Skip to main content
Log in

Stability and Localization of Deformation Delay in Finitely Strained Plates at Arbitrary Strain-Rates

  • Published:
Journal of Elasticity Aims and scope Submit manuscript

Abstract

Of interest here is the stability and associated deformation localization of structures when inertial effects are considered. Concerned primarily with the study of necking failure patterns, the prevailing approach in the relevant literature uses the “modal analysis” method to find the wavelength of the structure’s fastest growing eigenmode, an approach that often uses a rate-dependent material response. However, the experimental studies of (Zhang and Ravi-Chandar in Int. J. Fracture 142: 183–217, 2006; Zhang and Ravi-Chandar in Int. J. Fracture 163: 41–65, 2010) on the high strain-rate expansion of thin rings and tubes, show no evidence of a dominant wavelength in their failure mode and no influence of strain-rate sensitivity on the necking strains. Moreover, modal analysis assumes that at all times the entire structure sees the applied eigenmode perturbation in spite of the physical limitation of a finite wave propagation speed. In addition, the closely related problem of stability in dynamically loaded structures, i.e., the time evolution of perturbations introduced at different times during loading, does not seem to have attracted attention.

Based on the above-mentioned experimental and theoretical observations, (Ravi-Chandar and Triantafyllidis in Int. J. Solids Struct. 58: 301–308, 2015) proposed a “localized perturbation” approach to study the dynamic stability of an incompressible, nonlinearly elastic bar at different strain-rates by following the evolution of spatially localized small perturbations introduced at different times. The goal of the present work is to study the dynamic stability – linear and nonlinear – of rate-independent biaxially strained thin plates by following the evolution of spatially localized perturbations introduced at different times, to understand the initiation of the corresponding failure mechanisms. Our 2D linearized analysis of a thin plate under plane stress state, shows that these plates are stable until \(\tau _{L}\), the dimensionless limit time corresponding to the loss of the uniformly strained plate’s stability. This result is supported by fully nonlinear calculations.

Our nonlinear numerical calculations also show an imperfection amplitude-dependent and biaxiality-dependent delay in the appearance of localization patterns in dynamically loaded plates for dimensionless times well beyond \(\tau _{m}\), corresponding to the onset of loss of ellipticity in the constitutive law. Moreover, the failure patterns of these plates are studied numerically by following the time evolution of randomly distributed imperfections of different amplitudes.

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
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

Notes

  1. Strain rates investigated were of order up to \(10^{4} \sec^{-1}\).

  2. Plate is unstable under the “transient growth criterion” used in fluid mechanics, e.g., see [25].

  3. Henceforth, Greek indexes range from 1 to 2 while Latin indexes range from 1 to 3.

  4. The reference mass density \(\rho _{0}\) is per unit reference area.

  5. Henceforth a superimposed dot denotes time differentiation at a fixed material point initially at \(\mathbf{X}\), i.e., \(\dot{f} \equiv \partial f(\mathbf{X},t) / \partial t\).

  6. From symmetry considerations \(A_{\alpha \gamma} (\mathbf{n}) = A_{\alpha \gamma} (-\mathbf{n})\).

  7. In the problem investigated here we also have \(\delta _{-} = \int _{0}^{\tau _{m}} \nu _{-}\left ( \phi _{m}, \tau \right ) \text{d}\tau \).

  8. Hereinafter, for simplicity the same symbols, \(u\) and \(v\), are used for the dimensionless quantities as for their dimensioned counterparts while a superimposed dot denotes differentiation with respect to \(\tau \).

  9. Hereinafter a superscript \((^{0})\) or subscript \((_{0})\) denotes evaluation of a quantity on the uniform strain principal solution.

  10. Linearized stability is studied here only for a hyperelastic material; for the other constitutive laws only numerical calculations are available, which nevertheless confirm our findings for the hyperelastic case.

  11. The matrix \({\mathbf{C}}\) has all eigenvalues with negative real part iff \((\omega /\eta )^{2}{\mathbf{A}}-{\dot {\mathbf {A}}}{\mathbf {A}}^{-1}\) is positive definite (e.g., see [2]).

  12. This inequality is verified by the numerical calculations in Appendix C.

  13. Best approximation of uniaxial stress under strain control, corresponding to an initial uniaxial stress \(\sigma _{1}\ne 0,\; \sigma _{2}=0\). From (A.16) one has \(\epsilon _{2}=-\epsilon _{1}/2\); for small strain \(\epsilon _{i} \approx \lambda _{i} -1\), resulting in \(\psi =-\pi /12\) according to (2.1).

  14. A norm choice involving both displacements and velocities could be used but the choice made here is judged more physically appropriate.

  15. A deformation theory type of plasticity based on logarithmic strain in the absence of unloading.

  16. Localization of deformation requires that the constitutive law looses ellipticity at some strain level.

  17. Note: for a uniaxial stress state \(\epsilon _{2} = -\epsilon _{1} / 2\) and \(\epsilon _{e} = \epsilon \); Moreover \(\epsilon _{y}\) and \(\sigma _{y} = E \epsilon _{y}\) are the yield strain and stress respectively in a uniaxial loading path.

  18. For the value of \(\epsilon _{y}=0.002\) used in the calculations of Sect. 3, the equivalent strain verifies \(\epsilon _{e}(\tau _{m}) > \epsilon _{y}\).

References

  1. Abeyaratne, R., Triantafyllidis, N.: The emergence of shear bands in plane strain. Int. J. Solids Struct. 17, 1113–1134 (1981)

    Article  MATH  Google Scholar 

  2. Bernstein, D.S.: Matrix Mathematics; Theory, Facts, and Formulas. Princeton University Press, Princeton (2009)

    Book  MATH  Google Scholar 

  3. Brush, D., Almroth, B.: Buckling of Bars, Plates, and Shells. McGraw-Hill, New York (1975)

    Book  MATH  Google Scholar 

  4. Budiansky, B., Hutchinson, J.: Dynamic buckling of imperfection sensitive structures. Proc. XI Int. Cong. Appl. Mech., pp. 636–651. (1964)

    Google Scholar 

  5. Clifton, R.J., Duffy, J., Hartley, K.A., Shawki, T.G.: On critical conditions for shear bar formation at high strain rates. Scr. Metall. 18, 443–448 (1984)

    Article  Google Scholar 

  6. Dequiedt, J.L.: Statistics of dynamic fragmentation for a necking instability. Int. J. Solids Struct. 75, 32–44 (2015)

    Article  Google Scholar 

  7. El Mai, S., Mercier, S., Petit, J., Molinari, A.: An extension of the linear stability analysis for the prediction of multiple necking during dynamic extension of round bar. Int. J. Solids Struct. 51, 3491–3507 (2014)

    Article  Google Scholar 

  8. Fressengeas, C., Molinari, A.: Inertia and thermal effects on the localization of plastic flow. Acta Metall. Mater. 33, 387–396 (1985)

    Article  Google Scholar 

  9. Fressengeas, C., Molinari, A.: Fragmentation of rapidly stretching sheets. Eur. J. Mech. A, Solids 13(2), 251 (1994)

    MATH  Google Scholar 

  10. Guduru, P., Freund, L.: The dynamics of multiple neck formation and fragmentation in high rate extension of ductile materials. Int. J. Solids Struct. 39, 5615–5632 (2002)

    Article  MATH  Google Scholar 

  11. Guduru, P.R., Bharathi, M.S., Freund, L.: The influence of a surface coating on the high-rate fragmentation of a ductile material. Int. J. Fract. 137(1–4), 89–108 (2006)

    Article  MATH  Google Scholar 

  12. Jouve, D.: Analytic study of the onset of plastic necking instabilities during biaxial tension tests on metallic plates. Eur. J. Mech. A, Solids 50, 59–69 (2015)

    Article  MATH  Google Scholar 

  13. Koiter, W.T.: The Stability of Elastic Equilibrium (in Dutch) English Translation: Technical Report No. AFFDL-TR-70-25, sponsored by Lockheed Missiles and Space Co., February 1970 Ph.D. thesis. Technische Hooge School Delft (1945)

  14. Koning, C., Taub, J.: Impact buckling of thin bars in the elastic range hinged at both ends. Luftfahrt-Forsch. 10, 55–64 (1933)

    Google Scholar 

  15. Mainy, A.: Dynamic buckling of thin metallic rings under external pressure. Master’s thesis, University of Texas (2012)

  16. Mercier, S., Granier, N., Molinari, A., Liorca, F., Buy, F.: Multiple necking during the dynamic expansion of hemispherical metallic shells, from experiments to modelling. J. Mech. Phys. Solids 58, 955–982 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  17. Mercier, S., Molinari, A.: Predictions of bifurcations and instabilities during dynamic extensions. Int. J. Solids Struct. 40, 1995–2016 (2003)

    Article  MATH  Google Scholar 

  18. Molnár, G., Gravouil, A., Seghir, R., Réthoré, J.: An open-source abaqus implementation of the phase-field method to study the effect of plasticity on the instantaneous fracture toughness in dynamic crack propagation. Comput. Methods Appl. Mech. Eng. 364, 11304 (2020)

    MathSciNet  MATH  Google Scholar 

  19. Putelat, T., Triantafyllidis, N.: Dynamic stability of externally pressurized elastic rings subjected to high rates of loading. Int. J. Solids Struct. 51, 1–12 (2014)

    Article  Google Scholar 

  20. Ravi-Chandar, K., Triantafyllidis, N.: Dynamic stability of a bar under high loading rate: response to local perturbations. Int. J. Solids Struct. 58, 301–308 (2015)

    Article  Google Scholar 

  21. Rodrigues-Martinez, J.A., Vadillo, G., Fernandez-Saez, J., Molinari, A.: Identification of the critical wavelength responsible for the fragmentation of ductile rings expanding at very high strain rates. J. Mech. Phys. Solids 61, 1357–1376 (2013)

    Article  MathSciNet  Google Scholar 

  22. Rodríguez-Martínez, J.A., Molinari, A., Zaera, R., Fernández-Sáez, J.: The critical neck spacing in ductile plates subjected to dynamic biaxial loading: on the interplay between loading path and inertia effects. Int. J. Solids Struct. 108, 74–84 (2017)

    Article  Google Scholar 

  23. Shenoy, V., Freund, L.: Necking bifurcations during high strain rate extension. J. Mech. Phys. Solids 47, 2209–2233 (1999)

    Article  MATH  Google Scholar 

  24. Stören, S., Rice, J.R.: Localized necking in thin sheets. J. Mech. Phys. Solids 23, 421–441 (1975)

    Article  MATH  Google Scholar 

  25. Trefethen, L.N., Trefethen, A.E., Reddy, S.C., Driscoll, T.A.: Hydrodynamic stability without eigenvalues. Science 261, 578–584 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  26. Triantafyllidis, N.: Bifurcation phenomena in pure bending. J. Mech. Phys. Solids 28, 221–245 (1980)

    Article  MathSciNet  MATH  Google Scholar 

  27. Vaz-Romero, A., Mercier, S., Rodrigues-Martinez, J.A., Molinari, A.: A comparative study of the dynamic fragmentation of non-linear elastic and elasto-plastic rings: the roles of stored elastic energy and plastic dissipation. Mech. Mater. 132, 134–148 (2019)

    Article  Google Scholar 

  28. Vaz-Romero, A., Rodríguez-Martínez, J.A., Mercier, S., Molinari, A.: Multiple necking pattern in nonlinear elastic bars subjected to dynamic stretching: the role of defects and inertia. Int. J. Solids Struct. 125, 232–243 (2017)

    Article  Google Scholar 

  29. Wen, G., Triantafyllidis, N.: Dynamic stability of biaxially strained thin sheets under high strain-rates: response to local perturbations. Int. J. Fract. 200, 99–113 (2016)

    Article  Google Scholar 

  30. Xue, Z., Vaziri, A., Hutchinson, J.: Material aspects of dynamic neck retardation. J. Mech. Phys. Solids 56, 93–113 (2008)

    Article  MATH  Google Scholar 

  31. Zaera, R.: Collective behaviour and spacing of necks in ductile plates subjected to dynamic biaxial loading. J. Mech. Phys. Solids 85, 245–269 (2015)

    Article  MathSciNet  Google Scholar 

  32. Zhang, H., Ravi-Chandar, K.: On the dynamics of necking and fragmentation – I. Real-time and post-mortem observations in Al 6061-O. Int. J. Fract. 142, 183–217 (2006)

    Article  Google Scholar 

  33. Zhang, H., Ravi-Chandar, K.: On the dynamics of necking and fragmentation – II. Effect of material properties, geometrical constraints and absolute size. Int. J. Fract. 150, 3–36 (2008)

    Article  MATH  Google Scholar 

  34. Zhang, H., Ravi-Chandar, K.: On the dynamics of necking and fragmentation – IV. Expansion of al 6061-o tubes. Int. J. Fract. 163, 41–65 (2010)

    Article  MATH  Google Scholar 

  35. Zhou, F., Molinari, J., Ramesh, K.T.: An elastic–visco-plastic analysis of ductile expanding ring. Int. J. Impact Eng. 33, 880–891 (2006)

    Article  Google Scholar 

Download references

Acknowledgements

This paper is dedicated to the memory Prof. J. Ericksen. The work was initiated in part by grants from École Polytechnique and the CNRS (Centre National de Recherche Scientifique) during the AY 2011-2012, during the second author’s sabbatical leave as a Distinguished Visiting Professor of the École Polytechnique, in residence at the École’s Solid Mechanics Laboratory (LMS). This work continued during several subsequent visits, made possible through the generosity of LMS and the École Polytechnique and was finalized during the June 2022 visit of Prof. R. S. Elliott, also supported by the LMS. We also acknowledge some very helpful discussions on dynamic systems stability with applications in fluid mechanics with Prof. L. Guin from École Polytechnique.

Author information

Authors and Affiliations

Authors

Contributions

G.W. performed the numerical simulations and prepared the figures. N.T., R.E., and K.R. wrote the main manuscript text and contributed to the problem formulation and results analysis. All authors reviewed the manuscript.

Corresponding author

Correspondence to R. S. Elliott.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

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

Appendices

Appendix A: Constitutive Laws Adopted and Their Plane Stress Incremental Moduli

Although the expressions for the 2D plane stress incremental moduli are stated in [29], their derivation from 3D considerations is interesting and hence, for the sake of completeness of the presentation, the details of the corresponding derivations are given in this Appendix.

The analysis presented in Sect. 2 is general; any rate-independent constitutive lawFootnote 16 which can be put in the form of (2.5), can be accommodated. Results presented here correspond to the three such models: a hyperelastic deformation theory model of plasticity, the \(J_{2}\) deformation theory model of [24] and a finite strain generalization of the \(J_{2}\) flow theory. All models are fitted to the same power law uniaxial stress-strain curve and thus share the same principal solution. For further details on this construction, the interested reader is referred to [26]. For the small perturbations considered here, deviations from proportional loading are initially small and no unloading occurs in the perturbed plate until well after \(\tau _{m}\) (the time corresponding to the loss of ellipticity onset in the unperturbed – uniform strain – solution), thus justifying their use for analyzing the plate’s stability.

Assuming material incompressibility, the [24] \(J_{2}\) deformation and the \(J_{2}\) flow theory models that can be put in the following 3D rate form

$$ \overset{\triangledown}{\boldsymbol{\sigma}} = \boldsymbol{C} : \boldsymbol{D} - \dot{p} \boldsymbol{I} , $$
(A.1)

where \(\overset{\triangledown}{(\phantom{A})}\) denotes the Jaumann rate of the Cauchy stress tensor (\(\overset{\triangledown}{\boldsymbol{\sigma}} = \dot{\boldsymbol{\sigma}} - \boldsymbol{\Omega} \cdot \boldsymbol{\sigma} + \boldsymbol{\sigma} \cdot \boldsymbol{\Omega}\) with \(\boldsymbol{\Omega}\) the spin tensor), \(\boldsymbol{D}\) is the strain rate tensor and \(\dot{p}\) the hydrostatic pressure rate.

For the case of finite strains, the above 3D current configuration relation can be transformed into its reference configuration counterpart

$$ \dot{S}_{\mathit{ij}} = L_{\mathit{ijkl}} \dot{E}_{\mathit{kl}} - \dot{p} C_{\mathit{ij}}^{-1} ;\quad E_{\mathit{ij}} = \frac{1}{2} \big( C_{\mathit{ij}} - \delta _{\mathit{ij}} \big) , $$
(A.2)

where \(\boldsymbol{S}\) is the second Piola-Kirchhoff stress, \(\boldsymbol{E}\) its work-conjugate Green-Lagrange strain and the reference configuration components of the incremental moduli tensor \(\boldsymbol{L}\) are

$$ \begin{aligned}[b] L_{\mathit{ijkl}} \equiv \,&\frac{2}{3} E^{*} \Big[\frac{1}{2}\big(C^{-1}_{\mathit{ik}} C^{-1}_{\mathit{jl}} + C^{-1}_{\mathit{il}} C^{-1}_{\mathit{jk}}\big) - \frac{3}{2} \Big(1 - \frac{E_{t}}{E^{*}}\Big) \frac{S^{\prime}_{\mathit{ij}}S^{\prime}_{\mathit{kl}}}{\sigma _{e}^{2}}\Big] \\ &{}- \frac{1}{2} \Big[C^{-1}_{\mathit{ik}} S_{\mathit{jl}} + C^{-1}_{\mathit{jk}} S_{\mathit{il}} + C^{-1}_{\mathit{il}} S_{\mathit{jk}} + C^{-1}_{\mathit{jl}} S_{\mathit{ik}} \Big] , \end{aligned} $$
(A.3)

where \(\boldsymbol{S}^{\prime}\) is the deviatoric part of the stress tensor \(\boldsymbol{S}\) and \(\sigma _{e}\) the von Mises equivalent stress, namely

$$ S^{\prime}_{\mathit{ij}} = S_{\mathit{ij}} - \frac{1}{3} C^{-1}_{\mathit{ij}} C_{\mathit{kl}} S_{\mathit{kl}}, \quad \sigma _{e}^{2} = \frac{3}{2} C_{\mathit{ik}} C_{\mathit{jl}} S^{\prime}_{\mathit{ij}} S^{ \prime}_{\mathit{kl}} . $$
(A.4)

In the above expressions \(E^{*} = E\) for the \(J_{2}\) flow theory while for the \(J_{2}\) deformation theory of [24] \(E^{*} = E_{s} \equiv \sigma _{e} / \epsilon _{e}\) is the secant modulus of the uniaxial stress-strain curve. In both models \(E_{t}\) is the tangent modulus of the uniaxial stress-strain curve \(E_{t} \equiv \text{d}\sigma _{e} / \text{d} \epsilon _{e}\). For the \(J_{2}\) hyperelastic model the general expression for \(E^{*}\) is more complicated and will be given below for the special case of the biaxially stretched membrane.

The principal axes expressions in three dimensions for the (von Mises) equivalent stress \(\sigma _{e}\) and the equivalent strain \(\epsilon _{e}\), required for the calculation of \(E_{s}\) and \(E_{t}\) are

$$ \sigma _{e} = \big( \sigma _{1}^{2} + \sigma _{2}^{2} + \sigma _{3}^{2} - \sigma _{1} \sigma _{2} - \sigma _{2} \sigma _{3} - \sigma _{3} \sigma _{1} \big) ^{1/2} ,\quad \epsilon _{e} = \frac{2}{3} \big( \epsilon _{1}^{2} + \epsilon _{2}^{2} + \epsilon _{3}^{2} - \epsilon _{1} \epsilon _{2} - \epsilon _{2} \epsilon _{3} - \epsilon _{3} \epsilon _{1} \big) ^{1/2} . $$
(A.5)

Due to plane stress loading conditions

$$ \dot{S}_{3i} = 0 ,\quad \dot{E}_{\alpha 3} = 0 ,\quad \dot{E}_{33} = - C_{33} C^{-1}_{\gamma \delta} \dot{E}_{\gamma \delta} , $$
(A.6)

which substituted in (A.2) give the following relation between \(\dot{S}_{\alpha \beta}\) and its work conjugate \(\dot{E}_{\alpha \beta}\)

$$ \dot{S}_{\alpha \beta} = {\mathcal{M}}_{\alpha \beta \gamma \delta} \dot{E}_{\gamma \delta} ; \quad {\mathcal{M}}_{\alpha \beta \gamma \delta} = L_{\alpha \beta \gamma \delta} - C_{33} \big( L_{\alpha \beta 33} C^{-1}_{ \gamma \delta} + C^{-1}_{\alpha \beta} L_{33 \gamma \delta} \big) + C^{-1}_{ \alpha \beta} C^{-1}_{\gamma \delta} C_{33}^{2} L_{3333} . $$
(A.7)

The above result, combined with (A.3) gives the following expression for \(M_{\alpha \beta \gamma \delta}\)

$$ \begin{aligned}[b] {\mathcal{M}}_{\alpha \beta \gamma \delta} = \,&\frac{2}{3} E^{*} \Big[ \frac{1}{2} \big(C^{-1}_{\alpha \gamma} C^{-1}_{\beta \delta} + C^{-1}_{ \alpha \delta} C^{-1}_{\beta \gamma} \big) + C^{-1}_{\alpha \beta} C^{-1}_{ \gamma \delta} - \frac{3}{2} \Big(1 - \frac{E_{t}}{E^{*}}\Big) \frac{S^{\prime}_{\alpha \beta}S^{\prime}_{\gamma \delta}}{\sigma _{e}^{2}} \Big] \\ &{}- \frac{1}{2} \Big[C^{-1}_{\alpha \gamma} S_{\beta \delta} + C^{-1}_{ \beta \gamma} S_{\alpha \delta} + C^{-1}_{\alpha \delta} S_{\beta \gamma} + C^{-1}_{\beta \delta} S_{\alpha \gamma} \Big] . \end{aligned} $$
(A.8)

Recalling the relations between \(\boldsymbol{S}\) & \(\boldsymbol{N}\) and \(\boldsymbol{E}\) & \(\boldsymbol{F}\), the plane stress incremental moduli \(\boldsymbol{\mathcal{L}}\) in (2.5) are found to be

$$ \mathcal{L}_{\alpha \beta \gamma \delta} = {\mathcal {M}}_{\epsilon \beta \zeta \delta} F_{\alpha \epsilon} F_{\gamma \zeta} + S_{\delta \beta} \delta _{\alpha \gamma} . $$
(A.9)

For the case of biaxial loading of interest here \(F_{\alpha \beta} = \operatorname{diag} \big( \lambda _{\gamma }\big)\) and thus

$$ C_{11} = \lambda _{1}^{2} , \quad C_{22} = \lambda _{2}^{2} , \quad C_{12} = 0 ; \quad S_{11} = \sigma _{1} / \lambda _{1}^{2} , \quad S_{22} = \sigma _{2} / \lambda _{2}^{2} , \quad S_{12} = 0 , $$
(A.10)

which upon substitution into (A.8), (A.9) gives the incremental moduli expressions in (A.16), (A.17) for \(J_{2}\) flow and \(J_{2}\) deformation theories.

Calculations of the moduli for the hyperelastic model use (A.9), (A.10) and \(\boldsymbol{\mathcal{M}}\) is derived from the strain energy potential \(W\)

$$ {\mathcal {M}}_{\alpha \beta \gamma \delta} = \frac{\partial ^{2} W}{\partial E_{\alpha \beta} \partial E_{\gamma \delta}} = 4 \frac{\partial ^{2} W}{\partial C_{\alpha \beta} \partial C_{\gamma \delta}} , $$
(A.11)

obtained by successive application of chain rule of differentiation using \(W(\epsilon _{e})\) where the equivalent strain \(\epsilon _{e} = \epsilon _{e}(I_{1}, I_{2})\) is expressed in terms of the invariants of \(\boldsymbol{C}\), which in turn depend on the principal stretch ratios \(\lambda _{\alpha}\) through

$$ I_{1} = \operatorname{tr} \boldsymbol{C} = \lambda _{1}^{2} + \lambda _{2}^{2} ,\quad I_{2} = \det \boldsymbol{C} = \big( \lambda _{1} \lambda _{2} \big)^{2} . $$
(A.12)

The 3D strain energy density \(W\) used here is a function of the equivalent strain \(\epsilon _{e}\); the isotropic model is fitted to a piecewise power law uniaxial stress-strain curveFootnote 17 resulting in the following expressions for \(W\)

$$ \frac{\sigma _{e}}{\sigma _{y}} = \Big( \frac{\epsilon _{e}}{\epsilon _{y}} \Big)^{m} \; ; \quad W = E\big( \epsilon _{y}\big)^{2} \bigg[ \frac{1}{1+m} \Big( \frac{\epsilon _{e}}{\epsilon _{y}} \Big)^{m+1} + \frac{1}{2} \Big( \frac{m-1}{m+1} \Big) \bigg] , \quad \left \{ \begin{aligned} & m= 1 \; \text{for} \; \epsilon _{e} \leq \epsilon _{y} , \\ & m = n \; \text{for} \; \epsilon _{e} > \epsilon _{y} , \end{aligned} \right . $$
(A.13)

where the equivalent strain \(\epsilon _{e}\) is given in terms of the principal logarithmic strain components \(\epsilon _{\alpha}\) by (A.5). Accounting for the incompressibility constraint \(\epsilon _{1}+\epsilon _{2}+\epsilon _{3}=0\) (A.5) one obtains

$$ \sigma _{e} = \big(\sigma _{1}^{2} + \sigma _{2}^{2} - \sigma _{1} \sigma _{2} \big)^{1/2} \; , \quad \epsilon _{e} = \frac{2}{\sqrt{3}} \Big[ \epsilon _{1}^{2} + \epsilon _{2}^{2} + \epsilon _{1} \epsilon _{2} \Big] ^{1/2} \; , $$
(A.14)

where the equivalent stress \(\sigma _{e} \) expression accounts for the plane stress condition \(\sigma _{3}=0\) and the equivalent strain accounts for incompressibility \(\epsilon _{1}+\epsilon _{2}+\epsilon _{3}=0\). Since the principal solution is biaxial straining, the principal Cauchy stresses \(\sigma _{\alpha}\) are related to their work-conjugate the principal logarithmic strains \(\epsilon _{\alpha}\) by:

$$ \sigma _{\alpha }= \frac{\partial W}{\partial \epsilon _{\alpha}} \; , \ \sigma _{1} = \frac{2}{3} E_{s} \big(2 \epsilon _{1} + \epsilon _{2} \big)\; , \ \sigma _{2} = \frac{2}{3} E_{s} \big( \epsilon _{1} + 2 \epsilon _{2} \big)\; ; \quad \epsilon _{\alpha}=\ln \lambda _{\alpha }\; , $$
(A.15)

where the secant \(E_{s}\) and tangent \(E_{t}\) moduli of the uniaxial stress-strain curve that appear in (A.3).

After some lengthy algebra we end in the following expressions for the non-zero components of the plane stress moduli in (2.6), given below in two groups; the normal moduli components are:

$$ \textstyle\begin{array}{rl} \mathcal{L}_{1111} & =\displaystyle \frac{1}{\lambda _{1}^{2}} \Big[ \frac{4}{3} E^{*} + \big( E_{t} - E^{*} \big) \Big( \frac{\sigma _{1}}{\sigma _{e}} \Big)^{2} - \sigma _{1} \Big] , \\ \mathcal{L}_{1122} = \mathcal{L}_{2211} & =\displaystyle \frac{1}{\lambda _{1} \lambda _{2}} \Big[ \frac{2}{3} E^{*} + \big( E_{t} - E^{*} \big) \frac{\sigma _{1} \sigma _{2}}{\sigma _{e}^{2}} \Big] , \\ \mathcal{L}_{2222} & =\displaystyle \frac{1}{\lambda _{2}^{2}} \Big[ \frac{4}{3} E^{*} + \big( E_{t} - E^{*} \big) \Big( \frac{\sigma _{2}}{\sigma _{e}} \Big)^{2} - \sigma _{2} \Big] , \end{array} $$
(A.16)

and the shear moduli components are given by:

$$ \textstyle\begin{array}{rl} \mathcal{L}_{1212} & = \displaystyle \frac{1}{\lambda _{2}^{2}} \Big[ \frac{E^{*}}{3} + \frac{\sigma _{2} - \sigma _{1}}{2} \Big] , \\ \mathcal{L}_{2121} & =\displaystyle \frac{1}{\lambda _{1}^{2}} \Big[ \frac{E^{*}}{3} + \frac{\sigma _{1} - \sigma _{2}}{2} \Big] , \\ \mathcal{L}_{1221} = \mathcal{L}_{2112} & = \displaystyle \frac{1}{\lambda _{1} \lambda _{2}} \Big[ \frac{E^{*}}{3} - \frac{\sigma _{1} + \sigma _{2}}{2} \Big] , \end{array} $$
(A.17)

where for the \(J_{2}\) flow theory \(E^{*} = E\), the \(J_{2}\) deformation theory \(E^{*} = E_{s}\) and the hyperelastic model \(E^{*} = E_{s} [(\lambda _{1}^{2} + \lambda _{2}^{2})/(\lambda _{1}^{2} - \lambda _{2}^{2})](\ln \lambda _{1} - \ln \lambda _{2})\) where these quantities were defined immediately after (A.4). Derivation details for the hyperelastic model can be found in [26]. The principal stresses \(\sigma _{\alpha}\) for all three models are identical and given by (A.15).

In the above expressions the material energy density \(W\) is the 3D version. We use here in Appendix A the same symbol as for its 2D counterpart in Sect. 2 to avoid extra symbols. In the same spirit, the dimensionless version of the stresses \(\sigma _{\alpha}\) and moduli \({\mathcal {L}}_{\alpha \beta \gamma \delta}\) in (A.15), (A.16) and (A.17), are found by taking in all the above expressions the initial Young modulus to be \(E=1\).

Appendix B: Influence Discs as Functions of Constitutive Law and Load Orientation

The minimum and maximum influence disc sizes \(\delta _{-}\), defined in (2.12), and \(\delta _{+}\),defined in (2.13), at the onset of loss of ellipticity \(\tau _{m}\), have been presented in [29]. However, for the sake of completeness of the presentation and also in view of a different definition of \(\delta _{+}\) adopted in this paper, these results are presented again here in Fig. 13 with the disc sizes given as functions of the load orientation angle \(\psi \) for the three different constitutive models considered in Appendix A and for two different power-law hardening exponents, \(n=0.22\) (typical of Al alloys, plotted in solid lines) and \(n=0.40\) (typical of steel alloys, plotted in dotted lines). Curves in the \(\psi < 0\) range are terminated at \(\psi =\tan ^{-1}(-0.5)\) below which one stress becomes compressive. Moreover, \(J_{2}\) flow theory curves are only plotted for \(\psi \le 0\), in view of their well-known – see [24] – unrealistic predictions for \(\psi > 0\).

Fig. 13
figure 13

Minimum \(\delta _{-}\) and maximum \(\delta _{+}\) influence disc sizes at the onset of loss of ellipticity \(\tau _{m}\), as functions of the load orientation angle \(\psi \) for the three different constitutive models considered: hyperelastic, \(J_{2}\)-deformation theory and \(J_{2}\)-flow theory, all sharing the same uniaxial stress-strain law with \(\epsilon _{y} = 0.002\) and \(n=0.22\) (solid line) and \(n=0.40\) (dotted line)

As expected, for a given material and load orientation \(\psi \), both \(\delta _{-}\) and \(\delta _{+}\) are increasing functions of the hardening exponent \(n\). For the lower hardening exponent \(n=0.22\), there is practically no difference for either the minimum or maximum influence disc sizes, \(\delta _{-}\) and \(\delta _{-}\) respectively, between the \(J_{2}\) deformation and hyperelastic theory models over the entire range of load orientations considered. For the higher hardening exponent \(n=0.4\), the \(\delta _{-}\) predictions of the \(J_{2}\) deformation and hyperelastic theory models are practically the same over the entire range of loading angles, while the maximum influence disc size \(\delta _{+}\) predictions for these two constitutive models start diverging at about \(\psi < -\pi /12\).

Setting aside the fact that \(\delta _{-}\) and \(\delta _{+}\) exist for a significantly smaller range of load orientations for the much stiffer \(J_{2}\) flow theory, one can observe that for \(n=0.22\) there is no noticeable increase in the \(J_{2}\) flow theory value of \(\delta _{-}\) compared to its hyperelastic and deformation theory counterparts, but there is a large increase in the corresponding value of \(\delta _{+}\), as compared to the other two constitutive models with the same uniaxial curve. The same trend holds true for \(n=0.40\), but the differences between constitutive models, especially for \(\delta _{+}\), are considerably larger. Moreover, the \(J_{2}\) flow theory gives unrealistically high values for \(\delta _{+}\), save for a small range near \(\psi = 0\).

The small difference of the minimum \(\delta _{-}\) and maximum \(\delta _{+}\) influence disc sizes predicted by the hyperelastic and \(J_{2}\) deformation theory constitutive models is the reason for using the simpler to implement hyperelastic model in most of our numerical calculations.

Appendix C: Critical (\(\tau _{m}\)) and Stability Limit (\(\tau _{L}\)) Dimensionless Times

Having established the incremental moduli tensor \(\boldsymbol{\mathcal{L}}\) in Appendix A for the constitutive laws adopted in this work, we calculate next the dimensionless critical time \(\tau _{m}\) for the onset of loss of ellipticity, the corresponding localization angle \(\phi _{m}\) introduced in (2.11) and the dimensionless Lyapunov stability time \(\tau _{L}\) in order to establish the range of validity of the plate’s linearized stability, as presented in Sect. 2.2.

From the definition of \(\mathbf{A}\) in (2.7) we have

$$ \begin{aligned}[b] & {\mathbf{A}}(\tau , \phi ) \ = \left [ \textstyle\begin{array}{c@{\quad}c} {\mathcal {L}}_{1111}(\tau )(n_{1})^{2} + {\mathcal {L}}_{1212}(\tau )(n_{2})^{2} & ({\mathcal {L}}_{1122}(\tau ) + {\mathcal {L}}_{1221}(\tau ))(n_{1}n_{2}) \\ ({\mathcal {L}}_{2211}(\tau ) + {\mathcal {L}}_{2112}(\tau ))(n_{1}n_{2}) & {\mathcal {L}}_{2121}(\tau )(n_{1})^{2} + {\mathcal {L}}_{2222}(\tau )(n_{2})^{2} \end{array}\displaystyle \right ] \; ; \\ &\quad n_{1}=\cos (\phi ), \; n_{2}=\sin (\phi ) , \end{aligned} $$
(C.1)

and from the definitions in (2.11) and (2.10), \(\tau _{m}\) is the lowest positive \(\tau \) root of the characteristic equation of \(\boldsymbol{A}(\tau , \phi ), \forall \phi \in [0 , \pi /2]\)

$$ \begin{aligned}[b] &{\mathcal {L}}_{1111}(\tau ){\mathcal {L}}_{2121}(\tau ) (n_{1})^{4} + [{ \mathcal {L}}_{1111}(\tau ){\mathcal {L}}_{2222}(\tau ) + {\mathcal {L}}_{1212}( \tau ){\mathcal {L}}_{2121}(\tau ) \\ & \quad - ({\mathcal {L}}_{1122}(\tau ){ \mathcal {L}}_{1221}(\tau ))^{2}](n_{1}n_{2})^{2} + {\mathcal {L}}_{1212}( \tau ){\mathcal {L}}_{2222}(\tau ) (n_{2})^{4} . \end{aligned} $$
(C.2)

For the loading angles \(\psi \) considered in Sect. 3, we find

$$ \textstyle\begin{array}{ll} \displaystyle \psi = - {\frac{\pi}{12}} & : \quad \begin{aligned}[t] &{}[{\mathcal {L}}_{1111}( \tau _{m}){\mathcal {L}}_{2222}(\tau _{m})]^{1/2} + [{\mathcal {L}}_{1212}( \tau _{m}){\mathcal {L}}_{2121}(\tau _{m})]^{1/2} \\& \quad - [{\mathcal {L}}_{1122}( \tau _{m})+{\mathcal {L}}_{1221}(\tau _{m})] = 0 , \end{aligned} \\ \displaystyle \psi = 0 & : \quad {\mathcal {L}}_{1111}(\tau _{m})= 0 \Longrightarrow \ \tau _{m} = \exp (n)-1 , \\ \displaystyle \psi = {\frac{\pi}{4}} & \displaystyle : \quad { \mathcal {L}}_{1111}(\tau _{m})= 0 \Longrightarrow \ \tau _{m} = \left [\exp ((1 + 3n)/6) -1\right ]\sqrt {2} , \end{array} $$
(C.3)

where the corresponding angles in the interval \([0 , \pi /2]\) are

$$ \begin{aligned}[b] &\displaystyle \psi = - {\frac{\pi}{12}} \ : \ \phi _{m}= \tan ^{-1} \left [{ \frac{{\mathcal {L}}_{1111}(\tau _{m}){\mathcal {L}}_{2121}(\tau _{m})}{{{\mathcal {L}}_{1212}(\tau _{m}){\mathcal {L}}}_{2222}(\tau _{m})}} \right ]^{1/4} , \quad \psi = 0 \ : \ \phi _{m} = 0, \quad \\ &\psi = \frac{\pi}{4} \ : \ \forall \phi _{m} \in \left [0 , \frac{\pi}{2} \right ] . \end{aligned} $$
(C.4)

Notice that for loading angles \(\psi = 0, \pi /4\), the critical time is independent of the particular version of the model used (hyperelastic, \(J_{2}\) deformation or \(J_{2}\) flow theory) and depends solely on the exponent \(n\)Footnote 18 of the power law-type uniaxial stress-strain curve in (A.14).

The Lyapunov (linearized) stability of the system, as discussed in Sect. 2.2, is guaranteed by the positive definiteness of \(\boldsymbol{A}(\tau , \phi )\) and \(-\dot{\boldsymbol{A}}(\tau , \phi )\). By construction \(\boldsymbol{A}(\tau , \phi )\) is positive definite for \(\forall \tau \in [0,\tau _{m})\), \(\forall \phi \in [0,\pi /2]\). Finding the value of \(\tau _{L} \le \tau _{m}\) that guarantees the positive definiteness of \(-\dot{\boldsymbol{A}}(\tau , \phi )\) for \(\forall \tau \in [0,\tau _{L})\), \(\forall \phi \in [0,\pi /2]\) is based on numerical calculations using a symbolic manipulator.

For the cases investigated in Sect. 3 the following values are found for the dimensionless critical time \(\tau _{m}\) and the corresponding linearized (Lyapunov) stability time \(\tau _{L}\)

$$\begin{aligned} {\mathrm{hyperelastic}}\ ; & \ n=0.22 \quad \left \{ \textstyle\begin{array}{l@{\quad}l@{\quad}l} \displaystyle \psi = \tan ^{-1}\left (-{\frac{1}{2}}\right ), & \tau _{m} = 0.508, & \tau _{L} = 0.38 \\ \displaystyle \psi = -{\frac{\pi}{12}}, & \tau _{m} = 0.352, & \tau _{L} = \tau _{m} \\ \displaystyle \psi = 0, & \tau _{m} = 0.246, & \tau _{L} = \tau _{m} \\ \displaystyle \psi = {\frac{\pi}{4}}, & \tau _{m} = 0.451, & \tau _{L} = \tau _{m} \end{array}\displaystyle \right . \\ & \ n=0.40 \quad \left \{ \textstyle\begin{array}{l@{\quad}l@{\quad}l} \displaystyle \psi = \tan ^{-1}\left (-{\frac{1}{2}}\right ), & \tau _{m} = 0.900, & \tau _{L} = 0.33 \\ \displaystyle \psi = -{\frac{\pi}{12}}, & \tau _{m} = 0.715, & \tau _{L} =0.475 \\ \displaystyle \psi = 0, & \tau _{m} = 0.492, & \tau _{L} = \tau _{m} \\ \displaystyle \psi = {\frac{\pi}{4}}, & \tau _{m} = 0.626, & \tau _{L} = \tau _{m} \end{array}\displaystyle \right . \\ J_{2}\ {\mathrm{deformation}}\ ; & \ n=0.22 \quad \left \{ \textstyle\begin{array}{l@{\quad}l@{\quad}l} \displaystyle \psi = \tan ^{-1}\left (-{\frac{1}{2}}\right ), & \tau _{m} = 0.482, & \tau _{L} = 0.44 \\ \displaystyle \psi = -{\frac{\pi}{12}}, & \tau _{m} = 0.351, & \tau _{L} = \tau _{m} \\ \displaystyle \psi = 0, & \tau _{m} = 0.246, & \tau _{L} = \tau _{m} \\ \displaystyle \psi = {\frac{\pi}{4}}, & \tau _{m} = 0.451, & \tau _{L} = \tau _{m} \end{array}\displaystyle \right . \\ & \ n=0.40 \quad \left \{ \textstyle\begin{array}{l@{\quad}l@{\quad}l} \displaystyle \psi = \tan ^{-1}\left (-{\frac{1}{2}}\right ), & \tau _{m} = 0.723, & \tau _{L} = 0.38 \\ \displaystyle \psi = -{\frac{\pi}{12}}, & \tau _{m} = 0.689, & \tau _{L} =0.620 \\ \displaystyle \psi = 0, & \tau _{m} = 0.492, & \tau _{L} = \tau _{m} \\ \displaystyle \psi = {\frac{\pi}{4}}, & \tau _{m} = 0.626, & \tau _{L} = \tau _{m} \end{array}\displaystyle \right . \\ J_{2}\ {\mathrm{flow}}\ ; & \ n=0.22 \quad \left \{ \textstyle\begin{array}{l@{\quad}l@{\quad}l} \displaystyle \psi = \tan ^{-1}\left (-{\frac{1}{2}}\right ), & \tau _{m} = 0.756, & \tau _{L} = 0 \\ \displaystyle \psi = -{\frac{\pi}{12}}, & \tau _{m} = 0.405, & \tau _{L} = 0 \\ \displaystyle \psi = 0, & \tau _{m} = 0.246, & \tau _{L} = 0 \end{array}\displaystyle \right . \\ & \ n=0.40 \quad \left \{ \textstyle\begin{array}{l@{\quad}l@{\quad}l} \displaystyle \psi = \tan ^{-1}\left (-{\frac{1}{2}}\right ), & \tau _{m} = 0.924, & \tau _{L} = 0 \\ \displaystyle \psi = -{\frac{\pi}{12}}, & \tau _{m} = 1.131, & \tau _{L} = 0 \\ \displaystyle \psi = 0, & \tau _{m} = 0.492, & \tau _{L} = 0 \end{array}\displaystyle \right . \end{aligned}$$
(C.5)

It is worth noticing in (C.5) that – for the same hardening exponent \(n\) – the results for the hyperelastic and \(J_{2}\) deformation theory models are either very close (or coincident for \(\psi = 0, \pi /4\)), while the much stiffer \(J_{2}\) flow theory is not linearly stable since \(-\dot{\boldsymbol{A}}(\tau , \phi )\) looses positive definiteness for certain values of \(\phi \) at all dimensionless times \(\tau \) (and hence \(\tau _{L}=0\)).

Rights and permissions

Springer Nature or its licensor (e.g., a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wen, G., Ravi-Chandar, K., Elliott, R.S. et al. Stability and Localization of Deformation Delay in Finitely Strained Plates at Arbitrary Strain-Rates. J Elast (2022). https://doi.org/10.1007/s10659-022-09962-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10659-022-09962-9

Keywords

Mathematics Subject Classification (2010)

Navigation