Log in

Eigenvalue Curves for Generalized MIT Bag Models

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

Abstract

We study spectral properties of Dirac operators on bounded domains \(\Omega \subset {\mathbb {R}}^3\) with boundary conditions of electrostatic and Lorentz scalar type and which depend on a parameter \(\tau \in \mathbb {R}\); the case \(\tau = 0\) corresponds to the MIT bag model. We show that the eigenvalues are parametrized as increasing functions of \(\tau \), and we exploit this monotonicity to study the limits as \(\tau \rightarrow \pm \infty \). We prove that if \(\Omega \) is not a ball then the first positive eigenvalue is greater than the one of a ball with the same volume for all \(\tau \) large enough. Moreover, we show that the first positive eigenvalue converges to the mass of the particle as \(\tau \downarrow -\infty \), and we also analyze its first order asymptotics.

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

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (Brazil)

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. From the proof of [52, Theorem VII.1.3], it is enough to have \(({\mathcal {H}}_\tau - \zeta )^{-1}\) holomorphic in \(\tau \) for a single \(\zeta \in \rho ({\mathcal {H}}_0)\cap \rho ({\mathcal {H}}_\tau )\), thus we can take any \(\zeta \in \mathbb {C}\setminus \mathbb {R}\).

  2. Note that the kernel \(k_m\) of \(K_m\) does not depend on m. Hence, we can assume here that \(m>0\) to cover as well the case \(m=0\).

References

  1. Abramowitz, M., Stegun, I.A. (eds.): Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, vol. 55. US Government Printing Office, Washington, DC (1964)

    MATH  Google Scholar 

  2. Agricola, I., Friedrich, T.: Upper bounds for the first eigenvalue of the Dirac operator on surfaces. J. Geom. Phys. 30, 1–22 (1999)

    ADS  MATH  Google Scholar 

  3. Akhmerov, A.R., Beenakker, C.W.J.: Boundary conditions for Dirac fermions on a terminated honeycomb lattice. Phys. Rev. B 77, 085423 (2008)

    ADS  Google Scholar 

  4. Ammann, B., Bär, C.: Dirac eigenvalue estimates on surfaces. Math. Z. 240, 423–449 (2002)

    MATH  Google Scholar 

  5. Antunes, P.R.S., Benguria, R.D., Lotoreichik, V., Ourmières-Bonafos, T.: A variational formulation for Dirac operators in bounded domains. Applications to spectral geometric inequalities. Commun. Math. Phys. 386, 781–818 (2021)

    ADS  MATH  Google Scholar 

  6. 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)

    MATH  Google Scholar 

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

    ADS  MATH  Google Scholar 

  8. Arrizabalaga, N., Le Treust, L., Raymond, N.: Extension operator for the MIT bag model. Ann. Fac. Sci. Toulouse Math. (6) 29, 135–147 (2020)

    MATH  Google Scholar 

  9. Arrizabalaga, N., Mas, A., Vega, L.: Shell interactions for Dirac operators. J. Math. Pures Appl. (9) 102, 617–639 (2014)

    MATH  Google Scholar 

  10. Arrizabalaga, N., Mas, A., Vega, L.: Shell interactions for Dirac operators: on the point spectrum and the confinement. SIAM J. Math. Anal. 47, 1044–1069 (2015)

    MATH  Google Scholar 

  11. Arrizabalaga, N., Mas, A., Vega, L.: An isoperimetric-type inequality for electrostatic shell interactions for Dirac operators. Commun. Math. Phys. 344, 483–505 (2016)

    ADS  MATH  Google Scholar 

  12. Bär, C.: Lower eigenvalue estimates for Dirac operators. Math. Ann. 293, 39–46 (1992)

    MATH  Google Scholar 

  13. Bär, C.: Extrinsic bounds for eigenvalues of the Dirac operator. Ann. Glob. Anal. Geom. 16, 573–596 (1998)

    MATH  Google Scholar 

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

    ADS  MATH  Google Scholar 

  15. Barbaroux, J.-M., Le Treust, L., Raymond, N., Stockmeyer, E.: On the Dirac bag model in strong magnetic fields. ar**v:2007.03242 (2020)

  16. Behrndt, J., Exner, P., Holzmann, M., Lotoreichik, V.: On the spectral properties of Dirac operators with electrostatic \(\delta \)-shell interactions. J. Math. Pures Appl. (9) 111, 47–78 (2018)

    MATH  Google Scholar 

  17. Behrndt, J., Exner, P., Holzmann, M., Lotoreichik, V.: On Dirac operators in \(\mathbb{R} ^3\) with electrostatic and Lorentz scalar \(\delta \)-shell interactions. Quantum Stud. Math. Found. 6, 295–314 (2019)

    MATH  Google Scholar 

  18. Behrndt, J., Gesztesy, F., Mitrea, M.: Sharp boundary trace theory and Schrödinger operators on bounded Lipschitz domains, arxiv:2209.09230 (2021)

  19. Behrndt, J., Holzmann, M.: On Dirac operators with electrostatic \(\delta \)-shell interactions of critical strength. J. Spectr. Theory 10, 147–184 (2020)

    MATH  Google Scholar 

  20. Behrndt, J., Holzmann, M., Mantile, A., Posilicano, A.: Limiting absorption principle and scattering matrix for Dirac operators with \(\delta \)-shell interactions. J. Math. Phys. 61, 033504 (2020)

    ADS  MATH  Google Scholar 

  21. Behrndt, J., Holzmann, M., Mas, A.: Self-adjoint Dirac operators on domains in \({\mathbb{R} }^3\). Ann. Henri Poincaré 21, 2681–2735 (2020)

    ADS  MATH  Google Scholar 

  22. Behrndt, J., Holzmann, M., Ourmières-Bonafos, T., Pankrashkin, K.: Two-dimensional Dirac operators with singular interactions supported on closed curves. J. Funct. Anal. 279, 108700 (2020)

    MATH  Google Scholar 

  23. Benguria, R.D., Fournais, S., Stockmeyer, E., Van Den Bosch, H.: Self-adjointness of two-dimensional Dirac operators on domains. Ann. Henri Poincaré 18, 1371–1383 (2017)

    ADS  MATH  Google Scholar 

  24. Benguria, R.D., Fournais, S., Stockmeyer, E., Van Den Bosch, H.: Spectral gaps of Dirac operators describing graphene quantum dots. Math. Phys. Anal. Geom. 20, 12 (2017)

    MATH  Google Scholar 

  25. Benhellal, B.: Spectral asymptotic for the infinite mass Dirac operator in bounded domain. ar**v:1909.03769 (2019)

  26. Benhellal, B.: Spectral properties of the Dirac operator coupled with \(\delta \)-shell interactions. ar**v:2102.10207 (2021)

  27. Benhellal, B.: Spectral analysis of Dirac operators with delta interactions supported on the boundaries of rough domains. J. Math. Phys. 63, 011507 (2022)

    ADS  MATH  Google Scholar 

  28. Berry, M.V., Mondragon, R.J.: Neutrino billiards: time-reversal symmetry-breaking without magnetic fields. Proc. Roy. Soc. Lond. Ser. A 412, 53–74 (1987)

    ADS  Google Scholar 

  29. Bogolioubov, P.N.: Sur un modèle à quarks quasi-indépendants. Ann. Inst. H. Poincaré Sect. A 8, 163–189 (1968)

    Google Scholar 

  30. Briet, P., Krejčiřík, D.: Spectral optimisation of Dirac rectangles. J. Math. Phys. 63, 013502 (2022)

    ADS  MATH  Google Scholar 

  31. Bucur, D., Freitas, P., Kennedy, J.: The Robin problem. In: Henrot, A. (ed.) Shape Optimization and Spectral Theory, pp. 78–119. De Gruyter Open, Warsaw (2017)

    Google Scholar 

  32. Cassano, B., Lotoreichik, V.: Self-adjoint extensions of the two-valley Dirac operator with discontinuous infinite mass boundary conditions. Oper. Matrices 14, 667–678 (2020)

    MATH  Google Scholar 

  33. Cassano, B., Lotoreichik, V., Mas, A., Tušek, M.: General \(\delta \)-shell interactions for the two-dimensional Dirac operator: self-adjointness and approximation. Rev. Mat. Iberoam. (2022). https://doi.org/10.4171/RMI/1354

    Article  Google Scholar 

  34. Castro Neto, A.H., Guinea, F., Peres, N.M.R., Novoselov, K.S., Geim, A.K.: The electronic properties of graphene. Rev. Mod. Phys. 81, 109–162 (2009)

    ADS  Google Scholar 

  35. Chami, F.E., Ginoux, N., Habib, G.: New eigenvalue estimates involving Bessel functions. Publ. Mat. 65, 681–726 (2021)

    MATH  Google Scholar 

  36. Chipot, M.: Handbook of Differential Equations: Stationary Partial Differential Equations, vol. 4. Elsevier, Amsterdam (2011)

    MATH  Google Scholar 

  37. Chodos, A., Jaffe, R.L., Johnson, K., Thorn, C.B., Weisskopf, V.F.: New extended model of hadrons. Phys. Rev. D (3) 9, 3471–3495 (1974)

    ADS  Google Scholar 

  38. DeGrand, T., Jaffe, R.L., Johnson, K., Kiskis, J.: Masses and other parameters of the light hadrons. Phys. Rev. D 12, 2060–2076 (1975)

    ADS  Google Scholar 

  39. Dittrich, J., Exner, P., Šeba, P.: Dirac operators with a spherically symmetric \(\delta \) -shell interaction. J. Math. Phys. 30, 2875–2882 (1989)

    ADS  MATH  Google Scholar 

  40. Evans, L.: Partial Differential Equations, vol. 19, 2nd edn. American Mathematical Society, Providence (2010)

    MATH  Google Scholar 

  41. Folland, G.: Introduction to Partial Differential Equations, vol. 102. Princeton University Press, Princeton (1995)

    MATH  Google Scholar 

  42. Freitas, P., Siegl, P.: Spectra of graphene nanoribbons with armchair and zigzag boundary conditions. Rev. Math. Phys. 26, 1450018 (2014)

    MATH  Google Scholar 

  43. Güçlü, A.D., Potasz, P., Korkusinski, M., Hawrylak, P.: Graphene quantum dots. Springer, Berlin (2014)

    Google Scholar 

  44. Henrot, A.: Extremum Problems for Eigenvalues of Elliptic Operators. Frontiers in Mathematics. Birkhäuser, Basel (2006)

    MATH  Google Scholar 

  45. Hijazi, O.: A conformal lower bound for the smallest eigenvalue of the Dirac operator and Killing spinors. Commun. Math. Phys. 104, 151–162 (1986)

    ADS  MATH  Google Scholar 

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

    ADS  MATH  Google Scholar 

  47. Hijazi, O., Montiel, S., Zhang, X.: Eigenvalues of the Dirac operator on manifolds with boundary. Commun. Math. Phys. 221, 255–265 (2001)

    ADS  MATH  Google Scholar 

  48. Hofmann, S., Marmolejo-Olea, E., Mitrea, M., Pérez-Esteva, S., Taylor, M.: Hardy spaces, singular integrals and the geometry of Euclidean domains of locally finite perimeter. Geom. Funct. Anal. 19, 842–882 (2009)

    MATH  Google Scholar 

  49. Holzmann, M.: A note on the three dimensional Dirac operator with zigzag type boundary conditions. Complex Anal. Oper. Theory 15, 15 (2021)

    MATH  Google Scholar 

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

    MATH  Google Scholar 

  51. Johnson, K.: The MIT bag model. Acta Phys. Pol. B 6, 865–892 (1975)

    Google Scholar 

  52. Kato, T.: Perturbation theory for linear operators. In: Classics in Mathematics. Springer, Berlin (1995). Reprint of the 1980 edition

  53. Kramer, W., Semmelmann, U., Weingart, G.: The first eigenvalue of the Dirac operator on quaternionic Kähler manifolds. Commun. Math. Phys. 199, 327–349 (1998)

    ADS  MATH  Google Scholar 

  54. Krejčiřík, D., Larson, S., Lotoreichik, V. (eds.): Problem List of the AIM Workshop: Shape Optimization with surface Interactions, San Jose, USA, 2019. http://aimpl.org/shapesurface

  55. Le Treust, L., Ourmières-Bonafos, T.: Self-adjointness of Dirac operators with infinite mass boundary conditions in sectors. Ann. Henri Poincaré 19, 1465–1487 (2018)

    ADS  MATH  Google Scholar 

  56. Lotoreichik, V., Ourmières-Bonafos, T.: A sharp upper bound on the spectral gap for graphene quantum dots. Math. Phys. Anal. Geom. 22, 1–30 (2019)

    MATH  Google Scholar 

  57. Mas, A.: Dirac operators, shell interactions, and discontinuous gauge functions across the boundary. J. Math. Phys. 58, 022301 (2017)

    ADS  MATH  Google Scholar 

  58. Mas, A., Pizzichillo, F.: The relativistic spherical \(\delta \)-shell interaction in \({\mathbb{R} }^3\): spectrum and approximation. J. Math. Phys. 58, 082102 (2017)

    ADS  MATH  Google Scholar 

  59. Mas, A., Pizzichillo, F.: Klein’s paradox and the relativistic \(\delta \)-shell interaction in \({\mathbb{R}}^3\). Anal. PDE 11, 705–744 (2018)

    MATH  Google Scholar 

  60. McCann, E., Fal’ko, V.I.: Symmetry of boundary conditions of the Dirac equation for electrons in carbon nanotubes. J. Phys. Condens. Matter 16, 2371 (2004)

    ADS  Google Scholar 

  61. Moroianu, A., Ourmières-Bonafos, T., Pankrashkin, K.: Dirac operators on hypersurfaces as large mass limits. Commun. Math. Phys. 374, 1963–2013 (2020)

    ADS  MATH  Google Scholar 

  62. Nédélec, J.-C.: Acoustic and Electromagnetic Equations. Applied Mathematical Sciences, vol. 144. Springer, New York (2001)

    Google Scholar 

  63. Ourmières-Bonafos, T., Pizzichillo, F.: Dirac operators and Shell interactions: a survey. In: Mathematical Challenges of Zero-Range Physics. Springer INdAM Series, vol. 42, pp. 105–131. Springer, Cham (2021)

  64. Ourmières-Bonafos, T., Vega, L.: A strategy for self-adjointness of Dirac operators: applications to the MIT bag model and \(\delta \)-shell interactions. Publ. Mat. 62, 397–437 (2018)

    MATH  Google Scholar 

  65. Pizzichillo, F., Van Den Bosch, H.: Self-adjointness of two dimensional Dirac operators on corner domains. J. Spectr. Theory 11, 1043–1079 (2021)

  66. Ponomarenko, L.A., Schedin, F., Katsnelson, M.I., Yang, R., Hill, E.W., Novoselov, K.S., Geim, A.K.: Chaotic Dirac billiard in graphene quantum dots. Science 320, 356–358 (2008)

    ADS  Google Scholar 

  67. Rabinovich, V.S.: Boundary problems for three-dimensional Dirac operators and generalized MIT bag models for unbounded domains. Russ. J. Math. Phys. 27, 500–516 (2020)

    MATH  Google Scholar 

  68. Rabinovich, V.S.: Fredholm property and essential spectrum of 3-D Dirac operators with regular and singular potentials. Complex Var. Elliptic Equ. 67,1–4 (2020)

  69. Sauter, S.A., Schwab, C.: Boundary Element Methods. Springer Series in Computational Mathematics, vol. 39. Springer, Berlin (2011)

    Google Scholar 

  70. Stockmeyer, E., Vugalter, S.: Infinite mass boundary conditions for Dirac operators. J. Spectr. Theory 9, 569–600 (2019)

    MATH  Google Scholar 

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

    Google Scholar 

  72. Watson, G.N.: A treatise on the theory of Bessel functions. In: Cambridge Mathematical Library. Cambridge University Press, Cambridge (1995). Reprint of the second (1944) edition

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tomás Sanz-Perela.

Ethics declarations

Funding

All authors are supported by the ERC-2014-ADG project HADE Id. 669689 (European Research Council). N. A. is supported by the MINECO Grant PGC2018-094522-B-I00 (Spain) and IT1247-19 (Gobierno Vasco). A. M. is supported by Grants MTM2017-84214-C2-1-P and RED2018-102650-T funded by MCIN/AEI/10.13039/501100011033 and by “ERDF A way of making Europe”, by MINECO Grant MTM2017-83499-P (Spain), and by the Spanish State Research Agency, through the Severo Ochoa and María de Maeztu Program for Centers and Units of Excellence in R &D (CEX2020-001084-M). T. S.-P. is supported by Grants MTM2017-84214-C2-1-P and RED2018-102650-T funded by MCIN/AEI/10.13039/501100011033 and by “ERDF A way of making Europe”, AGAUR research group 2017-SGR-1392 (Catalunya), and EPSRC Grant EP/S03157X/1. L. V. is supported by the Basque Government through the BERC 2018-2021 program and by the Spanish State Research Agency through BCAM Severo Ochoa excellence accreditation SEV-2017-0718.

Conflict of interest

The authors have no conflicts of interest to declare that are relevant to the content of this article.

Availability of data and materials

Not applicable.

Code availability

Not applicable.

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

Appendix A. Properties of the Spectrum

Here we prove the properties of the spectrum of \({\mathcal {H}}_\tau \) collected in Lemma 1.2. For a shorter notation, we will use

$$\begin{aligned} {\mathcal {B}} := -i \beta (\alpha \cdot \nu ), \end{aligned}$$

which defines a self-adjoint operator in \(L^2(\partial \Omega )^4\) such that \({\mathcal {B}}^2 = I_4\). For every \(\varphi \in \text {Dom}({\mathcal {H}}_\tau )\) it holds \(\varphi = \sinh \tau {\mathcal {B}} \beta \varphi + \cosh \tau {\mathcal {B}} \varphi \) on \(\partial \Omega \).

Proof of Lemma 1.2

From [21, Proposition 5.15] we know that \({\mathcal {H}}_\tau \) is self-adjoint in \(L^2(\Omega )^4\). The proof of (i) follows from this and the compact embedding of \(H^1(\Omega )^4\) into \(L^2(\Omega )^4\).

Let us now show (ii). We first claim that for every \(\varphi = (u,v)^\intercal \in \text {Dom}({\mathcal {H}}_\tau )\) it holds

$$\begin{aligned} \left\| { {\mathcal {H}}\varphi } \right\| ^2_{L^2(\Omega )^4} = \left\| { \alpha \cdot \nabla \varphi } \right\| ^2_{L^2(\Omega )^4} + m^2 \left\| {\varphi } \right\| ^2_{L^2(\Omega )^4} + m e^\tau \left\| {u} \right\| ^2_{L^2(\partial \Omega )^2} + m e^{-\tau } \left\| {v} \right\| ^2_{L^2(\partial \Omega )^2}.\nonumber \\ \end{aligned}$$
(A.1)

To prove this formula, note first that expanding the square we have

$$\begin{aligned} \left\| { {\mathcal {H}}\varphi } \right\| ^2_{L^2(\Omega )^4}= & {} \langle -i \alpha \cdot \nabla \varphi , -i \alpha \cdot \nabla \varphi \rangle _{L^2(\Omega )^4} + m^2 \langle \beta \varphi , \beta \varphi \rangle _{L^2(\Omega )^4} \\{} & {} + 2 m \Re \langle \beta \varphi , -i \alpha \cdot \nabla \varphi \rangle _{L^2(\Omega )^4} \\= & {} \left\| { \alpha \cdot \nabla \varphi } \right\| ^2_{L^2(\Omega )^4} + m^2 \left\| {\varphi } \right\| ^2_{L^2(\Omega )^4} + 2 m \Re \langle \beta \varphi , -i \alpha \cdot \nabla \varphi \rangle _{L^2(\Omega )^4} . \end{aligned}$$

Integrating by parts we get

$$\begin{aligned} \begin{aligned} \langle \beta \varphi , -i \alpha \cdot \nabla \varphi \rangle _{L^2(\Omega )^4}&= \langle -i \alpha \cdot \nabla (\beta \varphi ), \varphi \rangle _{L^2(\Omega )^4} - \langle -i (\alpha \cdot \nu ) \beta \varphi , \varphi \rangle _{L^2(\partial \Omega )^4} \\&= - \langle \beta (-i \alpha \cdot \nabla )\varphi , \varphi \rangle _{L^2(\Omega )^4} + \langle {\mathcal {B}} \varphi , \varphi \rangle _{L^2(\partial \Omega )^4} , \end{aligned} \end{aligned}$$

and using that \(\beta \) is hermitian, we obtain \(2 \Re \langle \beta \varphi , -i \alpha \cdot \nabla \varphi \rangle _{L^2(\Omega )^4} = \langle {\mathcal {B}} \varphi , \varphi \rangle _{L^2(\partial \Omega )^4}\). Thus,

$$\begin{aligned} \left\| { {\mathcal {H}}\varphi } \right\| ^2_{L^2(\Omega )^4} = \left\| { \alpha \cdot \nabla \varphi } \right\| ^2_{L^2(\Omega )^4} + m^2 \left\| {\varphi } \right\| ^2_{L^2(\Omega )^4} + m \langle {\mathcal {B}} \varphi , \varphi \rangle _{L^2(\partial \Omega )^4}. \end{aligned}$$
(A.2)

Now, using that \(\varphi \in \text {Dom}({\mathcal {H}}_{\tau })\), we see that

$$\begin{aligned} \langle {\mathcal {B}} \varphi , \varphi \rangle _{L^2(\partial \Omega )^4}= & {} \langle {\mathcal {B}} \varphi , \sinh \tau {\mathcal {B}} \beta \varphi + \cosh \tau {\mathcal {B}} \varphi \rangle _{L^2(\partial \Omega )^4} \\= & {} \langle \varphi , \sinh \tau {\mathcal {B}}^2 \beta \varphi + \cosh \tau {\mathcal {B}}^2 \varphi \rangle _{L^2(\partial \Omega )^4} \\= & {} \langle \varphi , \sinh \tau \beta \varphi + \cosh \tau \varphi \rangle _{L^2(\partial \Omega )^4}, \end{aligned}$$

where we have used that \({\mathcal {B}}\) is self-adjoint and that \({\mathcal {B}}^2 = I_4\). From here, and writing \(\varphi = (u, v)^\intercal \), we easily see that

$$\begin{aligned} \langle {\mathcal {B}} \varphi , \varphi \rangle _{L^2(\partial \Omega )^4} = e^\tau \left\| {u} \right\| ^2_{L^2(\partial \Omega )^2} + e^{-\tau } \left\| {v} \right\| ^2_{L^2(\partial \Omega )^2}. \end{aligned}$$

Plugging this into (A.2) we obtain (A.1), proving the claim. Now, let \(\varphi \in \text {Dom}({\mathcal {H}}_\tau )\setminus \{0\}\) be such that \({\mathcal {H}}\varphi = \uplambda \varphi \) in \(\Omega \). Note that, by Lemma 2.4, \(\varphi \) cannot vanish identically on \(\partial \Omega \). Therefore, using (A.1) we obtain

$$\begin{aligned} \uplambda ^2 \left\| {\varphi } \right\| ^2_{L^2(\Omega )^4} = \left\| { {\mathcal {H}}\varphi } \right\| ^2_{L^2(\Omega )^4} > m^2 \left\| {\varphi } \right\| ^2_{L^2(\Omega )^4}, \end{aligned}$$

which yields \(|\uplambda | > m\).

We finally prove (iii) and (iv). First, by the compact embedding of \(H^1(\Omega )^4\) into \(L^2(\Omega )^4\) we have that the resolvent of \({\mathcal {H}}_\tau \) is a compact operator, which yields that every eigenvalue has finite multiplicity. Now, given \(\psi \in \mathbb {C}^4\), consider the charge conjugation operator

$$\begin{aligned} {\mathcal {C}} \psi := i \beta \alpha _{2} {\overline{\psi }} \end{aligned}$$

and the time reversal-symmetry operator

$$\begin{aligned} T \psi : =-i \gamma _{5} \alpha _{2} {\overline{\psi }}, \quad \text {where}\quad \gamma _{5}:= \begin{pmatrix}0&{}I_2\\ I_2&{}0\end{pmatrix}. \end{aligned}$$
(A.3)

Then, simple computations show that \({\mathcal {H}}T=T {\mathcal {H}}\), \({\mathcal {H}}{\mathcal {C}}=-{\mathcal {C}} {\mathcal {H}}\), and \( T {\mathcal {C}}= {\mathcal {C}} T\). In addition, setting

$$\begin{aligned} {\mathcal {B}}_\tau := \sinh \tau {\mathcal {B}} \beta + \cosh \tau {\mathcal {B}} = i (\sinh \tau - \cosh \tau \beta ) (\alpha \cdot \nu ), \end{aligned}$$

it is also easy to check that \({\mathcal {B}}_{\tau } T=T {\mathcal {B}}_{\tau }\) and \( {\mathcal {B}}_{\tau } {\mathcal {C}}={\mathcal {C}} {\mathcal {B}}_{-\tau }\). Note that for every function \(\varphi \in \text {Dom}({\mathcal {H}}_{\tau })\) it holds \(\varphi = {\mathcal {B}}_\tau \varphi \) on \(\partial \Omega \). As a consequence, given an eigenfunction \(\varphi \) of \({\mathcal {H}}_\tau \) with eigenvalue \(\uplambda \), \(T\varphi \) is also an eigenfunction of \({\mathcal {H}}_\tau \) with eigenvalue \(\uplambda \). Furthermore, \({\mathcal {C}}\varphi \) and \(T{\mathcal {C}} \varphi \) are eigenfunctions of \({\mathcal {H}}_{-\tau }\) with eigenvalue \(-\uplambda \). \(\square \)

To conclude this section, we establish a formula which relates the \(L^2(\Omega )^4\)-norms of \(\nabla \varphi \) and \(\alpha \cdot \nabla \varphi \) for functions \(\varphi \in \text {Dom} ({\mathcal {H}}_\tau )\). Although we do not use this formula in this article, we think that it has its own interest, and it may be useful to present it here for future reference. The formula is a generalization of [7, formula (1.3)], in which the case \(\tau = 0\) is considered, and the proof follows the same lines.

Lemma A.1

Let \(\tau \in \mathbb {R}\) and \(\varphi \in \text {Dom} ({\mathcal {H}}_\tau ) \cap H^1(\partial \Omega )^4\). Then,

$$\begin{aligned}{} & {} \left\| { \alpha \cdot \nabla \varphi } \right\| ^2_{L^2(\Omega )^4} \\{} & {} = \left\| { \nabla \varphi } \right\| ^2_{L^2(\Omega )^4} + \dfrac{1}{2} \int _{\partial \Omega } \kappa |\varphi |^2 d \upsigma + \sinh \tau \langle \gamma _5 (\alpha \cdot \nu ) \varphi , \alpha \cdot (\nu \times \nabla ) \varphi \rangle _{L^2(\partial \Omega )^4}, \end{aligned}$$

where \(\kappa \) denotes the mean curvature of \(\partial \Omega \).

Proof

First, for every \(\varphi \in H^2(\Omega )^4\), it holds

$$\begin{aligned} \left\| { \alpha \cdot \nabla \varphi } \right\| ^2_{L^2(\Omega )^4} = \left\| { \nabla \varphi } \right\| ^2_{L^2(\Omega )^4} - \langle \gamma _5 \varphi , -i \alpha \cdot (\nu \times \nabla ) \varphi \rangle _{L^2(\partial \Omega )^4}, \end{aligned}$$
(A.4)

where \(\gamma _5\) is defined in (A.3). This is proved in [7, Appendix A.2]. By density, it also holds for all \(\varphi \in H^1(\Omega )^4 \cap H^1(\partial \Omega )^4\).

Let us now investigate the boundary term in the above expression. The crucial point is to use that the mean curvature of \(\partial \Omega \) arises in our context through the formula

$$\begin{aligned}{}[-i \alpha \cdot (\nu \times \nabla ), {\mathcal {B}}]= - \kappa \gamma _5 {\mathcal {B}}, \end{aligned}$$

where \([\cdot , \cdot ]\) denotes the commutator of two operators, i.e., \([S,T] := ST - TS\); see [7, Lemma A.3]. Using this and the boundary condition for \(\varphi \in \text {Dom}({\mathcal {H}}_\tau )\cap H^1(\partial \Omega )^4\) we get

$$\begin{aligned} \begin{aligned} \langle \gamma _5 \varphi&,-i \alpha \cdot (\nu \times \nabla ) \varphi \rangle _{L^2(\partial \Omega )^4} \\&= \sinh \tau \langle \gamma _5 \varphi , -i \alpha \cdot (\nu \times \nabla ) {\mathcal {B}} \beta \varphi \rangle _{L^2(\partial \Omega )^4} + \cosh \tau \langle \gamma _5 \varphi , -i \alpha \cdot (\nu \times \nabla ) {\mathcal {B}}\varphi \rangle _{L^2(\partial \Omega )^4} \\&= \sinh \tau \langle \gamma _5 \varphi , [-i \alpha \cdot (\nu \times \nabla ), {\mathcal {B}}] \beta \varphi \rangle _{L^2(\partial \Omega )^4} + \sinh \tau \langle \gamma _5 \varphi , {\mathcal {B}} ( -i \alpha \cdot (\nu \times \nabla ) ) \beta \varphi \rangle _{L^2(\partial \Omega )^4} \\&\quad \quad + \cosh \tau \langle \gamma _5 \varphi , [-i \alpha \cdot (\nu \times \nabla ), {\mathcal {B}}]\varphi \rangle _{L^2(\partial \Omega )^4} + \cosh \tau \langle \gamma _5 \varphi , {\mathcal {B}} (-i \alpha \cdot (\nu \times \nabla ))\varphi \rangle _{L^2(\partial \Omega )^4} \\&= -\sinh \tau \langle \gamma _5 \varphi , \kappa \gamma _5 {\mathcal {B}} \beta \varphi \rangle _{L^2(\partial \Omega )^4}+ \sinh \tau \langle {\mathcal {B}} \gamma _5 \varphi , -i \alpha \cdot (\nu \times \nabla ) \beta \varphi \rangle _{L^2(\partial \Omega )^4} \\&\quad \quad - \cosh \tau \langle \gamma _5 \varphi , \kappa \gamma _5 {\mathcal {B}} \varphi \rangle _{L^2(\partial \Omega )^4} + \cosh \tau \langle {\mathcal {B}} \gamma _5 \varphi , -i \alpha \cdot (\nu \times \nabla ) \varphi \rangle _{L^2(\partial \Omega )^4} \\&= - \langle \gamma _5 \varphi , \kappa \gamma _5 (\sinh \tau {\mathcal {B}} \beta + \cosh \tau {\mathcal {B}} )\varphi \rangle _{L^2(\partial \Omega )^4} + \sinh \tau \langle {\mathcal {B}} \gamma _5 \varphi , -i \alpha \cdot (\nu \times \nabla ) \beta \varphi \rangle _{L^2(\partial \Omega )^4} \\&\quad \quad + \cosh \tau \langle {\mathcal {B}} \gamma _5 \varphi , -i \alpha \cdot (\nu \times \nabla ) \varphi \rangle _{L^2(\partial \Omega )^4} \\&= - \langle \gamma _5 \varphi , \kappa \gamma _5 \varphi \rangle _{L^2(\partial \Omega )^4} - \sinh \tau \langle \gamma _5 {\mathcal {B}} \varphi , -i \alpha \cdot (\nu \times \nabla ) \beta \varphi \rangle _{L^2(\partial \Omega )^4} \\&\quad \quad - \cosh \tau \langle \gamma _5 {\mathcal {B}}\varphi , -i \alpha \cdot (\nu \times \nabla ) \varphi \rangle _{L^2(\partial \Omega )^4} .\\ \end{aligned} \end{aligned}$$

Here we have used that \((\alpha \cdot x) \gamma _5 = \gamma _5 (\alpha \cdot x)\) for all \(x\in \mathbb {R}^3\) (see [7, Lemma A.1]) and that \(\beta \) anticommutes with \(\gamma _5\), thus \( {\mathcal {B}} \gamma _5 = - \gamma _5 {\mathcal {B}} \). Now, using that \((\alpha \cdot x) \beta = - \beta (\alpha \cdot x)\) for every \(x\in \mathbb {R}^3\), and that \(\beta \) anticommutes with \({\mathcal {B}}\) and \(\gamma _5\), we have

$$\begin{aligned} \begin{aligned}&\langle \gamma _5 {\mathcal {B}} \varphi , -i \alpha \cdot (\nu \times \nabla ) \beta \varphi \rangle _{L^2(\partial \Omega )^4} \\&= - \langle \gamma _5 {\mathcal {B}} \varphi , \beta (-i \alpha \cdot (\nu \times \nabla ) ) \varphi \rangle _{L^2(\partial \Omega )^4} \\&= - \langle \beta \gamma _5{\mathcal {B}} \varphi , -i \alpha \cdot (\nu \times \nabla ) \varphi \rangle _{L^2(\partial \Omega )^4} = -\langle \gamma _5 {\mathcal {B}} \beta \varphi , -i \alpha \cdot (\nu \times \nabla ) \varphi \rangle _{L^2(\partial \Omega )^4}. \end{aligned} \end{aligned}$$

Hence,

$$\begin{aligned} \begin{aligned}&\langle \gamma _5 \varphi , -i \alpha \cdot (\nu \times \nabla ) \varphi \rangle _{L^2(\partial \Omega )^4} \\&= - \langle \gamma _5 \varphi , \kappa \gamma _5 \varphi \rangle _{L^2(\partial \Omega )^4} + \sinh \tau \langle \gamma _5 {\mathcal {B}} \beta \varphi , -i \alpha \cdot (\nu \times \nabla ) \varphi \rangle _{L^2(\partial \Omega )^4} \\&\quad \quad - \cosh \tau \langle \gamma _5 {\mathcal {B}}\varphi , -i \alpha \cdot (\nu \times \nabla ) \varphi \rangle _{L^2(\partial \Omega )^4} \\&= - \langle \varphi , \kappa \varphi \rangle _{L^2(\partial \Omega )^4} + 2\sinh \tau \langle \gamma _5 {\mathcal {B}} \beta \varphi , -i \alpha \cdot (\nu \times \nabla ) \varphi \rangle _{L^2(\partial \Omega )^4} \\&\quad \quad - \langle \gamma _5 (\sinh \tau {\mathcal {B}} \beta + \cosh \tau {\mathcal {B}})\varphi , -i \alpha \cdot (\nu \times \nabla ) \varphi \rangle _{L^2(\partial \Omega )^4} \end{aligned} \end{aligned}$$

and thus, using the boundary condition, we get

$$\begin{aligned}{} & {} \langle \gamma _5 \varphi , -i \alpha \cdot (\nu \times \nabla ) \varphi \rangle _{L^2(\partial \Omega )^4} \\{} & {} = - \dfrac{1}{2} \int _{\partial \Omega } \kappa |\varphi |^2 d \upsigma + \sinh \tau \langle \gamma _5 {\mathcal {B}} \beta \varphi , -i \alpha \cdot (\nu \times \nabla ) \varphi \rangle _{L^2(\partial \Omega )^4}, \end{aligned}$$

which combined with (A.4) gives

$$\begin{aligned}{} & {} \left\| { \alpha \cdot \nabla \varphi } \right\| ^2_{L^2(\Omega )^4} \\{} & {} = \left\| { \nabla \varphi } \right\| ^2_{L^2(\Omega )^4} + \dfrac{1}{2} \int _{\partial \Omega } \kappa |\varphi |^2 d \upsigma - \sinh \tau \langle \gamma _5 {\mathcal {B}} \beta \varphi , -i \alpha \cdot (\nu \times \nabla ) \varphi \rangle _{L^2(\partial \Omega )^4}. \end{aligned}$$

Finally, using again that \((\alpha \cdot x) \beta = - \beta (\alpha \cdot x)\) for every \(x\in \mathbb {R}^3\), we have \({\mathcal {B}} \beta = i \alpha \cdot \nu \), and inserting this into the above identity we conclude the proof. \(\square \)

Appendix B. The Ball

In this appendix we present a more explicit spectral analysis in the case that \(\Omega \subset \mathbb {R}^3\) is a ball of radius \(R>0\) centered at the origin, which will be denoted by \(B_R\). To study this radially symmetric case we introduce spherical coordinates: if \(x\in \mathbb {R}^3\) we write \(x=r\theta \) with \(r = |x| \in [0,+\infty )\) and \(\theta = x/|x| \in \mathbb {S}^2\). Using separation of variables and the spherical harmonic spinors, we give the explicit equations for the eigenvalues and eigenfunctions of \({\mathcal {H}}_\tau \).

1.1 B.1 Decomposition using spherical harmonic spinors

Let \(Y^\ell _n\) be the usual spherical harmonics on \(\mathbb {S}^2\); here \(n = 0, 1, 2, \ldots \) and \(\ell = -n, -n+1, \ldots , n + 1, n\). They satisfy \(\Delta _{\mathbb {S}^2} Y^\ell _n = -n(n + 1)Y^\ell _n\), where \(\Delta _{\mathbb {S}^2}\) denotes the usual spherical Laplacian. Moreover, \(Y^\ell _n\) form a complete orthonormal set in \(L^2(\mathbb {S}^2)\).

Following [71, Section 4.6.4], the spherical harmonic spinors are defined as follows: for \(j=1/2, 3/2, \ldots \) and \(\mu _j = -j, -j+1, \ldots , j-1, j\), set

$$\begin{aligned} \psi _{j-1 / 2}^{\mu _j}= & {} \frac{1}{\sqrt{2 j}}\begin{pmatrix} \sqrt{j+\mu _j} Y_{j-1 / 2}^{\mu _j -1 / 2} \\ \sqrt{j-\mu _j} Y_{j-1 / 2}^{\mu _j +1 / 2} \end{pmatrix} \quad \text { and } \\ \quad \psi _{j+1 / 2}^{\mu _j}= & {} \frac{1}{\sqrt{2 j+2}}\begin{pmatrix} \sqrt{j+1-\mu _j} Y_{j+1 / 2}^{\mu _j -1 / 2} \\ -\sqrt{j+1+\mu _j} Y_{j+1 / 2}^{\mu _j +1 / 2} \end{pmatrix}. \end{aligned}$$

As shown in [71, Section 4.6.5], one can decompose the space \(L^2(\mathbb {R}^3)^4\)—and analogously \(L^2(B_R)^4\)—as

$$\begin{aligned} L^2(\mathbb {R}^3)^4 = \bigoplus _{j=1/2}^{+\infty } \, \bigoplus _{\mu _j=-j}^j L_{j, \mu _j}^+ \oplus L_{j, \mu _j}^-, \end{aligned}$$

where

$$\begin{aligned} L_{j, \mu _j}^\pm := \left\{ \varphi \in L^2(\mathbb {R}^3)^4 : \, \varphi (r\theta ) = \begin{pmatrix} i {\tilde{f}}(r) \psi _{j \pm 1/2}^{\mu _j} (\theta ) \\ {\tilde{g}}(r) \psi _{j {\mp } 1 / 2}^{\mu _j} (\theta ) \end{pmatrix} \text { with } {\tilde{f}}, {\tilde{g}} \in L^2(\mathbb {R}_+, r^2 d r) \right\} . \end{aligned}$$

In each subspace define the map** \(U_{j,\mu _j}^\pm : L^\pm _{j, \mu _j} \rightarrow L^2(\mathbb {R}_+)^2\) by

$$\begin{aligned} (U_{j,\mu _j}^\pm \varphi ) (r) = \begin{pmatrix} r {\tilde{f}}(r)\\ r {\tilde{g}}(r) \end{pmatrix} =: \begin{pmatrix} f(r)\\ g(r) \end{pmatrix}, \end{aligned}$$

and also define the differential operator (see [71, equation (4.129)])

$$\begin{aligned} {\widehat{H}}_{j,\pm } := \begin{pmatrix} m &{} - \partial _r + \kappa _{j,\pm }/r\\ \partial _r + \kappa _{j,\pm }/r &{} -m \end{pmatrix}, \quad \text {where } \kappa _{j,\pm } := \pm (j + 1/2). \end{aligned}$$

Then, the differential operator \({\mathcal {H}}= -i \alpha \cdot \nabla + m \beta \) decomposes into the orthogonal sum of the operators \((U_{j,\mu _j}^\pm )^{-1} {\widehat{H}}_{j,\pm } U_{j,\mu _j}^\pm \). In particular, if \(\phi = (f,g)^\intercal \) satisfies \({\widehat{H}}_{j,\pm } \phi = \uplambda \phi \) in (0, R), then

$$\begin{aligned} \varphi = \begin{pmatrix} u\\ v \end{pmatrix} = \begin{pmatrix} \dfrac{ i f(r)}{r} \psi ^{\mu _j}_{j \pm 1/2}\\ \dfrac{g(r)}{r} \psi ^{\mu _j}_{j {\mp } 1/2} \end{pmatrix} \end{aligned}$$
(B.1)

satisfies \({\mathcal {H}}\varphi = \uplambda \varphi \) in \(B_R\setminus \{0\}\). As we will see, by further imposing that f(0) is finite, we can guarantee that \({\mathcal {H}}\varphi = \uplambda \varphi \) holds across the origin.

1.2 B.2. Eigenvalue equations

Our first goal is to find solutions to \({\widehat{H}}_{j,\pm } (f, g)^\intercal = \uplambda (f,g)^\intercal \). This equation rewrites as the system of ODE

$$\begin{aligned} \left\{ \begin{array}{ll} - g' + \frac{\kappa }{r} g &{} = (\uplambda - m) f, \\ f' + \frac{\kappa }{r} f &{} = (\uplambda + m) g, \end{array} \right. \end{aligned}$$

where \(\kappa := \kappa _{j,\pm } := \pm (j + 1/2)\). For simplicity, let us assume first that \(\kappa = j + 1/2\). To solve the system, note that from the second ODE we get

$$\begin{aligned} g = \dfrac{1}{\uplambda + m} \Big (f' + \frac{\kappa }{r} f \Big ) \end{aligned}$$
(B.2)

and, thus, inserting this into the first one we get the Bessel-type ODE

$$\begin{aligned} f'' + \Big ( \uplambda ^2 - m^2 - \frac{\kappa ^2 + \kappa }{r^2} \Big ) f = 0. \end{aligned}$$

Therefore, f is of the form

$$\begin{aligned} f(r) = b_1 \sqrt{r} J_{\kappa + 1/2}(\sqrt{\uplambda ^2 - m^2} r) + b_2 \sqrt{r} Y_{\kappa + 1/2}( \sqrt{\uplambda ^2 - m^2} r), \end{aligned}$$

where \(b_1,b_2\in \mathbb {C}\), and \(J_{\kappa + 1/2}\) and \(Y_{\kappa + 1/2}\) denote the Bessel functions of the first and second kind of order \({\kappa + 1/2}\); see [1, Chapters 9 and 10]. Since the eigenfunctions are not allowed to be singular at \(r=0\) (as the corresponding \(\varphi \) given by (B.1) must solve an elliptic equation across the origin), we deduce that \(b_2=0\), and thus f is of the form

$$\begin{aligned} f(r) = b_1 \sqrt{r} J_{\kappa + 1/2}(\sqrt{\uplambda ^2 - m^2} r). \end{aligned}$$

Now, note that for every real index p, one has the relation

$$\begin{aligned} \partial _r [J_{p}(\sqrt{\uplambda ^2 - m^2} r) ]= \sqrt{\uplambda ^2 - m^2} J_{p - 1}(\sqrt{\uplambda ^2 - m^2} r) - \dfrac{p}{r} J_{p}(\sqrt{\uplambda ^2 - m^2} r); \end{aligned}$$

see [1, formula (9.1.27)]. Using this and (B.2), we see that

$$\begin{aligned} g(r) = b_1 \dfrac{\sqrt{\uplambda ^2 - m^2}}{\uplambda + m}\sqrt{r} J_{\kappa - 1/2}(\sqrt{\uplambda ^2 - m^2} r). \end{aligned}$$

The case \(\kappa = - j - 1/2\) follows by similar arguments. One isolates f instead of g and uses that, for a positive integer p, \(J_{-(p + 1/2)} = (-1)^{p+1} Y_{p + 1/2}\) and \(Y_{-(p+ 1/2)} = (-1)^{p} J_{p + 1/2}\).

As a conclusion, we obtain that every eigenfunction of \({\widehat{H}}_{j,\pm }\) with eigenvalue \(\uplambda \) is, up to a multiplicative constant, of the form

$$\begin{aligned} \begin{pmatrix} f(r)\\ g(r) \end{pmatrix} = \sqrt{r} \begin{pmatrix} J_{\ell + 1/2}(\sqrt{\uplambda ^2 - m^2} r) \\ \pm \dfrac{\sqrt{\uplambda ^2 - m^2}}{\uplambda + m} J_{\ell '+1/2}(\sqrt{\uplambda ^2 - m^2} r) \end{pmatrix}, \end{aligned}$$

where \(\ell = j \pm 1/2\) and \(\ell ' = j {\mp } 1/2\).

To obtain the equation (1.4) that relates \(\uplambda \) and \(\tau \) by means of Bessel functions, it only remains to impose the boundary condition \(v = i e^\tau (\sigma \cdot \nu ) u\) on \(\partial B_R\) for \(\varphi =(u,v)^\intercal \) as in (B.1) and satisfying \({\mathcal {H}}\varphi =\uplambda \varphi \) in \(B_R\). Since

$$\begin{aligned} (\sigma \cdot \nu ) \psi _{j \pm 1/2}^{\mu _j} = \psi _{j {\mp } 1/2}^{\mu _j}, \end{aligned}$$

by [71, equation (4.121)], it follows from (B.1) that the boundary condition relating f and g is

$$\begin{aligned} g(R) = - e^\tau f(R). \end{aligned}$$

Therefore, for each \(j = 1/2, 3/2, \ldots \), each \(\mu _j = -j, -j + 1,\ldots , j\), and each subspace \(L^\pm _{j, \mu _j}\), we obtain the eigenvalue equation

$$\begin{aligned} e^\tau J_{\ell + 1/2}(\sqrt{\uplambda ^2 - m^2} R) \pm \dfrac{\sqrt{\uplambda ^2 - m^2}}{\uplambda + m} J_{\ell '+1/2}(\sqrt{\uplambda ^2 - m^2} R) = 0, \end{aligned}$$
(B.3)

where \(\ell = j \pm 1/2\) and \(\ell ' = j {\mp } 1/2\). This corresponds to (1.4). Note that the equation is independent of the indexes \(\mu _j\), accounting for the multiplicity of the eigenvalues.

1.3 B.3. Parametrization of the eigenvalues

Our goal now is to exploit the eigenvalue equations given by (B.3) to prove that the eigenvalues of \({\mathcal {H}}_{\tau }\) can be parametrized in terms of \(\tau \in \mathbb {R}\), obtaining a family of increasing curves whose limits as \(\tau \rightarrow \pm \infty \) are related with the zeroes of the Bessel functions (and thus with the eigenvalues of the Dirichlet Laplacian). In the following lemma we collect the results on the Bessel functions that we will use.

Lemma B.1

Let \(J_p\) be the Bessel function of the first kind of order \(p >0\), and denote the k-th positive zero of this function by \(z_{p, k}\).

Then,

  1. (i)

    the positive zeroes of \(J_p\) are simple and form an infinite increasing sequence,

  2. (ii)

    the zeroes of two consecutive Bessel functions are interlaced, meaning that

    $$\begin{aligned} 0< z_{p, 1}< z_{p + 1, 1}< z_{p, 2}< z_{p + 1, 2} < \ldots , \end{aligned}$$
  3. (iii)

    the quotient of two consecutive Bessel functions can be expressed as

    $$\begin{aligned} \dfrac{J_{p + 1}(x)}{J_{p}(x)} = \sum _{k \ge 1} \dfrac{2x}{z_{p, k}^2 - x^2} \quad \text { for } x \in \mathbb {R}\setminus \{z_{p, k}\}_{k\in \mathbb {N}}. \end{aligned}$$
    (B.4)

    As a consequence, \(J_{p + 1}/J_{p}\) is odd, strictly increasing in each interval contained in \(\mathbb {R}\setminus \{z_{p, k}\}_{k\in \mathbb {N}}\), it is positive in the intervals \((0,z_{p, 1})\) and \((z_{p + 1, k},z_{p, k + 1})\) for \(k\ge 1\), and negative in the intervals \((z_{p, k},z_{p + 1, k})\) for \(k\ge 1\).

Proof

The first two statements are shown in [72, Chapter XV]. Note that for \(p \ge -1\) the zeroes of \(J_p\) are real, and thus we can order them; see also [1, p. 372]. Last, (B.4) follows from formula (1) in [72, p. 498]. Note that this yields that \(J_{p + 1}/J_{p}\) is an infinite sum of functions with singularities at \(\pm z_{p, k}\) which have strictly positive derivative in their domain of definition. As a consequence, in each interval of the form \((z_{p, k},z_{p, k + 1})\) for \(k\in \mathbb {N}\), the function \(J_{p + 1}/J_{p}\) is well defined, smooth, and strictly increasing, and therefore has a unique zero which necessarily is \(z_{p + 1, k}\). \(\square \)

With the help of the previous lemma we can now establish the following result on the parametrization of the eigenvalues.

Proposition B.2

For each index \(j = 1/2, 3/2, \ldots \) there exists an infinite number of smooth and strictly increasing functions \(\{\tau \mapsto \uplambda _{j, k}^\pm (\tau ) \}_{k \in \mathbb {Z}}\) such that, for each \(\tau \in \mathbb {R}\), the real number \(\uplambda _{j, k}^\pm (\tau )\) is an eigenvalue of \({\mathcal {H}}_{\tau }\). The functions \(\uplambda _{j, k}^\pm \) are surjectively defined by

$$\begin{aligned} \begin{matrix} \uplambda _{j, k}^\pm : &{}\mathbb {R}&{}\rightarrow &{}I_{j, k}^\pm \\ &{}\tau &{}\mapsto &{} \uplambda _{j, k}^\pm (\tau ), \end{matrix} \end{aligned}$$

with

  • \(I_{j,0}^-= \big (m, \sqrt{(z_{j,1}/R)^2 + m^2}\,\big )\),

  • \(I_{j,k}^- = \big (\sqrt{(z_{j + 1,k}/R)^2 + m^2}, \sqrt{(z_{j,k + 1}/R)^2 + m^2}\,\big )\) for \(k = 1, 2, \ldots \),

  • \(I_{j,k}^- = \big (\!-\!\sqrt{(z_{j + 1,|k|}/R)^2 + m^2}, - \sqrt{(z_{j,|k|}/R)^2 + m^2}\,\big )\) for \(k=-1, -2, \ldots \),

and

  • \(I_{j,0}^+= \big (\!-\!\sqrt{(z_{j,1}/R)^2 + m^2}, -m\big )\),

  • \(I_{j,k}^+ = \big (\!-\!\sqrt{(z_{j,k + 1}/R)^2 + m^2}, -\sqrt{(z_{j + 1,k}/R)^2 + m^2}\,\big )\) for \(k = 1, 2, \ldots \),

  • \(I_{j,k}^+ = \big (\sqrt{(z_{j,|k|}/R)^2 + m^2}, \sqrt{(z_{j + 1,|k|}/R)^2 + m^2}\,\big )\) for \(k=-1, -2, \ldots \),

where \(z_{p, k}\) denotes the k-th positive zero of \(J_p\), the Bessel function of order p. As a consequence, for every \(\tau \in \mathbb {R}\), the function

$$\begin{aligned} \varphi _\tau := \begin{pmatrix} u_\tau \\ v_\tau \end{pmatrix} = \dfrac{1}{\sqrt{r}} \begin{pmatrix} i J_{\ell + 1/2}\left( \sqrt{\uplambda _{j, k}^\pm (\tau )^2 - m^2} r \right) \, \psi ^{\mu _j}_{j \pm 1/2}(\theta ) \\ \pm \dfrac{\sqrt{\uplambda _{j, k}^\pm (\tau )^2 - m^2}}{\uplambda _{j, k}^\pm (\tau ) + m} J_{\ell '+1/2} \left( \sqrt{\uplambda _{j, k}^\pm (\tau )^2 - m^2} r\right) \, \psi ^{\mu _j}_{j {\mp } 1/2} (\theta ) \end{pmatrix} \end{aligned}$$

with \(j = 1/2, 3/2, \ldots \), \(\ell = j \pm 1/2\), \(\ell ' = j {\mp } 1/2\), \(\mu _j = -j, -j + 1,\ldots , j\), and \(k\in \mathbb {Z}\), belongs to \(L^\pm _{j, \mu _j}\) and is an eigenfunction of \({\mathcal {H}}_{\tau }\) with eigenvalue \(\uplambda _{j, k}^\pm (\tau )\).

Note that the superindex in \(\uplambda _{j, k}^\pm \) indicates to which invariant subspace belongs the associated eigenfunction. It should not be confused with the sign of the eigenvalue (as the superindex in \(\uplambda ^\pm _1\) denotes).

Proof of Proposition B.2

From the arguments already presented in Appendix B.1 it only remains to show that from (B.3) we can obtain the aforementioned infinite number of parametrizations of \(\uplambda \) in terms of \(\tau \in \mathbb {R}\). To do it, the idea is to rewrite the eigenvalue equation (B.3) as

$$\begin{aligned} e^\tau = {\mp } \dfrac{\sqrt{\uplambda ^2-m^2}}{\uplambda +m} \dfrac{ J_{\ell ' + 1/2} ( \sqrt{\uplambda ^2-m^2} R ) }{J_{\ell + 1/2} ( \sqrt{\uplambda ^2-m^2} R ) } =: h(\uplambda ). \end{aligned}$$

Then, our goal will be to invert h in suitable intervals to get \(\uplambda =\uplambda (\tau ):=h^{-1}(e^\tau )\).

First, note that we can restrict ourselves to the subspaces \(L^-_{j, \mu _j}\), thanks to the odd symmetry mentioned in Remark 1.1. In this case the eigenvalue equation is written as

$$\begin{aligned} e^\tau = {{\,\textrm{sign}\,}}(\uplambda +m ) \sqrt{\dfrac{\uplambda -m}{\uplambda +m}} \dfrac{ J_{j + 1} ( \sqrt{\uplambda ^2-m^2} R ) }{J_{j} ( \sqrt{\uplambda ^2-m^2} R ) } = h(\uplambda ). \end{aligned}$$

Due to the fact that \(e^\tau >0\) for all \(\tau \in \mathbb {R}\), we are forced to work with h only on intervals \(I\subset \mathbb {R}\) such that \(h(I)\subset (0,+\infty )\) and where h is invertible. Since \(\frac{\uplambda -m}{\uplambda +m}\) is positive and strictly increasing for \(\uplambda \in (m, +\infty )\) and for \(\uplambda \in (- \infty , -m)\), I must be such that

$$\begin{aligned} {{\,\textrm{sign}\,}}(\uplambda +m ) \dfrac{ J_{j + 1} ( \sqrt{\uplambda ^2-m^2} R ) }{J_{j} ( \sqrt{\uplambda ^2-m^2} R ) } \text { is positive and strictly increasing for } \uplambda \in I. \end{aligned}$$

Then, Lemma B.1 yields that all the possible intervals I are:

  • \(I= (m, \sqrt{(z_{j,1}/R)^2 + m^2})\),

  • \(I = (\sqrt{(z_{j + 1,k}/R)^2 + m^2}, \sqrt{(z_{j,k + 1}/R)^2 + m^2})\) for \(k\ge 1\),

  • \(I = (-\sqrt{(z_{j + 1,k}/R)^2 + m^2}, - \sqrt{(z_{j,k}/R)^2 + m^2})\) for \(k\ge 1\).

In each of these intervals I the function \(h: I \rightarrow (0,+\infty )\) is of class \(C^\infty \), strictly increasing, and surjective. Therefore, it can be inverted, obtaining a \(C^\infty \) function \(\tau \mapsto \uplambda (\tau ) := h^{-1}(e^\tau )\) which maps \(\mathbb {R}\) into I surjectively and corresponds, for each \(\tau \in \mathbb {R}\), to an eigenvalue of \({\mathcal {H}}_\tau \). In addition, the monotonicity of \(\uplambda \mapsto \tau =\tau (\uplambda ):=\log (h(\uplambda ))\) yields that \(\tau \mapsto \uplambda (\tau )\) is also strictly increasing. \(\square \)

The previous result yields that, for any given eigenvalue curve \(\tau \mapsto \uplambda (\tau )\), it holds that \(\lim _{\tau \rightarrow \pm \infty }|\uplambda (\tau )|\) is either m or a positive zero of the function \(J_{k + 1/2} ( \sqrt{(\cdot )^2-m^2} R )\) for some \(k=0,1,2,\ldots \); note that each of these zeroes corresponds to the square root of a Dirichlet eigenvalue of \(-\Delta + m^2\) in \(B_R\). The monotonicity and limiting values of \(\tau \mapsto \uplambda (\tau )\) was observed in the curves plotted in Fig. 1. Furthermore, the alternation (with respect to the zeroes of the Bessel function) between positive and negative eigenvalue curves, which is given by the intervals \(I_{j,k}^\pm \) defined in Proposition B.2, was already shown numerically in Figs. 2 and 3.

1.4 B.4. The first positive eigenvalue

In this section we focus on the first positive eigenvalue of \({\mathcal {H}}_{\tau }\) when \(\Omega = B_R\). We provide a fine description of the associated eigenvalue curve, whose main properties are summarized in the following proposition. The reader may compare it with Theorem 1.5, which is the analogous result for general domains; see also Theorem 1.7 regarding (B.5).

Proposition B.3

The function \(\tau \mapsto \uplambda _1^+(\tau )=\min (\sigma ({\mathcal {H}}_\tau )\cap (m,+\infty ))\) is of class \(C^\infty \) in \(\mathbb {R}\), and satisfies

$$\begin{aligned} \lim _{\tau \downarrow - \infty } \uplambda _1^+ (\tau ) = m \quad \text {and} \quad \lim _{\tau \uparrow + \infty } \uplambda _1^+ (\tau ) = \sqrt{\pi ^2/R^2 + m^2} = \sqrt{ \min \sigma (-\Delta _D) + m^2}, \end{aligned}$$

where \(\min \sigma (-\Delta _D)\) denotes the first Dirichlet eigenvalue of \(-\Delta \) in \(B_R\). In addition, the corresponding eigenspace associated to \(\uplambda _1^+(\tau )\) has always dimension 2.

Furthermore,

$$\begin{aligned} L^\star _{B_R} := \lim _{\tau \downarrow -\infty } (\uplambda _1^+(\tau )-m)e^{-\tau } = \frac{3}{R}=\frac{1}{{\mathscr {R}}_{B_R}}, \end{aligned}$$
(B.5)

where \({\mathscr {R}}_{B_R}\) is defined in (1.14).

Proof

Let \(\uplambda : \mathbb {R}\rightarrow \big (m, \sqrt{\pi ^2/R^2 + m^2}\big )\) be the eigenvalue curve corresponding to \(\uplambda _{1/2, 0}^-\) in the notation of Proposition B.2. This eigenfunction is associated to the two subspaces \(L^-_{1/2, \mu _{1/2}}\), with either \(\mu _{1/2} = 1/2\) or \(\mu _{1/2} = -1/2\), and solves the implicit equation

$$\begin{aligned} e^\tau = \dfrac{\sqrt{\uplambda ^2-m^2}}{\uplambda +m} \dfrac{ J_{3/2} ( \sqrt{\uplambda ^2-m^2} R ) }{J_{1/2} ( \sqrt{\uplambda ^2-m^2} R ) }. \end{aligned}$$
(B.6)

We will show that \(\uplambda = \uplambda _1^+\). The upper bound for \(\uplambda (\tau )\) (and limit as \(\tau \rightarrow +\infty \)) is given by the fact that \(z_{1/2, 1} = \pi \), which follows from the explicit expression of the Bessel functions involved in the above equation:

$$\begin{aligned} J_{1/2} (x) = \sqrt{\dfrac{2}{\pi x}} \sin (x) \quad \text { and } \quad J_{3/2} (x) = \sqrt{\dfrac{2}{\pi x}} \left( \dfrac{\sin (x)}{x} - \cos (x) \right) . \end{aligned}$$

As a matter of fact, these expressions can be used to show—after a tedious computation and using that \(x>\sin (x)\) for all \(x\in (0,\pi )\)—that the right-hand side of (B.6) is a strictly increasing function of \(\uplambda \) for \(\uplambda \in \big (m, \sqrt{\pi ^2/R^2 + m^2}\big )\) without making use of Lemma B.1 (as done in the proof of Proposition B.2).

Since \(\pi = z_{1/2, 1}\) is the smallest positive zero among all the positive zeroes of the Bessel functions of half-integer index—as shown by Lemma B.1 (ii)—, it follows that \(\uplambda (\tau ) \) coincides with \( \uplambda ^+_1(\tau )\) at least for big enough values of \(\tau \). To show that indeed \(\uplambda (\tau )\) is the first positive eigenvalue \(\uplambda _1^+ (\tau )\) for all \(\tau \in \mathbb {R}\), it suffices to show that \(\uplambda \) cannot cross any other eigenvalue curve. On the one hand, taking into account the possible intervals \(I_{j,k}^+\) given in Proposition B.2, it follows that any positive eigenvalue curve associated to the spaces \(L^+_{j, \mu _{j}}\) must lie above \(\sqrt{\pi ^2/R^2 + m^2}\), and thus it cannot cross \(\uplambda (\tau )\). On the other hand, if there was a crossing between \(\uplambda (\tau )\) and another positive eigenvalue curve associated to \(L^-_{j_\circ , \mu _{j_\circ }}\) for some half-integer \(j_\circ > 1/2\), then by (B.6) there would exist a point \(x_\circ \in (0, z_{1/2, 1})\) such that

$$\begin{aligned} \dfrac{ J_{3/2} (x_\circ ) }{J_{1/2} (x_\circ ) } = \dfrac{ J_{j_\circ + 1} (x_\circ ) }{J_{j_\circ } (x_\circ ) }. \end{aligned}$$
(B.7)

However, since the zeroes of the Bessel functions are ordered (see Lemma B.1), for every half-integer \(j\ge 1/2\) it follows that \(z_{j, k} < z_{j + 1,k}\) for all \(k \ge 1\), and therefore

$$\begin{aligned} \dfrac{2x}{z_{j, k}^2 - x^2} > \dfrac{2x}{z_{j + 1, k}^2 - x^2} \quad \text { for every } x \in (0, z_{1/2, 1}). \end{aligned}$$

Hence, by (B.4) it follows that

$$\begin{aligned} \dfrac{ J_{j + 1} (x) }{J_{j} (x ) } > \dfrac{ J_{j+2} (x) }{J_{j+1} (x ) } \quad \text { for every } x \in (0, z_{1/2, 1}) \text { and for every half-integer } j \ge 1/2, \end{aligned}$$

thus there cannot exist such a \(x_\circ \in (0, z_{1/2, 1})\) satisfying (B.7). In conclusion, \(\uplambda \) does not cross any other eigenvalue curve. Therefore, \(\uplambda (\tau )\) is the first positive eigenvalue of \({\mathcal {H}}_\tau \) for all \(\tau \in \mathbb {R}\). As a byproduct, since \(\uplambda (\tau )\) is associated to \(L^-_{1/2, \mu _{1/2}}\) with either \(\mu _{1/2} = 1/2\) or \(\mu _{1/2} = -1/2\), it follows that the first positive eigenvalue of \({\mathcal {H}}_\tau \) has multiplicity 2 for all \(\tau \in \mathbb {R}\).

To conclude the proof, we are only left to show that \(L_{B_R}^* = 3/R=1/{\mathscr {R}}_{B_R}\). First, note that from the rescaling properties of the operators \(K_m\) and \(W_m\) defined in (2.12), it follows readily that \({\mathscr {R}}_{B_R} = R{\mathscr {R}}_{B_1}\). Moreover, \(L_{B_R}^\star = L_{B_1}^\star /R\). To show this second equality, one notices that if \(u_1\) and \(u_R\) denote the boundary values of the upper component of the first eigenfunction in \(B_1\) and \(B_R\) respectively, both associated to the same subspace \(L^-_{1/2, \mu _{1/2}}\), then after a normalization one can choose them in such a way that \(u_R(\cdot ) = u_1(\cdot /R)\). Hence, from (2.19) and taking the limit \(\tau \downarrow -\infty \), using again the scaling of \(K_m\), and that \(\{W_m, \sigma \cdot \nu \} = 0\) by Lemma 4.7, it follows that \(L_{B_R}^\star = L_{B_1}^\star /R\); see the argument to get to (B.9) below for more details. As a consequence of all this, it is enough to prove the result for \(R=1\).

Recall that \(\uplambda (\tau )\) is associated, for all \(\tau \in \mathbb {R}\), to the two subspaces \(L^-_{1/2, \mu _{1/2}}\) with either \(\mu _{1/2}= 1/2\) or \(\mu _{1/2}= -1/2\). We will work in one of these two subspaces (the precise choice will be completely irrelevant for the rest of the argument), and therefore, after a normalization, we can take an eigenfunction \(\varphi _\tau = (u_\tau , v_\tau )^\intercal \) associated to \(\uplambda (\tau )\) such that \(u_\tau \) at the boundary of \(B_1\) is given by \( \psi ^{\mu _{1/2}}_0\) for all \(\tau \in \mathbb {R}\). By Lemma 2.9, it holds

$$\begin{aligned} \begin{aligned} \Big ( \frac{1}{2} -i W_\uplambda ({\sigma }\cdot \nu ) \Big )\psi ^{\mu _{1/2}}_0&= - (\uplambda +m) e^\tau K_\uplambda \psi ^{\mu _{1/2}}_0\quad \text {and}\\ \Big ( \frac{1}{2} -i ({\sigma }\cdot \nu ) W_\uplambda \Big )\psi ^{\mu _{1/2}}_0&= (\uplambda -m)e^{-\tau } ({\sigma }\cdot \nu ) \, K_\uplambda ({\sigma }\cdot \nu ) \psi ^{\mu _{1/2}}_0 \end{aligned} \end{aligned}$$
(B.8)

in \(L^2(\mathbb {S}^2)^2\). We will take the limit \(\tau \downarrow -\infty \) in (B.8), taking into account that \(\uplambda \downarrow m\) as \(\tau \downarrow -\infty \). Letting \(\tau \downarrow -\infty \) in (B.8), and using that then \(K_\uplambda \rightarrow K_m\) and \(W_\uplambda \rightarrow W_m\) as bounded operators in \(L^2(\mathbb {S}^2)^2\) (which follows from the explicit expressions of the operators), we obtain

$$\begin{aligned} \begin{aligned} P_- \, \psi ^{\mu _{1/2}}_0 = 0 \quad \text {and} \quad (P_+)^*\, \psi ^{\mu _{1/2}}_0 = L_{B_1}^\star ({\sigma }\cdot \nu ) \, K_m ({\sigma }\cdot \nu ) \psi ^{\mu _{1/2}}_0 \quad \text { in } L^2(\mathbb {S}^2)^2, \end{aligned}\nonumber \\ \end{aligned}$$
(B.9)

where \(L^\star _{B_1} =\lim _{\tau \downarrow -\infty } (\uplambda (\tau )-m)e^{-\tau }\), and \(P_- = \frac{1}{2} -i W_m ({\sigma }\cdot \nu )\) and \((P_+)^*= \frac{1}{2} -i ({\sigma }\cdot \nu ) W_m \), as defined in Sect. 4.2. In particular, (B.9) shows that \(L^\star _{B_1}\) is finite, since all the involved operators are bounded and \(K_m\) is injective. It is worth pointing out that this last argument leading to \(L_{B_1}^\star < +\infty \) works thanks to the fact that, on \(\partial \Omega \), the eigenfunction \(u_\tau = \psi ^{\mu _{1/2}}_0\) is indeed independent of \(\tau \), something that may not be guaranteed on a general domain \(\Omega \).

We will use (B.9) to establish (B.5) for \(R=1\). First, let \(L\in {\mathcal {L}}_{B_1}\)—recall that \({\mathcal {L}}_{B_1}\) is defined in (1.11), see also (4.20). We claim that 1/L is an eigenvalue of \(K_m\). This claim can be proven by adding the two equations from which \({\mathcal {L}}_{B_1}\) is defined and using that \(\{W_m, \sigma \cdot \nu \} = 0\) as operators in \(L^2(\mathbb {S}^2)^2\); see Lemma 4.7. As a consequence, and since \(1/{\mathscr {R}}_\Omega = \min {\mathcal {L}}_{\Omega }\) for every \(\Omega \subset \mathbb {R}^3\) by Theorem 1.7, \( {\mathscr {R}}_{B_1}\) is the maximum of the eigenvalues of \(K_m\) among eigenfunctions of the form \((\sigma \cdot \nu )u\) with \(u\in P_+(L^2(\mathbb {S}^2)^2)\), where \(P_+=\frac{1}{2}+ iW_m(\sigma \cdot \nu )\).

Let us now compute explicitly the eigenvalues of \(K_m\) as an operator in \(L^2(\mathbb {S}^2)^2\). Using [10, Lemma 4.3] and [58, Theorem 3.6] on \(K_a\) for \(a>0\) and lettingFootnote 2\(a \uparrow m\), it follows that the spectrum of \(K_m\) is given by a sequence \(\{ d_{j\pm 1/2} \}_{j = 1/2, \ 3/2, \ldots }\) whose associated eigenfunctions are \(\psi _{j \pm 1 / 2}^{\mu _j}\), i.e.,

$$\begin{aligned} K_m \psi _{j \pm 1 / 2}^{\mu _j} = d_{j\pm 1/2}\, \psi _{j \pm 1 / 2}^{\mu _j}, \end{aligned}$$
(B.10)

and the eigenvalues \( d_{j\pm 1/2}\) are given by the expression

$$\begin{aligned} d_{j\pm 1/2} = \lim _{t\downarrow 0} {\mathcal {I}}_{j \pm 1/2 +1/2}(t) {\mathcal {K}}_{j \pm 1/2 +1/2}(t). \end{aligned}$$

Here \({\mathcal {I}}_\kappa \) and \({\mathcal {K}}_\kappa \) are the modified Bessel functions of the first and second kind of order \(\kappa \), respectively. By [1, formulas 9.6.7 and 9.6.9] we have

$$\begin{aligned} \lim _{t\downarrow 0} {\mathcal {I}}_{\kappa }(t) {\mathcal {K}}_{\kappa }(t) = \dfrac{\Gamma (\kappa )}{2 \Gamma (\kappa + 1)} = \dfrac{1}{2 \kappa } \quad \text {for } \kappa > 0. \end{aligned}$$

Hence, for \(k= 0, 1, \ldots ,\) we have

$$\begin{aligned} d_{k} = \dfrac{1}{2k +1}. \end{aligned}$$
(B.11)

We claim that \((\sigma \cdot \nu ) \psi _0^{\mu _j} = \psi _1^{\mu _j} \) is orthogonal to \(P_+(L^2(\mathbb {S}^2)^2)\). Once this is proved, it follows that \(1/d_0 \notin {\mathcal {L}}_{B_1}\) and, as a consequence, \({\mathscr {R}}_{B_1} \le d_1 = 1/3\). To prove the claim, notice that for all \(\tau \in \mathbb {R}\), after a normalization, \(\psi _1^{\mu _j}\) is the upper component of the eigenfunction of \({\mathcal {H}}_{\tau }\) associated to \(-\uplambda (-\tau )\), i.e., the first (larger) negative eigenvalue of \({\mathcal {H}}_{\tau }\). Indeed, this can be shown with exactly the same arguments as we did at the beginning of the proof of the proposition for the first positive eigenvalue, considering in this case the subspace \(L^+_{1/2, \mu _{1/2}}\); see also Remark 1.1. Therefore, since such eigenvalue converges to \(-m\) as \(\tau \uparrow +\infty \), using the second equation in (2.18) with \(u= \psi _1^{\mu _j}\) and taking the limit \(\tau \uparrow +\infty \)—analogously as how we proceeded before to show (B.9)—, it follows that \((P_+)^* \psi _1^{\mu _j} = 0\), establishing our claim and, as a byproduct, the inequality \({\mathscr {R}}_{B_1} \le d_1 = 1/3\).

We shall now prove that \(L_{B_1}^\star = 1/ d_1 = 3\), which will finish the proof of (B.5) since \(1/{\mathscr {R}}_{B_1} \le L_{B_1}^*\) by Lemmas 4.10 and 4.11. By adding the two equations in (B.9) and using again that \(\{W_m,\sigma \cdot \nu \} = 0\) (or equivalently, that \((P_+)^* = P_+\) by Lemma 4.7 since the underlying domain is a ball), we obtain

$$\begin{aligned} \psi ^{\mu _{1/2}}_0 = L_{B_1}^\star ({\sigma }\cdot \nu ) \, K_m ({\sigma }\cdot \nu ) \psi ^{\mu _{1/2}}_0 \quad \text {in } L^2(\mathbb {S}^2)^2. \end{aligned}$$

Finally, using that \(({\sigma }\cdot \nu ) \psi ^{\mu _{1/2}}_{j\pm 1/2} = \psi ^{\mu _{1/2}}_{j {\mp } 1/2}\) for all half-integers j (see [71, equation (4.121)]), and that \(\psi ^{\mu _{1/2}}_1\) is an eigenfunction of \(K_m\) with eigenvalue \(d_1\) by (B.10), we get

$$\begin{aligned} \psi ^{\mu _{1/2}}_0 = L_{B_1}^\star d_1 \psi ^{\mu _{1/2}}_0 \quad \text { in } L^2(\mathbb {S}^2)^2. \end{aligned}$$

Therefore, we have \(L_{B_1}^\star = 1/d_1 = 3\) by (B.11). This concludes the proof. \(\square \)

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Arrizabalaga, N., Mas, A., Sanz-Perela, T. et al. Eigenvalue Curves for Generalized MIT Bag Models. Commun. Math. Phys. 397, 337–392 (2023). https://doi.org/10.1007/s00220-022-04526-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-022-04526-3

Navigation