Skip to main content
Log in

Nonlinear and Linear Elastodynamic Transformation Cloaking

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

Abstract

In this paper we formulate the problems of nonlinear and linear elastodynamic transformation cloaking in a geometric framework. In particular, it is noted that a cloaking transformation is neither a spatial nor a referential change of frame (coordinates); a cloaking transformation maps the boundary-value problem of an isotropic and homogeneous elastic body (virtual problem) to that of an anisotropic and inhomogeneous elastic body with a hole surrounded by a cloak that is to be designed (physical problem). The virtual body has a desired mechanical response while the physical body is designed to mimic the same response outside the cloak using a cloaking transformation. We show that nonlinear elastodynamic transformation cloaking is not possible while nonlinear elastostatic transformation cloaking may be possible for special deformations, e.g., radial deformations in a body with either a cylindrical or a spherical cavity. In the case of classical linear elastodynamics, in agreement with the previous observations in the literature, we show that the elastic constants in the cloak are not fully symmetric; they do not possess the minor symmetries. We prove that elastodynamic transformation cloaking is not possible regardless of the shape of the hole and the cloak. It is shown that the small-on-large theory, i.e., linearized elasticity with respect to a pre-stressed configuration, does not allow for transformation cloaking either. However, elastodynamic cloaking of a cylindrical hole is possible for in-plane deformations while it is not possible for anti-plane deformations. We next show that for a cavity of any shape elastodynamic transformation cloaking cannot be achieved for linear gradient elastic solids; similar to classical linear elasticity the balance of angular momentum is the obstruction to transformation cloaking. We finally prove that transformation cloaking is not possible for linear elastic generalized Cosserat solids in dimension two for any shape of the hole and the cloak. In particular, in dimension two transformation cloaking cannot be achieved in linear Cosserat elasticity. We show that transformation cloaking for a spherical cavity covered by a spherical cloak is not possible in the setting of linear elastic generalized Cosserat elasticity. We conjecture that this result is true for a cavity of any shape. It should be emphasized that in this paper we do not consider the so-called metamaterials [70, 72].

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
Fig. 4
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

Notes

  1. One should note the frequency limitation in the existing electromagnetic cloaking works; the existing works have been limited to microwaves frequencies.

  2. We should mention that there are recent geometric developments using non-Euclidean reference configurations that allow for sources of residual stress [23, 33,34,35,36, 86, 99, 101, 106, 107, 116,117,118,119, 123].

  3. There is a typo in their transformed mass density. The second term does not have the correct physical dimension. The similar expression in [4] has the correct dimension with a factor \(1/\omega ^2\) but has the opposite sign.

  4. Surprisingly, in none of the works that accept non-symmetric Cauchy stresses in the cloak is there any mention of the balance of angular momentum and the distribution of couple stresses. There is also no discussion on what the extra elastic constants of the Cosserat cloak should be.

  5. Note that \(\mathbf {C}^\flat \) agrees with the pull-back of the ambient space metric by \(\varphi _t\), i.e., \(\mathbf {C}^\flat =\varphi _t^*\mathbf {g}.\)

  6. This definition is not to be confused with a Lodge transformation.

  7. Note that if a connection \(\nabla \) is \(\mathbf {G}\)-compatible, then , where \(D_t\) is the covariant time derivative.

  8. See [109] for a detailed discussion on integrity basis for a finite set of tensors.

  9. This is also known as the principle of material objectivity (see, e.g., [77]).

  10. The response of a simple material at any material point depends only on the first deformation gradient (and its evolution) at that point [77].

  11. Here we assume that \(\mathscr {R}\) is the energy function. Response function may be any measure of stress as well.

  12. Note that Orth\((\mathbf {G})=\left\{ \mathbf {Q}: T_X\mathcal {B}\rightarrow T_X\mathcal {B}\quad |\quad \mathbf {Q}^\top =\mathbf {Q}^{-1}\right\} \). We use the notation \(\mathscr {G}\leqslant \mathscr {H}\) when \(\mathscr {G}\) is a subgroup of \(\mathscr {H}\).

  13. The normalizer group \(\mathcal {N}_{\mathscr {G}}(\mathcal {Q})\) of a subgroup \(\mathcal {Q}\) of \(\mathscr {G}\) (\(\mathcal {Q}\leqslant \mathscr {G}\)) is defined as \(\mathcal {N}_{\mathscr {G}}(\mathcal {Q})=\{g_i\in \mathscr {G}:g_i\mathcal {Q}g_i^{-1}=\mathcal {Q}\}\).

  14. Note that such a collection forms a basis for the space of tensors that are invariant under the action of \(\mathcal {G}\).

  15. Note that \(\left<\mathbf {Q}\right>_{m}\left( \mathbf {v}_1\otimes \dots \otimes \mathbf {v}_m\right) =\mathbf {Q}\mathbf {v}_1\otimes \dots \otimes \mathbf {Q}\mathbf {v}_m\), where \(\mathbf {v}_i\in T_X\mathcal {B}\), \(i=1,\dots ,m\), are arbitrary vectors.

  16. For a flat ambient space this is identical to the corresponding equation in [100]. However, note that even for a flat ambient space this is not identical to what Marsden and Hughes [68] obtained; they do not have the term \({\text {div}}_{\mathbf {g}}\left( \nabla ^{\mathbf {g}}\mathbf {u}\cdot \mathring{\varvec{\sigma }}\right) \).

  17. Note that (3.75) and (3.76) imply that two elasticity tensors are equivalent (represent the same type of material anisotropy) if and only if there exists an orthogonal transformation such that under its action the two elasticity tensors (or their symmetry groups) coincide.

  18. Also, see [82] for calculation of the symmetry classes of an even-order tensor space.

  19. The isomorphism is trivially given by the conjugacy relations as follows. Let and let \(\mathbf {H}\in \mathrm {Sym}_{\mathbf {G}}(\mathring{\varvec{\mathsf {C}}})\). Then, . It is straightforward to see that \(\phi (\mathbf {H}_1\mathbf {H}_2)=\phi (\mathbf {H}_1)\phi (\mathbf {H}_2)\) for \(\mathbf {H}_1, \mathbf {H}_2\in \mathrm {Sym}_{\mathbf {G}}(\mathring{\varvec{\mathsf {C}}})\). Also, it is straightforward to see that \(\phi \) is one-to-one, and thus, an isomorphism.

  20. Note that the Piola identity can be written as

    figure a

    Thus

    $$\begin{aligned} \left[ J_{\varXi }^{-1}\mathop {F}\limits ^{\varXi }{}{}^{\tilde{A}}{}_A P^{aA}\right] _{|\tilde{A}} =\left[ J_{\varXi }^{-1}\mathop {F}\limits ^{\varXi }{}{}^{\tilde{A}}{}_A\right] _{|\tilde{A}}P^{aA} +J_{\varXi }^{-1}\mathop {F}\limits ^{\varXi }{}{}^{\tilde{A}}{}_A {P^{aA}}_{|\tilde{A}}\,. \end{aligned}$$

    Therefore, using (4.7), one can write

    $$\begin{aligned} J_{\varXi }\left[ J_{\varXi }^{-1}\mathop {F}\limits ^{\varXi }{}{}^{\tilde{A}}{}_AP^{aA}\right] _{|\tilde{A}} =\mathop {F}\limits ^{\varXi }{}{}^{\tilde{A}}{}_A P^{aA}{}_{|\tilde{A}}=P^{aA}{}_{|A}\,. \end{aligned}$$
  21. Note that we do not know a priori how many structural tensors are needed and what they are.

  22. The kinematic relationship in linear elastodynamic cloaking turns out to be \(\tilde{\mathbf {U}}\circ \varXi =\mathbf {U}\).

  23. In the case of elastodynamic cloaking the kinematic relation between the physical and virtual problems was the equality (up to a shift) of acceleration vectors. In elastostatics there is no such constraint and one has freedom in choosing a kinematic relation between the two problems. Note that (4.29) is just one choice and one may assume

    $$\begin{aligned} \frac{\tilde{r}(\tilde{R})}{\tilde{R}}=h\left( \frac{r(R)}{R}\right) , \end{aligned}$$

    for any positive and strictly increasing function h such that \(h(1)=1\). Note also that (4.29) is a nonlinear analogue of Olsson and Wall [83]’s kinematic assumption.

  24. Note that transformation of the elastic constants under a cloaking map is different from that under a material change of coordinates (3.72).

  25. When Cartesian coordinates are used in both the physical and virtual bodies one writes \(\tilde{\mathsf {C}}^{\tilde{A}\tilde{B}\tilde{C}\tilde{D}} =J_{\varXi }^{-1}\mathop {F}\limits ^{\varXi }{}{}^{\tilde{A}}{}_A~\!\mathop {F}\limits ^{\varXi }{}{}^{\tilde{C}}{}_C~\!\delta ^{\tilde{B}}_B\delta ^{\tilde{D}}_D ~\!\mathsf {C}^{ABCD}\). This is identical to Norris and Parnell [79]’s Eq.(2.6). See also Al-Attar and Crawford [1]’s Eq.(129).

  26. Note that in the Voigt notation one has the following bijection between indices: \(\{11,22,33,23,13,12\}\rightarrow \{1,2,3,4,5,6\}\).

  27. Note that the physical components of the elasticity tensor \(\hat{C}^{ABCD}\) are related to the components of the elasticity tensor as \(\hat{C}^{ABCD}=\sqrt{G_{AA}}\sqrt{G_{BB}}\sqrt{G_{CC}}\sqrt{G_{DD}}C^{ABCD}\) (no summation) [114].

  28. Using (3.16), one obtains

    $$\begin{aligned} \left( \mathsf {A}^{aA}{}_z{}^C W^z{}_{|C}\right) _{|A}=\left( \mathsf {C}^{ABCN}\mathring{F}^a{}_B\mathring{F}^n{}_Ng_{zn}W^z{}_{|C}\right) _{|A}=\left( \mathsf {C}^{ABCZ}\delta ^a{}_B\delta ^z{}_Zg_{zz}W^z{}_{|C}\right) _{|A}\,. \end{aligned}$$

    Note that

    $$\begin{aligned} \begin{aligned} \left( \mathsf {C}^{AB}{}^C{}^Z W{}_{|C}\right) _{|A}&=\left( \mathsf {C}^{RB}{}^R{}^Z W{}_{|R}+\mathsf {C}^{RB}{}^\varTheta {}^Z W{}_{|\varTheta }\right) _{|R} \\&\quad +\left( \mathsf {C}^{\varTheta B}{}^R{}^Z W{}_{|R}+\mathsf {C}^{\varTheta B}{}^\varTheta {}^Z W{}_{|\varTheta }\right) _{|\varTheta }\\ {}&=\left( \mathsf {C}^{RZ}{}^R{}^Z W{}_{|R}\right) _{|R}+\left( \mathsf {C}^{\varTheta Z}{}^\varTheta {}^Z W{}_{|\varTheta }\right) _{|\varTheta }\,. \end{aligned} \end{aligned}$$
  29. This map does not satisfy the required traction continuity condition, i.e., \(f'(R_o)=1\), but nevertheless has been extensively used in the literature (see Section 5.2).

  30. Knowing that \(\mathsf {F}_{ij}=\mathsf {F}_{ji}\), from (4.88) and (4.89), one obtains

    figure b
  31. The inner boundary of the virtual and physical holes must be traction-free. This condition for the virtual hole reads \(\delta \tilde{{\mathbf {t}}}=\delta \tilde{P}^{{\tilde{a}}\tilde{A}}\tilde{N}_{{\tilde{A}}}|_{\partial \tilde{\mathcal {H}}}=0\). However, note that \(\varvec{\mathsf {s}}\,\delta \mathbf {t}\,dA=\delta \tilde{\mathbf t}\,d\tilde{A}=\mathbf {0}\), i.e., the linearized traction in the physical body vanishes, and thus, in order for the hole to be traction-free in the physical body, the initial traction must vanish, i.e., \(\mathring{P}^{aA}N_A|_{\partial {\mathcal {H}}}=0\).

  32. We use the standard notation \(\varPi ^{[ab]}=\frac{1}{2}\left( \varPi ^{ab}-\varPi ^{ba}\right) \).

  33. Note that only the virtual body is assumed to be centro-symmetric. There is no such constraint on the physical body; it can be both non-centro-symmetric and anisotropic.

  34. Note that

    $$\begin{aligned} \begin{aligned} \delta W&=\frac{1}{2}\frac{\partial W}{\partial {F^a}_A\partial F^b{}_B}U^a{}_{|A}U^b{}_{|B} +\frac{\partial W}{\partial F^a{}_A\partial F^b{}_{B|C}}U^a{}_{|A}U^b{}_{|B|C} \\&\quad +\frac{1}{2}\frac{\partial W}{\partial F^a{}_{A|B}\partial F^b{}_{C|D}}U^a{}_{|A|B}U^b{}_{|C|D} \\&= \frac{1}{2}{\mathbb {A}}^{aAbB}U_{a|A}U_{b|B} +{\mathbb {B}}^{aAbBC}U_{a|A}U_{b|B|C} +\frac{1}{2}{\mathbb {C}}^{aABbCD}U_{a|A|B}U_{b|C|D}. \end{aligned} \end{aligned}$$

    Positive-definiteness of energy requires that \(\delta W>0\) for any pair \((U_{a|A},U_{a|A|B})\ne (0,0)\). In particular, when \(U_{a|A}\ne 0\), and \(U_{a|A|B}=0\), \({\mathbb {A}}^{aAbB}U_{a|A}U_{b|B}>0\), which implies that \(\varvec{{\mathbb {A}}}\) must be positive-definite. In the case of isotropic solids this is equivalent to \(\mu >0\), and \(3\lambda +2\mu >0\).

  35. Note that the energy function can have an explicit dependence on the director field, i.e., \(W=W(X,\mathbf {F},\mathop {\varvec{\mathsf {d}}}\limits _{\mathfrak {a}}{},\mathop {\varvec{\mathsf {F}}}\limits _{\mathfrak {a}}{},\mathbf {G},\mathbf {g})\). The partial derivative \(\frac{\partial W}{\partial \mathop {\varvec{\mathsf {d}}}\limits _{\mathfrak {a}}{}}\) is a micro body force that we do not consider in this paper.

  36. More specifically, \(\{11,12,\cdots ,1n,21,22,\cdots ,nn\}\rightarrow \{1,2,\cdots ,n^2\}\) and .

  37. This condition has been ignored in the existing works on elastodynamic transformation cloaking. In particular, borrowing the cloaking transformation of Pendry et al. [89] from electromagnetism is not acceptable as it does not satisfy this condition.

  38. Note that the physical components of the elasticity tensor \(\hat{\mathbb A}^{aAbB}\) are related to the components of the elasticity tensor as \(\hat{\mathbb A}^{aAbB}=\sqrt{g_{aa}}\sqrt{G_{AA}}\sqrt{g_{bb}}\sqrt{G_{BB}}{\mathbb A}^{aAbB}\) (no summation).

References

  1. Al-Attar, D., Crawford, O.: Particle relabelling transformations in elastodynamics. Geophys. J. Int. 205(1), 575–593, 2016

    Article  ADS  Google Scholar 

  2. Amirkhizi, A.V., Tehranian, A., Nemat-Nasser, S.: Stress-wave energy management through material anisotropy. Wave Motion 47(8), 519–536, 2010

    Article  MATH  Google Scholar 

  3. Auffray, N., Le Quang, H., He, Q.-C.: Matrix representations for 3D strain-gradient elasticity. J. Mech. Phys. Solids 61(5), 1202–1223, 2013

    Article  ADS  MathSciNet  MATH  Google Scholar 

  4. Banerjee, B.: An Introduction to Metamaterials and Waves in Composites. CRC Press, Boca Raton 2011

    Book  Google Scholar 

  5. Benveniste, Y., Milton, G.W.: New exact results for the effective electric, elastic, piezoelectric and other properties of composite ellipsoid assemblages. J. Mech. Phys. Solids 51(10), 1773–1813, 2003

    Article  ADS  MathSciNet  MATH  Google Scholar 

  6. Boehler, J.-P.: A simple derivation of representations for non-polynomial constitutive equations in some cases of anisotropy. Zeitschrift für Angewandte Mathematik und Mechanik (ZAMM) 59(4), 157–167, 1979

    Article  ADS  MathSciNet  MATH  Google Scholar 

  7. Boehler, J.-P.: Applications of Tensor Functions in Solid Mechanics, vol. 292. Springer, New York 1987

    Book  MATH  Google Scholar 

  8. Brun, M., Guenneau, S., Movchan, A.B.: Achieving control of in-plane elastic waves. Appl. Phys. Lett. 94(6), 061903, 2009

    Article  ADS  Google Scholar 

  9. Bryan , K., Leise , T.: Impedance imaging, inverse problems, and Harry Potter’s cloak. SIAM Rev. 52(2), 359–377, 2010

    Article  MathSciNet  MATH  Google Scholar 

  10. Capriz , G.: Continua with Microstructure, vol. 35. Springer, New York 2013

    MATH  Google Scholar 

  11. Chadwick , P., Vianello , M., Cowin , S.C.: A new proof that the number of linear elastic symmetries is eight. J. Mech. Phys. Solids 49(11), 2471–2492, 2001

    Article  ADS  MathSciNet  MATH  Google Scholar 

  12. Chang , Z., Guo , H.-Y., Li , B., Feng , X.-Q.: Disentangling longitudinal and shear elastic waves by neo-Hookean soft devices. Appl. Phys. Lett. 106(16), 161903, 2015

    Article  ADS  Google Scholar 

  13. Chen , H., Chan , C.T.: Acoustic cloaking in three dimensions using acoustic metamaterials. Appl. Phys. Lett. 91(18), 183518, 2007

    Article  ADS  Google Scholar 

  14. Coleman , B.D., Noll , W.: On the thermostatics of continuous media. Arch. Ration. Mech. Anal. 4(1), 97–128, 1959

    Article  MathSciNet  MATH  Google Scholar 

  15. Coleman , B.D., Noll , W.: The thermodynamics of elastic materials with heat conduction and viscosity. Arch. Ration. Mech. Anal. 13(1), 167–178, 1963

    Article  MathSciNet  MATH  Google Scholar 

  16. Coleman , B.D., Noll , W.: Material symmetry and thermostatic inequalities in finite elastic deformations. Arch. Ration. Mech. Anal. 15(2), 87–111, 1964

    Article  MathSciNet  MATH  Google Scholar 

  17. Cosserat, Eugène: François Cosserat. Théorie des corps déformables, 1909

  18. Cummer , S.A., Popa , B.-I., Schurig , D., Smith , D.R., Pendry , J., Rahm , M., Starr , A.: Scattering theory derivation of a 3D acoustic cloaking shell. Phys. Rev. Lett. 100(2), 024301, 2008a

    Article  ADS  Google Scholar 

  19. Cummer , S.A., Rahm , M., Schurig , D.: Material parameters and vector scaling in transformation acoustics. N. J. Phys. 10(11), 115025, 2008b

    Article  Google Scholar 

  20. De Klerk , E.: Aspects of Semidefinite Programming: Interior Point Algorithms and Selected Applications, vol. 65. Springer, New York 2006

    MATH  Google Scholar 

  21. DiVincenzo, D.P.: Dispersive corrections to continuum elastic theory in cubic crystals. Phys. Rev. B 34(8), 5450, 1986

    Article  ADS  Google Scholar 

  22. Doyle, T.C., Ericksen, J.L.: Nonlinear elasticity. Adv. Appl. Mech. 4, 53–115, 1956

    Article  MathSciNet  Google Scholar 

  23. Efrati, E., Sharon, E., Kupferman, R.: The metric description of elasticity in residually stressed soft materials. Soft Matter 9(34), 8187–8197, 2013

    Article  ADS  Google Scholar 

  24. Ehret, A.E., Itskov, M.: Modeling of anisotropic softening phenomena: application to soft biological tissues. Int. J. Plast. 25(5), 901–919, 2009

    Article  MATH  Google Scholar 

  25. Ergin, T., Stenger, N., Brenner, P., Pendry, J.B., Wegener, M.: Three-dimensional invisibility cloak at optical wavelengths. Science 328(5976), 337–339, 2010

    Article  ADS  Google Scholar 

  26. Ericksen, J.L.: Deformations possible in every compressible, isotropic, perfectly elastic material. Stud. Appl. Math. 34(1–4), 126–128, 1955

    MathSciNet  MATH  Google Scholar 

  27. Ericksen, J.L., Truesdell, C.: Exact theory of stress and strain in rods and shells. Arch. Ration. Mech. Anal. 1(1), 295–323, 1957

    Article  MathSciNet  MATH  Google Scholar 

  28. Eringen, A.C.: Microcontinuum Field Theories: I. Foundations and Solids. Springer, New York 2012

    MATH  Google Scholar 

  29. Farhat, M., Enoch, S., Guenneau, S., Movchan, A.B.: Broadband cylindrical acoustic cloak for linear surface waves in a fluid. Phys. Rev. Lett. 101(13), 134501, 2008

    Article  ADS  Google Scholar 

  30. Farhat, M., Guenneau, S., Enoch, S.: Ultrabroadband elastic cloaking in thin plates. Phys. Rev. Lett. 103(2), 024301, 2009a

    Article  ADS  Google Scholar 

  31. Farhat, M., Guenneau, S., Enoch, S., Movchan, A.B.: Cloaking bending waves propagating in thin elastic plates. Phys. Rev. B 79(3), 033102, 2009b

    Article  ADS  Google Scholar 

  32. Forte, S., Vianello, M.: Symmetry classes for elasticity tensors. J. Elast. 43(2), 81–108, 1996

    Article  MathSciNet  MATH  Google Scholar 

  33. Golgoon, A., Yavari, A.: On the stress field of a nonlinear elastic solid torus with a toroidal inclusion. J. Elast. 128(1), 115–145, 2017

    Article  MathSciNet  MATH  Google Scholar 

  34. Golgoon, A., Yavari, A.: Line and point defects in nonlinear anisotropic solids. Zeitschrift für angewandte Mathematik und Physik 69(3), 81, 2018a

    Article  ADS  MathSciNet  MATH  Google Scholar 

  35. Golgoon, A., Yavari, A.: Nonlinear elastic inclusions in anisotropic solids. J. Elast. 130(2), 239–269, 2018b

    Article  MathSciNet  MATH  Google Scholar 

  36. Golgoon, A., Sadik, S., Yavari, A.: Circumferentially-symmetric finite eigenstrains in incompressible isotropic nonlinear elastic wedges. Int. J. Non-Linear Mech. 84, 116–129, 2016

    Article  ADS  Google Scholar 

  37. Goriely, A.: The Mathematics and Mechanics of Biological Growth, vol. 45. Springer, New York 2017

    Book  MATH  Google Scholar 

  38. Green, A.E., Rivlin, R.S., Shield, R.T.: General theory of small elastic deformations superposed on finite elastic deformations. Proc. R. Soc. Lond. A 211(1104), 128–154, 1952

    Article  ADS  MathSciNet  MATH  Google Scholar 

  39. Green, A.E., Rivlin, R.S.: On Cauchy’s equations of motion. Zeitschrift für Angewandte Mathematik und Physik (ZAMP) 15(3), 290–292, 1964

    Article  ADS  MathSciNet  MATH  Google Scholar 

  40. Greenleaf, A., Lassas, M., Uhlmann, G.: On nonuniqueness for Calderón’s inverse problem. Math. Res. Lett. 10, 2003a

  41. Greenleaf, A., Lassas, M., Uhlmann, G.: Anisotropic conductivities that cannot be detected by EIT. Physiol. Meas. 24(2), 413, 2003b

    Article  Google Scholar 

  42. Greenleaf, A., Kurylev, Y., Lassas, M., Uhlmann, G.: Full-wave invisibility of active devices at all frequencies. Commun. Math. Phys. 275(3), 749–789, 2007

    Article  ADS  MathSciNet  MATH  Google Scholar 

  43. Greenleaf, A., Kurylev, Y., Lassas, M., Uhlmann, G.: Cloaking devices, electromagnetic wormholes, and transformation optics. SIAM Rev. 51(1), 3–33, 2009a

    Article  ADS  MathSciNet  MATH  Google Scholar 

  44. Greenleaf, A., Kurylev, Y., Lassas, M., Uhlmann, G.: Invisibility and inverse problems. Bull. Am. Math. Soc. 46(1), 55–97, 2009b

    Article  MathSciNet  MATH  Google Scholar 

  45. Guenneau, S., Puvirajesinghe, T.M.: Fick’s second law transformed: one path to cloaking in mass diffusion. J. R. Soc. Interface 10(83), 20130106, 2013

    Article  Google Scholar 

  46. Guenneau, S., Amra, C., Veynante, D.: Transformation thermodynamics: cloaking and concentrating heat flux. Opt. Express 20(7), 8207–8218, 2012

    Article  ADS  Google Scholar 

  47. Gurney, C.: An Analysis of the Stresses in a Flat Plate with a Reinforced Circular Hole Under Edge Forces. Reports and Memoranda. H.M, Stationery Office, 1938

  48. Hashin, Z.: The elastic moduli of heterogeneous materials. J. Appl. Mech. 29(1), 143–150, 1962

    Article  ADS  MathSciNet  MATH  Google Scholar 

  49. Hashin, Z.: Large isotropic elastic deformation of composites and porous media. Int. J. Solids Struct. 21(7), 711–720, 1985

    Article  MATH  Google Scholar 

  50. Hashin, Z., Rosen, B.W.: The elastic moduli of fiber-reinforced materials. J. Appl. Mech. 31(2), 223–232, 1964

    Article  ADS  Google Scholar 

  51. Hashin, Z., Shtrikman, S.: A variational approach to the theory of the elastic behaviour of multiphase materials. J. Mech. Phys. Solids 11(2), 127–140, 1963

    Article  ADS  MathSciNet  MATH  Google Scholar 

  52. Hughes, T.J.R., Marsden, J.E.: Some applications of geometry is continuum mechanics. Rep. Math. Phys. 12(1), 35–44, 1977

    Article  ADS  MATH  Google Scholar 

  53. Kadic, M., Bückmann, T., Schittny, R., Wegener, M.: Metamaterials beyond electromagnetism. Rep. Prog. Phys. 76(12), 126501, 2013

    Article  ADS  Google Scholar 

  54. Kadic, M., Bückmann, T., Schittny, R., Wegener, M.: Experiments on cloaking in optics, thermodynamics and mechanics. Philos. Trans. A 373(2049), 20140357, 2015

    Article  ADS  Google Scholar 

  55. Khlopotin, A., Olsson, P., Larsson, F.: Transformational cloaking from seismic surface waves by micropolar metamaterials with finite couple stiffness. Wave Motion 58, 53–67, 2015

    Article  MathSciNet  MATH  Google Scholar 

  56. Kohn, R.V., Shen, H., Vogelius, M.S., Weinstein, M.I.: Cloaking via change of variables in electric impedance tomography. Inverse Probl. 24(1), 015016, 2008

    Article  ADS  MathSciNet  MATH  Google Scholar 

  57. Lang, H.A.: The affine transformation for orthotropic plane-stress and plane-strain problems. J. Appl. Mech. 23(1), 1–6, 1956

    MathSciNet  MATH  Google Scholar 

  58. Lee, J.M.: Riemannian Manifold An Introduction to Curves. Springer, New York 1997

    Book  Google Scholar 

  59. Leonhardt, U.: Optical conformal mapping. Science 312(5781), 1777–1780, 2006

    Article  ADS  MathSciNet  MATH  Google Scholar 

  60. Leonhardt, U., Philbin, T.: Geometry and Light: The Science of Invisibility. Courier Corporation, New York, 2012

  61. Liu, I.: On representations of anisotropic invariants. Int. J. Eng. Sci. 20(10), 1099–1109, 1982

    Article  MathSciNet  MATH  Google Scholar 

  62. Liu, R., Ji, C., Mock, J.J., Chin, J.Y., Cui, T.J., Smith, D.R.: Broadband ground-plane cloak. Science 323(5912), 366–369, 2009

    Article  ADS  Google Scholar 

  63. Lodge, A.S.: A new theorem in the classical theory of elasticity. Nature 169, 926–927, 1952

    Article  ADS  MathSciNet  MATH  Google Scholar 

  64. Lodge, A.S.: The transformation to isotropic form of the equilibrium equations for a class of anisotropic elastic solids. Quart. J. Mech. Appl. Math. 8(2), 211–225, 1955

    Article  MathSciNet  MATH  Google Scholar 

  65. Jia, L., Papadopoulos, P.: A covariant constitutive description of anisotropic non-linear elasticity. Zeitschrift für Angewandte Mathematik und Physik (ZAMP) 51(2), 204–217, 2000

    Article  ADS  MathSciNet  MATH  Google Scholar 

  66. Mansfield, E.H.: Neutral holes in plane sheet-reinforced holes which are elastically equivalent to the uncut sheet. Quart. J. Mech. Appl. Math. 6(3), 370, 1953

    Article  MathSciNet  MATH  Google Scholar 

  67. Maranganti, R., Sharma, P.: A novel atomistic approach to determine strain-gradient elasticity constants: tabulation and comparison for various metals, semiconductors, silica, polymers and their relevance for nanotechnologies. J. Mech. Phys. Solids 55(9), 1823–1852, 2007

    Article  ADS  MATH  Google Scholar 

  68. Marsden, J.E., Hughes, T.J.R.: Mathematical Foundations of Elasticity. Dover, New York 1983

    MATH  Google Scholar 

  69. Mazzucato, A.L., Rachele, L.V.: Partial uniqueness and obstruction to uniqueness in inverse problems for anisotropic elastic media. J. Elast. 83(3), 205–245, 2006

    Article  MathSciNet  MATH  Google Scholar 

  70. Milton, G.W.: New metamaterials with macroscopic behavior outside that of continuum elastodynamics. N. J. Phys. 9(10), 359, 2007

    Article  Google Scholar 

  71. Milton, G.W., Cherkaev, A.V.: Which elasticity tensors are realizable? J. Eng. Mater. Technol. 117(4), 483–493, 1995

    Article  Google Scholar 

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

    Article  ADS  MathSciNet  MATH  Google Scholar 

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

    Article  Google Scholar 

  74. Mindlin, R.D.: Micro-structure in linear elasticity. Arch. Ration Mech. Anal. 16(1), 51–78, 1964

    Article  MathSciNet  MATH  Google Scholar 

  75. Mindlin, R.D., Tiersten, H.F.: Effects of couple-stresses in linear elasticity. Arch. Ration Mech. Anal. 11(1), 415–448, 1962

    Article  MathSciNet  MATH  Google Scholar 

  76. Nishikawa, S.: Variational Problems in Geometry. American Mathematical Society, New York 2002

    Book  MATH  Google Scholar 

  77. Noll, W.: A mathematical theory of the mechanical behavior of continuous media. Arch. Ration Mech. Anal. 2(1), 197–226, 1958

    Article  MathSciNet  MATH  Google Scholar 

  78. Norris, A.N.: Acoustic cloaking theory. Proc. R. Soc. Lond. A 464(2097), 2411–2434, 2008

    Article  ADS  MathSciNet  MATH  Google Scholar 

  79. Norris, A.N., Parnell, W.J.: Hyperelastic cloaking theory: transformation elasticity with pre-stressed solids. Proc. R. Soc. A 468(2146), 2881–2903, 2012

    Article  ADS  MathSciNet  MATH  Google Scholar 

  80. Norris, A.N., Shuvalov, A.L.: Elastic cloaking theory. Wave Motion 48(6), 525–538, 2011

    Article  MathSciNet  MATH  Google Scholar 

  81. Ogden, R.W.: Non-linear Elastic Deformations. Courier Corporation, New York 1997

    Google Scholar 

  82. Olive, M., Auffray, N.: Symmetry classes for even-order tensors. Math. Mech. Complex Syst. 1(2), 177–210, 2013

    Article  MATH  Google Scholar 

  83. Olsson, P., Wall, D.J.N.: Partial elastodynamic cloaking by means of fiber-reinforced composites. Inverse Probl. 27(4), 045010, 2011

    Article  ADS  MathSciNet  MATH  Google Scholar 

  84. Olver, P.J.: Canonical elastic moduli. J. Elast. 19(3), 189–212, 1988

    Article  MathSciNet  MATH  Google Scholar 

  85. Ostrosablin, N.I.: Affine transformations of the equations of the linear theory of elasticity. J. Appl. Mech. Tech. Phys. 47(4), 564–572, 2006

    Article  ADS  MathSciNet  Google Scholar 

  86. Ozakin, A., Yavari, A.: A geometric theory of thermal stresses. J. Math. Phys. 51, 032902, 2010

    Article  ADS  MathSciNet  MATH  Google Scholar 

  87. Parnell, W.J.: Nonlinear pre-stress for cloaking from antiplane elastic waves. Proc. R. Soc. A 468(2138), 563–580, 2012

    Article  ADS  MathSciNet  MATH  Google Scholar 

  88. Parnell, W.J., Norris, A.N., Shearer, T.: Employing pre-stress to generate finite cloaks for antiplane elastic waves. Appl. Phys. Lett. 100(17), 171907, 2012

    Article  ADS  Google Scholar 

  89. Pendry, J.B., Schurig, D., Smith, D.R.: Controlling electromagnetic fields. Science 312(5781), 1780–1782, 2006

    Article  ADS  MathSciNet  MATH  Google Scholar 

  90. Post, E.J.: Formal Structure of Electromagnetics: General Covariance and Electromagnetics. Courier Corporation, New York 1997

    MATH  Google Scholar 

  91. Reissner, H., Morduchow, M.: Reinforced circular cutouts in plane sheets. Technical Report Technical Note No. 1852, 1949

  92. Rivlin, R.S.: Large elastic deformations of isotropic materials. I. Fundamental concepts. Philos. Trans. R. Soc. Lond. Ser. A 240, 459–490, 1948

    Article  ADS  MathSciNet  MATH  Google Scholar 

  93. Rivlin, R.S.: Large elastic deformations of isotropic materials. II. Some uniqueness theorems for pure, homogeneous deformation. Philos. Trans. R. Soc. Lond. A 240(822), 491–508, 1948a

    Article  ADS  MathSciNet  Google Scholar 

  94. Rivlin, R.S.: Large elastic deformations of isotropic materials. III. Some simple problems in cylindrical polar co-ordinates. Philos. Trans. R. Soc. Lond. A 240(823), 509–525, 1948b

    Article  ADS  MATH  Google Scholar 

  95. Rivlin, R.S.: Large elastic deformations of isotropic materials. IV. Further developments of the general theory. Philos. Trans. R. Soc. Lond. A 241(835), 379–397, 1948c

    Article  ADS  MathSciNet  MATH  Google Scholar 

  96. Rivlin, R.S.: Large elastic deformations of isotropic materials. V. The problem of flexure. Philos. Trans. R. Soc. Lond. A 195(1043), 463–473, 1949a

    MathSciNet  MATH  Google Scholar 

  97. Rivlin, R.S.: Large elastic deformations of isotropic materials. VI. Further results in the theory of torsion, shear and flexure. Philos. Trans. R. Soc. Lond. A 242(845), 173–195, 1949b

    Article  ADS  MathSciNet  MATH  Google Scholar 

  98. Rivlin, R.S., Saunders, D.W.: Large elastic deformations of isotropic materials. VII. Experiments on the deformation of rubber. Philos. Trans. R. Soc. Lond. A 243(865), 251–288, 1951

    Article  ADS  MATH  Google Scholar 

  99. Sadik, S., Yavari, A.: Geometric nonlinear thermoelasticity and the time evolution of thermal stresses. Math. Mech. Solids 1, 42, 2015

    MATH  Google Scholar 

  100. Sadik, S., Yavari, A.: Small-on-large geometric anelasticity. Proc. R. Soc. A 472(2195), 20160659, 2016

    Article  ADS  MathSciNet  MATH  Google Scholar 

  101. Sadik, S., Angoshtari, A., Goriely, A., Yavari, A.: A geometric theory of nonlinear morphoelastic shells. J. Nonlinear Sci. 26, 929–978, 2016

    Article  ADS  MathSciNet  MATH  Google Scholar 

  102. Schurig, D., Mock, J.J., Justice, B.J., Cummer, S.A., Pendry, J.B., Starr, A.F., Smith, D.R.: Metamaterial electromagnetic cloak at microwave frequencies. Science 314(5801), 977–980, 2006

    Article  ADS  Google Scholar 

  103. Simo, J.C., Marsden, J.E.: Stress tensors, Riemannian metrics and the alternative descriptions in elasticity. In: Trends and Applications of Pure Mathematics to Mechanics, pp. 369–383. Springer, 1984

  104. Simo, J.C., Marsden, J.E., Krishnaprasad, P.S.: The Hamiltonian structure of nonlinear elasticity: the material and convective representations of solids, rods, and plates. Arch. Ration. Mech. Anal. 104(2), 125–183, 1988

    Article  MathSciNet  MATH  Google Scholar 

  105. Sklan, S.R., Pak, R.Y.S., Li, B.: Seismic invisibility: elastic wave cloaking via symmetrized transformation media. N. J. Phys. 20(6), 063013, 2018

    Article  Google Scholar 

  106. Sozio, F., Yavari, A.: Nonlinear mechanics of surface growth for cylindrical and spherical elastic bodies. J. Mech. Phys. Solids 98, 12–48, 2017

    Article  ADS  MathSciNet  Google Scholar 

  107. Sozio, F., Yavari, A.: Nonlinear mechanics of accretion. J. Nonlinear Sci. 1–51, 2019

  108. Spencer, A.J.M.: The Formulation of Constitutive Equation for Anisotropic Solids, pp. 3–26. Springer, Dordrecht 1982

    Google Scholar 

  109. Spencer, A.J.M.: Part III. Theory of invariants. Contin. Phys. 1, 239–353, 1971

    Google Scholar 

  110. Steigmann, D.J.: On the frame invariance of linear elasticity theory. Zeitschrift für Angewandte Mathematik und Physik 58(1), 121–136, 2007

    Article  ADS  MathSciNet  MATH  Google Scholar 

  111. Stojanović, R.: Recent Developments in the Theory of Polar Continua. Springer, New York 1970

    Book  MATH  Google Scholar 

  112. Toupin, R.A.: Elastic materials with couple-stresses. Arch. Ration. Mech. Anal. 11(1), 385–414, 1962

    Article  MathSciNet  MATH  Google Scholar 

  113. Toupin, R.A.: Theories of elasticity with couple-stress. Arch. Ration. Mech. Anal. 17(2), 85–112, 1964

    Article  MathSciNet  MATH  Google Scholar 

  114. Truesdell, C.: The physical components of vectors and tensors. Zeitschrift für Angewandte Mathematik und Mechanik 33(10–11), 345–356, 1953

    Article  ADS  MathSciNet  MATH  Google Scholar 

  115. Yano, K.L.: The Theory of Lie Derivatives and its Applications. North-Holland, Amsterdam, 1957

  116. Yavari, A.: A geometric theory of growth mechanics. J. Nonlinear Sci. 20(6), 781–830, 2010

    Article  ADS  MathSciNet  MATH  Google Scholar 

  117. Yavari, A., Goriely, A.: Riemann–Cartan geometry of nonlinear dislocation mechanics. Arch. Ration. Mech. Anal. 205(1), 59–118, 2012a

    Article  MathSciNet  MATH  Google Scholar 

  118. Yavari, A., Goriely, A.: Weyl geometry and the nonlinear mechanics of distributed point defects. Proc. R. Soc. A 468(2148), 3902–3922, 2012b

    Article  ADS  MathSciNet  MATH  Google Scholar 

  119. Yavari, A., Goriely, A.: The geometry of discombinations and its applications to semi-inverse problems in anelasticity. Proc. R. Soc. A 470(2169), 20140403, 2014

    Article  ADS  MathSciNet  MATH  Google Scholar 

  120. Yavari, A., Goriely, A.: The anelastic Ericksen problem: universal eigenstrains and deformations in compressible isotropic elastic solids. Proc. R. Soc. A 472(2196), 20160690, 2016

    Article  ADS  MathSciNet  MATH  Google Scholar 

  121. Yavari, A.: On geometric discretization of elasticity. J. Math. Phys. 49(2), 022901, 2008

    Article  ADS  MathSciNet  MATH  Google Scholar 

  122. Yavari, A.: Compatibility equations of nonlinear elasticity for non-simply-connected bodies. Arch. Ration. Mech. Anal. 209(1), 237–253, 2013

    Article  MathSciNet  MATH  Google Scholar 

  123. Yavari, A., Goriely, A.: Riemann–Cartan geometry of nonlinear disclination mechanics. Math. Mech. Solids 18(1), 91–102, 2012c

    Article  MathSciNet  MATH  Google Scholar 

  124. Yavari, A., Marsden, J.E.: Covariant balance laws in continua with microstructure. Rep. Math. Phys. 63(1), 1–42, 2009a

    Article  ADS  MathSciNet  MATH  Google Scholar 

  125. Yavari, A., Marsden, J.E.: Energy balance invariance for interacting particle systems. Zeitschrift für Angewandte Mathematik und Physik 60(4), 723–738, 2009b

    Article  ADS  MathSciNet  MATH  Google Scholar 

  126. Yavari, A., Ozakin, A.: Covariance in linearized elasticity. Zeitschrift für Angewandte Mathematik und Physik 59(6), 1081–1110, 2008

    Article  ADS  MathSciNet  MATH  Google Scholar 

  127. Yavari, A., Marsden, J.E., Ortiz, M.: On spatial and material covariant balance laws in elasticity. J. Math. Phys. 47(4), 042903, 2006

    Article  ADS  MathSciNet  MATH  Google Scholar 

  128. Zhang, J.M., Rychlewski, J.: Structural tensors for anisotropic solids. Arch. Mech. 42(3), 267–277, 1990

    MathSciNet  MATH  Google Scholar 

  129. Zhang, S., Genov, D.A., Sun, C., Zhang, X.: Cloaking of matter waves. Phys. Rev. Lett. 100(12), 123002, 2008

    Article  ADS  Google Scholar 

  130. Zheng, Q.S.: Theory of representations for tensor functions. Appl. Mech. Rev. 47(11), 545–587, 1994

    Article  ADS  Google Scholar 

  131. Zheng, Q.-S., Spencer, A.J.M.: Tensors which characterize anisotropies. Int. J. Eng. Sci. 31(5), 679–693, 1993a

    Article  MathSciNet  MATH  Google Scholar 

  132. Zheng, Q.-S., Spencer, A.J.M.: On the canonical representations for Kronecker powers of orthogonal tensors with application to material symmetry problems. Int. J. Eng. Sci. 31(4), 617–635, 1993b

    Article  MathSciNet  MATH  Google Scholar 

  133. Zhou, X., Gengkai, H., Tianjian, L.: Elastic wave transparency of a solid sphere coated with metamaterials. Phys. Rev. B 77(2), 024101, 2008

    Article  ADS  Google Scholar 

Download references

Acknowledgements

This research was supported by ARO W911NF-16-1-0064 and ARO W911NF-18-1-0003 (Dr. David Stepp).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Arash Yavari.

Additional information

Communicated by M. Ortiz.

Dedicated to Professor Zdeněk P. Bažant on the occasion of his 80th birthday.

Publisher's Note

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

Appendix A Riemannian Geometry

Appendix A Riemannian Geometry

To make the paper self-contained, in this appendix some basic concepts of Riemannian geometry are tersely reviewed. It should be emphasized that only those geometric concepts that have been used in the paper are discussed here.

For a smooth n-dimensional manifold \(\mathcal {B}\), the tangent space of \(\mathcal {B}\) at a point \(X \in \mathcal {B}\) is denoted by \(T_X\mathcal {B}\). Assume that \(\mathcal {S}\) is another n-dimensional manifold and \(\varphi :\mathcal {B}\rightarrow \mathcal {S}\) is a diffeomorphism (smooth and invertible map with a smooth inverse) between the two manifolds. A smooth vector field \(\mathbf {W}\) on \(\mathcal {B}\) assigns a vector \(\mathbf {W}_X\in T_X\mathcal {B}\) for every \(X\in \mathcal {B}\) such that the mapping \(X\mapsto \mathbf {W}_X\) is smooth. If \(\mathbf {W}\) is a vector field on \(\mathcal {B}\), then the push-forward of \(\mathbf {W}\) by \(\varphi \) is a vector field on \(\varphi (\mathcal {B})\) defined as \(\varphi _*\mathbf {W} = T\varphi \cdot \mathbf {W} \circ \varphi ^{-1}\). Similarly, if \(\mathbf {w}\) is a vector field on \(\varphi (\mathcal {B}) \subset \mathcal {S}\), the pull-back of \(\mathbf {w}\) by \(\varphi \) is a vector field on \(\mathcal {B}\) defined as \(\varphi ^*\mathbf {w}=T(\varphi ^{-1}) \cdot \mathbf {w} \circ \varphi \). Let us denote the tangent map of \(\varphi \) by \(\mathbf {F}\), i.e., \(\mathbf {F}=T\varphi \). Let \(\{X^A\}\) and \(\{x^a\}\) be the local charts for \(\mathcal {B}\) and \(\mathcal {S}\), respectively. More specifically, a local chart for \(\mathcal {B}\) at \(X\in \mathcal {B}\) is a homeomorphism from an open subset \(\mathcal {U}\subset \mathcal {B}\) (\(X\in \mathcal {U}\)) to an open subset \(\mathcal {V}\subset {{\mathbb {R}}}^n\). \(\{X^A\}\) are components of this map. The derivative map \(\mathbf {F}\) is a two-point tensor with the following representation in the local charts: \(\mathbf {F}=F^a{}_A\frac{\partial }{\partial X^A}\otimes dx^a\), \(F^a{}_A=\frac{\partial \varphi ^a}{\partial X^A}\), where \(\{\frac{\partial }{\partial X^A}\}\) and \(\{dx^a\}\) are bases for \(T_X\mathcal {B}\) and \(T^*_{\varphi (X)}\varphi (\mathcal {B})\), respectively. Recall that \(T^*_{\varphi (X)}\varphi (\mathcal {B})\) denotes the cotangent space (or the dual space) of \(T_{\varphi (X)}\varphi (\mathcal {B})\). The push-forward and pull-back of vectors have the following coordinate representations: \((\varphi _*\mathbf {W})^a=F^a{}_A W^A\), and \((\varphi ^*\mathbf {w})^A=(F^{-1})^A{}_a w^a\).

A type \(({}^{0}_{2})\)-tensor at \(X \in \mathcal {B}\) is a bilinear map \(\mathbf {T}: T_{X}\mathcal {B}\times T_{X}\mathcal {B} \rightarrow {\mathbb {R}}\), where in a local coordinate chart \(\{X^A\}\) for \(\mathcal {B}\) it reads \(\mathbf {T}(\mathbf {U},\mathbf {V})= T_{AB}U^{A}V^{B}, ~\forall \, \mathbf {U},\mathbf {V} \in T_{X} \mathcal {B}\). A Riemannian manifold \((\mathcal {B},\mathbf {G})\) is a smooth manifold \(\mathcal {B}\) endowed with an inner product \(\mathbf {G}_X\) (a symmetric \(({}^{0}_{2})\)-tensor field) on the tangent space \(T_X\mathcal {B}\) that smoothly varies in the sense that if \(\mathbf {U}\) and \(\mathbf {V}\) are smooth vector fields on \(\mathcal {B}\), then \(X \mapsto \mathbf {G}_X(\mathbf {U}_X,\mathbf {V}_X) =: \langle \!\langle \mathbf {U}_X,\mathbf {V}_X \rangle \!\rangle _{\mathbf {G}_X}\), is a smooth function. Let \((\mathcal {B},\mathbf {G})\) and \((\mathcal {S},\mathbf {g})\) be Riemannian manifolds and let \(\varphi :\mathcal {B}\rightarrow \mathcal {S}\) be a diffeomorphism (smooth map with smooth inverse). The push-forward of the metric \(\mathbf {G}\) is a metric on \(\varphi (\mathcal {B})\subset \mathcal {S}\), which is denoted by \(\varphi _*\mathbf {G}\) defined as

$$\begin{aligned} (\varphi _*\mathbf {G})_{\varphi (X)}\left( \mathbf {u}_{\varphi (X)},\mathbf {v}_{\varphi (X)}\right) :=\mathbf {G}_X\left( (\varphi ^*\mathbf {u})_X,(\varphi ^*\mathbf {v})_X\right) . \end{aligned}$$
(A.1)

In components, \((\varphi _*\mathbf {G})_{ab} = (F^{-1})^A{}_a (F^{-1})^B{}_b G_{AB}\). Similarly, the pull-back of the metric \(\mathbf {g}\) is a metric in \(\mathcal {B}\), which is denoted by \(\varphi ^*\mathbf {g}\) defined as

$$\begin{aligned} (\varphi ^*\mathbf {g})_X(\mathbf {U}_X,\mathbf {V}_X) :=\mathbf {g}_{\varphi (X)}((\varphi _*\mathbf {U})_{\varphi (X)},(\varphi _*\mathbf {V})_{\varphi (X)}) . \end{aligned}$$
(A.2)

In components, \((\varphi ^*\mathbf {g})_{AB} = F^a{}_A F^b{}_B g_{ab}\). The diffeomorphism \(\varphi \) is an isometry between two Riemannian manifolds \((\mathcal {B},\mathbf {G})\) and \((\mathcal {S},\mathbf {g})\) if \(\mathbf {g}=\varphi _*\mathbf {G}\), or equivalently, \(\mathbf {G}=\varphi ^*\mathbf {g}\). An isometry, by definition, preserves distances.

Affine connections, and their torsion and curvature tensors.A linear (affine) connection on a manifold \(\mathcal {B}\) is an operation \(\nabla :\mathcal {X}(\mathcal {B})\times \mathcal {X}(\mathcal {B})\rightarrow \mathcal {X}(\mathcal {B})\), where \(\mathcal {X}(\mathcal {B})\) is the set of vector fields on \(\mathcal {B}\), such that \(\forall ~\mathbf {X},\mathbf {Y},\mathbf {X}_1,\mathbf {X}_2,\mathbf {Y}_1,\mathbf {Y}_2\in \mathcal {X}(\mathcal {B}),\forall ~f,f_1,f_2\in C^{\infty }(\mathcal {B}),\forall ~a_1,a_2\in {\mathbb {R}}\): i) \(\nabla _{f_1\mathbf {X}_1+f_2\mathbf {X}_2}\mathbf {Y}=f_1\nabla _{\mathbf {X}_1}\mathbf {Y} +f_2\nabla _{\mathbf {X}_2}\mathbf {Y}\), ii) \(\nabla _{\mathbf {X}}(a_1\mathbf {Y}_1+a_2\mathbf {Y}_2)=a_1\nabla _{\mathbf {X}}(\mathbf {Y}_1)+a_2\nabla _{\mathbf {X}}(\mathbf {Y}_2)\), and iii) \(\nabla _{\mathbf {X}}(f\mathbf {Y})=f\nabla _{\mathbf {X}}\mathbf {Y}+(\mathbf {X}f)\mathbf {Y}\). \(\nabla _{\mathbf {X}}\mathbf {Y}\) is called the covariant derivative of \(\mathbf {Y}\) along \(\mathbf {X}\). In a local coordinate chart \(\{X^A\}\), \(\nabla _{\partial _A}\partial _B=\varGamma ^C{}_{AB}\partial _C\), where \(\varGamma ^C{}_{AB}\) are Christoffel symbols of the connection, and \(\partial _A=\frac{\partial }{\partial x^A}\) are the natural bases for the tangent space corresponding to a coordinate chart \(\{x^A\}\). A linear connection is said to be compatible with a metric \(\mathbf {G}\) on the manifold if

(A.3)

where \(\left\langle \!\left\langle .,. \right\rangle \!\right\rangle _{\mathbf {G}}\) is the inner product induced by the metric \(\mathbf {G}\). It can be shown that \(\nabla \) is compatible with \(\mathbf {G}\) if and only if \(\nabla \mathbf {G}=\mathbf {0}\), or in components

$$\begin{aligned} G_{AB|C}=\frac{\partial G_{AB}}{\partial X^C}-\varGamma ^S{}_{CA}G_{SB}-\varGamma ^S{}_{CB}G_{AS}=0. \end{aligned}$$
(A.4)

Suppose \(\mathbf {V},\mathbf {W}\in \mathcal {X}(\mathcal {B})\) are vector fields and \(\alpha :I\rightarrow \mathcal {B}\) is a smooth curve. The restriction of the vector fields to \(\alpha \), i.e., \(\mathbf {V}\circ \alpha \) and \(\mathbf {W}\circ \alpha \) are called vector fields along the curve \(\alpha \). The set of all vector fields along \(\alpha \) is denoted by \(\mathcal {X}(\alpha )\). Covariant derivative along the curve \(\alpha \) is a map \(D_t:\mathcal {X}(\alpha )\rightarrow \mathcal {X}(\alpha )\) with the following properties: \(D_t(\mathbf {V}+\mathbf {W})=D_t\mathbf {V}+D_t\mathbf {W}\), and \(D_t(f\mathbf {W})=\frac{df}{dt}\mathbf {W}+fD_t\mathbf {W}\). If \(\mathbf {W}\in \mathcal {X}(\alpha )\) is the restriction of \(\widetilde{\mathbf {W}}\in \mathcal {X}(\mathcal {B})\) to \(\alpha \), then, \(D_t\mathbf {W}=\nabla _{\alpha '(t)}\widetilde{\mathbf {W}}\). If the connection \(\nabla \) is \(\mathbf {G}\)-compatible, then

(A.5)

The covariant derivative of a two-point tensor \(\mathbf {T}\) is given by

$$\begin{aligned}&T^{AB\cdots F}{}_{G\cdots Q}{}^{ab\cdots f}{}_{g\cdots q|K} =\frac{\partial }{\partial X^k}T^{AB\cdots F}{}_{G\cdots Q}{}^{ab\cdots f}{}_{g\cdots q}\nonumber \\&\qquad +T^{RB\cdots F}{}_{G\cdots Q}{}^{ab\cdots f}{}_{g\cdots q}\varGamma ^A{}_{RK} +\mathrm {(all\,\,upper\,\, referential\,\, indices)}\nonumber \\&\qquad -T^{AB\cdots F}{}_{R\cdots Q}{}^{ab\cdots f}{}_{g\cdots q}\varGamma ^R{}_{GK} -\mathrm {(all\,\,lower\,\, referential\,\, indices)}\nonumber \\&\qquad +T^{RB\cdots F}{}_{G\cdots Q}{}^{lb\cdots f}{}_{g\cdots q}\gamma ^a{}_{lr}F^r{}_K +\mathrm {(all\,\,upper\,\, spatial\,\, indices)}\nonumber \\&\qquad -T^{AB\cdots F}{}_{G\cdots Q}{}^{ab\cdots f}{}_{l\cdots q}\gamma ^l{}_{gr}F^r{}_K -\mathrm {(all\,\,lower\,\, spatial\,\, indices)}\,. \end{aligned}$$
(A.6)

The torsion of a connection is defined as \(\varvec{T}(\mathbf {X},\mathbf {Y})=\nabla _{\mathbf {X}}\mathbf {Y}-\nabla _{\mathbf {Y}}\mathbf {X}-[\mathbf {X},\mathbf {Y}]\), where \([\mathbf {X},\mathbf {Y}](F)=\mathbf {X}(\mathbf {Y}(F))-\mathbf {Y}(\mathbf {X}(F)),~\forall ~F\in C^{\infty }(\mathcal {S})\), is the commutator of \(\mathbf X \) and \(\mathbf Y \). In components, in a local chart \(\{X^A\}\), \(T^A{}_{BC}=\varGamma ^A{}_{BC}-\varGamma ^A{}_{CB}\), and \([\mathbf {X},\mathbf {Y}]^a=\frac{\partial Y^a}{\partial x^b}X^b-\frac{\partial X^a}{\partial x^b}Y^b\). \(\nabla \) is symmetric if it is torsion-free, i.e., \(\nabla _{\mathbf {X}}\mathbf {Y}-\nabla _{\mathbf {Y}}\mathbf {X}=[\mathbf {X},\mathbf {Y}]\). On any Riemannian manifold \((\mathcal {B},\mathbf {G})\) there is a unique linear connection \(\nabla ^{\mathbf {G}}\) that is compatible with \(\mathbf {G}\) and is torsion-free. This is the Levi-Civita connection. If the Levi-Civita connection \(\nabla ^{\mathbf {G}}\) is used, the covariant time derivative is denoted by \(D^{\mathbf {G}}_t\). In a manifold with a connection the curvature is a map \(\varvec{\mathcal {R}}:\mathcal {X}(\mathcal {B})\times \mathcal {X}(\mathcal {B})\times \mathcal {X}(\mathcal {B})\rightarrow \mathcal {X}(\mathcal {B})\) defined by \(\varvec{\mathcal {R}}(\mathbf {X},\mathbf {Y},\mathbf {Z})=\nabla _{\mathbf {X}}\nabla _{\mathbf {Y}}\mathbf {Z}-\nabla _{\mathbf {Y}}\nabla _{\mathbf {X}}\mathbf {Z}-\nabla _{[\mathbf {X},\mathbf {Y}]}\mathbf {Z}\), or in components \(\mathcal {R}^A{}_{BCD}=\frac{\partial \varGamma ^A{}_{CD}}{\partial X^B}-\frac{\partial \varGamma ^A{}_{BD}}{\partial X^C}+\varGamma ^A{}_{BM}\varGamma ^M{}_{CD}-\varGamma ^A{}_{CM}\varGamma ^M{}_{BD}\). The Riemannian curvature is the curvature tensor of the Levi-Civita connection \(\nabla ^{\mathbf {G}}\) and is denoted by \(\varvec{\mathcal {R}}_{\mathbf {G}}\). The Ricci identity for a vector field \(\mathbf {U}\) with components \(W^A\) reads \(U^A{}_{|BC}-U^A{}_{|CB}=\mathcal {R}^A{}_{BCD}U^D\). Ricci identity for a 1-form \(\varvec{\alpha }\) with components \(\alpha _A\) reads \(\alpha _A{}_{|BC}-\alpha _A{}_{|CB}=\mathcal {R}^D{}_{BCA}\alpha _D\). The Ricci curvature \(\varvec{\mathsf {Ric}}\) is defined as \(\mathsf {Ric}_{CD}=\mathcal {R}^A{}_{ACD}\), and is a symmetric tensor. The Ricci curvature of the Levi-Civita connection \(\nabla ^{\mathbf {G}}\) is denoted by \(\varvec{\mathsf {Ric}}_{\mathbf {G}}\).

Vector bundles.Suppose \(\mathcal {E}\) and \(\mathcal {B}\) are sets and consider a map \(\pi :\mathcal {E}\rightarrow \mathcal {B}\). The fiber over \(X\in \mathcal {B}\) is the set \(\mathcal {E}_X:=\pi ^{-1}(X)\subset \mathcal {E}\). For an onto map \(\pi \) fibers are non-empty and \(\mathcal {E}=\sqcup _{X\in \mathcal {B}}\mathcal {E}_X\), where \(\sqcup \) denoted disjoint union of sets. Now suppose \(\mathcal {E}\) and \(\mathcal {B}\) are manifolds and assume that for any \(X\in \mathcal {B}\), there exists a neighborhood \(\mathcal {U}\subset \mathcal {B}\) of X, a manifold \(\mathcal {F}\), and a diffeomorphism \(\psi :\pi ^{-1}(\mathcal {U})\rightarrow \mathcal {U}\times \mathcal {F}\) such that \(\pi ={\text {pr}}_1\circ \psi \), where \({\text {pr}}_1:\mathcal {U}\times \mathcal {F}\rightarrow \mathcal {U}\) is projection onto the first factor. \((\mathcal {E},\pi ,\mathcal {B})\) is called a fiber bundle and \(\mathcal {E}\), \(\pi \), and \(\mathcal {B}\) are called the total space, the projection, and the base space, respectively. If for any \(X\in \mathcal {B}\), \(\pi ^{-1}(X)\) is a vector space, \((\mathcal {E},\pi ,\mathcal {B})\) is called a vector bundle. The set of sections of this bundle \(\varGamma (\mathcal {E})\) is the set of all smooth maps \(\sigma :\mathcal {B}\rightarrow \mathcal {E}\) such that \(\sigma (X)\in \mathcal {E}_X,~\forall ~X\in \mathcal {B}\). An important example of a vector bundle is the tangent bundle of a manifold for which \(\mathcal {E}=T\mathcal {B}\).

Induced bundle and connection.Consider a map between Riemannian manifolds \(\varphi :\mathcal {B}\rightarrow \mathcal {S}\). The tangent bundles of \(\mathcal {B}\) and \(\mathcal {S}\) are denoted by \(T\mathcal {B}=\sqcup _{X\in \mathcal {B}}T_X\mathcal {B}\) and \(T\mathcal {S}=\sqcup _{x\in \mathcal {S}}T_x\mathcal {S}\), respectively. We define an induced vector bundle \(\varphi ^{-1}T\mathcal {S}\), which is a vector bundle over \(\mathcal {B}\) whose fiber over \(X\in \mathcal {B}\) is \(T_{\varphi (X)}\mathcal {S}\) [76]. The connection \(\nabla ^{\mathbf {g}}\) induces a unique connection \(\nabla ^{\varphi }\) on \(\varphi ^{-1}T\mathcal {S}\) defined as

$$\begin{aligned} \nabla ^{\varphi }_{\mathbf {W}}\mathbf {w}\circ \varphi =\nabla ^{\mathbf {g}}_{\varphi _*\mathbf {W}}\mathbf {w} ,~~~\mathbf {W}\in T_X\mathcal {B},~\mathbf {w}\in \varGamma (T\mathcal {S})\,. \end{aligned}$$
(A.7)

\(\nabla ^{\varphi }\) is called the induced connection. It can be shown that its connection coefficients with respect to the coordinate charts \(\{X^A\}\) and \(\{x^a\}\) of \(\mathcal {B}\) and \(\mathcal {S}\), respectively, are \(\frac{\partial \varphi ^b}{\partial X^A}\gamma ^a{}_{bc}\). In particular, the variation field \(\delta \varphi \) defined in §3 is a section of \(\varGamma (\varphi ^{-1}T\mathcal {S})\), i.e., \(\delta \varphi \) defines a vector field in \(\mathcal {S}\) along the map \(\varphi \). For a two-point tensor, e.g., deformation gradient, covariant derivative involves both \(\nabla ^{\mathbf {g}}\) and \(\nabla ^{\mathbf {G}}\): \(F^a{}_{A|B}=\frac{\partial F^a{}_A}{\partial X^B}+(F^b{}_B\gamma ^a{}_{bc})F^c{}_A-\varGamma ^C{}_{AB}F^a{}_C=\frac{\partial F^a{}_A}{\partial X^B}+\gamma ^a{}_{bc}F^b{}_BF^c{}_A-\varGamma ^C{}_{AB}F^a{}_C\). We denote the covariant derivative of the deformation gradient by \(\nabla \mathbf {F}=F^a{}_{A|B}dX^B\otimes dX^A\otimes \frac{\partial }{\partial x^a}\). It is straightforward to show that [76]

$$\begin{aligned} \nabla ^{\varphi }\mathbf {F}(\mathbf {X},\mathbf {Y})=\nabla ^{\varphi }_{\mathbf {X}}\varphi _*\mathbf {Y} -\varphi _*\nabla ^{\mathbf {G}}_{\mathbf {X}}\mathbf {Y},~~~ \nabla ^{\varphi }_{\mathbf {X}}\varphi _*\mathbf {Y}-\nabla ^{\varphi }_{\mathbf {Y}}\varphi _*\mathbf {X} =\varphi _*[\mathbf {X},\mathbf {Y}]. \end{aligned}$$
(A.8)

The metrics \(\mathbf {G}\) and \(\mathbf {g}\) induce an inner product \(\langle , \rangle _X\) in \(T_{\varphi (X)}\mathcal {S}\otimes T^*_X\mathcal {B}\). This is defined first for the basis \(\left\{ \frac{\partial }{\partial x^a}\otimes dX^A, ~ 1\le a\le n,~1\le A\le n \right\} \) as \(\langle \frac{\partial }{\partial x^a}\otimes dX^A,\frac{\partial }{\partial x^b}\otimes dX^B \rangle _X=g_{ab}G^{AB}\), and then one extends it linearly to arbitrary elements in \(T_{\varphi (X)}\mathcal {S}\otimes T^*_X\mathcal {B}\). \(\varphi ^{-1}T\mathcal {S}\otimes T^*\mathcal {B}\) is the vector bundle whose fiber at \(X\in \mathcal {B}\) is \(T_{\varphi (X)}\mathcal {S}\otimes T^*_X\mathcal {B}\). The two-point tensor \(\mathbf {F}=T\varphi :\mathcal {B}\rightarrow \varphi ^{-1}T\mathcal {S}\otimes T^*\mathcal {B}\), is a section of this bundle, i.e., \(\mathbf {F}\in \varGamma (\varphi ^{-1}T\mathcal {S}\otimes T^*\mathcal {B})\). One can define a fiber metric on \(\varphi ^{-1}T\mathcal {S}\otimes T^*\mathcal {B}\) using the inner product \(\langle , \rangle _X\) in \(T_{\varphi (X)}\mathcal {S}\otimes T_X^*\mathcal {B}\) as follows: for \(\sigma , \tau \in \varGamma (\varphi ^{-1}T\mathcal {S}\otimes T^*\mathcal {B})\), define . One can define a connection \(\nabla \) in \(\varphi ^{-1}T\mathcal {S}\otimes T^*\mathcal {B}\) using the Levi-Civita connections \(\nabla ^{\mathbf {G}}\) and \(\nabla ^{\mathbf {g}}\): consider a section \(\mathbf {W}\otimes \varvec{\alpha }\in \varGamma (\varphi ^{-1}T\mathcal {S}\otimes T^*\mathcal {B})\) and let \(\nabla (\mathbf {W}\otimes \varvec{\alpha })=\nabla ^{\varphi }\mathbf {W}\otimes \varvec{\alpha }+\mathbf {W}\otimes \nabla ^{\mathbf {G}}\varvec{\alpha }\). This connection is compatible with the fiber metric in \(\varphi ^{-1}T\mathcal {S}\otimes T^*\mathcal {B}\).

The Piola transform.The Piola transform of a vector \({\mathbf {w}}\in T_{\varphi (X)}\mathcal {S}\) is a vector \(\mathbf {W}\in T_X\mathcal {B}\) given by \(\mathbf {W}=J\varphi ^*\mathbf {w}=J\mathbf {F}^{-1}\mathbf {w}\). In coordinates, \(W^A=J(F^{-1})^A{}_bw^b\), where \(J=\sqrt{\frac{\det \mathbf {g}}{\det \mathbf {G}}}\det \mathbf {F}\) is the Jacobian of \(\varphi \) with \(\mathbf {G}\) and \(\mathbf {g}\) the Riemannian metrics of \(\mathcal {B}\) and \(\mathcal {S}\), respectively. It can be shown that \({\text {Div}}\mathbf {W}=J({\text {div}}\mathbf {w})\circ \varphi \). In coordinates, \(W^A{}_{|A}=Jw^a{}_{|a}\); this is also known as the Piola identity. Another way of writing the Piola identity is in terms of the unit normal vectors of a surface in \(\mathcal {B}\) and its corresponding surface in \(\mathcal {S}\) and the area elements. It is written as \(\hat{\mathbf {n}}da=J\mathbf {F}^{-\star }\hat{\mathbf {N}}dA\), or in components, \(n_ada=J(F^{-1})^A{}_aN_AdA\). In the literature of continuum mechanics, this is called Nanson’s formula.

Lie derivative.Let \(\mathbf {w}:\mathcal {U}\rightarrow T\mathcal {S}\) be a vector field, where \(\mathcal {U} \subset \mathcal {S}\) is open. A curve \(\alpha :I \rightarrow \mathcal {S}\), where I is an open interval, is an integral curve of \(\mathbf {w}\) if \(\frac{d \alpha (t)}{dt}=\mathbf {w}(\alpha (t)),~\forall ~t \in I\). For a time-dependent vector field \(\mathbf {w}:\mathcal {S}\times I\rightarrow T\mathcal {S}\), where I is some open interval, the collection of maps \(\psi _{\tau ,t}\) is the flow of \(\mathbf {w}\) if for each t and x, \(\tau \mapsto \psi _{\tau ,t}(x)\) is an integral curve of \(\mathbf {w}_t\), i.e., \(\frac{d}{d\tau }\psi _{\tau ,t}(x)=\mathbf {w}(\psi _{\tau ,t}(x),\tau )\), and \(\psi _{t,t}(x)=x\). Let \(\mathbf {t}\) be a time-dependent tensor field on \(\mathcal {S}\), i.e., \(\mathbf {t}_t(x)=\mathbf {t}(x,t)\) is a tensor. The Lie derivative of \(\mathbf {t}\) with respect to \(\mathbf {w}\) is defined as \(\mathbf {L}_{\mathbf {w}}\mathbf {t}=\frac{d}{d \tau }\psi _{\tau ,t}^* \mathbf {t}_{\tau } \big |_{\tau =t}\). Note that \(\psi _{\tau ,t}\) maps \(\mathbf {t}_t\) to \(\mathbf {t}_{\tau }\). Hence, to calculate the Lie derivative one drags \(\mathbf {t}\) along the flow of \(\mathbf {w}\) from \(\tau \) to t and then differentiates the Lie dragged tensor with respect to \(\tau \). The autonomous Lie derivative of \(\mathbf {t}\) with respect to \(\mathbf {w}\) is defined as \(\mathfrak {L}_{\mathbf {w}}\mathbf {t}=\frac{d}{d \tau }\psi _{\tau ,t}^* \mathbf {t}_{t} \big |_{\tau =t}\). Thus, \(\mathbf {L}_{\mathbf {w}}\mathbf {t}=\partial \mathbf {t}/\partial t+\mathfrak {L}_{\mathbf {w}}\mathbf {t}\). For a scalar f, \(\mathbf {L}_{\mathbf {w}}f=\partial f/\partial t+\mathbf {w}[f]\). In a coordinate chart \(\{x^a\}\) this reads, \(\mathbf {L}_{\mathbf {w}}f=\frac{\partial f}{\partial t}+\frac{\partial f}{\partial x^a}w^a\). For a vector \(\mathbf {u}\), one can show that \(\mathbf {L}_{\mathbf {w}}\mathbf {u}=\frac{\partial \mathbf {w}}{\partial t}+[\mathbf {w},\mathbf {u}]\). If \(\nabla \) is a torsion-free connection, then \([\mathbf {w},\mathbf {u}]=\nabla _{\mathbf {w}}\mathbf {u}-\nabla _{\mathbf {u}}\mathbf {w}\). Thus, \(\mathbf {L}_{\mathbf {w}}\mathbf {u}=\frac{\partial \mathbf {w}}{\partial t}+\nabla _{\mathbf {w}}\mathbf {u}-\nabla _{\mathbf {u}}\mathbf {w}\).

When linearizing nonlinear elasticity one starts with a one-parameter family of motions \(\varphi _{t,\epsilon }:\mathcal {B}\rightarrow \mathcal {S}\). By definition of the variation field \(\mathbf {U}_t=\delta \varphi _t\), \(\varphi _{t,\epsilon }\) is the flow of the variation field. Given a tensor field \(\mathbf {t}\) in \(\mathcal {S}\), \(\bar{\mathbf {T}}_{\epsilon }=\varphi _{t,\epsilon }^*\mathbf {t}\circ \varphi _{t,\epsilon }\) is a vector field on \(\mathcal {B}\). Its linearization is defined as

$$\begin{aligned} \begin{aligned} \delta \bar{\mathbf {T}}&=\frac{d}{d\epsilon }\bar{\mathbf {T}}_{\epsilon }\Big |_{\epsilon =0} =\left( \frac{d}{d\epsilon }\varphi _{t,\epsilon }^*\mathbf {t}\circ \varphi _{t,\epsilon }\right) \Big |_{\epsilon =0} \\&=\left( \varphi _{t,\epsilon }^*\mathbf {L}_{\mathbf {U}_{t,\epsilon }}\mathbf {t}\circ \varphi _{t,\epsilon }\right) \Big |_{\epsilon =0} =\mathring{\varphi }_t^*\left( \mathbf {L}_{\mathbf {U}_{t}}\mathbf {t}\circ \mathring{\varphi }_t\right) . \end{aligned} \end{aligned}$$
(A.9)

Thus, \(\delta \mathbf {t}=\mathbf {L}_{\mathbf {u}_t}\mathbf {t}\), where \(\mathbf {u}_t=\mathbf {U}_t\circ \mathring{\varphi }_t^{-1}\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yavari, A., Golgoon, A. Nonlinear and Linear Elastodynamic Transformation Cloaking. Arch Rational Mech Anal 234, 211–316 (2019). https://doi.org/10.1007/s00205-019-01389-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00205-019-01389-2

Navigation