Skip to main content
Log in

Dirac Operators on Hypersurfaces as Large Mass Limits

  • Published:
Communications in Mathematical Physics Aims and scope Submit manuscript

Abstract

We show that the eigenvalues of the intrinsic Dirac operator on the boundary of a Euclidean domain can be obtained as the limits of eigenvalues of Euclidean Dirac operators, either in the domain with a MIT-bag type boundary condition or in the whole space, with a suitably chosen zero order mass term.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. It was pointed out by one of the anonymous reviewers that some of the results of [14] on the self-adjointness of Dirac operators with boundary conditions depend on implicit assumptions, as the respective proofs of [14] use previous works which were found to contain flaws. We stress that we do not use any of the self-adjointness results stated in [14], and we only use some constructions of Sections 1–3 in [14], which are not concerned by the issues mentioned.

  2. In the language of [9, Theorem 15.19], the matrices \(\alpha _1,\dots ,\alpha _n,\mathrm {i}\alpha _{n+1}\) form a basis of a minimal representation of the Clifford algebra of \(\mathbb {R}^{1,n}\), while \(\theta \) is a real (if \(\theta ^2=1\)) or quaternionic (if \(\theta ^2=-1\)) charge conjugation, which is shown to exist for \(n-1\notin 4\mathbb {Z}+3\).

  3. See footnote 1 on page 10.

References

  1. Arrizabalaga, N., Le Treust, L., Mas, A., Raymond, N.: The MIT bag model as an infinite mass limit. J. Éc. Polytech. Math. 6, 329–365 (2019)

    Google Scholar 

  2. Arrizabalaga, N., Le Treust, L., Raymond, N.: On the MIT bag model in the non-relativistic limit. Commun. Math. Phys. 354, 641–669 (2017)

    Google Scholar 

  3. Bär, C., Ballmann, W.: Guide to boundary value problems for Dirac-type operators. In: Ballmann, W., et al. (eds.) Arbeitstagung Bonn 2013. Progress in Mathematics, vol. 319, pp. 43–80. Birkhäuser, Cham (2016)

    Google Scholar 

  4. Bär, C., Ballmann, W.: Boundary value problems for elliptic differential operators of first order. In: Cao, H.-D., Yau, S.-T. (eds.) Algebra and Geometry: In Memory of C. C. Hsiung. Surveys in Differential Geometry, vol. 17, pp. 1–78. International Press, Boston (2012)

    Google Scholar 

  5. Barbaroux, J.-M., Cornean, H.D., Le Treust, L., Stockmeyer, E.: Resolvent convergence to Dirac operators on planar domains. Ann. Henri Poincaré 20, 1877–1891 (2019)

    Google Scholar 

  6. Behrndt, J., Hassi, S., de Snoo, H., Wietsma, R.: Monotone convergence theorems for semibounded operators and forms with applications. Proc. R. Soc. Edinb. Sect. A 140, 927–951 (2010)

    Google Scholar 

  7. Bourguignon, J.-P., Hijazi, O., Milhorat, J.-L., Moroianu, A., Moroianu, S.: A Spinorial Approach to Riemannian and Conformal Geometry. EMS Monographs in Mathematics, vol. 6. EMS Publ. House, Zurich (2015)

    Google Scholar 

  8. de Wit, B., Smith, J.: Field Theory in Particle Physics, vol. 1. North-Holland, Amsterdam (1986)

    Google Scholar 

  9. Dereziński, J., Gérard, C.: Quantization and Quantum Fields. Cambridge University Press, Cambridge (2013)

    Google Scholar 

  10. Dolbeault, J., Esteban, M.J., Séré, E.: On the eigenvalues of operators with gaps. Application to Dirac operators. J. Funct. Anal. 174, 208–226 (2000)

    Google Scholar 

  11. Erdös, L., Solovej, J.P.: The kernel of Dirac operators on \(\mathbb{S}^3\) and \(\mathbb{R}^3\). Rev. Math. Phys. 13, 1247–1280 (2001)

    Google Scholar 

  12. Friedrich, T.: Dirac Operators in Riemannian Geometry. Graduate Studies in Mathematics, vol. 25. AMS, Providence, RI (2000)

    Google Scholar 

  13. Ginoux, N.: The Dirac Spectrum. Lecture Notes in Mathematics, vol. 1976. Springer, Berlin (2009)

    Google Scholar 

  14. Hijazi, O., Montiel, S., Zhang, X.: Dirac operator on embedded hypersurfaces. Math. Res. Lett. 8, 195–208 (2001)

    Google Scholar 

  15. Holzmann, M., Ourmières-Bonafos, T., Pankrashkin, K.: Dirac operators with Lorentz scalar shell interactions. Rev. Math. Phys. 30, 1850013 (2018)

    Google Scholar 

  16. Kato, T.: Perturbation Theory for Linear Operators, 2nd edn. Springer, Berlin (1980)

    Google Scholar 

  17. Lichnerowicz, A.: Spineurs harmoniques. C. R. Acad. Sci. Paris Sér. A 257, 7–9 (1963)

    Google Scholar 

  18. Lott, J.: Collapsing and Dirac-type operators. Geom. Dedicata 91, 175–196 (2002)

    Google Scholar 

  19. Reed, M., Simon, B.: Methods of Modern Mathematical Physics. I: Functional Analysis. Academic Press, New York (1972)

    Google Scholar 

  20. Reed, M., Simon, B.: Methods of Modern Mathematical Physics. IV: Analysis of Operators. Academic Press, New York (1978)

    Google Scholar 

  21. Roos, S.: The Dirac operator under collapse to a smooth limit space. Ann. Glob. Anal. Geom. (2019). https://doi.org/10.1007/s10455-019-09691-8

  22. Schrödinger, E.: Diracsches Elektron im Schwerefeld I. Sitzungsberichte Preuss. Akad. Wiss. Phys. Math. Kl. 11–12, 105–128 (1932) (Reprint available at http://edition-open-sources.org/sources/10/15/)

  23. Simon, B.: A canonical decomposition for quadratic forms with applications to monotone convergence theorems. J. Funct. Anal. 28, 377–385 (1978)

    Google Scholar 

  24. Stockmayer, E., Vugalter, S.: Infinite mass boundary conditions for Dirac operators. J. Spect. Theory 9, 569–600 (2019)

    Google Scholar 

  25. Thaller, B.: The Dirac Equation. Texts and Monographs in Physics. Springer, Berlin (1992)

    Google Scholar 

  26. Weidmann, J.: Stetige Abhängigkeit der Eigenwerte und Eigenfunktionen elliptischer Differentialoperatoren vom Gebiet. Math. Scand. 54, 51–69 (1984)

    Google Scholar 

Download references

Acknowledgements

The authors thank Christian Gérard for numerous useful discussions. A large part of this paper was written while Thomas Ourmières-Bonafos was supported by a public grant as part of the “Investissement d’avenir” project, reference ANR-11-LABX-0056-LMH, LabEx LMH, at the University Paris-Sud, and by the ANR "Défi des autres savoirs (DS10) 2017” programm, reference ANR-17-CE29-0004, project molQED. Konstantin Pankrashkin was in part supported by the PRC 1556 CNRS-RFBR 2017–2019 “Multi-dimensional semi-classical problems of condensed matter physics and quantum mechanics”. We also thank the two anonymous reviewers, whose valuable comments on the first version of the manuscript contributed to a significant improvement of the presentation.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Konstantin Pankrashkin.

Additional information

Communicated by R. Seiringer

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

A Basic Properties of Euclidean Dirac Operators

In the present section we would like to recall some key facts on the Euclidean Dirac operators \(B_{m,M}\) and \(A_m\). All the properties are very standard and well known for \(n\in \{2,3\}\), but we were not able to find a suitable single reference covering arbitrary dimensions.

The analysis of \(B_{m,M}\) will be provided in a slightly more general setting, with the hope that the constructions can be of use for other works. Denote

$$\begin{aligned} B_M:=B_{M,M}. \end{aligned}$$

Proposition A.1

Let \(V\in L^\infty (\mathbb {R}^n)\) be real-valued with compact support. Consider the linear operator \(C:=B_M+V\,\alpha _{n+1}\) in \(L^2(\mathbb {R}^n,\mathbb {C}^N)\) with domain \({\mathcal {D}}(C)=H^1(\mathbb {R}^n,\mathbb {C}^N)\), then:

  1. (a)

    The operator C is self-adjoint, and it is essentially self-adjoint on \(C^\infty _c(\mathbb {R}^n,\mathbb {C}^N)\).

  2. (b)

    The essential spectrum of C is \(\big (-\infty ,-|M|\,\big ]\,\cup \,\big [\,|M|,+\infty \big )\), and there are at most finitely many discrete eigenvalues in \(\big (-|M|,|M|\big )\).

  3. (c)

    Assume in addition that \(n\notin 4\mathbb {Z}\). If E is an eigenvalue of C, then \((-E)\) is also an eigenvalue of C of the same multiplicity.

In particular, the three assertions hold for \(C:=B_{m,M}\equiv B_M+(m-M)1_{\Omega }\alpha _{n+1}\).

Proof

  1. (a)

    As both C and \(B_0\) are symmetric and only differ by the bounded symmetric operator \((M+V)\alpha _{n+1}\), the (essential) self-adjointness of C is equivalent to that of \(B_0\). First, it is directly seen that \(B_0\) is symmetric. Let S be the restriction of \(B_0\) to \(C^\infty _c(\mathbb {R}^n,\mathbb {C}^N)\), then S is a densely defined symmetric operator, and by applying the definitions one sees that its adjoint \(S^*\) is given by

    $$\begin{aligned} S^*u=D_0 u, \quad {\mathcal {D}}(S^*)=\big \{u\in L^2(\mathbb {R}^n,\mathbb {C}^N):\, D_0 u\in L^2(\mathbb {R}^n,\mathbb {C}^N)\big \} \end{aligned}$$

with \(D_0\) acting in the sense of distributions. In order to complete the proof we simply need to show that \(S^*=B_0\). As both operators are given by the same differential expression \(D_0\) and \(S^*\) is clearly an extension of \(B_0\), one only needs to check the inclusion \({\mathcal {D}}(S^*)\subset {\mathcal {D}}(B_0)\). Let a function u belong to \({\mathcal {D}}(S^*)\), i.e.

$$\begin{aligned} u\in L^2(\mathbb {R}^n,\mathbb {C}^N) \text { and } -\mathrm {i}\sum _{j=1}^n \alpha _j \partial _j u\in L^2(\mathbb {R}^n,\mathbb {C}^N). \end{aligned}$$

We need to show that this implies \(u\in H^1(\mathbb {R}^n,\mathbb {C}^N)\). We remark first that the Fourier transform \(\widehat{u}\) of u satisfies \(\widehat{u}\in L^2(\mathbb {R}^n,\mathbb {C}^N)\) and \(\Gamma (\xi ) \widehat{u} \in L^2(\mathbb {R}^n,\mathbb {C}^N)\). The matrices \(\Gamma (\xi )\) are Hermitian, and, by construction, \(\Gamma (\xi )^2=|\xi |^2 I_N\). In particular, one can represent \(\Gamma (\xi )=|\xi | U(\xi )\) with the unitary matrices \(U(\xi ):=\Gamma (\xi /|\xi |)\). As the pointwise multiplication by \(U(\cdot )\) is an isomorphism of \(L^2(\mathbb {R}^n,\mathbb {C}^N)\), it follows that the condition \(\Gamma (\xi ) \widehat{u} \in L^2(\mathbb {R}^n,\mathbb {C}^N)\) yields \(|\xi | \widehat{u} \in L^2(\mathbb {R}^n,\mathbb {C}^N)\). Due to \(|\xi _j|\le |\xi |\) we obtain \(\xi _j \widehat{u} \in L^2(\mathbb {R}^n,\mathbb {C}^N)\) for each \(j=1,\dots n\). By applying the inverse Fourier transform this implies \(\partial _j u \in L^2(\mathbb {R}^n,\mathbb {C}^N)\) for each \(j=1,\dots n\), and then \(u\in H^1(\mathbb {R}^n,\mathbb {C}^N)\).

  1. (b)

    Let us start by showing the inclusion

    $$\begin{aligned} \big (-\infty ,-|M|\big ]\cup \big [|M|,+\infty \big )\subset {\mathrm {spec}}_{\mathrm{ess}}C. \end{aligned}$$
    (59)

Denote \(e_1:=(1,0,\dots ,0)\in \mathbb {R}^n\). For \(\rho >0\), denote during the proof

$$\begin{aligned} b_\rho :=\{x\in \mathbb {R}^n: |x|<\rho \}. \end{aligned}$$

Let \(\chi \in C^\infty _c(\mathbb {R}^n)\) with \(\chi =1\) in \(b_1\) and \(\chi =0\) outside \(b_2\), and \(E\in \mathbb {R}\) with \(|E|>|M|\). Let us choose \(\eta \in \mathbb {C}^N\) with

$$\begin{aligned} \varphi :=(\sqrt{E^2-M^2}\,\alpha _1+M\alpha _{n+1}+E\,I_N)\eta \ne 0, \end{aligned}$$

which is possible as the matrix \(\sqrt{E^2-M^2}\,\alpha _1+M\alpha _{n+1}+E\,I_N\) is non-zero (otherwise \(\alpha _1\) could not anticommute with \(\alpha _{n+1}\)). For large \(k\in \mathbb {N}\) consider the following functions \(u_k\in H^1(\mathbb {R}^n,\mathbb {C}^N)\):

$$\begin{aligned} u_k(x)&=\chi \Big (\dfrac{x}{k}-ke_1\Big ) e^{\mathrm {i}\sqrt{E^2-M^2}\, x_1} \varphi \\&\equiv \chi \Big (\dfrac{x}{k}-ke_1\Big ) e^{\mathrm {i}\sqrt{E^2-M^2}\, x_1} (\sqrt{E^2-M^2}\,\alpha _1+M\alpha _{n+1}+E\,I_N)\eta , \end{aligned}$$

then one easily checks that \(\Vert u_k\Vert _{L^2(\mathbb {R}^n,\mathbb {C}^N)}^2=c k^n>0\) with \(c=\Vert \chi \Vert _{L^2(\mathbb {R}^n,\mathbb {C}^N)}^2>0\) independent of k. On the other hand, for large k one has \(V=0\) on the support of \(u_k\), hence,

$$\begin{aligned} (C u_k)(x)&=-\mathrm {i}\sum _{j=1}^n \alpha _j\partial _j u_k(x) + M\alpha _{n+1} u_k (x)\\&= -\dfrac{\mathrm {i}}{k}\sum _{j=1}^n \alpha _j \partial _j \chi \Big (\dfrac{x}{k}-ke_1\Big ) e^{\mathrm {i}\sqrt{E^2-M^2}\, x_1} \varphi \\&\qquad +\chi \Big (\dfrac{x}{k}-ke_1\Big ) \Big (\sqrt{E^2-M^2}\,\alpha _1+M\alpha _{n+1}\Big ) e^{\mathrm {i}\sqrt{E^2-M^2}\, x_1}\varphi , \end{aligned}$$

and then \((C-E)u_k=v_k+w_k\) with

$$\begin{aligned} v_k(x)&=-\dfrac{\mathrm {i}}{k}\sum _{j=1}^n \alpha _j \partial _j \chi \Big (\dfrac{x}{k}-ke_1\Big ) e^{\mathrm {i}\sqrt{E^2-M^2}\, x_1} \varphi ,\\ w_k(x)&=\chi \Big (\dfrac{x}{k}-ke_1\Big ) e^{\mathrm {i}\sqrt{E^2-M^2}\, x_1}\Big (\sqrt{E^2-M^2}\,\alpha _1+M\alpha _{n+1}-E\,I_N\Big ) \varphi \\&\equiv \chi \Big (\dfrac{x}{k}-ke_1\Big )e^{\mathrm {i}\sqrt{E^2-M^2}\, x_1}\Big (\sqrt{E^2-M^2}\,\alpha _1+M\alpha _{n+1}-E\,I_N\Big )\\&\qquad \times \Big (\sqrt{E^2-M^2}\,\alpha _1+M\alpha _{n+1}+E\Big )\eta =0. \end{aligned}$$

One estimates easily \(\Vert v_k\Vert _{L^2(\mathbb {R}^n,\mathbb {C}^N)}^2=\mathcal {O}(k^{n-2})\) for \(k\rightarrow +\infty \), and this shows that \(\Vert (C-E)u_k\Vert _{L^2(\mathbb {R}^n,\mathbb {C}^N)}/\Vert u_k\Vert _{L^2(\mathbb {R}^n,\mathbb {C}^N)}=\mathcal {O}(1/k)\rightarrow 0\) and yields \(E\in {\mathrm {spec}}C\). It follows that \(\big (-\infty ,-|M|\big )\cup \big (|M|,+\infty \big )\subset {\mathrm {spec}}C\), and one can take the closure on the left-hand side as \({\mathrm {spec}}C\) is a closed set. Furthermore, as the set \(\big (-\infty ,-|M|\big ]\cup \big [|M|,+\infty \big )\) has no isolated points, it is included into the essential spectrum of C. Hence, the claim (59) is proved.

Now it remains to check that C has no essential spectrum in \(\big (-|M|,|M|\big )\) and that it has at most finitely many discrete eigenvalues, which will be done by an iterated application of the min–max principle. For \(E\in \mathbb {R}\) and a self-adjoint semibounded from below operator T we will denote

$$\begin{aligned} \mathcal {N}(E,T):=\#\big \{j\in \mathbb {N}: E_j(T)<E\big \}. \end{aligned}$$

In other words, if \({\mathrm {spec}}_{\mathrm{ess}}T\,\cap \, (-\infty ,E)\ne \emptyset \), then \(\mathcal {N}(E,T)=+\infty \), otherwise \(\mathcal {N}(E,T)\) is the number of eigenvalues of T in \((-\infty ,E)\), where each eigenvalue counted according to its multiplicity. In these terms, we simply need to show that \(\mathcal {N}(M^2,C^2)<+\infty \).

Choose \(r>0\) large such that the support of V is contained in \(b_r\), and then pick any \(R>r\) and real-valued functions \(\chi _1, \chi _2\in C^\infty (\mathbb {R}^2)\) with

$$\begin{aligned} \chi _1^2+\chi _2^2=1, \quad \chi _1=1 \text { in } b_r, \quad \chi _2=1 \text { in } \mathbb {R}^2{\setminus } b_R. \end{aligned}$$

Let \(u\in {\mathcal {D}}(C)\), then for each \(k\in \{1,2\}\) one also has \(\chi _k u \in {\mathcal {D}}(C)\) and \(C (\chi _k u)=\chi _k C u -\mathrm {i}\Gamma (\nabla \chi _k) u\), and

$$\begin{aligned} \big \Vert C(\chi _k u)\big \Vert ^2_{L^2(\mathbb {R}^n,\mathbb {C}^N)}&=\int _{\mathbb {R}^n} \Big (\chi _k^2|C u|^2+ \big |\mathrm {i}\Gamma (\nabla \chi _k) u\big |^2\Big )\,\mathrm {d}x + \mathcal {J}_k,\\ \mathcal {J}_k&=2 \mathfrak {R}\int _{\mathbb {R}^n} \, \big \langle \chi _k C u, -\mathrm {i}\Gamma (\nabla \chi _k) u\big \rangle _{\mathbb {C}^n}\,\mathrm {d}x\\&= \mathfrak {R}\int _{\mathbb {R}^n} \, \big \langle C u, -\mathrm {i}\Gamma (2\chi _k\nabla \chi _k) u\big \rangle _{\mathbb {C}^n}\,\mathrm {d}x = \mathfrak {R}\int _{\mathbb {R}^n} \, \big \langle C u,-\mathrm {i}\Gamma \big (\nabla (\chi _k^2)\big ) u\big \rangle _{\mathbb {C}^n}\,\mathrm {d}x. \end{aligned}$$

From \(\chi _1^2+\chi _2^2=1\) we infer \(\nabla (\chi _1^2+\chi _2^2)=0\) and then \(\mathcal {J}_1+\mathcal {J}_2=0\). Therefore,

$$\begin{aligned}&\big \Vert C(\chi _1 u)\big \Vert ^2_{L^2(\mathbb {R}^n,\mathbb {C}^N)}+\big \Vert C(\chi _2 u)\big \Vert ^2_{L^2(\mathbb {R}^n,\mathbb {C}^N)}\nonumber \\&\quad =\int _{\mathbb {R}^n} ( \chi _1^2+\chi _2^2) |C u|^2\,\mathrm {d}x +\int _{\mathbb {R}^2} \big (|\nabla \chi _1|^2+|\nabla \chi _2|^2\big )|u|^2\big )\,\mathrm {d}x\nonumber \\&\quad = \int _{\mathbb {R}^n} |Cu|^2\,\mathrm {d}x +\int _{\mathbb {R}^n} W|u|^2\,\mathrm {d}x, \quad W:=|\nabla \chi _1|^2+|\nabla \chi _2|^2. \end{aligned}$$
(60)

Recall that W is supported in \(\overline{b_R{\setminus } b_r}\), while the support of \(\chi _2 u\) does not intersect the support of V, which gives

$$\begin{aligned} \big \Vert C(\chi _2 u)\Vert _{L^2(\mathbb {R}^n,\mathbb {C}^N)}^2=\big \Vert B_M(\chi _2 u)\Vert _{L^2(\mathbb {R}^n,\mathbb {C}^N)}^2=\int _{\mathbb {R}^n} \big (|\nabla (\chi _2 u)|^2+M^2|\chi _2 u|^2\big )\,\mathrm {d}x \end{aligned}$$

(for \(u\in C^\infty _c(\mathbb {R}^n,\mathbb {C}^N)\) this is a simple integration by parts, and it is then extended by density to the whole of \({\mathcal {D}}(C)\) as \(C^\infty _c(\mathbb {R}^n,\mathbb {C}^N)\) is a domain of essential self-adjointness as shown above). This allows one to rewrite (60) as

$$\begin{aligned} \big \Vert C u\Vert _{L^2(\mathbb {R}^n,\mathbb {C}^N)}^2= & {} \int _{b_R} \big ( \big |C(\chi _1 u)\big |^2- W|\chi _1 u|^2\big )\,\mathrm {d}x\nonumber \\&+\int _{\mathbb {R}^n} \big (|\nabla (\chi _2 u)|^2+M^2|\chi _2 u|^2 - W |\chi _2 u|^2 \big )\,\mathrm {d}x. \end{aligned}$$
(61)

For \(v\in L^2(b_R,\mathbb {C}^N)\) let us denote by \(v'\) its extension by zero to the whole of \(\mathbb {R}^n\) and consider the following sesqulinear form \(s_1\) in \(L^2(b_R,\mathbb {C}^N)\):

$$\begin{aligned} s_1(v,v)= \int _{b_R} |C v'|^2\,\mathrm {d}x, \quad {\mathcal {D}}(s_1)=H^1_0(b_R,\mathbb {C}^N). \end{aligned}$$

The form \(s_1\) is clearly non-negative and densely defined. Let us show that it is also closed. Let \((v_k)\subset {\mathcal {D}}(S_1)\) and \(v\in L^2(b_R,\mathbb {C}^N)\) with \(\Vert v_k-v\Vert _{L^2(b_R,\mathbb {C}^N)}\rightarrow 0\) for \(k\rightarrow \infty \) and \(s_1(v_k-v_l,v_k-v_l)\equiv \Vert C(v'_k-v'_l)\Vert ^2_{L^2(\mathbb {R}^n,\mathbb {C}^N)}\rightarrow 0\) as \(k,l\rightarrow +\infty \). Then one has \(\Vert v'_k-v'\Vert _{L^2(\mathbb {R}^n,\mathbb {C}^N)}\rightarrow 0\) for \(k\rightarrow \infty \), and the closedness of the operator C implies that \(v'\in {\mathcal {D}}(C)\) with \(Cv'=\lim _{k\rightarrow \infty } Cv_k'\). Hence, the function v is such that its extension by zero belongs to \(H^1(\mathbb {R}^n,\mathbb {C}^N)\), and this shows \(v\in H^1_0(b_R,\mathbb {C}^N)\equiv {\mathcal {D}}(s_1)\) and then \(s_1(v_k-v,v_k-v)\equiv \Vert C(v'_k-v')\Vert ^2_{L^2(\mathbb {R}^n,\mathbb {C}^N)}\rightarrow 0\). The semiboundedness and closedness of \(s_1\) imply the existence of a self-adjoint operator \(S_1\) in \(L^2(b_R,\mathbb {C}^N)\) with \(S_1[v,v]=s_1(v,v)\) for \(v\in {\mathcal {Q}}(S_1)={\mathcal {D}}(s_1)\), and \(S_1\) has compact resolvent as \({\mathcal {Q}}(S_1)\) is compactly embedded into \(L^2(b_R,\mathbb {C}^N)\).

Denote by \(S_0\) the (scalar) Schrödinger operator \(-\Delta -W\) in \(L^2(\mathbb {R}^n)\), i.e.

$$\begin{aligned} S_0[u,u]=\int _{\mathbb {R}^n}\big ( |\nabla u|^2-W|u|^2\big )\,\mathrm {d}x, \quad {\mathcal {Q}}(S_0)=H^1(\mathbb {R}^n), \end{aligned}$$

and consider the linear map \(J:{\mathcal {D}}(C)\equiv {\mathcal {Q}}(C^2)\mapsto {\mathcal {Q}}(S_1)\times {\mathcal {Q}}(S_0\otimes I_N)\) given by \(J u=\big ((\chi _1 u)|_{b_R}, \chi _2 u\big )\). For any \(u\in {\mathcal {D}}(C)\) one has then \(\Vert Ju\Vert _{L^2(b_R,\mathbb {C}^N)\times L^2(\mathbb {R}^n,\mathbb {C}^N)}=\Vert u\Vert _{L^2(\mathbb {R}^n,\mathbb {C}^N)}\), and the above representation (61) can be rewritten as

$$\begin{aligned} C^2[u,u]= & {} \big \Vert C u\Vert _{L^2(\mathbb {R}^n,\mathbb {C}^N)}^2\\= & {} (S_1-W)[(\chi _1 u)|_{b_R},(\chi _1 u)|_{b_R}] +\big ((S_0+M^2)\otimes I_N\big )[\chi _2u,\chi _2 u]\\\equiv & {} \Big ((S_1-W) \,\oplus \big ((S_0+M^2)\otimes I_N\big )\Big ) [Ju,Ju]. \end{aligned}$$

A standard application of the min–max principle (see e.g. Proposition 1.9) yields the inequality

$$\begin{aligned} \mathcal {N}(M^2,C^2)&\le \mathcal {N}\Big (M^2,(S_1-W) \,\oplus \big ((S_0+M^2)\otimes I_N\big )\Big )\\&\equiv \mathcal {N}(M^2,S_1-W) + N \mathcal {N}(0,S_0). \end{aligned}$$

As \(S_1-W\) is semibounded from below with compact resolvent, one has \(\mathcal {N}(M^2,S_1-W)<\infty \), and in order to show \(\mathcal {N}(M^2,C^2)<\infty \) it is sufficient to prove that \(\mathcal {N}(0,S_0)<\infty \).

Recall that W vanishes outside \(b_R\). Let \(T_R\) be the self-adjoint operator in \(L^2(b_R)\) with

$$\begin{aligned} T_R[u,u]=\int _{b_R} \big ( |\nabla u|^2- W |u|^2\big )\,\mathrm {d}x, \quad {\mathcal {Q}}(T_R)=H^1(b_R), \end{aligned}$$

denote by \(O_R\) the zero operator in \(L^2(\mathbb {R}^2{\setminus } b_R)\), and consider the linear map

$$\begin{aligned} J':{\mathcal {Q}}(S_0)\rightarrow {\mathcal {Q}}(T_R)\times {\mathcal {Q}}(O_R), \quad J'u=(u|_{b_R}, u|_{\mathbb {R}^2{\setminus } b_R}). \end{aligned}$$

One has then

$$\begin{aligned} S_0[u,u]\ge \int _{b_R} \big ( |\nabla u|^2- W |u|^2\big )\,\mathrm {d}x\equiv (T_R \oplus O_R)[J'u,J'u] \end{aligned}$$

implying \(\mathcal {N}(0,S_0)\le \mathcal {N}(0,T_R\oplus O_R)\equiv \mathcal {N}(0,T_R)+\mathcal {N}(0,O_R)\equiv \mathcal {N}(0,T_R) <\infty \), as \(T_R\) is semibounded from below with compact resolvent. This shows that \(\mathcal {N}(M^2,C^2)<\infty \) and finishes the proof of(b).

  1. (c)

    According to the general theory of Clifford algebras, see e.g. Theorem 15.19 in [9], for \(n\notin 4\mathbb {Z}\) there exists an antilinear map \(\theta :\mathbb {C}^N\rightarrow \mathbb {C}^N\) with \(\theta ^2\in \{-1,+1\big \}\) andFootnote 2

    $$\begin{aligned} \theta \alpha _j = \alpha _j \theta \text { for } j=1,\dots , n, \quad \theta (\mathrm {i}\alpha _{n+1})=(\mathrm {i}\alpha _{n+1})\theta . \end{aligned}$$
    (62)

The pointwise map \(\Theta \) defined by \((\Theta u)(x)=\theta \big ( u(x)\big )\) is clearly an isomorphism of \(H^1(\mathbb {R}^n,\mathbb {C}^N)\). Let \(u\in \ker (C-E)\), \(E\in \mathbb {R}\), then \(\Theta u\in {\mathcal {D}}(C)\) and

$$\begin{aligned} E u=-\mathrm {i}\sum _{j=1}^n\alpha _j \dfrac{\partial u}{\partial x_j} + V\alpha _{n+1} u. \end{aligned}$$

The last equality in (62) rewrites as \(\theta \alpha _{n+1}= - \alpha _{n+1}\theta \). We compute then

$$\begin{aligned} -E\Theta u&=\Theta (-Eu)=\Theta \Big (\mathrm {i}\sum _{j=1}^n\alpha _j \dfrac{\partial u}{\partial x_j} - V\alpha _{n+1} u\Big )\\&=-\mathrm {i}\sum _{j=1}^n\theta \alpha _j \dfrac{\partial u}{\partial x_j} - V\theta \alpha _{n+1} u =-\mathrm {i}\sum _{j=1}^n\alpha _j \theta \dfrac{\partial u}{\partial x_j} + V\alpha _{n+1} \theta u=C\Theta u, \end{aligned}$$

which shows that \(\Theta u\in \ker (C+E)\). By construction one has \(\Theta ^2\in \{1,-1\}\), hence, \(\Theta \) is a bijection and the claim follows. \(\square \)

Let us now discuss the basic properties of the operator \(A_m\).

Proposition A.2

The following assertions hold:

  1. (a)

    The operator \(A_m\) is self-adjoint, and, in addition, it is essentially self-adjoint on \({\mathcal {D}}_\infty (A_m):= C^\infty (\overline{\Omega },\mathbb {C}^N)\cap {\mathcal {D}}(A_m)\),

  2. (b)

    The operator \(A_m\) has compact resolvent, and its eigenfunctions belong to \({\mathcal {D}}_\infty (A_m)\),

  3. (c)

    Assume in addition that \(n\notin 4\mathbb {Z}\), then the spectrum of \(A_m\) is symmetric with respect to zero, i.e. \(\dim \ker (A_{m}-E)=\dim \ker (A_{m}+E)\) for any \(E\in \mathbb {R}\).

The above assertions also hold with \(A_m\) replaced by \(A_m'\).

Proof

The assertions (a) and (b) are standard properties of elliptic Dirac operators. One can use e.g. [3, Ex. 4.20] or [4, Ex. 7.26], by noting that \(\mathcal {B}\) is a boundary chirality operator, which shows that \(A_m\) belongs to the class of Dirac operators with local elliptic boundary conditions. Then [3, Thm. 4.11] implies that the restriction R of \(A_m\) on \({\mathcal {D}}_\infty (A_m)\) is essentially self-adjoint, while the boundary regularity theorem [3, Thm. 4.9] implies that \(\overline{R}\subset A_m\). As \(A_m\) is clearly symmetric, this implies that \(A_m\) is self-adjoint and proves (a). The same boundary regularity theorem [3, Thm. 4.9] implies that all eigenfunctions are smooth up to the boundary, and the compactness of the resolvent of \(A_m\) follows then from the compactness of the embedding of \(H^1(\Omega )\) into \(L^2(\Omega )\). This proves (b).

The proof of (c) is very similar to the proof of Proposition A.1(c) with an additional attention given to the operator domain. Namely, for \(n\notin 4\mathbb {Z}\) there exist an antilinear map \(\theta :\mathbb {C}^N\rightarrow \mathbb {C}^N\) with \(\theta ^2\in \{-1,+1\big \}\) satisfying (62), and the pointwise map \(\Theta \) defined by \((\Theta u)(x)=\theta \big (u(x)\big )\) is again an isomorphism of \(H^1(\Omega ,\mathbb {C}^N)\). Furthermore, if \(v\in H^1(\Omega ,\mathbb {C}^N)\) satisfies the boundary condition \(v=\mathcal {B}v\) on \(\Sigma \), then due to

$$\begin{aligned} \Theta \mathcal {B}(s)=\Theta \big ( -\mathrm {i}\alpha _{n+1} \Gamma (s)\big )=-\mathrm {i}\alpha _{n+1} \Theta \Gamma (s)=-\mathrm {i}\alpha _{n+1}\Gamma (s)\Theta =\mathcal {B}\Theta (s) \end{aligned}$$

one also has \(\Theta v=\mathcal {B}\Theta v\) on \(\Sigma \). Therefore, if \(v\in {\mathcal {D}}(A_m)\), then also \(\Theta v\in {\mathcal {D}}(A_m)\). Finally, let \(v\in \ker (A_m-E)\), \(E\in \mathbb {R}\), then the same computation as in Proposition A.1(c) gives \(\Theta v\in \ker (A_m+E)\).

The preceding arguments apply to \(A_m'\) as well. First, \((-\mathcal {B})\) is still a boundary chirality operator in the sense of [3], which implies the assertions (a) and (b). The proof of (c) is the same by noting that \(v=-\mathcal {B}v\) on \(\Sigma \) yields \(\Theta v=-\mathcal {B}\Theta v\) on \(\Sigma \). \(\square \)

B Schrödinger–Lichnerowicz Formula for Euclidean Hypersurfaces

Let \(\Sigma \subset \mathbb {R}^n\) be a smooth compact hypersurface with the outer unit normal field \(\nu \) and endowed with the Riemannian metric induced by the embedding. Recall that the standard scalar product in \(\mathbb {R}^n\) gives rise to the induced scalar product in \(T\Sigma \), which we simply denote by \(\langle \cdot ,\cdot \rangle \) in this section. Denote by W the Weingarten operator, \(W X=\nabla _X \nu \) for \(X\in T\Sigma \), with \(\nabla \) being the gradient in \(\mathbb {R}^n\). Recall that the Levi-Civita connection \(\nabla '\) on \(\Sigma \) is given by the Gauss formula

$$\begin{aligned} \nabla '_X Y=\nabla _X Y+\langle WX,Y\rangle \nu , \quad X,Y\in T\Sigma . \end{aligned}$$

We denote

$$\begin{aligned} H_1:={{\,\mathrm{tr}\,}}W, \quad |W|^2:={{\,\mathrm{tr}\,}}(W^2), \quad H_2:=\dfrac{H_1^2-|W|^2}{2}, \end{aligned}$$

i.e. \(H_1\) is the mean curvature and \(H_2\) is the half of the scalar curvature of \(\Sigma \).

Let \(N\in \mathbb {N}\) and \(\gamma _1,\dots ,\gamma _n\) be \(N\times N\) anticommuting Hermitian matrices satisfying \(\gamma _j^2=I\), with I being the \(N\times N\) identity matrix, then the matrices

$$\begin{aligned} \gamma (x):=\sum _{j=1}^n x_j \gamma _j, \quad x=(x_1,\dots ,x_n)\in \mathbb {R}^n, \end{aligned}$$

satisfy the Clifford commutation relation \(\gamma (x)\gamma (y)+\gamma (y)\gamma (x)=2\langle x,y\rangle I\) for all \(x,y\in \mathbb {R}^n\). Let us recall the definition of the associated extrinsically defined Dirac operator \(D^\Sigma \) on \(\Sigma \) following [14, Sec. 1–3].Footnote 3 The induced spin connection \(\nabla ^\Sigma \) on \(\Sigma \) is defined by

$$\begin{aligned} \nabla ^\Sigma _X \psi =\nabla _X+\dfrac{1}{2}\,\gamma (\nu )\gamma (WX):\, C^\infty (\Sigma ,\mathbb {C}^N)\rightarrow C^\infty (\Sigma ,\mathbb {C}^N), \quad X\in T\Sigma , \end{aligned}$$

then \(D^\Sigma \) acts on functions \(\psi \in C^\infty (\Sigma ,\mathbb {C}^N)\) by

$$\begin{aligned} D^\Sigma \psi :=-\gamma (\nu ) \sum _{j=1}^{n-1} \gamma (e_j) \nabla ^\Sigma _{e_j} \psi \end{aligned}$$

with \((e_1,\dots ,e_{n-1})\) being an orthonormal frame of \(T\Sigma \). Recall that \(\gamma (e_j)\) anticommute with \(\gamma (\nu )\) and, furthermore,

$$\begin{aligned} \sum _{j=1}^{n-1} \gamma (e_j) \gamma (We_j)=H_1\, I \end{aligned}$$
(63)

(which is seen by testing on an eigenbasis of W), and we may rewrite

$$\begin{aligned} D^\Sigma \psi =\dfrac{H_1}{2}\,\psi -\gamma (\nu ) \sum _{j=1}^{n-1} \gamma (e_j) \nabla _{e_j}\psi , \end{aligned}$$

Being viewed as an operator in \(L^2(\Sigma ,\mathbb {C}^N)\), the operator \(D^\Sigma \) is known to be essentially self-adjoint on \(C^\infty (\Sigma ,\mathbb {R}^N)\). We would like to provide a elementary direct proof, adapted to the Euclidean setting, of the eminent Schrödinger–Lichnerowicz formula

$$\begin{aligned} (D^\Sigma )^2=(\nabla ^\Sigma )^*\nabla ^\Sigma +\dfrac{H_2}{2}\, I, \end{aligned}$$
(64)

where the first term on the right-hand side is the Bochner Laplacian associated with the above spin connection \(\nabla ^\Sigma \), which is a self-adjoint operator in \(L^2(\Sigma ,\mathbb {C}^N)\). (We refer to the original papers [17, 22] and other sources [7, 11,12,13] for a more general setting.)

In what follows we use the standard identification of \(T\Sigma \) and \(T^*\Sigma \) with the help of the musical isomorphism. For \(\psi \in C^\infty (\Sigma ,\mathbb {C}^N)\) we have the decomposition

$$\begin{aligned} \nabla ^\Sigma \psi =\sum _{j=1}^{n-1} e_j\otimes \nabla ^\Sigma _{e_j} \psi = \sum _{j=1}^{n-1} e_j\otimes \Big (\nabla _{e_j}+\dfrac{1}{2}\,\gamma (\nu )\gamma (W e_j)\Big ) \psi . \end{aligned}$$
(65)

To compute the adjoint \((\nabla ^\Sigma )^*:T^*\Sigma \otimes C^\infty (\Sigma ,\mathbb {C}^N)\rightarrow C^\infty (\Sigma ,\mathbb {C}^N)\), let \(X\in T\Sigma \simeq T^* \Sigma \) and \(\varphi ,\psi \in C^\infty (\Sigma ,\mathbb {C}^N)\) then

$$\begin{aligned} \big \langle (\nabla ^\Sigma )^*(X\otimes \varphi ),\psi \big \rangle _{L^2(\Sigma ,\mathbb {C}^N)}&=\langle X\otimes \varphi ,\nabla ^\Sigma \psi \rangle _{T^*\Sigma \otimes L^2(\Sigma ,\mathbb {C}^N)}\\&=\big \langle \varphi , \nabla _X \psi +\dfrac{1}{2}\,\gamma (\nu )\gamma (WX)\psi \big \rangle _{L^2(\Sigma ,\mathbb {C}^N)}\\&=\big \langle \varphi , \nabla _X \psi \big \rangle _{L^2(\Sigma ,\mathbb {C}^N)} +\Big \langle \dfrac{1}{2}\,\gamma (WX) \gamma (\nu )\varphi ,\psi \Big \rangle _{L^2(\Sigma ,\mathbb {C}^N)}. \end{aligned}$$

Using Leibniz rule and the divergence theorem we have

$$\begin{aligned} \langle \varphi , \nabla _X \psi \rangle _{L^2(\Sigma ,\mathbb {C}^N)}&= \int _\Sigma X\langle \varphi ,\psi \rangle _{\mathbb {C}^N} \,\mathrm {d}s-\langle \nabla _X \varphi , \psi \rangle _{L^2(\Sigma ,\mathbb {C}^N)}\\&=-\big \langle ({{\,\mathrm{div}\,}}_\Sigma X) \varphi +\nabla _X \varphi ,\psi \big \rangle _{L^2(\Sigma ,\mathbb {C}^N)}, \end{aligned}$$

where \({{\,\mathrm{div}\,}}_\Sigma \) is the divergence on \(\Sigma \),

$$\begin{aligned} {{\,\mathrm{div}\,}}_\Sigma X=\sum _{j=1}^{n-1} \langle e_j,\nabla '_{e_j} X\rangle . \end{aligned}$$

Therefore,

$$\begin{aligned} (\nabla ^\Sigma )^*(X\otimes \varphi )=-({{\,\mathrm{div}\,}}_\Sigma X)\,\varphi -\nabla _X \varphi + \frac{1}{2}\,\gamma (WX)\gamma (\nu )\varphi . \end{aligned}$$

By combining (65) with the last expression, for \(\psi \in C^\infty (\Sigma ,\mathbb {C}^N)\) one obtains

$$\begin{aligned} (\nabla ^\Sigma )^* \nabla ^\Sigma \psi&=\sum _{j=1}^{n-1} (\nabla ^\Sigma )^* \Big [e_j\otimes \Big (\nabla _{e_j}+\dfrac{1}{2}\,\gamma (\nu )\gamma (W e_j)\Big ) \psi \Big ]\nonumber \\&=-\sum _{j=1}^{n-1} ({{\,\mathrm{div}\,}}_\Sigma e_j) \Big (\nabla _{e_j}+\dfrac{1}{2}\,\gamma (\nu )\gamma (W e_j)\Big ) \psi \nonumber \\&\quad +\sum _{j=1}^{n-1}\bigg \{ -\nabla _{e_j}\Big (\nabla _{e_j}+\dfrac{1}{2}\,\gamma (\nu )\gamma (W e_j)\Big )\psi \nonumber \\&\quad +\dfrac{1}{2}\,\gamma (We_j)\gamma (\nu )\Big (\nabla _{e_j}+\dfrac{1}{2}\,\gamma (\nu )\gamma (W e_j)\Big )\psi \bigg \}=:S_1+S_2. \end{aligned}$$
(66)

To simplify \(S_1\) we first use the Leibniz rule and the orthogonality of \((e_j)\) to obtain

$$\begin{aligned} {{\,\mathrm{div}\,}}_\Sigma e_j&=\sum _{k=1}^{n-1} \langle e_k,\nabla '_{e_k} e_j\rangle =-\sum _{k=1}^{n-1} \langle \nabla '_{e_k} e_k, e_j\rangle ,\\ S_1&=\sum _{j,k=1}^{n-1} \langle \nabla '_{e_k} e_k, e_j\rangle \nabla _{e_j}\psi +\dfrac{1}{2}\sum _{j,k=1}^{n-1} \langle \nabla '_{e_k} e_k, e_j\rangle \gamma (\nu )\gamma (W e_j) \psi \\&=\sum _{k=1}^{n-1} \Big (\sum _{j=1}^{n-1}\langle \nabla '_{e_k} e_k, e_j\rangle \nabla _{e_j}\psi \Big ) +\dfrac{1}{2}\sum _{k=1}^{n-1} \gamma (\nu ) \gamma \bigg ( W\sum _{j=1}^{n-1}\langle \nabla '_{e_k} e_k, e_j\rangle e_j \bigg )\\&=\sum _{k=1}^{n-1} \nabla _{\nabla '_{e_k} e_k} \psi +\dfrac{1}{2}\sum _{k=1}^{n-1} \gamma (\nu )\gamma \Big ( W \nabla '_{e_k} e_k\Big ). \end{aligned}$$

Furthermore,

$$\begin{aligned} S_2&=\sum _{j=1}^{n-1} \bigg \{-\nabla _{e_j}\nabla _{e_j}\psi -\dfrac{1}{2}\,\gamma (We_j)\gamma (W e_j)\psi -\dfrac{1}{2}\,\gamma (\nu )\gamma \big (\nabla _{e_j}(W e_j)\big )\psi \\&\qquad -\dfrac{1}{2}\,\gamma (\nu )\gamma (We_j)\nabla _{e_j}\psi +\dfrac{1}{2}\,\gamma (We_j)\gamma (\nu )\nabla _{e_j}\psi + \dfrac{1}{4}\,\gamma (We_j)\gamma (\nu )\gamma (\nu )\gamma (W e_j)\psi \bigg \}\\&=\sum _{j=1}^{n-1} \bigg \{-\nabla _{e_j}\nabla _{e_j}\psi -\dfrac{1}{2}\,\gamma (\nu )\gamma \big (\nabla _{e_j}(W e_j)\big )\psi -\gamma (\nu )\gamma (We_j)\nabla _{e_j}\psi \bigg \}\psi - \dfrac{1}{4}\,|W|^2\psi , \end{aligned}$$

and then

$$\begin{aligned} (\nabla ^\Sigma )^* \nabla ^\Sigma \psi= & {} \sum _{j=1}^{n-1} \bigg [ \nabla _{\nabla '_{e_j} e_j} \psi -\nabla _{e_j}\nabla _{e_j}\psi \\&+\dfrac{1}{2}\gamma (\nu )\gamma \Big ( W \nabla '_{e_j} e_j-\nabla _{e_j}(W e_j)\Big )\psi -\gamma (\nu )\gamma (We_j)\nabla _{e_j}\psi \bigg ] - \dfrac{1}{4}\,|W|^2\psi . \end{aligned}$$

Using \(\nabla '_{e_j} (W e_j)=\nabla _{e_j} (W e_j)+|W e_j|^2\nu \) and Leibniz rule we have

$$\begin{aligned} W \nabla '_{e_j} e_j-\nabla _{e_j}(W e_j)=W\nabla '_{e_j} e_j-\nabla '_{e_j} (W e_j)+|W e_j|^2\nu =-(\nabla '_{e_j} W)e_j +|W e_j|^2\nu \end{aligned}$$

implying \(\gamma (\nu )\gamma \big ( W \nabla '_{e_j} e_j-\nabla _{e_j}(W e_j)\big )\psi =-\gamma (\nu )\gamma \big ((\nabla '_{e_j} W)e_j\big )\psi +|W e_j|^2\psi \), and then

$$\begin{aligned} (\nabla ^\Sigma )^* \nabla ^\Sigma \psi= & {} \sum _{j=1}^{n-1} \bigg [ \nabla _{\nabla '_{e_j} e_j} \psi -\nabla _{e_j}\nabla _{e_j}\psi \nonumber \\&-\dfrac{1}{2} \,\gamma (\nu )\,\gamma \big ((\nabla '_{e_j} W)e_j\big )\psi -\gamma (\nu )\,\gamma (We_j)\nabla _{e_j}\psi \bigg ] + \dfrac{1}{4}\,|W|^2\psi .\nonumber \\ \end{aligned}$$
(67)

On the other hand,

$$\begin{aligned} (D^\Sigma )^2 \psi&=\Big (\dfrac{H_1}{2}-\gamma (\nu ) \sum _{j=1}^{n-1} \gamma (e_j) \nabla _{e_j}\Big )\Big (\dfrac{H_1\,\psi }{2}-\gamma (\nu ) \sum _{j=1}^{n-1} \gamma (e_j) \nabla _{e_j}\psi \Big )\nonumber \\&=\dfrac{H_1^2}{4}\,\psi -\dfrac{1}{2} \Big ( \gamma (\nu ) \sum _{j=1}^{n-1} \gamma (e_j) \nabla _{e_j} H_1\Big )\, \psi - \dfrac{H_1}{2}\,\gamma (\nu ) \sum _{j=1}^{n-1} \gamma (e_j) \nabla _{e_j}\psi \nonumber \\&\quad -\dfrac{H_1}{2}\,\gamma (\nu ) \sum _{j=1}^{n-1} \gamma (e_j) \nabla _{e_j}\psi +\gamma (\nu )\sum _{j,k=1}^{n-1} \gamma (e_j)\gamma (We_j)\gamma (e_k)\nabla _{e_k}\psi \nonumber \\&\quad + \gamma (\nu )\sum _{j,k=1}^{n-1} \gamma (e_j)\gamma (\nu )\gamma (\nabla _{e_j}e_k)\nabla _{e_k}\psi +\gamma (\nu )\sum _{j,k=1}^{n-1}\gamma (e_j)\gamma (\nu )\gamma (e_k)\nabla _{e_j}\nabla _{e_k}\psi . \end{aligned}$$
(68)

The sum of the third, forth and fifth terms is zero, in fact,

$$\begin{aligned}&-\dfrac{H_1}{2}\,\gamma (\nu ) \sum _{j=1}^{n-1} \gamma (e_j) \nabla _{e_j}\psi -\dfrac{H_1}{2}\,\gamma (\nu ) \sum _{j=1}^{n-1} \gamma (e_j) \nabla _{e_j}\psi \\&\qquad +\gamma (\nu )\sum _{j,k=1}^{n-1} \gamma (e_j)\gamma (We_j)\gamma (e_k)\nabla _{e_k}\psi \\&\quad = -H_1\,\gamma (\nu ) \sum _{j=1}^{n-1} \gamma (e_k) \nabla _{e_k}\psi +\gamma (\nu )\sum _{j,k=1}^{n-1} \gamma (e_j)\gamma (We_j)\gamma (e_k)\nabla _{e_k}\psi \\&\quad =\gamma (\nu )\sum _{k=1}^{n-1} \Big ( -H_1 +\sum _{j=1}^{n-1}\gamma (e_j)\gamma (We_j)\Big )\gamma (e_k)\nabla _{e_k}\psi =0 \end{aligned}$$

as the term in the parentheses vanishes due to (63). Therefore, Eq. (68) reads as

$$\begin{aligned} (D^\Sigma )^2\psi&=\dfrac{H_1^2}{4}\,\psi -\dfrac{1}{2} \, \gamma (\nu ) \sum _{j=1}^{n-1} \gamma (e_j) (\nabla _{e_j} H_1)\, \psi - \sum _{j,k=1}^{n-1} \gamma (e_j)\gamma (\nabla _{e_j}e_k)\nabla _{e_k}\psi \nonumber \\&\quad -\sum _{j,k=1}^{n-1}\gamma (e_j)\gamma (e_k)\nabla _{e_j}\nabla _{e_k}\psi . \end{aligned}$$
(69)

We transform the last summand as follows:

$$\begin{aligned} \sum _{j,k=1}^{n-1}\gamma (e_j)\gamma (e_k)\nabla _{e_j}\nabla _{e_k}\psi&=\dfrac{1}{2}\,\sum _{j,k=1}^{n-1}\Big (\gamma (e_j)\gamma (e_k)\nabla _{e_j}\nabla _{e_k}\psi +\gamma (e_k)\gamma (e_j)\nabla _{e_k}\nabla _{e_j}\psi \Big )\\&=\dfrac{1}{2}\,\sum _{j,k=1}^{n-1}\Big (\gamma (e_j)\gamma (e_k)+\gamma (e_k)\gamma (e_j)\Big )\nabla _{e_j}\nabla _{e_k}\psi \\&\quad +\dfrac{1}{2}\sum _{j,k=1}^{n-1}\gamma (e_k)\gamma (e_j) \Big ( \nabla _{e_k}\nabla _{e_j}-\nabla _{e_j}\nabla _{e_k}\Big )\psi \\&=\sum _{j=1}^{n-1}\nabla _{e_j}\nabla _{e_j}\psi +\dfrac{1}{2}\,J, \end{aligned}$$

where

$$\begin{aligned} J:=\sum _{j,k=1}^{n-1}\gamma (e_j)\gamma (e_k) \big ( \nabla _{e_j}\nabla _{e_k}-\nabla _{e_k}\nabla _{e_j}\big )\psi \equiv \sum _{j,k=1}^{n-1}\gamma (e_j)\gamma (e_k) \nabla _{[e_j,e_k]}\psi . \end{aligned}$$

Representing \([e_j,e_k]=\sum _{l=1}^{n-1} \big \langle e_l,[e_j,e_k]\big \rangle e_k\) we have

$$\begin{aligned} J=\sum _{j,k,l=1}^{n-1} \gamma (e_j)\gamma (e_k)\Big [\langle e_l,\nabla '_{e_j}e_k\rangle -\langle e_l,\nabla '_{e_k}e_j\rangle \Big ]\nabla _{e_l}\psi , \end{aligned}$$

and using

$$\begin{aligned} \sum _{j=1}^{n-1} e_j\langle e_l,\nabla '_{e_k}e_j\rangle =-\sum _{j=1}^{n-1} e_j\langle \nabla '_{e_k} e_l,e_j\rangle =-\nabla '_{e_k} e_l \end{aligned}$$

we rewrite

$$\begin{aligned} J&=-\sum _{j,k=1}^{n-1}\gamma (e_j)\,\gamma (\nabla '_{e_j}e_k) \nabla _{e_k}\psi +\sum _{j,k=1}^{n-1}\gamma (\nabla '_{e_j}e_k)\,\gamma (e_j)\, \nabla _{e_k}\psi \\&=-2\sum _{j,k=1}^{n-1}\gamma (e_j)\,\gamma (\nabla '_{e_j}e_k) \nabla _{e_k}\psi \\&\quad +\sum _{j,k=1}^{n-1}\Big ( \gamma (e_j)\,\gamma (\nabla '_{e_j}e_k)+\gamma (\nabla '_{e_j}e_k)\,\gamma (e_j)\,\Big ) \nabla _{e_k}\psi \\&=-2\sum _{j,k=1}^{n-1}\gamma (e_j)\,\gamma (\nabla '_{e_j}e_k) \nabla _{e_k}\psi +2\sum _{j,k=1}^{n-1}\langle e_j,\nabla '_{e_j}e_k\rangle \nabla _{e_k}\psi \\&=-2\sum _{j,k=1}^{n-1}\gamma (e_j)\,\gamma (\nabla '_{e_j}e_k) \nabla _{e_k}\psi -2\sum _{j,k=1}^{n-1}\langle \nabla '_{e_j}e_j,e_k\rangle \nabla _{e_k}\psi \\&=-2\sum _{j,k=1}^{n-1}\gamma (e_j)\,\gamma (\nabla '_{e_j}e_k) \nabla _{e_k}\psi -2\sum _{j=1}^{n-1} \nabla _{\nabla '_{e_j}e_j}\psi . \end{aligned}$$

The substitution into (69) gives

$$\begin{aligned} (D^\Sigma )^2 \psi&=\dfrac{H_1^2}{4}\,\psi -\dfrac{1}{2} \, \gamma (\nu ) \sum _{j=1}^{n-1} \gamma (e_j) (\nabla _{e_j} H_1)\, \psi - \sum _{j,k=1}^{n-1} \gamma (e_j)\gamma (\nabla _{e_j}e_k)\nabla _{e_k}\psi \\&\quad -\sum _{j=1}^{n-1}\nabla _{e_j}\nabla _{e_j}\psi +\sum _{j,k=1}^{n-1}\gamma (e_j)\,\gamma (\nabla '_{e_j}e_k) \nabla _{e_k}\psi +\sum _{j=1}^{n-1} \nabla _{\nabla '_{e_j}e_j}\psi . \end{aligned}$$

The sum of the third and the fifth terms simplifies as

$$\begin{aligned}&- \sum _{j,k=1}^{n-1} \gamma (e_j)\gamma (\nabla _{e_j}e_k)\nabla _{e_k}\psi +\sum _{j,k=1}^{n-1}\gamma (e_j)\,\gamma (\nabla '_{e_j}e_k) \nabla _{e_k}\psi \\&\quad =\sum _{j,k=1}^{n-1} \gamma (e_j)\gamma (\nabla '_{e_j}e_k-\nabla _{e_j}e_k) \nabla _{e_k}\psi =\sum _{j,k=1}^{n-1} \gamma (e_j)\gamma \big (\langle W e_j,e_k\rangle \nu \big ) \nabla _{e_k}\psi \\&\quad =\sum _{j,k=1}^{n-1} \gamma \big (e_j \langle e_j, W e_k\rangle \big )\gamma (\nu ) \nabla _{e_k}\psi = \sum _{k=1}^{n-1}\gamma (W e_k) \gamma (\nu )\nabla _{e_k}\psi , \end{aligned}$$

hence,

$$\begin{aligned} (D^\Sigma )^2 \psi= & {} \dfrac{H_1^2}{4}\, \psi -\dfrac{1}{2} \, \gamma (\nu ) \sum _{j=1}^{n-1} \gamma (e_j) (\nabla _{e_j} H_1)\, \psi \\&+\sum _{j=1}^{n-1}\gamma (W e_j) \gamma (\nu )\nabla _{e_j}\psi -\sum _{j=1}^{n-1}\nabla _{e_j}\nabla _{e_j}\psi +\sum _{j=1}^{n-1} \nabla _{\nabla '_{e_j}e_j}\psi . \end{aligned}$$

By comparing the last expression with (67) we obtain

$$\begin{aligned}&D^2_\Sigma \psi - (\nabla ^\Sigma )^*\nabla ^\Sigma \psi \\&\quad = \dfrac{H_1^2}{4}\,\psi -\dfrac{1}{4} |W|^2\,\psi -\dfrac{1}{2} \, \gamma (\nu ) \sum _{j=1}^{n-1} \gamma (e_j) (\nabla _{e_j} H_1)\, \psi +\sum _{j=1}^{n-1}\gamma (W e_j) \gamma (\nu )\nabla _{e_j}\psi \\&\qquad +\dfrac{1}{2} \sum _{j=1}^{n-1}\,\gamma (\nu )\,\gamma \big ((\nabla '_{e_j} W)e_j\big )\psi +\sum _{j=1}^{n-1}\gamma (\nu )\,\gamma (We_j)\nabla _{e_j}\psi . \end{aligned}$$

The sum of the fourth and sixth term on the right hand is zero, hence,

$$\begin{aligned}&(D^\Sigma )^2 \psi - (\nabla ^\Sigma )^*\nabla ^\Sigma \psi \\&\quad = \dfrac{H_2}{2}\,\psi -\dfrac{1}{2} \, \gamma (\nu ) \sum _{j=1}^{n-1} \gamma (e_j) (\nabla _{e_j} H_1)\, \psi +\dfrac{1}{2} \sum _{j=1}^{n-1}\,\gamma (\nu )\,\gamma \big ((\nabla '_{e_j} W)e_j\big )\psi \\&\quad =\dfrac{H_2}{2}\,\psi + \dfrac{1}{2}\,\gamma (\nu ) \gamma \bigg (\sum _{j=1}^{n-1} (\nabla '_{e_j} W) e_j - \sum _{j=1}^{n-1} (\nabla _{e_j} H_1)e_j\bigg )\psi . \end{aligned}$$

Therefore, to show the sought identity (64) it is sufficient to prove the equality

$$\begin{aligned} \sum _{j=1}^{n-1} (\nabla '_{e_j} W)e_j=\sum _{j=1}^{n-1} (\nabla _{e_j} H_1)e_j. \end{aligned}$$
(70)

In order to check (70) let us remark that \(\nabla _X \nabla _Y Z-\nabla _Y \nabla _X Z - \nabla _{[X,Y]} Z=0\) for any \(X,Y,Z\in T\Sigma \). Using the definition of \(\nabla '\) we have

$$\begin{aligned} 0&= \nabla _X \big ( \nabla '_Y Z - \langle WY,Z\rangle \nu \big )-\nabla _Y\big (\nabla '_X Z -\langle WX,Z\rangle \nu \big )- \nabla '_{[X,Y]} Z +\big \langle W[X,Y],Z\big \rangle \nu \\&= \nabla '_X\big ( \nabla '_Y Z - \langle WY,Z\rangle \nu \big )- \big \langle WX, \nabla '_Y Z - \langle WY,Z\rangle \nu \big \rangle - \nabla '_Y\big (\nabla '_X Z -\langle WX,Z\rangle \nu \big )\\&\quad +\big \langle WY, \nabla '_X Z -\langle WX,Z\rangle \nu \big \rangle - \nabla '_{[X,Y]} Z +\big \langle W[X,Y],Z\big \rangle \nu . \end{aligned}$$

Using \(\nabla '_X \nu =\nabla _X \nu =WX\) we then arrive at

$$\begin{aligned} 0&= \nabla '_X\nabla '_Y Z - \big \langle (\nabla '_X W)Y,Z\big \rangle \nu -\big \langle W(\nabla '_X Y),Z\big \rangle \nu -\langle WY,\nabla '_X Z\rangle \nu -\langle WY,Z\rangle WX\\&\quad - \langle WX, \nabla '_Y Z\rangle \nu -\nabla '_Y \nabla '_X Z+ \big \langle (\nabla '_Y W)X,Z\big \rangle \nu +\langle W\nabla '_Y X,Z\rangle \nu \\&\quad +\langle WX,\nabla '_Y Z\rangle \nu +\langle WX,Z\rangle WY+\langle WY, \nabla '_X Z\rangle - \nabla '_{[X,Y]} Z +\big \langle W[X,Y],Z\big \rangle \nu \\&= \nabla '_X\nabla '_Y Z-\nabla '_Y \nabla '_X Z- \nabla '_{[X,Y]} Z -\langle WY,Z\rangle WX +\langle WX,Z\rangle WY\\&\quad +\big \langle (\nabla '_Y W)X,Z\big \rangle \nu - \big \langle (\nabla '_X W)Y,Z\big \rangle \nu \\&\quad +\langle W\nabla '_Y X,Z\rangle \nu -\big \langle W \nabla '_X Y,Z\big \rangle \nu +\big \langle W[X,Y],Z\big \rangle \nu . \end{aligned}$$

As \(\nabla '_X Y-\nabla '_Y X=[X,Y]\), the sum of the three terms in the last line vanishes. Considering the normal components of the remaining equality we obtain \(\big \langle (\nabla '_Y W)X,Z\big \rangle = \big \langle (\nabla '_X W)Y,Z\big \rangle \), and then \(\big \langle (\nabla '_Y W)X,Z\big \rangle =\big \langle Y,(\nabla '_X W) Z\big \rangle \). Taking \(Y=Z=e_k\) and summing over k we arrive at

$$\begin{aligned} \sum _{k=1}^{n-1}\big \langle (\nabla '_{e_k} W) X,e_k\big \rangle =\sum _{j=1}^{n-1} \big \langle e_k,(\nabla '_X W ) e_k\big \rangle , \text { i.e. } \sum _{k=1}^{n-1}\big \langle X, (\nabla '_{e_k} W) e_k\big \rangle =\nabla _X H_1. \end{aligned}$$

Using the last equality for \(X=e_j\) we obtain

$$\begin{aligned} \sum _{j=1}^{n-1}\sum _{k=1}^{n-1}\big \langle e_j, (\nabla '_{e_k} W) e_k\big \rangle e_j=\sum _{j=1}^{n-1} (\nabla _{e_j} H_1) e_j. \end{aligned}$$

The left-hand side of the last equality simplifies to \(\sum _{k=1}^{n-1} (\nabla '_{e_k} W) e_k\), which gives (70) and finishes the proof of (64).

C Dirac Operator on a Loop

Let \(\Sigma \subset \mathbb {R}^2\) be a smooth loop of length \(\ell >0\). We give an explicit computation for the eigenvalues of the intrinsic Dirac operator on \(\Sigma \). Consider first the associated extrinsically defined Dirac operator \(D^\Sigma \) using the notation introduced in Sect. 2.2. Denote by \(\mathbb {T}:=\mathbb {R}/(\ell \mathbb {Z})\) and let \(\gamma :\mathbb {T}\rightarrow \mathbb {R}^2\) be an arc-length parametrization of \(\Sigma \), which provides a global coordinate of \(\Sigma \). To be definite, assume that \(\gamma (s)\) runs through \(\Sigma \) in the anti-clockwise direction as s runs from 0 to \(\ell \), which amounts to the choice of an orientation. We take (e) with \(e=\tau :=\gamma '\) as an orthonormal frame tangent to \(\Sigma \) and denote by \(\nu \) the outer unit normal. In addition, let us make an explicit choice of \(2\times 2\) matrices \(\beta _1\) and \(\beta _2\) satisfying the Clifford commutation relation: we choose them as the Pauli matrices,

$$\begin{aligned} \beta _1=\begin{pmatrix} 0 &{} 1\\ 1 &{} 0 \end{pmatrix}, \quad \beta _2=\begin{pmatrix} 0 &{} -\mathrm {i}\\ \mathrm {i}&{} 0 \end{pmatrix}, \end{aligned}$$

then, by setting \(\mathcal {N}:=\nu _1+\mathrm {i}\nu _2\) and \(\mathcal {T}:=\tau _1+\mathrm {i}\tau _2\),

$$\begin{aligned} \beta (\nu )=\begin{pmatrix} 0 &{} \overline{\mathcal {N}}\\ \mathcal {N}&{} 0 \end{pmatrix}, \quad \beta (e)= \begin{pmatrix} 0 &{} \overline{\mathcal {T}}\\ \mathcal {T}&{} 0 \end{pmatrix} \end{aligned}$$

Using (11) we realize \(D^\Sigma \) as an operator in \(L^2(\mathbb {T},\mathbb {C}^2)\): for \(\psi \in C^\infty (\mathbb {T},\mathbb {C}^2)\) one has

$$\begin{aligned} D^\Sigma \psi =\Big (\dfrac{H_1}{2}-\beta (\nu ) \sum _{j=1}^{n-1} \beta (e_j) \nabla _{e_j}\Big ) \psi = \dfrac{\kappa }{2}\,\psi - \begin{pmatrix} 0 &{} \overline{\mathcal {N}}\\ \mathcal {N}&{} 0 \end{pmatrix} \begin{pmatrix} 0 &{} \overline{\mathcal {T}}\\ \mathcal {T}&{} 0 \end{pmatrix} \psi '. \end{aligned}$$

where \(\kappa \) is the curvature of \(\Sigma \). The above choice of orientation gives \(\mathcal {T}=\mathrm {i}\, \mathcal {N}\) and

$$\begin{aligned} D^\Sigma \psi = \dfrac{\kappa }{2}\,\psi +\mathrm {i}\begin{pmatrix} 1 &{} 0 \\ 0 &{} -1\end{pmatrix} \psi '. \end{aligned}$$

Consider the function \(K:\mathbb {R}\rightarrow \mathbb {R}\) given by

$$\begin{aligned} K(s):=\dfrac{1}{2}\int _0^s \kappa (t)\,\mathrm {d}t. \end{aligned}$$

Using the well-known identity

$$\begin{aligned} \int _0^\ell \kappa (t)\,\mathrm {d}t=2\pi \end{aligned}$$

we conclude that \(K(\cdot +\ell )=\pi + K\). Hence, using the unitary transform

$$\begin{aligned} U:L^2( (0,\ell ), \mathbb {C}^2)\phi \mapsto \begin{pmatrix} e^{\mathrm {i}K} &{} 0 \\ 0 &{} e^{-\mathrm {i}K} \end{pmatrix}\,\phi \in L^2(\mathbb {T},\mathbb {C}^2) \end{aligned}$$

we rewrite

$$\begin{aligned} U^{-1}D^\Sigma U \phi = \mathrm {i}\begin{pmatrix} 1 &{} 0 \\ 0 &{} -1\end{pmatrix} \phi ', \end{aligned}$$

and remark that \(U\phi \in C^\infty (\mathbb {T},\mathbb {C}^2)\) if and only if \(\phi \) extends to a function from \(C^\infty (\mathbb {R},\mathbb {C}^2)\) with \(\phi (\cdot +\ell )=-\phi \).

It follows that \(D^\Sigma \) is unitarily equivalent to \(D\oplus (-D)\), where D is the operator \(\phi \rightarrow -\mathrm {i}\phi '\) on \((0,\ell )\) with the antiperiodic boundary condition \(\phi (\ell )=-\phi (0)\), and one easily shows that the eigenvalues of D are \((2r-1)\pi /\ell \), \(r\in \mathbb {Z}\). In addition, as the dimension \(n=2\) is even, the operator \(D^\Sigma \) is also unitarily equivalent to , which means that has the same eigenvalues as D.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Moroianu, A., Ourmières-Bonafos, T. & Pankrashkin, K. Dirac Operators on Hypersurfaces as Large Mass Limits. Commun. Math. Phys. 374, 1963–2013 (2020). https://doi.org/10.1007/s00220-019-03642-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-019-03642-x

Navigation