes the three circle lemma (that works for annular domain) to the case of multi-connected domain (i.e. a disk with more than 22 smaller ones removed).

As an application, we improve a result in [Yin19]. For a harmonic map uu from a closed Riemannian surface Σ\Sigma to NN, let JuJ_{u} be the linearization of the tension field operator τ(u)\tau(u), Nul(u)Nul(u) be the dimension of its kernel and NI(u)NI(u) be the number of nonpositive eigenvalues of JuJ_{u} (counting multiplicity). A semi-continuity property of NINI was proved by studying the limit of eigenfunctions of JuiJ_{u_{i}}.

Theorem 1.4 (Theorem 1.6 of [Yin19]).

Let uiu_{i} be a sequence of harmonic maps from a closed Riemannian surface Σ\Sigma to a closed Riemannian manifold NN. Let uu_{\infty} be the weak limit and ω1,,ωl\omega_{1},\cdots,\omega_{l} be all the bubbles (ghost bubbles included) in the bubble tree. Then

(3) lim supiNI(ui)NI(u)+k=1lNI(ωj)\limsup_{i\to\infty}NI(u_{i})\leq NI(u_{\infty})+\sum_{k=1}^{l}NI(\omega_{j})

and

(4) lim supiNul(ui)Nul(u)+k=1lNul(ωj).\limsup_{i\to\infty}Nul(u_{i})\leq Nul(u_{\infty})+\sum_{k=1}^{l}Nul(\omega_{j}).

Here in the definition of NulNul and NINI of ωk\omega_{k}, we regard ωk\omega_{k} as a harmonic map from S2S^{2} to NN.

The proof of Theorem 1.4 was based on an upper bound of |ui|\left|\nabla u_{i}\right| on the (annular) neck domain. If ωk\omega_{k} is a ghost bubble (i.e. constant map), then

Nul(ωk)=NI(ωk)=dimN.Nul(\omega_{k})=NI(\omega_{k})=\dim N.

The possible existence of ghost bubbles weakens the result of Theorem 1.4. As an application of Theorem 1.2, we are able to show

Theorem 1.5.

Under the same assumptions as in Theorem 1.4, (3) and (4) hold for ω1,,ωl\omega_{1},\cdots,\omega_{l} being the real bubbles in the bubble tree.

The rest of the paper is organized as follows. In Section 2, we recall the construction of bubble tree and define the generalized neck domain. We also define the metric g¯i\bar{g}_{i} that is going to be used in Section 4. At the end of Section 2, we introduce the metric g~i\tilde{g}_{i} in Theorem 1.2 as the pullback metric of a specific sequence of uiu_{i}. In Section 3, we prove key lemmas that will be used in Section 4, where Theorem 1.2 is proved. In the final section, we discuss the application and prove Theorem 1.5.

Acknowledgement

The author thanks Professor Yuxiang Li for numerous discussions on conformal immersion and its relation to harmonic maps. This research is supported by NSFC11971451.

2. The generalized neck domain

In the literature, there are several different ways to construct the bubble tree(see [Par96, DT95]). In this section, we present the argument in two steps. We first obtain the set of (real) bubbles by an abstract maximizing argument. Given the set of real bubbles, we can then decompose the domain BB into the union of some bubble domains (one for each real bubble) and some generalized neck domains. By adding ghost bubbles, we show how the generalized neck domain is further decomposed into the union of ghost bubble domains and neck domains (i.e. annulus). Along with the decomposition, we define the metric g¯i\bar{g}_{i}.

Remark 2.1.

The material presented in this section is a re-formulation of very well known facts. Hence, the verification of elementary properties is left to the readers.

2.1. The tree of real bubbles

Definition 2.2.

A (real) bubble \mathcal{B} is a sequence of pairs (xi,λi)(x_{i},\lambda_{i}) such that the rescaled maps vi(x)=ui(xi+λix)v_{i}(x)=u_{i}(x_{i}+\lambda_{i}x) converge weakly in W1,2W^{1,2} to some nontrivial harmonic map ω\omega from 2\mathbb{R}^{2} to NN.

Intuitively, the word ’bubble’ may refer to the image of the limit map ω\omega. However, Definition 2.2 is good for technical reasons. It dictates a region (roughly, B(xi,Rλi)B(x_{i},R\lambda_{i}) for large RR) in the domain such that the maps restricted to this region converge to ω\omega. Very often, it is this region that matters.

Notice that different sequences may give the same region and lead to the same limit map ω\omega (up to reparametrization). Hence, any two sequences of pairs, (xi,λi)(x_{i},\lambda_{i}) and (yi,σi)(y_{i},\sigma_{i}) are said to be equivalent if and only if there is c>0c>0 such that for all ii,

cλiσi1cand|xiyi|1cλi.c\leq\frac{\lambda_{i}}{\sigma_{i}}\leq\frac{1}{c}\qquad\text{and}\qquad\left|x_{i}-y_{i}\right|\leq\frac{1}{c}\lambda_{i}.
Remark 2.3.

Rigorously speaking, a bubble should be defined as the equivalence class of the sequence of pairs (xi,λi)(x_{i},\lambda_{i}). However, for simplicity, we simply agree that equivalent sequences define the same bubble.

It follows from the gap theorem and the total energy bound that (by passing to a subsequence) there exists a unique set of bubbles 𝒯\mathcal{T} which is maximal in the sense that one can not add another (not equivalent) real bubble. This is a consequence of finite induction.

For convenience, we add a trivial sequence (0,1)(0,1) (xi=0x_{i}=0 and λi=1\lambda_{i}=1) to 𝒯\mathcal{T}. This ’bubble’ represents the weak limit of uiu_{i} and it is going to be the root of the bubble tree.

According to the size and position of the bubbles, we define a partial order which yields the tree structure in the set of real bubbles 𝒯\mathcal{T}.

Definition 2.4.

(1)A bubble 1=(xi,λi)\mathcal{B}_{1}=(x_{i},\lambda_{i}) is said to be on top of another bubble 2=(yi,σi)\mathcal{B}_{2}=(y_{i},\sigma_{i}) if and only if there is some constant c>0c>0 such that

limiλiσi=0and|xiyi|cσi.\lim_{i\to\infty}\frac{\lambda_{i}}{\sigma_{i}}=0\qquad\text{and}\qquad\left|x_{i}-y_{i}\right|\leq c\sigma_{i}.

(2) 1\mathcal{B}_{1} is said to be directly on top of 2\mathcal{B}_{2} if (i) 1\mathcal{B}_{1} is on top of 2\mathcal{B}_{2} and (ii) there is no other 𝒯\mathcal{B}\in\mathcal{T} satisfying 1\mathcal{B}_{1} is on top of \mathcal{B} and \mathcal{B} is on top of B2B_{2}.

By taking 𝒯\mathcal{T} as the set of vertices and taking the set of pairs (1,2)(\mathcal{B}_{1},\mathcal{B}_{2}) satisfying 1\mathcal{B}_{1} is directly on top of 2\mathcal{B}_{2} as the set of edges, we define a graph, which is obviously a tree, and which we also denote by 𝒯\mathcal{T} for simplicity. This is the tree of real bubbles.

2.2. The generalized neck domain

Given the bubble tree 𝒯\mathcal{T} above, we decompose the domain BB into the union of bubble domains and generalized neck domains. The decomposition depends on a parameter δ\delta, which is a small positive number depending on uiu_{i} and will be chosen in the constructure below.

For each bubble =(xi,λi)𝒯\mathcal{B}=(x_{i},\lambda_{i})\in\mathcal{T}, suppose that there are ll bubbles 1,,l\mathcal{B}_{1},\cdots,\mathcal{B}_{l} that are directly on top of \mathcal{B}. For j=1,,lj=1,\cdots,l, let j\mathcal{B}_{j} be represented by the sequence (yi(j),λi(j))(y_{i}^{(j)},\lambda_{i}^{(j)}). The concentration set is defined by

𝒞:={limiyi(j)xiλi|j=1,,l}.\mathcal{C}:=\left\{\lim_{i\to\infty}\frac{y_{i}^{(j)}-x_{i}}{\lambda_{i}}|\quad j=1,\cdots,l\right\}.

Notice that the number of elements in the concentration set may be strictly smaller than ll. Keeping Remark 2.3 in mind, we may assume that for all x𝒞x\in\mathcal{C}, we have |x|1\left|x\right|\leq 1 by choosing a larger λi\lambda_{i}.

For a fixed concentration point x𝒞x\in\mathcal{C}, assume that there are lxl_{x} bubbles (among 1\mathcal{B}_{1},…,l\mathcal{B}_{l}), say 1,,lx\mathcal{B}_{1},\cdots,\mathcal{B}_{l_{x}}, satisfying

limiyi(j)xiλi=x,j=1,,lx.\lim_{i\to\infty}\frac{y_{i}^{(j)}-x_{i}}{\lambda_{i}}=x,\qquad j=1,\cdots,l_{x}.

These 1,,lx\mathcal{B}_{1},\cdots,\mathcal{B}_{l_{x}} are said to be concentrated at xx.

At each concentration point x𝒞x\in\mathcal{C}, we define a center of mass position

ci=1lxj=1lxyi(j).c_{i}=\frac{1}{l_{x}}\sum_{j=1}^{l_{x}}y_{i}^{(j)}.

Obviously, limicixiλi=x\lim_{i\to\infty}\frac{c_{i}-x_{i}}{\lambda_{i}}=x. We write ci(x)c_{i}(x) if we need to emphasize its dependence on x𝒞x\in\mathcal{C}.

Definition 2.5.

The bubble domain of \mathcal{B} is

Ω=B(xi,δ1λi)x𝒞B(ci(x),δλi).\Omega_{\mathcal{B}}=B(x_{i},\delta^{-1}\lambda_{i})\setminus\bigcup_{x\in\mathcal{C}}B(c_{i}(x),\delta\lambda_{i}).

Here we assume that δ\delta is small so that the balls B(ci(x),δλi)B(c_{i}(x),\delta\lambda_{i}) for x𝒞x\in\mathcal{C} are disjoint. In case that 𝒞\mathcal{C} is empty set, namely, there is no bubble on top of \mathcal{B}, the bubble domain is just B(xi,δ1λi)B(x_{i},\delta^{-1}\lambda_{i}).

In case that 𝒞\mathcal{C} is not empty, we define

Definition 2.6.

The generalized neck domain at xx is

Ω,x=B(ci,δλi)j=1,,lxB(yi(j),δ1λi(j)).\Omega_{\mathcal{B},x}=B(c_{i},\delta\lambda_{i})\setminus\bigcup_{j=1,\cdots,l_{x}}B(y_{i}^{(j)},\delta^{-1}\lambda_{i}^{(j)}).

Here we omit the routine verification that the balls B(yi(j),δ1λi(j))B(y_{i}^{(j)},\delta^{-1}\lambda_{i}^{(j)}) are disjoint and contained in B(ci,δλi)B(c_{i},\delta\lambda_{i}) when ii is sufficiently large.

The topology of generalized neck domain is a disk with finitely many smaller ones removed. Notice that we have no control over the size and position of these removed disks. It is the main task of this paper to study the behavior of uiu_{i} in this domain.

2.3. Decomposition of generalized neck domain by adding ghost bubbles

Let xx be a concentration point of the bubble \mathcal{B} as in the previous subsection. When lx=1l_{x}=1, we have ci=yi(1)c_{i}=y_{i}^{(1)} and the generalized neck domain Ω,x\Omega_{\mathcal{B},x} takes the particular simple form

B(yi(1),δλi)B(yi(1),δ1λi(1)).B(y_{i}^{(1)},\delta\lambda_{i})\setminus B(y_{i}^{(1)},\delta^{-1}\lambda_{i}^{(1)}).

Such annulus type domain is called simple neck domain and known results on the neck analysis apply to this type of neck domain. In this simple case, we define

(5) w(z)=1|zci|2onB(ci,δλi)B(ci,δ1λi(1)).w(z)=\frac{1}{\left|z-c_{i}\right|^{2}}\qquad\text{on}\quad B(c_{i},\delta\lambda_{i})\setminus B(c_{i},\delta^{-1}\lambda_{i}^{(1)}).

This function ww is going to be used as a conformal factor in the definition of g¯i\bar{g}_{i}.

When lx>1l_{x}>1, we describe below an induction process, which by adding some more ghost bubbles, further decompose the generalized neck domain into the union of simple neck domains and (small) ghost bubble domains.

By taking subsequence, we may assume that

(6) σi:=2maxj1,j2=1,,lx|yi(j1)yi(j2)|=|yi(1)yi(2)|>0.\sigma_{i}:=2\max_{j_{1},j_{2}=1,\cdots,l_{x}}\left|y_{i}^{(j_{1})}-y_{i}^{(j_{2})}\right|=\left|y_{i}^{(1)}-y_{i}^{(2)}\right|>0.

The sequence (ci,σi)(c_{i},\sigma_{i}) (up to equivalence as in Remark 2.3) represents a ghost bubble ~\tilde{\mathcal{B}}. Due to the maximality of 𝒯\mathcal{T}, one can check that the limit of

vi(y)=ui(ci+yσi)v_{i}(y)=u_{i}(c_{i}+y\sigma_{i})

is constant map. Indeed, we have

(1) (ci,σi)(c_{i},\sigma_{i}) is not equivalent to any bubble in 𝒯\mathcal{T};

(2) B(ci,δλi)B(ci,2σi)B(c_{i},\delta\lambda_{i})\setminus B(c_{i},2\sigma_{i}) is a simple neck domain, on which we define

(7) w(z)=1|zci|2;w(z)=\frac{1}{\left|z-c_{i}\right|^{2}};

(3) 1,,lx{\mathcal{B}}_{1},\cdots,{\mathcal{B}}_{l_{x}} are directly on top of ~\tilde{\mathcal{B}};

(4) the concentration set

𝒞~={limiyi(j)ciσi|j=1,,lx}\tilde{\mathcal{C}}=\left\{\lim_{i\to\infty}\frac{y_{i}^{(j)}-c_{i}}{\sigma_{i}}|\,j=1,\cdots,l_{x}\right\}

contains at least two points and is a subset of B(0,1/2)B(0,1/2) (see (6)).

By choosing δ\delta small, we may assume that the minimal distance between any two points in 𝒞~\tilde{\mathcal{C}} is larger than 3δ3\delta. For each y𝒞~y\in\tilde{\mathcal{C}}, define the center of mass (as before)

ci(y)=1lyj=1lyyi(αj)c_{i}(y)=\frac{1}{l_{y}}\sum_{j=1}^{l_{y}}y_{i}^{(\alpha_{j})}

where (yi(αj),λi(αj))(j=1,,ly)(y_{i}^{(\alpha_{j})},\lambda_{i}^{(\alpha_{j})})(j=1,\cdots,l_{y}) are a choice of lyl_{y} bubbles among 1,,lx\mathcal{B}_{1},\cdots,\mathcal{B}_{l_{x}}.

Definition 2.7.

The ghost bubble domain is defined to be

B(ci,2σi)y𝒞~B(ci(y),δσi).B(c_{i},2\sigma_{i})\setminus\bigcup_{y\in\tilde{\mathcal{C}}}B(c_{i}(y),\delta\sigma_{i}).

On the above ghost bubble domain, we choose ww to be any smooth functions satisfying

(W1) there is some constant C>0C>0 such that

σi2Cw(z)Cσi2;\frac{\sigma_{i}^{2}}{C}\leq w(z)\leq C\sigma_{i}^{2};

(W2) ww is 1|zci|2\frac{1}{\left|z-c_{i}\right|^{2}} in a neighborhood of B(ci,2σi)\partial B(c_{i},2\sigma_{i});

(W2) ww is 1|zci(y)|2\frac{1}{\left|z-c_{i}(y)\right|^{2}} in a neighborhood of B(ci(y),δσi)\partial B(c_{i}(y),\delta\sigma_{i}).

For each y𝒞~y\in\tilde{\mathcal{C}}, it is a concentration point on the ghost bubble (ci,σi)(c_{i},\sigma_{i}). We repeat the construction above. Notice that the total number of real bubbles (directly on top) concentrated at yy becomes strictly smaller than lxl_{x}. Hence the induction stops after finitely many steps.

We conclude this subsection by setting

(8) g¯i=w(z)dzdz¯,onΩ,x\bar{g}_{i}=w(z)dz\wedge d\bar{z},\qquad\text{on}\quad\Omega_{\mathcal{B},x}

where ww is defined by (5), (7) and (W1-W3).

2.4. An example of bubble tree

This short section consists of two parts. The first part is an example which demonstrates that ghost bubbles and very complicated generalized neck domains do occur. The second part shows that the metric in (2) is the pullback of some holomorphic maps into n\mathbb{C}^{n}. This not only helps the understanding of Theorem 1.2, but also plays a role in the proof of Theorem 1.5.

Since this paper deals with ghost bubbles, it is natural to ask whether there exists a sequence of uiu_{i} as in Theorem 1.2 that the construction in the previous subsections leads to a ghost bubble. Further more, is there a generalized neck domain as constructed above such that the punctured disks shrink and approach each other at arbitrary speed?

Indeed, the following example shows that one can prescribe the number of bubbles, the position and the scale of each (real) bubble, so that the argument in Section 2.3 gives a generalized neck domain as complicated as one needs.

Precisely, let ll be the number of (real) bubbles. Let (xi(j),λi(j))(x_{i}^{(j)},\lambda_{i}^{(j)}) be the center and the scale of the jj-th bubble. In terms of the holomorphic coordinate zz on \mathbb{C}, we define a sequence of maps u~i\tilde{u}_{i} from BB to l+1\mathbb{C}^{l+1} to be

(9) u~i(z)=(z,λi(1)zxi(1),,λi(l)zxi(l)).\tilde{u}_{i}(z)=\left(z,\frac{\lambda_{i}^{(1)}}{z-x_{i}^{(1)}},\cdots,\frac{\lambda_{i}^{(l)}}{z-x_{i}^{(l)}}\right).

Using the homogeneous coordinates [z1,z2][z_{1},z_{2}] of P1\mathbb{C}P^{1}, we may regard \mathbb{C} as an open subset of P1\mathbb{C}P^{1}, via the identification,

z[z,1].z\mapsto[z,1].

Similarly, l+1\mathbb{C}^{l+1} is identified with an open subset of Pl+1\mathbb{C}P^{l+1} via

(z1,,zl+1)[z1,,zl+1,1].(z_{1},\cdots,z_{l+1})\mapsto[z_{1},\cdots,z_{l+1},1].

Hence u~i\tilde{u}_{i} can be extended in a unique way as a map uiu_{i} from P1\mathbb{C}P^{1}(S2S^{2}) to Pl+1\mathbb{C}P^{l+1} and the map uiu_{i} is holomorphic (hence harmonic). It is elementary to check that there are exactly ll bubbles occurring at the prescribed position and rate.

Next, we assume that

(E1) the ll bubbles concentrate at 0, i.e.

limixi(j)=0,j=1,,l.\lim_{i\to\infty}x_{i}^{(j)}=0,\qquad j=1,\cdots,l.

(E2) the center of mass is 0, i.e.

j=1lxi(j)=0.\sum_{j=1}^{l}x_{i}^{(j)}=0.

(E3) the ll bubbles are separated from each other, in other words, no one is on top of another. That is, for any R>0R>0 and j1j2j_{1}\neq j_{2},

B(xi(j1),Rλij1)B(xi(j2),Rλi(j2))=B(x_{i}^{(j_{1})},R\lambda_{i}^{j_{1}})\cap B(x_{i}^{(j_{2})},R\lambda_{i}^{(j_{2})})=\emptyset

for sufficiently large ii.

The generalized neck domain given in Definition 2.6 is

Ω=B(0,δ)j=1,,lB(xi(j),δ1λi(j)).\Omega=B(0,\delta)\setminus\bigcup_{j=1,\ldots,l}B(x_{i}^{(j)},\delta^{-1}\lambda_{i}^{(j)}).

For the u~i\tilde{u}_{i} in (9), we regard it as a holomorphic map from Ω\Omega to l+1\mathbb{C}^{l+1} and the pullback metric is

(10) g~i:=(1+j=1,,l(λi(j))2|zxi(j)|4)dzdz¯.\tilde{g}_{i}:=\left(1+\sum_{j=1,\ldots,l}\frac{(\lambda_{i}^{(j)})^{2}}{\left|z-x_{i}^{(j)}\right|^{4}}\right)dz\wedge d\bar{z}.

To conclude this section, we remark that u~i\tilde{u}_{i} parametrizes an embedded minimal surface in l+1\mathbb{C}^{l+1}. The surface has l+1l+1 ends and its tangent cone at the infinity is the union of l+1l+1 coordinate planes of l+1\mathbb{C}^{l+1}.

3. Various energy decay estimates

In this section, we prove two estimates about the decay/growth of energy density on different domains. The first one is a generalization of the well known three circle lemma. The second one is a sharp growth estimate on long cylinder.

Before we start, we note the following convention about the notations. We will use c~\tilde{c} for universal constants, cc for constants that depend on the target manifold NN, and CC for constants that depend both on NN and the particular sequence of maps uiu_{i}. In general, these constants may vary from line to line. However, subscripts will be added, if it is necessary to note the distinction between them.

Moreover, throughout this section, there is a small constant ε1\varepsilon_{1} appearing in the assumptions of the following results. We remark that it depends only on NN, not on the sequence uiu_{i}.

3.1. Generalized three circle estimate

The application of the three circle estimate to the study of harmonic maps has a long tradition(see [Sim83, QT97, LY16, AY17]). We start by recalling the following well known result.

Lemma 3.1.

There is some ε1(N)>0\varepsilon_{1}(N)>0. For any β>0\beta>0, there is L0>1L_{0}>1 (depending only on β\beta) such the following is true for any L>L0L>L_{0}. Assume that uu is a harmonic map defined on [0,3L]×S1[0,3L]\times S^{1} satisfying

(11) [t,t+1]×S1|u|2ε1,t[0,3L1]\int_{[t,t+1]\times S^{1}}\left|\nabla u\right|^{2}\leq\varepsilon_{1},\qquad\forall t\in[0,3L-1]

and

(12) {t}×S1|θu|2|tu|2=0,t[0,3L].\int_{\left\{t\right\}\times S^{1}}\left|\partial_{\theta}u\right|^{2}-\left|\partial_{t}u\right|^{2}=0,\qquad\forall t\in[0,3L].

Then

(13) [L,2L]×S1|u|2β([0,L]×S1+[L,3L]×S1)|u|2.\int_{[L,2L]\times S^{1}}\left|\nabla u\right|^{2}\leq\beta\left(\int_{[0,L]\times S^{1}}+\int_{[L,3L]\times S^{1}}\right)\left|\nabla u\right|^{2}.

Lemma 3.1 compares the energy on a piece of cylinder with the energy on adjacent pieces of the same length and concludes that (when the assumptions hold) at least one of the two pieces have (significantly) larger energy.

In this section, we prove a generalization of this fact. We compare the energy on a ghost bubble domain with the energy on cylinders that are directly connected to it. To be precise, we recall that for a sequence of harmonic maps uiu_{i}, the ghost bubble domain in Definition 2.7 is

B(yi,2σi)y𝒞~B(ci(y),δσi),B(y_{i},2\sigma_{i})\setminus\bigcup_{y\in\tilde{\mathcal{C}}}B(c_{i}(y),\delta\sigma_{i}),

where 𝒞~\tilde{\mathcal{C}} is the energy concentration set of the scaled sequence vi(y)=ui(yi+yσi)v_{i}(y)=u_{i}(y_{i}+y\sigma_{i}). Since the problem under investigation is scaling invariant, we may assume that the ghost bubble domain is

B(yi,2)y𝒞~B(ci(y),δ),B(y_{i},2)\setminus\bigcup_{y\in\tilde{\mathcal{C}}}B(c_{i}(y),\delta),

and

limiyi=0.\lim_{i\to\infty}y_{i}=0.

This domain varies with ii. However, the number of points in 𝒞\mathcal{C} is bounded, their distances to the origin are bounded and the distance between any pair is bounded from below. Hence, by passing to a subsequence if necessary, the ghost bubble domain approaches a limit,

Ω0=B(0,2)y𝒞~B(y,δ).\Omega_{0}=B(0,2)\setminus\bigcup_{y\in\tilde{\mathcal{C}}}B(y,\delta).

For some small η>0\eta>0 to be determined later, we set

(14) Ωj=B(0,2ηj)y𝒞~B(y,δηj).\Omega_{j}=B(0,2\eta^{-j})\setminus\bigcup_{y\in\tilde{\mathcal{C}}}B(y,\delta\eta^{j}).

In what follows, we prove estimates for the energy of harmonic map uu defined on Ωj\Omega_{j} (see Lemma 3.2 and Lemma 3.4). The constants appeared there depend on Ω0\Omega_{0}, or more precisely, depend on 𝒞~\tilde{\mathcal{C}} and δ\delta, which in turn depend on the particular sequence. These estimates hold for the original sequence uiu_{i} for sufficiently large ii with the same set of constants. It is in this sense that we say the estimates depend on the sequence uiu_{i}.

The following is a set of natural assumptions under which our estimates hold and they are verified easily in the construction of bubble tree.

(S1)

Ω0|u|2ε1;\int_{\Omega_{0}}\left|\nabla u\right|^{2}\leq\varepsilon_{1};

(S2) for any y𝒞~y\in\tilde{\mathcal{C}} and ρ(δη3,δ/2)\rho\in(\delta\eta^{3},\delta/2),

B(y,2ρ)B(y,ρ)|u|2ε1;\int_{B(y,2\rho)\setminus B(y,\rho)}\left|\nabla u\right|^{2}\leq\varepsilon_{1};

(S3) for any η3ρ2\eta^{-3}\geq\rho\geq 2,

B(0,2ρ)B(0,ρ)|u|2ε1.\int_{B(0,2\rho)\setminus B(0,\rho)}\left|\nabla u\right|^{2}\leq\varepsilon_{1}.

Our first result is the following lemma.

Lemma 3.2.

Let uu be a harmonic map defined on Ω3\Omega_{3} satisfying (S1)-(S3) for some small ε1\varepsilon_{1} depending only on NN. There is a constant C1C_{1} depending on NN and Ω0\Omega_{0} but not on η\eta such that

Ω1|u|2C1Ω2Ω1|u|2.\int_{\Omega_{1}}\left|\nabla u\right|^{2}\leq C_{1}\int_{\Omega_{2}\setminus\Omega_{1}}\left|\nabla u\right|^{2}.
Proof.

Assume that the lemma is false. Then, there is a sequence of ηi>0\eta_{i}>0 and a sequence of harmonic maps uiu_{i} satisfying (S1)-(S3) such that

(15) Ω1(ηi)|ui|2iΩ2(ηi)Ω1(ηi)|ui|2.\int_{\Omega_{1}(\eta_{i})}\left|\nabla u_{i}\right|^{2}\geq i\int_{\Omega_{2}(\eta_{i})\setminus\Omega_{1}(\eta_{i})}\left|\nabla u_{i}\right|^{2}.
Remark 3.3.

(1) We have used the notation Ω1(ηi)\Omega_{1}(\eta_{i}) and Ω2(ηi)\Omega_{2}(\eta_{i}) to emphasize the dependence on ηi\eta_{i}. When this dependence is clear from the context, we simply write Ω1\Omega_{1} and Ω2\Omega_{2}.

(2) The sequence uiu_{i} is not the sequence in the main theorem. We recycle the notation for simplicity and this usage is valid only in this proof.

To get a contradiction, we distinguish two cases.

Case 1: lim infiηi>0\liminf_{i\to\infty}\eta_{i}>0. By passing to a subsequence, we assume that ηiη\eta_{i}\to\eta.

The ε\varepsilon-regularity theorem of harmonic maps and (S1)-(S3) together imply the existence of a smooth limit uu_{\infty} of uiu_{i} defined on Ω2(η)\Omega_{2}(\eta), which is also a harmonic map. If Ω1(ηi)|ui|2\int_{\Omega_{1}(\eta_{i})}\left|\nabla u_{i}\right|^{2} has a positive lower bound, uu_{\infty} is nontrivial. However, (15) implies that uu_{\infty} is constant map on Ω2Ω1\Omega_{2}\setminus\Omega_{1}. This is a contradiction to the unique continuation theorem([Sam78]).

If Ω1|ui|20\int_{\Omega_{1}}\left|\nabla u_{i}\right|^{2}\to 0, then we scale the ambient space p\mathbb{R}^{p} in which NN is embedded and set

(16) u~i=ε11/2ui(Ω1|ui|2)1/2.\tilde{u}_{i}=\frac{\varepsilon_{1}^{1/2}u_{i}}{\left(\int_{\Omega_{1}}\left|\nabla u_{i}\right|^{2}\right)^{1/2}}.

After the scaling, we have

(17) Ω1|u~i|2=ε1,\int_{\Omega_{1}}\left|\nabla\tilde{u}_{i}\right|^{2}=\varepsilon_{1},

which together with (15) implies

(18) Ω2|u~i|22ε1.\int_{\Omega_{2}}\left|\nabla\tilde{u}_{i}\right|^{2}\leq 2\varepsilon_{1}.

Notice that u~i\tilde{u}_{i} is now a harmonic map into a different target manifold NiN_{i}, which converges to a linear subspace of p\mathbb{R}^{p} as ii\to\infty. Since the small constant in the ε\varepsilon-regularity theorem is uniform for all NiN_{i}, (18) provides the uniform estimate that yields a limit u~\tilde{u}_{\infty} defined on Ω2\Omega_{2}. Due to the scaling, u~\tilde{u}_{\infty} is a harmonic function. By (17), the limit u~\tilde{u}_{\infty} is not trivial. However, its restriction to Ω2Ω1\Omega_{2}\setminus\Omega_{1} is constant. This is impossible and we get a contradiction.

Case 2: ηi0\eta_{i}\to 0. We define u~i\tilde{u}_{i} as in (16). The same argument as above gives a limit u~\tilde{u}_{\infty}, which is a harmonic map if Ω1|ui|2\int_{\Omega_{1}}\left|\nabla u_{i}\right|^{2} has a positive lower bound and is a harmonic function if otherwise. Since ηi0\eta_{i}\to 0, the domain Ω1(ηi)\Omega_{1}(\eta_{i}) converges to 2𝒞~\mathbb{R}^{2}\setminus\tilde{\mathcal{C}} and

2𝒞~|u~|2ε1.\int_{\mathbb{R}^{2}\setminus\tilde{\mathcal{C}}}\left|\nabla\tilde{u}_{\infty}\right|^{2}\leq\varepsilon_{1}.

Due to the removable singularity theorem and the gap theorem of harmonic map, or the fact that there is no nontrivial harmonic function on 2\mathbb{R}^{2} with bounded Dirichlet energy, u~\tilde{u}_{\infty} must be constant map/function (if ε1\varepsilon_{1} is small). To get a contradiction, it suffices to prove

(19) B(0,4)(y𝒞~B(y,δ/2))|u~i|21c~1ε1.\int_{B(0,4)\setminus\left(\bigcup_{y\in\tilde{\mathcal{C}}}B(y,\delta/2)\right)}\left|\nabla\tilde{u}_{i}\right|^{2}\geq\frac{1}{\tilde{c}_{1}}\varepsilon_{1}.

Here c~1\tilde{c}_{1} is a universal constant that will be made clear in a minute.

If (19) is not true,

(20) (B(0,4)B(0,2))(y𝒞~B(y,δ)B(y,δ/2))|u~i|21c~1ε1.\int_{(B(0,4)\setminus B(0,2))\cup\left(\bigcup_{y\in\tilde{\mathcal{C}}}B(y,\delta)\setminus B(y,\delta/2)\right)}\left|\nabla\tilde{u}_{i}\right|^{2}\leq\frac{1}{\tilde{c}_{1}}\varepsilon_{1}.

By (15) and (17), we have

ε1=Ω1(ηi)|u~i|2iΩ2(ηi)Ω1(ηi)|u~i|2,\varepsilon_{1}=\int_{\Omega_{1}(\eta_{i})}\left|\nabla\tilde{u}_{i}\right|^{2}\geq i\int_{\Omega_{2}(\eta_{i})\setminus\Omega_{1}(\eta_{i})}\left|\nabla\tilde{u}_{i}\right|^{2},

which implies that for i>c~1i>\tilde{c}_{1},

(21) B(0,ηi2)B(0,ηi2/2)(y𝒞~B(y,2ηi2δ)B(y,ηi2δ))|u~i|21c~1ε1.\int_{B(0,\eta_{i}^{-2})\setminus B(0,\eta_{i}^{-2}/2)\cup\left(\bigcup_{y\in\tilde{\mathcal{C}}}B(y,2\eta^{2}_{i}\delta)\setminus B(y,\eta^{2}_{i}\delta)\right)}\left|\nabla\tilde{u}_{i}\right|^{2}\leq\frac{1}{\tilde{c}_{1}}\varepsilon_{1}.

If mm is the number of points in 𝒞~\tilde{\mathcal{C}}, we now have m+1m+1 annular domains,

B(0,ηi2)B(0,2)andB(y,δ)B(y,ηi2δ)for eachy𝒞~.B(0,\eta_{i}^{-2})\setminus B(0,2)\quad\mbox{and}\quad B(y,\delta)\setminus B(y,\eta_{i}^{2}\delta)\quad\mbox{for each}\,y\in\tilde{\mathcal{C}}.

As in Case 1, we still have (18), which allows us to apply the energy identity theorem to these m+1m+1 annular domains simultaneously. Due to (20) and (21), by choosing c~1\tilde{c}_{1} large, we can have

(B(0,ηi1)B(0,4))(y𝒞~B(y,δ/2)B(y,ηiδ))|u~i|2<12ε1.\int_{(B(0,\eta_{i}^{-1})\setminus B(0,4))\cup\left(\bigcup_{y\in\tilde{\mathcal{C}}}B(y,\delta/2)\setminus B(y,\eta_{i}\delta)\right)}\left|\nabla\tilde{u}_{i}\right|^{2}<\frac{1}{2}\varepsilon_{1}.

Since we may choose c~1>2\tilde{c}_{1}>2, the above inequality and the assumed falsity of (19) imply that

Ω1|u~i|2<ε1,\int_{\Omega_{1}}\left|\nabla\tilde{u}_{i}\right|^{2}<\varepsilon_{1},

which is a contradiction to (17). ∎

An unfavorable aspect of the above lemma is that we have no control on the size of the constant C1C_{1}, because the proof is by contradiction. On the other hand, C1C_{1} does not depend on η\eta. By choosing η\eta small, we obtain the following counterpart of Lemma 3.1.

Lemma 3.4.

Suppose that uu satisfies (S1)-(S3). For any β>0\beta>0, there is η0>0\eta_{0}>0 small such that for all η<η0\eta<\eta_{0},

Ω2(η)|u|2βΩ3(η)Ω2(η)|u|2.\int_{\Omega_{2}(\eta)}\left|\nabla u\right|^{2}\leq\beta\int_{\Omega_{3}(\eta)\setminus\Omega_{2}(\eta)}\left|\nabla u\right|^{2}.
Proof.

By Lemma 3.2,

(22) Ω1Ω0|u|2C1Ω2Ω1|u|2.\int_{\Omega_{1}\setminus\Omega_{0}}\left|\nabla u\right|^{2}\leq C_{1}\int_{\Omega_{2}\setminus\Omega_{1}}\left|\nabla u\right|^{2}.

For β~\tilde{\beta} to be determined, Lemma 3.1 implies the existence of some η\eta such that

(23) Ω2Ω1|u|2β~(Ω1Ω0|u|2+Ω3Ω2|u|2).\int_{\Omega_{2}\setminus\Omega_{1}}\left|\nabla u\right|^{2}\leq\tilde{\beta}\left(\int_{\Omega_{1}\setminus\Omega_{0}}\left|\nabla u\right|^{2}+\int_{\Omega_{3}\setminus\Omega_{2}}\left|\nabla u\right|^{2}\right).

By (22) and (23), we have

Ω2Ω1|u|2β~1C1β~Ω3Ω2|u|2,\int_{\Omega_{2}\setminus\Omega_{1}}\left|\nabla u\right|^{2}\leq\frac{\tilde{\beta}}{1-C_{1}\tilde{\beta}}\int_{\Omega_{3}\setminus\Omega_{2}}\left|\nabla u\right|^{2},

which implies (using Lemma 3.2 again)

Ω2|u|2(1+C1)β~1C1β~Ω3Ω2|u|2.\int_{\Omega_{2}}\left|\nabla u\right|^{2}\leq\frac{(1+C_{1})\tilde{\beta}}{1-C_{1}\tilde{\beta}}\int_{\Omega_{3}\setminus\Omega_{2}}\left|\nabla u\right|^{2}.

It suffices to choose β~\tilde{\beta} small so that

(1+C1)β~1C1β~<β.\frac{(1+C_{1})\tilde{\beta}}{1-C_{1}\tilde{\beta}}<\beta.

Taking β~\tilde{\beta} as the β\beta is Lemma 3.1, we obtain an η\eta such that the above computation works. ∎

3.2. Optimal decay estimate

In this section, we are interested in a long cylinder [0,L~]×S1[0,\tilde{L}]\times S^{1}. Assume that L~\tilde{L} is a multiple of some L>2L>2 such that

[0,L~]×S1=i=1mWi,[0,\tilde{L}]\times S^{1}=\bigcup_{i=1}^{m}W_{i},

where Wi=[(i1)L,iL]×S1W_{i}=[(i-1)L,iL]\times S^{1}.

The aim of this section is to prove the following.

Lemma 3.5.

There exists some ε1(N)>0\varepsilon_{1}(N)>0. If uu is a harmonic map from [0,L~]×S1[0,\tilde{L}]\times S^{1} to NN satisfying that

(24) Wi|u|2<ε1(N),{t}×S1|tu|2|θu|2=0\int_{W_{i}}\left|\nabla u\right|^{2}<\varepsilon_{1}(N),\qquad\int_{\left\{t\right\}\times S^{1}}\left|\partial_{t}u\right|^{2}-\left|\partial_{\theta}u\right|^{2}=0

and

(25) Wi|u|212Wi+1|u|2,i=1,2,,m1,\int_{W_{i}}\left|\nabla u\right|^{2}\leq\frac{1}{2}\int_{W_{i+1}}\left|\nabla u\right|^{2},\qquad\forall i=1,2,\cdots,m-1,

then

(26) W1|u|2C(L)e2L~Wm|u|2.\int_{W_{1}}\left|\nabla u\right|^{2}\leq C(L)e^{-2\tilde{L}}\int_{W_{m}}\left|\nabla u\right|^{2}.

By some simple arguments, we justify some further assumptions that helps in the proof.

(A1) Obviously, by (25), it is enough to show

(27) W2|u|2C(L)e2L~Wm|u|2.\int_{W_{2}}\left|\nabla u\right|^{2}\leq C(L)e^{-2\tilde{L}}\int_{W_{m}}\left|\nabla u\right|^{2}.

Due to (25) again, this is trivial if L~5L\tilde{L}\leq 5L. Hence, we can argue by induction and assume that (27) is proved for L~=(m1)L\tilde{L}=(m-1)L. Notice that the constant C(L)C(L) in (27) should not depend on mm. We may assume further that

(28) W3|u|2e2LW2|u|2.\int_{W_{3}}\left|\nabla u\right|^{2}\leq e^{2L}\int_{W_{2}}\left|\nabla u\right|^{2}.

If otherwise, we may apply the induction hypothesis to the cylinder [L,L~]×S1[L,\tilde{L}]\times S^{1} to see

W3|u|2C(L)e2(L~L)Wm|u|2,\int_{W_{3}}\left|\nabla u\right|^{2}\leq C(L)e^{-2(\tilde{L}-L)}\int_{W_{m}}\left|\nabla u\right|^{2},

from which (27) follows. Using the elliptic estimate, (25) and (28), we have

(29) supt[L,2L]sup[t,t+1]×S1|θ2u|2C1(L)W2|u|2.\sup_{t\in[L,2L]}\sup_{[t,t+1]\times S^{1}}\left|\partial_{\theta}^{2}u\right|^{2}\leq C_{1}(L)\int_{W_{2}}\left|\nabla u\right|^{2}.

(A2) Near the other end of the cylinder, we consider a natural number m1m_{1} such that mm1m-m_{1} is bounded by a constant depending on LL and

(30) Wm1|u|2ε1C1(L)L.\int_{W_{m_{1}}}\left|\nabla u\right|^{2}\leq\frac{\varepsilon_{1}}{C_{1}(L)\cdot L}.

(A3) By a similar argument as in (A1), we may assume that

(31) Wm1|u|2e2LWm11|u|2.\int_{W_{m_{1}}}\left|\nabla u\right|^{2}\leq e^{2L}\int_{W_{m_{1}-1}}\left|\nabla u\right|^{2}.

Together with (25), it implies that

(32) supt[(m12)L,(m11)L]sup[t1,t]×S1|θ2u|2C1(L)Wm11|u|2.\sup_{t\in[(m_{1}-2)L,(m_{1}-1)L]}\sup_{[t-1,t]\times S^{1}}\left|\partial_{\theta}^{2}u\right|^{2}\leq C_{1}(L)\int_{W_{m_{1}-1}}\left|\nabla u\right|^{2}.

For the proof of Lemma 3.5, it suffices to show

(33) W2|u|2C(L)e2m1LWm11|u|2.\int_{W_{2}}\left|\nabla u\right|^{2}\leq C(L)e^{-2m_{1}L}\int_{W_{m_{1}-1}}\left|\nabla u\right|^{2}.

By the mean value theorem, there is ti[(i1)L,iL]t_{i}\in[(i-1)L,iL] such that

Wi|u|2=L{ti}×S1|u|2.\int_{W_{i}}\left|\nabla u\right|^{2}=L\cdot\int_{\left\{t_{i}\right\}\times S^{1}}\left|\nabla u\right|^{2}.

Hence, finally, the proof of Lemma 3.5 is reduced to proving

(34) {t2}×S1|u|2C(L)e2(tm11t2){tm11}×S1|u|2.\int_{\left\{t_{2}\right\}\times S^{1}}\left|\nabla u\right|^{2}\leq C(L)e^{-2(t_{m_{1}-1}-t_{2})}\int_{\left\{t_{m_{1}-1}\right\}\times S^{1}}\left|\nabla u\right|^{2}.

(A1-A3) above implies that u|[t2,tm11]×S1u|_{[t_{2},t_{m_{1}-1}]\times S^{1}} satsfies the assumption of the following proposition with C2(L)=(C1(L)L)1/2C_{2}(L)=(C_{1}(L)\cdot L)^{1/2}.

Proposition 3.6.

There is some ε1(N)>0\varepsilon_{1}(N)>0. Assume that uu is a harmonic map defined on [0,T]×S1[0,T]\times S^{1} satisfying

(35) sup{t}×S1|u|+|2u|ε11/2,t[0,T]\sup_{\left\{t\right\}\times S^{1}}\left|\nabla u\right|+\left|\nabla^{2}u\right|\leq\varepsilon_{1}^{1/2},\qquad\forall t\in[0,T]

and

(36) {t}×S1|θu|2|tu|2=0,t[0,T].\int_{\left\{t\right\}\times S^{1}}\left|\partial_{\theta}u\right|^{2}-\left|\partial_{t}u\right|^{2}=0,\qquad\forall t\in[0,T].

If

({0}×S1|u|2)1/2=aand({T}×S1|u|2)1/2=b\left(\int_{\left\{0\right\}\times S^{1}}\left|\nabla u\right|^{2}\right)^{1/2}=a\quad\text{and}\quad\left(\int_{\left\{T\right\}\times S^{1}}\left|\nabla u\right|^{2}\right)^{1/2}=b

and

(37) sup[0,1]×S1|θ2u|C2(L)a<ε11/2,sup[T1,T]×S1|θ2u|C2(L)bε11/2,\sup_{[0,1]\times S^{1}}\left|\partial_{\theta}^{2}u\right|\leq C_{2}(L)a<\varepsilon_{1}^{1/2},\quad\sup_{[T-1,T]\times S^{1}}\left|\partial_{\theta}^{2}u\right|\leq C_{2}(L)b\leq\varepsilon_{1}^{1/2},

then for any t[0,T]t\in[0,T],

(38) ({t}×S1|u|2)1/22E1et+2E2et\left(\int_{\left\{t\right\}\times S^{1}}\left|\nabla u\right|^{2}\right)^{1/2}\leq 2E_{1}e^{-t}+2E_{2}e^{t}

where

E1=ae2TbeTe2T1andE2=beTae2T1.E_{1}=\frac{ae^{2T}-be^{T}}{e^{2T}-1}\quad\text{and}\quad E_{2}=\frac{be^{T}-a}{e^{2T}-1}.
Remark 3.7.

Notice that the right hand side of (38) is the solution of the ODE

g′′=gwithg(0)=2a,g(T)=2b.g^{\prime\prime}=g\qquad\text{with}\quad g(0)=2a,\quad g(T)=2b.
Proof.

Due to (36), we set

f(t)=12{t}×S1|u|2={t}×S1|θu|2.f(t)=\frac{1}{2}\int_{\left\{t\right\}\times S^{1}}\left|\nabla u\right|^{2}=\int_{\left\{t\right\}\times S^{1}}\left|\partial_{\theta}u\right|^{2}.

A computation (following the Lemma 2.1 of [LW98]) yields

(39) f′′(t)=2{t}×S1(θt2u,θu)+2(tθ2u,tθ2u)=2{t}×S1|tθ2u|2+2{t}×S1|θ2u|22{t}×S1(θ2u,A(u)(u,u)).\begin{split}f^{\prime\prime}(t)&=2\int_{\left\{t\right\}\times S^{1}}(\partial_{\theta}\partial_{t}^{2}u,\partial_{\theta}u)+2(\partial^{2}_{t\theta}u,\partial^{2}_{t\theta}u)\\ &=2\int_{\left\{t\right\}\times S^{1}}\left|\partial^{2}_{t\theta}u\right|^{2}+2\int_{\left\{t\right\}\times S^{1}}\left|\partial_{\theta}^{2}u\right|^{2}-2\int_{\left\{t\right\}\times S^{1}}(\partial_{\theta}^{2}u,A(u)(\nabla u,\nabla u)).\end{split}

By setting γ=f1/2\gamma=f^{1/2} (see [CS19]), we have

2γγ=f=2{t}×S1(θu,tθ2u)2γ({t}×S1|tθ2u|2)1/2,2\gamma\gamma^{\prime}=f^{\prime}=2\int_{\left\{t\right\}\times S^{1}}(\partial_{\theta}u,\partial^{2}_{t\theta}u)\leq 2\gamma\left(\int_{\left\{t\right\}\times S^{1}}\left|\partial^{2}_{t\theta}u\right|^{2}\right)^{1/2},

which implies that

(40) (γ)2S1|tθ2u|2.(\gamma^{\prime})^{2}\leq\int_{S^{1}}\left|\partial^{2}_{t\theta}u\right|^{2}.

Together with

f′′(t)=2γγ′′+2(γ)2,f^{\prime\prime}(t)=2\gamma\gamma^{\prime\prime}+2(\gamma^{\prime})^{2},

(39) and (40) imply that

(41) 2γγ′′2{t}×S1|θ2u|2c1{t}×S1|θ2u||u|2.2\gamma\gamma^{\prime\prime}\geq 2\int_{\left\{t\right\}\times S^{1}}\left|\partial^{2}_{\theta}u\right|^{2}-c_{1}\int_{\left\{t\right\}\times S^{1}}\left|\partial^{2}_{\theta}u\right|\left|\nabla u\right|^{2}.

By the Poincaré inequality

{t}×S1|θ2u|2{t}×S1|θu|2=γ2\int_{\left\{t\right\}\times S^{1}}\left|\partial_{\theta}^{2}u\right|^{2}\geq\int_{\left\{t\right\}\times S^{1}}\left|\partial_{\theta}u\right|^{2}=\gamma^{2}

and (35), we obtain from (41) that

γ′′γc2ε11/2γ.\gamma^{\prime\prime}\geq\gamma-c_{2}\varepsilon_{1}^{1/2}\gamma.

We assume that ε1\varepsilon_{1} is small so that 1c2ε11/29101-c_{2}\varepsilon_{1}^{1/2}\geq\frac{9}{10} so that

γ′′910γ.\gamma^{\prime\prime}\geq\frac{9}{10}\gamma.

Let hh be the solution of the ODE

h′′=910h,h(0)=a2,h(T)=b2.h^{\prime\prime}=\frac{9}{10}h,\qquad h(0)=\frac{a}{\sqrt{2}},\qquad h(T)=\frac{b}{\sqrt{2}}.

Then ODE comparison shows

(42) γh.\gamma\leq h.

This implies some decay of γ\gamma along the neck. However, the decay rate is not optimal. To improve it, we would like to use (41) again. More precisely, elliptic estimate implies that for any s[1,T1]s\in[1,T-1], we have

(43) max{s}×S1|θ2u|c3max[s1,s+1]γc3max[s1,s+1]hc3c~h(s).\max_{\left\{s\right\}\times S^{1}}\left|\partial^{2}_{\theta}u\right|\leq c_{3}\max_{[s-1,s+1]}\gamma\leq c_{3}\max_{[s-1,s+1]}h\leq c_{3}\tilde{c}h(s).

Here in the last inequality above, we have used (67) of Lemma A.1. By (67) of Lemma A.1 again and (37), we obtain

max{s}×S1|θ2u|c~C2(L)h(s),s[T1,T].\max_{\left\{s\right\}\times S^{1}}\left|\partial^{2}_{\theta}u\right|\leq\tilde{c}C_{2}(L)h(s),\qquad\forall s\in[T-1,T].

The same inequality holds for s[0,1]s\in[0,1], because of (37) and (68) of Lemma A.1.

With the new upper bound of |θ2u|\left|\partial_{\theta}^{2}u\right|, we derive from (41)

(44) γ′′γc4C2(L)h2s[0,T].\gamma^{\prime\prime}\geq\gamma-c_{4}C_{2}(L)h^{2}\qquad\forall s\in[0,T].

Let HH be the solution of the ODE

H′′=95H,H(0)=a22,h(T)=b22.H^{\prime\prime}=\frac{9}{5}H,\qquad H(0)=\frac{a^{2}}{2},\qquad h(T)=\frac{b^{2}}{2}.

We claim that

(45) h2Hon[0,T].h^{2}\leq H\qquad\text{on}\quad[0,T].

In fact,

(h2)′′=2(h)2+95h295h2,h(0)=a22,h(T)=b22.(h^{2})^{\prime\prime}=2(h^{\prime})^{2}+\frac{9}{5}h^{2}\geq\frac{9}{5}h^{2},\qquad h(0)=\frac{a^{2}}{2},\qquad h(T)=\frac{b^{2}}{2}.

The claim follows from ODE comparison again.

Combining (44) and (45), we obtain

γ′′γc4C2(L)Hon[0,T].\gamma^{\prime\prime}\geq\gamma-c_{4}C_{2}(L)H\qquad\text{on}\qquad[0,T].

Hence, if c5=54c4c_{5}=\frac{5}{4}c_{4}, we have

(γ+c5C2(L)H)′′(γ+c5C2(L)H)on[0,T].(\gamma+c_{5}C_{2}(L)H)^{\prime\prime}\geq(\gamma+c_{5}C_{2}(L)H)\qquad\text{on}\qquad[0,T].

Moreover, the assumption (37) implies that

(γ+c5C2(L)H)(0)\displaystyle(\gamma+c_{5}C_{2}(L)H)(0) \displaystyle\leq (h+c5C2(L)H)(0)\displaystyle(h+c_{5}C_{2}(L)H)(0)
=\displaystyle= (a2+c5C2(L)a22)\displaystyle(\frac{a}{\sqrt{2}}+c_{5}C_{2}(L)\frac{a^{2}}{2})
\displaystyle\leq 2a,\displaystyle\sqrt{2}a,

if we require ε11/2c5\varepsilon_{1}^{1/2}c_{5} to be small. Similarly,

(γ+c5C2(L)H)(T)\displaystyle(\gamma+c_{5}C_{2}(L)H)(T) \displaystyle\leq 2b.\displaystyle\sqrt{2}b.

ODE comparison again gives that

γγ+c6Hg2on[0,T]\gamma\leq\gamma+c_{6}H\leq\frac{g}{\sqrt{2}}\qquad\text{on}\qquad[0,T]

where gg is the solution to

g′′=g,g(0)=2a,g(T)=2b.g^{\prime\prime}=g,\qquad g(0)=2{a},\qquad g(T)=2{b}.

With Proposition 3.6, we are now ready to finish the proof of Lemma 3.5. The growth condition (25) implies that

{t4}×S1|u|2{t2}×S1|u|2.\int_{\left\{t_{4}\right\}\times S^{1}}\left|\nabla u\right|^{2}\geq\int_{\left\{t_{2}\right\}\times S^{1}}\left|\nabla u\right|^{2}.

Setting L=(t4t2)L^{\prime}=(t_{4}-t_{2}) and noticing that L(L,3L)L^{\prime}\in(L,3L), we can derive (34) (hence finish the proof of Lemma 3.5) from the following corollary.

Corollary 3.8.

Assume that uu satisfies all assumptions of Proposition 3.6 with L2L\geq 2 and T>4LT>4L. If

(46) {L}×S1|u|2{0}×S1|u|2\int_{\left\{L^{\prime}\right\}\times S^{1}}\left|\nabla u\right|^{2}\geq\int_{\left\{0\right\}\times S^{1}}\left|\nabla u\right|^{2}

for some L[L,3L]L^{\prime}\in[L,3L], then

{0}×S1|u|2C3(L)e2T{T}×S1|u|2.\int_{\left\{0\right\}\times S^{1}}\left|\nabla u\right|^{2}\leq C_{3}(L)e^{-2T}\int_{\left\{T\right\}\times S^{1}}\left|\nabla u\right|^{2}.
Proof.

Proposition 3.6 and (46) imply that

a2ae2TbeTe2T1eL+2beTae2T1eL.a\leq 2\frac{ae^{2T}-be^{T}}{e^{2T}-1}e^{-L^{\prime}}+2\frac{be^{T}-a}{e^{2T}-1}e^{L^{\prime}}.

Hence,

beT+LeTLe2T1a(12e2TLeLe2T1).b\frac{e^{T+L^{\prime}}-e^{T-L^{\prime}}}{e^{2T}-1}\geq a\left(\frac{1}{2}-\frac{e^{2T-L^{\prime}}-e^{L^{\prime}}}{e^{2T}-1}\right).

Recalling that L(L,3L)L^{\prime}\in(L,3L) and that L>log100L>\log 100, we get

e3LeTb1e4L1e2Ta(12eL1e2T+2L1e2T).e^{3L}e^{-T}b\frac{1-e^{-4L}}{1-e^{-2T}}\geq a\left(\frac{1}{2}-e^{-L}\frac{1-e^{-2T+2L}}{1-e^{-2T}}\right).

The proof is done by taking C3(L)=16e6LC_{3}(L)=16e^{6L} and noticing that T>8T>8. ∎

4. Proof of the main theorem

The proof of Theorem 1.2 consists of three steps. First, by studying the relation between g~i\tilde{g}_{i} and g¯i\bar{g}_{i}, we give an equivalent form of the main theorem. Then, we prove a weak decay estimate by using the three circle lemma (Lemma 3.1) and its generalization (Lemma 3.4). Finally, we use the results in Section 3.2 to improve the weak decay into a sharp one, which is exactly our main theorem.

4.1. An equivalent form of the main theorem

As before,

Ω=B(0,δ)j=1,,lB(xi(j),δ1λi(j)).\Omega=B(0,\delta)\setminus\bigcup_{j=1,\cdots,l}B(x_{i}^{(j)},\delta^{-1}\lambda_{i}^{(j)}).

Recall that we have defined two metrics g¯i\bar{g}_{i} and g~i\tilde{g}_{i} on Ω\Omega. The following lemma compares them.

Lemma 4.1.

There is a constant C>0C>0 such that for any zΩz\in\Omega,

1Cg¯ie2d(z)g~iC,\frac{1}{C}\leq\frac{\bar{g}_{i}}{e^{2d(z)}\tilde{g}_{i}}\leq C,

where d(z)d(z) is the distance from zz to Ω\partial\Omega with respect to g¯i\bar{g}_{i}.

Before the proof, we notice that it implies that the following theorem is equivalent to Theorem 1.2.

Theorem 4.2.

Suppose that uiu_{i} is a sequence of harmonic maps satisfying (U1) and (U2) and that Ωi\Omega_{i} is a generalized neck domain, then there is some constant CC such that

(47) |ui|g¯iCed(z)onΩi.\left|\nabla u_{i}\right|_{\bar{g}_{i}}\leq Ce^{-d(z)}\qquad\text{on}\quad\Omega_{i}.

By definition(see (8) and (2)), for the proof of Lemma 4.1, it suffices to show that there exists C>0C>0 such that

(48) 1Cw(z)e2d(z)ω(z)Cw(z),zΩ\frac{1}{C}w(z)\leq e^{2d(z)}\omega(z)\leq Cw(z),\qquad\forall z\in\Omega

where

(49) ω(z)=1+j=1l(λi(j))2|zxi(j)|4.\omega(z)=1+\sum_{j=1\ldots l}\frac{(\lambda_{i}^{(j)})^{2}}{\left|z-x_{i}^{(j)}\right|^{4}}.

For the proof, it is essential to understand the meaning of d(z)d(z) and the contribution of each term in the sum of (49). Figure 2 illustrates an example and it is helpful in understanding the proof that follows. Here R1,R2,R3R_{1},R_{2},R_{3} represent three real bubbles; G1,G2G_{1},G_{2} two ghost bubbles; for a point zΩz\in\Omega, we denote by p1,,p4p_{1},\cdots,p_{4} the four paths from zz to the components of Ω\partial\Omega. In this case, d(z)d(z) is going to be the minimal length of p1,,p4p_{1},\cdots,p_{4}.

Refer to caption
Figure 2. Distance function to the boundary

In the following proof, we write aibia_{i}\sim b_{i} if there is C>0C>0 such that 1CaibiCai\frac{1}{C}a_{i}\leq b_{i}\leq Ca_{i}.

Proof.

By taking logarithm, it suffices to show

(50) |12logω(z)w(z)d(z)|CzΩ.\left|-\frac{1}{2}\log\frac{\omega(z)}{w(z)}-d(z)\right|\leq C\qquad\forall z\in\Omega.

The rest of the proof deals with two cases separately: zz lies in a ghost bubble domain, or in a simple neck domain.

Case 1. Assume that zz is in the ghost bubble domain

B(ci,2σi)y𝒞B(y,δσi).B(c_{i},2\sigma_{i})\setminus\bigcup_{y\in\mathcal{C}}B(y,\delta\sigma_{i}).

Take any zB(ci,2σi)z^{\prime}\in\partial B(c_{i},2\sigma_{i}). We claim that

|(12logω(z)w(z)d(z))(12logω(z)w(z)d(z))|C.\left|\left(-\frac{1}{2}\log\frac{\omega(z^{\prime})}{w(z^{\prime})}-d(z^{\prime})\right)-\left(-\frac{1}{2}\log\frac{\omega(z)}{w(z)}-d(z)\right)\right|\leq C.

Hence, it suffices to prove (50) for zB(ci,2σi)z\in\partial B(c_{i},2\sigma_{i}), which is Case 2. To show the claim, recall (W1) in the definition of ww on ghost bubble domain, which implies that

w(z)w(z)and|d(z)d(z)|C.w(z)\sim w(z^{\prime})\quad\text{and}\quad\left|d(z)-d(z^{\prime})\right|\leq C.

Next, we study the difference between ω(z)\omega(z) and ω(z)\omega(z^{\prime}). For each jj, the bubble (xi(j),λi(j))(x_{i}^{(j)},\lambda_{i}^{(j)}) is either on top of (ci,σi)(c_{i},\sigma_{i}), or is separate from (ci,σi)(c_{i},\sigma_{i}). In the first case, we have

limixi(j)ciσi𝒞.\lim_{i\to\infty}\frac{x_{i}^{(j)}-c_{i}}{\sigma_{i}}\in\mathcal{C}.

Hence, for ii sufficiently large,

|zxi(j)|,|zxi(j)|[δ/2σi,4σi],\left|z-x_{i}^{(j)}\right|,\left|z^{\prime}-x_{i}^{(j)}\right|\in[\delta/2\sigma_{i},4\sigma_{i}],

which implies that

(51) (λi(j))2|zxi(j)|4(λi(j))2|zxi(j)|4.\frac{(\lambda_{i}^{(j)})^{2}}{\left|z-x_{i}^{(j)}\right|^{4}}\sim\frac{(\lambda_{i}^{(j)})^{2}}{\left|z^{\prime}-x_{i}^{(j)}\right|^{4}}.

In the second case, we have

limi|cixi(j)|σi=.\lim_{i\to\infty}\frac{\left|c_{i}-x_{i}^{(j)}\right|}{\sigma_{i}}=\infty.

Together with the fact that

|ciz|,|ciz|2σi,\left|c_{i}-z\right|,\left|c_{i}-z^{\prime}\right|\leq 2\sigma_{i},

we obtain (51) again. In summary, we have ω(z)ω(z)\omega(z)\sim\omega(z^{\prime}) and our claim is proved.

Case 2. Assume that zz is in a simple neck domain

B(ci,δσi)B(ci,λi).B(c_{i},\delta\sigma_{i})\setminus B(c_{i},\lambda_{i}).

First, we study the distance from zz to B(0,δ)\partial B(0,\delta) with respect to g¯i\bar{g}_{i}. In general, the path from zz to B(0,δ)\partial B(0,\delta) may pass several (or no) ghost bubble domains. For simplicity, we assume that there is only one ghost bubble domain. This is a situation illustrated by p4p_{4} in Figure 2. In this case, the ghost bubble G2G_{2} is represented by a sequence (xi,σi)(x_{i},\sigma_{i}) and

limicixiσi\lim_{i\to\infty}\frac{c_{i}-x_{i}}{\sigma_{i}}

is the concentration point at which the real bubbles R1R_{1} and R2R_{2} hide. Since the diameter of the ghost bubble domain measured by g¯i\bar{g}_{i} is bounded, we have

(52) d(z,B(0,δ))=d(z,B(ci,δσi))+C+d(B(xi,σi),B(0,δ))=logσi|zci|+log1σi+C=log|zci|+C.\begin{split}d(z,\partial B(0,\delta))&=d(z,\partial B(c_{i},\delta\sigma_{i}))+C+d(\partial B(x_{i},\sigma_{i}),\partial B(0,\delta))\\ &=\log\frac{\sigma_{i}}{\left|z-c_{i}\right|}+\log\frac{1}{\sigma_{i}}+C\\ &=-\log\left|z-c_{i}\right|+C.\end{split}

Next, we study the distance from zz to B(xi(j),δ1λi(j))\partial B(x_{i}^{(j)},\delta^{-1}\lambda_{i}^{(j)}) for j=1,,lj=1,\cdots,l. There are two possibilities:

Case A. The real bubble (xi(j),λi(j))(x_{i}^{(j)},\lambda_{i}^{(j)}) sits ’on top of’ the neck containing zz, in the sense that

(53) limixi(j)ci|zci|<;\lim_{i\to\infty}\frac{x_{i}^{(j)}-c_{i}}{\left|z-c_{i}\right|}<\infty;

Case B. The real bubble is separate from the neck, in the sense that

(54) limixi(j)ci|zci|=.\lim_{i\to\infty}\frac{x_{i}^{(j)}-c_{i}}{\left|z-c_{i}\right|}=\infty.

In Figure 2, the real bubbles R1R_{1} and R2R_{2} are Case A and the bubble R3R_{3} is Case B. For case A, we assume again that the path from zz to B(xi(j),λi(j))\partial B(x_{i}^{(j)},\lambda_{i}^{(j)}) passes only one ghost bubble (see G1G_{1} in Figure 2), (ci,λi)(c_{i},\lambda_{i}), so that

limixi(j)ciλi\lim_{i\to\infty}\frac{x_{i}^{(j)}-c_{i}}{\lambda_{i}}

is where (xi(j),λi(j))(x_{i}^{(j)},\lambda_{i}^{(j)}) concentrates. As in Case 1, we have

(55) d(z,B(xi(j),δ1λi(j)))=d(z,B(ci,2λi))+C+d(B(xi(j),δλi),B(xi(j),δ1λi(j)))=log|zci|λi+logλiλi(j)+C=logλi(j)|zci|+C.\begin{split}d(z,\partial B(x_{i}^{(j)},\delta^{-1}\lambda_{i}^{(j)}))&=d(z,\partial B(c_{i},2\lambda_{i}))+C+d(\partial B(x_{i}^{(j)},\delta\lambda_{i}),\partial B(x_{i}^{(j)},\delta^{-1}\lambda_{i}^{(j)}))\\ &=\log\frac{\left|z-c_{i}\right|}{\lambda_{i}}+\log\frac{\lambda_{i}}{\lambda_{i}^{(j)}}+C\\ &=-\log\frac{\lambda_{i}^{(j)}}{\left|z-c_{i}\right|}+C.\end{split}

For Case B, there is a ghost bubble (see G2G_{2} in Figure 2) (ci,σi)(c^{\prime}_{i},\sigma^{\prime}_{i}) such that

limizciσilimixi(j)ciσi.\lim_{i\to\infty}\frac{z-c^{\prime}_{i}}{\sigma^{\prime}_{i}}\neq\lim_{i\to\infty}\frac{x_{i}^{(j)}-c^{\prime}_{i}}{\sigma^{\prime}_{i}}.

This case is illustrated by p3p_{3} in Figure 2. In general, the path from zz to G2G_{2} and from G2G_{2} to R3R_{3} may pass more ghost bubble domains. However, the proof remains the same by similar argument above. In this case,

(56) d(z,B(xi(j),δ1λi(j)))=d(z,B(ci,δσi))+C+d(B(xi(j),δσi),B(xi(j),δ1λi(j)))=logσi|zci|+logσiλi(j)+C=log(σi)2λi(j)|zci|+C.\begin{split}d(z,\partial B(x_{i}^{(j)},\delta^{-1}\lambda_{i}^{(j)}))&=d(z,\partial B(c_{i},\delta\sigma^{\prime}_{i}))+C+d(\partial B(x_{i}^{(j)},\delta\sigma^{\prime}_{i}),\partial B(x_{i}^{(j)},\delta^{-1}\lambda_{i}^{(j)}))\\ &=\log\frac{\sigma^{\prime}_{i}}{\left|z-c_{i}\right|}+\log\frac{\sigma^{\prime}_{i}}{\lambda_{i}^{(j)}}+C\\ &=-\log\frac{(\sigma_{i}^{\prime})^{2}}{\lambda_{i}^{(j)}\cdot\left|z-c_{i}\right|}+C.\end{split}

We go back to the proof of (50) by computing

(57) 12logω(z)w(z)=12log(|zci|2+j(λi(j))2|zxi(j)|2|zci|2|zxi(j)|2).-\frac{1}{2}\log\frac{\omega(z)}{w(z)}=-\frac{1}{2}\log\left(\left|z-c_{i}\right|^{2}+\sum_{j}\frac{(\lambda_{i}^{(j)})^{2}}{\left|z-x_{i}^{(j)}\right|^{2}}\cdot\frac{\left|z-c_{i}\right|^{2}}{\left|z-x_{i}^{(j)}\right|^{2}}\right).

Here in the parenthesis, it is the sum of l+1l+1 positive terms. If we denote the largest one by MM, we have

|12logω(z)w(z)(12logM)|c~.\left|-\frac{1}{2}\log\frac{\omega(z)}{w(z)}-(-\frac{1}{2}\log M)\right|\leq\tilde{c}.

There are exactly l+1l+1 boundary components of Ω\partial\Omega. We will show that the minus logarithm of each positive term in the parenthesis is (up to a constant) the distance from zz to a boundary component.

First, by (52), the first term in the parenthesis correpsonds to the distance from zz to B(0,δ)\partial B(0,\delta).

For j=1,,lj=1,\cdots,l, in Case A, we have

|zxi(j)||zci|,\left|z-x_{i}^{(j)}\right|\sim\left|z-c_{i}\right|,

which implies that

(λi(j))2|zxi(j)|2|zci|2|zxi(j)|2(λi(j))2|zci|2.\frac{(\lambda_{i}^{(j)})^{2}}{\left|z-x_{i}^{(j)}\right|^{2}}\cdot\frac{\left|z-c_{i}\right|^{2}}{\left|z-x_{i}^{(j)}\right|^{2}}\sim\frac{(\lambda_{i}^{(j)})^{2}}{\left|z-c_{i}\right|^{2}}.

It is related to the distance from zz to B(xi(j),δ1λi(j))\partial B(x^{(j)}_{i},\delta^{-1}\lambda_{i}^{(j)}) by (55). In case B, we have

|zxi(j)|σi,\left|z-x_{i}^{(j)}\right|\sim\sigma^{\prime}_{i},

which implies that

(λi(j))2|zxi(j)|2|zci|2|zxi(j)|2(λi(j))2|zci|2(σi)4.\frac{(\lambda_{i}^{(j)})^{2}}{\left|z-x_{i}^{(j)}\right|^{2}}\cdot\frac{\left|z-c_{i}\right|^{2}}{\left|z-x_{i}^{(j)}\right|^{2}}\sim\frac{(\lambda_{i}^{(j)})^{2}\left|z-c_{i}\right|^{2}}{(\sigma^{\prime}_{i})^{4}}.

It is related to the distance from zz to B(xi(j),δ1λi(j))\partial B(x^{(j)}_{i},\delta^{-1}\lambda_{i}^{(j)}) by (56). ∎

4.2. A weak decay estimate

The aim of this subsection is to prove the following inequality that is weaker than (47),

(58) |ui|g¯iCeαd(z)onΩi\left|\nabla u_{i}\right|_{\bar{g}_{i}}\leq Ce^{-\alpha d(z)}\qquad\text{on}\quad\Omega_{i}

for some α(0,1)\alpha\in(0,1).

Let m1m_{1} be the total number of ghost bubbles and m2m_{2} be the maximal number of boundary components of the ghost bubble domains. Setting β=12m2\beta=\frac{1}{2m_{2}}, a constant LL is determined by Lemma 3.1 and Lemma 3.4 (with η=eL\eta=e^{-L}). As ii goes to infinity, so do the lengths of simple neck domains. Assume without loss of generality that these lengths are integer multiples of LL. Then the generalized neck domain becomes the union of many cylinder pieces

W=[0,L]×S1W=[0,L]\times S^{1}

and one ghost bubble piece(domain)

W=B(xi,2σi)y𝒞B(ci(y),δσi).W=B(x_{i},2\sigma_{i})\setminus\bigcup_{y\in\mathcal{C}}B(c_{i}(y),\delta\sigma_{i}).

In either case, we write E(W)E(W) for the integral W|u|g¯i2\int_{W}\left|\nabla u\right|_{\bar{g}_{i}}^{2}.

Since the diameter of each piece (w.r.t. g¯i\bar{g}_{i}) is bounded, the weak decay estimate (58) follows from the next lemma.

Lemma 4.3.

For any piece WW as above, there are ss pieces W1,,WsW_{1},\cdots,W_{s} such that

(i) W=W1W=W_{1};

(ii) WsW_{s} touches one component of Ω\partial\Omega;

(iii) j=1,,sWj\bigcup_{j=1,\cdots,s}W_{j} is connected;

(iv) The number of index kk from 22 to ss not satisfying

2E(Wk1)E(Wk)2E(W_{k-1})\leq E(W_{k})

is bounded by 4m14m_{1};

(v) For any j=1,,sj=1,\cdots,s,

E(Wj)C2j.E(W_{j})\leq C2^{-j}.
Refer to caption
Figure 3. From left to right: Case 1, Case 2A and Case 2B
Proof.

The proof is by induction. Set W1=WW_{1}=W.

There are three cases depending on the position of WW.

Case 1: WW is a ghost bubble domain(see Case 1 of Figure 3). By Lemma 3.4,

Ω2|u|2βΩ3Ω2|u|2.\int_{\Omega_{2}}\left|\nabla u\right|^{2}\leq\beta\int_{\Omega_{3}\setminus\Omega_{2}}\left|\nabla u\right|^{2}.

Since β=12m2\beta=\frac{1}{2m_{2}}, there is a component of Ω3Ω2\Omega_{3}\setminus\Omega_{2}, which is a cylinder piece denoted by W4W_{4}, such that

Ω2|u|212E(W4).\int_{\Omega_{2}}\left|\nabla u\right|^{2}\leq\frac{1}{2}E(W_{4}).

Let W2W_{2} and W3W_{3} be the two pieces connecting WW and W4W_{4}. Notice that W3W_{3} is a cylinder and we have E(W4)2E(W3)E(W_{4})\geq 2E(W_{3}).

Case 2: If WW is a cylinder next to a ghost bubble domain, we use Lemma 3.4 as above to get a component WW^{\prime} of Ω3Ω2\Omega_{3}\setminus\Omega_{2} satisfying

Ω2|u|212E(W).\int_{\Omega_{2}}\left|\nabla u\right|^{2}\leq\frac{1}{2}E(W^{\prime}).

If WW^{\prime} and WW is in the same component of Ω3Ω0\Omega_{3}\setminus\Omega_{0} (see Case 2A of Figure 3), then we set W=W3W^{\prime}=W_{3} and let W2W_{2} be the piece between W3W_{3} and W1W_{1}. Notice that W2W_{2} is a cylinder and that E(W3)2E(W2)E(W_{3})\geq 2E(W_{2}).

If WW^{\prime} and WW are not in the same component of Ω3Ω0\Omega_{3}\setminus\Omega_{0} (see Case 2B of Figure 3), then we set W2W_{2} to be the ghost bubble domain, W5=WW_{5}=W^{\prime} and W3,W4W_{3},W_{4} be the two pieces between W2W_{2} and W5W_{5}.

In this case, it is also true that W4W_{4} and W5W_{5} are cylinders and E(W5)2E(W4)E(W_{5})\geq 2E(W_{4}).

Case 3: WW is a cylinder and the two adjacent pieces are also cylinders. By Lemma 3.1, since β14\beta\leq\frac{1}{4}, there is at least one adjacent piece, which we denote by W2W_{2} satisfying E(W2)2E(W1)E(W_{2})\geq 2E(W_{1}).

Assume that we have found W1,,WjW_{1},\cdots,W_{j} such that

(H1) Wk1W_{k-1} is adjacent to WkW_{k} for any k=2,,jk=2,\cdots,j.

(H2) Wj1W_{j-1} and WjW_{j} are both cylinders satisfying E(Wj)2E(Wj1)E(W_{j})\geq 2E(W_{j-1});

(H3) the number of index kk from 22 to jj not satisfying

2E(Wk1)E(Wk)2E(W_{k-1})\leq E(W_{k})

is bounded by 4m14m_{1}.

If WjW_{j} touches the boundary of Ω\partial\Omega, then the induction is complete. To conclude the proof, it remains to justify (iv) and (v). Notice that (iv) is just (H3) and (v) follows from (iv) and the fact that for any piece WW^{\prime} touching the boundary, we have E(W)Cε1E(W^{\prime})\leq C\varepsilon_{1}.

If WjW_{j} does not touch the boundary, we define Wj+1W_{j+1} as follows. By (H2), WjW_{j} is a cylinder. Hence, there are two adjacent pieces and one of them is Wj1W_{j-1}. Denote the other by WW^{\prime}.

If WW^{\prime} is a cylinder, then by Lemma 3.1, we conclude that E(W)2E(Wj)E(W^{\prime})\geq 2E(W_{j}) and denote WW^{\prime} by Wj+1W_{j+1}.

Refer to caption
Figure 4. passing the ghost bubble domain

If WW^{\prime} is a ghost bubble domain(see Figure 4), we choose W′′W^{\prime\prime} to be the component in Ω3Ω2\Omega_{3}\setminus\Omega_{2} satisfying

(59) Ω2|u|212E(W′′).\int_{\Omega_{2}}\left|\nabla u\right|^{2}\leq\frac{1}{2}E(W^{\prime\prime}).

Notice that W′′W^{\prime\prime} and WjW_{j} can not be in the same component of Ω3Ω0\Omega_{3}\setminus\Omega_{0}, otherwise W′′W^{\prime\prime} would be Wj2W_{j-2}, which contradicts (59). Then we set Wj+1=WW_{j+1}=W^{\prime}, Wj+4=W′′W_{j+4}=W^{\prime\prime} and let Wj+2,Wj+3W_{j+2},W_{j+3} be the two pieces between WW^{\prime} and W′′W^{\prime\prime}.

It is easy to check that the induction hypothesis (H1)-(H3) hold. We repeat the induction construction until the proof is done. ∎

4.3. The optimal decay estimate

The aim of this subsection is to prove (47). The proof is based on Lemma 4.3 and Lemma 3.5.

Recall that in the proof of Lemma 4.3, Ω\Omega is decomposed into cylinder pieces of length LL and ghost bubble piece. Our first step in the proof of Theorem 4.2 is to show that it suffices to prove (47) for zΩz\in\Omega satisfying

d(z,W)>15Ld(z,W)>15L

for any ghost bubble piece WW. Here d(z,W)d(z,W) is measured with respect to g¯i\bar{g}_{i}.

Assume this is true and let z~\tilde{z} be any point satisfying d(z~,W)15Ld(\tilde{z},W)\leq 15L for some ghost bubble domain WW. With W=Ω0W=\Omega_{0} in mind, we recall the definition of Ω1,Ω2,\Omega_{1},\Omega_{2},\cdots in (14). By Lemma 3.4 and Lemma 3.1, we have

E(Ω2)12E(Ω3Ω2)14E(Ω4Ω3).E(\Omega_{2})\leq\frac{1}{2}E(\Omega_{3}\setminus\Omega_{2})\leq\frac{1}{4}E(\Omega_{4}\setminus\Omega_{3})\leq\cdots.

Together with elliptic estimates, the above inequality implies that

|u|g¯i2(z~)CΩ16|u|g¯i2CΩ17Ω16|u|g¯i2.\left|\nabla u\right|_{\bar{g}_{i}}^{2}(\tilde{z})\leq C\int_{\Omega_{16}}\left|\nabla u\right|_{\bar{g}_{i}}^{2}\leq C\int_{\Omega_{17}\setminus\Omega_{16}}\left|\nabla u\right|_{\bar{g}_{i}}^{2}.

By our assumption, for any zΩ17Ω16z^{\prime}\in\Omega_{17}\setminus\Omega_{16}, we have d(z,W)>15Ld(z^{\prime},W)>15L and hence

|u|g¯i2(z~)CsupzΩ17Ω16e2d(z)Ce2d(z~),\left|\nabla u\right|_{\bar{g}_{i}}^{2}(\tilde{z})\leq C\sup_{z^{\prime}\in\Omega_{17}\setminus\Omega_{16}}e^{-2d(z^{\prime})}\leq Ce^{-2d(\tilde{z})},

because d(z~,z)Cd(\tilde{z},z^{\prime})\leq C.

Hence, for the rest of the proof we assume that d(z,W)>15Ld(z,W)>15L for any ghost bubble piece. Let WzW_{z} be the cylinder piece containing zz and WW_{-} and W+W_{+} be the two adjacent pieces. By elliptic estimates, we have

|u|g¯i2(z)C(E(W)+E(Wz)+E(W+)).\left|\nabla u\right|^{2}_{\bar{g}_{i}}(z)\leq C\left(E(W_{-})+E(W_{z})+E(W_{+})\right).

Therefore, it suffices to show

E(W)Ce2d(W,Ω)E(W)\leq Ce^{-2d(W,\partial\Omega)}

for any cylinder piece WW whose distance to any ghost bubble piece is larger than 12L12L.

For this WW, Lemma 4.3 gives a sequence W1,,WsW_{1},\cdots,W_{s}. For simplicy, we assume only one of them, say WlW_{l}, is a ghost bubble piece. The proof of Lemma 4.3 shows that

2E(Wk1)E(Wk)2E(W_{k-1})\leq E(W_{k})

for k=2,3,,l1k=2,3,\cdots,l-1 and k=l+3,l+4,,sk=l+3,l+4,\cdots,s (see Figure 4). Hence, we can apply Lemma 3.5 to see

E(W1)C(L)e2lLE(Wl1)E(W_{1})\leq C(L)e^{-2lL}E(W_{l-1})

and

E(Wl+3)C(L)e2(sl)LE(Ws).E(W_{l+3})\leq C(L)e^{-2(s-l)L}E(W_{s}).

Moreover, we have E(Wl+3)E(Wl1)E(W_{l+3})\geq E(W_{l-1}) as a consequence of Lemma 3.4. This finishes the proof of Theorem 4.2.

5. An application

In this section, we prove Theorem 1.5. To simplify the notations, we assume that there is only one energy concentration point pΣp\in\Sigma and that there are several real bubbles concentrated at pp and these bubbles are all separated. Hence, the real bubbles and the weak limit are connected with only one generalized neck domain. More precisely, take a conformal coordinate centered at pp and assume that the real bubbles are

j:(xi(j),λi(j)),j=1,,l.\mathcal{B}_{j}:\quad(x_{i}^{(j)},\lambda_{i}^{(j)}),\qquad j=1,\cdots,l.

By setting

ci=1lj=1lxi(j),c_{i}=\frac{1}{l}\sum_{j=1}^{l}x_{i}^{(j)},

for some small δ0>0\delta_{0}>0, the generalized neck domain is

Ωi=B(ci,δ0)j=1,,lB(xi(j),δ01λi(j)).\Omega_{i}=B(c_{i},\delta_{0})\setminus\bigcup_{j=1,\cdots,l}B(x_{i}^{(j)},\delta_{0}^{-1}\lambda_{i}^{(j)}).

We choose this δ0\delta_{0} to be small so that all results proved in previous sections hold. In what follows, we shall need another parameter δ(0,δ0)\delta\in(0,\delta_{0}) and set

(60) Ωi(δ)=B(ci,δ)j=1,,lB(xi(j),δ1λi(j)).\Omega_{i}(\delta)=B(c_{i},\delta)\setminus\bigcup_{j=1,\cdots,l}B(x_{i}^{(j)},\delta^{-1}\lambda_{i}^{(j)}).

The outline of the proof is the same as in [Yin19], which we recall below.

5.1. Outline of proof

Let uiu_{i} be the sequence in Theorem 1.4. We shall define a sequence of conformal metrics gig_{i} on Σ\Sigma. While NI(ui)NI(u_{i}) is conformally invariant, the eigenvalues and eigenfunctions of the operator JuiJ_{u_{i}} do depend on gig_{i}. By taking a subsequence if necessary, we assume

m=limiNI(ui).m=\lim_{i\to\infty}NI(u_{i}).

Suppose that βi,1,,βi,m\beta_{i,1},\cdots,\beta_{i,m} are the nonpositive eigenvalues (counting multiplicities) of JuiJ_{u_{i}} and that vi,1,,vi,mv_{i,1},\cdots,v_{i,m} are the corresponding eigenfunctions, i.e.

(61) Jui(vi,k)=βi,kvi,kk=1,,m.J_{u_{i}}(v_{i,k})=\beta_{i,k}v_{i,k}\qquad k=1,\cdots,m.

Here vi,kv_{i,k} are the sections of the pullback bundle uiTNu_{i}^{*}TN, which are normalized so that

(62) Mvi,k,vi,k𝑑Vgi=δk,k.\int_{M}\langle v_{i,k},v_{i,k^{\prime}}\rangle dV_{g_{i}}=\delta_{k,k^{\prime}}.

Notice that we have embedded NN into p\mathbb{R}^{p} and hence vi,kv_{i,k} are also regarded as p\mathbb{R}^{p}-valued functions that are perpendicular to the tangent space of NN at uiu_{i}.

We study the limit of (Σ,gi,ui,βi,k,vi,k)(\Sigma,g_{i},u_{i},\beta_{i,k},v_{i,k}). By our choice of gig_{i} (see below), we shall obtain a limit

(Σ,g,u,βk,vk).(\Sigma,g,u_{\infty},\beta_{k},v_{k}).

Here uu_{\infty} is the weak limit of uiu_{i} and βk\beta_{k} and vkv_{k} are the eigenvalues and eigenfunctions of JuJ_{u_{\infty}} respectively. For now, we do not know if they are linearly independent or not. This is a key issue that will be addressed later.

For each real bubble j(j=1,,l)\mathcal{B}_{j}(j=1,\cdots,l), we obtain a limit

(2,gb,ωj,β~k(j),v~k(j)).(\mathbb{R}^{2},g_{b},\omega_{j},\tilde{\beta}_{k}^{(j)},\tilde{v}_{k}^{(j)}).

For the definition of gbg_{b}, see (64) in the next subsection. Again, the eigenfunctions v~k(j)\tilde{v}_{k}^{(j)} for JωjJ_{\omega_{j}} may be linearly dependent.

To proved the desired inequality in Theorem 1.5, we claim that

(63) Σvk,vk𝑑Vg+j=1lS2v~k(j),v~k(j)𝑑Vgb=δk,k.\int_{\Sigma}\langle v_{k},v_{k^{\prime}}\rangle dV_{g}+\sum_{j=1}^{l}\int_{S^{2}}\langle\tilde{v}_{k}^{(j)},\tilde{v}_{k^{\prime}}^{(j)}\rangle dV_{g_{b}}=\delta_{k,k^{\prime}}.

In a linear space with inner product, the dimension of the linear subspace spanned by α1,,αm\alpha_{1},\cdots,\alpha_{m} is the rank of the matrix

(αk,αk)k,k=1,,m.\left(\langle\alpha_{k},\alpha_{k^{\prime}}\rangle\right)_{k,k^{\prime}=1,\cdots,m}.

Hence, Theorem 1.5 follows from (63).

Intuitively, (63) is a consequence of (62). For any fixed δ\delta, while the convergence on ΣB(0,δ)\Sigma\setminus B(0,\delta) and B(xi(j),δ1λi(j))B(x_{i}^{(j)},\delta^{-1}\lambda_{i}^{(j)}) is nice (see [Yin19] for details), there is no control over the integral

Ωi(δ)vi,k,vi,k𝑑Vgi.\int_{\Omega_{i}(\delta)}\langle v_{i,k},v_{i,k^{\prime}}\rangle dV_{g_{i}}.

Therefore, the most important step in the proof of Theorem 1.5 is to show that

limδ0limiΩi(δ)vi,k,vi,k𝑑Vgi=0,\lim_{\delta\to 0}\lim_{i\to\infty}\int_{\Omega_{i}(\delta)}\langle v_{i,k},v_{i,k^{\prime}}\rangle dV_{g_{i}}=0,

which is a consequence of

(T1) the volume of Ωi(δ)\Omega_{i}(\delta) with respect to gig_{i} goes to zero when δ0\delta\to 0;

(T2) there is some constant C>0C>0 independent of ii such that

supΩi(δ0/16)|vi,k|C.\sup_{\Omega_{i}(\delta_{0}/16)}\left|v_{i,k}\right|\leq C.

5.2. Metric gig_{i} on the generalized neck domain

For the definition of gig_{i}, we first define a metric on 2\mathbb{R}^{2} as follows

(64) gb=f(r)(dr2+r2dθ2)g_{b}=f(r)(dr^{2}+r^{2}d\theta^{2})

where (r,θ)(r,\theta) is the polar coordinates and

f(r)={(11+r2)2r11r4r>2.f(r)=\left\{\begin{array}[]{ll}\left(\frac{1}{1+r^{2}}\right)^{2}&r\leq 1\\ \frac{1}{r^{4}}&r>2.\end{array}\right.

This is supposed to be the limit of gig_{i} on each real bubble domain and it is to be connected to the Σ\Sigma by g~i\tilde{g}_{i} defined on the generalized neck domain, which is defined as (see (2))

g~i=(1+j=1l(λi(j))2|zxi(j)|4)dzdz¯.\tilde{g}_{i}=\left(1+\sum_{j=1}^{l}\frac{(\lambda_{i}^{(j)})^{2}}{\left|z-x_{i}^{(j)}\right|^{4}}\right)dz\wedge d\bar{z}.

With a cut-off function φ:[0,+)[0,1]\varphi:[0,+\infty)\to[0,1] satisfying φ(x)0\varphi(x)\equiv 0 for s1s\leq 1 and φ(s)1\varphi(s)\equiv 1 for s2s\geq 2, we define

(65) gi(z)={g(z)onΣB(ci,δ0/2)φ(4|zci|δ0)g+(1φ(4|zci|δ0))g~ionB(ci,δ0/2)B(ci,δ0/4)g~i(z)onΩi(δ0/4)φ(|zxi(j)|δ02λi(j))g~i+(1φ(|zxi(j)|δ02λi(j)))(Ll(j))gbonB(xi(j),4δ01λi(j))B(xi(j),2δ01λi(j))(Li(j))gbonB(xi(j),2δ01λi(j))g_{i}(z)=\left\{\begin{array}[]{ll}g(z)&\text{on}\quad\Sigma\setminus B(c_{i},\delta_{0}/2)\\ \varphi(\frac{4\left|z-c_{i}\right|}{\delta_{0}})g+(1-\varphi(\frac{4\left|z-c_{i}\right|}{\delta_{0}}))\tilde{g}_{i}&\text{on}\quad B(c_{i},\delta_{0}/2)\setminus B(c_{i},{\delta_{0}/4})\\ \tilde{g}_{i}(z)&\text{on}\quad\Omega_{i}(\delta_{0}/4)\\ \varphi(\frac{\left|z-x_{i}^{(j)}\right|\delta_{0}}{2\lambda_{i}^{(j)}})\tilde{g}_{i}+(1-\varphi(\frac{\left|z-x_{i}^{(j)}\right|\delta_{0}}{2\lambda_{i}^{(j)}}))(L_{l}^{(j)})^{*}g_{b}&\text{on}\quad B(x_{i}^{(j)},4\delta_{0}^{-1}\lambda_{i}^{(j)})\setminus B(x_{i}^{(j)},2\delta_{0}^{-1}\lambda_{i}^{(j)})\\ (L_{i}^{(j)})^{*}g_{b}&\text{on}\quad B(x_{i}^{(j)},2\delta_{0}^{-1}\lambda_{i}^{(j)})\end{array}\right.

where j=1,,lj=1,\cdots,l and Li(j):22L_{i}^{(j)}:\mathbb{R}^{2}\to\mathbb{R}^{2} maps zz to zxi(j)λi(j)\frac{z-x_{i}^{(j)}}{\lambda_{i}^{(j)}}.

With this definition, it is straight forward to show an analog of Lemma 5.2 in [Yin19], from which (T1) follows.

Lemma 5.1.

For any δ(0,δ0)\delta\in(0,\delta_{0}), we have, when ii\to\infty,

(1) gig_{i} converges to gg on ΣB(ci,δ)\Sigma\setminus B(c_{i},\delta);

(2) for each j=1,,lj=1,\cdots,l, ((Li(j))1)gi((L_{i}^{(j)})^{-1})^{*}g_{i} converges to gbg_{b} on B(0,δ1)B(0,\delta^{-1});

(3) The volume of Ωi(δ)\Omega_{i}(\delta) with respect to gig_{i} is bounded by Cδ2C\delta^{2} for some universal constant C>0C>0.

As explained in Section 2.4, g~i\tilde{g}_{i} is the pullback metric of some minimal embedding. In [Yin19], on a simple neck domain (or a cylinder), an explicit parametrization of the catenoid in 3\mathbb{R}^{3} was used. Here, we use the sequence uiu_{i} given in (9). The following mean value inequality is a generalization of Lemma 5.3 of [Yin19].

Lemma 5.2.

For any positive number C1>0C_{1}>0, there is C2C_{2} depending on C1C_{1} but not ii such that if a nonnegative function ww satisfies

giwC1w,on Ωi(δ0/8)\triangle g_{i}w\geq-C_{1}w,\qquad\text{on \quad}\Omega_{i}(\delta_{0}/8)

then for sufficiently large ii,

supΩi(δ0/16)wC2Ωi(δ0/8)w𝑑Vgi.\sup_{\Omega_{i}(\delta_{0}/16)}w\leq C_{2}\int_{\Omega_{i}(\delta_{0}/8)}wdV_{g_{i}}.
Proof.

Recall that the metric gig_{i} restricted to Ωi(δ0/8)\Omega_{i}(\delta_{0}/8) is the same as g~i\tilde{g}_{i} and g~i\tilde{g}_{i} is the pullback metric by u~i\tilde{u}_{i} defined in (9). Since u~i\tilde{u}_{i} parametrizes a minimal surface in l+1\mathbb{C}^{l+1}, the classical mean value inequality (see Corollary 1.16 of [CM11]) implies that for any yu~i(Ωi(δ0/16))y\in\tilde{u}_{i}(\Omega_{i}(\delta_{0}/16)),

w(y)C2B^(y,δ032)u~i(Ωi(δ0/8))w𝑑VΣ,w(y)\leq C_{2}\int_{\hat{B}(y,\frac{\delta_{0}}{32})\cap\tilde{u}_{i}(\Omega_{i}(\delta_{0}/8))}wdV_{\Sigma},

as long as we verify that

(66) B^(y,δ032)u~i(Ωi(δ0/8))=.\hat{B}(y,\frac{\delta_{0}}{32})\cap\tilde{u}_{i}(\partial\Omega_{i}(\delta_{0}/8))=\emptyset.

Here dVΣdV_{\Sigma} is the induced metric on the image of u~i\tilde{u}_{i} and B^\hat{B} is the metric ball in l+1\mathbb{C}^{l+1}. To see that (66) holds for large ii, we consider the limit of u~i(Ωi(δ))\tilde{u}_{i}(\Omega_{i}(\delta)) as a subset in l+1\mathbb{C}^{l+1}, which by (9) and (60) is

Ω~(δ):=j=1,,l+1{(z1,,zl+1)||zj|δ,zk=0forkj}.\tilde{\Omega}(\delta):=\bigcup_{j=1,\cdots,l+1}\left\{(z_{1},\cdots,z_{l+1})|\quad\left|z_{j}\right|\leq\delta,\quad z_{k}=0\quad\text{for}\quad k\neq j\right\}.

Hence, the boundary of Ω~(δ0/8)\tilde{\Omega}(\delta_{0}/8) is

j=1,,l+1{(z1,,zl+1)||zj|=δ0/8,zk=0forkj},\bigcup_{j=1,\cdots,l+1}\left\{(z_{1},\cdots,z_{l+1})|\quad\left|z_{j}\right|=\delta_{0}/8,\quad z_{k}=0\quad\text{for}\quad k\neq j\right\},

whose distance to u~i(Ωi(δ0/16))\tilde{u}_{i}(\Omega_{i}(\delta_{0}/16)) is greater than δ0/32\delta_{0}/32. ∎

With Lemma 5.2, we may prove (T2) as Lemma 5.7 [Yin19]. Notice that in this proof, we used Theorem 1.2 in the form that

supΩi(δ0)uig~iC.\sup_{\Omega_{i}(\delta_{0})}\left\|\nabla u_{i}\right\|_{\tilde{g}_{i}}\leq C.

With (T1) and (T2), the rest of the proof is the same as in [Yin19].

Appendix A Some properties of an ODE solution

In this appendix, we show some elementary properties of the solution g(t)g(t) to the ordinary differential equation with boundary values

g′′(t)=γ2g(t),g(0)=aandg(T)=b.g^{\prime\prime}(t)=\gamma^{2}g(t),\qquad g(0)=a\quad\text{and}\quad g(T)=b.

We assume that γ>1/2\gamma>1/2 and T>5T>5. They are not essential and we assume these for simplicity.

Lemma A.1.

There is a universal constant c~\tilde{c} such that for any positive constants aa and bb with bab\geq a, we have

(67) sup[1,T]|(logg)|c~\sup_{[1,T]}\left|(\log g)^{\prime}\right|\leq\tilde{c}

and

(68) ac~inf[0,1]g.a\leq\tilde{c}\inf_{[0,1]}g.
Remark A.2.

In general, since bb may be very large, even a lot larger than eγTae^{\gamma T}a, we can not expect an upper bound of (logg)(\log g)^{\prime} over [0,T][0,T]. The observation is that such an upper bound holds for [1,T][1,T] regardless of the size of aa, bb and TT.

The proof follows from explicit computation, since we have the following formula for the solution

(69) g(t)=ae2γTbeγTe2γT1eγt+beγTae2γT1eγt.g(t)=\frac{ae^{2\gamma T}-be^{\gamma T}}{e^{2\gamma T}-1}e^{-\gamma t}+\frac{be^{\gamma T}-a}{e^{2\gamma T}-1}e^{\gamma t}.

Direct computation shows

(70) g(t)=γae2γTbeγTe2γT1eγt+γbeγTae2γT1eγtg^{\prime}(t)=-\gamma\frac{ae^{2\gamma T}-be^{\gamma T}}{e^{2\gamma T}-1}e^{-\gamma t}+\gamma\frac{be^{\gamma T}-a}{e^{2\gamma T}-1}e^{\gamma t}

and

(71) g(t)g(t)=γ(beγTa)eγt(ae2γTbeγT)eγt(beγTa)eγt+(ae2γTbeγT)eγt.\frac{g^{\prime}(t)}{g(t)}=\gamma\frac{(be^{\gamma T}-a)e^{\gamma t}-(ae^{2\gamma T}-be^{\gamma T})e^{-\gamma t}}{(be^{\gamma T}-a)e^{\gamma t}+(ae^{2\gamma T}-be^{\gamma T})e^{-\gamma t}}.

Taking one more derivative, we get

(72) (gg)=γ24(beγTa)(ae2γTbeγT)((beγTa)eγt+(ae2γTbeγT)eγt)2.\left(\frac{g^{\prime}}{g}\right)^{\prime}=\gamma^{2}\frac{4(be^{\gamma T}-a)(ae^{2\gamma T}-be^{\gamma T})}{((be^{\gamma T}-a)e^{\gamma t}+(ae^{2\gamma T}-be^{\gamma T})e^{-\gamma t})^{2}}.

(i) If aa and bb are comparable in the sense that abeγTaa\leq b\leq e^{\gamma T}a, the lemma holds with c~=γ\tilde{c}=\gamma. To see this, we notice that in this case

ae2γTbeγT,beγTa0.{ae^{2\gamma T}-be^{\gamma T}},{be^{\gamma T}-a}\geq 0.

Hence, by (71), we have |(logg)|γ\left|(\log g)^{\prime}\right|\leq\gamma for all t[0,T]t\in[0,T], from which both (67) and (68) follow.

(ii) If beγTab\geq e^{\gamma T}a, (70) implies that g0g^{\prime}\geq 0. Hence, gg is increasing and (68) follows.

Moreover, (72) implies that (logg)′′0(\log g)^{\prime\prime}\leq 0. Together with the observation

(logg)(T)=γb(e2γT+1)2aeγTb(e2γT1)(0,2γ),(\log g)^{\prime}(T)=\gamma\frac{b(e^{2\gamma T}+1)-2ae^{\gamma T}}{b(e^{2\gamma T}-1)}\in(0,2\gamma),

it suffices to bound (logg)(1)(\log g)^{\prime}(1), which we compute

(logg)(1)\displaystyle(\log g)^{\prime}(1) =\displaystyle= γ(beγTa)eγ(ae2γTbeγT)eγ(beγTa)eγ+(ae2γTbeγT)eγ\displaystyle\gamma\frac{(be^{\gamma T}-a)e^{\gamma}-(ae^{2\gamma T}-be^{\gamma T})e^{-\gamma}}{(be^{\gamma T}-a)e^{\gamma}+(ae^{2\gamma T}-be^{\gamma T})e^{-\gamma}}
=\displaystyle= γbeγ(T+1)+beγ(T1)aeγaeγ(2T1)beγ(T+1)beγ(T1)aeγ+aeγ(2T1)\displaystyle\gamma\frac{be^{\gamma(T+1)}+be^{\gamma(T-1)}-ae^{\gamma}-ae^{\gamma(2T-1)}}{be^{\gamma(T+1)}-be^{\gamma(T-1)}-ae^{\gamma}+ae^{\gamma(2T-1)}}
\displaystyle\leq γbeγ(T+1)+beγ(T1)beγ(T+1)beγ(T1)aeγ.\displaystyle\gamma\frac{be^{\gamma(T+1)}+be^{\gamma(T-1)}}{be^{\gamma(T+1)}-be^{\gamma(T-1)}-ae^{\gamma}}.

Finally, we notice that

aeγbeγT+γ,ae^{\gamma}\leq be^{-\gamma T+\gamma},

which implies that

(logg)(1)2γ1e2γe2γT4γ.(\log g)^{\prime}(1)\leq\frac{2\gamma}{1-e^{-2\gamma}-e^{-2\gamma T}}\leq 4\gamma.

This concludes the proof of the lemma.

References

  • [AY17] Wanjun Ai and Hao Yin. Neck analysis of extrinsic polyharmonic maps. Annals of Global Analysis and Geometry, 52(2):129–156, 2017.
  • [CM11] Tobias H Colding and William P Minicozzi. A course in minimal surfaces, volume 121. American Mathematical Soc., 2011.
  • [CS19] Bo Chen and Chong Song. Isolated Singularities of Yang-Mills-Higgs fields on surfaces. arXiv preprint arXiv:1907.07092, 2019.
  • [DT95] Weiyue Ding and Gang Tian. Energy identity for a class of approximate harmonic maps from surfaces. Communications in analysis and geometry, 3(4):543–554, 1995.
  • [LW98] Fanghua Lin and Changyou Wang. Energy identity of harmonic map flows from surfaces at finite singular time. Calc. Var. Partial Differential Equations, 6(4):369–380, 1998.
  • [LY16] Lei Liu and Hao Yin. Neck analysis for biharmonic maps. Math. Z., 283(3-4):807–834, 2016.
  • [Par96] Thomas H. Parker. Bubble tree convergence for harmonic maps. J. Differential Geom., 44(3):595–633, 1996.
  • [QT97] Jie Qing and Gang Tian. Bubbling of the heat flows for harmonic maps from surfaces. Comm. Pure Appl. Math., 50(4):295–310, 1997.
  • [Rad93] Johan Rade. Decay estimates for yang-mills fields: two new proofs. Global analysis in modern mathematics (Orono, 1991, Waltham, 1992), Publish or Perish, Houston, pages 91–105, 1993.
  • [Sam78] J.H. Sampson. Some properties and applications of harmonic mappings. Ann. Sci. École Norm. Sup., 11(4):211–228, 1978.
  • [Sim83] Leon Simon. Asymptotics for a class of nonlinear evolution equations, with applications to geometric problems. Ann. of Math. (2), 118(3):525–571, 1983.
  • [Yin19] Hao Yin. Higher order neck analysis of harmonic maps and its applications. arXiv preprint arXiv:1904.07354, 2019.