HTML conversions sometimes display errors due to content that did not convert correctly from the source. This paper uses the following packages that are not yet supported by the HTML conversion tool. Feedback on these issues are not necessary; they are known and are being worked on.

  • failed: yhmath
  • failed: leftidx
  • failed: stackengine

Authors: achieve the best HTML results from your LaTeX submissions by following these best practices.

License: arXiv.org perpetual non-exclusive license
arXiv:2402.00368v1 [cond-mat.mtrl-sci] 01 Feb 2024

Emergence of tension-compression asymmetry from a complete phase-field approach to brittle fracture

Chang Liu [email protected] Aditya Kumar [email protected] School of Civil and Environmental Engineering, Georgia Institute of Technology, Atlanta, GA 30332, USA
Abstract

The classical variational approach to brittle fracture propagation does not distinguish between strain energy accumulation in tension versus compression and consequently results in physically unrealistic cracking under compression. A variety of energy splits have been proposed as a possible remedy. However, a unique energy split that can describe this asymmetry for general loading conditions has not been found. The main objective of this paper is to show that a complete phase-field theory of brittle fracture nucleation and propagation, one that accounts for the material strength at large, can naturally capture the tension-compression asymmetry without an energy split. One such theory has been recently proposed by Kumar et al. (2018). Over the past few years, several studies have shown that this theory is capable of accurately describing fracture nucleation and propagation for materials soft and hard under arbitrary monotonic loading conditions. However, a systematic study of the tension-compression asymmetry that emerges from this theory has not yet been reported. This paper does precisely that. In particular, this paper reports a comprehensive study of crack propagation in two problems, one involving a symmetric tension-compression state and the other involving larger compressive stresses at the crack tip. The results are compared with popular energy splits used in literature. The results show that, remarkably, for the second problem, only the complete theory is able to produce experimentally consistent results.

keywords:
Tension-compression asymmetry; Brittle materials; Phase-field regularization; Fracture nucleation; Strength

1 Introduction

The mathematical reformulation of Griffith’s Griffith (1921) original idea for fracture by Francfort and Marigo Francfort and Marigo (1998) has led to an elegant theory of fracture propagation called the variational theory of brittle fracture. In this theory, the calculation of the fracture state is viewed as a minimization problem for the sum of the bulk strain energy and surface fracture energies with arbitrary add-cracks as one of the test fields. The variational theory can predict both when a large pre-existing crack starts to propagate and how it propagates. Phase-field models Bourdin et al. (2000), developed as a regularization of the variational theory to make it more amenable to numerical studies, retained the elegancy of the sharp theory and have proved to be a successful approach to simulate fracture propagation. However, they also inherited a significant shortcoming of the variational theory – they do not distinguish between strain energy accumulation in tension and compression. Consequently, they can not describe the tension-compression asymmetry in fracture response, resulting in cracks under compression.

The popular remedial approach is to split or decompose the strain energy into a positive (opening) and negative (closing) part where only the positive energy drives the evolution of the scalar phase field variable. The energy split approach keeps the variational nature of the classical phase-field model. However, this approach suffers from several shortcomings. First, it is not apparent how a unique split can be constructed for a linear elastic isotropic material Amor et al. (2009); Miehe et al. (2010), much less for an anisotropic material van Dijk et al. (2020); Vu et al. (2022) or a finite elastic material Borden et al. (2016); Tang et al. (2019). Second, many of the popular methods of energy splits are prone to having non-zero residual stiffness in the cracked regions Vicentini et al. (2023). But perhaps most importantly, no single energy split has been yet shown in the literature to be able to comprehensively describe the tension-compression asymmetry for all loading conditions, static or dynamic Strobl and Seelig (2020); Zhang et al. (2022); Vicentini et al. (2023).

An alternative and more natural solution may arise from the generalized approach of Kumar et al. Kumar et al. (2018, 2020) to phase-field modeling of nucleation and propagation of brittle fracture. In a nutshell, this theory generalizes the classical phase-field theory for fracture propagation by accounting for nucleation in general through the material’s strength surface, while kee** undisturbed the ability of the classical phase-field regularization to model crack propagation according to Griffith’s fracture postulate. A string of recent works Kumar et al. (2018, 2020); Kumar and Lopez-Pamies (2020, 2021); Kumar et al. (2022, 2024) have provided a wide range of validation results for a broad spectrum of materials (silicone, titania, graphite, polyurethane, PMMA, alumina, natural rubber, glass, rock), specimen geometries (with large and small pre-existing cracks, V notches, U notches, and smooth boundaries), and loading conditions suggesting that this theory may indeed provide a complete framework for elastic brittle materials. Because the theory accounts for the strength of the material at large, it has been conjectured that it can account for tension-compression asymmetry naturally Kumar et al. (2020).

Some evidence in this direction has been presented before in previous works for crack nucleation in the absence of pre-existing cracks. The theory was able to provide qualitative and quantitative agreement with the experiments for the indentation test Kumar et al. (2022) and Brazilian fracture test Kumar et al. (2024) – problems that both involve large compressive stresses. However, it remains understudied whether the theory can describe the asymmetry under general loading conditions and whether it depends on any model parameters. It is also understudied whether it performs better than the energy split methods for problems involving large pre-existing cracks. The main purpose of this paper is to address these questions. We do so by studying two prominent examples of fracture propagation in which the correct modeling of asymmetry plays an important role.

In the first example, we focus on the mode II fracture of a plate under a quasi-static simple shear loading. This is a benchmark problem for tension-compression asymmetry studied in many previous works Bourdin et al. (2008); Miehe et al. (2010). It involves a symmetric tension-compression state in front of the crack. All energy split methods proposed in the literature can predict the correct crack path for this problem. In the second example, we study the propagation from two inclined cracks in a plate under uniaxial compression. The rationale for our choice to study this problem is threefold. First, this problem is of broad interest in rock mechanics and has been extensively studied. Second, it involves an asymmetric tension-compression state at the crack tip with significantly larger compressive stresses. Third, the analysis of this problem with the phase-field method utilizing the conventional energy splits has proved to be technically challenging Zhang et al. (2017). It has been claimed in the literature that the experimental crack path can only be obtained by assuming a mode-dependent fracture toughness for the material Zhang et al. (2017); Bryant and Sun (2018); Steinke et al. (2022) and more involved splits have been proposed as a solution. We begin in Section 2 by summarizing the general fracture theory of Kumar et al. (2018, 2022). We then present in Sections 3 and 4 a comprehensive study of the two examples. Lastly, some final comments are recorded in Section 5.

2 The complete phase-field theory of brittle fracture of Kumar et al. (2018, 2020)

Consider a structure made of an isotropic linear elastic brittle material occupying an open bounded domain Ω3Ωsuperscript3\mathrm{\Omega}\subset\mathbb{R}^{3}roman_Ω ⊂ blackboard_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, with boundary ΩΩ\partial\mathrm{\Omega}∂ roman_Ω in its undeformed and stress-free configuration at time t=0𝑡0t=0italic_t = 0. At a later time t(0,T]𝑡0𝑇t\in(0,T]italic_t ∈ ( 0 , italic_T ], due to an externally applied displacement 𝐮¯(𝐗,t)¯𝐮𝐗𝑡\overline{{\bf u}}({\bf X},t)over¯ start_ARG bold_u end_ARG ( bold_X , italic_t ) on a part Ω𝒟subscriptΩ𝒟\partial\mathrm{\Omega}_{\mathcal{D}}∂ roman_Ω start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT of the boundary and a traction 𝐭¯(𝐗,t)¯𝐭𝐗𝑡\overline{\textbf{t}}({\bf X},t)over¯ start_ARG t end_ARG ( bold_X , italic_t ) on the complementary part Ω𝒩=ΩΩ𝒟subscriptΩ𝒩ΩsubscriptΩ𝒟\partial\mathrm{\Omega}_{\mathcal{N}}=\partial\mathrm{\Omega}\setminus\partial% \mathrm{\Omega}_{\mathcal{D}}∂ roman_Ω start_POSTSUBSCRIPT caligraphic_N end_POSTSUBSCRIPT = ∂ roman_Ω ∖ ∂ roman_Ω start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT, the structure experiences a deformation field 𝐮(𝐗,t)𝐮𝐗𝑡{\bf u}({\bf X},t)bold_u ( bold_X , italic_t ). We write the infinitesimal strain tensor as

𝐄(𝐮)=12(𝐮+𝐮T).𝐄𝐮12𝐮superscript𝐮𝑇{\bf E}({\bf u})=\dfrac{1}{2}(\nabla{\bf u}+\nabla{\bf u}^{T}).bold_E ( bold_u ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( ∇ bold_u + ∇ bold_u start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ) .

Non-interpenetration constraint implies that det(𝐈+𝐮)>0det𝐈𝐮0{\rm det}({\bf I}+\nabla{\bf u})>0roman_det ( bold_I + ∇ bold_u ) > 0. In response to the externally applied mechanical stimuli, cracks can also nucleate and propagate in the structure. Those are described in a regularized way by the phase field

v=v(𝐗,t)𝑣𝑣𝐗𝑡v=v({\bf X},t)italic_v = italic_v ( bold_X , italic_t )

taking values in [0,1]01[0,1][ 0 , 1 ]. Precisely, v=1𝑣1v=1italic_v = 1 identifies regions of the sound material, whereas v<1𝑣1v<1italic_v < 1 identifies regions of the material that have been fractured.

2.1 Constitutive behavior of the material

For isotropic brittle materials, the mechanical behavior is assumed to be completely characterized by three intrinsic properties of the material: (i) elasticity, (ii) strength, and (iii) critical energy release rate.

Elasticity

The elastic behavior for an isotropic linear elastic material is characterized by the stored-energy function

W(𝐄(𝐮))=μ𝐄𝐄+λ2(tr𝐄)2,𝑊𝐄𝐮𝜇𝐄𝐄𝜆2superscripttr𝐄2W({\bf E}({\bf u}))=\mu{\bf E}\cdot{\bf E}+\dfrac{\lambda}{2}({\rm tr}\,{\bf E% })^{2},italic_W ( bold_E ( bold_u ) ) = italic_μ bold_E ⋅ bold_E + divide start_ARG italic_λ end_ARG start_ARG 2 end_ARG ( roman_tr bold_E ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (1)

where μ>0𝜇0\mu>0italic_μ > 0 and λ>2/3μ𝜆23𝜇\lambda>-2/3\muitalic_λ > - 2 / 3 italic_μ are the Lamé constants. Recall the basic relations μ=E/(2(1+ν))𝜇𝐸21𝜈\mu=E/(2(1+\nu))italic_μ = italic_E / ( 2 ( 1 + italic_ν ) ) and λ=Eν/((1+ν)(12ν))𝜆𝐸𝜈1𝜈12𝜈\lambda=E\nu/((1+\nu)(1-2\nu))italic_λ = italic_E italic_ν / ( ( 1 + italic_ν ) ( 1 - 2 italic_ν ) ), where E𝐸Eitalic_E is the Young’s modulus and ν𝜈\nuitalic_ν is the Poisson’s ratio. The stress-strain relation is given by

𝝈(𝐗,t)=E1+ν𝐄+Eν(1+ν)(12ν)(tr𝐄)𝐈𝝈𝐗𝑡𝐸1𝜈𝐄𝐸𝜈1𝜈12𝜈tr𝐄𝐈\bm{\sigma}({\bf X},t)=\dfrac{E}{1+\nu}{\bf E}+\dfrac{E\,\nu}{(1+\nu)(1-2\nu)}% ({\rm tr}\,{\bf E}){\bf I}bold_italic_σ ( bold_X , italic_t ) = divide start_ARG italic_E end_ARG start_ARG 1 + italic_ν end_ARG bold_E + divide start_ARG italic_E italic_ν end_ARG start_ARG ( 1 + italic_ν ) ( 1 - 2 italic_ν ) end_ARG ( roman_tr bold_E ) bold_I

Strength

When a macroscopic piece of an elastic brittle material is subjected to an arbitrary but uniform state of stress 𝝈𝝈\bm{\sigma}bold_italic_σ, fracture nucleates at a critical value of the applied stress. The set of all such stresses defines a surface in the stress space – called a strength surface

(𝝈)=0,𝝈0\mathcal{F}(\bm{\sigma})=0,caligraphic_F ( bold_italic_σ ) = 0 , (2)

where 𝝈𝝈\bm{\sigma}bold_italic_σ stands for the Cauchy stress tensor. A popular, but not general, choice for the strength surface of the material is the Drucker-Prager strength surface

(𝝈)=J2+γ1I1+γ0=0with{γ0=2σ𝚌𝚜σ𝚝𝚜3(σ𝚌𝚜+σ𝚝𝚜)γ1=σ𝚌𝚜σ𝚝𝚜3(σ𝚌𝚜+σ𝚝𝚜),formulae-sequence𝝈subscript𝐽2subscript𝛾1subscript𝐼1subscript𝛾00withcasessubscript𝛾02subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜3subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜subscript𝛾1subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜3subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜\mathcal{F}(\bm{\sigma})=\sqrt{J_{2}}+\gamma_{1}I_{1}+\gamma_{0}=0\qquad{\rm with% }\qquad\left\{\begin{array}[]{l}\gamma_{0}=-\dfrac{2\sigma_{\texttt{cs}}\sigma% _{\texttt{ts}}}{\sqrt{3}\left(\sigma_{\texttt{cs}}+\sigma_{\texttt{ts}}\right)% }\\ \gamma_{1}=\dfrac{\sigma_{\texttt{cs}}-\sigma_{\texttt{ts}}}{\sqrt{3}\left(% \sigma_{\texttt{cs}}+\sigma_{\texttt{ts}}\right)}\end{array}\right.,caligraphic_F ( bold_italic_σ ) = square-root start_ARG italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG + italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0 roman_with { start_ARRAY start_ROW start_CELL italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = - divide start_ARG 2 italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT end_ARG start_ARG square-root start_ARG 3 end_ARG ( italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT + italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT ) end_ARG end_CELL end_ROW start_ROW start_CELL italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = divide start_ARG italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT - italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT end_ARG start_ARG square-root start_ARG 3 end_ARG ( italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT + italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT ) end_ARG end_CELL end_ROW end_ARRAY , (3)

where

I1=tr𝝈andJ2=12tr𝝈D2with𝝈D=𝝈13(tr𝝈)𝐈formulae-sequencesubscript𝐼1tr𝝈andformulae-sequencesubscript𝐽212trsubscriptsuperscript𝝈2𝐷withsubscript𝝈𝐷𝝈13tr𝝈𝐈I_{1}={\rm tr}\,\bm{\sigma}\qquad{\rm and}\qquad J_{2}=\dfrac{1}{2}{\rm tr}\,% \bm{\sigma}^{2}_{D}\qquad{\rm with}\quad\bm{\sigma}_{D}=\bm{\sigma}-\dfrac{1}{% 3}({\rm tr}\,\bm{\sigma}){\bf I}italic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = roman_tr bold_italic_σ roman_and italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_tr bold_italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT roman_with bold_italic_σ start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT = bold_italic_σ - divide start_ARG 1 end_ARG start_ARG 3 end_ARG ( roman_tr bold_italic_σ ) bold_I (4)

stand for two of the standard invariants of the stress tensor 𝝈𝝈\bm{\sigma}bold_italic_σ, while the constants σ𝚝𝚜>0subscript𝜎𝚝𝚜0\sigma_{\texttt{ts}}>0italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT > 0 and σ𝚌𝚜>0subscript𝜎𝚌𝚜0\sigma_{\texttt{cs}}>0italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT > 0 denote the uniaxial tensile and compressive strengths of the material. This two-material-parameter Drucker-Prager strength surface (1952) is arguably the simplest model that has proven capable of describing reasonably well the strength of many nominally brittle materials and has been used extensively in previous phase-field studies Kumar et al. (2020); Kumar and Lopez-Pamies (2020); Kumar et al. (2022, 2024). We emphasize that the strength surface only defines the crack nucleation under uniform stress states. More precisely, it defines when the phase field v𝑣vitalic_v ceases to be 1 under uniform stress states. Under non-uniform stress states, the violation of the strength surface is not a sufficient condition for crack nucleation and requires the contribution of the third intrinsic property: critical energy release rate.

Critical energy release rate

The critical energy release rate (or intrinsic fracture toughness), Gcsubscript𝐺𝑐G_{c}italic_G start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT describes the total energy expended in creating unit fracture surface area. It captures the resistance to the growth of a pre-existing crack. For an isotropic brittle material, Gcsubscript𝐺𝑐G_{c}italic_G start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is a scalar material constant.

2.2 The governing equations of deformation and fracture

According to the theory of Kumar et al. (2018), the displacement field 𝐮k(𝐗)=𝐮(𝐗,tk)subscript𝐮𝑘𝐗𝐮𝐗subscript𝑡𝑘{\bf u}_{k}({\bf X})={\bf u}({\bf X},t_{k})bold_u start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( bold_X ) = bold_u ( bold_X , italic_t start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) and phase field vk(𝐗)=v(𝐗,tk)subscript𝑣𝑘𝐗𝑣𝐗subscript𝑡𝑘v_{k}({\bf X})=v({\bf X},t_{k})italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( bold_X ) = italic_v ( bold_X , italic_t start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) at any material point 𝐗Ω¯𝐗¯Ω{\bf X}\in\overline{\mathrm{\Omega}}bold_X ∈ over¯ start_ARG roman_Ω end_ARG and discrete time tk{0=t0,t1,,tm,t_{k}\in\{0=t_{0},t_{1},...,t_{m},italic_t start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∈ { 0 = italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_t start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT , tm+1,,subscript𝑡𝑚1t_{m+1},...,italic_t start_POSTSUBSCRIPT italic_m + 1 end_POSTSUBSCRIPT , … , tM=T}t_{M}=T\}italic_t start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT = italic_T } are determined by the system of coupled partial differential equations (PDEs)

{Div[vk2W𝐄(𝐄(𝐮k))]=𝟎,𝐗Ω,𝐮k=𝐮¯(𝐗,tk),𝐗Ω𝒟,[vk2W𝐄(𝐄(𝐮k))]𝐍=𝐭¯(𝐗,tk),𝐗Ω𝒩casesformulae-sequenceDivdelimited-[]superscriptsubscript𝑣𝑘2𝑊𝐄𝐄subscript𝐮𝑘0𝐗Ωmissing-subexpressionformulae-sequencesubscript𝐮𝑘¯𝐮𝐗subscript𝑡𝑘𝐗subscriptΩ𝒟missing-subexpressionformulae-sequencedelimited-[]superscriptsubscript𝑣𝑘2𝑊𝐄𝐄subscript𝐮𝑘𝐍¯𝐭𝐗subscript𝑡𝑘𝐗subscriptΩ𝒩missing-subexpression\left\{\begin{array}[]{ll}{\rm Div}\left[v_{k}^{2}\dfrac{\partial W}{\partial{% \bf E}}({\bf E}({\bf u}_{k}))\right]={\bf 0},\quad{\bf X}\in\mathrm{\Omega},\\% [10.0pt] {\bf u}_{k}=\overline{{\bf u}}({\bf X},t_{k}),\quad{\bf X}\in\partial\mathrm{% \Omega}_{\mathcal{D}},\\[10.0pt] \left[v_{k}^{2}\dfrac{\partial W}{\partial{\bf E}}({\bf E}({\bf u}_{k}))\right% ]{\bf N}=\overline{\textbf{t}}({\bf X},t_{k}),\quad{\bf X}\in\partial\mathrm{% \Omega}_{\mathcal{N}}\end{array}\right.{ start_ARRAY start_ROW start_CELL roman_Div [ italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG ∂ italic_W end_ARG start_ARG ∂ bold_E end_ARG ( bold_E ( bold_u start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) ) ] = bold_0 , bold_X ∈ roman_Ω , end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL bold_u start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = over¯ start_ARG bold_u end_ARG ( bold_X , italic_t start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) , bold_X ∈ ∂ roman_Ω start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT , end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL [ italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG ∂ italic_W end_ARG start_ARG ∂ bold_E end_ARG ( bold_E ( bold_u start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) ) ] bold_N = over¯ start_ARG t end_ARG ( bold_X , italic_t start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) , bold_X ∈ ∂ roman_Ω start_POSTSUBSCRIPT caligraphic_N end_POSTSUBSCRIPT end_CELL start_CELL end_CELL end_ROW end_ARRAY (5)

and

{εGcvk=83vkW(𝐄(𝐮k))43c𝚎(𝐗,tk)Gc2ε,if vk(𝐗)<vk1(𝐗),𝐗ΩεGcvk83vkW(𝐄(𝐮k))43c𝚎(𝐗,tk)Gc2ε,if vk(𝐗)=1 or vk(𝐗)=vk1(𝐗)>0,𝐗Ωvk(𝐗)=0, if vk1(𝐗)=0,𝐗Ωvk𝐍=0,𝐗Ωcasesformulae-sequence𝜀subscript𝐺𝑐subscript𝑣𝑘83subscript𝑣𝑘𝑊𝐄subscript𝐮𝑘43subscript𝑐𝚎𝐗subscript𝑡𝑘subscript𝐺𝑐2𝜀formulae-sequenceif subscript𝑣𝑘𝐗subscript𝑣𝑘1𝐗𝐗Ωformulae-sequenceformulae-sequence𝜀subscript𝐺𝑐subscript𝑣𝑘83subscript𝑣𝑘𝑊𝐄subscript𝐮𝑘43subscript𝑐𝚎𝐗subscript𝑡𝑘subscript𝐺𝑐2𝜀if subscript𝑣𝑘𝐗1 or subscript𝑣𝑘𝐗subscript𝑣𝑘1𝐗0𝐗Ωformulae-sequencesubscript𝑣𝑘𝐗0formulae-sequence if subscript𝑣𝑘1𝐗0𝐗Ωformulae-sequencesubscript𝑣𝑘𝐍0𝐗Ω\left\{\begin{array}[]{l}\varepsilon\,G_{c}\triangle v_{k}=\dfrac{8}{3}v_{k}W(% {\bf E}({\bf u}_{k}))-\dfrac{4}{3}c_{\texttt{e}}({\bf X},t_{k})-\dfrac{G_{c}}{% 2\varepsilon},\mbox{if }v_{k}({\bf X})<v_{k-1}({\bf X}),\quad{\bf X}\in\mathrm% {\Omega}\\[10.0pt] \varepsilon\,G_{c}\triangle v_{k}\geq\dfrac{8}{3}v_{k}W({\bf E}({\bf u}_{k}))-% \dfrac{4}{3}c_{\texttt{e}}({\bf X},t_{k})-\dfrac{G_{c}}{2\varepsilon},\mbox{if% }v_{k}({\bf X})=1\;\mbox{ or }\;v_{k}({\bf X})=v_{k-1}({\bf X})>0,\quad{\bf X% }\in\mathrm{\Omega}\\[10.0pt] v_{k}({\bf X})=0,\quad\mbox{ if }v_{k-1}({\bf X})=0,\quad{\bf X}\in\mathrm{% \Omega}\\[10.0pt] \nabla v_{k}\cdot{\bf N}=0,\quad{\bf X}\in\partial\mathrm{\Omega}\end{array}\right.{ start_ARRAY start_ROW start_CELL italic_ε italic_G start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT △ italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = divide start_ARG 8 end_ARG start_ARG 3 end_ARG italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_W ( bold_E ( bold_u start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) ) - divide start_ARG 4 end_ARG start_ARG 3 end_ARG italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT ( bold_X , italic_t start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) - divide start_ARG italic_G start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ε end_ARG , if italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( bold_X ) < italic_v start_POSTSUBSCRIPT italic_k - 1 end_POSTSUBSCRIPT ( bold_X ) , bold_X ∈ roman_Ω end_CELL end_ROW start_ROW start_CELL italic_ε italic_G start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT △ italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ≥ divide start_ARG 8 end_ARG start_ARG 3 end_ARG italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_W ( bold_E ( bold_u start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) ) - divide start_ARG 4 end_ARG start_ARG 3 end_ARG italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT ( bold_X , italic_t start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) - divide start_ARG italic_G start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ε end_ARG , if italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( bold_X ) = 1 or italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( bold_X ) = italic_v start_POSTSUBSCRIPT italic_k - 1 end_POSTSUBSCRIPT ( bold_X ) > 0 , bold_X ∈ roman_Ω end_CELL end_ROW start_ROW start_CELL italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( bold_X ) = 0 , if italic_v start_POSTSUBSCRIPT italic_k - 1 end_POSTSUBSCRIPT ( bold_X ) = 0 , bold_X ∈ roman_Ω end_CELL end_ROW start_ROW start_CELL ∇ italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ⋅ bold_N = 0 , bold_X ∈ ∂ roman_Ω end_CELL end_ROW end_ARRAY (6)

with 𝐮(𝐗,0)𝟎𝐮𝐗0𝟎{\bf u}({\bf X},0)\equiv\textbf{0}bold_u ( bold_X , 0 ) ≡ 0 and v(𝐗,0)1𝑣𝐗01v({\bf X},0)\equiv 1italic_v ( bold_X , 0 ) ≡ 1, where 𝐮k(𝐗)=𝐮(𝐗,tk)subscript𝐮𝑘𝐗𝐮𝐗subscript𝑡𝑘\nabla{\bf u}_{k}({\bf X})=\nabla{\bf u}({\bf X},t_{k})∇ bold_u start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( bold_X ) = ∇ bold_u ( bold_X , italic_t start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ), vk(𝐗)=v(𝐗,tk)subscript𝑣𝑘𝐗𝑣𝐗subscript𝑡𝑘\nabla v_{k}({\bf X})=\nabla v({\bf X},t_{k})∇ italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( bold_X ) = ∇ italic_v ( bold_X , italic_t start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ), vk(𝐗)=v(𝐗,\triangle v_{k}({\bf X})=\triangle v({\bf X},△ italic_v start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( bold_X ) = △ italic_v ( bold_X , tk)t_{k})italic_t start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ), and where ε>0𝜀0\varepsilon>0italic_ε > 0 is a regularization or localization length and c𝚎(𝐗,t)subscript𝑐𝚎𝐗𝑡c_{\texttt{e}}({\bf X},t)italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT ( bold_X , italic_t ) is a driving force containing information about material’s strength. The specific constitutive prescription for c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT depends on the particular form of the strength surface and is spelled out below for the case of Drucker-Prager strength surfaces.

Remark 1.

The inequalities in (6) describe the constraint that phase field is constrained between 0 and 1 and the classical assumption that fracture is an irreversible process. Moreover, the localization length ε𝜀\varepsilonitalic_ε in (6) is purely a regularization parameter that is set to be smaller than the smallest characteristic length scale in the structural problem at hand. In practice, it is also set to be no larger than the material’s characteristic fracture length scale.

Remark 2.

The computational implementation of the equations (5)–(6) differs from the implementation of the classical variational model only through the presence of the term c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT on the right-hand side of the PDE for the phase-field (6). Accordingly, the standard finite-element staggered scheme can be utilized to solve these equations. In absence of the c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT term in (6), the equations (5)-(6) represent the classical AT11{}_{1}start_FLOATSUBSCRIPT 1 end_FLOATSUBSCRIPT phase-field model Pham et al. (2011); Tanné et al. (2018).

Remark 3.

The solution of the equations (5)–(6) may be suspect to material interpenetration under compression, unless geometrical nonlinearity through finite kinematics is utilized Kumar et al. (2018). In this work, we assume geometrical linearity and only check a posteriori that the non-interpenetration constraint is satisfied.

2.3 The model for the external driving force c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT

The external driving force is a manifestation of the presence of the inherent defects in the material, that is, its strength surface at large. Hence, the information about the strength surface enters the governing PDEs through c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT. A blueprint for constructing c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT was outlined in Kumar et al. (2020). The construction process specifies that c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT takes the same functional form as the strength surface, in this case, the Drucker-Prager strength surface (3), but with ε𝜀\varepsilonitalic_ε-dependent coefficients. Precisely, it is defined as

c𝚎(𝐗,t)=c^𝚎(I1,J2;ε)=β2εJ2+β1εI1+β0ε,subscript𝑐𝚎𝐗𝑡subscript^𝑐𝚎subscript𝐼1subscript𝐽2𝜀superscriptsubscript𝛽2𝜀subscript𝐽2superscriptsubscript𝛽1𝜀subscript𝐼1superscriptsubscript𝛽0𝜀\displaystyle c_{\texttt{e}}({\bf X},t)=\widehat{c}_{\texttt{e}}(I_{1},J_{2};% \varepsilon)=\beta_{2}^{\varepsilon}\sqrt{J_{2}}+\beta_{1}^{\varepsilon}I_{1}+% \beta_{0}^{\varepsilon},italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT ( bold_X , italic_t ) = over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT e end_POSTSUBSCRIPT ( italic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ; italic_ε ) = italic_β start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT square-root start_ARG italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG + italic_β start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT italic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT , (7)

where

{β0ε=δε3Gc8εβ1ε=((1+δε)(σ𝚌𝚜σ𝚝𝚜)2σ𝚌𝚜σ𝚝𝚜)3Gc8ε+σ𝚌𝚜σ𝚝𝚜2Eβ2ε=(3(1+δε)(σ𝚌𝚜+σ𝚝𝚜)2σ𝚌𝚜σ𝚝𝚜)3Gc8ε+3(σ𝚌𝚜+σ𝚝𝚜)2E,casessuperscriptsubscript𝛽0𝜀superscript𝛿𝜀3subscript𝐺𝑐8𝜀subscriptsuperscript𝛽𝜀11superscript𝛿𝜀subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜2subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜3subscript𝐺𝑐8𝜀subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜2𝐸subscriptsuperscript𝛽𝜀231superscript𝛿𝜀subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜2subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜3subscript𝐺𝑐8𝜀3subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜2𝐸\left\{\begin{array}[]{l}\beta_{0}^{\varepsilon}=\delta^{\varepsilon}\dfrac{3G% _{c}}{8\varepsilon}\\[12.0pt] \beta^{\varepsilon}_{1}=-\left(\dfrac{(1+\delta^{\varepsilon})(\sigma_{\texttt% {cs}}-\sigma_{\texttt{ts}})}{2\sigma_{\texttt{cs}}\sigma_{\texttt{ts}}}\right)% \dfrac{3G_{c}}{8\varepsilon}+\dfrac{\sigma_{\texttt{cs}}-\sigma_{\texttt{ts}}}% {2E}\\[12.0pt] \beta^{\varepsilon}_{2}=-\left(\dfrac{\sqrt{3}(1+\delta^{\varepsilon})(\sigma_% {\texttt{cs}}+\sigma_{\texttt{ts}})}{2\sigma_{\texttt{cs}}\sigma_{\texttt{ts}}% }\right)\dfrac{3G_{c}}{8\varepsilon}+\dfrac{\sqrt{3}(\sigma_{\texttt{cs}}+% \sigma_{\texttt{ts}})}{2E}\end{array}\right.,{ start_ARRAY start_ROW start_CELL italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT = italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT divide start_ARG 3 italic_G start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG start_ARG 8 italic_ε end_ARG end_CELL end_ROW start_ROW start_CELL italic_β start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = - ( divide start_ARG ( 1 + italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT ) ( italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT - italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT ) end_ARG start_ARG 2 italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT end_ARG ) divide start_ARG 3 italic_G start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG start_ARG 8 italic_ε end_ARG + divide start_ARG italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT - italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_E end_ARG end_CELL end_ROW start_ROW start_CELL italic_β start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = - ( divide start_ARG square-root start_ARG 3 end_ARG ( 1 + italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT ) ( italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT + italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT ) end_ARG start_ARG 2 italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT end_ARG ) divide start_ARG 3 italic_G start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG start_ARG 8 italic_ε end_ARG + divide start_ARG square-root start_ARG 3 end_ARG ( italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT + italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT ) end_ARG start_ARG 2 italic_E end_ARG end_CELL end_ROW end_ARRAY , (8)

I1subscript𝐼1I_{1}italic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and J2subscript𝐽2J_{2}italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT stand for the invariants (4) of the Cauchy stress

𝝈(𝐗,t)=v2W𝐄(𝐄(𝐮))𝝈𝐗𝑡superscript𝑣2𝑊𝐄𝐄𝐮\bm{\sigma}({\bf X},t)=v^{2}\dfrac{\partial W}{\partial{\bf E}}({\bf E}({\bf u% }))bold_italic_σ ( bold_X , italic_t ) = italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG ∂ italic_W end_ARG start_ARG ∂ bold_E end_ARG ( bold_E ( bold_u ) )

and, hence, read as

I1=(3λ+2μ)v2tr𝐄(𝐮)andJ2=2μ2v4tr𝐄D2(𝐮)formulae-sequencesubscript𝐼13𝜆2𝜇superscript𝑣2tr𝐄𝐮andsubscript𝐽22superscript𝜇2superscript𝑣4trsubscriptsuperscript𝐄2𝐷𝐮I_{1}=(3\lambda+2\mu)v^{2}{\rm tr}\,{\bf E}({\bf u})\quad{\rm and}\quad J_{2}=% 2\mu^{2}v^{4}{\rm tr}\,{\bf E}^{2}_{D}({\bf u})italic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = ( 3 italic_λ + 2 italic_μ ) italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_tr bold_E ( bold_u ) roman_and italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 2 italic_μ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_tr bold_E start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT ( bold_u )

with 𝐄D(𝐮)=𝐄(𝐮)1/3(tr𝐄(𝐮))𝐈subscript𝐄𝐷𝐮𝐄𝐮13tr𝐄𝐮𝐈{\bf E}_{D}({\bf u})={\bf E}({\bf u})-1/3\left({\rm tr}\,{\bf E}({\bf u})% \right){\bf I}bold_E start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT ( bold_u ) = bold_E ( bold_u ) - 1 / 3 ( roman_tr bold_E ( bold_u ) ) bold_I in terms of the displacement field 𝐮𝐮{\bf u}bold_u and phase field v𝑣vitalic_v, and where δεsuperscript𝛿𝜀\delta^{\varepsilon}italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT is a unitless ε𝜀\varepsilonitalic_ε-dependent coefficient. The value of δεsuperscript𝛿𝜀\delta^{\varepsilon}italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT needs to be calibrated to retain Griffith-fracture propagation with the equations (5)–(6). Currently, the simplest procedure to calibrate δεsuperscript𝛿𝜀\delta^{\varepsilon}italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT is to solve, for a given set of material constants E𝐸Eitalic_E, ν𝜈\nuitalic_ν, Gcsubscript𝐺𝑐G_{c}italic_G start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, σ𝚝𝚜subscript𝜎𝚝𝚜\sigma_{\texttt{ts}}italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT, σ𝚌𝚜subscript𝜎𝚌𝚜\sigma_{\texttt{cs}}italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT, and a given finite localization length ε𝜀\varepsilonitalic_ε, a single boundary-value problem of choice for which the nucleation from a large pre-existing crack can be determined analytically and then adjust δεsuperscript𝛿𝜀\delta^{\varepsilon}italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT so that the phase-field theory matches the analytical solution. An analytical formula for δεsuperscript𝛿𝜀\delta^{\varepsilon}italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT is under development.

Refer to caption
Figure 1: The strength surfaces in the (σ1subscript𝜎1\sigma_{1}italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, σ2subscript𝜎2\sigma_{2}italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT)-space for stress states with σ3=0subscript𝜎30\sigma_{3}=0italic_σ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 0 for gypsum (see Section 4) and localization length ε=1𝜀1\varepsilon=1italic_ε = 1 mm, as predicted by the driving force introduced in Kumar et al. (2020), named Asymptotic Model, and the driving force introduced in Kumar et al. (2022), named Finite-ε𝜀\varepsilonitalic_ε Model. For direct comparison, the Drucker–Prager strength surface (6) is also included.

A correction to the form for c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT was presented in Kumar et al. Kumar et al. (2022) to improve the description of the compressive part of the strength surface from the governing PDEs (5)–(6) for large values of localization lengths ε𝜀\varepsilonitalic_ε. Specifically, this form of c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT is defined as

c𝚎(𝐗,t)=c^𝚎(I1,J2;ε)=β2εJ2+β1εI1+β0ε+1v3(1I12I1)(J2(1+ν)E+I12(12ν)6E),subscript𝑐𝚎𝐗𝑡subscript^𝑐𝚎subscript𝐼1subscript𝐽2𝜀superscriptsubscript𝛽2𝜀subscript𝐽2superscriptsubscript𝛽1𝜀subscript𝐼1superscriptsubscript𝛽0𝜀1superscript𝑣31superscriptsubscript𝐼12subscript𝐼1subscript𝐽21𝜈𝐸superscriptsubscript𝐼1212𝜈6𝐸\displaystyle c_{\texttt{e}}({\bf X},t)=\widehat{c}_{\texttt{e}}(I_{1},J_{2};% \varepsilon)=\beta_{2}^{\varepsilon}\sqrt{J_{2}}+\beta_{1}^{\varepsilon}I_{1}+% \beta_{0}^{\varepsilon}+\dfrac{1}{v^{3}}\left(1-\dfrac{\sqrt{I_{1}^{2}}}{I_{1}% }\right)\left(\dfrac{J_{2}(1+\nu)}{E}+\dfrac{I_{1}^{2}(1-2\nu)}{6E}\right),italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT ( bold_X , italic_t ) = over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT e end_POSTSUBSCRIPT ( italic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ; italic_ε ) = italic_β start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT square-root start_ARG italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG + italic_β start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT italic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT + divide start_ARG 1 end_ARG start_ARG italic_v start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ( 1 - divide start_ARG square-root start_ARG italic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG start_ARG italic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) ( divide start_ARG italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( 1 + italic_ν ) end_ARG start_ARG italic_E end_ARG + divide start_ARG italic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 - 2 italic_ν ) end_ARG start_ARG 6 italic_E end_ARG ) , (9)

where

{β0ε=δε3Gc8εβ1ε=((1+δε)(σ𝚌𝚜σ𝚝𝚜)2σ𝚌𝚜σ𝚝𝚜)3Gc8ε+σ𝚝𝚜2Eβ2ε=(3(1+δε)(σ𝚌𝚜+σ𝚝𝚜)2σ𝚌𝚜σ𝚝𝚜)3Gc8ε+3σ𝚝𝚜2E,casessuperscriptsubscript𝛽0𝜀superscript𝛿𝜀3subscript𝐺𝑐8𝜀subscriptsuperscript𝛽𝜀11superscript𝛿𝜀subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜2subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜3subscript𝐺𝑐8𝜀subscript𝜎𝚝𝚜2𝐸subscriptsuperscript𝛽𝜀231superscript𝛿𝜀subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜2subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜3subscript𝐺𝑐8𝜀3subscript𝜎𝚝𝚜2𝐸\left\{\begin{array}[]{l}\beta_{0}^{\varepsilon}=\delta^{\varepsilon}\dfrac{3G% _{c}}{8\varepsilon}\\[12.0pt] \beta^{\varepsilon}_{1}=-\left(\dfrac{(1+\delta^{\varepsilon})(\sigma_{\texttt% {cs}}-\sigma_{\texttt{ts}})}{2\sigma_{\texttt{cs}}\sigma_{\texttt{ts}}}\right)% \dfrac{3G_{c}}{8\varepsilon}+\dfrac{\sigma_{\texttt{ts}}}{2E}\\[12.0pt] \beta^{\varepsilon}_{2}=-\left(\dfrac{\sqrt{3}(1+\delta^{\varepsilon})(\sigma_% {\texttt{cs}}+\sigma_{\texttt{ts}})}{2\sigma_{\texttt{cs}}\sigma_{\texttt{ts}}% }\right)\dfrac{3G_{c}}{8\varepsilon}+\dfrac{\sqrt{3}\sigma_{\texttt{ts}}}{2E}% \end{array}\right.,{ start_ARRAY start_ROW start_CELL italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT = italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT divide start_ARG 3 italic_G start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG start_ARG 8 italic_ε end_ARG end_CELL end_ROW start_ROW start_CELL italic_β start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = - ( divide start_ARG ( 1 + italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT ) ( italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT - italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT ) end_ARG start_ARG 2 italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT end_ARG ) divide start_ARG 3 italic_G start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG start_ARG 8 italic_ε end_ARG + divide start_ARG italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_E end_ARG end_CELL end_ROW start_ROW start_CELL italic_β start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = - ( divide start_ARG square-root start_ARG 3 end_ARG ( 1 + italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT ) ( italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT + italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT ) end_ARG start_ARG 2 italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT end_ARG ) divide start_ARG 3 italic_G start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG start_ARG 8 italic_ε end_ARG + divide start_ARG square-root start_ARG 3 end_ARG italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_E end_ARG end_CELL end_ROW end_ARRAY , (10)

The last term is the correction term that is zero when I1>0subscript𝐼10I_{1}>0italic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT > 0 and non-zero when I1<0subscript𝐼10I_{1}<0italic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < 0. In the limit of ε0𝜀0\varepsilon\searrow 0italic_ε ↘ 0, the correction term is negligible as it is O(ε0)𝑂superscript𝜀0O(\varepsilon^{0})italic_O ( italic_ε start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ) and the original model (7) is asymptotically obtained. We refer to the form (7) for c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT as the ‘Asymptotic Model’ and the form (9) as the ‘Finite-ε𝜀\varepsilonitalic_ε Model’ in the text below. A direct comparison between the strength surface implied from the two models is shown in Figure 1. The results are shown for gypsum studied in Section 4 with ε=1𝜀1\varepsilon=1italic_ε = 1 mm. The figure also includes the Drucker-Prager strength surface 3 for comparison. It can be clearly observed that the Finite-ε𝜀\varepsilonitalic_ε Model provides a better approximation of Drucker-Prager strength surface than the Asymptotic Model for this value of ε𝜀\varepsilonitalic_ε.

3 The simple shear problem

In this section, we deploy the phase-field theory (5)–(6) to do a numerical study of a rectangular specimen with a pre-existing crack, subjected to simple shear loading. This is considered a benchmark problem to investigate tension-compression asymmetry of phase-field models Bourdin et al. (2008); Miehe et al. (2010). Fig 2 shows a schematic of the initial geometry of the specimen and applied boundary conditions, as well as the expected crack path normal to the direction of tensile principal stress. The classical phase-field model without an energy split Bourdin et al. (2000, 2008) predicts an invalid branching of the pre-existing crack into a tension crack and a compression crack. However, models with energy splits are all able to capture the correct crack path.

Refer to caption
Figure 2: Schematics of the initial specimen geometry and boundary conditions for a pre-notched sample under simple shear. The expected crack path is shown in red.

Following Miehe et al. (2010), we set Young’s modulus E𝐸Eitalic_E and Poisson’s ratio ν𝜈\nuitalic_ν to be 210210210210 GPa and 0.30.30.30.3 respectively, while the critical energy release rate Gcsubscript𝐺𝑐G_{c}italic_G start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is set to 2700270027002700 N/m. The tensile strength σ𝚝𝚜subscript𝜎𝚝𝚜\sigma_{\texttt{ts}}italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT is fixed to be 3.263.263.263.26 GPa and the compressive strength σ𝚌𝚜subscript𝜎𝚌𝚜\sigma_{\texttt{cs}}italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT is varied. With these material parameters, the value of the classical material’s characteristic length for the AT11{}_{1}start_FLOATSUBSCRIPT 1 end_FLOATSUBSCRIPT model is (3EGc)/(8σ𝚝𝚜2)=0.023𝐸subscript𝐺𝑐8superscriptsubscript𝜎𝚝𝚜20.02(3E\,G_{c})/(8\sigma_{\texttt{ts}}^{2})=0.02( 3 italic_E italic_G start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) / ( 8 italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) = 0.02 mm Tanné et al. (2018). Table 1 provides the values of the regularization length ε𝜀\varepsilonitalic_ε, the finite element mesh size hhitalic_h, and the corresponding values of the parameter δεsuperscript𝛿𝜀\delta^{\varepsilon}italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT for all the results presented below with the Asymptotic Model (7) and Finite-ε𝜀\varepsilonitalic_ε Model (9) for the driving force c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT.

Table 1: Values of the regularization length ε𝜀\varepsilonitalic_ε, FE mesh size hhitalic_h, and the parameter δεsuperscript𝛿𝜀\delta^{\varepsilon}italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT utilized in the simulations of simple shear problem.
Model for c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT σ𝚌𝚜/σ𝚝𝚜subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT ε𝜀\varepsilonitalic_ε(mm) Mesh size hhitalic_h(mm) δεsuperscript𝛿𝜀\delta^{\varepsilon}italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT
Asymptotic Model (7) 3 0.015 0.0020 1.1
3 0.01 0.0013 2.15
1.5 0.015 0.0020 0.4
1.5 0.01 0.0013 1.15
1.5 0.0075 0.00075 2.6
Finite-ε𝜀\varepsilonitalic_ε Model (9) 1.5 0.015 0.002 -0.03
1.5 0.0075 0.00075 2.1

Brittle materials typically have a ratio of compressive strength to tensile strength greater than 3. Hence, with the Asymptotic Model, we first investigate the case when σ𝚌𝚜/σ𝚝𝚜=3subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜3\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}=3italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT = 3. Figure 3 presents snapshots of the phase field v𝑣vitalic_v predicted by the theory at different values of applied displacement for two values of localization lengths ε=0.015𝜀0.015\varepsilon=0.015italic_ε = 0.015 mm and ε=0.01𝜀0.01\varepsilon=0.01italic_ε = 0.01 mm. We observe that the theory predicts the physically expected crack path for both values of ε𝜀\varepsilonitalic_ε and the results are independent of ε𝜀\varepsilonitalic_ε. The compression crack is suppressed naturally in this case due to the tension-compression asymmetry built into the theory by accounting for the entire strength surface of the material. Similar results are obtained for all cases with σ𝚌𝚜/σ𝚝𝚜>3subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜3\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}>3italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT > 3.

Refer to caption
Figure 3: Contour plots of the phase field v𝑣vitalic_v predicted by the theory with the Asymptotic Model for driving force c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT for σ𝚌𝚜/σ𝚝𝚜=3subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜3\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}=3italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT = 3 and two values of localization length, ε=0.015,0.01𝜀0.0150.01\varepsilon=0.015,0.01italic_ε = 0.015 , 0.01 mm. Results are shown for four values of applied displacement: (a) u=0.007𝑢0.007u=0.007italic_u = 0.007 mm, (b) u=0.008𝑢0.008u=0.008italic_u = 0.008 mm, (c) u=0.010𝑢0.010u=0.010italic_u = 0.010 mm, and (d) u=0.015𝑢0.015u=0.015italic_u = 0.015 mm.
Refer to caption
Figure 4: Contour plots of the phase field v𝑣vitalic_v predicted by the Asymptotic Model for driving force c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT for σ𝚌𝚜/σ𝚝𝚜=3/2subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜32\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}=3/2italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT = 3 / 2 at displacements just smaller and just larger than the critical displacement, usubscript𝑢u_{\star}italic_u start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT at which the compression crack nucleates. Results are shown for three localization lengths: (a) ε=0.015𝜀0.015\varepsilon=0.015italic_ε = 0.015 mm, (b) ε=0.01𝜀0.01\varepsilon=0.01italic_ε = 0.01 mm, and (c) ε=0.0075𝜀0.0075\varepsilon=0.0075italic_ε = 0.0075 mm.

Next, we study a case where compressive strength is of similar value as tensile strength to investigate the limits of the Asymptotic Model. We set σ𝚌𝚜/σ𝚝𝚜=1.5subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜1.5\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}=1.5italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT = 1.5 and carry out simulations for three values of localization length ε=0.015,0.01𝜀0.0150.01\varepsilon=0.015,0.01italic_ε = 0.015 , 0.01 and 0.00750.00750.00750.0075 mm. We observe that while initially, only a tension crack nucleates and propagates like the previous case, a compression crack eventually nucleates from the pre-existing crack. Figure 4 shows the contour plots of phase field v𝑣vitalic_v for all three values of ε𝜀\varepsilonitalic_ε at two instants: shortly before and shortly after the compression crack nucleates. The compression crack nucleates for all three cases, however, it is observed visually that the tension crack propagates more before the compression crack nucleates as the value of ε𝜀\varepsilonitalic_ε is decreased. Further evidence for this observation is provided by a plot of reaction force P𝑃Pitalic_P as a function of applied displacement in Figure 5(a). It is plain from this plot that lowering the value of ϵitalic-ϵ\epsilonitalic_ϵ has a substantial impact on the critical applied displacement at which compression crack nucleates. The plot of the implied strength surface for the three values of ε𝜀\varepsilonitalic_ε also shows non-convergence. A prohibitively small value of localization length is likely required to get a converged and physically realistic solution.

Refer to caption
Figure 5: (a) Load-displacement curves and (b) strength surface with the Asymptotic Model for driving force for three values of localization lengths ε=0.015,0.01𝜀0.0150.01\varepsilon=0.015,0.01italic_ε = 0.015 , 0.01 and 0.00750.00750.00750.0075 mm.

Our previous results for indentation and Brazilian fracture problems have shown that the Finite-ε𝜀\varepsilonitalic_ε Model for driving force does not suffer from the same limitations. To verify, we carry out the simulations with that model for σ𝚌𝚜/σ𝚝𝚜=1.5subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜1.5\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}=1.5italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT = 1.5 and the lowest value of localization length, ε=0.0075𝜀0.0075\varepsilon=0.0075italic_ε = 0.0075 mm, adopted in the simulations with the Asymptotic Model. The contour plots for the phase-field v𝑣vitalic_v are plotted in Figure 6 (a). No compression crack is observed to nucleate. A simulation is also carried out for a higher value of localization length, ε=0.015𝜀0.015\varepsilon=0.015italic_ε = 0.015 mm, as shown in Figure 6 (b). The absence of a compression crack in that result indicates that the Finite-ε𝜀\varepsilonitalic_ε Model can even be used with larger values of ε𝜀\varepsilonitalic_ε. The results with larger values of σ𝚌𝚜/σ𝚝𝚜subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT with the Finite-ε𝜀\varepsilonitalic_ε Model are not included here, but the same conclusions can be drawn from them.

Refer to caption
Figure 6: Contour plots of the phase field v𝑣vitalic_v predicted by the theory with the Finite-ε𝜀\varepsilonitalic_ε Model for driving force c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT for σ𝚌𝚜/σ𝚝𝚜=3/2subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜32\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}=3/2italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT = 3 / 2 and two values of localization length: (a) ε=0.0075𝜀0.0075\varepsilon=0.0075italic_ε = 0.0075 mm and (b) ε=0.015𝜀0.015\varepsilon=0.015italic_ε = 0.015 mm. Load-displacement curves for both values of ε𝜀\varepsilonitalic_ε are shown on the right.

It is plain from these simulations that while the Asymptotic Model can capture tension-compression asymmetry naturally in problems involving symmetric tension and compression states for realistic values of σ𝚌𝚜/σ𝚝𝚜subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT for brittle materials, it fails to achieve that for comparable values of σ𝚌𝚜subscript𝜎𝚌𝚜\sigma_{\texttt{cs}}italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT and σ𝚝𝚜subscript𝜎𝚝𝚜\sigma_{\texttt{ts}}italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT with finite values of ε𝜀\varepsilonitalic_ε. The Finite-ε𝜀\varepsilonitalic_ε Model in contrast is more robust and can describe the asymmetry for any values of σ𝚌𝚜/σ𝚝𝚜subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT and ε𝜀\varepsilonitalic_ε. Recall that this model is identical to the Asymptotic Model in the limit ε0𝜀0\varepsilon\searrow 0italic_ε ↘ 0.

4 The wing cracks problem

Next, we confront the theory (5)–(6) and the two models for driving force c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT (7, 9) with a problem of crack nucleation from pre-existing cracks involving large global compressive forces. Fig 7(a) shows a schematic of the initial geometry of the specimen and applied boundary conditions. In this popular test for rock-type materials, a rectangular specimen with two (one or more) inclined cracks is subjected to uniaxial compression. Experimental observations Bobet and Einstein (1998); Sagong and Bobet (2002); Lee and Jeon (2011); Zhou et al. (2021) show that the so-called wing cracks (inner and outer) nucleate from existing cracks under this loading; see Fig 7(b). The wing crack nucleation has been extensively studied in the literature and several authors have theoretically and experimentally shown that when and how wing cracks nucleate is significantly affected by the friction between the crack faces Nemat-Nasser and Horii (1982); Horii and Nemat-Nasser (1985); Steif (1984); Bobet and Einstein (1998). In this work, for computational ease, we assume no friction and no normal contact between the crack faces, consistent with an open flaw assumption. The experimental observations have also indicated that secondary shear cracks may nucleate later resulting in coalescence of the two initial cracks.

Refer to caption
Figure 7: Schematics of the initial specimen geometry and boundary conditions for the wing cracks evolution problem. The expected inner and outer wing crack paths based on experimental observations are shown on the right.

Several authors have also investigated this problem with the classical phase-field method Zhang et al. (2017); Bryant and Sun (2018); Zhou et al. (2018); You et al. (2020); Steinke et al. (2022) and have noted the inability of the classical energy split techniques, in particular, the two most popular energy split techniques – volumetric-deviatoric and spectral, to accurately describe the crack paths observed in the experiments. For direct comparison, we first simulated this problem with the classical model and the typical energy splits. We adopt representative values of material parameters for gypsum Steinke et al. (2022). Young’s modulus E𝐸Eitalic_E, Poisson’s ratio ν𝜈\nuitalic_ν, and critical energy release rate Gcsubscript𝐺𝑐G_{c}italic_G start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT are set to 5.965.965.965.96 GPa, 0.150.150.150.15 and 10101010 N/m respectively. The localization length ε𝜀\varepsilonitalic_ε is chosen to be 1 mm. The separation between the cracks, w𝑤witalic_w, is taken as 6.35 mm.

The results are demonstrated in Figure 8(a)-(c). It is plain from these results these classical energy splits perform poorly for this problem. Volumetric-deviatoric split leads to compression cracks. The recently introduced star-convex model Vicentini et al. (2023) as a variation on the volumetric-deviatoric split would show similar compression cracks. The results with spectral split show outer wing cracks, but they appear as artificially wider and distorted presumably due to non-zero residual stresses in the cracked region. Inner wing cracks are not observed at all. These deficiencies of the classical splits under large compressive forces have been noticed before for the problem of indentation of glass plates with flat-ended indenters Strobl and Seelig (2020); Kumar et al. (2022). Alternative energy splits have been proposed to resolve these deficiencies. One such split is the modal split of Bryant and Sun (2018), in which the energy is first split based on the different kinematic modes – opening mode and shearing mode. The opening mode energy is taken to be non-zero only when a scalar mode I strain is greater than 0. This energy split, however, predicts a primary shear crack instead of wing cracks when the fracture toughness is mode-independent, see Fig. 8(d) where a result reproduced from Bryant and Sun (2018) is shown. The authors claimed that primary wing cracks can only be obtained when the mode II fracture toughness is much higher than the mode I fracture toughness.

Refer to caption
Figure 8: Contour plots of the phase-field v𝑣vitalic_v predicted by the (a) classical phase-field (no-split) model, (b) classical phase-field model with volumetric-deviatoric energy split, (c) classical phase-field model with spectral split, and (d) phase-field model with modal split reproduced from Bryant and Sun Bryant and Sun (2018).

To simulate this problem with the theory of Kumar et al. (5)–(6) and the two models for driving force c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT, we fix tensile strength σ𝚝𝚜subscript𝜎𝚝𝚜\sigma_{\texttt{ts}}italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT of gypsum to 3333 MPa and we choose the compressive strength as 20 MPa or 40 MPa. The fracture toughness is considered mode-independent. Table 2 provides the values of the regularization length ε𝜀\varepsilonitalic_ε, the finite element mesh size hhitalic_h, and the corresponding values of the parameter δεsuperscript𝛿𝜀\delta^{\varepsilon}italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT for the results with the theory of Kumar et al. To avoid the possibility of crack face interpenetration, we model the pre-existing cracks as phase-field cracks by setting phase-field v𝑣vitalic_v equal to 0 on the crack surfaces. Even so, modeling the cracks as open flaws with a small distance between the crack surfaces yields similar results.

Table 2: Values of the regularization length ε𝜀\varepsilonitalic_ε, FE mesh size hhitalic_h, and the parameter δεsuperscript𝛿𝜀\delta^{\varepsilon}italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT utilized in the simulations of wing crack problem.
Model for c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT σ𝚌𝚜/σ𝚝𝚜subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT ε𝜀\varepsilonitalic_ε(mm) Mesh size hhitalic_h(mm) δεsuperscript𝛿𝜀\delta^{\varepsilon}italic_δ start_POSTSUPERSCRIPT italic_ε end_POSTSUPERSCRIPT
Asymptotic Model (7) 40/3 1 0.1 7.5
Finite-ε𝜀\varepsilonitalic_ε Model (9) 20/3 1 0.1 3.4
40/3 1 0.1 3.75
Refer to caption
Figure 9: Contour plots of the phase field v𝑣vitalic_v predicted by the theory with the Asymptotic Model for driving force c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT for wing crack problem with w=6.35𝑤6.35w=6.35italic_w = 6.35 mm, σ𝚌𝚜/σ𝚝𝚜=40/3subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜403\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}=40/3italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT = 40 / 3 and ε=1𝜀1\varepsilon=1italic_ε = 1 mm.

Figure 9 presents results with the Asymptotic Model for c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT with w=6.35𝑤6.35w=6.35italic_w = 6.35 mm. Fig. 9(a) shows that both inner and outer wing cracks initially nucleate in the specimen. However, soon after, compressive cracks nucleate and propagate fast toward the lateral boundary as seen in Fig. 9(b)-(c). Unlike the simple shear example in Section 3, a larger value of compressive strength does not suppress the nucleation of compressive cracks, however it does delay it.

Refer to caption
Figure 10: Contour plots of the phase field v𝑣vitalic_v predicted by the theory with the Finite-ε𝜀\varepsilonitalic_ε Model for driving force c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT for wing crack problem w=6.35𝑤6.35w=6.35italic_w = 6.35 mm, ε=1𝜀1\varepsilon=1italic_ε = 1 mm and two values for compressive strength to tensile strength ratio: (a) σ𝚌𝚜/σ𝚝𝚜=20/3subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜203\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}=20/3italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT = 20 / 3 and (b) σ𝚌𝚜/σ𝚝𝚜=40/3subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜403\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}=40/3italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT = 40 / 3, at four applied displacements u𝑢uitalic_u.

Figure 10(a) presents results with the Finite-ε𝜀\varepsilonitalic_ε Model for c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT with σ𝚌𝚜/σ𝚝𝚜=20/3subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜203\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}=20/3italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT = 20 / 3. The contour plots of the phase field v𝑣vitalic_v are shown at four values of applied displacements u𝑢uitalic_u. Two comments are in order for these results. First, the Finite-ε𝜀\varepsilonitalic_ε Model for the driving force is able to capture the experimentally observed inner and outer wing crack paths. We observe that the inner and outer wing cracks nucleate at an applied displacement of u=0.12𝑢0.12u=0.12italic_u = 0.12 mm and stably propagate in the direction of applied loading. Second, the model can also capture the formation of secondary shear cracks and the coalescence of initial cracks. At around u=0.7𝑢0.7u=0.7italic_u = 0.7 mm, a secondary shear crack forms in between the inner wing cracks. However, the coalescence is a brutal crack event, and we experience numerical convergence issues beyond this displacement.

Fig. 10(b) shows the results obtained for σ𝚌𝚜/σ𝚝𝚜=40/3subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜403\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}=40/3italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT = 40 / 3. The results are qualitatively and quantitatively similar to the case with σ𝚌𝚜/σ𝚝𝚜=20/3subscript𝜎𝚌𝚜subscript𝜎𝚝𝚜203\sigma_{\texttt{cs}}/\sigma_{\texttt{ts}}=20/3italic_σ start_POSTSUBSCRIPT cs end_POSTSUBSCRIPT / italic_σ start_POSTSUBSCRIPT ts end_POSTSUBSCRIPT = 20 / 3, except that because of the higher compressive strength and consequently higher shear strength, the secondary shear crack supposedly nucleates at a larger displacement than 0.7 mm. For completeness, we also carried out simulations for three different separations between the two pre-existing cracks as done in experimental studies. In each case, we observe inner and outer wing cracks nucleating at roughly the same angle as the base case with w=6.35𝑤6.35w=6.35italic_w = 6.35 mm and propagating in the loading direction towards the center of the left and right boundaries.

Refer to caption
Figure 11: Contour plots of the phase field v𝑣vitalic_v for wing crack problem with three different values of separation between the two pre-existing cracks – (a) w=0𝑤0w=0italic_w = 0 mm, (b) w=12.7𝑤12.7w=12.7italic_w = 12.7 mm, and (c) w=25.4𝑤25.4w=25.4italic_w = 25.4 mm – predicted with the Finite-ε𝜀\varepsilonitalic_ε Model for driving force c𝚎subscript𝑐𝚎c_{\texttt{e}}italic_c start_POSTSUBSCRIPT e end_POSTSUBSCRIPT.

5 Final Comments

The two examples presented in this paper add to the validation results presented in previous works on the indentation problem Kumar et al. (2022) and the Brazilian fracture test Kumar et al. (2024) which demonstrate that the theory of Kumar et al. (2020) for linear elastic brittle fracture can describe the tension-compression asymmetry in fracture nucleation and propagation under a wide variety of loading and geometrical situations. This work in particular has highlighted the differences between the results from the two models for the driving force that describe the material’s strength surface. The first model proposed in Kumar et al. (2020) successfully suppresses compression cracks for realistic brittle materials when the problem involves a symmetric tension-compression state, like the simple shear problem in Section 3. However, for localization lengths ε𝜀\varepsilonitalic_ε that are not prohibitively small, it fails in problems involving large compressive forces. The second model proposed in Kumar et al. (2022) to improve the representation of strength surface for finite values of localization length under compressive states proves to be more robust and reliably provides physically consistent results for a wide range of values for ε𝜀\varepsilonitalic_ε and compressive strength to tensile strength ratio.

The results presented in Section 4 for the wing cracks problem have also made it plain that a complete model of tension-compression asymmetry with the phase-field models is needed to consistently and uniquely interpret the fracture toughness from experimental results involving large compression and mixed mode fracture. Future work will investigate further the effect of crack-parallel compressive stresses on fracture nucleation from large pre-existing cracks.

Acknowledgements

The author A. Kumar would like to acknowledge the financial support from the start-up fund provided by the Georgia Institute of Technology.

References

  • Griffith (1921) A. A. Griffith, Vi. the phenomena of rupture and flow in solids, Philosophical transactions of the royal society of london. Series A, containing papers of a mathematical or physical character 221 (1921) 163–198.
  • Francfort and Marigo (1998) G. A. Francfort, J.-J. Marigo, Revisiting brittle fracture as an energy minimization problem, Journal of the Mechanics and Physics of Solids 46 (1998) 1319–1342.
  • Bourdin et al. (2000) B. Bourdin, G. A. Francfort, J.-J. Marigo, Numerical experiments in revisited brittle fracture, Journal of the Mechanics and Physics of Solids 48 (2000) 797–826.
  • Amor et al. (2009) H. Amor, J.-J. Marigo, C. Maurini, Regularized formulation of the variational brittle fracture with unilateral contact: Numerical experiments, Journal of the Mechanics and Physics of Solids 57 (2009) 1209–1229.
  • Miehe et al. (2010) C. Miehe, F. Welschinger, M. Hofacker, Thermodynamically consistent phase-field models of fracture: Variational principles and multi-field fe implementations, International journal for numerical methods in engineering 83 (2010) 1273–1311.
  • van Dijk et al. (2020) N. P. van Dijk, J. J. Espadas-Escalante, P. Isaksson, Strain energy density decompositions in phase-field fracture theories for orthotropy and anisotropy, International Journal of Solids and Structures 196 (2020) 140–153.
  • Vu et al. (2022) B.-T. Vu, H. Le Quang, Q.-C. He, Modelling and simulation of fracture in anisotropic brittle materials by the phase-field method with novel strain decompositions, Mechanics Research Communications 124 (2022) 103936.
  • Borden et al. (2016) M. J. Borden, T. J. Hughes, C. M. Landis, A. Anvari, I. J. Lee, A phase-field formulation for fracture in ductile materials: Finite deformation balance law derivation, plastic degradation, and stress triaxiality effects, Computer Methods in Applied Mechanics and Engineering 312 (2016) 130–166.
  • Tang et al. (2019) S. Tang, G. Zhang, T. F. Guo, X. Guo, W. K. Liu, Phase field modeling of fracture in nonlinearly elastic solids via energy decomposition, Computer Methods in Applied Mechanics and Engineering 347 (2019) 477–494.
  • Vicentini et al. (2023) F. Vicentini, C. Zolesi, P. Carrara, C. Maurini, L. de Lorenzis, On the energy decomposition in variational phase-field models for brittle fracture under multi-axial stress states (2023).
  • Strobl and Seelig (2020) M. Strobl, T. Seelig, Phase field modeling of hertzian indentation fracture, Journal of the Mechanics and Physics of Solids 143 (2020) 104026.
  • Zhang et al. (2022) S. Zhang, W. Jiang, M. R. Tonks, Assessment of four strain energy decomposition methods for phase field fracture models using quasi-static and dynamic benchmark cases, Materials Theory 6 (2022) 6.
  • Kumar et al. (2018) A. Kumar, G. A. Francfort, O. Lopez-Pamies, Fracture and healing of elastomers: A phase-transition theory and numerical implementation, Journal of the Mechanics and Physics of Solids 112 (2018) 523–551.
  • Kumar et al. (2020) A. Kumar, B. Bourdin, G. A. Francfort, O. Lopez-Pamies, Revisiting nucleation in the phase-field approach to brittle fracture, Journal of the Mechanics and Physics of Solids 142 (2020) 104027.
  • Kumar et al. (2018) A. Kumar, K. Ravi-Chandar, O. Lopez-Pamies, The configurational-forces view of the nucleation and propagation of fracture and healing in elastomers as a phase transition, International Journal of Fracture 213 (2018) 1–16.
  • Kumar and Lopez-Pamies (2020) A. Kumar, O. Lopez-Pamies, The phase-field approach to self-healable fracture of elastomers: A model accounting for fracture nucleation at large, with application to a class of conspicuous experiments, Theoretical and Applied Fracture Mechanics 107 (2020) 102550.
  • Kumar and Lopez-Pamies (2021) A. Kumar, O. Lopez-Pamies, The poker-chip experiments of gent and lindley (1959) explained, Journal of the Mechanics and Physics of Solids 150 (2021) 104359.
  • Kumar et al. (2022) A. Kumar, K. Ravi-Chandar, O. Lopez-Pamies, The revisited phase-field approach to brittle fracture: application to indentation and notch problems, International Journal of Fracture 237 (2022) 83–100.
  • Kumar et al. (2024) A. Kumar, Y. Liu, J. E. Dolbow, O. Lopez-Pamies, The strength of the brazilian fracture test, Journal of the Mechanics and Physics of Solids 182 (2024) 105473.
  • Bourdin et al. (2008) B. Bourdin, G. A. Francfort, J.-J. Marigo, The variational approach to fracture, Journal of elasticity 91 (2008) 5–148.
  • Zhang et al. (2017) X. Zhang, S. W. Sloan, C. Vignes, D. Sheng, A modification of the phase-field model for mixed mode crack propagation in rock-like materials, Computer Methods in Applied Mechanics and Engineering 322 (2017) 123–136.
  • Bryant and Sun (2018) E. C. Bryant, W. Sun, A mixed-mode phase field fracture model in anisotropic rocks with consistent kinematics, Computer Methods in Applied Mechanics and Engineering 342 (2018) 561–584.
  • Steinke et al. (2022) C. Steinke, J. Storm, M. Kaliske, Energetically motivated crack orientation vector for phase-field fracture with a directional split, International Journal of Fracture 237 (2022) 15–46.
  • Pham et al. (2011) K. Pham, H. Amor, J.-J. Marigo, C. Maurini, Gradient damage models and their use to approximate brittle fracture, International Journal of Damage Mechanics 20 (2011) 618–652.
  • Tanné et al. (2018) E. Tanné, T. Li, B. Bourdin, J.-J. Marigo, C. Maurini, Crack nucleation in variational phase-field models of brittle fracture, Journal of the Mechanics and Physics of Solids 110 (2018) 80–99.
  • Bobet and Einstein (1998) A. Bobet, H. Einstein, Fracture coalescence in rock-type materials under uniaxial and biaxial compression, International Journal of Rock Mechanics and Mining Sciences 35 (1998) 863–888.
  • Sagong and Bobet (2002) M. Sagong, A. Bobet, Coalescence of multiple flaws in a rock-model material in uniaxial compression, International Journal of Rock Mechanics and Mining Sciences 39 (2002) 229–241.
  • Lee and Jeon (2011) H. Lee, S. Jeon, An experimental and numerical study of fracture coalescence in pre-cracked specimens under uniaxial compression, International Journal of Solids and Structures 48 (2011) 979–999.
  • Zhou et al. (2021) X.-P. Zhou, J.-Z. Zhang, S.-Q. Yang, F. Berto, Compression-induced crack initiation and growth in flawed rocks: a review, Fatigue & Fracture of Engineering Materials & Structures 44 (2021) 1681–1707.
  • Nemat-Nasser and Horii (1982) S. Nemat-Nasser, H. Horii, Compression-induced nonplanar crack extension with application to splitting, exfoliation, and rockburst, Journal of Geophysical Research: Solid Earth 87 (1982) 6805–6821.
  • Horii and Nemat-Nasser (1985) H. Horii, S. Nemat-Nasser, Compression-induced microcrack growth in brittle solids: Axial splitting and shear failure, Journal of Geophysical Research: Solid Earth 90 (1985) 3105–3125.
  • Steif (1984) P. S. Steif, Crack extension under compressive loading, Engineering Fracture Mechanics 20 (1984) 463–473.
  • Zhou et al. (2018) S. Zhou, X. Zhuang, H. Zhu, T. Rabczuk, Phase field modelling of crack propagation, branching and coalescence in rocks, Theoretical and Applied Fracture Mechanics 96 (2018) 174–192.
  • You et al. (2020) T. You, Q.-Z. Zhu, P.-F. Li, J.-F. Shao, Incorporation of tension-compression asymmetry into plastic damage phase-field modeling of quasi brittle geomaterials, International Journal of Plasticity 124 (2020) 71–95.