Skip to main content

Advertisement

Log in

Macroscopic and Microscopic Behavior of Narrow Elastic Ribbons

  • Published:
Journal of Elasticity Aims and scope Submit manuscript

Abstract

A one-dimensional model for a narrow ribbon is derived from the plate theory of Kirchhoff by means of a power expansion in the width variable. The energy found coincides with the corrected Sadowsky’s energy. Furthermore, we derive the Euler-Lagrange equations and use them to study an equilibrium configuration of a twisted ribbon. Within this example we also describe how to construct the fine scale oscillations that develop in the deformed configuration.

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

Similar content being viewed by others

References

  1. Agostiniani, V., DeSimone, A., Koumatos, K.: Shape programming for narrow ribbons of nematic elastomers. J. Elast. 127, 1–24 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  2. Agostiniani, V., DeSimone, A.: Dimension reduction via \(\varGamma \)-convergence for soft active materials. Meccanica 52, 3457–3470 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  3. Antman, S.S.: Nonlinear Problems of Elasticity, 2nd edn. Applied Mathematical Sciences, vol. 107. Springer, New York (2005)

    MATH  Google Scholar 

  4. Audoly, B., Seffen, K.A.: Buckling of naturally curved elastic strips: the ribbon model makes a difference. J. Elast. 119, 293–320 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  5. Chen, Y.C., Fosdick, R., Fried, E.: Representation for a smooth isometric mapping from a connected planar domain to a surface. J. Elast. 119, 335–350 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  6. Chen, Y.C., Fosdick, R., Fried, E.: Representation of a smooth isometric deformation of a planar material region into a curved surface. J. Elast. 130, 145–195 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  7. Chen, Y.C., Fosdick, R., Fried, E.: Issues concerning isometric deformations of planar regions to curved surfaces. J. Elast. 132, 1–42 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  8. Chen, Y.C., Fried, E.: Möbius bands, unstretchable material sheets, and developable surfaces. Proc. R. Soc., Math. Phys. Eng. Sci. 472, 20150760 (2016)

    Article  ADS  MATH  Google Scholar 

  9. Chopin, J., Kudrolli, A.: Helicoids, wrinkles, and loops in twisted ribbons. Phys. Rev. Lett. 111, 174302 (2013)

    Article  ADS  Google Scholar 

  10. Chopin, J., Démery, V., Davidovitch, B.: Roadmap to the morphological instabilities of a stretched twisted ribbon. J. Elast. 119, 137–189 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  11. Dacorogna, B.: Direct Methods in the Calculus of Variations. Applied Math. Sciences, vol. 78. Springer, Berlin (1989)

    MATH  Google Scholar 

  12. Dias, M.A., Audoly, B.: “Wunderlich, Meet Kirchhoff”: a general and unified description of elastic ribbons and thin rods. J. Elast. 119, 49–66 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  13. Efrati, E.: Non-Euclidean ribbons. J. Elast. 119, 251–261 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  14. Fonseca, I., Leoni, G.: Modern Methods in the Calculus of Variations: \(L^{p}\) Spaces. Springer Monographs in Mathematics. Springer, New York (2007)

    MATH  Google Scholar 

  15. Fosdick, R., Fried, E.: The Mechanics of Ribbons and Möbius Bands. Springer, Dordrecht (2016)

    Book  MATH  Google Scholar 

  16. Freddi, L., Hornung, P., Mora, M.G., Paroni, R.: A corrected Sadowsky functional for inextensible elastic ribbons. J. Elast. 123, 125–136 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  17. Freddi, L., Hornung, P., Mora, M.G., Paroni, R.: A variational model for anisotropic and naturally twisted ribbons. SIAM J. Math. Anal. 48, 3883–3906 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  18. Freddi, L., Hornung, P., Mora, M.G., Paroni, R.: One-dimensional von Kármán models for elastic ribbons. Meccanica 53, 659–670 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  19. Freddi, L., Hornung, P., Mora, M.G., Paroni, R.: Generalised Sadowsky theories for ribbons from three-dimensional nonlinear elasticity. Submitted

  20. Freddi, L., Mora, M.G., Paroni, R.: Nonlinear thin-walled beams with a rectangular cross-section—Part I. Math. Models Methods Appl. Sci. 22, 1150016 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  21. Freddi, L., Mora, M.G., Paroni, R.: Nonlinear thin-walled beams with a rectangular cross-section—Part II. Math. Models Methods Appl. Sci. 23, 743–775 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  22. Hinz, D.F., Fried, E.: Translation of Michael Sadowsky’s paper “An elementary proof for the existence of a developable Möbius band and the attribution of the geometric problem to a variational problem”. J. Elast. 119, 3–6 (2015)

    Article  MATH  Google Scholar 

  23. Kirby, N.O., Fried, E.: Gamma-limit of a model for the elastic energy of an inextensible ribbon. J. Elast. 119, 35–47 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  24. Kirchhoff, G.: Über das Gleichgewicht und die Bewegung einer elastischen Scheibe. J. Reine Angew. Math. 40, 51–88 (1850)

    Article  MathSciNet  Google Scholar 

  25. Kohn, R.V., O’Brien, E.: The wrinkling of a twisted ribbon. J. Nonlinear Sci. 28, 1221–1249 (2018)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  26. Moore, A., Healey, T.: Computation of elastic equilibria of complete Möbius bands and their stability. Math. Mech. Solids (2018). https://doi.org/10.1177/1081286518761789

    Google Scholar 

  27. Nardinocchi, P., Pezzulla, M., Teresi, L.: Anisotropic swelling of thin gel sheets. Soft Matter 11, 1492–1499 (2015)

    Article  ADS  Google Scholar 

  28. Paroni, R.: An existence theorem for inextensible nets with slack. Math. Mech. Solids 17, 460–472 (2012)

    Article  MathSciNet  Google Scholar 

  29. Pipkin, A.C.: Inextensible networks with slack. Q. Appl. Math. 40, 63–71 (1982/1983)

    Article  MathSciNet  MATH  Google Scholar 

  30. Rivlin, R.S.: Plane strain of a net formed by inextensible cords. J. Ration. Mech. Anal. 4, 951–974 (1955)

    MathSciNet  MATH  Google Scholar 

  31. Rivlin, R.S.: The deformation of a membrane formed by inextensible cords. Arch. Ration. Mech. Anal. 2, 447–476 (1958/1959)

    Article  MathSciNet  MATH  Google Scholar 

  32. Sadowsky, M.: Ein elementarer Beweis für die Existenz eines abwickelbaren Möbiuschen Bandes und die Zurückführung des geometrischen Problems auf ein Variationsproblem. Sitz.ber. Preuss. Akad. Wiss. 412–415 (1930)

  33. Starostin, E.L., van der Heijden, G.H.M.: The equilibrium shape of an elastic developable Möbius strip. Proc. Appl. Math. Mech. 7, 2020115–2020116 (2007)

    Article  Google Scholar 

  34. Starostin, E.L., van der Heijden, G.H.M.: Equilibrium shapes with stress localisation for inextensible elastic Möbius and other strips. J. Elast. 119, 67–112 (2015)

    Article  MATH  Google Scholar 

  35. Tomassetti, G., Varano, V.: Capturing the helical to spiral transitions in thin ribbons of nematic elastomers. Meccanica 52, 3431–3441 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  36. Todres, R.E.: Translation of W. Wunderlich’s “On a developable Möbius band”. J. Elast. 119, 23–34 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  37. Teresi, L., Varano, V.: Modeling helicoid to spiral-ribbon transitions of twist-nematic elastomers. Soft Matter 9, 3081–3088 (2013)

    Article  ADS  Google Scholar 

  38. Sawa, Y., Urayama, K., Takigawa, T., Gimenez-Pinto, V., Mbanga, B.L., Ye, F., Selinger, J.V., Selinger, R.L.: Shape and chirality transitions in off-axis twist nematic elastomer ribbons. Phys. Rev. E 88, 022502 (2013)

    Article  ADS  Google Scholar 

  39. Wan, G., Jin, C., Trase, I., Zhao, S., Chen, Z.: Helical structures mimicking chiral seedpod opening and tendril coiling. Sensors 18, 2973 (2018)

    Article  Google Scholar 

  40. Wunderlich, W.: Über ein abwickelbares Möbiusband. Monatshefte Math. 66, 276–289 (1962)

    Article  MATH  Google Scholar 

Download references

Acknowledgements

R. Paroni acknowledges support from the Università di Pisa through the project PRA_2018_61 “Modellazione multi-scala in ingegneria strutturale”.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Roberto Paroni.

Additional information

Dedicated to the memory of Walter Noll

Publisher’s Note

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

Appendix

Appendix

1.1 A.1 Convexification

The aim of this appendix is to determine the convex envelope of

$$ W(\mathbf{L})=\left \{ \textstyle\begin{array}{l@{\quad }l} Q(\mathbf{L}) & \mbox{if }\det \mathbf{L}=0 , \\ +\infty & \mbox{else}, \end{array}\displaystyle \right . $$

where \(Q\) is a positive definite quadratic form defined on 2-by-2 symmetric matrices.

Consider the function

$$ Q_{\mathrm{det}}(\mathbf{L};\alpha )=Q(\mathbf{L})+\alpha \det \mathbf{L} $$

with \(\alpha \) a real number. For fixed \(\alpha \), \(Q_{\mathrm{det}}( \cdot ;\alpha )\) is a quadratic function—recall that \(\mathbf{L}\) may be represented by a \(2\times 2\) matrix—and hence it is convex if and only if \(Q_{\mathrm{det}}(\mathbf{L};\alpha )\ge 0\) for every \(\mathbf{L}\). This inequality is satisfied for all \(\mathbf{L}\) with determinant equal to zero, since \(Q\) is positive definite. If \(\det \mathbf{L}>0\), the condition \(Q_{\mathrm{det}}(\mathbf{L};\alpha ) \ge 0\) for every \(\mathbf{L}\) holds if and only if

$$ Q(\mathbf{L}/\sqrt{\det \mathbf{L}})+\alpha \ge 0\quad \mbox{for every } \mathbf{L}, $$

that is equivalent to

$$ \alpha \ge - Q(\mathbf{L})\quad \mbox{for every }\mathbf{L}\mbox{ with }\det \mathbf{L}=1, $$

or

$$ \alpha \ge - \min_{\det \mathbf{L}=1} Q(\mathbf{L})=:-\alpha ^{-}. $$
(60)

Similarly, if \(\det \mathbf{L}<0\), the condition \(Q_{\mathrm{det}}( \mathbf{L};\alpha )\ge 0\) for every \(\mathbf{L}\) holds if and only if

$$ \alpha \le \min_{\det \mathbf{L}=-1} Q(\mathbf{L})=:\alpha ^{+}. $$
(61)

Note that \(\alpha ^{-}, \alpha ^{+}>0\), since \(Q\) is positive definite. Hence, \(Q_{\mathrm{det}}(\cdot ;\alpha )\) is convex if and only if \(-\alpha ^{-}\le \alpha \le \alpha ^{+}\), and

$$ W_{c}(\mathbf{L})=\sup_{-\alpha ^{-}\le \alpha \le \alpha ^{+}}Q_{\mathrm{det}}( \mathbf{L};\alpha ) $$
(62)

is also convex, for the point-wise supremum of convex functions is convex. It immediately follows that

$$ W_{c}(\mathbf{L})=Q(\mathbf{L})+\alpha ^{+}(\det \mathbf{L})^{+}+\alpha ^{-}( \det \mathbf{L})^{-}, $$
(63)

where \((\det \mathbf{L})^{+}\) is the positive part of \(\det \mathbf{L}\), and \((\det \mathbf{L})^{-}\) is the negative part, i.e., \((\det \mathbf{L})^{-}=- \det \mathbf{L}\) if \(\det \mathbf{L}<0\) and zero otherwise.

Clearly, \(W(\mathbf{L})\ge Q_{\mathrm{det}}(\mathbf{L};\alpha )\) for any \(\mathbf{L}\) and for any \(\alpha \), hence \(W(\mathbf{L})\ge W_{c}( \mathbf{L})\). Since \(W_{c}\) is convex it also follows that

$$ W^{**}(\mathbf{L})\ge W_{c}(\mathbf{L}). $$

In the rest of the section we will show that the inequality above is indeed an equality, and that the lower bound can be achieved by a single lamination. More precisely, we show that given \(\mathbf{L}\) we can find two symmetric tensors \(\mathbf{A}\) and \(\mathbf{B}\), and \(0< t<1\) such that

$$ (1-t)\mathbf{A}+t\mathbf{B}=\mathbf{L}, \quad \mbox{and} \quad (1-t)W( \mathbf{A})+t W(\mathbf{B})=W_{c}(\mathbf{L}). $$

By the definition of \(W\), the statement above is equivalent to: given \(\mathbf{L}\) there exist two symmetric tensors \(\mathbf{A}\) and \(\mathbf{B}\) with \(\det \mathbf{A}=\det \mathbf{B}=0\), and \(0< t<1\) such that

$$ (1-t)\mathbf{A}+t\mathbf{B}=\mathbf{L}, \quad \mbox{and} \quad (1-t)Q( \mathbf{A})+t Q(\mathbf{B})=W_{c}(\mathbf{L}). $$

In this statement we may eliminate the tensor \(\mathbf{A}\) by noticing that: given \(\mathbf{L},\mathbf{B}\) and \(t\), as above, there exists a symmetric tensor \(\mathbf{A}\) with \(\det \mathbf{A}=0\) and satisfying the equation

$$ (1-t)\mathbf{A}=\mathbf{L}-t\mathbf{B} $$

if and only if \(\det (\mathbf{L}- t\mathbf{B})=0\). Since

$$\begin{aligned} (1-t)Q(\mathbf{A})+t Q(\mathbf{B}) &=\frac{1}{1-t} Q\bigl((1-t)\mathbf{A} \bigr)+t Q( \mathbf{B}) \\ &=\frac{1}{1-t} Q(\mathbf{L}-t\mathbf{B})+t Q(\mathbf{B}) \\ &=Q(\mathbf{L})+\frac{t}{1-t} Q(\mathbf{L}-\mathbf{B}), \end{aligned}$$

the statement to be proven is: given \(\mathbf{L}\) there exist a symmetric tensor \(\mathbf{B}\) with \(\det \mathbf{B}=0\), and \(0< t<1\) such that

$$ \det (\mathbf{L}- t\mathbf{B})=0, \quad \mbox{and} \quad Q( \mathbf{L})+\frac{t}{1-t} Q(\mathbf{L}-\mathbf{B})=W_{c}( \mathbf{L}). $$
(64)

Recalling (60) and (61), let \(\mathbf{D}^{-}, \mathbf{D}^{+}\) be the two symmetric matrices with \(\det \mathbf{D}^{-}=-1\) and \(\det \mathbf{D}^{+}=1\) such that

$$ \alpha ^{-}=Q\bigl(\mathbf{D}^{+}\bigr), \qquad \alpha ^{+}=Q\bigl(\mathbf{D}^{-}\bigr). $$
(65)

Note that the tensors \(\mathbf{D}^{-}\) and \(\mathbf{D}^{+}\) only depend on the bending stiffness tensor \(\mathbb{D}\). We now show that if \(\det \mathbf{L}>0\) there exist a symmetric tensor \(\mathbf{B}\) with null determinant and \(0< t<1\) such that

$$ \det (\mathbf{L}- t\mathbf{B})=0, \quad \quad \det (\mathbf{L}- \mathbf{B})=- \frac{1-t}{t} \det \mathbf{L}, $$
(66)

and

$$ \mathbf{D}^{-}=\frac{\mathbf{L}-\mathbf{B}}{\sqrt{-\det (\mathbf{L}- \mathbf{B})}}. $$
(67)

Assuming, for the moment, the validity of this statement we deduce (64) in the case \(\det \mathbf{L}>0\):

$$\begin{aligned} Q(\mathbf{L})+\frac{t}{1-t} Q(\mathbf{L}-\mathbf{B}) &=Q(\mathbf{L})+ \frac{t}{1-t} \bigl(-\det (\mathbf{L}-\mathbf{B})\bigr) Q\bigl( \mathbf{D}^{-}\bigr) \\ &=Q(\mathbf{L})+Q\bigl(\mathbf{D}^{-}\bigr)\det \mathbf{L} \\ &=Q(\mathbf{L})+\alpha ^{+}\det \mathbf{L}=W_{c}( \mathbf{L}), \end{aligned}$$

where we have used, in order, (67), (66), (65), and (63). Hence, assuming (66) and (67) we have shown that if \(\det \mathbf{L}>0\) then \(W^{**}(\mathbf{L})= W_{c}(\mathbf{L})\). A similar proof can be made for \(\det \mathbf{L}<0\), while the result is trivial for \(\det \mathbf{L}=0\).

We therefore only need to show the validity of (66) and (67). In doing it, we shall provide formulae that deliver \(\mathbf{B}\) and \(t\).

For \(\det \mathbf{L}>0\), set \(\mathbf{B}=\mathbf{L}-\eta \mathbf{D}^{-}\) where the real number \(\eta \) is chosen by imposing that \(\det \mathbf{B}=0\). With (1), this amount to solve the second order equation

$$ \det \mathbf{L}-\eta \mathbf{L}^{*}\cdot \mathbf{D}^{-} -\eta ^{2}=0, $$
(68)

where we used the fact that \(\det \mathbf{D}^{-}=-1\). Of the two solutions we choose

$$ \eta =\frac{-\mathbf{L}^{*}\cdot \mathbf{D}^{-} + \sqrt{(\mathbf{L}^{*} \cdot \mathbf{D}^{-})^{2}+4\det \mathbf{L}}}{2}. $$

To keep the notation compact, we set

$$ \lambda =\frac{\mathbf{L}^{*}\cdot \mathbf{D}^{-}}{2 \sqrt{\det \mathbf{L}}} $$

so that

$$ \eta =\sqrt{\det \mathbf{L}}\bigl(\sqrt{\lambda ^{2}+1}-\lambda \bigr). $$

Now that \(\mathbf{B}\) has been fixed, we find \(t\) by solving (66)1. We first write

$$ \mathbf{L}- t\mathbf{B}=\mathbf{L}- t\bigl(\mathbf{L}-\eta \mathbf{D}^{-} \bigr)=(1-t) \mathbf{L}+t \eta \mathbf{D}^{-}, $$

and by using (1), we write (66)1 as

$$ (1-t)^{2}\det \mathbf{L}+(1-t)t \eta \mathbf{L}^{*}\cdot \mathbf{D}^{-}-t ^{2} \eta ^{2}=0, $$

which simplifies, thanks to (68), to

$$ \det \mathbf{L}+t\bigl(\eta \mathbf{L}^{*}\cdot \mathbf{D}^{-}-2 \det \mathbf{L}\bigr)=0. $$

Thus

$$ t=\frac{-\det \mathbf{L}}{\eta \mathbf{L}^{*}\cdot \mathbf{D}^{-}-2\det \mathbf{L}}=\frac{1}{2-2\lambda (\sqrt{\lambda ^{2}+1}-\lambda )}. $$

We may easily verify that \(0< t<1\). Indeed, \(t>0\) if and only if \(2\lambda (\sqrt{\lambda ^{2}+1}-\lambda )<2\), which clearly holds if \(\lambda \le 0\), while for \(\lambda >0\) is equivalent to \(\sqrt{ \lambda ^{2}+1}<\frac{1}{\lambda }+\lambda \) which is trivially verified. Similarly we may check that \(t<1\).

From the definition of \(\mathbf{B}\), we have that \(\eta \mathbf{D}^{-}= \mathbf{L}-\mathbf{B}\). Taking the determinant, we find \(-\eta ^{2}=\det ( \mathbf{L}-\mathbf{B})\) from which (67) follows. Finally, by applying (1) twice, we obtain that

$$\begin{aligned} \det (\mathbf{L}-\mathbf{B}) &=\det \mathbf{L}-\mathbf{L}^{*}\cdot \mathbf{B}= \det \mathbf{L}-\frac{\det \mathbf{L}}{t} + \frac{\det \mathbf{L}-\mathbf{L} ^{*}\cdot (t\mathbf{B})}{t} \\ &=\frac{t-1}{t}\det \mathbf{L}+\frac{\det (\mathbf{L}-t\mathbf{B})}{t}, \end{aligned}$$

and recalling (66)1 we deduce (66)2.

We close this appendix by summarizing the results found:

$$ W^{**}(\mathbf{L})=Q(\mathbf{L})+\alpha ^{+}(\det \mathbf{L})^{+}+\alpha ^{-}( \det \mathbf{L})^{-}, $$

where

$$ \alpha ^{-}=\min_{\det \mathbf{L}=1}Q(\mathbf{L})=Q\bigl( \mathbf{D}^{+}\bigr), \qquad \alpha ^{+}=\min _{\det \mathbf{L}=-1}Q(\mathbf{L})=Q\bigl(\mathbf{D}^{-}\bigr). $$

If \(\det \mathbf{L}>0\), set

$$ \begin{aligned} \lambda &=\frac{\mathbf{L}^{*}\cdot \mathbf{D}^{-}}{2 \sqrt{\det \mathbf{L}}}, \quad \quad t=\frac{1}{2-2\lambda \bigl(\sqrt{\lambda ^{2}+1}-\lambda \bigr)}, \\ \eta &=\sqrt{\det \mathbf{L}}\bigl(\sqrt{\lambda ^{2}+1}-\lambda \bigr),\quad \quad \mathbf{B}=\mathbf{L}-\eta \mathbf{D}^{-}; \end{aligned} $$

while, if \(\det \mathbf{L}<0\), set

$$ \begin{aligned} \lambda &=\frac{\mathbf{L}^{*}\cdot \mathbf{D}^{+}}{2\sqrt{-\det \mathbf{L}}}, \quad \quad t=\frac{1}{2+2\lambda \bigl(\lambda -\sqrt{\lambda ^{2}+1}\bigr)}, \\ \eta &=\sqrt{-\det \mathbf{L}}\bigl(\lambda -\sqrt{\lambda ^{2}+1}\bigr),\quad \quad \mathbf{B}=\mathbf{L}-\eta \mathbf{D}^{+}. \end{aligned} $$

Then, with

$$ \mathbf{A}=\frac{\mathbf{L}-t\mathbf{B}}{1-t}, $$

we have that

$$ (1-t)\mathbf{A}+t\mathbf{B}=\mathbf{L}, \quad \mbox{and} \quad (1-t)Q( \mathbf{A})+t Q(\mathbf{B})=W^{**}(\mathbf{L}). $$

1.2 A.2 Evaluation of \(\alpha ^{\pm }\) for Various Symmetries

We first turn our attention to the computation of the constants \(\alpha ^{+}\) and \(\alpha ^{-}\) in the definition of the convexification \(W^{**}\) in (24). According to (25) we must minimize the quadratic form \(Q(\mathbf{L})\) over the manifolds \(\{\mathbf{L}: \det \mathbf{L}=\pm 1\}\). We accomplish this task through the method of Lagrange multipliers by seeking stationary points of the augmented functional\((\mathbf{L},\beta )\mapsto Q(\mathbf{L})+\beta (\det \mathbf{L}\pm 1)\). We note that the determinant of a 2-by-2 matrix is a quadratic form. Thus, there exists a fourth-order tensor \(\mathbb{E}\) such that

$$ \det \mathbf{L}=L_{11}L_{22}-L_{12}^{2}= \frac{1}{2}\mathbb{E}\mathbf{L} \cdot \mathbf{L}. $$
(69)

Thus, the augmented functional is

$$ (\mathbf{L},\beta )\mapsto \frac{1}{2} (\mathbb{D}+\beta \mathbb{E}) \mathbf{L}\cdot \mathbf{L}\mp \beta . $$
(70)

We now argue that the minima \(\alpha ^{\pm }\) have the following property:

$$ \mathbb{D} \mp \alpha ^{\mp }\mathbb{E} \text{ is singular}, $$
(71)

that is \(\alpha ^{\mp }\) are generalized eigenvalues of\(\mathbb{D}\)with respect to\(\mathbb{E}\). To verify the last statement, let us recall the definition of \(\mathbf{D}^{\pm }\), see (27), that is

$$ \mathbf{D}^{\pm }=\operatorname{argmin}_{\det \mathbf{L}=\pm 1} Q( \mathbf{L}). $$
(72)

Then there exist \(\beta ^{\mp }\) such that the pair \((\mathbf{D}^{\pm }, \beta ^{\mp })\) is a stationary point of the augmented functional defined in (70). The stationarity conditions for such pair are

$$\begin{aligned} &\bigl(\mathbb{D}+\beta ^{\mp }\mathbb{E}\bigr) \mathbf{D}^{\pm }=\mathbf{0}, \end{aligned}$$
(73a)
$$\begin{aligned} &\frac{1}{2}\mathbb{E}\mathbf{D}^{\pm }\cdot \mathbf{D}^{\pm }\mp 1=0. \end{aligned}$$
(73b)

It follows from the second stationarity condition that \(\mathbf{D}^{ \pm }\neq \mathbf{0}\). This implies that the tensor \(\mathbb{D}+ \beta ^{\mp }\mathbb{E}\) is singular. Moreover, the minimum \(\alpha ^{\mp }\) coincides with the Lagrange multiplier \(\beta ^{\mp }\), up to sign:

$$ \alpha ^{\mp }=Q\bigl(\mathbf{D}^{\pm }\bigr)=\frac{1}{2} \mathbb{D}\mathbf{D}^{ \pm }\cdot \mathbf{D}^{\pm }\stackrel{ \text{(73a)}}{=}-\frac{ \beta ^{\mp }}{2}\mathbb{E}\mathbf{D}^{\pm } \cdot \mathbf{D}^{\pm }\stackrel{ \text{(73b)}}{=}\mp \beta ^{\mp }. $$
(74)

To carry our calculation further on, we observe that

$$ \mathbb{D}\mathbf{L}\cdot \mathbf{L}= \begin{pmatrix} \mathbb{D}_{1111} & \mathbb{D}_{1122} & \mathbb{D}_{1112} \\ \mathbb{D}_{1122} & \mathbb{D}_{2222} & \mathbb{D}_{1222} \\ \mathbb{D}_{1112} & \mathbb{D}_{1222} & \mathbb{D}_{1212} \end{pmatrix} \begin{pmatrix} L_{11} \\ L_{22} \\ 2L_{12} \end{pmatrix} \cdot \begin{pmatrix} L_{11} \\ L_{22} \\ 2L_{12} \end{pmatrix} $$
(75)

and that

$$ \mathbb{E}\mathbf{L}\cdot \mathbf{L}= \begin{pmatrix} 0 & 1 & 0 \\ 1 & 0 & 0 \\ 0 & 0 & -1/2 \end{pmatrix} \begin{pmatrix} L_{11} \\ L_{22} \\ 2L_{12} \end{pmatrix} \cdot \begin{pmatrix} L_{11} \\ L_{22} \\ 2L_{12} \end{pmatrix} . $$
(76)

Thus, on denoting by d and e the matrices appearing in the definitions of the quadratic forms (75) and (76), we can write (71) as

det ( d α e ) =0.
(77)

Orthotropic Response

In the special case, when the material is orthotropic with respect to the basis \((\mathbf{e}_{1},\mathbf{e}_{2})\), we have

$$ \mathbb{D}_{1112}=0, \quad \text{and} \quad \mathbb{D}_{1222}=0 $$

(the other vanishing components of \(\mathbb{D}\) are obtained by application of the standard symmetries: \(\mathbb{D}_{\alpha \beta \gamma \delta }=\mathbb{D}_{\beta \alpha \gamma \delta }=\mathbb{D} _{\gamma \delta \alpha \beta }\)). In this case, the characteristic polynomial

p(β):=det(d+βe)= ( D 1212 β 2 ) ( D 1111 D 2222 ( D 1122 + β ) 2 )

has the following set of roots:

$$ \{-\sqrt{\mathbb{D}_{1111}\mathbb{D}_{2222}}- \mathbb{D}_{1122},\sqrt{ \mathbb{D}_{1111}\mathbb{D}_{2222}}- \mathbb{D}_{1122}, 2\mathbb{D} _{1212}\}. $$

The positive definiteness of \(\mathbb{D}\) implies that the first root is negative, and that the remaining roots are positive. The negative root must coincide with \(\beta ^{-}=-\alpha ^{-}\), that is

$$ \alpha ^{-}=\sqrt{\mathbb{D}_{1111}\mathbb{D}_{2222}}+ \mathbb{D}_{1122}. $$

On the other hand, the minimality of \(\alpha ^{+}\) yields

$$ \alpha ^{+}=\min \{\sqrt{\mathbb{D}_{1111} \mathbb{D}_{2222}}- \mathbb{D}_{1122}, 2\mathbb{D}_{1212} \}. $$

We note that the foregoing calculation can be used also when the plate is orthotropic with respect to a basis \((\mathbf{e}'_{1},\mathbf{e}'_{2})\), not necessarily coincident with \((\mathbf{e}_{1},\mathbf{e}_{2})\). In this case, the relevant strain energy is \(W'(\mathbf{L})=W(\mathbf{R}\mathbf{L} \mathbf{R}^{T})\), where \(\mathbf{R}=\mathbf{e}'_{\alpha }\otimes \mathbf{e} _{\alpha }\) is a rotation matrix. Then it can be easily shown that the convexification of \(Q'\) is given by

$$ \bigl(W'\bigr)^{**}(\mathbf{L})=W^{**}\bigl( \mathbf{R}\mathbf{L}\mathbf{R}^{T}\bigr). $$

Isotropic Response

For the isotropic strain energy \(Q(\mathbf{L})=d_{\mu }|\mathbf{L}|^{2}+\frac{d _{\lambda }}{2} (L_{11}+L_{22})^{2}\) the relevant components of \(\mathbb{D}\) are \(\mathbb{D}_{1111}=\mathbb{D}_{2222}=2d_{\mu }+d_{ \lambda }\), \(\mathbb{D}_{1122}=d_{\lambda }\), \(\mathbb{D}_{1212}=d _{\mu }\). Then

$$ \alpha ^{-}=2(d_{\mu }+d_{\lambda }), \qquad \alpha ^{+}=2d_{\mu }, $$

and so

$$\begin{aligned} W^{**}(\mathbf{L}) =& d_{\mu }|\mathbf{L}|^{2}+ \frac{d_{\lambda }}{2}(L_{11}+L_{22})^{2}+2d _{\mu }(\det \mathbf{L})^{+}+2(d_{\mu }+d_{\lambda }) (\det \mathbf{L})^{-} \\ =& d_{\mu }|\mathbf{L}|^{2}+\frac{d_{\lambda }}{2} | \mathbf{L}|^{2} +d_{ \lambda }\det \mathbf{L}+2d_{\mu }(\det \mathbf{L})^{+}+2(d_{\mu }+d_{ \lambda }) (\det \mathbf{L})^{-} \\ =& \biggl(d_{\mu }+\frac{d_{\lambda }}{2} \biggr) |\mathbf{L}|^{2} +2 d _{\mu }|\det \mathbf{L}| +d_{\lambda }\bigl(\det \mathbf{L}+2( \det \mathbf{L})^{-}\bigr) \\ =& \biggl(d_{\mu }+\frac{d_{\lambda }}{2} \biggr) \bigl(| \mathbf{L}|^{2} +2 |\det \mathbf{L}| \bigr). \end{aligned}$$

1.3 A.3 Compatibility Among Curvatures

The aim of this Appendix is to prove the following statement.

Let \(\omega , \omega _{\mathbf{A}}\) and \(\omega _{\mathbf{B}}\) be three two-dimensional open sets, with \(\omega = \omega _{\mathbf{A}}\cup \omega _{\mathbf{B}}\cup \varGamma \), where \(\varGamma =\overline{\omega }_{ \mathbf{A}}\cap \overline{\omega }_{\mathbf{B}}\) is a smooth curve and \(\overline{\omega }_{\mathbf{A}}\) denotes the closure of \(\omega _{ \mathbf{A}}\). Let \(\mathbf{A}\) and \(\mathbf{B}\), with \(\mathbf{A}\ne \mathbf{B}\), be two 2-by-2 symmetric matrices with null determinant. If there exists a deformation \(\mathbf{y}\) continuously differentiable on \(\omega \) and twice differentiable on \(\overline{\omega }_{\mathbf{A}}\) and on \(\overline{\omega }_{\mathbf{B}}\) such that

$$ \mathbf{K}_{\mathbf{y}}=\mathbf{A}\mbox{ on }\overline{\omega }_{\mathbf{A}}, \quad \mbox{and}\quad \mathbf{K}_{\mathbf{y}}=\mathbf{B} \mbox{ on }\overline{ \omega }_{\mathbf{B}}, $$

where \(\mathbf{K}_{\mathbf{y}}\) denotes the second fundamental form of \(\mathbf{y}\), then the curve \(\varGamma \) is straight and the matrices \(\mathbf{A}\) and \(\mathbf{B}\) are proportional, that is, there exists a constant \(\sigma \) for which either \(\mathbf{A}=\sigma \mathbf{B}\) or \(\mathbf{B}=\sigma \mathbf{A}\).

We assume that a deformation \(\mathbf{y}\) as described in the statement above exists and we study the consequences.

Since \(\mathbf{A}\ne \mathbf{B}\), we may assume without loss of generality that \(\mathbf{A}\neq \mathbf{0}\). Since \(\mathbf{y}\) is continuously differentiable we have that \(\nabla \mathbf{y}\) and the normal \(\boldsymbol{\nu }= \partial _{1}\mathbf{y}\wedge \partial _{2}\mathbf{y}\) are continuous on \(\omega \). Let \(\mathbf{t}\) be the unit tangent to the curve \(\varGamma \). Since \((\nabla \boldsymbol{\nu })\mathbf{t}= \partial _{\mathbf{t}}\boldsymbol{\nu }\) and since \(\mathbf{K}=- (\nabla \mathbf{y})^{\top }\nabla \boldsymbol{\nu }\) we have that

$$ \mathbf{A}\mathbf{t}= \mathbf{B}\mathbf{t},\quad \mbox{on } \varGamma . $$

Thus \(\det (\mathbf{A}-\mathbf{B})=0\), and the identity \(\det (\mathbf{A}- \mathbf{B})=\det \mathbf{A}-\mathbf{A}^{*}\cdot \mathbf{B}+ \det \mathbf{B}\) implies that

$$ \mathbf{A}^{*}\cdot \mathbf{B}=0. $$

Let \((\lambda ,\mathbf{a}_{1})\) and \((0,\mathbf{a}_{2})\) be the pairs of eigenvalues and eigenvectors of \(\mathbf{A}\), so that

$$ \mathbf{A}=\lambda \mathbf{a}_{1}\otimes \mathbf{a}_{1}, $$

with \(\lambda \neq 0\). Then \(\mathbf{A}^{*}= \lambda \mathbf{a}_{2}\otimes \mathbf{a}_{2}\) and \(\mathbf{A}^{*}\cdot \mathbf{B}=0 \) implies that \(\mathbf{B}\mathbf{a}_{2}\cdot \mathbf{a}_{2}=0\). We may therefore write

$$ \mathbf{B}=\beta \mathbf{a}_{1}\otimes \mathbf{a}_{1}+ \tilde{\beta }( \mathbf{a}_{1}\otimes \mathbf{a}_{2}+ \mathbf{a}_{2}\otimes \mathbf{a}_{1}), $$

and, since \(\det \mathbf{B}=0\), conclude that

$$ \mathbf{B}=\beta \mathbf{a}_{1}\otimes \mathbf{a}_{1}. $$

Setting \(\sigma =\beta /\lambda \) we obtain \(\mathbf{B}= \sigma \mathbf{A}\). Since \(\mathbf{A}\ne \mathbf{B}\), from the identity \(\mathbf{A}\mathbf{t}= \mathbf{B}\mathbf{t}\) we deduce that \(\mathbf{a}_{1}\cdot \mathbf{t}=0\), which implies that \(\mathbf{t}\) is constant and, in turn, that the curve \(\varGamma \) is a straight segment.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Paroni, R., Tomassetti, G. Macroscopic and Microscopic Behavior of Narrow Elastic Ribbons. J Elast 135, 409–433 (2019). https://doi.org/10.1007/s10659-018-09712-w

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10659-018-09712-w

Keywords

Mathematics Subject Classification

Navigation