Anda di halaman 1dari 14

Electrochimica Acta 55 (2010) 89608973

Contents lists available at ScienceDirect

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

Thermodynamically consistent modeling of elementary


electrochemistry in lithium-ion batteries
Andrew M. Colclasure, Robert J. Kee
Engineering Division, Colorado School of Mines, 1500 Illinois Street, Golden, CO 80401, USA

a r t i c l e

i n f o

Article history:
Received 29 April 2010
Received in revised form 29 June 2010
Accepted 5 August 2010
Available online 12 August 2010
Keywords:
Li-ion battery
Elementary charge transfer
Non-ideal thermodynamics
Modeling

a b s t r a c t
This paper is particularly concerned with the elementary reactions and transport processes that are
responsible for Li-ion battery performance. The model generally follows the widely practiced approach
developed by Newman and co-workers (e.g., Doyle et al., J. Electrochem. Soc. 140 (1993) 1526 [1]).
However, there are signicant departures, especially in modeling electrochemical charge transfer. The
present approach introduces systems of microscopically reversible reactions, including both heterogeneous thermal reactions and electrochemical charge-transfer reactions. All reaction rates are evaluated in
elementary form, providing a powerful alternative to a ButlerVolmer formalism for the charge-transfer
reactions. The paper is particularly concerned with the inuence of non-ideal thermodynamics for evaluating reversible potentials as well as charge-transfer rates. The theory and modeling approach establishes
a framework for extending chemistry models to incorporate detailed reaction mechanisms that represent
multiple competitive reaction pathways.
2010 Elsevier Ltd. All rights reserved.

1. Introduction
Lithium-ion batteries are certainly making signicant impacts
on a range of applications from small portable devices to automotive traction. Fig. 1 illustrates the structure of a cylindrical Li-ion
cell over a range of length scales. The objective of the present
effort is to develop models at the electrode-particle scale that
incorporate fundamental thermodynamics and chemical kinetics.
Charge-transfer processes within lithium-ion batteries are usually
modeled using a ButlerVolmer representation [13]. The present
approach provides a potentially powerful alternative to representing electrochemical charge-transfer rates via global ButlerVolmer
expressions. Representing chemistry in terms of elementary reactions enables the study of competitive reaction pathways. For
example, potentially deleterious Li plating on anode surfaces competes with desired Li-ion intercalation into the electrode particle.
Details of the chemistry associated with solid electrolyte interfaces
(SEI) can be incorporated in the reaction mechanism [4,5]. Although
the present effort concentrates upon intercalation electrodes, the
fundamental approach to modeling kinetics is well suited to modeling conversion electrodes.
Before developing the details of the model, it is helpful to
review the broad concepts of Li-ion battery function. Consider the
electrode-particle scale (Fig. 1), with a composite anode on the

Corresponding author. Tel.: +1 303 273 3379; fax: +1 303 273 3602.
E-mail address: rjkee@mines.edu (R.J. Kee).
0013-4686/$ see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.electacta.2010.08.018

left, a separator in the center, and a composite cathode on the


right. The composite electrodes consist of active particles, binder,
and conductive connectors, all of which are surrounded by the
lithium-ion-conducting electrolyte phase. Lithiated graphite (Lix C6 ,
where 0.1 < x < 0.8) is commonly used as the active material in the
anode (negative electrode). The cathode (positive electrode) active
material is often a metal oxide (e.g., Liy CoO2 where 0.4 < y < 1.0). A
microporous separator lm electrically isolates the cathode from
the anode. The electrolyte phase is contained within the pores of
the separator, which enables ion conduction across the separator.
During discharge intercalated Li leaves anode particles and
enters the electrolyte as Li+ . This charge-transfer process delivers
electrons to a current collector and an external load. The Li ions
are transported via diffusion and migration through the electrically
insulating electrolyte solution toward the cathode. Reacting with
electrons from the external load, a charge-transfer process delivers Li into the cathode particles. During charging the processes are
reversed.
The reversible charge-transfer process at a graphite anode may
be represented globally as
Li(C6 )  V(C6 ) + Li+ (e) + e (a),

(1)

where Li+ (e) is a positively charged lithium ion in the electrolyte


phase, and V(C6 ) is an intercalation vacancy within the graphite.
The electron is in the anode phase, consisting of the graphite, conductive ller that surrounds the graphite, and the current collector.
The reversible charge-transfer process at the cathode (e.g., CoO2 )

A.M. Colclasure, R.J. Kee / Electrochimica Acta 55 (2010) 89608973

8961

chemistry that is primarily responsible for battery function. For


example, Li plating on anode particles can limit charging rates. The
chemistry responsible for solid electrolyte interface (SEI) formation
and degradation can also be incorporated [4,5].
2.1. Electrolyte charge conservation
Charge-conservation equations are solved for the electric potentials of the electrode and electrolyte phases throughout the cell. The
transport equations are derived using concentrated solution theory
which considers both solutesolvent and solutesolute interactions [9]. However, because formulation of the transport equations
follows established practice, the conservation equations are only
summarized here.
Charge conservation within the electrolyte phase is represented
as

(eff e ) + (Deff ln CLi+ (e) ) + As s Li+ (e) F = 0,

Fig. 1. Structure of a Li-ion battery, showing cell-scale and particle-scale structures.

may be represented globally as


V(CoO2 ) + Li+ (e) + e (c)  Li(CoO2 ).

(2)

Here V(CoO2 ) is an intercalation vacancy within a cathode particle and Li(CoO2 ) is an intercalated Li. The electron e (c) is in the
electronically conducting phases of the cathode.
2. Model formulation
The modeling approach builds upon the foundation established
by Newman and co-workers [1,2,6]. Species- and chargeconservation equations are solved to predict temporal and spatial
proles of lithium concentration and electric potentials. Transport
through the composite electrodes is modeled using porouselectrode theory. Thus, the electrolyte and active electrode material
are treated as a superimposed continua [1]. Concentrated solution
theory is used to describe transport within the binary liquid electrolyte.
The present effort incorporates detailed reaction mechanisms
into Newmans framework to describe the charge-transfer kinetics, providing an alternative to the ButlerVolmer representation.
The approach to formulating charge-transfer chemistry builds
upon earlier work in modeling elementary charge-transfer reactions in fuel-cell electrodes [7,8]. Generally speaking, elementary
reactions involve species in three phaseswithin the electrolyte,
on the electrode surfaces, and within the bulk of the electrode. Evaluating reaction rates depends upon the activities of
species and, for charge-transfer reactions, the electric potentials
of the participating phases. To enforce microscopic reversibility,
self-consistent thermodynamic properties are required for all participating species.
The elementary-reaction formulation facilitates the incorporation of reactions among surface-adsorbed species. Such
heterogeneous reactions may compete with the charge-transfer

(3)

where the dependent variables are the electric potential of the


electrolyte phase e and the lithium-ion concentration CLi+ (e) . The
specic surface area As represents the surface area of active electrode particles per unit volume of composite electrode. The net
production rate of lithium ions via charge-transfer reactions is
represented as s Li+ (e) and F is the Faraday constant. Because bulk
velocity of the electrolyte is neglected there is no convective term.
Charge enters and leaves the electrolyte phase via surface
reactions (third term in Eq. (3)). Within composite electrodes an
electric double layer is formed between the electrode particles and
electrolyte. The charge-transfer rates depend upon the electricpotential differences between the electrode and electrolyte phases.
Because the characteristic time for charging and discharging the
double layer is on the order of milliseconds or shorter [6], double layer charging is neglected in the present model. Within the
separator there are no electrode phases, and thus there are no
charge-transfer reactions.
Despite the fact that battery charging and discharging is an
inherently transient problem, Eq. (3) does not include a transient term. In the absence of very large electric elds, electrolyte
charge conservation is maintained by an instantaneously steadystate equation (i.e., Eq. (3)). That is to say there is no capacity to store
charge locally within the electrolyte solution. All charge storage is
within very thin double layers at electrodeelectrolyte interfaces,
which charge and discharge at very high rates.
Eq. (3) is derived assuming that the electrolyte solution maintains charge neutrality. That is,

Ck zk = 0,

(4)

where Ck and zk are the concentration and charge of species k within


the electrolyte phase. The model does not explicitly represent the
electrolytesolution anion concentration. Rather, the anion concentration can be derived from z C = CLi+ (e) , where the subscript
refers to the anion species in the electrolyte phase. In the presence of sufciently large electric elds charge neutrality may not
be preserved. However, such effects are not included in the present
model and are generally not relevant for battery applications.
The effective ionic conductivity of the electrolyte phase eff and
eff are evaluated as
the effective diffusional conductivity D
p

eff = e ,
eff
=
D

2RTeff
F


(t 0 +

Li (e)

1)

1+

d(ln f )
d(ln CLi+ (e) )

(5)


.

(6)

Ion conductivity of the pure electrolyte phase is represented as ,


the volume fraction of the electrolyte phase is e , and p the Bruggeman porosity exponent. To maintain charge neutrality the lithium

8962

A.M. Colclasure, R.J. Kee / Electrochimica Acta 55 (2010) 89608973

ions positive charge must be compensated by a negatively charged


species. Both positively and negatively charged species can produce
ion current. The transference number tk0 is the fraction of current
carried by species k relative to the total current carried by all ionic
species in an electrolyte solution with uniform composition [9,10].
For example, if the lithium-ion transference number t 0 + is unity,

more than one surface site [12]. The site fractions are determined
by solving

charge movement is entirely the result of electric-potential gradients. The mean molar activity coefcient of the electrolyte is f ,
which is assumed to be constant in the present paper.

2.5. Boundary conditions

n k
= s k .
k (n) t

(10)

Li (e)

2.2. Electrolyte species conservation


Lithium-ion concentration within the binary electrolyte solution is represented as
e CLi+ (e)
t

= (Deeff CLi+ (e) ) + (1 t 0 +

Li (e)

)As s Li+ (e) .

(7)

The effective diffusion coefcient is calculated using the Bruggeman expression


p

Deeff = De e .

(8)

The electrolyte diffusion coefcient De is a function of local concentrations, species charges, and binary diffusion coefcients within
the electrolyte phase. The last term represents the effects of lithium
ions entering the electrolyte via heterogeneous charge-transfer
chemistry at the electrode surfaces. In the limit of the lithium-ion
transference being unity, the charge-transfer chemistry does not
affect the electrolyte concentration. Hence the electrolyte composition would be spatially uniform. Conversely, if the transference
number is zero, then the charge-transfer chemistry causes signicant lithium-ion variations throughout the electrolyte [2].
Unlike the charge-conservation equation (Eq. (3)), the speciesconservation equation (Eq. (7)) retains the transient term. Species
transport through the electrolyte proceeds on time scales that are
commensurate with charge and discharge times.

Eqs. (3), (7), and (9) are spatially second order and thus
each require boundary conditions. Fig. 2 summarizes the conservation equations and the boundary conditions, illustrated in a
one-dimensional conguration. In the present model, the same
electrolyte solution is presumed to be used throughout the cell.
Therefore, Eqs. (3) and (7) span the entire cell, although with
different parameters (e.g., diffusion coefcients, conductivities,
porosities, reaction rates, etc.) in the anode, separator, and cathode regions. Composition and ux continuity must be preserved
at the material interfaces. Assuming the current collectors are pure
electronic conductors, the charge and concentration uxes through
the electrolyte must vanish at the current-collector interfaces.
Charge conservation within the solid electrode-particle network
is described by Eq. (9). At the anode current collector, the anodephase electric potential is set to a reference voltage a,cc = Vref . At
the cathode current collector, the cathode-phase electric-potential
gradient is set to meet the specied current density
ie = ceff c .

(11)

The current density ie is positive when the cell is discharging and


negative when the cell is being charged. The total current produced
by the cell is Ie = ie A, where A is the area of the electrode currentcollection surface (e.g., the foil in Fig. 1). The cell operating voltage
is equal to the difference in potential at the current-collector
interfaces Vcell = c,cc a,cc . Assuming that the separator is a pure
ion conductor, the electric current must vanish at the interfaces
between the electrode-particle networks and the separator (e.g.,
n a and n c ).

2.3. Electrode charge conservation


Charge conservation within the solid-phase electrodes is represented as

( eff b ) As ie = 0,

(9)

where b is the bulk electrode-phase electric potential and  eff


is the effective solid-phase conductivity. The current density ie is
the result of charge-transfer reactions delivering electrons into the
electrode phase. The rst term in Eq. (9) represents charge transport via electronic conduction. The second term represents a source
(sink) of charge associated with electrons entering (leaving) the
solid electrode phase via charge-transfer reactions. The effective
solid-phase conductivity can be represented in a variety of ways.
Presently, the effective conductivity is evaluated with the simple
expression  eff =  b b , where b is the solid-phase volume fraction. However, percolation theory may provide the basis for a more
accurate approach [11].
2.4. Surface site conservation
Because site fractions are needed to evaluate species activities,
species-conservation equations must be solved for the local surface
coverages on electrode particles. The surface concentration Ck,s is
related to the surface site fraction k as Ck,s =  k  n / k (n), where
 n is the available site density for phase n and  k (n) is the siteoccupancy number for species k on phase n. For example, organic
compounds that adsorb onto the electrode surfaces can occupy

3. Particle intercalation
For the purposes of modeling lithium intercalation, lithium
transport between particles is neglected. That is, intercalatedlithium diffuses within a particle, but not from particle to particle.
Additionally the electrode particles are presumed to be spherical.
With these assumptions, the intercalated-lithium transport within
an electrode particle can be represented as
CLiI
t

1
= 2
r r


2

r DLiI

CLiI
r


,

(12)

where CLiI is the concentration of intercalated lithium. The diffusion


coefcient of intercalated lithium DLiI can be a strong function of
the concentration of intercalated lithium [13], which renders the
diffusion problem to be nonlinear. The transient term is retained
because the intercalation time scales are commensurate with the
charge/discharge times. In fact, intercalation is often the ratelimiting process in battery functioning.
Solution of Eq. (12) requires an initial condition (concentration
prole) and two boundary conditions. At the center of a spherical
particle the radial gradient vanishes (i.e., CLiI /r = 0). At the particle surface the internal diffusion is balanced by the net production
rate of intercalated lithium via surface reaction,
DLiI

CLiI
r

= s LiI .

(13)

A.M. Colclasure, R.J. Kee / Electrochimica Acta 55 (2010) 89608973

8963

Fig. 2. Section of a lithium-ion cell, showing cell-scale conservation equations and boundary conditions.

4. Chemical kinetics

as

Although charge transfer is central to battery function, the overall charge-transfer process is comprised of both charge-neutral
heterogeneous reactions and charge-transfer reactions. Thus, the
detailed reaction mechanisms that describe the process must
accommodate both types of reactions. Species that participate in
the reactions are presumed to reside in one of three possible phases,
the electrolyte phase, the electrode-surface phase, or the electrodebulk phase. Evaluation of thermodynamic properties, including
chemical activities, depends upon the phase in which the species
resides. In general there are Ke species in the electrolyte phase,
Ks species in the electrode-surface phase, and Kb species in the
electrode-bulk phase.
The species activities are evaluated differently depending upon
the phase in which they reside. It is assumed here that the effective concentration for a species in the electrolyte solution is its
molar volumetric concentration. The activity of an adsorbed species
on an electrode particle is evaluated as the fraction of available
sites occupied (i.e., the site fraction). Evaluating the activities of
the intercalated species and intercalation vacancies is a somewhat
more complex task, which is discussed in subsequent sections.

= (a e ) (a e )

4.1. ButlerVolmer formulation


The rate of a charge-transfer process is usually modeled with
a ButlerVolmer equation, which can apply to an elementary or
a global reaction. In ButlerVolmer form, the current density is
evaluated as

ie = i0 exp

 F 
a
RT

 F 
c

exp

RT

(14)

where i0 is the exchange current density, and a and c are the


anodic and cathodic symmetry factors, respectively. For an elementary reaction a + c = 1. However, for a global reaction there
is no such restriction. The rst exponential term represents the
anodic rate (i.e., producing electrons), and the second represents
the cathodic rate (i.e., consuming electrons). The exchange current
density is a function of temperature and the activities of participating species. For anode particles the overpotential is dened

eq

eq

= (a e ) Ea ,

(15)

where a is the electric potential at the surface of an anode


particle and e is the local electric potential of the electrolyte surrounding the particle. The equilibrium electric-potential difference
eq
eq
Ea = (a e ) is the electric-potential difference at which the
charge-transfer chemistry proceeds at equal and opposite rates
in the anodic and cathodic directions, producing no net current
(i.e., ie = 0). The equilibrium electric-potential difference is usually a function of the intercalated-lithium concentration at the
surface of the electrode particles. However, the expressions used
eq
for Ea usually neglect variations in electrolyte concentration. The
overpotentials within the cathode structure are evaluated with an
analogous expression as
= (c e ) (c e )

eq

eq

= (c e ) Ec ,

(16)

where c is the electric potential at the surface of a cathode particle.


Electrons are produced when > 0 (anodic current) and electrons
are consumed when < 0 (cathodic current). The local electric
potentials are determined from the solutions of Eqs. (3) and (9).
Despite the widespread use of the ButlerVolmer equation to
represent charge-transfer chemistry, there are some limitations
[7]. For example, because ButlerVolmer formulation assumes a
single rate-limiting step, it can be difcult to model competitive
parallel reactions. If the phase surrounding the electrode particle is not in chemical equilibrium, then there is no explicit way
to dene a single reversible potential. Modeling charge-transfer
chemistry in terms of elementary reactions enables the incorporation of multiple parallel pathways. The elementary formulation
can also provide more physical insight. For instance, important
information about the solid electrolyte interface (SEI) can be used
directly in an elementary-reaction mechanism. Such a mechanism
can also incorporate parallel, possibly deleterious reactions, such
as lithium plating that limits charging rate. A drawback, however,
is that it can be difcult to establish all the needed physical and
chemical properties and rate expressions.

8964

A.M. Colclasure, R.J. Kee / Electrochimica Acta 55 (2010) 89608973

4.2. Elementary charge-transfer reactions

4.5. Net rate of progress

Consider a generic elementary charge-transfer reaction as

A(b) + B (e)  C(b) + D(e) + e (b),

(17)

where (b) means the species is in the bulk electrode phase and
(e) means the species is in the electrolyte phase. The bulk electrode phase and surface of the electrode are modeled as having
the same electric potential. By denition a charge-transfer reaction
moves charge from one phase to another. An elementary chargetransfer reaction transfers a single electron. For example, Reaction
(17) moves the charge from the anion B (e) in the electrolyte to
an electron e (b) in the electrode phase. Reaction (17) is written
in the anodic direction, meaning the forward reaction delivers a
negative charge into the electrode phase. The usual case produces
an electron in the electrode phase. However, a negatively charged
species moving from the electrolyte phase to the electrode phase
(e.g., A (e) A (b)) or a positively charged species moving from
the electrode to the electrolyte phase (e.g., B+ (B) B+ (e)) are both
anodic reactions.
The anodic and cathodic rates of progress q depend upon
thermal rate expressions, species activities, and electric-potential
differences as

qa = ka CA aA CB aB exp


qc = kc CC aC CD aD exp

a F(b e )
RT

c F(b e )
RT

(18)


,

(19)

where ak is the (dimensionless) activity of the participating species


k and Ck is its standard-state concentration. Here, it is assumed
that C I = C V = Cmax and C + is the average Li-ion concentraLi

Li (e)

Li

tion within the electrolyte solution (Table 3). The product ak Ck


produces the effective concentration of species k. The forward
(anodic) rate expression is usually written in Arrhenius form as
ka (T) = ATn exp( E/RT). The electric potentials of the electrode
phase (anode or cathode) and the electrolyte phase are represented
as b and e , respectively. The anodic and cathodic symmetry factors are represented as a and c . For an elementary reaction (i.e.,
transferring a single charge), 0 < < 1 and a + c = 1.
4.3. Elementary charge-neutral reactions
The rates of reactions that do not transfer charge do not depend
upon the electric-potential differences between phases. For example, again generically,
A(b) + B(e)  C(b) + D(e).

(20)

The net production rates for species k are found by summing the
rates of progress qi over all reactions I
s k =

I


ki qi ,

(24)

i=1

where ki is the stoichiometric coefcient for species k in the ith


reaction. The net rate of progress for each reaction is evaluated
as the difference between anodic (or forward) and cathodic (or
reverse) rates of progress as
qi = qa qc ,

qi = qf qr ,

For charge-transfer reactions that produce (consume) electrons, the


current density is related to the net production rates as
ie = F s e = F

I


e ,i qi ,

where e ,i is the stoichiometric coefcient of the electron in reaction i.


5. Numerical implementation
Spatial derivatives in the governing conservation equations are
discretized using the nite-volume method. The discretized equations form a set of differential algebraic equations (DAE). The
software implementation is written in C++ and uses the DAE solver
contained in Sundials [14]. For the results shown later in the paper,
the composite-electrode structures each contain 50 nite volumes
and the separator is represented with 5 control volumes. Within
each electrode control volume, a representative spherical intercalation particle is discretized with 10 radial control volumes.
The model uses the Cantera [15] software to evaluate thermodynamic properties and reaction rates. A full constant-current
discharge/charge cycle can be simulated in a few tens of seconds
on a personal computer.
6. Thermodynamics of intercalated lithium
The reaction mechanism requires thermodynamic properties,
and the approach followed here was developed by Karthikeyan et
al. [13].
6.1. Reversible potential
A species chemical potential may be written generally as
k = k + RT ln ak ,

qf = kf CA aA CB aB ,

(21)

where k

qr = kr CC aC CD aD .

(22)

To preserve microscopic reversibility, the reverse (cathodic) rate


expression is related to the forward rate expression through an
equilibrium constant Keq as
CC
ka
= Keq = C D exp
CB CA
kc


G 
RT

(23)

This relationship is valid for both charge-transfer and chargeneutral reactions. Evaluating the equilibrium constant requires
species thermodynamic properties such that
G can be evaluated.

(26)

i=1

In this case, the forward and reverse rates of progress may be written as

4.4. Microscopic reversibility

(25)

(27)

is the standard-state chemical potential. The species electrochemical potential may be written as

k = k + zk F,

(28)

where zk is the species charge and  is the electric potential of the


phase associated with the species.
Measurements of reversible electrode potential as a function of
intercalation fraction can be used to develop thermodynamic properties for intercalated lithium. Consider a generic lithium-insertion
reaction,
LiI (b)  LiV (b) + Li+ (e) + e (b).
LiI (b),

LiV (b),

(29)
Li+ (e)

The species
and
represent: an intercalated
lithium within the bulk phase of the electrode, a lithium vacancy
within the bulk phase of the electrode, and a lithium ion within the

A.M. Colclasure, R.J. Kee / Electrochimica Acta 55 (2010) 89608973

electrolyte phase. At equilibrium the Gibbs free-energy change for


the reaction is zero (
G = 0). Thus,
I + RT ln aLiI = V + RT ln aLiV + +
Li

Li (e)

Li

+ e

+ RT ln aLi+ (e) + Fe

+ RT ln ae Fb .

(30)

By convention, the chemical potential of an electron inside the bulk


electrode phase is set to zero. Based upon these assumptions, the
equilibrium electric-potential difference is
E eq = b e =

G29

RT
ln
F

aLiV aLi+ (e)


aLiI

(31)

where
G is the change in Gibbs free energy for Reaction (29).

The electric potential of the electrolyte phase is not easily measured. Therefore, the potential of the intercalation electrode is
usually measured relative to a reference pure lithium-metal electrode. The global reaction at the reference lithium electrode is
Li  Li+ (e) + e ,

(32)

where Li is bulk lithium metal. The overall reaction occuring when


the reversible potential for an intercalation electrode is measured
is
LiI  LiV + Li.

(33)

At equilibrium the free-energy change for Reaction (32) is zero.


Therefore,
Li + RT ln aLi = +

Li (e)

+ RT ln aLi+ (e) + Fe Fref .

(34)

The activity of the lithium metal is unity. After some algebraic


manipulation, the reversible potential at the lithium reference electrode can be written as
+ Li
RT
Li (e)
ELi,ref = ref e =
+
(35)
ln(aLi+ (e) ).
F
F
This expression for ELi,ref is substituted into Eq. (31) to obtain an
expression for the reversible potential of a lithium intercalation
electrode, referenced to a lithium reference electrode
eq

Eref = b ref =

G33

Experimental data for the reversible potential as a function of


intercalation fraction XLiI can be used to determine the relationship between activity coefcients and intercalation fraction. Such
data also provide the standard-state chemical potential of an intercalated lithium relative to a lithium vacancy and lithium metal
(
G33 ). For ideal solid solutions, the activity coefcients are unity.
However, intercalating electrodes can be complex solid solutions
with many phase transitions [13,16].
The two activity coefcients are related to one another by the
GibbsDuhem equation [17,13]. Therefore, a single expression is
developed for the excess Gibbs free energy GE of the solution. The
activity coefcient is related to the excess Gibbs free energy of the
solution as
RT ln LiI =

6.2. Reference electrode

RT
ln
F

aLiV

RT ln LiV =

aLiI

(36)

The activities for an intercalated lithium and a lithium vacancy


can be written in terms of a mole fraction Xk and activity coefcient k (ak = Xk k ). The mole fractions for intercalated lithium and
the vacancy Xk are dened as the ratio of the species concentration
and the maximum intercalation concentration Xk = Ck /Cmax . Consider, for example, a graphite electrode in which there can at most
be one mole of intercalated lithium for every mole of graphite. The
maximum intercalation concentration for a graphite electrode is
equal to the molar concentration of graphite within the electrode.
The mole fractions are evaluated based on the intercalated lithium
and lithium-vacancy concentrations at the surface of a particle. The
reversible potential of the intercalation electrode with respect to a
reference electrode in terms of mole fractions and activity coefcients is
eq

RT
ln
F

(1 XLiI ) LiV
XLiI LiI

E
GLiV

(37)

(nt GE )
nLiV

(38)

(39)

T,P,n V
Li

T,P,n I
Li

where nk represents the moles of species k and nt represents the


E

GE = XLiI XLiV

N


Am (XLiI XLiV )m ,

(40)

m=0

where Am are coefcients that can be t from equilibrium electricpotential measurements. After substituting Eq. (40) into Eqs. (38)
and (39) and some algebraic manipulation, the following relationships are obtained for the activity coefcients:
RT ln LiI = (1 XLiI )

N



m

Am (2XLiI 1)

m=0

RT ln LiV = X 2 I

N



Am (2XLiI 1)m

m=0

1+

2mXLiI


,

2XLiI 1

2m(1 XLiI )
2XLiI 1

(41)


.

(42)

The activity expressions (Eqs. (41) and (42)) can be substituted into
the expression for the reversible electric-potential difference (Eq.
(37)) and simplied to give
eq

Eref =

(nt GE )
nLiI

total moles. The overbar (Gk ) indicates the partial molar excess free
energy. The excess Gibbs free energy must vanish when XLiI = 0 or
XLiV = 0.
Alternative models have been proposed to represent the excess
Gibbs free energy [13,16]. The semi-theoretical RedlichKister
expansion is used here to model the complexity of a lithium intercalation electrode as

6.3. Activity of intercalated lithium

Li

Thus, the reversible potential referenced to a lithium electrode does


not depend on the activity or standard chemical potential of lithium
ions in the electrolyte phase.

G33

E
GLiI

8965

Eref =

G33

RT
ln
F

1 XLiI

XLiI

1
+
Am (2XLiI 1)m+1
F
N

m=0

2mXLiI (1 XLiI )
(2XLiI 1)1m

(43)

This expression provides a basis for evaluating thermodynamic


properties of intercalated lithium and lithium vacancies as functions of the intercalated mole fraction XLiI [13]. If the intercalation
electrode exhibits ideal behavior, then Eq. (43) simplies to
eq

Eref =

G33

RT
ln
F

1 XLiI
XLiI

(44)

6.4. RedlichKister coefcients


Tables 1 and 2 list RedlichKister coefcients that are t to
represent several intercalation electrodes. Table 1 is taken from

8966

A.M. Colclasure, R.J. Kee / Electrochimica Acta 55 (2010) 89608973

Table 1
RedlichKister coefcients for mesocarbon microbeads (MCMB) and cobalt-oxide
intercalation electrodes. All of the coefcients have units of J kmol1 . These coefcients are taken from Karthikeyan et al. [13].
Parameter

MCMB anode

LiCoO2 cathode

1.660 108

2.857 109

A0
A1
A2
A3
A4
A5
A6
A7
A8
A9
A10

3.580 108
3.501 108
3.525 108
3.569 108
3.863 108
3.591 108
2.879 108
1.498 108
3.991 108
9.617 108
6.326 108

6.483 109
6.517 109
6.566 109
6.579 109
6.302 109
5.047 109
2.711 109
6.905 108
N/A
N/A
N/A

Li

Karthikeyan et al. [13], who reported results for mesocarbonmicrobead (MCMB) anodes and lithium cobalt-oxide (Lix CoO2 )
cathodes. These coefcients were established by measuring
the reversible potential as a function of intercalation fraction. Table 2 provides coefcients for a graphite anode and a
nickelcobaltaluminum-oxide (NCAO) cathode with chemical
composition Lix Niy Coz AL1yz o2 . These coefcients were t using
empirical expressions for reversible potential reported by Doyle
and Fuentes [18] and Smith and Wang [3]. In both tables, the
standard-state chemical potential of intercalated lithium I is relLi

ative to the chemical potential of a lithium vacancy V and lithium


Li

metal Li .
Figs. 3 and 4 illustrate the relatively complex behavior of lithium
intercalation electrodes. These gures are produced using Eq. (43)
and the parameters contained in Tables 1 and 2. Fig. 3 shows that
the MCMB electrode has three slight plateaus in the reversible
potential, which are likely caused by a mixture of solid phases
[16,19]. The graphite reversible potential (Fig. 4) is very close to
that of the MCMB electrode, but does not show the plateau features. Perhaps the graphite electrode has similar plateaus, but the
empirical expression provided by Doyle et al. does not capture such
features. The Liy CoO2 has a reversible-potential plateau between
0.8 < x < 0.9. The NCAO cathode has a reversible potential that is

Fig. 3. Reversible potentials as functions of lithium intercalation fraction for a


mesocarbon-microbead (MCMB) anode and a Liy CoCo2 cathode. These graphs are
produced from the RedlichKister ts reported by Karthikeyan et al. [13].

generally 200300 mV lower than the lithiumcobalt-oxide electrode.


Accurately modeling the fall-off behavior of intercalation electrodes at high intercalation fractions (e.g., the cathode in Fig. 4)
requires a large number of coefcients. Also, a large number of
terms must be used for the chemical potential of individual species
to be reasonable. For instance, the chemical potential for the NCAO

Table 2
RedlichKister coefcients for a graphite anode and a NCAO cathode. All of the coefcients have units of J kmol1 . The graphite parameters are established by tting
an empirical expression for reversible potential [18]. The parameters for the NCAO
cathode are based upon an empirical expression reported by Smith and Wang [3].
Parameter

Graphite anode

NCAO cathode

1.165 107

3.955 108

A0
A1
A2
A3
A4
A5
A6
A7
A8
A9
A10
A11
A12
A13
A14
A15
A16
A17
A18
A19

3.268 106
3.955 106
4.573 106
6.147 106
3.339 106
1.117 107
2.997 105
4.866 107
1.362 105
1.373 108
2.129 107
1.722 108
3.956 107
9.302 107
3.280 107
N/A
N/A
N/A
N/A
N/A

7.676 107
3.799 107
2.873 107
1.169 107
1.451 107
8.938 107
1.671 108
7.236 107
1.746 108
4.067 108
9.534 108
5.897 108
7.455 108
1.102 109
2.927 108
7.214 108
9.029 108
1.599 108
6.658 108
1.084 109

Li

Fig. 4. Reversible potentials as functions of lithium intercalation fraction for a


graphite anode and a NCAO cathode. The RedlichKister ts are based upon empirical expressions reported by Smith and Wang [3] and Doyle and Fuentes [18]. Note
the ideal activity reversible potentials approach + or innity as X approaches 0
and 1, respectively.

A.M. Colclasure, R.J. Kee / Electrochimica Acta 55 (2010) 89608973

8967

cathode with 20 coefcients is approximately 400 kJ mol1 and


with 8 coefcients it is around 2800 kJ mol1 , which is physically
unreasonable. With a low number of coefcients, only the chemical potential difference between vacant and intercalated lithium is
physically reasonable.
There are similarities, as well as signicant differences, between
the present model and the approach reported by Verbrugge and
Koch who developed a non-ideal thermodynamics approach to
model reversible potentials for a carbon-ber electrode [23].
Verbrugge and Koch evaluated excess Gibbs free energy of the electrode using a series expansion that is analogous to Eq. (40) and they
modeled the charge-transfer process using an global ButlerVolmer
equation. However, they evaluated the exchange current density using concentrations alone (i.e., ideal thermodynamics). The
present model uses self-consistent non-ideal thermodynamics to
evaluate charge-transfer rates. Moreover, the current model is
written in terms of elementary reactions, thus facilitating the incorporation of increasingly complex chemistry.
7. Two-step reaction mechanism
An long-term objective of the modeling approach is to develop
detailed, multi-step, reaction mechanisms. As a rst step consider
a relatively straightforward reaction mechanism for lithium charge
transfer:
LiI + (s)  Li(s) + LiV ,

(45)

Li(s)  Li+ (e) + e + (s).

(46)

The rst reaction is a thermal reaction in which an intercalated


lithium moves from the bulk onto an open surface site. The second
reaction is a charge-transfer reaction in which the surface lithium
gives up an electron, moves into the electrolyte phase, and opens a
surface site.
Ideally, independent experiments would be designed and used
to identify elementary-reaction pathways and to determine the
associated rate expressions. However, results from such experiments are not generally available. As a rst approximation,
the kinetic parameters and surface-species thermodynamics are
adjusted to match results from the ButlerVolmer expression used
by Smith and Wang [3] to model a 276 volt hybrid-electric-vehicle
(HEV) battery pack. Table 3 lists the physical parameters for this
system. The ButlerVolmer exchange current density (Eq. (14)) is
written as
i0 = kct F(CLi+ (e) )a (Cmax CLiI )a (CLiI )c ,

(47)

where the concentrations are evaluated at the interface between


active particles and the electrolyte phase. The symmetry factors are assumed to be a = c = 1/2 for both the electrodes.
The rate constant for the charge-transfer reaction kct (Eq.
(46)) at the negative and positive electrode are 4.52 108
and 2.07 108 m2.5 kmol0.5 s1 , respectively. The concentration
dependence (i.e., reaction orders) for the exchange current density (Eq. (47)) was proposed by Newman and co-workers [1]. This
expression was derived by assuming that the global reaction rate
is rst-order with respect to effective concentrations, the symmetry factors sum to unity, and activity coefcients for intercalated
lithium and lithium vacancies are unity.
Table 4 shows specic two-step reaction mechanisms that
were developed to match the modeling results reported by Smith
and Wang [3]. The thermal reactions (equivalent to Eq. (45)) are
assumed to be considerably faster than the rate-limiting chargetransfer reactions (equivalent to Eq. (46)). Therefore, the thermal
reaction rates are set to be sufciently high as to maintain the reactions near partial equilibrium. A consistent set of thermodynamic

Fig. 5. Cell voltage as a function of state-of-charge for various charging and discharging rates computed using ButlerVolmer equations and the two-step reaction
mechanisms.

properties is needed for all species to assure microscopic reversibility. Table 5 provides the standard-state free energies for all the
species in Table 4.
The internal energy of Li(NCAOs ) for a surface that is almost completely covered with lithium and the internal energy of Li(NCAO)
for a intercalation fraction close to unity are approximately 368
and 395 kJ mol1 , respectively. Conversely, the internal energy of
Li(Cs6 ) for a surface that is almost completely covered with lithium is
nearly equal to the internal energy of Li(C6 ) for a intercalation fraction close to unity (i.e., approximately 10 kJ mol1 ). The chemical
potential of an intercalated-lithium species (i.e., Li(NCAO) or Li(C6 ))
is a stronger function of the intercalation fraction than is the chemical potential of a surface lithium species (i.e., Li(NCAOs ) or Li(Cs6 ))
as a function of the adsorbed lithium site fraction.
Fig. 5 compares model-predicted results for cell voltage as
functions of state-of-charge, using ButlerVolmer formulations
and two-step charge-transfer mechanisms (Table 4). Because the
overall charge-transfer process remains close to equilibrium, the
voltage curves for the two chemistry formulations are virtually indistinguishable. Even for charge/discharge rates of 10C, the
highest activation overpotential is only around 4 mV. Such low
polarization is consistent with the high efciency of lithium-ion
batteries. Setting the elementary-reaction rates to be consistent
with the ButlerVolmer equation is equivalent to setting individual reaction rates sufciently high as to maintain reactions near
partial equilibrium.
8. Exchange current density and reversible potential
The exchange current density and reversible potential for
lithium intercalation both depend upon the activities of intercalated lithium and lithium vacancies. The functional form of the
exchange current density is usually derived by assuming that the
intercalation process depends upon species activity in an elementary fashion. By further assuming ideal activity, the exchange

8968

A.M. Colclasure, R.J. Kee / Electrochimica Acta 55 (2010) 89608973

Table 3
Model parameters for FreedomCAR cell as derived from Smith and Wang [3].
Graphite

NCAO cathode

Anode

Separator

Cell geometry
Thickness (m)
Particle radius, rp (m)
Electrode volume fraction, s
Electrolyte volume fraction, elyt
Active specic area, As = 3s /rp (m1 )
Electrode area, A (m2 )

50
1
0.58
0.332
1.74 106
1.045

25.4

Lithium intercalation
Cmax (kmol m3 )
XLiI at 0% SOC
XLiI at 100% SOC
CLi+ (e) initial (kmol m3 )

16.1
0.126
0.676
1.2

Electrode surface
Available sites,  (kmol m2 )

2.6 108

2.6 108

100

10

Transport parameters
Electrode conductivity,  b (S m1 )

36.4
1
0.50
0.330
1.5 106
1.045

0.5

23.9
0.936
0.442
1.2

1.2

Electrolyte conductivity,  = ACLi+ (e) exp(k(CLi+ (e) ) )


A (S m2 kmol1 )
k (m4.2 kmol1.4 )
d

1.58
0.85
1.4

1.58
0.85
1.4

1.58
0.85
1.4

Intercalation diff. coeff., DLiI (m2 s1 )


Electrolyte diff. coeff., De (m2 s1 )
Bruggemann exponent, p
Transference number, t 0 +

2.0 1016
2.6 1010
1.5
0.363

2.6 101
1.5
0.363

3.7 1016
2.6 101
1.5
0.36

Li (e)

current density simplies into the form that is typically used in


the literature (i.e., Eq. (47)) [3,2022]. However, measurements of
the reversible electric-potential difference for lithium intercalation
electrodes show that non-ideal activities must be used for modeling the intercalation process (Figs. 3 and 4). It is for this reason
that relatively complex empirical functions (e.g., RedlichKister)
are used to represent the reversible potential. However, as rst
discussed by Zhang et al. [19], there is an inherent inconsistency
in using non-ideal activities for evaluating reversible potentials
but using ideal activities to evaluate exchange current densities. The following discussion explores the consequences of this
inconsistency.

where ak and Ck are the activity and standard-state concentration


for species k, and kf and kr are forward and reverse rate expressions, respectively. The electric-potential difference between the
electrode and electrolyte phases is E = b e . Assuming an ideal
mixture, species activities are evaluated as mole fractions, and the
standard-state concentration is the total mixture concentration.
Assuming an ideal surface, the adsorbate activities are evaluated
as the surface site fractions. The standard-state concentration for
a surface species is the concentration of total surface sites. The
standard-state concentration for intercalated lithium and lithium
vacancies are evaluated as the lithium host site concentration
Cmax = CLiV + CLiI . With these assumptions, the net electron production rate can b rewritten as

8.1. Exchange current density for lithium intercalation

s e = Cmax kf,29 aLiI exp

If the overall lithium intercalation process (Eq. (29)) is assumed


to be rst-order with respect to species activities, then the net
electron production rate (kmol m2 s1 ) takes the form


s e = kf,29 C I aLiI exp
Li

a EF
RT

kr,29 C V aLiV C + aLi+ (e)


Li (e)
Li


exp

c EF
RT

kr,29 aLiV aLi+ (e) C +

a EF
RT

Li (e)

exp

c EF
RT


.

(49)

To assure microscopic reversibility, the reverse rate expression kr is


written in terms of the forward rate expression kf and the reactions
equilibrium constant K as


,

(48)

K29

kf,29
=
= C + exp
Li (e)
kr,29

G29

RT

(50)

Table 4
Two-step reaction mechanisms for the anode and cathode of a cell with a graphite anode and a NCAO cathode. The available site density for both electrode surfaces is
 = 2.6 108 kmol m2 . The superscript s denotes a surface site. The rate expressions and thermodynamic properties are adjusted to match results using the ButlerVolmer
expressions given by Smith and Wang [3].
kf (kmol, m, s)

 Li(Cs6 ) + (C6 )
(e) + (Cs6 ) + e (C6 )

2.3 10

N/A

4.74

0.5

Li(NCAO) + (NCAOs )  Li(NCAOs ) + V(NCAO)

2.1 109

N/A

Li(NCAOs )  Li+ (e) + (NCAOs ) + e (NCAO)

1.6 1029

0.5

Reaction
Li(C6 ) +

(Cs6 )
+

Li(Cs6 )  Li

G (kJ mol1 )
2.45
9.2
27.5
368.0

A.M. Colclasure, R.J. Kee / Electrochimica Acta 55 (2010) 89608973

8969

Table 5
Thermodynamic properties for species that participate in the
two-step reaction mechanisms shown in Table 4.
G (kJ mol1 )

Surface species
Anode
Li(Cs6 )
(Cs6 )
Li(C6 )
V(C6 )
Li+ (e)
e (C6 )

9.2
0.0
11.65
0.0
0.0
0.0

Cathode
Li(NCAOs )
(NCAOs )
Li(NCAO)
V(NCAO)
Li+ (e)
e (NCAO)

368.00
0.0
3.95.5
0.0
0.0
0.0

Fig. 6. Activities of intercalated lithium and lithium vacancies within graphite as


functions of intercalation fraction. The activities are evaluated using ideal activity
coefcients and using the RedlichKister expansion.

Substituting Eq. (50) into Eq. (49) yields

s e = Cmax kf,29 aLiI exp


exp

G29

a EF
RT

8.2. Reversible potentials with ideal and non-ideal activities

aLiV aLi+ (e) exp

RT

c EF
RT


.

(51)

The electric-potential difference may be written in terms of an


equilibrium electric-potential difference and an overpotential as
E = Eeq + . Writing the electron production rate in terms of overpotential and equilibrium electric-potential yields

s e = Cmax kf,29 aLiI exp


exp


G 
29

RT

a E eq F
RT


exp

aLiV aLi+ (e) exp

a F
RT

c E eq F
RT


exp

c F
RT


.

(52)

After substituting Eq. (31) for the equilibrium electric potential and
some further algebraic manipulation, the net electron production
rate becomes

s e = Cmax kf,29 exp


exp

a F
RT

G29
a

RT

a aa aa
a1
+
I
V

Li

exp

c 1c ac exp
a1
V a +
I

Li (e) Li

Li

Li

8.3. Functional form of the exchange current density

Li (e)

(1 c a )
G29

If a charge-transfer reaction is rst-order with respect to species


activities and the symmetry factors sum to unity, then the exchange
current density takes the form

RT

c F
RT

(53)

Assuming for an elementary reaction that the symmetry factors


sum to unity (a + c = 1), the net electron production rate can be
written as

s e = Cmax kf,29 exp


exp

a F
RT

G29
a

RT

aaV aa+
Li

exp

ac
Li (e) LiI

c F
RT


.

(54)

The current density (A m2 ) is the product of the net electron production rate and Faradays constant. Thus, in ButlerVolmer form,
the current density may be written as

i = i0 exp

a F
RT

exp

c F
RT

(55)

ac .
Li (e) LiI

(56)

where the exchange current density is

i0 = Cmax Fkf,29 exp

G29
a

RT

aaV aa+
Li

As illustrated in Fig. 4, reversible potentials can vary greatly


from ideal behavior. Fig. 6 shows the activities of intercalated
lithium and lithium vacancies within a graphite electrode, comparing ideal activity coefcients (i.e., = 1) and the RedlichKister
expansion. Ideal activities are a linear function of the intercalated mole fraction, with actual activities being functionally quite
different. By denition, both ideal and non-ideal activities for intercalated species must be zero at zero intercalation fraction and
unity at unity intercalation fraction. For intercalation fractions
below XLiI < 0.1, the lithium-vacancy activity decreases sharply
as a function of the intercalation fraction. Below XLiI < 0.1 the
intercalated-lithium activity appears to be nearly zero. However,
despite being small (compared to unity) in this region the the value
of aLiI is varying by orders of magnitude. These variations signicantly affect the equilibrium electric-potential difference, which
varies as Eeq ln(aLiV /aLiI ).

i0 = Q (T )aaV aa+
Li

ac ,
Li (e) LiI

(57)

where Q(T) is a function of temperature alone. The exchangecurrent-density expression that is typically used in the literature
(i.e., Eq. (47)) can be derived from Eq. (57) by assuming that the
activity coefcients for intercalated lithium and lithium vacancies
are unity and that the effective Li-ion concentration is evaluated as
its actual concentration [3,2022]. Thus, the most frequently used
formulation for the exchange current density is based upon a defacto assumption of ideal activities. However, as can be seen from
the reversible-potential data, the intercalation process is not ideal.
Although self consistency requires the use of non-ideal activities,
most lithium-ion battery models use non-ideal empirical functions
for the reversible potential, but assume ideal activities to evaluate charge-transfer rates. Zhang et al. investigated the effect of this
inconsistency [19]. An analogous approach is used here to evaluate
the inuence of using RedlichKister expansions to model activity
coefcients.
Written in terms of mole fractions and activity coefcients, the
exchange current density is
i0 = L(T )C a+

C a C c a c ,
Li (e) LiV LiI LiV LiI

(58)

8970

A.M. Colclasure, R.J. Kee / Electrochimica Acta 55 (2010) 89608973

Fig. 8. Cell voltage as a function of state-of-charge for various discharge rates, comparing the use of ideal activities and using the RedlichKister expansion to evaluate
the exchange current density for a global one-step charge-transfer reaction.

Fig. 7. Exchange current density for a graphite electrode and a NCAO cathode as
functions of Li intercalation fraction [3]. The exchange current densities are calculated using ideal activities and using the RedlichKister expansion. The exchange
current densities have been normalized so that the maximum current densities are
the same as those reported by Smith and Wang [3]. Also, the electrolyte concentra3
tion is set to be CLi+ (e) = 1.2 kmol m .

where L(T) is a function of temperature alone. Substituting


the RedlichKister expansion, which represents the reversiblepotential accurately, for the activity coefcients leads to

L(T ) a
c
C + C a C c exp
(1 XLiI )2
i0 =
RT Li (e) LiV LiI
RT

m=0

exp

Am (2XLiI 1)m

1+

2XLiI 1

a 2 
X
Am (2XLiI 1)m
RT LiI
N

m=0

8.4. Exchange current density for a two-step mechanism

2mXLiI


1

approach full intercalation, the exchange current density calculated with the RedlichKister expansion is signicantly lower than
the exchange current density calculated with ideal activities. The
lower exchange current density with non-ideal activities causes a
signicant increase in the cathode activation overpotentials at high
discharge rates. At low discharge rates (i.e., low current density),
only a small overpotential is required to compensate for the lower
exchange current density.
Lithium diffusion within the electrode particles usually limits
battery performance. Thus, only a small overpotential is required to
drive charge-transfer processes. As a result, under ordinary operating conditions, the functional form of the exchange current density
is not too important. At low to moderate discharge rates, the modeling error introduced by assuming ideal activity for the exchange
current density is small. However, the error is more signicant for
high discharge rates. However, under off-nominal operating conditions (e.g., very low temperature) the charge-transfer process
could become more rate limiting. In these cases, using the correct
functional form for the exchange current density is increasingly
important.

2m(1 XLiI )
2XLiI 1

.
(59)

Fig. 7 illustrates the relationship between exchange current density and intercalated-lithium fraction using a graphite anode and
NCAO cathode that represent the cell modeled by Smith and Wang
[3]. The exchange current densities are evaluated with two different assumptionsone assuming that the activity coefcients are
unity (Eq. (47)) and the other using RedlichKister expansions (Eq.
(59)). It is evident that non-ideal activities produce a functionally
different exchange current density, especially for the cathode.
To evaluate the importance of the non-ideal activities, the
battery model is implemented with the exchange current density calculated using ideal activity and using the RedlichKister
expansion. Fig. 8 compares constant-current discharge proles as
calculated using the two methods. At low to moderate discharge
rates, the two exchange-current-density formulations produce
nearly identical results. At discharge rates greater than 5C, the
two approaches begin to produce small, but noticeable, differences. At the higher rates the differences are greatest as the
cathode intercalation fraction nears unity. As the cathode particles

Assuming that the charge-transfer reactions are rate limiting


(Table 4), the net charge-transfer rate can be evaluated using the
ButlerVolmer formulation. The exchange current density is written as

i0 = Fkf,46  exp

a I
Li

RT

1
ac aa aa+ .
aLiI + (aLiV /Keq,45 ) LiI LiV Li (e)
(60)

The exchange-current-density formulation proposed by Newman


and co-workers [1] (Eq. (47)) can be recovered by making two further assumptions. First, the activities of intercalated lithium and
lithium vacancies are evaluated as the vacancy mole fractions. Second, the standard-state Gibbs free-energy change for the thermal
reaction is small (i.e., Keq,45 1). However, it should be recognized
that these assumptions may not be generally valid.
Fig. 9 compares exchange current densities as functions of intercalation fraction for a graphite anode and a NCAO cathode. The
exchange current densities are calculated assuming a one-step
mechanism with ideal activities (Eq. (47)) and for a two-step mechanism with non-ideal activities for the bulk electrode species (Eq.
(60)). The exchange current density for graphite based upon oneand two-step mechanisms are relatively close to one another over
the region of battery operation (0.1 < XLiI < 0.8). However, over a
nominal operating range (0.4 < XLiI < 0.97), the cathode exchange
current densities are signicantly different. For the one-step mech-

A.M. Colclasure, R.J. Kee / Electrochimica Acta 55 (2010) 89608973

8971

Fig. 10. Lithium-ion concentration along the length of the cell at the end of the
discharge cycle.

Fig. 9. Exchange current densities for a graphite electrode and the NCAO cathode as
a functions of intercalation fraction [3]. The exchange current density is calculated
based upon a one-step ButlerVolmer reaction with ideal activities (Eq. (47)) and
based upon the proposed two-step mechanism with non-ideal activities(Eq. (60)).
3
The electrolyte concentration is set to be CLi+ (e) = 1.2 kmol m .

anism, the exchange current density is relatively uniform over this


region. The exchange current density for the two-step mechanism
for the cathode falls off quickly at high intercalation fractions. However, because the rate of Eq. (46) for the cathode is high the cathode
charge-transfer process is not rate limiting, even at high intercalation fractions. Thus, the discharge-charge performance is similar
for the two approaches (Fig. 5), even though the exchange current
densities can be quite different (Fig. 9).

concentration of 1.2 kmol m3 . Changes in electrolyte concentration can affect the net rate of the charge-transfer reaction and
slightly affect the reversible potential. The reverse rate of progress
for Eq. (46) is proportional to the lithium-ion concentration within
the electrolyte phase. However, the 20% change in the electrolytephase concentration for a discharge rate of 10C only is overcome
by a slight change in the potential difference between electrode
and electrolyte phases. For instance, if the diffusion coefcient
for lithium ions within the electrolyte phase is increased by an
order of magnitude, then the maximum and minimum electrolyte
concentrations for the 10C discharge are 1.22 and 1.18 kmol m3 .
Even with this more uniform electrolyte concentration, the voltage discharge curve calculated with the higher electrolyte diffusion
coefcient is only slightly above the curve calculated with the nominal electrolyte diffusion coefcient (the maximum difference is
only 15 mV).
9.2. Electrode intercalation proles

9. Spatial proles
As the discharge/charge rates increase, the battery efciency
decreases, due primarily to relatively slower diffusion of intercalated lithium within electrode particles. As previously discussed,
the activation overpotentials are usually small, even for high currents. Therefore, the efciency loss is attributed primarily to lithium
diffusion within the electrolyte and/or solid electrode phase. The
diffusion coefcient for intercalated lithium within graphite and
NCAO particles are taken to be 2 1016 and 3.7 1016 m2 s1 ,
respectively [3]. The effective diffusion coefcient for lithium ions
within the electrolyte phase is around 5 1011 m2 s1 . Thus, the
diffusion coefcient for lithium ions in the electrolyte phase is
ve orders of magnitude higher than the diffusion coefcient for
intercalated lithium within solid electrode particles. However, the
average diffusion length for intercalated lithium is usually around
0.5 m and for lithium ions is around 70 m. Based upon these
diffusion coefcients and diffusion lengths, the polarization losses
are expected to be dominated by diffusion of intercalated lithium
within electrode particles.

Fig. 11 shows radial concentration proles of intercalated


lithium within anode and cathode particles near the end of the
discharge cycles. The particles are located roughly in the middle
of the composite-electrode structure. However, the concentration
proles within the particles do not vary signicantly as functions
of position within the electrode structure. Ideally, the intercalatedlithium concentration would vary little within an active particle.
This condition is met for low discharge rates of 0.5C and 1C.
However, due to diffusion limitations, LiI concentration variations
become signicant for discharge rates equal to or greater than 2C.
At a discharge rate of 10C, the LiI concentration varies from 10.7 to
0.9 kmol m3 within the negative electrode. With a discharge rate
of 0.5C, the concentration only varies from 4.1 to 3.2 kmol m3 . At
10C, if the intercalated diffusion coefcients are increased by an
order of magnitude, then the cell efciency improves substantially.
In summary, the low diffusion coefcient for LiI in the negative and
positive electrode particles causes the cell efciency to decrease
with higher charge/discharge rates.
9.3. Surface-coverage proles

9.1. Lithium-ion concentration proles


Fig. 10 shows predicted lithium-ion concentration proles
within the electrolyte phase near the end of a discharge cycle
(cell voltage in Fig. 5). For the highest discharge rate, the highest
and lowest electrolyte concentrations are 1.42 and 0.97 kmol m3 ,
respectively. This is roughly a 20% change from the initial uniform

During discharge/charge cycles, the surface coverages on electrode particles change considerably. Fig. 12 shows proles of the
the fractional surface coverage of adsorbed lithium  Li(s) at different states of charge (SOC) during the 5C discharge cycle. During
discharge, intercalated lithium moves from the graphite particles
into the NCAO particles. Thus, the intercalated-lithium fraction

8972

A.M. Colclasure, R.J. Kee / Electrochimica Acta 55 (2010) 89608973

electrode/separator interface, and Li(NCAOs ) slightly increasing near


the positive electrode/separator interface. Over the course of a discharge, the concentration of lithium ions in the electrolyte phase
increases inside the negative electrode and decreases inside the
positive electrode. This ion concentration behavior should lead to
an increase in surface lithium on graphite particles and a decrease
on NCAO particles. However, the surface coverage shows the exact
opposite behavior. Thus, the surface coverage is more dependent
on the electric-potential phase difference and intercalation fraction
than the concentration of lithium ions in the electrolyte phase.
The surface coverage varies more during a discharge/charge
cycle on the NCAO particles compared to the graphite particles
(Fig. 12). A particle surface site is either open (s) or lled by an
adsorbed lithium Li(s). At the beginning of discharge, the cobaltoxide cathode surfaces are almost completely void of adsorbed
lithium (Li(NCAOs ) 105 ). However, as the SOC reaches 50%,
the NCAO surfaces are nearly saturated with adsorbed lithium.
Conversely, at 100% SOC the graphite anode surfaces have significant levels of vacant sites (Li(C s ) 0.62). Even at the end of the
6

discharge, Li(C s ) 0.01. The differences in coverage behavior is


6

related to chemical potentials of the bulk electrode species (i.e., the


intercalated Li). The chemical potential of intercalated lithium and
lithium vacancies vary more as a function of intercalation fraction
in the NCAO than in the graphite.

Fig. 11. Intercalated-lithium concentration within anode and cathode particles near
the end of a discharge cycle. The representative particles are located near the middle
of the composite-electrode structure.

decreases inside graphite particles and increases inside NCAO particles. This process leads to a reduction in Li(C s ) and a rise in Li(NCAOs ) .
6

The faradaic current density is highest near the electrode/separator


interfaces. This leads to Li(C s ) slightly decreasing near the negative
6

9.4. Electrolyte-phase potential


At the end of a 5C discharge, the anode and cathode electricpotential proles are essentially uniform at 0.0 and 3.11 V,
respectively. The electrolyte-phase electric-potential prole is
essentially uniform at 0.22 V. Intuitively, one might expect the
value of the electric potential of the electrolyte phase to be between
the electric potentials of the electrodes. However, the electrolytephase electric potential is affected by the value selected for the
reversible potential of the reference lithium electrode ELi,ref (Eq.
(35)). By convention, ELi,ref = 0.0 V. Thus, the standard chemical
potential of a lithium ion with respect to lithium is + =
Li (e)

0.0 kJ mol1 . The charge-transfer rates given in Table 4 depend


upon this convention, and the electrolyte-phase electric potential
typically falls below the electric potentials of both electrodes.
The exchange current densities for the elementary reactions
(Table 4) are proportional to G + as


i0 exp

a G +

Li (e)

RT

Li (e)

(61)

If the standard-state chemical potential of a lithium ion is varied,


then the charge-transfer reaction rates must be adjusted to maintain the same exchange current densities. Consider, for example,
setting ELi,ref = 1.00 V. For this choice, + = 96.5 kJ mol1 .
Li (e)

The electrolyte-phase electric potential would be located between


the electric potentials of the cathode and anode. To recover the
same results as those presented in this paper, except for the
electrolyte-phase electric potential, the forward charge-transfer
reaction rate constants for the graphite and NCAO are 1.34 109
and 4.51 1021 s1 , respectively.
10. Conclusion

Fig. 12. Lithium surface-coverage proles on graphite anode particles and NCAO
cathode particles as functions of state-of-charge (SOC) during a 5C discharge cycle.

A one-dimensional model of a lithium-ion cell has been derived


and implemented to predict concentration and electric-potential
variations within a cell. Transport through the electrode phase is
modeled using composite-electrode theory. Concentrated solution
theory is used to model the binary electrolyte mixture. The model

A.M. Colclasure, R.J. Kee / Electrochimica Acta 55 (2010) 89608973

provides a new capability to evaluate the rate of charge-transfer


using elementary-reaction mechanisms as an alternative to the
widely used ButlerVolmer equation. However, such detailed reaction mechanisms require the evaluation of several thermodynamic
and kinetic parameters, many of which are difcult to measure
or estimate. In this paper, parameters were assigned for a relatively simple two-step mechanism. The rates of the elementary
reactions depend upon the surface coverages of electrode particles. The present framework is able to incorporate detailed reaction
mechanisms leading to SEI growth and lithium plating. Non-ideal
thermodynamic properties for the electrode are modeled using
a RedlichKister formulation. Compared to ideal activities, the
non-ideal thermodymanics has a large inuence upon reversible
potentials. However, using consistent non-ideal thermodynamics
in the evaluation of charge-transfer rates has a relatively small
effect. For the batteries modeled in this paper, the performance
is generally limited by the diffusion of intercalated lithium within
electrode particles.

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]

[15]

Acknowledgement
This effort was supported by the Ofce of Naval Research via
an RTC grant (N00014-05-1-03339). We gratefully acknowledge
many insightful discussions with Prof. David Goodwin (Caltech)
concerning the elementary-reaction representation and software
implementations of complex electrochemical charge-transfer processes. We also gratefully acknowledge benecial discussions with
Drs. Graham Goldin (ANSYS, Inc.), Harry Moffat (Sandia Natl.
Labs), Kandler Smith (NREL), and Huayang Zhu (Colorado School
of Mines).

8973

[16]
[17]
[18]
[19]
[20]
[21]
[22]

[23]

M. Doyle, T.F. Fuller, J. Newman, J. Electrochem. Soc. 140 (1993) 1526.


M. Doyle, T.F. Fuller, J. Newman, Electrochim. Acta 39 (1994) 2073.
K. Smith, C. Wang, J. Power Sources 161 (2006) 628.
D. Aurbach, J. Power Sources 89 (2000) 206.
B. Markovsky, A. Nimberger, Y. Talyosef, A. Rodkin, A.M. Belostotskii, G. Salitra,
D. Aurbach, H. Kim, J. Power Sources 136 (2004) 296.
M. Doyle, T.F. Fuller, J. Newman, J. Electrochem. Soc. 141 (1994) 982.
D.G. Goodwin, H. Zhu, A.M. Colclasure, R.J. Kee, J. Electrochem. Soc. 156 (2009)
B1004.
W.G. Bessler, S. Gewies, M. Vogler, Electrochim. Acta 53 (2007) 1782.
J. Newman, K.E. Thomas-Alyea, Electrochemical Systems, John Wiley and Sons,
2004.
A.J. Bard, L.R. Faulkner, Electrochemical Methods Fundamentals and Applications, John Wiley and Sons, 2001.
D. Chen, Z. Lina, H. Zhu, R.J. Kee, J. Power Sources 191 (2009) 240.
J. Christensen, J. Newman, J. Electrochem. Soc. 151 (2004) A1977.
D.K. Karthikeyan, G. Sikha, R.E. White, J. Power Sources 185 (2008) 1398.
A.C. Hindmarsh, P.N. Brown, K.E. Grant, S.L. Lee, R. Serban, D.E. Shumaker, C.S. Woodward, ACM Transactions on Mathematical Software, 31,
no. 3, Electrochemical Society, 2005, p. 363 https://computation.llnl.gov/
casc/sundials/main.html.
D.G. Goodwin, in: M. Allendorf, F. Maury, F. Teyssandier (Eds.), Chemical Vapor
Deposition XVI and EUROCVD 14, vol. PV 20032008, Electrochemical Society,
2003, p. 155 http://code.google.com/p/cantera/.
H. Lee, C. Wan, Y. Wang, J. Power Sources 114 (2003) 285.
J.P. OConnell, J.M. Haile, Thermodynamics Fundamentals and Applications,
Cambridge University Press, Cambridge, 2005.
M. Doyle, Y. Fuentes, J. Electrochem. Soc. 150 (2003) A706.
Q. Zhang, Q. Guo, R.E. White, J. Electrochem. Soc. 153 (2006) A301.
P. Arora, M. Doyle, R.E. White, J. Electrochem. Soc. 146 (1999) 3453.
P. Ramadass, B. Haran, P.M. Gomadam, R.E. White, B.N. Popov, J. Electrochem.
Soc. 151 (1999) A196.
K.E. Thomas, J. Newman, R.M. Darling, Mathematical modeling of lithium batteries, in: W.A. van Schalkwijk, B. Scrosati (Eds.), Advances in Lithium-ion
Batteries, Kluwer Academic Publishers, 2002, p. 345.
M.W. Verbrugge, B.J. Koch, J. Electrochem. Soc. 143 (1996) 600.

Anda mungkin juga menyukai