Energy methods for Dirac-type equations in two-dimensional ...

1 downloads 0 Views 269KB Size Report
Jun 19, 2017 - in two-dimensional Minkowski space in the next subsection, we will focus ...... Faculty of Mathematics, Oskar-Morgenstern-Platz 1, 1090 Vienna, ...
ENERGY METHODS FOR DIRAC-TYPE EQUATIONS IN TWO-DIMENSIONAL MINKOWSKI SPACE VOLKER BRANDING

arXiv:1706.05971v1 [math.DG] 19 Jun 2017

Abstract. In this article we develop energy methods for a large class of linear and nonlinear Dirac-type equations in two-dimensional Minkowski space. We will derive existence results for several Dirac-type equations originating in quantum field theory, in particular for Dirac-wave maps to compact Riemannian manifolds.

1. Introduction and results In quantum field theory spinors are used to describe fermions, which are elementary particles of half-integer spin. The equations that govern their behavior are both linear and nonlinear Dirac equations. Linear Dirac equations are employed to model free fermions. However, to model the interaction of fermions one has to take into account nonlinearities. In mathematical terms spinors are sections in the spinor bundle, which is a vector bundle defined on the underlying manifold. Its existence requires the vanishing of the second Stiefel-Whitney class, which is a topological condition. The natural operator acting on spinors is the Dirac operator, which is a first-order differential operator. If the underlying manifold is Riemannian the Dirac operator is elliptic, if the manifold is Lorentzian the Dirac operator is hyperbolic. The fact that the Dirac operator is of first order usually leads to technical problems since there are less tools available compared to second order operators such as the Laplacian. In the case of a Riemannian manifold many results on the qualitative behavior of nonlinear Dirac equations have been obtained recently, see [22, 15, 17]. However, obtaining an existence result for nonlinear Dirac equations in the Riemannian case is rather complicated. This article is supposed to be the first step to develop energy methods for linear and nonlinear Dirac equations on Lorentzian manifolds in a geometric framework. For an introduction to linear geometric wave equations on globally hyperbolic manifolds we refer to [6]. As a starting point we will stick to the case of two-dimensional Minkowski space, the generalization to higher dimensional globally hyperbolic manifolds will be treated in a subsequent work. Most of the analytic results on nonlinear Dirac equation in Minkowski space make use of a global trivialization of the spinor bundle and investigate the resulting coupled system of partial differential equations of complex-valued functions. In our approach we do not make use of a global trivialization of the spinor bundle, but derive energy estimates for the spinor itself. This approach seems to be the natural one from a geometric point of view. This article is organized as follows: After presenting the necessary background on spin geometry in two-dimensional Minkowski space in the next subsection, we will focus on the analysis of linear Dirac equations in section 2. Afterwards, in section 3 we will consider several models from quantum field theory that involve nonlinear Dirac equations. Making use of the energy methods developed before we derive existence results for some of these models. The last section is devoted to the study of Dirac-wave maps from two-dimensional Minkowski space taking values in a compact Riemannian manifold. Again, by application of suitable energy methods, we are able to derive an existence result for the latter. Date: June 20, 2017. 2010 Mathematics Subject Classification. 35L02; 35L60; 58J45; 53C27. Key words and phrases. nonlinear Dirac equations; two-dimensional Minkowski space; energy methods; Diracwave maps. 1

2

VOLKER BRANDING

1.1. Spin geometry in two-dimensional Minkowski space. First, let us fix the notations that will be used throughout this article. In addition, we want to recall several facts on spin geometry in the setting of a Lorentzian manifold. In this article, we will make use of the Einstein summation convention, that is we sum over repeated indices. In the following we will consider two-dimensional Minkowski space R1,1 with metric (+, −) and global coordinates (t, x). The tangent vectors of R1,1 will be denoted by ∂t , ∂x . The spinor bundle over R1,1 will be denoted by ΣR1,1 and sections in this bundle will be called spinors. Note that ΣR1,1 can be globally trivialized, however we will not often make use of this fact. On ΣR1,1 there exists a metric connection and we have a hermitian, but indefinite, scalar product denoted by h·, ·i. We will use the convention that this scalar product is linear in the first and anti-linear in the second slot. We denote Clifford multiplication of a spinor ψ with a tangent vector X by X · ψ. Note that in contrast to the Riemannian case, Clifford multiplication is symmetric, that is hX · ξ, ψiΣR1,1 = hξ, X · ψiΣR1,1 for all ψ, ξ ∈ Γ(ΣR1,1 ) and all X ∈ T R1,1 . In addition, the Clifford relations X · Y · ψ + Y · X · ψ = −2g(X, Y )ψ hold for all X, Y ∈ T R1,1 , where g is the metric of R1,1 . The Dirac operator on two-dimensional Minkowski is defined as (with εj = g(ej , ej ) = ±1) D :=

2 X

εj ej · ∇ej = ∂t · ∇∂t − ∂x · ∇∂x .

j=1

Here, ∇ denotes the connection on ΣR1,1 and ei , i = 1, 2 is a pseudo-orthonormal basis of T R1,1 . Note that, in contrast to the Riemannian case, the Dirac operator defined above is not formally self-adjoint with respect to the L2 -norm but satisfies Z Z hDξ, ψidµ hξ, Dψidµ = − R1,1

R1,1

ΣR1,1 .

for all ψ, ξ ∈ For this reason, we will mostly consider the operator iD, since this combination is formally self-adjoint with respect to the L2 -norm. For many of the analytic questions discussed in this article it will be necessary to have a positive-definite scalar product on the spinor bundle in order to establish energy estimates. For this reason we consider the positive definite scalar product h∂t ·, i, where ∂t denotes the globally-defined timelike vector field. The resulting norm will be denoted by ||β , that is 0 ≤ |ψ|2β := h∂t · ψ, ψi for ψ ∈ Γ(ΣR1,1 ). For more details on spin geometry on Lorentzian manifolds we refer to [8] and also [5, 4, 7]. Remark 1.1. Note that we have two kinds of natural scalar products on the spinor bundle in the semi-Riemannian case. On the one hand we have the (geometric) scalar product that is invariant under the Spin group, but indefinite. On the other hand we have the (analytic) scalar product, which is positive definite but breaks the geometric invariance. 2. Linear Dirac equations in two-dimensional Minkowski space In this section we derive conserved energies for solutions of linear Dirac-type equations. Later on, we will generalize these methods to the non-linear case. We start be analyzing solutions of Dψ = 0.

(2.1)

ENERGY METHODS FOR DIRAC-TYPE EQUATIONS IN TWO-DIMENSIONAL MINKOWSKI SPACE

3

Note that, for a solution of (2.1), we have the following identities ∂ ∂ h∂t · ψ, ψi − h∂x · ψ, ψi =hDψ, ψi + hψ, Dψi = 0, (2.2) ∂t ∂x ∂ ∂ h∂x · ψ, ψi − h∂t · ψ, ψi =h(∂t · ∇∂x − ∂x · ∇∂t )ψ, ψi − hψ, (∂t · ∇∂x + ∂x · ∇∂t )ψi = 0. ∂t ∂x (2.3) We will use these identities to derive several conservation laws. Lemma 2.1. Let ψ ∈ Γ(ΣR1,1 ) be a solution of (2.1). Then the energies Z 1 E1 (t) = |ψ|2 dx, 2 R β Z  ∂ 1 |ψ|2 2 + ∂ |ψ|2 2 dx, E2 (t) = β β 2 R ∂x ∂t Z 1 E3 (t) = (|∇∂t ψ|2β + |∇∂x ψ|2β )dx 2 R are conserved. Proof. The first claim follows from integrating (2.2). Differentiating (2.2) with respect to t and (2.3) with respect to x we find ∂2 ∂2 ∂2 h∂ · ψ, ψi = h∂ · ψ, ψi = h∂t · ψ, ψi. t x ∂t2 ∂x∂t ∂x2 Hence, |ψ|2β solves the one-dimensional wave equation, which yields the second statement. The  third assertion follows since ψ solves ∇2∂t ψ = ∇2∂x ψ. Remark 2.2. Note that E1 (t) and E3 (t) are also conserved in higher-dimensional Minkowski space. We can use (2.2) and (2.3) to find conserved energies that involve higher Lp norms of ψ. Lemma 2.3. Let ψ ∈ Γ(ΣR1,1 ) be a solution of (2.1). Then the energy Z E4 (t) = (|ψ|4β + |h∂x · ψ, ψi|2 )dx R

is conserved. Proof. Making use of (2.2) and (2.3) we calculate Z Z d 1 ∂ 4 |ψ|β dx = |ψ|2β h∂t · ψ, ψidx dt 2 R ∂t ZR ∂ = |ψ|2β h∂x · ψ, ψidx ∂x R Z ∂ = − h∂x · ψ, ψi h∂t · ψ, ψidx ∂x ZR ∂ = − h∂x · ψ, ψi h∂x · ψ, ψidx ∂t R Z d1 =− |h∂x · ψ, ψi|2β dx, dt 2 R which proves the claim.



Remark 2.4. It is straightforward to generalize the previous conservation law to any Lp norm of ψ making use of (2.2) and (2.3). Proposition 2.5. Let ψ ∈ Γ(ΣR1,1 ) be a solution of (2.1). Then the following energy Z  ∂ 1 e(ψ) 2 + ∂ e(ψ) 2 dx E5 (t) = 2 R ∂x ∂t

4

VOLKER BRANDING

is conserved, where

1 (|∇∂t ψ|2β + |∇∂x ψ|2β ). 2 Proof. By a direct calculation we find that e(ψ) solves the one-dimensional wave equation, which yields the statement.  e(ψ) :=

Remark 2.6. By the Sobolev embedding H 1 ֒→ L∞ this yields a pointwise bound on e(ψ). It is straightforward to also bound higher derivatives of ψ. As a next step we investigate if the same conservation laws still hold if ψ solves a linear Dirac equation with a right hand side. To this end, we consider the linear Dirac equation iDψ = λψ,

λ ∈ R.

(2.4)

Note that (2.4) arises as critical point of the functional Z (hψ, iDψi − λ|ψ|2 )dµ, S(ψ) = R1,1

which leads to the prefactor of i in front of the Dirac operator. For a solution of (2.4), we have the following identities ∂ ∂ h∂t · ψ, ψi − h∂x · ψ, ψi = − λ(hiψ, ψi + hψ, iψi = 0, ∂t ∂x ∂ ∂ h∂x · ψ, ψi − h∂t · ψ, ψi =h(∂t · ∇∂x − ∂x · ∇∂t )ψ, ψi − hψ, (∂t · ∇∂x + ∂x · ∇∂t )ψi ∂t ∂x =2λhi∂x · ∂t · ψ, ψi.

(2.5) (2.6)

Lemma 2.7. Let ψ ∈ Γ(ΣR1,1 ) be a solution of (2.4). Then the energies Z 1 |ψ|2 dx, E1 (t) = 2 R β Z 1 E5 (t) = (|∇∂t ψ|2β + |∇∂x ψ|2β − λ2 |ψ|2β )dx 2 R are conserved. Proof. The first statement follows from integrating (2.5). Regarding the second claim, we note that due to the prefactor of i we obtain the following wave-type equation when squaring the Dirac operator ∇2∂t ψ − ∇2∂x ψ = λ2 ψ, which yields the second statement.  Similar to the case of λ = 0 we can also control higher Lp norms of ψ. Lemma 2.8. Let ψ ∈ Γ(ΣR1,1 ) be a solution of (2.1). Then the following inequality holds Z (|ψ|4β + |h∂x · ψ, ψi|2 )dx ≤ Ce|λ|t , R

where the positive constant C depends on ψ0 . Proof. Making use of (2.5) and (2.6) we calculate Z Z d 1 ∂ |ψ|4β dx = h∂t · ψ, ψi h∂t · ψ, ψidx dt 2 R ∂t ZR ∂ = h∂t · ψ, ψi h∂x · ψ, ψidx ∂x R Z ∂ = − h∂x · ψ, ψi h∂t · ψ, ψidx ∂x ZR ∂ = − h∂x · ψ, ψi( h∂x · ψ, ψi − 2λhi∂x · ∂t · ψ, ψi)dx ∂t R Z Z d 1 2 |h∂x · ψ, ψi|β dx + 2λ h∂x · ψ, ψihi∂x · ∂t · ψ, ψidx. =− dt 2 R R

ENERGY METHODS FOR DIRAC-TYPE EQUATIONS IN TWO-DIMENSIONAL MINKOWSKI SPACE

5

Note that the last term on the right hand side can be estimated as Z Z 2 h∂x · ψ, ψihi∂x · ∂t · ψ, ψidx ≤ (|ψ|4β + |h∂x · ψ, ψi|2β )dx R

R

giving the inequality Z Z d (|ψ|4β + |h∂x · ψ, ψi|2β )dx ≤ |λ| (|ψ|4β + |h∂x · ψ, ψi|2β )dx, dt R R which completes the proof.



Proposition 2.9. For given initial data ψ(0, x) = ψ0 (x) the solution of (2.4) exists globally on R1,1 . Proof. Making use of the conserved energy E5 (t) we obtain the following energy inequality Z Z Z 2 2 2 |ψ|2β dx + E5 (t) ≤ C |∇∂x ψ|β dx ≤ λ |∂x |ψ|β | dx ≤ R

R

R

for a uniform constant C. By the Sobolev embedding H 1 ֒→ L∞ this yields a pointwise bound on |ψ|β . Consequently, the solution of (2.4) exists globally.  2.1. Twisted spinors. In this subsection we want to discuss if the previous results still hold when we consider twisted spinors, which are sections in the spinor bundle that is twisted by some additional vector bundle F . To this end let F be a hermitian vector bundle with a metric connection. Moreover, we will assume that we have a positive definite scalar product on F . On the twisted bundle ΣR1,1 ⊗ F we obtain a metric connection induced from the connections on ΣR1,1 and F , which we will ˜ via setting denote by ∇, ˜ := ∇ΣR1,1 ⊗ 1F + 1ΣR1,1 ⊗ ∇F . ∇ The twisted Dirac operator D F : Γ(ΣR1,1 ⊗ F ) → Γ(ΣR1,1 ⊗ F ) is defined by ˜∂ . ˜ ∂ − ∂x · ∇ ˜ e = ∂t · ∇ D F := εj ej · ∇ x t j In contrast to the spinor bundle ΣR1,1 over Minkowski space the vector bundle F is not supposed to be trivial such that it may have non-vanishing curvature. We will denote its curvature endomorphism by RF (·, ·). Lemma 2.10. The square of the twisted Dirac operator D F satisfies the following Weitzenboeck formula ˜ 2 − ∂t · ∂x · RF (∂t , ∂x ), ˜2 + ∇ (D F )2 = −∇ ∂x ∂t

(2.7)

where RF denotes the curvature of the vector bundle F . Proof. We calculate ˜∂ ) ˜ ∂ − ∂x · ∇ ˜ ∂ )(∂t · ∇ ˜ ∂ − ∂x · ∇ (D F )2 =(∂t · ∇ x t x t 2 2 ˜ ˜ ˜ ˜∂ ∇ ˜ ˜∂ = − ∇∂t + ∇∂x − ∂t · ∂x · ∇∂t ∇∂x − ∂x · ∂t · ∇ x t F 2 2 ˜ ˜ = − ∇ + ∇ − ∂t · ∂x · R (∂t , ∂x ), ∂t

which completes the proof.

∂x



Note that, compared to the Riemannian case, we have a different sign in front of the curvature term of the vector bundle F . In addition, we do not get a scalar curvature contribution in (2.7) since we are restricting ourselves to two-dimensional Minkowski space. Remark 2.11. Most of the Dirac type equations studied in quantum field theory involve twisted Dirac operators [30]. In particular, the spinors that are considered in the standard model of elementary particle physics are sections in the spinor bundle twisted by some vector bundle.

6

VOLKER BRANDING

Again, we start by deriving several energy estimates for solutions of D F ψ = 0.

(2.8)

For solutions of (2.8) we obtain the following identities ∂ ∂ h∂t · ψ, ψi − h∂x · ψ, ψi =hD F ψ, ψi + hψ, D F ψi = 0, (2.9) ∂t ∂x ∂ ∂ ˜ ∂ )ψi = 0. ˜ ∂ + ∂x · ∇ ˜ ∂ )ψ, ψi − hψ, (∂t · ∇ ˜ ∂ − ∂x · ∇ h∂x · ψ, ψi − h∂t · ψ, ψi =h(∂t · ∇ t x t x ∂t ∂x (2.10) Proposition 2.12. Let ψ ∈ Γ(ΣR1,1 ⊗ F ) be a solution of (2.8). Then the energies Z 1 ˜ |ψ|2 dx, E1 (t) = 2 R β Z  ∂ 1 ˜ |ψ|2 2 + ∂ |ψ|2 2 dx E2 (t) = β β 2 R ∂x ∂t

are conserved.

Proof. This follows as in the proof of Lemma 2.1.



Lemma 2.13. Let ψ ∈ Γ(ΣR1,1 ⊗ F ) be a solution of (2.8). Then the energy Z E4 (t) = (|ψ|4β + |h∂x · ψ, ψi|2 )dx R

is conserved. Proof. This follows as in the proof of Lemma 2.3 making use of (2.9) and (2.10).



Remark 2.14. Again, it is straightforward to also find conserved energies involving higher Lp norms of ψ for solutions of (2.8). We set ˜3 (t) := 1 E 2

Z

R

˜ ∂x ψ|2 )dx. ˜ ∂t ψ|2 + |∇ (|∇ β β

When we try to control derivatives of solutions of (2.8) it will be necessary to control the curvature of the vector bundle F . ˜3 (t) satisfies Lemma 2.15. Let ψ ∈ Γ(ΣR1,1 ⊗ F ) be a solution of (2.8). Then E Z d ˜ ˜3 (t) + 1 E3 (t) ≤ E |RF (∂t , ∂x )|2 dx. dt 2 R Proof. By assumption we have D F ψ = 0 and consequently also (D F )2 ψ = 0. Now, we calculate Z d ˜ ˜∂ ∇ ˜ ∂ ψi)dx ˜ ∂ ψi + h∂t · ∇ ˜ ∂ ψ, ∇ ˜ 2 ψ, ∇ E3 (t) = (h∂t · ∇ t ∂t x t x dt ZR ˜2 − ∇ ˜ 2 )ψ, ∇ ˜ ∂ ψi)dx ˜ ∂ ψi + h∂t · RF (∂t , ∂x )ψ, ∇ = (h∂t · (∇ ZR

∂t

∂x

x

t

˜ ∂ ψi)dx ˜ ∂ ψi + h∂t · RF (∂t , ∂x )ψ, ∇ h∂x · RF (∂t , ∂x )ψ, ∇ x t Z ˜3 (t) + 1 ≤E |RF (∂t , ∂x )|2 dx 2 R =

R

yielding the result.



As a next step we discuss if the previous methods can still be employed if ψ solves a linear Dirac equation with non-trivial right-hand side, that is iD F ψ = λψ,

λ ∈ R.

(2.11)

ENERGY METHODS FOR DIRAC-TYPE EQUATIONS IN TWO-DIMENSIONAL MINKOWSKI SPACE

For a solution of (2.11), we have the following identities ∂ ∂ h∂t · ψ, ψi − h∂x · ψ, ψi =0, ∂t ∂x ∂ ∂ h∂x · ψ, ψi − h∂t · ψ, ψi =2λhi∂x · ∂t · ψ, ψi. ∂t ∂x

7

(2.12) (2.13)

Lemma 2.16. Let ψ ∈ Γ(ΣR1,1 ⊗ F ) be a solution of (2.11). Then the energy Z ˜1 (t) = 1 |ψ|2 dx E 2 R β is conserved. Proof. This follows by integrating (2.12).



Lemma 2.17. Let ψ ∈ Γ(ΣR1,1 ⊗ F ) be a solution of (2.11). Then the following inequality holds Z (|ψ|4β + |h∂x · ψ, ψi|2 )dx ≤ Ce|λ|t , R

where the positive constant C depends on ψ0 . Proof. The proof is the same as for Lemma 2.8 making use of (2.12) and (2.13).



Lemma 2.18. Let ψ ∈ Γ(ΣR1,1 ⊗ F ) be a solution of (2.11). Then the following inequality holds Z Z Z d 2 2 2 2 2 2 ˜ ˜ ˜ ˜ |ψ|2β |RF (∂t , ∂x )|2 dx. (|∇∂t ψ|β + |∇∂x ψ|β − λ |ψ|β )dx ≤ (|∇∂t ψ|β + |∇∂x ψ|β )dx + dt R R R Proof. By the Weitzenboeck formula (2.7) we find ˜2 ψ −∇ ˜ 2 ψ = λ2 ψ − ∂t · ∂x · RF (∂t , ∂x )ψ. ∇ ∂t ∂x In addition, we calculate Z d 1 ˜ ∂ ψ|2 − λ2 |ψ|2 ˜ ∂ ψ|2 +|∇ (|∇ x β β t β dt 2 R Z ˜ ∂ ψi)dx ˜ ∂ ψi + h∂t · RF (∂t , ∂x )ψ, ∇ = (h∂x · RF (∂t , ∂x )ψ, ∇ x t R

yielding the result. Remark 2.19. We have to impose a bound of the form to deduce an energy estimate for solutions of (2.11).

 R

R

|ψ|2β |RF (∂t , ∂x )|2 dx ≤ C if we want

3. Nonlinear Dirac equations in two-dimensional Minkowski space In this section we want to investigate if the energy methods developed for linear Dirac equations in the previous section can also be applied to the nonlinear case. The equations we will study mostly arise in quantum field theory, however, some of them also are connected to problems in differential geometry, see for example [28]. Up to now there exist many analytic results on nonlinear Dirac equations in two-dimensional Minkowski space. A general framework for semilinear hyperbolic systems was developed in [3], for a recent survey on nonlinear Dirac equations see [26] and references therein. 3.1. The Thirring Model. First, we will focus on a famous model from quantum field theory, the Thirring model. This model was introduced in [33] to describe the self-interaction of a Dirac field in two-dimensional Minkowski space. In the physics literature there exists a huge number of results on the Thirring model and also in the mathematical literature many results have been established. Most of the methods employed so far make use of a global trivialization of the spinor bundle over two-dimensional Minkowski space yielding existence results for the Thirring model, see for example [25, 29, 23, 36]. Making use of our energy methods we will also provide an existence result for the Thirring model.

8

VOLKER BRANDING

The action functional for the Thirring model is given by Z κ (hψ, iDψi − λ|ψ|2 − εj hψ, ej · ψihψ, ej · ψi)dµ S(ψ) = 2 R1,1

(3.1)

with real parameters κ and λ. In physics, λ is usually interpreted as mass, whereas κ describes the strength of interaction. The critical points of (3.1) are given by iDψ = λψ + κεj hψ, ej · ψiej · ψ.

(3.2)

Remark 3.1. Note that for λ = 0 solutions of (3.2) are invariant under scaling, that is if ψ is a solution of (3.2), then ψ(t, x) → rψ(r 2 t, r 2 x), where r is a positive number, is also a solution. For a solution of (3.2) the following identity holds ∂x · ∇∂t ψ − ∂t · ∇∂x =iλ∂x · ∂t · ψ + iεj κhψ, ej · ψi∂x · ∂t · ej · ψ =iλ∂x · ∂t · ψ − iκhψ, ∂t · ψi∂x · ψ + iκhψ, ∂x · ψi∂t · ψ leading to the two equations ∂ ∂ h∂t · ψ, ψi − h∂x · ψ, ψi =λ(hiψ, ψi + hψ, iψi) − κhψ, ej · ψi(hiej · ψ, ψi + hψ, iej · ψi) = 0, ∂t ∂x (3.3) ∂ ∂ h∂x · ψ, ψi − h∂t · ψ, ψi =2λhi∂x · ∂t · ψ, ψi. (3.4) ∂t ∂x Hence, for a solution of (3.2) the energy Z 1 E1 (t) = |ψ|2 dx (3.5) 2 R β is conserved again. By combining (3.3) and (3.4) we find ∂ hi∂x · ∂t · ψ, ψi = 4λ(hi∇∂t ψ, ψi + 2|ψ|2β (λ − κ|ψ|2 )). ∂x This identity turns out to be very useful in the case of the massless Thirring model, that is for solutions of (3.2) with λ = 0. More precisely, we find |ψ|2β = 2λ

Proposition 3.2. Let ψ ∈ Γ(ΣR1,1 ) be a solution of (3.2) with λ = 0. Then |ψ|2β ≤ C

(3.6)

for a positive constant C. Proof. Since |ψ|2β solves the one-dimensional wave equation, we again get a conserved energy and the result follows from the Sobolev embedding H 1 ֒→ L∞ .  In order to treat the massive Thirring model we need several auxiliary lemmata. Lemma 3.3. Let ψ ∈ Γ(ΣR1,1 ) be a solution of (3.2). Then the following wave type equation holds ∇2∂t ψ − ∇2∂x ψ =λ2 ψ + εj λκhψ, ej · ψiej · ψ + 2κλhi∂x · ∂t · ψ, ψii∂x · ∂t · ψ

(3.7)

+ κhψ, ∂t · ψi(−2i∇∂t ψ − λ∂t · ψ) + κhψ, ∂x · ψi(−2i∂t · ∂x · ∇∂t ψ − λ∂x · ψ) + κ2 (|hψ, ∂t · ψi|2 ψ + |hψ, ∂x · ψi|2 ψ − 2hψ, ∂x · ψihψ, ∂t · ψi∂x · ∂t · ψ). Proof. Applying iD to (3.2) we find ∇2∂t ψ − ∇2∂x ψ = λ2 ψ + εj λκhψ, ej · ψiej · ψ + iκD(εj hψ, ej · ψiej · ψ).

ENERGY METHODS FOR DIRAC-TYPE EQUATIONS IN TWO-DIMENSIONAL MINKOWSKI SPACE

9

In order to manipulate the last contribution on the right hand side we calculate (∂t · ∇∂t − ∂x · ∇∂x )(hψ, ∂t · ψi∂t · ψ − hψ, ∂x · ψi∂x · ψ) =2λhi∂x · ∂t · ψ, ψi∂x · ∂t · ψ − hψ, ∂t · ψi(∇∂t ψ + ∂x · ∂t · ∇∂x ψ) + hψ, ∂x · ψi(∇∂x ψ + ∂x · ∂t · ∇∂t ψ), where we made use of (3.3) and (3.4). Rewriting (3.2) as ∇∂x ψ = ∂x · ∂t · ∇∂t ψ + iλ∂x · ψ + iκεj hψ, ej · ψi∂x · ej · ψ and using the identity εj hψ, ∂t · ψihψ, ej · ψi∂t · ej · ψ + εj hψ, ∂x · ψihψ, ej · ψi∂x · ej · ψ = − |hψ, ∂t · ψi|2 ψ − |hψ, ∂x · ψi|2 ψ + 2hψ, ∂t · ψihψ, ∂x · ψi∂x · ∂t · ψ yields the claim.



Lemma 3.4. Let ψ ∈ Γ(ΣR1,1 ) be a solution of (3.2). Then the following inequality holds Z Z d 1 (3.8) ( |ψ|6β + |h∂x · ψ, ψi|2 |ψ|2β )dx = 4λ h∂t · ψ, ψih∂x · ψ, ψihi∂x · ∂t · ψ, ψidx. dt R 3 R Proof. This follows by a direct calculation using (3.3) and (3.4).



Proposition 3.5. Let ψ ∈ Γ(ΣR1,1 ) be a solution of (3.2) with λ 6= 0. Then the following inequality holds Z (|∇∂t ψ|2β + |∇∂x ψ|2β )dx ≤ CeCt , (3.9) R

where the constant C depends on λ, κ and the initial data. Proof. From (3.7) and a direct calculation we obtain Z Z Z d 1 d 1 |ψ|3β |∇ψ|β dx (|∇∂t ψ|2β + |∇∂x ψ|2β )dx ≤ κ2 ( |ψ|6β + |h∂x · ψ, ψi|2 |ψ|2β )dx + C dt 2 R dt 3 R Z Z R 2 6 |∇ψ|β dx, |ψ|β dx + C ≤C R

R

where we used (3.8) in the last step. In order to estimate the L6 -norm of ψ we make use of the Sobolev embedding theorem in one dimension Z Z 2 3 1 6 2 ∂x |ψ|3 3 dx 2 |ψ|β dx ≤ C β R ZR 2 4 3 |∇∂x ψ|β3 |ψ|β3 dx 2 ≤C ZR Z 1 |∇∂x ψ|2β dx 2 ≤C |ψ|2β dx. R

Since the

R

L2 -norm

of ψ is conserved for a solution of (3.2) we obtain Z Z d 1 2 2 (|∇∂t ψ|β + |∇∂x ψ|β )dx ≤C (|∇∂t ψ|2β + |∇∂x ψ|2β )dx dt 2 R R

and the result follows by integration of the differential inequality.



Corollary 3.6. Let ψ ∈ Γ(ΣR1,1 ) be a solution of (3.2) with λ 6= 0. Then the following estimate holds |ψ|β ≤ CeCt , where the positive constant C depends on λ, κ and the initial data.

(3.10)

10

VOLKER BRANDING

Theorem 3.7 (Existence of a global solution). For any given initial data of the regularity ψ(0, x) =ψ0 (x) ∈ H 1 (R, ΣR1,1 ) the equation (3.2) admits a global weak solution in H 1 (R1,1 , ΣR1,1 ), which is uniquely determined by the initial data. Proof. The existence of a global solution follows directly since we have a uniform bound on ψ for the massless case λ = 0. Moreover, in the massive case λ 6= 0 we have an exponential bound on ψ. These bounds ensure that the solution cannot blow up and has to exist globally. To achieve uniqueness let us consider two solutions ψ, ξ of (3.2) that coincide at t = 0. Set η := ψ − ξ. Then η satisfies ∂t · ∇∂t η = ∂x · ∇∂x η − iλη − iκεj (hη, ej · ψiej · ψ + hξ, ej · ηiej · ψ − hξ, ej · ξiej · η). Thus, we find Z Z d 2 |η| dx =2κ (εj Rehη, ej · ψihiej · ψ, ηi) + εj Rehξ, ej · ηihiej · ψ, ηi)dx dt R β ZR |η|2β (|ψ|2β + |χ|β |ψ|β )dx ≤C ZR |η|2β dx, ≤C R

where we used the pointwise bound on ψ, ξ in the last step. Consequently, we find Z Z Ct 2 |η|2β |t=0 dx |η|β dx ≤ e R

R

such that if ψ = ξ at t = 0 then ψ = ξ for all times.



Remark 3.8. The existence of a global solution for the Thirring model is due to the algebraic structure of the right hand side of (3.2). More generally, we could consider iDψ = κV (ψ)εj hej · ψ, ψiej · ψ, where V (ψ) is supposed to be a real-valued potential. It can again be checked that |ψ|2β = 0 such that we get a global bound on ψ. Remark 3.9. In the physics literature the Thirring model is usually formulated as Z ¯ iDψi − λ|ψ|2 − κ εj hψ, ¯ = ¯ ej · ψihψ, ¯ ej · ψi)dµ, (hψ, S(ψ, ψ) β 2 R1,1 where it is assumed that ψ and ψ¯ are independent. Hence, the critical points consist of two equations ¯ ej · ψiej · ψ, iDψ = λψ + κhψ,

¯ j · ψ. ¯ iD ψ¯ = λψ¯ + κhψ, ej · ψie

These equations have to be considered as independent as can be checked by a direct calculation. 3.2. The Soler model. In this subsection we want to briefly analyze if the energy methods from above can also be applied to another famous model from quantum field theory, the so-called Soler model [32]. Then energy functional of the Soler model is given by Z κ (hψ, iDψi − λ|ψ|2 − |ψ|4 )dµ, S(ψ) = (3.11) 2 1,1 R where λ and κ are real parameters. Again, λ is usually interpreted as mass and κ as strength of interaction. The critical points of (3.11) are given by iDψ = λψ + µ|ψ|2 ψ.

(3.12)

ENERGY METHODS FOR DIRAC-TYPE EQUATIONS IN TWO-DIMENSIONAL MINKOWSKI SPACE

11

For a solution of (3.12) we obtain the following identities ∂ ∂ h∂t · ψ, ψi − h∂x · ψ, ψi = 0, ∂t ∂x ∂ ∂ h∂x · ψ, ψi − h∂t · ψ, ψi = 2(λ + µ|ψ|2 )hψ, i∂x · ∂t · ψi ∂t ∂x and also the following wave-type equation ∇2∂t ψ − ∇2∂x ψ = λ2 ψ + 2λµ|ψ|2 ψ + µ2 |ψ|4 ψ + iµ(

∂ ∂ |ψ|2 ∂t · ψ − |ψ|2 ∂x · ψ). ∂t ∂x

(3.13) (3.14)

(3.15)

By integrating (3.13) we directly obtain that 1 E1 (t) = 2

Z

|ψ|2β dx R

is conserved. Lemma 3.10. Let ψ ∈ Γ(ΣR1,1 ) be a solution of (3.12). Then the following equations hold |ψ|2 =0, |ψ|2β =2

∂ (λ + µ|ψ|2 )hψ, i∂x · ∂t · ψi. ∂x

Proof. In order to prove the first assertion we note that (3.12) leads to ∇∂t ψ = −∂t · ∂x · ∇∂x ψ + i∂t · (λ + µ|ψ|2 )ψ,

∇∂x ψ = ∂x · ∂t · ∇∂x ψ + i∂x · (λ + µ|ψ|2 )ψ.

Making use of these equations we find ∂ |ψ|2 = −h∂t · ∂x · ∇∂x ψ, ψi − hψ, ∂t · ∂x · ∇∂x ψi, ∂t ∂ |ψ|2 = h∂x · ∂t · ∇∂t ψ, ψi + hψ, ∂x · ∂t · ∇∂t ψi. ∂x This directly yields |ψ|2 = 0. For the second assertion we use (3.13) and (3.14).



Note that we do not get control over ψ from |ψ|2 = 0 since the scalar product used here is non-definite. Due to the same reason we also do not obtain a useful conserved energy from (3.15). Remark 3.11. Let us briefly discuss the Soler model how it is usually formulated in the physics literature. Here, one to assumes that ψ and ψ¯ := ∂t · ψ are independent fields, such that the critical points consist of two independent equations. Z ¯ ψi|2 )dµ. ¯ iDψi − λhψ, ¯ ψi − κ |hψ, ¯ = (3.16) (hψ, S(ψ, ψ) 2 1,1 R The critical points of (3.16) are given by the two equations iDψ = λψ + κ|ψ|2β ψ,

iD ψ¯ = λψ¯ + κ|ψ|2β ψ¯

(3.17)

and it can easily be checked that these two equations have to be considered as independent. For a solution of (3.17) we have ∂ ∂ h∂t · ψ, ψi − h∂x · ψ, ψi = 0, ∂t ∂x ∂ ∂ h∂x · ψ, ψi − h∂t · ψ, ψi = 2(λ + µ|ψ|2β )hψ, i∂x · ∂t · ψi ∂t ∂x ¯ 2 . Hence, for a solution of (3.17) the energy and similarly for |ψ| β Z Z 1 1 2 ¯ 2 dx E1 (t) = |ψ| dx = |ψ| 2 R β 2 R β is conserved again.

12

VOLKER BRANDING

4. Dirac-wave maps from two-dimensional Minkowski space Dirac-wave maps arise as a mathematical version of the supersymmetric nonlinear sigma model studied in quantum field theory, see for example [1] for the physics background. The central object of the supersymmetric nonlinear sigma model is an energy functional that consists of a map between two manifolds and so-called vector spinors. We want to analyze this model with the methods from geometric analysis, hence in contrast to the physics literature we will consider standard instead of Grassmann-valued spinors. In order to define Dirac-wave maps from two-dimensional Minkowski space we choose the following setup. Let (N, h) be a compact Riemannian manifold and let φ : R1,1 → N be a map. We consider the pullback of the tangent bundle from the target, which will be denoted by φ∗ T N . As discussed in section 2.1 we form the twisted spinor bundle ΣM ⊗ φ∗ T N . Sections in ΣM ⊗ φ∗ T N will be called vector spinors. Most of the results that have been obtained in the mathematical literature on the supersymmetric nonlinear sigma model consider the case where both domain and target manifolds are Riemannian. This study was initiated in [20], where the notion of Dirac-harmonic maps was introduced. Dirac-harmonic maps form a semilinear elliptic system for a map between two Riemannian manifolds and a spinor along that map. For a given Dirac-harmonic map many analytic and geometric results have been established, as for example the regularity of weak solutions [34]. Motivated from the physics literature there exist several extensions of the Dirac-harmonic map system such as Dirac-harmonic maps with curvature term [13], [19] and Dirac-harmonic maps with torsion [14] Making use of the Atiyah-Singer index theorem uncoupled solutions to the equations for Diracharmonic maps have been constructed in [2]. Here, uncoupled refers to the fact that the map part is harmonic. In addition, several approaches to the existence problem that make use of the heat-flow method have been studied in [16, 21, 27, 11, 35] and [10]. Up to now there is only one reference investigating Dirac-wave maps [24], that is critical points of the supersymmetric nonlinear sigma model with the domain being two-dimensional Minkowski space. Expressing the Dirac-wave map system in characteristic coordinates an existence result for smooth initial data could be obtained. In this section we will extend the analysis of Dirac-wave maps and derive an existence result that also includes distributional initial data. The methods we use here are partly inspired from the analysis of wave maps, see [31] for an introduction to the latter. A problem similar to the one studied in this section, namely the full bosonic string from twodimensional Minkowski space to Riemannian manifolds was treated in [18]. The energy functional for Dirac-wave maps is given by 1 S(φ, ψ) = 2

Z

(|dφ|2 + hψ, iD φ

∗T N

(4.1)

ψi)dµ.

R1,1

Here, D φ T N is the twisted Dirac operator acting on vector spinors. Note that iD φ T N is self-adjoint with respect to the L2 -norm such that the energy functional is real-valued. Whenever choosing local coordinates we will use Latin indices to denote coordinates on twodimensional Minkowski space and Greek indices to denote coordinates on the target manifold. For the sake of completeness we will give a short derivation of the critical points of (4.1). ∗



Proposition 4.1. The Euler-Lagrange equations of (4.1) read 1 τ (φ) = RN (ψ, iej · ψ)εj dφ(ej ), 2 ∗ D φ T N ψ =0.

(4.2) (4.3)

Here, RN denotes the curvature tensor of the target N and ej , j = 1, 2 is an pseudo-orthonormal basis of T R1,1 .

ENERGY METHODS FOR DIRAC-TYPE EQUATIONS IN TWO-DIMENSIONAL MINKOWSKI SPACE

13

˜ Proof. First, we consider a variation of ψ, while keeping φ fixed, satisfying ∇ψ ∂s s=0 = ξ. We calculate Z Z d 1 ∗ φ∗ T N φ∗ T N Rehξ, iD φ T N ψidµ (hξ, iD ψi + hψ, iD ξi)dµ = S(φ, ψ) = s=0 ds 2 R1,1 R1,1 yielding the equation for the vector spinor ψ. To obtain the equation for the map φ we consider = η. It is well-known that a variation of φ, while keeping ψ fixed, that is ∂φ ∂s s=0 Z Z d 1 hτ (φ), ηidµ, |dφ|2 dµ = − s=0 ds 2 R1,1 R1,1

where τ (φ) := ∇eφj T N dφ(ej ) is the wave-map operator. In addition, we find Z Z 1 1 d φ∗ T N hψ, iD ψidµ = hψ, RN (dφ(∂s ), dφ(ej ))iεj ej · ψi|s=0 dµ ds s=0 2 R1,1 2 R1,1 Z 1 = hψ, RN (ψ, iεj ej · ψ)dφ(ej ), ηidµ, 2 R1,1 completing the proof. ∗



R1,1

Solutions of the system (4.2), (4.3) are called Dirac-wave maps from → N. The system (4.2), (4.3) can be expanded as ∇ ∇ 1 1 dφ(∂t ) − dφ(∂x ) = RN (ψ, i∂t · ψ)dφ(∂t ) − RN (ψ, i∂x · ψ)dφ(∂x ), ∂t ∂x 2 2 ˜ ˜ ∂t · ∇∂t ψ =∂x · ∇∂x ψ.

(4.4) (4.5)

Choosing local coordinates on the target N , the Euler-Lagrange equations acquire the form (α = 1, . . . , dim N ) γ β β β ∂φβ ∂φγ  ∂ 2 φα ∂ 2 φα α ∂φ ∂φ γ δ ∂φ γ δ ∂φ α − + Γ − − hψ , i∂ · ψ i ), (hψ , i∂ · ψ i =R x t βγ βγδ ∂t2 ∂x2 ∂t ∂t ∂x ∂x ∂t ∂x ∂φβ εj ej · ψ γ , Dψ α = − Γαβγ ∂xj

where Γαβγ are the Christoffel symbols and Rαβγδ the components of the curvature tensor on the target N . In order to treat a weak version of the system (4.2), (4.3) it will be necessary to embed N isometrically into some Rq making use of the Nash embedding theorem. Then we have φ : R1,1 → Rq and ψ ∈ Γ(ΣR1,1 ⊗ Rq ). In this case the equations for Dirac-wave maps acquire the form ∂2φ ∂2φ − =II(dφ, dφ) + P (II(ψ, dφ(∂t )), i∂t · ψ) − P (II(ψ, dφ(∂x )), i∂x · ψ), (4.6) ∂t2 ∂x2 ∂φ ∂φ ∂t · ∇∂t ψ − ∂x · ∇∂x ψ =II( , ∂t · ψ) − II( , ∂x · ψ), (4.7) ∂t ∂x where II denotes the second fundamental form of the embedding and P the shape operator defined by hP (ξ, X), Y )i = hII(X, Y ), ξi for X, Y ∈ Γ(T N ) and ξ ∈ T ⊥ N . For solutions of (4.3) we obtain the following identities ∂ ∂ ∗ ∗ h∂t · ψ, ψi − h∂x · ψ, ψi =hD φ T N ψ, ψi + hψ, D φ T N ψi = 0, (4.8) ∂t ∂x ∂ ∂ ˜ ∂ )ψi = 0. ˜ ∂ + ∂x · ∇ ˜ ∂ )ψ, ψi − hψ, (∂t · ∇ ˜ ∂ − ∂x · ∇ h∂x · ψ, ψi − h∂t · ψ, ψi =h(∂t · ∇ t x t x ∂t ∂x (4.9) Moreover, for a solution of (4.3) the following wave type equation holds ˜2 ψ − ∇ ˜ 2 ψ = −∂t · ∂x · RN (dφ(∂t ), dφ(∂x ))ψ ∇ ∂t ∂x due to the Weitzenboeck formula (2.7) with F = φ∗ T N .

14

VOLKER BRANDING

Remark 4.2. In the physics literature [1] the energy functional for the supersymmetric nonlinear sigma model is defined as follows Z 1 ¯ ¯ iD φ∗ T N ψi)dµ, S(φ, ψ, ψ) = (|dφ|2 + hψ, 2 R1,1 where ψ¯ := ∂t · ψ. Treating ψ, ψ¯ as independent fields, we obtain the Euler-Lagrange equations ∗ ∗ 1 ¯ iej · ψ)εj dφ(ej ), D φ T N ψ = 0, D φ T N ψ¯ = 0. τ (φ) = RN (ψ, 2 Note that the spinors ψ and ψ¯ have to be considered as independent since the two equations ∗ ∗ D φ T N ψ = 0, D φ T N (∂t · ψ) = 0 are not compatible. In addition, the equation for the map φ would acquire the form 1 τ (φ) = − RN (ψ, i∂t · ∂x · ψ)dφ(∂x ), |2 {z }

=0

when inserting ψ¯ := ∂t · ψ. It turns out that the curvature term on the right hand side vanishes due to symmetry reasons. In the physics literature one usually considers anticommuting spinors such that this term does not vanish. In the following we will only analyze the system (4.2), (4.3). Let us demonstrate how to construct an explicit solution to the Euler-Lagrange equations (4.2), (4.3), where we follow the ideas used for the construction of an explicit solution to the Diracharmonic map system in [20, Proposition 2.2]. To this end we recall the following

Definition 4.3. Let M be a n-dimensional Lorentzian spin manifold. A spinor ψ ∈ Γ(ΣM ) is called twistor spinor if it satisfies 1 (4.10) PX ψ := ∇ΣM X ψ + X · Dψ = 0 n for all vector fields X. In two-dimensional Minkowski space twistor spinors are of the form ψ(x) = ψ1 + x · ψ2 , where ψ1 , ψ2 are constant spinors [9]. Proposition 4.4. Let φ : R1,1 → N be a wave map, that is a solution of τ (φ) = 0. We set ψ := εj ej · χ ⊗ dφ(ej ), where χ is a twistor spinor. Then the pair (φ, ψ) is a Dirac-wave map, that is uncoupled: 1 ∗ τ (φ) = 0 = εj RN (ψ, iej · ψ)dφ(ej ), D φ T N ψ = 0. 2 Proof. First, we check that the equation for the vector spinor ψ is satisfied. To this end, we calculate   ∗ ˜ ∂ ∂t · χ ⊗ dφ(∂t ) − ∂x · χ ⊗ dφ(∂x ) ˜ ∂ − ∂x · ∇ D φ T N ψ = ∂t · ∇ x t   ΣM ΣM ΣM = − ∇ΣM ∂t χ + ∂x · ∂t · ∇∂x χ ⊗ dφ(∂t ) + ∇∂x χ − ∂t · ∂x · ∇∂t χ ⊗ dφ(∂x ) ∇ ∇ dφ(∂t )) − χ ⊗ τ (φ) − ∂t · ∂x · χ ⊗ ( dφ(∂x ) − ∂t ∂x =0, the first two terms vanish since χ is a twistor spinor by assumption. As a second step we check that the curvature term on the right hand side of (4.2) vanishes. Using the local expression of (4.2) εj RN (ψ, iej · ψ)dφ(ej ) =Rαβγδ

β β ∂ γ δ ∂φ γ δ ∂φ (hψ , i∂ · ψ i − hψ , i∂ · ψ i ) t x ∂y α ∂t ∂x

ENERGY METHODS FOR DIRAC-TYPE EQUATIONS IN TWO-DIMENSIONAL MINKOWSKI SPACE

15

we find (the second term vanishes for the same reason) ∂φγ ∂φδ ∂φγ ∂φδ  + ∂t ∂t ∂x ∂x γ ∂φδ  γ ∂φδ ∂φ ∂φ + =0 + Rαβγδ hχ, i∂x · χi ∂t ∂x ∂x ∂t due to the symmetries of the curvature tensor on N . Rαβγδ hψ γ , i∂t · ψ δ i = − Rαβγδ hχ, i∂t · χi



4.1. Conserved Energies. In this subsection we give several conserved energies for solutions of the system (4.2), (4.3). By integrating (4.8) we directly get that Z |ψ|2β dx E1 (t) = R

is conserved for a solution of (4.3). Again, it is straightforward to also control higher Lp norms of ψ as discussed in 2.1. Moreover, for a solution of (4.3) we have ˜ ∂ ψ|2 . ˜ ∂ ψ|2 = |∇ |∇ β x β t Proposition 4.5. Let (φ, ψ) : R1,1 → N be a Dirac-wave map. Then the energy Z  1 ˜ ∂ ψi dx E(t) := |dφ(∂t )|2 + |dφ(∂x )|2 − hψ, i∂t · ∇ t 2 R

(4.11)

is conserved. Proof. We calculate Z Z d 1 2 2 (|dφ(∂t )| + |dφ(∂x )| )dx = hdφ(∂t ), τ (φ)idx dt 2 R R Z 1 =− hdφ(∂t ), RN (ψ, i∂x · ψ)dφ(∂x )idx. 2 R Differentiating (4.3) with respect to t we obtain the following identity ˜ 2 ψ = ∂x · ∇ ˜∂ ∇ ˜∂ ∇ ˜ ∂ ψ = ∂x · RN (dφ(∂t ), dφ(∂x ))ψ + ∂x · ∇ ˜ ∂ ψ. ∂t · ∇ ∂t t x x t Thus, we find d 1 dt 2

Z

˜ ∂ ψidx = 1 hψ, i∂t · ∇ t 2 R

Z

˜ ∂ )ψ , i∇ ˜ ∂ − ∂x · ∇ ˜ ∂ ψi (h(∂t · ∇ x t t | {z } R =0 N

+ hψ, i∂x · R (dφ(∂t ), dφ(∂x ))ψi)dx,

where we used integration by parts in the last step. The assertion then follows by adding up both contributions.  Having gained control over ψ we now establish a bound on the derivatives of φ. To this end we set 1 e(φ) := (|dφ(∂t )|2 + |dφ(∂x )|2 ). 2 Then we obtain the following Proposition 4.6. Let (φ, ψ) : R1,1 → N be a Dirac-wave map. Then the following formula holds e(φ) = where  :=

∂2 ∂t2



∂2 . ∂x2

∂2 ˜ ∂2 ˜ h∇∂x ψ, i∂x · ψi − 2 h∇ ∂t ψ, i∂t · ψi, 2 ∂t ∂x

(4.12)

16

VOLKER BRANDING

Proof. Testing (4.4) with dφ(∂t ) and dφ(∂x ) we obtain the two equations ∂ ∂ e(φ) − hdφ(∂t ), dφ(∂x )i = −hRN (ψ, i∂x · ψ)dφ(∂x ), dφ(∂t )i, ∂t ∂x ∂ ∂ − e(φ) + hdφ(∂t ), dφ(∂x )i = hRN (ψ, i∂t · ψ)dφ(∂t ), dφ(∂x )i. ∂x ∂t Differentiating the first equation with respect to t, the second one with respect to x and adding up both contributions we find e(φ) =

∂ ∂ hRN (ψ, i∂t · ψ)dφ(∂t ), dφ(∂x )i − hRN (ψ, i∂x · ψ)dφ(∂x ), dφ(∂t )i. ∂x ∂t

We may rewrite the right-hand side as follows ∂ ∂ hRN (ψ, i∂t · ψ)dφ(∂t ), dφ(∂x )i − hRN (ψ, i∂x · ψ)dφ(∂x ), dφ(∂t )i ∂x ∂t ∂ ˜ ˜ ∂ ˜ ˜ ∂ ˜ ˜ ∂ ˜ ˜ h∇∂x ∇∂t ψ, i∂t · ψi − h∇ h∇∂t ∇∂x ψ, i∂x · ψi = h∇ ∂t ∇∂x ψ, i∂t · ψi − ∂x ∇∂t ψ, i∂x · ψi + ∂x ∂x ∂t ∂t 2 ∂ ˜ ∂2 ˜ ˜ ∂ ψi − ∂ h∇ ˜ ∂ ψi ˜ ∂ ψ, i∂t · ∇ ˜ ∂ ψ, i∂t · ψi + ∂ h∇ h∇∂x ψ, i∂t · ψi − h∇∂x ψ, i∂t · ∇ = t x t t 2 ∂x∂t ∂x ∂x ∂x 2 2 ∂ ˜ ∂ ψi + ∂ h∇ ˜ ∂ ψi ˜ ∂ ψ, i∂x · ψi + ∂ h∇ ˜ ∂ ψ, i∂x · ∇ ˜ ∂ ψ, i∂x · ∇ ˜ ∂ ψ, i∂x · ψi − ∂ h∇ − h∇ x t t t x x 2 ∂x∂t ∂t ∂t ∂t 2  ∂ ˜ ∂ )ψ , iψi + ∂ h∇ ˜ ∂ − ∂x · ∇ ˜ ∂ ψi ˜ ∂ ψ, i∂t · ∇ ˜ ∂ ψi − h∇ ˜ ∂ ψ, i∂t · ∇ = h(∂t · ∇ t x t x x t {z } {z } ∂x∂t | ∂x | =0

=0

2 2  ∂ ˜ ∂ ψ, i∂x · ψi − ∂ h∇ ˜ ∂ ψ, i∂t · ψi ˜ ∂ ψi + ∂ h∇ ˜ ∂ ψ, i∂x · ∇ ˜ ∂ ψi − h∇ ˜ ∂ ψ, i∂x · ∇ + h∇ x t t x x {z } ∂t | t ∂t2 ∂x2 =0

=

∂2

∂t2

˜ ∂ ψ, i∂x · ψi − h∇ x

∂2

∂x2

˜ ∂ ψ, i∂t · ψi, h∇ t

where we used that ψ is a solution of (4.5) several times completing the proof.



Remark 4.7. The conserved energies that we have presented so far all reflect the hyperbolic nature of the Dirac-wave map system (4.2), (4.3). In addition, as in the Riemannian case, we also have the energy-momentum tensor, which is conserved for a Dirac-wave map. More precisely, the symmetric 2-tensor Tij defined by 1 ˜ e + ej · ∇ ˜ e )ψi Tij := 2hdφ(ei ), dφ(ej )i − gij |dφ|2 − hψ, (ei · ∇ j i 2 is divergence free for a solution of (4.2), (4.3). 4.2. An existence result via energy methods. In this subsection we will derive an existence result for the Cauchy problem associated to (4.2), (4.3). Since we want to be able to treat initial data of low regularity, we have to use the extrinsic version of the Dirac-wave map system (4.6), (4.7). Making use of the conserved energies from the last subsection we find Proposition 4.8. Let (φ, ψ) : R1,1 → Rq be a Dirac-wave map. Then the following energy is conserved Z  ∂φ 2 ∂φ 2 1 (4.13) | + | − hψ, i∂t · ∇∂t ψi dx. EDW (t) := 2 R ∂t ∂x Proof. By (4.7) we have a conserved energy for the intrinsic version of the Dirac-wave map system. To obtain the conserved energy for the extrinsic version we consider the isometric embedding ι : N → Rq . Since ι is an isometry we may apply its differential to all terms in (4.13), which gives the statement. 

ENERGY METHODS FOR DIRAC-TYPE EQUATIONS IN TWO-DIMENSIONAL MINKOWSKI SPACE

Corollary 4.9. Let (φ, ψ) : R1,1 → Rq be a Dirac-wave map. Then the energy Z  ∂ (e(φ) − hi∂t · ∇∂ ψ, ψi) 2 + | ∂ (e(φ) − hi∂t · ∇∂ ψ, ψi) 2 dx E(t) = t t ∂x ∂t R

17

(4.14)

is conserved.

Proof. We obtain a conserved energy from (4.12). Applying the isometric embedding ι again yields the claim.  Proposition 4.10. Let ψ be a solution of (4.3). Then the following pointwise bound holds |ψ|β ≤ C,

(4.15)

where the positive constant C depends on ψ0 . Proof. This follows from Proposition 2.12 with F = φ∗ T N and the Sobolev embedding H 1 ֒→ L∞ .  Lemma 4.11. Let (φ, ψ) : R1,1 → Rq be a Dirac-wave map. Then the following estimate holds ∂φ 2 ∂φ 2 | + | ≤ C(1 + |∇ψ|β ), (4.16) ∂t ∂x where the positive constant C depends on EDW (0) and ψ0 . Proof. From the conserved energy (4.14) and the Sobolev embedding H 1 ֒→ L∞ we obtain the following bound |e(φ) − hi∂t · ∇∂t ψ, ψi)| ≤ EDW (0). Using the pointwise bound on ψ we obtain the claim.



In order to obtain control over the derivatives of ψ we turn (4.7) into a wave-type equation. Applying the Dirac operator D on both sides of (4.7) we obtain the wave-type equation ∂φ ∂φ , ∇∂t ψ + ∂x · ∂t · ∇∂x ψ) + II( , −∇∂x ψ + ∂t · ∂x · ∇∂t ψ) ∂t ∂x ∂φ ∂φ + (∇dφ(∂t ) II)( , ψ) + (∇dφ(∂t ) II)( , ∂t · ∂x · ψ) ∂t ∂x ∂φ ∂φ − (∇dφ(∂x ) II)( , ∂t · ∂x · ψ) + (∇dφ(∂x ) II)( , ψ). (4.17) ∂x ∂t

∇2∂t ψ − ∇2∂x ψ =II(φ, ψ) + II(

We set E(ψ,1,2) (t) :=

1 2

Z

(|∇∂t ψ|2β + |∇∂x ψ|2β )dx.

R

The energy E(ψ,1,2) (t) satisfies the following differential inequality Lemma 4.12. Let (φ, ψ) : R1,1 → Rq be a Dirac-wave map. Then the following inequality holds 1 d E(ψ,1,2) (t) ≤ C E(ψ,1,2) (t) + (E(ψ,1,2) (t)) 2 , (4.18) dt where the positive constant C depends on N and the initial data. Proof. Making use of (4.17) a direct calculation yields Z ∂φ d h∂t · ∇∂t ψ, II(φ, ψ)i + h∂t · ∇∂t ψ, II( , ∇∂t ψ + ∂x · ∂t · ∇∂x ψ)i E (t) = dt (ψ,1,2) ∂t R ∂φ ∂φ + h∂t · ∇∂t ψ, II( , −∇∂x ψ + ∂t · ∂x · ∇∂t ψ)i + h∂t · ∇∂t ψ, (∇dφ(∂t ) II)( , ψ)i ∂x ∂t ∂φ ∂φ + h∂t · ∇∂t ψ, (∇dφ(∂t ) II)( , ∂t · ∂x · ψ)i + h∂t · ∇∂t ψ, (∇dφ(∂x ) II)( , ψ)i ∂x ∂t  ∂φ − h∂t · ∇∂t ψ, (∇dφ(∂x ) II)( , ∂t · ∂x · ψ)i dx. ∂x

18

VOLKER BRANDING

Note that second and third term on the right hand side can be rewritten as ∂φ ∂φ h∂t · ∇∂t ψ, II( , ∇∂t ψ + ∂x · ∂t · ∇∂x ψ)i = −h∂t · ψ, (∇dφ(∂t ) II)( , ∇∂t ψ + ∂x · ∂t · ∇∂x ψ)i ∂t ∂t since ψ ⊥ II. Consequently, we get the following inequality Z d E (t) ≤C (|ψ|β |∇ψ|β |φ| + |ψ|β |∇ψ|β |dφ|2 )dx dt (ψ,1,2) ZR ≤C (|ψ|β |∇ψ|β |dφ|2 + |ψ|5β |∇ψ|β )dx Z ZR 1  2 (|∇ψ|2β dx 2 , (|∇ψ|β dx + ≤C R

R

where we used (4.16) in the last step, completing the proof.



Corollary 4.13. Let (φ, ψ) : R1,1 → Rq be a Dirac-wave map. Then the following estimate holds E(ψ,1,2) (t) ≤ CeCt , where the positive constant C depends on N and the initial data. Proof. This follows from the last Lemma and the Gronwall inequality.



Corollary 4.14. Let (φ, ψ) : R1,1 → Rq be a Dirac-wave map. Then the following estimate holds Z ∂φ 2  ∂φ 2 | + | dx ≤ C(1 + eCt ), ∂t ∂x R where the positive constant C depends on N and the initial data. Proof. From the conserved energy (4.13) we obtain Z Z Z 1 1 ∂φ 2  ∂φ 2 2 2 |ψ|β dx +| dx ≤ EDW (0) + C |∇ψ|2β )dx 2 ≤ EDW (0) + CeCt , | ∂t ∂x R R R yielding the result. It turns out that we need to gain control over the following fact.

L4 -norm



of ∇ψ. To this end let us recall the

Remark 4.15. Suppose that f : R1,1 → R is a solution of the scalar wave equation, that is f = 0. Then we have the following conservation law Z  ∂f 4 ∂f 4 d + + 6 ∂f 2 ∂f 2 dx = 0. dt R ∂t ∂x ∂t ∂x

Motivated from the conserved energy for solutions of the scalar wave equation we define Z E(ψ,1,4) (t) := (|∇∂t ψ|4β + |∇∂x ψ|4β + 2|∇∂x ψ|2β |∇∂t ψ|2β + 4|h∂t · ∇∂x ψ, ∇∂t ψi|2 )dx. R

It can be checked by a direct calculation that E(ψ,1,4) (t) is conserved when ψ ∈ Γ(ΣR1,1 ) is a solution of ∇2∂t ψ = ∇2∂x ψ. Lemma 4.16. Let (φ, ψ) : R1,1 → Rq be a Dirac-wave map. Then the following inequality holds 1 d E(ψ,1,4) (t) ≤ C E(ψ,1,4) (t) + (E(ψ,1,4) (t)) 2 , (4.19) dt where the positive constant C depends on N and the initial data. Proof. Let ψ be a solution of ∇2∂t ψ − ∇2∂x ψ = f . Then a direct, but lengthy calculation yields Z d E(ψ,1,4) (t) = (4|∇∂t ψ|2β h∂t · ∇∂t ψ, f i − 4|∇∂x ψ|2β h∂t · ∇∂t ψ, f i dt R + 8h∂t · ∇∂x ψ, f ih∂t · ∇∂t ψ, ∇∂x ψi)dx.

ENERGY METHODS FOR DIRAC-TYPE EQUATIONS IN TWO-DIMENSIONAL MINKOWSKI SPACE

19

At this point we choose ∂φ ∂φ , ∇∂t ψ + ∂x · ∂t · ∇∂x ψ) + II( , −∇∂x ψ + ∂t · ∂x · ∇∂t ψ) ∂t ∂x ∂φ ∂φ + (∇dφ(∂t ) II)( , ψ) + (∇dφ(∂t ) II)( , ∂t · ∂x · ψ) ∂t ∂x ∂φ ∂φ − (∇dφ(∂x ) II)( , ∂t · ∂x · ψ) + (∇dφ(∂x ) II)( , ψ). ∂x ∂t By the same reasoning as in the proof of Lemma 4.12 we find Z d E(ψ,1,4) (t) ≤C (|ψ|β ||∇ψ|3β |φ| + |ψ|β |∇ψ|3β |dφ|2 )dx dt Z ZR 1 4 |∇ψ|4β dx 2 , |∇ψ|β dx + C ≤C f =II(φ, ψ) + II(

R

R

which completes the proof.



Corollary 4.17. Let (φ, ψ) : R1,1 → Rq be a Dirac-wave map. Then the following estimate holds E(ψ,1,4) (t) ≤ CeCt ,

(4.20)

where the positive constant C depends on N and the initial data. Proof. This follows from the last lemma and the Gronwall inequality.



After having gained control over the derivatives of ψ we now control the second derivative of φ. To this end we set Z ∂ 2 φ 2 ∂ 2 φ 2  1 + dx. E(φ,2,2) (t) := 2 R ∂x2 ∂x∂t Proposition 4.18. Let (φ, ψ) : R1,1 → Rq be a Dirac-wave map. Then the following inequality holds Z Z  d (4.21) E(φ,2,2) (t) ≤ C E(φ,2,2) (t) e(φ)dx + E(φ,2,2) (t) + (|ψ|2β + |∇ψ|4β )dx , dt R R where the positive constant C depends on N and the initial data. Proof. Using the extrinsic version for Dirac-wave maps (4.6) we calculate Z d 2 φidx E (t) = h∂x (φ), ∂xt dt (φ,2,2) R Z Z 2 2 φidx = h∂x (II(dφ, dφ)), ∂xt φidx + h∂x (P (II(ψ, dφ(∂t )), i∂t · ψ)), ∂xt R R Z 2 φidx. − h∂x (P (II(ψ, dφ(∂x )), i∂x · ψ)), ∂xt R

It is well-known that the contribution containing the second fundamental form can be estimated as 2 |h∂x (II(dφ, dφ)), ∂xt φi| ≤ C|dφ|3 |d2 φ| such that

|dφ|3 |d2 φ|

L1 (R)



≤ |dφ|3 L2 (R) |d2 φ| L2 (R)



≤C d|dφ|3 32 |d2 φ| L2 (R) L (R)



2 2

≤C |dφ| |d φ| 23 |d2 φ| L2 (R) L (R)

2 2

2 ≤C |dφ| L2 (R) |d φ| L2 (R) ,

20

VOLKER BRANDING

where we used the Sobolev embedding theorem in one dimension in the second step. Regarding the other two terms we note 2 |h∂x (P (II(ψ, dφ(∂t )), i∂t · ψ)), ∂xt φi| ≤C(|ψ|2β |dφ|2 |d2 φ| + |∇ψ|β |ψ|β |dφ||d2 φ| + |ψ|2β |d2 φ|2 )

≤C(|d2 φ|2 + |ψ|4β |dφ|4 + |∇ψ|2β |ψ|2β |dφ|2 ) ≤C(|d2 φ|2 + |ψ|2β + |∇ψ|4β ) and the remaining term can be estimated by the same method. By adding up the various contributions we obtain the claim.  Corollary 4.19. Let (φ, ψ) : R1,1 → Rq be a Dirac-wave map. Then the following estimate holds E(φ,2,2) (t) ≤ Cf (t),

(4.22)

where the positive constant C depends on N and the initial data and the function f (t) is finite for finite values of t. Proof. By the last Proposition we obtain an inequality of the form d E (t) ≤ C(eCt E(φ,2,2) (t) + E(φ,2,2) (t) + eCt ). dt (φ,2,2) The claim follows by the Gronwall-Lemma.



At this point we have obtained sufficient control over the pair (φ, ψ) and its derivatives, which is what we need to obtain a global solution of the system (4.2), (4.3). In order to obtain a unique solution we will need the following Proposition 4.20 (Uniqueness). Suppose we have two solutions of the Dirac-wave map system (4.2), (4.3). If their initial data coincides, then they coincide for all times. Proof. We make use of the extrinsic version of the Dirac-wave map system (4.6), (4.7). Suppose that u, v : R1,1 → Rq both solve (4.6) and ψ, ξ ∈ Γ(ΣR1,1 ⊗ Rq ) both solve (4.7), where ψ is a vector spinor along u and ξ a vector spinor along v. Since we have gained control over the L2 -norm of the second derivatives of φ, we have a pointwise bound on its first derivatives. We set w := u − v, η := ψ − ξ and calculate Z Z ∂w d1 2 |dw| dx = h , widx dt 2 R R ∂t Z ∂w = h , II(u)(du, du) − II(v)(dv, dv)idx R ∂t Z ∂w , P (u)(II(u)(ej · ψ, du(ej )), iψ) − P (v)(II(v)(ej · ξ, dv(ej )), iξ)idx. εj h + ∂t R We rewrite the first term on the right hand side as follows II(u)(du, du) − II(v)(dv, dv) = II(u)(du, du) − II(v)(du, du) + II(v)(du + dv, dw). This yields ∂w , (II(u) − II(v))(du, du)i ≤ C|du|2 |w||dw|. ∂t Using the orthogonality ∂t ⊥ II we find h

∂w ∂u ∂u , II(v)(du, dw)i = h , II(v)(du, dw)i = h , (II(v) − II(u))(du, dw)i ≤ C|du|2 |w||dw|. ∂t ∂t ∂t The same argument also holds for the term involving dv. We rewrite h

P (u)(II(u)(ej · ψ, du(ej )), iψ) − P (v)(II(v)(ej · ξ, dv(ej )), iξ) =(P (u) − P (v))(II(u)(ej · ψ, du(ej )), iψ) + P (v)((II(u) − II(v))(ej · ψ, du(ej )), iψ) + P (v)(II(v)(ej · η, du(ej )), iψ) + P (v)(II(v)(ej · ξ, dw(ej )), iψ) + P (v)(II(v)(ej · ψ, du(ej )), iη)

ENERGY METHODS FOR DIRAC-TYPE EQUATIONS IN TWO-DIMENSIONAL MINKOWSKI SPACE

21

such that we may estimate |h

∂w ,P (u)(II(u)(ej · ψ, du(ej )), iψ) − P (v)(II(v)(ej · ξ, dv(ej )), iξ)i| ∂t  ≤ C |w||dw||ψ|2β |du| + |η|β |w||du||ψ|β + |w||dv||η|β |ψ|β ≤ C(|η|2β + |w|2 + |dw|2 ).

Regarding the spinors ψ, ξ we calculate Z Z d1 εj hII(u)(ej · ψ, du(ej )) − II(v)(ej · ξ, dv(ej )), ηi)dx. (h∂t · η, ηidx = dt 2 R R We may rewrite the right hand side as follows

II(u)(ej · ψ, du(ej )) − II(v)(ej · ξ, dv(ej )) =(II(u) − II(v))(ej · ψ, du(ej )) + II(v)(ej · η, dv(ej )) + II(v)(ej · ξ, dw(ej )). This allows us to obtain the following inequality Z Z Z d 2 2 |η| dx ≤ C (|η|β |w| + |η|β |dv| + |η|β |dw||ξ|)dx ≤ C (|η|2β + |w|2 + |dw|2 )dx. dt R β R R In total, we find d dt

Z

R

(|η|2β

2

2

+ |w| + |dw| )dx ≤ C

Z

R

(|η|2β + |w|2 + |dw|2 )dx

such that Z

(|η|2β + |w|2 + |dw|2 )dx ≤ eCt

R

Z

(|η0 |2β + |w0 |2 + |dw0 |2 )dx. R

Consequently, we have u = v and ψ = ξ for all t, whenever the initial data coincides.



In the end we obtain the following existence result Theorem 4.21. Let R1,1 be two-dimensional Minkowski space and (N, g) be a compact Riemannian manifold. Then for any given initial data of the regularity φ(0, x) =φ0 (x) ∈ H 2 (R, N ), ∂φ (0, x) =φ1 (x) ∈ H 1 (R, T N ), ∂t ψ(0, x) =ψ0 (x) ∈ H 1 (R, ΣR1,1 ⊗ φ∗ T N ) ∩ W 1,4 (R, ΣR1,1 ⊗ φ∗ T N ) the equations for Dirac-wave maps (4.2), (4.3) admit a global weak solution of the class φ ∈ H 2 (R1,1 , N ),

ψ ∈ H 1 (R1,1 , ΣR1,1 ⊗ φ∗ T N ) ∩ W 1,4 (R, ΣR1,1 ⊗ φ∗ T N ),

which is uniquely determined by the initial data. Proof. By the energy estimates derived throughout this section we have enough control on the right hand sides of (4.6) and (4.7) to extend the solution of the Dirac-wave map system for all times. The uniqueness follows from Proposition 4.20.  Remark 4.22. It is very desirable to get rid of the requirement ψ0 (x) ∈ W 1,4 (R, ΣR1,1 ⊗φ∗ T N ) in Theorem 4.21. However, in order to control the energy E(φ,2,2) (t) it seems to be necessary to control Lp norms of ψ with p ≥ 2. Remark 4.23. The statement of Theorem 4.21 still holds if we take into account an additional two-form as was done in [12] for the case of the domain being a closed Riemannian surface. Acknowledgements: The author gratefully acknowledges the support of the Austrian Science Fund (FWF) through the START-Project Y963-N35 of Michael Eichmair.

22

VOLKER BRANDING

References [1] Luis Alvarez-Gaum´e and Daniel Z. Freedman. Geometrical structure and ultraviolet finiteness in the supersymmetric σ-model. Comm. Math. Phys., 80(3):443–451, 1981. [2] Bernd Ammann and Nicolas Ginoux. Dirac-harmonic maps from index theory. Calc. Var. Partial Differential Equations, 47(3-4):739–762, 2013. [3] Alain Bachelot. Global Cauchy problem for semilinear hyperbolic systems with nonlocal interactions. Applications to Dirac equations. J. Math. Pures Appl. (9), 86(3):201–236, 2006. [4] Christian B¨ ar, Paul Gauduchon, and Andrei Moroianu. Generalized cylinders in semi-Riemannian and Spin geometry. Math. Z., 249(3):545–580, 2005. [5] Christian B¨ ar and Nicolas Ginoux. CCR- versus CAR-quantization on curved spacetimes. In Quantum field theory and gravity, pages 183–206. Birkh¨ auser/Springer Basel AG, Basel, 2012. [6] Christian B¨ ar, Nicolas Ginoux, and Frank Pf¨ affle. Wave equations on Lorentzian manifolds and quantization. ESI Lectures in Mathematics and Physics. European Mathematical Society (EMS), Z¨ urich, 2007. [7] Christian B¨ ar and Alexander Strohmaier. A rigorous geometric derivation of the chiral anomaly in curved backgrounds. Comm. Math. Phys., 347(3):703–721, 2016. [8] Helga Baum. Spin-Strukturen und Dirac-Operatoren u ¨ber pseudoriemannschen Mannigfaltigkeiten, volume 41 of Teubner-Texte zur Mathematik [Teubner Texts in Mathematics]. BSB B. G. Teubner Verlagsgesellschaft, Leipzig, 1981. With English, French and Russian summaries. [9] Helga Baum. Twistor spinors on Lorentzian manifolds, CR-geometry and Fefferman spaces. In Differential geometry and applications (Brno, 1998), pages 29–37. Masaryk Univ., Brno, 1999. [10] Volker Branding. The evolution equations for Dirac-harmonic maps. PhD thesis, 2012. [11] Volker Branding. On the evolution of regularized Dirac-harmonic maps from closed surfaces. arXiv:1406.6274, 2014. [12] Volker Branding. Magnetic Dirac-harmonic maps. Anal. Math. Phys., 5(1):23–37, 2015. [13] Volker Branding. Some aspects of Dirac-harmonic maps with curvature term. Differential Geom. Appl., 40:1–13, 2015. [14] Volker Branding. Dirac-harmonic maps with torsion. Commun. Contemp. Math., 18(4):1550064, 19, 2016. [15] Volker Branding. Energy estimates for the supersymmetric nonlinear sigma model and applications. Potential Anal., 45(4):737–754, 2016. [16] Volker Branding. The evolution equations for regularized Dirac-geodesics. J. Geom. Phys., 100:1–19, 2016. [17] Volker Branding. Nonlinear Dirac equations, monotonicity formulas and Liouville theorems. arXiv preprint arXiv:1605.03453, 2016. [18] Volker Branding. On the full bosonic string from Minkowski space to Riemannian manifolds. J. Math. Anal. Appl., 451(2):858–872, 2017. [19] Q. Chen, J. Jost, and G. Wang. Liouville theorems for Dirac-harmonic maps. J. Math. Phys., 48(11):113517, 13, 2007. [20] Qun Chen, J¨ urgen Jost, Jiayu Li, and Guofang Wang. Dirac-harmonic maps. Math. Z., 254(2):409–432, 2006. [21] Qun Chen, J¨ urgen Jost, Linlin Sun, and Miaomiao Zhu. Dirac-geodesics and their heat flows. Calc. Var. Partial Differential Equations, 54(3):2615–2635, 2015. [22] Qun Chen, J¨ urgen Jost, and Guofang Wang. Nonlinear Dirac equations on Riemann surfaces. Ann. Global Anal. Geom., 33(3):253–270, 2008. [23] Jo˜ ao-Paulo Dias and M´ ario Figueira. Time decay for the solutions of a nonlinear Dirac equation in one space dimension. Ricerche Mat., 35(2):309–316, 1986. [24] Xiaoli Han. Dirac-wave maps. Calc. Var. Partial Differential Equations, 23(2):193–204, 2005. [25] Hyungjin Huh. Global strong solution to the Thirring model in critical space. J. Math. Anal. Appl., 381(2):513–520, 2011. [26] Hyungjin Huh. Remarks on nonlinear Dirac equations in one space dimension. Commun. Korean Math. Soc., 30(3):201–208, 2015. [27] J¨ urgen Jost, Lei Liu, and Miaomiao Zhu. A global weak solution of the Dirac-harmonic map flow. Annales de l’Institut Henri Poincare (C) Non Linear Analysis, 2017. [28] Marie-Am´elie Lawn. Immersions of Lorentzian surfaces in R2,1 . J. Geom. Phys., 58(6):683–700, 2008. [29] Shuji Machihara and Takayuki Omoso. The explicit solutions to the nonlinear Dirac equation and DiracKlein-Gordon equation. Ric. Mat., 56(1):19–30, 2007. [30] Thomas H. Parker. Gauge theories on four-dimensional Riemannian manifolds. Comm. Math. Phys., 85(4):563–602, 1982. [31] Jalal Shatah and Michael Struwe. Geometric wave equations, volume 2 of Courant Lecture Notes in Mathematics. New York University, Courant Institute of Mathematical Sciences, New York; American Mathematical Society, Providence, RI, 1998. [32] Mario Soler. Classical, stable, nonlinear spinor field with positive rest energy. Phys. Rev. D, 1:2766–2769, May 1970. [33] Walter E Thirring. A soluble relativistic field theory. Annals of Physics, 3(1):91–112, 1958.

ENERGY METHODS FOR DIRAC-TYPE EQUATIONS IN TWO-DIMENSIONAL MINKOWSKI SPACE

23

[34] Changyou Wang and Deliang Xu. Regularity of Dirac-harmonic maps. Int. Math. Res. Not. IMRN, (20):3759– 3792, 2009. [35] Johannes Wittmann. Short time existence of the heat flow for Dirac-harmonic maps on closed manifolds. arXiv:1705.08935, 2017. [36] Yongqian Zhang and Qin Zhao. Global solution to nonlinear Dirac equation for Gross-Neveu model in 1 + 1 dimensions. Nonlinear Anal., 118:82–96, 2015. University of Vienna, Faculty of Mathematics, Oskar-Morgenstern-Platz 1, 1090 Vienna, Austria, E-mail address: [email protected]