You are on page 1of 11

Electrochimica Acta 130 (2014) 707717

Contents lists available at ScienceDirect

Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta

A Computational Model of the Mechanical Behavior within


Reconstructed Lix CoO2 Li-ion Battery Cathode Particles
Veruska Malav, J.R. Berger , Huayang Zhu, Robert J. Kee
Department of Mechanical Engineering, Colorado School ofMines, Golden, CO 80401, USA

a r t i c l e

i n f o

Article history:
Received 31 December 2013
Received in revised form 18 March 2014
Accepted 24 March 2014
Available online 1 April 2014
Keywords:
Lithium ion battery
Diffusion induced stress
Intercalation process
Phase transformations
Anisotropic stress-strain
Discharge rate
Particle morphology

a b s t r a c t
A coupled electrochemical-mechanical model is developed and applied to predict transient threedimensional stress elds within reconstructed Lix CoO2 cathode particles from commercial Li-ion
batteries. The reconstructed particle geometries are derived from focused-ion-beamscanning-electronmicroscopy (FIB-SEM) experiments. The study uses three individual particles, representing typical sizes
and shapes. The mechanical model incorporates measured anisotropic strain within the Lix CoO2 lattice
and includes strains due to phase transformations. The stresses are generally found to be compressive in
the particle interiors and tensile near the surfaces. Small-scale surface morphology, high Li concentration
gradients, and phase transformations are found to have a major inuence on the stresses, with particularly high tensile stresses near small protuberances and concave notch-like features on the electrode
surfaces. The study considers 1C and 5C discharge rates. The qualitative behaviors are similar at different
discharge rates, but the stress magnitudes are higher at higher discharge rates.
2014 Elsevier Ltd. All rights reserved.

1. Introduction
This paper reports the development of a micro-scale threedimensional (3D) nite element (FE) linear elastic approach to
predict the mechanical behavior within reconstructed Lix CoO2
Li-ion battery (LIB) cathode particles during discharge. The
mechanical model, which is coupled directly to an electrochemistry model, includes the effects of crystal anisotropy and phase
transformations. The study is particularly concerned with predicting the inuence of particle size and surface morphology, as well
as discharge rates. Results show that tensile stresses, especially on
the particle surfaces, can be sufciently high as to suggest particle
fracture.
Representative cathode particles are reconstructed from a
commercial battery using focused-ion-beamscanning-electronmicroscopy (FIB-SEM) [1]. The FIB-SEM experiments typically
produce reconstructions for an assembly of particles within a 3D
rectangular domain measuring a few tens of microns on a side [2].
Individual particles can be extracted from the assembly of particles,
and the present study uses three individual reconstructed particles
with different sizes and shapes.
The particle mechanical behavior is closely coupled with
the transient Li-concentration eld within the cathode particles.

Corresponding author. Tel.: 1 303 273 3682; fax:+1 303 2733602.


http://dx.doi.org/10.1016/j.electacta.2014.03.113
0013-4686/ 2014 Elsevier Ltd. All rights reserved.

Thus, the approach depends upon coupling an electrochemical


simulation with the mechanical simulation. The electrochemical
simulation is accomplished in a nite-volume (FV) setting using
extensions of the ANSYS Fluent software [2]. The mechanical simulation is accomplished in an FE setting using extensions of the
ANSYS Mechanical software.1 At each time step during a transient
discharge simulation, the Li-concentration eld within the particles must be communicated from the electrochemical simulation
to the mechanical simulation.
The mechanical simulations depend upon constitutive relationships between the stress and strain tensors. The present study
uses data published by Reimers, et al. [3], considering the effects
of both isotropic and anisotropic stress-strain relationships. The
Lix CoO2 lattice experiences signicant volume changes and phase
transformations during the lithiation (discharge) process [3]. The
diffusion-induced stresses (DIS) can be very high, potentially
exceeding the material strength and leading to electrode degradation and particle fracture. The present results show that stresses
can be particularly high in the vicinity of notch-like features on
the particle surfaces. Results also show that high discharge rates
and phase transformation occurring during the Li intercalation also
contribute to high stresses.

ANSYS, Inc., Canonsburg, PA 15317; www.ansys.com

708

V. Malav et al. / Electrochimica Acta 130 (2014) 707717

The study focuses on isolated particles that are not constrained


by neighboring particles (i.e., under free-expansion conditions).
The present study also considers isothermal behavior. Thus the
results reveal behaviors that are solely attributed to internal
diffusion-induced stresses.
The electrode particles from actual batteries can vary significantly and randomly in size, shape, and surface morphology.
The physical connections between particles within the electrode
assembly and the electrical contacts vary randomly within the
porous electrode assembly. Moreover, the crystallographic orientations within particles are random. Thus, although the simulations
presented here are quantitative using actual reconstructed particles, the results must be understood in a qualitative context. The
broad objective is to glean observations and trends that are generally applicable.

2. Prior literature
There is signicant foregoing research concerning the mechanical behavior of Li-ion battery cathodes. Reconstructed LIB electrode
particles have been extracted from electrodes using a variety of
microscopic techniques [46,2,7]. Lim et al. [5] developed computational simulations to show that lithiation-induced stresses
depend on geometric characteristics, with the stresses being much
higher in reconstructed Liy C6 and Lix CoO2 particles than in idealized, spherical particles. Likewise, Seo et al. [4] and Chung et al. [6]
used reconstructed particles of LiMn2 O4 compounds in an FE solid
mechanics simulation. Chung et al. [6] reported that the DIS are
much greater in actual particles that in spherical particles. These
investigations [46] were based on elastic, isotropic, single-phase
individual particles and made use of the thermal analogy to compute diffusion strains [8,9]. The present investigation develops an
analogous approach, but additionally considers phase transformation and crystal anisotropy as well as particle surface morphology.
Although a few recent studies have considered anisotropic Li
diffusion in polycrystalline Lix CoO2 particles [1012], the effects of
anisotropy on the stress response in geometrically complex electrode particles has not been reported. Additionally, most analytical
and numerical investigations of DIS in LIBs have not incorporated
the effects of volumetric and/or lattice strains that result from
phase transformations. However, Park et al. [13] have incorporated
phase-transformation-induced stresses in a 3D numerical model
of spherical LiMn2 O4 particles. Their results showed that stresses
associated with phase transformations were greater than those
developed when considering the intercalation process alone. Analytical methods, such as moving boundary and porous electrode
theory, have been developed to investigate the effects of phase
transition and/or phase coexistence during Li intercalation [1417].
These approaches, however, were limited by the following assumptions: a) isotropic elastic behavior in smooth, idealistic particles, b)
two phases concentrically coexisting, c) Li transport is decoupled
from intercalation-induced stress phenomena, and d) phase coexistence modeled as Li-poor or Li-rich phases. Regarding the latter,
no published literature suggests that either of the two hexagonal
phases in the phase coexistence region of Lix CoO2 is richer in Li than
the other. In addition, smooth and spherical particles are unable
to capture stress concentrations that can develop in local concave
regions of actual cathode particles. Understanding and predicting
the mechanical behavior of electrodes is practically important. For
example, capacity fade can be associated with diffusion-induced
stress [4]. Even under normal operating conditions, particles can
fracture and thus degrade battery performance. Particle fracture
can originate from locally high stresses leading to the formation
and growth of microcracks [18]. Such processes are known to be
intensied at high discharge rates [19]. If a fractured particle looses

electrical contact with neighboring particles or current collection


foils, it can no longer participate electrochemically and the battery resistance increases and capacity fades [13,20]. Additionally,
fragmentation exposes fresh electrode surfaces to the electrolyte
solvent, thus promoting the growth of new surface solid electrolyte
interface (SEI) lms [20,4].
Hydrostatic stress gradients are known to inuence Li diffusion with electrode particles [21]. Thus, in addition to mechanical
degradation associated with diffusion-induced stress, the stress
state couples back into the electrochemistry problem. This effect
is neglected in the present study, but is the subject of active model
development.
3. Particle reconstruction and computational discretization
Fig. 1 illustrates the process used to dene the single particles used in the present study. A commercial cell (here, Lishen2
LR18650AH) is disassembled and a small portion of a cathode is
prepared for FIB-SEM imaging [1]. The raw data from the FIB-SEM
consists of approximately 200 two-dimensional SEM slices, with
each slice being separated by approximately 60 nm. In Fig. 1 the
white areas in the FIB-SEM slices represent the cathode particles
and the black space represents the pore space that would be lled
by electrolyte solvent. As discussed by Wiedemann et al. [2], the 3D
cube is reconstructed using the Mimics software.3 The 3D geometry
is represented in STL (STereoLithography) format, which is a computational denition of the particle surfaces. For the purposes of the
present study, individual particles are extracted from the rectangular assembly of many particles. The individual particle geometry
is also represented in STL format, which is used as the basis for
computational discretization.
Fig. 2 illustrates the three reconstructed cathode particles used
in the present study. The particles, labeled P1 , P2 , and P3 , are rendered at the same scale to show the relative particle sizes. In
addition to the reconstructed particles, a perfectly smooth spherical particle (labeled Ps ) is also modeled. Cathode particles may be
polycrystalline with a few grains or be composed of a single grain
[22]. The present model assumes that each particle is composed of
a single crystal for the anisotropic studies.
To be electrochemically active, the particles must be connected
to other particles and ultimately to current-collection foils in the
battery. The yellow patches on particles (Fig. 2) indicate the surface positions through which electrical current enters the particle
during discharge. Lithium enters the particles via charge-transfer
reactions on the surfaces that are in contact with the electrolyte
solution. The anisotropic Lix CoO2 crystal grain orientation is illustrated by the xyz axes which correspond to the abc crystallographic
axes. In the present study, the electrical contact areas (yellow
patches) and crystal orientation are assigned somewhat arbitrarily.
However, in all cases the electrical contacts are essentially aligned
with the c-axis of the Lix CoO2 lattice.
The electrochemical model, which has been described previously [2], uses an FV mesh that is generated using the
TGRID algorithm. The structural model uses the ve degree-offreedom (DOF) element SOLID 227 that is implemented in ANSYS
Mechanical v14.5. This element consists of a 3D 10-node tetrahedron. The tetrahedral elements were selected for a variety of
reasons. First, coupled-eld elements are required in order to couple the mechanical response to the chemical diffusions. These
particular elements in ANSYS Mechanical are tetrahedral or hexahedral elements. However, the tetrahedral elements provide a
lower error when numerical solutions are compared to analytic

2
3

Tianjin Lishen Battery Co., Ltd., Tianjin, China, http://en.lishen.com.cn


Materialize, NV; Leuven, Belgium; http://www.materialise.com

V. Malav et al. / Electrochimica Acta 130 (2014) 707717

709

Fig. 1. Individual particles are extracted from FIB-SEM cathode reconstructions.

as much as a 2.6% contraction upon full lithiation. There is also signicant discontinuity between the lattice volumes associated with
the two hexagonal phases. Figs. 3c-d show signicant anisotropy
between the a and c lattice parameters. The a and b lattice parameters are equal in the hexagonal phases. However, in the monoclinic

Fig. 2. Three reconstructed cathode particles and a 2.7-m spherical particle. The
electrical contact areas are shown as yellow patches. As indicated by the Cartesian
axis, the z axis of the Lix CoC2 is essentially normal to the electrical contact area.

solutions, and provide a good representation of the DIS state for


the reconstructed surfaces based on the original STL image format.
Because two different discretizations are used, the Li-concentration
eld from the FV simulation of the electrochemistry must be interpolated onto the FE mesh at each time step during the simulation.
Table 1 provides some summary information about each of the
particles.
4. Phase transformation and strain
Fig. 3a illustrates a pseudo phase diagram, which shows the
possibility of three phase transformations as a function of Li intercalation [3,18,2326]. Fig. 3b shows that the lattice can experience
Table 1
Particle geometric and mesh characteristics and current densities at 1C
Particle

Volume
(m3 )

Surface Area
(m2 )

Size
(m)

FE Nodes

i at 1C
(A m2 )

P1
P2
P3
Ps

4.10
16.50
82.34
82.34

15.0
37.1
116.7
91.6

2.6
3.6
10.2
2.7

86,015
106,506
316,624
65,637

7.56
7.92
7.58
9.14

Fig. 3. Li/Lix CoO2 pseudo phase diagram (adapted from Reimers, et al. [3]). The
monoclinic phase is labeled as M1. Two hexagonal phases are labeled as H1 and
H2.

710

V. Malav et al. / Electrochimica Acta 130 (2014) 707717

phase, where b =
/ a, the average lattice parameter is expressed as
[3]
1
aM1 = aM1 + bM1
3
Note that there is a very small increase in the a lattice parameter
as the Li fraction increases. However, there is a much more significant contraction in the c axis as Li fraction increases (i.e., during
battery discharge).
The coincidence of two hexagonal phases (Fig. 3, 0.75 < x < 0.94)
presents a challenge to mechanical modeling. It is assumed that
the volumetric and/or lattice dimensions are affected equally by
contributions from the H1 and H2 phases, thus representing the
overall mist strain as an average. The present study is the rst to
numerically predict the mechanical behavior associated with the
two coexisting hexagonal phases.
Linear functions can be used to approximate the strain associated with lithiation (straight lines in Fig. 3b-d). The present
study compares predictions based upon isotropic and anisotropic
structural behavior. The isotropic case uses the linear functions
representing volumetric strain as illustrated in Fig. 3b. The linear
function is somewhat less accurate in the monoclinic phase than it
is in the hexagonal phases. The anisotropic case uses the linear functions to approximate the lattice parameters as shown in Figs. 3c-d.
Note that intercalation-induced variations in the c-axis strain are
much greater than they are in the a axis.
5. Electrochemical model
The electrochemical model in the present study makes two
important assumptions [2]. First, the Li diffusion within the
cathode particle is isotropic. Second, the Li diffusion is entirely
driven by concentration gradients, thus neglecting any contribution associated with hydrostatic stress gradients. Based upon these
assumptions, the Li-concentration eld within the particle may be
represented as

C
= (D C) ,
t

(1)

where C is the Li concentration, J = D C is the Li diffusion ux,


and D is the diffusion coefcient of Li within the Lix CoO2 lattice.
Although D is typically assumed to be a constant, a wide range of
values for the Li diffusion coefcient have been reported [27]. The
present study uses D = 5.39 1015 m2 s1 .
The electrochemical model must also account for the electrostatic potential  within the electrode particle, which is
represented by the Gauss equation as

( ) = 0,

(2)
m1 .

where the conductivity is taken as  = 10 S


Because Eqs. 1 and 2 both involve elliptic spatial operators,
boundary conditions are required everywhere on the particle surface. Everywhere on the surface, except the electrical contact area
(yellow patches, Fig. 2), the Li ux is established via a chargetransfer reaction (Li+ + e  Li) governed by a ButlerVolmer rate
expression as

s e = i0 exp

 F 
a
RT

 F 
c

exp

RT

(3)

where s e is the Faradaic current density, i0 is the exchange current density, a and c are anodic and cathodic symmetry factors,
respectively, F is the Faraday constant, R is the gas constant, T is the
absolute temperature, and  is the activation overpotential. The
overpotential is dened as
eq

 = ed el Eed ,

(4)

Fig. 4. Cathode particle, illustrating electrochemical conservation equations and


interface boundary conditions.

where ed and el represent electrostatic potentials of the electrode


eq
and electrolyte phases, respectively, and Eed is the equilibrium
eq
electric-potential difference. The present study represents Eed as
an empirical expression that depends upon temperature and Li
concentration [28]. The exchange current density is modeled as

io = kr FCLi+ a C max C

a

C c ,

(5)

where kr = 2.07 108 m2.5 kmol0.5 s1 is a constant, CLi+ is the


Li-ion concentration in the electrolyte solution, C is the intercalated Li concentration at the outer edge of the cathode (i.e., interface
between electrode and electrolyte solution), and Cmax is the saturated (maximum) intercalated Li concentration within the cathode.
The present study assumes a constant and uniform Li+ concentration in the electrolyte as CLi+ = 1.2 kmol m3 .
As illustrated in Fig. 4, at the interfaces between the cathode
particle and the electrolyte solution, the electrostatic-potential gradient and the Li ux are governed by the charge-transfer chemistry
as
  n = s e ,

D C n =

s e
.
F

(6)

The outward-pointing surface-normal unit vector n represents


the local spatial orientation of the cathode surface. The local current
density s e is evaluated from the ButlerVolmer equation (Eq. 3),
based upon local conditions at the cathode surfaces.
At the interfaces between the cathode particle and electrical
contact, the Li uxes must vanish and the electronic current density
i is specied by the discharge rate.
  n = i,

D C n = 0.

(7)

Assuming a specied C rate, the current density i = I/A (i.e., current per unit area) at the electrical-contact interface is evaluated
from the total current I and the contact area A (yellow patches,
Fig. 2).
6. Mechanical model
The FE mechanical model within the particles depends upon
the local Li concentration elds that are imported from the
electrochemical model at each timestep during a discharge. In
contrast to Li diffusion within the particle solids, which is a timedependent process (Eq. 1), swelling strains in the elastic media
occur signicantly faster, achieving nearly instantaneous mechanical equilibrium [29]. Thus, a quasi-static approach is appropriate
for modeling an elastostatic problem at each instant during the
discharge. The local stress eld  ij within an elastic solid can be

V. Malav et al. / Electrochimica Acta 130 (2014) 707717


Table 2
Li/Lix CoO2 cathode model parameters used in the FE structural analysis
Name

Symbol

Value

Reference

Density
Youngs modulus
Elastic stiffness


E
C11
C12
C13
C33
C44

94%
CLi
50%
CLi
DLi

2328.5 kg m3
370.0 GPa
596.0 GPa
200.0 GPa
133.0 GPa
375.0 GPa
124.0 GPa
0.20
22.37 kmol m3
11.95 kmol m3
5.387 1015 m2 s1

[2]
Current study
[10], [30]

Poissons ratio
Li saturation concentration
Initial Li concentration
Li diffusivity

Current study
[2]
Current study
[2]

711

The value of for the isotropic case can be extracted


from the lattice-volume data shown in Fig. 3b. For a given
volumetric strain (represented in Cartesian coordinates), the volume at a given intercalation fraction x may be evaluated as
V(x) = (da + da)(db + db)(dc + dc). Expanding this relationship,
and neglecting higher-order contributions, the normal strain due
to both intercalation and phase-transformation can be represented
as
ij =

1
3

 V (x) V 
0

V0

daa = dbb = dcc =

ij ,

(11)

1 j

C = j
C,
3 V0

(12)

where
expressed assuming mechanical equilibrium in the absence of body
forces as

ij
= 0,
xj

(8)

where xj are spatial coordinates. The total elastic strain tensor, Tij ,
and chemical diffusion
can be decomposed into the mechanical m
ij
dij contributions [19],
Tij = m
+ dij .
ij

(9)

The present model neglects external loads (i.e., the surface of


the electrode particle is traction free). Thus, the cathode particle is
considered to be in a free-expansion, quasi-static, state wherein
equilibrium (Eq. 8) is satised everywhere within the particle.
However, the spatially non-uniform Li concentration eld creates
non-uniform elastic strain and stress elds within the particle (irrespective of external constraints).
As a basis for comparison, the material properties within cathode particles are rst assumed to be isotropic. Then, in subsequent
simulations, the Lix CoO2 lattice is assumed to be anisotropic. A
thermal analogy is used to compute the strains resulting from Li
diffusion [8,9]. A single stress-strain relationship is used to represent the strains that are induced by Li intercalation as well as
phase transformations. This approach is facilitated by dening a
concentration-dependent chemical expansion tensor, ij . This tensor species the changes of the crystal volume in the isotropic
case (Fig. 3b), and the changes in the lattice parameters for the
anisotropic case (Fig. 3c-d). Thus, the diffusion-induced strain is
both structure and composition sensitive.
6.1. Isotropic strains
The elastic strains in the isotropic case, where ij = ij , are
expressed as
Tij =

1
[(1 + )ij kk ij ] + ij
C,
E

(10)

where E is the Young modulus,  is the Poisson ratio, ij is the


Kronecker delta,  ij is the stress tensor, and
C is the concentration difference between the current and initial Li composition (i.e.,

C = C C0 ). In the isotropic case, it is assumed that the particle


contains a sufcient number of grains to use effective isotropic elastic parameters. Reuss averages are used, with an effective Youngs
modulus E and an effective Poisson ratio  evaluated from the hexagonal elastic-constant data reported by Hart and Bates [30]. To the
authors best knowledge, no data have been reported regarding
the Lix CoO2 monoclinic-phase material properties. Therefore, the
anisotropic models use the hexagonal elastic constants. Previous
studies have used a wide range of values for the Youngs modulus
of Lix CoO2 [16,5,31,30]. The value used here, E = 370 GPa (Table 2),
lies within previously reported ranges.

j =

1 j
.
3 V0

(13)

In these expressions j is the slope of the linear function corresponding to the specic Li intercalation fraction and V0 is the initial
crystal volume (evaluated at x 0.5).
6.2. Anisotropic strains
Most previous studies of mechanical response in LIB cathodes
have used isotropic elastic properties, even for Lix CoO2 . However, among the commonly used cathode materials, anisotropic
lattice strains are especially signicant for Lix CoO2 [32]. The
anisotropy is particularly important because individual particles
can be monocrystalline or polycrystalline with only a few grains
[22].
Anisotropic stress analysis is based upon directional lattice
strains instead of overall volumetric strains. The diffusion strain is
determined from the anisotropic crystal lattice behavior (Fig. 3c-d).
The anisotropic elastic stress-strain relationships may be represented as
Tij = Sijk k + ij
C,

(14)

where ij is the anisotropic expansion coefcient and Sijk is the


elastic-compliance tensor. The components of Sijk can be determined from the elastic stiffness values provided in Table 2. The
concentration-dependent ij is evaluated from the linear ts to
the lattice expansion data (Fig. 3c-d). The present model incorporates the anisotropic ij tensor into the normal strain relationship.
These anisotropic relationships capture the combined effect of
the lattice-contraction and phase-transformation strains that are
induced during battery discharge. The anisotropic strains may be
written as
daa = dbb =

a
,
a0

dcc =

c
,
c0

(15)

where a0 and c0 are the corresponding a and c lattice-constant values at x 0.5 (i.e., beginning of the discharge process). The resulting
anisotropic stress-strain relationship is then decomposed as
Tij = Sijk k + dij ,

(16)

where dij is the strain tensor expressed in Eq. 15.


6.3. Effect of Li intercalation
Fig. 5 plots the approximate isotropic and anisotropic strain
functions that are derived from the Reimers et al. [3] data and
used in the model. Fig. 5a shows slightly swelling anisotropic strain
in the a and b lattice directions as the Li fraction increases. Conversely, Fig. 5b shows signicanty stronger anisotropic contraction
in the c lattice direction as Li fraction increases. By denition, the
isotropic strain must be the same in all directions. Figs. 5a-b show

712

V. Malav et al. / Electrochimica Acta 130 (2014) 707717

1000

a) a and b axes

500
0

Anisotropic model

Chemical Strain (strains)

-500
-1000

Isotropic model

-1500

Isotropic model

-2000
-2500
2000

b) c axis

Isotropic model

-2000

-6000

Fig. 7. Diffusion-induced stress elds in fully lithiated particles Ps and P3 at the


end of a 1C discharge (i.e., t = 3600 s). Both simulations are based on the anisotropic
model. These two particles have comparable volumes (Table 1).

Anisotropic model

-4000

Anisotropic model

-8000
0.5

0.55

0.6

0.65

0.7

0.75

0.8

0.85

0.9

x in LixCoO2
Fig. 5. Approximate linear strains as functions of Li intercalation, a) aligned with a
and b lattice directions, daa and dbb and b) aligned with c lattice direction, dcc . The
isotropic strains are the same in both panels, but plotted using different scales for
the ordinates.

the same isotropic strains, but plotted on different ordinate scales.


The discontinuities at x 0.75 is the result of phase transformation
between the H1 and H2 hexagonal phases.
Fig. 5 shows isotropic and anisotropic strain functions to be
signicantly different. Thus, it should be anticipated stress elds
predicted by isotropic models will be quite different from those
that assume the more-realistic anisotropic behavior.
7. Results and discussion
Table 2 lists physical parameters that are used in the present
study. Other parameters and implementation details for the
electrochemical model can be found elsewhere [2]. The results presented here express the DIS as maximum principal stresses, which
are especially signicant for predicting the failure of ceramic, nonductile, materials.
7.1. Isotropic and anisotropic comparison
Fig. 6 shows stress elds in particle P2 at t = 1820 s during a
1C discharge, using the isotropic and anisotropic models. In both
cases the stresses are compressive in the particle interior, but tensile in the outer regions near the particle surfaces. The anisotropic

model predicts signicantly higher tensile stresses. As indicated by


the arrow in Fig. 6b, concave or notched features on the particle
surface show particularly high tensile stresses. Such high tensile
stresses are likely to cause crack initiation and fracture. Although
different particles with different crystallographic orientations and
different electrical contacts will lead to different stress elds, this
result suggests strongly that using an isotropic model may signicantly under-predict deleterious tensile stresses.
7.2. Surface morphology
Fig. 7 compares predicted stress elds in the ideal spherical particle Ps with the reconstructed particle P3 that has a comparable
volume (Table 1). Both cases use the anisotropic model, but the
spherical particle has a smooth convex surface compared to the
relative rough features of the reconstructed particle. At the end
of a 1C discharge, the maximum predicted tensile stress in Ps is
approximately 53.8 MPa. By comparison, the maximum predicted
tensile stress in P3 is approximately 79.9 MPa. In both cases, stresses
in the core of the particles are compressive, with tensile stresses
in the outer regions. The non-smooth surface morphology of the
actual particle leads to much higher local stress variations than are
predicted in the smooth-surface spherical particle.
The stress eld within the spherical particle tends to be concentric, which is the result of an essentially concentric Li concentration
eld. However, because the electrical contact occupies a portion
of the particle surface, the elds are not exactly concentric. In
any case, the shapes of actual particles are usually very different
from spheres. Consequently, although the predicted elds may look
qualitatively concentric (i.e., more compressive toward the interior), the actual elds can be quite different from those in a spherical
particle.
These results demonstrate that surface features dominate the
mechanical response of the cathode particle. Local small-scale protrusions in the surface topology create local areas that lithiate more
quickly than the surrounding areas. Consequently, these locations
undergo phase transformations more rapidly than the surrounding
material, leading to high, localized, stresses. Such behavior is not
captured by simulations using smooth spherical particles.
7.3. Particle shape and size

Fig. 6. Diffusion-induced stress elds in particle P2 using a) the isotropic model


and b) the anisotropic model. In both cases the battery is discharging at 1C and the
stress elds are shown at t = 1820 s, which is approximately midway through the
discharge. The arrow points to a notched region on the surface with high tensile
stress.

Figs. 810 show predicted maximum-principal-stress elds in


the three reconstructed particles at t = 1650 s into a 1C discharge. In
all cases, the stress patterns are qualitatively similar. That is, compressive in the interior and tensile near the outside edges. Concave
features on the surfaces tend to be regions of relatively high tensile stress. Particles P1 and P2 , which are comparable in particle
size, show comparable stress distributions. Particle P3 is larger and
shows a wider range of DIS. However, as discussed subsequently,

V. Malav et al. / Electrochimica Acta 130 (2014) 707717

713

a) Variation in Li

0.14

1C discharge

0.12
0.1

P3

0.08
0.06
0.04

P1

P2

0.02

Fig. 8. Maximum principal stress eld in P1 at t = 1650 s under a 1C discharge. a)


stress contours on the surface and b) stress contours on a cross-section cut.

Max. Prin. Stress (MPa)

0
700
600
500
400
300
200
100
0

b) Maximum DIS

P3
P2

P1

100

600

1100

1600

2100

2600

3100

3600

Discharge time (s)


Fig. 11. a) Variation of intercalated Li fraction within the three reconstructed cathode particles during 1C discharge. b) Maximum principal tensile stress within the
reconstructed particles throughout a 1C discharge.

7.4. Maximum principal stress

Fig. 9. Maximum principal stress eld in P2 at t = 1650 s under a 1C discharge. a)


stress contours on the surface and b) stress contours on a cross-section cut.

the particle size itself may not be the dominant factor in governing
the maximum stress.
Although not shown in Figs. 810, the model reveal that the
magnitude of the principal stresses depend upon the magnitude of
the Li-concentration gradients. In P3 the maximum Li concentration
gradients are on the order of 108 kmol m4 , while in P1 and P2 the
maximum Li concentration gradients were on the order of 106 kmol
m4 .
Although the predicted results are based upon quantitative simulations, they should be understood in terms of qualitative trends.
Individual particle shapes and sizes are random within the full electrode. The electrical contacts and crystallographic orientations are
also random. Thus, any particular simulation cannot produce a fully
general result.

Fig. 11a shows the maximum variation in Li fraction


x within
the three reconstructed particles. The relatively small particles (P1
and P2 ) show relatively small variations that are essentially constant at
x 0.02. The variations in the larger particle (P3 ) are much
greater, reaching a peak at
x 0.14 about midway through the
discharge. Fig. 11b shows the transient maximum principal tensile
stresses during a 1C discharge. The maximum-stress proles are
qualitatively similar for all the particles, but the magnitudes are
greater for P3 . The stress peaks at approximately midway through
the discharge are caused by the phase transformations between the
two hexagonal phases (H1 to H1+H2) that occur at x 0.75 (Figs. 3
and 5). These sharp increases in tensile stress, which tend to be concentrated around features such as small protuberances or notches
on the particles surfaces, can serve to damage, crack, or fracture the
cathode particles.
Fig. 12, which shows contours of maximum principal stress,
strain, and Li concentration on the surface of P3 at t = 1820 s during
a 1C discharge, helps to explain the proles shown in Fig. 11. The
dominant behaviors are localized around a small protruding surface
feature that is also shown at an expanded scale. The peak compressive and tensile stresses are highly localized around this surface
feature (Fig. 12a). The strain is relatively small on the tip of the
feature, but increases greatly near the base of the feature where it
joins the bulk of the particle (Fig. 12b). Because the feature is small,
the Li intercalation fraction is highest near the tip of the feature
(Fig. 12c). Because the small feature lithiates so rapidly relative
to its nearby surroundings, a highly localized phase transformation (H1 to H1+H2) contributes to locally high stresses. The local
phase transformation results in a relaxation of the local diffusion
strains.
7.5. Discharge rate

Fig. 10. Maximum principal stress eld in P3 at t = 1650 s under a 1C discharge. a)


stress contours on the surface and b) stress contours on a cross-section cut.

Increasing the discharge rate increase the stresses, but the


general behaviors remain qualitatively similar. The present study
compares mechanical behaviors at 1C and 5C. At the higher discharge rate, the resulting Li concentration eld is found to be more
non-uniformly distributed spatially. As a consequence, there is an
increase in the concentration gradients of at least one order of

714

V. Malav et al. / Electrochimica Acta 130 (2014) 707717

Fig. 12. Localized elds within P3 at t = 1820 s during a 1C discharge. a) Maximum principal stress; b) equivalent diffusion strain; c) local Li fraction x.

magnitude. The increases in maximum stresses are approximately


proportional to the increases in Li concentration gradients. In large
measure, the stresses increase at higher discharge rates because the
Li-concentration variations within the particle are greater at high
rates.
Figs. 13, 14, and 15 show maximum principal stress maps for for
P1 , P2 , and P3 , respectively. As is the case at lower discharge rates,
stresses are generally compressive in the particle interiors and tensile near the surfaces. Also, locally high stresses are evident around
small protuberances and concave features on the surfaces. These
observations are consistent with the previously reported results
[19,5].
Fig. 16a shows that after an initial transient, the maximum variations in intercalated Li are approximately steady at
x 0.07
Fig. 14. DIS eld in P2 at t = 335 s during a 5C discharge condition. a) Particle surface
and b) Particle cross section.

Fig. 13. DIS eld in P1 at t = 335 s during a 5C discharge condition. a) Particle surface
and b) Particle cross section.

Fig. 15. DIS eld in P3 at t = 335 s during a 5C discharge condition. a) Particle surface
and b) Particle cross section.

V. Malav et al. / Electrochimica Acta 130 (2014) 707717

0.25

5C discharge

a) Normalized Li

0.20

715

P3

0.15

P2

0.10
0.05

P1

Max. Prin. Stress (MPa)

800

b) Maximum DIS

600

P3
P2

400

P1

200
0

20

120

220

320

420

520

620

720

Discharge time (s)


Fig. 16. a) Maximum difference in intercalation fraction
x within particles during
a 5C discharge. b) Maximum principal stress histories for three particles during a 5C
discharge.

for P1 and
x 0.10 P2 . The Li variation is greater for P3 , with
larger variations during the course of the lithiation. The peak
variation
x 0.23 occurs during the H2-to-(H2+H1) phase transition (t = 413 s) and gradually declines for the remainder of the
lithiation process. As was the case in the 1C discharge, the small
feature illustrated in Fig. 12 plays a signicant role.
Fig. 16b shows the history of the DIS for the particles during a
5C discharge. Comparison with Fig. 11 shows that maximum principal stresses in the M1 and H1 phases are generally greater than
they are during a 1C discharge. The H2-to-(H2+H1) phase transition occurs at t 400 s, leading to the highest stresses. Because
the intercalation-induced strains decline in the later stages of
the discharge where the H1 and H2 phases coexist, the stress
levels decrease. The stress state depends upon the Li intercalation fraction, x, its variation,
x, and the concentration gradient
throughout the particle. This agrees with the observations reported
in [33]. Both of these factors are highly dependent on the discharge rate as well as particle size and morphology. The stresses
are generally higher at higher discharge rates because
x is
greater.
Fig. 17 illustrates maximum tensile stress histories at four
selected points on the surface of P3 with discharge rates of 1C and
5C. The particular locations are chosen near topological features
that would tend to create high local stresses. To assist comparison between 1C and 5C the abscissa is a normalized time, t = t/tD ,
where tD is the time required for full discharge (e.g., tD = 3600
s at 1C). At most points the stress histories are similar but the
magnitudes are greater at high C rates. The most striking behavior occurs at t 0.5, where the H2-to-(H2+H1) phase transition
takes place at pronounced concave and convex locations. These
sites are accommodating regions under stress peaks and decays.
Entering the coexistence region, a comparable stress evolution pattern occurs at all sites evaluated in the particle throughout the
remainder of the lithiation.
Fig. 18 illustrates a matrix of cross-sectional images for P3 , showing principal stresses, strain, and lithium intercalation fractions at
three times during a 5C discharge. The row at t = 85 s corresponds
to the M1-to-H2 phase transition. The row at t = 413 s corresponds
to the H2-to-(H2+H1) phase transition. The row at t = 720 s corresponds to the end of the 5C discharge. Fig. 18a shows surface
tensile stresses around 550 MPa. Fig. 18d shows that during the H2to-(H2+H1) phase transition the surface tensile stresses are even
higher, in the range of 670 MPa. Even within the particle interior,
tensile stresses can exceed 300 MPa. At t = 413 s, Fig. 18d show

Fig. 17. Maximum principal stress histories at selected sites on the surface of P3
when discharged at 1C and 5C. The time is normalized between 0 t 1 to accommodate the different C rates (t = 1 corresponds to full discharge).

a nominally concentric ring of stresses that are relatively more


compressive than the surrounding regions outside the ring. The
ring corresponds to a chemical strain mist at x 0.75 as the Li
intercalation proceeds inward (Fig. 18e). Fig. 18g shows that at
the end of the lithiation process (t = 720 s) the stress magnitude
decreases, which is caused by the increasingly uniform Li intercalation. Although the stress magnitudes decrease near the end of
discharge, the general pattern of being tensile near the surface and
compressive in the interior is preserved.

716

V. Malav et al. / Electrochimica Acta 130 (2014) 707717

Fig. 18. Mechanical and chemical response of particle P3 at three instants in time during a 5C discharge.

8. Summary and conclusions


The present study is based upon the computational simulation of reconstructed Lix CoO2 cathode particles during discharge.
Three individual particles were selected from FIB-SEM reconstructions of a commercial battery, with the particles being selected to
represent a range of sizes and shapes. The approach couples threedimensional nite-volume simulations of the electrochemistry and
Li transport with three-dimensional nite-element simulations of
the mechanical response. The mechanical models compare both
isotropic and anisotropic behavior, considering phase transitions
(monoclinic and two hexagonal phases) associated with the extent
of lithiation in the Lix CoO2 lattice.
A number of general observations can be drawn from the study.
During discharge, the stresses are generally compressive in the particle interiors and tensile near the particles surfaces. By comparing
the behavior of a smooth spherical particle with the reconstructed
particles, it is evident that the surface morphology plays a strong
role on the maximum tensile stresses. Especially high stresses are
found near concave, notch-like, features on the surfaces. Surface
features such as small protuberances that lithiate rapidly relative
to the surrounding surface also promote high stresses.
The anisotropic strain characteristics of the Lix CoO2 lattice
are important. Comparison with equivalent isotropic stress-strain
models show that that stresses are signicantly higher when
the anisotropic models are used. Thus, predictions based on
isotropic assumptions are likely to under-estimate the propensity
for electrode degradation, crack formation, and particle fracture.
Additionally, a major factor contributing to the stress state in
the individual Lix CoO2 particles was the phase transformation the

crystal structure undergoes after half lithiation. This phase transformation, combined with large concentration gradients, allow
dramatic changes in the stress that evolve in the particle.
The mechanical behavior depends signicantly upon discharge
rate. The present study compared 1C and 5C discharges. Although
the qualitative behaviors are similar at different discharge rates, the
stresses are signicantly higher at higher discharge rates. Larger
Li-concentration gradients at high discharge rates contributes to
the higher stresses. Spatially varying lithiation rates, especially
near small-scale surface features, also contributes to high stresses.
Assuming that one is concerned with mechanical damage and particle fracture, the study suggest some considerations for electrode
design and operation. As much as practical, nominal spherical particles with smooth surface morphologies are desirable. Because
crystal anisotropy is important, it may be desirable to fabricate
polycrystalline particles with numerous randomly oriented grains,
leading to overall more isotropic behavior. Limiting high discharge
rates is also desirable from the perspective of limiting mechanical
degradation.

Acknowledgements
We are grateful to Prof. Scott Barnett (Northwestern University), who provided data from FIB-SEM experiments and to Andreas
Wiedemann who assisted with the generation of the model geometry. The authors also gratefully acknowledge the assistance of Drs.
Graham Goldin and Bill Bulat (ANSYS, Inc.) in the development of
the computational models. This effort was supported by the Ofce
of Naval Research via Grant N00014-08-1-0539.

V. Malav et al. / Electrochimica Acta 130 (2014) 707717

References
[1] J.R. Wilson, J.S. Cronin, S.A. Barnett, S.J. Harris, Measurement of threedimensional microstructure in a LiCoO2 positive electrode, J. Power Sources
196 (2011) 4433447.
[2] A.H. Wiedemann, G.M. Goldin, S.A. Barnett, H. Zhu, R.J. Kee, Effects of threedimensional cathode microstructure on the performance of lithium-ion battery
cathodes, Electrochim. Acta 88 (2013) 580588.
[3] J.N. Reimers, N. Jan, J.R. Dahn, Electrochemical and in situ x-ray diffraction
studies of lithium intercalation in Lix CoO2 , J. Electrochem. Soc. 139 (1992)
20912097.
[4] J.H. Seo, M. Chung, J.M. Park, S.W. Han, X. Zhang, A.M. Sastry, Generation of
realistic particle structures and simulations of internal stress: A numerical/afm
study of LiMn2 O4 particles, J. Electrochem. Soc. 158 (2011) A434A442.
[5] C. Lim, B. Yan, L. Yin, L. Zhu, Simulation of diffusion-induced stress using
reconstructed electrodes particle structures generated by micro/nano-CT, Electrochim. Acta 75 (2012) 279287.
[6] M.D. Chung, J.H. Seo, X.C. Zhang, A.M. Sastry, Implementing realistic geometry and measured diffusion coefcients into single particle electrode modeling
based on experiments with single LiMn2 O4 spinel particles, J. Electrochem. Soc.
158 (2011) A371A378.
[7] T. Hutzenlaub, S. Thiele, R. Zengerle, C. Ziegler, Three-dimensional reconstruction of a LiCoO2 Li-ion battery cathode, Electrochem. Solid-State Lett. 15 (2011)
A33A36.
[8] X. Zhang, W. Shyy, A.M. Sastry, Numerical simulation of intercalation-induced
stress in Li-ion battery electrode particles, J. Electrochem. Soc. 154 (2007)
A910A916.
[9] X. Zhang, A.M. Sastry, W. Shyy, Intercalation-induced stress and heat generation
within single lithium-ion battery cathode particles, J. Electrochem. Soc. 155
(2008) A542A552.
[10] D. Chung, N. Balke, S.V. Kalinin, R.E. Garcia, Virtual electrochemical strain
microscopy of polycrystalline LiCoO2 lms, J. Electrochem. Soc. 158 (2011)
A1083A1089.
[11] N. Balke, E.A. Eliseev, S. Jesse, S. Kalnaus, C. Daniel, N.J. Dudney, A.N.
Morozovska, S.V. Kalinin, Three-dimensional vector electrochemical strain
microscopy, J. Appl. Phys. 112 (2012) 052020.
[12] S. Yamakawa, H. Yamasaki, T. Koyama, R. Asahi, Numerical study of Li diffusion
in polycrystalline LiCoO2 , J. Power Sources 223 (2013) 199205.
[13] J. Park, W. Lu, A.M. Sastry, Numerical simulation of stress evolution in lithium
manganese dioxide particles due to coupled phase transition and intercalation,
J. Electrochem. Soc. 158 (2011) A201A206.
[14] R.T. Purkayastha, R.M. McMeeking, An integrated 2-D model of a lithium ion
battery: The effect of material parameters and morphology on storage particle
stress, Comput. Mechanics 50 (2012) 209227.
[15] Q. Zhang, R. White, Moving boundary model for the discharge of a LiCoO2
electrode, J. Electrochem. Soc. 154 (2007) A587A596.

717

[16] S. Renganathan, G. Sikha, S. Santhanagopalan, R.E. White, Theoretical analysis


of stresses in a lithium ion cell, J. Electrochem. Soc. 157 (2010) A155A163.
[17] S. Golmon, K. Maute, M.L. Dunn, Numerical modeling of electrochemical
mechanical interactions in lithium polymer batteries, Comput. Struct. 87 (2009)
15671579.
[18] J. Park, S. Kalnaus, S. Han, Y.K. Lee, G.B. Less, N.J. Dudney, C. Daniel, A.M. Sastry,
In situ atomic force microscopy studies on lithium (de)intercalation-induced
morphology changes in Lix CoO2 micro-machined thin lm electrodes, J. Power
Sources 222 (2013) 417425.
[19] R.E. Garcia, Y. Chiang, W.C. Carter, P. Limthongkul, C.M. Bishop, Microstructural
modeling and design of rechargeable lithium-ion batteries, J. Electrochem. Soc.
152 (2005) A255A263.
[20] M. Zhu, J. Park, A.M. Sastry, Fracture analysis of the cathode in Li-ion batteries:
A simulation study, J. Electrochem. Soc. 159 (2012) A492A498.
[21] Y.M. Choi, S.I. Pyun, Effects of intercalation-induced stress on lithium transport
through porous LiCoO2 electrodes, Solid State Ionics 99 (1997) 173184.
[22] K. Zhao, M. Pharr, J.J. Vlassak, Z. Suo, Fracture of electrodes in lithium-ion batteries caused by fast charging, J. Appl. Phys. 108 (2010) 073517.
[23] H. Wang, Y.I. Jang, B. Huang, D.R. Sadoway, Y.M. Chiang, Batteries and energy
conversion - TEM study of electrochemical cycling-induced damage and disorder in LiCoO2 cathodes for rechargeable lithium batteries, J. Electrochem. Soc.
146 (1999) 473.
[24] T. Bak, J. Nowotny, M. Rekas, C.C. Sorrell, S. Sugihara, Properties of the electrode
material LiCoO2 , Ionics 6 (2000) 92106.
[25] G.A. Nazri, G. Pistoia, Lithium Batteries: Science and Technology, Springer, New
York, 2009.
[26] S.H. Choi, J. Kim, Y.S. Yoon, A TEM study of cycled nano-crystalline HT-LiCoO2
cathodes for rechargeable lithium batteries, J. Power Sources 135 (2004)
286290.
[27] M. Park, M.X. Zhang, M. Chung, G.B. Less, A.M. Sastry, A review of conduction
phenomena in Li-ion batteries, J. Power Sources 195 (2010) 79047929.
[28] M. Guo, R.E. White, Thermal model for lithium ion battery pack with mixed
parallel and series conguration, J. Electrochem. Soc. 158 (2011) A1166A1176.
[29] F. Yang, Interaction between diffusion and chemical stresses, Mater. Sci. Eng. A
409 (2005) 153159.
[30] F.X. Hart, J.B. Bates, Lattice model calculation of the strain energy density and
other properties of crystalline LiCoO2 , J. Appl. Phys. 83 (1998) 75607566.
[31] M. Qu, W.H. Woodford, J.M. Maloney, W. Carter, Y.M. Chiang, K.J. van Vliet,
Nanomechanical quantication of elastic, plastic, and fracture properties of
LiCoO2 , Adv. Energy Mater. 2 (2012) 940944.
[32] W.H. Woodford, W.C. Carter, Y.M. Chiang, Design criteria for electrochemical shock resistant battery electrodes, Energy Environ.l Sci. 5 (2012)
80148024.
[33] Y. Song, X. Shao, Z. Guo, J. Zhang, Role of material properties and mechanical constraint on stress-assisted diffusion in plate electrodes of lithium ion
batteries, J. Phys. D: Appl. Phys. 46 (2013) 105307.

You might also like