Skip to main content
Log in

A micromechanical analysis of inter-fiber failure in long reinforced composites based on the phase field approach of fracture combined with the cohesive zone model

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

Abstract

The nature of failure in long fiber-reinforced composites is strongly affected by damage at the micro-scale. The presence of different phases at different length scales leads to a significant complexity in the failure progression. At the micro-scale, the complexity is due to the presence of different points of initiation for damage and the presence of cracks propagating both in the matrix and along the fiber–matrix interfaces. This scenario gives also the opportunity to improve the material design by modifying the properties of the different constituents in order to inhibit or delay some failure mechanisms. In view of this complexity, the development of predictive numerical tools with high capabilities in terms of reliability becomes of notable importance. In order to address this aspect, the combination of the phase-field approach for fracture and the cohesive zone model is herein exploited to demonstrate its capability and accuracy for the study of failure initiation at the micro-scale. Single-fiber problems subjected to transverse loading are considered as benchmark for the prediction of the sequence of stages of failure initiation, the size effect of the fiber radius on the apparent strength, the effect of a secondary tensile transverse load and the effect of a secondary neighbouring fiber. Numerical predictions are found to be in very good agreement with experimental trends and finite fracture mechanics predictions available in the literature.

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
Fig. 16
Fig. 17
Fig. 18
Fig. 19

Similar content being viewed by others

References

  • Ambati M, Gerasimov T, De Lorenzis L (2015) Phase-field modeling of ductile fracture. Comput Mech 55(5):1017–1040

    Google Scholar 

  • Arefi A, van der Meer FP (2018) Formulation of a consistent pressure-dependent damage model with fracture energy as input. Unpublished manuscript under review, TU Delft, Delft 201, pp 1–26

  • Arteiro A, Catalanotti G, Melro AR, Linde P, Camanho PP (2014) Micro-mechanical analysis of the in situ effect in polymer composite laminates. Compos Struct 116(1):827–840

    Google Scholar 

  • Arteiro A, Catalanotti G, Melro AR, Linde P, Camanho PP (2015) Micro-mechanical analysis of the effect of ply thickness on the transverse compressive strength of polymer composites. Compos Part A 79:127–137

    CAS  Google Scholar 

  • Bažant ZP (2005) Scaling of structural strength. Elsevier, Amsterdam

    Google Scholar 

  • Bourdin B, Francfort GA, Marigo J-J (2000) Numerical experiments in revisited brittle fracture. J Mech Phys Solids 48(4):797–826

    Google Scholar 

  • Carollo V, Reinoso J, Paggi M (2017) A 3D finite strain model for intralayer and interlayer crack simulation coupling the phase field approach and cohesive zone model. Compos Struct 182:636–651

    Google Scholar 

  • Carollo V, Reinoso J, Paggi M (2018) Modeling complex crack paths in ceramic laminates: A novel variational framework combining the phase field method of fracture and the cohesive zone model. J Eur Ceram Soc 38(8):2994–3003

    CAS  Google Scholar 

  • Comi C (1999) Computational modelling of gradient-enhanced damage in quasi-brittle materials. Mech Cohes Frict Mater 4(1):17–36

    Google Scholar 

  • Comi C, Perego U (2001) Fracture energy based bi-dissipative damage model for concrete. Int J Solids Struct 38(36):6427–6454

    Google Scholar 

  • Correa E, Mantič V, París F (2008) Numerical characterisation of the fibre-matrix interface crack growth in composites under transverse compression. Eng Fract Mech 75(14):4085–4103

    Google Scholar 

  • Correa E, París F, Mantič V (2014) Effect of a secondary transverse load on the inter-fibre failure under compression. Compos Part B 65:57–68

    CAS  Google Scholar 

  • Correa E, París F, Mantič V (2013) Effect of the presence of a secondary transverse load on the inter-fibre failure under tension. Eng Fract Mech 103:174–189

    Google Scholar 

  • Danzi F, Fanteria D, Panettieri E, Palermo M (2017) A numerical micro-mechanical study of the influence of fibermatrix interphase failure on carbon/epoxy material properties. Compos Struct 159:625–635

    Google Scholar 

  • García IG, Mantič V, Graciani E (2015) Debonding at the fibre-matrix interface under remote transverse tension. One debond or two symmetric debonds? Eur J Mech A Solids 53:75–88

    Google Scholar 

  • García IG, Paggi M, Mantič V (2014) Fiber-size effects on the onset of fiber-matrix debonding under transverse tension: a comparison between cohesive zone and finite fracture mechanics models. Eng Fract Mech 115:96–110

    Google Scholar 

  • Griffith AA (1921) The phenomena of rupture and flow in solids’, Philosophical transaction of the royal society of London. Series A containing papers of a Mathematical or physical character 221:163–198

    Google Scholar 

  • Hansen-Dörr AC, de Borst R, Hennig P, Kästner M (2019) Phase-field modelling of interface failure in brittle materials. Comput Methods Appl Mech Eng 346:25–42

    Google Scholar 

  • Herráez M, Mora D, Naya F, Lopes CS, González C, Llorca J (2015) Transverse cracking of cross-ply laminates: A computational micromechanics perspective. Compos Sci Technol 110:196–204

    Google Scholar 

  • Legarth BN, Yang Q (2016) Micromechanical analyses of debonding and matrix cracking in dual-phase materials. J Appl Mech 83(5):051006

    Google Scholar 

  • Linse T, Hennig P, de Kästner M, Borst R (2017) A convergence study of phase-field models for brittle fracture. Eng Fract Mech 184:307–318

    Google Scholar 

  • Maimí P, Camanho PP, Mayugo JA, Dávila CG (2007) A continuum damage model for composite laminates: Part I—constitutive model. Mech Mater 39:897–908

    Google Scholar 

  • Mantič V (2009) Interface crack onset at a circular cylindrical inclusion under a remote transverse tension. Application of a coupled stress and energy criterion. Int J Solids Struct 46(6):1287–1304

    Google Scholar 

  • Mantič V, García IG (2012) Crack onset and growth at the fibre-matrix interface under a remote biaxial transverse load. Application of a coupled stress and energy criterion. Int J Solids Struct 49(17):2273–2290

    Google Scholar 

  • Melro AR, Camanho PP, Andrade Pires FM, Pinho ST (2013) Micromechanical analysis of polymer composites reinforced by unidirectional fibres: Part II-micromechanical analyses. Int J Solids Struct 50(11–12):1906–1915

    CAS  Google Scholar 

  • Miehe C, Hofacker M, Welschinger F (2010a) A phase field model for rate-independent crack propagation: Robust algorithmic implementation based on operator splits. Comput Methods Appl Mech Eng 199(45):2765–2778

    Google Scholar 

  • Miehe C, Welschinger F, Hofacker M (2010b) Thermodynamically consistent phase-field models of fracture: variational principles and multi-field fe implementations. Int J Numer Methods Eng 83(10):1273–1311

    Google Scholar 

  • Msekh M-A, Sargado J-M, Jamshidian M, Areias P, Rabczuk T (2015) Abaqus implementation of phase-field model for brittle fracture. Comput Mater Sci 96:472–484

    Google Scholar 

  • Nguyen T-T, Yvonnet J, Bornert M, Chateau C (2016a) Initiation and propagation of complex 3D networks of cracks in heterogeneous quasi-brittle materials: direct comparison between in situ testing-microct experiments and phase field simulations. J Mech Phys Solids 95:320–350

    CAS  Google Scholar 

  • Nguyen T-T, Yvonnet J, Zhu Q-Z, Bornert M, Chateau C (2016b) A phase-field method for computational modeling of interfacial damage interacting with crack propagation in realistic microstructures obtained by microtomography. Comput Methods Appl Mech Eng 312:567–595

    Google Scholar 

  • Ortiz M, Pandolfi A (1999) Finite deformation irreversible cohesive elements for three-dimensional crack-propagation analysis. Int J Numer Methods Eng 44:1267–1282

    Google Scholar 

  • Paggi M, Corrado M, Reinoso J (2018) Fracture of solar-grade anisotropic polycrystalline silicon: A combined phase field-cohesive zone model approach. Comput Methods Appl Mech Eng 330:123–148

    Google Scholar 

  • Paggi M, Reinoso J (2017) Revisiting the problem of a crack impinging on an interface: a modeling framework for the interaction between the phase field approach for brittle fracture and the interface cohesive zone model. Comput Methods Appl Mech Eng 321:145–172

    Google Scholar 

  • Paggi M, Wriggers P (2012) Stiffness and strength of hierarchical polycrystalline materials with imperfect interfaces. J Mech Phys Solids 60(4):557–572

    Google Scholar 

  • Puck A, Schürmann H (2004) Failure analysis of FRP laminates by means of physically based phenomenological models. Fail Criteria Fibre Reinf Polym Compos 3538(96):832–876

    Google Scholar 

  • Quintanas-Corominas A, Maimí P, Casoni E, Turon A, Mayugo JA, Guillamet G, Vázquez M (2018) A 3D transversally isotropic constitutive model for advanced composites implemented in a high performance computing code. Eur J Mech A 71:278–291

    Google Scholar 

  • Reinoso J, Arteiro A, Paggi M, Camanho P (2017a) Strength prediction of notched thin ply laminates using finite fracture mechanics and the phase field approach. Compos Sci Technol 150:205–216

    CAS  Google Scholar 

  • Reinoso J, Catalanotti G, Blázquez A, Areias P, Camanho P, París F (2017b) A consistent anisotropic damage model for laminated fiber-reinforced composites using the 3D-version of the puck failure criterion. Int J Solids Struct 126–127:37–53

    Google Scholar 

  • Reinoso J, Paggi M (2014) A consistent interface element formulation for geometrical and material nonlinearities. Comput Mech 54(6):1569–1581

    Google Scholar 

  • Reinoso J, Paggi M, Linder C (2017c) Phase field modeling of brittle fracture for enhanced assumed strain shells at large deformations: formulation and finite element implementation. Comput Mech 59(6):1–21

    Google Scholar 

  • Rodríguez M, Molina-Aldareguía JM, González C, Llorca J (2012) A methodology to measure the interface shear strength by means of the fiber push-in test. Compos Sci Technol 72(15):1924–1932

    Google Scholar 

  • Steinke C, Kaliske M (2018) A phase-field crack model based on directional stress decomposition. Comput Mech 63(5):1019–1046

    Google Scholar 

  • Strobl M, Seelig T (2015) A novel treatment of crack boundary conditions in phase field models of fracture. PAMM 15(1):155–156

    Google Scholar 

  • Tan W, Naya F, Yang L, Chang T, Falzon BG, Zhan L, Molina-Aldareguía JM, González C, Llorca J (2018) The role of interfacial properties on the intralaminar and interlaminar damage behaviour of unidirectional composite laminates: experimental characterization and multiscale modelling. Compos Part B 138:206–221

    CAS  Google Scholar 

  • Tanné E, Li T, Bourdin B, Marigo J-J, Maurini C (2018) Crack nucleation in variational phase-field models of brittle fracture. J Mech Phys Solids 110:80–99

    Google Scholar 

  • Távara L, Mantič V, Graciani E, París F (2011) BEM analysis of crack onset and propagation along fiber-matrix interface under transverse tension using a linear elastic-brittle interface model. Eng Anal Bound Elem 35(2):207–222

    Google Scholar 

  • Távara L, Mantič V, Graciani E, París F (2016) Modelling interfacial debonds in unidirectional fibre-reinforced composites under biaxial transverse loads. Compos Struct 136:305–312

    Google Scholar 

  • Távara L, Reinoso J, Castillo D, Mantič V (2018) Mixed-mode failure of interfaces studied by the 2D linear elastic-brittle interface model: macro-and micro-mechanical finite-element applications in composites. J Adhesion 94(8):627–656

    Google Scholar 

  • Totry E, Molina-Aldareguía JM, González C, LLorca J (2010) Effect of fiber, matrix and interface properties on the in-plane shear deformation of carbon-fiber reinforced composites. Compos Sci Technol 70(6):970–980

    CAS  Google Scholar 

  • Turon A, Camanho P, Costa J, Renart J (2010) Accurate simulation of delamination growth under mixed-mode loading using cohesive elements: definition of interlaminar strengths and elastic stiffness. Compos Struct 92(8):1857–1864

    Google Scholar 

  • Van Der Meer FP (2016) Micromechanical validation of a mesomodel for plasticity in composites. Eur J Mech A 60:58–69

    Google Scholar 

  • Verhoosel CV, de Borst R (2013) A phase-field model for cohesive fracture. Int J Numer Methods Eng 96(1):43–62

    Google Scholar 

  • Xu X-P, Needleman A (1994) Numerical simulations of fast crack growth in brittle solids. J Mech Phys Solids 42(9):1397–1434

    Google Scholar 

  • Zhuang L, Pupurs A, Varna J, Talreja R, Ayadi Z (2018a) Effects of inter-fiber spacing on fiber-matrix debond crack growth in unidirectional composites under transverse loading. Compos Part A 109:463–471

    CAS  Google Scholar 

  • Zhuang L, Talreja R, Varna J (2018b) Transverse crack formation in unidirectional composites by linking of fibre/matrix debond cracks. Compos Part A 107:294–303

    CAS  Google Scholar 

Download references

Acknowledgements

IGG and JR acknowledges the support of the projects funded by the Spanish Ministry of Economy and Competitiveness/FEDER (Projects MAT2015-71036-P and MAT2015-71309-P) and the Andalusian Government (Projects of Excellence Nos. TEP-7093 and P12-TEP-1050).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to J. Reinoso.

Additional information

Publisher's Note

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

Appendices

Appendix A: Numerical implementation of the phase field approach of fracture in the bulk

In the sequel, the main aspects of the numerical implementation of the PF approach of fracture for triggering failure events within the bulk are described within the FEM framework.

The displacements and the phase field variable are interpolated using an isoparametric approach using standard Lagrangian shape functions \(N^{I}({\varvec{\xi }})\) defined in the parametric domain \({\varvec{\xi }} = \{\xi ^{1}, \xi ^{2} \}\) and arranged in the vector \({\mathbf {N}}\). Therefore, the interpolation of these fields, their variation and their linearization adopt the following scheme:

$$\begin{aligned} {\mathbf {x}}\cong & {} \sum _{I=1}^{n} N^{I} \widetilde{{\mathbf {x}}}_{I} = {\mathbf {N}} \widetilde{{\mathbf {x}}}; \quad {\mathbf {u}} \cong \sum _{I=1}^{n} N^{I} {\mathbf {d}}_{I} = {\mathbf {N}} {\mathbf {d}}; \nonumber \\ \delta {\mathbf {u}}\cong & {} \sum _{I=1}^{n} N^{I} \delta {\mathbf {d}}_{I} = {\mathbf {N}} \delta {\mathbf {d}}; \nonumber \\ \varDelta {\mathbf {u}}\cong & {} \sum _{I=1}^{n} N^{I} \varDelta {\mathbf {d}}_{I} = {\mathbf {N}} \delta {\mathbf {d}}, \end{aligned}$$
(A.17)
$$\begin{aligned} {\mathfrak {d}}\cong & {} \sum _{I=1}^{n} N^{I} {\mathfrak {\overline{d}}}_{I} = {\mathbf {N}} {\mathfrak {\overline{d}}}; \quad \delta {\mathfrak {d}} \cong \sum _{I=1}^{n} N^{I} \delta {\mathfrak {\overline{d}}}_{I} = {\mathbf {N}} \delta {\mathfrak {\overline{d}}}; \nonumber \\ \varDelta {\mathfrak {d}}\cong & {} \sum _{I=1}^{n} N^{I} \varDelta {\mathfrak {\overline{d}}}_{I} = {\mathbf {N}} \varDelta {\mathfrak {\overline{d}}}. \end{aligned}$$
(A.18)

In Eqs. (A.17), (A.18), \({\mathbf {d}}_{I}\) and \({\mathfrak {\overline{d}}}_{I}\) are the nodal displacement vector and phase field values, respectively, which are arranged at the element level in the vectors \({\mathbf {d}}\) and \({\mathfrak {\overline{d}}}\)

$$\begin{aligned} {\varvec{\varepsilon }} \cong {\mathbf {B}}_{{\mathbf {d}}} {\mathbf {d}}; \quad \delta {\varvec{\varepsilon }} \cong {\mathbf {B}}_{{\mathbf {d}}} \delta {\mathbf {d}}; \quad \varDelta {\varvec{\varepsilon }} \cong {\mathbf {B}}_{{\mathbf {d}}} \varDelta {\mathbf {d}} \end{aligned}$$
(A.19)

Moreover, the spatial gradient of the phase field variable can be approximated as follows:

$$\begin{aligned} \nabla _{{\mathbf {x}}} {\mathfrak {d}} \cong {\mathbf {B}}_{{\mathfrak {d}}} {\mathfrak {\overline{d}}}; \quad \nabla _{{\mathbf {x}}} (\delta {\mathfrak {d}}) \cong {\mathbf {B}}_{{\mathfrak {d}}} \delta {\mathfrak {\overline{d}}}; \quad \nabla _{{\mathbf {x}}} (\varDelta {\mathfrak {d}}) \cong {\mathbf {B}}_{{\mathfrak {d}}} \varDelta {\mathfrak {\overline{d}}},\nonumber \\ \end{aligned}$$
(A.20)

being \({\mathbf {B}}_{{\mathfrak {d}}}\) a phase field compatibility-like operator.

The discrete form of Eq. (6) at the element level, being denoted by the superscript el, renders

$$\begin{aligned}&\delta {\tilde{\varPi }}_{b}^{el} ({\mathbf {d}}, \delta {\mathbf {d}}, {\mathfrak {\overline{d}}} , \delta {\mathfrak {\overline{d}}}) \nonumber \\&\quad = \delta {\mathbf {d}}^{\text {T}} \left\{ \int _{{\mathcal {B}}^{el}} \left[ \left( \left( 1 - {\mathfrak {d}} \right) ^{2} + {\mathcal {K}} \right) {\mathbf {B}}_{{\mathbf {d}}}^{\text {T}} {\varvec{\sigma }} \right. \right. \nonumber \\&\qquad \left. \left. + \, {\mathbf {B}}_{{\mathbf {d}}}^{\text {T}} {\varvec{\sigma }}_{-} \right] \,\mathrm {d}\varOmega - \int _{ \partial {\mathcal {B}}^{el}} {\mathbf {N}}^{\text {T}} \overline{{\mathbf {t}}} \,\mathrm {d} \partial \varOmega - \int _{{\mathcal {B}}^{el}} {\mathbf {N}}^{\text {T}} {\mathbf {f}}_{v} \,\mathrm {d}\varOmega \right\} \nonumber \\&\qquad +\delta {\mathfrak {\overline{d}}}^{\text {T}} \left\{ \int _{ {\mathcal {B}}^{el}} - 2(1- {\mathfrak {d}}) {\mathbf {N}}^{\text {T}} \psi ^{e}_{+} ({\varvec{\varepsilon }}) \,\mathrm {d}\varOmega \right. \nonumber \\&\qquad \left. + \int _{ {\mathcal {B}}^{el}} {\mathcal {G}}_{c}^{b} l\left( {\mathbf {B}}_{{\mathfrak {d}}}^{\text {T}} \nabla _{{\mathbf {x}}}{\mathfrak {d}} + \frac{1}{l^{2}} {\mathbf {N}}^{\text {T}} {\mathfrak {d}} \right) \,\mathrm {d}\varOmega \right\} \nonumber \\&\quad = \delta {\mathbf {d}}^{\text {T}} {\mathbf {f}}_{{\mathbf {d}}}^{b} + \delta {\mathfrak {\overline{d}}}^{\text {T}} {\mathbf {f}}_{{\mathfrak {d}}}^{b} \end{aligned}$$
(A.21)

where

$$\begin{aligned} {\mathbf {f}}_{{\mathbf {d}},\text {int}}^{b}= & {} \int _{{\mathcal {B}}^{el}} \left[ \left( \left( 1 - {\mathfrak {d}} \right) ^{2} + {\mathcal {K}} \right) {\mathbf {B}}_{{\mathbf {d}}}^{\text {T}} {\varvec{\sigma }}_{+} + {\mathbf {B}}_{{\mathbf {d}}}^{\text {T}} {\varvec{\sigma }}_{-}\right] \,\mathrm {d}\varOmega ,\nonumber \\ \end{aligned}$$
(A.22)
$$\begin{aligned} {\mathbf {f}}_{{\mathbf {d}},\text {ext}}^{b}= & {} \int _{ \partial {\mathcal {B}}^{el}} {\mathbf {N}}^{\text {T}} \overline{{\mathbf {t}}} \,\mathrm {d} \partial \varOmega + \int _{{\mathcal {B}}} {\mathbf {N}}^{\text {T}} {\mathbf {f}}_{v} \,\mathrm {d}\varOmega , \end{aligned}$$
(A.23)
$$\begin{aligned} {\mathbf {f}}_{{\mathfrak {d}}}^{b}= & {} \int _{ {\mathcal {B}}^{el}} - 2(1- {\mathfrak {d}}) {\mathbf {N}}^{\text {T}} \psi ^{e}_{+} ({\varvec{\varepsilon }}) \,\mathrm {d}\varOmega \nonumber \\&+ \int _{ {\mathcal {B}}^{el}} {\mathcal {G}}_{c}^{b} l\left[ {\mathbf {B}}_{{\mathfrak {d}}}^{\text {T}} \nabla _{{\mathbf {x}}} + {\mathfrak {d}} \frac{1}{l^{2}} {\mathbf {N}}^{\text {T}} {\mathfrak {d}} \right] \,\mathrm {d}\varOmega . \end{aligned}$$
(A.24)

The vectors \({\mathbf {f}}_{{\mathbf {d}},\text {int}}^{b}\) and \({\mathbf {f}}_{{\mathbf {d}},\text {ext}}^{b}\) correspond to the internal and residual vectors associated with the displacement field, whilst \({\mathbf {f}}_{{\mathfrak {d}}}^{b}\) stands for the residual vector of the phase field.

The linearization of the previous residual vector yields to the following system (Nguyen et al. 2016b, a; Ambati et al. 2015):

$$\begin{aligned} \begin{bmatrix} {\mathbf {K}}_{{\mathbf {d}}{\mathbf {d}}}^{b}&{\mathbf {0}} \\ {\mathbf {0}}&{\mathbf {K}}_{ {\mathfrak {d}} {\mathfrak {d}}}^{b} \end{bmatrix} \begin{bmatrix} \varDelta {\mathbf {d}} \\ \varDelta {\mathfrak {d}} \end{bmatrix} =\begin{bmatrix} {\mathbf {f}}_{{\mathbf {d}},\text {ext}}^{b} \\ 0 \end{bmatrix} - \begin{bmatrix} {\mathbf {f}}_{{\mathbf {d}},\text {int}}^{b} \\ {\mathbf {f}}_{ {\mathfrak {d}}}^{b} \end{bmatrix}. \end{aligned}$$
(A.25)

The system given in Eq. (A.25) is solved in Jacobi-type solution scheme in line with (Miehe et al. 2010a), though the fully coupled system can be computed via the incorporation of a viscous term in order to guarantee the convexity of the corresponding formulation. It is worth noting that the modification of the formulation in order to incorporate this viscous term is not addressed in the current paper since a staggered solution scheme has been adopted. Moreover, the previous formulation is equipped with a history variable into a modified version of the residual vector of the phase field in line with (Miehe et al. 2010b; Reinoso et al. 2017c):

$$\begin{aligned} {\mathbf {f}}_{{\mathfrak {d}}}^{b}= & {} \int _{ {\mathcal {B}}^{el}} - 2(1- {\mathfrak {d}}) {\mathbf {N}}^{\text {T}} \left[ \psi ^{e}_{+} ({\varvec{\varepsilon }})\right. \nonumber \\&\left. - \frac{\eta }{n \varDelta t} \left\langle {\dot{\overline{\mathfrak {d}}}} \right\rangle _{-}^{n}\right] \,\mathrm {d}\varOmega \nonumber \\&+ \int _{ {\mathcal {B}}^{el}} {\mathcal {G}}_{c}^{b} l\left[ {\mathbf {B}}_{{\mathfrak {d}}}^{\text {T}} \nabla _{{\mathbf {x}}} + {\mathfrak {d}} \frac{1}{l^{2}} {\mathbf {N}}^{\text {T}} {\mathfrak {d}} \right] \,\mathrm {d}\varOmega . \end{aligned}$$
(A.26)

where \(\eta \) is a penalty parameter which is set \(n=2\) (Msekh et al. 2015), which was previously examined by the authors in Reinoso et al. (2017c, 2017a) with respect to experimental data leading to satisfactory predictions, see (Linse et al. 2017) for a very comprehensive and detailed assessment regarding the convergence of PF methods; \(\left\langle {\dot{\overline{\mathfrak {d}}}}\right\rangle _{-}^{n}= \left\langle {\overline{\mathfrak {d}}}^{t+1} - {\overline{\mathfrak {d}}}^{t}\right\rangle _{-}^{n} \) and the operator \(\left\langle \right\rangle _{-}\) for a variable a renders:

$$\begin{aligned} \left\langle a\right\rangle _{-} = \left\{ \begin{aligned} - a\text {, }&\text { } a < 0, \\ 0\text {, }&\text { } a \ge 0. \end{aligned} \right. \end{aligned}$$
(A.27)

Finally the particular form of the tangent matrices \({\mathbf {K}}_{{\mathbf {d}}{\mathbf {d}}}^{b}\) and \({\mathbf {K}}_{ {\mathfrak {d}}{\mathfrak {d}}}^{b}\) are not included in the present manuscript for conciseness reasons.

Appendix B: Variational formulation and numerical implementation of the interface element compatible with the PF approach of fracture

This Appendix presents the variational formulation regarding the interface contribution to the total functional of the system given in Eq. (10). Complying with a standard Galerkin procedure, the weak form of the interface contribution reads

$$\begin{aligned}&\delta \varPi _{\varGamma _{i}} ({\mathbf {u}}, \delta {\mathbf {u}}, {\mathfrak {d}}, \delta {\mathfrak {d}}) \nonumber \\&\quad = \int _{\varGamma _{i}} \left( \dfrac{\partial {\mathcal {G}}^{i}({\mathbf {u}},{\mathfrak {d}})}{\partial {\mathbf {u}}}\delta {\mathbf {u}}+\dfrac{\partial {\mathcal {G}}^{i}({\mathbf {u}},{\mathfrak {d}})}{\partial {\mathfrak {d}}}\delta {\mathfrak {d}} \right) \,\mathrm {d}\varGamma ,\nonumber \\ \end{aligned}$$
(B.28)

In Eq. (B.28), \(\delta {\mathbf {u}}\) denotes the admissible kinematic field vector function (\({\mathfrak {V}}^{u} = \big \{\delta {\mathbf {u}}\, | \, {\mathbf {u}} = \overline{{\mathbf {u}}} \text{ on } \partial {\mathcal {B}}_{u} , {\mathbf {u}} \in {\mathcal {H}}^{1} \big \}\)), whereas \(\delta {\mathfrak {d}}\) identifies the approximation functions of the phase field variable (\({\mathfrak {V}}^{{\mathfrak {d}}} = \big \{\delta {\mathfrak {d}} \, | \, \delta {\mathfrak {d}} = 0 \text{ on } \varGamma _{b} , {\mathfrak {d}} \in {\mathcal {H}}^{0} \big \}\)).

The discretization of the previous variational form can be carried out within the context of the FEM through the use of standard isoparametric elements. Without loss of generality, we adopt a first-order interpolation scheme for the kinematic and the phase field variables. Thus, \({\mathbf {d}}\) represents the nodal-based displacement field and \(\bar{{\mathfrak {d}}}\) denotes the nodal-based phase field, both vectors being defined at the element level for the corresponding numerical implementation.

Accordingly, the discrete form of Eq. (B.28) at the element level \(\varGamma _{i}^{el}\) (\(\varGamma _i \sim \bigcup \varGamma _{i}^{el}\)) adopts the form:

$$\begin{aligned}&\delta {\tilde{\varPi }}^{el}_{\varGamma _i} ({\mathbf {d}}, \delta {\mathbf {d}}, \bar{{\mathfrak {d}}}, \delta \bar{{\mathfrak {d}}})\nonumber \\&\quad =\int _{\varGamma _i^{el}} \left( \dfrac{\partial {\mathcal {G}}^{i} ({\mathbf {d}},\bar{{\mathfrak {d}}})}{\partial {\mathbf {d}}}\delta {\mathbf {d}}+\dfrac{\partial {\mathcal {G}}^{i} ({\mathbf {d}},\bar{{\mathfrak {d}}})}{\partial \bar{{\mathfrak {d}}}}\delta \bar{{\mathfrak {d}}} \right) \,\mathrm {d}\varGamma ,\nonumber \\ \end{aligned}$$
(B.29)

where \({\mathcal {G}}^{i}={\mathcal {G}}_{I}^{i}+{\mathcal {G}}_{II}^{i}\):

The displacement vector between the crack flanks along the interface is characterized by the gap vector, \({\mathbf {g}}\). The discrete form of the gap vector for each point of \(\varGamma _i^{el}\) can be computed as the difference between the displacements of opposite points along the interface, which can be obtained from the nodal displacements \({\mathbf {d}}\) multiplied by the average matrix \({\mathbf {L}}\) and the interpolation matrix \({\mathbf {N}}\):

$$\begin{aligned} {\mathbf {g}}={\mathbf {N}}{\mathbf {L}}{\mathbf {d}}= \hat{{\mathbf {B}}}_{{\mathbf {d}}} {\mathbf {d}}, \end{aligned}$$
(B.30)

identifying \(\hat{{\mathbf {B}}}_{{\mathbf {d}}}={\mathbf {N}}{\mathbf {L}}\) the interface compatibility operator.

The constitutive law at the interface requires the consideration of a local setting, which is defined by the normal and tangential unit vectors (Paggi and Wriggers 2012). Correspondingly, the gap vector in the global setting (Eq. (B.30)) is multiplied by rotation matrix \({\mathbf {R}}\) in order to obtain the gap vector in the local setting \({\mathbf {g}}_{\text {loc}}\):

$$\begin{aligned} {\mathbf {g}}_{\text {loc}} \cong {\mathbf {R}}{\mathbf {g}}={\mathbf {R}}\hat{{\mathbf {B}}}_{{\mathbf {b}}} {\mathbf {d}}. \end{aligned}$$
(B.31)

In a similar manner, the discrete average phase field variable \({\mathfrak {d}}\) at the interface (\(\varGamma ^{el}_i\)) can be computed at the element level as follows:

$$\begin{aligned} {\mathfrak {d}} \cong {\mathbf {N}}_{{\mathfrak {d}}} {\mathbf {M}}_{{\mathfrak {d}}} \bar{{\mathfrak {d}}} ={\hat{{\mathbf {B}}}}_{{\mathfrak {d}}} \bar{{\mathfrak {d}}}, \end{aligned}$$
(B.32)

where \({\mathbf {M}}_{{\mathfrak {d}}}\) is an matrix for determining the average value of the phase field variable between the interface flanks, and \({\hat{{\mathbf {B}}}}_{{\mathfrak {d}}}={\mathbf {N}}_{{\mathfrak {d}}} {\mathbf {M}}_{{\mathfrak {d}}}\) identifies the compatibility operator for the phase field. The particular form of such matrices are derived in Reinoso and Paggi (2014) and Paggi and Reinoso (2017), being omitted here for the sake of brevity.

Then, the discrete variational form renders

$$\begin{aligned} \begin{aligned}&\delta {\tilde{\varPi }}_{\varGamma _i}^{el} ({\mathbf {d}}, \delta {\mathbf {d}}, \bar{{\mathfrak {d}}}, \delta \bar{{\mathfrak {d}}}) \\&\quad =\delta {\mathbf {d}}^{\mathrm {T}}\int _{\varGamma _i^{el}} \left( \dfrac{\partial {\mathcal {G}}^{i} ({\mathbf {d}}, \bar{{\mathfrak {d}}})}{\partial {\mathbf {d}}}\right) ^{\mathrm {T}}\,\mathrm {d}\varGamma \\&\qquad + \delta \bar{{\mathfrak {d}}}^{\mathrm {T}}\int _{\varGamma _i^{el}} \left( \dfrac{\partial {\mathcal {G}}^{i} ({\mathbf {d}}, \bar{{\mathfrak {d}}})}{\partial \bar{{\mathfrak {d}}}}\right) ^{\mathrm {T}}\,\mathrm {d}\varGamma \\&\quad =\delta {\mathbf {d}}^{\mathrm {T}}\int _{\varGamma _i^{el}} \hat{{\mathbf {B}}}_{{\mathbf {d}}}^{\mathrm {T}}{\mathbf {R}}^{\mathrm {T}}\left( \dfrac{\partial {\mathcal {G}}^{i} ({\mathbf {d}}, \bar{{\mathfrak {d}}}) }{\partial {\mathbf {g}}_{\text {loc}}}\right) ^{\mathrm {T}}\,\mathrm {d}\varGamma \\&\qquad + \delta \bar{{\mathfrak {d}}}^{\mathrm {T}}\int _{\varGamma _i^{el}} {\hat{{\mathbf {B}}}}_{{\mathfrak {d}}}^{\mathrm {T}}\left( \dfrac{\partial {\mathcal {G}}^{i}({\mathbf {d}}, \bar{{\mathfrak {d}}})}{\partial \bar{{\mathfrak {d}}}}\right) ^{\mathrm {T}}\,\mathrm {d}\varGamma \end{aligned} \end{aligned}$$
(B.33)

leading to the residual vectors

$$\begin{aligned} {\mathbf {f}}_{{\mathbf {d}}}^{i}&=\int _{\varGamma _i^{el}}\hat{{\mathbf {B}}}_{{\mathbf {d}}}^{\mathrm {T}}{\mathbf {R}}^{\mathrm {T}}\left( \dfrac{\partial {\mathcal {G}}^{i} ({\mathbf {d}}, \bar{{\mathfrak {d}}}) }{\partial {\mathbf {g}}_{\text {loc}}}\right) ^{\mathrm {T}}\,\mathrm {d}\varGamma , \end{aligned}$$
(B.34a)
$$\begin{aligned} {\mathbf {f}}_{{\mathfrak {d}}}^{i}&=\int _{\varGamma _i^{el}}{\hat{{\mathbf {B}}}}_{{\mathfrak {d}}}^{\mathrm {T}}\left( \dfrac{\partial {\mathcal {G}}^{i}({\mathbf {d}}, \bar{{\mathfrak {d}}})}{\partial {\mathfrak {d}}}\right) ^{\mathrm {T}}\,\mathrm {d}\varGamma . \end{aligned}$$
(B.34b)

Finally, the consistent linearization of the previous residual vectors allows the computation of the element tangents of the proposed interface formulation:

$$\begin{aligned} {\mathbf {K}}_{{\mathbf {dd}}}^{i}&=\dfrac{\partial {\mathbf {f}}_{{\mathbf {d}}}}{\partial {\mathbf {d}}}=\int _{\varGamma _i^{el}}\hat{{\mathbf {B}}}_{{\mathbf {d}}}^{\mathrm {T}}{\mathbf {R}}^{\mathrm {T}}{\mathbb {C}}_{{\mathbf {dd}}}^{i}{\mathbf {R}}\hat{{\mathbf {B}}}_{{\mathbf {d}}}\,\mathrm {d}\varGamma , \end{aligned}$$
(B.35a)
$$\begin{aligned} {\mathbf {K}}_{{\mathbf {d}}{\mathfrak {d}}}^{i}&=\dfrac{\partial {\mathbf {f}}_{{\mathbf {d}}}}{\partial {\mathfrak {d}}}=\int _{\varGamma _i^{el}}\hat{{\mathbf {B}}}_{{\mathbf {d}}} ^{\mathrm {T}}{\mathbf {R}}^{\mathrm {T}}{\mathbb {C}}_{{\mathbf {d}}{\mathfrak {d}}}^{i}{\hat{{\mathbf {B}}}}_{{\mathfrak {d}}}\,\mathrm {d}\varGamma , \end{aligned}$$
(B.35b)
$$\begin{aligned} {\mathbf {K}}_{{\mathfrak {d}}{\mathbf {d}}}^{i}&=\dfrac{\partial {\mathbf {f}}_{\mathfrak {d}}}{\partial {\mathbf {d}}}=\int _{\varGamma _i^{el}}{\hat{{\mathbf {B}}}}_{{\mathfrak {d}}}^{\mathrm {T}}{\mathbb {C}}_{{\mathfrak {d}}d}^{i}{\mathbf {R}}\hat{{\mathbf {B}}}_{{\mathbf {d}}}\,\mathrm {d}\varGamma , \end{aligned}$$
(B.35c)
$$\begin{aligned} {\mathbf {K}}_{{\mathfrak {d}}{\mathfrak {d}}}^{i}&=\dfrac{\partial {\mathbf {f}}_{\mathfrak {d}}}{\partial {\mathfrak {d}}}=\int _{\varGamma _i^{el}}{\hat{{\mathbf {B}}}}_{{\mathfrak {d}}} ^{\mathrm {T}}{\mathbb {C}}_{{\mathfrak {d}}{\mathfrak {d}}}^{i}{\hat{{\mathbf {B}}}}_{{\mathfrak {d}}}\,\mathrm {d}\varGamma , \end{aligned}$$
(B.35d)

where the interface operators are given by

$$\begin{aligned} {\mathbb {C}}_{dd}^{i}&=\left[ \begin{array}{cc} {\hat{\alpha }} k_{n} &{} 0 \\ 0 &{} {\hat{\beta }} k_{t} \\ \end{array} \right] , \end{aligned}$$
(B.36a)
$$\begin{aligned} {\mathbb {C}}_{{\mathbf {d}}{\mathfrak {d}}}^{i}&=\left[ g_n k_{n}\dfrac{\partial {\hat{\alpha }}}{\partial {\mathfrak {d}}},g_t k_{t}\dfrac{\partial {\hat{\beta }}}{\partial {\mathfrak {d}}}\right] , \end{aligned}$$
(B.36b)
$$\begin{aligned} {\mathbb {C}}_{{\mathfrak {d}}{\mathbf {d}}}^{i}&=\left[ \begin{array}{c} g_n k_{n}\dfrac{\partial {\hat{\alpha }}}{\partial {\mathfrak {d}}} \\ g_t k_{t}\dfrac{\partial {\hat{\beta }}}{\partial {\mathfrak {d}}}\\ \end{array} \right] , \end{aligned}$$
(B.36c)
$$\begin{aligned} {\mathbb {C}}_{{\mathfrak {d}}{\mathfrak {d}}}^{i}&=\dfrac{1}{2}g_{n}^2 k_{n}\dfrac{\partial ^2{\hat{\alpha }}}{\partial {\mathfrak {d}}^2}+\dfrac{1}{2}g_{t}^2k_{t}\dfrac{\partial ^2 {\hat{\beta }}}{\partial {\mathfrak {d}}^2}. \end{aligned}$$
(B.36d)

In the previous expressions \({\hat{\alpha }}\) y \({\hat{\beta }}\) are particularized as follows:

$$\begin{aligned} {\hat{\alpha }}&= \dfrac{g_{nc,0}^2}{\left[ (1-{\mathfrak {d}})g_{nc,0}+ {\mathfrak {d}} g_{nc,1}\right] ^2}, \end{aligned}$$
(B.37a)
$$\begin{aligned} {\hat{\beta }}&= \dfrac{g_{tc,0}^2}{\left[ (1-{\mathfrak {d}})g_{tc,0}+ {\mathfrak {d}} g_{tc,1}\right] ^2}. \end{aligned}$$
(B.37b)

In line with Eq. (A.25), the final derivation endows a coupled system of equations.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Guillén-Hernández, T., García, I.G., Reinoso, J. et al. A micromechanical analysis of inter-fiber failure in long reinforced composites based on the phase field approach of fracture combined with the cohesive zone model. Int J Fract 220, 181–203 (2019). https://doi.org/10.1007/s10704-019-00384-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10704-019-00384-8

Keywords

Navigation