Skip to main content
Springer logoLink to Springer
. 2018 Sep 24;374(1):133–177. doi: 10.1007/s00208-018-1759-8

The qualitative behavior at the free boundary for approximate harmonic maps from surfaces

Jürgen Jost 1,2,, Lei Liu 1,3, Miaomiao Zhu 4
PMCID: PMC6560013  PMID: 31258185

Abstract

Let {un} be a sequence of maps from a compact Riemann surface M with smooth boundary to a general compact Riemannian manifold N with free boundary on a smooth submanifold KN satisfying

supnunL2(M)+τ(un)L2(M)Λ,

where τ(un) is the tension field of the map un. We show that the energy identity and the no neck property hold during a blow-up process. The assumptions are such that this result also applies to the harmonic map heat flow with free boundary, to prove the energy identity at finite singular time as well as at infinity time. Also, the no neck property holds at infinity time.

Mathematics Subject Classification: 53C43 58E20

Introduction

Let (Mg) be a compact Riemannian manifold with smooth boundary and (Nh) be a compact Riemannian manifold of dimension n. Let KN be a k-dimensional closed submanifold where 1kn. For a mapping uC2(M,N), the energy density of u is defined by

e(u)=12|u|2=Traceguh,

where uh is the pull-back of the metric tensor h.

The energy of the mapping u is defined as

E(u)=Me(u)dvolg.

Define

C(K)=uC2(M,N);u(M)K.

A critical point of the energy E over C(K) is a harmonic map with free boundary u(M) on K. The problem of the existence, uniqueness and regularity of such harmonic maps with a free boundary was first systematically investigated in [8].

By Nash’s embedding theorem, (Nh) can be isometrically embedded into some Euclidean space RN. Then we can get the Euler-Lagrange equation

Δgu=A(u)(u,u),

where A is the second fundamental form of NRN and Δg is the Laplace-Beltrami operator on M which is defined by

Δg:=-1gxβggαβxα.

Moreover, for 1kn-1, u has free boundary u(M) on K, that is

u(x)K,du(x)(n)Tu(x)K,a.e.xM, 1.1

where n is the outward unite normal vector on M and means orthogonal.

Specially, for k=n, u satisfies a homogeneous Neumann condition on K, that is

u(x)K,du(x)(n)=0,a.e.xM. 1.2

The tension field τ(u) is defined by

τ(u)=-Δgu+A(u)(u,u). 1.3

Thus, u is a harmonic map if and only if τ(u)=0.

When we consider a limit of a sequence of maps with uniformly L2-bounded tension fields, the domain may decompose into several pieces (a phenomenon called bubbling or blow-up), and the limit map satisfies the equations or bounds on each piece. The question is whether the sum of the energies of the limit map on those pieces equals the limit of the energies of the approximating maps. Affirmative results are called energy identity and no neck property, and the approach is called blow-up theory; the precise definitions will be given below. Because the problem is conformally invariant only in dimension 2, the analysis usually needs to be restricted to that case, and this will also apply to this paper.

When M is a closed surface, the compactness problem and the blow-up theory (energy identity and no neck property) for a sequence of maps {un} from M to N with uniformly L2-bounded tension fields τ(un) and uniformly bounded energy has been extensively studied (see e.g. [6, 13, 29, 31, 32, 48]), since the fundamental work of Sacks-Uhlenbeck [38]. For sequences of general bounded tension fields, see [20, 21, 26, 49]. For sequences of solutions of more general elliptic systems with an antisymmetric structure, we refer to [16, 18]. For corresponding results about harmonic map flows, see e.g. [24, 31, 32, 44, 47]. For results of other types of approximate sequences for harmonic maps, see e.g. [4, 11, 13, 15, 23]. For the energy identity of harmonic maps from higher dimensional domains, see [25].

In this paper, we shall study the blow-up analysis for a sequence of maps {un} from a compact Riemann surface M with smooth boundary M to a compact Riemannian manifold N with uniformly L2-bounded tension fields τ(un), uniformly bounded energy and with free boundary un(M) on K. Since the interior case is already well understood, we shall focus on the case where the energy concentration occurs at the free boundary and complete the blow-up theory at the free boundary for a bubbling sequence. When boundary blow-up occurs, the corresponding neck domains are in general not simply half annuli and hence a finer decomposition of the neck domains would be necessary in order to carry out the neck analysis (see Sect. 5).

In fact, we shall first address the regularity problem at the free boundary for weak solutions (see Sect. 3) of

-Δgu+A(u)(u,u)=FinM 1.4

for some FLp(M), p>1 and under the free boundary constraint (1.1), as it provides some necessary elliptic estimates at the free boundary, which form the analytical foundation of the blow-up theory for the sequence {un} (see Sect. 4). We would like to remark that the regularity at the free boundary for weak solutions of (1.4) can be proved by applying the classical reflection methods for the harmonic map case by Gulliver-Jost [8] and Scheven [39] or a modified reflection method in [3] and [43] which combines Hélein’s moving frame method [10] and Scheven’s reflection method [39] so that the technique of Rivière-Struwe in [35] (which holds true also in dimension 2) can be applied. The latter was developed for Dirac-harmonic maps which includes harmonic maps as a special case. In this paper, we shall present an alternative approach without using moving frames (see Sect. 3).

Now, we state our first main result:

Theorem 1.1

Let un:MN be a sequence of W2,2 maps with free boundary un(M) on K(1kn), satisfying

E(un)+τ(un)L2(M)Λ<,

where τ(un) is the tension field of un. We define the blow-up set

S:=r>0xM|lim infnDrM(x)|dun|2dvolϵ¯2, 1.5

where DrM(x)={yM|dist(x,y)r} denotes the geodesic ball in M and ϵ¯>0 is a constant whose value will be given in (5.3). Then S is a finite set {p1,...,pI}. By taking subsequences, {un} converges in Wloc2,2(M\S) to some limit map u0W2,2(M,N) with free boundary on K and there are finitely many bubbles: a finite set of harmonic spheres wil:S2N, l=1,...,li, and a finite set of harmonic disks wik:D1(0)N, k=1,...,ki with free boundaries on K, where li,ki0 and li+ki1, i=1,...,I, such that

limnE(un)=E(u0)+i=1Il=1liE(wil)+i=1Ik=1kiE(wik), 1.6

and the image u0(M)i=1I(l=1li(wil(S2))k=1ki(wik(D1(0)))) is a connected set. Here, harmonic spheres are minimal spheres and harmonic disks with free boundary on K are minimal disks with free boundary on K.

In contrast to the Dirichlet problem where, due to the pointwise boundary condition, no blow-up at the boundary is possible. Here, a blow-up may occur at the boundary and produce one or more harmonic disks with the same free boundary K as the original maps. We should also mention that the Plateau boundary condition for minimal surfaces can also be seen as a free boundary condition where the target set K is a Jordan curve. Here, the monotonicity condition and the three-point normalization that are usually imposed prevent a boundary blow-up, however, see [8] and the systematic discussion in [13].

Our results in the above theorem apply to some classical problems like minimal surfaces in Riemannian manifolds with free boundaries, harmonic functions with free boundary (c.f. [17]) as well as to pseudo holomorphic curves in sympletic manifolds with totally real boundary conditions and Lagrangian boundary conditions, c.f. [7, 12, 28, 51, 53] and to string theory where the free boundary represents a D-brane, c.f. [14].

The reason why we work with a sequence of maps with uniformly L2-bounded tension fields and with free boundary is that we want to apply our results in Theorem 1.1 to the following heat flow for harmonic maps with free boundary:

tu(x,t)=τ(u(x))(x,t)M×(0,T); 1.7
u(·,0)=u0(x)xM; 1.8
u(x,t)K,a.e.xM,t0; 1.9
du(x)(n)Tu(x)K,(x,t)M×(0,T). 1.10

The existence of a global weak solution of (1.71.10) with finitely many singularities was considered by Ma [27], following the pioneering works by Struwe [44, 45]. For higher dimensional cases, we refer to [2, 46]. For other work on the harmonic map flow with free boundary, see [19]. For the harmonic map flow with Dirichlet boundary condition, we refer to Chang [1].

Let u:M×(0,)N be a global weak solution to (1.71.10), which is smooth away from a finite number of singular points {(xi,ti)}M×(0,). In this paper, we shall complete the qualitative picture at the singularities of this flow, where bubbles (nontrivial harmonic spheres or nontrivial harmonic disks with free boundary) split off.

At infinite time, we have

Theorem 1.2

There exist a harmonic map u:MN with free boundary in K, a finite number of bubbles {ωi}i=1m and sequences {xni}i=1mM, {λni}i=1mR+ and {tn}R+ such that

limtE(u(·,t),M)=E(u,M)+i=1mE(ωi) 1.11

and

u(·,tn)-u(·)-i=1mωni(·)L(M)0 1.12

as n, where ωni(·)=ωi·-xniλni-ωi(). Here, (1.12) is equivalent to say that the image of weak limit u and bubbles {ωi}i=1m is a connected set as in Theorem 1.1.

For finite time blow-ups, we have

Theorem 1.3

For T0<, let uC(M×(0,T0),N) be a solution to (1.71.10) with T0 as its singular time. Then there exist finite many bubbles {ωi}i=1l such that

limtT0E(u(·,t),M)=E(u(·,T0),M)+i=1lE(ωi). 1.13

To study the regularity or the qualitative behavior at the free boundary for approximate harmonic maps in this paper, we need some new observations. Firstly, we need to extend the solution across the free boundary as in the harmonic map case done by Scheven [39] and the main difficulty is to write the equation of the extended map into an elliptic system with an antisymmetric potential up to some transformation (see Proposition 3.3). Secondly, thanks to the free boundary condition, we can apply the Pohozaev’s argument which was firstly introduced by Lin-Wang [24] for approximate harmonic maps, in the local region as Dr(x)M with xM. See Lemma 4.3. This is crucial when we estimate the energy concentration in the neck domain. Thirdly, we have a finer observation of the neck domain. For the boundary blow-up point, the neck domains consist of some irregular half annulus. We will decompose these irregular neck domains into three parts as: interior parts, regular half annulus with the center points living on the boundary and the remaining parts. The first and third parts are easy to control due to the classical blow-up theory of (approximate) harmonic maps with interior blow-up points. In this paper, we focus on the energy concentration in the domains of the second parts.

Since the extended map satisfies an elliptic system with an antisymmetric potential up to some transformation and with some error term F (see Proposition 3.3), one can utilize the idea in [18] (with F=0) with some modifications to get the energy identity. Here in the present paper, we shall adapt the methods in [5] developed for the interior bubbling case to get the energy identity and the no neck property in the free boundary case. To show the no neck property, namely, bubble tree convergence, we shall get the exponential decay of the energy by deriving a differential inequality on the neck region.

This paper is organized as follows. In Sect. 2, we recall some classical results which will be used in this paper. In Sect. 3, we derive a new form of the elliptic system for the extended map after involution across the boundary which will allow us to turn the boundary regularity problem into an interior regularity problem. As a corollary of this boundary regularity result, we prove a removability theorem for singularities at the free boundary. In Sect. 4, using the new equation of the involuted map, we obtain the small energy regularity in the free boundary case. The gap theorem and Pohozaev’s identity in the free boundary case will also be established. In Sect. 5, we prove the energy identity and no neck property at the free boundary by decomposing the neck domain into several parts including a half annulus centered at the boundary and then using the involuted map’s equation. Combining this with the interior blow-up theory, we complete the proof of Theorem 1.1. In Sect. 6, we apply Theorem 1.1 to the harmonic map flow with free boundary and prove Theorem 1.2 and Theorem 1.3.

Notation:Dr(x0) denotes the closed ball of radius r and center x0 in R2. Denote

Dr+(x0):=x=(x1,x2)Dr(x0)|x20,Dr-(x0):=x=(x1,x2)Dr(x0)|x20,+Dr(x0):=x=(x1,x2)Dr(x0)|x20,-Dr(x0):=x=(x1,x2)Dr(x0)|x20,0Dr+(x0)=0Dr-(x0):=Dr+(x0)\+Dr(x0).

Let a0 be a constant, denote

Ra2:=(x1,x2)|x2-aandRa2+:=(x1,x2)|x2>-a.

For convenience, we denote Dr=Dr(0), D=D1(0) and R+2=Ra2 when a=0.

Let TM be a smooth boundary portion, denote

Wk,p(T)=gL1(T):g=G|TforsomeGWk,p(M)

with norm

gWk,p(T)=infGWk,p(M),G|T=gGWk,p(M).

In this paper, we use the notation Δg (or ΔM) to denote the Laplace-Beltrami operator on the Riemannian manifold (Mg) and use Δ:=x2+y2 to denote the usual Laplace operator on R2.

Preliminary results

In this section, we will recall some well known results that are useful for our problem.

Firstly, we recall the interior small energy regularity result (see [6, 20]) which is firstly introduced in [38].

Lemma 2.1

Let uW2,p(D,N) for some 1<p2. There exist constants ϵ1=ϵ1(p,N)>0 and C=C(p,N)>0, such that if uL2(D)ϵ1, then

u-1πDu(x)dxW2,p(D1/2)C(p,N)(uLp(D)+τ(u)Lp(D)), 2.1

where τ(u) is the tension field of u.

Moreover, by the Sobolev embedding W2,p(R2)C0(R2), we have

uOsc(D1/2)=supx,yD1/2|u(x)-u(y)|C(p,N)(uLp(D)+τ(u)Lp(D)). 2.2

Secondly, we recall a gap theorem for the case of a closed domain.

Lemma 2.2

([5]) There exists a constant ϵ0=ϵ0(M,N)>0 such that if u is a smooth harmonic map from a closed Riemann surface M to a compact Riemannian manifold N and satisfying

M|u|2dvolϵ0,

then u is a constant map.

Thirdly, we state an interior removable singularity result.

Theorem 2.3

([22]) Let u:D\{0}N be a Wloc2,2(D\{0}) map with finite energy that satisfies

τ(u)=gL2(D,TN),xD\{0}.

Then u can be extended to a map in W2,2(D,N).

Next, combining the regularity results for critical elliptical systems with an antisymmetric structure developed by Rivière [33] and Rivière-Struwe [35] with various extensions in e.g. [34, 36, 37, 4042, 54], we state the following

Theorem 2.4

Let d2, 0sd, 0<Λ< and 1<p<2. For any ALW1,2(D,GL(d)), ΩL2(D,so(d)1Rm), fLp(D,Rd) and any uW1,2(D,Rd) weakly solving

d(Adu)=Ω,Adu+finD, 2.3

with A satisfying

Λ-1|ξ||A(x)ξ|Λ|ξ|for a.e.xD,for allξRd, 2.4

we have uWloc2,p(D) and there exist ϵ=ϵ(d,Λ,p)>0 and C=C(d,Λ,p)>0 such that whenever ΩL2(D)+AL2(D)ϵ then

2uLpD12+uL2p2-pD12C(uL1(D)+fLp(D)).

It is well known that the harmonic map equation can be written as a critical elliptical system with an antisymmetric structure and hence we have the following (which can also be proved by using classical methods developed for the harmonic map case, see e.g. [10])

Theorem 2.5

For every p(1,) there exists an ϵ>0 with the following property. Suppose that uW1,2(D;N) and fLp(D;RN) satisfy

τ(u)=finD

weakly, then uWloc2,p(D).

Finally, we recall the classical boundary estimates for the Laplace operator under Neumann boundary condition.

Lemma 2.6

(see e.g. [50]) Let fWk,p(M) and gWk,p(M) for some kN0, 1<p<. Assume that uW1,p(M) weakly solves

ΔMu=finM;un=gonM.

Then uWk+2,p(M) is a strong solution. Moreover, there exist constants C=C(M)>0 and C=C(M)>0 such that for all uWk+2,p(M)

uWk+2,pMCΔMuWk,pM+unWk+1,pM+uLpM;uWk+2,pMCΔMuWk,pM+unWk+1,pM,ifMu=0.

Regularity at the free boundary

In this section, we will prove a regularity theorem for weak solutions of (1.4) and (1.1) with FLp(M,RN) for some p>1 where F(x)Tu(x)N for a.e.xM. As an application, we derive the removability theorem for a local singularity at the free boundary.

We first need to define weak solutions of (1.4) and (1.1).

Definition 3.1

uH1(M,N) is called a weak solution to (1.4) and (1.1) if u(M)K a.e. and

-Mu·φdvol=MF·φdvol

for any vector field φLH1(M,TN) that is tangential along u and satisfies the boundary condition φ(x)Tu(x)K for a.e. xM. We also say uH1(M,N) is a weak solution of (1.4) with free boundary u(M) on K.

For a weakly harmonic map with free boundary (i.e.F=0), it is shown that the image of the map is contained in a small tubular neighborhood of K if the energy of the map is small, see Lemma 3.1 in [39]. The proof there requires the interior L-estimate for the gradient of the map. Here, we extend this localization property to the more general case of weak solutions of (1.4) with FLp(D+) for some 1<p2 and derive certain oscillation estimate for the solution. In our case, there is in general no interior L-estimate for the gradient of the map.

Lemma 3.2

Let FLp(D+) for some 1<p2 and uW1,2(D+,N) be a weak solution of (1.4) with free boundary u(0D+) on K. Then there exists positive constants C=C(p,N), ϵ2=ϵ2(p,N), such that if uL2(D+)ϵ2, then

dist(u(x),K)C(p,N)(uL2(D+)+FLp(D+))forallxD1/2+. 3.1

Moreover, we have

OscD14+u:=supx,yD14+|u(x)-u(y)|C(p,N)uL2(D+)+FLp(D+). 3.2

Proof

We shall follow the scheme of the proof of Lemma 3.1 in [39]. Take ϵ2=min{ϵ1,ϵ} where ϵ1 and ϵ are the corresponding constants in Lemma 2.1 and Theorem 2.5. By the interior regularity result Theorem 2.5, we know uWloc2,p(D+\D+). For any x0D1/2+\0D+, set R=13dist(x0,0D+) and suppose x10D+ is the nearest point to x0, i.e. |x0-x1|=dist(x0,0D+)=3R. Let Gx0 be the fundamental solution of the Laplace operator with singularity at x0 which satisfies

|Gx0|C(n)|x-x0|-1forallxR2.

Setting w(x)=u(x)-u¯ where u¯:=1|D5R+(x1)|D5R+(x1)udx and choosing a cut-off function ηC0(D2R(x0)) such that 0η1, η|DR(x0)1 and |η|CR, by Green’s representation theorem and integrating by parts, we have

|w(x0)|2=-D2R(x0)Gx0(x)(|w|2η2)dxCD2R(x0)|Gx0(x)||ww|η2dx+CD2R(x0)\DR(x0)|Gx0(x)||w|2|η|dxCwL(D2R(x0))D2R(x0)|Gx0(x)||u|dx+CR-2D2R(x0)\DR(x0)|w|2dxCwL(D2R(x0))Gx0(x)Lqq-1(D2R(x0))uLq(D2R(x0))+CR-2D2R(x0)|w|2dx:=I+II, 3.3

where 2<q=p2-p<2p2-p if 1<p<2 and q=4 if p=2.

According to Lemma 2.1, we have

R1-2suLs(D2R(x0))+uOsc(D2R(x0))C(s,p,N)uL2(D3R(x0))+R1-1pFLp(D3R(x0))C(s,p,N)uL2(D+)+FLp(D+) 3.4

for any 2<s<2p2-p. Thus, we obtain

IC(p,N)uL2(D+)+FLp(D+)R1-2/qwL(D2R(x0))1|x-x0|Lqq-1(D2R(x0))C(p,N)(uL2(D+)+FLp(D+))wL(D2R(x0))C(p,N)(uL2(D+)+FLp(D+))(|w(x0)|+uOsc(D2R(x0)))12|w(x0)|2+C(p,N)(uL2(D+)+FLp(D+))2.

Combining the Poincaré inequality with the fact D2R(x0)D5R+(x1)D+, we get

IICR-2D5R+(x1)|w|2dxCD5R+(x1)|u|2dx.

So, we have

|u(x0)-u¯|C(p,N)(uL2(D+)+FLp(D+)). 3.5

Set d(y):=dist(y,K) for yN, then we have

d(u¯)d(u(x))+|u(x)-u¯|.

Integrating the above inequality, we get

du¯1|D5R+x1|D5R+x1duxdx+1|D5R+x1|D5R+x1|ux-u¯|dxCD5R+x1|dux|2dx1/2+CD5R+x1|u|2dx1/2CD5R+x1|u|2dx1/2CuL2D+,

where the second inequality follows from the Poincaré inequality since d(u(x))=0 on 0D5R+(x1) and the third inequality follows from the fact that Lip(d)=1.

Then, we have

dist(u(x0),K)dist(u¯,K)+|u(x0)-u¯|C(p,N)(uL2(D+)+FLp(D+)),

which implies (3.1) holds.

For (3.2), taking x0=0,12D12+\0D+ in (3.5), then x1=0, R=13|x0-x1|=16 and we get

u0,12-1D56+(0)D56+(0)udxC(p,N)(uL2(D+)+FLp(D+)). 3.6

For any y0D14+\0D+, set Ry0=13dist(y0,0D+) and suppose y10D+ is the nearest point to y0, i.e. |y0-y1|=dist(y0,0D+)=3Ry0. Combing (3.5) with (3.6), we obtain that

uy0-u0,12uy0-1D5Ry0+y1D5Ry0+y1udx+u0,12-1D56+0D56+0udx+1D5Ry0+y1D5Ry0+y1udx-1D56+0D56+0udxCp,NuL2D++FLpD++1D5Ry0+y1D5Ry0+y1udx-1D56+0D56+0udx.

Noting that D5Ry0+(y1)D56+(0), by a variant of the classical Poincaré inequality, we have

1D5Ry0+y1D5Ry0+y1udx-1D56+0D56+0udx1D56+0D56+0u-1D5Ry0+y1D5Ry0+y1udxdxCuL2D56+0CuL2D1+0.

Therefore,

OscD14+u:=supx,yD14+|ux-uy|ux-u0,12+uy-u0,12Cp,NuL2D++FLpD+.

Thus, the lemma follows immediately.

With the help of Lemma 3.2, we can extend the map to the whole disc D by involuting. Firstly, we consider 1kn-1. Without loss of generality, we may assume KN= in this paper. In fact, if KN, we extend the target manifold N smoothly across the boundary to another compact Riemannian manifold N, such that NN and KN=. Then we can consider N as a new target manifold.

Denote by Kδ0 the δ0-tubular neighborhood of K in N. Taking δ0>0 small enough, then for any yKδ0, there exists a unique projection yK. Set y¯=expy{-expy-1y}. So we may define an involution σ, i.e. σ2=Id as in [8, 9, 39] by

σ(y)=y¯foryKδ0.

Then it is easy to check that the linear operator Dσ:TN|Kδ0TN|Kδ0 satisfies Dσ(V)=V for VTK and Dσ(ξ)=-ξ for ξTK.

Let FLp(D2+) for some 1<p2 and uW1,2(D2+,N) be a weak solution of (1.4) with free boundary u(0D2+) on K. If uL2(D2+)+FLp(D2+)ϵ3 where ϵ3=ϵ3(p,N,δ0)>0 is small, by the oscillation estimate (3.2) in Lemma 3.2, we know

uD+BCϵ3Nu0,12Kδ0, 3.7

where BCϵ3N(u(0,12)) is the geodesic ball in N with the center point u(0,12) and radius Cϵ3. Then we can define an extension of u to D1(0) that

u^(x)=u(x),ifxD+;σ(u(ρ(x))),ifxD-, 3.8

where ρ(x)=(x1,-x2) for x=(x1,x2)D1(0). For k=n, we also use the above extension by replacing σ=Id. In the following part of this paper, we always state the argument for 1kn-1, since k=n is similar and easier.

At this point, one can derive the regularity at the free boundary for weak solutions of (1.4) by applying classical methods in [8, 39] for harmonic maps or the method in [3, 43] which combines the method of moving frame and some modification of Rivière-Struwe’s method in [35]. Now, we shall give our alternative approach which is also based on some extension of Rivière-Struwe’s result.

In order to derive the equation of the involuted map u^, we shall first define

P:Bδ1Nu0,12Kδ0GLRN,RN=GLTRN,TRN

by

P(y)ξ=Dσ(y)ξ(y)+l=n+1Nξ,νl(y)νl(σ(y)), 3.9

where δ1=δ1(N) is small such that B4δ1N(u(0,12))Kδ0 and there exists a local orthonormal basis {νl}l=n+1N of the normal bundle TN|B4δ1Nu0,12, ξ(y) is the projection map of RNTyN. On one hand, Lemma 3.2 tells us that dist(u(0,12),K)Cϵ3 which implies σBδ1N(u(0,12))B4δ1Nu0,12 if we take ϵ3 small enough (e.g. Cϵ3δ1). Thus, (3.9) is well defined. On the other hand, noting that since (3.7) holds, if ϵ3 is small enough (e.g. 4Cϵ3δ1), then we know that u^(D)B4Cϵ3Nu0,12Bδ1Nu0,12 and the notations P(u^(x)), O(u^(x)) in the sequel (see below) are well defined. It is easy to check that P(y) is invertible linear operator for any yBδ1Nu0,12, since the linear operator Dσ(y) is invertible. For simplicity, we still denote by P(y) the matrix corresponding to the linear operator P(y) under the standard orthonormal basis of RN. Moreover, the matrix P(y) and its inverse matrix P-1(y) are smooth for yBδ1Nu0,12. So, there exists an orthogonal matrix O(y) which is also smooth, such that

OTPTPO=Ξ:=λ1(y)000000λN(y)

where PT is the transposed matrix and λi(y), i=1,...,N is the eigenvalues of the positive symmetric matrix PT(y)P(y). It is easy to see that λi(y)=1 for yK, i=1,...,N.

Define

ρ(x)=x,xD+;ρ(x),xD-,andσ(u^(x))=u^(x),xD+;σ(u^(x)),xD-,

and the matrixes

Q=Q(x)=IdN×N,xD+;P(u^(x)),xD-,andQ~=Q~(x)=IdN×N,xD+;O(u^)Ξ(u^)OT(u^),xD-,

where

Ξ(y)=λ1(y)000000λN(y).

One can easily find that Q~LW1,2(D,RN) and is invertible.

The involuted map satisfies the following proposition:

Proposition 3.3

Let FLp(D2+) for some 1<p2 and u(x)W1,2(D2+) be a weak solution of (1.4) with free boundary u(0D2+) on K. There exists a positive constant ϵ3=ϵ3(p,N), such that if uL2(D2+)+FLp(D2+)ϵ3 and u^ is defined as above, then u^W1,2(D) is a weak solution of

div(Q~·u^(x))=Ω·Q~·u^(x)+Q~-1·QT·F(ρ(x)),xD, 3.10

where

Ω(x)=Ω2(x),xD+;Ω1(u^(x))+Ω2(x)-Q~-1·12(QTQ-QTQ)·Q~-1,xD-,

and

Ω1=(Ω1)AB:=OOT+12OΞOTOΞ-1OT-12OΞ-1OTOΞOT,Ω2=(Ω2)AB:=Q~·Q-1·(νl(σ(u^)))·νlT(u^)·Q~-1-Q~-1·νl(u^)·(νlT(σ(u^)))·(Q-1)T·Q~,

in the distribution sense. Here, Ω(x), Ω1(x) and Ω2(x) are antisymmetric matrices in L2(D).

Moreover, if uW2,p(D+), 1<p2, then u^W2,p(D) and satisfies

u^+Υu^(u^,u^)=F^inD, 3.11

where Υu^(·,·) is a bounded bilinear form and F^Lp(D) which are defined by (3.21), satisfying

|Υu^(u^,u^)|C(N)|u^|2andF^Lp(D)C(N)FLp(D+).

Proof

Step 1 Firstly, it is easy to see that u^W1,2(D). Secondly, we prove that for any arbitrary test vector field VLW01,2(D,TN) with V(x)Tu^(x)N for a.e. xD, there holds

-DQ·u^(x)·(Q·V)dx=DF(ρ(x))·Q·Vdx. 3.12

Set Σ(x):=Dσ|u^(x) for xD. We decompose V into the symmetric and anti-symmetric part with respect to σ as in [39], i.e. V=Ve+Va, where

Ve(x):=12{V(x)+Σ(ρ(x))V(ρ(x))},Va(x):=12{V(x)-Σ(ρ(x))V(ρ(x))}.

Since σ2=Id, we have Σ(x)Σ(ρ(x))=Id. Then,

Ve(ρ(x))=Σ(x)Ve(x)andVa(ρ(x))=-Σ(x)Va(x).

Noting Dσ:TN|Kδ0TN|Kδ0 satisfying Dσ(V)=V for VTK and Dσ(ξ)=-ξ for ξTK, for any x0D+, we know

Ve(x)=12{V(x)+Σ(x)V(x)}=ΠTKV(x)TK

where ΠTK:TNTK is the orthogonal projection.

Since u is a weak solution of (1.4) in D+, we have

-D+u(x)Ve(x)dx=D+F(x)·Ve(x)dx. 3.13

Thus,

-D-Q·u^(x)·(Q·Ve(x))dx=-D-Dσ|u^·u^(x)·(Dσ|u^·Ve(x))dx=-D-(u(ρ(x)))·(Σ(x)·Ve(x))dx=-D-(u(ρ(x)))·(Ve(ρ(x)))dx=-D+u(x)Ve(x)dx=D+F(x)·Ve(x)dx=D-F(ρ(x))·Q·Ve(x)dx. 3.14

Moreover, there holds

-D-Q·u^(x)·(Q·Va(x))dx=-D-Dσ|u^·u^(x)·(Dσ|u^·Va(x))dx=D-(u(ρ(x)))·(Va(ρ(x)))dx=D+u(x)Va(x)dx, 3.15

and

DF(ρ(x))·Q·Va(x)dx=D+F(x)·Va(x)dx+D-F(ρ(x))·Q·Va(x)dx=D+F(x)·Va(x)dx-D-F(ρ(x))·Va(ρ(x))dx=D+F(x)·Va(x)dx-D+F(x)·Va(x)dx=0. 3.16

Then (3.13), (3.14), (3.15) and (3.16) imply (3.12) immediately.

Step 2 We claim: for any VLW01,2(D,RN), there holds

-DQ·u^(x)·(Q·V)dx=-DQ·u^(x),(νl(σ(u^)))·νl(u^),Vdx+DF(ρ(x))·Q·Vdx. 3.17

In fact, on the one hand, by (3.12), we get

-DQ·u^(x)·(Q·V)dx=-DQ·u^(x)·(Q·V)dx-DQ·u^(x)·(Q·V)dx=DF(ρ(x))·Q·Vdx-DQ·u^(x)·(Q·V)dx.

On the other hand, we have

-DQ·u^(x)·(Q·V)dx=-D+u(x)·Vdx-D-Q·u^(x)·(Q·V)dx=I+II.

Computing directly, we have

I=-D+u(x)·(V,νlνl)dx=-D+u(x)·V,νlνldx=-D+Q·u^(x),(νl(σ(u^)))·V,νl(u^)dx

and

II=-D-Q·u^(x)·(Q·V)dx=-D-Q·u^(x)·(Q·V,νl(u^)νl(u^))dx=-D-Q·u^(x)·(V,νl(u^)νl(σ(u^)))dx=-D-Q·u^(x),(νl(σ(u^)))·V,νl(u^)dx.

Combining these equations, we obtain

-DQ·u^(x)·(Q·V)dx=-DQ·u^(x),(νl(σ(u^)))·νl(u^),Vdx. 3.18

Thus, we have

-DQ·u^(x)·(Q·V)dx=-DQ·u^(x),(νl(σ(u^)))·νl(u^),Vdx+DF(ρ(x))·Q·Vdx,

where the equality follows from that F(ρ(x))Tu(ρ(x))N=Tσ(u^)N. This is (3.17).

Step 3 In order to prove u^ is a weak solution of (3.10), take an arbitrary test vector field VLW01,2(D,RN), since the matrix Q~,Q~-1LW1,2(D,RN), it is sufficient to prove

-DQ~·u^(x)·(Q~·V)dx=DΩ·Q~·u^(x)+Q~-1·QT·F(ρ(x)),Q~·Vdx=-DQ~·u^(x),Ω·Q~·Vdx+DF(ρ(x))·Q·Vdx. 3.19

Computing directly, we get

-D-Q·u^(x)·(Q·V)dx=-D-QTQ·u^(x),Vdx-D-u^(x),QTQ·Vdx=-D-OΞOT·u^(x),OΞOT·Vdx-D-u^(x),12(QTQ)·Vdx-D-u^(x),12(QTQ-QTQ)·Vdx=-D-OΞOT·u^(x),(OΞOT·V)dx-D-u^(x),12(QTQ-QTQ)·Vdx+D-Q~·u^(x),(OΞOT)-Q~-1·12(QTQ)·Vdx,

and

(OΞOT)-Q~-1·12(QTQ)·Q~-1=OOT+12OΞOTOΞ-1OT-12OΞ-1OTOΞOT:=Ω1,

where Ω1 is an antisymmetric matrix since OTO=OOT=Id.

Noting that Q(x)=Q~(x), xD+, thus, we have

-DQ·u^(x)·(Q·V)dx=-DQ~·u^(x)·(Q~·V)dx-D-u^(x),12(QTQ-QTQ)·Vdx+D-Q~·u^(x),Ω1·Q~·Vdx.

By (3.17), we get

-DQ~·u^(x)·(Q~·V)dx=D-u^(x),12(QTQ-QTQ)·Vdx-D-Q~·u^(x),Ω1·Q~·Vdx-DQTQ·u^(x),Q-1(νl(σ(u^)))·νl(u^),Vdx+DF(ρ(x))·Q·Vdx. 3.20

Noting that Q~T=Q~ and

Q~·u^(x),Q~-1·νl(u^)=0,

we have

-DQTQ·u^(x),Q-1(νl(σ(u^)))·νl(u^),Vdx=-DQ~·u^(x),Q~·Q-1(νl(σ(u^)))·Q~-1·νl(u^),Q~·Vdx=-DQ~·u^(x),Ω2·Q~·Vdx.

Thus, (3.20) implies

-DQ~·u^(x)·(Q~·V)dx=D-Q~·u^(x),12Q~-1·(QTQ-QTQ)·Q~-1·Q~·Vdx-D-Q~·u^(x),Ω1·Q~·Vdx-DQ~·u^(x),Ω2·Q~·Vdx+DF(ρ(x))·Q·Vdx.

This is (3.19). We proved the first result of the lemma.

Step 4 If uW2,p(D+), according to the property of Dσ, it is easy to see u^W2,p(D) since u satisfies free boundary condition. Computing directly, we have

eαu^(x)=Dσ|u(ρ(x))Du|ρ(x)Dρ|x(eα)=Dσ|u(ρ(x))DΠN|u(ρ(x))Du|ρ(x)Dρ|x(eα),xD-,

where ΠN:Nδ0N is the nearest projection map for some δ0-neighborhood of N in RN.

By direct computing, we obtain

Δu^(x)=D2(σΠN)|σ(u^)((uρ),(uρ))+Dσ(σ(u^))·F(ρ(x))=D2(σΠN)|σ(u^)(P(u^)·u^(x),P(u^)·u^(x))+P(σ(u^))·F(ρ(x)).

Combining this with the fact that u^ satisfies Eq. (1.4) in D+, the Eq. (3.11) follows immediately by taking

Υu^(·,·)=A(u^)(·,·)inD+,D2(σΠN)|σ(u^)(P(u^)·,P(u^)·)inD-;andF^=F(x)inD+,P(σ(u^))·F(ρ(x))inD-. 3.21

Now, applying Theorem 2.4, we derive the following

Theorem 3.4

Let FLp(D2+) for some p>1 and uW1,2(D2+,N) be a weak solution of (1.4) with free boundary u(0D2+) on K. Suppose uL2(D2+)+τ(u)Lp(D2+)ϵ3, then u(x)W2,p(D12+).

Proof

By Proposition 3.3, the extended u^W1,2(D,RN) is a weak solution to a system (2.3) with A satisfying (2.4) and with Ω satisfying |Ω|C|u^|. Then we can apply Theorem 2.4 (taking ϵ3 smaller if necessary) for 1<p<2 and bootstrap for p2 to prove the theorem.

Moreover, we have

Theorem 3.5

Let M be a compact Riemann surface with smooth boundary M, N a compact Riemannian manifold, and KN a smooth submanifold. Let FLp(M) for some p>1, and uH1(M,N) be a weak solution of (1.4) with free boundary u(M) on K, then uW2,p(M).

To end this section, we derive the removability of a local singularity at the free boundary (see Theorem 2.3 for the interior case).

Theorem 3.6

Let uWloc2,p(D+\{0},N), p>1 be a map with finite energy that satisfies

τ(u)=gLp(D+,TN),a.e.xD+, 3.22
u(x)K,du(x)(n)Tu(x)K,a.e.x0D+, 3.23

then u can be extended to a map belonging to W2,p(D+,N).

Proof

Applying a similar argument as in Lemma A.2 in [13], it is easy to see that u is a weak solution of (1.4) with F=g and with free boundary u(0D+) on K. By Theorem 3.4, we know uW2,p(Dr+) for some small r>0. Thus, uW2,p(D+).

Some basic analytic properties for the free boundary case

In this section, we will prove some basic lemmas for the free boundary case, such as small energy regularity (near the boundary), gap theorem, Pohozaev identity.

Firstly, we prove a small energy regularity lemma near the boundary.

Lemma 4.1

Let uW2,p(D2+,N), 1<p2 be a map with tension field τ(u)Lp(D2+) and with free boundary u(0D2+) on K. There exists ϵ4=ϵ4(p,N)>0, such that if uL2(D2+)+τ(u)Lp(D2+)ϵ4, then

u-1|D+|D+udxW2,p(D1/2+)C(p,N)(uLp(D+)+τ(u)Lp(D+)). 4.1

Moreover, by the Sobolev embedding W2,p(R2)C0(R2), we have

uOsc(D1/2+)=supx,yD1/2+|u(x)-u(y)|C(p,N)(uLp(D+)+τ(u)Lp(D+)). 4.2

Proof

By Proposition 3.3, we can extend u to u^W2,p(D) which is defined in D and satisfies

u^+Υu^(u^,u^)=F^inD. 4.3

where F=τ(u) in D+ and Υu^(·,·), F^(x) are defined by (3.21).

Firstly, we let 1<p<2. Take a cut-off function ηC0(D), such that 0η1, η|D3/41 and |η|C. Then, we have

Δ(ηu^)=ηΔu^+2ηu^+u^ΔηC(N)|u^||(ηu^)|+C(N)(|u^|+|u^|+|F^|).

Without loss of generality, we may assume 1|D+|D+u^dx=1|D+|D+udx=0. By the standard elliptic estimates, Sobolev’s embedding, Poincaré’s inequality and Proposition 3.3, we have

ηu^W2,p(D)C(p,N)|u^||(ηu^)|Lp(D)+C(p,N)(u^Lp(D)+u^Lp(D)+F^Lp(D))C(p,N)u^L2(D)(ηu^)L2p2-p(D)+C(p,N)(u^Lp(D)+τ(u)Lp(D+))C(p,N)ϵ4ηu^W2,p(D)+C(p,N)(uLp(D+)+τ(u)Lp(D+)),

where we also used the fact that u^Lp(D)C(N)uLp(D+), 1<p2.

Taking ϵ4 sufficiently small, we have

uW2,p(D1/2+)ηu^W2,p(D)C(p,N)(uLp(D+)+τ(u)Lp(D+)).

So, we have proved the lemma in the case 1<p<2. Next, if p=2, one can first derive the above estimate with p=43. Such an estimate gives a L4(D3/4+)- bound for u. Then one can apply the W2,2-boundary estimate to the equation and get the conclusion of lemma with p=2.

The gap theorem still holds for harmonic maps with free boundary.

Lemma 4.2

There exists a constant ϵ5=ϵ5(M,N)>0 such that if u is a smooth harmonic map from M to N with free boundary on K and satisfying

M|u|2dvolϵ5,

then u is a constant map.

Proof

By Lemmas 2.1, 3.2, and 4.1, take any x0M, then we may assume the image of u is contained in a Fermi-coordinate chart (BR0N(u(x0)),yi) of N. Thus, we can rewrite the equation in the new coordinate as follows:

-ΔMu+Γ(u)(u,u)=0,inM;ui(x)n=0,1ik,uj(x)=0,k+1jn,xM.

where Γ(u)(u,u)=gαβΓjki(u)ujxαukxβyi and Γjki are the Christoffel symbol of N in local coordinates {yi}i=1n.

Without loss of generality, we may assume Mui=0, 1ik. By standard elliptic estimates with Dirichlet boundary condition and Neumann boundary condition (see Lemma 2.6), we have

uW1,4/3(M)C(M)ΔMuL4/3(M)C(M,N)uL2(M)uL4(M)C(M,N)ϵ5uL4(M)C(M,N)ϵ5uW1,4/3(M).

If ϵ5 is small, then u is a constant map.

Next, we compute the Pohozaev identity which is similar to [24].

Lemma 4.3

For x00D+, let u(x)W2,2(D+(x0),N) be a map with tension field τ(u)L2(D+(x0)) and with free boundary u(0D+) on K. Then, for any 0<t<1, there holds

+Dt+x0rur2-12u2=Dt+x0rurτdx 4.4

where (r,θ)(0,1)×(0,π) are the polar coordinates at x0.

Proof

Since u(x) satisfies the equation

τ=Δu+A(u)(u,u)a.e.xD+(x0)

with the free boundary u(0D+) on K, multiplying (x-x0)u to both sides of the above equation and integrating by parts, for any 0<t<1, we get

Dt+(x0)τ·((x-x0)u)dx=Dt+(x0)Δu·((x-x0)u)dx=(Dt+(x0))un·((x-x0)u)-Dt+(x0)eαu·eα((x-x0)u)dx=+(Dt+(x0))un·((x-x0)u)-Dt+(x0)|u|2dx-12Dt+(x0)(x-x0)·|u|2dx=+(Dt+(x0))un·((x-x0)u)-12(Dt+(x0))x-x0,n|u|2=+(Dt+(x0))un·((x-x0)u)-12+(Dt+(x0))x-x0,n|u|2=+(Dt+(x0))rur2-12|u|2,

where the last second equality follows from the fact that x-x0,n=0 on 0Dt+(x0). Then the conclusion of lemma follows immediately.

Corollary 4.4

Under the assumptions of Lemma 4.3, we have

D2t+x0\Dt+x0ur2-12u2dxtuL2D+x0τL2D+x0.

Proof

From Lemma 4.3, we have

+Dt+x0ur2-12|u|2=Dt+x0rturτdxuL2D+x0τL2D+x0.

Integrating from t to 2t, we will get the conclusion of the corollary.

Energy identity and no neck property

In this section, we shall prove our main Theorem 1.1.

We first consider the following simpler case of a single boundary blow-up point.

Theorem 5.1

Let unW2,2(D1+(0),N) be a sequence of maps with tension fields τ(un) and with free boundaries un(0D+) on K and satisfying

  1. unW1,2(D+)+τ(un)L2(D+)Λ,

  2. unustronglyinWloc1,2(D+\{0},RN)asn,

  3. un(x)K,dun(x)(n)Tun(x)K,x0D+.

Then there exist a subsequence of un (still denoted by un) and a nonnegative integer L such that, for any i=1,...,L, there exist a point xni, positive numbers λni and a nonconstant harmonic sphere wi or a nonnegative constant ai and a nonconstant harmonic disk wi (which we view as a map from Rai2{}N) with free boundary wi(Rai2) on K such that:

  1. xni0,λni0, as n;

  2. dist(xni,0D+)λniai or dist(xni,0D+)λni(i.e.ai=), as n;

  3. limn(λniλnj+λnjλni+|xni-xnj|λni+λnj)= for any ij;

  4. wi is the weak limit of un(xni+λnix) in Wloc1,2(R2), if dist(xni,0D+)λni or wi is the weak limit of un(xni+λnix) in Wloc1,2(Rai2+), if dist(xni,0D+)λniai;

  5. Energy identity: we have
    limnE(un,D+)=E(u,D+)+i=1LE(wi). 5.1
  6. No neck property: The image
    u(D+)i=1Lwi(Rai2) 5.2
    is a connected set, where wi(Rai2)=wi(R2), if dist(xni,0D+)λni.

Proof of Theorem 5.1

Assume 0 is the only blow-up point of the sequence {un} in D+, i.e.

lim infnE(un;Dr+)ϵ¯28forallr>0 5.3

where ϵ¯=min{ϵ1,ϵ3,ϵ4}. By the standard argument of blow-up analysis we can assume that, for any n, there exist sequences xn0 and rn0 such that

E(un;Drn+(xn))=supxD+,rrnDr+(x)D+E(un;Dr+(x))=ϵ¯232. 5.4

Denoting dn=dist(xn,0D+), we have the following two cases:

Case 1lim supndnrn<.

Set

vn(x):=un(xn+rnx)

and

Bn:={xR2|xn+rnxD+}.

After taking a subsequence, we may assume limndnrn=a0. Then

BnRa2:={(x1,x2)|x2-a}.

It is easy to see that vn(x) is defined in Bn and satisfies

τ(vn(x))=Δvn(x)+A(vn(x))(vn(x),vn(x))xBn; 5.5
vn(x)K,dvn(x)(n)Tvn(x)K,ifxn+rnx0D+, 5.6

where τ(vn(x))=rn2τ(un(xn+rnx)).

Noting that for any x0Bn:={xR2|xn+rnx0D+} on the boundary,

vn(x)K,dvn(x)(n)Tvn(x)K,

since τ(vn)L2(Bn)rnτ(un)L2(D+)ϵ¯24 when n is big enough, by (5.4) and Lemma 4.1, we have

vnW2,2(D4R(0)Bn)C(R,N) 5.7

for any D4R(0)R2. Then there exist a subsequence of vn (also denoted by vn) and a nontrivial harmonic map v~1W2,2(Ra2) with free boundary v~1(Ra2) on K such that for any R>0, there hold

limnvn(x)-v~1(x)W1,2(DR(0)BnRa2)=0 5.8

and

limnvn(x)W1,2(DR(0)Bn)=v~1(x)W1,2(DR(0)Ra2). 5.9

In fact, by (5.7), we have

vnx-0,dnrnW2,2(D3R+(0))C(R,N) 5.10

when n is big enough. Then there exist a subsequence of vn (also denoted by vn) and a harmonic map v~W2,2(D3R+(0)) such that

limnvnx-0,dnrn-v~(x)W1,2(D3R+(0))=0

and vnx-0,dnrnv~(x), dvnx-0,dnrndndv~dn(x), a.e.x0D3R+(0) as n. Set

v~1(x):=v~(x+(0,a)),

then v~1W2,2(Ra2D2R(0)) is a harmonic map with free boundary v~1(Ra2D2R(0)) on K such that

limnvn(x)-v~1(x)W1,2(D2R(0)BnRa2)=0.

Lastly, (5.9) follows from (5.7), (5.8), Sobolev embedding, Young’s inequality and the fact that the measure of D2R(0)Bn\Ra2 goes to zero as n.

In addition, E(v~1;D1(0)Ra2)=ϵ¯232. By the conformal invariance of harmonic maps and the removable singularity Theorem 3.6, v~1(x) can be extended to a nontrivial harmonic disk.

Case 2lim supndnrn=.

In this case, we can see that vn(x) is defined in Bn which tends to R2 as n. Moreover, for any xR2, when n is sufficiently large, by (5.4), we have

E(vn;D1(x))ϵ¯232. 5.11

According to Lemma 2.1, there exist a subsequence of vn (we still denote it by vn) and a harmonic map v1(x)W1,2(R2,N) such that

limnvn(x)=v1(x)inWloc1,2(R2).

Besides, we know E(v1;D1(0))=ϵ¯232. By the standard theory of harmonic maps, v1(x) can be extended to a nontrivial harmonic sphere. We call the above harmonic sphere v1(x) or harmonic disk v~1(x) the first bubble.

We will split the proof of Theorem 5.1 into two parts, energy identity and no neck result. Now, we begin to prove the energy identity.

Energy identity: By the standard induction argument in [6], we only need to prove the theorem in the case where there is only one bubble.

By Lemmas 2.1 and 4.1, there exist a subsequence of un (still denoted by un) and a weak limit uW2,2(D+) such that

limδ0limnE(un;D+\Dδ+(xn))=E(u;D+).

So, in both cases, the energy identity is equivalent to

limRlimδ0limnE(un;Dδ+(xn)\DrnR+(xn))=0. 5.12

To prove the no neck property, i.e. that the sets u(D+) and v(R2) or v(Ra2) are connected, it is enough to show that

limRlimδ0limnunOsc(Dδ+(xn)\DrnR+(xn))=0. 5.13

Step 1 We prove the energy identity for Case 1, i.e., limndnrn=a<.

Under the “one bubble” assumption, we first make the following:

Claim: for any ϵ>0, there exist δ>0 and R>0 such that

D8t+(xn)\Dt+(xn)|un|2dxϵ2foranyt12rnR,2δ 5.14

when n is large enough.

In fact, if (5.14) is not true, then we can find tn0, such that limntnrn= and

D8tn+(xn)\Dtn+(xn)|un|2dxϵ6>0. 5.15

Then we have

limndntn=0.

We set

wn(x):=un(xn+tnx)

and

Bn:={xR2|xn+tnxD+}.

Then wn(x) lives in Bn which tends to R+2 as n. It is easy to see that 0 is an energy concentration point for wn. We have to consider the following two cases:

(a)wn has no other energy concentration points except 0.

By Lemmas 2.1, 4.1 and the process of constructing the first bubble, passing to a subsequence, we may assume that wn converges to a harmonic map w(x):R+2N with free boundary w(R+2) on K satisfying, for any R>0,

supλ>0limnwn(x)-w(x)W1,2((DR(0)Bn)\Dλ(0))=0.

Noting that (5.15) implies

(D8\D1)Bn|w|2dx=limn(D8\D1)Bn|wn|2dxϵ6>0. 5.16

By the conformal invariance of harmonic map and Theorem 3.6, w(x) is a nontrivial harmonic disk which can be seen as the second bubble. This contradicts the “one bubble” assumption.

(b)wn has another energy concentration point p0.

Without loss of generality, we may assume p is the only energy concentration point in Dr0+(p) for some r0>0. Similar to the process of constructing the first bubble, there exist xnp and rn0 such that

E(wn;Drn+(xn)Bn)=supxDr0+(p),rrnDr+(x)Dr0+(p)E(wn;Dr+(x)Bn)=ϵ¯232. 5.17

By (5.4), we know rntnrn. Then, passing to a subsequence we may assume limndnrntn=d[0,a]. Moreover, there exists a nontrivial harmonic map v~2(x):Rd2N with free boundary v~2(Rd2) on K satisfying, for any R>0,

limnwn(xn+rnx)-v~2(x)W1,2(DR(0)Bn)=0.

where Bn:={xR2|xn+rnxBn}. That is

limnun(xn+tnxn+tnrnx)-v~2(x)W1,2(DR(0)Bn)=0. 5.18

Therefore, v~2(x) is also a bubble for the sequence un. This is also contradiction to the ”one bubble” assumption. Thus, we proved Claim (5.14).

Let xn0D+ be the projection of xn, i.e. dn=dist(xn,0D+)=|xn-xn|. Firstly, we decompose the neck domain Dδ+(xn)\DrnR+(xn) as follows

Dδ+(xn)\DrnR+(xn)=Dδ+(xn)\Dδ2+(xn)Dδ2+(xn)\D2rnR+(xn)D2rnR+(xn)\DrnR+(xn):=Ω1Ω2Ω3,

when n and R are large.

Since limndnrn=a, when n and R are large enough, it is easy to see that

Ω1Dδ+(xn)\Dδ4+(xn)andΩ3D4rnR+(xn)\DrnR+(xn).

Moreover, for any 2rnRt12δ, there holds

D2t+(xn)\Dt+(xn)D4t+(xn)\Dt/2+(xn).

By assumption (5.14), we have

E(un;Ω1)+E(un;Ω3)ϵ2 5.19

and

D2t+(xn)\Dt+(xn)|un|2dxϵ2foranyt(2rnR,12δ). 5.20

By a scaling argument, we may assume

unL2(D4t+(xn)\Dt/2+(xn))+τ(un)Lp(D4t+(xn)\Dt/2+(xn))ϵ¯.

According to the small energy regularity theory Lemmas 2.1 and 4.1, we obtain

OscD2t+(xn)\Dt+(xn)unC(unL2(D4t+(xn)\Dt/2+(xn))+tτ(un)L2(D4t+(xn)\Dt/2+(xn))) 5.21

for any t(2rnR,12δ). Thus, un(Ω2)Kδ0 and we can extend the definition of un to the domain Ω^2:=Dδ2(xn)\D2rnR(xn) by defining u^n as (3.8). Then u^nW2,2(Ω^2) and satisfies Eq. (3.11) where we take Fn(x)=τ(un)(x) and define Υun^(·,·), Fn^(x) as in (3.21).

Define

u^n(r):=12πrDr(xn)u^n.

Then by (5.21), we have

u^n(x)-u^n(x)L(Ω^2)sup2rnRtδ2u^n(x)-u^n(x)L(D2t(xn)\Dt(xn))C(1+DσL)OscD2t+(xn)\Dt+(xn)unC(N)(ϵ+δ).

We have

Ω^2u^n(u^n-u^n)dx=Ω^2u^nn(u^n-u^n)-Ω^2Δu^n(u^n-u^n)dx.

On the one hand, by Jessen’s inequality, we have

Ω^2u^nu^n-u^ndx=Ω^2u^n2dx-Ω^2u^nru^nrdxΩ^2u^n2dx-Ω^2u^nr2dx1/2Ω^212π02πu^nrr,θdθ2dx1/2Ω^2u^n2dx-Ω^2u^nr2dx=12Ω^2u^n2dx-Ω^2u^nr2-12u^n2dx.

On the other hand, using Eq. (3.11), we get

-Ω^2Δu^n(u^n-u^n)dxΩ^2|Υu^n(u^n,u^n)+Fn^||u^n-u^n|dxC(ϵ+δ)Ω^2|u^n|2dx+C(ϵ+δ)Ω^2|Fn^|dxC(ϵ+δ)Ω^2|u^n|2dx+C(ϵ+δ)τnL2(Ω2).

Thus,

12-Cϵ+δΩ^2|u^n|2dxΩ^2u^nnu^n-u^n+Ω^2u^nr2-12|u^n|2dx+Cϵ+δ. 5.22

By the definition of u^n (see (3.8)), we obtain

Ω^2u^nr2-12u^n2dx=Ω2unr2-12un2dx+Ω^2\Ω2Dσ·unρxr2-12Dσ·unρx2dx=Ω2unr2-12un2dx+Ω2Dσ·unxr2-12Dσ·unx2dx.

Note that

Dσ·unxr2=Punx·unxr,Punx·unxr=PTP·unxr,unxr=PTP-Idunxr,unxr+unxr2,

where P is the matrix corresponding to the linear operator defined by (3.9) under the orthonormal basis of RN.

Similarly,

|Dσ·un(x)|2=(PTP-Id)un(x),un(x)+|un(x)|2.

Noting that Ξ|K=Id, by the continuity of eigenvalues of PTP, we have that for any δ>0, there exists a constant δ1=δ1(δ)>0, such that for any ξRn and yKδ1, there holds

PT(y)P(y)ξ,ξ(1+δ)|ξ|2.

By (5.21), we have dist(un,K)L(Ω2)C(ϵ+δ). Thus, for any δ>0, ξRn, there holds

(PT(un(x))P(un(x))-Id)ξ,ξδ|ξ|2

when ϵ and δ are small enough.

Thus,

Ω^2u^nr2-12u^n2dx2Ω2unr2-12un2dx+CδΩ2unx2dx=2i=1mnD2i2rnR+xn\D2i-12rnR+xnunr2-12un2dx+CδΩ2unx2dxCi=1mn2irnR+CδΩ2unx2dxCδ+CδΩ2unx2dx, 5.23

where the last second inequality follows from Corollary 4.4.

Combining inequality (5.22) with (5.23), we have

12-C(ϵ+δ+δ)Ω^2|u^n|2dxΩ^2u^nn(u^n-u^n)+C(ϵ+δ). 5.24

As for the boundary term, by trace theory, we have

D12δ(xn)u^nn(u^n-u^n)C(ϵ+δ)+D12δ(xn)|un|C(ϵ+δ)unL2(D12δ+(xn)\D14δ+(xn))+δ2unL2(D12δ+(xn)\D14δ+(xn))C(ϵ+δ)unL2(Dδ+(xn)\D18δ+(xn))+δτnL2(Dδ+(xn)\D18δ+(xn))C(ϵ+δ),

where the last second inequality can be derived from Lemmas 2.1 and 4.1.

Also, there holds

D2rnR(xn)u^nn(u^n-u^n)C(ϵ+δ).

Therefore, combining these results and taking ϵ and δ in (5.24) sufficiently small (then δ is small), we have

Ω2|un|2dxΩ^2|u^n|2dxC(δ+ϵ). 5.25

Then the equality (5.12) follows from (5.19) and (5.25). We proved the energy identity for the Case 1.

Step 2 We prove the energy identity for Case 2, i.e., lim supndnrn=.

The proof is similar to the Case 1. Firstly, we need to show the Claim (5.14) also holds in this case.

In fact, if (5.14) is not true, then we can find tn0, such that limntnrn= and

D8tn+(xn)\Dtn+(xn)|un|2dxϵ6>0. 5.26

Then passing to a subsequence, we may assume

limndntn=b[0,].

We set

wn(x):=un(xn+tnx)

and

Bn:={xR2|xn+tnxD+}.

Then wn(x) lives in Bn and 0 is an energy concentration point for wn. We have to consider the following two cases:

(c)b<.

Then Bn tends to Rb2 as n. Here, we also need to consider two cases.

(i)wn has no other energy concentration points except 0. It is almost the same as Case(a) in Step 1 where by passing to a subsequence, wn converges to a nontrivial harmonic map w(x):Rb2N with free boundary w(Rb2) on K which can be seen as the second bubble. This is a contradiction to the ”one bubble” assumption.

(ii)wn has another energy concentration point p0. Similar to the process of Case(b) in Step 1, there exist xnp and rn0 such that (5.17) holds. Then, passing to a subsequence, we may assume

limndnrntn=d[0,].

Moreover, if d[0,), then there exists a nontrivial harmonic map v~2(x):Rd2N with free boundary v~2(Rd2) on K satisfying (5.18) as in Case(b). If d=, by the process of constructing the first bubble in Case 2, there exists v2(x):R2N is a nontrivial harmonic map such that

wn(xn+rnx)v2(x)inWloc1,2(R2),

that is

un(xn+tnxn+tnrnx)v2(x)inWloc1,2(R2).

In both cases, we will get the second bubble v2(x) or v~2(x). This contradicts the ”one bubble” assumption.

(d)b=.

Then Bn tends to R2 as n. Again, we need to consider two cases.

(iii)wn has no other energy concentration points except 0. By Lemma 2.1, Theorem 2.3 and (5.26), there exists v2(x):R2N is a nontrivial harmonic map such that

wn(x)v2(x)inWloc1,2(R2\{0}).

Then, we get the second bubble v2(x) which contradicts the ”one bubble” assumption.

(iv)wn has another energy concentration point p0. Similar to Case(b) in Step 1, there exist xnp and rn0 such that (5.17) holds and passing to a subsequence, we have

limndnrntn=.

Moreover, by the process of constructing the first bubble in Case 2, there exists a nontrivial harmonic map v2(x):R2N such that

wn(xn+rnx)v2(x)inWloc1,2(R2),

that is

un(xn+tnxn+tnrnx)v2(x)inWloc1,2(R2).

This is also a contradiction to the ”one bubble” assumption. Thus, we proved our Claim (5.14).

Secondly, we decompose the neck domain Dδ+(xn)\DrnR+(xn) as follows

Dδ+(xn)\DrnR+(xn)=Dδ+(xn)\Dδ2+(xn)Dδ2+(xn)\D2dn+(xn)D2dn+(xn)\Ddn+(xn)Ddn+(xn)\DrnR+(xn):=Ω1Ω2Ω3Ω4,

when n is large.

Since limndn=0 and limndnrn=, when n is large enough, it is easy to see that

Ω1Dδ+(xn)\Dδ4+(xn),andΩ3D4dn+(xn)\Ddn+(xn).

Moreover, for any 2dnt12δ, there holds

D2t+(xn)\Dt+(xn)D4t+(xn)\Dt/2+(xn).

By assumption (5.14), we have

E(un;Ω1)+E(un;Ω3)ϵ2 5.27

and

D2t+(xn)\Dt+(xn)|un|2dxϵ2foranyt2dn,12δ. 5.28

Noting that Ω4=Ddn+(xn)\DrnR+(xn)=Ddn(xn)\DrnR(xn), by the well-known blow-up analysis theory of harmonic maps with interior blow-up points (also a sequence of maps with uniformly Lp bounded tension fields for some p65), there holds

limRlimn0E(un;Ddn(xn)\DrnR(xn))=0. 5.29

and

limRlimn0Osc(un)Ddn(xn)\DrnR(xn)=0. 5.30

See [6, 20, 32] for details.

Lastly, to estimate the energy concentration in Ω2, we can use the same argument as in the previous Case 1 to get

Ω2|un|2dxC(δ+ϵ). 5.31

Combining (5.27), (5.29) with (5.31), it is easy to obtain (5.12). We proved the energy identity.

Next, we prove the no neck property in Theorem 5.1, i.e., the base map and the bubbles are connected in the target manifold.

No neck property: Here, we also need to consider two cases. But, for Case 2, we use the same argument as in the previous reasoning where we split the neck domain into two parts, an interior domain and a boundary domain. Then, with the help of the no neck results in [20, 32] for a sequence of maps with uniformly L2-bounded tension fields, we just need to prove (5.13) for Case 1.

We may assume limndnrn=a and decompose the neck domain Dδ+(xn)\DrnR+(xn)=Ω1Ω2Ω3, when n and R are large.

By assumption (5.14) and small energy regularity (Lemmas 2.1 and 4.1), we have

unOscDδ+xn\Dδ4+xnunOscDδ+xn\Dδ5+xnCunL2D4δ3+xn\Dδ6+xn+δτnL2D4δ3+xn\Dδ6+xnCϵ+δ 5.32

and

unOscD4rnR+xn\DrnR+xnunOscD5rnR+xn\DrnR+xnCunL2D6rnR+xn\D3rnR4+xn+rnRτnL2D6rnR+xn\D3rnR4+xnCϵ+δ, 5.33

when n, R are large and δ is small.

Without loss of generality, we may assume 12δ=2mn(2rnR) where mn as n. Inspired by a technique by Ding [5] for the interior bubbling case, we set Q(t):=D2t0+t2rnR+(xn)\D2t0-t2rnR+(xn), Q^(t):=D2t0+t2rnR(xn)\D2t0-t2rnR(xn) and define

f(t):=Q(t)|un|2dx,

where 0t0mn and 0tmin{t0,mn-t0}.

Similar to the proof of (5.22) and (5.23), we have

12-Cϵ+δQ^t|u^n|2dxQ^tu^nnu^n-u^n+Q^tu^nr2-12|u^n|2dx+Cϵ+δQt|τn|dx 5.34

and

Q^tu^nr2-12|u^n|2dx2Qtunr2-12|un|2dx+CδQt|unx|2dxC2t0+trnR+CδQt|unx|2dx 5.35

where the last inequality follows from Corollary 4.4.

As for the boundary, by Poincaré’s inequality, we have

D2t0+t2rnRxnu^nnu^n-u^nD2t0+t2rnRxnu^nr212D2t0+t2rnRxnu^n-u^n212CD2t0+t2rnRxnu^nr2122t0+t2rnR02πu^nθ212C2t0+t2rnRD2t0+t2rnRxnu^n2C2t0+t2rnR+D2t0+t2rnR+xnun2.

Similarly, we get

D2t0-t2rnRxnu^nnu^n-u^nC2t0-t2rnR+D2t0-t2rnR+xnun2.

Using these together, we have

12-Cϵ+δ+δQ^tu^n2dxC2t0+t2rnR+D2t0+t2rnR+xnun2+C2t0-t2rnR+D2t0-t2rnR+xnun2+C2t0+trnR+Cϵ+δQtτndx.

Taking ϵ and δ sufficiently small, we get

Q(t)|un|2dxC2t0+t2rnR+(D2t0+t2rnR+(xn))|un|2+C2t0-t2rnR+(D2t0-t2rnR+(xn))|un|2+C2t0+trnR.

Therefore,

f(t)Clog2f(t)+C2t0+trnR. 5.36

Thus,

2-1Ctf(t)-C2t0+(1-1/C)trnR.

Integrating from 2 to L, we arrive at

f(2)C2-1CLf(L)+C2t0rnR2L2(1-1/C)tdtC2-1CLf(L)+C2t0rnR2(1-1/C)L.

Now, let t0=i and L=Li:=min{i,mn-i}. Then, we have Q(Li)Dδ/2+(xn)\D2rnR+(xn)Dδ+(xn)\DrnR+(xn) and

D2i+22rnR+xn\D2i-22rnR+xn|un|2dxCEun,Dδ+xn\DrnR+xn2-1CLi+C2irnR21-1/CLiCEun,Dδ+xn\DrnR+xn2-1CLi+C2irnR21-1/Cmn-iCEun,Dδ+xn\DrnR+xn2-1CLi+Cδ2-1/Cmn-iCϵ2-1CLi+Cδ2-1/Cmn-i,

where we used the energy identity (5.12).

By Lemmas 2.1 and 4.1, we obtain

OscD2i+12rnR+(xn)\D2i-12rnR+(xn)unCunL2(D2i+22rnR+(xn)\D2i-22rnR+(xn))+(2i+22rnR)τnL2(D2i+22rnR+(xn)\D2i-22rnR+(xn))CunL2(D2i+22rnR+(xn)\D2i-22rnR+(xn))+2irnR.

Summing over i from 2 to mn-2, we have

unOsc(Dδ/4+(xn)\D4rnR+(xn))i=2mn-2unOsc(D2i+12rnR+(xn)\D2i-12rnR+(xn))Ci=2mn-2ϵ2-1CLi+δ2(-1/C)(mn-i)+2irnRCi=2mn-22-1Ci(ϵ+δ)+CδC(ϵ+δ).

This inequality and (5.32), (5.33) imply (5.13) and we have proved there is no neck during the blow-up process.

Now, we can prove Theorem 1.1.

Proof of Theorem 1.1

Combining the blow-up theory of a sequence of maps with uniformly L2-bounded tension fields from a closed Riemann surface (see [6, 20, 24, 26, 32]) and Theorem 5.1, we can easily get the conclusion of Theorem 1.1 by following the standard blow-up scheme in [6].

On the other hand, it is well known that harmonic spheres are minimal spheres and harmonic disks with free boundary on K are minimal disks with free boundary on K (see e.g. the proof of Theorem B in [27], page 300).

Application to the harmonic map flow with free boundary

In this section, we will apply the results in Theorem 1.1 to the harmonic map flow with free boundary and prove Theorem 1.2 and Theorem 1.3.

Firstly, we have

Lemma 6.1

Let u:M×(0,)N be a global weak solution to (1.7-1.10), which is smooth away from a finite number of singular points. There holds the estimate

0M|tu|2dxdtE(u0). 6.1

Moreover, E(u(·,t)) is continuous on [0,) and non-increasing.

Proof

The proof is similar to Lemma 3.4 in [44]. Multiply the equation (1.7) by tu and integrate by parts, for any 0t1t2, to get

t1t2M|tu|2dxdt=t1t2M-Δgu·tudxdt=t1t2Mun·tu-t1t2Mu·(tu)dxdt=-t1t2M12t|u|2dxdt=E(u(·,t1))-E(u(·,t2)),

where n is the outward unit normal vector field on M and we used the free boundary condition that untu. Then the conclusion of the lemma follows immediately.

Similar to the case of a closed domain (see Lemma 2.5 in [24]), we have

Lemma 6.2

Let uC(M×(0,T0),N) be a solution to (1.71.10). Then there exists a constant R0>0 such that, for any x0M, 0<ts<T0 and 0<RR0, there hold:

E(u(s);BRM(x0))E(u(t);B2RM(x0))+Cs-tR2E(u0), 6.2

and

E(u(t);BRM(x0))E(u(s);B2RM(x0))+CtsM|tu|2dxdt+Cs-tR2E(u0). 6.3

Proof

Let ηC0(B2RM(x0)) be such that 0η1, η|BRM(x0)1 and |η|CR. Multiplying (1.7) by η2tu and integrating by parts, we get

M|tu|2η2dx+ddt(12M|u|2η2dx)=Mun·tuη2-2Mtuuηηdx=-2Mtuuηηdx,

where we used the free boundary condition that untu.

Since

|2Mtuuηηdx|12M|tu|2η2dx+2M|u|2|η|2dx,

we have

-32M|tu|2η2dx-2M|u|2|η|2dxddt12M|u|2η2dx2M|u|2|η|2dx.

Integrating the above inequality from t to s, we will get the conclusion of the lemma.

With the help of Lemma 6.2, we can apply the standard argument for the closed case (see Lemma 4.1 in [24]) to obtain the following:

Lemma 6.3

Let uC(M×(0,T0),N) be a solution to (1.71.10). Suppose x0M is the only singular point at time T0. Then there exists a positive number m>0 such that

|u|2(x,t)dxmδx0+|u|2(x,T0)dx, 6.4

for tT0, as Radon measures. Here, δx0 denotes the δ-mass at x0.

Now, we begin to prove Theorems 1.2 and 1.3. Firstly, it is easy to see that Lemmas 6.1, 6.3 and Theorem 1.1 imply Theorem 1.2. In fact,

Proof of Theorem 1.2

By Lemma 6.1, we can find a sequence tn such that

limnM|tu|2(·,tn)dx=0andE(u(·,tn))E(u0).

Take the sequence un=u(·,tn), τ(un)=tu(·,tn) in Theorem 1.1. Combining this with Lemma 6.3, the conclusion of Theorem 1.2 follows immediately.

Proof of Theorem 1.3

It is sufficient to consider the case that (x0,T0) with x0M being the only singular point at time T0. For the case of an interior singularity x0M\M, one can refer to [24]. Without loss of generality, we may assume M=D1+(0) and x0=0. By Lemma 6.3, there exist sequences tnT0 and λn0 such that

limnDλn+(0)|u|2(·,tn)dx=m.

Let un(x,t)=u(λnx,tn+λn2t). Without loss of generality, we may assume tn-2λn2>0. Then un is defined in Dλn-1+(0)×[-2,0] satisfying (1.7) and

-20Dλn-1+(0)|tun|2dxdt=tn-2λn2tnD1+(0)|tu|2dxdt0

as n. By Fubini’s theorem, there exists sn(-1,-12) such that

limnDλn-1+(0)|tun|2(·,sn)dx=0. 6.5

For the sequence {un(·,sn)}, there holds

limRlimnDR+(0)|un|2(·,sn)dx=m. 6.6

In fact, on the one hand, by (6.2), we have

DR+(0)|un|2(·,sn)dx=DλnR+(0)|u|2(·,tn+λn2sn)dxDλn+(0)|u|2(·,tn)dx-C1R2E(u0).

Thus,

limRlimnDR+(0)|un|2(·,sn)dxm. 6.7

On the other hand, by (6.4), for any R>0 and σ>0, we have

limnDλnR+(0)|u|2(·,tn+λn2sn)dxlimnDσ+(0)|u|2(·,tn+λn2sn)dx=m+Dσ+(0)|u|2(·,T0)dx.

Letting σ0, we obtain

limnDR+(0)|un|2(·,sn)dxm 6.8

and (6.6) follows immediately.

Fixing R>0, we consider the sequence {un(·,sn)}n=1 which is defined in DR+(0). By (6.8) and (6.5), we know it is a sequence of maps from DR+(0) to N with finite energy and tension fields

τnL2(DR+(0))=tun(·,sn)L2(DR+(0))0

as n. Moreover, for each R>0, un(·,sn) weakly converges to a constant map. In fact, by Lemma 6.3, for any σ>0, we have

limnE(un(·,sn),DR+\Dσ+)=limnE(u(·,tn+λn2sn),DλnR+\Dλnσ+)limnE(u(·,T0),DλnR)=0.

According to Theorem 5.1, we know there exist LR nontrivial bubbles {wRi}i=1LR such that

limnE(un(·,sn),DR+)=i=1LRE(wRi). 6.9

Since the energy of the bubble has a lower bound, i.e. E(w)ϵ0¯:=min{ϵ0,ϵ5}, we have 1LRmϵ0¯+1. Therefore, there exist a subsequence R and a constant L[1,mϵ0¯+1] such that LR=L and

m=limRlimnE(un(·,sn),DR+)=limRi=1LE(wRi). 6.10

Using Theorem 1.1 with M=S2 or M=D and τ0, there exist Li bubbles {wj}j=1Li such that

limRE(wRi)=j=1LiE(wj).

Then

m=limRlimnE(un(·,sn),DR+)=limRi=1LE(wRi)=i=1Lj=1LiE(wj). 6.11

Combining with Lemma 6.3, we obtain the conclusion of Theorem 1.3.

Acknowledgements

Open access funding provided by Max Planck Society.

Footnotes

The research leading to these results has received funding from the European Research Council under the European Union’s Seventh Framework Program (FP7/2007-2013)/ERC Grant agreement no. 267087. Miaomiao Zhu was supported in part by National Natural Science Foundation of China (No. 11601325). We would like to thank the referee for careful comments and useful suggestions in improving the presentation of the paper.

Contributor Information

Jürgen Jost, Email: jost@mis.mpg.de.

Lei Liu, Email: leiliu@mis.mpg.de, Email: llei1988@mail.ustc.edu.cn.

Miaomiao Zhu, Email: mizhu@sjtu.edu.cn.

References

  • 1.Chang KC. Heat flow and boundary value problem for harmonic maps. Ann. Inst. Henri. Poincaré, Anal. Non Lineaire. 1989;6(5):363–395. doi: 10.1016/S0294-1449(16)30316-X. [DOI] [Google Scholar]
  • 2.Chen Y, Lin F. Evolution equations with a free boundary condition. J. Geom. Anal. 1998;8(2):179–197. doi: 10.1007/BF02921640. [DOI] [Google Scholar]
  • 3.Chen Q, Jost J, Wang G, Zhu M. The boundary value problem for Dirac-harmonic maps. J. Eur. Math. Soc. (JEMS) 2013;15(3):997–1031. doi: 10.4171/JEMS/384. [DOI] [Google Scholar]
  • 4.Colding T, Minicozzi W. Width and finite extinction time of Ricci flow. Geom. Topol. 2008;12(5):2537–2586. doi: 10.2140/gt.2008.12.2537. [DOI] [Google Scholar]
  • 5.Ding, W.: Lectures on Heat Flow of Harmonic Maps. Lecture notes at CTS, NTHU, Taiwan (1998)
  • 6.Ding W, Tian G. Energy identity for a class of approximate harmonic maps from surfaces. Comm. Anal. Geom. 1995;3(3–4):543–554. doi: 10.4310/CAG.1995.v3.n4.a1. [DOI] [Google Scholar]
  • 7.Frauenfelder U. Gromov convergence of pseudoholomorphic disks. J. Fixed Point Theory Appl. 2008;3(2):215–271. doi: 10.1007/s11784-008-0078-1. [DOI] [Google Scholar]
  • 8.Gulliver R, Jost J. Harmonic maps which solve a free-boundary problem. J. Reine Angew. Math. 1987;381:61–89. [Google Scholar]
  • 9.Hamilton, R.: Harmonic Maps of Manofolds with Boundary. L. N. in Math. 471 Springer, New York (1975)
  • 10.Hélein, F.: Harmonic maps, conservation laws and moving frames, volume 150 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, second edition, 2002. Translated from the 1996 French original, With a foreword by James Eells. (1983)
  • 11.Hong M, Yin H. On the Sacks-Uhlenbeck flow of Riemannian surfaces. Comm. Anal. Geom. 2013;21(5):917–955. doi: 10.4310/CAG.2013.v21.n5.a3. [DOI] [Google Scholar]
  • 12.Ivashkovich, S., Shevchishin, V.: Gromov compactness theorem for J-complex curves with boundary. Int. Math. Res. Notices (22), 1167–1206 (2000)
  • 13.Jost J. Two-Dimensional Geometric Variational Problems. New York: Wiley; 1991. [Google Scholar]
  • 14.Jost J. Geometry and Physics. New York: Springer; 2009. [Google Scholar]
  • 15.Lamm T. Energy identity for approximations of harmonic maps from surfaces. Trans. Am. Math. Soc. 2010;362:4077–4097. doi: 10.1090/S0002-9947-10-04912-3. [DOI] [Google Scholar]
  • 16.Lamm T, Sharp B. Global estimates and energy identities for elliptic systems with antisymmetric potentials. Comm. Part. Differ. Equ. 2016;41:579–608. doi: 10.1080/03605302.2015.1116559. [DOI] [Google Scholar]
  • 17.Laurain P, Petrides R. Regularity and quantification for harmonic maps with free boundary. Adv. Calc. Var. 2017;10(1):69–82. doi: 10.1515/acv-2015-0026. [DOI] [Google Scholar]
  • 18.Laurain P, Rivière T. Angular energy quantization for linear elliptic systems with antisymmetric potentials and applications. Anal. PDE. 2014;7(1):1–41. doi: 10.2140/apde.2014.7.1. [DOI] [Google Scholar]
  • 19.Li J. Heat flows and harmonic maps with a free boundary. Math. Z. 1994;217(3):487–495. doi: 10.1007/BF02571957. [DOI] [Google Scholar]
  • 20.Li J, Zhu X. Energy identity for the maps from a surface with tension field bounded in Lp. Pac. J. Math. 2012;260(1):181–195. doi: 10.2140/pjm.2012.260.181. [DOI] [Google Scholar]
  • 21.Li J, Zhu X. Small energy compactness for approximate harmonic mappings. Commun. Contemp. Math. 2011;13(5):741–763. doi: 10.1142/S0219199711004427. [DOI] [Google Scholar]
  • 22.Li Y, Wang Y. Bubbling location for sequences of approximated f-harmonic maps from surfaces. Internat. J. Math. 2010;21(4):475–495. doi: 10.1142/S0129167X10006136. [DOI] [Google Scholar]
  • 23.Li Y, Wang Y. A weak energy identity and the length of necks for a sequence of Sacks-Uhlenbeck α-harmonic maps. Adv. Math. 2010;225(3):1134–1184. doi: 10.1016/j.aim.2010.03.020. [DOI] [Google Scholar]
  • 24.Lin F, Wang C. Energy identity of harmonic map flow from surfaces at finite singular time. Calc. Var. Par. Differ. Equ. 1998;6:369–380. doi: 10.1007/s005260050095. [DOI] [Google Scholar]
  • 25.Lin F, Rivière T. Energy quantization for harmonic maps. Duke Math. J. 2002;111(1):177–193. doi: 10.1215/S0012-7094-02-11116-8. [DOI] [Google Scholar]
  • 26.Luo Y. Energy identity and removable singularities of maps from a Riemannian surface with tension field unbounded in L2. Pac. J. Math. 2012;256(2):365–380. doi: 10.2140/pjm.2012.256.365. [DOI] [Google Scholar]
  • 27.Ma L. Harmonic map heat flow with free boundary. Comm. Math. Hel. 1991;66:279–301. doi: 10.1007/BF02566648. [DOI] [Google Scholar]
  • 28.McDuff D, Salamon D. J-Holomorphic Curves and Symplectic Topology. New York: AMS Colloquium Publications; 2004. [Google Scholar]
  • 29.Parker T. Bubble tree convergence for harmonic maps. J. Diff. Geom. 1996;44(3):595–633. doi: 10.4310/jdg/1214459224. [DOI] [Google Scholar]
  • 30.Parker T, Wolfson J. Pseudo-holomorpohic maps and bubble trees. J. Geometr. Anal. 1993;3(1):63–98. doi: 10.1007/BF02921330. [DOI] [Google Scholar]
  • 31.Qing J. On singularities of the heat flow for harmonic maps from surface into spheres. Comm. Anal. Geom. 1995;3(1–2):297–315. doi: 10.4310/CAG.1995.v3.n2.a4. [DOI] [Google Scholar]
  • 32.Qing J, Tian G. Bubbling of the heat flows for harmonic maps from surfaces. Commun. Pure Appl. Math. 1997;50(4):295–310. doi: 10.1002/(SICI)1097-0312(199704)50:4&#x0003c;295::AID-CPA1&#x0003e;3.0.CO;2-5. [DOI] [Google Scholar]
  • 33.Rivière T. Conservation laws for conformally invariant variational problems. Invent. Math. 2007;168:1–22. doi: 10.1007/s00222-006-0023-0. [DOI] [Google Scholar]
  • 34.Rivière, T.: Conformally Invariant 2-Dimensional Variational Problems Cours joint de l’Institut Henri Poincaré—Paris XII Creteil (2010)
  • 35.Rivière T, Struwe M. Partial regularity for harmonic maps and related problems. Comm. Pure Appl. Math. 2008;61(4):451–463. doi: 10.1002/cpa.20205. [DOI] [Google Scholar]
  • 36.Roger M. An Lp regularity theory for harmonic maps. Trans. Am. Math. Soc. 2015;367(1):1–30. [Google Scholar]
  • 37.Rupflin M. An improved uniqueness result for the harmonic map flow in two dimensions. Calc. Var. Par. Differ. Equ. 2008;33(3):329–341. doi: 10.1007/s00526-008-0164-7. [DOI] [Google Scholar]
  • 38.Sacks J, Uhlenbeck K. The existence of minimal immersions of 2-spheres. Ann. Math. 1981;113:1–24. doi: 10.2307/1971131. [DOI] [Google Scholar]
  • 39.Scheven C. Partial regularity for stationary harmonic maps at a free boundary. Math. Z. 2006;253(1):135–157. doi: 10.1007/s00209-005-0891-9. [DOI] [Google Scholar]
  • 40.Schikorra A. A remark on gauge transformations and the moving frame method. Ann. Inst. H. Poincaré Anal. Non Linéaire. 2010;27(2):503–515. doi: 10.1016/j.anihpc.2009.09.004. [DOI] [Google Scholar]
  • 41.Sharp B. Higher integrability for solutions to a system of critical elliptic PDE. Methods. Appl. Anal. 2014;21(2):221–240. [Google Scholar]
  • 42.Sharp B, Topping P. Decay estimates for Rivière’s equation, with applications to regularity and compactness. Trans. Am. Math. Soc. 2013;365(5):2317–2339. doi: 10.1090/S0002-9947-2012-05671-6. [DOI] [Google Scholar]
  • 43.Sharp B, Zhu M. Regularity at the free boundary for Dirac-harmonic maps from surfaces. Calc. Var. Par. Differ. Equ. 2016;55(2):55:27. [Google Scholar]
  • 44.Struwe M. On the evolution of harmonic mappings of Riemannian surfaces. Comm. Math. Helv. 1985;60:558–581. doi: 10.1007/BF02567432. [DOI] [Google Scholar]
  • 45.Struwe M. The existence of surfaces of constant mean curvature with free boundaries. Acta Math. 1988;160:19–64. doi: 10.1007/BF02392272. [DOI] [Google Scholar]
  • 46.Struwe M. The evolution of harmonic mappings with free boundaries. Manuscripta Math. 1991;70:373–384. doi: 10.1007/BF02568385. [DOI] [Google Scholar]
  • 47.Topping P. Repulsion and quantization in almost-harmonic maps, and asymptotics of the harmonic flow. Ann. Math. (2) 2004;159(2):465–534. doi: 10.4007/annals.2004.159.465. [DOI] [Google Scholar]
  • 48.Wang C. Bubbling phenomena of certain Palais-Smale sequences from surfaces to general targets. Houston J. Math. 1996;V22:N3. [Google Scholar]
  • 49.Wang W, Wei D, Zhang Z. Energy identity for approximate harmonic maps from surface to general targets. J. Funct. Anal. 2017;272(2):776–803. doi: 10.1016/j.jfa.2016.09.018. [DOI] [Google Scholar]
  • 50.Wehrheim, K.: Uhlenbeck Compactness. EMS Series of lectures in mathematics, European Mathematics Society (EMS), Zurich (2004)
  • 51.Wehrheim K. Energy quantization and mean value inequalities for nonlinear boundary value problems. J. Eur. Math. Soc. (JEMS) 2005;7(3):305–318. doi: 10.4171/JEMS/30. [DOI] [Google Scholar]
  • 52.Wolfson J. Gromov’s compactness of pseudo-holomorphic curves and symplectic geometry. J. Differ. Geom. 1988;28(3):383–405. doi: 10.4310/jdg/1214442470. [DOI] [Google Scholar]
  • 53.Ye R. Gromov’s compactness theorem for pseudo-holomorphic curves. Trans. Am. Math. Soc. 1994;342(2):671–694. [Google Scholar]
  • 54.Zhu M. Regularity for harmonic maps into certain Pseudo-Riemannian manifolds. J. Math. Pures Appl. 2013;99(1):106–123. doi: 10.1016/j.matpur.2012.06.006. [DOI] [Google Scholar]

Articles from Mathematische Annalen are provided here courtesy of Springer

RESOURCES