You are on page 1of 8

Journal of The Electrochemical

Society

Numerical Simulation of Intercalation-Induced Stress in Li-Ion Battery


Electrode Particles
To cite this article: Xiangchun Zhang et al 2007 J. Electrochem. Soc. 154 A910

View the article online for updates and enhancements.

This content was downloaded from IP address 203.199.213.6 on 18/11/2020 at 06:21


A910 Journal of The Electrochemical Society, 154 共10兲 A910-A916 共2007兲
0013-4651/2007/154共10兲/A910/7/$20.00 © The Electrochemical Society

Numerical Simulation of Intercalation-Induced Stress in Li-Ion


Battery Electrode Particles
Xiangchun Zhang,a,* Wei Shyy,b and Ann Marie Sastrya,c,d,**,z
a
Department of Mechanical Engineering, bDepartment of Aerospace Engineering, cDepartment of
Biomedical Engineering, and dDepartment of Material Science and Engineering, University of Michigan,
Ann Arbor, Michigan 48109-2125, USA

Severe, particle-level strains induced during both production and cycling have been putatively linked to lifetime limiting damage
in lithium-ion cells. Because of the presently unknown contributions of manufacturing and intercalation induced stresses in Li
cells, this correlation is critical in determining optimal materials and manufacturing methods for these cells. Both global and
localized loads must be estimated, in order to select materials able to resist fracture. Here, we select the LiMn2O4 system for study.
We present results of a set of simulation techniques, ranging from one-dimensional finite difference simulations of spherical
particles, to fully three-dimensional 共3D兲 simulations of ellipsoidal particles, to systematically study the intercalation-induced
stresses developed in particles of various shapes and sizes, with the latter 3D calculations performed using a commercial finite
element code. Simulations of spherical particles show that larger particle sizes and larger discharge current densities give larger
intercalation-induced stresses. Simulations of ellipsoidal particles show that large aspect ratios are preferred to reduce the
intercalation-induced stresses. In total, these results suggest that it is desirable to synthesize electrode particles with smaller sizes
and larger aspect ratios to reduce intercalation-induced stress during cycling of lithium-ion batteries.
© 2007 The Electrochemical Society. 关DOI: 10.1149/1.2759840兴 All rights reserved.

Manuscript submitted February 19, 2007; revised manuscript received May 28, 2007. Available electronically July 31, 2007.

Severe, particle-level strains induced during both production and fracture. Further, the role of localized particle fracture in capacity
cycling have been putatively linked to lifetime limiting damage in fade has been implied, but not quantified, given the general lack of
lithium-ion cells. Intercalation and deintercalation of Li ions into understanding of localized loads in batteries.
cathodic lattices, including LiCoO2,1 LiMn2O4,2 and LiFePO4,3 Thus, the present work is focused on determination of localized
have been postulated to result in fraction inside the particles, as particle stresses in cathode particles. Here, we selected the LiMn2O4
determined by experimentation on model systems. In LiMn2O4, for system, following Ref. 14-18 on battery performance modeling, Ref.
19 and 20 on atomic scale simulations of structure, and Ref. 5 on
example, a 6.5% percent volume change has been reported when
intercalation-induced stress simulation because of the low cost and
Mn2O4 is lithiated into LiMn2O4;4 simulations of LiMn2O4 sug- environmental safety of LiMn2O4. We have the following objectives
gested that intercalation-induced stress may exceed the ultimate in this study:
strength of the material.5 Also, stress generation due to cell-scale
loads by compression during manufacturing has been shown to re- 1. To determine diffusion-induced stresses according to an anal-
sult in localized particle stresses that are much higher in the graphite ogy to thermal stress, following Ref. 11-13 for single particles, and
anode material6 共the ratio between local and global stresses is determine the correspondence with prior work5 in Li cells;
around 25 to 140兲. Indeed, stresses of these orders exceed the known 2. To verify the implementation of a single-particle model nu-
strengths of these materials including the most commonly used, and merically, using a finite difference scheme and validation of simple
most promising, cathode materials 关Table I 共Ref. 7, 8, 4, and 9兲兴. results; and
Stress generation due to Li-intercalation, and more generally in 3. To implement this model into a full finite element scheme,
other processes, has been modeled in prior work at the particle scale. and simulate stresses induced by intercalation in particles of non-
spherical geometry.
Christensen and Newman estimated stress generation in lithium in-
sertion into the carbon anode10 and LiMn2O4 cathode5 particles.
More broadly, stresses induced by species diffusion have been stud- Simulation
ied in other fields, including metal oxidation and semiconductor
Stress-strain relations.— For intercalation processes, the lattice
doping. Prussin11 first treated diffusion-induced stress by analogy to constants of the material may be assumed to change linearly4 with
thermal stress. In this study, stress generation during doping of bo- the volume of ions inserted, which results in stresses. Therefore, one
ron and phosphorus into silicon wafer was studied. Li12 studied can calculate intercalation-induced stress by analogy to thermal
diffusion-induced stress or chemical stress in elastic media of simple stress. Prussin11 previously treated concentration gradients analo-
geometries following this method as well.12 Yang13 studied the evo- gously to those generated by temperature gradients in an otherwise
lution of chemical stress in a thin plate considering the interaction unstressed body.
between chemical stress and diffusion following the thermal stress Stress-strain relations including thermal effects are written clas-
analogy by Prussin.11 sically for an elastic body21 as
Though these sets of efforts offer a means of stress estimation at 1
the particle scale, by different physical assumptions, the implemen- ␧xx − ␣T = 关␴xx − ␯共␴yy + ␴zz兲兴 关1a兴
tations to date have not been applied to the problem of three- E
dimensional stresses. Because of the presently unknown contribu-
tions of manufacturing and intercalation-induced stresses in Li cells,
this correlation is critical in determining optimal materials and Table I. Stress and strain in cathode materials in the intercala-
manufacturing methods for these cells. Both global and localized tion process.
loads must be estimated in order to select materials able to resist
Material Measurement technique Stress or strain
LiCoO2 film 共Ref. 7兲 Laser beam deflection ⬃1 GPa 共stress兲
* Electrochemical Society Student Member. LiMn2O4 film 共Ref. 8兲 Laser beam deflection ⬃0.64 GPa 共stress兲
** Electrochemical Society Active Member. LiMn2O4 共Ref. 4兲 Neutron-diffraction 0.027 共strain兲
z
E-mail: amsastry@umich.edu LiFePO4 共Ref. 9兲 X-ray diffraction 0.022 共strain兲
Journal of The Electrochemical Society, 154 共10兲 A910-A916 共2007兲 A911

1 The electrochemical potential in an ideal solid solution can be


␧yy − ␣T = 关␴yy − ␯共␴xx + ␴zz兲兴 关1b兴 expressed as13,24
E
␮ = ␮0 + RT ln X − ⍀␴h 关11兴
1
␧zz − ␣T = 关␴zz − ␯共␴xx + ␴yy兲兴 关1c兴 where ␮0 is a constant, R is gas constant, T is absolute temperature,
E X is the molar fraction of lithium ion, ⍀ is partial molar volume of
lithium ion, and ␴h is the hydrostatic stress, which is defined as
␴xy ␴yz ␴xz
␧xy = ␧yz = ␧xz = 关1d兴 ␴h = 共␴11 + ␴22 + ␴33兲/3 共where ␴ij are the elements in stress ten-
2G 2G 2G sor兲. Equations 10 and 11 show that the diffusion flux depends on
where ␧ij are strain components, ␴ij are stress components, E is concentration, temperature, and stress field. Substitution of Eq. 11
Young’s modulus, ␯ is Poisson’s ratio, G is modulus of elasticity in into Eq. 10, assuming temperature is uniform, and noting that
shear, ␣ is thermal expansion coefficient, and T is the temperature 1 1
change from the original value. Analogously, the stress-strain rela- ⵜ共RT ln X兲 = RT ⵜ X = RT ⵜ c 关12兴
tion with the existing of concentration gradients can be written as13 X c
an expression of species flux 共when there is no temperature gradient
1 c̃⍀
␧ij = 关共1 + ␯兲␴ij − ␯␴kk␦ij兴 + ␦ij 关2兴 inside the particle兲 can be obtained as

冉 冊
E 3
⍀c
where c̃ = c − c0 is the concentration change of the diffusion species J = − D ⵜc − ⵜ ␴h 关13兴
RT
from the original 共stress-free兲 value, and ⍀ is partial molar volume
of solute. Equation 2 can be rewritten to obtain the expression for where D = MRT is diffusion coefficient. Conservation of species
the components of stresses gives
␴ij = 2␮␧ij + 共␭␧kk − ␤c̃兲␦ij 关3兴 ⳵c
+ ⵜ ·J=0 关14兴
where ␮ = E/2共1 + ␯兲, ␭ = 2␯␮/共1 − 2␯兲, and ␤ = ⍀共3␭ + 2␮兲/3. ⳵t
As usual in elasticity, the strain tensor is related to displacement u Then, substituting Eq. 13 into Eq. 14 gives finally
as21

␧ij = 冉
1 ⳵ ui ⳵ u j
+ 冊 关4兴
⳵c
⳵t

= D ⵜ 2c −

RT
ⵜ c · ⵜ ␴h −
⍀c 2
RT
ⵜ ␴h 冊 关15兴
2 ⳵ x j ⳵ xi
as the governing equation for the diffusion process. The initial con-
and the equilibrium equation, neglecting body forces, is21 dition is c = c0, with the boundary condition
␴ij,i = 0 共j = 1,2,3兲
Substitution of Eq. 3 and 4 into Eq. 5 leads to the displacement
关5兴
J = − D ⵜc −冉 ⍀c
RT
ⵜ ␴h =
in
F
冊 关16兴
equations22
where in is the current density on the particle surface 共which is
␮ⵜ2ui + 共␭ + ␮兲uk,ki − ␤c̃,i = 0 共i = 1,2,3兲 关6兴 assumed to be a constant, known value in this study兲, and F is
The boundary condition for the case of a single particle is that the Faraday’s constant.
particle surface is traction-free. This condition can be expressed as22
pnx = ␴xxl + ␴yxm + ␴zxn = 0 关7a兴 Numerical methods.— Finite difference method for 1D prob-
lem.— For the case of a spherical particle, the above equations be-
pny = ␴xyl + ␴yym + ␴zyn = 0 关7b兴 come one-dimensional. The stress tensor contains two independent
components, radial stress ␴r and tangential stress ␴t. The equilib-
pnz = ␴xzl + ␴yzm + ␴zzn = 0 关7c兴 rium equation 共refer to Eq. 5兲 for this case is simply
where l,m,n denote the direction cosines between the external nor- d␴r 2
mal and each axis. Substitution of Eq. 3 and 4 into boundary con- + 共␴r − ␴t兲 = 0 关17兴
dr r
ditions Eq. 7 yields
and the stress-strain relations 共referring to Eq. 2兲 are
␮共ui,j + u j,i兲n j + 共␭uk,k − ␤c兲ni = 0 i = 1,2,3 关8兴
1 ⍀
where n1 = l, n2 = m and n3 = n. Therefore, we are left to solve Eq. ␧r = 共␴r − 2␯␴t兲 + c̃ 关18兴
6, with the boundary condition of Eq. 8. E 3

1 ⍀
Diffusion equation.— As shown in Eq. 2 and 3, concentrations ␧t = 关␴t − ␯共␴r + ␴t兲兴 + c̃ 关19兴
E 3
are needed to calculate intercalation-induced stresses. To obtain a
concentration profile, the insertion and extraction of ions are mod- The strain-displacement relations 共referring to Eq. 4兲 are
eled as a diffusion process. The effect of existing electrons in the
solid on the species flux of lithium can be neglected, because elec- du u
␧r = ␧t = 关20兴
trons are much more mobile than intercalated atoms.23 The chemical dr r
potential gradient is the driving force for the movement of lithium and the displacement equation 共refer to Eq. 6兲 is
ions. The velocity of lithium ions can be expressed as
d2u 2 du 2u 1 + ␯ ⍀ dc̃
v=−Mⵜ␮ 关9兴 + − 2 = 关21兴
dr2 r dr r 1 − ␯ 3 dr
where M is the mobility of lithium ions, and ␮ is the chemical
potential. The species flux can then be written as23 Integration of this equation yields a solution for u, from which
stresses may be obtained. Noting that stresses are finite at the center
J = cv = − Mc ⵜ ␮ 关10兴 of the sphere 共r = 0兲, and that radial stresses are zero, ␴r = 0, at the
where c is the concentration of the diffusion component 共lithium particle surface 共r = r0兲, the two constants in the solution can be
ions兲. determined as
A912 Journal of The Electrochemical Society, 154 共10兲 A910-A916 共2007兲

␴r =
2⍀E 1
3共1 − ␯兲 r30 冉冕 r0

0
c̃r2dr −
1
r3
冕 冊 0
r
c̃r2dr 关22兴
Table II. Material properties of Mn2O4.

Parameter Symbol and dimensions Value


and Young’s modulus E 共GPa兲 10 共Ref. 25兲

␴t =
⍀E 2
3共1 − ␯兲 r30 冉冕 r0

0
c̃r2dr +
1
r3
冕0
r
c̃r2dr − c̃ 冊 关23兴
Poisson’s ratio
Diffusion coefficient
Partial molar volume
Stoichiometric maximum

D 共m2 /s兲
⍀ 共m3 /mol兲
cmax 共mol/m3兲
0.3 共Ref. 25兲
7.08 ⫻ 10−15 共Ref. 5兲
3.497 ⫻ 10−6
2.29 ⫻ 104
Equation 22 shows that radial stress actually depends upon the dif- concentration
ference between the global and local averages of concentration.
The diffusion equation is 共referring to Eq. 15兲
⳵c
⳵t
=D
⳵ 2c
⳵r

2 +
2⳵ c

⍀ ⳵ c ⳵ ␴h ⍀c ⳵2␴h 2 ⳵ ␴h
r ⳵ r RT ⳵ r ⳵ r

RT ⳵ r2
+
r ⳵r
冉 冊册 step for the terms in the two parentheses on the right side. The
Crank-Nicolson method is used for other terms. The difference
equation obtained is
关24兴
ĉin+1 − ĉin 共ĉi+1
n+1 n+1
+ ĉi−1 − 2ĉin+1兲 + 共ĉi+1
n n
+ ĉi−1 − 2ĉin兲
Equations 22 and 23 allow calculation of hydrostatic stress as = 共1 + ␪ˆ ĉin兲

冉冕 冊
2
r0 ⌬t̂ 2共⌬r̂兲


2⍀E 3
␴h = 共␴r + 2␴t兲/3 = 2
c̃r dr − c̃ 关25兴 2 n
ĉi+1 − ĉi−1 n
9共1 − ␯兲 r30 0 + + ␪ˆ
r̂i 2⌬r̂


By assuming that the characteristic time for elastic deformation of
solids is much smaller than that for atomic diffusion, the elastic 2 ˆ n 共ĉi+1
n+1 n+1
− ĉi−1 兲 + 共ĉi+1
n n
− ĉi−1 兲
+ ␪ĉi 关30兴
deformation can be treated as quasistatic.13 Therefore, Eq. 25 can be r̂i 2共2⌬r̂兲
substituted into Eq. 24 to obtain

冋 冉 冊 冉 冊册
Terms including 1/r̂ will be singular at the particle center r̂ = 0. To
⳵c ⳵ 2c 2⳵ c ⳵c 2
⳵ 2c 2⳵ c solve this difficulty, noting that
=D + +␪ + ␪c + 关26兴
⳵t ⳵ r2 r ⳵ r ⳵r ⳵ r2 r ⳵ r ⳵ ĉ
where ␪ = 共⍀/RT兲关共2⍀E兲/9共1 − ␯兲兴. =0 when r̂ = 0 关31兴
⳵ r̂
Substituting Eq. 25 into boundary conditions Eq. 16, one has
L’Hopital’s rule
⳵ c in
J = − D共1 + ␪c兲 = at r = r0 关27兴 1 ⳵ ĉ ⳵2ĉ
⳵r F lim = 关32兴
r̂→0 r̂ ⳵ r̂ ⳵ r̂2
In this way, the two variables, concentration and stress, are decou-
pled. can be used to eliminate the 1/r̂ factor. Thus, Eq. 28 becomes
To solve the above equation numerically, it is, along with bound-
ary and initial condition, transformed into dimensionless form first
as
⳵ ĉ
⳵ t̂
⳵2ĉ
= 共3 + 3␪ˆ ĉ兲 2 + ␪ˆ
⳵ r̂
⳵ ĉ ⳵ ĉ
⳵ r̂ ⳵ r̂
冉 冊 关33兴

⳵ ĉ
⳵ t̂
=
⳵2ĉ
⳵ r̂ 2
+
2 ⳵ ĉ ˆ ⳵ ĉ
r ⳵ r̂
+␪
⳵ r̂
冉 冊 冉 2
+ ␪ˆ ĉ
⳵2ĉ
⳵ r̂ 2
+
2 ⳵ ĉ
r̂ ⳵ r̂
冊 关28兴
which has no singularity at r̂ = 0. Therefore, Eq. 33 will be solved at
r̂ = 0 while Eq. 29 is solved elsewhere.
At two boundary points, imaginary points 共out of the boundary兲
are used to discretize the governing equation; the concentration val-
0 ⱕ r̂ ⱕ 1 0 ⱕ t̂ ⱕ T̂ 关where T̂ satisfies ĉ共r̂ = 1,t̂ = T̂兲 = 1兴 ues of these imaginary points are obtained by central differencing of
the flux boundary condition at the boundary points.
⳵ ĉ ⳵ ĉ The Thomas algorithm is used to solve the tridiagonal system of
r̂ = 1 − 共1 + ␪ˆ ĉ兲 =I r̂ = 0 =0
⳵ r̂ ⳵ r̂ the difference equations. The simulation is halted when the concen-
tration on the particle surface r̂ = 1 reaches the stoichiometric maxi-
t̂ = 0 ĉ = c0 /cmax mum.
where dimensionless variables are defined as Finite element method for 3D problem.— The 3D problem was
simulated using FEMLAB 共COMSOL Multiphysics兲. Two models are
r tD c i nr 0 included in the multiphysics simulation, PDE 共partial differential
r̂ = t̂ = ĉ = ␪ˆ = ␪cmax I=
r0 r20 cmax DcmaxF equation兲 model 共general form兲 and solid stress-strain model. In
PDE model, the diffusion process is described by the generalized
In the above equations, cmax is the stoichiometric maximum concen- form
tration and c0 is the initial concentration. It may be seen that the
effect of discharge current density, particle radius, and diffusion co- ⳵c
+ ⵜ ·⌫=0 关34兴
efficient are all combined into the dimensionless current density I. ⳵t
The numerical procedure is as follows. For each time step, con-
where
centration distribution is solved first by Eq. 28. Then, the concen-
tration is substituted into Eq. 22 and 23 to calculate stresses. Equa-
tion 28 is a nonlinear, parabolic partial differential equation. The
finite difference method is used here to solve the equation.
⌫ = − D ⵜc − 冉 ⍀c
RT
ⵜ ␴h 冊 关35兴

First, Eq. 28 is rewritten as In the solid stress-strain model, “thermal expansion” is included as a

冉 冊
load based on the variable of concentration c instead of temperature
⳵ ĉ ⳵2ĉ 2 ⳵ ĉ 2␪ˆ ĉ ⳵ ĉ in thermal stress calculation.
= 共1 + ␪ĉ兲 + + ␪ˆ + 关29兴
⳵ t̂ ⳵ r̂2 r̂ ⳵ r̂ r̂ ⳵ r̂
Material properties.— All the material properties used in the
To discretize the differential equation into difference equations, the simulation for Mn2O4 are listed in Table II.5,25 From Eq. 2, we see
problem is linearized by taking the value from the previous time that that partial molar volume plays a role analogous to a thermal
Journal of The Electrochemical Society, 154 共10兲 A910-A916 共2007兲 A913

Figure 2. Maximum dimensionless radial stress vs dimensionless current


Figure 1. Comparison of simulation results of two models. density.

expansion coefficient, in calculating intercalation-induced stress. To teries, because it reduces the utilization of the material. Therefore,
obtain the value for this property, the volume change of 6.5% for only the increasing branch of the curve is actually feasible. The
y = 0.2 to y = 0.995 of LiyMn2O4 is used.5 The volume change of increasing branch shows that increase of discharge current density
6.5% gives a strain of 0.0212, which corresponds to the concentra- and particle radius will increase the intercalation-induced stress. In
tion change as y = 0.2 to y = 0.995. Therefore, partial molar volume other words, smaller particles should be used to reduce intercalation-
is, by noting the analogy between thermal expansion coefficient and induced stresses.
⍀/3 As mentioned earlier, the model used here to simulate the
intercalation-induced stress is a diffusion-stress coupling model. The
0.0212 ⫻ 3 effect of stress on diffusion will be discussed briefly using the 1D
⍀= = 3.497 ⫻ 10−6 m3 /mol
共0.995 − 0.2兲cmax equations for a spherical particle. Substituting Eq. 25 into Eq. 13,
we obtain
Results and Discussion ⳵c
J = − D共1 + ␪c兲 关37兴
1D finite difference simulations.— Christensen and Newman 5 ⳵r
modeled the stress generated in LiyMn2O4 during lithium intercala- In Eq. 37, ␪c is always a positive number, and the effective diffusion
tion on the 4 V plateau 共0.2 ⬍ y ⬍ 1兲. The same parameters and coefficient is essentially D共1 + ␪c兲 ⬎ D. Therefore, the diffusion is
properties are used here, except for the diffusion coefficient. In their enhanced due to the extra term ␪c, which basically comes from the
simulation, they used a state of charge-dependent diffusion coeffi- hydrostatic stress gradient term in Eq. 13. In other words, stress
cient, which includes a binary interaction parameter and a thermo- enhances the diffusion. This stress enhancement effect is also dem-
dynamic factor. Here, a constant diffusion coefficient, taking the onstrated numerically, as shown in Fig. 3. It shows the concentration
value of the reference binary interaction parameter in their paper, is profile at t = 1000 s with discharge current density i = 2 A/m2 on
used. The simulation results from the thermal stress analogy model the surface. The profile including the effect of stress has a smaller
and the Christensen and Newman model are shown in Fig. 1. Al- gradient than that excluding the stress effect, confirming that stress
though different approaches are applied to calculate the enhances diffusion.
intercalation-induced stress, the results qualitatively show the same Substituting the material properties into the expression of ␪
trend. = 2⍀2E/关9共1 − ␯兲RT兴, we obtain ␪ = 1.557 ⫻ 10−5 m3 /mol. If the
We used the 1D model to simulate cycling of the active material
maximum concentration is used, ␪cmax = 0.356, which is not negli-
between y = 0 and y = 1, giving an initial condition for Eq. 26 of
c0 = 0. Results show that that maximum radial stress is located at
the center of the particle. The magnitude of the spatial maximum
dimensionless radial stress is given by

␴ˆ r,max =
␴r,max
E
=
2⍀cmax
3共1 − ␯兲 冉冕0
1
ĉr̂2dr̂ − 冏 冏冊
1
3

r̂=0
关36兴

Figure 2 shows how dimensionless maximum radial stress ␴ˆ r,max


共both temporally and spatially during the discharge process兲 varies
with dimensionless current density 共or dimensionless boundary flux兲
I.
As shown in Fig. 2, maximum radial stress 共spatially and tem-
porally兲 inside an electrode particle during the discharge process
increases with increasing dimensionless current density when 0
⬍ I ⬍ 2.7. However, maximum radial stress decreases with increas-
ing dimensionless current density when it is larger than 2.7. The
decrease in stress is attributable to the fact that the concentration
profile is not fully developed, so that the global average term 共first
term in the parentheses兲 in Eq. 36 decreases with dimensionless
current density, while the local average 共second term in the paren-
thesis兲 remains constant. This is not desirable in the cycling of bat- Figure 3. Numerical results for the effects of stress.
A914 Journal of The Electrochemical Society, 154 共10兲 A910-A916 共2007兲

Figure 5. Schematic of an ellipsoidal particle, with coordinate system.

Characteristic solution profiles of concentration, von Mises stress


and shear stress ␴yz, are shown in Fig. 6 at the end of the discharge
process 共when the surface concentration reaches the stoichiometric
maximum兲 for an ellipsoid with aspect ratio 1.953. Figure 6 shows
that 共i兲 the concentration is higher around the poles, 共ii兲 the von
Mises stress is larger around the equator, and 共iii兲 shear stress has its
maximum on the surface. The solution profiles have the same pat-
terns for other ellipsoids with different aspect ratios.
Figure 7 shows how the maximum von Mises stress inside the
particle varies during the discharge process, for particles with dif-
ferent aspect ratios. It takes less time for particles with larger aspect
ratios to completely discharge. Also, during discharge, von Mises
stress increases first, and then drops. In Fig. 7, it can be observed
that when aspect ratio increases, the stress increases first 共for aspect
ratios from 1.0 to 1.37兲 and then decreases 共for aspect ratios from
1.37 to 3.81兲. For ellipsoids with aspect ratio 2.92 and 3.81, the
intercalation-induced stress is smaller than that inside a sphere 共as-
pect ratio 1.0兲.
Figure 8 shows how aspect ratio affects 共a兲 peak value of maxi-
mum von Mises stress, and 共b兲 peak value of maximum shear when
the volumes of particles are fixed. Figure 8a shows that peak value
of maximum von Mises stress inside the particle increases first and
then decreases, as aspect ratio increases. For aspect ratios larger than
Figure 4. Convergence plot of finite element solutions for 共a兲 hydrostatic 2.2, maximum von Mises stress decreases to less than that inside
stress and 共b兲 concentration. spherical particles 共aspect ratio 1兲. Figure 8b shows that peak value
of maximum shear stress decreases as aspect ratio increases. The
results of Fig. 8 show that larger aspect ratios reduce the
intercalation-induced stresses, over particles of lower aspect ratio
when volume is preserved.
gible compared to unity. Therefore, stress effect cannot be neglected
The peak values of maximum von Mises stresses are shown in
here for the case of LiMn2O4. From the expression for ␪, it can be
Fig. 8a. Maximum stress first increases, then decreases with aspect
observed that ␪ has smaller magnitude when the material has
ratio. This is due to two competing effects. When particle volume is
smaller modulus E and smaller partial molar volume ⍀. Thus, the preserved, increased aspect ratios result in increase of the longer
stress effect on diffusion may be negligible when the material is soft
semiaxis c, and reduction of shorter semiaxes a and b. Elongation of
共i.e., having a low modulus兲.
the longer semiaxis tends to increase maximum stress, while reduc-
tion of the shorter semiaxes tends to decrease the maximum stress.
3D finite element simulation results.— The 1D finite difference This competition results in a global maximum of stress at an aspect
simulation, with 4001 grid points and a time step of 0.001 s, was ratio of ⬃1.37.
used as the reference solution to study the convergence of the finite To further illustrate the effect of semiaxes on maximum stress, an
element method. Figure 4 shows the 2-norm errors 共differences兲 additional set of simulations was performed in which the shorter
between the finite element solutions and finite difference reference semiaxes a and b were fixed, and aspect ratio ␣ was increased by
solutions at t = 1000 s. The parameters used in the simulations are elongation of the longer semiaxis, c. Results obtained with a dis-
current density i = 2 A/m2, and particle radius r0 = 5 ␮m. The fi- charge current density i = 2 A/m2 are shown in Fig. 9. Stress first
nite element solutions converged to the reference solution as the increases with aspect ratio because of the increase of the longer
number of elements used increased. At the same time, Fig. 4 also semiaxis, and then decreases slightly until asymptotically approach-
shows that solutions from 1D finite difference method and 3D finite ing the cylinder-fiber limit, i.e., ␣ → ⬁. As represented by the
element method were consistent, because the nondimensionalized dashed line in Fig. 9, no physically relevant solutions are obtained
errors of concentration and stress from 17,359 elements simulation for ␣ ⬎ 7.9, because the discharge process stops when the concen-
were 6.5 ⫻ 10−7 and 1.5 ⫻ 10−5, respectively 共if nondimensional- tration on the particle surface reaches stoichiometric maximum, be-
ized by the maximum values at t = 1000 s inside the particle兲. fore the maximum stress actually reaches the peak value. To quan-
To study the effect of aspect ratios on the intercalation-induced titatively illustrate this point, for an ellipsoid with aspect ratio 10,
stress, ellipsoids with different aspect ratios were studied. The cur- the stress reaches its peak value at t = 721 s. However, the simula-
rent density on the surface is fixed at i = 2 A/m2. For the ellipsoid, tion should terminate at t = 617.5 s, when the surface concentration
the lengths of three semiaxes a, b, and c satisfy a = b, and the already reaches stoichiometric maximum. The maximum stress at
aspect ratio is defined as ␣ = c/a, as sketched in Fig. 5. The vol- t = 721 s is 52.47 MPa, and the stress at t = 617.5 s is 52.26 MPa.
umes of the ellipsoids were fixed at V = 4␲ ⫻ 53 /3 ␮m3. A set of Therefore, the stress when the process is terminated is only slightly
simulations, with different aspect ratios, was run by FEMLAB. smaller than the peak value.
Journal of The Electrochemical Society, 154 共10兲 A910-A916 共2007兲 A915

Figure 7. Maximum von Mises stress during discharge, for various ellip-
soids.

that large aspect ratios are preferred to reduce the intercalation-


induced stresses. In total, these results suggest that it is desirable to
synthesize electrode particles with smaller sizes and larger aspect
ratios to reduce intercalation-induced stress during cycling of
lithium-ion batteries.
Acknowledgments
The authors gratefully acknowledge the support of our sponsors,
including the U.S. Department of Energy through the BATT pro-
gram 共Dr. Tien Duong, Program Manager兲, the Army Research Of-

Figure 6. Solutions at the end of discharge for an ellipsoid of aspect ratio


1.953. 共a兲 Concentration, 共b兲 von Mises stress, and 共c兲 shear stress ␴yz. 共Di-
mensions in µm.兲

Conclusion
Intercalation-induced stresses during the discharge process were
simulated in this study using a stress-diffusion coupling model.
Intercalation-induced stresses were simulated by analogy to thermal
stress. Simulations of spherical particles show that larger particle
sizes and larger discharge current densities give larger intercalation-
induced stresses. Furthermore, internal stress gradients significantly
enhance diffusion. Simulation results for ellipsoidal particles show Figure 8. The effect of aspect ratio, for fixed particle volume.
A916 Journal of The Electrochemical Society, 154 共10兲 A910-A916 共2007兲

Greek

␣ aspect ratio
␧ij strain
␮ chemical potential, J/mol
␯ Poisson’s ratio
␴ij stress, Pa
⍀ partial molar volume, m3 /mol
Subscript

h hydrostatic 共stress兲
max maximum
r radial direction
t tangential direction

Others

∧ dimensionless variables

References
1. H. Wang, Y.-I. Jang, B. Huang, D. R. Sadoway, and Y.-M. Chiang, J. Electrochem.
Figure 9. The effect of aspect ratio, for fixed shorter semiaxes. Soc., 146, 473 共1999兲.
2. M.-R. Lim, W.-I. Cho, and K.-B. Kim, J. Power Sources, 144, 3496 共1997兲.
3. D. Wang, X. Wu, Z. Wang, and L. Chen, J. Power Sources, 140, 125 共2005兲.
4. W. I. F. David, M. M. Thackeray, L. A. de Picciotto, and J. B. Goodenough, J.
fice 共Dr. Bruce LaMattina, Program Manager兲, and the Ford Motor Solid State Chem., 67, 316 共1987兲.
5. J. Christensen and J. Newman, J. Electrochem. Soc., 153, A1019 共2006兲.
Company 共Ted Miller and Kent Snyder, Program Managers兲. The 6. Y.-B. Yi, C.-W. Wang, and A. M. Sastry, J. Eng. Mater. Technol., 128, 73 共2006兲.
authors also thank Professor James R. Barber and Dr. Chia-Wei 7. S.-I. Pyun, J.-Y. Go, and T.-S. Jang, Electrochim. Acta, 49, 4477 共2004兲.
Wang, both of the Department of Mechanical Engineering, Univer- 8. Y.-H. Kim, S.-I. Pyun, and J.-Y. Go, Electrochim. Acta, 51, 441 共2005兲.
sity of Michigan, Ann Arbor, for helpful discussion and remarks. 9. A. K. Padhi, K. S. Nanjundaswamy, and J. B. Goodenough, J. Electrochem. Soc.,
144, 1188 共1997兲.
University of Michigan assisted in meeting the publication costs of this 10. J. Christensen and J. Newman, J. Solid State Electrochem., 10, 293 共2006兲.
article. 11. S. Prussin, J. Appl. Phys., 32, 1876 共1961兲.
12. J. C. M. Li, Metall. Trans. A, 9A, 1353 共1978兲.
13. F. Yang, Mater. Sci. Eng., A, 409, 153 共2005兲.
14. M. Doyle, J. Newman, A. S. Gozdz, C. N. Schmutz, and C. M. Tarascon, J.
List of Symbols Electrochem. Soc., 143, 1890 共1996兲.
15. R. Darling and J. Newman, J. Electrochem. Soc., 144, 4201 共1997兲.
a,b,c lengths of the three semi-axes of ellipsoid, m 16. P. Arora, M. Doyle, A. S. Gozdz, R. E. White, and J. Newman, J. Power Sources,
c concentration of lithium ions, mol/m3 88, 219 共2000兲.
c̃ concentration change from initial value, mol/m3 17. E. Deiss, D. Haringer, P. Novak, and O. Haas, Electrochim. Acta, 46, 4185 共2001兲.
D lithium diffusion coefficient, m2 /s 18. R. Darling and J. Newman, J. Electrochem. Soc., 145, 990 共1998兲.
E Young’s modulus, GPa 19. C. Y. Ouyang, S. Q. Shi, Z. X. Wang, H. Li, X. J. Huang, and L. Q. Chen,
F Faraday’s constant, 96,487 C/mol Europhys. Lett., 67, 28 共2004兲.
I dimensionless current density 20. B. Ammundsen, J Roziere, and M. S. Islam, J. Phys. Chem. B, 101, 8156 共1997兲.
in current density, A/m2 21. S. P. Timoshenko and J. N. Goodier, Theory of Elasticity, McGraw-Hill, New York
共1970兲.
J species flux, mol/共m2 s兲
22. N. Noda, R. B. Hetnarski, and Y. Tanigawa, Thermal Stresses, 2nd ed., Taylor &
M mobility, m2 mol/共J s兲 Francis, New York 共2003兲.
R gas constant, 8.314 J共mol K兲 23. W. R. McKinnon and R. R. Haering, in Modern Aspects of Electrochemistry, No.
r0 particle radius, ␮m 15, R. E. White, J. O’M. Bockris, and B. E. Conway, Editors, Plenum Press, New
T temperature, K York 共1983兲.
u displacement, m 24. W. L. Wang, S. Lee, and J. R. Chen, J. Appl. Phys., 91, 9584 共2002兲.
v ion movement velocity inside solid particles, m/s 25. A. Paolone, R. Cantelli, G. Rousse, and C. Masquelier, J. Phys.: Condens. Matter,
X molar fraction of lithium in the electrode 15, 457 共2003兲.

You might also like