Abstract

Fracture in solid solutions, such as electrodes for lithium-ion batteries and fuel cells, is mediated by intricate interactions between solid-state diffusion and crack propagation. In this work, we developed a composition-dependent cohesive zone model and integrated it with a chemo-mechanical coupling constitutive model to study the fracture mechanisms of solid solutions. The computational framework was used to investigate the effective fracture properties of chemo-mechanically coupled solid solutions over a wide range of crack growth velocities and compositional dependence of intrinsic fracture energy. The results revealed an important characteristic crack velocity, which is set by the ratio of the diffusivity to the intrinsic fracture energy and dictates the effective fracture resistance of the material. We also applied the model to study the fracture behavior of two-phase lithiated silicon (Si) and germanium (Ge) nanostructures as candidate high-capacity anodes for next-generation lithium-ion batteries, and showed that Ge nanostructures are more fracture resistant than their Si counterparts. The computational study presented here provides important insights for the rational design, operation, and mechanical testing of chemo-mechanically active material systems for their use in energy storage and conversion.

1 Introduction

Solid-state diffusion occurs ubiquitously in many material systems, such as rechargeable battery electrodes, fuel cells for energy conversion, and metal alloys. In these solid solutions, mass transport and diffusion often generate substantial stress, which can lead to the fracture and ultimate failure of the materials. For the rational design and development of these material systems, the combined effects of chemical and mechanical driving forces on fracture must be thoroughly understood.

A general thermodynamic framework accounting for stress–diffusion coupling in multicomponent solids was established by Larché and Cahn [1]. Using a stress-dependent chemical potential as one of the key ingredients, this framework was initially applied to metallic systems [25] and later extended to study ionic crystals [6,7], nonstoichiometric oxide ceramics [811], and oxide-metal interfaces [12,13]. More recently, the chemomechanics of lithium-ion battery (LIB) electrodes has been a topic of intensive research, driven by the escalating electrification in various sectors. The operation of a LIB involves repeated insertion and extraction of mobile lithium (Li) ions in electrode materials, which causes appreciable volume changes. In particular, many alloying- and conversion-type electrode materials, such as silicon (Si), germanium (Ge), tin, and oxide and sulfide compounds, undergo large-volume-change transformations (up to 300%) during lithiation [1417]. If left unaddressed, this large deformation can lead to mechanical failure and consequently rapid capacity fading of the electrodes which compromises their high-capacity advantage [1820].

Extensive experimental and modeling work has recently been conducted to gain the fundamental understanding of chemo-mechanical interplay in the context of LIB electrodes. The multibeam optical sensor (MOS) and nanoindentation techniques have been employed to characterize the stress evolution [21], stress-potential coupling [22], and stress–diffusion coupling of lithiated thin-film electrodes [23,24]. Christensen and Newman are among the first to model the coupling between mechanical stresses and Li diffusion using a mathematical model for diffusion-induced stresses [25,26]. A variety of computational continuum models were subsequently developed to account for more general material behaviors, such as finite deformation, elastoplasticity, and phase segregation [2735]. Atomistic mechanisms underpinning diffusive transport and plastic deformation of LixSi nanostructures have also been revealed using ab initio calculations [36,37] and molecular dynamics (MD) simulations [3840]. In addition, analytical approaches have been taken to derive closed-form relationships to characterize stress buildup in electrode materials [4143].

Fracture caused by diffusion-induced stresses within solid electrodes has been analyzed with several classical fracture criteria, including Griffith’s criterion [44,45], stress intensity factor criterion [4648], and nonlinear energy release rate (ERR) (J-integral) criterion [4951]. Grantab et al. [52] and Bower et al. [53] used a cohesive zone approach to model crack nucleation and propagation in electrode materials by assuming composition-independent fracture properties. Employing large-scale atomistic simulations, Ding et al. investigated the fracture behavior of amorphous lithiated Si [54]. Their work revealed a brittle fracture at low Li concentrations due to void nucleation and coalescence and ductile fracture at high Li concentrations through extensive shear banding. The fracture toughness of lithiated Si versus composition and charging/discharging history was further studied by Khosrownejad et al. through K-controlled MD simulations [55] and was found to be consistent with the Gurson model for ductile fracture. More recently, Xu et al. investigated the corrosive fracture of electrodes in LIBs by implementing a theory of coupled diffusion and large deformation [56]. The theory was applied to model intergranular fracture in LiNixMnyCozO2 (NMC) cathode particles during delithiation and Li embrittlement was found to accelerate the initiation and growth of intergranular cracks in NMC particles.

In a solid solution, the coupling of crack-tip mechanical and chemical fields can have a profound influence on fracture. Stress concentration near a crack tip modifies the gradient of chemical potential which drives solute diffusion. The resulting change in solute concentration, in turn, can alter the crack-tip deformation and stress fields. Despite its importance, the intriguing interaction between mass diffusion and crack growth in solid solutions remains understudied and poorly understood. In this work, we develop a comprehensive computational framework for modeling the fracture of chemo-mechanically coupled solids. In Sec. 2, we present this framework which features an integration of a composition-dependent cohesive zone model with a two-way, chemo-mechanical coupling continuum model. Numerical implementation of this framework is explained in Sec. 3. In Sec. 4 we apply our computational framework to investigate the crack growth resistance of solid solutions and the fracture behavior of lithiated nanostructures. Finally, we conclude in Sec. 5.

2 Theoretical Framework

2.1 Kinematics.

As illustrated in Fig. 1(a), we consider the propagation of a crack in a solid solution AξB that consists of guest mobile species A and host species B. The molar concentration of A in AξB, i.e., the solvability of A in B, varies from ξ = 0 to ξ = ξmax. A dimensionless concentration of A is defined as c = ξ/ξmax (0 ≤ c ≤ 1). The diffusivity of the host atoms is assumed to be negligible compared with the guest ones, which allows the definition of continuum deformation [1]. When subjected to mechanical or chemical loadings, such as the transport and diffusion of species A, the solid may change its shape. The position of a material point X in the reference (Lagrangian) configuration changes to a new position x in the deformed (Eulerian) configuration. In what follows, we assume that all field quantities are functions of X, unless stated explicitly otherwise. The displacement of the material point can be then calculated as
(1)
We employ a cohesive zone approach to model the crack growth. The kinematics of the cohesive zone is described by the crack opening displacement
(2)
in which u+ and u are the displacements of two initially coincident points on the opposite crack surfaces.The change in the shape of a material element within the solid is characterized by the deformation gradient tensor
(3)
where denotes the gradient operator with respect to the Lagrangian coordinates and I is the identity matrix. The Lagrangian finite strain tensor is defined in terms of the deformation gradient
(4)
The total deformation of the material element is attributed to three processes: an irreversible plastic deformation Fp, a stress-free volume change FSF due to the expansion of the host material during guest species insertion, and a reversible elastic deformation Fe. The three contributions are assumed to be multiplicative through
(5)
Combining Eqs. (4) and (5), the total Lagrangian strain can be written as
(6)
where Ee, ESF, and Ep are defined based on Eq. (4) and refer to the elastic, compositional, and plastic strains, respectively. The velocity gradient of the deformation follows as
(7)
where overdot denotes time derivative, and x is the gradient operator with respect to the Eulerian coordinates. The three terms, Le=F˙eFe1, LSF=FeF˙SFFSF1Fe1, and Lp=FeFSFF˙pF1, can be interpreted as the velocity gradients due to elastic stretching, the insertion of guest species, and plastic flow, respectively. If the deformation gradient due to the insertion of guest species is isotropic, FSF takes the form of
(8)
where β(c) measures the volume expansion when a concentration of guest species c is inserted into the solid network. Assuming the atomic volumes of the host and guest species, Ωh and Ωg, are additive and remain unchanged during insertion, we have the following linear relationship
(9)
in which Ω is a dimensionless parameter determined by Ω = ξmaxΩgh.
Fig. 1
(a) Schematic illustration of crack growth within a chemo-mechanically coupled solid solution. (b) A chemical-composition-dependent bilinear cohesive zone law adopted in this study. (c) Schematic of a mixed finite element framework used for handling chemo-mechanical coupling.
Fig. 1
(a) Schematic illustration of crack growth within a chemo-mechanically coupled solid solution. (b) A chemical-composition-dependent bilinear cohesive zone law adopted in this study. (c) Schematic of a mixed finite element framework used for handling chemo-mechanical coupling.
Close modal

2.2 Variational Formulation.

Classical state-type variational principles are applicable to conservative systems only and cannot be used for dissipative systems undergoing irreversible processes. Here we present a rate-type variational approach for formulating the full coupling between diffusion, finite elastoplastic deformation, and fracture in chemo-mechanically active solids. As will be shown, the variational framework provides a complete, physics-based description of the key governing equations that are thermodynamically consistent. We start by defining the isothermal rate potential of the system as
(10)
Here, φbulk is the free energy density of the bulk material, ψpl and ψchem are the dissipation potentials due to plastic deformation and diffusion, respectively, Pext is the power of the external work, and φcoh is the free energy density of the cohesive crack that depends on both the crack opening displacement and the concentration of the mobile species. V()dV and Scoh()dS denote volume integration over the material domain and surface integration over the cohesive zone, respectively, in the reference (Lagrangian) configuration. We assume φbulk has three components, including an elastic energy density function φel, a plastic work density function φpl depending on a set of internal strain-hardening variables α, and a concentration-dependent chemical energy function φchem. The three contributions are assumed to be additive according to
(11)
Invoking the principle of maximum plastic dissipation and assuming the associated flow rule, we have the following plastic dissipation function
(12)
Here, LP is the plastic velocity gradient chosen as a measure of the plastic strain rate, λ˙0 is the plastic multiplier, and ΣP and sP are the driving and resistance forces for plastic flow that are work-conjugate to the plastic strain rate and hardening variables, respectively. 〈f〉 ≡ max (0, f) is the Macaulay operator of a convex yield function f. Following Miehe et al. [57], we postulate a concave form of the chemical dissipation function as
(13)
where μ is the chemical potential, D is the diffusivity, and F1FT is the inverse of the right Cauchy–Green tensor used to pull back the chemical potential gradient to the reference space. Assuming quasi-static deformation, the power of the external work is calculated as
(14)
in which b is the body force, and t and Q are the traction and surface flux prescribed at boundaries ST and SQ, respectively. With the definitions of Eqs. (11)(14), the rate potential Π˙ is a function that is convex in {u˙,c˙,LP,α˙} and concave in {μ,ΣP,sP}. A mixed multifield variational analysis can be performed by maximizing the plastic and chemical dissipations with respect to {μ,ΣP,sP} and minimizing the total rate potential with respect to {u˙,c˙,LP,α˙}. Specifically, variation of Π˙ with respect to u˙ gives the weak form of stress equilibrium or the principle of virtual work:
(15)
in which σPK1 is the first Piola–Kirchhoff stress tensor, and T=φcoh/Δ|c is the cohesive traction. Variation of Π˙ with respect to μ gives the weak form of mass balance:
(16)
where J=(cD/kBT)F1FTμ is the flux in the Lagrangian frame. Variation of Π˙ with respect to c˙ leads to the weak form of chemical microforce balance:
(17)
in which μchem = ∂φchem/∂c is the stress-free chemical potential in absence of mechanical constraints and can be expressed as
(18)
Here, μ0 is the reference chemical potential, and γ is the activity coefficient. The term φel/c|F,Fp in Eq. (17) modifies the total chemical potential and is responsible for the coupling between mechanical stress and diffusion. It can be shown, in the most general case, that
(19)
Assuming isotropic elastic properties, small elastic deformation, and volume-preserving plastic deformation (i.e., det(Fp)=1), we can express the elastic energy density in terms of the bulk modulus K and shear modulus S as
(20)
Here, tr(Ee) and dev(Ee) are the trace and deviatoric components of the elastic strain. With further assumption of linear, isotropic stress-free expansion, Eqs. (8), (9), and (20) can be substituted into Eq. (19) to yield
(21)
in which σ=det(F)1FσPK1 is the Cauchy stress, and σh=tr(σ)/3 is the hydrostatic stress in the Eulerian description. Finally, variation with respect to the remaining field variables provides all the evolution and governing equations for plasticity:
(22a)
(22b)
(22c)
(22d)
In Eq. (22a), H(f) is the Heaviside step function: H(f)=1iff0, 0 otherwise. In the simple case of isotropic hardening, only a single state variable, namely the equivalent plastic strain, is needed: α={εp}. The corresponding work-conjugate becomes the equivalent stress: sP={σp}. In this work, we adopt the classical von Mises yield function with linear strain hardening defined by
(23)
(24)
where σY is the yield stress, and h is the hardening modulus. Consequently, Eqs. (22a)(22c) can be explicitly written as
(25a)
(25b)
(25c)

2.3 Composition-Dependent Cohesive Zone Law.

A chemical-composition-dependent, bilinear traction-separation law is adopted to describe the cohesive zone behavior, as illustrated in Fig. 1(b). σ0 quantifies the strength of the interface where the crack propagates along, σ0/δ0 controls the initial elastic stiffness, and δc is the critical opening beyond which the cohesive zone completely loses its load carrying capability. kc is a large contact stiffness that resists the inter-penetration of crack surfaces under compressive loading. To limit artificial compliance introduced by the cohesive zone, the ratio of δ0 to δc is taken to be small at 0.05. δc and σ0 together determine the intrinsic fracture energy through Gc = σ0δc/2, which is in general a function of guest species concentration. In a finite element simulation, the opening of a discretized cohesive zone is known to be dictated by the constrained deformation kinematics associated with the shape functions. In this work, we consider the fracture process of the bulk material to be localized at the atomic scale, resulting in an intrinsic cohesive zone size much smaller than the mesh size. In this case, the cohesive strength and critical opening can be deduced from the fracture energy and finite element size according to a scaling law [58]:
(26a)
(26b)
where ν is Poisson’s ratio of the bulk material and L is the finite element size.

2.4 Evaluation of Energy Release Rate.

For quantitative evaluation of the ERR during crack propagation, we employ a chemo-mechanical J-integral method. The classical J-integral [59] for elastic materials is defined over a closed contour Γ,
(27)
where φel is the elastic energy density, t is the traction vector on the contour, and N1 is the first component of the unit outward normal N to the contour. The J-integral represents a balance between the total energy inside the closed contour and the energy passing through the contour line. In the presence of solid-state diffusion, the total energy inside the closed contour has a significant contribution from the chemical energy that drives diffusion. Considering mechanical energy alone hence no longer satisfies the energy balance [60]. Instead, the crack-tip region needs to be treated as a thermodynamic system, which can exchange mass, energy, and entropy with the surrounding continuum body. In this work, we adopt a path-independent J-integral accounting for coupled chemo-mechanical driving forces [49]:
(28)
in which Jmech and Jchem are the mechanical and chemical contributions to the modified J-integral, φel and φpl are the elastic strain energy density and plastic work density inheriting the same definition in Eq. (11), and VΓ is the crack-tip domain bounded by Γ. The domain integral Jchem quantifies the change in chemical energy of VΓ and renders the path independence of the chemo-mechanical J-integral Jcm. We now consider two different contours enclosing the tip of a propagating crack (Fig. 1(a)), one in the far field (Γ) and the other at the crack tip (Γtip). Because the domain integral Jchem vanishes for Γtip, Jcmtip) and Jmechtip) are identical and calculate the crack-tip ERR Jtip. On the other hand, Jmech) can be interpreted as the applied ERR in the far field (J). Invoking the path independence of Jcm, we have
(29)
where ΔJ=Jchem(VΓ)=VΓμ(c/X1)dV is the diffusion-induced energy dissipation within the region enclosed by Γ. Griffith’s criterion for crack growth requires Jtip = Gc and gives the far-field, effective fracture resistance as
(30)

3 Finite Element Implementation

The governing equations in their variational forms (Eqs. (15)(17)) are discretized in space and time and then solved numerically using a finite element method. The field variables at time t = tn are interpolated as follows:
(31a)
(31b)
(31c)
(31d)
where ai(i=u,Δ,c,μ) are the nodal degrees of freedom (DOFs) of displacement, cohesive separation, concentration, and chemical potential, respectively, and Ni are the corresponding shape functions defined in the reference frame. Discretizing the variational governing equations by substituting Eqs. (31a)(31d) and requiring the discretized equations to hold for arbitrary variations in nodal DOFs result in the following residual force equations:
(32a)
(32b)
(32c)
Here, Bi=Ni are the spatial derivatives of the shape functions, and M is a transformation matrix that relates the local nodal cohesive separations to the global nodal displacements. The unconditionally stable backward Euler approximation of the diffusion rate term, c˙, is used in Eq. (32b) for time discretization. The set of nonlinear equations are solved for the current time tn from the previous time using the iterative Newton–Raphson method. The Jacobians needed for iterations are obtained by differentiating the residual forces (Eqs. (32a)(32c)) with respect to the nodal variables.

The above finite element scheme is implemented through the user element (UEL) interface of ABAQUS. The bulk material is modeled using three-dimensional (3D) 20-node brick elements (Fig. 1(c)). The eight corner nodes have displacement, concentration, and chemical potential DOFs, and the 12 edge nodes only have displacement DOFs. Second-order isoparametric shape functions are chosen for displacement interpolation, and linear isoparametric shape functions are used to interpolate the concentration and chemical potential fields. The cohesive zone is modeled using 16-node cohesive zone elements, which have zero initial thickness and share nodes with the adjacent bulk elements. Like the bulk elements, the cohesive zone elements combine quadratic interpolation for the crack opening displacement and linear interpolation for the other two fields. This hybrid interpolation scheme is used here to provide an optimal balance between computational accuracy and efficiency.

4 Results and Discussion

4.1 Crack Growth Resistance of a Model Solid Solution.

The computational framework outlined in the preceding sections has been applied to investigate a model problem, where a plane strain specimen containing an initial edge crack is subjected to mode I loading (Fig. 1(a)). Through this model problem, we seek to gain fundamental insights into how the fracture behavior of a solid solution is controlled by the interplay between the mechanical and chemical fields around a crack tip. The specimen has a width of W, a height of H = 0.3W, an initial crack length of a0 = 0.3W, and a uniform mesh size of L = W/200. The solid solution is initially stress free with a homogeneous distribution of guest species concentration c0, which corresponds to an initial intrinsic fracture energy of Gc0=Gc(c0). For simplicity, an isotropic linear elastic material behavior is assumed and Young’s modulus and Poisson’s ratio are taken to be independent of the concentration. Due to the geometrical and loading symmetry about the crack plane, only the upper half of the specimen is simulated. Mode I fracture loading is applied by prescribing a pair of opening displacements with constant speed vapp on the upper and lower vertical surfaces to the left of the crack. As the fracture loading increases, a cohesive crack initiates and propagates with a nonlinear fracture process zone at the crack tip, whose constitutive behavior is governed by the traction-separation law described earlier. The path-independent chemo-mechanical J-integral calculation (Sec. 2.4) is performed on the cracked specimen to evaluate the energy release rates at various stages of loading, which are then used to construct the fracture resistance curve (R-curve) of J versus Δa of the material.

In quasi-static fracture simulations, snap-back instability in the load–displacement path is often observed because of the excessive rate of change in crack-tip ERR (i.e., dJtip/da > dGc/da), leading to numerical convergence issues. With the displacement-controlled loading condition in the current model, the ERR decreases with crack growth, which helps maintain crack stability. However, as will be shown later, the resistance curve of the material is strongly influenced by chemo-mechanical coupling, which may lead to crack instability even under displacement control. To overcome this issue, a small artificial viscosity is introduced in the cohesive zone law [61,62]. This viscosity term provides an energy dissipation mechanism for capturing post-instability behaviors, while making a negligible contribution to the fracture resistance during stable crack propagation.

Systematic parametric studies have been carried out to study the effects of two controlling parameters, the crack growth velocity vc and the concentration dependence of intrinsic fracture energy dGc/dc. Our dimensional analysis yields a characteristic length l0=Gc0/E, a characteristic time t0=(Gc0)2/(E2D), and a characteristic velocity v0=ED/Gc0, which are used for dimensionless representation of the relevant quantities. Also, the concentration dependence of intrinsic fracture energy is normalized by its initial value Gc0 to be expressed as a dimensionless parameter χ=(1/Gc0)(dGc/dc).

Figure 2(a) shows a plot of crack advance versus time for three distinct loading velocities (vapp = 0.000265, 0.0265, and 2.65) and a representative value of χ = 3.33. The crack length is seen to increase linearly with time in the low- and high-velocity cases, but this linear relationship does not hold very well for the intermediate velocity. Linear fitting of each curve yields an average crack growth velocity vc, which takes the place of vapp as a more intrinsic fracture condition. The R-curves for the three cases are shown in Fig. 2(b). For the high and low crack velocities (vc/v0 = 24.7 and 0.0019), J is found to increase abruptly and reach a plateau (maximum) value with little crack extension. In sharp contrast, the J curve for the intermediate crack velocity (vc/v0 = 0.19) fluctuates as the crack first starts to grow, and eventually converges to a plateau. In the following, the plateau value of each J curve will be interpreted as the effective fracture resistance Gceff under the specific fracture condition.

Fig. 2
(a) Crack extension curves under different loading velocities. (b) Normalized crack growth resistance curves (R-curves) under various loading velocities.
Fig. 2
(a) Crack extension curves under different loading velocities. (b) Normalized crack growth resistance curves (R-curves) under various loading velocities.
Close modal

To elucidate the fracture mechanism underlying the different resistance curves in Fig. 2(b), we resort to the steady-state mechanical and chemical fields around the crack tip, which are obtained when the plateau values of J are reached. Figure 3 shows the spatial distributions of concentration change Δc = cc0, hydrostatic stress σh, and crack opening stress component σyy. When the crack propagates at the low velocity (vc/v0 = 0.0019), the gradient in hydrostatic stress near the crack tip drives the guest species to diffuse toward the crack tip and cause solute accumulation around the tip, as evidenced by the contour plot a1 in Fig. 3(a) and the line plot in Fig. 3(b). The influx of the guest species can influence the effective fracture resistance in two separate ways. First, it causes the local area in the wake of the crack tip to expand. Such dilatational deformation lowers the local stress level of the crack-tip region, thus producing an effective shielding from the far-field mechanical loading (see a2 in Figs. 3(a) and 3(c)). We note that this toughening mechanism closely resembles the concept of stress-induced phase transformation in shape memory alloys [63] and zirconia-containing ceramics [64]. From an energetic point of view, this toughening effect can also be attributed to the energy dissipation by diffusion in the near-crack-tip region. Second, since the fracture energy of the solid solution is in general dependent on the chemical composition, the change in crack-tip concentration directly alters the intrinsic resistance to fracture.

Fig. 3
(a) Contour plots of the change in solute concentration Δc = c − c0 (a1)–(c1), hydrostatic stress σh (a2)–(c2), and in-plane stress component σyy (a3)–(c3) for three different loading velocities (vapp = 0.000265, 0.0265, and 2.65). (b)–(d) Line plots of Δc, σh, and σyy along the crack line corresponding to the contour plots in (a).
Fig. 3
(a) Contour plots of the change in solute concentration Δc = c − c0 (a1)–(c1), hydrostatic stress σh (a2)–(c2), and in-plane stress component σyy (a3)–(c3) for three different loading velocities (vapp = 0.000265, 0.0265, and 2.65). (b)–(d) Line plots of Δc, σh, and σyy along the crack line corresponding to the contour plots in (a).
Close modal

For fast crack growth (FCG) (vc/v0 = 24.7), the time taken for crack growth is much smaller than the characteristic time for solute diffusion. Therefore, no appreciable amount of diffusion takes place in the crack-tip region, leading to no significant change in solute concentration (see c1 in Figs. 3(a) and 3(b)). In this case, the solute is essentially immobile and the fracture process is purely mechanically driven. This renders an effective fracture resistance equal to the intrinsic fracture energy of the material.

For the intermediate crack velocity (vc/v0 = 0.19), the crack growth and diffusion processes occur on comparable time scales, which give rise to a more complex fracture behavior than the two previous cases. Figure 4 shows the contour and line plots of Δc and σh at four successive time instants that are marked as A–D on the R-curve in Fig. 2(b). Upon initial loading, the stress singularity at the crack tip creates a localized region of high solute concentration (instant A). When the crack starts to propagate, the crack growth velocity is so large that solute diffusion cannot catch up, causing the high-concentration site to lag behind the crack tip (instant B). This hysteretic phenomenon is more evident as the crack further propagates, and a second high-concentration site forms at the moving crack tip and the original one starts to fade away (instant C). The negotiation of the propagating crack with solute diffusion results in the nonconstant crack velocity shown in Fig. 2(a) and the oscillating crack growth resistance curve as seen in Fig. 2(b). Beyond time instant D, a dynamic equilibrium between the chemical and mechanical driving forces for diffusion is reached and maintained, giving rise to a plateau fracture resistance value between those of the low- and high-crack-velocity cases.

Fig. 4
(a) Contour plots of the change in solute concentration Δc = c − c0 (a1)–(d1) and hydrostatic stress σh (a2)–(d2) corresponding to the four time instants (a)–(d) marked on the R-curve in Fig. 2. (b) and (c) Line plots of Δc and σh along the crack line corresponding to the contour plots in (a).
Fig. 4
(a) Contour plots of the change in solute concentration Δc = c − c0 (a1)–(d1) and hydrostatic stress σh (a2)–(d2) corresponding to the four time instants (a)–(d) marked on the R-curve in Fig. 2. (b) and (c) Line plots of Δc and σh along the crack line corresponding to the contour plots in (a).
Close modal

In Fig. 5(a) we present a parametric study of the dependence of effective fracture resistance Gceff on crack growth velocity and χ. The characteristic velocity v0, which is set by the ratio of diffusivity to intrinsic fracture energy, serves as an important parameter for dictating the effective fracture resistance. Depending on the crack velocity relative to this characteristic velocity, there are three distinct fracture mechanism regimes: the slow crack growth (SCG) regime, the FCG regime, and the intermediate crack growth (ICG) regime in between. For a fixed value of χ, Gceff is found to maintain a relatively constant value in the SCG regime (vc/v0 < 0.002). As vc/v0 increases, all the Gceff curves first rise in the ICG regime, and then converge nonmonotonically to the same asymptotic value in the FCG regime (vc/v0 > 2) that is equal to the initial intrinsic fracture energy Gc0. To understand the complex behavior of these Gceff curves, we consider the decomposition of Gceff into the crack-tip fracture energy Gc and the diffusion-induced energy dissipation ΔJ, which are presented in Figs. 5(b) and 5(c). In the SCG and ICG regimes, the solute atoms have sufficient time to diffuse and redistribute around the crack tip. The diffusion may lead to enhanced (Gc>Gc0) or reduced (Gc<Gc0) crack-tip fracture energy, depending on the sign of the χ value. However, the diffusion-induced ΔJ is always positive and therefore contributes to fracture toughening as discussed earlier. It is worth noting that the ΔJ curves in Fig. 5(c) are maximized in the ICG regime and decay toward zero for both low- and high-crack velocities, giving rise to the nonmonotonic dependence of Gceff on crack velocity. The trend of ΔJ → 0 as vc → 0 may seem counterintuitive, considering that a low crack velocity promotes diffusion in the near-crack-tip region. Nevertheless, this result can be rationalized by the fact that diffusion-induced dilatational eigenstrains around a stationary crack do not affect the energy release rate (i.e., ΔJ = 0). Such assertion has previously been proved in the context of dilatant transformation toughening in ceramics [65].

Fig. 5
(a) Effective fracture energy, (b) crack-tip fracture energy, and (c) J-integral imbalance as functions of crack growth velocity for five different χ values. (a) and (b) share the same legend as (c).
Fig. 5
(a) Effective fracture energy, (b) crack-tip fracture energy, and (c) J-integral imbalance as functions of crack growth velocity for five different χ values. (a) and (b) share the same legend as (c).
Close modal

The fracture mechanism map in Fig. 5(a) has profound implications for materials design and characterization of chemo-mechanically coupled solid solutions. In a solid solution where the solute concentration induces toughening, crack growth velocities in the ICG regime are favorable for attaining high effective damage tolerance of the material. On the other hand, if a material is susceptible to solute embrittlement, crack velocities beyond the SCG regime can mitigate the embrittlement effect. For fracture testing of solid solutions, crack growth velocities maintained within the FCG regime can prevent an appreciable amount of solute diffusion around the crack tip and therefore allow for the accurate measurement of intrinsic fracture properties. For example, typical diffusivity values for most solid solutions range from 10−12 to 10−16 m2/s. Assuming representative values of E = 100 GPa and Gc0=10J/m2, we obtain characteristic crack velocities ranging from 0.001 to 10 mm/s. While crack growth velocities higher than these are relatively easy to attain for macroscopic fracture specimens, fracture testing at the nano- and micro-scale is usually performed at fairly slow loading rates due to instrumental constraints. Therefore, care must be taken in these fracture experiments regarding the FCG requirement to properly interpret measurement results.

4.2 Fracture of Nanostructured Battery Electrodes.

To further illustrate the utility of our computational framework, we have also applied it to study the fracture behavior of Si nanoparticles (SiNPs) and Ge nanoparticles (GeNPs) upon lithiation. In our earlier works [66,67], the fracture properties of electrochemically lithiated Si and Ge at various Li concentrations were quantitatively measured by nanoindentation. Electrodes for nanoindentation testing were fabricated in a thin-film form and electrochemically lithiated to various Li concentrations. To measure the fracture toughness, the LiξSi and LiξGe electrodes were fracture tested with a cube-corner indenter tip. The indentation load and crack sizes were used to evaluate the fracture toughness of LiξSi and LiξGe with the Morris nanoindentation model [68,69]. Figure 6(a) shows the measured fracture energies of LiξSi and LiξGe as functions of Li concentration, which clearly reveal that the fracture energies of LiξSi and LiξGe alloys depend sensitively on the Li concentration. As the Li concentration increases, the fracture energy of LiξSi first decreases slightly, indicating lithiation-induced embrittlement. However, upon further lithiation beyond ξ = 0.31, the measured fracture energy increases substantially with increasing Li concentration. Unlike LiξSi, the obtained fracture energy of LiξGe is shown to increase monotonically with the increase of Li concentration. These experimentally characterized fracture properties of LiξSi and LiξGe are incorporated into the computational model in this work to improve its predictive capability.

Fig. 6
(a) Fracture energies of lithiated Si and Ge as functions of Li concentration [66,67]. (b) TEM image of two-phase lithiation in amorphous Si [70]. (c) Schematic of a double-well potential used for modeling two-phase lithiation.
Fig. 6
(a) Fracture energies of lithiated Si and Ge as functions of Li concentration [66,67]. (b) TEM image of two-phase lithiation in amorphous Si [70]. (c) Schematic of a double-well potential used for modeling two-phase lithiation.
Close modal

Figure 7 presents the simulated fracture behavior of lithiated Si and Ge nanoparticles using our computational framework. An initial crack is placed at the edge of each nanoparticle configuration. Three particle diameters, 100 nm, 250 nm, and 500 nm, are used for both Si and Ge nanoparticles in the comparative study, with the initial crack length being 4% of the particle diameter in each simulation. Due to the geometry and loading symmetry about the crack plane, only the upper half of each nanoparticle is modeled. A linear rule of mixtures is used to determine the composition-dependent mechanical properties, including the yield strength and elastic modulus, of lithiated Si and Ge. The lithiation of Si and Ge has been experimentally observed to proceed with the movement of a sharp phase boundary between pristine and lithiated regions [7072], as shown in Fig. 6(b). This two-phase lithiation mechanism is implemented into our model by adding a double-well potential [73] to the chemical potential described in Eq. (18) and assuming ξ1 = 0 and ξ2 = 2.33 as two favorable states (Fig. 6(c)). As shown in Fig. 7, SiNPs exhibit size-dependent fracture upon lithiation, with no cracking observed for a small particle diameter of 100 nm (Fig. 7(a)) but obvious fracture for particle diameters of 250 nm and 500 nm (Figs. 7(b) and 7(c)). In distinct contrast to SiNPs, GeNPs undergo no visible cracking for all the three diameters investigated (Figs. 7(d)7(f)). These simulation results conform quantitatively to the in situ transmission electron microscopy (TEM) observations reported by previous works [74,75], indicating that GeNPs are more robust than their Si counterparts and can be a promising electrode candidate for durable, high-rate rechargeable batteries.

Fig. 7
Simulated fracture behavior of lithiated Si (a)–(c) and Ge (d)–(f) nanoparticles
Fig. 7
Simulated fracture behavior of lithiated Si (a)–(c) and Ge (d)–(f) nanoparticles
Close modal

5 Conclusions

In this work, we integrate a composition-dependent cohesive zone model with a diffusion–deformation coupling continuum model to investigate the fracture mechanics of chemo-mechanically coupled solid solutions. We apply our computational framework to analyze the crack growth resistance of a model solid-solution system. We show that the dynamic interaction between stress-induced diffusion and crack propagation dictates the fracture response of the system. A chemo-mechanical J-integral method allows the effective and crack-tip fracture energies to be calculated separately. For sufficiently high-crack velocities, both energy terms converge to the intrinsic fracture energy. However, when the crack grows at lower velocities, stress-induced diffusion can lead to enhancement or weakening in the crack-tip fracture resistance depending on the composition dependence of the intrinsic fracture energy. Diffusion-induced dilatational eigenstrains are found to contribute positively to the effective fracture resistance. Such a contribution is most significant when the crack velocity is comparable to a characteristic velocity set by the ratio of diffusivity to intrinsic fracture energy. We also use our computational method to study the fracture behavior of lithiated Si and Ge nanoparticles. The results reveal that SiNPs exhibit size-dependent fracture, while GeNPs experience no obvious cracking under similar lithiation conditions. The demonstrated mechanical robustness of GeNPs make them a promising electrode candidate for durable rechargeable batteries.

Acknowledgment

The authors gratefully acknowledge the support of the National Science Foundation (CMMI-1300458 and CMMI-1554393).

Conflict of Interest

There are no conflicts of interest.

Data Availability Statement

The datasets generated and supporting the findings of this article are obtainable from the corresponding author upon reasonable request.

References

1.
Larché
,
F.
, and
Cahn
,
J. W.
,
1973
, “
A Linear Theory of Thermochemical Equilibrium of Solids Under Stress
,”
Acta Metall.
,
21
(
8
), pp.
1051
1063
.
2.
Larche
,
F. C.
, and
Cahn
,
J. W.
,
1982
, “
The Effect of Self-Stress on Diffusion in Solids
,”
Acta Metall.
,
30
(
10
), pp.
1835
1845
.
3.
Larche
,
F. C.
, and
Cahn
,
J. W.
,
1985
, “
The Interactions of Composition and Stress in Crystalline Solids
,”
Acta Metall.
,
33
(
3
), pp.
331
357
.
4.
Larche
,
F. C.
, and
Cahn
,
J. W.
,
1987
, “
Stress Effects on III–V Solid–Liquid Equilibria
,”
J. Appl. Phys.
,
62
(
4
), pp.
1232
1239
.
5.
Larche
,
F. C.
, and
Cahn
,
J. W.
,
1992
, “
Phase-Changes in a Thin Plate With Nonlocal Self-Stress Effects
,”
Acta Metall. Mater.
,
40
(
5
), pp.
947
955
.
6.
Johnson
,
W. C.
,
1994
, “
Thermodynamic Equilibria in 2-Phase, Elastically Stressed Ionic-Crystals
,”
J. Am. Ceram. Soc.
,
77
(
6
), pp.
1581
1591
.
7.
Johnson
,
W. C.
, and
Schmalzried
,
H.
,
1993
, “
Phenomenological Thermodynamic Treatment of Elastically Stressed Ionic-Crystals
,”
J. Am. Ceram. Soc.
,
76
(
7
), pp.
1713
1719
.
8.
Krishnamurthy
,
R.
, and
Sheldon
,
B. W.
,
2004
, “
Stresses Due to Oxygen Potential Gradients in Non-stoichiometric Oxides
,”
Acta Mater.
,
52
(
7
), pp.
1807
1822
.
9.
Swaminathan
,
N.
, and
Qu
,
J.
,
2007
, “
Interactions Between Non-stoichiometric Stresses and Defect Transport in a Tubular Electrolyte
,”
Fuel Cells
,
7
(
6
), pp.
453
462
.
10.
Swaminathan
,
N.
,
Qu
,
J.
, and
Sun
,
Y.
,
2007
, “
An Electrochemomechanical Theory of Defects in Ionic Solids. I. Theory
,”
Philos. Mag.
,
87
(
11
), pp.
1705
1721
.
11.
Swaminathan
,
N.
,
Qu
,
J.
, and
Sun
,
Y.
,
2007
, “
An Electrochemomechanical Theory of Defects in Ionic Solids. Part II. Examples
,”
Philos. Mag.
,
87
(
11
), pp.
1723
1742
.
12.
Zhou
,
H. G.
,
Qu
,
J. M.
, and
Cherkaoui
,
M.
,
2010
, “
Finite Element Analysis of Oxidation Induced Metal Depletion at Oxide-Metal Interface
,”
Comput. Mater. Sci.
,
48
(
4
), pp.
842
847
.
13.
Zhou
,
H. G.
,
Qu
,
J. M.
, and
Cherkaoui
,
M.
,
2010
, “
Stress-Oxidation Interaction in Selective Oxidation of Cr–Fe Alloys
,”
Mech. Mater.
,
42
(
1
), pp.
63
71
.
14.
Obrovac
,
M. N.
, and
Christensen
,
L.
,
2004
, “
Structural Changes in Silicon Anodes During Lithium Insertion/Extraction
,”
Electrochem. Solid State Lett.
,
7
(
5
), pp.
A93
A96
.
15.
Obrovac
,
M. N.
, and
Krause
,
L. J.
,
2007
, “
Reversible Cycling of Crystalline Silicon Powder
,”
J. Electrochem. Soc.
,
154
(
2
), pp.
A103
A108
.
16.
Beaulieu
,
L. Y.
,
Hatchard
,
T. D.
,
Bonakdarpour
,
A.
,
Fleischauer
,
M. D.
, and
Dahn
,
J. R.
,
2003
, “
Reaction of Li With Alloy Thin Films Studied by In Situ AFM
,”
J. Electrochem. Soc.
,
150
(
11
), pp.
A1457
A1464
.
17.
Liu
,
X. H.
,
Liu
,
Y.
,
Kushima
,
A.
,
Zhang
,
S. L.
,
Zhu
,
T.
,
Li
,
J.
, and
Huang
,
J. Y.
,
2012
, “
In Situ TEM Experiments of Electrochemical Lithiation and Delithiation of Individual Nanostructures
,”
Adv. Energy Mater.
,
2
(
7
), pp.
722
741
.
18.
Xiao
,
X.
,
Liu
,
P.
,
Verbrugge
,
M. W.
,
Haftbaradaran
,
H.
, and
Gao
,
H.
,
2011
, “
Improved Cycling Stability of Silicon Thin Film Electrodes Through Patterning for High Energy Density Lithium Batteries
,”
J. Power Sources
,
196
(
3
), pp.
1409
1416
.
19.
Choi
,
S.
,
Kwon
,
T. W.
,
Coskun
,
A.
, and
Choi
,
J. W.
,
2017
, “
Highly Elastic Binders Integrating Polyrotaxanes for Silicon Microparticle Anodes in Lithium Ion Batteries
,”
Science
,
357
(
6348
), pp.
279
283
.
20.
McDowell
,
M. T.
,
Xia
,
S.
, and
Zhu
,
T.
,
2016
, “
The Mechanics of Large-Volume-Change Transformations in High-Capacity Battery Materials
,”
Extreme Mech. Lett.
,
9
(
3
), pp.
480
494
.
21.
Sethuraman
,
V. A.
,
Chon
,
M. J.
,
Shimshak
,
M.
,
Srinivasan
,
V.
, and
Guduru
,
P. R.
,
2010
, “
In Situ Measurements of Stress Evolution in Silicon Thin Films During Electrochemical Lithiation and Delithiation
,”
J. Power Sources
,
195
(
15
), pp.
5062
5066
.
22.
Sethuraman
,
V. A.
,
Srinivasan
,
V.
,
Bower
,
A. F.
, and
Guduru
,
P. R.
,
2010
, “
In Situ Measurements of Stress-Potential Coupling in Lithiated Silicon
,”
J. Electrochem. Soc.
,
157
(
11
), p.
A1253
.
23.
Papakyriakou
,
M.
,
Wang
,
X.
, and
Xia
,
S.
,
2018
, “
Characterization of Stress–Diffusion Coupling in Lithiated Germanium by Nanoindentation
,”
Exp. Mech.
,
58
(
4
), pp.
613
625
.
24.
Papakyriakou
,
M.
,
Lu
,
M.
, and
Xia
,
S.
,
2022
, “
Nanoindentation Size Effects in Lithiated and Sodiated Battery Electrode Materials
,”
ASME J. Appl. Mech.
,
89
(
7
), p.
071007
.
25.
Christensen
,
J.
, and
Newman
,
J.
,
2006
, “
A Mathematical Model of Stress Generation and Fracture in Lithium Manganese Oxide
,”
J. Electrochem. Soc.
,
153
(
6
), pp.
A1019
A1030
.
26.
Christensen
,
J.
, and
Newman
,
J.
,
2006
, “
Stress Generation and Fracture in Lithium Insertion Materials
,”
J. Solid State Electrochem.
,
10
(
5
), pp.
293
319
.
27.
Zhang
,
X. C.
,
Shyy
,
W.
, and
Sastry
,
A. M.
,
2007
, “
Numerical Simulation of Intercalation-Induced Stress in Li-Ion Battery Electrode Particles
,”
J. Electrochem. Soc.
,
154
(
10
), pp.
A910
A916
.
28.
Zhao
,
K. J.
,
Pharr
,
M.
,
Cai
,
S. Q.
,
Vlassak
,
J. J.
, and
Suo
,
Z. G.
,
2011
, “
Large Plastic Deformation in High-Capacity Lithium-Ion Batteries Caused by Charge and Discharge
,”
J. Am. Ceram. Soc.
,
94
(
1
), pp.
S226
S235
.
29.
Zhao
,
K. J.
,
Pharr
,
M.
,
Vlassak
,
J. J.
, and
Suo
,
Z. G.
,
2011
, “
Inelastic Hosts as Electrodes for High-Capacity Lithium-Ion Batteries
,”
J. Appl. Phys.
,
109
(
1
), p.
016110
.
30.
Anand
,
L.
,
2012
, “
A Cahn–Hilliard-Type Theory for Species Diffusion Coupled With Large Elastic–Plastic Deformations
,”
J. Mech. Phys. Solids
,
60
(
12
), pp.
1983
2002
.
31.
Bower
,
A. F.
,
Guduru
,
P. R.
, and
Sethuraman
,
V. A.
,
2011
, “
A Finite Strain Model of Stress, Diffusion, Plastic Flow, and Electrochemical Reactions in a Lithium-Ion Half-Cell
,”
J. Mech. Phys. Solids
,
59
(
4
), pp.
804
828
.
32.
Cui
,
Z. W.
,
Gao
,
F.
, and
Qu
,
J. M.
,
2012
, “
A Finite Deformation Stress-Dependent Chemical Potential and Its Applications to Lithium Ion Batteries
,”
J. Mech. Phys. Solids
,
60
(
7
), pp.
1280
1295
.
33.
Gao
,
Y. F.
, and
Zhou
,
M.
,
2011
, “
Strong Stress-Enhanced Diffusion in Amorphous Lithium Alloy Nanowire Electrodes
,”
J. Appl. Phys.
,
109
(
1
), p.
014310
.
34.
Gao
,
Y. F.
,
Cho
,
M.
, and
Zhou
,
M.
,
2013
, “
Stress Relaxation Through Interdiffusion in Amorphous Lithium Alloy Electrodes
,”
J. Mech. Phys. Solids
,
61
(
2
), pp.
579
596
.
35.
Huang
,
S.
,
Fan
,
F.
,
Li
,
J.
,
Zhang
,
S.
, and
Zhu
,
T.
,
2013
, “
Stress Generation During Lithiation of High-Capacity Electrode Particles in Lithium ion Batteries
,”
Acta Mater.
,
61
(
12
), pp.
4354
4364
.
36.
Zhao
,
K.
,
Wang
,
W. L.
,
Gregoire
,
J.
,
Pharr
,
M.
,
Suo
,
Z.
,
Vlassak
,
J. J.
, and
Kaxiras
,
E.
,
2011
, “
Lithium-Assisted Plastic Deformation of Silicon Electrodes in Lithium-Ion Batteries: A First-Principles Theoretical Study
,”
Nano Lett.
,
11
(
7
), pp.
2962
2967
.
37.
Wang
,
H.
,
Wang
,
X.
,
Xia
,
S.
, and
Chew
,
H. B.
,
2015
, “
Brittle-to-Ductile Transition of Lithiated Silicon Electrodes: Crazing to Stable Nanopore Growth
,”
J. Chem. Phys.
,
143
(
10
), p.
104703
.
38.
Yan
,
X.
,
Gouissem
,
A.
, and
Sharma
,
P.
,
2015
, “
Atomistic Insights Into Li-Ion Diffusion in Amorphous Silicon
,”
Mech. Mater.
,
91
(
2
), pp.
306
312
.
39.
Yan
,
X.
,
Gouissem
,
A.
,
Guduru
,
P. R.
, and
Sharma
,
P.
,
2017
, “
Elucidating the Atomistic Mechanisms Underpinning Plasticity in Li–Si Nanostructures
,”
Phys. Rev. Mater.
,
1
(
5
), p.
055401
.
40.
Ding
,
B.
,
Wu
,
H.
,
Xu
,
Z. P.
,
Li
,
X. Y.
, and
Gao
,
H. J.
,
2017
, “
Stress Effects on Lithiation in Silicon
,”
Nano Energy
,
38
, pp.
486
493
.
41.
Cheng
,
Y. T.
, and
Verbrugge
,
M. W.
,
2008
, “
The Influence of Surface Mechanics on Diffusion Induced Stresses Within Spherical Nanoparticles
,”
J. Appl. Phys.
,
104
(
8
), p.
083521
.
42.
Cheng
,
Y. T.
, and
Verbrugge
,
M. W.
,
2009
, “
Evolution of Stress Within a Spherical Insertion Electrode Particle Under Potentiostatic and Galvanostatic Operation
,”
J. Power Sources
,
190
(
2
), pp.
453
460
.
43.
Purkayastha
,
R.
, and
McMeeking
,
R. M.
,
2012
, “
A Linearized Model for Lithium Ion Batteries and Maps for Their Performance and Failure
,”
ASME J. Appl. Mech.
,
79
(
3
), p.
031021
.
44.
Aifantis
,
K. E.
,
Hackney
,
S. A.
, and
Dempsey
,
J. P.
,
2007
, “
Design Criteria for Nanostructured Li-Ion Batteries
,”
J. Power Sources
,
165
(
2
), pp.
874
879
.
45.
Zhao
,
K. J.
,
Pharr
,
M.
,
Vlassak
,
J. J.
, and
Suo
,
Z. G.
,
2010
, “
Fracture of Electrodes in Lithium-Ion Batteries Caused by Fast Charging
,”
J. Appl. Phys.
,
108
(
7
), p.
073517
.
46.
Woodford
,
W. H.
,
Carter
,
W. C.
, and
Chiang
,
Y. M.
,
2012
, “
Design Criteria for Electrochemical Shock Resistant Battery Electrodes
,”
Energy Environ. Sci.
,
5
(
7
), pp.
8014
8024
.
47.
Woodford
,
W. H.
,
Chiang
,
Y. M.
, and
Carter
,
W. C.
,
2013
, “
Electrochemical Shock in Ion-Intercalation Materials With Limited Solid-Solubility
,”
J. Electrochem. Soc.
,
160
(
8
), pp.
A1286
A1292
.
48.
Woodford
,
W. H.
,
Chiang
,
Y. M.
, and
Carter
,
W. C.
,
2010
, “
Electrochemical Shock” of Intercalation Electrodes: A Fracture Mechanics Analysis
,”
J. Electrochem. Soc.
,
157
(
10
), pp.
A1052
A1059
.
49.
Gao
,
Y. F.
, and
Zhou
,
M.
,
2013
, “
Coupled Mechano-Diffusional Driving Forces for Fracture in Electrode Materials
,”
J. Power Sources
,
230
, pp.
176
193
.
50.
Zhang
,
M.
,
Qu
,
J.
, and
Rice
,
J. R.
,
2017
, “
Path Independent Integrals in Equilibrium Electro-Chemo-Elasticity
,”
J. Mech. Phys. Solids
,
107
, pp.
525
541
.
51.
Yang
,
H.
, and
Qu
,
J.
,
2019
, “
Fracture Toughness of LixSi Alloys in Lithium Ion Battery
,”
Extreme Mech. Lett.
,
32
, p.
100555
.
52.
Grantab
,
R.
, and
Shenoy
,
V. B.
,
2012
, “
Pressure-Gradient Dependent Diffusion and Crack Propagation in Lithiated Silicon Nanowires
,”
J. Electrochem. Soc.
,
159
(
5
), pp.
A584
A591
.
53.
Bower
,
A. F.
, and
Guduru
,
P. R.
,
2012
, “
A Simple Finite Element Model of Diffusion, Finite Deformation, Plasticity and Fracture in Lithium Ion Insertion Electrode Materials
,”
Modell. Simul. Mater. Sci. Eng.
,
20
(
4
), p.
045004
.
54.
Ding
,
B.
,
Li
,
X. Y.
,
Zhang
,
X.
,
Wu
,
H.
,
Xu
,
Z. P.
, and
Gao
,
H. J.
,
2015
, “
Brittle Versus Ductile Fracture Mechanism Transition in Amorphous Lithiated Silicon: From Intrinsic Nanoscale Cavitation to Shear Banding
,”
Nano Energy
,
18
, pp.
89
96
.
55.
Khosrownejad
,
S. M.
, and
Curtin
,
W. A.
,
2017
, “
Crack Growth and Fracture Toughness of Amorphous Li–Si Anodes: Mechanisms and Role of Charging/Discharging Studied by Atomistic Simulations
,”
J. Mech. Phys. Solids.
,
107
, pp.
542
559
.
56.
Xu
,
R.
, and
Zhao
,
K. J.
,
2018
, “
Corrosive Fracture of Electrodes in Li-Ion Batteries
,”
J. Mech. Phys. Solids
,
121
, pp.
258
280
.
57.
Miehe
,
C.
,
Mauthe
,
S.
, and
Ulmer
,
H.
,
2014
, “
Formulation and Numerical Exploitation of Mixed Variational Principles for Coupled Problems of Cahn–Hilliard-Type and Standard Diffusion in Elastic Solids
,”
Int. J. Numer. Methods Eng.
,
99
(
10
), pp.
737
762
.
58.
Xia
,
S. M.
,
Qi
,
Y.
,
Perry
,
T.
, and
Kim
,
K. S.
,
2009
, “
Strength Characterization of Al/Si Interfaces: A Hybrid Method of Nanoindentation and Finite Element Analysis
,”
Acta Mater.
,
57
(
3
), pp.
695
707
.
59.
Rice
,
J. R.
,
1968
, “
A Path Independent Integral and Approximate Analysis of Strain Concentration by Notches and Cracks
,”
ASME J. Appl. Mech.
,
35
(
2
), pp.
379
386.
60.
Haftbaradaran
,
H.
, and
Qu
,
J. M.
,
2014
, “
A Path-Independent Integral for Fracture of Solids Under Combined Electrochemical and Mechanical Loadings
,”
J. Mech. Phys. Solids
,
71
, pp.
1
14
.
61.
Xia
,
S.
,
Gao
,
Y.
,
Bower
,
A. F.
,
Lev
,
L. C.
, and
Cheng
,
Y.-T.
,
2007
, “
Delamination Mechanism Maps for a Strong Elastic Coating on an Elastic–Plastic Substrate Subjected to Contact Loading
,”
Int. J. Solids Struct.
,
44
(
11
), pp.
3685
3699
.
62.
Gao
,
Y. F.
, and
Bower
,
A. F.
,
2004
, “
A Simple Technique for Avoiding Convergence Problems in Finite Element Simulations of Crack Nucleation and Growth on Cohesive Interfaces
,”
Modell. Simul. Mater. Sci. Eng.
,
12
(
3
), pp.
453
463
.
63.
Robertson
,
S. W.
, and
Ritchie
,
R. O.
,
2007
, “
In Vitro Fatigue-Crack Growth and Fracture Toughness Behavior of Thin-Walled Superelastic Nitinol Tube for Endovascular Stents: A Basis for Defining the Effect of Crack-Like Defects
,”
Biomaterials
,
28
(
4
), pp.
700
709
.
64.
Evans
,
A.
, and
Heuer
,
A.
,
1980
, “
Transformation Toughening in Ceramics: Martensitic Transformations in Crack-Tip Stress Fields
,”
J. Am. Ceram. Soc.
,
63
(
5–6
), pp.
241
248
.
65.
Budiansky
,
B.
,
Hutchinson
,
J.
, and
Lambropoulos
,
J.
,
1983
, “
Continuum Theory of Dilatant Transformation Toughening in Ceramics
,”
Int. J. Solids Struct.
,
19
(
4
), pp.
337
355
.
66.
Wang
,
X. J.
,
Fan
,
F. F.
,
Wang
,
J. W.
,
Wang
,
H. R.
,
Tao
,
S. Y.
,
Yang
,
A.
,
Liu
,
Y.
, et al
,
2015
, “
High Damage Tolerance of Electrochemically Lithiated Silicon
,”
Nat. Commun.
,
6
, p.
8417
.
67.
Wang
,
X.
,
Yang
,
A.
, and
Xia
,
S.
,
2016
, “
Fracture Toughness Characterization of Lithiated Germanium as an Anode Material for Lithium-Ion Batteries
,”
J. Electrochem. Soc.
,
163
(
2
), pp.
A90
A95
.
68.
Morris
,
D. J.
, and
Cook
,
R. F.
,
2008
, “
Indentation Fracture of Low-Dielectric Constant Films: Part II. Indentation Fracture Mechanics Model
,”
J. Mater. Res.
,
23
(
9
), pp.
2443
2457
.
69.
Morris
,
D. J.
, and
Cook
,
R. F.
,
2008
, “
Indentation Fracture of Low-Dielectric Constant Films: Part I. Experiments and Observations
,”
J. Mater. Res.
,
23
(
9
), pp.
2429
2442
.
70.
Wang
,
X.
,
Pan
,
Z.
,
Fan
,
F.
,
Wang
,
J.
,
Liu
,
Y.
,
Mao
,
S. X.
,
Zhu
,
T.
, and
Xia
,
S.
,
2015
, “
Nanoscale Deformation Analysis With High-Resolution Transmission Electron Microscopy and Digital Image Correlation
,”
ASME J. Appl. Mech.
,
82
(
12
), p.
121001
.
71.
Wang
,
J. W.
,
He
,
Y.
,
Fan
,
F. F.
,
Liu
,
X. H.
,
Xia
,
S. M.
,
Liu
,
Y.
,
Harris
,
C. T.
, et al
,
2013
, “
Two-Phase Electrochemical Lithiation in Amorphous Silicon
,”
Nano Lett.
,
13
(
2
), pp.
709
715
.
72.
Liu
,
X. H.
,
Wang
,
J. W.
,
Huang
,
S.
,
Fan
,
F. F.
,
Huang
,
X.
,
Liu
,
Y.
,
Krylyuk
,
S.
, et al
,
2012
, “
In Situ Atomic-Scale Imaging of Electrochemical Lithiation in Silicon
,”
Nat. Nanotechnol.
,
7
(
11
), pp.
749
756
.
73.
Gao
,
F.
, and
Hong
,
W.
,
2016
, “
Phase-Field Model for the Two-Phase Lithiation of Silicon
,”
J. Mech. Phys. Solids
,
94
, pp.
18
32
.
74.
Liu
,
X. H.
,
Zhong
,
L.
,
Huang
,
S.
,
Mao
,
S. X.
,
Zhu
,
T.
, and
Huang
,
J. Y.
,
2012
, “
Size-Dependent Fracture of Silicon Nanoparticles During Lithiation
,”
ACS Nano
,
6
(
2
), pp.
1522
1531
.
75.
Liang
,
W. T.
,
Yang
,
H.
,
Fan
,
F. F.
,
Liu
,
Y.
,
Liu
,
X. H.
,
Huang
,
J. Y.
,
Zhu
,
T.
, and
Zhang
,
S. L.
,
2013
, “
Tough Germanium Nanoparticles Under Electrochemical Cycling
,”
ACS Nano
,
7
(
4
), pp.
3427
3433
.