Skip to main content
Log in

Coherent energetic interfaces accounting for in-plane degradation

  • ISDMM15
  • Published:
International Journal of Fracture Aims and scope Submit manuscript

Abstract

Interfaces can play a dominant role in the overall response of a body. The importance of interfaces is particularly appreciated at small length scales due to large area to volume ratios. From the mechanical point of view, this scale dependent characteristic can be captured by endowing a coherent interface with its own elastic resistance as proposed by the interface elasticity theory. This theory proves to be an extremely powerful tool to explain size effects and to predict the behavior of nano-materials. To date, interface elasticity theory only accounts for the elastic response of coherent interfaces and obviously lacks an explanation for inelastic interface behavior such as damage or plasticity. The objective of this contribution is to extend interface elasticity theory to account for damage of coherent interfaces. To this end, a thermodynamically consistent interface elasticity theory with damage is proposed. A local damage model for the interface is presented and is extended towards a non-local damage model. The non-linear governing equations and the weak forms thereof are derived. The numerical implementation is carried out using the finite element method and consistent tangents are listed. The computational algorithms are given in detail. Finally, a series of numerical examples is studied to provide further insight into the problem and to carefully elucidate key features of the proposed theory.

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
Fig. 13
Fig. 14
Fig. 15

Similar content being viewed by others

Notes

  1. The tangential deformation of the interface and shear/sliding displacement jump across the non-coherent interface (see Tvergaard 1990, for instance) are two very different phenomena. The former is measured in terms of a second-order superficial deformation gradient and the latter in terms of the displacement jump vector. The former then causes interface stress on the tangential plane of the interface resulting in the superficial second-order Piola stress tensor while the latter causes traction, a vector quantity across the interface. To induce stress on the tangential plane of the interface one needs to apply some form of deformation on the elastic interface, whereas a cohesive interface is existent if and only if there is some form of opening (normal or shear) across the interface.

  2. The superficiality of the interface Piola stress tensor is a classical assumption of interface elasticity theory. Recently, Javili et al. (2013a) have proven that this condition is the consequence of a first-order continuum theory.

  3. Note that the same notation \(\overline{F}_\text {max}\) is used for both the local and non-local versions. Nevertheless, they are clearly distinguished by their definition. The implementation of this contribution focuses only on the non-local version. Clearly, the non-local theory boils down to the local theory in the limit case of \(\overline{\omega }\) being the Dirac delta distribution. This can be achieved by setting \(\overline{R} = 0\). The same discussion holds for the bulk as well.

  4. The derivation is only carried out for the interface tangent stiffness matrix. Analogous derivations for the bulk are standard and are omitted for the sake of conciseness.

  5. A remark on the legends of the graphs is necessary. The legends “case1”, “case2” and “case3” represent the locations where damage initiates and evolves, which are in the bulk, on the interface and in both the bulk and interface respectively. The “-on-interfaceNode” and “on-bulkNode” part of all the legends stand for the location of the nodes on which the measurements are done to draw the graphs (see Fig. 6). The “-on-interfaceNode\(_\text {b}\)” means a bulk quantity is measured at the interface node.

  6. The term “volumetric” has a different meaning on the interface. As opposed to a volumetric deformation in the bulk, which changes the volume uniformly, a volumetric interface deformation changes the area uniformly.

  7. In Table 7, \(10\times 10\times 1\) for instance indicates 10 elements in x, y and 1 element in z direction.

References

  • Alfano G, Crisfield MA (2001) Finite element interface models for the delamination analysis of laminated composites: Mechanical and computational issues. Int J Numer Methods Eng 50:1701–1736

    Article  Google Scholar 

  • Alfano G, Sacco E (2006) Combining interface damage and friction in a cohesive-zone model. Int J Numer Methods Eng 68(5):542–582

    Article  Google Scholar 

  • Allix O, Ladevéze P, Corigliano A (1995) Damage analysis of interlaminar fracture specimens. Compos Struct 31(1):61–74

    Article  Google Scholar 

  • Allix O, Corigliano A (1996) Modeling and simulation of crack propagation in mixed-modes interlaminar fracture specimens. lnt J Fract 77(2):111–140

    Article  Google Scholar 

  • Andrade FXC, Sá JMACD, Pires FMA (2011) A ductile damage nonlocal model of integral-type at finite strains: formulation and numerical issues. Int J Damage Mech 20(4):515–557

    Article  Google Scholar 

  • Andrade F, Sá JCD, Pires FA (2014) Assessment and comparison of non-local integral models for ductile damage. Int J Damage Mech 23(2):261–296

    Article  Google Scholar 

  • Aragón AM, Soghrati S, Geubelle PH (2013) Effect of in-plane deformation on the cohesive failure of heterogeneous adhesives. J Mech Phys Solids 61(7):1600–1611

    Article  Google Scholar 

  • Askes H, Sluys LJ (2000) Remeshing strategies for adaptive ALE analysis of strain localisation. Eur J Mech A Solids 19(3):447–467

    Article  Google Scholar 

  • Barenblatt GI (1962) The mathematical theory of equilibrium cracks in brittle fracture. Advances in applied mechanics 7(1):55–129

    Article  Google Scholar 

  • Bažant ZP (1994) Nonlocal damage theory based on micromechanics of crack interactions. J Eng Mech 120(3):593–617

    Article  Google Scholar 

  • Bažant ZP, Cabot GP (1989) Measurement of characteristic length of nonlocal continuum. J Eng Mech 115(4):755–767

    Article  Google Scholar 

  • Bažant ZP, Jirásek M (2002) Nonlocal integral formulations of plasticity and damage: survey of progress. J Eng Mech 128(11):1119–1149

    Article  Google Scholar 

  • Bažant ZP, Xi Y (1991) Statistical size effect in quasibrittle structures: II. Nonlocal theory. J Eng Mech 117(11):2623–2640

    Article  Google Scholar 

  • Benveniste Y (2006) A general interface model for a three-dimensional curved thin anisotropic interphase between two anisotropic media. J Mech Phys Solids 54(4):708–734

    Article  Google Scholar 

  • Benveniste Y (2013) Models of thin interphases and the effective medium approximation in composite media with curvilinearly anisotropic coated inclusions. Int J Eng Sci 72:140–154

    Article  Google Scholar 

  • Benveniste Y, Miloh T (2001) Imperfect soft and stiff interfaces in two-dimensional elasticity. Mech Mater 33(6):309–323

    Article  Google Scholar 

  • Bolzon G, Corigliano A (1997) A discrete formulation for elastic solids with damaging interfaces. Comput Methods Appl Mech Eng 140(3–4):329–359

    Article  Google Scholar 

  • Cabot GP, Bažant ZP (1987) Nonlocal damage theory. J Eng Mech 113(10):1512–1533

    Article  Google Scholar 

  • Cammarata RC (1997) Surface and interface stress effects on interfacial and nanostructured materials. Mater Sci Eng A 237(2):180–184

    Article  Google Scholar 

  • Carmeliet J (1999) Optimal estimation of gradient damage parameters from localization phenomena in quasi-brittle materials. Mech Cohes Frict Mater 4(1):1–16

    Article  Google Scholar 

  • Cazes F, Coret M, Combescure A, Gravouil A (2009) A thermodynamic method for the construction of a cohesive law from a nonlocal damage model. Int J Solids Struct 46(6):1476–1490

    Article  Google Scholar 

  • Chaboche JL (1981) Continuous damage mechanics—a tool to describe phenomena before crack initiation. Nucl Eng Des 64(2):233–247

    Article  Google Scholar 

  • Chaboche JL (1984) Anisotropic creep damage in the framework of continuum damage mechanics. Nucl Eng Des 79(3):309–319

    Article  Google Scholar 

  • Chaboche JL, Girard R, Schaff A (1997) Numerical analysis of composite systems by using interphase/interface models. Comput Mech 20:3–11

    Article  Google Scholar 

  • Chatzigeorgiou G, Javili A, Steinmann P (2013) Multiscale modelling for composites with energetic interfaces at the micro- or nanoscale. Math Mech Solids

  • Chen J, Crisfield M, Kinloch AJ, Busso EP, Matthews FL, Qiu Y (1999) Predicting progressive delamination of composite material specimens via interface elements. Mech Compos Mater Struct 6(4):301–317

    Article  Google Scholar 

  • Cordero NM, Forest S, Busso EP (2015) Second strain gradient elasticity of nano-objects. J Mech Phys Solids. doi:10.1016/j.jmps.2015.07.012

  • Corigliano A (1993) Formulation, identification and use of interface models in the numerical analysis of composite delamination. Int J Solids Struct 30(20):2779–2811

    Article  Google Scholar 

  • Daher N, Maugin GA (1986) The method of virtual power in continuum mechanics application to media presenting singular surfaces and interfaces. Acta Mech 60(3–4):217–240

    Article  Google Scholar 

  • Davydov D, Javili A, Steinmann P (2013) On molecular statics and surface-enhanced continuum modeling of nano-structures. Comput Mater Sci 69:510–519

    Article  Google Scholar 

  • de Souza Neto EA, Perić D, Owen DRJ (1994a) A model for elastoplastic damage at finite strains: algorithmic issues and applications. Eng Comput 11(3):257–281

    Article  Google Scholar 

  • de Souza Neto EA, Perić D, Owen DRJ (1994b) A phenomenological three-dimensional rate-idependent continuum damage model for highly filled polymers: formulation and computational aspects. J Mech Phys Solids 42(10):1533–1550

    Article  Google Scholar 

  • de Souza Neto EA, Perić D, Owen DRJ (1998) Continuum modelling and numerical simulation of material damage at finite strains. Arch Comput Methods Eng 5(4):311–384

    Article  Google Scholar 

  • de Souza Neto EA, Peric D, Owen DRJ (2011) Computational methods for plasticity: theory and applications. Wiley, London

    Google Scholar 

  • de Souza Neto EA, Perić D (1996) A computational framework for a class of fully coupled models for elastoplastic damage at finite strains with reference to the linearization aspects. Comput Methods Appl Mech Eng 130(1–2):179–193

    Article  Google Scholar 

  • dell’Isola F, Romano A (1987) On the derivation of thermomechanical balance equations for continuous systems with a nonmaterial interface. Int J Eng Sci 25:1459–1468

    Article  Google Scholar 

  • Dingreville R, Qu J, Cherkaoui M (2005) Surface free energy and its effect on the elastic behavior of nano-sized particles, wires and films. J Mech Phys Solids 53(8):1827–1854

    Article  Google Scholar 

  • Duan HL, Karihaloo BL (2007) Effective thermal conductivities of heterogeneous media containing multiple imperfectly bonded inclusions. Phys Rev B 75(6):064,206

  • Duan HL, Wang J, Huang ZP, Karihaloo BL (2005a) Eshelby formalism for nano-inhomogeneities. Proc R Soc A Math Phys Eng Sci 461(2062):3335–3353

    Article  Google Scholar 

  • Duan HL, Wang J, Huang ZP, Karihaloo BL (2005b) Size-dependent effective elastic constants of solids containing nano-inhomogeneities with interface stress. J Mech Phys Solids 53(7):1574–1596

    Article  Google Scholar 

  • Duan HL, Wang J, Karihaloo BL (2009) Theory of elasticity at the nanoscale. Adv Appl Mech 42:1–68

    Article  Google Scholar 

  • Eringen AC (1966) A unified theory of thermomechanical materials. Int J Eng Sci 4(2):179–202

    Article  Google Scholar 

  • Fagerström M, Larsson R (2008) A thermo-mechanical cohesive zone formulation for ductile fracture. J Mech Phys Solids 56(10):3037–3058

    Article  Google Scholar 

  • Fischer FD, Waitz T, Vollath D, Simha NK (2008) On the role of surface energy and surface stress in phase-transforming nanoparticles. Prog Mater Sci 53(3):481–527

  • Fischer FD, Svoboda J (2010) Stresses in hollow nanoparticles. Int J Solids Struct 47(20):2799–2805

    Article  Google Scholar 

  • Fleischhauer R, Behnke R, Kaliske M (2013) A thermomechanical interface element formulation for finite deformations. Comput Mech 52(5):1039–1058

    Article  Google Scholar 

  • Fried E, Todres R (2005) Mind the gap: the shape of the free surface of a rubber-like material in proximity to a rigid contactor. J Elast 80(1–3):97–151

    Article  Google Scholar 

  • Gurtin ME, Weissmüller J, Larché F (1998) A general theory of curved deformable interfaces in solids at equilibrium. Philos Mag A 78(5):1093–1109

    Article  Google Scholar 

  • Gurtin ME, Murdoch AI (1975) A continuum theory of elastic material surfaces. Arch Ration Mech Anal 57(4):291–323

    Article  Google Scholar 

  • Haiss W (2001) Surface stress of clean and adsorbate-covered solids. Rep Prog Phys 64(5):591

    Article  Google Scholar 

  • Holzapfel GA (2000) Nonlinear Solid mechanics: a continuum approach for engineering. Wiley, London

    Google Scholar 

  • Huang ZP, Sun L (2007) Size-dependent effective properties of a heterogeneous material with interface energy effect: From finite deformation theory to infinitesimal strain analysis. Acta Mech 190(1–4):151–163

    Article  Google Scholar 

  • Ijaz H, Asad M, Gornet L, Alam SY (2014) Prediction of delamination crack growth in carbon/fiber epoxy composite laminates using non-local interface damage model. Mech Ind 15(4):293–300

    Article  Google Scholar 

  • Javili A, McBride A, Steinmann P (2012) Numerical modelling of thermomechanical solids with mechanically energetic (generalised) Kapitza interfaces. Comput Mater Sci 65:542–551

    Article  Google Scholar 

  • Javili A, dell’Isola F, Steinmann P (2013a) Geometrically nonlinear higher-gradient elasticity with energetic boundaries. J Mech Phys Solids 61(12):2381–2401

    Article  Google Scholar 

  • Javili A, McBride A, Steinmann P (2013b) Numerical modelling of thermomechanical solids with highly conductive energetic interfaces. Int J Numer Meth Eng 93(5):551–574

    Article  Google Scholar 

  • Javili A, McBride A, Steinmann P (2013c) Thermomechanics of solids with lower-dimensional energetics: on the importance of surface, interface, and curve structures at the nanoscale. A unifying review. Appl Mech Rev 65(1):010802

    Article  Google Scholar 

  • Javili A, Kaessmair S, Steinmann P (2014a) General imperfect interfaces. Comput Methods Appl Mech Eng 275:76–97

    Article  Google Scholar 

  • Javili A, McBride A, Steinmann P, Reddy BD (2014b) A unified computational framework for bulk and surface elasticity theory: a curvilinear-coordinate-based finite element methodology. Comput Mech 54(3):745–762

    Article  Google Scholar 

  • Javili A, Steinmann P (2009) A finite element framework for continua with boundary energies. Part I: the two-dimensional case. Comput Methods Appl Mech Eng 198(27–29):2198–2208

    Article  Google Scholar 

  • Javili A, Steinmann P (2010a) A finite element framework for continua with boundary energies. Part II: the three-dimensional case. Comput Methods Appl Mech Eng 199(9–12):755–765

    Article  Google Scholar 

  • Javili A, Steinmann P (2010b) On thermomechanical solids with boundary structures. Int J Solids Struct 47(24):3245–3253

    Article  Google Scholar 

  • Jirasek M (1998) Nonlocal models for damage and fracture: comparison of approaches. Int J Solids Struct 35(31):4133–4145

    Article  Google Scholar 

  • Jirásek M, Patzák B (2002) Consistent tangent stiffness for nonlocal damage models. Comput Struct 80(14–15):1279–1293

  • Kachanov LM (1958) Time of the rupture process under creep conditions. Izv Akad Nauk SSR Otd Tech Nauk 8:26–31

    Google Scholar 

  • Krajcinovic D, Fonseka GU (1981) Continuous damage theory of brittle materials—1. General theory. J Appl Mech Trans ASME 48(4):809–815

    Article  Google Scholar 

  • Krayani A, Pijaudier-Cabot G, Dufour F (2009) Boundary effect on weight function in nonlocal damage model. Eng Fract Mech 76(14):2217–2231

    Article  Google Scholar 

  • Ladevèze P, Allix O, Gornet L, Lévêque D, Perret L (1998) A computational damage mechanics approach for laminates: identification and comparison with experimental results. Stud Appl Mech 46:481–500

    Article  Google Scholar 

  • Lemaitre J (1984) Three-dimensional ductile damage model applied to deep-drawing forming limits. In: Proceedings of the \(4^{{\rm th}}\) international conference on the mechanical behaviour of materials vol 2, pp 1047–1053

  • Levitas VI, Javanbakht M (2010) Surface tension and energy in multivariant martensitic transformations: phase-field theory, simulations, and model of coherent interface. Phys Rev Lett 105(16):165–701

    Article  Google Scholar 

  • Lin G, Geubelle PH, Sottos NR (2001) Simulation of fiber debonding with friction in a model composite pushout test. Int J Solids Struct 38(46–47):8547–8562

    Article  Google Scholar 

  • Mazars J, Berthaud Y, Ramtani S (1990) The unilateral behaviour of damaged concrete. Eng Fract Mech 35(4):629–635

    Article  Google Scholar 

  • McBride AT, Javili A, Steinmann P, Bargmann S (2011) Geometrically nonlinear continuum thermomechanics with surface energies coupled to diffusion. J Mech Phys Solids 59(10):2116–2133

    Article  Google Scholar 

  • Mi Y, Crisfield MA, Davies GAO, Hellweg HB (1998) Progressive delamination using interface elements. J Compos Mater 32(14):1246–1272

    Article  Google Scholar 

  • Moeckel GP (1975) Thermodynamics of an interface. Arch Ration Mech Anal 57(3):255–280

    Article  Google Scholar 

  • Mosler J, Scheider I (2011) A thermodynamically and variationally consistent class of damage-type cohesive models. J Mech Phys Solids 59(8):1647–1668

    Article  Google Scholar 

  • Murakami S, Ohno N (1981) A continuum theory of creep and creep damage. In: Ponter ARS, Hayhurst DR (eds) Creep in structures, international union of theoretical and applied mechanics. Springer, Berlin, pp 422–444

    Google Scholar 

  • Murdoch AI (1976) A thermodynamical theory of elastic material interfaces. Q J Mech Appl Math 29(3):245–275

    Article  Google Scholar 

  • Needleman A (1990) An analysis of tensile decohesion along an interface. J Mech Phys Solids 38(3):289–324

    Article  Google Scholar 

  • Needleman A (1992) Micromechanical modelling of interfacial decohesion. Ultramicroscopy 40:203–214

    Article  Google Scholar 

  • Needleman A (2014) Some issues in cohesive surface modeling. Procedia IUTAM 10:221–246

    Article  Google Scholar 

  • Özdemir I, Brekelmans WAM, Geers MGD (2010) A thermo-mechanical cohesive zone model. Comput Mech 46:735–745

    Article  Google Scholar 

  • Park HS, Klein PA (2007) Surface Cauchy–Born analysis of surface stress effects on metallic nanowires. Phys Rev B 75(085):408

    Google Scholar 

  • Parrinello F, Failla B, Borino G (2009) Cohesive–frictional interface constitutive model. Int J Solids Struct 46(13):2680–2692

  • Pijaudier-Cabot G, Dufour F (2010) Non local damage model: boundary and evolving boundary effects. Eur J Environ Civil Eng 14(6–7):729–749

    Article  Google Scholar 

  • Pijaudier-Cabot G, Grégoire D (2014) A review of non local continuum damage: modelling of failure? Netw Heterog Media 9(4):575–597

    Article  Google Scholar 

  • Pijaudier-Cabot G, Huerta A (1991) Finite element analysis of bifurcation in nonlocal strain softening solids. Comput Methods Appl Mech Eng 90(1–3):905–919

    Article  Google Scholar 

  • Rabotnov YN (1963) On the equation of state of creep. Proc Inst Mech Eng Conf Proc 178(1):2-117–2-122

    Google Scholar 

  • Raous M (2011) Interface models coupling adhesion and friction. Comptes Rendus Mecanique 339(7–8):491–501

    Article  Google Scholar 

  • Saanouni K, Chaboche JL, Lesne PM (1989) On the creep crack-growth prediction by a non local damage formulation. Eur J Mech A Solids 8(6):437–459

    Google Scholar 

  • Saouridis C, Mazars J (1992) Prediction of the failure and size effect in concrete via a bi-scale damage approach. Eng Comput 9(3):329–344

    Article  Google Scholar 

  • Schellekens JCJ, Borst RD (1993) A non-linear finite element approach for the analysis of mode-I free edge delamination in composites. Int J Solids Struct 30(9):1239–1253

    Article  Google Scholar 

  • Sharma P, Ganti S, Bhate N (2003) Effect of surfaces on the size-dependent elastic state of nano-inhomogeneities. Appl Phys Lett 82(4):535–537

    Article  Google Scholar 

  • Sharma P, Ganti S (2004) Size-dependent Eshelbys tensor for embedded nano-inclusions incorporating surface/interface energies. J Appl Mech 71(5):663–671

    Article  Google Scholar 

  • Sharma P, Wheeler LT (2007) Size-dependent elastic state of ellipsoidal nano-inclusions incorporating surface/ interface tension. J Appl Mech 74(3):447–454

    Article  Google Scholar 

  • Simo JC, Hughes TJR (1998) Comput Inelast. Springer, New York

    Google Scholar 

  • Simo JC, Ju JW (1987) Strain- and stress-based continuum damage models I: formulation. Int J Solids Struct 23(7):821–840

    Article  Google Scholar 

  • Steigmann DJ, Ogden RW (1999) Elastic surface–substrate interactions. Proc R Soc Lond A Math Phys Eng Sci 455(1982):437–474

    Article  Google Scholar 

  • Steinmann P, Miehe C, Stein E (1994) Comparison of different finite deformation inelastic damage models within multiplicative elastoplasticity for ductile materials. Comput Mech 13(6):458–474

    Article  Google Scholar 

  • Steinmann P (1999) Formulation and computation of geometrically non-linear gradient damage. Int J Numer Meth Eng 46(5):757–779

    Article  Google Scholar 

  • Steinmann P (2008) On boundary potential energies in deformational and configurational mechanics. J Mech Phys Solids 56(3):772–800

    Article  Google Scholar 

  • Tvergaard V (1990) Effect of fibre debonding in a whisker-reinforced metal. Mater Sci Eng A 125(2):203–213

    Article  Google Scholar 

  • van den Bosch MJ, Schreurs PJG, Geers MGD (2006) An improved description of the exponential Xu and Needleman cohesive zone law for mixed-mode decohesion. Eng Fract Mech 73:1220–1234

  • van den Bosch MJ, Schreurs PJG, Geers MGD (2007) A cohesive zone model with a large displacement formulation accounting for interfacial fibrilation. Eur J Mech A/Solids 26:1–19

  • Willam K, Rhee I, Shing B (2004) Interface damage model for thermomechanical degradation of heterogeneous materials. Comput Methods Appl Mech Eng 193(30):3327–3350

  • Wu L, Becker G, Noels L (2014) Elastic damage to crack transition in a coupled non-local implicit discontinuous Galerkin/extrinsic cohesive law framework. Comput Methods Appl Mech Eng 279:379–409

    Article  Google Scholar 

  • Yvonnet J, He QC, Zhu QZ, Shao JF (2011a) A general and efficient computational procedure for modelling the Kapitza thermal resistance based on XFEM. Comput Mater Sci 50(4):1220–1224

    Article  Google Scholar 

  • Yvonnet J, Mitrushchenkov A, Chambaud G, He QC (2011b) Finite element model of ionic nanowires with size-dependent mechanical properties determined by ab initio calculations. Comput Methods Appl Mech Eng 200(5–8):614–625

    Article  Google Scholar 

Download references

Acknowledgments

This research is performed as part of the Energie Campus Nuremberg and supported by funding through the “Bavaria on the Move” initiative of the state of Bavaria. The authors also gratefully acknowledge the support by the Cluster of Excellence “Engineering of Advanced Materials”.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Paul Steinmann.

Appendices

Appendix 1: Some mathematical relations and derivations

In this section we present the derivation of the balance of forces on the interface and the corresponding weak form. Subsequently, the decoupled form of the interface energy, Piola stress and elasticity tensor are provided. Some useful identities and relations used in the derivations are also given without proof.

1.1 Extended divergence theorem

The extended forms of divergence theorem in the material configuration for a bulk tensor field \(\{\bullet \}\) and a tensorial quantity on the interface \(\left\{ \overline{\bullet }\right\} \) are

$$\begin{aligned}&\int _{{\mathcal {B}}_0} {\mathrm {Div}}\{\bullet \} \mathrm {d}V = \int _{\partial {\mathcal {B}}_0} \left\{ \bullet \right\} \cdot {\varvec{N}} \mathrm {d}A - \int _{ {\mathcal {I}}_0} \llbracket \bullet \rrbracket \cdot \overline{{\varvec{N}}} \mathrm {d}A\;, \end{aligned}$$
(37)
$$\begin{aligned}&\int _{{\mathcal {I}}_0} \overline{{\mathrm {Div}}}\left\{ \overline{\bullet }\right\} \mathrm {d}A = \int _{\partial {\mathcal {I}}_0\setminus \partial {\mathcal {I}}_0^{\mathrm {N}}} \left\{ \overline{\bullet }\right\} \cdot \widetilde{{\varvec{N}}} \mathrm {d}L\nonumber \\&\quad + \int _{\partial {\mathcal {I}}_0^{\mathrm {N}}} \left\{ \overline{\bullet }\right\} \cdot \widetilde{{\varvec{N}}} \mathrm {d}L- \int _{{\mathcal {I}}_0} \overline{C} \left\{ \overline{\bullet }\right\} \cdot \overline{{\varvec{N}}} \mathrm {d}A \;, \end{aligned}$$
(38)

where the curvature of the interface is denoted by \(\overline{C}\). Note that \(\partial {\mathcal {I}}_0^{\mathrm {N}}\) is the portion of the interface boundary that intersects with the bulk’s boundary, thus \(\partial {\mathcal {I}}_0\setminus \partial {\mathcal {I}}_0^{\mathrm {N}} \cap \partial {\mathcal {B}}_0 = \emptyset \).

1.2 Balance of forces on interface

The global form of the balance of forces both in the bulk and on the interface is given as (see, Javili and Steinmann 2010b, for further details concerning thermomechanical solids with surface energy only)

$$\begin{aligned}&\int _{{\mathcal {B}}_0} {\varvec{B}}^{\mathrm {p}} \mathrm {d}V + \int _{{\mathcal {I}}_0} \overline{{\varvec{B}}}{}^{\mathrm {p}} \mathrm {d}A + \int _{\partial {\mathcal {B}}_0^{\mathrm {N}}} \widehat{{\varvec{B}}}{}^{\mathrm {p}}_{\mathrm {N}} \mathrm {d}A + \int _{\partial {\mathcal {I}}_0^{\mathrm {N}}} \widetilde{{\varvec{B}}}{}^{\mathrm {p}}_{\mathrm {N}} \mathrm {d}L\nonumber \\&\quad + \int _{\partial {\mathcal {B}}_0 \setminus \partial {\mathcal {B}}_0^{\mathrm {N}}} {\varvec{P}} \cdot {\varvec{N}} \mathrm {d}A + \int _{\partial {\mathcal {I}}_0 \setminus \partial {\mathcal {I}}_0^{\mathrm {N}}} \overline{{\varvec{P}}} \cdot \widetilde{{\varvec{N}}} \mathrm {d}L= {\varvec{0}}.\nonumber \\ \end{aligned}$$
(39)

Taking the limit \({\mathcal {B}}_0 \rightarrow \emptyset \), and consequently \({\partial {\mathcal {B}}_0} = {{\mathcal {I}}_0}\), with \({\varvec{N}} = \overline{{\varvec{N}}}\) on \({{\mathcal {I}}_0^+}\), \({\varvec{N}} = -\overline{{\varvec{N}}}\) on \({{\mathcal {I}}_0^-}\), \({\partial {\mathcal {I}}_0^{\mathrm {N}}} = \emptyset \), \({\partial {\mathcal {B}}_0^{\mathrm {N}}} = \emptyset \), and taking into account the extended forms of the divergence theorem (37) and (38), one obtains the local balance of forces on the interface as

$$\begin{aligned}&\int _{\partial {\mathcal {I}}_0} \overline{{\varvec{P}}} \cdot \widetilde{{\varvec{N}}} \mathrm {d}L + \int _{{\mathcal {I}}_0 } {\varvec{P}} \cdot {\varvec{N}} \mathrm {d}A + \int _{{\mathcal {I}}_0} \overline{{\varvec{B}}}{}^{\mathrm {p}} \mathrm {d}A = {\varvec{0}} \nonumber \\&\quad \Longrightarrow \int _{{\mathcal {I}}_0} \overline{{\mathrm {Div}}}~\overline{\varvec{P}} + \llbracket \varvec{P}\rrbracket \cdot \overline{{\varvec{N}}} + \overline{{\varvec{B}}}{}^{\mathrm {p}} \mathrm {d}A = {\varvec{0}}\;. \end{aligned}$$
(40)

From arbitrariness of \({\mathcal {B}}_0\) and thus \({\mathcal {I}}_0\), the balance of force on the interface listed in Table 2 then follows. In the case that the interface is not energetic i.e. \(\overline{\varvec{P}} = {\varvec{0}}\), and in the absence of interface body force (\(\overline{{\varvec{B}}}{}^{\mathrm {p}} = {\varvec{0}}\)), the classical traction continuity condition is recovered.

1.3 Weak form of the balance of forces

The localized balance equations in the bulk and on the interface, given in Table 2 are tested from the left with vector valued functions \(\delta {\varvec{\varphi }}\) and \(\delta \overline{{\varvec{\varphi }}}\), respectively as follows

$$\begin{aligned}&\int _{{\mathcal {B}}_0} \delta {\varvec{\varphi }} \cdot [~ {\mathrm {Div}}{\varvec{P}} + {\varvec{B}}^{\mathrm {p}} ~] \mathrm {d}V \nonumber \\&\quad + \int _{{\mathcal {I}}_0} \delta \overline{{\varvec{\varphi }}} \cdot [~ \overline{{\mathrm {Div}}}\, \overline{{\varvec{P}}} + \overline{{\varvec{B}}}{}^{\mathrm {p}} + \llbracket {\varvec{P}}\rrbracket \cdot \overline{{\varvec{N}}} ~] \mathrm {d}A = 0,\nonumber \\ \end{aligned}$$
(41)

which can be alternatively written as

$$\begin{aligned}&\int _{{\mathcal {B}}_0} -{\varvec{P}} : {\mathrm {Grad}}\delta {\varvec{\varphi }} + {\mathrm {Div}}( \delta {\varvec{\varphi }} \cdot {\varvec{P}}) + \delta {\varvec{\varphi }} \cdot {\varvec{B}}^{\mathrm {p}} \mathrm {d}V \nonumber \\&\quad + \int _{{\mathcal {I}}_0} -\overline{{\varvec{P}}} : \overline{{\mathrm {Grad}}}\delta \overline{{\varvec{\varphi }}} + \overline{{\mathrm {Div}}}( \delta \overline{{\varvec{\varphi }}} \cdot \overline{{\varvec{P}}}) \mathrm {d}A\nonumber \\&\quad + \int _{{\mathcal {I}}_0} \delta \overline{{\varvec{\varphi }}} \cdot \overline{{\varvec{B}}}{}^{\mathrm {p}} + \delta \overline{{\varvec{\varphi }}} \cdot [~\llbracket {\varvec{P}}\rrbracket \cdot \overline{{\varvec{N}}} ~] \mathrm {d}A = 0\;, \end{aligned}$$
(42)

and using the extended forms of divergence theorem (37) and (38), for various parts of the body results in

$$\begin{aligned}&\int _{{\mathcal {B}}_0} {\varvec{P}} : {\mathrm {Grad}}\delta {\varvec{\varphi }} \mathrm {d}V - \int _{\partial {\mathcal {B}}_0^{\mathrm {N}}} \delta {\varvec{\varphi }} \cdot [~ {\varvec{P}} \cdot {\varvec{N}} ~] \mathrm {d}A \nonumber \\&\quad + \int _{{\mathcal {I}}_0} \llbracket \delta {{\varvec{\varphi }}} \cdot {\varvec{P}}\rrbracket \cdot \overline{{\varvec{N}}}\mathrm {d}A - \int _{{\mathcal {B}}_0} \delta {\varvec{\varphi }} \cdot {\varvec{B}}^{\mathrm {p}} \mathrm {d}V \nonumber \\&\quad + \int _{{\mathcal {I}}_0} \overline{{\varvec{P}}} : \overline{{\mathrm {Grad}}}\delta \overline{{\varvec{\varphi }}}\mathrm {d}A - \int _{\partial {\mathcal {I}}_0^{{\mathrm {N}}}} \delta \overline{{\varvec{\varphi }}} \cdot [~\overline{{\varvec{P}}} \cdot \widetilde{{\varvec{N}}}~]\mathrm {d}L\nonumber \\&\quad - \int _{{\mathcal {I}}_0} \delta \overline{{\varvec{\varphi }}} \cdot \overline{{\varvec{B}}}{}^{\mathrm {p}} + \delta \overline{{\varvec{\varphi }}} \cdot [~\llbracket {\varvec{P}}\rrbracket \cdot \overline{{\varvec{N}}} ~] \mathrm {d}A = 0 . \end{aligned}$$
(43)

On the Neumann boundaries of the bulk and interface, \({\varvec{P}} \cdot {\varvec{N}} = \widehat{{\varvec{B}}}_{\mathrm {N}}^{\mathrm {p}} \) and \(\overline{{\varvec{P}}} \cdot \widetilde{{\varvec{N}}} = \widetilde{{\varvec{B}}}{}^{\mathrm {p}} _{\mathrm {N}} \), respectively. Noting , \(\llbracket \delta {{\varvec{\varphi }}} \rrbracket = {\varvec{0}}\), for coherent interfaces, and , Eq. (43) simplifies to the weak form Eq. (18).

1.4 Decoupled form of stress and elasticity tensor

It is sometimes useful to decouple the bulk deformation into the volumetric and isochoric part. In analogy, the volumetricFootnote 6 and isochoric part of the interface deformation read

$$\begin{aligned}&\overline{{\varvec{F}}} = \overline{{\varvec{F}}}{}^\text {vol}\overline{{\varvec{F}}}{}^\text {iso}\quad \text {with}\quad \overline{{\varvec{F}}}{}^\text {vol}= \overline{J}{}^{1/2}\overline{{\varvec{I}}}\nonumber \\&\quad \text {and}\quad \overline{{\varvec{F}}}{}^\text {iso}= \overline{J}{}^{-1/2} \overline{{\varvec{F}}} \;. \end{aligned}$$
(44)

Furthermore, the Helmholtz energy can be written as

$$\begin{aligned} \overline{\varPsi }\left( \overline{{\varvec{F}}}\right) = \overline{\varPsi }\left( \overline{{\varvec{F}}}{}^\text {iso},\,\overline{{\varvec{F}}}{}^\text {vol}\right) \;. \end{aligned}$$
(45)

The Piola stress reads

$$\begin{aligned} \begin{aligned}&\overline{{\varvec{P}}} {:}{=} \frac{\partial \overline{\varPsi }}{\partial \overline{{\varvec{F}}}} = \frac{\partial \overline{\varPsi }}{\partial \overline{{\varvec{F}}}{}^\text {iso}}:\frac{\partial \overline{{\varvec{F}}}{}^\text {iso}}{\partial \overline{{\varvec{F}}}} + \frac{\partial \overline{\varPsi }}{\partial \overline{{\varvec{F}}}{}^\text {vol}}:\frac{\partial \overline{{\varvec{F}}}{}^\text {vol}}{\partial \overline{{\varvec{F}}}}&\text {or}&\\&\overline{{\varvec{P}}} = \overline{{\varvec{P}}}{}^\text {iso}+\overline{{\varvec{P}}}{}^\text {vol}\;, \end{aligned} \end{aligned}$$
(46)

with

$$\begin{aligned} \overline{{\varvec{P}}}{}^\text {vol}= & {} \frac{\partial \overline{\varPsi }}{\partial \overline{{\varvec{F}}}{}^\text {vol}} : \frac{\partial \overline{{\varvec{F}}}{}^\text {vol}}{\partial \overline{{\varvec{F}}}} = \frac{\partial \overline{\varPsi }}{\partial \overline{{\varvec{F}}}{}^\text {vol}} : \left[ \frac{1}{2}\overline{J}{}^{1/2} \overline{{\varvec{I}}} \otimes \overline{{\varvec{F}}}{}^{-{\mathrm {t}}}\right] \;, \end{aligned}$$
(47)
$$\begin{aligned} {\overline{{\varvec{P}}}}{}^\text {iso}= & {} \frac{\partial \overline{\varPsi }}{\partial \overline{{\varvec{F}}}{}^\text {iso}} : \frac{\partial \overline{{\varvec{F}}}{}^\text {iso}}{\partial \overline{{\varvec{F}}}} = \overline{J}{}^{-1/2} \frac{\partial \overline{\varPsi }}{\partial \overline{{\varvec{F}}}{}^\text {iso}} :\left[ \overline{{\mathbb {I}}}-\displaystyle \frac{1}{2} \overline{{\varvec{F}}} \otimes \overline{{\varvec{F}}}{}^{-{\mathrm {t}}}\right] .\nonumber \\ \end{aligned}$$
(48)

The Piola stress tangent follows

$$\begin{aligned} \overline{{\mathbb {A}}} {:=} \frac{\partial \overline{{\varvec{P}}}}{\partial \overline{{\varvec{F}}}} = \frac{\partial \overline{{\varvec{P}}}{}^\text {iso}}{\partial \overline{{\varvec{F}}}} + \frac{\partial \overline{{\varvec{P}}}{}^\text {vol}}{\partial \overline{{\varvec{F}}}}&\text {or}&\overline{{\mathbb {A}}} = \overline{{\mathbb {A}}}{}^\text {iso}+ \overline{{\mathbb {A}}}{}^\text {vol}\;, \end{aligned}$$
(49)

with

$$\begin{aligned} \overline{{\mathbb {A}}}{}^\text {vol}= & {} \frac{\partial \overline{{\varvec{P}}}{}^\text {vol}}{\partial \overline{{\varvec{F}}}} = \frac{1}{4}\overline{{\varvec{F}}}{}^{-{\mathrm {t}}}\otimes \left[ \left[ \frac{\partial ^2\overline{\varPsi }}{\partial \overline{{\varvec{F}}}_\text {vol}\partial \overline{{\varvec{F}}}_\text {vol}} :\left[ \overline{{\varvec{I}}}\otimes \overline{{\varvec{F}}}{}^{-{\mathrm {t}}} \right] \right] :\overline{{\varvec{I}}} \right] \nonumber \\&- \frac{1}{4}\overline{{\varvec{F}}}{}^{-{\mathrm {t}}}\otimes \left[ \overline{J}{}^{-1/2}\frac{\partial \overline{\varPsi }}{\partial \overline{{\varvec{F}}}_\text {vol}}:\left[ \overline{{\varvec{I}}}\otimes \overline{{\varvec{F}}}{}^{-{\mathrm {t}}}\right] \right] \nonumber \\&+\frac{1}{2}\overline{J}{}^{-1/2}\left[ \frac{\partial \overline{\varPsi }}{\partial \overline{{\varvec{F}}}_\text {vol}} :\overline{{\varvec{I}}}\right] \overline{{\mathbb {D}}}\;, \end{aligned}$$
(50)
$$\begin{aligned} \overline{{\mathbb {A}}}{}^\text {iso}= & {} \frac{\partial \overline{{\varvec{P}}}{}^\text {iso}}{\partial \overline{{\varvec{F}}}}= -\frac{1}{2}\overline{J}{}^{-1/2}\left[ \frac{\partial \overline{\varPsi }}{\partial \overline{{\varvec{F}}}{}^\text {iso}} : \left[ \overline{{\mathbb {I}}}-\displaystyle \frac{1}{2} \overline{{\varvec{F}}} \otimes \overline{{\varvec{F}}}{}^{-{\mathrm {t}}}\right] \right] \otimes \overline{{\varvec{F}}}{}^{-{\mathrm {t}}} \nonumber \\&+ \overline{J}{}^{-1/2} \left[ \overline{{\mathbb {A}}}_1 - \frac{1}{2}\left[ \overline{{\mathbb {A}}}_2 + \overline{{\mathbb {A}}}_3 \right] \right] \;, \end{aligned}$$
(51)

where

$$\begin{aligned}&\overline{{\mathbb {A}}}_1 = \left[ \overline{{\varvec{I}}} ~ \overline{\otimes }~ {\varvec{i}}\right] :\frac{\partial ^2 \overline{\varPsi }}{\partial \overline{{\varvec{F}}}{}^\text {iso}\partial \overline{{\varvec{F}}}{}^\text {iso}}:\left[ \overline{{\mathbb {I}}}-\displaystyle \frac{1}{2} \overline{{\varvec{F}}} \otimes \overline{{\varvec{F}}}{}^{-{\mathrm {t}}}\right] \;,\nonumber \\&\overline{{\mathbb {A}}}_2 = \left[ \frac{\partial \overline{\varPsi }}{\partial \overline{{\varvec{F}}}{}^\text {iso}}: \overline{{\varvec{F}}}\right] \overline{{\mathbb {D}}}\;,\nonumber \\&\overline{{\mathbb {A}}}_3 = \overline{{\varvec{F}}}{}^{-{\mathrm {t}}}\otimes \left[ \frac{\partial \overline{\varPsi }}{\partial \overline{{\varvec{F}}}{}^\text {iso}}:\overline{{\mathbb {I}}} + \frac{\partial ^2 \overline{\varPsi }}{\partial \overline{{\varvec{F}}}{}^\text {iso}\partial \overline{{\varvec{F}}}{}^\text {iso}} :\left[ \overline{{\mathbb {I}}}-\displaystyle \frac{1}{2} \overline{{\varvec{F}}} \otimes \overline{{\varvec{F}}}{}^{-{\mathrm {t}}}\right] : \overline{{\varvec{F}}}\right] \;. \end{aligned}$$
(52)

Appendix 2: Differential geometry of two-dimensional manifolds embedded in three-dimensional space

In this section we briefly review some common terminologies in differential geometry of two-dimensional manifolds frequently used in this work to represent the interface elasticity theory. Finding the shortest distance (minimal geodesic) on a curved two-dimensional manifold (interface) embedded in a three-dimensional Euclidean space is presented subsequently, which is employed in the calculation of the non-local coefficients on the interface.

Fig. 16
figure 16

Illustration of the Cartesian basis \({\mathcal {E}}\) and an arbitrary basis \({\mathcal {A}}\) (a), and a cylindrical interface with its minimal geodesic (helix) along with the polar coordinate system (b). The red portion of the helix on the cylinder is the shortest arc-length of all the curves connecting the point \(\overline{\varvec{x}}_\text {r}\) and \(\overline{\varvec{x}}_\text {s}\). In (a), the \({\varvec{a}}_1\) and \({\varvec{a}}_2\) in \({\mathcal {A}}\) are not necessarily orthogonal to each other, whereas the Cartesian coordinate system \({\mathcal {E}}\) is composed of three orthogonal axes. In (b), every point on the surface of the cylinder is characterized by the coordinates \(\rho \) the radius, \(\theta \) the sweeping angle and z the height

A parametric interface \({\mathscr {I}}\) in \({\mathbb {E}}^3\) (three-dimensional embedding Euclidean space) is a map \({\mathscr {I}} : {\mathcal {I}} \rightarrow {\mathbb {E}}^3\) (with \({\mathcal {I}} \in {\mathbb {E}}^2 \)) such that its differential has rank 2 at all points \(\eta ^\alpha \in {\mathcal {I}}\) with \(\alpha = 1 \,,\, 2\). The interface can be defined by a parametric equation \( \overline{{\varvec{x}}} : {\mathcal {I}} \rightarrow {\mathbb {E}}^3\) as \( \overline{{\varvec{x}}} = \overline{{\varvec{x}}}(\eta ^\alpha )\), where \( \overline{{\varvec{x}}}\) is a vector-valued function of the scalar-valued parameters \(\eta ^\alpha \). The tangent space to \( \overline{{\varvec{x}}}\) at \(\eta ^\alpha \) is the linear map \(\mathrm {D}\overline{{\varvec{x}}}(\eta ^\alpha ) : {\mathbb {E}}^2 \rightarrow {\mathbb {E}}^3\) denoted by \(T {\mathscr {I}}\) where \(\mathrm {D}\overline{{\varvec{x}}}(\eta ^\alpha )\) is the differential of the interface. The tangent vectors \({\varvec{a}}_\alpha \in T {\mathscr {I}}\), i.e. the covariant interface basis vectors (see Fig. 16a) are then given by \({\varvec{a}}_\alpha = \partial _{\eta ^\alpha } \overline{{\varvec{x}}}(\eta ^\alpha )\).

The corresponding contravariant (dual) interface basis vectors \( {\varvec{a}}^\alpha \) are related to the covariant interface basis vectors by means of the co- and contravariant interface metric coefficients \(a_{\alpha \beta }\) (first fundamental form for the interface) and \(a^{\alpha \beta }\), respectively, as

$$\begin{aligned}&{\varvec{a}}^\alpha = a^{\alpha \beta }{\varvec{a}}_\beta \qquad \text {and}\qquad {\varvec{a}}_\alpha = a_{\alpha \beta }{\varvec{a}}^\beta \qquad \text {where}\nonumber \\&a_{\alpha \beta } = {\varvec{a}}_{\alpha }\cdot {\varvec{a}}_{\beta } \qquad \text {and} \qquad a^{\alpha \beta } = {\varvec{a}}^{\alpha } \cdot {\varvec{a}}^{\beta } \;. \end{aligned}$$
(53)

Note that the two metrics are inverse to each other, i.e. \([a^{\alpha \beta }] = [a_{\alpha \beta }]^{-1}\). The contra- and covariant base vectors \({\varvec{a}}_3\) and \({\varvec{a}}^3\), normal to \(T{\mathscr {I}}\), are defined by

$$\begin{aligned} {\varvec{a}}^3 {:=} {\varvec{a}}_1 \times {\varvec{a}}_2&\text {and}&{\varvec{a}}_3 {:=} [a^{33}]^{-1} {\varvec{a}}^3\;, \end{aligned}$$
(54)

such that \({\varvec{a}}^3 \cdot {\varvec{a}}_3 = 1\). Correspondingly the unit normal to the interface, parallel to \({\varvec{a}}^3\) and \({\varvec{a}}_3\), can be calculated as \(\overline{{\varvec{n}}} = {\varvec{a}}_3/|{\varvec{a}}_3| = {\varvec{a}}^3/|{\varvec{a}}^3|\). The interface identity tensor is defined as \(\overline{{\varvec{i}}} {:=} {\varvec{i}} - {\varvec{a}}_3 \otimes {\varvec{a}}^3 = {\varvec{i}} - \overline{{\varvec{n}}} \otimes \overline{{\varvec{n}}}\), where \({\varvec{i}}\) denotes the ordinary mixed-variant unit tensor of the embedding Euclidean space. The interface gradient, divergence and determinant operators in a general curvilinear coordinate are defined as

$$\begin{aligned}&\overline{{\mathrm {grad}}}\left\{ \overline{\bullet }\right\} {:=}\frac{\partial \left\{ \overline{\bullet }\right\} }{\partial \eta ^\alpha }\otimes {\varvec{a}}_\alpha , \quad \overline{{\mathrm {div}}}\left\{ \overline{\bullet }\right\} {:=}\frac{\partial \left\{ \overline{\bullet }\right\} }{\partial \eta ^\alpha }\cdot {\varvec{a}}_\alpha&\text {and}&\nonumber \\&\overline{{\mathrm {det}}}\left\{ \overline{\bullet }\right\} {:=}\frac{\left[ \left\{ \overline{\bullet }\right\} \cdot {\varvec{a}}_1\right] \times \left[ \left\{ \overline{\bullet }\right\} \cdot {\varvec{a}}_2\right] }{|{\varvec{a}}_1\times {\varvec{a}}_2|}\;. \end{aligned}$$
(55)

Having obtained the normal to the interface, co- and contravariant basis vectors, the interface curvature tensor \(\overline{{\varvec{k}}}\), second-order superficial deformation gradient tensor \(\overline{{\varvec{F}}}\) and its inverse \(\overline{{\varvec{f}}}\) are defined, respectively, as

$$\begin{aligned}&\overline{{\varvec{k}}} {:=} -\overline{{\mathrm {grad}}}\overline{{\varvec{n}}},\qquad \overline{{\varvec{F}}} {:=} \overline{{\mathrm {Grad}}}\overline{{\varvec{x}}} = {\varvec{a}}_\alpha \otimes {\varvec{A}}^\alpha&\text {and}&\nonumber \\&\overline{{\varvec{f}}} {:=} \overline{{\mathrm {grad}}}\overline{{\varvec{X}}} = {\varvec{A}}_\alpha \otimes {\varvec{a}}^\alpha \;. \end{aligned}$$
(56)

Note that due to the superficiality (rank deficiency) of the interface deformation gradient \(\overline{{\varvec{F}}}\), its inverse \(\overline{{\varvec{f}}}\) must be computed using Eq. (56)\(_3\).

Table 6 \(L_2\) norm of the residual for three scenarios: bulk and interface damaged, only bulk damaged and only interface damaged

The geodesics are the general form of straight lines when applied to curved, three-dimensional interfaces. The minimal geodesics in differential geometry are the shortest distance paths between two points on an interface. Clearly, minimal geodesics on the interfaces are curves of minimum arc-lengths. To find the minimal geodesics, first we introduce the parameter t on which the interface parameters \(\eta ^\alpha \) are dependent. The arc-length of the curve connecting any two points \(\overline{{\varvec{x}}}_\text {r}(t_1)\) and \(\overline{{\varvec{x}}}_\text {s}(t_2)\) on the curved interface can be written as

$$\begin{aligned} I = \int _{t_1}^{t_2} \sqrt{a_{\alpha \beta }\,\frac{\mathrm {d}\eta ^\alpha (t)}{\mathrm {d}t}\,\frac{\mathrm {d}\eta ^\beta (t)}{\mathrm {d}t}} ~\mathrm {d}t \;, \end{aligned}$$
(57)

where the integrand is the infinitesimal line element. Next, to find the shortest arc-length of the curves connecting the two points (minimal geodesic), the functional I is minimized. In doing so, the integrand in Eq. (57) (the Lagrangian), denoted by \(L(\eta ^\alpha (t), \dot{\eta }^\alpha (t))\) must satisfy the Euler-Lagrange equations of the form

$$\begin{aligned} \frac{\mathrm {d}}{\mathrm {d}t}\frac{\partial L}{\partial \dot{\eta }^\alpha } - \frac{\partial L}{\partial \eta ^\alpha }=0 \;, \end{aligned}$$
(58)

where \(\dot{\{\bullet \}}\) signifies a differentiation with respect to the parameter t.

Convergence behavior and run-time analysis

In this section we present some data on the convergence behavior and run-time of the computational problem at hand. Firstly, the \(L_2\) norms of the residual of few increments for all the three examples discussed in Sect. 6 are given in Table 6. As mentioned before, due to the consistent linearization, the asymptotic quadratic rate of convergence associated with the Newton–Raphson scheme is achieved. Secondly, to study how the interactive radius would change the memory consumption and run-time, we only allow the damage initiation in the bulk, increase both number of elementsFootnote 7 and interactive radius R and measure the run-time for one iteration per increment. These measurements are carried out on a machine with the following specifications:

  • processors: Intel Core i7-4770 CPU 3.40GHz \(\times \) 8,

  • memory: 15.6GB,

  • OS type: 64-bit.

The run-time measurements per iteration together with the memory usage for a serial and parallel code are given in Table 7. The parallel implementation is carried out using the MPI library. Note that the total run-time for a mesh size of \(100\times 100\times 1\) with \(R=0.01\) mm is approximately 39.3 h. A parallel implementation, using 4 processors, results in a speed-up of 3.5 and consequently a run-time of 7.75 h. As indicated by the presented data in Table 7, increasing the interactive radius R causes a substantial increase in the run-time. The main source of the time consumption, as expected, is in the stiffness assembly since firstly the non-localization is integral-type and secondly the linearization is consistent (introducing the double integrals in the stiffness formulation). However, as mentioned in Jirásek and Patzák (2002), if this type of non-localization is chosen, the extra time spent on the stiffness assembly due to consistent linearization in every iteration might be compensated by fewer iterations required to meet the convergence criterion. This matter becomes critical if a more stringent convergence criterion is necessary. For further details on the implementation issues of the consistent linearization of non-local damage problems of integral-type we refer to Jirásek and Patzák (2002).

Table 7 Serial- and parallel-code run-time measurements per iteration together with the memory consumption for various mesh sizes and values of interactive radius

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Esmaeili, A., Javili, A. & Steinmann, P. Coherent energetic interfaces accounting for in-plane degradation. Int J Fract 202, 135–165 (2016). https://doi.org/10.1007/s10704-016-0160-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10704-016-0160-4

Keywords

Navigation