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

Dirac-harmonic maps with the trivial index

Jürgen Jost Max Planck Institute for Mathematics in the Sciences, Inselstrasse 22, 04103 Leipzig, Germany jost@mis.mpg.de Linlin Sun School of Mathematics and Statistics, Wuhan University, Wuhan 430072, China sunll@whu.edu.cn  and  Jingyong Zhu College of Mathematics, Sichuan University, Chengdu 610065, China jzhu@scu.edu.cn
Abstract.

For a homotopy class [u][u] of maps between a closed Riemannian manifold MM and a general manifold NN, we want to find a Dirac-harmonic map with the map component in the given homotopy class. Most known results require the index to be nontrivial. When the index is trivial, the few known results are all constructive and produce uncoupled solutions. In this paper, we define a new quantity. As a byproduct of proving the homotopy invariance of this new quantity, we find a new simple proof for the fact that all Dirac-harmonic spheres in surfaces are uncoupled. More importantly, by using the homotopy invariance of this new quantity, we prove the existence of Dirac-harmonic maps from manifolds in the trivial index case. In particular, when the domain is a closed Riemann surface, we prove the short-time existence of the α\alpha-Dirac-harmonic map flow in the trivial index case. Together with the density of the minimal kernel, we get an existence result for Dirac-harmonic maps from closed Riemann surfaces to Kähler manifolds, which extends the previous result of the first and third authors. This establishes a general existence theory for Dirac-harmonic maps in the context of trivial index.

Key words and phrases:
Dirac-harmonic map; α\alpha-Dirac-harmonic map flow; minimal kernel; existence; Kähler manifolds.
2010 Mathematics Subject Classification:
53C43; 58E20
The third author wants to thank the support by the National Natural Science Foundation of China (Grant No. 12201440) and the Fundamental Research Funds for the Central Universities.

1. Introduction

Motivated by the supersymmetric nonlinear sigma model from quantum field theory, see [6], Dirac-harmonic maps from Riemann surfaces (with a fixed spin structure) into Riemannian manifolds were introduced in [2]. They generalize harmonic maps and harmonic spinors. From the variational point of view, they are critical points of a conformally invariant action functional. The Euler-Lagrange equation then is an elliptic system coupling a harmonic map type equation with a Dirac type equation.

The existence of Dirac-harmonic maps from closed spin manifolds is a very difficult problem. So far, there are only a few results in this direction. Most solutions found so far are uncoupled in the sense that the map part is harmonic. The existence result of [1] for uncoupled solutions depends on the index I(M,u)I(M,u) being non-zero (see Definition 2.2). But when the domain and target are both closed Riemann surfaces, the index I(M,u)I(M,u) always vanishes. In this case, an existence result about uncoupled Dirac-harmonic maps was proved in [3] by the Riemann-Roch formula. Later, this result was generalized to Kähler manifolds in [16]. A general existence result for Dirac-harmonic maps from closed Riemann surfaces to compact manifolds was first established in [9]. This implies the existence of Dirac-harmonic maps when the index I(M,u)I(M,u) is nontrivial.

This then naturally raises the question of the existence of Dirac-harmonic maps in a given homotopy class [u][u] between manifolds M,NM,N with trivial index I(M,u)I(M,u). Of course, we should first identify conditions under which this index vanishes. When the domain MM is a closed Riemann surface with positive genus and the target NN is an odd-dimensional oriented manifold, there is always a spin structure on MM such that the index I(M,u)I(M,u) is nontrivial. When, in contrast, MM is a closed Riemann surface and NN is an even-dimensional spin manifold, the index I(M,u)I(M,u) is always zero. Therefore, here we consider the case where the target manifold NN is a Kähler spin manifold. In this case, it is necessary to use a new quantity that can replace the index I(M,u)I(M,u). For this purpose, we introduce a candidate that uses the complex structure of the target manifold. More precisely, we first decompose the twisted Dirac operator as

u=1,0u+0,1u\not{D}^{u}=\not{D}_{1,0}^{u}+\not{D}_{0,1}^{u}

according to the decomposition (uTN)=uT1,0NuT1,0N(u^{*}TN)^{\mathbb{C}}=u^{*}T_{1,0}N\oplus u^{*}T_{1,0}N. Then we just consider the kernel of one of the two operators, such as 1,0\not{D}_{1,0}. We define

(M,uT1,0N):=[12dimker1,0u]2,\mathcal{I}(M,u^{*}T_{1,0}N):=\left[\frac{1}{2}{\rm dim}_{\mathbb{C}}{\rm ker}\not{D}_{1,0}^{u}\right]_{\mathbb{Z}_{2}},

for an even dimensional spin manifold MM whenever the complex dimension of the kernel of 1,0u\not{D}_{1,0}^{u} is even.

In order to be useful for our purposes, this should be homotopy invariant. Let us first look at an example. Suppose M=P1M=\mathbb{C}{\rm P}^{1} and NN is a compact surface. Consider any map u:MNu:M\to N, the spinor bundle ΣP1\Sigma\mathbb{C}{\rm P}^{1} and the twisted bundle ΣP1uT1,0N\Sigma\mathbb{C}{\rm P}^{1}\otimes u^{*}T_{1,0}N. Let gNg_{N} be the genus of NN and c1(uT1,0N)=aγ,a=2deg(u)(1gN)c_{1}\left(u^{*}T_{1,0}N\right)=a\gamma,a=2\deg(u)(1-g_{N}), where γ\gamma is the tautological bundle of P1\mathbb{C}{\rm P}^{1}. The unique spin structure of P1\mathbb{C}{\rm P}^{1} is determined by γ\gamma since Λ1,0P1=γ2\Lambda^{1,0}\mathbb{C}{\rm P}^{1}=\gamma^{2}. Then as a holomorphic bundle, we have

ΣP1uT1,0N=(γΛ0,1P1γ)γa=γa+1Λ0,1P1γa+1.\displaystyle\Sigma\mathbb{C}{\rm P}^{1}\otimes u^{*}T_{1,0}N=\left(\gamma\oplus\Lambda^{0,1}\mathbb{C}{\rm P}^{1}\otimes\gamma\right)\otimes\gamma^{a}=\gamma^{a+1}\oplus\Lambda^{0,1}\mathbb{C}{\rm P}^{1}\otimes\gamma^{a+1}.

Since

dimH0(P1,γm)={0,m>0,1m,m0,\displaystyle\dim_{\mathbb{C}}H^{0}\left(\mathbb{C}{\rm P}^{1},\gamma^{m}\right)=\begin{cases}0,&m>0,\\ 1-m,&m\leq 0,\end{cases}

we conclude that

dimker1,0u=\displaystyle\dim_{\mathbb{C}}\ker\not{D}^{u}_{1,0}= dimH0(P1,γa+1)+dimH1(P1,γa+1)\displaystyle\dim_{\mathbb{C}}H^{0}\left(\mathbb{C}{\rm P}^{1},\gamma^{a+1}\right)+\dim_{\mathbb{C}}H^{1}\left(\mathbb{C}{\rm P}^{1},\gamma^{a+1}\right)
=\displaystyle= dimH0(P1,γa+1)+dimH0(P1,γ1a)\displaystyle\dim_{\mathbb{C}}H^{0}\left(\mathbb{C}{\rm P}^{1},\gamma^{a+1}\right)+\dim_{\mathbb{C}}H^{0}\left(\mathbb{C}{\rm P}^{1},\gamma^{1-a}\right)
=\displaystyle= |a|=2|deg(u)(gN1)|.\displaystyle\left|a\right|=2|{\rm deg}(u)(g_{N}-1)|.

Therefore, dimker1,0u\dim_{\mathbb{C}}\ker\not{D}^{u}_{1,0} is invariant in the homotopy class [u][u]. This implies the homotopy invariance of (P1,uT1,0N)\mathcal{I}(\mathbb{C}{\rm P}^{1},u^{*}T_{1,0}N), which is equal to [|deg(u)(gN1)|]2\left[|{\rm deg}(u)(g_{N}-1)|\right]_{\mathbb{Z}_{2}}. Moreover, the dimension of the kernel of the Dirac operator is a constant in a given homotopy class. Then the following well-known fact follows from the first variational formula.

Proposition 1.1 ([17]).

There is no coupled Dirac-harmonic map from the 22-sphere into a compact Riemann surface.

In general, we can give two different sufficient conditions to guarantee the homotopy invariance of \mathcal{I}.

Theorem 1.2.

Suppose that MM is an even dimensional spin Riemannian manifold and (N,i)(N,i) is a Kähler manifold. If one of the following holds:

  • (1)

    the complex spinor bundle (ΣM,i1)(\Sigma M,i_{1}) over MM admits a commuting real structure jj, i.e. a real structure (j2=idΣM,ji1=i1jj^{2}=id_{\Sigma M},\ ji_{1}=-i_{1}j) commutes with Clifford multiplication and NN is hyperKähler;

  • (2)

    the complex spinor bundle ΣM\Sigma M over MM admits a commuting quaternionic structure j1j_{1} and there exists a parallel real structure j2j_{2} on T1,0NT_{1,0}N, i.e.

    j22=idT1,0N,j2i=ij2andj2=0.j_{2}^{2}=id_{T_{1,0}N},\ j_{2}i=-ij_{2}\ \text{and}\ \nabla{j_{2}}=0.

Then all the eigenspaces of 1,0u\not{D}^{u}_{1,0} are quaternionic vector spaces for any map u:MNu:M\to N and (M,uT1,0N)\mathcal{I}(M,u^{*}T_{1,0}N) is invariant in the homotopy class [u][u].

Moreover, if (M,uT1,0N)0\mathcal{I}(M,u^{*}T_{1,0}N)\neq 0, then there is a real vector space of real dimension 4\geq 4 such that all (u~,ψ)(\tilde{u},\psi) are uncoupled α\alpha-Dirac-harmonic maps as long as there is an α\alpha-harmonic map u~[u]\tilde{u}\in[u] for α1\alpha\geq 1.

Here α\alpha-Dirac-harmonic maps are the critical points of the following functional

Lα(u,ψ)=12M(1+|du|2)α+12Mψ,uψΣMuTN,α1.L^{\alpha}(u,\psi)=\frac{1}{2}\int_{M}(1+|du|^{2})^{\alpha}+\frac{1}{2}\int_{M}\langle\psi,\not{D}^{u}\psi\rangle_{\Sigma M\otimes u^{*}TN},\ \ \forall\alpha\geq 1.

They are generalizations of Dirac-harmonic maps (i.e. the case of α=1\alpha=1). As generalizations of harmonic maps, α\alpha-harmonic maps are the critical points of the following functional

Eα(u)=12M(1+|du|2)α,α1,E_{\alpha}(u)=\frac{1}{2}\int_{M}(1+|du|^{2})^{\alpha},\ \ \forall\alpha\geq 1,

which was introduced by Sacks and Uhlenbeck in [14].

By the statement in [5, Theorem 2.2.2], such a commuting real structure in Theorem 1.2 always exists on MM if m=0,6,7(mod 8)m=0,6,7({\rm mod}\ 8). In particular, when m=0,6(mod 8)m=0,6({\rm mod}\ 8), we can get the existence of uncoupled Dirac-harmonic maps.

Corollary 1.3.

Let mm be the dimension of MM. Suppose one of the following holds:

  • (a)

    m=0,6(mod 8)m=0,6({\rm mod}\ 8), NN is a hyperKähler manifold and a homotopy class [u][u] satisfies (M,uT1,0N)0\mathcal{I}(M,u^{*}T_{1,0}N)\neq 0;

  • (b)

    m=2,4(mod 8)m=2,4({\rm mod}\ 8), NN is a Kähler manifold with a parallel real structure j2j_{2} defined in Theorem 1.2 and a homotopy class [u][u] satisfies (M,uT1,0N)0\mathcal{I}(M,u^{*}T_{1,0}N)\neq 0.

Then there is a real vector space of real dimension 4\geq 4 such that all (u~,ψ)(\tilde{u},\psi) are uncoupled α\alpha-Dirac-harmonic maps as long as there is an α\alpha-harmonic map u~[u]\tilde{u}\in[u] for α1\alpha\geq 1.

Remark 1.4.

Note that the case m=6(mod 8)m=6({\rm mod}\ 8) is not included in [1] due to the definition of index α(M,u)\alpha(M,u). When m=0(mod 4)m=0({\rm mod}\ 4), the triviality of the index α(M,u)\alpha(M,u) implies that of the index ind(+){\rm ind}(\not{D}^{+}), where +\not{D}^{+} comes from the decomposition of the Dirac operator according to that of the spinor bundle (see the Section 2). When m=2km=2k and kk is odd, the triviality of ind(+){\rm ind}(\not{D}^{+}) does not imply that of ind(1,0+){\rm ind}(\not{D}^{+}_{1,0}). And the second author used the non-trivial index ind(1,0+){\rm ind}(\not{D}^{+}_{1,0}) to get an existence result in [16]. Our corollary is still valid even if ind(1,0+)=0{\rm ind}(\not{D}^{+}_{1,0})=0. For example, our result applies to the case when the dimensions of the kernels in those four subspaces in the decomposition (2.8) are all equal to one, which is never considered in literatures.

When MM is a closed Riemann surface, we can prove the short-time existence of α\alpha-Dirac-harmonic map flow into a Kähler manifold, which generalizes the result in [10].

The rest of the paper is organized as follows: In Section 2, we recall some facts about Dirac-harmonic maps as well as the Dirac operator. In Section 3, we prove Theorem 1.2 and end this section by showing the density of minimal kernel. In Section 4, under the minimality assumption on the kernel of 1,0u0\not{D}_{1,0}^{u_{0}}, we prove the short time existence of α\alpha-Dirac-harmonic map flow (Theorem 4.2) and the existence of Dirac-harmonic maps (Theorem 4.7). In the Appendix, we solve the constraint equation and prove Lipschitz continuity of the solution with respect to the map.

2. Preliminaries

Let (M,g)(M,g) be a compact Riemann surface with a fixed spin structure χ\chi. On the complex spinor bundle ΣM\Sigma M, we denote the Hermitian inner product by ,ΣM\langle\cdot,\cdot\rangle_{\Sigma M}. For any XΓ(TM)X\in\Gamma(TM) and ξΓ(ΣM)\xi\in\Gamma(\Sigma M), the Clifford multiplication satisfies the following skew-adjointness:

Xξ,ηΣM=ξ,XηΣM.\langle X\cdot\xi,\eta\rangle_{\Sigma M}=-\langle\xi,X\cdot\eta\rangle_{\Sigma M}.

Let \nabla be the Levi-Civita connection on (M,g)(M,g). There is a unique connection (also denoted by \nabla) on ΣM\Sigma M compatible with ,ΣM\langle\cdot,\cdot\rangle_{\Sigma M}. Choosing a local orthonormal basis {eβ}β=1,2\{e_{\beta}\}_{\beta=1,2} on MM, the usual Dirac operator is defined as ∂̸:=eββ\not{\partial}:=e_{\beta}\cdot\nabla_{\beta}, where β=1,2\beta=1,2. Here and in the sequel, we use the Einstein summation convention. One can find more about spin geometry in [11].

Let uu be a smooth map from MM to another compact Riemannian manifold (N,h)(N,h) of dimension n2n\geq 2. Let uTNu^{*}TN be the pull-back bundle of TNTN by uu and consider the twisted bundle ΣMuTN\Sigma M\otimes_{\mathbb{R}}u^{*}TN. On this bundle there is a metric ,ΣMuTN\langle\cdot,\cdot\rangle_{\Sigma M\otimes u^{*}TN} induced from the metric on ΣM\Sigma M and uTNu^{*}TN. Also, we have a connection ~\tilde{\nabla} on this twisted bundle naturally induced from those on ΣM\Sigma M and uTNu^{*}TN. In local coordinates {yi}i=1,,n\{y^{i}\}_{i=1,\dots,n}, the section ψ\psi of ΣMuTN\Sigma M\otimes_{\mathbb{R}}u^{*}TN is written as

ψ=ψiyi(u),\psi=\psi_{i}\otimes\partial_{y^{i}}(u),

where each ψi\psi^{i} is a usual spinor on MM. We also have the following local expression of ~\tilde{\nabla}

~ψ=(ψi+Γjki(u)ujψk)yi(u),\tilde{\nabla}\psi=\left(\nabla\psi^{i}+\Gamma_{jk}^{i}(u)\nabla u^{j}\psi^{k}\right)\otimes\partial_{y^{i}}(u),

where Γjki\Gamma^{i}_{jk} are the Christoffel symbols of the Levi-Civita connection of NN. The Dirac operator along the map uu is defined as

(2.1) :=eα~eαψ=(∂̸ψi+Γjki(u)eαuj(eαψk))yi(u),\not{D}:=e_{\alpha}\cdot\tilde{\nabla}_{e_{\alpha}}\psi=\left(\not{\partial}\psi^{i}+\Gamma_{jk}^{i}(u)\nabla_{e_{\alpha}}u^{j}(e_{\alpha}\cdot\psi^{k})\right)\otimes\partial_{y^{i}}(u),

which is self-adjoint [7]. Sometimes, we use u\not{D}^{u} to distinguish the Dirac operators defined on different maps. In [2], the authors introduced the functional

L(u,ψ):=12M(|du|2+ψ,ψΣMuTN)=12M(hij(u)gαβuixαujxβ+hij(u)ψi,ψjΣM).\begin{split}L(u,\psi)&:=\frac{1}{2}\int_{M}\left(|du|^{2}+\langle\psi,\not{D}\psi\rangle_{\Sigma M\otimes u^{*}TN}\right)\\ &=\frac{1}{2}\int_{M}\left(h_{ij}(u)g^{\alpha\beta}\frac{\partial u^{i}}{\partial x^{\alpha}}\frac{\partial u^{j}}{\partial x^{\beta}}+h_{ij}(u)\langle\psi^{i},\not{D}\psi^{j}\rangle_{\Sigma M}\right).\end{split}

They computed the Euler-Lagrange equations of LL:

(2.2) τm(u)12Rlijmψi,ulψjΣM=0,\displaystyle\tau^{m}(u)-\frac{1}{2}R^{m}_{lij}\langle\psi^{i},\nabla u^{l}\cdot\psi^{j}\rangle_{\Sigma M}=0,
(2.3) ψi:=∂̸ψi+Γjki(u)eαuj(eαψk)=0,\displaystyle\not{D}\psi^{i}:=\not{\partial}\psi^{i}+\Gamma_{jk}^{i}(u)\nabla_{e_{\alpha}}u^{j}(e_{\alpha}\cdot\psi^{k})=0,

where τm(u)\tau^{m}(u) is the mm-th component of the tension field [7] of the map uu with respect to the coordinates on NN, ulψj\nabla u^{l}\cdot\psi^{j} denotes the Clifford multiplication of the vector field ul\nabla u^{l} with the spinor ψj\psi^{j}, and RlijmR^{m}_{lij} stands for the components of the Riemann curvature tensor of the target manifold NN. Denote

(u,ψ):=12Rlijmψi,ulψjΣMym.\mathcal{R}(u,\psi):=\frac{1}{2}R^{m}_{lij}\langle\psi^{i},\nabla u^{l}\cdot\psi^{j}\rangle_{\Sigma M}\partial_{y^{m}}.

We can write (2.2) and (2.3) in the following global form:

(2.4) τ(u)=(u,ψ),\displaystyle\tau(u)=\mathcal{R}(u,\psi),
(2.5) ψ=0,\displaystyle\not{D}\psi=0,

and call the solutions (u,ψ)(u,\psi) Dirac-harmonic maps from MM to NN.

With the aim to get a general existence scheme for Dirac-harmonic maps, the following heat flow for Dirac-harmonic maps was introduced in [4]:

(2.6) tu=τ(u)(u,ψ),\displaystyle\partial_{t}u=\tau(u)-\mathcal{R}(u,\psi),\ on(0,T)×M,\displaystyle\text{on}\ (0,T)\times M,
(2.7) ψ=0,\displaystyle\not{D}\psi=0,\ on[0,T]×M.\displaystyle\text{on}\ [0,T]\times M.

When MM has boundary, the short time existence and uniqueness of (2.6)-(2.7) was shown in [4].

For a closed manifold MM, the situation is more complicated because one cannot uniquely solve the second equation (2.7) and the kernel of the Dirac operator may jump along the flow. As we stated in the introduction, the short-time existence is only known in the minimal kernel case, i.e. dimker=1{\rm dim}_{\mathbb{H}}{\rm ker}\not{D}=1. However, when the target manifold NN is an even-dimensional spin manifold, the index α(M,u)\alpha(M,u) always vanishes for any map uu between MM and NN. In order to deal with this case, we utilize the complex structure on NN. We denote the complexification of uTNu^{*}TN by (uTN)(u^{*}TN)^{\mathbb{C}}. Then we have

ΣMuTN=(ΣM)uTN=ΣM(uTN).\Sigma M\otimes_{\mathbb{R}}u^{*}TN=(\Sigma M\otimes_{\mathbb{C}}{\mathbb{C}})\otimes_{\mathbb{R}}u^{*}TN=\Sigma M\otimes_{\mathbb{C}}(u^{*}TN)^{\mathbb{C}}.

The pull-back metric ugu^{*}g on uTNu^{*}TN could be naturally extended to a Hermitian product on (uTN)(u^{*}TN)^{\mathbb{C}}. Moreover, there is a natural Hermitian product on ΣM(uTN)\Sigma M\otimes(u^{*}TN)^{\mathbb{C}} induced from those on ΣM\Sigma M and (uTN)(u^{*}TN)^{\mathbb{C}}, which is denoted by ,ΣM(uTN)\langle\cdot,\cdot\rangle_{\Sigma M\otimes(u^{*}TN)^{\mathbb{C}}}.

For a general even-dimensional spin Riemannian manifold MM, there is a parallel 2{\mathbb{Z}}_{2}-grading GEnd(ΣM)G\in{\rm End}(\Sigma M) given by G(ψ)=(1)m/2e1e2emψG(\psi)=\left(\sqrt{-1}\right)^{m/2}e_{1}\cdot e_{2}\cdot\dotsc\cdot e_{m}\cdot\psi for a positively oriented orthonormal local frame {e1,e2,,em}\{e_{1},e_{2},\dotsc,e_{m}\} where m=dimMm=\dim M. Thus the spinor bundle can be decomposed as

ΣM=Σ+MΣM,\Sigma M=\Sigma^{+}M\oplus\Sigma^{-}M,

where Σ±M\Sigma^{\pm}M are the eigenspaces of GG associated to the ±1\pm 1, respectively. As GG is Hermitian and parallel, the decomposition is orthogonal in the complex sense and parallel. Consequently, we have

(2.8) ΣM(uTN)=(Σ+MuT1,0N)(ΣMuT1,0N)(Σ+MuT0,1N)(ΣMuT0,1N),\begin{split}\Sigma M\otimes_{\mathbb{C}}(u^{*}TN)^{\mathbb{C}}=&(\Sigma^{+}M\otimes_{\mathbb{C}}u^{*}T_{1,0}N)\oplus(\Sigma^{-}M\otimes_{\mathbb{C}}u^{*}T_{1,0}N)\\ &\oplus(\Sigma^{+}M\otimes_{\mathbb{C}}u^{*}T_{0,1}N)\oplus(\Sigma^{-}M\otimes_{\mathbb{C}}u^{*}T_{0,1}N),\end{split}

where we used (uTN)=uT1,0NuT1,0N(u^{*}TN)^{\mathbb{C}}=u^{*}T_{1,0}N\oplus u^{*}T_{1,0}N. Moreover, we also have the following decomposition for the Dirac operator:

=1,0+0,11,0=1,0++1,00,1=0,1++0,1,\begin{split}\not{D}=&\not{D}_{1,0}+\not{D}_{0,1}\\ \not{D}_{1,0}=&\not{D}_{1,0}^{+}+\not{D}_{1,0}^{-}\\ \not{D}_{0,1}=&\not{D}_{0,1}^{+}+\not{D}_{0,1}^{-},\end{split}

where 1,0±\not{D}_{1,0}^{\pm} (resp. 0,1±\not{D}_{0,1}^{\pm}) is obtained by restricting \not{D} on Σ±MuT1,0N\Sigma^{\pm}M\otimes_{\mathbb{C}}u^{*}T_{1,0}N (resp. Σ±MuT0,1N\Sigma^{\pm}M\otimes_{\mathbb{C}}u^{*}T_{0,1}N).

By [13], we can isometrically embed NN into q\mathbb{R}^{q}. Then (2.4)-(2.5) is equivalent to the following system:

{Δgu=II(du,du)+Re(P(𝒮(du(eβ),eβψ);ψ)),∂̸ψ=𝒮(du(eβ),eβψ),\begin{cases}\Delta_{g}{u}=II(du,du)+Re(P(\mathcal{S}(du(e_{\beta}),e_{\beta}\cdot\psi);\psi)),\\ \not{\partial}\psi=\mathcal{S}(du(e_{\beta}),e_{\beta}\cdot\psi),\end{cases}

where IIII is the second fundamental form of NN in q\mathbb{R}^{q}, and

𝒮(du(eβ),eβψ):=(uAψB)II(zA,zB),\mathcal{S}(du(e_{\beta}),e_{\beta}\cdot\psi):=(\nabla{u^{A}}\cdot\psi^{B})\otimes II(\partial_{z^{A}},\partial_{z^{B}}),
Re(P(𝒮(du(eβ),eβψ);ψ)):=P(S(zC,zB);zA)Re(ψA,duCψB).Re(P(\mathcal{S}(du(e_{\beta}),e_{\beta}\cdot\psi);\psi)):=P(S(\partial_{z^{C}},\partial_{z^{B}});\partial_{z^{A}})Re(\langle\psi^{A},du^{C}\cdot\psi^{B}\rangle).

Here P(ξ;)P(\xi;\cdot) denotes the shape operator, defined by P(ξ;X),Y=A(X,Y),ξ\langle P(\xi;X),Y\rangle=\langle A(X,Y),\xi\rangle for X,YΓ(TN)X,Y\in\Gamma(TN) and Re(z)Re(z) denotes the real part of zz\in\mathbb{C}. Together with the nearest point projection:

π:NδN,\pi:\ N_{\delta}\to N,

where Nδ:={zq|d(z,N)δ}N_{\delta}:=\{z\in\mathbb{R}^{q}|d(z,N)\leq\delta\}, we can rewrite the evolution equation (2.6) as an equation in q\mathbb{R}^{q}.

Lemma 2.1.

[4] A tuple (u,ψ)(u,\psi), where u:[0,T]×MNu:[0,T]\times M\to N and ψΓ(ΣMuTN)\psi\in\Gamma(\Sigma M\otimes u^{*}TN), is a solution of (2.6) if and only if

tuAΔuA=πBCA(u)uB,uCπBA(u)πBDC(u)πEFC(ψD,uEψF)\partial_{t}u^{A}-\Delta u^{A}=-\pi^{A}_{BC}(u)\langle\nabla u^{B},\nabla u^{C}\rangle-\pi^{A}_{B}(u)\pi^{C}_{BD}(u)\pi^{C}_{EF}(\psi^{D},\nabla u^{E}\cdot\psi^{F})

on (0,T)×M(0,T)\times M, for A=1,,qA=1,\dots,q. Here we denote the AA-th component function of u:[0,T]×MNqu:[0,T]\times M\to N\subset\mathbb{R}^{q} by uA:Mu^{A}:M\to\mathbb{R}, write πBA(z)\pi^{A}_{B}(z) for the BB-th partial derivative of the AA-th component function of π:qq\pi:\mathbb{R}^{q}\to\mathbb{R}^{q} and the global sections ψAΓ(ΣM)\psi^{A}\in\Gamma(\Sigma M) are defined by ψ=ψA(Au)\psi=\psi^{A}\otimes(\partial_{A}\circ u), where (A)A=1,,q(\partial_{A})_{A=1,\dots,q} is the standard basis of TqT\mathbb{R}^{q}. Moreover, \nabla and ,\langle\cdot,\cdot\rangle denote the gradient and the Riemannian metric on MM, respectively.

For future reference, we define

(2.9) F1A(u):=πBCA(u)uB,uC,F_{1}^{A}(u):=-\pi^{A}_{BC}(u)\langle\nabla u^{B},\nabla u^{C}\rangle,
(2.10) F2A(u,ψ):=πBA(u)πBDC(u)πEFC(ψD,uEψF).F_{2}^{A}(u,\psi):=-\pi^{A}_{B}(u)\pi^{C}_{BD}(u)\pi^{C}_{EF}(\psi^{D},\nabla u^{E}\cdot\psi^{F}).

Note that for uC1(M,N)u\in C^{1}(M,N) and ψΓ(ΣMuTN)\psi\in\Gamma(\Sigma M\otimes u^{*}TN) we have

II(dup(eα),dup(eα)))=F1A(u)|pA|u(p),II(du_{p}(e_{\alpha}),du_{p}(e_{\alpha})))=-F_{1}^{A}(u)|_{p}\partial_{A}|_{u(p)},
(u,ψ)|p=F2A(u,ψ)|pA|u(p)\mathcal{R}(u,\psi)|_{p}=-F_{2}^{A}(u,\psi)|_{p}\partial_{A}|_{u(p)}

for all pMp\in M, where {eα}\{e_{\alpha}\} is an orthonormal basis of TpMT_{p}M.

Next, let us fix some notations, which will be used in the Section 4 and Appendix. For every T>0T>0, we denote by XTX_{T} the Banach space of bounded maps:

XT:=B([0,T];C1(M,q)),X_{T}:=B([0,T];C^{1}(M,\mathbb{R}^{q})),
uXT:=maxA=1,,qsupt[0,T](uA(t,)C0(M)+uA(t,)C0(M)).\|u\|_{X_{T}}:=\max\limits_{A=1,\dots,q}\sup\limits_{t\in[0,T]}(\|u^{A}(t,\cdot)\|_{C^{0}(M)}+\|\nabla u^{A}(t,\cdot)\|_{C^{0}(M)}).

For any map vXTv\in X_{T}, the closed ball with center vv and radius RR in XTX_{T} is defined by

BRT(v):={uXT|uvR}.B_{R}^{T}(v):=\{u\in X_{T}|\|u-v\|\leq R\}.

We denote by Put,vs=Put,vs(x)P^{u_{t},v_{s}}=P^{u_{t},v_{s}}(x) the parallel transport of NN along the unique shortest geodesic from π(u(x,t))\pi(u(x,t)) to π(v(x,s))\pi(v(x,s)). We also denote by Put,vsP^{u_{t},v_{s}} the inducing mappings

(πut)TN(πvs)TN,(\pi\circ u_{t})^{*}TN\to(\pi\circ v_{s})^{*}TN,
ΣM(πut)TNΣM(πvs)TN\Sigma M\otimes(\pi\circ u_{t})^{*}TN\to\Sigma M\otimes(\pi\circ v_{s})^{*}TN

and

ΓC1(ΣM(πut)TN)ΓC1(ΣM(πvs)TN).\Gamma_{C^{1}}(\Sigma M\otimes(\pi\circ u_{t})^{*}TN)\to\Gamma_{C^{1}}(\Sigma M\otimes(\pi\circ v_{s})^{*}TN).

We also define

Λ(ut)=sup{Λ~|spec(πut){0}(Λ~(ut),Λ~(ut))}\Lambda(u_{t})=\sup\{\tilde{\Lambda}|{\rm spec}(\not{D}^{\pi\circ u_{t}})\setminus\{0\}\subset\mathbb{R}\setminus(-\tilde{\Lambda}(u_{t}),\tilde{\Lambda}(u_{t}))\}

and γt(x):[0,2π]\gamma_{t}(x):[0,2\pi]\to\mathbb{C} as

γt(x):=Λ(ut)2eix.\gamma_{t}(x):=\frac{\Lambda(u_{t})}{2}e^{ix}.

In general, we also denote by γ\gamma the curve γ(x):[0,2π]\gamma(x):[0,2\pi]\to\mathbb{C} as

(2.11) γ(x):=Λ2eix\gamma(x):=\frac{\Lambda}{2}e^{ix}

for some constant Λ\Lambda to be determined. Then the orthogonal projection onto ker(πut){\rm ker}(\not{D}^{\pi\circ u_{t}}), which is the mapping

ΓL2(ΣM(πut)TN)ΓL2(ΣM(πut)TN),\Gamma_{L^{2}}(\Sigma M\otimes(\pi\circ u_{t})^{*}TN)\longrightarrow\Gamma_{L^{2}}(\Sigma M\otimes(\pi\circ u_{t})^{*}TN),

can be written via the resolvent by

s12πiγtR(λ,πut)s𝑑λ,s\mapsto-\frac{1}{2\pi i}\int_{\gamma_{t}}R(\lambda,\not{D}^{\pi\circ u_{t}})sd\lambda,

where R(λ,πut):ΓL2ΓL2R(\lambda,\not{D}^{\pi\circ u_{t}}):\Gamma_{L^{2}}\to\Gamma_{L^{2}} is the resolvent of πut:ΓW1,2ΓL2\not{D}^{\pi\circ u_{t}}:\Gamma_{W^{1,2}}\to\Gamma_{L^{2}}.

In the end of this section, we recall the definition of the index.

Definition 2.2.

Let EME\to M be a Riemannian real vector bundle with metric connection. Then one can associate to the twisted Dirac operator E:C(M,ΣME)C(M,ΣME)\not{D}^{E}:C^{\infty}(M,\Sigma M\otimes E)\to C^{\infty}(M,\Sigma M\otimes E) an index I(M,χ,E)KOm(pt)I(M,\chi,E)\in KO_{m}({\rm pt}), where

KOm(pt){,ifm=0(4),2,ifm=1,2(8),0,otherwise.KO_{m}({\rm pt})\cong\begin{cases}\mathbb{Z},&\text{if}\ m=0\ (4),\\ \mathbb{Z}_{2},&\text{if}\ m=1,2\ (8),\\ 0,&\text{otherwise}.\end{cases}

The index I(M,χ,E)I(M,\chi,E) can be determined out of ker(E){\rm ker}(\not{D}^{E}) using the following formula:

I(M,χ,E)={{ch(E)A^(M)}[M],ifm=0(8),[dim(ker(E)]2,ifm=1(8),[dim(ker(E)2]2,ifm=2(8),12{ch(E)A^(M)}[M],ifm=4(8).I(M,\chi,E)=\begin{cases}\{{\rm ch}(E)\cdot\hat{A}(M)\}[M],&\text{if}\ m=0\ (8),\\ [{\rm dim}_{\mathbb{C}}({\rm ker}(\not{D}^{E})]_{\mathbb{Z}_{2}},&\text{if}\ m=1\ (8),\\ \left[\frac{{\rm dim}_{\mathbb{C}}({\rm ker}(\not{D}^{E})}{2}\right]_{\mathbb{Z}_{2}},&\text{if}\ m=2\ (8),\\ \frac{1}{2}\{{\rm ch}(E)\cdot\hat{A}(M)\}[M],&\text{if}\ m=4\ (8).\end{cases}

In particular, when E=uTNE=u^{*}TN and χ\chi is fixed, we denote I(M,χ,E)I(M,\chi,E) by I(M,u)I(M,u).

3. Quaternionic structure on the twisted bundle

In this section, we will prove Theorem 1.2 by constructing a commuting quaternionic structure on the twisted bundle ΣMuT1,0N\Sigma M\otimes_{\mathbb{C}}u^{*}T_{1,0}N and show the density of the minimal kernel.

Proof of Theorem 1.2.

Let ρ:lmEnd(Σm)\rho:\mathbb{C}{\rm l}_{m}\to{\rm End}_{\mathbb{C}}(\Sigma_{m}) be an irreducible complex representation of the complex Clifford algebra lm\mathbb{C}{\rm l}_{m}. Suppose the condition (2) holds. Then every fibre of the complex spinor bundle ΣM=Spin(M)×ρΣm\Sigma M={\rm Spin}(M)\times_{\rho}\Sigma_{m} turns into a quaternionic vector space by defining

[p,v]h:=[p,vh][p,v]h:=[p,vh]

for all pSpin(M)p\in{\rm Spin}(M), vΣmv\in\Sigma_{m} and hh\in\mathbb{H}.

Since the tensor product of the twisted bundle ΣMuT1,0N\Sigma M\otimes_{\mathbb{C}}u^{*}T_{1,0}N is taken over \mathbb{C}, there is a natural complex structure II on ΣMuT1,0N\Sigma M\otimes_{\mathbb{C}}u^{*}T_{1,0}N defined by

I(ψkθk):=i(ψk)θk=ψki(θk).I(\psi^{k}\otimes_{\mathbb{C}}\theta_{k}):=i(\psi^{k})\otimes_{\mathbb{C}}\theta_{k}=\psi^{k}\otimes_{\mathbb{C}}i(\theta_{k}).

However, the quaternionic structure on ΣM\Sigma M cannot directly extend to the twisted bundle. To overcome this problem, we need an extra structure on uT1,0Nu^{*}T_{1,0}N. By our assumption, we define J:ΣMuT1,0NΣMuT1,0NJ:\Sigma M\otimes_{\mathbb{C}}u^{*}T_{1,0}N\to\Sigma M\otimes_{\mathbb{C}}u^{*}T_{1,0}N by

J(ψiθi):=j1(ψk)j2(θk).J(\psi^{i}\otimes_{\mathbb{C}}\theta_{i}):=j_{1}(\psi^{k})\otimes_{\mathbb{C}}j_{2}(\theta_{k}).

Since both j1j_{1} and j2j_{2} anti-commute with the complex structure ii, JJ is well-defined on ΣMuT1,0N\Sigma M\otimes_{\mathbb{C}}u^{*}T_{1,0}N. By the definitions of j1j_{1} and j2j_{2}, JJ anti-commutes with II and J2=1J^{2}=-1. Moreover, JJ also commutes with the Clifford multiplication and hence the Dirac operator 1,0u\not{D}^{u}_{1,0} (see also [5]), i.e.

1,0uJ=J1,0u.\not{D}^{u}_{1,0}\circ J=J\circ\not{D}^{u}_{1,0}.

Therefore, we conclude that all the eigenspaces of 1,0u\not{D}^{u}_{1,0} are quaternionic vector space with two complex structures II and JJ, which are anti-commuting with each other.

If condition (1) holds, i.e. j1j_{1} is a commuting real structure and j2j_{2} is a quaternionic structure, then it follows from the argument above that the conclusion is also true.

When m3(mod 4)m\neq 3({\rm mod}\ 4), the eigenvalues are symmetric with respect to the origin (see Remark 2.2.3 in [5]). For any two maps in [u][u], there is a piecewise smooth curve connecting them with parameter t[0,1]t\in[0,1]. Along this curve, the eigenvalues of the Dirac operator are continuous functions of tt. Suppose there is an eigenvalue λ1(t)\lambda_{1}(t) that decreases to zero as tTt\to T. By the symmetry of the eigenvalues, there is another eigenvalue λ1(t)\lambda_{-1}(t) such that λ1(t)=λ1(t)\lambda_{-1}(t)=-\lambda_{1}(t). Therefore, the difference of the quaternionic dimension of the kernel of the corresponding Dirac operator is always an even number.

When mm is even, we have a parallel 2\mathbb{Z}_{2}-grading GG described in the previous section. From the orthogonality of the splitting, we have

1,0uψ+,ψ+=1,0uψ,ψ=0\langle\not{D}^{u}_{1,0}\psi^{+},\psi^{+}\rangle=\langle\not{D}^{u}_{1,0}\psi^{-},\psi^{-}\rangle=0

for all ψ±C(M,Σ±MuT1,0N)\psi^{\pm}\in C^{\infty}(M,\Sigma^{\pm}M\otimes u^{*}T_{1,0}N). Thus,

(3.1) (1,0uψ+,ψ+)L2=(1,0uψ,ψ)L2=0.(\not{D}^{u}_{1,0}\psi^{+},\psi^{+})_{L^{2}}=(\not{D}^{u}_{1,0}\psi^{-},\psi^{-})_{L^{2}}=0.

Now, for any smooth variation (us)s(ϵ,ϵ)(u_{s})_{s\in(-\epsilon,\epsilon)} of the α\alpha-harmonic map u~[u]\tilde{u}\in[u] with us|s=0=u~u_{s}|_{s=0}=\tilde{u}, we split the bundle ΣMusT1,0N\Sigma M\otimes u_{s}^{*}T_{1,0}N into

ΣMusT1,0N=(Σ+MusT1,0N)(ΣMusT1,0N),\Sigma M\otimes u_{s}^{*}T_{1,0}N=(\Sigma^{+}M\otimes u_{s}^{*}T_{1,0}N)\oplus(\Sigma^{-}M\otimes u_{s}^{*}T_{1,0}N),

which is orthogonal in the complex sense and parallel. Since (M,uT1,0N)0\mathcal{I}(M,u^{*}T_{1,0}N)\neq 0, there exists Ψker1,0u~\Psi\in{\rm ker}{\not{D}^{\tilde{u}}_{1,0}} which can be written as Ψ=Ψ++Ψ\Psi=\Psi^{+}+\Psi^{-}, where Ψ±Γ(Σ±Mu~T1,0N)\Psi^{\pm}\in\Gamma(\Sigma^{\pm}M\otimes\tilde{u}^{*}T_{1,0}N). Then there always exists a variation Ψs\Psi_{s} of Ψ\Psi such that Ψs±Γ(Σ±MusT1,0N)\Psi_{s}^{\pm}\in\Gamma(\Sigma^{\pm}M\otimes u_{s}^{*}T_{1,0}N) are smooth variations of Ψ±\Psi^{\pm}, respectively. Moreover, (3.1) implies that

ddt|s=0(usΨs±,Ψs±)L2=0.\frac{d}{dt}\bigg{|}_{s=0}(\not{D}^{u_{s}}\Psi_{s}^{\pm},\Psi_{s}^{\pm})_{L^{2}}=0.

Therefore, for the α\alpha-harmonic map u~\tilde{u}, we have

ddt|s=0Lα(us,Ψs±)=ddt|s=0M(1+|dus|2)α=0.\frac{d}{dt}\bigg{|}_{s=0}L^{\alpha}(u_{s},\Psi_{s}^{\pm})=\frac{d}{dt}\bigg{|}_{s=0}\int_{M}(1+|du_{s}|^{2})^{\alpha}=0.

Hence, we get α\alpha-Dirac-harmonic maps (u~,Ψ±)(\tilde{u},\Psi^{\pm}).

In the rest of this section, we will show the density of the minimal kernel. By the definition of (M,uT1,0N)\mathcal{I}(M,u^{*}T_{1,0}N), we have

dimker(1,0u){0,ifinduT1,0N(M)=0;1,ifinduT1,0N(M)0.{\rm dim}_{\mathbb{H}}{\rm ker}(\not{D}_{1,0}^{u})\geq\begin{cases}0,&\text{if}\ {\rm ind}_{u^{*}T_{1,0}N}(M)=0;\\ 1,&\text{if}\ {\rm ind}_{u^{*}T_{1,0}N}(M)\neq 0.\end{cases}

If equality holds above, then we say that 1,0u\not{D}_{1,0}^{u} has minimal kernel. Using the analyticity of NN, one can prove the following density result for the minimal kernel.

Lemma 3.1.

If 1,0u\not{D}_{1,0}^{u} has minimal kernel, then 1,0u\not{D}_{1,0}^{u^{\prime}} also has minimal kernel for a generic map u[u]u^{\prime}\in[u].

Proof.

Let u[u]u^{\prime}\in[u] and HH be any homotopy between uu^{\prime} and uu. More precisely, H:[0,1]C(M,N)H:[0,1]\to C^{\infty}(M,N) with H(0)=uH(0)=u and H(1)=uH(1)=u^{\prime}. We can cover the image of HH by finitely many balls {Vl}l=1L\{V_{l}\}_{l=1}^{L} of radius less than 12inj(N)\frac{1}{2}{\rm inj}(N) such that

VlVl+1fori=1,,L1V_{l}\cap V_{l+1}\neq\emptyset\ \text{for}\ i=1,\cdots,L-1

and

uV1,uVL.u\in V_{1},\ \ \ u^{\prime}\in V_{L}.

We choose u1V1V2u_{1}\in V_{1}\cap V_{2} arbitrarily and define a homotopy Ht1H^{1}_{t} by

Ht1(x):=expu(x)(texpu(x)1u1(x)),H^{1}_{t}(x):=\exp_{u(x)}(t\exp^{-1}_{u(x)}u_{1}(x)),

where xMx\in M and exp\exp is the exponential map on NN. We denote by

Pt=Pt(x):T1,0N|u(x)T1,0N|Ht1(x)P_{t}=P_{t}(x):T_{1,0}N|_{u(x)}\to T_{1,0}N|_{H^{1}_{t}(x)}

the parallel transport along the unique shortest geodesic of NN connecting u(x)u(x) and Ht1(x)H^{1}_{t}(x) and consider

t:=Pt11,0Ht1Pt.\not{D}_{t}:=P_{t}^{-1}\circ\not{D}_{1,0}^{H^{1}_{t}}\circ P_{t}.

Since t\not{D}_{t} depends analytically on tt by the analyticity of NN, 1,0ut\not{D}^{u_{t}}_{1,0} has minimal kernel for all but finitely many t[0,1]t\in[0,1]. Therefore, we can assume 1,0u1\not{D}^{u_{1}}_{1,0} has minimal kernel. Continuing this procedure, we can get uL1VL1VLu_{L-1}\in V_{L-1}\cap V_{L} such that 1,0uL1\not{D}^{u_{L-1}}_{1,0} also has minimal kernel and a homotopy HtL1H^{L-1}_{t} between uL1u_{L-1} and uu^{\prime} such that 1,0HtL1\not{D}^{H^{L-1}_{t}}_{1,0} has minimal kernel for all but finitely many t[0,1]t\in[0,1]. Hence the set of maps along which the (1,0)(1,0)-part of the Dirac operator has minimal kernel is CC^{\infty}-dense in [u][u]. Its C1C^{1}-openness directly follows from the continuity of the eigenvalues.

4. The heat flow for α\alpha-Dirac-harmonic maps

In this section, we will prove the short-time existence of the heat flow for α\alpha-Dirac-harmonic maps. Since we are working on a closed surface MM, we cannot uniquely solve the Dirac equation in the following system:

(4.1) tu=1(1+|u|2)α1(τα(u)1α(u,ψ)),\displaystyle\partial_{t}u=\frac{1}{(1+|\nabla u|^{2})^{\alpha-1}}\bigg{(}\tau^{\alpha}(u)-\frac{1}{\alpha}\mathcal{R}(u,\psi)\bigg{)},
(4.2) uψ=0.\displaystyle\not{D}^{u}\psi=0.

The short time existence and its extension are the obstacles. This system (if it converges) leads to an α\alpha-Dirac-harmonic map which is a solution of the system

{τα(u):=τ((1+|du|2)α)=1α(u,ψ),uψ=0,\displaystyle\begin{cases}\tau^{\alpha}(u):=\tau((1+|du|^{2})^{\alpha})=\frac{1}{\alpha}\mathcal{R}(u,\psi),\\ \not{D}^{u}\psi=0,\end{cases}

where τ\tau is the tension field.

4.1. Short time existence

As in Section 2, we now embed NN into q\mathbb{R}^{q}. Let u:MNu:M\to N with u=(uA)u=(u^{A}) and denote the spinor along the map uu by ψ=ψA(Au)\psi=\psi^{A}\otimes(\partial_{A}\circ u), where ψA\psi^{A} are spinors over MM. For any smooth map ηC0(M,q)\eta\in C^{\infty}_{0}(M,\mathbb{R}^{q}) and any smooth spinor field ξC0(ΣMq)\xi\in C^{\infty}_{0}(\Sigma M\otimes\mathbb{R}^{q}), we consider the variation

(4.3) ut=π(u+tη),ψtA=πBA(ut)(ψB+tξB),u_{t}=\pi(u+t\eta),\ \ \ \psi^{A}_{t}=\pi^{A}_{B}(u_{t})(\psi^{B}+t\xi^{B}),

where π\pi is the nearest point projection as in Section 2. Then we have

Lemma 4.1 ([10]).

The Euler-Lagrange equations for LαL^{\alpha} are

ΔuA=2(α1)βγ2uBβuBγuA1+|u|2+πBCA(u)uB,uC+πBA(u)πBDC(u)πEFC(u)ψD,uEψFα(1+|u|2)α1\begin{split}\Delta u^{A}&=-2(\alpha-1)\frac{\nabla^{2}_{\beta\gamma}u^{B}\nabla_{\beta}u^{B}\nabla_{\gamma}u^{A}}{1+|\nabla u|^{2}}+\pi^{A}_{BC}(u)\langle\nabla{u^{B}},\nabla{u^{C}}\rangle\\ &\quad+\frac{\pi^{A}_{B}(u)\pi^{C}_{BD}(u)\pi^{C}_{EF}(u)\langle\psi^{D},\nabla{u}^{E}\cdot\psi^{F}\rangle}{\alpha(1+|\nabla{u}|^{2})^{\alpha-1}}\end{split}

and

∂̸ψA=πBCA(u)uBψC.\not{\partial}\psi^{A}=\pi^{A}_{BC}(u)\nabla{u}^{B}\cdot\psi^{C}.

Lemma 4.1 implies that (4.1)-(4.2) is equivalent to

(4.4) tuA=ΔuA+2(α1)βγ2uBβuBγuA1+|u|2πBCA(u)uB,uCπBA(u)πBDC(u)πEFC(u)ψD,uEψFα(1+|u|2)α1,\displaystyle\begin{aligned} \partial_{t}u^{A}&=\Delta{u^{A}}+2(\alpha-1)\frac{\nabla^{2}_{\beta\gamma}u^{B}\nabla_{\beta}u^{B}\nabla_{\gamma}u^{A}}{1+|\nabla u|^{2}}-\pi^{A}_{BC}(u)\langle\nabla{u^{B}},\nabla{u^{C}}\rangle\\ &\quad-\frac{\pi^{A}_{B}(u)\pi^{C}_{BD}(u)\pi^{C}_{EF}(u)\langle\psi^{D},\nabla{u}^{E}\cdot\psi^{F}\rangle}{\alpha(1+|\nabla{u}|^{2})^{\alpha-1}},\end{aligned}
(4.5) πuψ=0,\displaystyle\not{D}^{\pi\circ u}\psi=0,

Now, let us state the main result of this subsection.

Theorem 4.2.

Let MM be a closed surface, and NN a closed nn-dimensional Riemannian manifold. Let u0C2+μ(M,N)u_{0}\in C^{2+\mu}(M,N) for some 0<μ<10<\mu<1 with dimker(1,0u0)=1{\rm dim}_{\mathbb{H}}{\rm ker}(\not{D}_{1,0}^{u_{0}})=1 and ψ0ker(1,0u0)\psi_{0}\in{\rm ker}(\not{D}_{1,0}^{u_{0}}) with ψ0L2=1\|\psi_{0}\|_{L^{2}}=1. Then there exists ϵ1=ϵ1(M,N)>0\epsilon_{1}=\epsilon_{1}(M,N)>0 such that, for any α(1,1+ϵ1)\alpha\in(1,1+\epsilon_{1}), the problem (4.1)-(4.2) has a solution (u,ψ)(u,\psi) with

(4.6) {ψtL2=1,t[0,T],u|t=0=u0,ψ|t=0=ψ0.\begin{cases}\|\psi_{t}\|_{L^{2}}=1,&\ \forall t\in[0,T],\\ u|_{t=0}=u_{0},\ \psi|_{t=0}=\psi_{0}.\end{cases}

satisfying

uC2+μ,1+μ/2(M×[0,T],N)u\in C^{2+\mu,1+\mu/2}(M\times[0,T],N)

and

ψCμ,μ/2(M×[0,T],ΣMuTN)L([0,T];C1+μ(M)).\psi\in C^{\mu,\mu/2}(M\times[0,T],\Sigma M\otimes u^{*}TN)\cap L^{\infty}([0,T];C^{1+\mu}(M)).

for some T>0T>0.

Proof.

We will prove the theorem in two steps. In Step 1, we will find a solution u:M×[0,T]qu:M\times[0,T]\to\mathbb{R}^{q} and ψt:MΣM(πut)TN\psi_{t}:M\to\Sigma M\otimes(\pi\circ{u_{t}})^{*}TN of (4.4)-(4.5) with the initial values (4.6). Since ψt\psi_{t} takes the value along the projection πut\pi\circ{u_{t}}, it remains to show uu takes the value in NN, which will be proved in Step 2.

Step 1: Solving (4.4)-(4.5) in q\mathbb{R}^{q}.

We first give a solution to (4.5) in a neighborhood of u0u_{0}. For any T>0T>0, we can choose ϵ\epsilon, δ\delta and RR as in the Appendix such that

u(x,t)Nδu(x,t)\in N_{\delta}

and

dN((πu)(x,t),(πv)(x,s))<ϵ<12inj(N)d^{N}((\pi\circ u)(x,t),(\pi\circ v)(x,s))<\epsilon<\frac{1}{2}{\rm inj}(N)

for all u,vBRT:=BRT(u¯0)={uXT|uu¯0XTR}{u|t=0=u0}u,v\in B^{T}_{R}:=B^{T}_{R}(\bar{u}_{0})=\{u\in X_{T}|\|u-\bar{u}_{0}\|_{X_{T}}\leq R\}\cap\{u|_{t=0}=u_{0}\}, xMx\in M and t,s[0,T]t,s\in[0,T], where u¯0(x,t)=u0(x)\bar{u}_{0}(x,t)=u_{0}(x) for any t[0,T]t\in[0,T]. If RR is small enough, then by Lemma 5.5, we have

dimker(1,0πut)=1{\rm dim}_{\mathbb{H}}{\rm ker}(\not{D}_{1,0}^{\pi\circ u_{t}})=1

and there exists Λ=12Λ(u0)\Lambda=\frac{1}{2}\Lambda(u_{0}) such that

#{spec(1,0πut)[Λ,Λ]}=1\#\{{\rm spec}(\not{D}_{1,0}^{\pi\circ u_{t}})\cap[-\Lambda,\Lambda]\}=1

for any uBRTu\in B_{R}^{T} and t[0,T]t\in[0,T], where Λ(u0)\Lambda(u_{0}) is a constant such that spec(1,0u0){0}[Λ(u0),Λ(u0)]{\rm spec}(\not{D}_{1,0}^{u_{0}})\setminus\{0\}\subset\mathbb{R}\setminus[-\Lambda(u_{0}),\Lambda(u_{0})]. Furthermore, for ψ0ker(1,0u0)\psi_{0}\in{\rm ker}(\not{D}_{1,0}^{u_{0}}) with ψ0L2=1\|\psi_{0}\|_{L^{2}}=1, Lemma 5.7 implies that

34ψ~1utL21\sqrt{\frac{3}{4}}\leq\|\tilde{\psi}_{1}^{u_{t}}\|_{L^{2}}\leq 1

for any uBR1Tu\in B_{R_{1}}^{T} and t[0,T]t\in[0,T], where ψ~ut=Pu0,utψ=ψ~1ut+ψ~2ut\tilde{\psi}^{u_{t}}=P^{u_{0},u_{t}}\psi=\tilde{\psi}^{u_{t}}_{1}+\tilde{\psi}^{u_{t}}_{2} with respect to the decomposition ΓL2=ker(1,0πut)(ker(1,0πut))\Gamma_{L^{2}}={\rm ker}(\not{D}_{1,0}^{\pi\circ u_{t}})\oplus({\rm ker}(\not{D}_{1,0}^{\pi\circ u_{t}}))^{\bot} and R1=R1(R,ϵ,u0)>0R_{1}=R_{1}(R,\epsilon,u_{0})>0.

Now, for any T>0T>0 and κ>0\kappa>0, we define

VκT:={vC1+μ,1+μ2(M×[0,T])|vC1+μ,1+μ2κ,v|M×{0}=0}.V^{T}_{\kappa}:=\{v\in C^{1+\mu,\frac{1+\mu}{2}}(M\times[0,T])|\|v\|_{C^{1+\mu,\frac{1+\mu}{2}}}\leq\kappa,\ v|_{M\times\{0\}}=0\}.

Then, there exists κR1:=κ(R1)>0\kappa_{R_{1}}:=\kappa(R_{1})>0 such that

u0+vBR1T,vVκT,κκR1.u_{0}+v\in B^{T}_{R_{1}},\ \forall v\in V_{\kappa}^{T},\ \forall\kappa\leq\kappa_{R_{1}}.

Now, we denote κ0:=κR1\kappa_{0}:=\kappa_{R_{1}} and VT:=Vκ0TV^{T}:=V^{T}_{\kappa_{0}}.

For every vVTv\in V^{T}, u0+vBR1Tu_{0}+v\in B^{T}_{R_{1}}, Lemma 5.8 gives us a solution ψ(v+u0)\psi(v+u_{0}) to the constraint equation. Since v+u0C1+μ(M)v+u_{0}\in C^{1+\mu}(M), by LpL^{p} regularity and Schauder estimates in [4], we have

(4.7) ψ(v+u0)C1+μ(M)C(μ,M,N,κ0,u0C1+μ(M)).\|\psi(v+u_{0})\|_{C^{1+\mu}(M)}\leq C(\mu,M,N,\kappa_{0},\|u_{0}\|_{C^{1+\mu}(M)}).

For any 0<t,s<T0<t,s<T, we also have

∂̸(ψ(v+u0)(t)ψ(v+u0)(s))=Γ(π(v+u0)(t))#(π(v+u0)(t))#ψ(v+u0)(t)+Γ(π(v+u0)(s))#(π(v+u0)(s))#ψ(v+u0)(s)=Γ(π(v+u0)(t))#(π(v+u0)(t))#(ψv(t)ψ(v+u0)(s))Γ(π(v+u0)(t))#((π(v+u0)(t))(π(v+u0)(s)))#ψ(v+u0)(t)(Γ(π(v+u0)(t))Γ(π(v+u0)(s)))#(π(v+u0)(s))#ψ(v+u0)(s),\begin{split}&\quad\not{\partial}(\psi(v+u_{0})(t)-\psi(v+u_{0})(s))\\ &=-\Gamma(\pi\circ(v+u_{0})(t))\#\nabla(\pi\circ(v+u_{0})(t))\#\psi(v+u_{0})(t)\\ &\quad+\Gamma(\pi\circ(v+u_{0})(s))\#\nabla(\pi\circ(v+u_{0})(s))\#\psi(v+u_{0})(s)\\ &=-\Gamma(\pi\circ(v+u_{0})(t))\#\nabla(\pi\circ(v+u_{0})(t))\#(\psi^{v}(t)-\psi(v+u_{0})(s))\\ &\quad-\Gamma(\pi\circ(v+u_{0})(t))\#(\nabla(\pi\circ(v+u_{0})(t))-\nabla(\pi\circ(v+u_{0})(s)))\#\psi(v+u_{0})(t)\\ &\quad-(\Gamma(\pi\circ(v+u_{0})(t))-\Gamma(\pi\circ(v+u_{0})(s)))\#\nabla(\pi\circ(v+u_{0})(s))\#\psi(v+u_{0})(s),\end{split}

that is,

πv(t)(ψ(v+u0)(t)ψ(v+u0)(s))=Γ(π(v+u0)(t))#((π(v+u0)(t))(π(v+u0)(s)))#ψ(v+u0)(t)(Γ(π(v+u0)(t))Γ(π(v+u0)(s)))#(π(v+u0)(s))#ψ(v+u0)(s),\begin{split}&\quad\not{D}^{\pi\circ v(t)}(\psi(v+u_{0})(t)-\psi(v+u_{0})(s))\\ &=-\Gamma(\pi\circ(v+u_{0})(t))\#(\nabla(\pi\circ(v+u_{0})(t))-\nabla(\pi\circ(v+u_{0})(s)))\#\psi(v+u_{0})(t)\\ &\quad-(\Gamma(\pi\circ(v+u_{0})(t))-\Gamma(\pi\circ(v+u_{0})(s)))\#\nabla(\pi\circ(v+u_{0})(s))\#\psi(v+u_{0})(s),\end{split}

where #\# denotes a multi-linear map with smooth coefficients. For any λ(0,1)\lambda\in(0,1), by the Sobolev embedding, LpL^{p} regularity and Lemma 5.8, we have

ψ(v+u0)(t)ψ(v+u0)(s)Cλ(M)C(λ,M,N,κ0,u0C1(M))(v(t)v(s)L(M)+dv(t)dv(sL))C(λ,M,N,κ0,u0C1(M))|ts|μ/2.\begin{split}&\quad\|\psi(v+u_{0})(t)-\psi(v+u_{0})(s)\|_{C^{\lambda}(M)}\\ &\leq C(\lambda,M,N,\kappa_{0},\|u_{0}\|_{C^{1}(M)})(\|v(t)-v(s)\|_{L^{\infty}(M)}+\|dv(t)-dv(s\|_{L^{\infty}}))\\ &\leq C(\lambda,M,N,\kappa_{0},\|u_{0}\|_{C^{1}(M)})|t-s|^{\mu/2}.\end{split}

Therefore,

(4.8) ψ(v+u0)Cμ,μ/2(M)C(μ,M,N,κ0,u0C1(M)).\|\psi(v+u_{0})\|_{C^{\mu,\mu/2}(M)}\leq C(\mu,M,N,\kappa_{0},\|u_{0}\|_{C^{1}(M)}).

Now, when α1\alpha-1 is sufficiently small, for the (v,ψv)(v,\psi^{v}) above, the standard theory of linear parabolic systems (see [15]) implies that there exists a unique solution v1C2+μ,1+μ/2(M×[0,T],q)v_{1}\in C^{2+\mu,1+\mu/2}(M\times[0,T],\mathbb{R}^{q}) to the following Dirichlet problem:

(4.9) twA=ΔgwA+2(α1)βγ2wBβ(v+u0)Bγ(v+u0)A1+|(v+u0)|2+πBCA(v+u0)(v+u0)B,(v+u0)C+(πBAπBDCπEFC)(v+u0)ψD(v+u0),(v+u0)EψF(v+u0)α(1+|(v+u0)|2)α1,+Δgu0A+2(α1)βγ2u0Bβ(v+u0)Bγ(v+u0)A1+|(v+u0)|2,\begin{split}\partial_{t}w^{A}&=\Delta_{g}w^{A}+2(\alpha-1)\frac{\nabla^{2}_{\beta\gamma}w^{B}\nabla_{\beta}(v+u_{0})^{B}\nabla_{\gamma}(v+u_{0})^{A}}{1+|\nabla(v+u_{0})|^{2}}\\ &\quad+\pi^{A}_{BC}(v+u_{0})\langle\nabla{(v+u_{0})^{B}},\nabla{(v+u_{0})^{C}}\rangle\\ &\quad+\frac{(\pi^{A}_{B}\pi^{C}_{BD}\pi^{C}_{EF})(v+u_{0})\langle\psi^{D}(v+u_{0}),\nabla{(v+u_{0})}^{E}\cdot\psi^{F}(v+u_{0})\rangle}{\alpha(1+|\nabla{(v+u_{0})}|^{2})^{\alpha-1}},\\ &\quad+\Delta_{g}u_{0}^{A}+2(\alpha-1)\frac{\nabla^{2}_{\beta\gamma}u_{0}^{B}\nabla_{\beta}(v+u_{0})^{B}\nabla_{\gamma}(v+u_{0})^{A}}{1+|\nabla(v+u_{0})|^{2}},\end{split}
(4.10) w(,0)=0,w(\cdot,0)=0,

satisfying

v1C2+μ,1+μ/2(M×[0,T])C(μ,M,N)(v1C0(M×[0,T])+u0C2+ν(M)+κ0).\|v_{1}\|_{C^{2+\mu,1+\mu/2}(M\times[0,T])}\leq C(\mu,M,N)(\|v_{1}\|_{C^{0}(M\times[0,T])}+\|u_{0}\|_{C^{2+\nu}(M)}+\kappa_{0}).

Since v1(,0)=0v_{1}(\cdot,0)=0, we have

v1C0(M×[0,T])C(μ,M,N)T(v1C0(M×[0,T])+u0C2+ν(M)+κ0).\|v_{1}\|_{C^{0}(M\times[0,T])}\leq C(\mu,M,N)T(\|v_{1}\|_{C^{0}(M\times[0,T])}+\|u_{0}\|_{C^{2+\nu}(M)}+\kappa_{0}).

By taking T>0T>0 small enough, we get

v1C0(M×[0,T])C(μ,M,N)T(u0C2+ν(M)+κ0).\|v_{1}\|_{C^{0}(M\times[0,T])}\leq C(\mu,M,N)T(\|u_{0}\|_{C^{2+\nu}(M)}+\kappa_{0}).

Then the interpolation inequality in [12] implies that v1VTv_{1}\in V^{T} for T>0T>0 sufficiently small. For such v1v_{1}, we have ψ(v1+u0)\psi(v_{1}+u_{0}) satisfying (4.7) and (4.8). Replacing (v,ψ(v+u0))(v,\psi(v+u_{0})) in (4.9)-(4.10) by (v1,ψ(v1+u0))(v_{1},\psi(v_{1}+u_{0})), then we get v2VTv_{2}\in V^{T}. Iterating this procedure, we get a solution vk+1v_{k+1} of (4.9)-(4.10) with (v,ψ(v+u0))(v,\psi(v+u_{0})) replacing by (vk,ψ(vk+u0))(v_{k},\psi(v_{k}+u_{0})), which satisfies

ψ(vk+1+u0)Cμ,μ/2(M)C(μ,M,N,κ0,u0C1(M)).\|\psi(v_{k+1}+u_{0})\|_{C^{\mu,\mu/2}(M)}\leq C(\mu,M,N,\kappa_{0},\|u_{0}\|_{C^{1}(M)}).

and

vk+1C2+μ,1+μ/2(M×[0,T])C(μ,M,N)(u0C2+ν(M)+κ0).\|v_{k+1}\|_{C^{2+\mu,1+\mu/2}(M\times[0,T])}\leq C(\mu,M,N)(\|u_{0}\|_{C^{2+\nu}(M)}+\kappa_{0}).

By passing to a subsequence, we know that vkv_{k} converges to some uu in C2,1(M×[0,T])C^{2,1}(M\times[0,T]) and ψvk+u0\psi^{v_{k}+u_{0}} converges to some ψ\psi in C0(M×[0,T])C^{0}(M\times[0,T]). Then it is easy to see that (u,ψ)(u,\psi) is a solution of (4.4)-(4.5) with u(,0)=u0u(\cdot,0)=u_{0} and ψ(,0)=ψ0\psi(\cdot,0)=\psi_{0}.

Step 2: u(x,t)u(x,t) takes value in NN for any (x,t)M×[0,T](x,t)\in M\times[0,T].

Suppose uC2,1(M×[0,T],q)u\in C^{2,1}(M\times[0,T],\mathbb{R}^{q}) and ψCμ,μ/2(M×[0,T],ΣM(πu)TN)L([0,T];C1+μ(M))\psi\in C^{\mu,\mu/2}(M\times[0,T],\Sigma M\otimes(\pi\circ u)^{*}TN)\cap L^{\infty}([0,T];C^{1+\mu}(M)) satisfy (4.4)-(4.5). In the following, we write ||||||\cdot|| and ,\langle\cdot,\cdot\rangle for the Euclidean norm and scalar product, respectively. Similarly, we write ||||g||\cdot||_{g} and ,g\langle\cdot,\cdot\rangle_{g} for the norm and inner product of (M,g)(M,g), respectively. We define

ρ:qq\rho:\mathbb{R}^{q}\to\mathbb{R}^{q}

by ρ(z)=zπ(z)\rho(z)=z-\pi(z) and

φ:M×[0,T]\varphi:M\times[0,T]\to\mathbb{R}

by φ(x,t)=ρ(u(x,t))2=A=1q|ρA(u(x,t))|2\varphi(x,t)=||\rho(u(x,t))||^{2}=\sum\limits_{A=1}^{q}|\rho^{A}(u(x,t))|^{2}. A direct computation yields

(tΔ)φ(x,t)=2A=1q(ρAu)(x,t)g2+2ρu,πBA(u)F1B(u)+2α(1+|u|2)α1ρu,ρBA(u)F2B(u,ψ)+4(α1)1+|u|2ρu,βγ2uCβuCγuBρBA(u),\begin{split}(\frac{\partial}{\partial{t}}-\Delta)\varphi(x,t)&=-2\sum\limits_{A=1}^{q}||\nabla(\rho^{A}\circ u)(x,t)||_{g}^{2}\\ &\quad+2\langle\rho\circ u,-\pi^{A}_{B}(u)F^{B}_{1}(u)\rangle\\ &\quad+\frac{2}{\alpha(1+|\nabla{u}|^{2})^{\alpha-1}}\langle\rho\circ u,\rho^{A}_{B}(u)F^{B}_{2}(u,\psi)\rangle\\ &\quad+\frac{4(\alpha-1)}{1+|\nabla{u}|^{2}}\langle\rho\circ u,\nabla^{2}_{\beta\gamma}u^{C}\nabla_{\beta}u^{C}\nabla_{\gamma}u^{B}\rho_{B}^{A}(u)\rangle,\end{split}

where F1AF_{1}^{A} and F2AF_{2}^{A} are defined in (2.9) and (2.10), respectively.

Since ρuTπuN\rho\circ u\in T^{\perp}_{\pi\circ{u}}N and (dπ)u:qTπuN(d\pi)_{u}:\mathbb{R}^{q}\to T_{\pi\circ{u}}N, we have

ρu,πBA(u)F1B=ρu,ρBA(u)F2B=0.\langle\rho\circ u,-\pi^{A}_{B}(u)F^{B}_{1}\rangle=\langle\rho\circ u,\rho^{A}_{B}(u)F^{B}_{2}\rangle=0.

Together with

4(α1)1+|u|2ρu,βγ2uCβuCγuBρBA(u)4(α1)uC2(M)ρu(ρu)2(α1)(uC2(M)2φ+(ρu)2),\begin{split}&\frac{4(\alpha-1)}{1+|\nabla{u}|^{2}}\langle\rho\circ u,\nabla^{2}_{\beta\gamma}u^{C}\nabla_{\beta}u^{C}\nabla_{\gamma}u^{B}\rho_{B}^{A}(u)\rangle\\ &\leq 4(\alpha-1)||u||_{C^{2}(M)}||\rho\circ{u}||||\nabla(\rho\circ{u})||\\ &\leq 2(\alpha-1)(||u||^{2}_{C^{2}(M)}\varphi+||\nabla(\rho\circ{u})||^{2}),\end{split}

we get (tΔ)φ(x,t)Cφ,(\frac{\partial}{\partial{t}}-\Delta)\varphi(x,t)\leq C\varphi, where C=C(uC2,1(M×[0,T]))C=C(\|u\|_{C^{2,1}(M\times[0,T])}). Since φ(x,t)0\varphi(x,t)\geq 0 and φ(x,0)=0\varphi(x,0)=0 for any (x,t)M×[0,T](x,t)\in M\times[0,T], we conclude φ=0\varphi=0 on M×[0,T]M\times[0,T]. We have shown that u(x,t)Nu(x,t)\in N for all (x,t)M×[0,T](x,t)\in M\times[0,T].

Finally, by using the ϵ\epsilon-regularity (see Lemma 4.5 below), we conclude that

uC2+μ,1+μ/2(M×[0,T],N)u\in C^{2+\mu,1+\mu/2}(M\times[0,T],N)

and

ψCμ,μ/2(M×[0,T],ΣMuTN)L([0,T];C1+μ(M)).\psi\in C^{\mu,\mu/2}(M\times[0,T],\Sigma M\otimes u^{*}TN)\cap L^{\infty}([0,T];C^{1+\mu}(M)).

4.2. Regularity of the flow

In this subsection, we will give some estimates for the regularity of the flow. The proofs can be found in [10] and the references therein. Let us start with the following estimate of the energy of the map part.

Lemma 4.3.

Suppose (u,ψ)(u,\psi) is a solution of (4.1)-(4.2) with the initial values (4.6). Then there holds

Eα(u(t))+2α0tM(1+|u|2)α1|tu|2=Eα(u0).E_{\alpha}(u(t))+2\alpha\int_{0}^{t}\int_{M}(1+|\nabla{u}|^{2})^{\alpha-1}|\partial_{t}u|^{2}=E_{\alpha}(u_{0}).

Moreover, Eα(u(t))E_{\alpha}(u(t)) is absolutely continuous on [0,T][0,T] and non-increasing.

Consequently, we can also control the spinor part along the heat flow of the α\alpha-Dirac-harmonic map.

Lemma 4.4.

Suppose (u,ψ)(u,\psi) is a solution of (4.1)-(4.2) with the initial values (4.6). Then for any p(1,2)p\in(1,2), there holds

ψ(,t)W1,p(M)C,t[0,T],||\psi(\cdot,t)||_{W^{1,p}(M)}\leq C,\ \forall t\in[0,T],

where C=C(p,M,N,Eα(u0))C=C(p,M,N,E_{\alpha}(u_{0})).

To get the convergence of the flow, we also need the following ϵ\epsilon-regularity.

Lemma 4.5.

Suppose (u,ψ)(u,\psi) is a solution of (4.1)-(4.2) with the initial values (4.6). Given ω0=(x0,t0)M×(0,T]\omega_{0}=(x_{0},t_{0})\in M\times(0,T], denote

PR(ω0):=BR(x0)×[t0R2,t0].P_{R}(\omega_{0}):=B_{R}(x_{0})\times[t_{0}-R^{2},t_{0}].

Then there exist three constants ϵ2=ϵ2(M,N)>0\epsilon_{2}=\epsilon_{2}(M,N)>0, ϵ3=ϵ3(M,N,u0)>0\epsilon_{3}=\epsilon_{3}(M,N,u_{0})>0 and C=C(μ,R,M,N,Eα(u0))>0C=C(\mu,R,M,N,E_{\alpha}(u_{0}))>0 such that if

1<α<1+ϵ2,andsup[t04R2,t0]E(u(t);B2R(ω0))ϵ3,1<\alpha<1+\epsilon_{2},\ \text{and}\sup_{[t_{0}-4R^{2},t_{0}]}E(u(t);B_{2R}(\omega_{0}))\leq\epsilon_{3},

then

RψL(PR(ω0))+RuL(PR(ω0))C\sqrt{R}||\psi||_{L^{\infty}(P_{R}(\omega_{0}))}+R||\nabla{u}||_{L^{\infty}(P_{R}(\omega_{0}))}\leq C

and for any 0<β<10<\beta<1,

sup[t0R24,t0]ψ(t)C1+μ(BR/2(x0))+uCβ,β/2(PR/2(ω0))C(β).\sup_{[t_{0}-\frac{R^{2}}{4},t_{0}]}||\psi(t)||_{C^{1+\mu}(B_{R/2}(x_{0}))}+||\nabla{u}||_{C^{\beta,\beta/2}(P_{R/2}(\omega_{0}))}\leq C(\beta).

Moreover, if

supMsup[t04R2,t0]E(u(t);B2R(ω0))ϵ3,\sup_{M}\sup_{[t_{0}-4R^{2},t_{0}]}E(u(t);B_{2R}(\omega_{0}))\leq\epsilon_{3},

then

uC2+μ,1+μ/2(M×[t0R28,t0])+ψCμ,μ/2(M×[t0R28,t0])+sup[t0R28,t0]ψ(t)C1+μ(M)C.||u||_{C^{2+\mu,1+\mu/2}(M\times[t_{0}-\frac{R^{2}}{8},t_{0}])}+||\psi||_{C^{\mu,\mu/2}(M\times[t_{0}-\frac{R^{2}}{8},t_{0}])}+\sup_{[t_{0}-\frac{R^{2}}{8},t_{0}]}||\psi(t)||_{C^{1+\mu}(M)}\leq C.

4.3. Existence of Dirac-harmonic maps

In this section, we will prove Theorem 4.7 by the short-time existence of α\alpha-Dirac-harmonic map flow. First, we will prove the following existence result about the α\alpha-Dirac-harmonic maps for α>1\alpha>1. Then, by the compactness, we get a Dirac-harmonic map as the limit of these α\alpha-Dirac-harmonic maps. Last, we prove that the bubbles only can be harmonic spheres, and finish the proof of Theorem 4.7.

Theorem 4.6.

Let MM be a closed spin surface and (N,h)(N,h) a closed Kähler manifold. Suppose there exists a map u0C2+μ(M,N)u_{0}\in C^{2+\mu}(M,N) for some μ(0,1)\mu\in(0,1) such that dimker1,0u0=1{\rm dim}_{\mathbb{H}}{\rm ker}\not{D}_{1,0}^{u_{0}}=1. Then for any α(1,1+ϵ1)\alpha\in(1,1+\epsilon_{1}), there exists a nontrivial smooth α\alpha-Dirac-harmonic map (uα,ψα)(u_{\alpha},\psi_{\alpha}) such that the map part uαu_{\alpha} stays in the same homotopy class as u0u_{0} and ψαL2=1\|\psi_{\alpha}\|_{L^{2}}=1.

Proof of Theorem 4.6.

Let us define

m0α:=inf{Eα(u)|uW1,2α(M,N)[u0]},m^{\alpha}_{0}:=\inf\{E_{\alpha}(u)|u\in W^{1,2\alpha}(M,N)\cap[u_{0}]\},

where [u0][u_{0}] denotes the homotopy class of u0u_{0}. If u0u_{0} is a minimizing α\alpha-harmonic map, it follows from Lemma 4.3 that (u0,ψ0)(u_{0},\psi_{0}) is an α\alpha-Dirac-harmonic map for any ψ0ker1,0u0\psi_{0}\in{\rm ker}\not{D}_{1,0}^{u_{0}}. If Eα(u0)>m0αE_{\alpha}(u_{0})>m^{\alpha}_{0}, then Theorem 4.2 gives us a solution

uC2+μ,1+μ/2(M×[0,T),N)u\in C^{2+\mu,1+\mu/2}(M\times[0,T),N)

and

ψCμ,μ/2(M×[0,T),ΣMuTN)0<s<TL([0,s];C1+μ(M)).\psi\in C^{\mu,\mu/2}(M\times[0,T),\Sigma M\otimes u^{*}TN)\cap{\cap_{0<s<T}}L^{\infty}([0,s];C^{1+\mu}(M)).

to the problem (4.1)-(4.2) with the initial values (4.6).

By Lemma 4.3, we know

M(1+|u|2)αEα(u0).\int_{M}(1+|\nabla{u}|^{2})^{\alpha}\leq E_{\alpha}(u_{0}).

Then it is easy to see that, for any 0<ϵ<ϵ30<\epsilon<\epsilon_{3}, there exists a positive constant r0=r0(ϵ,α,Eα(u0))r_{0}=r_{0}(\epsilon,\alpha,E_{\alpha}(u_{0})) such that for all (x,t)M×[0,T)(x,t)\in M\times[0,T), there holds

Br0(x)|u|2CEα(u0)1/αr011αϵ.\int_{B_{r_{0}}(x)}|\nabla{u}|^{2}\leq CE_{\alpha}(u_{0})^{1/\alpha}r_{0}^{1-\frac{1}{\alpha}}\leq\epsilon.

Therefore, by Theorem 4.2 and Lemma 4.5, we know that the singular time can be characterized as

Z={T|limtiTdimker1,0u(ti)>1}Z=\{T\in\mathbb{R}|\lim\limits_{t_{i}\nearrow T}{\rm dim}_{\mathbb{H}}{\rm ker}\not{D}_{1,0}^{u(t_{i})}>1\}

and there exists a sequence {ti}T\{t_{i}\}\nearrow T such that

(u(,ti),ψ(,ti))(u(,T),ψ(,T))inC2+μ(M)×C1+μ/2(M)(u(\cdot,t_{i}),\psi(\cdot,t_{i}))\to(u(\cdot,T),\psi(\cdot,T))\ \text{in}\ C^{2+\mu}(M)\times C^{1+\mu/2}(M)

and

ψ(,T)L2=1.\|\psi(\cdot,T)\|_{L^{2}}=1.

If Z=Z=\emptyset, then, by Theorem 4.2, we can extend the solution (u,ψ)(u,\psi) beyond the time TT by using (u(,T),ψ(,T))(u(\cdot,T),\psi(\cdot,T)) as new initial values. Thus, we have the global existence of the flow. For the limit behavior as tt\to\infty, Lemma 4.3 implies that there exists a sequence {ti}\{t_{i}\}\to\infty such that

(4.11) M|tu|2(,ti)0.\int_{M}|\partial_{t}u|^{2}(\cdot,t_{i})\to 0.

Together with Lemma 4.5, there is a subsequence, still denoted by {ti}\{t_{i}\}, and an α\alpha-Dirac-harmonic map (uα,ψα)C(M,N)×C(M,ΣM(uα)TN)(u_{\alpha},\psi_{\alpha})\in C^{\infty}(M,N)\times C^{\infty}(M,\Sigma M\otimes(u_{\alpha})^{*}TN) such that (u(,ti),ψ(,ti))(u(\cdot,t_{i}),\psi(\cdot,t_{i})) converges to (uα,ψα)(u_{\alpha},\psi_{\alpha}) in C2(M)×C1(M)C^{2}(M)\times C^{1}(M) and ψαL2=1\|\psi_{\alpha}\|_{L^{2}}=1.

If ZZ\neq\emptyset and TZT\in{Z}, let us assume that Eα(u(,T))>m0αE_{\alpha}(u(\cdot,T))>m^{\alpha}_{0} and (u(,T),ψ(,T))(u(\cdot,T),\psi(\cdot,T)) is not already an α\alpha-Dirac-harmonic map. We extend the flow as follows: By Lemma 3.1, there is a map u1C2+μ(M,N)u_{1}\in C^{2+\mu}(M,N) such that

(4.12) m0α<Eα(u1)<Eα(u(,T))m^{\alpha}_{0}<E_{\alpha}(u_{1})<E_{\alpha}(u(\cdot,T))

and

(4.13) dimker1,0u1=1.{\rm dim}_{\mathbb{H}}{\rm ker}\not{D}_{1,0}^{u_{1}}=1.

Thus, picking any ψ1keru1\psi_{1}\in{\rm ker}\not{D}^{u_{1}} with ψ1L2=1\|\psi_{1}\|_{L^{2}}=1, we can restart the flow from the new initial values (u1,ψ1)(u_{1},\psi_{1}). If there is no singular time along the flow started from (u1,ψ1)(u_{1},\psi_{1}), then we get an α\alpha-Dirac-harmonic map as in the case of Z=Z=\emptyset. Otherwise, we use again the procedure above to choose (u2,ψ2)(u_{2},\psi_{2}) as initial values and restart the flow. This procedure will stop in finitely or infinitely many steps.

If infinitely many steps are required, then there exist infinitely many flow pieces {ui(x,t)}i=1,,\{u_{i}(x,t)\}_{i=1,\dots,\infty} and {Ti}i=1,,\{T_{i}\}_{i=1,\dots,\infty} such that

Eα(ui(t))+2α0tM(1+|ui|2)α1|tui|2=Eα(ui),t(0,Ti),E_{\alpha}(u_{i}(t))+2\alpha\int_{0}^{t}\int_{M}(1+|\nabla{u}_{i}|^{2})^{\alpha-1}|\partial_{t}u_{i}|^{2}=E_{\alpha}(u_{i}),\ \forall t\in(0,T_{i}),

where ui(,0)=uiC2+μ(M,N)u_{i}(\cdot,0)=u_{i}\in C^{2+\mu}(M,N). If the TiT_{i} are bounded away from zero, then there is {ti}\{t_{i}\} such that (4.11) holds for ti(0,Ti)t_{i}\in(0,T_{i}). Therefore, we have an α\alpha-Dirac-harmonic map as before. If Ti0T_{i}\to 0, then we look at the limit of Eα(ui)E_{\alpha}(u_{i}). If the limit is strictly bigger than m0αm^{\alpha}_{0}, we again choose another map satisfying (4.12) and (4.13) as a new starting point. If the limit is exactly m0αm^{\alpha}_{0}, then we choose {ti}\{t_{i}\} such that ti(0,Ti)t_{i}\in(0,T_{i}) for each ii. By Lemma 4.5, ui(ti)u_{i}(t_{i}) converges in C2(M)×C1(M)C^{2}(M)\times C^{1}(M) to a minimizing α\alpha-harmonic map uαu_{\alpha}. If 1,0uα\not{D}_{1,0}^{u_{\alpha}} has minimal kernel, then for any ψker1,0uα\psi\in{\rm ker}{\not{D}_{1,0}^{u_{\alpha}}}, (uα,ψ)(u_{\alpha},\psi) is an α\alpha-Dirac-harmonic map as we showed in the beginning of the proof. If 1,0uα\not{D}_{1,0}^{u_{\alpha}} has non-minimal kernel, by using the 2\mathbb{Z}_{2}-grading GidG\otimes id as in the proof of Theorem 1.2, we get α\alpha-Dirac-harmonic maps (uα,ψα±)(u_{\alpha},\psi_{\alpha}^{\pm}) for any keruαψα=ψα++ψα{\rm ker}{\not{D}^{u_{\alpha}}}\ni\psi_{\alpha}=\psi_{\alpha}^{+}+\psi_{\alpha}^{-}. In particular, we can choose ψα\psi_{\alpha} such that ψα+L2=1\|\psi_{\alpha}^{+}\|_{L^{2}}=1 or ψαL2=1\|\psi_{\alpha}^{-}\|_{L^{2}}=1. By this procedure, we either get an α\alpha-Dirac-harmonic map or keep on choosing new maps satisfying (4.12) and (4.13). In the latter case, since the energies of the initial maps are bounded and decreasing, they converge to the minimizing energy m0αm^{\alpha}_{0}. (Otherwise, suppose the constant is C>m0αC>m^{\alpha}_{0}. Then one can choose a new map with a lower energy such that the limit is not CC.) Therefore, we also get an α\alpha-Dirac-harmonic map in the latter case as before.

If it stops in finitely many steps, there exists a sequence {ti}\{t_{i}\} and some 0<Tk+0<T_{k}\leq+\infty such that

limtiT(u(,ti),ψ(,ti))(uα,ψα)inC2(M)×C1(M),\lim\limits_{t_{i}\nearrow T}(u(\cdot,t_{i}),\psi(\cdot,t_{i}))\to(u_{\alpha},\psi_{\alpha})\ \text{in}\ C^{2}(M)\times C^{1}(M),

where (uα,ψα)(u_{\alpha},\psi_{\alpha}) either is an α\alpha-Dirac-harmonic map or satisfies Eα(uα)=m0αE_{\alpha}(u_{\alpha})=m^{\alpha}_{0}. And in the latter case, uαu_{\alpha} is a minimizing α\alpha-harmonic map. Then we can again get a nontrivial α\alpha-Dirac-harmonic map as above.

By Theorem 4.6, for any α>1\alpha>1 sufficiently close to 11, there exists an α\alpha-Dirac-harmonic map (uα,ψα)(u_{\alpha},\psi_{\alpha}) with the properties

(4.14) Eα(uα)Eα(u0),ψαL2=1E_{\alpha}(u_{\alpha})\leq E_{\alpha}(u_{0}),\ \ \|\psi_{\alpha}\|_{L^{2}}=1

and

(4.15) ψαW1,p(M)C(p,M,N,Eα(u0))||\psi_{\alpha}||_{W^{1,p}(M)}\leq C(p,M,N,E_{\alpha}(u_{0}))

for any 1<p<21<p<2. Then it is natural to consider the limit behavior when α\alpha decreases to 11. Together with the blow-up analysis in [8], we have the following existence result.

Theorem 4.7.

Let MM be a closed Riemann surface and NN a complex nn-dimensional analytic Kähler manifold and a parallel real structure be j2j_{2} defined in Theorem 1.2. Suppose there exists a map u0C2+μ(M,N)u_{0}\in C^{2+\mu}(M,N) for some μ(0,1)\mu\in(0,1) such that dimker1,0u0=1{\rm dim}_{\mathbb{H}}{\rm ker}\not{D}_{1,0}^{u_{0}}=1. Then there exists a nontrivial (i.e. Ψ0\Psi\neq 0) smooth Dirac-harmonic map (Φ,Ψ)(\Phi,\Psi) with ΨL2=1\|\Psi\|_{L^{2}}=1. In particular, if NN has nonpositive curvature, then the map Φ\Phi stays in the same homotopy class as u0u_{0}.

Proof.

By Theorem 4.6, we have a sequence of smooth α\alpha-Dirac-harmonic maps (uαk,ψαk)(u_{\alpha_{k}},\psi_{\alpha_{k}}) with (4.14) and (4.15), where αk1\alpha_{k}\searrow 1 as kk\to\infty. Then, by the compactness theorem in [8], there is a constant ϵ0>0\epsilon_{0}>0 and a Dirac-harmonic map

(Φ,Ψ)C(M,N)×C(M,ΣMΦTN)(\Phi,\Psi)\in C^{\infty}(M,N)\times C^{\infty}(M,\Sigma M\otimes\Phi^{*}TN)

such that

(uαk,ψαk)(Φ,Ψ)inCloc2(M𝒮)×Cloc1(M𝒮),(u_{\alpha_{k}},\psi_{\alpha_{k}})\to(\Phi,\Psi)\ \text{in}\ C^{2}_{loc}(M\setminus{\mathcal{S}})\times C^{1}_{loc}(M\setminus{\mathcal{S}}),

where

𝒮:={xM|lim infαk1E(uαk;Br(x))ϵ02,r>0}\mathcal{S}:=\{x\in M|\liminf_{\alpha_{k}\to 1}E(u_{\alpha_{k}};B_{r}(x))\geq\frac{\epsilon_{0}}{2},\forall r>0\}

is a finite set.

Now, taking x0𝒮x_{0}\in\mathcal{S}, there exists a sequence xαkx0x_{\alpha_{k}}\to x_{0}, λαk0\lambda_{\alpha_{k}}\to 0 and a nontrivial Dirac-harmonic map (ϕ,ξ):2N(\phi,\xi):\mathbb{R}^{2}\to N such that

(uαk(xαk+λαkx),λαkαk1λαkψαk(xαk+λαkx))(ϕ,ξ)inCloc2(2),(u_{\alpha_{k}}(x_{\alpha_{k}}+\lambda_{\alpha_{k}}x),\lambda_{\alpha_{k}}^{{\alpha_{k}}-1}\sqrt{\lambda_{\alpha_{k}}}\psi_{\alpha_{k}}(x_{\alpha_{k}}+\lambda_{\alpha_{k}}x))\to(\phi,\xi)\ \text{in}\ C^{2}_{loc}(\mathbb{R}^{2}),

as α1\alpha\to 1. Choose any p>4p^{*}>4, by taking p=2p2+pp=\frac{2p^{*}}{2+p^{*}} in (4.15), we get

ψαkLp(M)C(p,M,N,Eαk(u0))||\psi_{\alpha_{k}}||_{L^{p^{*}}(M)}\leq C(p^{*},M,N,E^{\alpha_{k}}(u_{0}))

and

ξL4(DR(0))=limαk1λαkαk1ψαkL4(DλαkR(xαk))limαk1CψαkLp(M)(λαkR)2(141p)=0.\begin{split}||\xi||_{L^{4}(D_{R}(0))}&=\lim\limits_{{\alpha_{k}}\to 1}\lambda^{{\alpha_{k}}-1}_{\alpha_{k}}||\psi_{\alpha_{k}}||_{L^{4}(D_{\lambda_{\alpha_{k}}{R}}(x_{\alpha_{k}}))}\\ &\leq\lim\limits_{{\alpha_{k}}\to 1}C||\psi_{\alpha_{k}}||_{L^{p^{*}}(M)}(\lambda_{\alpha_{k}}{R})^{2(\frac{1}{4}-\frac{1}{p^{*}})}=0.\end{split}

Thus, ξ=0\xi=0 and ϕ\phi can be extended to a nontrivial smooth harmonic sphere. Since ψαL2=1||\psi_{\alpha}||_{L^{2}}=1, the Sobolev embedding implies that ΨL2(M)=limαk1ψαL2(M)=1||\Psi||_{L^{2}(M)}=\lim\limits_{{\alpha_{k}}\to 1}||\psi_{\alpha}||_{L^{2}(M)}=1. Therefore, (Φ,Ψ)(\Phi,\Psi) is nontrivial. Furthermore, if (N,h)(N,h) does not admit any nontrivial harmonic sphere, then

(uαk,ψαk)(Φ,Ψ)inC2(M)×C1(M).(u_{\alpha_{k}},\psi_{\alpha_{k}})\to(\Phi,\Psi)\ \text{in}\ C^{2}(M)\times C^{1}(M).

Therefore, Φ\Phi is in the same homotopy class as u0u_{0}. ∎

5. Appendix

We will use the parallel construction in [10] to construct the solution to the constraint equation for spinors under a different pull-back bundle uT1,0Nu^{*}T_{1,0}N. Since the only thing changed is the bundle we twisted, the proofs of those nice properties are parallel to those in [10]. For completeness, we give the details in this appendix.

For every T>0T>0, we consider the space BRT(u¯0):={uXT|uu¯0XTR}{u|t=0=u0}B^{T}_{R}(\bar{u}_{0}):=\{u\in X_{T}|\|u-\bar{u}_{0}\|_{X_{T}}\leq R\}\cap\{u|_{t=0}=u_{0}\} where u¯0(x,t)=u0(x)\bar{u}_{0}(x,t)=u_{0}(x) for any t[0,T]t\in[0,T]. To get the necessary estimate for the solution of the constraint equation, we will use the parallel transport along the unique shortest geodesic between u0(x)u_{0}(x) and πut(x)\pi\circ{u_{t}}(x) in N. To do this, we need the following lemma which tells us that the distances in NN can be locally controlled by the distances in q\mathbb{R}^{q}.

Lemma 5.1.

Let NqN\subset\mathbb{R}^{q} be a closed embedded submanifold of q\mathbb{R}^{q} with the induced Riemannian metric. Denote by AA its Weingarten map. Choose C>0C>0 such that AC||A||\leq C, where

A:=sup{AvX|vTpN,XTpN,v=1,X=1,pN}.||A||:=\sup\{||A_{v}X|||\ v\in T_{p}^{\perp}{N},\ X\in T_{p}N,\ ||v||=1,\ ||X||=1,\ p\in N\}.

Then there exists 0<δ0<1C0<\delta_{0}<\frac{1}{C} such that for all 0<δδ00<\delta\leq\delta_{0} and for all p,qNp,q\in N with pq2<δ||p-q||_{2}<\delta, it holds that

dN(p,q)11δCpq2,d^{N}(p,q)\leq\frac{1}{1-\delta{C}}||p-q||_{2},

where we denote the Euclidean norm by ||||2||\cdot||_{2} in this section.

In the following, we will choose δ\delta and RR to ensure the existence of the unique shortest geodesics between the projections of any two elements in BRT(u¯0)B^{T}_{R}(\bar{u}_{0}). By the definition of BRT(u¯0)B^{T}_{R}(\bar{u}_{0}), we have

u(x,t)u¯0(x,t)2=u(x,t)u0(x)2R||u(x,t)-\bar{u}_{0}(x,t)||_{2}=||u(x,t)-u_{0}(x)||_{2}\leq R

for all (x,t)M×[0,T](x,t)\in M\times[0,T]. Then taking any RδR\leq\delta, we get

d(u(x,t),N)u(x,t)u0(x)2δd(u(x,t),N)\leq||u(x,t)-u_{0}(x)||_{2}\leq\delta

for all (x,t)M×[0,T](x,t)\in M\times[0,T]. Therefore, u(x,t)Nδu(x,t)\in N_{\delta}. In particular, πu\pi\circ{u} is NN-valued, and

(5.1) (πu)(x,t)u0(x)2(πu)(x,t)u(x,t)2+u(x,t)u0(x)22δ.||(\pi\circ{u})(x,t)-u_{0}(x)||_{2}\leq||(\pi\circ{u})(x,t)-u(x,t)||_{2}+||u(x,t)-u_{0}(x)||_{2}\leq 2\delta.

Now, we choose ϵ>0\epsilon>0 with 2ϵ<inj(N)2\epsilon<{\rm inj}(N) and δ\delta such that

(5.2) δ<min{14δ0,14ϵ(1δ0C)},\displaystyle\delta<\min\left\{\frac{1}{4}\delta_{0},\frac{1}{4}\epsilon(1-\delta_{0}C)\right\},

where δ0,C>0\delta_{0},C>0 are as in Lemma 5.1. From (5.1), we know that for all u,vBRT(u¯0)u,v\in B^{T}_{R}(\bar{u}_{0}), it holds that

(πu)(x,t)(πv)(x,s)24δ<δ0.||(\pi\circ{u})(x,t)-(\pi\circ{v})(x,s)||_{2}\leq 4\delta<\delta_{0}.

Then Lemma 5.1 and (5.2) imply that

(5.3) dN((πu)(x,t),(πv)(x,s))11δ0C(πu)(x,t)(πv)(x,s)211δ0C4δ<ϵ<12inj(N).\begin{split}d^{N}((\pi\circ{u})(x,t),(\pi\circ{v})(x,s))&\leq\frac{1}{1-\delta_{0}C}||(\pi\circ{u})(x,t)-(\pi\circ{v})(x,s)||_{2}\\ &\leq\frac{1}{1-\delta_{0}C}4\delta<\epsilon<\frac{1}{2}{\rm inj}(N).\end{split}

To summarize, under the choice of constants as follows:

(5.4) {ϵ>0,s.t. 2ϵ<inj(N),δ>0,s.t.δ<min{14δ0,14ϵ(1δ0C)},Rδ,\begin{cases}\epsilon>0,&\text{s.t.}\ 2\epsilon<{\rm inj}(N),\\ \delta>0,&\text{s.t.}\ \delta<\min\{\frac{1}{4}\delta_{0},\frac{1}{4}\epsilon(1-\delta_{0}C)\},\\ R\leq\delta,\end{cases}

we have shown that

(5.5) u(x,t)Nδu(x,t)\in N_{\delta}

and

(5.6) dN((πu)(x,t),(πv)(x,s))<ϵ<12inj(N)d^{N}((\pi\circ u)(x,t),(\pi\circ v)(x,s))<\epsilon<\frac{1}{2}{\rm inj}(N)

for all u,vBRT(u¯0)u,v\in B_{R}^{T}(\bar{u}_{0}), xMx\in M and t,s[0,T]t,s\in[0,T].

Using the properties (5.5) and (5.6), we can prove two important estimates. One is for the Dirac operators along maps.

Lemma 5.2.

Choose ϵ\epsilon, δ\delta and RR as in (5.4). If ϵ>0\epsilon>0 is small enough, then there exists C=C(R)>0C=C(R)>0 such that

((Pvs,ut)11,0πutPvs,ut1,0πvs)ψ(x)CutvsC0(M,q)ψ(x)||((P^{v_{s},u_{t}})^{-1}\not{D}_{1,0}^{\pi\circ{u_{t}}}P^{v_{s},u_{t}}-\not{D}_{1,0}^{\pi\circ{v_{s}}})\psi(x)||\leq C||u_{t}-v_{s}||_{C^{0}(M,\mathbb{R}^{q})}||\psi(x)||

for any u,vBRT(u¯0)u,v\in B^{T}_{R}(\bar{u}_{0}), ψΓC1(ΣM(πvs)T1,0N)\psi\in\Gamma_{C^{1}}(\Sigma M\otimes(\pi\circ{v_{s}})^{*}T_{1,0}N), xMx\in M and t,s[0,T]t,s\in[0,T].

Proof.

We write f0:=πvsf_{0}:=\pi\circ{v_{s}}, f1:=πutf_{1}:=\pi\circ{u_{t}} and define the C1C^{1} map F:M×[0,1]NF:M\times[0,1]\to N by

F(x,t):=expf0(x)(texpf0(x)1f1(x))F(x,t):=\exp_{f_{0}(x)}(t\exp^{-1}_{f_{0}(x)}f_{1}(x))

where exp\exp denotes the exponential map of the Riemannian manifold NN. Note that F(,0)=f0F(\cdot,0)=f_{0}, F(,1)=f1F(\cdot,1)=f_{1} and tF(x,t)t\mapsto F(x,t) is the unique shortest geodesic from f0(x)f_{0}(x) to f1(x)f_{1}(x). We denote by

𝒫t1,t2=𝒫t1,t2(x):T1,0N|F(x,t1)T1,0N|F(x,t2)\mathcal{P}_{t_{1},t_{2}}=\mathcal{P}_{t_{1},t_{2}}(x):T_{1,0}N|_{F(x,t_{1})}\to T_{1,0}N|_{F(x,t_{2})}

the parallel transport in FT1,0NF^{*}T_{1,0}N with respect to FT1,0N\nabla^{F^{*}T_{1,0}N} (pullback of the connection on T1,0NT_{1,0}N) along the curve γx(t):=(x,t)\gamma_{x}(t):=(x,t) from γx(t1)\gamma_{x}(t_{1}) to γx(t2)\gamma_{x}(t_{2}), xMx\in M, t1,t2[0,1]t_{1},t_{2}\in[0,1]. In particular, 𝒫0,1=Pvs,ut\mathcal{P}_{0,1}=P^{v_{s},u_{t}}. Let ψΓC1(ΣM(f0)T1,0N)\psi\in\Gamma_{C^{1}}(\Sigma M\otimes(f_{0})^{*}T_{1,0}N). We have

(5.7) ((𝒫0,1)1f1𝒫0,1f0)ψ=(eαψi)(((𝒫0,1)1eαf1T1,0N𝒫0,1eαf0T1,0N)(bif0))\begin{split}&((\mathcal{P}_{0,1})^{-1}\not{D}^{f_{1}}\mathcal{P}_{0,1}-\not{D}^{f_{0}})\psi\\ &=(e_{\alpha}\cdot\psi^{i})\otimes(((\mathcal{P}_{0,1})^{-1}\nabla_{e_{\alpha}}^{f_{1}^{*}T_{1,0}N}\mathcal{P}_{0,1}-\nabla_{e_{\alpha}}^{f_{0}^{*}T_{1,0}N})(b_{i}\circ{f_{0}}))\end{split}

where ψ=ψi(bif0)\psi=\psi^{i}\otimes(b_{i}\circ{f_{0}}), {bi}\{b_{i}\} is an orthonormal frame of T1,0NT_{1,0}N, ψi\psi^{i} are local C1C^{1} sections of ΣM\Sigma M, and {eα}\{e_{\alpha}\} is an orthonormal frame of TMTM.

We define local C1C^{1} sections Θi\Theta_{i} of FT1,0NF^{*}T_{1,0}N by

Θi(x,t):=𝒫0,t(x)(bif0)(x).\Theta_{i}(x,t):=\mathcal{P}_{0,t}(x)(b_{i}\circ{f_{0}})(x).

For each t[0,1]t\in[0,1] we define the functions Tij(,t):=Tijα(,t)T_{ij}(\cdot,t):=T_{ij}^{\alpha}(\cdot,t) by

(5.8) (𝒫0,t)1((eαFT1,0NΘi)(x,t))=jTijα(x,t)(bjf0)(x).(\mathcal{P}_{0,t})^{-1}((\nabla_{e_{\alpha}}^{F^{*}T_{1,0}N}\Theta_{i})(x,t))=\sum_{j}T_{ij}^{\alpha}(x,t)(b_{j}\circ{f_{0}})(x).

So far, we only know that the TijT_{ij} are continuous. In the following, we will perform some formal calculations and justify them afterwards. By a straightforward computation, we have

(5.9) ((𝒫0,1)1eαf1T1,0N𝒫0,1eαf0T1,0N)(bif0)(x)2=(𝒫0,1)1((eαFT1,0NΘi)(x,1))(𝒫0,0)1((eαFT1,0NΘi)(x,0))2=jTij(x,1)(bjf0)(x)jTij(x,0)(bjf0)(x)2=j(Tij(x,1)Tij(x,0))2=j(01ddt|t=rTij(x,t)dr)2.\begin{split}&||((\mathcal{P}_{0,1})^{-1}\nabla_{e_{\alpha}}^{f_{1}^{*}T_{1,0}N}\mathcal{P}_{0,1}-\nabla_{e_{\alpha}}^{f_{0}^{*}T_{1,0}N})(b_{i}\circ{f_{0}})(x)||^{2}\\ &=||(\mathcal{P}_{0,1})^{-1}((\nabla_{e_{\alpha}}^{F^{*}T_{1,0}N}\Theta_{i})(x,1))-(\mathcal{P}_{0,0})^{-1}((\nabla_{e_{\alpha}}^{F^{*}T_{1,0}N}\Theta_{i})(x,0))||^{2}\\ &=||\sum_{j}T_{ij}(x,1)(b_{j}\circ{f_{0}})(x)-\sum_{j}T_{ij}(x,0)(b_{j}\circ{f_{0}})(x)||^{2}\\ &=\sum_{j}(T_{ij}(x,1)-T_{ij}(x,0))^{2}\\ &=\sum_{j}\left(\int_{0}^{1}\frac{d}{dt}\bigg{|}_{t=r}T_{ij}(x,t)dr\right)^{2}.\end{split}

Therefore we want to control the first time-derivative of the TijT_{ij}. Equation (5.8) implies that these time-derivatives are related to the curvature of FT1,0NF^{*}T_{1,0}N. More precisely, for all XΓ(TM)X\in\Gamma(TM) we have

(5.10) ddt|t=r((𝒫0,t)1((XFT1,0NΘi)(x,t)))=ddt|t=0((𝒫0,t+r)1((XFT1,0NΘi)(x,t+r)))=ddt|t=0((𝒫0,r)1(𝒫r,r+t)1((XFT1,0NΘi)(x,t+r)))=(𝒫0,r)1ddt|t=0((𝒫r,r+t)1((XFT1,0NΘi)(x,t+r)))=(𝒫0,r)1((tFT1,0NXFT1,0NΘi)(x,r)).\begin{split}&\frac{d}{dt}\bigg{|}_{t=r}\left((\mathcal{P}_{0,t})^{-1}\left((\nabla_{X}^{F^{*}T_{1,0}N}\Theta_{i})(x,t)\right)\right)\\ &=\frac{d}{dt}\bigg{|}_{t=0}\left((\mathcal{P}_{0,t+r})^{-1}\left((\nabla_{X}^{F^{*}T_{1,0}N}\Theta_{i})(x,t+r)\right)\right)\\ &=\frac{d}{dt}\bigg{|}_{t=0}\left((\mathcal{P}_{0,r})^{-1}(\mathcal{P}_{r,r+t})^{-1}\left((\nabla_{X}^{F^{*}T_{1,0}N}\Theta_{i})(x,t+r)\right)\right)\\ &=(\mathcal{P}_{0,r})^{-1}\frac{d}{dt}\bigg{|}_{t=0}\left((\mathcal{P}_{r,r+t})^{-1}\left((\nabla_{X}^{F^{*}T_{1,0}N}\Theta_{i})(x,t+r)\right)\right)\\ &=(\mathcal{P}_{0,r})^{-1}\left((\nabla_{\frac{\partial}{\partial t}}^{F^{*}T_{1,0}N}\nabla_{X}^{F^{*}T_{1,0}N}\Theta_{i})(x,r)\right).\end{split}

Now, let us justify the formal calculations (5.9) and (5.10). Combining the definition of Θi\Theta_{i} as parallel transport and a careful examination of the regularity of F we deduce that (tFT1,0NXFT1,0NΘi)(x,r)(\nabla_{\frac{\partial}{\partial t}}^{F^{*}T_{1,0}N}\nabla_{X}^{F^{*}T_{1,0}N}\Theta_{i})(x,r) exists. Then (5.10) holds. Together with (5.8), we know that the TijT_{ij} are differentiable in tt. Therefore (5.9) also holds. We further get

tFT1,0NXFT1,0NΘi=RFT1,0N(t,X)Θi+XFT1,0NtFT1,0NΘi[t,X]FT1,0NΘi=RFT1,0N(t,X)Θi=RT1,0N(dF(t),dF(X))Θi,\begin{split}&\nabla_{\frac{\partial}{\partial t}}^{F^{*}T_{1,0}N}\nabla_{X}^{F^{*}T_{1,0}N}\Theta_{i}\\ &=R^{F^{*}T_{1,0}N}(\frac{\partial}{\partial t},X)\Theta_{i}+\nabla_{X}^{F^{*}T_{1,0}N}\nabla_{\frac{\partial}{\partial t}}^{F^{*}T_{1,0}N}\Theta_{i}-\nabla_{[\frac{\partial}{\partial t},X]}^{F^{*}T_{1,0}N}\Theta_{i}\\ &=R^{F^{*}T_{1,0}N}(\frac{\partial}{\partial t},X)\Theta_{i}=R^{T_{1,0}N}(dF(\frac{\partial}{\partial t}),dF(X))\Theta_{i},\end{split}

since tFT1,0NΘi=0\nabla_{\frac{\partial}{\partial t}}^{F^{*}T_{1,0}N}\Theta_{i}=0 by the definition of Θi\Theta_{i} and [t,X]=0[\frac{\partial}{\partial t},X]=0.

This implies

j(ddt|t=rTij(x,t))2=||ddt|t=r((𝒫0,t)1((eαFT1,0NΘi)(x,t)))||2=(tFT1,0NeαFT1,0NΘi)(x,r)2=RT1,0N(dF(x,r)(t),dF(x,r)(eα))Θi(x,r)2C1||dF(x,r)(t)||2||dF(x,r)(eα))||2,\begin{split}\sum_{j}\left(\frac{d}{dt}\bigg{|}_{t=r}T_{ij}(x,t)\right)^{2}&=||\frac{d}{dt}\bigg{|}_{t=r}\left((\mathcal{P}_{0,t})^{-1}((\nabla_{e_{\alpha}}^{F^{*}T_{1,0}N}\Theta_{i})(x,t))\right)||^{2}\\ &=||\left(\nabla_{\frac{\partial}{\partial t}}^{F^{*}T_{1,0}N}\nabla_{e_{\alpha}}^{F^{*}T_{1,0}N}\Theta_{i}\right)(x,r)||^{2}\\ &=||R^{T_{1,0}N}(dF_{(x,r)}(\frac{\partial}{\partial t}),dF_{(x,r)}(e_{\alpha}))\Theta_{i}(x,r)||^{2}\\ &\leq C_{1}||dF_{(x,r)}({\partial_{t}})||^{2}||dF_{(x,r)}(e_{\alpha}))||^{2},\end{split}

where C1C_{1} only depends on NN.

In the following we estimate dF(x,r)(t)||dF_{(x,r)}({\partial_{t}})|| and ||dF(x,r)(eα))||||dF_{(x,r)}(e_{\alpha}))||. We have

dF(x,r)(t|(x,r))=t|t=r(expf0(x)(texpf0(x)1f1(x)))=c(r),dF_{(x,r)}({\partial_{t}}|_{(x,r)})=\frac{\partial}{\partial t}\bigg{|}_{t=r}(\exp_{f_{0}(x)}(t\exp^{-1}_{f_{0}(x)}f_{1}(x)))=c^{\prime}(r),

where c(t):=expf0(x)(texpf0(x)1f1(x))c(t):=\exp_{f_{0}(x)}(t\exp^{-1}_{f_{0}(x)}f_{1}(x)) is a geodesic in NN. In particular, cc^{\prime} is parallel along cc and thus c(r)=c(0)=expf0(x)1f1(x)||c^{\prime}(r)||=||c^{\prime}(0)||=||\exp^{-1}_{f_{0}(x)}f_{1}(x)||. Therefore, we get

dF(x,r)(t)=expf0(x)1f1(x)dN(f0(x),f1(x))C2utvsC0(M,q),||dF_{(x,r)}({\partial_{t}})||=||\exp^{-1}_{f_{0}(x)}f_{1}(x)||\leq d^{N}(f_{0}(x),f_{1}(x))\leq C_{2}||u_{t}-v_{s}||_{C^{0}(M,\mathbb{R}^{q})},

where we have used Lemma 5.1 and the Lipschitz continuity of π\pi. Moreover, there exists C3(R)>0C_{3}(R)>0 such that ||dF(x,r)(eα))||C3(R)||dF_{(x,r)}(e_{\alpha}))||\leq C_{3}(R) for all (x,r)M×[0,1](x,r)\in M\times[0,1]. Thus, we have shown

j(ddt|t=rTij(x,t))2C1C22C3(R)2utvsC0(M,q)2\sum_{j}\left(\frac{d}{dt}\bigg{|}_{t=r}T_{ij}(x,t)\right)^{2}\leq C_{1}C_{2}^{2}C_{3}(R)^{2}||u_{t}-v_{s}||_{C^{0}(M,\mathbb{R}^{q})}^{2}

for all (x,t)(x,t). Combining this with (5.7) and (5.9), we complete the proof. ∎

The other one is for the parallel transport.

Lemma 5.3.

Choose ϵ\epsilon, δ\delta and RR as in (5.4). If ϵ>0\epsilon>0 is small enough, then there exists C=C(ϵ)>0C=C(\epsilon)>0 such that

Pvs,u0Put,vsPu0,utZZCutvsC0(M,q)Z||P^{v_{s},u_{0}}P^{u_{t},v_{s}}P^{u_{0},u_{t}}Z-Z||\leq C||u_{t}-v_{s}||_{C^{0}(M,\mathbb{R}^{q})}||Z||

for all ZT1,0N|u0(x)Z\in T_{1,0}N|_{u_{0}(x)}, u,vBRT(u¯0)u,v\in B^{T}_{R}(\bar{u}_{0}), xMx\in M and t,s[0,T]t,s\in[0,T].

Consequently, we also have

Lemma 5.4.

Choose ϵ\epsilon, δ\delta and RR as in (5.4). For u,vBRT(u¯0)u,v\in B^{T}_{R}(\bar{u}_{0}), s,t[0,T]s,t\in[0,T], the operator norm of the isomorphism of Banach spaces

Pvs,ut:ΓW1,p(ΣM(πvs)T1,0N)ΓW1,p(ΣM(πut)T1,0N)P^{v_{s},u_{t}}:\Gamma_{W^{1,p}}(\Sigma M\otimes(\pi\circ{v_{s}})^{*}T_{1,0}N)\to\Gamma_{W^{1,p}}(\Sigma M\otimes(\pi\circ{u_{t}})^{*}T_{1,0}N)

is uniformly bounded, i.e. there exists C=C(R,p)C=C(R,p) such that

Pvs,utL(W1,p,W1,p)C||P^{v_{s},u_{t}}||_{L(W^{1,p},W^{1,p})}\leq C

for all u,vBRT(u¯0)u,v\in B^{T}_{R}(\bar{u}_{0}), xMx\in M and t,s[0,T]t,s\in[0,T].

The proofs of these two lemmas only depend on the existence of the unique shortest geodesic between any two maps in BRT(u¯0)B^{T}_{R}(\bar{u}_{0}), which was already shown in (5.6). So we omit them here. Besides, by Lemma 5.2, one can immediately prove the following Lemma by the Min-Max principle.

Lemma 5.5.

Assume that dimker(1,0u0)=1{\rm dim}_{\mathbb{H}}{\rm ker}(\not{D}_{1,0}^{u_{0}})=1. Choose ϵ\epsilon, δ\delta and RR as in Lemma 5.2. If RR is small enough, then

dimker(1,0πut)=1{\rm dim}_{\mathbb{H}}{\rm ker}(\not{D}_{1,0}^{\pi\circ u_{t}})=1

and there exists Λ=12Λ(u0)\Lambda=\frac{1}{2}\Lambda(u_{0}) such that

#{spec(1,0πut)[Λ,Λ]}=1\#\{{\rm spec}(\not{D}_{1,0}^{\pi\circ u_{t}})\cap[-\Lambda,\Lambda]\}=1

for any uBRT(u¯0)u\in B^{T}_{R}(\bar{u}_{0}) and t[0,T]t\in[0,T], where Λ(u0)\Lambda(u_{0}) is a constant such that spec(1,0u0){0}(Λ(u0),Λ(u0)){\rm spec}(\not{D}_{1,0}^{u_{0}})\setminus\{0\}\subset\mathbb{R}\setminus(-\Lambda(u_{0}),\Lambda(u_{0})).

Once we have the minimality of the kernel in Lemma 5.5, we can prove the following uniform bounds for the resolvents, which are important for the Lipschitz continuity of the solution to the Dirac equation.

Lemma 5.6.

Assume we are in the situation of Lemma 5.5. We consider the resolvent R(λ,1,0πut):ΓL2ΓL2R(\lambda,\not{D}_{1,0}^{\pi\circ{u_{t}}}):\Gamma_{L^{2}}\to\Gamma_{L^{2}} of 1,0πut:ΓW1,2ΓL2\not{D}_{1,0}^{\pi\circ{u_{t}}}:\Gamma_{W^{1,2}}\to\Gamma_{L^{2}}. By the LpL^{p} estimate (see Lemma 3.3 in [4]), we know the restriction

R(λ,1,0πut):ΓLpΓW1,pR(\lambda,\not{D}_{1,0}^{\pi\circ{u_{t}}}):\Gamma_{L^{p}}\to\Gamma_{W^{1,p}}

is well-defined and bounded for any 2p<2\leq p<\infty. If R>0R>0 is small enough, then there exists C=C(p,R)>0C=C(p,R)>0 such that

sup|λ|=Λ2R(λ,1,0πut)L(Lp,W1,p)<C\sup_{|\lambda|=\frac{\Lambda}{2}}||R(\lambda,\not{D}_{1,0}^{\pi\circ{u_{t}}})||_{L(L^{p},W^{1,p})}<C

for any uBRT(u¯0)u\in B^{T}_{R}(\bar{u}_{0}), t[0,T]t\in[0,T].

Now, by the projector of the Dirac operator, we can construct a solution to the constraint equation whose nontriviality follows from the following lemma.

Lemma 5.7.

In the situation of Lemma 5.5, for any fixed uBRT(u¯0)u\in B^{T}_{R}(\bar{u}_{0}) and any ψ0ker(u0)\psi_{0}\in{\rm ker}(\not{D}^{u_{0}}) with ψ0L2=1\|\psi_{0}\|_{L^{2}}=1, we have

12ψ~1utL21,\sqrt{\frac{1}{2}}\leq\|\tilde{\psi}_{1}^{u_{t}}\|_{L^{2}}\leq 1,

where ψ~ut=Pu0,utψ0=ψ~1ut+ψ~2ut\tilde{\psi}^{u_{t}}=P^{u_{0},u_{t}}\psi_{0}=\tilde{\psi}_{1}^{u_{t}}+\tilde{\psi}_{2}^{u_{t}} with respect to the decomposition ΓL2=ker(1,0πut)(ker(1,0πut))\Gamma_{L^{2}}={\rm ker}(\not{D}_{1,0}^{\pi\circ u_{t}})\oplus({\rm ker}(\not{D}_{1,0}^{\pi\circ u_{t}}))^{\bot}

In Section 3, to show the short-time existence of the heat flow for α\alpha-Dirac-harmonic maps, we need the following Lipschitz estimate.

Lemma 5.8.

Choose δ\delta as in (5.4), ϵ\epsilon as in Lemma 5.2 and Lemma 5.3, RR as in Lemma 5.5 and Lemma 5.6. For any harmonic spinor ψ0ker(1,0u0)\psi_{0}\in{\rm ker}(\not{D}_{1,0}^{u_{0}}), we define

ψ¯(ut):=ψ~1ut=12πiγR(λ,1,0πut)Pu0,utψ0𝑑λ\bar{\psi}(u_{t}):=\tilde{\psi}^{u_{t}}_{1}=-\frac{1}{2\pi i}\int_{\gamma}R(\lambda,\not{D}_{1,0}^{\pi\circ u_{t}})P^{u_{0},u_{t}}\psi_{0}d\lambda

for any uBRT(u¯0)u\in B^{T}_{R}(\bar{u}_{0}), where γ\gamma is defined in the Section 22 with Λ=12Λ(u0)\Lambda=\frac{1}{2}\Lambda(u_{0}). In particular, ψ¯(ut)ker(1,0πut)ΓC0(ΣM(πut)T1,0N)\bar{\psi}(u_{t})\in{\rm ker}(\not{D}_{1,0}^{\pi\circ u_{t}})\subset\Gamma_{C^{0}}(\Sigma M\otimes(\pi\circ u_{t})^{*}T_{1,0}N). We write

ψ(ut):=ψ(u(,t))=ψ¯(ut)ψ¯(ut)L2.\psi(u_{t}):=\psi(u(\cdot,t))=\frac{\bar{\psi}(u_{t})}{\|\bar{\psi}(u_{t})\|_{L^{2}}}.

Let ψA(ut)\psi^{A}(u_{t}) be the sections of ΣM\Sigma M such that

ψ(ut)=ψA(ut)(Aπut)\psi(u_{t})=\psi^{A}(u_{t})\otimes(\partial_{A}\circ\pi\circ u_{t})

for A=1,,qA=1,\cdots,q. Then there exists C=C(R,ϵ,ψ0)>0C=C(R,\epsilon,\psi_{0})>0 such that

(5.11) Put,vsψ¯(ut)(x)ψ¯(ut)(x)CutvsC0(M,q)\|P^{u_{t},v_{s}}\bar{\psi}(u_{t})(x)-\bar{\psi}(u_{t})(x)\|\leq C\|u_{t}-v_{s}\|_{C^{0}(M,\mathbb{R}^{q})}

and

(5.12) ψA(ut)(x)ψA(vs)(x)CutvsC0(M,q)\|\psi^{A}(u_{t})(x)-\psi^{A}(v_{s})(x)\|\leq C\|u_{t}-v_{s}\|_{C^{0}(M,\mathbb{R}^{q})}

for all u,vBRT(u¯0)u,v\in B^{T}_{R}(\bar{u}_{0}), A=1,,qA=1,\cdots,q, xMx\in M and s,t[0,T]s,t\in[0,T].

Proof.

Using the following resolvent identity for two operators D1,D2D_{1},D_{2}

R(λ,D1)R(λ,D2)=R(λ,D1)(D1D2)R(λ,D2),R(\lambda,D_{1})-R(\lambda,D_{2})=R(\lambda,D_{1})\circ(D_{1}-D_{2})\circ R(\lambda,D_{2}),

we have

Put,vsψ¯(ut)ψ¯(vs)=12πiγR(λ,Put,vs1,0πut(Put,vs)1)(Put,vsPu0,utψ0Pu0,vsψ0)12πiγ(R(λ,Put,vs1,0πut(Put,vs)1)R(λ,1,0πvs))Pu0,vsψ0=12πiγR(λ,Put,vs1,0πut(Put,vs)1)(Put,vsPu0,utψ0Pu0,vsψ0)12πiγ(R(λ,Put,vs1,0πut(Put,vs)1)(Put,vs1,0πut(Put,vs)11,0πvs)R(λ,1,0πvs))Pu0,vsψ0,\begin{split}&P^{u_{t},v_{s}}\bar{\psi}(u_{t})-\bar{\psi}(v_{s})\\ &=-\frac{1}{2\pi i}\int_{\gamma}R(\lambda,P^{u_{t},v_{s}}\not{D}_{1,0}^{\pi\circ u_{t}}(P^{u_{t},v_{s}})^{-1})\bigg{(}P^{u_{t},v_{s}}P^{u_{0},u_{t}}\psi_{0}-P^{u_{0},v_{s}}\psi_{0}\bigg{)}\\ &\quad-\frac{1}{2\pi i}\int_{\gamma}\bigg{(}R(\lambda,P^{u_{t},v_{s}}\not{D}_{1,0}^{\pi\circ u_{t}}(P^{u_{t},v_{s}})^{-1})-R(\lambda,\not{D}_{1,0}^{\pi\circ v_{s}})\bigg{)}P^{u_{0},v_{s}}\psi_{0}\\ &=-\frac{1}{2\pi i}\int_{\gamma}R(\lambda,P^{u_{t},v_{s}}\not{D}_{1,0}^{\pi\circ u_{t}}(P^{u_{t},v_{s}})^{-1})\bigg{(}P^{u_{t},v_{s}}P^{u_{0},u_{t}}\psi_{0}-P^{u_{0},v_{s}}\psi_{0}\bigg{)}\\ &\begin{split}\quad-\frac{1}{2\pi i}\int_{\gamma}&\bigg{(}R(\lambda,P^{u_{t},v_{s}}\not{D}_{1,0}^{\pi\circ u_{t}}(P^{u_{t},v_{s}})^{-1})\circ\left(P^{u_{t},v_{s}}\not{D}_{1,0}^{\pi\circ u_{t}}(P^{u_{t},v_{s}})^{-1}-\not{D}_{1,0}^{\pi\circ v_{s}}\right)\circ\\ &\quad R(\lambda,\not{D}_{1,0}^{\pi\circ v_{s}})\bigg{)}P^{u_{0},v_{s}}\psi_{0},\end{split}\end{split}

where γ\gamma is defined in (2.11) with Λ=12Λ(u0)\Lambda=\frac{1}{2}\Lambda(u_{0}). Therefore, for pp large enough, we get

Put,vsψ¯(ut)(x)ψ¯(vs)(x)C1Put,vsψ¯utψ¯vsW1,p(M)C2γR(λ,Put,vsπut(Put,vs)1)(Put,vsPu0,utψ0Pu0,vsψ0)W1,p(M)+C2γ(R(λ,Put,vsπut(Put,vs)1)(Put,vsπut(Put,vs)1πvs)R(λ,πvs))Pu0,vsψ0W1,p(M)C2γR(λ,Put,vsπut(Put,vs)1)(Put,vsPu0,utψ0Pu0,vsψ0)W1,p(M)+C2γ(R(λ,Put,vsπut(Put,vs)1)(Put,vsπut(Put,vs)1πvs)R(λ,πvs))Pu0,vsψ0W1,p(M)C3supIm(γ)R(λ,Put,vsπut(Put,vs)1)L(Lp,W1,p)Put,vsPu0,utψ0Pu0,vsψ0Lp+C3supIm(γ)R(λ,Put,vsπut(Put,vs)1)L(Lp,W1,p)supIm(γ)R(λ,πvs)L(Lp,W1,p)Put,vsπut(Put,vs)1πvsL(W1,p,Lp)Pu0,vsψ0Lp.\begin{split}&||P^{u_{t},v_{s}}\bar{\psi}(u_{t})(x)-\bar{\psi}(v_{s})(x)||\leq C_{1}||P^{u_{t},v_{s}}\bar{\psi}^{u_{t}}-\bar{\psi}^{v_{s}}||_{W^{1,p}(M)}\\ &\leq C_{2}\bigg{\|}\int_{\gamma}R(\lambda,P^{u_{t},v_{s}}\not{D}^{\pi\circ u_{t}}(P^{u_{t},v_{s}})^{-1})\bigg{(}P^{u_{t},v_{s}}P^{u_{0},u_{t}}\psi_{0}-P^{u_{0},v_{s}}\psi_{0}\bigg{)}\bigg{\|}_{W^{1,p}(M)}\\ &\begin{split}+C_{2}\bigg{\|}\int_{\gamma}&\bigg{(}R(\lambda,P^{u_{t},v_{s}}\not{D}^{\pi\circ u_{t}}(P^{u_{t},v_{s}})^{-1})\circ\left(P^{u_{t},v_{s}}\not{D}^{\pi\circ u_{t}}(P^{u_{t},v_{s}})^{-1}-\not{D}^{\pi\circ v_{s}}\right)\circ\\ &\quad R(\lambda,\not{D}^{\pi\circ v_{s}})\bigg{)}P^{u_{0},v_{s}}\psi_{0}\bigg{\|}_{W^{1,p}(M)}\end{split}\\ &\leq C_{2}\int_{\gamma}\bigg{\|}R(\lambda,P^{u_{t},v_{s}}\not{D}^{\pi\circ u_{t}}(P^{u_{t},v_{s}})^{-1})\bigg{(}P^{u_{t},v_{s}}P^{u_{0},u_{t}}\psi_{0}-P^{u_{0},v_{s}}\psi_{0}\bigg{)}\bigg{\|}_{W^{1,p}(M)}\\ &\begin{split}+C_{2}\int_{\gamma}&\bigg{\|}\bigg{(}R(\lambda,P^{u_{t},v_{s}}\not{D}^{\pi\circ u_{t}}(P^{u_{t},v_{s}})^{-1})\circ\left(P^{u_{t},v_{s}}\not{D}^{\pi\circ u_{t}}(P^{u_{t},v_{s}})^{-1}-\not{D}^{\pi\circ v_{s}}\right)\circ\\ &\quad R(\lambda,\not{D}^{\pi\circ v_{s}})\bigg{)}P^{u_{0},v_{s}}\psi_{0}\bigg{\|}_{W^{1,p}(M)}\end{split}\\ &\leq C_{3}\sup\limits_{{\rm Im}(\gamma)}\|R(\lambda,P^{u_{t},v_{s}}\not{D}^{\pi\circ u_{t}}(P^{u_{t},v_{s}})^{-1})\|_{L(L^{p},W^{1,p})}\|P^{u_{t},v_{s}}P^{u_{0},u_{t}}\psi_{0}-P^{u_{0},v_{s}}\psi_{0}\|_{L^{p}}\\ &\quad+C_{3}\sup\limits_{{\rm Im}(\gamma)}\|R(\lambda,P^{u_{t},v_{s}}\not{D}^{\pi\circ u_{t}}(P^{u_{t},v_{s}})^{-1})\|_{L(L^{p},W^{1,p})}\sup\limits_{{\rm Im}(\gamma)}\|R(\lambda,\not{D}^{\pi\circ v_{s}})\|_{L(L^{p},W^{1,p})}\\ &\quad\quad\|P^{u_{t},v_{s}}\not{D}^{\pi\circ u_{t}}(P^{u_{t},v_{s}})^{-1}-\not{D}^{\pi\circ v_{s}}\|_{L(W^{1,p},L^{p})}\|P^{u_{0},v_{s}}\psi_{0}\|_{L^{p}}.\end{split}

Now, we estimate all the terms in the right-hand side of the inequality above. First, by Lemma 5.6 and Lemma 5.4, we know that all the resolvents above are uniformly bounded. Next, by Lemma 5.2, we have

Put,vsπut(Put,vs)1πvsL(W1,p,Lp)C(R)utvsC0(M,q).\|P^{u_{t},v_{s}}\not{D}^{\pi\circ u_{t}}(P^{u_{t},v_{s}})^{-1}-\not{D}^{\pi\circ v_{s}}\|_{L(W^{1,p},L^{p})}\leq C(R)\|u_{t}-v_{s}\|_{C^{0}(M,\mathbb{R}^{q})}.

Finally, by Lemma 5.3, we obtain

Put,vsPu0,utψ0Pu0,vsψ0LpC(ϵ,ψ0)utvsC0(M,q).\|P^{u_{t},v_{s}}P^{u_{0},u_{t}}\psi_{0}-P^{u_{0},v_{s}}\psi_{0}\|_{L^{p}}\leq C(\epsilon,\psi_{0})\|u_{t}-v_{s}\|_{C^{0}(M,\mathbb{R}^{q})}.

Putting these together, we get (5.11).

Next, we want to show the following estimate which is very close to (5.12).

(5.13) ψ¯A(ut)(x)ψ¯A(vs)(x)C(R,ϵ,ψ0)utvsC0(M,q).\|\bar{\psi}^{A}(u_{t})(x)-\bar{\psi}^{A}(v_{s})(x)\|\leq C(R,\epsilon,\psi_{0})\|u_{t}-v_{s}\|_{C^{0}(M,\mathbb{R}^{q})}.

In fact, we have

ψ¯A(ut)(x)ψ¯A(vs)(x)ψ¯(ut)(x)ψ¯(vs)(x)ΣxMqPut,vsψ¯(ut)(x)ψ¯(vs)(x)ΣxMq+Put,vsψ¯(ut)(x)ψ¯(ut)(x)ΣxMq=Put,vsψ¯(ut)(x)ψ¯(vs)(x)ΣxMT(πvs(x))N+Put,vsψ¯(ut)(x)ψ¯(ut)(x)ΣxMqC(R,ϵ,ψ0)utvsC0(M,q)+Put,vsψ¯(ut)(x)ψ¯(ut)(x)ΣxMq.\begin{split}&\|\bar{\psi}^{A}(u_{t})(x)-\bar{\psi}^{A}(v_{s})(x)\|\\ &\leq\|\bar{\psi}(u_{t})(x)-\bar{\psi}(v_{s})(x)\|_{\Sigma_{x}M\otimes\mathbb{R}^{q}}\\ &\leq\|P^{u_{t},v_{s}}\bar{\psi}(u_{t})(x)-\bar{\psi}(v_{s})(x)\|_{\Sigma_{x}M\otimes\mathbb{R}^{q}}+\|P^{u_{t},v_{s}}\bar{\psi}(u_{t})(x)-\bar{\psi}(u_{t})(x)\|_{\Sigma_{x}M\otimes\mathbb{R}^{q}}\\ &=\|P^{u_{t},v_{s}}\bar{\psi}(u_{t})(x)-\bar{\psi}(v_{s})(x)\|_{\Sigma_{x}M\otimes T_{(\pi\circ v_{s}(x))}N}\\ &\quad+\|P^{u_{t},v_{s}}\bar{\psi}(u_{t})(x)-\bar{\psi}(u_{t})(x)\|_{\Sigma_{x}M\otimes\mathbb{R}^{q}}\\ &\leq C(R,\epsilon,\psi_{0})\|u_{t}-v_{s}\|_{C^{0}(M,\mathbb{R}^{q})}+\|P^{u_{t},v_{s}}\bar{\psi}(u_{t})(x)-\bar{\psi}(u_{t})(x)\|_{\Sigma_{x}M\otimes\mathbb{R}^{q}}.\end{split}

It remains to estimate the last term in the inequality above. To that end, let γ(r):=exp(πut)(x)(rexp(πut)(x)1(πvs(x)))\gamma(r):=\exp_{(\pi\circ u_{t})(x)}(r\exp^{-1}_{(\pi\circ u_{t})(x)}(\pi\circ v_{s}(x))), r[0,1]r\in[0,1], be the unique shortest geodesic of NN from (πut)(x)(\pi\circ u_{t})(x) to (πvs)(x)(\pi\circ v_{s})(x). Let XTγ(0)NX\in T_{\gamma(0)}N be given and denote by X(r)X(r) the unique parallel vector field along γ\gamma with X(0)=XX(0)=X. Then we have

Put,vsXX=X(1)X(0)=01dXdr|r=ξdξ=01II(γ(r),X(r))𝑑r.P^{u_{t},v_{s}}X-X=X(1)-X(0)=\int_{0}^{1}\frac{dX}{dr}\bigg{|}_{r=\xi}d\xi=\int_{0}^{1}II(\gamma^{\prime}(r),X(r))dr.

Therefore,

Put,vsXXqC1supr[0,1]γ(r)Nsupr[0,1]X(r)N=C1γ(0)NXN\|P^{u_{t},v_{s}}X-X\|_{\mathbb{R}^{q}}\leq C_{1}\sup\limits_{r\in[0,1]}\|\gamma^{\prime}(r)\|_{N}\sup\limits_{r\in[0,1]}\|X(r)\|_{N}=C_{1}\|\gamma^{\prime}(0)\|_{N}\|X\|_{N}

where IIII is the second fundamental form of NN in q\mathbb{R}^{q} and C1C_{1} only depends on NN. Using (5.3) and the Lipschitz continuity of π\pi we get

γ(0)NdN((πut)(x),(πvs)(x))C2ut(x)vs(x)|q\|\gamma^{\prime}(0)\|_{N}\leq d^{N}((\pi\circ u_{t})(x),(\pi\circ v_{s})(x))\leq C_{2}\|u_{t}(x)-v_{s}(x)|\|_{\mathbb{R}^{q}}

and

Put,vsXXqC3ut(x)vs(x)|qXN.\|P^{u_{t},v_{s}}X-X\|_{\mathbb{R}^{q}}\leq C_{3}\|u_{t}(x)-v_{s}(x)|\|_{\mathbb{R}^{q}}\|X\|_{N}.

This implies

Put,vsψ¯(ut)(x)ψ¯(ut)(x)ΣxMqC(R,ϵ,ψ0)ut(x)vs(x)|q.\|P^{u_{t},v_{s}}\bar{\psi}(u_{t})(x)-\bar{\psi}(u_{t})(x)\|_{\Sigma_{x}M\otimes\mathbb{R}^{q}}\leq C(R,\epsilon,\psi_{0})\|u_{t}(x)-v_{s}(x)|\|_{\mathbb{R}^{q}}.

Hence, (5.13) holds.

Now, using (5.11) and (5.13), we get

ψA(ut)(x)ψA(vs)(x)=ψ¯A(ut)(x)ψ¯(ut)L2ψ¯A(ut)(x)ψ¯(vs)L2+ψ¯A(ut)(x)ψ¯(vs)L2ψ¯A(vs)(x)ψ¯(vs)L2ψ¯A(ut)(x)ψ¯(ut)L2ψ¯(vs)L2|ψ¯(vs)L2ψ¯(ut)L2|+1ψ¯(vs)L2ψ¯A(ut)(x)ψ¯A(vs)(x)=ψ¯A(ut)(x)ψ¯(ut)L2ψ¯(vs)L2|ψ¯(vs)L2Put,vsψ¯(ut)L2|+1ψ¯(vs)L2ψ¯A(ut)(x)ψ¯A(vs)(x)ψ¯A(ut)(x)ψ¯(ut)L2ψ¯(vs)L2Put,vsψ¯(ut)ψ¯(vs)L2+1ψ¯(vs)L2ψ¯A(ut)(x)ψ¯A(vs)(x)(ψ¯A(ut)(x)ψ¯(ut)L2ψ¯(vs)L2+1ψ¯(vs)L2)C(R,ϵ,ψ0)utvsC0(M,q).\begin{split}&\|\psi^{A}(u_{t})(x)-\psi^{A}(v_{s})(x)\|\\ &=\bigg{\|}\frac{\bar{\psi}^{A}(u_{t})(x)}{\|\bar{\psi}(u_{t})\|_{L^{2}}}-\frac{\bar{\psi}^{A}(u_{t})(x)}{\|\bar{\psi}(v_{s})\|_{L^{2}}}+\frac{\bar{\psi}^{A}(u_{t})(x)}{\|\bar{\psi}(v_{s})\|_{L^{2}}}-\frac{\bar{\psi}^{A}(v_{s})(x)}{\|\bar{\psi}(v_{s})\|_{L^{2}}}\bigg{\|}\\ &\leq\frac{\bar{\psi}^{A}(u_{t})(x)}{\|\bar{\psi}(u_{t})\|_{L^{2}}\|\bar{\psi}(v_{s})\|_{L^{2}}}\bigg{|}\|\bar{\psi}(v_{s})\|_{L^{2}}-\|\bar{\psi}(u_{t})\|_{L^{2}}\bigg{|}\\ &\quad+\frac{1}{\|\bar{\psi}(v_{s})\|_{L^{2}}}\|\bar{\psi}^{A}(u_{t})(x)-\bar{\psi}^{A}(v_{s})(x)\|\\ &=\frac{\bar{\psi}^{A}(u_{t})(x)}{\|\bar{\psi}(u_{t})\|_{L^{2}}\|\bar{\psi}(v_{s})\|_{L^{2}}}\bigg{|}\|\bar{\psi}(v_{s})\|_{L^{2}}-\|P^{u_{t},v_{s}}\bar{\psi}(u_{t})\|_{L^{2}}\bigg{|}\\ &\quad+\frac{1}{\|\bar{\psi}(v_{s})\|_{L^{2}}}\|\bar{\psi}^{A}(u_{t})(x)-\bar{\psi}^{A}(v_{s})(x)\|\\ &\leq\frac{\bar{\psi}^{A}(u_{t})(x)}{\|\bar{\psi}(u_{t})\|_{L^{2}}\|\bar{\psi}(v_{s})\|_{L^{2}}}\|P^{u_{t},v_{s}}\bar{\psi}(u_{t})-\bar{\psi}(v_{s})\|_{L^{2}}\\ &\quad+\frac{1}{\|\bar{\psi}(v_{s})\|_{L^{2}}}\|\bar{\psi}^{A}(u_{t})(x)-\bar{\psi}^{A}(v_{s})(x)\|\\ &\leq\bigg{(}\frac{\bar{\psi}^{A}(u_{t})(x)}{\|\bar{\psi}(u_{t})\|_{L^{2}}\|\bar{\psi}(v_{s})\|_{L^{2}}}+\frac{1}{\|\bar{\psi}(v_{s})\|_{L^{2}}}\bigg{)}C(R,\epsilon,\psi_{0})\|u_{t}-v_{s}\|_{C^{0}(M,\mathbb{R}^{q})}.\end{split}

Then the inequality (5.12) follows from Lemma 5.7 and (5.13). This completes the proof.

References

  • [1] Bernd Ammann and Nicolas Ginoux, Dirac-harmonic maps from index theory, Calculus of Variations and Partial Differential Equations 47 (2013), no. 3, 739–762.
  • [2] Qun Chen, Jürgen Jost, Jiayu Li, and Guofang Wang, Dirac-harmonic maps, Mathematische Zeitschrift 254 (2006), no. 2, 409–432.
  • [3] Qun Chen, Jürgen Jost, Linlin Sun, and Miaomiao Zhu, Dirac-harmonic maps between Riemann surfaces, Asian Journal of Mathematics 23 (2019), no. 1, 107–126.
  • [4] by same author, Estimates for solutions of Dirac equations and an application to a geometric elliptic-parabolic problem, Journal of the European Mathematical Society 21 (2019), no. 3, 665–707.
  • [5] Andreas Hermann, Dirac eigenspinors for generic metrics, arXiv preprint arXiv:1201.5771 (2012).
  • [6] Jürgen Jost, Geometry and physics, Springer Science & Business Media, 2009.
  • [7] Jürgen Jost, Riemannian geometry and geometric analysis, seventh ed., Universitext, Springer, 2017.
  • [8] Jürgen Jost, Lei Liu, and Miaomiao Zhu, A mixed elliptic-parabolic boundary value problem coupling a harmonic-like map with a nonlinear spinor, Journal für die reine und angewandte Mathematik (Crelles Journal) 2022 (2022), no. 785, 81–116.
  • [9] Jürgen Jost and Jingyong Zhu, α\alpha-Dirac-harmonic maps from closed surfaces, Calculus of variations and partial differential equations 60 (2021), no. 3.
  • [10] Jürgen Jost and Jingyong Zhu, Short-time existence of the α\alpha-Dirac-harmonic map flow and applications, Communications in Partial Differential Equations 46 (2021), no. 3, 442–469.
  • [11] H Blaine Lawson and Marie Louise Michelsohn, Spin geometry, Princeton Univ. Press, 1989.
  • [12] Gary M Lieberman, Second order parabolic differential equations, World scientific, 1996.
  • [13] John Nash, The imbedding problem for Riemannian manifolds, Annals of Mathematics (1956), 20–63.
  • [14] Jonathan Sacks and Karen Uhlenbeck, The existence of minimal immersions of 22-spheres, Annals of Mathematics (1981), 1–24.
  • [15] Wilhem Schlag, Schauder and LpL^{p} estimates for parabolic systems via Campanato spaces, Communications in Partial Differential Equations 21 (1996), no. 7-8, 1141–1175.
  • [16] Linlin Sun, A note on the uncoupled Dirac-harmonic maps from Kähler spin manifolds to Kähler manifolds, Manuscripta Mathematica 155 (2018), no. 1, 197–208.
  • [17] Ling Yang, A structure theorem of Dirac-harmonic maps between spheres, Calculus of Variations and Partial Differential Equations 35 (2009), no. 4, 409–420.