Skip to main content

Guide to Elliptic Boundary Value Problems for Dirac-Type Operators

  • Chapter
  • First Online:
Arbeitstagung Bonn 2013

Part of the book series: Progress in Mathematics ((PM,volume 319))

Abstract

We present an introduction to boundary value problems for Dirac-type operators on complete Riemannian manifolds with compact boundary. We introduce a very general class of boundary conditions which contain local elliptic boundary conditions in the sense of Lopatinski and Shapiro as well as the Atiyah–Patodi–Singer boundary conditions. We discuss boundary regularity of solutions and also spectral and index theory. The emphasis is on providing the reader with a working knowledge.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 99.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 129.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 129.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Here \([D,f] = D \circ (f \cdot \mathop{\mathrm{id}}\nolimits _{E}) - (f \cdot \mathop{\mathrm{id}}\nolimits _{F}) \circ D\).

  2. 2.

    The η-invariant of A is defined as the value of the meromorphic extension of η(s) =  λ ≠ 0sign(λ) | λ | s at s = 0, see [APS]. Here the sum is taken over all nonzero eigenvalues of A taking multiplicities into account. Hence the η-invariant is a measure for the asymmetry of the spectrum.

  3. 3.

    If χ commutes with σ D (ν ), then B χ and B χ are adjoint to each other.

References

  1. T. Ackermann, J. Tolksdorf, The generalized Lichnerowicz formula and analysis of Dirac operators. J. Reine Angew. Math. 471, 23–42 (1996)

    MathSciNet  MATH  Google Scholar 

  2. M. Atiyah, V. Patodi, I. Singer, Spectral asymmetry and Riemannian Geometry. I. Math. Proc. Camb. Philos. Soc. 77, 43–69 (1975)

    Article  MathSciNet  MATH  Google Scholar 

  3. C. Bär, Real killing spinors and holonomy. Commun. Math. Phys. 154, 509–521 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  4. C. Bär, W. Ballmann, Boundary value problems for elliptic differential operators of first order, in Survey in Differential Geometry, vol. 17, ed. by H.-D. Cao, S.-T. Yau (International Press, Somerville, 2012), pp. 1–78

    Google Scholar 

  5. W. Ballmann, J. Brüning, G. Carron, Regularity and index theory for Dirac-Schrödinger systems with Lipschitz coefficients. J. Math. Pures Appl. 89, 429–476 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  6. B. Booss–Bavnbek, K. Wojciechowski, Elliptic Boundary Problems for Dirac Operators (Birkhäuser, Boston, 1993)

    Google Scholar 

  7. D. Freed, Two index theorems in odd dimensions. Commun. Anal. Geom. 6, 317–329 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  8. T. Friedrich, Der erste Eigenwert des Dirac-Operators einer kompakten, Riemannschen Mannigfaltigkeit nichtnegativer Skalarkrümmung. Math. Nachr. 97, 117–146 (1980)

    Article  MathSciNet  MATH  Google Scholar 

  9. P. Gilkey, Invariance Theory, the Heat Equation, and the Atiyah-Singer Index Theorem (Publish or Perish, Wilmington, 1984)

    MATH  Google Scholar 

  10. M. Gromov, H.B. Lawson, Positive scalar curvature and the Dirac operator on complete Riemannian manifolds. Inst. Hautes Études Sci. Publ. Math. 58 (1983), 83–196 (1984)

    MathSciNet  MATH  Google Scholar 

  11. N. Große, R. Nakad, Boundary value problems for noncompact boundaries of spinc manifolds and spectral estimates. arxiv:1207.4568[v2] Proc. Lond. Math. Soc. (3)109 (2014), 4, 946–974

    Google Scholar 

  12. N. Higson, A note on the cobordism invariance of the index. Topology 30, 439–443 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  13. O. Hijazi, S. Montiel, A. Roldán, Eigenvalue boundary problems for the Dirac operator. Commun. Math. Phys. 231, 375–390 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  14. T. Kato, Perturbation Theory for Linear Operators. 2nd edn. (Springer, Berlin, 1986)

    Google Scholar 

  15. H.B. Lawson, M.-L. Michelsohn, Spin Geometry (Princeton University Press, Princeton, 1989)

    MATH  Google Scholar 

  16. R.S. Palais, Seminar on the Atiyah-Singer Index Theorem. With contributions by M. F. Atiyah, A. Borel, E. E. Floyd, R. T. Seeley, W. Shih, and R. Solovay (Princeton University Press, Princeton, 1965)

    Google Scholar 

  17. J. Roe, Partitioning noncompact manifolds and the dual Toeplitz problem, in Operator Algebras and Applications, vol. 1. London Mathematical Society Lecture Note Series, vol. 135 (Cambridge University Press, Cambridge, 1988), pp. 187–228

    Google Scholar 

  18. R.T. Seeley, Complex powers of an elliptic operator. Proc. Symp. Pure Math. 10, 288–307 (1967)

    Article  MathSciNet  MATH  Google Scholar 

  19. M.E. Taylor, Partial Differential Equations. I. Basic Theory (Springer, New York, 1996)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Werner Ballmann .

Editor information

Editors and Affiliations

Additional information

Dedicated to the memory of Friedrich Hirzebruch

Appendices

Appendix 1: Dirac Operators in the Sense of Gromov and Lawson

Here we discuss an important subclass of Dirac-type operators. Note that the connection in Corollary 2.4 is not metric, in general.

Definition A.1.

A formally self-adjoint operator D: C (M, E) → C (M, E) of Dirac type is called a Dirac operator in the sense of Gromov and Lawson if there exists a metric connection ∇ on E such that

  1. (1)

    \(D =\sum \nolimits _{j}\sigma _{D}(e_{j}^{{\ast}}) \circ \nabla _{e_{j}}\), for any local orthonormal tangent frame (e 1, , e n );

  2. (2)

    the principal symbol σ D of D is parallel with respect to ∇ and to the Levi-Civita connection.

This is equivalent to the definition of generalized Dirac operators in [GL, Sect. 1] or to Dirac operators on Dirac bundles in [LM, Chap. II, § 5].

Remark A.2.

For a Dirac operator in the sense of Gromov and Lawson, the connection ∇ in Definition A.1 and the connection in the Weitzenböck formula (9) coincide and are uniquely determined by these properties. We will call ∇ the connection associated with the Dirac operator D. Moreover, the endomorphism field \(\mathcal{K}\) in the Weitzenböck formula takes the form

$$\displaystyle{ \mathcal{K} = \frac{1} {2}\sum _{i,j}\sigma _{D}(e_{i}^{{\ast}}) \circ \sigma _{ D}(e_{j}^{{\ast}}) \circ R^{\nabla }(e_{ i},e_{j}) }$$

where R is the curvature tensor of ∇. See [GL, Proposition 2.5] for a proof.

Next, we show how to explicitly construct an adapted operator on the boundary satisfying (13) for a Dirac operator in the sense of Gromov and Lawson. Let ∇ be the associated connection. Along the boundary we define

$$\displaystyle{ A_{0}:=\sigma _{D}(\nu ^{\flat })^{-1}D -\nabla _{\nu } =\sigma _{ D}(\nu ^{\flat })^{-1}\sum _{ j=2}^{n}\sigma _{ D}(e_{j}^{{\ast}})\nabla _{ e_{j}}. }$$
(23)

Here (e 2, , e n ) is any local tangent frame for ∂ M. Then A 0 is a first-order differential operator acting on section of E |  ∂ M  → ∂ M with principal symbol \(\sigma _{A_{0}}(\xi ) =\sigma _{D}(\nu ^{\flat })^{-1}\sigma _{D}(\xi )\) as required for an adapted boundary operator. From the Weitzenböck formula (9) we get, using Proposition 2.1 twice, once for D and once for ∇, for all \(\Phi,\Psi \in C_{c}^{\infty }(M,E)\):

$$\displaystyle\begin{array}{rcl} 0& =& \int _{M}\big(\langle D^{2}\Phi,\Psi \rangle -\langle \nabla ^{{\ast}}\nabla \Phi,\Psi \rangle -\langle \mathcal{K}\Phi,\Psi \rangle \big)\mathop{\mathrm{dV}}\nolimits \\ & =& \int _{M}\big(\langle D\Phi,D\Psi \rangle -\langle \nabla \Phi,\nabla \Psi \rangle -\langle \mathcal{K}\Phi,\Psi \rangle \big)\mathop{\mathrm{dV}}\nolimits \\ & & \quad +\int _{\partial M}\big(-\langle \sigma _{D}(\nu ^{\flat })D\Phi,\Psi \rangle +\langle \sigma _{ \nabla ^{{\ast}}}(\nu ^{\flat })\nabla \Phi,\Psi \rangle \big)\mathop{\mathrm{dS}}\nolimits.{}\end{array}$$
(24)

For the boundary contribution we have

$$\displaystyle\begin{array}{rcl} -\langle \sigma _{D}(\nu ^{\flat })D\Phi,\Psi \rangle +\langle \sigma _{ \nabla ^{{\ast}}}(\nu ^{\flat })\nabla \Phi,\Psi \rangle & =& \langle \sigma _{ D}(\nu ^{\flat })^{-1}D\Phi,\Psi \rangle -\langle \nabla \Phi,\sigma _{ \nabla }(\nu ^{\flat })\Psi \rangle \\ & =& \langle \sigma _{D}(\nu ^{\flat })^{-1}D\Phi,\Psi \rangle -\langle \nabla \Phi,\nu ^{\flat } \otimes \Psi \rangle \\ & =& \langle \sigma _{D}(\nu ^{\flat })^{-1}D\Phi,\Psi \rangle -\langle \nabla _{\nu }\Phi,\Psi \rangle \\ & =& \langle A_{0}\Phi,\Psi \rangle. {}\end{array}$$
(25)

Inserting (25) into (24) we get

$$\displaystyle{ \int _{M}\big(\langle D\Phi,D\Psi \rangle -\langle \nabla \Phi,\nabla \Psi \rangle -\langle \mathcal{K}\Phi,\Psi \rangle \big)\mathop{\mathrm{dV}}\nolimits = -\int _{\partial M}\langle A_{0}\phi,\psi \rangle \mathop{\mathrm{dS}}\nolimits }$$
(26)

where \(\phi:= \Phi \vert _{\partial M}\) and \(\psi:= \Psi \vert _{\partial M}\). Since the left-hand side of (26) is symmetric in \(\Phi \) and \(\Psi \), the right-hand side is symmetric as well, hence A 0 is formally self-adjoint. This shows that A 0 is an adapted boundary operator for D.

In general, A 0 does not anticommute with σ D (ν ) however. We will rectify this by adding a suitable zero-order term. First, let us compute the anticommutator of A 0 and σ D (ν ):

$$\displaystyle\begin{array}{rcl} \{\sigma _{D}(\nu ^{\flat }),A_{ 0}\}\phi & =& \sum _{j=2}^{n}\sigma _{ D}(e_{j}^{{\ast}})\nabla _{ e_{j}}\phi +\sigma _{D}(\nu ^{\flat })^{-1}\sum _{ j=2}^{n}\sigma _{ D}(e_{j}^{{\ast}})\nabla _{ e_{j}}(\sigma _{D}(\nu ^{\flat })\phi ) {}\\ & =& \sum _{j=2}^{n}\Big(\sigma _{ D}(e_{j}^{{\ast}})\nabla _{ e_{j}}\phi +\sigma _{D}(\nu ^{\flat })^{-1}\sigma _{ D}(e_{j}^{{\ast}})\sigma _{ D}(\nu ^{\flat })\nabla _{ e_{j}}\phi {}\\ & & \quad +\sigma _{D}(\nu ^{\flat })^{-1}\sigma _{ D}(e_{j}^{{\ast}})\sigma _{ D}(\nabla _{e_{j}}\nu ^{\flat })\phi \Big) {}\\ & =& \sigma _{D}(\nu ^{\flat })^{-1}\sum _{ j=2}^{n}\sigma _{ D}(e_{j}^{{\ast}})\sigma _{ D}(\nabla _{e_{j}}\nu ^{\flat })\phi. {}\\ \end{array}$$

Now ∇⋅  ν is the negative of the Weingarten map of the boundary with respect to the normal field ν. We choose the orthonormal tangent frame (e 2, , e n ) to consist of eigenvectors of the Weingarten map. The corresponding eigenvalues κ 2, , κ n are the principal curvatures of ∂ M. We get

$$\displaystyle{ \sum _{j=2}^{n}\sigma _{ D}(e_{j}^{\flat })\sigma _{ D}(\nabla _{e_{j}}\nu ^{\flat }) = -\sum _{ j=2}^{n}\sigma _{ D}(e_{j}^{\flat })\sigma _{ D}(\kappa _{j}e_{j}^{\flat }) =\sum _{ j=2}^{n}\kappa _{ j} = (n - 1)H, }$$

where H is the mean curvature of ∂ M with respect to ν. Therefore,

$$\displaystyle{ \{\sigma _{D}(\nu ^{\flat }),A_{ 0}\} = (n - 1)H\sigma _{D}(\nu ^{\flat })^{-1} = -(n - 1)H\sigma _{ D}(\nu ^{\flat }). }$$

Since clearly

$$\displaystyle{\{\sigma _{D}(\nu ^{\flat }),(n - 1)H\} = 2(n - 1)H\sigma _{ D}(\nu ^{\flat }),}$$

the operator

$$\displaystyle{ A:= A_{0} + \frac{n - 1} {2} H =\sigma _{D}(\nu ^{\flat })^{-1}D -\nabla _{\nu } + \frac{n - 1} {2} H }$$

is an adapted boundary operator for D satisfying (13). From (26) we also have

$$\displaystyle{ \int _{M}\big(\langle D\Phi,D\Psi \rangle -\langle \nabla \Phi,\nabla \Psi \rangle -\langle \mathcal{K}\Phi,\Psi \rangle \big)\mathop{\mathrm{dV}}\nolimits =\int _{\partial M}\langle (\tfrac{n-1} {2} H - A)\phi,\psi \rangle \mathop{\mathrm{dS}}\nolimits. }$$
(27)

Definition A.3.

For a Dirac operator D in the sense of Gromov and Lawson as above, we call A the canonical boundary operator for D.

Remark A.4.

The canonical boundary operator A is again a Dirac operator in the sense of Gromov and Lawson. Namely, define a connection on E |  ∂ M by

$$\displaystyle{\nabla _{X}^{\partial }\phi:= \nabla _{ X}\phi + \frac{1} {2}\sigma _{D}(\nu ^{\flat })^{-1}\sigma _{ D}(\nabla _{X}\nu ^{\flat })\phi.}$$

The Clifford relations (6) show that the term σ D (ν )−1 σ D (∇ X ν ) = σ D (ν ) σ D (∇ X ν ) is skewhermitian, hence ∇ is a metric connection. By (23), \(A_{0} =\sum _{ j=2}^{n}\sigma _{A_{0}}(e_{j}^{{\ast}}) \circ \nabla _{e_{j}}\). This, \(\sigma _{A_{0}} =\sigma _{A}\), and

$$\displaystyle{\sum _{j=2}^{n}\sigma _{ A_{0}}(e_{j}^{{\ast}})\sigma _{ D}(\nu ^{\flat })^{{\ast}}\sigma _{ D}(\nabla _{e_{j}}\nu ^{\flat }) = \frac{n - 1} {2} \,H}$$

show that

$$\displaystyle{A =\sum _{ j=2}^{n}\sigma _{ A}(e_{j}^{{\ast}}) \circ \nabla _{ e_{j}}^{\partial }.}$$

Moreover, a straightforward computation using the Gauss equation for the Levi-Civita connections ∇ X ξ = ∇ X ξξ(∇ X ν)ν shows that σ A is parallel with respect to the boundary connections ∇.

Remark A.5.

The triangle inequality and the Cauchy–Schwarz inequality show

$$\displaystyle\begin{array}{rcl} \vert D\Phi \vert ^{2}& =& \big\vert \sum _{ j=1}^{n}\sigma _{ D}(e_{j}^{\flat })\nabla _{ e_{j}}\Phi \big\vert ^{2} \leq \big (\sum _{ j=1}^{n}\vert \sigma _{ D}(e_{j}^{\flat })\nabla _{ e_{j}}\Phi \vert \big)^{2} \\ & \leq & n \cdot \sum _{j=1}^{n}\vert \sigma _{ D}(e_{j}^{\flat })\nabla _{ e_{j}}\Phi \vert ^{2} = n \cdot \sum _{ j=1}^{n}\langle \sigma _{ D}(e_{j}^{\flat })^{{\ast}}\sigma _{ D}(e_{j}^{\flat })\nabla _{ e_{j}}\Phi,\nabla _{e_{j}}\Phi \rangle \\ & =& n \cdot \sum _{j=1}^{n}\vert \nabla _{ e_{j}}\Phi \vert ^{2} = n \cdot \vert \nabla \Phi \vert ^{2}, {}\end{array}$$
(28)

for any orthonormal tangent frame (e 1, , e n ) and all \(\Phi \in C^{\infty }(M,E)\).

When does equality hold? Equality in the Cauchy–Schwarz inequality implies that all summands \(\vert \sigma _{D}(e_{j}^{\flat })\nabla _{e_{j}}\Phi \vert \) are equal, i.e., \(\vert \sigma _{D}(e_{j}^{\flat })\nabla _{e_{j}}\Phi \vert = \vert \sigma _{D}(e_{1}^{\flat })\nabla _{e_{1}}\Phi \vert \). Equality in the triangle inequality then implies \(\sigma _{D}(e_{j}^{\flat })\nabla _{e_{j}}\Phi =\sigma _{D}(e_{1}^{\flat })\nabla _{e_{1}}\Phi \) for all j. Thus

$$\displaystyle{\sigma _{D}(e_{1}^{\flat })\nabla _{ e_{1}}\Phi = \frac{1} {n}\sum _{j=1}^{n}\sigma _{ D}(e_{j}^{\flat })\nabla _{ e_{j}}\Phi = \frac{1} {n}D\Phi,}$$

hence \(\nabla _{e_{1}}\Phi = \frac{1} {n}\sigma _{D}(e_{1}^{\flat })^{{\ast}}D\Phi \). Since e 1 is arbitrary, this shows the twistor equation

$$\displaystyle{ \nabla _{X}\Phi = \frac{1} {n}\sigma _{D}(X^{\flat })^{{\ast}}D\Phi, }$$
(29)

for all vector fields X on M. Conversely, if \(\Phi \) solves the twistor equation, one sees directly that equality holds in (28).

Inserting (28) into (27) yields

$$\displaystyle{ \tfrac{n-1} {n} \int _{M}\vert D\Phi \vert ^{2}\mathop{\mathrm{dV}}\nolimits \geq \int _{ M}\langle \mathcal{K}\Phi,\Phi \rangle \mathop{\mathrm{dV}}\nolimits +\int _{\partial M}\langle (\tfrac{n-1} {2} H - A)\phi,\phi \rangle \mathop{\mathrm{dS}}\nolimits, }$$

for all \(\Phi \in C^{\infty }_{c}(M,E)\), where \(\phi:= \Phi \vert _{\partial M}\). Moreover, equality holds if and only if \(\Phi \) solves the twistor equation (29).

Appendix 2: Proofs of Some Auxiliary Results

In this section we collect the proofs of some of the auxiliary results.

Proof of Proposition 2.3.

We start by choosing an arbitrary connection \(\bar{\nabla }\) on E and define

$$\displaystyle{ \bar{D}: C^{\infty }(M,E) \rightarrow C^{\infty }(M,F),\quad \bar{D}\Phi:=\sum \nolimits _{ j}\sigma _{D}(e_{j}^{{\ast}})\bar{\nabla }_{ e_{j}}\Phi. }$$

Then \(\bar{D}\) has the same principal symbol as D and, therefore, the difference \(S:= D -\bar{ D}\) is of order 0. In other words, S is a field of homomorphisms from E to F.

Since \(\mathcal{A}_{D}\) is onto, the restriction \(\mathcal{A}\) of \(\mathcal{A}_{D}\) to the orthogonal complement of the kernel of \(\mathcal{A}_{D}\) is a fiberwise isomorphism. We put \(V:= \mathcal{A}^{-1}(S)\) and define a new connection by

$$\displaystyle{\nabla:=\bar{ \nabla } + V.}$$

We compute

$$\displaystyle\begin{array}{rcl} \sum \nolimits _{j}\sigma _{D}(e_{j}^{{\ast}}) \circ \nabla _{ e_{j}}& =& \sum \nolimits _{j}\sigma _{D}(e_{j}^{{\ast}}) \circ \bar{\nabla }_{ e_{j}} +\sum \nolimits _{j}\sigma _{D}(e_{j}^{{\ast}}) \circ V (e_{ j}) {}\\ & =& \bar{D} + \mathcal{A}_{D}(V ) {}\\ & =& \bar{D} + S = D. {}\\ \end{array}$$

 □ 

Proof of Proposition 3.1.

Let \(\widetilde{\nabla }\) be any metric connection on E. Then \(F:= D^{{\ast}}D -\widetilde{\nabla }^{{\ast}}\widetilde{\nabla }\) is formally self-adjoint. Since both, D D and \(\widetilde{\nabla }^{{\ast}}\widetilde{\nabla }\), have the same principal symbol \(-\vert \xi \vert ^{2} \cdot \mathop{\mathrm{id}}\nolimits\), the operator F is of order at most one. Any other metric connection ∇ on E is of the form \(\nabla =\widetilde{ \nabla } + B\) where B is a 1-form with values in skewhermitian endomorphisms of E. Hence

$$\displaystyle{ D^{{\ast}}D = (\nabla - B)^{{\ast}}(\nabla - B) + F = \nabla ^{{\ast}}\nabla \mathop{\underbrace{-\nabla ^{{\ast}}B - B^{{\ast}}\nabla + B^{{\ast}}B + F}}\limits _{ =:\,\mathcal{K}}. }$$

In general, \(\mathcal{K}\) is of first order and we need to show that there is a unique B such that \(\mathcal{K}\) is of order zero. Since B B is of order zero, \(\mathcal{K}\) is of order zero if and only if F −∇ BB ∇ is of order zero, i.e., if and only if \(\sigma _{F}(\xi ) =\sigma _{\nabla ^{{\ast}}B+B^{{\ast}}\nabla }(\xi )\) for all ξ ∈ T M. We compute, using a local tangent frame e 1, , e n ,

$$\displaystyle\begin{array}{rcl} \left \langle \sigma _{\nabla ^{{\ast}}B+B^{{\ast}}\nabla }(\xi )\varphi,\psi \right \rangle & =& \left \langle \big(\sigma _{\nabla ^{{\ast}}}(\xi ) \circ B + B^{{\ast}}\circ \sigma _{ \nabla }(\xi )\big)\varphi,\psi \right \rangle {}\\ & =& -\left \langle B\varphi,\sigma _{\nabla }(\xi )\psi \right \rangle + \left \langle \sigma _{\nabla }(\xi )\varphi,B\psi \right \rangle {}\\ & =& -\left \langle B\varphi,\xi \otimes \psi \right \rangle + \left \langle \xi \otimes \varphi,B\psi \right \rangle {}\\ & =& -\big\langle \sum _{i}e_{i}^{{\ast}}\otimes B_{ e_{i}}\varphi,\xi \otimes \psi \big\rangle +\big\langle \xi \otimes \varphi,\sum _{i}e_{i}^{{\ast}}\otimes B_{ e_{i}}\psi \big\rangle {}\\ & =& -\sum _{i}\langle e_{i}^{{\ast}},\xi \rangle \langle B_{ e_{i}}\varphi,\psi \rangle +\sum _{i}\langle e_{i}^{{\ast}},\xi \rangle \left \langle \varphi,B_{ e_{i}}\psi \right \rangle {}\\ & =& -\left \langle B_{\xi ^{\sharp }}\varphi,\psi \right \rangle + \left \langle \varphi,B_{\xi ^{\sharp }}\psi \right \rangle {}\\ & =& -2\left \langle B_{\xi ^{\sharp }}\varphi,\psi \right \rangle. {}\\ \end{array}$$

Hence, \(\sigma _{\nabla ^{{\ast}}B+B^{{\ast}}\nabla }(\xi ) = -2B_{\xi ^{\sharp }}\). Thus, \(\mathcal{K}\) is of order 0 if and only if

$$\displaystyle{B_{X} = -\tfrac{1} {2}\,\sigma _{F}(X^{b})}$$

for all X ∈ TM. Note that σ F (ξ) is indeed skewhermitian because F is formally self-adjoint. □ 

Proof of Lemma 3.2.

Since D is formally self-adjoint and of Dirac type,

$$\displaystyle{ -\sigma _{D}(\nu ^{\flat }) =\sigma _{ D}(\nu ^{\flat })^{{\ast}} =\sigma _{ D}(\nu ^{\flat })^{-1}, }$$
(30)

by (1) and (8). Let A 0 be adapted to D along ∂ M and ξ ∈ T x ∂ M, as usual extended to T x M by ξ(ν(x)) = 0. Then, again using (6) and (11),

$$\displaystyle\begin{array}{rcl} \sigma _{A_{0}}(\xi ) +\sigma _{D}(\nu (x)^{\flat })& \sigma _{ A_{0}}& (\xi )\sigma _{D}(\nu (x)^{\flat })^{{\ast}} {}\\ & =& \sigma _{D}(\nu (x)^{\flat })^{-1}\sigma _{ D}(\xi ) +\sigma _{D}(\xi )\sigma _{D}(\nu (x)^{\flat })^{{\ast}} {}\\ & =& \sigma _{D}(\nu (x)^{\flat })^{{\ast}}\sigma _{ D}(\xi ) +\sigma _{D}(\xi )^{{\ast}}\sigma _{ D}(\nu (x)^{\flat }) {}\\ & =& 2\langle \nu (x)^{\flat },\xi \rangle \cdot \mathop{\mathrm{id}}\nolimits _{ E} {}\\ & =& 0. {}\\ \end{array}$$

Hence 2S: = A 0 +σ D (ν )A 0 σ D (ν ) is of order 0, that is, S is a field of endomorphisms of E along ∂ M. Since A 0 is formally self-adjoint so is S and, by (30),

$$\displaystyle{ \sigma _{D}(\nu ^{\flat })2S =\sigma _{ D}(\nu ^{\flat })A_{ 0} + A_{0}\sigma _{D}(\nu ^{\flat }) = 2S\sigma _{ D}(\nu ^{\flat }). }$$

Hence A: = A 0S is adapted to D along ∂ M and

$$\displaystyle\begin{array}{rcl} \sigma _{D}(\nu ^{\flat })A + A\sigma _{ D}(\nu ^{\flat })& =& \sigma _{ D}(\nu ^{\flat })A_{ 0} + A_{0}\sigma _{D}(\nu ^{\flat }) -\sigma _{ D}(\nu ^{\flat })S - S\sigma _{ D}(\nu ^{\flat }) {}\\ & =& \sigma _{D}(\nu ^{\flat })\big(A_{ 0} -\sigma _{D}(\nu ^{\flat })A_{ 0}\sigma _{D}(\nu ^{\flat }) - 2S\big) {}\\ & =& \sigma _{D}(\nu ^{\flat })\big(A_{ 0} +\sigma _{D}(\nu ^{\flat })A_{ 0}\sigma _{D}(\nu ^{\flat })^{{\ast}}- 2S\big) {}\\ & =& 0. {}\\ \end{array}$$

 □ 

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Bär, C., Ballmann, W. (2016). Guide to Elliptic Boundary Value Problems for Dirac-Type Operators. In: Ballmann, W., Blohmann, C., Faltings, G., Teichner, P., Zagier, D. (eds) Arbeitstagung Bonn 2013. Progress in Mathematics, vol 319. Birkhäuser, Cham. https://doi.org/10.1007/978-3-319-43648-7_3

Download citation

Publish with us

Policies and ethics