You are on page 1of 9

Journal of

Materials Chemistry A
View Article Online
PAPER View Journal | View Issue

Modeling of internal mechanical failure of all-solid-


Cite this: J. Mater. Chem. A, 2017, 5,
state batteries during electrochemical cycling, and
Published on 23 August 2017. Downloaded by University of Kansas on 9/3/2022 4:27:56 PM.

19422 implications for battery design


Giovanna Bucci, Tushar Swamy, Yet-Ming Chiang and W. Craig Carter *

This is the first quantitative analysis of mechanical reliability of all-solid state batteries. Mechanical
degradation of the solid electrolyte (SE) is caused by intercalation-induced expansion of the electrode
particles, within the constrains of a dense microstructure. A coupled electro-chemo-mechanical model
was implemented to quantify the material properties that cause an SE to fracture. The treatment of
microstructural details is essential to the understanding of stress-localization phenomena and fracture. A
cohesive zone model is employed to simulate the evolution of damage. In the numerical tests, fracture is
prevented when electrode-particle's expansion is lower than 7.5% (typical for most Li-intercalating
compounds) and the solid-electrolyte's fracture energy higher than Gc ¼ 4 J m2. Perhaps counter-
intuitively, the analyses show that compliant solid electrolytes (with Young's modulus in the order of ESE
Received 12th April 2017
Accepted 18th August 2017
¼ 15 GPa) are more prone to micro-cracking. This result, captured by our non-linear kinematics model,
contradicts the speculation that sulfide SEs are more suitable for the design of bulk-type batteries than
DOI: 10.1039/c7ta03199h
oxide SEs. Mechanical degradation is linked to the battery power-density. Fracture in solid Li-ion
rsc.li/materials-a conductors represents a barrier for Li transport, and accelerates the decay of rate performance.

been shown to penetrate even dense ceramic solid electrolytes if


1 Introduction the SE-surface contains aws.9 Predicting whether micro-cracks
Li-ion batteries that use solid electrolyte materials (SEs) in place form during battery operation is, therefore, critical towards pre-
of traditional liquid electrolytes could achieve high energy venting cell shorting.
density while avoiding safety issues surrounding liquid electro- We employ a fully coupled electro-chemo-mechanical model
lyte ammability. Solid electrolytes with conductivity approach- to investigate fracture mechanisms in composite solid-state
ing that of liquid electrolyte have recently been discovered.1–7 electrodes. Treatment of microstructural details and local vari-
Despite fast growing interest in all-solid-state batteries (ASSBs), ability10 enables the study of stress localization caused by
many challenges remain in both manufacturing and reliability of particle misalignment and non-smooth features. A cohesive
the technology. zone model (CZM) is employed to simulate the evolution of
In ASSBs, the solid-electrolyte is responsible for binding the damage.11 The detrimental effect of fracture on Li-ion ux is
active material and establishing conductive paths for Li ions. also taken into account by the CZM. We quantify the conditions
However, the formation of micro-cracks within the solid elec- under which fracture occurs, caused by the chemical expansion
trolyte is expected to reduce its effective ionic conductivity. of electrode particles. Fracture is the result of regions of shear
Additionally, low porosity solid-state systems are expected to be and tensile stresses formed during electrochemical cycling.
more prone to mechanical degradation if not designed to The role of intercalation-induced stress (also called Vegard's
accommodate intercalation-induced deformations. As fracture stress) on the mechanical failure of electrode particles has been
degrades the microstructure, paths for lithium diffusion become previously studied.18–51 Among these studies, only Bower and
more tortuous and the battery power-density decreases. Micro- Guduru20 employed a fully coupled chemo-mechanical model to
scale defects and inhomogeneities in battery microstructures simulate fracture of a simplied electrode microstructure. To
would interfere with Li transport and accelerate the decay of our knowledge, ours is the rst model to quantitatively assess
battery performance. Furthermore, at sufficiently high current mechanical reliability of all-solid state batteries, and predict the
densities, micro-cracks may provide a pathway for Li dendrite extension of fracture caused by electrochemical cycling.
growth, eventually causing the cell to short.8 Li-protrusions have The mechanical properties of solid electrolyte materials have
not received much attention and very limited experimental
chemo-mechanical properties are available. Measurements
Massachusetts Institute of Technology, Department of Materials Science and
Engineering, 77 Massachusetts Avenue, Cambridge, MA 02139-4307, USA. E-mail:
collected in Table 1 reveal a wide range of values for Young's
bucci@mit.edu modulus. In particular, sulde SEs tend to be much more

19422 | J. Mater. Chem. A, 2017, 5, 19422–19430 This journal is © The Royal Society of Chemistry 2017
Published on 23 August 2017. Downloaded by University of Kansas on 9/3/2022 4:27:56 PM.

Paper

Table 1 Mechanical properties of solid electrolyte materials

Fracture Conductivity (room


Compound Processing method Young's modulus toughness Testing method Reference temperature)

LiPON LixPOyNz Amorphous LiPON 77 GPa Nanoindentation 12 2  106 S cm1

This journal is © The Royal Society of Chemistry 2017


lms magnetron
sputtered
Perovskite Li0.33La0.57TiO3 – solid Hot-pressing at 1000  C, 186  4 0.890–1.34 MPa Nanoindentation 13 Bulk: 103 S cm1
state relative density >95% m0.5
Li0.33La0.57TiO3 – sol gel 200  3 GPa 0.890–1.31 MPa Nanoindentation Total: 105 S cm1
m0.5
Garnet Li6.24La3Zr2Al0.24O11.98 Hot pressed 150 GPa (porosity ¼ Resonant 14 0.2  103 S cm1
(LLZO) 0.03); 132.5 GPa ultrasound
(porosity ¼ 0.06) spectroscopy
Cubic Li7La3Zr2O12 Relative density of 9% 150 GPa Resonant 3  104 S cm1
ultrasound
spectroscopy
Sulde Li2S–P2S5 – hot pressed Sintering at 360 MPa 18–25 GPa Ultrasound velocity 15 3  104 S cm1 (for
and temperatures 20– measurement and fully dense material)
190  C compression test
Li2S–P2S5 – cold pressed Sintering at 180– 14–17 GPa
360 MPa and room
temperature
Li2S–P2S5 18.5  0.9 GPa 0.23  0.04 MPa Nanoindentation 16
m0.5
Li10GeP2S12 37.19 GPa Atomistic 17 1.2  102 S cm1
simulation
View Article Online

J. Mater. Chem. A, 2017, 5, 19422–19430 | 19423


Journal of Materials Chemistry A
View Article Online

Journal of Materials Chemistry A Paper

compliant than oxide electrolytes. The Young's modulus of We believe the assumption of a perfectly coherent and stable
Li2S–P2S5 sulde solid electrolytes has been estimated to be in interface to be appropriate for the scope of this study.
the range of 14–25 GPa.15,16 In the following section we illustrate the model and discuss
Such a low stiffness has been regarded as favorable for the the results in detail.
design of bulk-type batteries.52 However, we show that
compliant solid electrolytes (with Young's moduli on the order
of ESE ¼ 15 GPa) are more prone to micro-cracking. Solid elec- 2 Modeling of fracture in all-solid-
trolytes deform by stretching and shearing in response to the state battery electrodes
particles' volume change. The nonlinear formulation of the
mechanical equilibrium quanties the difference in deforma- Methods
Published on 23 August 2017. Downloaded by University of Kansas on 9/3/2022 4:27:56 PM.

tion and stress patterns associated with varying the SE's stiff- In one common design of ASSBs, the positive and negative
ness. A linear model, which would predict that stress scales with electrodes are composites of active electrode-particles
Young's modulus, would not capture the microstructural effects embedded in a solid-matrix admixture of ionic and electronic
that we describe below. conducting materials. Negative electrodes are, in most cases,
We compare the evolution of damage for several values of produced and assembled in the delithiated state. During the
electrolyte's fracture energy (Gc ¼ 0.25–4.0 J m2) and volume rst charge, the anode particles tend to expand as they inter-
expansion of the active material (7.5%, 15% and 30%). A calate Li (experimentally measured Vegard's strains for many Li-
cohesive model postulates that fracture energy is released storage compounds are summarized in Tables 1.1 and 1.2 of
gradually as the crack opens. The CZM differs from the Griffith Woodford's thesis,59 and in Table 1 of Mukhopadhyay & Shel-
model wherein energy is released instantaneously. The gradual don60). Constrained by the surrounding SE matrix, the particles
release presumes some cohesion between the separating anks will be in compression. The mechanical stress and degradation
of a crack. Generally, the traction decays with increasing sepa- caused by swelling of electrode particles is modeled as follows.
ration until it vanishes at a critical opening displacement. The A nite element (FE) code was implemented according to the
fracture energy represents the integral of the traction–separa- theoretical continuum model described by Bucci et al.10 A
tion curve and it is treated as a model-parameter. In our anal- representative microstructural arrangement of a distribution of
yses, cracking is prevented only in those cases for which the particle sizes and its FE discretization with linear quadrilateral
electrolyte's fracture energy Gc $ 4.0 J m2 and the particles' elements are represented in Fig. 1. The FE mesh is representative
total volumetric expansion is DV # 7.5%. In all the other cases, of a portion of the composite negative electrode (square high-
the model predicts some extension of mechanical degradation. lighted in Fig. 1). The section extends from the current collector
Recent studies have analyzed the properties of the interface to the separator. The direction parallel to the Li ux is marked
between solid electrolytes and electrode materials.53–58 In with y in Fig. 1. Because of the large computational cost of a full
particular, Zhu et al.53 have identied voltage-stability ranges for 3D analysis, the system is modeled in 2D under the assumption
many SE materials and predicted decomposition products of of plane strain. According to the 2D plane strain model, the
the interface reaction. The formation of an interface layer is electrode particles are allowed to expand in the xy plane, but not
expected to affect lithium transfer kinetics. Thick interface in the z direction. Plane strain is typically employed for a thin
layers may also modify the local mechanical response. An plate embedded in a thicker sample. This is a realistic assump-
interface layer that is stiffer than the bulk solid electrolyte is tion for a typical bulk-type electrode. The 2D model is expected to
expected to provide better fracture resistance. (A coherent shell correctly capture trends in stress and fracture.
surrounding the particle would constrain the particle's expan- The grid is managed with the deal.II nite-element
sion and reduce shearing and fracturing of the SE matrix.) library.61,62 The microstructure includes 36 randomly oriented
For thin SEI layers or large intercalation-induced deforma- square particles in a region of dimensions 11 mm  11 mm. The
tions (as in the case of Si, Al and Sn), we expect the interface particle positions and size distribution follow from a centroidal
layer to undergo large tensile stress and eventually fracture. For Voronoi tessellation. The average particle size is 1 mm. The area
instance, Wenzel et al. measured a stable 2–3 nm thick SEI ratio of active material is about 50%—a typical volume ratio for
between Li7P3S11 and Li metal.55 The decomposition products commercial Li-ion batteries is about 50–60%. We consider
have larger Young's modulus than Li2S–P2S5 glass ceramics. shapes with sharp corners (the squares chosen here) to be
However, the layer is not thick enough to constrain the particle's a more realistic representation of particles than circles—see for
expansion, particularly on the rst cycles when most micro- instance Fig. 6 of Sakuda et al.52 and 3D image reconstruction of
cracks form. LiCoO2 particles of Harris et al.63 Flaws and stresses are more
In general, the thickness of the interface layer depends on its likely to accumulate near sharp corners.
conducting properties. Mixed conducting layers tends to grow At a given time step, the electrochemical-mechanical problem
thicker than electronically insulating layers. Thick interface is solved employing a Newton–Raphson iterative algorithm. At
layers are undesirable, because they degrade the charge transfer each iteration, displacements, Li concentration and diffusion
kinetics (e.g., in the case of Li10GeP2S12 reacting with Li metal56). potential are calculated as the solution to three equations that
Further studies are needed to characterize the interfacial couple the electro-, chemo-, and mechanical elds. We model the
properties for a variety of electrode–electrolyte combinations. matrix as a homogenized admixture of ionic and electronic
conductors. The constitutive behavior for the electrode and the

19424 | J. Mater. Chem. A, 2017, 5, 19422–19430 This journal is © The Royal Society of Chemistry 2017
View Article Online

Paper Journal of Materials Chemistry A


Published on 23 August 2017. Downloaded by University of Kansas on 9/3/2022 4:27:56 PM.

Fig. 1 Geometry, discretization and boundary conditions of a finite element model of a composite electrode. Electrode particles are embedded
in a mixed conductor, consisting of a solid electrolyte and an electronically conductive additive.

electrolyte material is assumed to be elastically and diffusively


isotropic. Materials are assumed to have a linear elastic consti- Table 2 Material parameters for problems in Section 2
tutive behavior. The solid electrolyte material is considered to
have zero Vegard's strain. Input parameters for this problem are Input value Description
summarized in Table 2 (the variables listed appear in the equa- F ¼ 96 485.3365 C mol1 Faraday's constant
tions discussed in Bucci et al.10). R ¼ 8.314 J K1 mol1 Gas constant
The CZM of fracture is based on an intrinsic history- T ¼ 298 K Temperature
dependent constitutive behavior.11 The ux across the interface MEl ¼ 1015 m2 s1 Mobility of Li in the
is irreversibly set to zero at the onset of fracture. The “intrinsic” electrode material
MSE ¼ 1013 m2 s1 Mobility of Li in the solid
CZM approach is based on the pre-insertion of cohesive elements electrolyte material
along potential crack paths. Therefore, fracture is allowed to cmaxAM ¼ 1 Maximum relative number of moles
propagate along a subset of nite element interfaces. Those of Li per mole of electrode compound
interfaces lie along crack patterns that run between particles. The cmaxSE ¼ 0.25 Maximum relative number of moles of
pre-insertion of cohesive elements along potential crack paths Li per mole of solid electrolyte
i ¼ 10 A m2 Maximum relative number of moles
restrict the propagation of fracture. An alternative approach, of Li per mole of solid electrolyte
called “extrinsic”, is based on the insertion of cohesive elements gLi ¼ 1 Activity coefficient
on the y, only at the interfaces where the fracture criterion is n ¼ 0.3 Poisson's ratio for both materials
met. However, this method requires changes in the mesh bAM ¼ 0.1 Relative lattice constant for Li in
topology, and it is not suitable for parallel computing. We electrode material
bSE ¼ 0 Relative lattice constant for Li in
assume that the placement of the CZ elements in regions of high the solid electrolyte material
shear would not differ from the extrinsic approach. EAM ¼ 100 GPa Young's modulus of the electrode material
Galvanostatic tests are performed by applying a constant and ESE ¼ 15 GPa Young's modulus of the solid
uniform lithium ux (corresponding to a constant current electrolyte material
density) at the separator interface (top edge in Fig. 1). The spatial Gc ¼ 1.0 J m2 Fracture energy of the bulk solid
electrolyte material
variability of the current density at the separator depends on d0 ¼ 5 nm Opening displacement at the
specic features of the electrochemical cell, such as electrode onset of damage
thickness and tortuosity. Polarization effects in proximity of the dcr ¼ 20d0 Critical opening displacement
separator could be modeled by treating both sides of the elec- (complete interface separation)
trochemical cell. A zero ux is assumed on the remaining edges.

This journal is © The Royal Society of Chemistry 2017 J. Mater. Chem. A, 2017, 5, 19422–19430 | 19425
View Article Online

Journal of Materials Chemistry A Paper

For the mechanical problem, zero horizontal displacement


Dirichlet boundary conditions are applied on the le and right
boundaries. The displacement is considered fully constrained (in
the vertical direction) on the top edge by the presence of the SE,
and at the bottom edge by packaging or neighbor cells. We expect
the volume change of the active material to be accommodated by
the deformation of the SE matrix in ASSBs. This is similar to what
Harris et al.63 observed (by digital image correlation techniques)
in graphite electrodes, where most of graphite's swelling was
compensated by reducing the electrode porosity. Harris et al.
Published on 23 August 2017. Downloaded by University of Kansas on 9/3/2022 4:27:56 PM.

showed that the average strain of a graphite electrode was only


about 0.2% during lithiation, an order of magnitude smaller than
graphite's chemical expansion.

3 Results
For the baseline example, we choose a solid electrolyte material,
having Young's modulus ESE ¼ 15 GPa and bulk fracture energy
Gc ¼ 1.0 J m2. These are representative values for a sulde SE
material. In order to reveal behavior over a wide range of
Vegard's expansion, the intercalation compound was allowed to
have up to 30% volumetric expansion at full lithiation. Such
large swelling may arise from the concomitant intercalation of
Li and other species and from a growing interface layer. Our
simulation shows that fracture initiates when particles have
changed their volume by only 3%, a value that encompasses the
behavior of many intercalation compounds.
A sequence of snapshots in Fig. 2 illustrate the state of
charge (le column), the hydrostatic Cauchy stress (right
column). Cracks propagating within the solid electrolyte mate-
rial are represented as black lines. Thickening of the black lines
represents progressive interface separation and accumulation
of damage.
On average, compressive stress arises as a consequence of
the particles' chemical expansion because the entire system is
constrained by the surrounding material. The simulation shows
small regions of tensile stress developing in the area near the
particles' corners (rust-colored areas in the contour plots of
Fig. 2). The misalignment of these corners creates matrix shear-
and tensile-stresses. Fig. 2 Evolution of damage in the solid electrolyte material at subse-
quent states of charge. With increasing Li content, the active material
Each snapshot in Fig. 2 represents the lithiation and stress
undergoes chemical expansion. As the particles' stress-free strain
state at subsequent states of charge. For all the numerical tests, increases, compressive stress develops in most of the microstructure.
the current density is held constant at 1 mA cm2 (a typical However, a few regions in proximity of the particles corner are under
current density for commercial Li-ion batteries). Time is indic- tensile stress (rust-colored regions in the right contour plots above). This
ative of state of charge, unless the evolution of stress varies tension grows large enough to initiate fracture in the solid electrolyte
matrix, this phenomenon is captured in the simulation by the cohesive
signicantly among tests. This a consequence of the electro-
elements. Cracks (marked with black lines) propagate from corner to
chemo-mechanical coupling.64 corner, cutting off diffusion paths for Li within the electrolyte.
As the state of charge progresses, the pressure in the parti-
cles increases. In Fig. 2c, the compressive stress in the active
We performed a series of numerical tests by varying the frac-
material is higher than 1 GPa (for particles that have stored
ture energy of the SE material in the range Gc ¼ 0.25–4.0 J m2. To
approximately 50% of their total Li capacity). The stress–strain
our knowledge, the only experimental data on the fracture
curve measured by Sakuda et al.15 shows a linear-elastic
toughness of a sulde SE material is reported by McGrogan et al.16
behavior for Li2S–P2S5 in compression. Our simulations
McGrogan and coauthors measured the toughness KIc ¼ 0.23 
predict the compressive stress in the electrolyte to lie within the
0.04 MPa m0.5 via nano-indentation of a glassy Li2S–P2S5 sample.
linear elastic range of 0–200 MPa, similar to those measured by
Such a fracture toughness corresponds to the fracture energy Gc ¼
Sakuda et al.

19426 | J. Mater. Chem. A, 2017, 5, 19422–19430 This journal is © The Royal Society of Chemistry 2017
View Article Online

Paper Journal of Materials Chemistry A


Published on 23 August 2017. Downloaded by University of Kansas on 9/3/2022 4:27:56 PM.

Fig. 3 The extension of micro-cracks within the solid electrolyte Fig. 4 The extension of crack (normalized with respect to the elec-
material has been computed by a mesoscale finite element model. The trode thickness) is plotted for three cases characterized by different
curves represent five cases of SE fracture energies in the range Gc ¼ Young's moduli of the solid electrolyte material, i.e., ESE ¼ 15, 25, 50,
0.25–4.0 J m2. The propagation rate (slope of the curve in stage b) 150 GPa. Other input parameters are the same of the baseline case
increases with decreasing fracture energy. (see Table 2), and the Young's modulus of the active material remain
fixed at ESE ¼ 100 GPa. As the stiffness of the SE increases, a lower
velocity of fracture propagation is predicted. A stiffer SE tends to
2.8  1.8 J m2, given the Young's modulus ESE ¼ 18.5  0.9 GPa contain the chemical expansion of the active material. Shearing and
measured by McGrogan et al. tension—responsible for crack growth—are less likely to arise in a less
The relative crack length (extension of fracture normalized deformable material.
with respect to electrode thickness) is illustrated in Fig. 3 as it
evolves with respect to time. For each curve in Fig. 3 it is
possible to identify three stages: (a) onset of fracture, (b) displacements give rise to tensile and shear stresses which are
approximately constant propagation rate, and (c) decreasing particularly large in the case of compliant SEs.
propagation rate up to saturation. As expected, crack nucleation In order to identify conditions that prevent mechanical
is delayed in tougher materials. The propagation rate (slope of degradation, we consider active materials with lower volume
the curve in stage b) increases with decreasing fracture energy. change associated with changes in Li stoichiometry. Further-
In all the examples fracture propagates in a stable fashion more, we raised the solid electrolytes fracture energy up to Gc ¼
(rather than sudden failure). A plateau in the curves of Fig. 3 4.0 J m2. The model predicts that fracture is suppressed, for
indicates crack-growth saturation. In the cases with lower active materials with 7.5% Vegard's expansion and solid elec-
fracture energy, stage c may be biased by the availability of crack trolytes with the fracture energy Gc ¼ 4.0 J m2 and Young's
patterns—even if about 10% of the pre-inserted cohesive modulus ESE ¼ 15 GPa. Thus, mechanical damage is predomi-
interfaces remain unfractured. The pre-insertion of cohesive nantly dependent on SE fracture properties, as most Li-storage
elements in specied locations is a possible shortcoming of this compounds have Vegard's expansion below 7.5%.59,60
model. As shown in Fig. 3, solid electrolytes with fracture energy Results from several tests are collected in Fig. 5, where we
up to 4 J m2 are predicted to fracture when electrode particles observe the overlapping of two groups of curves. These curves
undergo 30% of intercalation-induced swelling. are marked with a and b in Fig. 5, and represent the tests with
We explore the dependence of the predicted damage on the SE fracture energy Gc ¼ 1.0 J m2 and Gc ¼ 0.25 J m2. The
electrolyte elastic properties. We consider electrolytes with results can be explained by referring to the system's energy
Young's moduli ESE ¼ 25 GPa and ESE ¼ 50 GPa and ESE ¼ balance. As the particle's volume increases upon lithiation, the
150 GPa in addition to the baseline case (ESE ¼ 15 GPa). The solid electrolyte volume is forced to shrink because the elec-
Young's modulus ESE ¼ 150 GPa is representative of a garnet trode's volume-averaged deformation is zero. Hydrostatic pres-
solid electrolyte material (see Table 1). The results in Fig. 4 sure developing in the electrode and electrolyte materials scales
illustrate the inverse relationship between the velocity of crack linearly with volume change. The pressure is also proportional
propagation and the electrolyte's stiffness. A more compliant to the materials' bulk moduli. It follows that the elastic energy
solid electrolyte tends to deform more by stretching and stored in the electrolyte becomes four times larger when the
shearing in response to the particles' volume change. Regions of active material's expansion is doubled. If the fracture energy
tensile-stress form in the SE matrix where fracture is promoted. required to open new cracks is also four times larger, fracture
An electrolyte with stiffness closer to that of the active material propagation rate remains the same. In order to characterize the
(here the Young's modulus of the electrode material is set to dependence on the total energy, both stored elastic and fracture
EAM ¼ 100 GPa) tends to develop higher compressive stress, but energy, we dene a new dimensionless parameter for all-solid-
undergoes lower tension. If the problem is considered from state electrodes: G ¼ 0.5kSE(3bAMAAM)2/(HGc). In the denition
a linear elasticity perspective, these results may seem counter- of G, kSE is the SE bulk modulus, bAM the Vegard's parameter
intuitive. However in the non-linear formulation, the large (3bAM is the volumetric expansion rate) of the active material, H

This journal is © The Royal Society of Chemistry 2017 J. Mater. Chem. A, 2017, 5, 19422–19430 | 19427
View Article Online

Journal of Materials Chemistry A Paper

and particle asperities are sufficient to cause tensile and shear


stress in the solid electrolyte matrix.
The simulations predict fracture to propagate in a stable
fashion (rather than abruptly). As expected, crack nucleation is
delayed in tougher materials. The propagation rate and the nal
extension of cracks also decrease with increasing electrolyte
fracture energy.
Perhaps counter-intuitively, the analyses show that
compliant solid electrolytes (with Young's modulus in the order
of ESE ¼ 15 GPa) are more prone to micro-cracking. Shearing
Published on 23 August 2017. Downloaded by University of Kansas on 9/3/2022 4:27:56 PM.

and stress arise if the particles have surface asperities or the


stress-elds of nearby particles interact. Because compliant SEs
allow for large deformations they are more likely to develop
Fig. 5 Various combinations of Vegard's parameters of the active
localized tension and fracture. A non-linear kinematics model is
material and SE fracture energy are analyzed. The overlap of two sets required to predict this effect which contradicts the speculation
of results is observed. They correspond to the two cases marked as that sulde SEs are more suitable for the design of bulk-type
a ((a) 30% volume change of the active material and SE fracture energy batteries than oxide SEs.52
Gc ¼ 1.0 J m2, and (a) 15% volume change of the active material and To our knowledge, this work is the rst to investigate
SE fracture energy Gc ¼ 0.25 J m2), and other two cases, marked as
b ((b) 15% volume change of the active material and SE fracture energy
mechanical reliability of all-solid state batteries. The results
Gc ¼ 1.0 J m2, and (b) 7.5% volume change of the active material and presented have implications for the battery power-density.
SE fracture energy Gc ¼ 0.25 J m2). The interpretation of this Fracture in solid Li-ion conductors represents a barrier for Li
outcome is based on the system's energy balance. As the particles transport and accelerate the decay of rate performance.
volume increases upon lithiation, the solid electrolyte volume is forced Reliability of ASSBs will depend on the elastic energy associated
to shrink. Here we assume the volume-averaged deformation of the
entire region to be zero. A less restrictive assumption would allow for
with intercalation-induced strain, the solid-electrolyte fracture
the electrode's thickness to evolve with state of charge. In linear energy and the geometry of the microstructure. Therefore,
elasticity, the hydrostatic pressure developing in both materials scales a simple design rule can be based on the dimensionless parameter
linearly with their volume change and in proportion to their bulk G ¼ 0.5kSE(3bAMAAM)2/(HGc) representing the ratio between elastic
modulus. It follows that the elastic energy stored in the SE material and fracture energies. We predict the integrity of elastic-
become four times larger, when the particles' expansion is doubled. If
the fracture energy required to open new cracks is also four times
brittle solid-state electrolytes to be preserved when the condition
larger, fracture propagates at the same rate. G < 1000 is met.

Conflicts of interest
is the electrode's thickness, and AAM is the area of active
material. The two overlapping curves for the case a and the case There are no conicts of interest to declare.
b in Fig. 5 are characterized by the same value of G. Therefore,
the dimensionless parameter G can be used to generalize the
results presented here. Acknowledgements
The authors are very grateful to Brian W. Sheldon and Frank
McGrogan for their valuable comments. The work was sup-
4 Conclusions ported by the grant DE-SC0002633 funded by the U.S. Depart-
Electro-chemo-mechanical FEM simulations capture the onset ment of Energy, Office of Science.
and propagation of damage in a solid-state composite electrode.
Fracture is prevented if electrode-particle's expansion is lower References
than 7.5% and the solid-electrolyte's fracture energy higher than
Gc ¼ 4 J m2 (under the assumption of SE Young's modulus ESE ¼ 1 K. Takada, Progress and prospective of solid-state lithium
15 GPa). This condition restricts the choice of electrolyte based batteries, Acta Mater., 2013, 61(3), 759–770.
on its fracture properties, while most intercalation oxides have 2 J. Li, C. Ma, M. Chi, C. Liang and N. J. Dudney, Solid
volume expansion below 7.5%.59 We refer here to the average electrolyte: The key for high-voltage lithium batteries, Adv.
volume change of a poly-crystalline material—Vegard's parame- Energy Mater., 2015, 5(4), 1–6.
ters can be largely anisotropic, this is for instance the case of 3 J. G. Kim, B. Son, S. Mukherjee, N. Schuppert, A. Bates,
graphite. Intercalation-induced expansion of the active material O. Kwon, M. J. Choi, H. Y. Chung and S. Park, A review of
is constrained in dense solid-state electrodes and electrolytes are lithium and non-lithium based solid state batteries, J.
prone to mechanical degradation. The particles and SE together Power Sources, 2015, 282, 299–322.
create a microstructure—the shape of particles and their prox- 4 V. Thangadurai, S. Narayanan and D. Pinzaru, Garnet-type
imity within the microstructure determine fracture. Microstruc- solid-state fast Li ion conductors for Li batteries: critical
tural inhomogeneities, such as particle-to-particle misalignment review, Chem. Soc. Rev., 2014, 43, 4714–4727.

19428 | J. Mater. Chem. A, 2017, 5, 19422–19430 This journal is © The Royal Society of Chemistry 2017
View Article Online

Paper Journal of Materials Chemistry A

5 E. Quartarone and P. Mustarelli, Electrolytes for solid-state Implications on the critical size for aw tolerant battery
lithium rechargeable batteries: recent advances and electrodes, Int. J. Solids Struct., 2010, 47(10), 1424–1434.
perspectives, Chem. Soc. Rev., 2011, 40, 2525–2540. 20 A. F. Bower and P. R. Guduru, A simple nite element model
6 K. Takada, N. Ohta and Y. Tateyama, Recent Progress in of diffusion, nite deformation, plasticity and fracture in
Interfacial Nanoarchitectonics in Solid-State Batteries, J. lithium ion insertion electrode materials, Modelling and
Inorg. Organomet. Polym. Mater., 2014, 205–213. Simulation in Materials Science and Engineering, 2012, 20(4),
7 M. Tatsumisago, M. Nagao and A. Hayashi, Recent 045004.
development of sulde solid electrolytes and interfacial 21 S. Renganathan, G. Sikha, S. Santhanagopalan and
modication for all-solid-state rechargeable lithium R. E. White, Theoretical analysis of stresses in a lithium
batteries, Journal of Asian Ceramic Societies, 2013, 1, 17–25. ion cell, J. Electrochem. Soc., 2010, 157(2), A155–A163.
Published on 23 August 2017. Downloaded by University of Kansas on 9/3/2022 4:27:56 PM.

8 Y. Suzuki, K. Kami, K. Watanabe, A. Watanabe, N. Saito, 22 R. E. Garcı́a, Y.-M. Chiang, W. Craig Carter, P. Limthongkul
T. Ohnishi, K. Takada, R. Sudo and N. Imanishi, and C. M. Bishop, Microstructural modeling and design of
Transparent cubic garnet-type solid electrolyte of Al2O3- rechargeable lithium-ion batteries, J. Electrochem. Soc.,
doped Li7La3Zr2O12, Solid State Ionics, 2015, 278, 172–176. 2005, 152(1), A255–A263.
9 L. Porz, T. Swamy, B. W. Sheldon, D. Rettenwander, 23 C.-W. Wang and A. M. Sastry, Mesoscale modeling of a Li-ion
T. Frömling, H. L. Thaman, S. Berendts, R. Uecker, polymer cell, J. Electrochem. Soc., 2007, 154(11), A1035–
W. C. Carter and Y.-M. Chiang, Mechanism of lithium A1047.
metal penetration through inorganic solid electrolytes, Adv. 24 S. Golmon, K. Maute and M. L. Dunn, Numerical modeling
Energy Mater., 2017, 1701003. of electrochemical–mechanical interactions in lithium
10 G. Bucci, Y.-M. Chiang and W. Carter, Formulation of the polymer batteries, Comput. Struct., 2009, 87(23–24), 1567–
coupled electrochemical–mechanical boundary-value 1579.
problem, with applications to transport of multiple 25 M. Zhu, J. Park and A. M. Sastry, Fracture analysis of the
charged species, Acta Mater., 2016, 62, 33–51. cathode in Li-ion batteries: A simulation study, J.
11 G. Bucci and W. Carter, Mechanics of Materials. Micro- Electrochem. Soc., 2012, 159(4), A492–A498.
mechanics in electrochemical systems (in production), 26 R. T. Purkayastha and R. M. McMeeking, An integrated 2-
Springer, 2016. d model of a lithium ion battery: the effect of material
12 E. G. Herbert, W. E. Tenhaeff, N. J. Dudney and G. M. Pharr, parameters and morphology on storage particle stress,
Mechanical characterization of LiPON lms using Comput. Mech., 2012, 50(2), 209–227.
nanoindentation, Thin Solid Films, 2011, 520(1), 413–418. 27 C. Miehe, H. Dal, L.-M. Schänzel and A. Raina, A phase-eld
13 Y.-H. Cho, J. Wolfenstine, E. Rangasamy, H. Kim, H. Choe model for chemo-mechanical induced fracture in lithium-
and J. Sakamoto, Mechanical properties of the solid Li-ion ion battery electrode particles, Int. J. Numer. Meth. Eng.,
conducting electrolyte: Li0.33La0.57TiO3, J. Mater. Sci., 2012, 2015, 106(9), 683–711.
47, 5970–5977. 28 K. Aifantis and J. Dempsey, Stable crack growth in
14 J. E. Ni, E. D. Case, J. S. Sakamoto, E. Rangasamy and nanostructured Li-batteries, J. Power Sources, 2005, 143(12),
J. B. Wolfenstine, Room temperature elastic moduli and 203–211.
Vickers hardness of hot-pressed LLZO cubic garnet, J. 29 I. Ryu, J. W. Choi, Y. Cui and W. D. Nix, Size-dependent
Mater. Sci., 2012, 47, 7978–7985. fracture of Si nanowire battery anodes, J. Mech. Phys.
15 A. Sakuda, A. Hayashi and Y. Takigawa, Evaluation of elastic Solids, 2011, 59(9), 1717–1730.
modulus of Li2S–P2S5 glassy solid electrolyte by ultrasonic 30 K. Aifantis and S. Hackney, Mechanical stability for
sound velocity measurement and compression test, J. nanostructured Sn- and Si-based anodes, J. Power Sources,
Ceram. Soc. Jpn., 2013, 121(11), 946–949. 2011, 196(4), 2122–2127.
16 F. P. McGrogan, T. Swamy, S. R. Bishop, E. Eggleton, L. Porz, 31 S. Kalnaus, K. Rhodes and C. Daniel, A study of lithium ion
X. Chen, Y.-M. Chiang and K. J. V. Vliet, Compliant yet brittle intercalation induced fracture of silicon particles used as
mechanical behavior of Li2S–P2S5 lithium-ion conducting anode material in Li-ion battery, J. Power Sources, 2011,
solid electrolyte, Adv. Energy Mater., 2017, 1602011. 196(19), 8116–8124.
17 Z. Q. Wang, M. S. Wu, G. Liu, X. L. Lei, B. Xu and 32 Y. Xia, T. Wierzbicki, E. Sahraei and X. Zhang, Damage of
C. Y. Ouyang, Elastic properties of new solid state cells and battery packs due to ground impact, J. Power
electrolyte material Li10GeP2S12: A study from rst- Sources, 2014, 267, 78–97.
principles calculations, Int. J. Electrochem. Sci., 2014, 9(2), 33 X. Zhang and T. Wierzbicki, Characterization of plasticity
562–568. and fracture of shell casing of lithium-ion cylindrical
18 W. H. Woodford, Y.-M. Chiang and W. C. Carter, battery, J. Power Sources, 2015, 280, 47–56.
Electrochemical shock of intercalation electrodes: 34 F. Hao and D. Fang, Reducing diffusion-induced stresses of
a fracture mechanics analysis, J. Electrochem. Soc., 2010, electrode-collector bilayer in lithium-ion battery by pre-
157(10), A1052–A1059. strain, J. Power Sources, 2013, 242, 415–420.
19 T. K. Bhandakkar and H. Gao, Cohesive modeling of crack 35 H. B. Chew, B. Hou, X. Wang and S. Xia, Cracking
nucleation under diffusion induced stresses in a thin strip: mechanisms in lithiated silicon thin lm electrodes, Int. J.
Solids Struct., 2014, 51(23–24), 4176–4187.

This journal is © The Royal Society of Chemistry 2017 J. Mater. Chem. A, 2017, 5, 19422–19430 | 19429
View Article Online

Journal of Materials Chemistry A Paper

36 A. Drozdov, Constitutive equations for self-limiting 50 B. Dimitrijevic, K. Aifantis and K. Hackl, The inuence of
lithiation of electrode nanoparticles in Li-ion batteries, particle size and spacing on the fragmentation of
Mech. Res. Commun., 2014, 57, 67–73. nanocomposite anodes for Li batteries, J. Power Sources,
37 J. Ye, Y. An, T. Heo, M. Biener, R. Nikolic, M. Tang, H. Jiang 2012, 206, 343–348.
and Y. Wang, Enhanced lithiation and fracture behavior of 51 S. Lee, J. Yang and W. Lu, Debonding at the interface
silicon mesoscale pillars via atomic layer coatings and between active particles and {PVDF} binder in Li-ion
geometry design, J. Power Sources, 2014, 248, 447–456. batteries, Extreme Mech. Lett., 2016, 6, 37–44.
38 L. Greve and C. Fehrenbach, Mechanical testing and macro- 52 A. Sakuda, A. Hayashi and M. Tatsumisago, Sulde Solid
mechanical nite element simulation of the deformation, Electrolyte with Favorable Mechanical Property for All-
fracture, and short circuit initiation of cylindrical lithium Solid-State Lithium Battery, Sci. Rep., 2013, 3, 2261.
Published on 23 August 2017. Downloaded by University of Kansas on 9/3/2022 4:27:56 PM.

ion battery cells, J. Power Sources, 2012, 214, 377–385. 53 Y. Zhu, X. He and Y. Mo, First principles study on
39 Y. Wang, Y. He, R. Xiao, H. Li, K. Aifantis and X. Huang, electrochemical and chemical stability of solid electrolyte–
Investigation of crack patterns and cyclic performance of electrode interfaces in all-solid-state Li-ion batteries, J.
Ti–Si nanocomposite thin lm anodes for lithium ion Mater. Chem. A, 2016, 4, 3253–3266.
batteries, J. Power Sources, 2012, 202, 236–245. 54 W. D. Richards, L. J. Miara, Y. Wang, J. C. Kim and G. Ceder,
40 M. Pharr, Z. Suo and J. J. Vlassak, Variation of stress with Interface stability in solid-state batteries, Chem. Mater.,
charging rate due to strain-rate sensitivity of silicon 2016, 28(1), 266–273.
electrodes of Li-ion batteries, J. Power Sources, 2014, 270, 55 S. Wenzel, D. A. Weber, T. Leichtweiss, M. R. Busche, J. Sann
569–575. and J. Janek, Interphase formation and degradation of
41 J. K. Min, M. Stackpool, C. H. Shin and C.-H. Lee, Cell safety charge transfer kinetics between a lithium metal anode
analysis of a molten sodium–sulfur battery under failure and highly crystalline Li7P3S11 solid electrolyte, Solid State
mode from a fracture in the solid electrolyte, J. Power Ionics, 2016, 286, 24–33.
Sources, 2015, 293, 835–845. 56 S. Wenzel, S. Randau, T. Leichtweiß, D. A. Weber, J. Sann,
42 C. Zhang, S. Santhanagopalan, M. A. Sprague and W. G. Zeier and J. Janek, Direct observation of the
A. A. Pesaran, Coupled mechanical–electrical–thermal interfacial instability of the fast ionic conductor
modeling for short-circuit prediction in a lithium-ion cell Li10GeP2S12 at the lithium metal anode, Chem. Mater.,
under mechanical abuse, J. Power Sources, 2015, 290, 102– 2016, 28(7), 2400–2407.
113. 57 L. Sang, R. T. Haasch, A. A. Gewirth and R. G. Nuzzo,
43 Z. Ma, Z. Xie, Y. Wang, P. Zhang, Y. Pan, Y. Zhou and C. Lu, Evolution at the solid electrolyte/gold electrode interface
Failure modes of hollow core–shell structural active during lithium deposition and stripping, Chem. Mater.,
materials during the lithiation–delithiation process, J. 2017, 29(7), 3029–3037.
Power Sources, 2015, 290, 114–122. 58 R. Koerver, I. Aygün, T. Leichtweiß, C. Dietrich, W. Zhang,
44 C. Zhang, S. Santhanagopalan, M. A. Sprague and J. O. Binder, P. Hartmann, W. G. Zeier and J. Janek,
A. A. Pesaran, A representative-sandwich model for Capacity fade in solid-state batteries: Interphase formation
simultaneously coupled mechanical–electrical–thermal and chemomechanical processes in nickel-rich layered
simulation of a lithium-ion cell under quasi-static oxide cathodes and lithium thiophosphate solid
indentation tests, J. Power Sources, 2015, 298, 309–321. electrolytes, Chem. Mater., 2017, 29(13), 5574–5582.
45 S. S. Damle, S. Pal, P. N. Kumta and S. Maiti, Effect of silicon 59 W. H. Woodford, Electrochemical Shock: Mechanical
congurations on the mechanical integrity of silicon–carbon Degradation of Ion-Intercalation Materials, PhD thesis,
nanotube heterostructured anode for lithium ion battery: A Massachusetts Institute of Technology, 2013.
computational study, J. Power Sources, 2016, 304, 373–383. 60 A. Mukhopadhyay and B. W. Sheldon, Deformation and
46 I. Laresgoiti, S. Kabitz, M. Ecker and D. U. Sauer, Modeling stress in electrode materials for Li-ion batteries, Prog.
mechanical degradation in lithium ion batteries during Mater. Sci., 2014, 63, 58–116.
cycling: Solid electrolyte interphase fracture, J. Power 61 W. Bangerth, R. Hartmann and G. Kanschat, deal.II—
Sources, 2015, 300, 112–122. a general purpose object oriented nite element library,
47 I. Ryu, S. W. Lee, H. Gao, Y. Cui and W. D. Nix, Microscopic ACM Trans. Math. Sow., vol. 33, no. 4, pp. 24/1–24/27, 2007.
model for fracture of crystalline Si nanopillars during 62 W. Bangerth, T. Heister, L. Heltai, G. Kanschat,
lithiation, J. Power Sources, 2014, 255, 274–282. M. Kronbichler, M. Maier, B. Turcksin and T. D. Young,
48 Y. Dai, L. Cai and R. E. White, Simulation and analysis of The deal.II library, version 8.2, Archive of Numerical
stress in a Li-ion battery with a blended LiMn2O4 and Soware, vol. 3, 2015.
LiNi0.8Co0.15Al0.05O2 cathode, J. Power Sources, 2014, 247, 63 S. J. Harris and P. Lu, Effects of Inhomogeneities—
365–376. Nanoscale to Mesoscale—on the Durability of Li-ion
49 H. Yang, F. Fan, W. Liang, X. Guo, T. Zhu and S. Zhang, A Batteries, J. Phys. Chem. C, 2013, 117(13), 6481–6492.
chemo-mechanical model of lithiation in silicon, J. Mech. 64 G. Bucci, T. Swamy, S. Bishop, B. W. Sheldon, Y.-M. Chiang
Phys. Solids, 2014, 70, 349–361. and W. C. Carter, The effect of stress on battery-electrode
capacity, J. Electrochem. Soc., 2017, 164(4), A645–A654.

19430 | J. Mater. Chem. A, 2017, 5, 19422–19430 This journal is © The Royal Society of Chemistry 2017

You might also like