835
Views
6
CrossRef citations to date
0
Altmetric
Research Article

Short-time existence of the α-Dirac-harmonic map flow and applications

&
Pages 442-469 | Received 23 Dec 2019, Accepted 19 Oct 2020, Published online: 10 Dec 2020

Abstract

In this paper, we discuss the general existence theory of Dirac-harmonic maps from closed surfaces via the heat flow for α-Dirac-harmonic maps and blow-up analysis. More precisely, given any initial map along which the Dirac operator has nontrivial minimal kernel, we first prove the short time existence of the heat flow for α-Dirac-harmonic maps. The obstacle to the global existence is the singular time when the kernel of the Dirac operator no longer stays minimal along the flow. In this case, the kernel may not be continuous even if the map is smooth with respect to time. To overcome this issue, we use the analyticity of the target manifold to obtain the density of the maps along which the Dirac operator has minimal kernel in the homotopy class of the given initial map. Then, when we arrive at the singular time, this density allows us to pick another map which has lower energy to restart the flow. Thus, we get a flow which may not be continuous at a set of isolated points. Furthermore, with the help of small energy regularity and blow-up analysis, we finally get the existence of nontrivial α-Dirac-harmonic maps (α1) from closed surfaces. Moreover, if the target manifold does not admit any nontrivial harmonic sphere, then the map part stays in the same homotopy class as the given initial map.

2010 MATHEMATICS SUBJECT CLASSIFICATION:

1. Introduction

Motivated by the supersymmetric nonlinear sigma model from quantum field theory, see [Citation1], Dirac-harmonic maps from spin Riemann surfaces into Riemannian manifolds were introduced in [Citation2]. They are generalizations of the classical harmonic maps and harmonic spinors. From the variational point of view, they are critical points of a conformal invariant action functional whose Euler-Lagrange equations are a coupled elliptic system consisting of a second order equation and a Dirac equation.

It turns out that the existence of Dirac-harmonic maps from closed surfaces is a very difficult problem. Different from the Dirichlet problem, even if there is no bubble, the nontriviality of the limit is also an issue. Here, a solution is considered trivial if the spinor part ψ vanishes identically. So far, there are only a few results about Dirac-harmonic maps from closed surfaces, see [Citation3–5] for uncoupled Dirac-harmonic maps (here uncoupled means that the map part is harmonic; by an observation of Bernd Ammann and Johannes Wittmann, this is the typical case) based on index theory and the Riemann-Roch theorem, respectively. In an important contribution [Citation6], Wittmann investigated the heat flow introduced in [Citation7] and showed the short-time existence of this flow; for reasons that will become apparent below this is not as easy as for other parabolic systems. The problem has also been approached by linking and Morse-Floer theory. See [Citation8, Citation9] for one dimension and [Citation10] for the two dimensional case.

In critical point theory, the Palais-Smale condition is a very strong and useful tool. It fails, however, for many of the basic problems in geometric analysis, and in particular for the energy functional of harmonic maps from spheres [Citation11]. Therefore, it is not expected to be true for Dirac-harmonic maps. To overcome this problem for harmonic maps, Sacks-Uhlenbeck [Citation12] introduced the notion of α-harmonic maps where the integrand in the energy functional is raised to a power α>1. These α-harmonic maps then satisfy the Palais-Smale condition. However, when we analogously introduce α-Dirac-harmonic maps, the Palais-Smale condition fails due to the following existence result for uncoupled α-Dirac-harmonic maps, which directly follows from the proof of Theorem 4.1.

Theorem 1.1.

For a closed spin surface M and a closed manifold N, consider a homotopy class [ϕ] of maps ϕ:MmNn for which [dimH(ker/Dϕ)]Z2 is non-trivial. Assume that ϕ0[ϕ] is an α-harmonic map. Then there is a real vector space V of real dimension 4 such that all (ϕ0,ψ),ψV, are α-Dirac-harmonic maps.

To overcome this issue, in [Citation8, Citation9], the authors add an extra nonlinear term to the action functional of Dirac-geodesics. As for the two dimensional case [Citation10], we even cannot directly prove the Palais-Smale condition for the action functional of perturbed Dirac-harmonic maps into non-flat target manifolds. Instead, we are only able to prove it for perturbed α-Dirac-harmonic maps, and then approximate the α-Dirac-harmonic map by a sequence of perturbed α-Dirac-harmonic maps. However, in this approach, it is not easy to control the energies of the perturbed α-Dirac-harmonic maps, which are constructed by a Min-Max method over increasingly large domains in the configuration space.

Due to these two problems, in this paper, we would like to use the heat flow method to get the existence of Dirac-harmonic maps from closed surfaces to general manifolds where the harmonic map type equation is parabolized and the first order Dirac equation is carried along as an elliptic side constraint [Citation7]. As already mentioned, the short-time existence of the heat flow for Dirac-harmonic map was proved by Wittmann [Citation6]. He constructed the solution to the constraint Dirac equation by the projector of the Dirac operator along maps. By assuming that the Dirac operator along the initial map has nontrivial minimal kernel, he showed that the kernel would stay minimal for small time in the homotopy class of the initial map. This minimality implies a uniform bound for the resolvents and the Lipschitz continuity of the normalized Dirac kernel along the flow. This Lipschitz continuity makes the Banach fixed point theorem available. If one follows this approach, the first issue is how to deal with the kernel jumping problem. Observe that if the Dirac operators converge at the jumping time, the symmetry of the spectrum of Dirac operator guarantees that the limiting Dirac operator has odd dimensional kernel. Therefore, it is natural to try to extend Wittmann’s short time existence to the odd dimensional case. However, the eigenvalues in this case may split at time t = 0. Then the projector may not be continuous even if the Dirac operator is smooth with respect to time along the flow (see [Citation13]), which means that the Lipschitz continuity of the kernel is not available in general. To overcome this issue, we need the density mentioned in the abstract, which gives us a piecewise smooth flow.

As for the convergence, it is sufficient to control the energy of the spinor field because the energy of the map decreases along the flow. To do so, one can impose a restriction on the energy of the initial map as in [Citation14] and get the existence of Dirac-harmonic maps when the initial map has small energy. Alternatively, we use another type flow, that is, the heat flow for α-Dirac-harmonic maps (also called α-Dirac-harmonic map flow in the literatures). Our motivation comes from the successful application of this flow to the Dirichlet problem [Citation15]. Different from there, we cannot uniquely solve the constraint equation. Moreover, our equations of the flow are different. We never write the constraint equation in the Euclidean space Rq. Instead, we just solve it in the target manifold N. Last, our flow is not unique due to the absent of a boundary. Instead, only a weak uniqueness is available. Consequently, we need prove the fact that the flow takes value in the target manifold N in a different way. Eventually, we shall obtain the following results on the general existence of Dirac-harmonic maps.

Theorem 1.2.

Let M be a closed spin surface and (N, h) a real analytic closed manifold. Suppose there exists a map u0C2+μ(M,N) for some μ(0,1) such that dimHker/Du0=1. Then there exists a nontrivial smooth Dirac-harmonic map (Φ,Ψ) satisfying E(Φ)E(u0) and ||Ψ||L2=1.

Furthermore, if (N, h) does not admit any nontrivial harmonic sphere, then the map part Φ is in the same homotopy class as u0 and (Φ,Ψ) is coupled if the energy of the map is strictly bigger than the energy minimizer in the homotopy class [u0].

Remark 1.3.

Our result can at least keep the possibility of the existence of coupled solutions while other solutions produced in the literatures are all uncoupled. The analyticity of the target manifold is a sufficient condition which is used to get the density mentioned in the abstract. In fact, it is easy to see from the proof that we only need the density of the following set

(1.1) Y:={e(m0αi,+)|there exists at least one mapusuch thatdimHker/Du=1andEαi(u)=e}(1.1)

at the αi-energy minimizer m0αi in the homotopy class [u0] for a sequence αi1 as i.

In [Citation16], Wittmann discussed the density of those maps along which all the Dirac operators have minimal kernel. In particular, we have the following corollary.

Corollary 1.4.

Let M be a closed spin surface and (N, h) a real analytic closed manifold. We also assume that

  1. M is connected, oriented and of positive genus;

  2. N is connected. If N is even-dimensional, then we assume that it is non-orientable.

Then there exists a nontrivial smooth Dirac-harmonic map.

The rest of paper is organized as follows: In Section 2, we recall some definitions, notations and lemmas about Dirac-harmonic maps and the kernel of Dirac operator. In Section 3, under the minimality assumption on the kernel of the Dirac operator along the initial map, we prove the short time existence, weak uniqueness and regularity of the heat flow for α-Dirac-harmonic maps. In Section 4, we prove the existence of α-Dirac-harmonic maps and Theorem 1.2. 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) be a compact surface with a fixed spin structure. On the spinor bundle ΣM, we denote the Hermitian inner product by ·,·ΣM. For any XΓ(TM) and ξΓ(ΣM), the Clifford multiplication satisfies the following skew-adjointness: (2.1) X·ξ,ηΣM=ξ,X·ηΣM.(2.1)

Let be the Levi-Civita connection on (M, g). There is a connection (also denoted by ) on ΣM compatible with ·,·ΣM. Choosing a local orthonormal basis {eβ}β=1,2 on M, the usual Dirac operator is defined as /:=eβ·β, where β=1,2. Here and in the sequel, we use the Einstein summation convention. One can find more about spin geometry in [Citation17].

Let ϕ be a smooth map from M to another compact Riemannian manifold (N, h) of dimension n2. Let ϕ*TN be the pull-back bundle of TN by ϕ and consider the twisted bundle ΣMϕ*TN. On this bundle there is a metric ·,·ΣMϕ*TN induced from the metric on ΣM and ϕ*TN. Also, we have a connection ˜ on this twisted bundle naturally induced from those on ΣM and ϕ*TN. In local coordinates {yi}i=1,,n, the section ψ of ΣMϕ*TN is written as ψ=ψiyi(ϕ), where each ψi is a usual spinor on M. We also have the following local expression of ˜ ˜ψ=ψiyi(ϕ)+Γjki(ϕ)ϕjψkyi(ϕ), where Γjki are the Christoffel symbols of the Levi-Civita connection of N. The Dirac operator along the map ϕ is defined as (2.2) /D:=eα·˜eαψ=/ψiyi(ϕ)+Γjki(ϕ)eαϕj(eα·ψk)yi(ϕ),(2.2) which is self-adjoint [Citation11]. Sometimes, we use /Dϕ to distinguish the Dirac operators defined on different maps. In [Citation2], the authors introduced the functional (2.3) L(ϕ,ψ):=12M(|dϕ|2+ψ,/DψΣMϕ*TN)=12Mhij(ϕ)gαβϕixαϕjxβ+hij(ϕ)ψi,/DψjΣM.(2.3)

They computed the Euler-Lagrange equations of L: (2.4) τm(ϕ)12Rlijmψi,ϕl·ψjΣM=0,(2.4) (2.5) /Dψi=/ψi+Γjki(ϕ)eαϕj(eα·ψk)=0,(2.5) where τm(ϕ) is the m-th component of the tension field [Citation11] of the map ϕ with respect to the coordinates on N, ϕl·ψj denotes the Clifford multiplication of the vector field ϕl with the spinor ψj, and Rlijm stands for the component of the Riemann curvature tensor of the target manifold N. Denote R(ϕ,ψ):=12Rlijmψi,ϕl·ψjΣMym.

We can write (Equation2.4) and (Equation2.5) in the following global form: (2.6) {τ(ϕ)=R(ϕ,ψ),/Dψ=0,(2.6) and call the solutions (ϕ,ψ) Dirac-harmonic maps from M to N.

With the aim to get a general existence scheme for Dirac-harmonic maps, the following heat flow for Dirac-harmonic maps was introduced in [Citation7]: (2.7) {tu=τ(u)R(u,ψ), on (0,T)×M,/Duψ=0, on [0,T]×M.(2.7)

When M has boundary, the short time existence and uniqueness of (Equation2.8)–(Equation2.9) was also shown in [Citation7]. Furthermore, the existence of a global weak solution to this flow in dimension two under some boundary-initial constraint was obtained in [Citation14]. In [Citation15], to remove the restriction on the initial maps, the authors refined an estimate about the spinor in [Citation7] as follows:

Lemma 2.1.

[Citation15] Let M be a compact spin Riemann surface with boundary M, N be a compact Riemann manifold. Let uW1,2α(M,N) for some α>1 and ψW1,p(M,ΣMu*TN) for 1<p<2, then there exists a positive constant C=C(p,M,N,||u||L2α) such that

(2.8) ||ψ||W1,p(M)C(||/Dψ||Lp(M)+||Bψ||W11/p,p(M)).(2.8)

Motivated by this lemma, they considered the α-Dirac-harmonic flow and got the existence of Dirac-harmonic maps. For a closed manifold M, the situation is much more complicated because the kernel of the Dirac operator is a linear space. If the Dirac operator along the initial map has one dimensional kernel, Wittmann proved the short time existence on M whose dimension is m0,1,2,4(mod8).

By [Citation18], we can isometrically embed N into Rq. Then (Equation2.6)–(Equation2.7) is equivalent to following system: (2.9) {Δgu=II(du,du)+Re(P(S(du(eβ),eβ·ψ);ψ)),/ψ=S(du(eβ),eβ·ψ),(2.9) where II is the second fundamental form of N in Rq, and (2.10) S(du(eβ),eβ·ψ):=(uA·ψB)II(zA,zB),(2.10) (2.11) Re(P(S(du(eβ),eβ·ψ);ψ)):=P(S(zC,zB);zA)Re(ψA,duC·ψB).(2.11)

Here P(ξ;·) denotes the shape operator, defined by P(ξ;X),Y=A(X,Y),ξ for X,YΓ(TN) and Re(z) denotes the real part of zC. Together with the nearest point projection: (2.12) π:NδN,(2.12) where Nδ:={zRq|d(z,N)δ}, we can rewrite the evolution Equationequation (Equation2.8) as an equation in Rq.

Lemma 2.2.

[Citation6, Citation7] A tuple (u,ψ), where u:[0,T]×MN and ψΓ(ΣMu*TN), is a solution of (Equation2.8) if and only if

(2.13) tuAΔuA=πBCA(u)uB,uCπBA(u)πBDC(u)πEFC(ψD,uE·ψF)(2.13)

on (0,T)×M, for A=1,,q. Here we denote the A-th component function of u:[0,T]×MNRq by uA:MR, write πBA(z) for the B-th partial derivative of the A-th component function of π:RqRq and the global sections ψAΓ(ΣM) are defined by ψ=ψA(A°u), where (A)A=1,,q is the standard basis of TRq. Moreover, and ·,· denote the gradient and the Riemannian metric on M, respectively.

For future reference, we define (2.14) F1A(u):=πBCA(u)uB,uC,(2.14) (2.15) F2A(u,ψ):=πBA(u)πBDC(u)πEFC(ψD,uE·ψF).(2.15)

Note that for uC1(M,N) and ψΓ(ΣMu*TN) we have (2.16) II(dup(eα),dup(eα)))=F1A(u)|pA|u(p),(2.16) (2.17) R(ϕ,ψ)|p=F2A(u,ψ)|pA|u(p)(2.17) for all pM, where {eα} is an orthonormal basis of TpM.

Next, for every T > 0, we denote by XT the Banach space of bounded maps: (2.18) XT:=B([0,T];C1(M,Rq)),(2.18) (2.19) ||u||XT:=maxA=1,,qsupt[0,T](||uA(t,·)||C0(M)+||uA(t,·)||C0(M)).(2.19)

For any map vXT, the closed ball with center v and radius R in XT is defined by (2.20) BRT(v):={uXT|||uv||R}.(2.20)

We denote by Put,vs=Put,vs(x) the parallel transport of N along the unique shortest geodesic from π(u(x,t)) to π(v(x,s)). We also denote by Put,vs the inducing mappings (2.21) (π°ut)*TN(π°vs)*TN,(2.21) (2.22) ΣM(π°ut)*TNΣM(π°vs)*TN(2.22) and (2.23) ΓC1(ΣM(π°ut)*TN)ΓC1(ΣM(π°vs)*TN).(2.23)

Now, let us define (2.24) Λ(ut)=sup{Λ˜|spec(/Dπ°ut){0}R(Λ˜(ut),Λ˜(ut))}(2.24) and γt(x):[0,2π]C as (2.25) γt(x):=Λ(ut)2eix.(2.25)

In general, we also denote by γ the curve γ(x):[0,2π]C as (2.26) γ(x):=Λ2eix(2.26) for some constant Λ to be determined. Then the orthogonal projection onto ker(/Dπ°ut), which is the mapping (2.27) ΓL2(ΣM(π°ut)*TN)ΓL2(ΣM(π°ut)*TN),(2.27) can be written by the resolvent by (2.28) s12πiγtR(λ,/Dπ°ut)sdλ,(2.28) where R(λ,/Dπ°ut):ΓL2ΓL2 is the resolvent of /Dπ°ut:ΓW1,2ΓL2.

Finally, the following density lemma is very useful for us to extend the flow beyond the singular time.

Lemma 2.3.

[Citation16] Let M be a closed spin surface and (N, h) a real analytic closed manifold. Suppose there exists a map u0C2+μ(M,N) for some μ(0,1) such that dimHker/Du0=1. Then the kernel of /Du is minimal for generic u[u0], i.e., for a C-dense and C1-open subset of [u0].

3. The heat flow for α-Dirac-harmonic maps

In this section, we will prove the short-time existence of the heat flow for α-Dirac-harmonic maps. Since we are working on a closed surface M, we cannot uniquely solve the Dirac equation in the following system: (3.1) {tu=1(1+|u|2)α1(τα(u)1αR(u,ψ)),/Duψ=0.(3.1)

The short time existence and its extension are the obstacles. This system (if it converges) leads to a α-Dirac-harmonic map which is a solution of the system (3.2) τα(u):=τ((1+|du|2)α)=1αR(u,ψ)/Duψ=0.(3.2) and equivalently a critical point of functional (3.3) Lα(u,ψ)=12M(1+|du|2)α+12Mψ,/DuψΣMϕ*TN,(3.3) where τ is the tension field.

3.1. Short time existence

As in Section 2, we now embed N into Rq. Let u:MN with u=(uA) and denote the spinor along the map u by ψ=ψA(A°u), where ψA are spinors over M. For any smooth map ηC0(M,Rq) and any smooth spinor field ξC0(ΣMRq), we consider the variation (3.4) ut=π(u+tη),ψtA=πBA(ut)(ψB+tξB),(3.4) where π is the nearest point projection as in Section 2. Then we have

Lemma 3.1.

The Euler-Lagrange equations for Lα are

(3.5) ΔuA=2(α1)βγ2uBβuBγuA1+|u|2+πBCA(u)uB,uC+πBA(u)πBDC(u)πEFC(u)ψD,uE·ψFα(1+|u|2)α1(3.5)

and (3.6) /ψA=πBCA(u)uB·ψC.(3.6)

Proof.

Suppose (u,ψ) is a critical point of Lα, then for the variation (Equation3.5) we have (3.7) dLα(ut,ψt)dt|t=0=αM(1+|u|2)α1uA,πBAηB+πBCAuCηB+M/ψA,πBAξB+πBCAπDCψBηD,=:I+II.(3.7)

Then the lemma directly follows from the following computations. I=αM(1+|u|2)α1uA,ηA+αM(1+|u|2)α1πBCAuB,uCηA=αM(1+|u|2)α1ΔuAηAα(α1)M(1+|u|2)α2|u|2,uAηA=αM(1+|u|2)α1(ΔuA+2(α1)βγ2uBβuBγuA1+|u|2πBCA(u)uB,uC)ηA.II=M/ψAπBCAuB·ψC,ξA+MπBAπBDCψD,/ψCηA=M/ψAπBCAuB·ψC,ξA+MπBAπBDCψD,/ψCπEFCuE·ψFηA+MπBAπBDCψD,πEFCuE·ψFηA.

Lemma 3.1 implies that (Equation3.1)–(Equation3.2) is equivalent to (3.8) {tuA=ΔuA+2(α1)βγ2uBβuBγuA1+|u|2πBCA(u)uB,uCπBA(u)πBDC(u)πEFC(u)ψD,uE·ψFα(1+|u|2)α1/Dπ°uψ=0,(3.8)

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

Theorem 3.2.

Let M be a closed surface, and N a closed n-dimensional Riemannian manifold. Let u0C2+μ(M,N) for some 0<μ<1 with dimHker(/Du0)=1 and ψ0ker(/Du0) with ||ψ0||L2=1. Then there exists ϵ1=ϵ1(M,N)>0 such that, for any α(1,1+ϵ1), the problem (Equation3.1)–(Equation3.2) has a solution (u,ψ) with

(3.9) {||ψt||L2=1,t[0,T],u|t=0=u0,ψ|t=0=ψ0.(3.9)

satisfying (3.10) uC2+μ,1+μ/2(M×[0,T],N)(3.10) and (3.11) ψCμ,μ/2(M×[0,T],ΣMu*TN)L([0,T];C1+μ(M)).(3.11) for some T > 0.

Proof.

  • Step 1: Solving (Equation3.9)–(Equation3.10) in Rq.

In this step, we want to find a solution u:M×[0,T]Rq and ψt:MΣM(π°ut)*TN of (Equation3.9)–(Equation3.10) with the initial values (Equation3.11). We first give a solution to (Equation3.10) in a neighborhood of u0. For any T > 0, we can choose ϵ, δ and R as in the Appendix such that (3.12) u(x,t)Nδ(3.12) and (3.13) dN((π°u)(x,t),(π°v)(x,s))<ϵ<12inj(N)(3.13) for all u,vBRT:=BRT(u¯0)={uXT|||uu¯0||XTR}{u|t=0=u0},xM and t,s[0,T], where u¯0(x,t)=u0(x) for any t[0,T]. If R is small enough, then by Lemma 5.5, we have (3.14) dimKker(/Dπ°ut)=1(3.14) and there exists Λ=12Λ(u0) such that (3.15) #{spec(/Dπ°ut)[Λ,Λ]}=1(3.15) for any uBRT and t[0,T], where Λ(u0) is a constant such that spec(/Du0){0}R[Λ(u0),Λ(u0)]. Furthermore, for ψ0ker(/Du0) with ||ψ0||L2=1, Lemma 5.7 implies that (3.16) 34||ψ˜1ut||L21(3.16) for any uBR1T and t[0,T], where ψ˜ut=Pu0,utψ=ψ˜1ut+ψ˜2ut with respect to the decomposition ΓL2=ker(/Dπ°ut)(ker(/Dπ°ut)) and R1=R1(R,ϵ,u0)>0.

Now, for any T > 0 and κ>0, we define VκT:={vC1+μ,1+μ2(M×[0,T])|||v||C1+μ,1+μ2κ,v|M×{0}=0}.

Then, there exists κR1:=κ(R1)>0 such that (3.17) u0+vBR1T,vVκT,κκR1.(3.17)

Now, we denote κ0:=κR1 and VT:=Vκ0T.

For every vVT,u0+vBR1T, Lemma 5.8 gives us a solution ψ(v+u0) to the constraint equation. Since v+u0C1+μ(M), by Lp regularity [Citation6] and Schauder estimate [Citation7], we have (3.18) ||ψ(v+u0)||C1+μ(M)C(μ,M,N,κ0,||u0||C1+μ(M)).(3.18)

For any 0<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), that is, /Dπ°v(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), where # denotes a multi-linear map with smooth coefficients. For any λ(0,1), by the Sobolev embedding, Lp-regularity in [Citation6] and Lemma 5.8, we have (3.19) ||ψ(v+u0)(t)ψ(v+u0)(s)||Cλ(M)C(λ,M,N,κ0,||u0||C1(M))(||v(t)v(s)||L(M)+||dv(t)dv(s||L))C(λ,M,N,κ0,||u0||C1(M))|ts|μ/2.(3.19)

Therefore, (3.20) ||ψ(v+u0)||Cμ,μ/2(M)C(μ,M,N,κ0,||u0||C1(M)).(3.20)

Now, when α1 is sufficiently small, for the (v+u0,ψ(v+u0)) above, the standard theory of linear parabolic systems (see [Citation19]) implies that there exists a unique solution v1C2+μ,1+μ/2(M×[0,T],Rq) to the following Dirichlet problem: (3.21) {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)E·ψF(v+u0)α(1+|(v+u0)|2)α1,+Δgu0A+2(α1)βγ2u0Bβ(v+u0)Bγ(v+u0)A1+|(v+u0)|2,w(·,0)=0.(3.21) satisfying (3.22) ||v1||C2+μ,1+μ/2(M×[0,T])C(μ,M,N)(||v1||C0(M×[0,T])+||u0||C2+ν(M)+κ0).(3.22)

Since v1(·,0)=0, we have (3.23) ||v1||C0(M×[0,T])C(μ,M,N)T(||v1||C0(M×[0,T])+||u0||C2+ν(M)+κ0).(3.23)

By taking T > 0 small enough, we get (3.24) ||v1||C0(M×[0,T])C(μ,M,N)T(||u0||C2+ν(M)+κ0).(3.24)

Then the interpolation inequality in [Citation20] implies that v1VT for T > 0 sufficiently small. For such v1, we have ψ(v1+u0) satisfying (Equation3.20) and (Equation3.22). Replacing (v,ψ(v+u0)) in (Equation3.23)–(Equation3.24) by (v1,ψ(v1+u0)), then we get v2VT. Iterating this procedure, we get a solution vk+1 of (Equation3.23)–(Equation3.24) with (v,ψ(v+u0)) replacing by (vk,ψ(vk+u0)), which satisfies (3.25) ||ψ(vk+1+u0)||Cμ,μ/2(M)C(μ,M,N,κ0,||u0||C1(M)).(3.25) and (3.26) ||vk+1||C2+μ,1+μ/2(M×[0,T])C(μ,M,N)(||u0||C2+ν(M)+κ0).(3.26)

By passing to a subsequence, we know that vk converges to some u in C2,1(M×[0,T]) and ψvk+u0 converges to some ψ in C0(M×[0,T]). Then it is easy to see that (u,ψ) is a solution of (Equation3.9)–(Equation3.10) with u(·,0)=u0 and ψ(·,0)=ψ0.

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

Suppose uC2,1(M×[0,T],Rq) and ψCμ,μ/2(M×[0,T],ΣM(π°u)*TN)L([0,T];C1+μ(M)) satisfy (Equation3.9)–(Equation3.10). In the following, we write ||·|| and ·,· for the Euclidean norm and scalar product, respectively. Similarly, we write ||·||g and ·,·g for the norm and inner product of (M, g), respectively. We define (3.27) ρ:RqRq(3.27) by ρ(z)=zπ(z) and (3.28) φ:M×[0,T]R(3.28) by φ(x,t)=||ρ(u(x,t))||2=A=1q|ρA(u(x,t))|2. A direct computation yields (3.29) (tΔ)φ(x,t)=2A=1q||(ρA°u)(x,t)||g2+2ρ°u,πBA(u)F1B(u)+2α(1+|u|2)α1ρ°u,ρBA(u)F2B(u,ψ)+4(α1)1+|u|2ρ°u,βγ2uCβuCγuBρBA(u),(3.29) where F1A and F2A are defined in (Equation2.17) and (Equation2.18), respectively.

Since ρ°uTπ°uN and (dπ)u:RqTπ°uN, we have (3.30) ρ°u,πBA(u)F1B=ρ°u,ρBA(u)F2B=0.(3.30)

Together with (3.31) 4(α1)1+|u|2ρ°u,βγ2uCβuCγuBρBA(u)4(α1)||u||C2(M)||ρ°u||||(ρ°u)||2(α1)(||u||C2(M)2φ+||(ρ°u)||2),(3.31) we get (tΔ)φ(x,t)Cφ, where C=C(||u||C2,1(M×[0,T])). Since φ(x,t)0 and φ(x,0)=0 for any (x,t)M×[0,T], we conclude φ=0 on M×[0,T]. We have shown that u(x,t)N for all (x,t)M×[0,T].

Finally, by using the ϵ-regularity (see Lemma 3.7 below), we conclude that (3.32) uC2+μ,1+μ/2(M×[0,T],N)(3.32) and (3.33) ψCμ,μ/2(M×[0,T],ΣM(π°u)*TN)L([0,T];C1+μ(M)).(3.33)

Since the equations for α-Dirac-harmonic maps are invariant under multiplying the spinor by elements of H with unit norm, by uniqueness we always mean uniqueness up to multiplication of the spinor by such elements. This kind of uniqueness for the Dirac-harmonic map flow was proved by the Banach fixed point theorem in [Citation6]. However, we cannot apply the fixed point theorem to the α-Dirac-harmonic map flow. Therefore, it is interesting to consider the uniqueness of the α-Dirac-harmonic map flow from closed surfaces. By considering the evolution inequality of ||u1u2||C0(M), we can prove the following uniqueness which is weaker than that in [Citation6] because when the quaternions ha are different, we can no longer bound the C0-norm of the difference of the maps.

Theorem 3.3.

For any given T > 0, let (u1,ψ1) and (u2,ψ2) be two solutions to (Equation3.1)- with the constraint (Equation3.11) and u1,u2C2+μ,1+μ/2(M×[0,T],N). Then there exists a time T1>0, which depends on R and the C1+μ,1+μ2 norms of u1 and u2, such that u1,u2BRT1 and

(3.34) ψ1(x,t)=h1(t)ψ(ua(x,t)),ψ2(x,t)=h2(t)ψ(ua(x,t))(3.34)

for some h1(t),h1(t)H with unit length, where ψ(u(x,t)) is defined by (Equation5.36). Furthermore, if h1(t)=h2(t) on [0,T2] for some T2T1, then (u1,ψ1)(u2,ψ2) on M×[0,T2].

Proof.

By the assumptions, we have (3.35) ||ua(·,t)u0||C0(M)0,||ua(·,t)u0||C0(M)0(3.35) for a = 1, 2. Therefore, for small enough T1, u1,u2BRT2(u¯0). Since dimH(/Dua)=1 for a = 1, 2, there exist ha(t)H such that (3.36) ψa(x,t)=ψ(ua(x,t))ha(t)(3.36) for all t[0,T˜], where ψ(u(x,t)) is defined by (Equation5.36). Moreover, ha(t) is of unit length since ||ψa||L2(M)=||ψ(ua)||L2=1.

Now, let us consider the uniqueness of the flow. First, by subtracting the equations of u1 and u2 and multiplying by u1u2, we have (3.37) 12t|u1u2|212Δ|u1u2|2+|(u1u2)|2=2(α1)βγ2u1iβu1iγu11+|u1|2βγ2u2jβu2jγu21+|u2|2,u1u2II(u1,u1)II(u2,u2),u1u2R(ψ1,u1·ψ1)R(ψ2,u2·ψ2),u1u2.(3.37)

In the sequel, we will estimate the terms on the right-hand side of the inequality (Equation3.40). (3.38) 2(α1)βγ2u1iβu1iγu11+|u1|2βγ2u2jβu2jγu21+|u2|2,u1u2=2(α1)βγ2(u1iu2i)βu1iγu11+|u1|2,u1u2+2(α1)βγ2u2iβu1iγu1(11+|u1|211+|u2|2),u1u2+2(α1)βγ2u2iγu11+|u2|2(βu1iβu2i),u1u2+2(α1)βγ2u2iβu2i1+|u2|2(γu1γu2),u1u22(α1)βγ2(u1iu2i)βu1iγu11+|u1|2,u1u2+C(α1)|(u1u2)||u1u2|,(3.38) where we used u1,u2C2+μ,1+μ/2(M×[0,T],N). Similar, by the triangle inequality, we get (3.39) |II(u1,u1)II(u2,u2),u1u2|C|u1u2|2+C|(u1u2)||u1u2|(3.39) and (3.40) |R(ψ1,u1·ψ1)R(ψ2,u2·ψ2),u1u2|C|u1u2|2+C|(u1u2)||u1u2|+C|ψ1ψ2||u1u2|.(3.40)

Based on these estimates, (Equation3.40) becomes (3.41) 12t|u1u2|212Δ|u1u2|22(α1)βγ2(u1iu2i)βu1iγu11+|u1|2,u1u2|(u1u2)|2+C|u1u2|2+C|(u1u2)||u1u2|+C|ψ1ψ2||u1u2|.(3.41)

Next, we want to bound those terms in the right-hand side of (Equation3.44) by |u1u2|2 and |u1u2|2. Since u1,u2BRT2(u¯0), there is a unique geodesic between u1(x,t) and u2(x,t) for any (x,t)M×[0,T2]. Now, for any (x,t)P:={xM×[0,T2]|u1(x,t)u2(x,t)}, we define (3.42) us(x,t):=expu1(x,t)(sv(x,t))=expu1(x)(sV(x,t)/|V(x,t)|)(3.42) where s[0,|V(x,t)|],V(x,t):=expu1(x,t)1u2(x,t) and |V(x,t)| denotes the norm of V(x, t) in the tangent space Tu1(x,t)N. Then we can estimate 2(u1u2) as follows: (3.43) βγ2(u2u1)(x,t)=βγ2u|V(x,t)|(x,t)βγ2u0(x,t)=0|V(x,t)|ddsβγ2us(x,t)sup[0,|V(x,t)|]×P|dds2us|dN(u1(x,t),u2(x,t))C|u1(x,t)u2(x,t)|,(3.43) where we used the Lemma 5.1 in the Appendix. Hence, we can rewrite (Equation3.44) as (3.44) 12t|u1u2|212Δ|u1u2|22(α1)βγ2(u1iu2i)βu1iγu11+|u1|2,u1u2|(u1u2)|2+C|u1u2|2+C|(u1u2)||u1u2|+C|ψ1ψ2||u1u2|C|u1u2|2+C|ψ1ψ2||u1u2|,(3.44) where we used Young’s inequality. It remains to bound |ψ1ψ2| by |u1u2|. To that end, we use the Lemma 5.8 and (Equation3.39) as follows: (3.45) |ψ1ψ2|=|h1ψ(u1)h2ψ2(u2)|=|ψ(u1)ψ(u2)|||u1u2||C0(M),(3.45) where we used h1 = h2 in the second equality.

Last, it is easy to see (u1ψ1)(u2,ψ2) by considering the following evolution inequality (3.46) t||u1u2||C0(M)2C||u1u2||C0(M)2(3.46) with u1(·,0)=u2(·,0).

3.2. Regularity of the flow

In this subsection, we will give some estimates on the regularity of the flow. Let us start with the following estimate of the energy of the map part.

Lemma 3.4.

Suppose (u,ψ) is a solution of (Equation3.1)–(Equation3.2) with the initial values (Equation3.11). Then there holds (3.47) Eα(u(t))+2α0tM(1+|u|2)α1|tu|2=Eα(u0),(3.47) where Eα(u):=12M(1+|u|2)α. Moreover, Eα(u(t)) is absolutely continuous on [0,T] and non-increasing.

Proof. N

ote that (Equation3.1) can be written as: (3.48) (1+|u|2)α1tu=div((1+|u|2)α1u)(1+|gu|2)α1A(du,du)1αRe(P(A(du(eβ),eβ·ψ);ψ)).(3.48)

Multiplying the inequality above by tu and using (3.49) 0=0tMψ,ddt/Dψ=0tMψ,/D(tψ)+eγ·ψiRijkmtujduk(eγ))ym=0tMRmijkψm,uk·ψituj=0tM[S(ym,yj),S(yi,yk)RqS(ym,yk),S(yi,yj)Rq]ψm,uk·ψituj=20tMS(ym,yj),S(yi,yk)RqRe(ψm,uk·ψi)tuj=20tMRe(P(A(du(eβ),eβ·ψ);ψ)),tuj,(3.49) we get (3.50) 0tM(1+|u|2)α1|tu|2=0tMdiv((1+|u|2)α1u),tu=0tM(1+|gu|2)α1u,tu=12α0tddtM(1+|u|2)α,(3.50) which directly gives us the lemma. □

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

Lemma 3.5.

Suppose (u,ψ) is a solution of (Equation3.1)–(Equation3.2) with the initial values (Equation3.11). Then for any p(1,2), there holds (3.51) ||ψ(·,t)||W1,p(M)C,t[0,T],(3.51) where C=C(p,M,N,Eα(u0)).

Proof.

The lemma directly follows from Lemma 3.4 and the following lemma:

Lemma 3.6.

Let M be a closed spin Riemann surface, N be a compact Riemann manifold. Let uW1,2α(M,N) for some α>1 and ψW1,p(M,ΣMu*TN) for 1<p<2, then there exists a positive constant C=C(p,M,N,||u||L2α) such that (3.52) ||ψ||W1,p(M)C(||/Dψ||Lp(M)+||ψ||Lp(M)).(3.52)

This lemma follows from applying Lemma 2.1 to ηψ, where η is a cutoff function. □

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

Lemma 3.7.

Suppose (u,ψ) is a solution of (Equation3.1)–(Equation3.2) with the initial values (Equation3.11). Given ω0=(x0,t0)M×(0,T], denote (3.53) PR(ω0):=BR(x0)×[t0R2,t0].(3.53)

Then there exist three constants ϵ2=ϵ2(M,N)>0,ϵ3=ϵ3(M,N,u0)>0 and C=C(μ,R,M,N,Eα(u0))>0 such that if (3.54) 1<α<1+ϵ2,andsup[t04R2,t0]E(u(t);B2R(ω0))ϵ3,(3.54) then (3.55) R||ψ||L(PR(ω0))+R||u||L(PR(ω0))C(3.55) and for any 0<β<1, (3.56) sup[t0R24,t0]||ψ(t)||C1+μ(BR/2(x0))+||u||Cβ,β/2(PR/2(ω0))C(β).(3.56)

Moreover, if (3.57) supMsup[t04R2,t0]E(u(t);B2R(ω0))ϵ3,(3.57) then (3.58) ||u||C2+μ,1+μ/2(M×[t0R28,t0])+||ψ||Cμ,μ/2(M×[t0R28,t0])+sup[t0R28,t0]||ψ(t)||C1+μ(M)C.(3.58)

Since M is closed, x0 has to be an interior point of M. Therefore, our Lemma is just a special case of the Lemma 3.4 in [Citation15]. So we omit the proof here.

4. Existence of α-Dirac-harmonic maps

In this section, we will prove Theorem 1.2 by the following theorem on the existence of α-Dirac-harmonic maps for α>1.

Theorem 4.1.

Let M be a closed spin surface and (N, h) a real analytic closed manifold. Suppose there exists a map u0C2+μ(M,N) for some μ(0,1) such that dimHker/Du0=1. Then for any α(1,1+ϵ1), there exists a nontrivial smooth α-Dirac-harmonic map (uα,ψα) such that the map part uα stays in the same homotopy class as u0 and ||ψα||L2=1.

Proof

of Theorem 4.1. By Theorem 2.3 in [Citation21], all the following α-Dirac-harmonic maps are smooth. Let us denote the energy minimizer by

(4.1) m0α:=inf{Eα(u)|uW1,2α(M,N)[u0]},(4.1)

where [u0] denotes the homotopy class of u0. If u0 is a minimizing α-harmonic map, it follows from Lemma 3.4 that (u0,ψ0) is an α-Dirac-harmonic map for any ψ0ker/Du0. If Eα(u0)>m0α, then Theorem 3.2 gives us a solution (4.2) uC2+μ,1+μ/2(M×[0,T),N)(4.2) and (4.3) ψCμ,μ/2(M×[0,T),ΣMu*TN)0<s<TL([0,s];C1+μ(M)).(4.3) to the problem (Equation3.1)–(Equation3.2) with the initial values (Equation3.11).

By Lemma 3.4, we know (4.4) M(1+|u|2)αEα(u0).(4.4)

Then it is easy to see that, for any 0<ϵ<ϵ3, there exists a positive constant r0=r0(ϵ,α,Eα(u0)) such that for all (x,t)M×[0,T), there holds (4.5) Br0(x)|u|2CEα(u0)1/αr011αϵ.(4.5)

Therefore, by Theorem 3.2 and Lemma 3.7, we know that the singular time can be characterized as (4.6) Z={TR|limtiTdimHker/Duti>1}(4.6) and there exists a sequence {ti}T such that (4.7) (u(·,ti),ψ(·,ti))(u(·,T),ψ(·,T))inC2+μ(M)×C1+μ/2(M)(4.7) and (4.8) ||ψ(·,T)||L2=1.(4.8)

If Z=, then, by Theorem 3.2, we can extend the solution (u,ψ) beyond the time T by using (u(·,T),ψ(·,T)) as new initial values. Thus, we have the global existence of the flow. For the limit behavior as t, Lemma 3.4 implies that there exists a sequence {ti} such that (4.9) M|tu|2(·,ti)0.(4.9)

Together with Lemma 3.7, there is a subsequence, still denoted by {ti}, and an α-Dirac-harmonic map (uα,ψα) such that (u(·,ti),ψ(·,ti)) converges to (uα,ψα) in C2(M)×C1(M) and ||ψα||L2=1.

If Z and TZ, let us assume that Eα(u(·,T))>m0α and (u(·,T),ψ(·,T)) is not already an α-Dirac-harmonic map. We extend the flow as follows: By Lemma 2.3, there is a map u1C2+μ(M,N) such that (4.10) m0α<Eα(u1)<Eα(u(·,T))(4.10) and (4.11) dimHker/Du1=1.(4.11)

Thus, picking any ψ1ker/Du1 with ||ψ1||L2=1, we can restart the flow from the new initial values (u1,ψ1). If there is no singular time along the flow started from (u1,ψ1), then we get an α-Dirac-harmonic map as in the case of Z=. Otherwise, we use again the procedure above to choose (u2,ψ2) 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,, and {Ti}i=1,, such that (4.12) Eα(ui(t))+2α0tM(1+|u|2)α1|tu|2=Eα(ui),t(0,Ti),(4.12) where ui(·,0)=uiC2+μ(M,N). If the Ti are bounded away from zero, then there is {ti} such that (Equation4.9) hold for ti(0,Ti). Therefore, we have an α-Dirac-harmonic map as before. If Ti0, then we look at the limit of Eα(ui). If the limit is strictly bigger than m0α, we again choose another map satisfying (Equation4.10) and (Equation4.11) as a new starting point. If the limit is exactly m0α, then we choose {ti} such that ti(0,Ti) for each i. By Lemma 3.7, ui(ti) converges in C2(M)×C1(M) to a minimizing α-harmonic map uα. If /Duα has minimal kernel, then for any ψker/Duα,(uα,ψ) is an α-Dirac-harmonic map as we showed in the beginning of the proof. If /Duα has non-minimal kernel, we use the decomposition of the twisted spinor bundle through the Z2-grading Gid (see [Citation3]). More precisely, for any smooth variation (us)s(ϵ,ϵ) of u0, we split the bundle ΣMus*TN into ΣMus*TN=Σ+Mus*TNΣMus*TN, which is orthogonal in the complex sense and parallel. Consequently, for any ψ0ker/Du0, we have (4.13) (/Du0ψ0+,ψ0+)L2=(/Du0ψ0,ψ0)L2=0(4.13) for ψ0=ψ0++ψ0, where ψ0±=ψ±u0*TN and ψ±Σ±. Therefore, ψs±:=ψ±us*TN are smooth variations of ψ0±, respectively, such that (4.14) ddt|t=0(/Dusψs±,ψs±)L2=0.(4.14)

By taking u0=uα and ψ0=ψαker/Duα, the first variation formula of Lα implies that (uα,ψα±) are α-Dirac-harmonic maps (see Corollary 5.2 in [Citation3]). In particular, we can choose ψα such that ||ψα+||L2=1 or ||ψα||=1.

If it stops in finitely many steps, there exists a sequence {ti} and some 0<Tk+ such that (4.15) limtiT(u(·,ti),ψ(·,ti))(uα,ψα)inC2(M)×C1(M),(4.15) where (uα,ψα) either is an α-Dirac-harmonic map or satisfies Eα(uα)=m0α. And in the latter case, uα is a minimizing α-harmonic map. Then we can again get a nontrivial α-Dirac-harmonic map as above. □

By Theorem 4.1, for any α>1 sufficiently close to 1, there exists an α-Dirac-harmonic map (uα,ψα) with the properties (4.16) Eα(uα)Eα(u0),||ψα||L2=1(4.16) and (4.17) ||ψα||W1,p(M)C(p,M,N,Eα(u0))(4.17) for any 1<p<2. Then it is natural to consider the limit behavior when α decreases to 1. Since the blow-up analysis was already well studied in [Citation15], we can directly prove Theorem 1.2.

Proof

of Theorem 1.2. By Theorem 4.1, we have a sequence of smooth α-Dirac-harmonic maps (uαk,ψαk) with (Equation4.16) and (Equation4.17), where αk1 as k. Then, by Theorem 2.1 in [Citation15], there is a constant ϵ0>0 and a Dirac-harmonic map (Φ,Ψ)C(M,N)×C(M,ΣMΦ*TN) such that (4.18) (uαk,ψαk)(Φ,Ψ)inCloc2(MS)×Cloc1(MS),(4.18) where (4.19) S:={xM|liminfαk1E(uαk;Br(x))ϵ02,r>0}(4.19) is a finite set.

Now, taking x0S, there exists a sequence xαkx0,λαk0 and a nontrivial Dirac-harmonic map (ϕ,ξ):R2N such that (4.20) (uαk(xαk+λαkx),λαkαk1λαkψαk(xαk+λαkx))(ϕ,ξ)inCloc2(R2),(4.20) as α1. Choose any p*>4, by taking p=2p*2+p* in (Equation4.17), we get (4.21) ||ψαk||Lp*(M)C(p*,M,N,Eαk(u0))(4.21) and (4.22) ||ξ||L4(DR(0))=limαk1λαkαk1||ψαk||L4(DλαkR(xαk))limαk1C||ψαk||Lp*(M)(λαkR)2(141p*)=0.(4.22)

Thus, ξ = 0 and ϕ can be extended to a nontrivial smooth harmonic sphere. Since ||ψα||L2=1, the Sobolev embedding implies that ||Ψ||L2(M)=limαk1||ψα||L2(M)=1. Therefore, (Φ,Ψ) is nontrivial. Furthermore, if (N, h) does not admit any nontrivial harmonic sphere, then (4.23) (uαk,ψαk)(Φ,Ψ)inC2(M)×C1(M).(4.23)

Therefore, Φ is in the same homotopy class as u0. □

5. Appendix

In Section 3, we used some convenient properties of the elements in BRT(u¯0). Those properties were already discussed in [Citation6]. However, the function space used there is BRT(v0), where v0(x,t)=Mp(x,y,t)u0(y)dV(y), because the solution there is the unique fixed point of the following integral representation over BRT(v0) (5.1) Lu(x,t):=v0(x,t)+0tMp(x,y,tτ)(F1(uτ)+F2(uτ,ψ(uτ)))dV(y)dτ(5.1) where p is the heat kernel of M, F1 and F2 are defined as in (Equation2.17) and (Equation2.18), respectively. Our proof for the short-time existence is different from there, and the space BRT(u¯0) is more natural and convenient in our situation. Therefore, we cannot directly use the statement in [Citation6]. Although the space is changed, the proofs of those nice properties are parallel. In fact, one can see from the following that to make the elements in BRT(u¯0) satisfy nice properties (Equation5.11) and (Equation5.12), it is sufficient to choose R small, namely, T is independent of R. This is the biggest advantage. In the following, we will give the precise statement of the properties we need in Section 3 and proofs for the most important lemmas.

For every T > 0, we consider the space BRT(u¯0):={uXT|||uu¯0||XTR}{u|t=0=u0} where u¯0(x,t)=u0(x) for any t[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) and π°ut(x) in N. To do this, we need the following lemma which tells us that the distances in N can be locally controlled by the distances in Rq.

Lemma 5.1.

[Citation6] Let NRq be a closed embedded submanifold of Rq with the induced Riemannian metric. Denote by A its Weingarten map. Choose C > 0 such that ||A||C, where

(5.2) ||A||:=sup{||AvX|||vTpN,XTpN,||v||=1,||X||=1,pN}.(5.2)

Then there exists 0<δ0<1C such that for all 0<δδ0 and for all p,qN with ||pq||2<δ, it holds that (5.3) dN(p,q)11δC||pq||2,(5.3) where we denote the Euclidean norm by ||·||2 in this section.

In the following, we will choose δ and R to ensure the existence of the unique shortest geodesics between the projections of any two elements in BRT(u¯0). By the definition of BRT(u¯0), we have (5.4) ||u(x,t)u¯0(x,t)||2=||u(x,t)u0(x)||2R(5.4) for all (x,t)M×[0,T]. Then taking any Rδ, we get (5.5) d(u(x,t),N)||u(x,t)u0(x)||2δ(5.5) for all (x,t)M×[0,T]. Therefore, u(x,t)Nδ. In particular, π°u is N-valued, and (5.6) ||(π°u)(x,t)u0(x)||2||(π°u)(x,t)u(x,t)||2+||u(x,t)u0(x)||22δ.(5.6)

Now, we choose ϵ>0 with 2ϵ<inj(N) and δ such that (5.7) δ<min{14δ0,14ϵ(1δ0C)}(5.7) where δ0,C>0 are as in Lemma 5.1. From (Equation5.6), we know that for all u,vBRT(u¯0), it holds that (5.8) ||(π°u)(x,t)(π°v)(x,s)||24δ<δ0.(5.8)

Then Lemma 5.1 and (Equation5.7) imply that (5.9) dN((π°u)(x,t),(π°v)(x,s))11δ0C||(π°u)(x,t)(π°v)(x,s)||211δ0C4δ<ϵ<12inj(N).(5.9)

To summarize, under the choice of constants as follows: (5.10) {ϵ>0,s.t.2ϵ<inj(N),δ>0,s.t.δ<min{14δ0,14ϵ(1δ0C)},Rδ,(5.10) we have shown that (5.11) u(x,t)Nδ(5.11) and (5.12) dN((π°u)(x,t),(π°v)(x,s))<ϵ<12inj(N)(5.12) for all u,vBRT(u¯0),xM and t,s[0,T].

Using the properties (Equation5.11) and (Equation5.12), we can parallelly prove two important estimates as in [Citation6]. One is for the Dirac operators along maps.

Lemma 5.2.

Choose ϵ, δ and R as in (Equation5.10). If ϵ>0 is small enough, then there exists C=C(R)>0 such that (5.13) ||((Pvs,ut)1/Dπ°utPvs,ut/Dπ°vs)ψ(x)||C||utvs||C0(M,Rq)||ψ(x)||(5.13) for any u,vBRT(u¯0),ψΓC1(ΣM(π°vs)*TN),xM and t,s[0,T].

Proof.

We write f0:=π°vs,f1:=π°ut and define the C1 map F:M×[0,1]N by (5.14) F(x,t):=expf0(x)(texpf0(x)1f1(x))(5.14) where exp denotes the exponential map of the Riemannian manifold N. Note that F(·,0)=f0,F(·,1)=f1 and tF(x,t) is the unique shortest geodesic from f0(x) to f1(x). We denote by (5.15) Pt1,t2=Pt1,t2(x):TF(x,t1)NTF(x,t2)N(5.15) the parallel transport in F*TN with respect to F*TN (pullback of the Levi-Civita connection on N) along the curve γx(t):=(x,t) from γx(t1) to γx(t2),xM,t1,t2[0,1]. In particular, P0,1=Pvs,ut. Let ψΓC1(ΣM(f0)*TN). We have (5.16) ((P0,1)1/Df1P0,1/Df0)ψ=(eα·ψi)(((P0,1)1eαf1*TNP0,1eαf0*TN)(bi°f0))(5.16) where ψ=ψi(bi°f0),{bi} is an orthonormal frame of TN, ψi are local C1 sections of ΣM, and {eα} is an orthonormal frame of TM.

We define local C1 sections Θi of F*TN by (5.17) Θi(x,t):=P0,t(x)(bi°f0)(x).(5.17)

For each t[0,1] we define the functions Tij(·,t):=Tijα(·,t) by (5.18) (P0,t)1((eαF*TNΘi)(x,t))=jTijα(x,t)(bj°f0)(x).(5.18)

So far, we only know that the Tij are continuous. In the following, we will perform some formal calculations and justify them afterwards. By a straightforward computation, we have (5.19) ||((P0,1)1eαf1*TNP0,1eαf0*TN)(bi°f0)(x)||h2=||(P0,1)1((eαF*TNΘi)(x,1))(P0,0)1((eαF*TNΘi)(x,0))||h2=||jTij(x,1)(bj°f0)(x)jTij(x,0)(bj°f0)(x)||h2=j(Tij(x,1)Tij(x,0))2=j(01ddt|t=rTij(x,t)dr)2.(5.19)

Therefore we want to control the first time-derivative of the Tij. EquationEquation (Equation5.18) implies that these time-derivatives are related to the curvature of F*TN. More precisely, for all XΓ(TM) we have (5.20) ddt|t=r((P0,t)1((XF*TNΘi)(x,t)))=ddt|t=0((P0,t+r)1((XF*TNΘi)(x,t+r)))=ddt|t=0((P0,r)1(Pr,r+t)1((XF*TNΘi)(x,t+r)))=(P0,r)1ddt|t=0((Pr,r+t)1((XF*TNΘi)(x,t+r)))=(P0,r)1((tF*TNXF*TNΘi)(x,r)).(5.20)

Now, let us justify the formal calculations (Equation5.19) and (Equation5.20). Combining the definition of Θi as parallel transport and a careful examination of the regularity of F we deduce that (tF*TNXF*TNΘi)(x,r) exists. Then (Equation5.20) holds. Together with (Equation5.18), we know that the Tij are differentiable in t. Therefore (Equation5.19) also holds. We further get (5.21) tF*TNXF*TNΘi=RF*TN(t,X)Θi+XF*TNtF*TNΘi[t,X]F*TNΘi=RF*TN(t,X)Θi=RTN(dF(t),dF(X))Θi,(5.21) since tF*TNΘi=0 by the definition of Θi and [t,X]=0.

This implies (5.22) j(ddt|t=rTij(x,t))2=||ddt|t=r((P0,t)1((eαF*TNΘi)(x,t)))||h2=||(tF*TNeαF*TNΘi)(x,r)||h2=||RTN(dF(x,r)(t),dF(x,r)(eα))Θi(x,r)||h2C1||dF(x,r)(t)||h2||dF(x,r)(eα))||h2,(5.22) where C1 only depends on N.

In the following we estimate ||dF(x,r)(t)||h and ||dF(x,r)(eα))||h. We have (5.23) dF(x,r)(t|(x,r))=t|t=r(expf0(x)(texpf0(x)1f1(x)))=c(r),(5.23) where c(t):=expf0(x)(texpf0(x)1f1(x)) is a geodesic in N. In particular, c is parallel along c and thus ||c(r)||h=||c(0)||h=||expf0(x)1f1(x)||h. Therefore, we get (5.24) ||dF(x,r)(t)||h=||expf0(x)1f1(x)||hdN(f0(x),f1(x))C2||utvs||C0(M,Rq),(5.24) where we have used Lemma 5.1 and the Lipschitz continuity of π. Moreover, there exists C3(R)>0 such that ||dF(x,r)(eα))||hC3(R) for all (x,r)M×[0,1].

We have shown (5.25) j(ddt|t=rTij(x,t))2C1C22C3(R)2||utvs||C0(M,Rq)2(5.25) for all (x, t). Combining this with (Equation5.16) and (Equation5.19), we complete the proof. □

The other one is for the parallel transport.

Lemma 5.3.

Choose ϵ, δ and R as in (Equation5.10). If ϵ>0 is small enough, then there exists C=C(ϵ)>0 such that (5.26) ||Pvs,u0Put,vsPu0,utZZ||C||utvs||C0(M,Rq)||Z||(5.26) for all ZTu0(x)N,u,vBRT(u¯0),xM and t,s[0,T].

Consequently, we also have

Lemma 5.4.

Choose ϵ, δ and R as in (Equation5.10). For u,vBRT(u¯0),s,t[0,T], the operator norm of the isomorphism of Banach spaces (5.27) Pvs,ut:ΓW1,p(ΣM(π°vs)*TN)ΓW1,p(ΣM(π°ut)*TN)(5.27) is uniformly bounded, i.e. there exists C=C(R,p) such that (5.28) ||Pvs,ut||L(W1,p,W1,p)C(5.28) for all u,vBRT(u¯0),xM and t,s[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), which was already shown in (Equation5.12). Therefore, we omit the detailed proof here. Besides, by Lemma 5.2, one can immediately prove the following Lemma by the Min-Max principle as in [Citation6].

Lemma 5.5.

Assume that dimKker(/Du0)=2l1, where lN and (5.29) K={C,ifm=0,1(mod8),H,ifm=2,4(mod8).(5.29)

Choose ϵ, δ and R as in Lemma 5.2. If R is small enough, then (5.30) dimKker(/Dπ°ut)=1(5.30) and there exists Λ=12Λ(u0) such that (5.31) #{spec(/Dπ°ut)[Λ,Λ]}=1(5.31) for any uBRT(u¯0) and t[0,T], where Λ(u0) is a constant such that spec(/Du0){0}R(Λ(u0),Λ(u0)).

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,/Dπ°ut):ΓL2ΓL2 of /Dπ°ut:ΓW1,2ΓL2. By the Lp estimate (see Lemma 2.1 in [Citation6]), we know the restriction (5.32) R(λ,/Dπ°ut):ΓLpΓW1,p(5.32) is well-defined and bounded for any 2p<. If R > 0 is small enough, then there exists C=C(p,R)>0 such that (5.33) sup|λ|=Λ2||R(λ,@@Dπ°ut)||L(Lp,W1,p)<C(5.33) for any uBRT(u¯0),t[0,T].

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

Lemma 5.7.

In the situation of Lemma 5.5, for any fixed uBRT(u¯0) and any ψker(/Du0) with ||ψ||L2=1, we have (5.34) 12||ψ˜1ut||L21,(5.34) where ψ˜ut=Pu0,utψ=ψ˜1ut+ψ˜2ut with respect to the decomposition ΓL2=ker(/Dπ°ut)(ker(/Dπ°ut))

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

Lemma 5.8.

Choose δ as in (Equation5.10), ϵ as in Lemmas 5.2 and 5.3, R as in Lemmas 5.5 and 5.6. For any harmonic spinor ψker(/Du0), we define (5.35) ψ¯(ut):=ψ˜1ut=12πiγR(λ,/Dπ°ut)σ(ut)dλ(5.35) for any uBRT(u¯0), where γ is defined in the Section 2 with Λ=12Λ(u0). In particular, ψ¯(ut)ker(/Dπ°ut)ΓC0(ΣM(π°ut)*TN). We write (5.36) ψ(ut):=ψ(u(·,t))=ψ¯(ut)||ψ¯(ut)||L2.(5.36)

Let ψA(ut) be the sections of ΣM such that (5.37) ψ(ut)=ψA(ut)(A°π°ut)(5.37) for A=1,,q. Then there exists C=C(R,ϵ,ψ0)>0 such that (5.38) ||Put,vsψ¯(ut)(x)ψ¯(ut)(x)||C||utvs||C0(M,Rq)(5.38) and (5.39) ||ψA(ut)(x)ψA(vs)(x)||C||utvs||C0(M,Rq)(5.39) for all u,vBRT(u¯0),A=1,,q,xM and s,t[0,T].

Proof.

Using the following resolvent identity for two operators D1, D2 (5.40) R(λ,D1)R(λ,D2)=R(λ,D1)°(D1D2)°R(λ,D2),(5.40) we have (5.41) Put,vsψ¯(ut)ψ¯(ut)=12πi(γR(λ,Put,vs/Dπ°ut(Put,vs)1)Put,vsPu0,utψ0γR(λ,/Dπ°vs)Pu0,vsψ0)=12πiγR(λ,Put,vs/Dπ°ut(Put,vs)1)(Put,vsPu0,utψ0Pu0,vsψ0)12πiγ(R(λ,Put,vs/Dπ°ut(Put,vs)1)R(λ,/Dπ°vs))Pu0,vsψ0=12πiγR(λ,Put,vs/Dπ°ut(Put,vs)1)(Put,vsPu0,utψ0Pu0,vsψ0)12πiγ(R(λ,Put,vs/Dπ°ut(Put,vs)1)°(Put,vs/Dπ°ut(Put,vs)1/Dπ°vs)°R(λ,/Dπ°vs))Pu0,vsψ0,(5.41) where γ is defined in (2.29) with Λ=12Λ(u0). Therefore, for p large enough, we get ||Put,vsψ¯(ut)(x)ψ¯(ut)(x)||C1||Put,vsψ¯utψ¯vs||W1,p(M)C2||γR(λ,Put,vs/Dπ°ut(Put,vs)1)(Put,vsPu0,utψ0Pu0,vsψ0)||W1,p(M)+C2γ(R(λ,Put,vs/Dπ°ut(Put,vs)1)°(Put,vs/Dπ°ut(Put,vs)1/Dπ°vs)°R(λ,/Dπ°vs))Pu0,vsψ0W1,p(M)C2γ||R(λ,Put,vs/Dπ°ut(Put,vs)1)(Put,vsPu0,utψ0Pu0,vsψ0)||W1,p(M)+C2γ||(R(λ,Put,vsDπ°ut(Put,vs)1)°(Put,vs/Dπ°ut(Put,vs)1/Dπ°vs)°R(λ,Dπ°vs))Pu0,vsψ0||W1,p(M) (5.42) C3supIm(γ)||R(λ,Put,vs/Dπ°ut(Put,vs)1)||L(Lp,W1,p)||Put,vsPu0,utψ0Pu0,vsψ0||Lp+C3supIm(γ)||R(λ,Put,vs/Dπ°ut(Put,vs)1)||L(Lp,W1,p)supIm(γ)||R(λ,/Dπ°vs)||L(Lp,W1,p)||Put,vs/Dπ°ut(Put,vs)1/Dπ°vs||L(W1,p,Lp)||Pu0,vsψ0||Lp.(5.42)

Now, we estimate all the terms in the right-hand side of the inequality above. First, by Lemmas 5.6 and 5.4, we know that all the resolvents above are uniformly bounded. Next, by Lemma 5.2, we have (5.43) ||Put,vs/Dπ°ut(Put,vs)1/Dπ°vs||L(W1,p,Lp)C(R)||utvs||C0(M,Rq).(5.43)

Finally, by Lemma 5.3, we obtain (5.44) ||Put,vsPu0,utψ0Pu0,vsψ0||LpC(ϵ,ψ0)||utvs||C0(M,Rq).(5.44)

Putting these together, we get (Equation5.38).

Next, we want to show the following estimate which is very close to (Equation5.39). (5.45) ||ψ¯A(ut)(x)ψ¯A(vs)(x)||C(R,ϵ,ψ0)||utvs||C0(M,Rq).(5.45)

In fact, we have ||ψ¯A(ut)(x)ψ¯A(vs)(x)||||ψ¯(ut)(x)ψ¯(vs)(x)||ΣxMRq||Put,vsψ¯(ut)(x)ψ¯(vs)(x)||ΣxMRq+||Put,vsψ¯(ut)(x)ψ¯(ut)(x)||ΣxMRq=||Put,vsψ¯(ut)(x)ψ¯(vs)(x)||ΣxMT(π°vs(x))N+||Put,vsψ¯(ut)(x)ψ¯(ut)(x)||ΣxMRqC(R,ϵ,ψ0)||utvs||C0(M,Rq)+||Put,vsψ¯(ut)(x)ψ¯(ut)(x)||ΣxMRq.

It remains to estimate the last term in the inequality above. To that end, let γ(r):=exp(π°ut)(x)(rexp(π°ut)(x)1(π°ut(x))),r[0,1], be the unique shortest geodesic of N from (π°ut)(x) to (π°vs)(x). Let XTγ(0)N be given and denote by X(r) the unique parallel vector field along γ with X(0)=X. Then we have (5.46) Put,vsXX=X(1)X(0)=01dXdr|r=ξdξ=01II(γ(r),X(r))dr.(5.46)

Therefore, (5.47) ||Put,vsXX||RqC1supr[0,1]||γ(r)||Nsupr[0,1]||X(r)||N=C1||γ(0)||N||X||N(5.47) where II is the second fundamental form of N in Rq and C1 only depends on N. Using (Equation5.9) and the Lipschitz continuity of π we get (5.48) ||γ(0)||NdN((π°ut)(x),(π°vs)(x))C2||ut(x)vs(x)|||Rq(5.48) and (5.49) ||Put,vsXX||RqC3||ut(x)vs(x)|||Rq||X||N.(5.49)

This implies (5.50) ||Put,vsψ¯(ut)(x)ψ¯(ut)(x)||ΣxMRqC(R,ϵ,ψ0)||ut(x)vs(x)|||Rq.(5.50)

Hence, (Equation5.45) holds.

Now, using (Equation5.38) and (Equation5.45), 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)||L2||ψ¯A(ut)(x)||ψ¯(ut)||L2||ψ¯(vs)||L2|||ψ¯(vs)||L2||ψ¯(ut)||L2|+1||ψ¯(vs)||L2||ψ¯A(ut)(x)ψ¯A(vs)(x)||=ψ¯A(ut)(x)||ψ¯(ut)||L2||ψ¯(vs)||L2|||ψ¯(vs)||L2||Put,vsψ¯(ut)||L2|+1||ψ¯(vs)||L2||ψ¯A(ut)(x)ψ¯A(vs)(x)||ψ¯A(ut)(x)||ψ¯(ut)||L2||ψ¯(vs)||L2||Put,vsψ¯(ut)ψ¯(vs)||L2+1||ψ¯(vs)||L2||ψ¯A(ut)(x)ψ¯A(vs)(x)||(ψ¯A(ut)(x)||ψ¯(ut)||L2||ψ¯(vs)||L2+1||ψ¯(vs)||L2)C(R,ϵ,ψ0)||utvs||C0(M,Rq).

Then the inequality (Equation5.39) follows from Lemma 5.7 and (Equation5.45). This completes the proof. □

References

  • Jost, J. (2009). Geometry and Physics. Berlin, Heidelberg: Springer Science & Business Media.
  • Chen, Q., Jost, J., Li, J., Wang, G. (2006). Dirac-harmonic maps. Math. Z. 254(2):409–432. DOI: 10.1007/s00209-006-0961-7.
  • Ammann, B., Ginoux, N. (2013). Dirac-harmonic maps from index theory. Calc. Var. 47(3–4):739–762. DOI: 10.1007/s00526-012-0534-z.
  • Chen, Q., Jost, J., Sun, L., Zhu, M. (2015). Dirac-harmonic maps between Riemann surfaces. MPI MIS Preprint 76.
  • Yang, L. (2009). A structure theorem of Dirac-harmonic maps between spheres. Calc. Var. 35(4):409–420. DOI: 10.1007/s00526-008-0210-5.
  • Wittmann, J. (2017). Short time existence of the heat flow for Dirac-harmonic maps on closed manifolds. Calc. Var. 56(6):169. DOI: 10.1007/s00526-017-1270-1.
  • Chen, Q., Jost, J., Sun, L., Zhu, M. (2018). Estimates for solutions of Dirac equations and an application to a geometric elliptic-parabolic problem. J. Eur. Math. Soc. 21(3):665–707. DOI: 10.4171/JEMS/847.
  • Isobe, T. (2012). On the existence of nonlinear Dirac-geodesics on compact manifolds. Calc. Var. 43(1–2):83–121. DOI: 10.1007/s00526-011-0404-0.
  • Isobe, T., Maalaoui, A. (2017). Morse-Floer theory for super-quadratic Dirac-geodesics, arXiv preprint arXiv:1712.08960.
  • Jost, J., Zhu, J. (2019). α-Dirac-harmonic maps from closed surfaces, arXiv preprint arXiv:1903.07927.
  • Jost, J. (2017). Riemannian Geometry and Geometric Analysis. Berlin: Springer.
  • Sacks, J., Uhlenbeck, K. (1981). The existence of minimal immersions of 2-spheres. Ann. Math. 113(1):1–24. DOI: 10.2307/1971131.
  • Kato, T. (2013). Perturbation Theory for Linear Operators, Vol. 132. New York: Springer Science & Business Media.
  • Jost, J., Liu, L., Zhu, M. (2017). A global weak solution of the Dirac-harmonic map flow. Annales de L’Institut Henri Poincare (C) Non Linear Analysis. 34(7):1851–1882. DOI: 10.1016/j.anihpc.2017.01.002.
  • Jost, J., Liu, L., Zhu, M. (2018). Geometric analysis of a mixed elliptic-parabolic conformally invariant boundary value problem. MPI MIS Preprint 41.
  • Wittmann, J. (2019). Minimal kernels of Dirac operators along maps. Mathematische Nachrichten. 292(7):1–9.
  • Blaine, H., Jr, L., Michelsohn, M.-L. (1989). Spin Geometry, Volume 38 of Princeton Mathematical Series. Princeton, NJ: Princeton University Press.
  • Nash, J. (1956). The imbedding problem for Riemannian manifolds. Ann. Math. 63(1):20–63. DOI: 10.2307/1969989.
  • Schlag, W. (1996). Schauder and Lp estimates for parabolic systems via Campanato spaces. Commun. Partial Differ. Equations. 21(7–8):1141–1175.
  • Lieberman, G. M. (1996). Second Order Parabolic Differential Equations. Singapore: World Scientific.
  • Chen, Q., Jost, J., Li, J., Wang, G. (2005). Regularity theorems and energy identities for Dirac-harmonic maps. Math. Z. 251(1):61–84. DOI: 10.1007/s00209-005-0788-7.