A Weak Energy Identity and the Length of Necks for a Sequence of

43
A Weak Energy Identity and the Length of Necks for a Sequence of Sacks-Uhlenbeck α-Harmonic Maps Yuxiang Li * Youde Wang Abstract In this paper we discuss the convergence behavior of a sequence of α-harmonic maps u α with E α (u α ) <C from a closed surface (M,g) into a compact Riemannian manifold (N,h) without boundary. Generally, such a sequence converges weakly to a harmonic map in W 1,2 (M,N ) and strongly in C away from a finite set of points in M . If energy con- centration phenomena appears, we show a generalized energy identity and discover a direct convergence relation between the blow-up radius and the parameter α which ensures the energy identity and no-neck property. Moreover, a length formula for the necks is derived L = r E(v) π log ν, where ν = lim inf α1 (λ α ) - α-1 and v is the bubble. We also give an example of neck which shows the neck contains at least one geodesic of infinite length. Mathematics Subject Classification: 58E20, 35J60. 1 Introduction Let (M,g) be a smooth closed Riemann surface, and (N,h) be an n-dimensional smooth compact Riemannian manifold without boundary. By Nash’s isometric embedding theorem, we assume that N is isometrically embedded in some Euclidean space N, R K for the sequel. We define the Sobolev space of W 1,2 -maps from M into N , denoted by W 1,2 (M,N ), as W 1,2 (M,N )= {u W 1,2 (M, R K ): u(x) N for a. e. x M }. If u W 1,2 (M,N ), we define the energy density e(u) of u by e(u) = Trace g u * h = |∇ g u| 2 , where u * h is the pull-back of the metric tensor h. In a local coordinate system of x M , the energy density can be expressed as e(u)(x)= g ij (x)h αβ (u(x)) ∂u α ∂x i ∂u β ∂x j . * This paper was written while the first author was researching at Mathematisches Institut, Albert-Ludwigs- Universit¨ at Freiburg, supported by Alexander von Humboldt Foundation. Partially supported by 973 project of China, Grant No. 2006CB805902, and NSFC, Grant No. 10990013. 1

Transcript of A Weak Energy Identity and the Length of Necks for a Sequence of

A Weak Energy Identity and the Length of Necks for aSequence of Sacks-Uhlenbeck α-Harmonic Maps

Yuxiang Li∗ Youde Wang†

Abstract

In this paper we discuss the convergence behavior of a sequence of α-harmonic mapsuα with Eα(uα) < C from a closed surface (M, g) into a compact Riemannian manifold(N, h) without boundary. Generally, such a sequence converges weakly to a harmonic mapin W 1,2(M, N) and strongly in C∞ away from a finite set of points in M . If energy con-centration phenomena appears, we show a generalized energy identity and discover a directconvergence relation between the blow-up radius and the parameter α which ensures theenergy identity and no-neck property. Moreover, a length formula for the necks is derived

L =

√E(v)

πlog ν,

where ν = lim infα→1(λα)−√

α−1 and v is the bubble.We also give an example of neck which shows the neck contains at least one geodesic of

infinite length.

Mathematics Subject Classification: 58E20, 35J60.

1 Introduction

Let (M, g) be a smooth closed Riemann surface, and (N, h) be an n-dimensional smooth compactRiemannian manifold without boundary. By Nash’s isometric embedding theorem, we assumethat N is isometrically embedded in some Euclidean space N → RK for the sequel.

We define the Sobolev space of W 1,2-maps from M into N , denoted by W 1,2(M, N), as

W 1,2(M, N) = u ∈ W 1,2(M,RK) : u(x) ∈ N for a. e. x ∈ M.

If u ∈ W 1,2(M, N), we define the energy density e(u) of u by

e(u) = Tracegu∗h = |∇gu|2,

where u∗h is the pull-back of the metric tensor h. In a local coordinate system of x ∈ M , theenergy density can be expressed as

e(u)(x) = gij(x)hαβ(u(x))∂uα

∂xi

∂uβ

∂xj.

∗This paper was written while the first author was researching at Mathematisches Institut, Albert-Ludwigs-Universitat Freiburg, supported by Alexander von Humboldt Foundation.

†Partially supported by 973 project of China, Grant No. 2006CB805902, and NSFC, Grant No. 10990013.

1

The energy E(u) of u is defined by

E(u) =∫

Me(u)dVg,

and the critical points of E(u) are called harmonic maps. Hence, a harmonic map u satisfiesthe corresponding Euler-Lagrange equation:

τ(u) = ∆u + A(u)(∇u,∇u) = 0,

where A(·, ·) is the second fundamental form of N in RK .Eells and Sampson first employed the heat flow method to approach the existence problems

of harmonic maps and deformed successfully a map from a closed manifold into a manifold withnonpositive sectional curvature into a homotopic harmonic map. Concretely, they consideredthe heat flow of harmonic maps (or the negative gradient flow of the energy functional E(u)):

∂u

∂t= τ(u).

If one can establish the global existence of the above flow with respect to the time variable t,or roughly speaking, the flow flows to infinity smoothly, then we are able to find a sequenceuk = u(x, tk) such that uk converges to a harmonic map as tk → +∞ (see [E-S]). Generally,such a flow does not exist globally (see [D3] and [C-D-Y]).

From the view point of calculus of variation, it is also not easy to find a harmonic map,since E does not satisfy the well-known Palais-Smale condition when the dimensions of domainmanifold dim(M) ≥ 2. In particular, when dim(M) = 2, it is well-known that the energyfunctional E is of conformal invariance and the corresponding variational problem possess anoncompact invariance group and represents limiting cases where the Palais-Smale conditionjust fails. Therefore, harmonic maps from a surface are of special importance and interest. Infact, mathematicians have paid more attention to this case. To prove the existence of harmonicmaps from a closed surface Sacks and Uhlenbeck in their pioneering paper [S-U] introduced aperturbed energy functional which satisfies the Palais-Smale condition, hence obtained so calledα-harmonic maps, as critical points of perturbed functional, to approximate harmonic maps.More precisely, for every u ∈ W 1,2α(M, N) Sacks and Uhlenbeck defined the so called α-energyfunctional Eα as

Eα(u) =∫

M(1 + |∇u|2)αdVg,

which can be regarded as a perturbation of energy functional E, and considered α-harmonicmaps, i.e. the critical points of Eα in W 1,2α(M, N), which satisfy the following equation:

∆guα + (α− 1)∇g|∇guα|2∇guα

1 + |∇guα|2 + A(uα)(duα, duα) = 0. (1.1)

If there is a subsequence uk = uαkwhich converges smoothly as αk → 1, then uαk

will convergeto a harmonic map. Such smooth convergence fails generally, thus, Sacks and Uhlenbeck devel-oped a powerful method and some techniques to investigate the blow-up phenomena for such avariational problem.

Later, Struwe [St] used the heat flow method of Eells and Sampson to approach the existenceproblems for harmonic maps from a closed surface and he obtained almost the same results as

2

in [S-U]. Chang showed the same results as in [St] for the case where the domain manifold is acompact surface with smooth boundary (see [C]).

Generally, the blow-up phenomena might happen for Sacks-Uhlenbeck perturbation or heatflow approach of two dimensional harmonic maps. That is to say, we are only sure that theconvergence is smooth away from finitely many points (which are called blow-up points) to asmooth harmonic map u0, which might be a trivial map. Around a blow-up point p, the energywill concentrate, i.e.

limr→0

limk→+∞

Br(p)|∇uk|2dVg > 0.

And then, there are sequences limk→+∞

xik → p, lim

k→+∞λi

k → 0, i = 1, · · · ,m, such that

uk(xik + λi

kx) → wi in Ckloc(R2 \ Ai),

where all wi are non-trivial harmonic maps from S2 to N , and Ai is a finite set.Naturally, one pays attention to the following two problems: One is whether the energy

identity holds true or not, i.e.

limk→+∞

|∇uk|2dVg =∫

|∇u0|2dVg +m∑

i=1

E(wi)?

The other is what we can say about necks joining bubbles if they exist?

When uk = u(x, tk) is a subsequence of a heat flow for two dimensional harmonic maps, theabove two problems were deeply studied. The energy identities have been proved by Qing [Q](for the case N = Sn), Ding and Tian [D-T] for the general case. Lin-Wang provided anotherproof of the energy identity in [Lin-W]. Furthermore, Qing-Tian [Q-T] proved that there is noneck if blow-up phenomena happens at infinite time and Ding [D2] proved a more general result.It is worthy to point out that Topping [T] provided a surprising example of heat flow blowingup at finite time such that the weak limit is not continuous.

The energy identity for an α-harmonic sequence with bounded energy is still open. It seemsthat the methods used to show the identity for heat flow, or a sequence with tension fields τbounded in L2, are not powerful enough to prove the energy identity for an α-harmonic mapsequence. We think that the key difficulty lies on the identity (2.5) in this paper. Indeed, for asequence with tension fields τ bounded in L2, then from (2.5) we can see easily

∂Br

∣∣∣∣∂uk

∂r

∣∣∣∣2

ds0 − 12

∂Br

|∇0uk|2ds0 = O

(1r

Br

|τ(uk)||∇uk|dVg

)+ O(1),

then the right side of the above identity is bounded, therefore we can derive the energy identityfor heat flow or a sequence with tension fields τ bounded in L2. However, for a sequence ofα-harmonic maps a very bad term as following

α− 1r

Br

(1 + |∇uα|2)α−1|∇uα|2dVg

appears in (2.5). It is difficult to control such a term.The known energy identities for some special α-harmonic sequences were obtained by some

methods which are completely different from the one of [D-T]. Now we would like to mentionthe following cases.

3

If uα is a sequence of minimizing α-harmonic maps, i.e. every uα is the minimizer of Eα,which belongs to the same homotopy class, Chen and Tian [C-T] proved that the necks consistof some geodesics of finite length, and moreover this implies no loss of energy in necks for thesequence, i.e. the energy identity holds true.

Another important case is the energy identity for a minimax sequence. We suppose that Mbe a compact Riemann surface and A be a parameter manifold. Let h0 : M × A → N be acontinuous map. Assume H be the class of all maps which are homotopic to h0, and

βα(H) = infh∈H

supt∈A

Eα(h(·, t)).

We can deduce from Jost’s result [J] that there is at least one sequence uαkwhich attained

βαk(H) satisfies the energy identity as αk → 1. Recently, T. Lamm gave a simple proof of this

energy identity [L] (Also see [C-M] and [Mo]).

In this paper, we will adopt some methods and techniques in [D-T] and [D2] to discussthe energy identity for an α-harmonic map sequence, especially the behavior of necks joiningthe bubbles. However, we can not give a final proof on the energy identity for such a Sacks-Uhlenbeck sequence, instead, we only show a weaker energy identity and give some observationon this subject. On the other hand, we reveal the details of convergence behavior of the necks.Precisely we provide a new method to show that the necks converge to geodesics and obtainthe formula on the length of the geodesics. In particular, we reveal the relations between theblow-up radius and the parameter α which ensures the energy identity and no-neck property.

In order to state our main results, here we need to introduce some basic facts about harmonicmaps and some related notions which are needed for the sequel. For a more detailed discussionof the facts reviewed here, we refer the reader to various articles cited below.

We usually use Br to denote a small ball in R2 centered at the origin 0 and with radius r. Itis well-known that there always exists an isothermal coordinate system in a small neighborhoodof every p ∈ M since dim(M) = 2, i.e. there is a real function ϕ with ϕ(p) = 0 such that themetric can be locally expressed as

ds2M = eϕ((dx1)2 + (dx2)2).

As the energy functional E is invariant under conformal transformations, so, essentially we onlyneed to consider the blow-up phenomena in a small ball Br in Euclidean plane with a metricgiven by g = eϕ((dx1)2 + (dx2)2) with ϕ(0) = 0.

Now, we assume that uα is a sequence of α-harmonic maps from (M, g) to (N, h) withbounded α-energy, i.e.

Eα(uα) < Θ.

Then, by the theory of Sacks and Uhlenbeck, there exists a sequence of αk-harmonic maps uαk

converge to a harmonic map u0 : M → N smoothly away from a finitely many points xi asαk → 1. Furthermore, we suppose that there are n0 bubbles at singular point x1. Then, forj = 1, · · · , n0 there are sequences of points xj

αk ∈ M : k ∈ Z with xjαk → x1 as αk → 1, and

sequences of positive real numbers λjαk : k ∈ Z with λj

αk → 0 as αk → 1, such that maps givenby

vjαk

= uαk(xj

αk+ λj

αkx)

converge in Ckloc(R2 \ p1, p2, · · · , psj) to a non-trivial harmonic map for each j, denoted by

vj : S2 → N, j = 1, 2, · · · , n0.

4

Moreover, one of the following always holds true:

(H1). For any fixed R > 0, BRλiα(xi

α)∩BRλj

α(xj

α) = ∅ whenever (α−1) are sufficiently small.

(H2). λiα

λjα

+ λjα

λiα→ +∞ as α → 1.

Remark 1.1. One is easy to check that if (λiα, xi

α) and (λjα, xj

α) do not satisfy (H1) and (H2),then one is able to find subsequences of λi

α, xiα and λj

α, xjα such that

λiα

λjα

→ λ ∈ (0,∞) andxi

α − xjα

λjα

→ a ∈ R2.

Since

uα(xiα + λi

αx) = uα

(xj

α + λjα

(xi

α − xjα

λjα

+λi

α

λjα

x

)),

we havevi(x) = vj(a + λx),

and then vi and vj are in fact the same bubble.

In order to reveal the relations between the blow-up radius and the parameter α whichensures the energy identity and no-neck property, we need to introduce the following quantities:

µj ≡ lim infα→1

(λjα)2−2α, (1.2)

andνj ≡ lim inf

α→1(λj

α)−√

α−1. (1.3)

Obviously, νj ∈ [1,∞]. µj are bounded and we can infer that µj ∈ [1, µmax] for some positivenumber µmax ≥ 1 (see Remark 1.2). We will see that the quantities µj and νj characterizecompletely the energy identity and the limiting length of necks joining bubbles respectively.

When we encounter the situation at a point there are several bubbles of a sequence ofα-harmonic maps, we need to consider a more general α-energy which is of the following form:

Eα,εα(u) =∫

B(εα + |∇guα|2)αdVgα .

In the beginning of Section 2, it will be explained why one needs to consider such a kind offunctionals.

On the other hand, it is also of interest to consider a sequence of critical points correspondingsuch a family of functional. In fact, given a sequence of uα from (M, g) to (N, h) with Eα,εα(uα) ≤Θ, each of which is a critical point of the corresponding Eα,εα(u), we may also apply the theoryof Sacks and Uhlenbeck to conclude that there exists a sequence of αk-harmonic maps uαk

converge to a harmonic map u0 : M → N smoothly away from a finite point set xi as αk → 1.In this paper, basing on the above motivations we will focus on the following problems:

Given a sequence of maps uα from (Bσ, gα) to (N, h), each of which is a critical point of thecorresponding Eα,εα(u), and Eα,εα(uα) ≤ Θ with εα satisfying

1 ≥ limα→1

εαα−1 > β0 > 0.

5

Here Bσ is a ball in R2, and gα = eϕα((dx1)2 + (dx2)2) with ϕα(0) = 0 and ϕα → ϕ ∈ C∞(Bσ)smoothly. Moreover, without loss of generality we assume that uα → u0 in Ck

loc(Bσ \ 0). Forsuch a sequence uα, does the energy identity hold true? If the necks joining bubbles exist, dothe limiting necks consist of some geodesics? And if so, can we give the length formula of suchgeodesics?

Now we are the position to state our main results. The first task of this paper is to establishthe following weak energy identity:

Theorem 1.1. Let Bσ = Bσ(0) be a ball in R2, gα = eϕα(x)((dx1)2 + (dx2)2) and g0 =eϕ(x)((dx1)2 + (dx2)2) be a family of metrics on Bσ, where ϕα ∈ C∞(Bσ), ϕα(0) = 0, andϕα converges smoothly to ϕ ∈ C∞(Bσ). Assume uα : (Bσ, gα) → N is a sequence of maps withEα,εα(uα, Bσ) < Θ for any 0 < α ≤ 1, each of which is a critical point of the correspondingfunctional Eα,εα(u,Bσ), where εα satisfy

0 < β0 < limα→1

εαα−1 ≤ 1;

moreover, uα → u0 in C∞loc(Bσ \ 0, N) and there are n0 bubbles vj : S2 → N at point 0.

Here Θ and β0 are some positive constants. Then the following energy identity holds true

limδ→0

limα→1

Bδ(0)(εα + |∇gαuα|2)αdVgα =

n0∑

j=1

µ2jE(vj),

where µj is defined by (1.2).

This theorem tells us that the energy identity holds true if and only if µj = 1. It provides anew route to approach the problem whether the necks contain energy or not. Obviously, in thetheorem u0 is a harmonic map from (Bσ, g0) into N . As a direct corollary, we obtain

Theorem 1.2. Let M be a smooth closed Riemann surface and N be a smooth compact Rie-mannian manifold without boundary. Assume that uαk

∈ C∞(M, N) (αk → 1) is a sequenceof αk-harmonic maps with uniformly bounded α-energy, i.e. Eαk

(uαk) ≤ Θ, and x1 be the only

blow-up point of the sequence uαk in Bσ(x1) ⊂ M . Then, passing to a subsequence, there exist

u0 : M → N which is a smooth harmonic map and finitely many bubbles vj : S2 → N such thatuαk

→ u0 weakly in W 1,2(M, N), and in C∞loc(Bσ(x1)\x1, N); and the following identity holds

limk→+∞

Eαk(uαk

, Bσ(x1)) = E(u0, Bσ(x1)) + |Bσ(x1)|+n0∑

j=1

µ2jE(vj), (1.4)

where µj is defined by (1.2) and n0 is the number of bubbles at x1.

Remark 1.2. We claim that µj ∈ [1, µmax] for some positive number µmax ∈ [1,∞). Indeed,fixing an R > 0, we have

BRλ

jαk

(xjαk

)\(∪sji=1B

δλjαk

(xjαk

+λjαk

pi))|∇guαk

|2αkdVg

= (λjαk)2−2α

BR(xjαk

)\(∪sji=1Bδ(xj

αk+pi))

|∇gvjαk|2αkdV

g(xjαk

+λjαk

x).

(1.5)

6

Since∫

BR(xjαk

)\(∪sji=1Bδ(xj

αk+pi))

|∇gvjαk|2αkdV

g(xjαk

+λjαk

x)→

BR\(∪sji=1Bδ(pi))

|∇0vj |2dx,

andλj

αk< 1,

then we have for large R and small δ

µj ≤ limk→∞

BRλ

jαk

(xjαk

)\(∪sji=1B

δλjαk

(xjαk

+pi))|∇gu

jαk|2αkdVg

BR\(∪sji=1Bδ(pi))

|∇0vj |2dx

.

Therefore, there holds

µj ≤ limδ→0

limR→∞

limk→∞

BRλ

jαk

(xjαk

)\(∪sji=1B

δλjαk

(xjαk

+pi))|∇gu

jαk|2αkdVg

BR\(∪sji=1Bδ(pi))

|∇0vj |2dx

≤ Θ− |M | − E(u0)θ

, (1.6)

where |M | denotes the area of M and

θ = infE(u) : u is a nontrivial harmonic map from S2 to N.

Therefore, it follows

µj ∈[1,

Θ− |M | − E(u0)θ

].

It is our another purpose to study the behavior of the necks joining bubbles. Our mainresults are stated as follows:

Theorem 1.3. Let Bσ = Bσ(0) be a ball in R2, gα = eϕα(x)((dx1)2 + (dx2)2) and g0 =eϕ(x)((dx1)2 + (dx2)2) be a family of metrics on Bσ, where ϕα ∈ C∞(Bσ), ϕα(0) = 0, andϕα converges smoothly to ϕ ∈ C∞(Bσ). Assume uα : (Bσ, gα) → N is a sequence of α-harmonicmaps as stated in Theorem 1.1 and there is only one bubble v1 : S2 → N in Bσ. Then we have

1) when ν1 = 1, the set u0(Bσ) ∪ v1(S2) is a connected subset of N ;2) when ν1 ∈ (1,∞), the set u0(Bσ) and v1(S2) are connected by a geodesic with Length

L =

√E(v1)

πlog ν1;

3) when ν1 = +∞, the neck contains at least an infinite length geodesic.Here ν1 is defined by (1.3), i.e. ν1 = lim inf

α→1(λ1

α)−√

α−1.

As a direct corollary, we have

7

Theorem 1.4. Let M be a smooth closed Riemann surface and N be a smooth closed Rie-mannian manifold and uαk

∈ C∞(M, N) be a sequence of αk-harmonic maps with uniformlybounded energy, i.e. Eαk

(uαk) ≤ Θ, and uαk

converges to a smooth harmonic map u0 : M → Nin C∞

loc(Bσ(x1)\x1, N) as αk → 1. Assume there is only one bubble in Bσ(x1) ⊂ M for uαk

and v1 : S2 → N is the bubbling map. Let ν1 = lim infα→1

(λ1α)−

√α−1. Then we have

1) when ν1 = 1, the set u0(Bσ(x1)) ∪ v1(S2) is a connected subset of N ;2) when ν1 ∈ (1,∞), the set u0(Bσ(x1)) and v1(S2) are connected by a geodesic with Length

L =

√E(v1)

πlog ν1;

3) when ν1 = +∞, the neck contains at least an infinite length geodesic.

We should mention that after we completed the paper we found that Moore had proved thatif a neck is of finite length L and g ≥ 1 (the genus of M), then

L =

√E(v1)

πlog ν.

Note that E(u) is defined by 12

∫M |∇u|2dVg in [Mo]. However, the arguments to prove Theorem

1.3 in this paper is completely different from Moore’s proof. The key estimation of us is theProposition 4.1 in section 4, which gives the details of the necks.

We fail to find a sufficient condition such that νi < +∞, but we will show that there areindeed many cases that the necks contain at least one infinite length geodesic:

Corollary 1.5. Let αk → 1, and uk : M → N be a minimizer of Eαkin the homotopic class

containing uk. We assume for any i 6= j, ui and uj are not in the same homotopic class. If

supk

Eαk(uk) < +∞,

then uk will blow up, and the neck contains at least one infinite length geodesic.

Remark 1.3. In the last section, by constructing a manifold N we will give an example of aminimizing α-harmonic map sequence, which satisfies the condition in the above corollary. Thisindicates that there exists a neck joining bubbles which is a geodesic of infinite length.

We conclude this introduction with showing the following proposition as a consequence of The-orem 1.1, which implies the result due to Chen-Tian that, if the necks consist of some geodesicsof finite length, then the energy identity is true:

Proposition 1.6. The energy identity holds true for a subsequence of uα if and only if

lim infα→1

‖∇uα‖α−1C0(M)

= 1. (1.7)

The limit set of such subsequence has no neck if and only if

lim infα→1

‖∇uα‖√

α−1C0(M)

= 1.

The bubbles in limit set of such subsequence are joined by some geodesics of finite length, if andonly if

lim infα→1

‖∇uα‖√

α−1C0(M)

< +∞.

8

Proof. We only prove the first claim.First, we prove (1.7) implies µj = 1. We assume vj

α(x) = uα(xjα + λj

αx) converges to vj inC1

loc(Rn \ p1, p2, · · · , ps). Then we have

(λjα)1−α =

|∇uα(xjα + λj

αx)|α−1

|∇vjα(x)|α−1

for any x with |∇vj(x)| 6= 0. Hence we get µj ≤ 1 and then µj = 1.Now, we will prove “µj = 1 for all j” implies (1.7). Let xα to be the point such that

|∇uα(xα)| = max |∇uα|, and

λα =1

|∇uα(xα)| .

We set vα(x) = uα(xα + λαx). One is easy to check that vα will converge to a non-trivialharmonic map v0 locally. By H1 and H2 we must find a j, such that

BRλj

α(xj

α) ∩BRλα(xα) 6= ∅, and1C

λjα < λα < Cλj

α

for some C > 0. Hence we get |λα|α−1 → 1. 2

Acknowledgement: The authors is grateful to thank Professor W. Ding for his help and encour-agement, and the referee’s beneficial suggestions such that this paper is more readable. The first authorwould like to thank Prof. E. Kuwert for many helpful discussions.

2 Preliminary

In this section we intend to establish some integral formulas on α-harmonic maps from a closedsurfaces by the variations of domain. Of course, we need to choose some suitable variationalvector fields on M which generate the transformations of M . We will see that these integralrelations will play an important role on the proofs of main theorems.

First we give the reasons and motivations to study the functionals Eα,εα stated in Section1. Recall that the functional Eα is not of conformal invariance. For example, on an isothermalcoordinate system around a point p ∈ M , if the metric is given by

g = eϕ((dx1)2 + (dx2)2)

with ϕ(p) = 0 and uα(x) = uα(λx), then we have∫

Bδ(p)(1 + |∇guα|2)αdVg =

B δλ

(p)λ2−2α(λ2 + |∇g′ uα|2)αdVg′ ,

where g′ = eϕ(p+λx)((dx1)2 + (dx2)2). On the other hand, we should also note that a sequenceof α-harmonic maps uα may have several bubbles near a blowing up point, for example, thereare sequences λ1

α and λ2α such that

λ2α

λ1α

→ 0, λ1α → 0,

9

as α → 1, and

v1α(x) = uα(λ1

αx) → v1 in Ckloc(R2), v2

α(x) = uα(λ2αx) → v2 in Ck

loc(R2 \ 0),where v1, v2 are non-trivial harmonic maps from S2 to N. For this case, we have

v2α(x) = v1

α

(λ2

α

λ1α

x

),

i.e. v2(x) is in fact a bubble for the sequence v1α. Therefore, we need to consider the equation

of v2α. It is easy to check that v2

α is locally a critical point of the functional

F (v) =∫

((λ2α)2 + |∇v|2)αdVgα ,

where gα = eϕ(p+λαx)((dx1)2 + (dx2)2). For the above reasons, we need to consider a moregeneral α-energy which is of the following form:

Eα,εα(u) =∫

B(εα + |∇guα|2)αdVgα ,

where B = x : |x| < 1 is a unit ball in R2 and εα ∈ (0, 1]. If uα be the critical point of theabove functional, then, uα satisfies the following elliptic system which is also called α-harmonicmap equation:

∆gαuα + (α− 1)∇gα |∇gαuα|2∇gαuα

εα + |∇gαuα|2 + A(uα)(duα, duα) = 0. (2.1)

Here we always assume that the sequence εα (εα ≤ 1) satisfies

limα→1

εαα−1 > β0 > 0. (2.2)

From (1.6) we know that this assumption is reasonable.

Next, we recall the well-known ε-regularity theorem due to Sacks-Uhlenbeck [S-U]:

Theorem 2.1. Let uα : (B, gα) → N satisfies equation (2.1). Then, there exists ε > 0 andα0 > 1 such that if E(uα, B) < ε and 1 ≤ α ≤ α0, then for all smaller r < 1, we have

‖∇uα‖W 1,p(Br) ≤ C(p, r)E(uα, B),

here Br ⊂ B is a ball with radius r, 1 < p < ∞.

We also have

Lemma 2.2. Let (B, gα) be a unit disk in R2 with metric gα = eϕα((dx1)2 + (dx2)2), whereϕα(0) = 0 and ϕα → ϕ ∈ C∞(B) smoothly. If uα : (B, gα) → N is a critical point of Eα,εα(u,B)given by

Eα,εα(u,B) =∫

B(εα + |∇guα|2)αdVgα ,

where limα→1

εαα−1 > β0 > 0, and Eα,εα(uα) ≤ Θ, then we have

β0 < lim infα→1

‖(εα + |∇guα|2)α−1‖C0(B) ≤ lim supα→1

‖(εα + |∇guα|2)α−1‖C0(B) < β1,

where β1 is independent of α.

10

Proof. Obviously, we only need to prove ‖∇guα‖α−1C0(B)

< C. We assume that there is a sequence

of uαk such that ‖∇guαk

‖αk−1C0(B)

→ +∞ as αk → 1. Suppose that max|∇guαk| is attained at

xαk, i.e.

|∇guαk(xαk

)| = max|∇guαk|.

Let λk = 1|∇guαk

(xαk)| and vk(x) = uαk

(xαk+λkx). Obviously, we have Eαk

(vk) ≤ C < ∞. Then

there exists a subsequence of vkj which converges to a new nontrivial harmonic map from S2

to N . Hence, from (1.6) we can infer λ1−αkj

kj< C, which contradicts with the choice of αk. 2

Remark 2.1. It is worthy to point out that the constant C(p, r) in Theorem 2.1 does not dependon α, εα and gα if gα = eϕα((dx1)2 +(dx2)2), where ϕα(0) = 0 and ϕα → ϕ ∈ C∞(B) smoothly.

Next we need to exploit some general variational formulas for the functional Eα,εα(u,B).By virtue of these variational identities we can derive some important estimates on the energycontained in the necks of the critical points of Eα,εα .

Lemma 2.3. Let (B, gα) be a unit disk in R2 with metric gα = eϕα((dx1)2 + (dx2)2), whereϕα(0) = 0 and ϕα → ϕ ∈ C∞(B) smoothly. If uα is a critical point of Eα,εα(u,B) where1 ≥ lim

α→1εα

α−1 > β0 > 0, then, there holds true for any t < 1

(1− 12α

)∫

∂Bt

(εα + |∇gαuα|2)α−1∣∣∂uα

∂r

∣∣2ds0 − 12α

∂Bt

(εα + |∇gαuα|2)α−1∣∣uα,θ

∣∣2ds0

=(α− 1)

αt

Bt

(εα + |∇gαuα|2)α−1|∇0uα|2dx + O(t),

where ds0 is the volume element of ∂Bt with respect to the Euclidean metric.

Proof. Take an 1-parameter family of transformations φs which is generated by the vectorfield X. If we assume X is supported in B, then we have

Eα,εα(u φs) =∫

B(εα + |∇g(u φs)|2)αdVgα

=∫

B(εα +

β

|d(u φs)(eβ(x))|2)αdVgα

=∫

B(εα +

β

|du(φs∗(eβ(x)))|2)αdVgα

=∫

B(εα +

β

|du(φs∗(eβ(φ−1s (x))))|2)αJac(φ−1

s )dVgα ,

where eα is a local orthonormal basis of TB. Noting

d

dsJac(φ−1

s )dVgα |s=0 = −div(X)dVgα ,

11

by differentiating the above identity we obtain the formula

dEα,εα(u)(u∗(X)) = −∫

B(εα + |∇gαu|2)αdiv(X)dVgα

+2α∑β

B(εα + |∇gαu|2)α−1〈du(∇eβ

X), du(eβ)〉dVgα .

Now, we assume uα to be the critical point of Eα,εα . For any vector field X on B, we have

2α∑

β

B(εα + |∇guα|2)α−1〈duα(∇eβ

X), duα(eβ)〉dVgα =∫

B(εα + |∇guα|2)αdivXdVgα . (2.3)

Next, for 0 < t′ < t ≤ ρ, we choose a vector field X with compact support in Bρ byX = η(r)r ∂

∂r = η(|x|)xi ∂∂xi , where η is defined by

η(r) =

1 if r ≤ t′

t− r

t− t′if t′ ≤ r ≤ t

0 if r ≥ t,

where r =√

(x1)2 + (x2)2. By a direct computation we obtain

∇ ∂∂x1

X = η∂

∂x1+ η′

(x1)2

r

∂x1+ η′

x1x2

r

∂x2+ ηx1Γ1

11

∂x1

+ηx1Γ211

∂x2+ ηx2Γ1

12

∂x1+ ηx2Γ2

12

∂x2.

Then,

∑β〈duα(∇eβ

X), duα(eβ)〉dVgα =⟨

duα(∇ ∂∂x1

X), duα

(∂

∂x1

)⟩dx

+⟨

duα(∇ ∂∂x2

X), duα

(∂

∂x2

)⟩dx

=

(η|∇0uα|2 + η′r

∣∣∣∣∂uα

∂r

∣∣∣∣2

+ O(|x|)|∇0uα|2)

dx, (2.4)

where ∇0 is the Riemannian connection with respect to standard metric.It is also easy to check

div(X) = 2η + rη′ + rη∂ϕ

∂r.

Hence, by substituting the above identity and (2.4) into (2.3) we derive

0 = (2α− 2)∫

Bt

η(εα + |∇gαuα|2)α−1|∇0uα|2dx

+∫

Bt

O(|x|)(εα + |∇gαuα|2)α−1|∇0uα|2dx

12

−2εα

Bt

η(εα + |∇gαuα|2)α−1dVgα +εα

t− t′

Bt\Bt′r(εα + |∇gαu|2)α−1dVgα

+1

t− t′

Bt\Bt′(εα + |∇gαuα|2)α−1

[|∇0uα|2r − 2αr

∣∣∣∣∂uα

∂r

∣∣∣∣2]

dx

−∫

Bt

εα(εα + |∇gαuα|2)α−1rη∂ϕ

∂rdVgα .

Letting t′ → t in the above identity and using Lemma 2.2, we obtain the following∫

∂Bt

(εα + |∇gαuα|2)α−1|∂uα

∂r|2ds0 − 1

∂Bt

(εα + |∇gαuα|2)α−1|∇0uα|2ds0

=(α− 1)

αt

Bt

(εα + |∇gαuα|2)α−1|∇0uα|2dx + O(t), (2.5)

where ds0 is the volume element of ∂Bt with respect to the Euclidean metric.Note the metric gα can be written as gα = eϕα(dr2 + r2dθ2) in the polar coordinate system.

Setuα,θ =

1r

∂uα

∂θ.

Obviously

|∇0uα|2 =∣∣∣∣∂uα

∂r

∣∣∣∣2

+ |uα,θ|2.

Hence, we get from the above identity and (2.5)

(1− 12α

)∫

∂Bt

(εα + |∇gαuα|2)α−1∣∣∂uα

∂r

∣∣2ds0 − 12α

∂Bt

(εα + |∇gαuα|2)α−1∣∣uα,θ

∣∣2ds0

=(α− 1)

αt

Bt

(εα + |∇gαuα|2)α−1|∇0uα|2dx + O(t).(2.6)

Thus we finish the proof of the lemma. 2

We also need to infer a special kind of variational formula which is derived from the variationswith respect to radial direction, i.e. the following Pohozaev identity.

Lemma 2.4. Let (B, gα) be a unit disk in R2 with metric gα = eϕα((dx1)2 + (dx2)2), whereϕα(0) = 0. If uα is a critical point of Eα,εα(u,B), then there holds true for any t < 1

∂Bt

(∣∣∣∣∂uα

∂r

∣∣∣∣2

− |uα,θ|2)

ds0 = −2(α− 1)t

Bt

∇0|∇gαuα|2∇0uα

εα + |∇gαuα|2 r∂uα

∂rdx,

where ∂∂r denotes the radial derivative.

Proof. Denote ∆0 = ∂2

∂(x1)2+ ∂2

∂(x2)2. Since uα is a critical point of Eα,εα(u,B), for each α ≥ 1,

uα satisfies the Euler-Lagrange equation (2.1), i.e.

∆0uα + (α− 1)∇0|∇gαuα|2∇0uα

εα + |∇gαuα|2 + A(uα)(duα, duα) = 0.

13

As in [Lin-W], we multiply the both sides of the above equation with r ∂uα∂r to obtain

Bt

r∂uα

∂r∆0uαdx = −(α− 1)

Bt

∇0|∇gαuα|2∇0u

εα + |∇gαuα|2 r∂uα

∂rdx.

It is easy to see∫

Bt

r∂uα

∂r∆0uαdx =

∂Bt

r|∂uα

∂r|2ds0 −

Bt

∇0(r∂uα

∂r)∇0uαdx.

Since∫

Bt

∇0(r∂uα

∂r)∇0uαdx =

Bt

∇0

(xk ∂uα

∂xk

)∇0uαdx

=∫

Bt

|∇0uα|2dx +∫ t

0

∫ 2π

0

r

2∂(|∇0uα|2)

∂rrdθdr

=∫

Bt

|∇0uα|2dx +12

∂Bt

|∇0uα|2tds0 −∫

Bt

|∇0uα|2dx

=12

∂Bt

|∇0uα|2tds0,

then, we have

∂Bt

(∣∣∣∣∂uα

∂r

∣∣∣∣2

− 12|∇0uα|2

)ds0 = −α− 1

t

Bt

∇0|∇guα|2∇0uα

εα + |∇guα|2 r∂uα

∂rdx. (2.7)

Immediately, it follows

∂Bt

(∣∣∣∣∂uα

∂r

∣∣∣∣2

− |uα,θ|2)

ds0 = −2(α− 1)t

Bt

∇0|∇guα|2∇0uα

εα + |∇guα|2 r∂uα

∂rdx. (2.8)

Thus, we complete the proof. 2

The integral identities in the above two lemmas, ie. (2.6) and (2.8), are two key variationalidentities which will be used repeatedly in the following arguments.

3 The proof of Theorem 1.1

In this section, we make effort to establish a weak energy identity for a sequence of α-harmonicmaps. We will follow the idea of Ding and Tian in [D-T] and apply (2.6) and (2.8) to showTheorem 1.1. We lay emphasis on the arguments on the case only one bubble occurs for thesequence of α-harmonic maps.

14

3.1 The weak energy identity for the case of only one bubble

In this subsection we prove Theorem 1.1 in the case of n0 = 1, where n0 is the number of thebubbles. The general case will be explained in the next subsection.

By the assumptions in Theorem 1.1, we know that uα : (Bσ, gα) → N is a sequence ofα-harmonic maps with

Eα,εα(uα, Bσ) < Θ,

each uα satisfies the corresponding equation (2.1) where

0 < β0 < limα→1

εαα−1 ≤ 1.

Moreover, there exists a harmonic map u0 from (Bσ, g0) to N such that uα → u0 in Ckloc(Bσ\0)

and 0 is the only blow-up point of uα in Bσ. Therefore, by the discussions in Lemma 2.3and Lemma 2.4 we know that for any α > 1, uα satisfy (2.6) and (2.8).

We need to define some important quantities related to the energy on necks. In the sequel,we always pick xα such that

|∇gαuα(xα)| = maxBσ

|∇gαuα|.

Obviously, xα → 0 for uα in Theorem 1.1.Set

Λ = limR→+∞

limα→1

Λα(R),

whereΛα(R) =

BRλα (xα)|∇gαuα|2αdVgα .

In the present situation, µ is written by

µ = limα→1

λ2−2αα .

Then, from (1.6) we can see easily Λ and µ satisfy the following relation

Λ = µE(v).

Moreover, we also have the following fact

Λ = limR→+∞

limα→1

BRλα

(εα + |∇gαuα|2)α−1|∇gαuα|2dVgα . (3.1)

Indeed,

limR→+∞

limα→1

BRλα

(εα + |∇gαuα|2)α−1|∇gαuα|2dVg

= limR→+∞

limα→1

BR

(εαλ2α + |∇gαvα|2)α−1λ2−2α

α |∇0vα|2dx

= µ

R2

|∇0v|2dx = Λ.

15

Furthermore, we claim that for any ε > 0 there exist δ1 and R such that, ∀λ ∈ (Rλα2 , δ1)

where 4δ1 ≤ σ, as α− 1 is small enough there holds∫

B2λ\Bλ(xα)|∇gαuα|2dVgα ≤ ε. (3.2)

Suppose that the claim is false, then there exist αi → 1 and λ′i → 0 satisfying λ′iλαi

→ +∞ suchthat ∫

B2λ′i\Bλ′

i(xαi )

|∇gαuαi |2dVgα ≥ ε. (3.3)

Denote v′αi(x) = uαi(λ

′ix+xαi). Then, there exists v′ such that v′αi

→ v′ in Ckloc(R2\(0∪A), N),

where A is a finite set which does not contain 0. If A = ∅ then it follows from (3.3) that v′

is a nonconstant harmonic sphere which is different from v1. This contradicts the assumptionn0 = 1. Next, if there exists x1 ∈ A, then, by a similar argument with the previous argumentsto get v = v1, we can still obtain a sequence xi → x1, λi → 0, s.t. v′i(xi + λix) converges to aharmonic map v∗. Hence we get uαi(xαi + λi(λαix + xi)) converges to v∗ strongly, and then v∗

is the second harmonic sphere. This shows that the claim (3.2) must be true.

Set

u∗α =12π

∫ 2π

0uα(xα + reiθ)dθ.

One is easy to check that, for any a < b, the following inequality holds true

Bb\Ba(xα)

∣∣∣∣∂u∗α∂r

∣∣∣∣2

dx =∫ b

a

∫ 2π

0

∣∣∣∣12π

∫ 2π

0

∂uα

∂rdθ

∣∣∣∣2

dθrdr

≤ 12π

∫ b

a

(∫ 2π

0|∂uα

∂r|2dθ

∫ 2π

0dθ

)rdr

=∫ b

a

∫ 2π

0

∣∣∣∣∂uα

∂r

∣∣∣∣2

rdrdθ =∫

Bb\Ba(xα)

∣∣∣∣∂uα

∂r

∣∣∣∣2

dx.

(3.4)

By applying (3.2) and Sacks-Uhlenbeck ε-regularity theorem (Theorem 2.1), we have thefollowing

Lemma 3.1. Let Bσ = Bσ(0) is a ball in R2 with metrics gα = eϕα(x)((dx1)2 + (dx2)2), whereϕα ∈ C∞(Bσ), ϕα(0) = 0, and ϕα converges smoothly to a real function ϕ ∈ C∞(Bσ). Assumeuα : (Bσ, gα) → N be a sequence of α-harmonic maps each of which satisfies the correspondingequation (2.1), and Eα,εα(uα, Bσ) < Θ. Then, as α− 1 is small enough, we have for any a andb such that Rλα < a < b < δ ≤ σ

4

Bb\Ba(xα)|∇2

gαuα|r|∇gαuα|dVgα ≤ C

B4b\B a2(xα)

|∇gαuα|2dVgα .

and ∫

Bb\Ba(xα)|∇2

gαuα| · |uα − u∗α|dVgα ≤ C

B4b\B a2(xα)

|∇gαuα|2dVgα ,

where C does not rely on α.

16

Proof. First, we prove the first inequality in the above lemma. Since Eα,εα(uα, Bσ) < Θ, it iseasy to see that there exists a constant C0 such that

E(uα, Bσ) < C0.

We assume that 2Ka ∈ (b, 2b) and set

Di = B2ia \B2i−1a(xα).

We scale Di to B2 \B1, and uα to uα. By Theorem 2.1 (the ε-regularity theory), we have on Di

|∇guα| ≤ 12i−1a

|∇0uα|C0(B2\B1) ≤C1

2i−1a‖∇0uα‖L2(B4\B1/2)

=C1

2i−1a‖∇guα‖L2(Di+1∪Di∪Di−1).

Hence, it follows

‖r∇guα‖C0(Di) ≤ 2‖(2i−1a)∇guα‖C0(Di) ≤ C2‖∇guα‖L2(Di+1∪Di∪Di−1)

Similarly, we have‖r2∇2

guα‖C0(Di) ≤ C ′2‖∇guα‖L2(Di+1∪Di∪Di−1).

Then we have∫

Di

|∇2guα|r|∇guα|dVg ≤ C

Di+1∪Di∪Di−1

|∇guα|2dVg

Di

dVg

r2

≤ C ′∫

Di+1∪Di∪Di−1

|∇guα|2dVg.

Therefore, we get the first inequality in this Lemma. The proof of the second inequality goes toalmost the same. 2

To prove energy identity, it is crucial to estimate the energy contained in necks. We firstestimate the energy in θ- direction, i.e.

∫Bδ\BRλα (xα) |uα,θ|2dx.

Lemma 3.2. For α-harmonic map sequence uα (α → 1) stated in Theorem 1.1, there holdstrue

limδ→0

limR→+∞

limα→1

Bδ\BRλα (xα)|uα,θ|2dx = 0.

Here 4δ ≤ σ.

Proof. We adopt the technique of Sacks-Uhlenbeck [S-U] and [L-W] to show the lemma. Using(3.2) we have

|u∗α(r)− uα(r, θ)| ≤ ε1. (3.5)

We compute

I =∫

Bδ\BRλα (xα)|∇guα|2dVg.

I =∫

Bδ\BRλα (xα)∇0uα∇0(uα − u∗α)dx +

Bδ\BRλα (xα)∇guα∇gu

∗αdVg

17

= −∫

Bδ\BRλα (xα)∆0uα(uα − u∗α)dx +

Bδ\BRλα (xα)∇0uα∇0u

∗αdx

+∫

∂(Bδ\BRλα (xα))

∂uα

∂r(uα − u∗α)ds0

=∫

Bδ\BRλα (xα)A(uα)(duα, duα)(uα − u∗α)dVg

+(α− 1)∫

Bδ\BRλα (xα)

∇g|∇guα|2∇guα

εα + |∇guα|2 (uα − u∗α)dVg

+∫

∂(Bδ\BRλα (xα))

∂uα

∂r(uα − u∗α)ds0 +

Bδ\BRλα (xα)

∂uα

∂r

∂u∗α∂r

dx. (3.6)

On the other hand, noting (3.4) we have

Bδ\BRλα (xα)

∂uα

∂r

∂u∗α∂r

dx ≤√∫

Bδ\BRλα (xα)|∂uα

∂r|2dx

Bδ\BRλα (xα)|∂u∗α

∂r|2dx

≤∫

Bδ\BRλα (xα)|∂uα

∂r|2dx. (3.7)

Hence, by using Lemma 3.1, (3.5), (3.7) and noting the following fact∣∣∣∣∇g|∇guα|2∇guα

εα + |∇guα|2∣∣∣∣ ≤ |∇2

guα|,

we can infer from (3.6)∫

Bδ\BRλα (xα)|∇0uα|2dx ≤

Bδ\BRλα (xα)|∂uα

∂r|2dx + 3C(α− 1)

B4δ

|∇guα|2dVg

+∫

∂Bδ(xα)

∂uα

∂r(uα − u∗α)ds0 −

∂BRλα (xα)

∂uα

∂r(uα − u∗α)ds0

+ε′1

Bδ\BRλα (xα)|∇0uα|2dx,

where ε′1 = ε1‖A‖L∞(M).Since

|∇0uα|2 = |∂uα

∂r|2 + |uα,θ|2,

we have∫

Bδ\BRλα (xα)|uα,θ|2dx ≤ −

∂Bδ(xα)

∂uα

∂r(uα − u∗α)ds0 +

∂BRλα (xα)

∂uα

∂r(uα − u∗α)ds0

+C ′((α− 1) + ε).

18

Keeping (3.2) in mind, we have

limδ→0

limα→1

∂Bδ(xα)

∂uα

∂r(uα − u∗α)ds0 = 0,

andlim

R→+∞limα→1

∂BRλα (xα)

∂uα

∂r(uα − u∗α)ds0 = 0.

Hence, we can see the above inequality implies the conclusion of Lemma 3.2.2

Immediately we infer from Lemma 2.2 and Lemma 3.2

Corollary 3.3. For α-harmonic map sequence uα (α → 1) stated in Theorem 1.1, there holdstrue

limδ→0

limR→+∞

limα→1

Bδ\BRλα (xα)(εα + |∇gαuα|2)α−1|uα,θ|2dx = 0.

Next, we will be concerned with the total energy on necks of uα. We need to study somedifferential relations of some functionals which are related closely to α-energy Eα,εα . Now wedefine these functional as follows

Fα(t) =∫

Bλt

α(xα)

(εα + |∇gαuα|2)α−1|∇0uα|2dx,

Er,α(t) =∫

Bλt

α\B

λt0α

(xα)(εα + |∇gαuα|2)α−1

∣∣∣∣∂uα

∂r

∣∣∣∣2

dx,

andEθ,α(t) =

Bλt

α\B

λt0α

(xα)(εα + |∇gαuα|2)α−1|uα,θ|2dx.

By (2.6), for t ∈ [ε, t0], we have

(1− 12α

)E′r,α −

12α

E′θ,α =

α− 1α

log λαFα(t) + O(λtα log λα).

Then

(1− 12α

)Er,α(t)− 12α

Eθ,α(t) =12

∫ t

t0

[1α

log λ2(α−1)α Fα(t) + O(λt

α log λα)]dt. (3.8)

It is easy to check that the sequences (1− 12α)Er,α(t)− 1

2αEθ,α(t) and Fα(t) are compact inC0([ε, t0]) topology for any ε > 0. Therefore, there exist two functions F and Er which belongto C0([ε, t0]) such that, as α → 1,

Fα → F, Er,α → Er in C0([ε, t0]),

since we have from Corollary 3.3

Eθ,α(t) → 0 in C0([ε, t0]).

19

Lemma 3.4. The functional Er(t) and F (t), associated with sequence uα stated in Theorem1.1, satisfy the following identities

Er(t) = µt0−tF (t0)− F (t0), (3.9)

andlimt0→1

F (t0) = Λ. (3.10)

Proof. By letting α → 1 in (3.8) we infer

Er(t) = − log µ

∫ t

t0

Fdt. (3.11)

By the definitions of Fα, Er,α and Eθ,α we have

Fα(t) = F (t0) + Er,α(t) + Eθ,α(t).

Since Corollary 3.3 claims limα→1

Eθ,α(t) = 0, as α → 1 the above identity leads to

F (t) = Er(t) + F (t0).

Hence, combining the identity and (3.11) yields

Er(t) = − log µ

∫ t

t0

(Er(t) + F (t0))dt.

This implies that Er(t) ∈ C1 and

E′r = −(log µ)(Er + F (t0)).

It followsEr(t) = µt0−tF (t0)− F (t0).

Now we turn to prove (3.10). Integrating (2.5) with respect to t on the interval [Rλα, λt0α ]

we obtain

Fα(t0)−∫

BRλα (xα)(εα + |∇gαuα|2)α−1|∇0uα|2dx

≤ C

t0α\BRλα (xα)

(εα + |∇gαuα|2)α−1|uα,θ|2dx

+C

∫ λt0α

Rλα

α− 1r

dr + C(λt0α −Rλα) (3.12)

From Corollary 3.3 we have

limt0→1

limR→+∞

limα→1

t0α\BRλα (xα)

(εα + |∇gαuα|2)α−1|uα,θ|2dx = 0.

20

Also note the following hold true

limt0→1

limR→+∞

limα→1

∫ λt0α

Rλα

α− 1r

dr = limt0→1

(1− t0)2

log µ = 0.

Thus, taking limit with respect to α, R and t0 in (3.12) leads to

limt0→1

F (t0) = limt0→1

limR→+∞

limα→1

BRλα (xα)(εα + |∇gαuα|2)α−1|∇0uα|2dx.

In view of (3.1), (3.10) follows from the above identity. 2

On the other hand, we also have

Lemma 3.5. For any t ∈ (0, 1), the sequence uα stated in Theorem 1.1 satisfies

limα→1

Bλt

α(xα)

(εα + |∇gαuα|2)α−1|∇gαuα|2dVgα = µ1−tΛ.

Proof. It is easy to verify∫

Bλt

α(xα)

(εα + |∇gαuα|2)α−1|∇gαuα|2dVgα = Er,α(t) + Eθ,α(t) + Fα(t0).

By Corollary 3.3 we have limα→1

Eθ,α(t) = 0. Hence, in view of (3.9) we can deduce the following

limα→1

Bλt

α(xα)

(εα + |∇gαuα|2)α−1|∇gαuα|2dVgα = Er(t) + F (t0) = µt0−tF (t0).

2

Proof of Theorem 1.1. In the present situation (2.5) is written by(

1− 12α

)∫

∂Bt

(εα + |∇gαuα|2)α−1|∇0uα|2ds0 −∫

∂Bt

(εα + |∇gαuα|2)α−1|uα,θ|2dx

=(α− 1)

αt

Bt

(εα + |∇gαuα|2)α−1|∇0uα|2dx + O(t),

Integrating the above equality with respect to t on the interval [λtα, δ] gives

Bδ\Bλtα

(xα)(εα + |∇gαuα|2)α−1|∇0uα|2dx

≤ C

Bδ\Bλtα

(xα)(εα + |∇gαuα|2)α−1|uα,θ|2dx + C

∫ δ

λtα

α− 1r

dr + C(δ − λtα)

Taking the same argument as we proved (3.10), we find that

limδ→0

limt→0

limα→1

Bδ\Bλtα

(xα)(εα + |∇guα|2)α−1|∇gαuα|2dVgα = 0. (3.13)

21

By Lemma 3.5 we have

limt→0

limα→1

Bλt

α(xα)

(εα + |∇gαuα|2)α−1|∇gαuα|2dVgα = µΛ.

The above equality and (3.13) lead to

limδ→0

limα→1

Bδ(xα)(εα + |∇guα|2)α−1|∇gαuα|2dVgα = µΛ = µ2E(v).

Here we have used the fact Λ = µE(v). Noting Lemma 2.2, we conclude

limδ→0

limα→1

Bδ(xα)(εα + |∇gαuα|2)αdVgα = µ2E(v).

So we have completed the proof of Theorem 1.1 in the case that n0 = 1. 2

3.2 The weak energy identity for the case of several bubbles

For the general case that n0 > 1, the proof of Theorem 1.1 can be completed by induction inn0, the number of bubbles.

Proof of Theorem 1.1 in the case n0 > 1.

Proof. By the assumptions of Theorem 1.1, uα ∈ C∞(Bσ, N) is a sequence of α-harmonic maps,each uα of which satisfies

∆0uα + (α− 1)∇0|∇guα|2∇0uα

εα + |∇guα|2 + A(uα)(duα, duα) = 0,

where g = eϕα(dx1⊗dx1 +dx2⊗dx2) with ϕα(0) = 0 and ϕα converges smoothly to ϕ as α → 1.First we show how to pick the first bubble of uα. Let x1

α ∈ Bδ such that

|∇uα(x1α)| = max

|∇uα|,

and λ1α = 1

maxBδ|∇uα| . Then, there exists a harmonic map v1 such that in Ck

loc(R2)

uα(x1α + λ1

αx) → v1.

Set v1α = uα(xα + λ1

αx). Then

∆0v1α + (α− 1)e−ϕα(xα+λ1

αx) ∇0|∇gαv1α|2∇0v

εα(λ1α)2 + |∇g1

αv1α|2

+ A(v1α)(dv1

α, dv1α) = 0,

whereg1α = eϕα(xα+λ1

αx)(dx1 ⊗ dx1 + dx2 ⊗ dx2).

Since ‖∇v1α‖L∞ ≤ 1, v1

α converges to a harmonic map v1 from R2 to N locally and smoothly.As |∇v1(0)| = 1, we know that v1 is not trivial. v1 is the first bubble.

22

Now, we assume there exists another n0− 1 bubbles v2, · · · , vn0 , and sequences xiα, λi

α suchthat for each i

uα(xiα + λi

αx) → vi

in Ckloc(R2 \ Ai), where Ai are some finite sets. No doubt, we may choose

λ1α = min

i∈1,··· ,n0λi

α.

Moreover, for any i 6= j, one of (H1) and (H2) holds.

Setrn0α = max

i|xi

α − x1α|+ |λi

α|.

Without loss of generality, we assume x1α ≡ 0 and rn0

α is attained by the n0-th bubble, i.e.

rn0α = |xn0

α |+ |λn0α |.

Similar to the proof of (3.2), we have for any ε > 0, there are δ0 and R s.t.∫

B2λ\Bλ(xα)|∇gαuα|2dVgα ≤ ε, ∀λ ∈ (Rrn0

α , δ0). (3.14)

Then, using the same arguments as in the above subsection, we infer

limα→1

(εα + |∇gαuα|2)αdVgα = limR→+∞

limα→1

BR

(rn0α )2−2α(εα(rn0

α )2 + |∇gn0α

un0α |2)αdVg

n0α

+|Bδ| limα→1

εα +∫

|∇gαu0|2dVgα , (3.15)

where un0α (x) = uα(rn0

α x + x1α). Moreover, (3.14) implies that all the blow-up points lie in BR

for some R > 0.Denote v0 as the limiting function to which un0

α converges away from the concentration pointsfor un0

α . We have the following cases: (i) v0 is a non-trivial harmonic map. (ii) v0 is trivial.

For case (i), v0 is a bubble, we denote v0 = vn0 . Then λn0α = rn0

α , limα→1

(rn0α )2−2α = µn0 , and

E(v0) = E(vn0). Clearly,

un0α

(xi

α −λi

α

λn0α

x1α +

λiα

λn0α

x

)= un0

α (xiα + λi

αx) → vi.

Therefore, v1, · · · , vn0−1 are also bubbles for un0α . Now we consider the following functional

Eα,εα(λn0α )2(v) ≡

BR

((λn0

α )2εα + |∇gαv|2)αdVgα .

Since there is only n0 − 1 bubbles for the sequence un0α , by induction we have

limα→1

BR

((λn0

α )2εα + |∇0un0α |2

)α−1 |∇gn0α

un0α |2dVg

n0α

= E(v0, BR) +n0−1∑

i=1

limα→1

(λi

α

λn0α

)2−2α

E(vi).

23

Noting∫

BR

(λn0α )2−2α

((λn0

α )2εα + |∇0un0α |2

)α−1 |∇gn0α

un0α |2dVg

n0α

=∫

BRλ

n0α

(εα + |∇gαuα|2)α−1|∇gαuα|2dVgα .

it follows that

limR→+∞

limα→1

BRλ

n0α

(εα + |∇gαuα|2)α−1|∇gαuα|2dVgα = µ2mE(vm) +

m−1∑

i=1

µ2i E(vi). (3.16)

Combining (3.15) and (3.16) leads to

limα→1

(εα + |∇gαuα|2)αdVgα =n0∑

i=1

µ2i E(vi) + |Bδ| lim

α→1εα +

|∇gαu0|2dVgα .

For case (ii), there are at least two concentration points. So, at any concentration point forun0

α , there are at most m− 1 bubbles, and then we can use the induction. Hence, we will get

limα→1

BR

((rn0α )2εα + |∇vα|2)α−1|∇vα|2dVg =

m∑

i=1

µ2i

limα→1

(rn0α )2−2α

E(vi).

Thus, we complete the proof of Theorem 1.1. 2

4 Description and further analysis of the necks

In this section, we always assume there is only one bubble on some small ball Bδ.

4.1 The description of no neck property

In this subsection, we consider the case ν = 1. We will show that there is no neck for this case,i.e. the length of neck is zero. Following the arguments of Ding [D2], we need to give someestimations on neck Bδ \BRλα . A detailed analysis on an integral

Bδ\BRλα

|∇uα|2dx

will be made and an ordinary differential inequality on the integral is derived.

Now we give the proof of Theorem 1.3 for the case ν = 1.

Proof. Since ν = 1, then we have µ = 1 and

limδ→0

limR→+∞

limα→1

Bδ\BRλα (xα)|∇gαuα|2dVgα = 0. (4.1)

24

For simplicity, we assume

P =log δ − log Rλα

log 2

is an integer. For any integer k ∈ [1, P − 1], we set

Qk(t) = B2k+tRλα\B2k−tRλα

(xα),

where t + k ≤ P and k − t ≥ 0.Using the same approximate method as in Section 3.1, we can conclude that on Qk(t) the

following inequality holds∫

Qk(t)|∇0uα|2dx ≤

Qk(t)A(uα)(duα, duα)(uα − u∗α)dx

+C(α− 1)∫

Qk(t+2)|∇0uα|2dx

+∫

∂Qk(t)

∂uα

∂r(uα − u∗α)ds0 +

Qk(t)|∂uα

∂r|2dx.

(4.2)

Next, we will apply Pohozaev identity (2.7) to controll the last term in the above inequality, i.e.∫Qk(t) |∂uα

∂r |2dx. For the sake of convenience, we set

H(r) = −∫

Br(xα)

∇g|∇gαuα|2∇gαuα

εα + |∇gαuα|2 r∂uα

∂rdVgα = −

Br(xα)

∇0|∇gαuα|2∇0uα

εα + |∇gαuα|2 r∂uα

∂rdx.

Using Lemma 3.1, we have

|H(r)| ≤∫

Br\BRλα (xα)|∇2

gαuα|r|∂uα

∂r|dx + |H(Rλα)|

≤ C

B4δ(xα)|∇0uα|2dx + |H(Rλα)| < C ′, (4.3)

where we use the fact

limα→1

|H(Rλα)| ≤ limα→1

BRλα (xα)|∇2

gαuα|r|∇gαuα| =

BR

|∇20vα|r|∇0vα|dx < C(R).

Therefore, combining (4.2), (4.3) and (2.7) we obtain

Qk(t)

∣∣∂uα

∂r

∣∣2dx− 12

Qk(t)

∣∣∇0u∣∣2dx ≤ C

∫ 2k+tRλα

2k−tRλα

α− 1r

dr ≤ C(α− 1)t.

It follows (4.2) and the above inequality

(12− ε1)

Qk(t)|∇0uα|2dx ≤ C(α− 1)(t + 1) +

∂Qk(t)

∂uα

∂r(uα − u∗α)ds0. (4.4)

On the other hand, we have

∣∣∣∫

∂Qk(t)

∂uα

∂r(uα − u∗α)ds0

∣∣∣ ≤√∫

∂Qk(t)|∂uα

∂r|2ds0

∂Qk(t)|uα − u∗α|2ds0

25

≤√∫

∂Qk(t)|∂uα

∂r|2ds0

∂Qk(t)|uα,θ|2r2ds0

≤ 12

[∫

∂Qk(t)|∂uα

∂r|2rds0 +

∂Qk(t)|uα,θ|2rds0

]

=12

∂Qk(t)r|∇uα|2ds0

= 2t+k−1Rλα

∂B2t+kRλα

(xα)|∇0uα|2ds0

−2k−t−1Rλα

∂B2k−tRλα

(xα)|∇0uα|2ds0. (4.5)

Letfk(t) =

Qk(t)|∇uα|2dx.

From (4.5) we know ∫

∂Qk(t)

∂uα

∂r(uα − u∗α)ds0 ≤ 1

2 log 2f ′k(t).

Hence, by combining (4.4) and the above inequality we have

(1− 2ε1)fk(t) ≤ 1log 2

f ′k(t) + C(α− 1)(t + 1).

Multiplying the two sides of the above inequality by 2−(1−2ε1)t and integrating we obtain

fk(1) ≤ C2−(1−2ε1)t1fk(t1) + C(α− 1).

It is easy to check that, if we set

t1 = Lk =

k if 2k − 1 ≤ PP − k if 2k − 1 > P

then, we get√

E(uα, Qk(1)) ≤ C2−aLk√

E(uα, Bδ \BRλα(xα)) + C√

α− 1

for some positive a and C.Since uα satisfies (2.1), by the standard Lp estimate, we have

oscB2k+1Rλα

\B2k−1Rλα

(xα)uα ≤ C2−aLk√

E(uα, Bδ \BRλα(xα)) + C√

α− 1.

These inequalities imply

oscBδ\BRλα (xα)uα ≤ C√

E(uα, Bδ \BRλα(xα))∑

2−aLk + C√

α− 1P

≤ C√

E(uα, Bδ \BRλα(xα)) + C(R, δ)√

α− 1 + C log λ−√

α−1α .

26

Letting α → 1, and then R → +∞, δ → 0, we get

oscBδ\BRλα (xα)uα → 0.

Thus we proved Theorem 1.3 in the case ν = 1. 2

4.2 Length formula of the neck when ν > 1

The goal of this section is to show the neck converges to a geodesic in N and furthermorecalculate the length of the geodesic.

4.2.1 Some estimates on the neck

In order to analyze the convergence behavior of the neck, we need to consider the behaviors ofuα (in Theorem 1.3) on ∂Bλt

α(xα) with t ∈ [t2, t1], where 0 < t2 < t1 < 1. By the arguments

in Lemma 3.5, we have

limα→1

Bλt

α(xα)

(εα + |∇gαuα|2)α−1|∇gαuα|2dVgα → µ2−tE(v1)

in C0([t2, t1]). This implies

limα→1

B2λt

α\B 1

2 λtα

(εα + |∇gαuα|2)α−1|∇gαuα|2dVgα → 0.

Then, we apply Lemma 2.2 to the above equality to derive

limα→1

B2λt

α\B 1

2 λtα

(xα)|∇gαuα|2dVgα → 0

in C0([t2, t1]). Therefore, for any t ∈ [t2, t1], we have

osc∂Bλt

α(xα)uα ≤ C

B2λt

α\B 1

2 λtα

(xα)|∇gαuα|2dVgα → 0, (4.6)

i.e. uα|∂Bλt

α(xα) will subconverge to a point belonging to N . Especially, we have that, as α → 1,

uα(∂Bλ

t1α

) → y1 ∈ N and uα(∂Bλ

t2α

) → y2 ∈ N.

For simplicity, we will use “(r, θ)” to denote “xα + r(cos θ, sin θ)”. Now we can state themain results of this subsection as follows:

Proposition 4.1. Let uα be a sequence stated in Theorem 1.3. When ν > 1 and 0 < t2 ≤tα ≤ t1 < 1, we have

limα→1

1α− 1

BRλ

tαα\B 1

tαα

(xα)|uα,θ|2dx = 0 (4.7)

for any R > 0, and1√

α− 1

(uα(λtα

α r, θ)− uα(λtαα , 0)

)→ ~a log r

27

strongly in Ck(S1×[ 1R , R],Rn), where θ is the angle parameter of the ball centered at xα, ~a ∈ TyN

is a vector in Rn with

|~a| = µ1− lim

α→1tα

√E(v)

π,

and y = limα→1

uα(λtαα , θ).

To prove Proposition 4.1, first we need to prove the following

Lemma 4.2. When ν > 1 and 0 < t2 ≤ tα ≤ t1 < 1, we have

limα→1

1α− 1

BRλ

tαα\B 1

tαα

(xα)|uα,θ|2dx < C (4.8)

where C does not depend on R.

Proof. We setQ(t) = B2tλtα

α(xα) \B2−tλtα

α(xα).

Here we assume 2t ≤ λ−εα , where

ε < mint2, 1− t1.Applying (2.7), we get from (4.2) the following

(12− ε1)

Q(t)|∇0uα|2dx ≤ (α− 1)

∫ 2tλtαα

2−tλtαα

1rH(r)dr + C

B2t+2λ

tαα\B

2−t−1λtαα

(xα)|∇0uα|2dx

−∫

∂Q(t)

∂uα

∂r(uα − u∗α)ds. (4.9)

For any r ∈ [λtα+εα , λtα−ε

α ], it is easy to check that

|H(r)−H(λtαα )| ≤

tα−εα

\Bλ

tα+εα

(xα)|∇2

gαuα||r∂uα

∂r|dx.

Using Lemma 3.1, we can get∫

tα−εα

\Bλ

tα+εα

(xα)|∇2

gαuα||r∂uα

∂r|dx ≤ C

B2λ

tα−εα

\B 12 λ

tα+εα

(xα)|∇0uα|2dx.

By integrating (2.5) we obtain

(1− 12α

)∫ 2λtα−ε

α

12λtα+ε

α

ds

∂Bs(xα)(εα + |∇gαuα|2)α−1

∣∣∂uα

∂r

∣∣2ds0

− 12α

∫ 2λtα−εα

12λtα+ε

α

ds

∂Bs(xα)(εα + |∇gαuα|2)α−1 1

s2

∣∣∂uα

∂θ

∣∣2ds0

=∫ 2λtα−ε

α

12λtα+ε

α

((α− 1)

αs

Bs(xα)(εα + |∇gαuα|2)α−1|∇0uα|2dx

)ds + O(λ2(tα−ε)

α ). (4.10)

28

By Corollary 3.3 we know that the second term on the left hand side of the above inequalityvanishes as α → 0. On the other hand, noting the fact (εα + |∇gαuα|2)α−1 is bounded, we have

∫ 2λtα−εα

12λtα+ε

α

((α− 1)

αs

Bs(xα)(εα + |∇gαuα|2)α−1|∇0uα|2dx

)ds

≤ C

∫ 2λtα−εα

12λtα+ε

α

(α− 1)αs

ds =Cε

αlog λ2−2α

α .

Noting the assumption limα→1 εα−1α = β0 > 0, we can infer from (4.10) that, as α is close to 1

enough, ∫

B2λ

tα−εα

\B 12 λ

tα+εα

(xα)|∇0uα|2dx ≤ Cε(log λ2−2α

α + 1) + O(λ2(tα−ε)α ).

So, when α is close to 1 enough we can always choose ε such that

C

B2λ

tα−εα

\B 12 λ

tα+εα

(xα)|∇0uα|2dx ≤ Cε(log µ + 1) < ε1.

Hence, we haveH(λtα

α )− ε1 ≤ H(r) ≤ H(λtαα ) + ε1. (4.11)

Letf(t) =

Q(t)|∇gαuα|2dVgα =

Q(t)|∇0uα|2dx.

By using a similar estimate with (4.5) and (4.11), we infer that as α is close to 1 enough thereholds

(1− 2ε1)f(t) ≤ 1log 2

f ′(t) + (α− 1)(at + ε1),

wherea = 4 log 2H(λtα

α ) + ε1.

Then, it is easy to see

(2−(1−2ε1)tf)′ ≥ −(α− 1)(at + ε1)2−(1−2ε1)t log 2.

Hence, we get

f(t) ≤ 2−(1−2ε1)(τ−t)f(τ) +α− 1

1− 2ε1

(ε1 + at +

a

log 2− aτ2−(1−2ε1)(τ−t) − a

2−(1−2ε1)(τ−t)

(1− 2ε1) log 2

).

Then, it follows

f(k) ≤ C1(k)2−(1−2ε1)τf(τ) +α− 1

1− 2ε1(ε1 + ak +

a

log 2+ aC2(k)aτ2−(1−2ε1)τ ).

Let 2τ = λ−εα . Then

B2kλ

tαα\B 1

2kλ

tαα

(xα)|∇0uα|2dx ≤ C(k)λε(1−2ε1)

α +α− 1

1− 2ε1(H(λtα

α )4k log 2 +a

log 2

+C(k)λε(1−2ε1)α log λα).

(4.12)

29

On the other hand, by (2.8) and (4.11), we get∫

B2kλ

tαα\B 1

2kλ

tαα

(xα)(|∂uα

∂r|2 − |uα,θ|2)dx ≥ (α− 1)4k log 2(H(λtα

α )− ε1). (4.13)

Therefore, subtracting (4.12) by (4.13) we obtain

2∫

B2kλ

tαα\B 1

2kλ

tαα

(xα)|uα,θ|2dx ≤ C(k)λε(1−2ε1)

α +(α− 1)1− 2ε1

(2ε1H(λtαα )4k log 2 +

a

log 2

+C(k)λε(1−2ε1)α log λα) + ε1(α− 1)4k log 2.

(4.14)

Sinceν = lim

α→1λ−

√α−1

α > 1,

it is easy to see that, for any m > 0,

λε(1−2ε1)α = o((α− 1)m). (4.15)

Then, noting (4.15) and letting ε1 → 0 in the above inequality (4.14), we get

limα→1

1α− 1

B2kλ

tαα\B 1

2kλ

tαα

(xα)|uα,θ|2dx ≤ a′

2 log 2,

where a′ is a constant which does not depend on R. Thus, we finish the proof of the lemma. 2

Now, we are in the position to give the proof of Proposition 4.1.

Proof. First we show (4.7). Since lemma 4.2 says

B2kλ

tαα\B 1

2kλ

tαα

(xα)

|uα,θ|2α− 1

dx =∫ 2kλtα

α

2−kλtαα

1r

(∫ 2π

0

|∂uα∂θ |2

α− 1dθ

)dr ≤ a′

2 log 2,

for any ε > 0, we can always find k0, which is independent of α, and Lα ∈ [2k0 , 22k0 ] such that

1α− 1

∂BLαλ

tαα

(xα)|uα,θ|2rds =

1α− 1

∫ 2π

0

∣∣∣∣∂uα

∂θ(Lαλtα

α , θ)∣∣∣∣2

dθ < ε,

and1

α− 1

∂B 1Lα

λtαα

(xα)|uα,θ|2rds =

1α− 1

∫ 2π

0

∣∣∣∣∂uα

∂θ

∣∣∣∣2

dθ < ε.

Then

∣∣∣∫

∂Q(log Lα/ log 2)

∂uα

∂r(uα − u∗α)ds

∣∣∣ ≤√∫

∂Q(log Lα/ log 2)r

∣∣∣∣∂uα

∂r

∣∣∣∣2

ds

∫ 2π

0

∣∣∣∣∂uα

∂θ

∣∣∣∣2

≤√

ε(α− 1)∫

∂Q(log Lα/ log 2)r

∣∣∣∣∂uα

∂r

∣∣∣∣2

ds.

30

From Lemma 2.2 and (2.5), we get∫

∂Q(log Lα/ log 2)r∣∣∂uα

∂r

∣∣2ds ≤ C

∂Q(log Lα/ log 2)r|uα,θ|2ds + C(α− 1) + Cλtα

α

≤ (C + ε)(α− 1).

By (4.9) and (4.11), we get

(1− 2ε1)∫

BLαλ

tαα\B 1

Lαλ

tαα

(xα)|∇0uα|2dx ≤ ε1(α− 1) + 2(α− 1)(2H(λtα

α ) log Lα + ε). (4.16)

Noting (4.11) we can infer from (2.8)

1α− 1

BLαλ

tαα\B 1

Lαλ

tαα

(xα)

(∣∣∂uα

∂r

∣∣2 − |uα,θ|2)dx =

∫ Lαλtαα

1Lα

λtαα

2rH(r)dr,

which implies that

1α− 1

BLαλ

tαα\B 1

Lαλ

tαα

(xα)

(∣∣∂uα

∂r

∣∣2 − |uα,θ|2)dx ≥ 4 log Lα(H(λtα

α )− ε).

Combining (4.16) with the above inequality we conclude the following inequality holds true asα is close to 1 sufficiently

1α− 1

BLαλ

tαα\B 1

Lαλ

tαα

(xα)|uα,θ|2dx ≤ ε1 + Cε.

Thus, we have shown (4.7).

Next, we turn to proving the remaining assertions of Proposition 4.1.For tα ⊂ [t2, t1], we assume

uα(∂Bλtαα

) → y, α → 1.

As N is regarded as an embedded submanifold in RK , for simplicity, we may assume y = 0 ∈ Nand TyN = Rn, where RK = Rn × RK−n. We also let λ′α = λtα

α , x′α = (λ′α, 0) + xα and

u′α(x) = uα(λ′αx + xα), vα(x) =1√

α− 1[uα(λ′αx + xα)− uα(x′α)].

By (4.12) and Theorem 2.1, we get

‖∇u′α‖C0(B2k\B2−k ) + ‖∇2u′α‖C0(B

2k\B2−k ) < C(k)√

α− 1,

and then‖∇vα‖C0(B

2k\B2−k ) + ‖∇2vα‖C0(B2k\B2−k ) < C(k).

Noting that vα(1, 0) = 0, we get

‖vα‖C0(B2k\B2−k ) < C ′(k).

31

Obviously, we have the equation:

∆0vα +√

α− 1(A(y) + o(1))(dvα, dvα) + (α− 1)O(|∇2vα|) = 0,

hence, the sequencevα −→ v0 in Ck

loc(R2 \ 0)

where v0 satisfies∆0v0 = 0 with v0 = v0(|x|).

Setv = (a1, a2, · · · , an, 0, · · · , 0) log r.

We deduce from (2.6) that∫

∂Bt

(εα + |∇uα|2)α−1|∇0vα|2ds0 =2α

2α− 1

∂Bt

(εα + |∇uα|2)α−1|vα,θ|2ds0

+2

(2α− 1)t

Bt

(εα + |∇uα|2)α−1|∇0uα|2dx.

Recalling that

Fα(t) =∫

Bλt

α

(εα + |∇gαuα|2)α−1|∇0uα|2dx

and keeping (4.7) in our minds, we infer from the above identity and Lemma 2.2 that, as α → 1,∫

B2λ′α\Bλ′α

(εα + |∇gαuα|2)α−1|∇0vα|2dx =2α

2α− 1

∫ 2λtαα

λtαα

1tFα(logλα

t)dt + o(1)

=2α

2α− 1log 2Fα(tα) + o(1)

→ 2 log 2F ( limα→1

tα).

On the other hand, we have that, as α → 1, there holds∫

B2λ′α\Bλ′α

(εα + |∇gαuα|2)α−1|∇0vα|2dx =∫

B2\B1

(εα + |∇gαvα|2 α− 1

λ′2α

)α−1

|∇0vα|2dx

→ 2πµlim

α→1tα |a|2 log 2.

Hence, we getlimα→1

vα = (a1, · · · , an, 0, · · · , 0) log r

withm∑

i=1

a2i =

Λπ

µ1−2 lim

α→1tα

.

As v : S2 −→ N is the corresponding only bubble, then the above identity can be written as

|~a|2 =E(v, S2)

πµ

2−2 limα→1

tα.

Thus, we complete the proof of Proposition 4.1. 2

32

Corollary 4.3. Let uαkbe a subsequence satisfying

Eαk(uαk

, Bλtαk

(xα)) → µ2−tE(v)

in C0([t2, t1]) with respect to C0-norm. If ν > 1, then

∫ 2λtαk

λtαk

1√α− 1

∣∣∂uαk

∂r

∣∣dr → log 2µ1−t

√E(v)

πin C0([t2, t1]);

and1√

α− 1(r

∣∣∂uαk

∂r

∣∣)(λtα, θ) → µ1−t

√E(v)

πin C0([t2, t1]).

Proof. We need only to prove the first claim, since the proof of the second claim is similar. Ifthe first claim was not true, then we assumed that there was a subsequence αki

, ti → t0 s.t.∣∣∣∣∣∫ 2λ

tiαki

λtiαki

1√α− 1

∣∣∂uαki

∂r

∣∣dr − log 2µ1−ti

√E(v)

π

∣∣∣∣∣ ≥ ε > 0.

On the other hand, from the above arguments on Proposition 4.1 we know that, after passingto a subsequence, there holds

uαki(λαki

x)− uαki(λαki

, 0)√α− 1

→ ~a log r,

with |~a| =∣∣∣µ1−t0

√E(v)

π

∣∣∣. Hence we derive the following

limi→+∞

∫ 2λtiαki

λtiαki

1√α− 1

∣∣∣∣∂uαki

∂r

∣∣∣∣ dr = |~a|∫ 2

1

1rdr = log 2µ1−t0

√E(v)

π.

This is a contradiction. 2

4.2.2 The proof Theorem 1.3 for the case ν > 1

First, we need to show the neck for the α-harmonic map sequence converge to a geodesic in Nwhich join N and the bubble. For this goal, we define the following curve

ωα(r) ≡ 12π

∫ 2π

0uα(r, θ)dθ : [λt1

α , λt2α ] −→ RK ,

which is denoted by Γα. For the sake of convenience, we denote the derivatives ∂f(r,θ)∂r and ∂f(r,θ)

∂θ

of f(r, θ) by f and f,θ respectively.

Lemma 4.4. Assume that uα be the sequence stated in Theorem 1.3. Then, the curve familyΓα in RK , each of which is defined by ωα(r) and parameterized by its arc length, converges to ageodesic on (N, h).

33

Proof. From (4.6), it is easy to see that for any rk → 0, ωα(rk) converges to a point p ∈ N afterpassing to a subsequence. Moreover, for any fixed R, it is easy to see ωα(r) converges to thepoint p in [ 1

Rrk, Rrk] as r → 0. However, if we use the arc length of the curve Γα to parameterizethe curves ωα respectively, then ωα(·) will converge locally.

By computation we have

ωα =12π

∫ 2π

0uα(r, θ)dθ =

12π

∫ 2π

0(uα(r, θ) +

uα,θθ

r2)dθ

=12π

∫ 2π

0∆0uαdθ − 1

∫ 2π

0

rdθ

= − 12π

∫ 2π

0A(uα)(duα, duα)− α− 1

∫ 2π

0

∇0|∇guα|2∇0uα

ε2α + |∇guα|2 dθ − ωα

r

where we have used the fact ∫ 2π

0uθθ(r, θ)dθ = 0.

LetGα = −ωα − ωα

r.

Denote the induced metric of Γα in RK by hα, and define AΓα as

AΓα(ωα, ωα) = −(

ωα − 〈ωα, ωα〉|ω|2 ωα

).

Given λtαα ∈ [λt1

α , λt2α ]. As before, we always have

uα(λtαα r, θ)− uα(λtα

α , 0)√α− 1

→ ~a log r,

where ~a ∈ TyN and y = limα→1

uα(λtαα , θ). Therefore, we have

ωα(λtαα ) =

√α− 1λtα

α(~a + o(1)), hα

(d

dr,

d

dr

)= |ωα|2 =

α− 1λ2tα

α

(|~a|2 + o(1)). (4.17)

where o(1) → 0 as α → 1.Moreover, we have

Gα(λtαα ) =

12π

∫ 2π

0A(uα)(duα, duα)dθ +

α− 12π

∫ 2π

0

∇0|∇gαuα|2∇0uα

ε2α + |∇gαuα|2 dθ

=α− 1λ2tα

α

(12π

∫ 2π

0A(y)(~a,~a)dθ + o(1)

)+ (α− 1)

∫ 2π

0O(|∇2

gαuα|)dθ

=α− 1λ2tα

α

(A(y)(~a,~a) + o(1) + O(√

α− 1))

=α− 1λ2tα

α

(A(y)(~a,~a) + o(1)).

34

In the above identity we have made use of the fact ν > 1 which implies that for any m > 0

λ2tαα = o((α− 1)m).

Noting that 〈A(y)(~a,~a),~a〉 = 0, we get

−AΓα(dωα, dωα) = ωα − 〈ωα, ωα〉|ω|2 ωα = −Gα +

〈Gα, ωα〉ωα

|ωα|2

= −α− 1λ2tα

α

(A(y)(~a,~a) + o(1)).(4.18)

Hence, we get‖AΓα‖2

hα(λtα

α ) < C.

Similar to the proof of Corollary 4.3, we have, after passing to a subsequence,

‖AΓα‖2hα

(λtα) < C

for any t ∈ [t2, t1].Now, we fix y ∈ N , and let s to be the arc length parameter of ωα(t) with s(λtα

α ) = 0. Weassume ωα(λtα

α ) → y as α → 1. It is well-known that ‖AΓα‖2hα

(λtαα ) does not depend on the

choice of parameter, andd2ωα

ds2= −AΓα(ωα)

(dωα

ds,dωα

ds

),

then ωα(s) will converges locally to ω(s) in C1, where s is still the arc length parameter. Thisimplies that Γα|[λt1

α ,λt2α ]

converges locally to a curve Γ locally. We claim that

AΓα(ωα)(

dωα

ds,dωα

ds

)→ A(ω)

(dω

ds,dω

ds

)

strongly in C0([0, s1],Rn) for sufficiently small s1. If this was not true, then for any small s1 wecould find a subsequence of uα, still denoted by uα, such that

s′α =∫ λ

t′αα

λtαα

|ωα|dr → s′ ∈ (0, s1)

and ∣∣∣AΓα(ωα)(

dωα

ds,dωα

ds

)−A(ω)

(dω

ds,dω

ds

) ∣∣∣s=s′α

> ε.

To apply Proposition 4.1, we must ensure that t′α ∈ [ t22 , t1]. For simplicity, we may assume

λt22

α = 2P λtαα

where P is an integer. Then, applying Corollary 4.3 we have

∫ 2i+1λtαα

2iλtαα

|ωα|dr =√

α− 1µ1−(tα−i logλα 2)

(log 2

√E(v)

π+ oα(1)

).

35

Therefore, as α is close to 1 enough,

∫ λt22

α

λtαα

|ωα|dr =P−1∑

i=0

∫ 2i+1λtαα

2iλtαα

|ωα|dr ≥ P√

α− 1

(√E(v)

πlog 2 + oα(1)

)

≥ C(tα − t22

) log λ−√

α−1α

≥ Ct22

log ν > 0.

Therefore, we may always choose s1 to be very small, for example s1 < C t22 log ν, such that

t′α ∈ [ t22 , t1]. Then, as before there holds

uα(λt′αα r, θ)− uα(λt′α

α , 0)√α− 1

→ ~a′ log r.

Obviously,ωα(λt′α

α )

|ωα(λt′αα )|

=dωα

ds(s′α) → dω

ds(s′).

Applying (4.17) and (4.18), we get that, after passing a subsequence the following holds

AΓα(ωα)(

dωα

ds,dωα

ds

) ∣∣s=s′α

=1

|ωα(λt′αα )|2

AΓα(ωα)(ωα, ωα)|r=λ

t′αα→ A(ω)(

ds,dω

ds)|s=s′

which contradicts the choice of s′α. So, we infer that

ds(s)− dω

ds(0) = −

∫ s

0A(ω)

(dω

ds,dω

ds

)ds

holds near s = 0. This shows ω is smooth near 0 and satisfies

d2ω

ds2= −A(ω)

(dω

ds,dω

ds

).

Therefore, we obtain finally

∇Ndωds

ds=

d2ω

ds2+ A(ω)

(dω

ds,dω

ds

)= 0,

which means that Γ is a geodesic. 2

Next, we give the length formula of the neck, i.e. calculate the length of the geodesic Γ.

The proof of Theorem 1.3 for the case ν > 1:

Proof. For simplicity, we assume λt2α = 2P λt1

α for some integer P . Then we have

P =t2 − t1log 2

log λα.

36

When ν = +∞, by Corollary 4.3, we have

L(Γα|B2k+1λ

t1α \2kλ

t1α

(xα)) ≥√

α− 1(

√E(v)

πlog 2 + o(1)).

ThenL(Γα) ≥ CP

√α− 1 ≥ C log λ−

√α−1

α → +∞.

This impliesL(Γ) = +∞.

Now, we assume ν < +∞. By Corollary 4.3,

L(Γα|B2k+1λ

t1α \2kλ

t1α

(xα)) =√

α− 1(

√E(v)

πlog 2 + oα(1)),

where oα(1) → 0 as α → 1 uniformly. Hence

L(Γ) = limα→1

√α− 1P

√E(v)

πlog 2 = (t1 − t2)

√E(v)

πlog ν.

Now, it is easy to see that to complete the proof of Theorem 1.3 we only need to prove thefollowing:

oscBλt

α\BRλα (xα)uα → 0, as α → 1, then R → +∞ and t → 1 (4.19)

andoscBδ\Bλt

α(xα)uα → 0, as α → 1, then δ → 0 and t → 0. (4.20)

Since ν < +∞ implies µ = 1, from Theorem 1.1 we know

limt→1

limR→+∞

limα→1

Bλt

α\BRλα (xα)

|∇uα|2 = 0.

Therefore, we can use the same method as in Subsection 4.1 (we replace δ with λtα) to deduce

oscBλt

α\BRλα

uα ≤ C√

E(uα, Bλtα\BRλα(xα)) + C(1− t) log ν + C

√α− 1,

then (4.19) follows. Similarly, we can prove (4.20). Hence, we derive the length formula of thegeodesic Γ

L =

√E(v)

πlog ν.

Thus, we finish the proof of Theorem 1.3. 2

Remark 4.1 Although we state and prove Theorem 1.3 only for one bubble case, it is notdifficult to follow the steps in section 3.2 to prove the general case. However, the general caseis quite complicated, for example, if we have 2 bubbles:

uα(λ1αx + x1) → v1, and uα(λ2

αx + x1) → v2

37

which satisfy: λ1α/λ2

α → 0 and ν1, ν2 < ∞, then u0(Bδ(x1)), v2(S2) are connected by a geodesicwith length

L =

√E(v1) + E(v2)

πlog ν2,

and v1(S2), v2(S2) are connected by a geodesic with length

L =

√E(v1)

πlog

ν1

ν2.

Next, we give the proof of Corollary 1.5. However, to prove the corollary we only need toprove the following proposition:

Proposition 4.5. If ν < +∞, then when (α − 1) is sufficiently small, all the uα are in thesame homotopy class.

Proof. When αi − 1 and αj − 1 are sufficiently small, we have ‖ui − uj‖C0 ≤ i(N) where i(N)is the injective radius of N . Hence, by using exponential map we know that uαi and uαj arehomotopic in M \Bδ, B

λt2α\B

λt1α

and BRλα respectively.Let p = u0(0) and q = v(+∞). By (4.19), we know that ui(Bδ(xα) \B

λt2α

(xα)) is containedin a simply connected ball centered at p when α is close to 1 enough, δ and t2 are small enough.Similarly, by (4.20) we also have uj(Bλ

t1α\BRλα(xα)) is contained in a small simply connected

ball in N with center q when α − 1, δ and 1 − t1 are sufficiently small. Hence ui and uj arehomotopic in Bδ \B

λt2α

and Bλ

t1α\BRλα respectively. So ui and uj are homotopic. 2

5 Some comments and an example

In this paper we only consider the case uα is an α-harmonic maps when the conformal structureof M is fixed. Naturally, one will ask the following problems (i) what could we say in the caseuα is an α-harmonic maps and the conformal structure of M varies with α, (ii) whether themethods in this paper can be extended to a class of variational problem which is more generalthan α-energy or not. In a forthcoming paper we will further develop some tools to discuss someissues which relate to the above problems.

On the other hand, one want to know whether one can give an example to show there is aneck joining the bubbles in the limit of an α-harmonic map sequence is of infinite length or not.However, if we can construct a manifold N and find a minimizing α-harmonic map sequencewhich satisfies the condition of Corollary 1.5, then the corollary tells us that indeed there existsa necks in the limit which if of infinite length. By modifying the example of Duzaar and Kuwert(see page 304 of [D-K]) we can construct such example as following .

Example. Let Z3 act on R3 by τκ(x, y, z) = (x+4k1, y+4k2, z+4k3), where κ = (k1, k2, k3) ∈ Z3.Consider

X = R3 \ ∪κτκ(B1(0))

and X is the quotient of X. Then X is a compact manifold with boundary. Topologically, X isT 3 minus a small ball.

38

Let Φ be a smooth map from R2 to ∂B1(0) satisfying

Φ(x) =

(1, 0, 0), |x| > 2,(−1, 0, 0), |x| < 1;

and the degree of map Φ denoted by deg(Φ) is 1 if we consider Φ as a map from S2 to S2,i.e. deg(Φ) = 1. Moreover, we let γk : [0, 1] → X be a curve which connects (4k − 1, 0, 0) and(1, 0, 0). We define

vk =

Φ(x), |x| ≥ δ,

γk

(log r − log Rε

log δ − log Rε

), Rε < |x| < δ,

τ(k,0,0)

(x

ε

)),

|x|ε≤ R.

We denote π to be the projection from X to X, then π(vk) ∈ π2(X). We have

Bδ\BRε

|∇vk|2 = 2π

∫ δ

∣∣∣∣∂γk

∂r

∣∣∣∣2

rdr

<c‖γ‖2

L∞

(− log Rε + log δ)2

∫ δ

dr

r=

c‖γ‖2L∞

log δ − log Rε,

R2\Bδ

|∇vk|2 ≤ E(Φ),

and ∫

BRε

|∇vk|2 ≤ E(Φ).

So, we can find suitable R and ε, s.t.

E(π(uk)) = E(uk) ≤ 2E(Φ) + 1.

We claim that [π(vk)] are different homotopy classes. Assuming this is not true, we can finda continuous map

H(x, t) : S2 × [0, 1] → X

s.t.H(x, 0) = π(vi) and H(x, 1) = π(vj).

Since S2 × [0, 1] is simply connected, we are able to lift H to H which is a map from S2 × [0, 1]into X with H(x, 0) = vi. We assume that H(x, 1) = τκ(vj). Hence [vi] = [τκ(vj)]. Therefore

[∂B1(0) + ∂τ(i,0,0)(B1(0))] = [∂τκ(B1(0)) + ∂τ(j,0,0)τκ(B1(0))] in π2(X),

where π2(X) is the second homotopy group of X. However, it is easy to check that π1(X) = 1,then by Hurewicz Theorem, the above identity is not true.

Now, we proceed to construct N . Let f be a homeomorphism from X to Y = X. Weconsider the quotient space of X ∪ Y , obtained by gluing every point x ∈ ∂X with f(x) ∈ ∂Ytogether. In this way, we get a closed compact manifold N and a projection φ : N → X. Oneis easy to check that π(vk) can be also considered as a map from S2 to N with E(π(vk)) < C.We claim that [π(vk)] are some different homotopic classes with each other in π2(N). Assum-ing it is not true. Then, we can find a continuous map H(x, t) : S2 × [0, 1] → N such that

39

H(x, 0) = π(vi) and H(x, 1) = π(vj). Hence, φ(H(x, t)) is just a homotopic map of π(vi) andπ(vj) in X. A contradiction.

Finally, we would like to ask the following problems:

Problem 1. Suppose all α-harmonic maps uα belong to the same homotopic class and satisfythe energy identity as α → 1. Do the necks consist of some geodesics of finite length?

Problem 2. Could we find a sequence αk → 1, and αk harmonic maps uαk, s.t. 1)the Morse

index tends to infinite; 2) supk Eαk(uαk

) < ∞; 3) for any i 6= j, uαi and uαj are not homotopicto each other.

6 Appendix

First we demonstrate how to construct the bubble trees at an energy concentration point asequence of α-harmonic maps. Assume uα ∈ C∞(Bσ, N) is the sequence stated in Theorem 1.1.Then, each uα satisfies

∆0uα + (α− 1)∇0|∇guα|2∇0uα

εα + |∇guα|2 + A(uα)(duα, duα) = 0,

where g = eϕα(dx1⊗dx1 +dx2⊗dx2) with ϕα(0) = 0 and ϕα converges smoothly to ϕ as α → 1.Denote by S(uα, B) the set

x ∈ Bσ : limα→1

lim infr→0

Br(x)|∇0uα|2dx >

ε02,

where ε0 is the constant defined in Theorem 2.1. Usually, S(uα, Bσ) is called the set of energy-concentration points (or simply concentration points) for uα on Bσ. Obviously, S(uα, Bσ) isa finite set.

One is easy to see that for any x /∈ S(uα, Bσ), there holds∫

Br(x)|∇0uα|2dx < ε0,

for sufficiently small r and α− 1. So, for any Ω ⊂ Bσ \ S(uα, Bσ), uα sub-converges in Ω.

Now, we show how to construct the bubble trees at each concentration point of S(uα, Bσ).For simplicity we assume 0 is the only concentration point for uα.

Step 1. Set r1α = λ1

α = 1maxBσ/2

|uα| . Let maxBσ/2|uα| be attained at the point xα. Since uα

converges smoothly away the concentration point, we have xα → 0.Set v1

α = uα(xα + λ1αx). Then

∆0v1α + (α− 1)e−ϕα(xα+λ1

αx) ∇0|∇gαv1α|2∇0v

εα(λ1α)2 + |∇g1

αv1α|2

+ A(v1α)(dv1

α, dv1α) = 0,

40

whereg1α = eϕα(xα+λ1

αx)(dx1 ⊗ dx1 + dx2 ⊗ dx2).

Since ‖∇v1α‖L∞ ≤ 1, v1

α converges to a harmonic map v1 from R2 to N locally and smoothly.As |∇v1(0)| = 1, we know that v1 is not trivial. We call v1 is the first bubble.

Step 2. To find the second bubble, we need to consider the following two cases.

Case 1. If for any ε > 0, we can find a positive number δ0 and a positive R such that∫

B2ρ\Bρ(x1α)|∇uα|2 < ε (6.1)

holds for any ρ ∈ [Rr1α, δ0], then energy concentration phenomena disappears. This means that

there is only one bubble for the sequence uα.

Case 2. If (6.1) is not true, then we can find r2α → 0 such that

limα→1

r2α

r1α

= +∞,

and ∫

B2r2

α\B

r2α

(x1α)|∇uα|2 > ε1 > 0, (6.2)

i.e. there are still some points where energy concentration happens. We set u2α = uα(xk + r2

αx),and denote the set of energy concentration points of u2

α on R2 by S(u2α,R2). Obviously,

0 ∈ S(u2α,R2), and v1 is also a bubble for the sequence u2

α. Moreover, passing to a sub-sequence we have u2

α converges to a harmonic map u20 away from S2(u2

α,R2), where u20 is a

harmonic map from R2 into N . Then we have two cases:

(a). u20 is trivial, i.e. a constant map;

(b). u20 is a non-trivial harmonic map from R2 to N .

If u20 is trivial, (6.2) implies that S2(u2

α,R2) ∩ (B3 \B1/2) 6= ∅, i.e. there must exist anotherconcentration point p for u2

α lies in B3 \ B1/2. At this new point of energy concentration foru2

α, we can get a bubble for u2α by the same program as in Step 1. That is to say, we may

find x′α → p, λ′α → 0 such thatu2

α(x′α + λ′αx) → v2,

where v2 is a harmonic map from R2 to N . If we set x2α = x′α + λ1

αx1α, λ2

α = λ1αλ′α, we get

uα(x2α + λ2

αx) → v2,

therefore, v2 is also a bubble for the sequence uα. Next, we repeat the step 1 and then repeatstep 2 at each concentration point for u2

α.

When u20 is non-trivial, v2 = u2

0 is a harmonic map from R2 to N . v2 is regarded as thesecond bubble. At each point of S2(u2

α,R2), we repeat step 1, and then repeat step 2.

41

Step 3. We replace r1α in Step 1 with r2

α and continue step 2 again.

In this way, we can get xkα → 0, λk

α → 0 such that

uα(xkα + λk

αx) → vk, k = 1, 2, · · · ,m,

where vk are harmonic maps from R2 to N . One is easy to check that for any i 6= j one of the(H1) and (H2) always hold. Moreover, since

(εα + |∇uα|2)αdVg ≥m∑

i=1

S2

|∇vi|2,

our procedure of construction must stop after finite steps, i.e. we can find a sequence of rk0α

such that for any ε > 0, there exist δ0 and R, s.t.∫

B2λ\Bλ

|∇uα|2 < ε, (6.3)

for any λ ∈ [Rrk0α , δ0].

References

[C] Chang, K.C.; Heat flow and boundary value problem for harmonic maps. Ann. Inst. H. PoincareAnal. Non Lineaire 6 (1989), no. 5, 363–395.

[C-D-Y] Chang, K.C., Ding, W. Y. and Ye, R.; Finite time blow-up of the heat flow of harmonic mapsfrom surfaces. J. Diff. Geom. 36(1992), 507-515.

[C-T] Chen, J. and Tian, G.; Compactification of moduli space of harmonic mappings. Comment. Math.Helv. 74 (1999), 201-237.

[C-M] Colding, T. H. and Minicozzi II, W. P.; Width and finite extinction time of Ricci flow. Preprint.

[D1] Ding, W.; An Introduction to harmonic maps. Manuscript.

[D2] Ding, W.; Lectures on the heat flow of harmonic maps. Manuscript.

[D3] Ding, W.; Blow-up of solutions of heat flows for harmonic maps. Advance in Math. (China) 19(1990), no. 1, 80–92.

[D-T] Ding, W. and Tian, G.; Energy identity for a class of approximate harmonic maps from surfaces,Comm. Anal. Geom. 3 (1995), 543-554.

[D-K] Duzaar, F and Kuwert, E.; Minimization of conformally invariant energies in homotopy classes.Calc. Var. Partial. Differ. Equ., 6 (1998), 285-313.

[E-S] Eells, J. and Sampson, J. H.; Harmonic mappings of Riemannian manifolds. Amer. J. Math. 86(1964), 109–160.

[J] Jost, J.; Two-dimensional geometric variantional problems. John Wiley and Sons, Chichester, 1991.

[L] Lamm, T.; Energy identy for approximations of harmonic maps from surfaces. Preprint.

[L-W] Li, Y. and Wang, Y.; Bubbling location for a sequence of approximate f -harmonic maps fromsurfaces. Preprint.

42

[Lin-W] Lin, F. and Wang, C.; Energy identity of harmonic map flows from surfaces at finite singulartime. Calc. Var. Partial Differential Equations, 6 (1998), 369-380.

[Mo] Moore, J. D.; Energy growth in minimal surface bubbles. “http://www.math.ucsb.edu/ moore/growthrev.pdf”.

[P] Parker, T. H.; Bubble tree convergence for harmonic maps. J. Diff. Geom. 44 (1996), 595-633.

[Q] Qing, J.; On sighularities of the heat flow for harmonic maps from surfaces into spheres. Comm.Anal. Geom., 3 (1995), 297-315.

[Q-T] Qing, J. and Tian, G.; Bubbling of the heat flow for harmonic maps from surfaces. Comm. Pure.Apple. Math., 50 (1997), 295-310.

[S-U] Sacks, J. and Uhlenbeck, K.; The existence of minimal immersions of 2-spheres, Ann. of Math.,113 (1981), 1-24.

[St] Struwe, M.; On the evolution of harmonic maps of Riemannian surfaces, Comment. Math. Helv., 60(1985), 558-581.

[St2] Struwe, M.; Positive solutions of critical semilinear elliptic equations on non-contractible planardomains. J. Eur. Math. Soc., 2 (2000), 329-388.

[T] Topping, P.; Winding behaviour of finite-time singularities of the harmonic map heat flow. Math.Zeit., 247 (2004), 279-302.

Yuxiang LiDepartment of Mathematical Sciences, Tsinghua University,Beijing 100084, P.R.China.Email: [email protected]

Youde WangAcademy of Mathematics and Systematic Sciences,Chinese Academy of Sciences,Beijing 100190, P.R. China.Email: [email protected]

43