Skip to main content

Advertisement

Log in

On the Global Stability of Compressible Elastic Cylinders in Tension

  • Published:
Journal of Elasticity Aims and scope Submit manuscript

Abstract

Consider a three-dimensional, homogeneous, compressible, hyperelastic body that occupies a cylindrical domain in its reference configuration. We identify a variety of hypotheses on the structure of the stored-energy function under which there exists an axisymmetric, homogeneous deformation that globally minimizes the energy. For certain classes of energy functions the uniqueness of this minimizer is also established. The primary boundary condition considered is the extension of the cylinder via the prescription of its deformed axial length, but the biaxial extension of the curved surface is also briefly considered. In particular, the results contained in this paper give conditions on the stored-energy function under which material instabilities, such as necking or the formation of shear bands, are not energy favorable.

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

Access this article

Subscribe and save

Springer+ Basic
$34.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Explore related subjects

Discover the latest articles, news and stories from top researchers in related subjects.

Notes

  1. If we denote the gradient of this homogeneous minimizer by \(\mathbf{F}_{0}\in\mathrm{M}_{+}^{3\times3}\), then it follows that the stored-energy function is quasiconvex at the boundary at (F 0,n), in the sense of Ball and Marsden [7], for any choice of unit vector n which is normal to some part of the lateral boundary ∂Ω×[0,L] of the cylinder \(\mathcal{C}\).

  2. Hill also allows m i <1, but such values are not covered by our results. See, also, Storåkers [35].

  3. Strict rank-one convexity is itself a consequence of the global strong ellipticity of the linearized equations, i.e., \(\mathbf{a}\otimes\mathbf{b}:\mathbb{C}(\mathbf {F})[\mathbf{a}\otimes\mathbf{b}]>0\) for every \(\mathbf{F}\in\mathrm{M}^{3\times3}_{+}\), \(\mathbf{a}\in \mathbb{R}^{3}\), and \(\mathbf{b}\in\mathbb{R}^{3}\) with a0b. Here \(\mathbb{C}(\mathbf {F}):=\mathrm{d}^{2}W/\mathrm{d}\mathbf{F}^{2}\).

  4. See, e.g., [20, p. 220] or [31, p. 408].

  5. Φ(a,b,c)=Φ(b,a,c)=Φ(c,b,a) and hence Φ,1(a,b,c)=Φ,2(b,a,c)=Φ,3(c,b,a).

  6. The fact that (3.33) implies ∇v is constant for vW 1,3 was first proven by Reshetnyak [27].

  7. The flat surfaces of the cylinder, z=−L and z=L, are left free.

  8. The existence of such a δ follows from the continuity of W together with the growth conditions (2.8). The uniqueness follows either from the strict rank-one convexity of W or, more simply, from the strict tension-extension inequality Φ,33>0.

References

  1. Antman, S.S.: Ordinary differential equations of nonlinear elasticity. II. Existence and regularity theory for conservative boundary value problem. Arch. Ration. Mech. Anal. 61, 353–393 (1976)

    MATH  MathSciNet  Google Scholar 

  2. Baker, M., Ericksen, J.L.: Inequalities restricting the form of the stress-deformation relations for isotropic elastic solids and Reiner-Rivlin fluids. J. Wash. Acad. Sci. 44, 33–35 (1954)

    MathSciNet  Google Scholar 

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

    Article  Google Scholar 

  4. Ball, J.M.: Constitutive inequalities and existence theorems in nonlinear elastostatics. In: Knops, R.J. (ed.) Nonlinear Analysis and Mechanics: Heriot-Watt Symposium, vol. I, pp. 187–241. Pitman, London (1977)

    Google Scholar 

  5. Ball, J.M.: Discontinuous equilibrium solutions and cavitation in nonlinear elasticity. Philos. Trans. R. Soc. Lond. A 306, 557–611 (1982)

    Article  ADS  MATH  Google Scholar 

  6. Ball, J.M.: Differentiability properties of symmetric and isotropic functions. Duke Math. J. 51, 699–728 (1984)

    Article  MATH  MathSciNet  Google Scholar 

  7. Ball, J.M., Marsden, J.E.: Quasiconvexity at the boundary, positivity of the second variation and elastic stability. Arch. Ration. Mech. Anal. 86, 251–277 (1984)

    Article  MATH  MathSciNet  Google Scholar 

  8. Ciarlet, P.G., Nečas, J.: Injectivity and self-contact in non-linear elasticity. Arch. Ration. Mech. Anal. 97, 171–188 (1987)

    Article  Google Scholar 

  9. Ciarlet, P.G.: Mathematical Elasticity, Vol. I. Elsevier, Amsterdam (1988)

    Google Scholar 

  10. Del Piero, G., Rizzoni, R.: Weak local minimizers in finite elasticity. J. Elast. 93, 203–244 (2008)

    Article  MATH  Google Scholar 

  11. Evans, L.C.: Partial Differential Equations, 2nd edn. American Mathematical Society, Providence (2010)

    MATH  Google Scholar 

  12. Fosdick, R., Foti, P., Fraddosio, A., Piccioni, M.D.: A lower bound estimate of the critical load for compressible elastic solids. Contin. Mech. Thermodyn. 22, 77–97 (2010)

    Article  ADS  MathSciNet  Google Scholar 

  13. Friesecke, G., James, R.D., Müller, S.: A theorem on geometric rigidity and the derivation of nonlinear plate theory from three-dimensional elasticity. Commun. Pure Appl. Math. 55, 1461–1506 (2002)

    Article  MATH  Google Scholar 

  14. Green, A.E., Adkins, J.E.: Large Elastic Deformations, 2nd edn. Oxford Clarendon Press, Oxford (1970)

    Google Scholar 

  15. Gurtin, M.E.: An Introduction to Continuum Mechanics. Academic Press, New York (1981)

    MATH  Google Scholar 

  16. Gurtin, M.E., Fried, E., Anand, L.: The Mechanics and Thermodynamics of Continua. Cambridge University Press, Cambridge (2010)

    Book  Google Scholar 

  17. Hill, R.: Aspects of invariance in solid mechanics. In: Yih, C.-S. (ed.) Advances in Applied Mechanics, vol. 18, pp. 1–75. Academic Press, New York (1978)

    Google Scholar 

  18. Hill, R., Hutchinson, J.W.: Bifurcation phenomena in the plane tension test. J. Mech. Phys. Solids 23, 239–264 (1975)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  19. Horgan, C.O., Murphy, J.G.: Constitutive models for almost incompressible isotropic elastic rubber-like materials. J. Elast. 87, 133–146 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  20. Iwaniec, T., Martin, G.: Geometric Function Theory and Non-linear Analysis. Oxford University Press, Oxford (2001)

    MATH  Google Scholar 

  21. Iwaniec, T., Onninen, J.: Hyperelastic deformations of smallest total energy. Arch. Ration. Mech. Anal. 194, 927–986 (2009)

    Article  MATH  MathSciNet  Google Scholar 

  22. Kaniel, S.: On the initial value problem for an incompressible fluid with nonlinear viscosity. J. Math. Mech. 19, 681–707 (1970)

    MATH  MathSciNet  Google Scholar 

  23. Mizel, V.J.: On the ubiquity of fracture in nonlinear elasticity. J. Elast. 52, 257–266 (1998)

    Article  MathSciNet  Google Scholar 

  24. Müller, S., Tang, Q., Yan, B.S.: On a new class of elastic deformation not allowing for cavitation. Anal. Non Linéaire 11, 217–243 (1994)

    Google Scholar 

  25. Murphy, J.G.: Strain energy functions for a Poisson power law function in simple tension of compressible hyperelastic materials. J. Elast. 60, 151–168 (2000)

    Article  MATH  Google Scholar 

  26. Ogden, R.W.: Non-Linear Elastic Deformations. Dover, Mineola (1984)

    Google Scholar 

  27. Reshetnyak, Y.G.: Liouville’s conformal mapping theorem under minimal regularity hypotheses. Sib. Math. J. 8, 631–634 (1967)

    Article  Google Scholar 

  28. Šilhavý, M.: The Mechanics and Thermodynamics of Continuous Media. Springer, Berlin (1997)

    MATH  Google Scholar 

  29. Šilhavý, M.: Differentiability properties of isotropic functions. Duke Math. J. 104, 367–373 (2000)

    Article  MathSciNet  Google Scholar 

  30. Šilhavý, M.: Differentiability properties of rotationally invariant functions. J. Elast. 58, 225–232 (2000)

    Article  MATH  Google Scholar 

  31. Sivaloganathan, J., Spector, S.J.: On the symmetry of energy-minimising deformations in nonlinear elasticity II: Compressible materials. Arch. Ration. Mech. Anal. 196, 395–431 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  32. Sivaloganathan, J., Spector, S.J.: On the global stability of two-dimensional, incompressible, elastic bars in uniaxial extension. Proc. R. Soc. Lond. A 466, 1167–1176 (2010)

    Article  ADS  MathSciNet  Google Scholar 

  33. Sivaloganathan, J., Spector, S.J.: On the stability of incompressible, elastic cylinders in uniaxial extension. J. Elast. 105, 313–330 (2011)

    Article  MathSciNet  Google Scholar 

  34. Spector, S.J.: On the absence of bifurcation for elastic bars in uniaxial tension. Arch. Ration. Mech. Anal. 85, 171–199 (1984)

    Article  MATH  MathSciNet  Google Scholar 

  35. Storåkers, B.: On material representation and constitutive branching in finite compressible elasticity. J. Mech. Phys. Solids 34, 125–145 (1986)

    Article  ADS  MATH  Google Scholar 

  36. Šverák, V.: Regularity properties of deformations with finite energy. Arch. Ration. Mech. Anal. 100, 105–127 (1988)

    Article  MATH  Google Scholar 

  37. Tang, Q.: Almost-everywhere injectivity in nonlinear elasticity. Proc. R. Soc. Edinb. 109A, 79–95 (1988)

    Google Scholar 

  38. Truesdell, C., Noll, W.: The Non-Linear Field Theories of Mechanics. Springer, Berlin (1964)

    Google Scholar 

  39. Vogel, T.I.: Stability of a liquid drop trapped between two parallel planes. SIAM J. Appl. Math. 47, 516–525 (1987)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  40. Wesołowski, Z.: Stability in some cases of tension in the light of the theory of finite strain. Arch. Mech. Stosow. 14, 875–900 (1962)

    MATH  Google Scholar 

  41. Wesołowski, Z.: The axially symmetric problem of stability loss of an elastic bar subject to tension. Arch. Mech. Stosow. 15, 383–394 (1963)

    MATH  Google Scholar 

Download references

Acknowledgements

The authors thank the referees and R. Fosdick for their helpful comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Scott J. Spector.

Additional information

This material is based upon work supported by the National Science Foundation under Grant No. DMS-1107899.

Appendices

Appendix A

We here gather a couple of the technical results used in this manuscript.

Proposition A.1

Fix \(N\in\mathbb{N}\). Let \({\varPsi}:(0,\infty)^{N+1}\to\mathbb{R}\) be (strictly) convex with

$$t_i\mapsto{\varPsi} (t_1,t_2, \ldots,t_i,\ldots ,t_{N},t_{N+1} ) $$

increasing for i=1,2,…,N. Suppose that v:(0,∞)→(0,∞)N with v i convex for i=1,2,…,N. Then

$$\psi(t):={\varPsi} \bigl(\mathbf{v}(t),t \bigr) \quad \textit{is}\ (\textit{strictly})\ \textit{convex}. $$

Proof

Let s,t∈(0,∞) and λ∈(0,1). Then the convexity of each component of v implies that

$$ v_i\bigl(\lambda t + (1-\lambda)s\bigr) \le\lambda v_i(t)+ (1-\lambda)v_i(s) \quad \mbox{for } i=1,2,\ldots,N. $$
(A.1)

Next, (A.1) and the monotonicity of Ψ in its first N-arguments yields

$$ \begin{aligned}[b] \psi \bigl(\lambda t + (1-\lambda)s \bigr)&= {\varPsi} \bigl(\mathbf{v}\bigl(\lambda t + (1-\lambda)s\bigr),\lambda t + (1-\lambda)s \bigr) \\ & \le {\varPsi} \bigl(\lambda\mathbf{v}(t)+(1-\lambda)\mathbf {v}(s),\lambda t + (1-\lambda)s \bigr). \end{aligned} $$
(A.2)

Finally, the strict convexity of Ψ gives us

$$\begin{aligned} {\varPsi} \bigl(\lambda\mathbf{v}(t)+(1-\lambda) \mathbf {v}(s),\lambda t + (1-\lambda)s \bigr) &< \lambda{\varPsi} \bigl( \mathbf{v}(t),t \bigr)+ (1-\lambda){\varPsi} \bigl(\mathbf{v}(s),s \bigr) \\ &= \lambda\psi (t )+ (1-\lambda)\psi (s ), \end{aligned} $$

which together with (A.2) yields the strict convexity of ψ. □

Remark A.2

If N=1 and Ψ is independent of its second argument then Proposition A.1 reduces to the classical result that the composition of 2 convex functions is convex whenever the outer function is also increasing.

A proof of the following well-known result on convex functions can be found in, for example, [21, Lemma 2.1] or [31, Lemma A.1].

Lemma A.3

Let \(\vartheta:\mathbb{R}^{+}\to\mathbb{R}\) be (strictly) convex. Then

$$Q(s,t)= t\vartheta \biggl(\frac{s}{t} \biggr) $$

is (strictly) convex on \(\mathbb{R}^{+}\times\mathbb{R}^{+}\).

Appendix B: The Principal Stretches; Other Invariants

In this section we show that slight variants of our method will allow for a variety of constitutive relations that depend on the principal stretches. The key to these results is the following classical lemma, a proof of which can be found in the appendix of [33]. (See [22, p. 693] for a proof in the case ω(t)=t p.)

Lemma B.1

Let P∈M3×3 be symmetric and strictly positive definite with eigenvalues λ 1, λ 2, and λ 3. Suppose that \(\omega:(0,\infty)\to\mathbb{R}\) is convex. Then

$$\omega(\lambda_1)+\omega(\lambda_2)+\omega( \lambda_3) \ge \omega(\mathbf{f}_1\cdot\mathbf{P} \mathbf{f}_1)+ \omega(\mathbf {f}_2\cdot\mathbf{P} \mathbf{f}_2)+ \omega(\mathbf{f}_3\cdot\mathbf{P} \mathbf{f}_3) $$

for any orthonormal basis {f 1,f 2,f 3} of \(\mathbb{R}^{3}\).

We also made use of a slight generalization of the arithmetic-geometric mean inequality. Again see, e.g., the appendix of [33] for a proof.

Lemma B.2

Let \(\omega:(0,\infty)\to\mathbb{R}\) be monotone increasing and convex. Then, for every \(\mathbf{a}\in\mathbb{R}^{n}\) with a i >0, i=1,2,…,n,

$$\sum_{i=1}^n \omega \bigl([a_i]^n \bigr) \ge n\omega \Biggl(\prod _{i=1}^n a_i \Biggr). $$

Moreover, this inequality is strict if either ω is strictly monotone or strictly convex unless all of the a i are equal.

Let q≥1 and r≥1. Suppose that \(\phi,\psi:(0,\infty)\to\mathbb{R}\) are monotone increasing with ψ convex and \(s\mapsto\phi(\sqrt{s^{q}})\) convex. In this section we will consider the energy functions

$$\begin{aligned} {\varLambda}(\mathbf{F})& := \phi \bigl(\alpha^q \bigr) +\phi \bigl(\beta^q \bigr) +\phi \bigl(\gamma^q \bigr) \quad \mbox{with}\ p_{\mathrm{max}}:=q\le2, \end{aligned}$$
(B.1)
$$\begin{aligned} {\varLambda}(\mathbf{F})& := \psi \biggl( \biggl[\frac{\alpha\beta}{\gamma} \biggr]^{q} \biggr) +\psi \biggl( \biggl[\frac{\beta\gamma}{\alpha} \biggr]^{q} \biggr) +\psi \biggl( \biggl[\frac{\gamma\alpha}{\beta} \biggr]^{q} \biggr) \quad \mbox{with}\ p_{\mathrm{max}}:=q, \end{aligned}$$
(B.2)
$$\begin{aligned} {\varLambda}(\mathbf{F})& := \bigl(\alpha^{2r}+\beta^{2r} + \gamma^{2r} \bigr)^{q/(2r)} \quad \mbox{with}\ p_{\mathrm{max}}:=\min \{q, 3r\}, \end{aligned}$$
(B.3)
$$\begin{aligned} {\varLambda}(\mathbf{F})& := \biggl[ \biggl(\frac{\alpha\beta}{\gamma} \biggr)^{r} + \biggl(\frac{\beta\gamma}{\alpha} \biggr)^{r} + \biggl(\frac{\gamma\alpha}{\beta} \biggr)^{r} \biggr]^{q/r}\quad \mbox{with } p_{\mathrm{max}}:=\min\{q, 3r\}, \end{aligned}$$
(B.4)

where α, β, and γ are the eigenvalues of \(\mathbf{U}=\sqrt{\mathbf{F}^{\mathrm{T}}\mathbf{F}}\) and \(\mathbf{F}\in\mathrm{M}^{3\times3}_{+}\).

The main result of this section is the following. The proof is contained in the subsections that follow.

Proposition B.3

Suppose that Λ is given by (B.1), (B.2), (B.3), or (B.4). Let λ>0, \(\mathbf{u}\in{\mathcal{A}_{\lambda}}\), and p∈[1,p max]. Define μ=μ(λ,p,u)>0 to be the unique positive real number that satisfies (3.2). Then

(B.5)

Moreover, if μλ then

(B.6)

where \({\mathbf{u}_{\lambda}^{\mu}}\) is given by (3.1) with ν:=μ. If, in addition, ϕ is strictly increasing and p>1, then (B.6) will be a strict inequality unless u∈PS λ .

Remark B.4

  1. 1.

    Since \(\omega(s):=\phi(\sqrt{s^{q}})\) in (B.1) is convex and increasing with q∈[1,2], so is ϕ(t)=ω(t 2/q) (see Remark A.2).

  2. 2.

    For \(\phi\in C^{2}((0,\infty);\mathbb{R})\), the hypothesis \(s\mapsto\phi(\sqrt{s^{q}})\) is convex is equivalent to

    $$t\phi''(t)\ge \biggl(1-\frac{q}{2} \biggr)\phi'(t)\quad \mbox{for all } t>0. $$
  3. 3.

    Suppose that the energy is given by (B.1) with q>2. Then one can rewrite (B.1) as

    $${\varLambda}(\mathbf{F})=\hat{\phi} \bigl(\alpha^2 \bigr) +\hat{ \phi} \bigl(\beta^2 \bigr) +\hat{\phi} \bigl(\gamma^2 \bigr), \qquad\hat{\phi}(t):=\phi \bigl(t^{q/2} \bigr), $$

    where \(\omega(s):=\hat{\phi}(s^{2/q})=\phi(s)\) is increasing and convex. Thus the restriction q≤2 is not essential.

  4. 4.

    If r=1 the energy (B.3) is the same as the energy in Sect. 3.1.

  5. 5.

    If r=1 the energy (B.4) can also be written:

    $${\varLambda}(\mathbf{F})= \biggl[\frac{|\mathrm{adj}\,\mathbf{F}|^2}{\det \mathbf{F}} \biggr]^{q}. $$

In order to establish the existence of a homogeneous minimizer we will also need the following result.

Proposition B.5

Suppose that Λ is given by (B.1), (B.2), (B.3), or (B.4). Let p∈[1,p max]. Then \(\sigma(t):={\varLambda}(\sqrt[p]{t}\mathbf{I})\) is increasing and convex.

Proof

If Λ is given by (B.1), then σ(t):=3ϕ(t q/p), which is monotone increasing and convex since qp>0 and ϕ is increasing and convex (see Remark A.2). If Λ is given by (B.2), then σ(t):=3ψ(t q/p), which is increasing and convex since qp>0 and ψ is increasing and convex. If Λ is given by (B.3), then σ(t):=3q/(2r) t q/p, which is increasing and convex since qp>0. Finally, if Λ is given by (B.4), then σ(t):=3q/r t q/p, which is increasing and convex since qp>0. □

2.1 B.1 The Energy Λ(F)=ϕ(α q)+ϕ(β q)+ϕ(γ q) for q∈[1,2]

In this subsection we prove Proposition B.3 when Λ is given by (B.1).

Proof of Proposition B.3: Case 1

Let Λ be given by (B.1) with q∈[1,2], where \(\phi\in C^{1}(\mathbb{R}^{+};\mathbb{R})\) is increasing and \(s\mapsto\phi(\sqrt{s^{q}})\) is convex. Fix λ>0, \(\mathbf{u}\in{\mathcal{A}_{\lambda}}\), and p∈[1,q]. Let μ be given by (3.2). Define F=F(x,y,z):=∇u(x,y,z), J u :=det∇u(x,y,z), and \(\mathbf{U}:=\sqrt{\mathbf{F}^{\mathrm{T}}\mathbf{F}}\) with eigenvalues α, β, and γ. Then, since qp>0, Hölder’s inequality and (3.2) yield

(B.7)

Next, since ϕ is increasing and convex, Lemma B.2, Jensen’s inequality, (B.1), (B.7), and the identity J u =αβγ imply that

which establishes (B.5).

Now assume that μλ. Define \(\omega:\mathbb{R}^{+}\to\mathbb{R}\) by \(\omega(s):=\phi (\sqrt{s^{q}})\) so that ω is convex. If we then apply Lemma B.1 with P=F T F (with eigenvalues α 2, β 2, and γ 2) we conclude that, for any \((x,y,z)\in\mathcal{C}\),

$$\begin{aligned} \omega\bigl(\alpha^2\bigr)+\omega\bigl( \beta^2\bigr)+\omega\bigl(\gamma^2\bigr) & \ge \omega( \mathbf{e}_z\cdot\mathbf{P}\mathbf{e}_z)+ \omega( \mathbf {e}_x\cdot\mathbf{P}\mathbf{e}_x)+ \omega( \mathbf{e}_y\cdot\mathbf{P}\mathbf{e}_y) \\ & = \omega \bigl(|\mathbf{F}\mathbf{e}_z|^2 \bigr)+ \omega \bigl(|\mathbf{F}\mathbf{e}_x|^2 \bigr)+ \omega \bigl(|\mathbf{F}\mathbf{e}_y|^2 \bigr) \end{aligned} $$

or, equivalently,

$$ {\varLambda}(\nabla\mathbf{u})=\phi \bigl(\alpha^q \bigr) +\phi \bigl(\beta^q \bigr) +\phi \bigl(\gamma^q \bigr) \ge \phi \bigl(|\mathbf{u}_z|^q \bigr)+ \phi \bigl(|\mathbf {u}_x|^q \bigr)+ \phi \bigl(| \mathbf{u}_y|^q \bigr). $$
(B.8)

Next, (B.8) together with (3.6), Lemma B.2, and the monotonicity of ϕ yield

$$ \begin{aligned}[b] {\varLambda}(\nabla\mathbf{u}) &\ge \phi \bigl(|\mathbf{u}_z|^q \bigr)+ 2\phi \bigl(\sqrt{|\mathbf{u}_x|^q|\mathbf{u}_y|^q} \bigr) \\ &\ge \phi \bigl(|\mathbf{u}_z|^q \bigr)+ 2\phi \bigl(\sqrt{|\mathbf{u}_x\times\mathbf{u}_y|^q} \bigr) \\ &\ge \phi \bigl(|\mathbf{u}_z|^q \bigr)+ 2\phi \biggl( \sqrt{\frac{ [J_\mathbf{u} ]^q}{|\mathbf{u}_z|^q}} \biggr). \end{aligned} $$
(B.9)

Define

$$H(s,t):= \phi\bigl(s^{q/p}\bigr) +2\phi \biggl(\frac{t^{3q/(2p)}}{s^{q/(2p)}} \biggr) =\phi\bigl(s^{q/p}\bigr) +2\phi \biggl( \biggl[s \biggl( \frac{t}{s} \biggr)^{3/2} \biggr]^{q/p} \biggr). $$

Then, in view of (B.9),

$$\begin{aligned} {\varLambda}(\nabla\mathbf{u}) \ge H \bigl(| \mathbf{u}_z|^p, [J_\mathbf{u} ]^{p/3} \bigr). \end{aligned} $$

Since ϕ is convex (see Remark B.4.1) and increasing it follows that \(H:\mathbb{R}^{+}\times\mathbb{R}^{+}\to \mathbb{R}\) is convex (see Lemma A.3 and Remark A.2). Moreover, since ϕ′ is positive and increasing, it follows that if t<s, then

$$\begin{aligned} \frac{p}{q} s^{[1-q/p]}\frac{\partial H}{\partial s} &= \phi' \bigl(s^{q/p} \bigr) - \biggl[ \frac{t}{s} \biggr]^{3q/(2p)} \phi' \biggl( \frac{t^{3q/(2p)}}{s^{q/(2p)}} \biggr) \\ &\ge \phi' \bigl(s^{q/p} \bigr) - \phi' \biggl( \biggl[s \biggl( \frac{t}{s} \biggr)^{3/2} \biggr]^{q/p} \biggr) \\ &\ge \phi' \bigl(s^{q/p} \bigr) - \phi' \bigl(s^{q/p} \bigr)\ge0; \end{aligned} $$

thus, for any t>0, the mapping sH(s,t) is increasing on [t,∞). The remainder of the proof of (B.6), when μλ, is then similar to the proof of (3.3) in Proposition 3.1. □

2.2 B.2 The Energy Λ(F)=ψ(α q β q/γ q)+ψ(β q γ q/α q)+ψ(γ q α q/β q) with q≥1

In this subsection we prove Proposition B.3 when Λ is given by (B.2).

Proof of Proposition B.3: Case 2

Let Λ be given by (B.2) with q≥1, where \(\psi\in C^{1}(\mathbb{R}^{+};\mathbb{R})\) is increasing and convex. Fix λ>0, \(\mathbf{u}\in{\mathcal{A}_{\lambda}}\), and p∈[1,q]. Let μ be given by (3.2). Define F=F(x,y,z):=∇u(x,y,z), J u :=det∇u(x,y,z), and \(\mathbf {U}:=\sqrt{\mathbf{F}^{\mathrm{T}}\mathbf{F}}\), with eigenvalues α, β, and γ.

We first note that, since ψ is increasing and convex, Lemma B.2, Jensen’s inequality, (B.7), and (B.2) imply that

which establishes (B.5).

Now suppose that μλ. Note that \(\mathbf{P}:=(\operatorname {adj}\mathbf{F})(\operatorname {adj}\mathbf{F})^{\mathrm{T}}/(\det \mathbf{F})\) is symmetric and strictly positive definite with eigenvalues αβ/γ, βγ/α, and γα/β. Then, in view of (B.2), Lemma B.1 (with ω(t)=ψ(t q)) yields, for any \((x,y,z)\in\mathcal{C}\),

$$ \begin{aligned}[b] {\varLambda}(\nabla\mathbf{u})&= \psi \biggl( \biggl[\frac{\alpha\beta}{\gamma} \biggr]^{q} \biggr) +\psi \biggl( \biggl[\frac{\beta\gamma}{\alpha} \biggr]^{q} \biggr) +\psi \biggl( \biggl[\frac{\gamma\alpha}{\beta} \biggr]^{q} \biggr) \\ & \ge \psi \bigl([\mathbf{e}_z\cdot\mathbf{P} \mathbf{e}_z]^q \bigr) + \psi \bigl([ \mathbf{e}_x\cdot\mathbf{P}\mathbf{e}_x]^q \bigr) + \psi \bigl([\mathbf{e}_y\cdot\mathbf{P} \mathbf{e}_y]^q \bigr) \\ & = \psi \biggl(\frac{|(\operatorname {adj}\mathbf{F})^\mathrm{T}\mathbf {e}_z|^{2q}}{[J_\mathbf{u}]^q} \biggr)+ \psi \biggl(\frac{|(\operatorname {adj}\mathbf{F})^\mathrm{T}\mathbf {e}_x|^{2q}}{[J_\mathbf{u}]^q} \biggr)+ \psi \biggl(\frac{|(\operatorname {adj}\mathbf{F})^\mathrm{T}\mathbf {e}_y|^{2q}}{[J_\mathbf{u}]^q} \biggr), \end{aligned} $$
(B.10)

where J u :=det∇u=αβγ. As in the proof in Sect. B.1, Lemma B.2 now implies that

$$ \begin{aligned}[b] \psi \biggl(\frac{|(\operatorname {adj}\mathbf{F})^\mathrm{T}\mathbf {e}_x|^{2q}}{[J_\mathbf{u}]^q} \biggr)+ \psi \biggl(\frac{|(\operatorname {adj}\mathbf{F})^\mathrm{T}\mathbf {e}_y|^{2q}}{[J_\mathbf{u}]^q} \biggr) \ge 2\psi \biggl( \frac{|(\operatorname {adj}\mathbf{F})^\mathrm{T}\mathbf{e}_x \times (\operatorname {adj}\mathbf{F})^\mathrm{T}\mathbf{e}_y|^q}{[J_\mathbf {u}]^q} \biggr). \end{aligned} $$
(B.11)

However,

$$ \begin{aligned} &(\operatorname {adj}\mathbf{F})^\mathrm{T} \mathbf{e}_x\times(\operatorname {adj}\mathbf {F})^\mathrm{T} \mathbf{e}_y = (\det\mathbf{F}) \mathbf{F}(\mathbf{e}_x \times\mathbf{e}_y) = [J_\mathbf{u}] \mathbf{F} \mathbf{e}_z, \\ &J_\mathbf{u}=\bigl[(\operatorname {adj}\mathbf{F})\mathbf{F}\bigr]^\mathrm{T} \mathbf {e}_z\cdot\mathbf{e}_z = (\operatorname {adj}\mathbf{F})^\mathrm{T}\mathbf{e}_z\cdot\mathbf{F}\mathbf {e}_z\le\bigl|(\operatorname {adj}\mathbf{F})^\mathrm{T}\mathbf{e}_z\bigr|| \mathbf {F}\mathbf{e}_z|, \end{aligned} $$
(B.12)

and Fe z =u z .

If we now combine (B.10), (B.11), and (B.12) we find, with the aid of the monotonicity of ψ, that

$$ \begin{aligned}[b] {\varLambda}(\nabla\mathbf{u}) & \ge \psi \biggl(\frac{[J_\mathbf{u}]^q}{|\mathbf{u}_z|^{2q}} \biggr)+ 2 \psi \bigl(|\mathbf{u}_z|^q \bigr). \end{aligned} $$
(B.13)

Define

$$H(s,t):= \psi \biggl(\frac{t^{3q/p}}{s^{2q/p}} \biggr) +2\psi \bigl(s^{q/p} \bigr) =\psi \biggl( \biggl[ s \biggl(\frac{t}{s} \biggr)^{3} \biggr]^{q/p} \biggr) +2\psi \bigl(s^{q/p} \bigr). $$

Then, in view of (B.13),

$$\begin{aligned} {\varLambda}(\nabla\mathbf{u}) \ge H \bigl(| \mathbf{u}_z|^p, [J_\mathbf{u} ]^{p/3} \bigr). \end{aligned} $$

Since ψ is convex and increasing and qp>0, it follows that \(H:\mathbb{R}^{+}\times\mathbb{R}^{+}\to\mathbb{R}\) is convex (see Lemma A.3 and Remark A.2). Moreover, since ψ′ is positive and increasing, it follows that if t<s, then

$$\begin{aligned} \frac{p}{2q}s^{[1-p/q]}\frac{\partial H}{\partial s} &= \psi' \bigl(s^{q/p} \bigr) - \biggl[ \frac{t}{s} \biggr]^{3q/p} \psi' \biggl( \frac{t^{3q/p}}{s^{2q/p}} \biggr) \\ &\ge \psi' \bigl(s^{q/p} \bigr) - \psi' \biggl( \biggl[ s \biggl(\frac{t}{s} \biggr)^{3} \biggr]^{q/p} \biggr) \\ &\ge \psi' \bigl(s^{q/p} \bigr) - \psi' \bigl(s^{q/p} \bigr)\ge0; \end{aligned} $$

thus, for any t>0, the mapping sH(s,t) is increasing on [t,∞). The remainder of the proof of (B.6) is then similar to the proof of (3.3) in Proposition 3.1. □

2.3 B.3 The Energy Λ(F)=(α 2r+β 2r+γ 2r)q/(2r) with q≥1 and r≥1

In this subsection we prove Proposition B.3 when Λ is given by (B.3).

Let Λ be given by (B.3) with r≥1 and q≥1. Fix λ>0, \(\mathbf{u}\in{\mathcal{A}_{\lambda}}\), and p≥1 with pq and p≤3r. Let μ be given by (3.2). Define F=F(x,y,z):=∇u(x,y,z), J u :=det∇u(x,y,z), and \(\mathbf{U}:=\sqrt{\mathbf{F}^{\mathrm{T}}\mathbf{F}}\), with eigenvalues α, β, and γ. Then Lemma B.2, (B.7), and (B.3) yield

which establishes (B.5).

Now suppose that μλ. If we then apply Lemma B.1 with ω(t)=t r and P=F T F (with eigenvalues α 2, β 2, and γ 2) we conclude that, for any \((x,y,z)\in\mathcal{C}\),

$$\begin{aligned} \alpha^{2r}+\beta^{2r}+ \gamma^{2r} & \ge (\mathbf{e}_z\cdot\mathbf{P} \mathbf{e}_z)^{r}+ (\mathbf{e}_x\cdot \mathbf{P}\mathbf{e}_x)^{r}+ (\mathbf{e}_y \cdot\mathbf{P}\mathbf{e}_y)^{r} \\ & = |\mathbf{F}\mathbf{e}_z|^{2r}+ |\mathbf{F}\mathbf {e}_x|^{2r}+|\mathbf{F}\mathbf{e}_y|^{2r} \end{aligned} $$

or, equivalently,

$$ \bigl[{\varLambda}(\nabla\mathbf{u}) \bigr]^{2r/q} = \alpha^{2r}+\beta^{2r}+\gamma^{2r} \ge | \mathbf{u}_z|^{2r}+ |\mathbf{u}_x|^{2r}+| \mathbf{u}_y|^{2r}. $$
(B.14)

Next, (B.14) together with (3.6), the arithmetic-geometric mean inequality, and the monotonicity of tt r yield

$$\begin{aligned} \bigl[{\varLambda}(\nabla\mathbf{u}) \bigr]^{2r/q} &\ge |\mathbf{u}_z|^{2r}+ 2 \bigl(| \mathbf{u}_x||\mathbf{u}_y| \bigr)^{r} \\ &\ge |\mathbf{u}_z|^{2r}+ 2|\mathbf{u}_x \times\mathbf{u}_y|^{r} \\ &\ge |\mathbf{u}_z|^{2r}+ 2 \biggl(\frac{J_\mathbf{u}}{|\mathbf{u}_z|} \biggr)^{r} \\ &= |\mathbf{u}_z|^{2r} \biggl(1+2\frac{|J_\mathbf{u}|^{r}}{|\mathbf{u}_z|^{3r}} \biggr) \end{aligned} $$

and hence

$$ \begin{aligned}[b] \bigl[{\varLambda}(\nabla\mathbf{u}) \bigr]^{p/q} &\ge |\mathbf{u}_z|^p \biggl(1+2 \biggl[\frac{|J_\mathbf{u}|^{p/3}}{ |\mathbf{u}_z|^p} \biggr]^{3r/p} \biggr)^{p/(2r)}. \end{aligned} $$
(B.15)

Define

$$\theta(s):= \bigl(1+2s^{3r/p} \bigr)^{p/(2r)}, \qquad G(s,t):= s \theta \biggl(\frac{t}{s} \biggr), \qquad H(s,t):= G(s,t)^{q/p} $$

and note that (B.15) implies

$${\varLambda}(\nabla\mathbf{u}) \ge H \bigl(|\mathbf{u}_z|^p, |J_\mathbf{u}|^{p/3} \bigr). $$

Now

$$\theta''(s)=\frac{3}{p} s^{[-2+3r/p]} \bigl(1+2s^{3r/p} \bigr)^{[-2+p/(2r)]} \bigl[ps^{3r/p}+(3r-p) \bigr]>0, $$

for s>0 and 3rp. Consequently, Lemma A.3 implies that G and hence H=G q/p is strictly convex. Moreover,

$$\frac{\partial G}{\partial s} = \theta (\tau )-\tau\theta' (r ) = \bigl(1-\tau^{3r/p} \bigr) \bigl(1+2\tau^{3r/p} \bigr)^{[-1+p/(2r)]}> 0 $$

for τ:=t/s<1; thus for any t>0 the mappings sG(s,t) and sH(s,t) are strictly increasing on [t,∞). The remainder of the proof of (B.6) is then similar to the proof of (3.3) in Proposition 3.1.  □

2.4 B.4 The Energy Λ(F)=[(αβ/γ)r+(βγ/α)r+(γα/β)r]q/r with q≥1 and r≥1

In this subsection we prove Proposition B.3 when Λ is given by (B.4).

Proof of Proposition B.3: Case 4

Let Λ be given by (B.4) with r≥1 and q≥1. Fix λ>0, \(\mathbf{u}\in{\mathcal {A}_{\lambda}}\), and p≥1 with pq and p≤3r. Let μ be given by (3.2). Define F=F(x,y,z):=∇u(x,y,z), J u :=det∇u(x,y,z), and \(\mathbf {U}:=\sqrt{\mathbf{F}^{\mathrm{T}}\mathbf{F}}\), with eigenvalues α, β, and γ. Then Lemma B.2, (B.7), and (B.4) yield

which establishes (B.5).

Now suppose that μλ. Then, if we apply Lemma B.1 with ω(t)=t r and \(\mathbf{P}:=(\operatorname {adj}\mathbf{F})(\operatorname {adj}\mathbf{F})^{\mathrm{T}}/(\det \mathbf{F})\) (with eigenvalues αβ/γ, βγ/α, and γα/β) we conclude, with the aid of (B.4), that for any \((x,y,z)\in\mathcal{C}\)

$$ \begin{aligned}[b] \bigl[{\varLambda}(\nabla\mathbf{u}) \bigr]^{r/q} & \ge (\mathbf{e}_z\cdot\mathbf{P} \mathbf{e}_z)^{r}+ (\mathbf{e}_x\cdot \mathbf{P}\mathbf{e}_x)^{r}+ (\mathbf{e}_y \cdot\mathbf{P}\mathbf{e}_y)^{r} \\ & = \frac{|(\operatorname {adj}\mathbf{F})^\mathrm{T}\mathbf{e}_z|^{2r}}{|J_\mathbf{u}|^{r}} + \frac{|(\operatorname {adj}\mathbf{F})^\mathrm{T}\mathbf{e}_x|^{2r}}{|J_\mathbf{u}|^{r}}+ \frac{|(\operatorname {adj}\mathbf{F})^\mathrm{T}\mathbf{e}_y|^{2r}}{|J_\mathbf{u}|^{r}}, \end{aligned} $$
(B.16)

where J u :=det∇u=αβγ. The arithmetic-geometric mean inequality now implies that

$$ \begin{aligned}[b] \bigl|(\operatorname {adj}\mathbf{F})^\mathrm{T} \mathbf{e}_x\bigr|^{2r}+ \bigl|(\operatorname {adj}\mathbf{F})^\mathrm{T} \mathbf{e}_y\bigr|^{2r} \ge 2\bigl\vert (\operatorname {adj}\mathbf{F})^\mathrm{T}\mathbf{e}_x \times (\operatorname {adj}\mathbf{F})^\mathrm{T}\mathbf{e}_y\bigr\vert ^{r}. \end{aligned} $$
(B.17)

However,

$$ \begin{aligned} & (\operatorname {adj}\mathbf{F})^\mathrm{T} \mathbf{e}_x\times(\operatorname {adj}\mathbf {F})^\mathrm{T} \mathbf{e}_y = (\det\mathbf{F}) \mathbf{F}(\mathbf{e}_x \times\mathbf{e}_y) = [J_\mathbf{u}] \mathbf{F} \mathbf{e}_z, \\ &J_\mathbf{u}=\bigl[\mathbf{F}(\operatorname {adj}\mathbf{F})\bigr]^\mathrm{T} \mathbf {e}_z\cdot\mathbf{e}_z = (\operatorname {adj}\mathbf{F})^\mathrm{T}\mathbf{e}_z\cdot\mathbf{F}\mathbf {e}_z\le\bigl|(\operatorname {adj}\mathbf{F})^\mathrm{T}\mathbf{e}_z\bigr|| \mathbf {F}\mathbf{e}_z|, \end{aligned} $$
(B.18)

and Fe z =u z .

If we now combine (B.16), (B.17), and (B.18) we find, with the aid of the monotonicity of tt r, that

$$ \begin{aligned}[b] \bigl[{\varLambda}(\nabla\mathbf{u}) \bigr]^{p/q} & \ge \biggl(\frac{|J_\mathbf{u}|^{r}}{|\mathbf{u}_z|^{2r}}+ 2 |\mathbf{u}_z|^{r} \biggr)^{p/r} \\ &= |\mathbf{u}_z|^{p} \biggl(\frac{|J_\mathbf{u}|^{r}}{ |\mathbf{u}_z|^{3r}} +2 \biggr)^{p/r} \\ &= |\mathbf{u}_z|^{p} \biggl(2+ \biggl[ \frac{|J_\mathbf{u}|^{p/3}}{ |\mathbf{u}_z|^{p}} \biggr]^{3r/p} \biggr)^{p/r}. \end{aligned} $$
(B.19)

Define

$$\theta(s):= \bigl(2+s^{3r/p} \bigr)^{p/r}, \qquad G(s,t):= s \theta \biggl(\frac{t}{s} \biggr), \qquad H(s,t):=G(s,t)^{q/p} $$

so that (B.19) implies

$${\varLambda}(\nabla\mathbf{u}) \ge H \bigl(|\mathbf{u}_z|^p, |J_\mathbf{u}|^{p/3} \bigr). $$

Then

$$\theta''(s)=\frac{6}{p} s^{[-2+3r/p]} \bigl(2+s^{3r/p} \bigr)^{[-2+p/r]} \bigl[ps^{3r/p}+(3r-p) \bigr]>0 $$

for s>0 and 3rp and hence, in view of Lemma A.3, H is strictly convex. Moreover,

$$\frac{\partial H}{\partial s} = \theta (\tau )-\tau\theta' (\tau ) = 2 \bigl(1-\tau^{3r/p} \bigr) \bigl(2+\tau^{3r/p} \bigr)^{[-1+p/r]}> 0 $$

for τ:=t/s<1; thus for any t>0 the mappings sG(s,t) and hence sH(s,t) are strictly increasing on [t,∞). The remainder of the proof of (B.6) is then similar to the proof of (3.3) in Proposition 3.1. □

Appendix C: The Homogeneity of Energy-Minimizing Deformations II

Fix \(N\in\mathbb{N}\) and define \({\boldsymbol{\varLambda}}:\mathrm {M}^{3\times3}_{+}\to\mathbb{R}^{N}\) by

$$ {\boldsymbol{\varLambda}}(\mathbf{F}) := \bigl({ \varLambda}_1(\mathbf{F}),{\varLambda}_2(\mathbf {F}),{ \varLambda}_3(\mathbf{F}),\ldots,{\varLambda}_N( \mathbf{F}) \bigr), $$
(C.1)

where each component \({\varLambda}_{i}:\mathrm{M}^{3\times3}_{+}\to \mathbb{R}\) is one of the functions, Λ(F), given in this section and \(p^{i}_{\mathrm{max}}\) is the corresponding maximum value of the parameter p.

Lemma C.1

Let Λ be given by (C.1). Suppose that the stored energy \(W\in C^{1}( \mathrm{M}_{+}^{3\times3};[0,\infty))\) satisfies

$$W(\mathbf{F})={\varPsi}\bigl({\boldsymbol{\varLambda}}(\mathbf{F}), (\det \mathbf{F})^{p/3}\bigr), $$

where p∈[1,p max] with \(p_{\mathrm{max}}:=\min\{p^{i}_{\mathrm{max}}, 1\le i \le n \}\), \({\varPsi}: (0,\infty)^{N+1}\to\mathbb{R}\) is (strictly) convex, and Λ is given by (C.1). Then

$$t \mapsto W\bigl(\sqrt[p]{t}\,\mathbf{I}\bigr)\quad \textit{is}\ (\textit{strictly})\ \textit{convex on}\ \mathbb{R}^+. $$

Moreover, if in addition the reference configuration is stress free, then

$$\lambda\mapsto W(\lambda\mathbf{I})\quad \textit{is} (\textit{strictly})\ \textit{increasing on}\ [1, \infty). $$

Proof

To obtain convexity define

$$\xi(t) := W\bigl(\sqrt[p]{t}\,\mathbf{I}\bigr)={\varPsi} \bigl({\boldsymbol{ \varLambda }}\bigl(\sqrt[p]{t}\,\mathbf{I}\bigr),t \bigr). $$

Then Proposition B.5 together with Proposition A.1 yield the (strict) convexity of ξ.

Next, if we take the derivative of ξ we find that

$$\dot{\xi}(t)= \frac{\sqrt[p]{t}}{pt} \,\mathbf{S}\bigl(\sqrt[p]{t}\,\mathbf{I}\bigr) :\mathbf{I}, $$

where S is the Piola-Kirchhoff stress tensor (2.6). Thus, if the reference configuration is stress free, \(\dot{\xi}(1)=0\). The (strict) convexity of ξ then implies that ξ and hence the mapping λW(λ I) is (strictly) increasing on [1,∞). □

3.1 C.1 Energy Reduction by Symmetrization

Theorem C.2

Assume that the reference configuration is stress free. Suppose that the stored energy \(W\in C^{1}( \mathrm{M}_{+}^{3\times3};[0,\infty))\) satisfies

$$ W(\mathbf{F})={\varPsi} \bigl({\boldsymbol{\varLambda}}(\mathbf{F}), (\det\mathbf{F})^{p/3} \bigr), $$
(C.2)

where p∈[1,p max] with \(p_{\mathrm{max}}:=\min\{p^{1}_{\mathrm{max}},p^{2}_{\mathrm {max}},p^{3}_{\mathrm{max}},\ldots,p^{N}_{\mathrm{max}}\}\), Λ is given by (C.1), and \({\varPsi}:(0,\infty)^{N+1}\to\mathbb{R}\) is increasing in each of its first N-arguments and convex. Let λ≥1 and \(\mathbf{u}\in{\mathcal{A}_{\lambda}}\). Define μ=μ(λ,p,u)>0 to be the unique positive real number that satisfies (3.2). Then

(C.3)

where \({\mathbf{u}_{\lambda}^{\nu}}\) is given by (3.1) with ν:=μ if μλ and ν:=λ if μλ. Moreover, if μλ, p>1, and Ψ is strictly increasing in one of its first N-arguments, then (C.3) is a strict inequality unless u∈PS λ .

Proof

Fix λ≥1, \(\mathbf{u}\in{\mathcal{A}_{\lambda}}\), and define μ by (3.2). If we now let F=∇u in (C.2), integrate over the cylinder \(\mathcal{C}\), and apply Jensen’s inequality to the convex function Ψ we conclude, with the aid of (3.2), that

(C.4)

Now suppose that μλ. Then we can make use of Proposition B.3, the monotonicity of Ψ in each of its first N-arguments, (3.1), and (C.2), to get

(C.5)

which when combined with (C.4) yields the desired inequality, (C.3) with ν=μ. Moreover, if Ψ is strictly increasing in, say, its i-argument, then the inequality in (C.5) is a strict inequality unless

Proposition B.3 now yields u∈PS λ when p>1.

Finally, suppose that μλ so that ν=λ. Then we can make use of Proposition B.3, the monotonicity of Ψ in each of its first N-arguments, (3.1), and (C.2), to obtain

(C.6)

and we have made use of (C.2). However, Lemma C.1 then implies that W(η I)≥W(λ I), which together with (C.4) and (C.6) yields the desired result, (C.3) with ν=λ. □

3.2 C.2 Existence of a Homogeneous Minimizer

Theorem C.3

Let W, Ψ, Λ, and p satisfy the hypotheses of Theorem  C.2. Suppose, in addition, that

$$ \lim_{|\mathbf{F}|\to\infty}W(\mathbf{F}) = +\infty, \qquad \lim _{\det\mathbf{F}\to0^+}W(\mathbf{F}) = +\infty. $$
(C.7)

Then for each λ≥1 there exists a κ=κ(λ)>0 such that the deformation

$$ {\mathbf{u}_\lambda^{\kappa}}(x,y,z) := \begin{bmatrix} \kappa x \\ \kappa y \\ \lambda z \end{bmatrix} $$
(C.8)

is an absolute minimizer of the energy among deformations in \({\mathcal{A}_{\lambda}}\).

Proof

Fix λ≥1. Then (C.7) and Proposition 2.5 yield α>0 and β>0 such that

$$ \varPhi(s,t,\lambda) \ge \varPhi(\alpha,\beta,\lambda) \quad \mbox{for every}\ s>0\ \mbox{and}\ t>0. $$
(C.9)

Now apply Theorem C.2 with u=u , where

$$\mathbf{u}^*(x,y,z)= \begin{bmatrix} \alpha x \\ \beta y \\ \lambda z \end{bmatrix} , $$

to get a κ:=ν>0 such that

$$ \varPhi(\alpha,\beta,\lambda)=\bar{E}\bigl(\mathbf{u}^*\bigr) \ge \bar{E}\bigl({\mathbf{u}_\lambda^{\kappa}}\bigr)= \varPhi(\kappa, \kappa,\lambda), $$
(C.10)

where we have made use of (3.12) and (C.8).

We claim that the homogeneous deformation given by (C.8) with this value of κ is an absolute minimizer of E among deformations in \({\mathcal{A}_{\lambda}}\). To see this, let \(\mathbf{u}\in{\mathcal{A}_{\lambda}}\). Then, by Theorem C.2, there is ν>0 such that the homogeneous deformation \({\mathbf{u}_{\lambda}^{\nu}}\) given by (3.1) satisfies (see (3.12))

$$ \bar{E}(\mathbf{u})\ge\bar{E}\bigl({\mathbf{u}_\lambda^{\nu}} \bigr)= \varPhi(\nu,\nu,\lambda). $$
(C.11)

The desired minimality of \({\mathbf{u}_{\lambda}^{\kappa}}\) now follows from (C.10), (C.11), and (C.9) with s=t=ν. □

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sivaloganathan, J., Spector, S.J. On the Global Stability of Compressible Elastic Cylinders in Tension. J Elast 120, 161–195 (2015). https://doi.org/10.1007/s10659-014-9510-5

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10659-014-9510-5

Keywords

Mathematics Subject Classification (2010)

Navigation