Skip to main content
Log in

Homogenization of the Eigenvalues of the Neumann–Poincaré Operator

  • Published:
Archive for Rational Mechanics and Analysis Aims and scope Submit manuscript

Abstract

In this article, we investigate the spectrum of the Neumann–Poincaré operator \({{\mathcal {K}}}_\varepsilon ^*\) (or equivalently, that of the associated Poincaré variational operator \(T_\varepsilon \)) associated to a periodic distribution of small inclusions with size \(\varepsilon \), and its asymptotic behavior as the parameter \(\varepsilon \) vanishes. Combining techniques pertaining to the fields of homogenization and potential theory, we prove that the limit spectrum is composed of the ‘trivial’ eigenvalues 0 and 1, and of a subset which stays bounded away from 0 and 1 uniformly with respect to \(\varepsilon \). This non trivial part is the reunion of the Bloch spectrum, accounting for the collective resonances between collections of inclusions, and of the boundary layer spectrum, associated to eigenfunctions which spend a not too small part of their energies near the boundary of the macroscopic device. These results shed new light on the issue of the homogenization of the voltage potential \(u_\varepsilon \) caused by a given source in a medium composed of a periodic distribution of small inclusions with an arbitrary (possible negative) conductivity a, surrounded by a dielectric medium, with unit conductivity. In particular, we prove that the limit behavior of \(u_\varepsilon \) is strongly related to the (possibly ill-defined) homogenized diffusion matrix predicted by the homogenization theory in the standard elliptic case. Additionally, we prove that the homogenization of \(u_\varepsilon \) is always possible when a is either positive, or negative with a ‘small’ or ‘large’ modulus.

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.

Fig. 1

Similar content being viewed by others

References

  1. Adams, R.A., Fournier, J.F.: Sobolev Spaces, 2nd edn. Academic Press, London 2003

    MATH  Google Scholar 

  2. Aguirre, F., Conca, C.: Eigenfrequencies of a tube bundle immersed in a fluid. Appl. Math. Optim. 18, 1–38, 1988

    Article  MathSciNet  MATH  Google Scholar 

  3. Allaire, G.: Homogenization and two-scale convergence. SIAM J. Math. Anal. 23(6), 1482–1518, 1992

    Article  MathSciNet  MATH  Google Scholar 

  4. Allaire, G.: Shape Optimization by the Homogenization Method. Springer, New York 2001

    MATH  Google Scholar 

  5. Allaire, G., Briane, M., Vanninathan, M.: A comparison between two-scale asymptotic expansions and Bloch wave expansions for the homogenization of periodic structures. SEMA J. 73(3), 237–259, 2016

    Article  MathSciNet  MATH  Google Scholar 

  6. Allaire, G., Conca, C.: Bloch wave homogenization for a spectral problem in fluid–solid structures. Arch. Ration. Mech. Anal. 135, 197–257, 1996

    Article  MathSciNet  MATH  Google Scholar 

  7. Allaire, G., Conca, C.: Bloch wave homogenization and spectral asymptotic analysis. J. Math. Pures Appl. 77, 153–208, 1998

    Article  MathSciNet  MATH  Google Scholar 

  8. Ammari, H., Ciraolo, G., Kang, H., Lee, H., Milton, G.W.: Spectral theory of a Neumann–Poincaré-type operator and analysis of cloaking due to anomalous localized resonance. Arch. Ration. Mech. Anal. 208, 667–692, 2013

    Article  MathSciNet  MATH  Google Scholar 

  9. Ammari, H., Ciraolo, G., Kang, H., Lee, H., Milton, G.W.: Spectral analysis of a Neumann–Poincaré-type operator and analysis of cloaking due to anomalous localized resonance II. Contemp. Math. 615, 1–14, 2014

    Article  MATH  Google Scholar 

  10. Ammari, H., Deng, Y., Millien, P.: Surface plasmon resonance of nanoparticles and applications in imaging. Arch. Ration. Mech. Anal. 220, 109–153, 2016

    Article  MathSciNet  MATH  Google Scholar 

  11. Ammari, H., Kang, H.: Polarization and Moment Tensors; With Applications to Inverse Problems and Effective Medium Theory. Springer Applied Mathematical Sciences, vol. 162, 2007

  12. Ammari, H., Kang, H., Lee, H.: Layer Potential Techniques in Spectral Analysis, Mathematical Surveys and Monographs, vol. 153. American Mathematical Society, Providence RI 2009

    Book  Google Scholar 

  13. Ammari, H., Kang, H., Lim, M.: Gradient estimates for solutions to the conductivity problem. Math. Ann. 332, 277–286, 2005

    Article  MathSciNet  MATH  Google Scholar 

  14. Ammari, H., Millien, P., Ruiz, M., Zhang, H.: Mathematical analysis of plasmonic nanoparticles: the scalar case, 2016. arXiv:1506.00866

  15. Ammari, H., Ruiz, M., Yu, S., Zhang, H.: Mathematical analysis of plasmonic resonances for nanoparticles: the full Maxwell equations, 2016. arXiv:1511.06817

  16. Ammari, H., Kang, H., Touibi, K.: Boundary layer techniques for deriving the effective properties of composite materials. Asymptot. Anal. 41(2), 119–140, 2005

    MathSciNet  MATH  Google Scholar 

  17. Ammari, H., Seo, J.K.: An accurate formula for the reconstruction of conductivity inhomogeneities. Adv. Appl. Math. 30(4), 679–705, 2003

    Article  MathSciNet  MATH  Google Scholar 

  18. Ando, K., Kang, H.: Analysis of plasmon resonance on smooth domains using spectral properties of the Neumann–Poincaré operator. J. Math. Anal. Appl. 435, 162–178, 2016

    Article  MathSciNet  MATH  Google Scholar 

  19. Arbogast Jr., T., Douglas, J., Hornung, U.: Derivation of the double porosity model of single phase flow via homogenization theory. SIAM J. Math. Anal. 21(4), 823–836, 1990

    Article  MathSciNet  MATH  Google Scholar 

  20. Bao, E.S., Li, Y., Yin, B.: Gradient estimates for the perfect conductivity problem. Arch. Ration. Mech. Anal. 193, 195–226, 2009

    Article  MathSciNet  MATH  Google Scholar 

  21. Bensoussan, A., Lions, J.-L., Papanicolau, G.: Asymptotic Analysis of Periodic Structures. North Holland, Amsterdam 1978

    Google Scholar 

  22. Bonnet-Ben Dhia, A.-S., Ciarlet Jr., P., Zwölf, C.-M.: Time harmonic wave diffraction problems in materials with sign-shifting coefficients. J. Comput. Appl. Math. 234, 1912–1919, 2007. Corrigendum 2616 (2010)

  23. Bonnetier, E., Nguyen, H.-M.: Superlensing using hyperbolic metamaterials: the scalar case. J de l’École polytechnique - Mathématiques 4(2017), 973–1003, 2017

    Article  MathSciNet  MATH  Google Scholar 

  24. Bonnetier, E., Triki, F.: Pointwise bounds on the gradient and the spectrum of the Neumann–Poincaré operator: the case of 2 discs. Contemp. Math. 577, 81–92, 2012

    Article  MATH  Google Scholar 

  25. Bonnetier, E., Triki, F.: On the spectrum of the Poincaré variational problem for two close-to-touching inclusions in 2d. Arch. Ration. Mech. Anal. 209, 541–567, 2013

    Article  MathSciNet  MATH  Google Scholar 

  26. Bonnetier, E., Triki, F., Tsou, C.H.: Eigenvalues of the Neumann-Poincaré operator for two inclusions with contact of order m: a numerical study. J. Comput. Math. 36, 17–28, 2018

    Article  MathSciNet  MATH  Google Scholar 

  27. Bouchitté, G., Schweizer, B.: Cloaking of small objects by anomalous localized resonance. Q. J. Mech. Appl. Math. 63, 437–463, 2010

    Article  MathSciNet  MATH  Google Scholar 

  28. Braides, A., Briane, M., Casado-Diaz, J.: Homogenization of non-uniformly bounded periodic diffusion energies in dimension two. Nonlinearity 22, 1459–1480, 2009

    Article  ADS  MathSciNet  MATH  Google Scholar 

  29. Briane, M., Casado-Diaz, J.: Uniform convergence of sequences of solutions of two-dimensional linear elliptic equations with unbounded coefficients. J. Differ. Equ. 245, 2038–2054, 2008

    Article  ADS  MathSciNet  MATH  Google Scholar 

  30. Brezis, H.: Functional Analysis, Sobolev Spaces and Partial Differential Equations. Springer, Berlin 2000

    Google Scholar 

  31. Bunoiu, R., Ramdani, K.: Homogenization of materials with sign changing coefficients, 2015 (submitted)

  32. Castro, C., Zuazua, E.: Une remarque sur l’analyse asymptotique spectrale en homogénéisation. C. R. Acad. Sci. Paris Ser. I 335, 99–104, 2002

    Article  Google Scholar 

  33. Cioranescu, D., Damlamian, A., Griso, G.: Periodic unfolding and homogenization. C. R. Acad. Sci. Paris Ser. I 322, 1043–1047, 1996

    MATH  Google Scholar 

  34. Cioranescu, D., Damlamian, A., Griso, G.: The periodic unfolding method in homogenization. SIAM J. Math. Anal. 40(4), 1585–1620, 2008

    Article  MathSciNet  MATH  Google Scholar 

  35. Cioranescu, D., Damlamian, A., Li, T.: Periodic homogenization for inner boundary conditions with equi-valued surfaces: the unfolding approach. In: Partial Differential Equations: Theory, Control and Approximation. pp 183–209. Springer, Berlin, Heidelberg 2014

  36. Conca, C., Planchard, J., Vanninathan, M.: Fluids and Periodic Structures, RMA, 38. Wiley & Masson, London 1995

    MATH  Google Scholar 

  37. Conca, C., Vanninathan, M.: Homogenization of periodic structures via Bloch decomposition. SIAM J. Appl. Math. 57, 1639–1659, 1997

    Article  MathSciNet  MATH  Google Scholar 

  38. Costabel, M., Stephan, E.: A direct boundary integral equation method for transmission problems. J. Math. Anal. Appl. 106, 367–413, 1985

    Article  MathSciNet  MATH  Google Scholar 

  39. El-Sayed, I.H., Huang, X., El-Sayed, M.A.: Surface plasmon resonance scattering and absorption of anti-EGFR antibody conjugated gold nanoparticles in cancer diagnostics: applications in oral cancer. Nano Lett. 5(5), 829–834, 2005

    Article  ADS  Google Scholar 

  40. Folland, G.B.: Introduction to Partial Differential Equations, 2nd edn. Princeton University Press, Princeton 1995

    MATH  Google Scholar 

  41. Gérard, P.: Mesures semi-classiques et ondes de Bloch. Séminaire Équations aux Dérivées Partielles 1990–1991, volume 16, Ecole Polytechnique, Palaiseau, 1991

  42. Grieser, D.: The plasmonic eigenvalue problem. Rev. Math. Phys. 26, 1450005, 2014

    Article  MathSciNet  MATH  Google Scholar 

  43. Griso, G.: Analyse asymptotique de structures réticulées. Thèse de l’Université Pierre et Marie Curie (Paris VI), 1996

  44. Jikov, V.V., Kozlov, S.M., Oleinik, O.A.: Homogenization of Differential Operators and Integral Functionals. Springer, Berlin 1994

    Book  Google Scholar 

  45. John, F.: The Dirichlet problem for a hyperbolic equation. Am. J. Math. 63(1), 141–154, 1941

    Article  MathSciNet  MATH  Google Scholar 

  46. Kang, H.: Layer potential approaches to interface problems. Inverse Problems and Imaging: Panoramas et synthèses, 44. Société Mathématique de France, 2013

  47. Khavinson, D., Putinar, M., Shapiro, H.S.: On Poincaré’s variational problem in potential theory. Arch. Ration. Mech. Anal. 185, 143–184, 2007

    Article  MathSciNet  MATH  Google Scholar 

  48. Kohn, R.V., Milton, G.W.: On bounding the effective conductivity of anisotropic composites. Homogenization and Effective Moduli of Materials and Media, Vol. 1 IMA Volumes in Mathematics and Its Applications (Eds. Ericksen, J.L., Kinderlehrer, D., Kohn, R., Lions, J.-L.) Springer, Berlin, 97–125, 1986

  49. Kuchment, P.: Floquet Theory for Partial Differential Equations. Birkhäuser, Basel 1993

    Book  MATH  Google Scholar 

  50. Lipton, R., Viator, R.: Bloch waves in crystals and periodic high contrast media, 2016 (submitted)

  51. Maier, S.A.: Plasmonics: Fundamentals and Applications. Springer, Berlin 2007

    Book  Google Scholar 

  52. Manley, P., Burger, S., Schmidt, F., Schmid, M.: Design principles for plasmonic nanoparticle devices. Progress in Nonlinear Nano-Optics Part of the Series Nano-Optics and Nanophotonics, 223–247, 2015

  53. Mc Lean, W.: Strongly Elliptic Systems and Boundary Integral Equations. Cambridge University Press, Cambridge 2000

    Google Scholar 

  54. Moskow, S., Vogelius, M.S.: First-order corrections to the homogenised eigenvalues of a periodic composite medium. A convergence proof. Proc. R. Soc. Edinb. 127, 1263–1299, 1997

    Article  MathSciNet  MATH  Google Scholar 

  55. Nicorovici, N.A., McPhedran, R.C., Milton, G.M.: Optical and dielectric properties of partially resonant composites. Phys. Rev. B 49, 8479–8482, 1994

    Article  ADS  Google Scholar 

  56. Nguyen, H.-M.: Cloaking using complementary media in the quasistatic regime. Ann. I. H. Poincaré (C) Non Linear Anal. 32, 471–484, 2015

    Article  ADS  Google Scholar 

  57. Nguyen, H.-M.: Negative index materials and their applications: recent mathematics progress. Chin. Ann. Math. 2016. https://doi.org/10.1007/s11401-017-1086-5

    Google Scholar 

  58. Nguetseng, G.: A general convergence result for a functional related to the theory of homogenization. SIAM J. Math. Anal 20(3), 608–623, 1989

    Article  MathSciNet  MATH  Google Scholar 

  59. Planchard, J.: Global behaviour of large elastic tube bundles immersed in a fluid. Comput. Mech. 2, 105–118, 1987

    Article  MATH  Google Scholar 

  60. Otomori, M., Yamada, T., Izui, K., Nishiwaki, S., Andkjær, J.: Topology optimization of hyperbolic metamaterials for an optical hyperlens. Struct. Multidiscip. Optim. 2016. https://doi.org/10.1007/s00158-016-1543-x

  61. Patching, S.G.: Surface plasmon resonance spectroscopy for characterisation of membrane protein–ligand interactions and its potential for drug discovery. Biochim. Biophys. Acta (BBA) - Biomembr. 1838(1), 43–55, 2014

    Article  Google Scholar 

  62. Poddubny, A., Iorsh, I., Belov, P., Kivshar, Y.: Hyperbolic metamaterials. Nat. Photon. 7, 948–957, 2013

    Article  ADS  Google Scholar 

  63. Reed, M., Simon, B.: Methods of Modern Mathematical Physics, IV, Analysis of operators. Academic Press, New York 1978

    MATH  Google Scholar 

  64. Rudin, W.: Functional Analysis, 2nd edn. International Series in Pure and Applied MathematicsMcGraw-Hill, New York, NY 1991

    MATH  Google Scholar 

  65. Sauter, S.A., Schwab, C.: Boundary Element Methods. Springer Series in Computational Mathematics, Springer, GmbH & Co. K, Berlin, Heidelberg 2010

  66. Shekhar, P., Atkinson, J., Jacob, Z.: Hyperbolic metamaterials: fundamentals and applications. Nano Converg. 1, 1–14, 2014

    Article  Google Scholar 

  67. Triki, F., Vauthrin, M.: Mathematical modeling of the Photoacoustic effect generated by the heating of metallic nanoparticles. Q. Appl. Math. 76, 673–698, 2018

    Article  MathSciNet  MATH  Google Scholar 

  68. Wilcox, C.: Theory of Bloch waves. J. Anal. Math. 33, 146–167, 1978

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The authors were partially supported by the AGIR-HOMONIM grant from Université Grenoble-Alpes, and by the Labex PERSYVAL-Lab (ANR-11-LABX-0025-01).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Charles Dapogny.

Additional information

Communicated by F. Otto

Publisher's Note

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

A Closer Study of the Particular Case of Rank 1 Laminates

A Closer Study of the Particular Case of Rank 1 Laminates

In this appendix, we focus on an interesting particular geometry of microstructures \(\omega \subset Y\), that of rank 1 laminates, which is one of the few amenable to explicit calculations. Note that this situation violates some of the prevailing assumptions of this article, notably the fact that \(\omega \Subset Y\); it is therefore not surprising that some of the general results established in the previous sections do not hold in the present case.

The ‘macroscopic’ domain \({\varOmega }\) at stake is the two-dimensional square \((0,1)^2\); if is filled with \(N^2\) identical cells, homothetic to the unit periodicity cell \(Y = (0,1)^2\). The rescaled inclusion pattern \(\omega \) in each cell is

$$\begin{aligned}&\omega := \left\{ y=(y_1,y_2) \in Y, \, 0<y_1<\theta \right\} , \\&\quad \text { and so } Y{\setminus }\overline{\omega } = \left\{ y=(y_1,y_2) \in Y, \, \theta<y_1<1 \right\} , \end{aligned}$$

where \(\theta \in (0,1)\) is the volume of \(\omega \). Accordingly, the total set of inclusions \(\omega _N \subset {\varOmega }\) read as

$$\begin{aligned} \omega _N := \bigcup _{\begin{array}{c} j \in {\mathbb {N}}^2 \\ 0 \leqq j \leqq N-1 \end{array}}{\frac{1}{N}(j + \omega )}; \end{aligned}$$

see Fig. 2 for an illustration.

Fig. 2
figure 2

Situation where the inclusion \(\omega \subset Y\) is a rank 1 laminate

To avoid effects caused by the interactions between cells and the boundary \(\partial {\varOmega }\), we impose periodic boundary conditions on \(\partial {\varOmega }\). In this context, the voltage potential \(u_N \in H^1_\#({\varOmega })/{\mathbb {R}}\) associated to a source \(f \in (H^1_\#({\varOmega }) / {\mathbb {R}})^*\) and to the distribution of conductivity equal to \(a \in {\mathbb {C}}\) in \(\omega _N\) and 1 in \(Y{\setminus }\overline{\omega _N}\) is solution to

$$\begin{aligned} -\mathrm{div}(A_N \nabla u) = f \text { in } {\varOmega }, \text { where } A_N(x) = \left\{ \begin{array}{ll} a &{}\quad \text {if } x \in \omega _N, \\ 1 &{}\quad \text {otherwise} \end{array} \right. . \end{aligned}$$
(A.1)

We are interested in the spectrum of the Poincaré variational operator \(T_N\), which maps an arbitrary function \(u \in H_\#^1({\varOmega }) / {\mathbb {R}}\) to the unique element \(T_N u \in H_\#^1({\varOmega }) / {\mathbb {R}}\) satisfying

$$\begin{aligned} \forall v \in H^1_\#({\varOmega })/ {\mathbb {R}}, \,\, \int _{{\varOmega }}{\nabla (T_N u)\cdot \nabla v \,\mathrm{d}x} = \int _{\omega _N}{\nabla u \cdot \nabla v\,\mathrm{d}x}, \end{aligned}$$

and, in particular, in the identification of the limit spectrum

$$\begin{aligned} \lim \limits _{N \rightarrow \infty }{\sigma (T_N)} = \left\{ \lambda \in {\mathbb {C}}, \,\, \exists \, N_j \rightarrow \infty , \, \lambda _{N_j} \in \sigma (T_{N_j}),\,\, \lambda _{N_j} \xrightarrow {N_j \rightarrow \infty } \lambda \right\} .\nonumber \\ \end{aligned}$$
(A.2)

1.1 Study of the Homogenization Process for the Operator \(T_N\)

The main tool in our analysis is the discrete Bloch decomposition [2], which has already been used without proof several times in this article. Although it is quite classical, we sketch the proof for completeness.

Theorem A.1

Let u be a function in \(L^2_\#({\varOmega })\). Then, there exists a unique collection \(\left\{ u_j \right\} \), indexed by \(j \in {\mathbb {N}}^2\), \(0 \leqq j \leqq N-1\), composed of \(N^2\) complex-valued functions in \(L^2_\#(Y)\) such that the following identity holds:

$$\begin{aligned} u(x) = \sum \limits _{0\leqq j \leqq N-1}{u_j(Nx)\, e^{2i\pi j \cdot x}}, \text { almost everywhere } x \in {\varOmega }. \end{aligned}$$
(A.3)

Furthermore, the Parseval identity holds: for \(u,v \in L^2_\#({\varOmega })\), with coefficients \(\left\{ u_j \right\} _{0\leqq j \leqq N-1}\), \(\left\{ v_j \right\} _{0\leqq j \leqq N-1}\), one has

$$\begin{aligned} \int _{\varOmega }{u(x)\overline{v(x)}\,\mathrm{d}x}= \sum \limits _{0\leqq j\leqq N-1}{\int _Y{u_j(y)\overline{v_j(y)}\,\mathrm{d}y}}. \end{aligned}$$
(A.4)

Proof

Let \(u \in L^2_\#(Y)\) be given, and assume that there exist \(N^2\) functions \(u_j \in L^2_\#(Y)\) such that (A.3) holds. Then, for arbitrary \(j^\prime \in {\mathbb {N}}^2\), \(0 \leqq j^\prime \leqq N-1\),

$$\begin{aligned} u\left( x + \frac{j^\prime }{N}\right) = \sum \limits _{0 \leqq j \leqq N-1}{u_j\left( Nx\right) e^{2i\pi j \cdot \left( x + \frac{j^\prime }{N}\right) }}. \end{aligned}$$

So as to isolate a particular index \(0 \leqq j^0 \leqq N-1\), we multiply both sides in the previous identity by \(e^{-2i\pi j^0 \cdot (x + \frac{j^\prime }{N})}\), then sum over \(j^\prime \) to obtain

$$\begin{aligned} \sum \limits _{\begin{array}{c} 0 \leqq j \leqq N-1 \\ 0 \leqq j^\prime \leqq N-1 \end{array}}{u_j\left( Nx\right) e^{2i\pi \left( j- j^0\right) \cdot \left( x + \frac{j^\prime }{N}\right) }} = \sum \limits _{0 \leqq j \leqq N-1}{u\left( x + \frac{j^\prime }{N}\right) e^{-2i\pi j^0 \cdot \left( x + \frac{j^\prime }{N}\right) }}.\nonumber \\ \end{aligned}$$
(A.5)

Using the fact that, for any \(N^{\text {th}}\) root \(r = e^{2\pi i j/N}\) of 1,

$$\begin{aligned} \sum \limits _{0\leqq k \leqq N-1}{r^k } = \left\{ \begin{array}{cc} N &{} \text {if } r = 1, \\ 0 &{} \text {otherwise} \end{array} \right. . \end{aligned}$$
(A.6)

The relation (A.5) simplifies into

$$\begin{aligned} u_{j^0}\left( y\right) =\frac{1}{N^2}\sum \limits _{0\leqq j^\prime \leqq N-1}{u\left( \frac{y}{N} + \frac{j^\prime }{N}\right) e^{-2i\pi j^0 \cdot \left( \frac{y}{N} + \frac{j^\prime }{N}\right) } }, \end{aligned}$$
(A.7)

a formula which clearly defines a function in \(L^2_\#(Y)\). Conversely, one easily proves that the \(u_j\) defined in (A.7) satisfy (A.3), which proves the first statement.

To verify the Parseval identity (A.4), let \(u,v \in L^2_\#({\varOmega })\) be decomposed as

$$\begin{aligned} u(x) = \sum \limits _{0\leqq j \leqq N-1}{u_j(Nx)\, e^{2i\pi j \cdot x}}, \text { and } v(x) = \sum \limits _{0\leqq j^\prime \leqq N-1}{ v_{j^\prime } (Nx)\, e^{2i\pi j^\prime \cdot x}}. \end{aligned}$$

A simple calculation yields

$$\begin{aligned} \int _\Omega {u(x)\overline{v(x)}\,\mathrm{d}x}= & {} \sum \limits _{0 \leqq j, j^\prime \leqq N-1}{\int _{\Omega }{u_j(Nx)\overline{v_{j^\prime }(Nx)} \, e^{2i\pi (j-j^\prime )\cdot x}\,\mathrm{d}x}}\\= & {} \sum \limits _{0 \leqq j, j^\prime ,k \leqq N-1}{\int _{\frac{1}{N}(k+Y)}{u_j(Nx)\overline{v_{j^\prime }(Nx)} \, e^{2i\pi (j-j^\prime )\cdot x}\,\mathrm{d}x}}\\= & {} \frac{1}{N^2}\sum \limits _{0 \leqq j, j^\prime ,k \leqq N-1}{\int _{Y}{u_j(y)\overline{v_{j^\prime }(y)} \, e^{2i\pi (j-j^\prime ) \cdot (\frac{k}{N} + \frac{y}{N})}\,\mathrm{d}y}}\\= & {} \sum \limits _{0 \leqq j \leqq N-1}{\int _{Y}{u_j(y)\overline{v_{j}(y)}\,\mathrm{d}y}}, \end{aligned}$$

where we have again made use of (A.6) to pass from the third line to the last. \(\quad \square \)

Let us now consider the operator B, from \(L^2_\#(Y)^{N^2}\) into \(L^2_\#({\varOmega })\) which maps a collection \(\left\{ u_j \right\} _{0\leqq j \leqq N-1}\) of coefficients to the function \(u \in L^2_\#({\varOmega })\) defined by

$$\begin{aligned} u(x) = \sum \limits _{0\leqq j \leqq N-1}{u_j(Nx)\, e^{2i\pi j \cdot x}}, \text { almost everywhere } x \in {\varOmega }. \end{aligned}$$

Equipping both spaces with their natural inner products, Theorem A.1 states that B is a bijective isometry, whose inverse, \( {B}^{-1}:L^2_\#({\varOmega }) \rightarrow L^2_\#(Y)^{N^2}\), is also its adjoint operator.

If u belongs to \(H^1_\#({\varOmega })\), (A.7) implies that the coefficients \(u_j\) of its Bloch decomposition actually belong to \(H^1_\#(Y)\), and that the Bloch decomposition of \(\nabla u\) read as

$$\begin{aligned} \nabla u(x) = N \sum \limits _{0\leqq j \leqq N-1}{\left( \nabla _y + 2i\pi \frac{j}{N}\right) u_j(Nx)\, e^{2i\pi j \cdot x}}, \text { almost everywhere } x \in {\varOmega }. \end{aligned}$$

Using the fact that the Bloch decomposition of the constant function \(u \equiv 1 \in H^1_\#({\varOmega })\) has coefficients

$$\begin{aligned} u_0(y) = 1, \text { and } u_j(y) = 0 \text { if } j \ne 0, \end{aligned}$$

B induces an invertible operator (still denoted by B) \((H^1_\#(Y)/{\mathbb {C}}) \times H^1_\#(Y)^{N^2-1} \rightarrow H^1_\#({\varOmega }) /{\mathbb {C}}\).

The following proposition easily follows from the previous remarks, and in particular from the Parseval identity (A.4):

Proposition A.1

For any \(\eta \in \overline{Y}\), \(\eta \ne 0\), define \(T_{\eta }: H^1_\#(Y) \rightarrow H^1_\#(Y)\) by

$$\begin{aligned}&\forall v \in H^1_\#(Y), \,\, \int _{Y}{(\nabla _y + 2i\pi \eta )(T_{\eta } u) \cdot \overline{(\nabla _y + 2i\pi \eta ) v }\,\mathrm{d}y} \nonumber \\&\quad = \int _{\omega }{(\nabla _y + 2i\pi \eta ) u \cdot \overline{(\nabla _y + 2i\pi \eta ) v }\,\mathrm{d}y }, \end{aligned}$$
(A.8)

and define \(T_0 : H^1_\#(Y) / {\mathbb {C}} \rightarrow H^1_\#(Y) / {\mathbb {C}}\) by

$$\begin{aligned} \forall v \in H^1(Y) / {\mathbb {C}}, \,\, \int _Y{\nabla _y (T_0 u) \cdot \overline{\nabla _y v}\,\mathrm{d}y} = \int _\omega {\nabla _y u \cdot \overline{\nabla _y v}\,\mathrm{d}y}. \end{aligned}$$
(A.9)

The operator B diagonalizes \(T_N\), that is the self-adjoint operator \(B^* T_N B\) maps a collection \(\left\{ u_j \right\} _{0 \leqq j \leqq N-1} \in (H^1_\#(Y)/{\mathbb {C}}) \times H^1_\#(Y)^{N^2-1}\) to \(\left\{ T_{\frac{j}{N}} u_j \right\} _{0 \leqq j \leqq N-1} \in (H^1_\#(Y)/{\mathbb {C}}) \times H^1_\#(Y)^{N^2-1}\). As a consequence, the spectrum \(\sigma (T_N)\) is the union of the spectra \(\sigma (T_{\frac{j}{N}})\):

$$\begin{aligned} \sigma (T_N) = \bigcup _{0 \leqq j \leqq N-1}{\sigma \left( T_{\frac{j}{N}}\right) }. \end{aligned}$$

Therefore, the study of the spectrum of \(T_N\), boils down to that of the spectra of the operators \(T_\eta \) defined by (A.8) and (A.9). Let us now take advantage of the particular geometry of \(\omega \) to simplify the problem further. We decompose functions \(u \in H^1_\#(Y)\) (or \(u \in H^1_\#(Y) / {\mathbb {C}}\)) by using partial Fourier series in the variable \(y_2\):

$$\begin{aligned} u(y) = \sum \limits _{n=-\infty }^{+\infty }{a_n(y_1) e^{2i\pi ny_2}}, \text { almost everywhere } y \in Y, \end{aligned}$$

where the \(a_n \in H^1_\#(0,1)\) (and \(a_0 \in H^1_\#(0,1) / {\mathbb {C}}\) if \(u \in H^1_\#(Y) / {\mathbb {C}}\)).

After some elementary calculations, the operators \(T_\eta \) are diagonalized by this Fourier decomposition, that is the spectrum of \(T_N\) read as

$$\begin{aligned} \sigma (T_N) = \bigcup _{\begin{array}{c} 0 \leqq j \leqq N-1 \\ n \in {\mathbb {N}} \end{array}}{\sigma \left( T_{\frac{j}{N}}^n\right) }, \end{aligned}$$

where, for any \(\eta = (\eta _1,\eta _2) \in \overline{Y}\), \(T_\eta ^n : H^1_\#(0,1) \rightarrow H^1_\#(0,1)\) is defined by

$$\begin{aligned}&\forall v \in H^1_\#(0,1) \in , \,\, \int _{0}^1\left( ((T_\eta ^n u)^\prime + 2i\pi \eta _1 (T_\eta ^n u))\overline{(v^\prime + 2i\pi \eta _1 v)}\right. \nonumber \\&\qquad \left. + 4\pi ^2(n+\eta _2)^2 u\overline{v}\right) \,\mathrm{d}y \nonumber \\&\quad = \int _{0}^\theta {\left( (u^\prime + 2i\pi \eta _1 u)\overline{(v^\prime + 2i\pi \eta _1 v)} + 4\pi ^2(n+\eta _2)^2 u\overline{v}\right) \,\mathrm{d}y}, \end{aligned}$$
(A.10)

and \(T_0^0 : H^1_\#(0,1)/ {\mathbb {C}} \rightarrow H^1_\#(0,1) / {\mathbb {C}}\) is given by

$$\begin{aligned} \forall v \in H^1_\#(0,1)/{\mathbb {C}}, \,\, \int _{0}^1{((T_0^0 u)^\prime )\overline{v^\prime }\,\mathrm{d}y} = \int _{0}^\theta {u^\prime \overline{v^\prime }\,\mathrm{d}y}. \end{aligned}$$
(A.11)

It is proved in the same way as in Section 3.3 that the spectrum of each operator \(T_\eta ^n\) consists of a sequence of eigenvalues in [0, 1] with \(\frac{1}{2}\) as unique accumulation point, and we now proceed to identify these eigenvalues.

Let us first study the eigenvalues of the operator \(T_\eta ^n\) in the case where either \(\eta _2 \ne 0\) or \(n \ne 0\). A value \(\beta \in {\mathbb {C}}\) is an eigenvalue for \(T_\eta ^n\) as defined in (A.10) if there exists \(u \in H^1_\#(0,1)\), \(u\ne 0\) such that

$$\begin{aligned} -\left( \frac{\partial }{\partial y_1} + 2i\pi \eta _1 \right) \left( A_\beta (y_1)\left( \frac{\partial u}{\partial y_1} + 2i\pi \eta _1 u \right) \right) + 4\pi ^2 A_\beta (y_1)(n + \eta _2)^2 u =0,\nonumber \\ \end{aligned}$$
(A.12)

where

$$\begin{aligned} A_\beta (y) = \left\{ \begin{array}{ll} \beta -1 &{}\quad \text {if } y_1< \theta , \\ \beta &{}\quad \text {if } \theta< y_1 < 1. \\ \end{array} \right. \end{aligned}$$

Assuming \(\beta \notin \left\{ 0,1\right\} \), (A.12) is equivalent to

$$\begin{aligned} u^{\prime \prime }(z) + 4i \pi \eta _1 u^\prime (z) - 4\pi ^2(\eta _1^2 + (n+\eta _2)^2) u(z) = 0 \text { almost everywhere } z\in (0,\theta ) \text { and } z \in (\theta ,1),\nonumber \\ \end{aligned}$$
(A.13)

complemented with the transmission conditions at \(z=0\) and \(z = \theta \):

$$\begin{aligned} u(0^+) = u(1^-), \,\,\,\, \beta (u^\prime + 2i\pi \eta _1u)(1^-) = (\beta -1) (u^\prime + 2i\pi \eta _1u)(0^+), \end{aligned}$$
(A.14)
$$\begin{aligned} u(\theta ^-) = u(\theta ^+), \,\,\,\, (\beta -1) (u^\prime + 2i\pi \eta _1u)(\theta ^-) = \beta (u^\prime + 2i\pi \eta _1u)(\theta ^+).\nonumber \\ \end{aligned}$$
(A.15)

The ordinary differential equation (A.13) has discriminant \({\varDelta } = 16\pi ^2(n + \eta _2)^2\), and the associated characteristic equation has two solutions:

$$\begin{aligned} r_1 = -2i \pi \eta _1 - 2\pi (n + \eta _2), \,\, r_2 = -2i\pi \eta _1 + 2\pi (n + \eta _2), \end{aligned}$$

which are distinct, since \(n+\eta _2 \ne 0\). Therefore, there exist 4 coefficients \(A,B,C,D \in {\mathbb {C}}\) such that

$$\begin{aligned} u(z) = Ae^{r_1z} + Be^{r_2z} \text { for } z \in (0,\theta ), \,\, u(z) = Ce^{r_1z} + De^{r_2z} \text { for } z \in (\theta ,1). \end{aligned}$$

The fact that \(u \in H^1_\#(0,1)\) imposes on us that

$$\begin{aligned} A + B = Ce^{r_1} + De^{r_2}, \text { and } Ae^{r_1\theta } + Be^{r_2\theta } = Ce^{r_1\theta } + De^{r_2\theta }, \end{aligned}$$

to be complemented with the transmission conditions

$$\begin{aligned}&(\beta -1) (r_1+2i\pi \eta _1) A+ (\beta -1) (r_2+2i\pi \eta _1) B \\&\quad = \beta (r_1+2i\pi \eta _1)Ce^{r_1} + \beta (r_2+2i\pi \eta _1)De^{r_2} \\&(\beta -1) (r_1+2i\pi \eta _1) Ae^{r_1\theta } + (\beta -1) (r_2+2i\pi \eta _1) Be^{r_2\theta } \\&\quad = \beta (r_1+2i\pi \eta _1)Ce^{r_1\theta } + \beta (r_2+2i\pi \eta _1)De^{r_2\theta } \end{aligned}$$

As a consequence, (A.13,A.14,A.15) has a non-trivial solution provided the following determinant vanishes:

$$\begin{aligned} \left|\begin{array}{c@{\quad }c@{\quad }c@{\quad }c} 1 &{} 1 &{} -e^{r_1} &{} -e^{r_2} \\ e^{r_1\theta } &{} e^{r_2\theta } &{} -e^{r_1\theta } &{} -e^{r_2\theta } \\ (\beta -1) (r_1+2i\pi \eta _1) &{} (\beta -1) (r_2+2i\pi \eta _1) &{} -\beta (r_1+2i\pi \eta _1) e^{r_1} &{} -\beta (r_2+2i\pi \eta _1)e^{r_2}\\ (\beta -1) (r_1+2i\pi \eta _1) e^{r_1\theta } &{} (\beta -1) (r_2+2i\pi \eta _1) e^{r_2\theta } &{} -\beta (r_1+2i\pi \eta _1) e^{r_1\theta } &{} -\beta (r_2+2i\pi \eta _1)e^{r_2\theta } \end{array} \right|= 0,\nonumber \\ \end{aligned}$$
(A.16)

which is a quadratic equation for \(\beta \). After tedious calculations, (A.16) simplifies into

$$\begin{aligned} \beta ^2 - \beta + \gamma = 0, \text { where } \gamma := \frac{1}{4} \frac{\cosh (2\pi (n + \eta _2)) - \cosh ((2\pi (n + \eta _2))(2\theta -1))}{\cosh (2\pi (n + \eta _2)) - \cos (2\pi \eta _1)}. \end{aligned}$$

The discriminant of this second order equation read as

$$\begin{aligned} {\varDelta }_\eta ^n = \frac{\cosh \left( (2\pi (n + \eta _2))(2\theta -1)\right) -\cos (2\pi \eta _1)}{\cosh (2\pi (n + \eta _2)) - \cos (2\pi \eta _1)}. \end{aligned}$$
(A.17)

We observe that \({\varDelta }_\eta ^n \in (0,1)\), leading to two distinct eigenvalues \(\beta _\eta ^{n\pm } = (1 \pm \sqrt{{\varDelta }_\eta ^n})/2\), which are symmetric with respect to \(\frac{1}{2}\).

As for the eigenvalues of \(T_\eta ^n\) in the case where \(\eta _2 = 0\) and \(n=0\), simple calculations show that \(\sigma (T_\eta ^n) = \left\{ 0, 1\right\} \) if \(\eta _1 \ne 0\) and that \(\sigma (T_0^0) = \left\{ 0,1-\theta ,1\right\} \). All in all, we have proved that the spectrum of \(T_N\) is

$$\begin{aligned} \sigma (T_N) = \left\{ 0,1-\theta ,1\right\} \cup \left\{ \frac{1}{2}\left( 1\pm \sqrt{{\varDelta }_{\frac{j}{N}}^n}\right) \right\} _{\begin{array}{c} 0 < j \leqq N-1 \\ n \in {\mathbb {N}} \end{array}}, \end{aligned}$$

where \({\varDelta }_\eta ^n\) is defined by (A.17). This allows for the identification of the limit spectrum (A.2) as

$$\begin{aligned} \lim \limits _{N\rightarrow \infty }{\sigma (T_N)} = \left\{ 0,1-\theta ,1\right\} \cup \left\{ \frac{1}{2}\left( 1\pm \sqrt{{\varDelta }_{\eta }^n}\right) \right\} _{\eta \in \overline{Y}, n \in {\mathbb {N}}} = [0,1]. \end{aligned}$$

Hence, in the present situation of rank 1 laminates, the limit spectrum of \(T_N\) is the whole interval [0, 1], in sharp contrast with the situation where \(\omega \Subset Y\), as studied in Section 4.

1.2 Analysis of the Homogenized Tensor \(A^*\)

As we have seen in Section 5.2 (see in particular Propositions 5.5 and 5.6, which are straightforwardly adapted to this context), the asymptotic behavior of the voltage potential \(u_N\) associated to a conductivity \(a \in {\mathbb {C}}\) inside the set of inclusions \(\omega _N\), solution to (A.1), is partly driven by the formally homogenized tensor \(A^*\) whose components \(A^*_{ij}\) are given by (\(i,j=1,2\)):

$$\begin{aligned} A^*_{ij} = \int _Y{A(y) (e_i + \nabla _y \chi _i) \cdot (e_j + \nabla _y \chi _j) \,\mathrm{d}y}, \text { where } A(y) = \left\{ \begin{array}{ll} a &{}\quad \text {if } y \in \omega , \\ 1 &{}\quad \text {otherwise}, \end{array} \right. \nonumber \\ \end{aligned}$$
(A.18)

and the cell functions \(\chi _i \in H^1_\#(Y) / {\mathbb {R}}\) (\(i=1,2\)) are solutions to

$$\begin{aligned} -\mathrm{div}_y(A(y)(e_i + \nabla _y \chi _i)) = 0. \end{aligned}$$
(A.19)

As we have seen in Section 5.1, both cell problems (A.19) have a unique solution, provided the conductivity a lies outside the exceptional set \({\varSigma }_\omega \) given by

$$\begin{aligned} {\varSigma }_\omega = \left\{ a \in {\mathbb {C}}, \,\, \frac{1}{1-a} \in \sigma (T_0)\right\} , \end{aligned}$$

where \(T_0 : H^1_\#(Y) / {\mathbb {R}} \rightarrow H^1_\#(Y) / {\mathbb {R}}\) is (the real version of that ) defined by (A.9) as follows:

$$\begin{aligned} \forall v \in H^1_\#(Y) / {\mathbb {R}}, \,\, \int _Y{\nabla _y (T_0 u) \cdot \nabla _y v\,\mathrm{d}y} = \int _\omega {\nabla _y u \cdot \nabla _y v\,\mathrm{d}y}. \end{aligned}$$
(A.20)

On the contrary, in the case \(a \in {\varSigma }_\omega \), there may be multiple solutions to (A.19) (or none), but when this happens, it is easily seen that (A.18) is independent of which of these solutions is used.

Calculations similar to those performed in the previous section (based on a partial Fourier decomposition) lead to an explicit characterization of the set \({\varSigma }_\omega \) in the present context of rank 1 laminates:

$$\begin{aligned} {\varSigma }_\omega := \left\{ -\frac{\theta }{1-\theta }, 0\right\} \cup \left\{ a_n^\pm \right\} _{n \in {\mathbb {N}}^*}, \end{aligned}$$

where the \(a_n^\pm \) read as

$$\begin{aligned} a^n_{\pm } = \frac{2(1 \pm 2\sinh (\pi n)\sinh (\pi n (2\theta -1)) ) - \cosh (2\pi n (2\theta -1)) - \cosh (2\pi n)}{\cosh (2\pi n) - \cosh (2\pi n (2\theta -1))}. \end{aligned}$$

Let us now turn to the cell problems (A.19) in this setting for an arbitrary conductivity \(a \in {\mathbb {C}}\). Relying again on a partial Fourier decomposition in the \(y_2\) variable, it is easy to see that, if (A.19) has solutions, one of them is necessarily of the form

$$\begin{aligned} \chi _i(y) = u_i(y_1), \text { for some function } u_i \in H^1_\#(0,1) / {\mathbb {R}}, \, i = 1,2. \end{aligned}$$
(A.21)

Now, easy calculations reveal that if \(a \ne -\frac{\theta }{1-\theta }\), (A.19) has (possibly non-unique) solutions \(\chi _i\) for \(i=1,2\):

$$\begin{aligned} \chi _1(y) = \left\{ \begin{array}{ll} Ay_1 + B &{}\quad \text {if } y_1 < \theta , \\ Cy_1 + D &{}\quad \text {if } y_1 > \theta , \\ \end{array} \right. ,\text { and } \chi _2(y) = 0, \end{aligned}$$

where the coefficients A and C read as

$$\begin{aligned} A = \frac{1-a}{a - \frac{\theta }{1-\theta }}, \, C = A \frac{\theta }{\theta -1}. \end{aligned}$$

Then, the homogenized tensor (A.18) equals

$$\begin{aligned} A^* = \left( \begin{array}{cc} \lambda _{\theta ,a}^- &{} 0, \\ 0 &{} \lambda _{\theta ,a}^+ \end{array} \right) , \text { where } \lambda _{\theta ,a}^- = \left( \frac{\theta }{a} + 1-\theta \right) ^{-1}, \text { and }\lambda _{\theta ,a}^+ = a\theta + (1-\theta ). \end{aligned}$$

Note that \(A^*\) is invertible, and becomes degenerate when a gets close to \(-\frac{\theta }{1-\theta }\). The behaviors of the mappings \(a \mapsto \lambda _{\theta ,a}^\pm \) change depending on whether the volume fraction \(\theta \) is larger or smaller than \(\frac{1}{2}\), as can be seen on Fig. 3. In the case where \(\theta < \frac{1}{2}\), three regimes are to be distinguished:

  • When \(-\frac{\theta }{1-\theta }< a < 0 \), \(A^*\) has eigenvalues with opposite signs.

  • When \(-\frac{1-\theta }{\theta }<a<-\frac{\theta }{1-\theta }\), \(A^*\) is positive definite.

  • When \(a < -\frac{1-\theta }{\theta }\), \(A^*\) has again eigenvalues with opposite signs.

This behavior differs in the case where \(\theta > \frac{1}{2}\):

  • When \(-\frac{1-\theta }{\theta }< a < 0 \), \(A^*\) has eigenvalues with opposite signs.

  • When \(-\frac{\theta }{1-\theta }<a<-\frac{1-\theta }{\theta }\), \(A^*\) is negative definite.

  • When \(a < -\frac{\theta }{1-\theta }\), \(A^*\) has again eigenvalues with opposite signs. These results are in sharp contrast with the case of inclusions \(\omega \Subset Y\), dealt with in Section 5. In the case \(a = -\frac{\theta }{1-\theta }\), the cell problem (A.19) has infinitely many solutions of the form (A.21) for \(i=2\), and none for \(i=1\).

It is remarkable that the only value of a for which the cell problems do not have solutions is \(a = -\frac{\theta }{1-\theta }\), corresponding to the essential spectrum of the operator defined in (A.20). We do not know whether this fact holds for general inclusion patterns \(\omega \subset Y\) or if it is particular to rank 1 laminates.

Fig. 3
figure 3

Behavior of the eigenvalues \(a \mapsto \lambda _{\theta ,a}^\pm \) of the homogenized matrix \(A^*\) for the volume fractions (left) \(\theta = \frac{1}{3}\), and (right) \(\theta = \frac{2}{3}\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bonnetier, É., Dapogny, C. & Triki, F. Homogenization of the Eigenvalues of the Neumann–Poincaré Operator. Arch Rational Mech Anal 234, 777–855 (2019). https://doi.org/10.1007/s00205-019-01402-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00205-019-01402-8

Navigation