Skip to main content
Log in

Exact relations for Green’s functions in linear PDE and boundary field equalities: a generalization of conservation laws

  • Research
  • Published:
Research in the Mathematical Sciences Aims and scope Submit manuscript

Abstract

Many physical equations have the form \(\mathbf{J }(\mathbf{x })=\mathbf{L }(\mathbf{x })\mathbf{E }(\mathbf{x })-\mathbf{h }(\mathbf{x })\) with source \(\mathbf{h }(\mathbf{x })\) and fields \(\mathbf{E }\) and \(\mathbf{J }\) satisfying differential constraints, symbolized by \(\mathbf{E }\in \mathcal E\), \(\mathbf{J }\in \mathcal J\) where \(\mathcal E\), \(\mathcal J\) are orthogonal spaces. We show that if \(\mathbf{L }(\mathbf{x })\) takes values in certain nonlinear manifolds \(\mathcal M\), and coercivity and boundedness conditions hold, then the infinite body Green’s function (fundamental solution) satisfies exact identities. The theory also links Green’s functions of different problems. The analysis is based on the theory of exact relations for composites, but without assumptions about the length scales of variations in \(\mathbf{L }(\mathbf{x })\), and more general equations, such as for waves in lossy media, are allowed. For bodies \(\Omega \), inside which \(\mathbf{L }(\mathbf{x })\in \mathcal{M}\), the “Dirichlet-to-Neumann map” giving the response also satisfies exact relations. These boundary field equalities generalize the notion of conservation laws: the field inside \(\Omega \) satisfies certain constraints that leave a wide choice in these fields, but which give identities satisfied by the boundary fields, and moreover provide constraints on the fields inside the body. A consequence is the following: if a matrix-valued field \(\mathbf{Q }(\mathbf{x })\) with divergence-free columns takes values within \(\Omega \) in a set \(\mathcal B\) (independent of \(\mathbf{x }\)) that lies on a nonlinear manifold, we find conditions on the manifold, and on \(\mathcal B\), that with appropriate conditions on the boundary fluxes \(\mathbf{q }(\mathbf{x })=\mathbf{n }(\mathbf{x })\cdot \mathbf{Q }(\mathbf{x })\) (where \(\mathbf{n }(\mathbf{x })\) is the outward normal to \(\partial \Omega \)) force \(\mathbf{Q }(\mathbf{x })\) within \(\Omega \) to take values in a subspace \(\mathcal D\). This forces \(\mathbf{q }(\mathbf{x })\) to take values in \(\mathbf{n }(\mathbf{x })\cdot \mathcal D\). We find there are additional divergence-free fields inside \(\Omega \) that in turn generate additional boundary field equalities. Consequently, there exist partial null Lagrangians, functionals \(F(\mathbf{w },\nabla \mathbf{w })\) of a vector potential \(\mathbf{w }\) and its gradient that act as null Lagrangians when \(\nabla \mathbf{w }\) is constrained for \(\mathbf{x }\in \Omega \) to take values in certain sets \(\mathcal A\), of appropriate nonlinear manifolds, and when \(\mathbf{w }\) satisfies appropriate boundary conditions. The extension to certain nonlinear minimization problems is also sketched.

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
Fig. 2
Fig. 3

Similar content being viewed by others

References

  1. Allaire, G.: Shape Optimization by the Homogenization Method. Applied Mathematical Sciences, vol. 146. Springer, Berlin (2002)

    MATH  Google Scholar 

  2. Backus, G.E.: Long-wave elastic anisotropy produced by horizontal layering. J. Geophys. Res. 67(11), 4427–4440 (1962)

    Article  Google Scholar 

  3. Ball, J.M.: Convexity conditions and existence theorems in non-linear elasticity. Arch. Ration. Mech. Anal. 63(4), 337–403 (1977)

    Article  Google Scholar 

  4. Batchelor, G.K.: Transport properties of two-phase materials with random structure. Annu. Rev. Fluid Mech. 6, 227–255 (1974)

    Article  Google Scholar 

  5. Beněsová, B.: Kružík, Martin: weak lower semicontinuity of integral functionals and applications. SIAM Rev. 59(4), 703–766 (2017)

    Article  MathSciNet  Google Scholar 

  6. Cherkaev, A.V.: Variational Methods for Structural Optimization. Applied Mathematical Sciences. Springer, Berlin (2000)

    Book  Google Scholar 

  7. Cherkaev, A.V., Gibiansky, L.V.: Variational principles for complex conductivity, viscoelasticity, and similar problems in media with complex moduli. J. Math. Phys. 35(1), 127–145 (1994)

    Article  MathSciNet  Google Scholar 

  8. Clark, K.E., Milton, G.W.: Modeling the effective conductivity function of an arbitrary two-dimensional polycrystal using sequential laminates. Proc. R. Soc. Edinb. 124A(4), 757–783 (1994)

    Article  Google Scholar 

  9. Dell’Antonio, G.F., Figari, R., Orlandi, E.: An approach through orthogonal projections to the study of inhomogeneous or random media with linear response. Ann l’inst Henri Poincaré (A) Phys. théor. 44(1), 1–28 (1986)

    MathSciNet  MATH  Google Scholar 

  10. Dykhne, A. M.: Conductivity of a two-dimensional two-phase system. Zhurnal eksperimental’noi i teoreticheskoi fiziki / Akademiia Nauk SSSR, 59:110–115, July 1970. English translation in Soviet Physics JETP 32(1):63–65 (1971)

  11. Eyre, D.J., Milton, G.W.: A fast numerical scheme for computing the response of composites using grid refinement. Eur. Phys. J. Appl. Phys. 6(1), 41–47 (1999)

    Article  Google Scholar 

  12. Faraco, D., Székelyhidi, L.: Tartars conjecture and localization of the quasiconvex hull in \(\mathbb{R}^{2\times 2}\). Acta Math. 200(2), 279–305 (2008)

    Article  MathSciNet  Google Scholar 

  13. Grabovsky, Y.: Exact relations for effective tensors of polycrystals. I: necessary conditions. Arch. Ration. Mech. Anal. 143(4), 309–329 (1998)

    Article  MathSciNet  Google Scholar 

  14. Grabovsky, Y.: Algebra, geometry and computations of exact relations for effective moduli of composites. In: Gianfranco Capriz and Paolo Maria Mariano, editors, Advances in Multifield Theories of Continua with Substructure, Modelling and Simulation in Science, Engineering and Technology, pp. 167–197. Birkhäuser Verlag, Boston (2004)

    Chapter  Google Scholar 

  15. Grabovsky, Y.: Composite Materials: Mathematical Theory and Exact Relations. IOP Publishing, Bristol (2016)

    Book  Google Scholar 

  16. Grabovsky, Y.: From microstructure-independent formulas for composite materials to rank-one convex, non-quasiconvex functions. Arch. Ration. Mech. Anal. 227(2), 607–636 (2018)

    Article  MathSciNet  Google Scholar 

  17. Grabovsky, Y., Milton, G.W., Sage, D.S.: Exact relations for effective tensors of composites: Necessary conditions and sufficient conditions. Commun. Pure Appl. Math. (New York) 53(3), 300–353 (2000)

    Article  MathSciNet  Google Scholar 

  18. Grabovsky, Y., Sage, D.S.: Exact relations for effective tensors of polycrystals. II: applications to elasticity and piezoelectricity. Arch. Ration. Mech. Anal. 143(4), 331–356 (1998)

    Article  MathSciNet  Google Scholar 

  19. Kohler, W., Papanicolaou, G. C: Bounds for the effective conductivity of random media. In Burridge, R., Childress, S., Papanicolaou, G.C. (eds.) Macroscopic Properties of Disordered Media: Proceedings of a Conference Held at the Courant Institute, June 1–3, 1981, Lecture Notes in Physics, vol. 154, pp. 111–130. Springer, Berlin (1982)

  20. Kohn, R.V.: The relaxation of a double-well energy. Contin. Mech. Thermodyn. 3(3), 193–236 (1991)

    Article  MathSciNet  Google Scholar 

  21. Lax, P.D.: Functional Analysis. Wiley, New York (2002)

    MATH  Google Scholar 

  22. Levin, V. M.: Thermal expansion coefficients of heterogeneous materials. Inzhenernyi Zhurnal. Mekhanika Tverdogo Tela: MTT, 2(1):88–94 (1967). English translation in Mechanics of Solids 2(1):58–61 (1967)

  23. Milgrom, M.: Linear response of general composite systems to many coupled fields. Phys. Rev. B: Condens. Matter Mater. Phys. 41(18), 12484–12494 (1990)

    Article  Google Scholar 

  24. Milton, G.W.: Multicomponent composites, electrical networks and new types of continued fraction. I. Commun. Math. Phys. 111(2), 281–327 (1987)

    Article  MathSciNet  Google Scholar 

  25. Milton, G.W.: On characterizing the set of possible effective tensors of composites: the variational method and the translation method. Commun. Pure Appl. Math. (New York) 43(1), 63–125 (1990)

    Article  MathSciNet  Google Scholar 

  26. Milton G. W: The Theory of Composites. Cambridge Monographs on Applied and Computational Mathematics, vol. 6. In: Ciarlet, P.G, Iserles, A., Kohn, R.V., Wright, M.H. (eds.) Cambridge University Press, Cambridge (2002)

  27. Milton, G.W.: Sharp inequalities that generalize the divergence theorem: an extension of the notion of quasi-convexity. Proc. R. Soc. A: Math. Phys. Eng. Sci. 469(2157), 20130075 (2013). See addendum [29]

    Article  MathSciNet  Google Scholar 

  28. Milton, G.W.: Addendum to “Sharp inequalities that generalize the divergence theorem: an extension of the notion of quasi-convexity. Proc. R. Soc. A: Math. Phys. Eng. Sci. 471(2176), 20140886 (2015). See [28]

    Article  Google Scholar 

  29. Milton, G.W.: A new route to finding bounds on the generalized spectrum of many physical operators. J. Math. Phys. 59, 061508 (2018)

    Article  MathSciNet  Google Scholar 

  30. Milton, G.W., Briane, M., Willis, J.R.: On cloaking for elasticity and physical equations with a transformation invariant form. New J. Phys. 8(10), 248 (2006)

    Article  Google Scholar 

  31. Milton, G.W., Golden, K.M.: Representations for the conductivity functions of multicomponent composites. Commun. Pure Appl. Math. (New York) 43(5), 647–671 (1990)

    Article  MathSciNet  Google Scholar 

  32. Milton, G.W., Seppecher, P., Guy, B.: Minimization variational principles for acoustics, elastodynamics and electromagnetism in lossy inhomogeneous bodies at fixed frequency. Proc. R. Soc. A: Math. Phys. Eng. Sci. 465(2102), 367–396 (2009)

    Article  MathSciNet  Google Scholar 

  33. Milton, G.W., Willis, J.R.: On modifications of Newton’s second law and linear continuum elastodynamics. Proc. R. Soc. A: Math. Phys. Eng. Sci. 463(2079), 855–880 (2007)

    Article  MathSciNet  Google Scholar 

  34. Milton, G.W., Willis, J.R.: Minimum variational principles for time-harmonic waves in a dissipative medium and associated variational principles of Hashin–Shtrikman type. Proc. R. Soc. A: Math. Phys. Eng. Sci. 466(2122), 3013–3032 (2010)

    Article  MathSciNet  Google Scholar 

  35. Milton, G.W. (ed.): Extending the Theory of Composites to Other Areas of Science. Milton–Patton Publishers, Salt Lake City (2016)

    Google Scholar 

  36. Moulinec, H., Suquet, P.M.: A fast numerical method for computing the linear and non-linear properties of composites. C. r. Séances l’Acad. Sci. Sér. II 318, 1417–1423 (1994)

    MATH  Google Scholar 

  37. Müller, S., Šverák, V.: Convex integration for Lipschitz mappings and counterexamples to regularity. Ann. Math. 157(3), 715–742 (2003)

    Article  MathSciNet  Google Scholar 

  38. Murat, F.: Compacité par compensation. Ann. Sc. Norm. Super. Pisa, Classe Sci. Ser. IV 5(3), 489–507 (1978). (French) [Compactness through compensation]

    MATH  Google Scholar 

  39. Murat, F.: Compacité par compensation: condition nécessaire et suffisante de continuité faible sous une hypothése de rang constant. (French) [Compensated compactness: Necessary and sufficient conditions for weak continuity under a constant-rank hypothesis]. Ann. Sc. Norm. Super. Pisa, Classe Sci. Ser. IV

  40. Murat, F.: A survey on compensated compactness. In: Cesari, L. (ed.) Contributions to Modern Calculus of Variations. Pitman Research Notes in Mathematics Series, vol. 148, pp. 145–183, Harlow, Essex (1987). Longman Scientific and Technical. Papers from the symposium marking the centenary of the birth of Leonida Tonelli held in Bologna, May 13–14, (1985)

  41. Olver, P.J., Sivaloganathan, J.: The structure of null Lagrangians. Nonlinearity (Bristol) 1(2), 389–398 (1988)

    Article  MathSciNet  Google Scholar 

  42. Pablo, P.: Weak continuity and weak lower semicontinuity for some compensation operators. Proc. R. Soc. Edinb. Sect. A, Math. Phys. Sci. 113(3–4), 267–279 (1989)

    MathSciNet  MATH  Google Scholar 

  43. Raitums, U.E.: On the local representation of \(G\)-closure. Arch. Ration. Mech. Anal. 158(3), 213–234 (2001)

    Article  MathSciNet  Google Scholar 

  44. Schoenberg, M., Sen, P.N.: Properties of a periodically stratified acoustic half-space and its relation to a Biot fluid. J. Acoust. Soc. Am. 73(1), 61–67 (1983)

    Article  Google Scholar 

  45. Tartar, L.: Compensated compactness and applications to partial differential equations. In: Knops, R.J. (ed.) Nonlinear Analysis and Mechanics, Heriot–Watt Symposium, Volume IV. Research Notes in Mathematics, vol. 39, pp. 136–212. Pitman Publishing Ltd, London (1979)

    Google Scholar 

  46. Tartar, L.: Estimation de coefficients homogénéisés. (French) [Estimation of homogenization coefficients]. In: Glowinski, R., Lions, J.L. (eds.) Computing Methods in Applied Sciences and Engineering: Third International Symposium, Versailles, France, December 5–9, 1977. Lecture Notes in Mathematics, vol. 704 , pp. 364–373. Springer, Berlin (1979). English translation in Topics in the Mathematical Modelling of Composite Materials, pp. 9–20, Cherkaev, A., Kohn, R. (eds.). ISBN 0-8176-3662-5

  47. Tartar, L.: The General Theory of Homogenization: A Personalized Introduction. Lecture Notes of the Unione Matematica Italiana. Springer, Berlin (2009)

    MATH  Google Scholar 

  48. Thaler, A.E., Milton, G.W.: Exact determination of the volume of an inclusion in a body having constant shear modulus. Inverse Probl. 30(12), 125008 (2014)

    Article  MathSciNet  Google Scholar 

  49. Torquato, S.: Random Heterogeneous Materials: Microstructure and Macroscopic Properties. Interdisciplinary Applied Mathematics, vol. 16. Springer, Berlin (2002)

    MATH  Google Scholar 

  50. Willis, J.R.: The non-local influence of density variations in a composite. Int. J. Solids Struct. 21(7), 805–817 (1985)

    Article  Google Scholar 

  51. Zhang, K.: On the structure of quasiconvex hulls. Ann. l’Inst. Henri Poincaré. Anal. Linéaire 15(6), 663–686 (1998)

    Article  MathSciNet  Google Scholar 

  52. Zhikov, V. V.: Estimates for the homogenized matrix and the homogenized tensor. Uspekhi Matematicheskikh Nauk = Russ. Math. Surv. 46:49–109 (1991). English translation in Russ. Math. Surv. 46(3):65–136 (1991)

    Article  MathSciNet  Google Scholar 

Download references

Acknowlegements

GWM is grateful to the National Science Foundation for support through the Research Grants DMS-1211359 and DMS-1814854. Both authors thank the Institute for Mathematics and its Applications at the University of Minnesota for hosting their visit there during the Fall 2016 where this work was initiated as part of the program on Mathematics and Optics. Yury Grabovsky and the referees are thanked for their comments which led to significant improvements of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Graeme W. Milton.

Additional information

Publisher's Note

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

Appendix: Some physical equations that can be expressed in the required canonical form

Appendix: Some physical equations that can be expressed in the required canonical form

Here we give examples of some physical equations that can be expressed in the required canonical form. The examples are by no means comprehensive: for further examples, see [26] Chap. 2, [35] Chap. 1, [27, 28] and the appendix of [29]. In the equations that follow we omit the source terms. We emphasize that if one is interested in exact relations satisfied by the DtN map and the associated boundary field equalities, then it is not necessary that the source terms have a physical significance provided they are localized outside the body \(\Omega \).

The simplest canonical equations are those of electrical conductivity

$$\begin{aligned} \underbrace{\mathbf{j }(\mathbf{x })}_{\mathbf{J }(\mathbf{x })}=\underbrace{{\varvec{\sigma }}(\mathbf{x })}_{\mathbf{L }(\mathbf{x })}\underbrace{\mathbf{e }(\mathbf{x })}_{\mathbf{E }(\mathbf{x })},\quad \nabla \cdot \mathbf{j }=0, \quad \mathbf{e }=-\nabla V=0, \end{aligned}$$
(10.4)

where \(\mathbf{j }(\mathbf{x })\) and \(\mathbf{e }(\mathbf{x })\) are the electrical current and electric field and \(V(\mathbf{x })\) is the electrical potential. The boundary fields \(\partial \mathbf{J }\) and \(\partial \mathbf{E }\) are the flux \(\mathbf{n }\cdot \mathbf{j }(\mathbf{x })\) and boundary voltage \(V(\mathbf{x })\), respectively, with \(\mathbf{x }\in \partial \Omega \) and \(\mathbf{n }\) being the outward normal to \(\Omega \). As displayed in the table at the beginning of Sect. 2.1 in [26] (adapted from one of Batchelor [4]), the equations for dielectrics, magnetostatics, heat conduction, particle diffusion, flow in porous media, and antiplane elasticity all take the same form as (10.4) and so any analysis applicable to (10.4) applies to them as well.

Another important example is that of linear elasticity,

$$\begin{aligned} \underbrace{{\varvec{\sigma }}(\mathbf{x })}_{\mathbf{J }(\mathbf{x })}=\underbrace{{\varvec{\mathcal {C}}}(\mathbf{x })}_{\mathbf{L }(\mathbf{x })}\underbrace{{\varvec{\epsilon }}(\mathbf{x })}_{\mathbf{E }(\mathbf{x })},\quad \nabla \cdot {\varvec{\sigma }}=0,\quad {\varvec{\epsilon }}=[\nabla \mathbf{u }+(\nabla \mathbf{u })^T]/2, \end{aligned}$$
(10.5)

where \({\varvec{\sigma }}(\mathbf{x })\) (not to be confused for the conductivity tensor field) is the stress, \({\varvec{\epsilon }}(\mathbf{x })\) is the strain, \(\mathbf{u }(\mathbf{x })\) is the displacement, and \({\varvec{\mathcal {C}}}(\mathbf{x })\) is the elasticity tensor field. The boundary fields \(\partial \mathbf{J }(\mathbf{x })\) and \(\partial \mathbf{E }(\mathbf{x })\) are the traction \(\mathbf{n }\cdot {\varvec{\sigma }}(\mathbf{x })\) and boundary displacement field \(\mathbf{u }(\mathbf{x })\), respectively.

It is also possible to have equations that couple fields together, such as the magnetoelectric equations,

$$\begin{aligned} \underbrace{\begin{pmatrix}\mathbf{d }\\ \mathbf{b }\end{pmatrix}}_{\mathbf{J }(\mathbf{x })}=\underbrace{\begin{pmatrix}{\varvec{\varepsilon }}&{}\quad {\varvec{\beta }}\\ {\varvec{\beta }}^T &{}\quad {\varvec{\mu }}\end{pmatrix}}_{\mathbf{L }(\mathbf{x })}\underbrace{\begin{pmatrix}\mathbf{e }\\ \mathbf{h }\end{pmatrix}}_{\mathbf{E }(\mathbf{x })},\quad \nabla \cdot \mathbf{d }=\nabla \cdot \mathbf{b }=0,\quad \mathbf{e }=-\nabla V,\quad \mathbf{h }=-\nabla \psi , \end{aligned}$$
(10.6)

where \(\mathbf{d }\) and \(\mathbf{b }\) are the electric displacement field and magnetic induction, \(\mathbf{e }\) and \(\mathbf{h }\) are the electric and magnetic fields, V and \(\psi \) are the electric potential and magnetic scalar potential (assuming there are no free currents), \({\varvec{\varepsilon }}\)(x) is the free-body electrical permittivity (with \(\mathbf{h }=0\)), \({\varvec{\beta }}(\mathbf{x })\) is the second-order magnetoelectric coupling tensor, \({\varvec{\mu }}(\mathbf{x })\) is the free-body magnetic permeability (with \(\mathbf{e }=0\)). The boundary fields \(\partial \mathbf{J }(\mathbf{x })\) and \(\partial \mathbf{E }(\mathbf{x })\) are then the flux pair \((\mathbf{n }\cdot \mathbf{d }(\mathbf{x }),\mathbf{n }\cdot \mathbf{b }(\mathbf{x }))\) and the potential pair \((V(\mathbf{x }),\psi (\mathbf{x }))\), respectively, with \(\mathbf{x }\in \partial \Omega \).

Thermoelectricity also takes this form, but one has to be careful in defining the fields to ensure that the associated tensor \(\mathbf{L }(\mathbf{x })\) is symmetric (see, e.g., Sect. 2.4 in [26]).

Fields that are coupled together need not have the same tensorial rank, an example being the equations of piezoelectricity,

$$\begin{aligned} \underbrace{\begin{pmatrix}{\varvec{\epsilon }}\\ \mathbf{d }\end{pmatrix}}_{\mathbf{J }(\mathbf{x })}=\underbrace{\begin{pmatrix}{\varvec{\mathcal {S}}}&{}\quad {\varvec{\mathcal {D}}}\\ {\varvec{\mathcal {D}}}^T &{}\quad {\varvec{\varepsilon }}\end{pmatrix}}_{\mathbf{L }(\mathbf{x })} \underbrace{\begin{pmatrix}{\varvec{\sigma }}\\ \mathbf{e }\end{pmatrix}}_{\mathbf{E }(\mathbf{x })}, \end{aligned}$$
(10.7)

where \({\varvec{\mathcal {S}}}(\mathbf{x })\) is the compliance tensor under short-circuit boundary conditions (i.e., with \(\mathbf{e }=0\)), \({\varvec{\mathcal {D}}}(\mathbf{x })\) is the piezoelectric stress coupling tensor, and \({\varvec{\varepsilon }}(\mathbf{x })\) is the free-body dielectric tensor (i.e., with \({\varvec{\tau }}=0\)). The strain field \({\varvec{\epsilon }}\), electric displacement field \(\mathbf{d }\), stress field \({\varvec{\sigma }}\), and electric field \(\mathbf{e }\) satisfy the usual differential constraints:

$$\begin{aligned} {\varvec{\epsilon }}= [\nabla \mathbf{u }+(\nabla \mathbf{u })^T]/2,\quad \nabla \cdot \mathbf{d }= 0,\quad \nabla \cdot {\varvec{\sigma }}= 0,\quad \mathbf{e }= -\nabla V. \end{aligned}$$
(10.8)

Since the stresses and strains are symmetric matrices, \({\varvec{\mathcal {D}}}\) is a third-order tensor that maps vectors to symmetric matrices. The boundary fields \(\partial \mathbf{J }(\mathbf{x })\) and \(\partial \mathbf{E }(\mathbf{x })\) are the displacement, flux pair \((\mathbf{u }(\mathbf{x }),\mathbf{n }\cdot \mathbf{d }(\mathbf{x }))\) and the traction, voltage pair \((\mathbf{n }\cdot {\varvec{\sigma }},V)\), respectively.

Of course, more than two fields can be coupled together. Thus, by combining a piezoelectric material and a magnetostrictive material in a composite, we can obtain a material where there is coupling between electric fields, elastic fields, and magnetic fields.

For fields varying in time at constant frequency \(\omega \), with wavelengths and attenuation lengths much bigger than the size of the body under consideration, the quasistatic equations are applicable. For dielectrics these take the same form as (10.4):

$$\begin{aligned} \mathbf{d }(\mathbf{x })={\varvec{\varepsilon }}(\mathbf{x })\mathbf{e }(\mathbf{x }),\quad \nabla \cdot \mathbf{j }=0, \quad \mathbf{e }=-\nabla V=0, \end{aligned}$$
(10.9)

where everything is now complex-valued: \(\mathbf{d }(\mathbf{x })\) and \(\mathbf{e }(\mathbf{x })\) are the complex-valued electrical displacement field and electric field, \(V(\mathbf{x })\) is the complex-valued electrical potential, and \({\varvec{\varepsilon }}(\mathbf{x })\) is the complex-valued electrical permittivity. (The physical displacement field, electric field, and potential are the real parts of \(\mathbf{d }(\mathbf{x })e^{-i\omega t}\), \(\mathbf{e }(\mathbf{x })e^{-i\omega t}\), and \(V e^{-i\omega t}\), respectively). Let us set

$$\begin{aligned} \mathbf{d }=\mathbf{d }'+i\mathbf{d }'',\quad \mathbf{e }=\mathbf{e }'+i\mathbf{e }'',\quad V=V'+iV'',\quad {\varvec{\varepsilon }}={\varvec{\varepsilon }}'+{\varvec{\varepsilon }}'', \end{aligned}$$
(10.10)

where the primed fields denote the real parts, while the doubled primed fields denote the imaginary parts. Physically, \({\varvec{\varepsilon }}''(\mathbf{x })\) is associated with electrical energy loss into heat and is positive semidefinite. Assuming it is positive definite and that an inverse \([{\varvec{\varepsilon }}'']^{-1}\) exists, substitution of (10.10) in (10.9), followed by suitable manipulation, gives the equivalent coupled field equations of Gibiansky and Cherkaev [7]:

$$\begin{aligned} \underbrace{\begin{pmatrix}\mathbf{e }'' \\ \mathbf{d }'' \\ \end{pmatrix}}_{\mathbf{J }(\mathbf{x })}= \underbrace{\begin{pmatrix}[{\varvec{\varepsilon }}'']^{-1} &{}\quad [{\varvec{\varepsilon }}'']^{-1} {\varvec{\varepsilon }}' \\ {\varvec{\varepsilon }}'[{\varvec{\varepsilon }}'']^{-1} &{}\quad {\varvec{\varepsilon }}'' + {\varvec{\varepsilon }}'[{\varvec{\varepsilon }}'']^{-1} {\varvec{\varepsilon }}' \end{pmatrix}}_{\mathbf{L }(\mathbf{x })}\underbrace{\begin{pmatrix}-\mathbf{d }' \\ \mathbf{e }' \\ \end{pmatrix}}_{\mathbf{E }(\mathbf{x })}, \end{aligned}$$
(10.11)

Clearly \(\mathbf{L }\) is real and symmetric and by inspection of the quadratic form associated with \(\mathbf{L }\) one sees that it is positive definite. Now \(\partial \mathbf{J }\) consists of the voltage, flux pair \((V'',\mathbf{n }\cdot \mathbf{d }'')\) while \(\partial \mathbf{E }\) consists of the flux, voltage pair \((-\mathbf{n }\cdot \mathbf{d }',V')\). As Gibiansky and Cherkaev show, similar manipulations can be done for viscoelasticity in the quasistatic limit where the equations have the form (10.5), but with all fields being complex. More generally, the Gibiansky–Cherkaev approach can be applied to equations where the tensor entering the constitutive law is not self-adjoint, but its self-adjoint part is positive definite, to an equivalent form where the tensor \(\mathbf{L }(\mathbf{x })\) entering the constitutive law is self-adjoint and positive definite [25] (see also Sect. 13.4 of [26]): such manipulations can be applied, for example, to electrical conduction in the presence of a magnetic field where, due to the Hall effect, the conductivity tensor \({\varvec{\sigma }}(\mathbf{x })\) entering (10.4) is not symmetric.

Wave equations, can be expressed in the form (4.12) with an identity like (4.15) holding. For example, at fixed frequency \(\omega \) with a \(\mathrm{{e}}^{-i\omega t}\) time dependence, as recognized in [32] the acoustic equations, with \(P(\mathbf{x })\) the (complex) pressure, \(\mathbf{v }(\mathbf{x })\) the (complex) velocity, \({\varvec{\rho }}(\mathbf{x },\omega )\) the effective mass density matrix, and \(\kappa (\mathbf{x },\omega )\) the bulk modulus, take the form

$$\begin{aligned} \underbrace{\begin{pmatrix}-i\mathbf{v }\\ -i\nabla \cdot \mathbf{v }\end{pmatrix}}_{\mathbf{J }(\mathbf{x })} =\underbrace{\begin{pmatrix}-(\omega {\varvec{\rho }})^{-1} &{}\quad 0 \\ 0 &{}\quad \omega /\kappa \end{pmatrix}}_{\mathbf{L }(\mathbf{x })}\underbrace{\begin{pmatrix}\nabla P \\ P\end{pmatrix}}_{\mathbf{E }(\mathbf{x })}, \end{aligned}$$
(10.12)

(and \(\partial \mathbf{E }\) and \(\partial \mathbf{J }\) can be identified with the boundary values of \(P(\mathbf{x })\) and \(\mathbf{n }\cdot \mathbf{v }(\mathbf{x })\) at \(\partial \Omega \), respectively). Here we allow for effective mass density matrices that, at a given frequency, can be anisotropic and complex-valued as may be the case in metamaterials [30, 33, 44, 50]. Maxwell’s equations, with \(\mathbf{e }(\mathbf{x })\) the (complex) electric field, \(\mathbf{h }(\mathbf{x })\) the (complex) magnetizing field, \({\varvec{\mu }}(\mathbf{x },\omega )\) the magnetic permeability, \({\varvec{\varepsilon }}(\mathbf{x })\) the electric permittivity, take the form

$$\begin{aligned} \underbrace{\begin{pmatrix}-i\mathbf{h }\\ i\nabla \times \mathbf{h }\end{pmatrix}}_{\mathbf{J }(\mathbf{x })} =\underbrace{\begin{pmatrix}-{[\omega {\varvec{\mu }}]}^{-1} &{}\quad 0 \\ 0 &{}\quad \omega {\varvec{\varepsilon }}\end{pmatrix}}_{\mathbf{L }(\mathbf{x })} \underbrace{\begin{pmatrix}\nabla \times \mathbf{e }\\ \mathbf{e }\end{pmatrix}}_{\mathbf{E }(\mathbf{x })}, \end{aligned}$$
(10.13)

(and \(\partial \mathbf{E }\) and \(\partial \mathbf{J }\) can be identified with the tangential values of \(\mathbf{e }(\mathbf{x })\) and \(\mathbf{h }(\mathbf{x })\) at \(\partial \Omega \), respectively). The linear elastodynamic equations, with \(\mathbf{u }(\mathbf{x })\) the (complex) displacement, \({\varvec{\sigma }}(\mathbf{x })\) the (complex) stress, \({\varvec{\mathcal {C}}}(\mathbf{x },\omega )\) the elasticity tensor, \({\varvec{\rho }}(\mathbf{x },\omega )\) the effective mass density matrix, take the form

$$\begin{aligned} \underbrace{\begin{pmatrix} -{\varvec{\sigma }}/\omega \\ -\nabla \cdot {\varvec{\sigma }}/\omega \end{pmatrix}}_{\mathbf{J }(\mathbf{x })} =\underbrace{\begin{pmatrix}-{\varvec{\mathcal {C}}}/\omega &{}\quad 0 \\ 0 &{}\quad \omega {\varvec{\rho }}\end{pmatrix}}_{\mathbf{L }(\mathbf{x })}\underbrace{\begin{pmatrix}/2 \\ \mathbf{u }\end{pmatrix}}_{\mathbf{E }(\mathbf{x })}, \end{aligned}$$
(10.14)

(and \(\partial \mathbf{E }\) and \(\partial \mathbf{J }\) can be identified with the values of \(\mathbf{u }(\mathbf{x })\) and the traction \(\mathbf{n }\cdot {\varvec{\sigma }}(\mathbf{x })\) at \(\partial \Omega \), respectively). The preceeding three equations have been written in this form so \({\mathrm{Im}}\mathbf{L }(\mathbf{x })\ge 0\) when \({\mathrm{Im}}\omega \ge 0\), where complex frequencies have the physical meaning of the solution increasing exponentially in time. Under assumptions that the material moduli are lossy, or that the frequency \(\omega \) is complex with positive imaginary part, one can easily manipulate them into equivalent forms similar to the Gibiansky–Cherkaev form in (10.11) with a positive semidefinite tensor entering the constitutive law [32, 34]. Of course, the boundary fields \(\partial \mathbf{E }\) and \(\partial \mathbf{J }\) then need to be appropriately redefined.

For thin plates, the dynamic plate equations at constant frequency can be written in the form

$$\begin{aligned} \underbrace{\begin{pmatrix} i\mathbf{M }\\ \nabla \cdot (\nabla \cdot \mathbf{M }) \end{pmatrix}}_{\mathbf{J }(\mathbf{x })} = \underbrace{\begin{pmatrix} -{\varvec{\mathcal {D}}}(\mathbf{x })/\omega &{}\quad 0 \\ 0 &{}\quad h(\mathbf{x })\omega \rho (\mathbf{x }) \end{pmatrix}}_{\mathbf{L }(\mathbf{x })} \underbrace{\begin{pmatrix} \nabla \nabla v \\ i v \end{pmatrix}}_{\mathbf{E }(\mathbf{x })}. \end{aligned}$$
(10.15)

Here \(\mathbf{M }(\mathbf{x },t)\) is the (complex) bending moment tensor, \({\varvec{\mathcal {D}}}(\mathbf{x })\) is the fourth-order tensor of plate rigidity coefficients, \(h(\mathbf{x })\) is the plate thickness, \(\rho (\mathbf{x })\) is the density, and \(v=\partial w/\partial t\) is the velocity of the (complex) vertical deflection \(w(\mathbf{x },t)\) of the plate. Note that the matrix \(\mathbf{L }(\mathbf{x })\) has positive definite imaginary part when \(\omega \) has positive imaginary part. \(\partial \mathbf{E }\) can be identified with the boundary values of the pair \((\nabla v,v)\) while \(\partial \mathbf{J }\) can be identified with the boundary values of the pair \((\mathbf{M }\mathbf{n },(\nabla \cdot \mathbf{M })\cdot \mathbf{n })\), in which \(\mathbf{n }\) is the outward normal to \(\partial \Omega \). Again, when the material moduli are lossy, or the frequency \(\omega \) is complex with positive imaginary part, this can be manipulated into the Gibiansky–Cherkaev form in (10.11) with a positive semidefinite tensor entering the constitutive law, and with appropriately redefined boundary fields.

Further examples of wave equations at constant frequency that can be represented in the required form are given in the appendix of [29].

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Milton, G.W., Onofrei, D. Exact relations for Green’s functions in linear PDE and boundary field equalities: a generalization of conservation laws. Res Math Sci 6, 19 (2019). https://doi.org/10.1007/s40687-019-0179-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s40687-019-0179-z

Keywords

Mathematics Subject Classification 2000

Navigation