Sie sind auf Seite 1von 17

Terms of Use

Theory of Chemical Kinetics and Charge


Transfer based on Nonequilibrium
Thermodynamics
MARTIN Z. BAZANT*
Departments of Chemical Engineering and Mathematics, Massachusetts Institute
of Technology, Cambridge, Massachusetts 02139, United States
RECEIVED ON JULY 20, 2012

CONSPECTUS

A dvances in the fields of catalysis and electrochemical energy


conversion often involve nanoparticles, which can have
Downloaded by 83.97.180.182 at 10:40:57:319 on May 24, 2019

kinetics surprisingly different from the bulk material. Classical


theories of chemical kinetics assume independent reactions in dilute
from https://pubs.acs.org/doi/10.1021/ar300145c.

solutions, whose rates are determined by mean concentrations. In


condensed matter, strong interactions alter chemical activities and
create variations that can dramatically affect the reaction rate. The
extreme case is that of a reaction coupled to a phase transformation,
whose kinetics must depend not only on the order parameter but
also on its gradients at phase boundaries. Reaction-driven phase
transformations are common in electrochemistry, when charge
transfer is accompanied by ion intercalation or deposition in a solid
phase. Examples abound in Li-ion, metalair, and leadacid
batteries, as well as metal electrodepositiondissolution. Despite complex thermodynamics, however, the standard kinetic model
is the ButlerVolmer equation, based on a dilute solution approximation. The Marcus theory of charge transfer likewise considers
isolated reactants and neglects elastic stress, configurational entropy, and other nonidealities in condensed phases.
The limitations of existing theories recently became apparent for the Li-ion battery material LixFePO4 (LFP). It has a strong
tendency to separate into Li-rich and Li-poor solid phases, which scientists believe limits its performance. Chemists first modeled
phase separation in LFP as an isotropic “shrinking core” within each particle, but experiments later revealed striped phase
boundaries on the active crystal facet. This raised the question: What is the reaction rate at a surface undergoing a phase
transformation? Meanwhile, dramatic rate enhancement was attained with LFP nanoparticles, and classical battery models could
not predict the roles of phase separation and surface modification.
In this Account, I present a general theory of chemical kinetics, developed over the past 7 years, which is capable of answering
these questions. The reaction rate is a nonlinear function of the thermodynamic driving force, the free energy of reaction, expressed
in terms of variational chemical potentials. The theory unifies and extends the CahnHilliard and AllenCahn equations through a
master equation for nonequilibrium chemical thermodynamics. For electrochemistry, I have also generalized both Marcus and
ButlerVolmer kinetics for concentrated solutions and ionic solids.
This new theory provides a quantitative description of LFP phase behavior. Concentration gradients and elastic coherency strain
enhance the intercalation rate. At low currents, the charge-transfer rate is focused on exposed phase boundaries, which propagate as
“intercalation waves”, nucleated by surface wetting. Unexpectedly, homogeneous reactions are favored above a critical current and
below a critical size, which helps to explain the rate capability of LFP nanoparticles. Contrary to other mechanisms, elevated
temperatures and currents may enhance battery performance and lifetime by suppressing phase separation. The theory has also been
extended to porous electrodes and could be used for battery engineering with multiphase active materials.
More broadly, the theory describes nonequilibrium chemical systems at mesoscopic length and time scales, beyond the reach of
molecular simulations and bulk continuum models. The reaction rate is consistently defined for inhomogeneous, nonequilibrium
states, for example, with phase separation, large electric fields, or mechanical stresses. This research is also potentially applicable
to fluid extraction from nanoporous solids, pattern formation in electrophoretic deposition, and electrochemical dynamics in
biological cells.

1144 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1144–1160 ’ 2013 ’ Vol. 46, No. 5 Published on the Web 03/22/2013 www.pubs.acs.org/accounts
10.1021/ar300145c & 2013 American Chemical Society
Theory of Chemical Kinetics and Charge Transfer Bazant

Introduction boundary in order to preserve the stable phases, but this


Breakthroughs in catalysis and electrochemical energy con- could not be described by classical kinetics proportional to
version often involve nanoparticles, whose kinetics can concentrations. Somehow the reaction rate had to be sensi-
differ unexpectedly from the bulk material. Perhaps the tive to concentration gradients.
most remarkable case is lithium iron phosphate, LixFePO4 As luck would have it, I was working on models of charge
(LFP). In the seminal study of micrometer-sized LFP particles, relaxation in concentrated electrolytes using nonequili-
Padhi et al.1 concluded that “the material is very good for brium thermodynamics,35,41 and this seemed like a natural
low-power applications” but “at higher current densities starting point. Gerbrand Ceder suggested adapting the
there is a reversible decrease in capacity that... is associated CahnHilliard (CH) model for LFP,42 but it took several years
with the movement of a two-phase interface” between to achieve a consistent theory. Our initial manuscript43 was
LiFePO4 and FePO4. Ironically, over the next decade, in rejected in 2007, just after Gogi left MIT and I went on
nanoparticle form, LFP became the most popular high- sabbatical leave to ESPCI, faced with rewriting the paper.12
power cathode material for Li-ion batteries.24 Explaining The rejection was a blessing in disguise, since it made
this reversal of fortune turned out to be a major scientific me think harder about the foundations of chemical kinetics.
challenge, with important technological implications. The paper contained some new ideas, phase-field chemical
It is now understood that phase separation is strongly kinetics and intercalation waves, that, the reviewers
suppressed in LFP nanoparticles, to some extent in felt, contradicted the laws of electrochemistry. It turns out
equilibrium58 but especially under applied current,7,911 since the basic concepts were correct, but Ken Sekimoto and
reaction limitation,12 anisotropic lithium transport,1316 elastic David Lacoste at ESPCI helped me realize that my initial
coherency strain,7,1719 and interfacial energies8,9,20,21 are all CahnHilliard reaction (CHR) model did not uphold the
enhanced. At low currents, anisotropic nucleation and growth De Donder relation.37 In 2008 in Paris, I completed the
can also occur,79,12,22 as well as multiparticle mosaic theory, prepared lecture notes,33 published generalized
instabilities.2326 These complex phenomena cannot be ButlerVolmer kinetics35 (section 5.4.2) and formulated
described by traditional battery models,27,28 which assume a nonequilibrium thermodynamics for porous electrodes26
spherical “shrinking core” phase boundary.29,30 (see also Sekimoto37).
This Account summarizes my struggle to develop a phase- Phase-field kinetics represents a paradigm shift in chemi-
field theory of electrochemical kinetics69,12,19,26,3335 by cal physics, which my group has successfully applied to
combining charge-transfer theory36 with concepts from sta- Li-ion batteries. Damian Burch6 used the CHR model to study
tistical physics37 and nonequilibrium thermodynamics.3840 intercalation in nanoparticles, and his thesis25 included early
It all began in 2006 when my postdoctoral associate, Gogi simulations of “mosaic instability” in collections of bistable
Singh, found the paper of Chen et al.31 revealing striped particles.23,24 Simulations of galvanostatic discharge by
phase boundaries in LFP, looking nothing like a shrinking Peng Bai and Daniel Cogswell led to the unexpected predic-
core and suggesting phase boundary motion perpendicular tion of a critical current for the suppression of phase
to the lithium flux (Figure 1). It occurred to me that, at such a separation.9 Liam Stanton modeled anisotropic coherency
surface, intercalation reactions must be favored on the phase strain,19 which Dan added to our LFP model,7 along with

FIGURE 1. Motivation to generalize charge-transfer theory. Observations by (a) Chen et al.31 and (b) Ramana et al.32 of separated FePO4 and LiFePO4
phases on the active {010} facet, which suggest (c) focusing of lithium intercalation reactions on the phase boundary, so it propagates as an
“intercalation wave”12 (or “domino cascade”15). From refs 12, 31, and 32.

Vol. 46, No. 5 ’ 2013 ’ 1144–1160 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1145


Theory of Chemical Kinetics and Charge Transfer Bazant

surface wetting.8 Using material properties from ab initio


calculations, Dan predicted phase behavior in LFP7 and
the critical voltage for nucleation8 in excellent agreement
with experiments. Meanwhile, Todd Ferguson26 did the first
simulations of phase separation in porous electrodes, pav-
ing the way for engineering applications.
What follows is a general synthesis of the theory and a
summary of its key predictions. A thermodynamic frame-
work is developed for chemical kinetics, whose application
to charge transfer generalizes the classical ButlerVolmer
and Marcus equations. The theory is then unified with
phase-field models and applied to Li-ion batteries.

Reactions in Concentrated Solutions FIGURE 2. (a) Landscape of excess chemical potential explored by
the reaction S1 f S2. (b) Adsorption from a liquid, where the transition
Generalized Kinetics. The theory is based on chemical
state (TS) excludes multiple surface sites (s > 1) while shedding the
thermodynamics. In an open system, the chemical potential first-neighbor shell. (c) Solid diffusion on a lattice, where the transition
of species i (per particle), state excludes two sites.

μi ¼ kB T ln ai þ μΘ ~
i ¼ kB T ln c i þ μi
ex
(1) diffusivity.28 In eq 5, the first term represents random
fluctuations and the second drift in response to the thermo-
is defined relative to a standard state (Θ) of unit activity dynamic bias, rμex i .
(ai = 1) and concentration ci = cΘ c i = ci/cΘ
i , where ~ i is the In a consistent formulation of reaction kinetics,33,37 there-
dimensionless concentration. The activity coefficient, fore, the reaction complex explores a landscape of excess
chemical potential, μex(x) between local minima μex 1 and μ2
ex
ex
 μΘ with transitions over an activation barrier μ (Figure 2a). For
ex
γi ¼ e(μi i
)=(kB T)
(2)
‡  μ1,2 . kBT), the reaction rate (per site) is
rare transitions (μex ex

is a measure of nonideality (ai = γi~c i). In a dilute solution,


Θ
μex
i = μi and γi = 1. For the general chemical reaction, R ¼ k f ~c 1 e (μ‡
ex
 μex
1
)=(kB T)
 k r ~c 2 e (μ‡
ex
 μex
2
)=(kB T)
(6)

S1 ¼ ∑r sr Ar f ∑p sp Bp ¼ S2 (3) Enforcing detailed balance (R = 0) in equilibrium (μ1 = μ2)


yields the reaction rate consistent with nonequilibrium
the equilibrium constant is thermodynamics:
 eq
R ¼ k0 (e (μ‡  μ1 )=(kB T)
 e (μ‡  μ2 )=(kB T)
ex ex
Θ a2 Θ
 μΘ ) (7)
K ¼ ¼ e(μ1 2
)=(kB T)
(4)
a1
where k0 = kf = kr (for properly defined μ). Equation 7
Q Q
where a1 = rasr r, a2 = paspp, μΘ Θ Θ Θ
1 = ∑rsrμr and μ2 = ∑pspμp . upholds the De Donder relation,37
The theory assumes that departures from equilibrium obey
linear irreversible thermodynamics (LIT).38,39 The flux of spe-
Rf K Θ a1  μ2 )=(kB T)
cies i is proportional to the thermodynamic driving force, rμi: ¼ ¼ e(μ1 (8)
Rr a2
!
μex
Fi ¼ Mi ci rμi ¼ Di rci þ ci r i
kB T which describes the steady state of chemical reactions in
open systems.44
¼ Dchem
i rci (5)
The thermodynamic driving force is

where Mi is the mobility, Di = MikBT is the tracer diffusivity, a2


Δμ ¼ μ2  μ1 ¼ kB T ln ¼ ΔG (9)
and Dchem
i = Di(1 þ (∂ ln γi)/(∂ ln ci)) is the chemical K Θ a1

1146 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1144–1160 ’ 2013 ’ Vol. 46, No. 5


Theory of Chemical Kinetics and Charge Transfer Bazant

also denoted as ΔG, the free energy of reaction. The reaction hopping between adjacent minima of μex having slowly
rate eq 7 can be expressed as a nonlinear function of Δμ: varying chemical potential, |Δμ| , kBT and concentration
Δ~c , 1. Linearizing the hopping rate,
R ¼ R0 (e RΔμ=(kB T)  e(1  R)Δμ=(kB T)
) (10)
R0 Δμ k0 ~c γ
R , R0  (17)
where R, the symmetry factor or generalized Brønsted kB T γ‡
coefficient,36 is approximately constant with 0 < R < 1 for
over a distance Δx through an area ΔyΔz with ∂μ/∂x ≈
many reactions. Defining the activity coefficient of the
Δμ/Δx, we obtain eq 5 with
transition state, γ‡, by
D γ
¼ (18)
μex
‡ ¼ kB T ln γ‡ þ (1  R)μΘ
1 þ RμΘ
2 (11) D0 γ‡

where D0 = k0Δx/(cΘΔyΔz). Equation 18 can be used to


the exchange rate R0 takes the form
derive the tracer diffusivity in a concentrated solid solu-
! tion by estimating γ‡ consistent with γ. For example, for
k0 a1 1  R a2 R R γ1 1 R
γ2 R
R0 ¼ ¼ k0 ~c 1
1
~c 2 R (12) diffusion on a lattice (Figure 2c) with γ = (1  ~c )1, the
γ‡ γ‡
transition state excludes two sites, γ‡ = (1  ~c )2; the
tracer diffusivity, D = D0(1  ~c ), scales with the mean
where the term in parentheses is the thermodynamic
number of empty neighboring sites, but the chemical
correction for a concentrated solution.
diffusivity is constant, Dchem = D0 = D(0) (particle/hole
Example: Surface Adsorption. Let us apply the formal-
ism to Langmuir adsorption, A f Aads, from a liquid mixture
duality).
with μ1 = kBT ln a (Figure 2b). The surface is an ideal solution
Electrochemistry in Concentrated Solutions
of adatoms and vacancies,
Electrochemical Thermodynamics. Next we apply eq 7
~c
μ2 ¼ kB T ln þ Eads (13) to the general Faradaic reaction,
1  ~c

with coverage ~c = c/cs, site density cs, and adsorption


S1 ¼ ∑i si, O Ozi , i O
þ ne  f ∑j sj, R Rzj ,
j R
¼ S2 (19)

energy Eads = μΘ Θ
2  μ1 . Equilibrium yields the Langmuir
isotherm, converting the oxidized state OzO = ∑isi,OOzi i,O to the reduced
K Θa state RzR = ∑jsj,RRzj j,R while consuming n electrons. Let μ1 = μO þ
~c eq ¼ , K Θ ¼ e Eads =(kB T) (14)
1 þ K Θa nμe = ∑isi,Oμi,O þ nμe and μ2 = μR = ∑jsj,Rμj,R. Charge conserva-
tion implies zO  n = zR where zO = ∑isi,Ozi,O and zR = ∑jsj,Rzj,R.
If the transition state excludes s surface sites,
The electrostatic energy, zieφi, is added to μexi to define the

μex ~ electrochemical potential,


‡ ¼ skB T ln(1  c ) þ E‡ (15)
μi ¼ kB T ln ai þ μΘ ~
i þ zi eφi ¼ kB T ln c i þ μi
ex
(20)
then eq 7 yields
where zie is the charge and φi is the Coulomb potential of
R ¼ k1 (1  ~c ) [K Θ a(1  ~c )  ~c ]
s1
(16) mean force.
The electrostatic potential is φe in the electrode and φ in the
where k1 = k0 e(EadsE‡)/(kBT) and E‡ is the transition state
electrolyte. The difference is the interfacial voltage, Δφ = φe  φ.
energy relative to the bulk. With only configurational The mean electric field rφ at a point is unique, so φi = φe for
entropy, we recover standard kinetics of adsorption, Asol þ ions in the electrode and φi = φ for those in the electrolyte
sV f Asurf þ (s  1)V, involving s vacancies. With attractive solution. In the most general case of a mixed ion-electron
forces, however, eq 7 predicts novel kinetics for inhomoge- conductor, the reduced and oxidized states are split across the
neous surfaces undergoing condensation (below). interface (Figure 3a). Charge conservation implies zOe þ zOs  n =
Example: Solid Diffusion. We can also derive the LIT flux, zRe þ zRs, and the net charge nce transferred from the solution
eq 5, for macroscopic transport in a solid by activated to the electrode is given by nc = zOs  zRs = zRe  zOe þ n.

Vol. 46, No. 5 ’ 2013 ’ 1144–1160 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1147


Theory of Chemical Kinetics and Charge Transfer Bazant

FIGURE 4. Landscape of excess chemical potential explored by the


Faradaic reaction O þ ne f R in Nernst equilibrium (blue) and after a
negative overpotential η = (μ2  μ1)/(ne) is applied (red) to favor
reduction, as illustrated below.

current I = neR (per active site) is controlled by the


activation overpotential,
FIGURE 3. Types of Faradaic reactions O þ ne f R. (a) General mixed Δμ ΔG
η ¼ Δφ  Δφeq ¼ ¼ (26)
ion-electron conductor electrode/electrolyte interface. (b) Redox in ne ne
solution. (c) Ion intercalation or electrodeposition.
Specific models of charge transfer correspond to different
Let us assume that ions only exist in the electrolyte choices of μex
‡ .

(zRe = zOe = 0, nc = n) since the extension to mixed ion- Generalized ButlerVolmer Kinetics. The standard phe-
electron conductors is straightforward. For redox reactions nomenological model of electrode kinetics is the Butler
(Figure 3b), for example, Fe3þ þ e f Fe2þ, the reduced state Volmer equation,28,45
is in the solution at the same potential, φR = φO = φ. For
I ¼ I0 (e Rc neη=(kB T)  eRa neη=(kB T) ) (27)
electrodeposition (Figure 3c), for example, Cu2þ þ 2e f Cu,
or ion intercalation as a neutral polaron, for example, CoO2 þ where I0 is the exchange current. For a single-step charge-
Liþ þ e f LiCoO2, the reduced state is uncharged, zR = 0, so we
transfer reaction, the anodic and cathodic charge-transfer
can also set φR = φ, even though it is in the electrode. For this
coefficients Ra and Rc satisfy Ra = 1  R and Rc = R with a
broad class of Faradaic reactions, we have
symmetry factor 0 < R < 1. The exchange current is
μO ¼ kB T ln aO þ μΘ
O þ zO eφ (21) typically modeled as I0 µ cROacRR c, but this is a dilute solution
approximation.
μR ¼ kB T ln aR þ μΘ
R þ zR eφ (22)
In concentrated solutions, the exchange current is af-
μe ¼ kB T ln ae þ μΘ
e  eφe (23) fected by configurational entropy and enthalpy, electrostatic
Q correlations, coherency strain, and other nonidealities. For
(aO ¼
i
asi i , μΘ
O ¼ ∑i si μΘi ) where μe is the Fermi level, Li-ion batteries, only excluded volume has been considered,
which depends on φe and the electron activity ae = γe~c e. using27,28 I0(c) µ (cs  c)RccRa. For fuel cells, many phenom-
In equilibrium (μ1 = μ2), the interfacial voltage is given by enological models have been developed for electrocatalytic
the Nernst equation reactions with surface adsorption steps.4648 Electrocataly-
sis can also be treated by our formalism,33 but here we focus
kB T aO ae n
Δφeq ¼ EΘ þ ln (24) on the elementary charge-transfer step and its coupling to
nc e aR
phase transformations, which has no prior literature.
where nc = n and In order to generalize BV kinetics (Figure 4), we model the
μΘ Θ Θ
O þ nμe  μR
transition state
EΘ ¼ (25)
ne μex Θ Θ Θ
‡ ¼ kB T ln γ‡ þ (1  R)(zO eφ  neφe þ μO þ nμe ) þ R(zR eφ þ μR )

is the standard half-cell potential. Out of equilibrium, the (28)

1148 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1144–1160 ’ 2013 ’ Vol. 46, No. 5


Theory of Chemical Kinetics and Charge Transfer Bazant

by averaging the standard chemical potential and elec-


trostatic energy of the initial and final states, which
assumes a constant electric field across the reaction
coordinate x with R = (x‡  xR)/(xO  xR). Substituting
eq 28 into eq 7 using eq 24, we obtain eq 27 with
1 R
k0 ne(aO ae n ) aR R
I0 ¼
γ‡
" #
1 R
n 1 R R (γO γe n ) γR R
¼ k0 ne(~c O ~c e ) ~c R (29)
γ‡

The factor in brackets is the thermodynamic correction


for the exchange current.
Generalized BV kinetics (eqs 27 and 29) consistently applies
chemical kinetics in concentrated solutions (eqs 10 and 12,
respectively) to Faradaic reactions. In Li-ion battery models,
Δφeq(c) is fitted to the open circuit voltage, and I0(c) and Dchem(c)
are fitted to discharge curves,27,29,30 but these quantities
are related by nonequilibrium thermodynamics.9,26,35 Lai and
Ciucci49,50 also recognized this inconsistency and used eqs 5 and
24 in battery models, but they postulated a barrier of total
(not excess) chemical potential, in contrast to eq 7, eq 29, and
charge-transfer theory.
Generalized Marcus Kinetics. The microscopic theory of
charge transfer, initiated by Marcus52,53 and honored by the FIGURE 5. (top) The Faradaic reaction O þ ne f R in concentrated
Nobel Prize in Chemistry,54 provides justification for the BV solutions. Each state explores a landscape of excess chemical potential μex.
Charge transfer occurs where the curves overlap, or just below, by
equation and a means to estimate its parameters based on quantum tunneling (dashed curves). (bottom) Example of ion
solvent reorganization.45 Quantum mechanical formula- intercalation into a solid electrode from a liquid electrolyte.
tions pioneered by Levich, Dogonadze, Marcus, Kuznetsov,
Θ kR 2
and Ulstrup further account for Fermi statistics, band struc- 2 (x) ¼ μR þ kB T ln γR þ zR eφ þ
μex (x  xR ) (31)
2
ture, and electron tunneling.36 Most theories, however,
make the dilute solution approximation by considering an The Nernst equation, eq 24, follows by equating the total
isolated reaction complex. chemical potentials at the local minima, μ1(xO) = μ2(xR) in
In order to extend Marcus theory for concentrated solu- equilibrium. The free energy barrier is set by the intersection
tions, our basic postulate (Figure 5) is that the Faradaic of the excess chemical potential curves, μex
‡ = μ1 (x‡) = μ2 (x‡),
ex ex

reaction eq 19 occurs when the excess chemical potential which determines the barrier position, x = x‡, and implies
of the reduced state, deformed along the reaction coordi- kO 2 kR 2
ΔGex ¼ μex
2 (xR )  μ1 (xO ) ¼
ex
(x‡  xO )  (x‡  xR ) (32)
nate by statistical fluctuations, equals that of the oxidized 2 2
state (plus n electrons in the electrode) at the same point.
where ΔGex is the excess free energy change per reaction.
(More precisely, charge transfer occurs at slightly lower
From eq 26, the overpotential is the total free energy
energies due to quantum tunneling.36,45) Following Marcus,
change per charge transferred,
we assume harmonic restoring forces for structural relaxa-
~c R
tion (e.g., shedding of the solvation shell from a liquid or ion neη ¼ ΔG ¼ ΔGex þ kB T ln (33)
~c O ~c e n
extraction from a solid) along the reaction coordinate x from
the oxidized state at xO to the reduced state at xR: In classical Marcus theory,45,54 the overpotential is de-
kO
fined by neη = ΔGex without the concentration factors
Θ Θ 2
μex
1 (x) ¼ μO þ nμe þ kB T ln(γO γe ) þ zO eφ  neφe þ
n
(x  xO ) (30)
2 required by nonequilibrium thermodynamics, which is

Vol. 46, No. 5 ’ 2013 ’ 1144–1160 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1149


Theory of Chemical Kinetics and Charge Transfer Bazant

valid for charge-transfer reactions in bulk phases (A þ Comparing eq 38 with eq 29 for R ≈ 1/2, we can relate the
B f A þ B) because the initial and final concentrations are reorganization energy to the activity coefficients defined
the same, and thus ΔG = ΔGex = ΔG° (standard free energy above
γ‡
of reaction). For Faradaic reactions at interfaces, however, λ  4kB T ln 1=2
(40)
the concentrations of reactions and products are different, (γO γe n γR )
and eq 33 must be used. The missing “Nernst concentration For a dilute solution, the reorganization energy λ0 can
term” in eq 33 has also been noted by Kuznetsov and be estimated by the classical Marcus approximation,
Ulstrup (p 219 of ref 36). λ0 = λi þ λo, where λi is the “inner” or short-range
In order to relate μex
‡ to ΔG , we solve eq 32 for x‡. In the
ex
contribution from structural relaxation (sum over normal
simplest approximation, kO = kR = k, the barriers for the
modes) and λo is the “outer” or long-range contribution
cathodic and anodic reactions,
from the Born energy of solvent dielectric relaxation.54,45
 
λ ΔGex 2 For polar solvents at room temperature, the large Born
ΔGc ¼ μ‡  μ1 (xO ) ¼
ex ex ex
1þ (34)
4 λ energy, λo > 0.5n2 eV ≈ 20n2kBT (at room temperature),
  implies that single-electron (n = 1), symmetric (R ≈ 1/2)
λ ΔGex 2
a ¼ μ‡  μ2 (xR ) ¼
ΔGex ex ex
1 (35) charge transfer is favored. Quantum mechanical approxima-
4 λ
tions of λ0 are also available.36 For a concentrated solution,
are related to the reorganization energy, λ = (k/2)(xO  xR) . 2
we can estimate the thermodynamic correction, γc‡, for the
These formulas contain the famous “inverted region” entropy and enthalpy of the transition state and write
predicted by Marcus for isotopic exchange,54 where
γ‡ ¼ γc‡ eλ0 =(4kB T) (41)
(say) the cathodic rate, kc µ eΔGc /(kBT) reaches a mini-
ex

mum and increases again with decreasing driving force which can be used in either Marcus (eqs 3740) or BV
ΔG , for x‡ < xR in Figure 5a. This effect remains for
ex
(eqs 2729) kinetics. An example for ion intercalation is
charge transfer in concentrated bulk solutions, for exam- given below, eq 81, but first we need to develop a
ple, A þ B f A þ B. For Fardaic reactions, however, it is modeling framework for chemical potentials.
suppressed at metal electrodes, since electrons can tun-
nel through unoccupied conduction-band states, but can Nonequilibrium Chemical Thermodynamics
36,53,54 General Theory. In homogeneous bulk phases, activity
arise in narrow-band semiconductors.
Substituting μex
‡ into eq 7, we obtain
coefficients depend on concentrations, but for reactions at
ex 2
an interface, concentration gradients must also play a role
R ¼ k0 e λ=(4kB T) e (ΔG ) =(4kB Tλ)
(~c O ~c e e ΔG =(2kB T)
n ex
(Figure 1). The main contribution of this work has been to
 ~c R eΔG =(2kB T)
ex
) (36) formulate chemical kinetics for inhomogeneous, nonequilib-
rium systems. The most general theory appears here for the
Using eq 33, we can relate the current to the overpotential, first time, building on my lectures notes.33
2
I ¼ I0 e (neη) =(4kB Tλ)
(e Rneη=(kB T)  e(1  R)neη=(kB T)
) (37) The theory is based on the Gibbs free energy functional

via the exchange current, Z I


G[fci g] ¼ g dV þ γs dA ¼ Gbulk þ Gsurf (42)
I0 ¼ nek0 e λ=(4kB T) (~c O ~c e )
n (3  2R)=4
~c R (1 þ 2R)=4 (38) V A

and symmetry factor, with integrals over the bulk volume V and surface area A.
!
1 kB T ~c O ~c e
n The variational derivative,55
R¼ 1þ ln (39)
2 λ ~c R
δG G[ci (y) þ εδε (y  x)]  G[ci (y)]
(x) ¼ lim (43)
In the typical case λ . kBT, the current eq 37 is well δci εf0 ε
approximated by the BV equation with R = 1/2 at moderate
overpotentials, |η| > (kBT/(ne))(λ/(kBT))1/2, and nondepleted is the change in G to add a “continuum particle” δ(yx) of
concentrations, |ln ~c | , (λ/(kBT)). species i at point x, where δε(z) f δ(z) is a finite-size

1150 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1144–1160 ’ 2013 ’ Vol. 46, No. 5


Theory of Chemical Kinetics and Charge Transfer Bazant

approximation for a particle that converges weakly


(in the sense of distributions) to the Dirac delta function,
pffiffiffiffiffiffiffiffi
for example, δε(z) = ez /(2ε)/ 2πε. This is the consistent
2

definition of diffusional chemical potential,39,56


δG
μi ¼ (44)
δci

If g depends on {ci} and {rci}, then


Dg Dg
μi ¼ r3 (45)
Dci Drci
The continuity of μi at the surface yields the “natural
boundary condition”, FIGURE 6. Types of reactions (R) in nonequilibrium chemical
Dg Dγ thermodynamics. (a) Heterogeneous chemistry at a surface, eq 54.
^3
n ¼ s (46) (b) Homogeneous chemistry, eq 56, with diffusing species. (c) phase
Drci Dci
transformations or homogeneous reactions with immobile species, eq 58.
We can also express the activity variationally,
 
1 δGmix reaction kinetics that is consistent with nonequilibrium
ai ¼ exp (47)
kB T δci thermodynamics. The reaction rate is a nonlinear func-
tion of the thermodynamic driving force,
in terms of the free energy of mixing
Z δG δG
Gmix ¼ Gbulk  Θ
μi ∑ici dV (48)
Δμ ¼
p
sp
δcp

r

sr
δcr ∑ (52)
V

which we define relative to the standard states of This is the most general, variational definition of the free
each species. energy of reaction. For |Δμ| , kBT, the rate expression (51)
The simplest approximation for an inhomogeneous sys- can be linearized as
k0 eμ‡ =kB T
ex
tem is the CahnHilliard56 (or LandauGinzburg, or van der
R ∼  kΔμ, k ¼ (53)
Waals57) gradient expansion, kB T
0 1
1 but more generally, the forward and backward rates have
g ¼ g(fci g) þ ∑
i
@μΘ ci þ
i
2 j

r~c i 3 Kij r~c j A (49)
exponential, Arrhenius dependence on the chemical po-
tential barriers. The variational formulation of chemical
for which
kinetics, eq 51, can be applied to any type of reaction
Dg ~c j
μi  μΘ
i ¼ kB T ln ai ¼
Dci
 ∑j r 3 Kij r Θ
cj
(50) (Figure 6), as we now explain.
Heterogeneous Chemistry. At an interface, eq 51 pro-
vides a new reaction boundary condition6,9,12,35
where g is the homogeneous free energy of mixing    !
Di ci δG δG
and where κij is a 2nd rank anisotropic tensor penalizing ^3 B
si A r n uci  r ¼ (R (54)
kB T δci δcj
gradients in components i and j. (Higher-order derivative
terms can also be added.58,59) (þ for reactants,  for products; Ar = reaction site area) for
With these definitions, eq 7 takes the variational form, the CahnHilliard (CH) equation,39
" !  
sr δG Dci Di ci δG
R ¼ k0 e μex

=(kB T)
exp ∑
r kB T δcr
Dt
þB
u 3 rci ¼ r 3 r
kB T δci
(55)

!#
sp δG expressing mass conservation for the LIT flux, eq 5, with
 exp ∑ p kB T δcp
(51) convection in a mean flow uB. For thermodynamic con-
sistency, Di is given by eq 18, which reduces eq 55 to the
for the general reaction, eq 3, in a concentrated solution. “modified” CH equation58 in an ideal mixture.26 This is the
This is the fundamental expression of thermally activated “CahnHilliard reaction (CHR) model”.

Vol. 46, No. 5 ’ 2013 ’ 1144–1160 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1151


Theory of Chemical Kinetics and Charge Transfer Bazant

FIGURE 7. Surface adsorption with condensation when an empty surface is brought into contact with a reservoir (μres = μ1 = kBT ln a > Eads = kBT ln
KΘ). (left) Homogeneous chemical potential of the adsorbed species μ. (right) (A) Early stage uniform adsorption and (B) late-stage adsorption waves
nucleated at edges, where the reaction is focused on advancing boundaries of the condensed phase.

Homogeneous Chemistry. For bulk reactions, eq 51 pro- strong enough to drive adatom condensation (separation
vides a new source term for the CH equation, into high- and low-density phases) on the surface.33
   ! Applications may include water adsorption in concrete60
Dci Di ci δG cs δG
u 3 rci ¼ r 3
þB r - R (56) or colloidal deposition in electrophoretic displays.61 Fol-
Dt kB T δci si δcj
lowing Cahn and Hilliard,56 the simplest model is a regular
(cs = reaction sites/volume). The AllenCahn equation39 solution of adatoms and vacancies with pair interaction
(AC) corresponds to the special case of an immobile reactant energy Ω,
B = 0) evolving according to linear kinetics, eq (53),
(Di = 0, u g ¼ cs fkB T[~c ln ~c þ (1  ~c ) ln(1  ~c )] þ Ω~c (1  ~c )
K
although the exchange-rate prefactor k is usually taken to be þ Eads ~c g þ jr~c j2 (59)
2
constant, in contrast to our nonlinear theory. Equation 56 is
~c K 2
the fundamental equation of nonequilibrium chemical ther- μ ¼ kB T ln þ Ω(1  2~c ) þ Eads  r ~ ~c (60)
1  ~c cs
modynamics. It unifies and extends the CH and AC equations
via a consistent set of reaction-diffusion equations based on Below the critical point, T < Tc = Ω/(2kB), the enthalpy of
variational principles. Equation 54 is its integrated form for a adatom attraction (third term, favoring phase separation
reaction localized on a boundary. ~c = 0, 1) dominates the configurational entropy of ada-
Phase Transformations. In the case of an immobile toms and vacancies (first two terms, favoring mixing ~c =
reactant with two or more stable states, our general reaction- 1/2). The gradient term controls spinodal decomposition
diffusion equation, eq 56, also describes thermally activated and stabilizes phase boundaries of thickness λb =
phase transformations. The immobile reactant concentra- (κ/(csΩ))1/2 and interphasial tension γb = (κcsΩ)1/2. Using
tion acts as a non-conserved order parameter, or phase field, eq 15 to model the transition state with one excluded site,
representing different thermodynamic states of the same
s = 1, the ACR model, eq 58, takes the dimensionless form,
molecular substance. For example, if g(c) has two local
D~c ~
equilibrium states, cA and cB, then ¼ K Θ a(1  ~c )  ~c exp(Ω (1  2~c )  ~
Kr~ 2 ~c ) (61)
D~t
c  cA
ξ¼ (57) ~ = Ω/(k T) = 2T /T, ~
cA  cB where ~t = k1t, Ω B c κ = κ/(L2cskBT) and
~
r = Lr (with length scale L). This nonlinear PDE describes
is a phase field with minima at ξ = 0 and ξ = 1 satisfying
  phase separation coupled to adsorption at an interface
Dξ δG
¼R (58) (Figure 7), controlled by the reservoir activity a. It resem-
Dt δξ
bles a reaction-diffusion equation, but there is no diffu-
sion; instead, κ~r ~2~c is a gradient correction to the
This is the “AllenCahn reaction (ACR) model”, which is a
nonlinear generalization of the AC equation for chemical chemical potential, which nonlinearly affects the adsorp-
kinetics.7,9,12,33 tion reaction rate. With modifications for charge transfer
Example: Adsorption with Condensation. To illustrate and coherency strain, a similar PDE describes ion inter-
the theory, we revisit surface adsorption with attractive forces, calation in a solid host, driven by an applied voltage.

1152 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1144–1160 ’ 2013 ’ Vol. 46, No. 5


Theory of Chemical Kinetics and Charge Transfer Bazant

Nonequilibrium Thermodynamics of potential of mean force, φ; cB is the set of concentrations


Electrochemical Systems (including electrons for mixed ion/electron conductors);
Background. We thus return to our original motivation, f is the homogeneous Helmholtz free energy density, Fe
phase separation in Li-ion batteries (Figure 1). Three impor- and qs are the bulk and surface charge densities; εp is the
tant papers in 2004 set the stage: Garcia et al.62 formulated permittivity tensor; and σ and ε are the stress and strain
variational principles for electromagnetically active systems, tensors. The potential φ acts as a Lagrange multiplier
which unify the CH equation with Maxwell's equations; constraining the total ion densities7,62 while enforc-
Guyer et al.64 represented the metal/electrolyte interface
ing Maxwell's equations for a linear dielectric material
with a continuous phase field ξ evolving by AC kinetics;
(δG/δφ = 0),
Han et al.42 used the CH equation to model diffusion in LFP,
leading directly to this work.  r 3 εp rφ ¼ Fe ¼ ∑i zi eci (66)
When the time is ripe for a new idea, a number of scientists
naturally think along similar lines. As described in the Introduc- ^ 3 εp rφ ¼ qs
n (67)
tion, my group first reported phase-field kinetics (CHR and
ACR)12,43 and modified PoissonNernstPlanck (PNP) equa-
The permittivity
 can
 be a linear operator, for example,
tions41 in 2007, the generalized BV equation35 in 2009, and εp = ε0 1  l c r , to account for electrostatic correlations
2 2

the complete theory9,33 in 2011. Independently, Lai and Ciucci in ionic liquids59 and concentrated electrolytes35,68 (as first
also applied nonequilibrium thermodynamics to electroche- derived for counterion plasmas69,70). Modified PNP equa-
mical transport49 but did not develop a variational formulation. tions35,41,49,50 correspond to eqs 55 and 66.
They proceeded to generalize BV kinetics50,51 (citing Singh et al.12) For elastic solids, the stress is given by Hooke's
but used μ in place of μex and neglected γex ‡ . Tang et al.
65
law, σij = Cijklεkl, where C is the elastic constant tensor.
were the first to apply CHR to ion intercalation with coher- The coherency strain,
ency strain, but like Guyer et al.,64 they assumed linear AC !
1 Dui Duj
kinetics. Recently, Liang et al.66 published the BVACR εij ¼ þ
2 Dxj Dxi

m
ε0ijm ~c m ∑
(68)
equation, claiming that “in contrast to all existing phase-field
models, the rate of temporal phase-field evolution... is is the total strain due to compositional inhomogeneity
considered nonlinear with respect to the thermodynamic
(first term) relative to the stress-free inelastic strain
driving force”. They cited my work6,7,9,12 as a “boundary
(second term), which contributes to Gmix. In a mean-
condition for a fixed electrodeelectrolyte interface” (CHR)
field approximation (Vegard's law), each molecule
but overlooked the same BVACR equation for the depth-
of species m exerts an independent strain ε0m
averaged ion concentration,9,12 identified as a phase field
(lattice misfit between ~c m = 0, 1 with cΘ
m = c s ). Since
for an open system.7,9 They also set I0 = constant, which
contradicts chemical kinetics (see below). elastic relaxation (sound) is faster than diffusion
Variational Electrochemical Kinetics. We now apply and kinetics, we assume mechanical equilibrium,
phase-field kinetics to charged species. The Gibbs free δG/δuB = r 3 σ = 0.
energy of ionic materials can be modeled as:6,7,9,12,59,62,63,67 For Faradaic reactions, eq 19, the overpotential is the
Z thermodynamic driving force for charge transfer,
G ¼ Gmix þ Gelec þ Gsurf þ ∑i μΘ
i ci dV (62) δG δG δG
Z
V neη ¼ ∑j sj, R δcj, R  ∑i si, O δci, O  nδce (69)
Gmix ¼ B) dV þ Ggrad
f (c (63)
V determined by the electrochemical potentials, μi = δG/δci.
Z
1 For thermodynamic consistency, the diffusivities, eq 18,
Ggrad ¼ (r B~c 3 Kr B
~c  rφ 3 εp rφ þ σ : ε) dV (64)
2 V Nernst voltage, eq 24, and exchange current, eq 29, must
Z I depend on cB, rcB, and σ via the variational activities,
Gelec ¼ Fe φ dV þ qs φ dA (65)
V A
eq 47, given by

Df r Kr~c i þ σ : ε0i Dεp


 3
where Ggrad is the free energy associated with all gradi-
kB T ln ai ¼  rφ 3 rφ (70)
ents; Gelec is the energy of charges in the electrostatic Dci cs Dci

Vol. 46, No. 5 ’ 2013 ’ 1144–1160 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1153


Theory of Chemical Kinetics and Charge Transfer Bazant

for the ionic model above. The Faradaic current density is The ACR equation, eq 58, for ξ = ~c using eqs 7177 differs
  from prior phase-field models of electrodeposition.64,66
neη
I ¼ I0 F (71) Equation 76 has the thermodynamically consistent depen-
kB T
dence on reactant activities (rather than I0 = constant).
where Coupled with generalized PNP equations, eq 56, for ~c (, our
 ~ ~
~) ¼
F(η e Rη  e(1  R)η ButlerVolmer theory also describes Frumkin corrections to ButlerVolmer
~2 ~ ~ ~
e η =(4λ ) (e Rη  e(1  R)η
) Marcus kinetics72,73 and electro-osmotic flows35,71 associated with
(72) double-layer diffuse charge. The permittivity ε p (~c , ~c ()
depends on the metal concentration, interpolating between
and I0 is given by either eq 29 or eqs 3841, respectively
p ,~
metal (εM c = 1) and bulk electrolyte (εbp, ~c = 0) values63 with a
~
( λ = λ/(kBT )). The charge-transfer rate, R = I/(ne), defines minimum (εSp, ~c ≈ 1/2) for the Stern layer,72 as well as the
the CHR and ACR models, eqs 5458, for electrochemical ionic concentrations, εp µ ε0(1 þ Rþ~c þ þ R~c ) where R( < 0
systems. are the excess polarizabilities.74,75 Dielectric saturation in
Example: Metal Electrodeposition. In models of electro- large fields, εp(|rφ|), can also be included.35 The charge
deposition63,64 and electrokinetics,71 the solid/electrolyte density includes the conduction electrons and solid metal
interface is represented by a continuous phase field, ξ, for ions, Fe = ze(cþ  c) þ zMe(cM  ce). Neglecting band
numerical convenience (to avoid tracking a sharp interface). If structure variations, we can set ae = ~c e and aM = ~c M. The
the phase field evolves by reactions, however, it has physical theory thus predicts diffuse-charge profiles on both sides of
significance as a chemical concentration. For example, the interface, involving ions and electrons,76 during Faradaic
consider electrodeposition, Mnþ þ ne f M, of solid metal reactions.
M from a binary electrolyte MþA with dimensionless Example: Ion Intercalation. Hereafter, we neglect double
concentrations, ξ = ~c = c/cs and ~c ( = c(/c0, respectively. layers and focus on solid thermodynamics. Consider cation
(The classical “phase field” is ξ = ~c .) In order to separate the intercalation, Anþ þ B þ ne f AB, from an electro-
metal from the electrolyte, we postulate lyte reservoir (aO = constant) into a conducting solid B
(ae = constant) as a neutral polaron (cR = c(x,t), zR = 0).
f ¼ W[h(~c ) þ ~c (~c þ þ ~c  )] þ fion (~c þ , ~c  ) (73)
The overpotential eq 69 takes the simple form
with W . kBT, where h = ~c 2(1  ~c )2 is an arbitrary double-
δG δGmix
welled potential. For a dilute electrolyte, fion = kBT(~c þ ln neη ¼  (μO þ nμe ) ¼ þ neΔΦ (78)
δc δc
~c þ þ ~c  ln ~c ), without phase separation,67 we include
gradient energy only for the metal. The activities, eq 70, where
for reduced metal kB T
ΔΦ ¼ Δφ  EΘ  ln aO ane (79)
0 Dεp ne
cs kB T ln a ¼ W[h (~c ) þ ~c þ þ ~c  ]  Kr2 ~c  jrφj2
D~c is the interfacial voltage relative to the ionic standard
(74)
state. The equilibrium voltage is
and metal cations δGmix
Dεp neΔΦeq ¼ kB T ln a ¼  (80)
c0 kB T ln a þ ¼ W ~c þ kB T ln ~c þ  jrφj2 (75) δc
D~c þ
Note that potentials can be shifted for convenience: Bai
define the current density eq 71 via
et al.9 and Ferguson and Bazant26 set μΘ = 0 for ions, so
nek0 cs μ = kBT ln a = δGmix/δc; Cogswell and Bazant7 defined
I0 ¼ K0 aR (a þ ae n )
1R
, K0 ¼ (76)
γ‡ “Δφ” = ΔΦ and shifted g by cΔΦ, so eη = δG/δc.
Our surface adsorption model, eq 59, can be adapted for ion
kB T a
η¼ ln  EΘ (77) intercalation by setting Ea = eΔΦ. If the transition state excludes
ne a þ ae n
s sites (where s > 1 could account for the Anþ solvation shell)
Note that the local potential for electrons and ions is and has strain ε‡, then its activity coefficient, eq 41, is
unique (φ = φe, Δφ = 0), but integration across the diffuse
s
γ‡ ¼ (1  ~c ) ~ : ε‡ þ ~
exp(σ λ 0 =4) (81)
interface yields the interfacial voltage.

1154 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1144–1160 ’ 2013 ’ Vol. 46, No. 5


Theory of Chemical Kinetics and Charge Transfer Bazant

FIGURE 8. Suppression of phase separation at constant current in a Li-ion battery nanoparticle (ACR model without coherency strain or surface
wetting).9 (a) Linear stability diagram for the homogeneous state versus dimensionless current, ~I = I/I0 (~c = 0.5), and state of charge X. (b) Battery
voltage versus X with increasing ~I. (c) Concentration profiles, spinodal decomposition at ~I = 0.01 leading to intercalation waves (Figure 1(c)), quasi-solid
solution at ~I = 0.25, and homogeneous filling at ~I = 2.

where ~
λ 0 = λ0/(kBT) and ~
σ = σ/(cskBT). The exchange ~c ~
~ K~
þ Ω(1  2~c )  ~
2
μ ¼ ln r ~c þ ~
σ:ε (85)
current, eq 29, is 1  ~c
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  

I0 ¼ nek(~c )~c R (1  ~c )sR exp(σ ~


~ : Δε þ RΩ ~ ~~
(1  2~c )  Rr 3 ~
Krc ) ~I0 ¼ ~c (1  ~c ) exp 1 Ω ~
(1 2~c )  ~ ~
Kr ~c
2
(86)
2
(82)
~
~
1R λ where ΔΦ = neΔΦ/(kBT), ~t = Nskt. The total current
k(~c ) ¼ k0 cs (a þ ane (~c )) e 0 =4 (83)
integrated over the active facet
where aþ is the ionic activity in the electrolyte and Δε = ε‡  Z
~
~I(~t) ¼ Dc dx~ dy~ (87)
Rε0 is the activation strain.77 For semiconductors, the elec- ~ D~
A t
tron activity ae = eΔEf/(kBT) depends on ~c , if the intercalated ~
is controlled while solving for ΔΦ(~t) (as in Figure 8) or vice
ion shifts the Fermi level by donating an electron to the ~
versa (where A is the dimensionless surface area of the
conduction band, for example, ΔEf µ (1 þ β~c )2/d for free
active facet).
electrons in d dimensions (as in LiWO3 with d = 378).
Intercalation Waves and Quasi-Solid Solutions. The
Application to Li-Ion Battery Electrodes theory predicts a rich variety of new intercalation mechan-
AllenCahn Reaction Model. The three-dimensional isms. A special case of the CHR model12 is isotropic diffusion-
CHR model eqs 5455 with current density I = neR given limited intercalation27,28 with a shrinking-core phase
by eqs 71 and 82 describes ion intercalation in a solid particle boundary,29,30 but the reaction-limited ACR model also
from an electrolyte reservoir. In nanoparticles, solid diffusion predicts intercalation waves (or “domino cascades”15),
times (milliseconds to seconds) are much shorter than discharge sweeping across the active facet, filling the crystal layer by
times, so a reaction-limited ACR model is often appropriate. In layer (Figure 1c).7,9,12,34,65 Intercalation waves result from
the case of LFP nanoparticles, strong crystal anisotropy leads to a spinodal decomposition or nucleation at surfaces9 and trace
two-dimensional ACR model over the active (010) facet by depth out the voltage plateau at low current (Figure 8).
averaging over Ns sites in the [010] direction.9,12 For particle sizes The theory makes other surprising predictions about elec-
below 100 nm, the concentration tends to be uniform in [010] trochemically driven phase transformations. Singh et al.12
due to the fast diffusion13 (uninhibited by Fe anti-site showed that intercalation wave solutions of the ACR equation
defects16) and elastically unfavorable phase separation.7 only exist over a finite range of thermodynamic driving force.
Using eqs 71 and 82 with isotropic ~ κ, ae = constant, ε‡ = Based on bulk free energy calculations, Malik et al.10 argued for
Rε0, R = 1/2, and s = 1, the ACR equation, eq 58, takes the a “solid solution pathway” without phase separation under
simple dimensionless form,7,9 applied current, but Bai et al.9 used the BVACR model to show
D~c ~ ~) that phase separation is suppressed by activation overpoten-
¼ I0 F( ~μ þ ΔΦ (84)
D~t tial at high current (Figure 8), due to the reduced area for

Vol. 46, No. 5 ’ 2013 ’ 1144–1160 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1155


Theory of Chemical Kinetics and Charge Transfer Bazant

FIGURE 9. ACR simulations of galvanostatic discharge in a 100 nm LiXFePO4 nanoparticle.7 As the current is increased, transient quasi-solid solutions
(images from the shaded region) transition to homogeneous filling for ~I > 0.1, as phase separation is suppressed.

FIGURE 10. Phase separation of a 500 nm particle of Li0.5FePO4 into Li-rich (black) and Li-poor phases (white) at zero current in ACR simulations,7
compared with ex situ experimental images.31,32 (a) Coherent phase separation with [101] interfaces. (b) Semicoherent phase separation, consistent
with observed {100} microcracks.31

intercalation on the phase boundary (Figure 1c). Linear stability battery operation in LFP nanoparticles, which helps to ex-
analysis of homogeneous filling predicts a critical current, plain their high rate capability and extended lifetime.7,9
on the order of the exchange current, above which phase Phase separation occurs at low currents and can be observed
separation by spinodal decomposition is impossible. Below ex situ in partially filled particles (Figure 10). Crystal anisotropy
this current, the homogeneous state is unstable over a range of leads to striped phase patterns in equilibrium,1719 whose
concentrations (smaller than the zero-current spinodal gap), spacing is set by the balance of elastic energy (favoring short
but for large currents, the time spent in this region is too small wavelengths at a stress-free boundary) and interfacial energy
for complete phase separation. Instead, the particle passes (favoring long wavelengths to minimize interfacial area).7
through a transient “quasi-solid solution” state, where its Stanton and Bazant19 predicted that simultaneous positive
voltage and concentration profile resemble those of a and negative eigenvalues of ε0 make phase boundaries tilt
homogeneous solid solution. When nucleation is possible with respect to the crystal axes. In LFP, lithiation causes con-
(see below), a similar current dependence is also observed. traction in the [001] direction and expansion in the [100] and
For quantitative interpretation of experiments, it is essen- [010] directions.31 Depending on the degree of coherency,
tial to account for the elastic energy.7 Coherency strain is a Cogswell and Bazant7 predicted phase morphologies in ex-
barrier to phase separation (Figure 9), which tilts the voltage cellent agreement with experiments (Figure 10) and inferred
plateau (compared with Figure 8) and reduces the critical the gradient penalty κ and the LiFePO4/FePO4 interfacial
current, far below the exchange current. An unexpected tension (beyond the reach of molecular simulations) from
prediction is that phase separation rarely occurs in situ during the observed stripe spacing.

1156 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1144–1160 ’ 2013 ’ Vol. 46, No. 5


Theory of Chemical Kinetics and Charge Transfer Bazant

FIGURE 11. (a) ACR simulation of galvanostatic nucleation in a realistic LFP nanoparticle shape (C3)76 with a 150 nm  76 nm top (010) active facet.8
Surface “wetting” of the side facets by lithium nucleates intercalation waves that propagate across the particle (while bending from coherency strain)
after the voltage exceeds the coherent miscibility limit. (b) Discharge plot indicating nucleation by fluctuations in voltage or composition.8 (c) Collapse
of experimental data for the nucleation voltage by the theory, without any fitting parameters.8 (d) Size dependence of the miscibility gap, fitted by the
theory.7

Driven Nucleation and Growth. The theory can also The agreement between theory and experiment, without
quantitatively predict nucleation dynamics driven by che- fitting any parameters, is impressive (Figure 11). Using our
mical reactions. Nucleation is perhaps the least understood prior ACR model7 augmented only by ab initio calculated
phenomenon of thermodynamics. In thermal phase transi- surface energies (in eq 46), the theory is able to collapse Eb
tions, such as boiling or freezing, the critical nucleus is data for LFP versus A/V, which lie either on the predicted line or
controlled by random heterogeneities, and its energy is below (e.g., from heterogeneities, lowering Eb, or missing
overestimated by classical spherical droplet nucleation theory. the tiniest nanoparticles, lowering A/V).8 This resolves a
Phase-field models address this problem but often lack suffi- major controversy, since the data had seemed inconsistent
cient details to be predictive. (Eb = 2.037 mV), and some had argued for12,22,77 and others
For battery nanoparticles, nucleation turns out to be more against the possibility of nucleation (using classical droplet
tractable, in part because the current and voltage can be more theory).10 The new theory also predicts that the nucleation
precisely controlled than heat flux and temperature. More barrier (Figure 11c) and miscibility gap (Figure 11d) vanish at
importantly, the critical nucleus has a well-defined form, set the same critical size, dc ≈ 22 nm, consistent with separate
by the geometry, due to strong surface “wetting” of crystal Li-solubility experiments.11
facets by different phases. Cogswell and Bazant8 showed that Mosaic Instability and Porous Electrodes. These findings
nucleation in binary solids occurs at the coherent miscibility have important implications for porous battery electrodes,
limit, as a surface layer becomes unstable and propagates into consisting of many phase-separating nanoparticles. The
the bulk. The nucleation barrier, Eb = eΔΦ is set by coherency prediction that small particles transform before larger ones
strain energy (scaling with volume) in large particles and is counterintuitive (since larger particles have more nuclea-
reduced by surface energy (scaling with area) in nanoparticles. tion sites) and opposite to classical nucleation theory. The
The barrier thus decays with the wetted area-to-volume new theory could be used to predict mean nucleation and
ratio, A/V, and vanishes at a critical size, below which nano- growth rates in a simple statistical model77 that fits current
particles remain homogeneous in the phase of lowest surface transients in LFP22 and guide extensions to account for the
energy. particle size distribution.

Vol. 46, No. 5 ’ 2013 ’ 1144–1160 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1157


Theory of Chemical Kinetics and Charge Transfer Bazant

FIGURE 12. Finite-volume simulations of a porous LFP cathode (Ferguson and Bazant26). (a) Voltage versus state of charge at different rates with
profiles of the mean solid Li concentration (AC), separator on the left, current collector on the right. (b) SEM image of LFP nanoparticles represented by
three finite volumes (P. Bai). (c) Experiments revealing a zero-current gap between noisy charge and discharge voltage plateaus (From Dreyer et al.23).

Discrete, random transformations also affect voltage macroscopic transport), where the active particles remain
transients. Using the CHR model6 for a collection of particles homogeneous. Using our model for LFP nanoparticles,7 the
in a reservoir, Burch25 discovered the “mosaic instability”, porous electrode theory predicts the zero-current voltage
whereby particles switch from uniform to sequential filling gap, without any fitting (Figure 12). (Using the mean particle
after entering the miscibility gap. Around the same time, size, the gap is somewhat too large, but this can be corrected
Dreyer et al.23 published a simple theory of the same effect by size-dependent nucleation (Figure 11), implying that smal-
(neglecting phase separation within particles) supported by ler particles were preferentially cycled in the experiments.)
experimental observations of voltage gap between charge/ Voltage fluctuations at low current correspond to discrete sets
discharge cycles in LFP batteries (Figure 12c), as well as of transforming particles. For a narrow particle size distribu-
pressure hysteresis in ballon array.24 tion, mosaic instability sweeps across the electrode from the
The key ingredient missing in these models is the transport of separator as a narrow reaction front (Figure 12a, inset). As the
ions (in the electrolyte) and electrons (in the conducting matrix), current is increased, the front width grows, and the active
which mediates interactions between nanoparticles and be- material transforms more uniformly across the porous elec-
comes rate limiting at high current. Conversely, the classical trode, limited by electrolyte diffusion. A wide particle size
description of porous electrodes, pioneered by Newman,27,28 distribution also broadens the reaction front, as particles
focuses on transport but mostly neglects the thermodynamics of transform in order of increasing size. These examples illus-
the active materials,26,50 for example, fitting29 rather than trate the complexity of phase transformations in porous
deriving9,23,49,51 the voltage plateau in LFP. These approaches media driven by chemical reactions.
are unified by nonequilibrium chemical thermodynamics.26
Generalized porous electrode theory is constructed by formally Conclusion
volume averaging over the microstructure to obtain macro- This Account describes a journey along the “middle way”,78
scopic reaction-diffusion equations of the form eq 56 for three searching for organizing principles of the mesoscopic do-
overlapping continua, the electrolyte, conducting matrix, and main between individual atoms and bulk materials. The
active material, each containing a source/sink for Faradaic motivation to understand phase behavior in Li-ion battery
reactions, integrated over the internal surface of the active nanoparticles gradually led to a theory of collective kinetics
particles, described by the CHR or ACR model. at length and time scales in the “middle”, beyond the reach
The simplest case is the “pseudo-capacitor approxima- of both molecular simulations and macroscopic continu-
tion” of fast solid relaxation (compared with reactions and um models. The work leveraged advances in ab initio

1158 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1144–1160 ’ 2013 ’ Vol. 46, No. 5


Theory of Chemical Kinetics and Charge Transfer Bazant

quantum-mechanical calculations and nanoscale imaging my postdoctoral associates (D. A. Cogswell, G. Singh) and stu-
but also required some new theoretical ideas. dents (P. Bai, D. Burch, T. R. Ferguson, E. Khoo, R. Smith, Y. Zeng).
Besides telling the story, this Account synthesizes my P. Bai noted the Nernst factor in eq 39.
work as a general theory of chemical physics, which trans-
cends its origins in electrochemistry. The main result, eq 56, Note Added after ASAP Publication. This paper was
generalizes the CahnHilliard and AllenCahn equations posted to the WEB on March 22, 2013 with errors in several
for reaction-diffusion phenomena. The reaction rate is a equations. The revised version was reposted on April 11,
nonlinear function of the species activities and the free 2013.
energy of reaction (eq 7) via variational derivatives of the
Gibbs free energy functional (eq 51), which are consistently
BIOGRAPHICAL INFORMATION
defined for nonequilibrium states, for example, during a
Martin Z. Bazant received his B.S. (Physics, Mathematics, 1992)
phase separation. For charged species, the theory gener-
and M.S. (Applied Mathematics, 1993) from the University of
alizes the PoissonNernstPlanck equations of ion trans- Arizona and Ph.D. (Physics, 1997) from Harvard University. He
port, the ButlerVolmer equation of electrochemical joined the faculty at MIT in Mathematics in 1998 and Chemical
kinetics (eq 29), and the Marcus theory of charge transfer Engineering in 2008. His honors include an Early Career Award
(eq 37) for concentrated electrolytes and ionic solids. from the Department of Energy (2003), Brilliant Ten from Popular
As its first application, the theory has predicted new Science (2007), and Paris Sciences Chair (2002, 2007) and Joliot
Chair (2008, 2012) from ESPCI (Paris, France).
intercalation mechanisms in phase-separating battery ma-
terials, exemplified by LFP: intercalation waves in anisotro- FOOTNOTES
pic nanoparticles at low currents (Figure 8); quasi-solid *E-mail: bazant@mit.edu.
solutions and suppressed phase separation at high currents The author declares no competing financial interest.

(Figure 9); relaxation to striped phases in partially filled


REFERENCES
particles (Figure 10); size-dependent nucleation by surface
1 Padhi, A.; Nanjundaswamy, K.; Goodenough, J. Phospho-olivines as positive-electrode
wetting (Figure 11); and mosaic instabilities and reaction fronts materials for rechargeable lithium batteries. J. Electrochem. Soc. 1997, 144, 1188–1194.
in porous electrodes (Figure 12). These results have some 2 Tarascon, J.; Armand, M. Issues and challenges facing rechargeable lithium batteries.
Nature 2001, 414, 359–367.
unexpected implications, for example, that battery perfor- 3 Kang, B.; Ceder, G. Battery materials for ultrafast charging and discharging. Nature 2009,
mance may be improved with elevated currents and tempera- 458, 190–193.
4 Tang, M.; Carter, W. C.; Chiang, Y.-M. Electrochemically driven phase transitions in
tures, wider particle size distributions, and coatings to alter insertion electrodes for lithium-ion batteries: Examples in lithium metal phosphate olivines.
surface energies. The model successfully describes phase Annu. Rev. Mater. Res. 2010, 40, 501–529.
5 Meethong, N.; Huang, H.-Y. S.; Carter, W. C.; Chiang, Y.-M. Size-dependent lithium miscibility
behavior of LFP cathodes, and my group is extending it to gap in nanoscale Li1xFePO4. Electrochem. Solid-State Lett. 2007, 10, A134–A138.
graphite anodes (“staging” of Li intercalation with g3 stable 6 Burch, D.; Bazant, M. Z. Size-dependent spinodal and miscibility gaps for intercalation in
nanoparticles. Nano Lett. 2009, 9, 3795–3800.
phases) and air cathodes (electrochemical growth of Li2O2). 7 Cogswell, D. A.; Bazant, M. Z. Coherency strain and the kinetics of phase separation in
The general theory may find many other applications in LiFePO4 nanoparticles. ACS Nano 2012, 6, 2215–2225.
8 Cogswell, D. A.; Bazant, M. Z. Theory of nucleation in nanoparticles. Nano Letters, submitted.
chemistry and biology. For example, the adsorption model
9 Bai, P.; Cogswell, D.; Bazant, M. Z. Suppression of phase separation in LiFePO4
(Figure 7) could be adapted for the deposition of charged nanoparticles during battery discharge. Nano Lett. 2011, 11, 4890–4896.
colloids on transparent electrodes in electrophoretic displays. 10 Malik, R.; Zhou, F.; Ceder, G. Kinetics of non-equilibrium lithium incorporation in LiFePO4.
Nat. Mater. 2011, 10, 587–590.
The porous electrode model (Figure 12) could be adapted 11 Wagemaker, M.; Singh, D. P.; Borghols, W. J.; Lafont, U.; Haverkate, L.; Peterson, V. K.;
for sorption/desorption kinetics in nanoporous solids, for Mulder, F. M. Dynamic solubility limits in nanosized olivine LiFePO4. J. Am. Chem. Soc.
2011, 133, 10222–10228.
example, for drying cycles of cementitious materials, release 12 Singh, G.; Burch, D.; Bazant, M. Z. Intercalation dynamics in rechargeable battery materials:
General theory and phase-transformation waves in LiFePO4. Electrochim. Acta 2008, 53,
of shale gas by hydraulic fracturing, carbon sequestration in 7599–7613arXiv:0707.1858v1 [cond-mat.mtrl-sci] (2007).
zeolites, or ion adsorption and impulse propagation in bio- 13 Morgan, D.; der Ven, A. V.; Ceder, G. Li conductivity in LixMPO4 (M = Mn, Fe, Co, Ni) olivine
materials. Electrochem. Solid State Lett. 2004, 7, A30–A32.
logical cells. The common theme is the coupling of chemical
14 Laffont, L.; Delacourt, C.; Gibot, P.; Wu, M. Y.; Kooyman, P.; Masquelier, C.; Tarascon, J. M.
kinetics with nonequilibrium thermodynamics. Study of the LiFePO4/FePO4 two-phase system by high-resolution electron energy loss
spectroscopy. Chem. Mater. 2006, 18, 5520–5529.
15 Delmas, C.; Maccario, M.; Croguennec, L.; Cras, F. L.; Weill, F. Lithium deintercalation of
LiFePO4 nanoparticles via a domino-cascade model. Nat. Mater. 2008, 7, 665–671.
This work was supported by the National Science Foundation 16 Malik, R.; Burch, D.; Bazant, M.; Ceder, G. Particle size dependence of the ionic diffusivity.
Nano Lett. 2010, 10, 4123–4127.
under Contracts DMS-0842504 and DMS-0948071 and by the
17 Meethong, N.; Huang, H. Y. S.; Speakman, S. A.; Carter, W. C.; Chiang, Y. M. Strain
MIT Energy Initiative and would not have been possible without accommodation during phase transformations in olivine-based cathodes as a materials

Vol. 46, No. 5 ’ 2013 ’ 1144–1160 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1159


Theory of Chemical Kinetics and Charge Transfer Bazant

selection criterion for high-power rechargeable batteries. Adv. Funct. Mater. 2007, 17, 50 Lai, W.; Ciucci, F. Mathematical modeling of porous battery electrodes - Revisit of
1115–1123. Newman's model. Electrochim. Acta 2011, 56, 4369–4377.
18 der Ven, A. V.; Garikipati, K.; Kim, S.; Wagemaker, M. The role of coherency strains on 51 Lai, W. Electrochemical modeling of single particle intercalation battery materials with
phase stability in LixFePO4: Needle crystallites minimize coherency strain and overpotential. different thermodynamics. J. Power Sources 2011, 196, 6534–6553.
J. Electrochem. Soc. 2009, 156, A949–A957. 52 Marcus, R. A. On the theory of oxidation-reduction reactions involving electron transfer. I.
19 Stanton, L. G.; Bazant, M. Z. Phase separation with anisotropic coherency strain. arXiv J. Chem .Phys. 1956, 24, 966–978.
2012, No. arXiv:1202.1626v1. 53 Marcus, R. A. . On the theory of electron-transfer reactions. VI. unified treatment for
20 der Ven, A. V.; Wagemaker, M. Effect of surface energies and nano-particle size distribution homogeneous and electrode reactions. J. Chem .Phys. 1965, 43, 679–701.
on open circuit voltage of Li-electrodes. Electrochem. Commun. 2009, 11, 881–884. 54 Marcus, R. A. Electron transfer reactions in chemistry. Theory and experiment. Rev. Mod.
21 Wagemaker, M.; Mulder, F. M.; der Ven, A. V. The role of surface and interface energy on Phys. 1993, 65, 599–610.
phase stability of nanosized insertion compounds. Adv. Mater. 2009, 21, 2703–2709.
55 Gelfand, I. M.; Fomin, S. V. Calculus of Variations; Dover: New York, 2000.
22 Oyama, G.; Yamada, Y.; Natsui, R.; Nishimura, S.; Yamada, A. Kinetics of nucleation and
growth in two-phase electrochemical reaction of LixFePO4. J. Phys. Chem. C 2012, 116, 56 Cahn, J. W.; Hilliard, J. W. Free energy of a non-uniform system: I. Interfacial energy.
7306–7311. J. Chem. Phys. 1958, 28, 258–267.
23 Dreyer, W.; Jamnik, J.; Guhlke, C.; Huth, R.; Moskon, J.; Gaberscek, M. The thermo- 57 van der Waals, J. D. The thermodynamic theory of capillarity under the hypothesis of a
dynamic origin of hysteresis in insertion batteries. Nat. Mater. 2010, 9, 448–453. continuous variation of density. Verh. K. Akad. Wet., Afd. Natuurkd., Amsterdam (Sect. 1)
1893, 1, 8. Translation by Rowlinson, J. S. J. Stat. Phys. 1979, 20, 197–244.
24 Dreyer, D.; Guhlke, C.; Huth, R. The behavior of a many-particle electrode in a lithium-ion
battery. Phys. D 2011, 240, 1008–1019. 58 Nauman, E. B.; Heb, D. Q. Nonlinear diffusion and phase separation. Chem. Eng. Sci. 2001,
56, 1999–2018.
25 Burch, D. Intercalation Dynamics in Lithium-Ion Batteries, Ph.D. Thesis in Mathematics,
Massachusetts Institute of Technology, 2009. 59 Bazant, M. Z.; Storey, B. D.; Kornyshev, A. A. Double layer in ionic liquids: Overscreening
versus crowding. Phys. Rev. Lett. 2011, 106, No. 046102.
26 Ferguson, T. R.; Bazant, M. Z. Non-equilibrium thermodynamics of porous electrodes.
J. Electrochem. Soc. 2012, 159, A1967–A1985. 60 Bazant, M. Z.; Bazant, Z. P. Theory of sorption hysteresis in nanoporous solids: II. Molecular
27 Doyle, M.; Fuller, T. F.; Newman, J. Modeling of galvanostatic charge and discharge of the condensation. J. Mech. Phys. Solids 2012, 60, 1660–1675.
lithium/polymer/insertion cell. J. Electrochem. Soc. 1993, 140, 1526–1533. 61 Murau, P.; Singer, B. The understanding and elimination of some suspension instabilities in
28 Newman, J. Electrochemical Systems, 2nd ed.; Prentice-Hall, Inc.: Englewood Cliffs, NJ, 1991. an electrophoretic display. J. Appl. Phys. 1978, 49, 4820–4829.
29 Srinivasan, V.; Newman, J. Discharge model for the lithium iron-phosphate electrode. 62 Garcia, R. E.; Bishop, C. M.; Carter, W. C. Thermodynamically consistent variational
J. Electrochem. Soc. 2004, 151, A1517–A1529. principles with applications to electrically and magnetically active systems. Acta Mater.
2004, 52, 11–21.
30 Dargaville, S.; Farrell, T. Predicting active material utilization in LiFePO4 electrodes using a
multiscale mathematical model. J. Electrochem. Soc. 2010, 157, A830–A840. 63 Guyer, J. E.; Boettinger, W. J.; Warren, J. A.; McFadden, G. B. Phase field modeling of
electrochemistry I: Equilibrium. Phys. Rev. E 2004, 69, No. 021603.
31 Chen, G.; Song, X.; Richardson, T. Electron microscopy study of the LiFePO4 to FePO4
phase transition. Electrochem. Solid State Lett. 2006, 9, A295–A298. 64 Guyer, J. E.; Boettinger, W. J.; Warren, J. A.; McFadden, G. B. Phase field modeling of
electrochemistry II: Kinetics. Phys. Rev. E 2004, 69, No. 021604.
32 Ramana, C. V.; Mauger, A.; Gendron, F.; Julien, C. M.; Zaghib, K. Study of the Li-insertion/
extraction process in LiFePO4/FePO4. J. Power Sources 2009, 187, 555–564. 65 Tang, M.; Belak, J. F.; Dorr, M. R. Anisotropic phase boundary morphology in nanoscale
33 Bazant, M. Z. 10.626 Electrochemical Energy Systems; Massachusetts Institute of Technology, olivine electrode particles. J. Phys. Chem. C 2011, 115, 4922–4926.
MIT OpenCourseWare, http://ocw.mit.edu; license, Creative Commons BY-NC-SA, 2011. 66 Liang, L.; Qi, Y.; Xue, F.; Bhattacharya, S.; Harris, S. J.; Chen, L.-Q. Nonlinear
34 Burch, D.; Singh, G.; Ceder, G.; Bazant, M. Z. Phase-transformation wave dynamics phase-field model for electrode-electrolyte interface evolution. Phys. Rev. E 2012, 86, No.
LiFePO4. Solid State Phenom. 2008, 139, 95–100. 051609.
35 Bazant, M. Z.; Kilic, M. S.; Storey, B.; Ajdari, A. Towards an understanding of nonlinear 67 Samin, S.; Tsori, Y. Vapor-liquid equilibrium in electric field gradients. J. Phys. Chem. B
electrokinetics at large voltages in concentrated solutions. Adv. Colloid Interface Sci. 2009, 2011, 115, 75–83.
152, 48–88. 68 Storey, B. D.; Bazant, M. Z. Effects of electrostatic correlations on electrokinetic
36 Kuznetsov, A. M.; Ulstrup, J. Electron Transfer in Chemistry and Biology: An Introduction to phenomena. Phys. Rev. E 2012, 86, No. 056303.
the Theory; Wiley: Chichester, U.K., 1999. 69 Santangelo, C. D. Computing counterion densities at intermediate coupling. Phys. Rev. E
37 Sekimoto, K. Stochastic Energetics; Springer: Berlin, 2010. 2006, 73, No. 041512.
38 Groot, S. R. D.; Mazur, P. Non-equilibrium Thermodynamics; Interscience Publishers, Inc.: 70 Hatlo, M. M.; Lue, L. Electrostatic interactions of charged bodies from the weak to the strong
New York, 1962. coupling regime. Europhys. Lett. 2010, 89, No. 25002.
39 Balluffi, R. W.; Allen, S. M.; Carter, W. C. Kinetics of Materials; Wiley: Hoboken, NJ, 2005. 71 Gregersen, M. M.; Okkels, F.; Bazant, M. Z.; Bruus, H. Topology and shape optimization of
40 Prigogine, I.; Defay, R. Chemical Thermodynamics; John Wiley and Sons: New York, 1954. induced-charge electro-osmotic micropumps. New J. Phys. 2009, 11, 075016.
41 Kilic, M. S.; Bazant, M. Z.; Ajdari, A. Steric effects on the dynamics of electrolytes at large 72 Biesheuvel, P. M.; van Soestbergen, M.; Bazant, M. Z. Imposed currents in galvanic cells.
applied voltages: II Modified Nernst-Planck equations. Phys. Rev. E 2007, 75, No. 021503. Electrochim. Acta 2009, 54, 4857–4871.
42 Han, B.; der Ven, A. V.; Morgan, D.; Ceder, G. Electrochemical modeling of intercalation 73 Biesheuvel, P.; Fu, Y.; Bazant, M. Diffuse charge and Faradaic reactions in porous
processes with phase field models. Electrochim. Acta 2004, 49, 4691–4699. electrodes. Phys. Rev. E 2011, 83, No. 061507.
43 Singh, G. K.; Bazant, M. Z.; Ceder, G. Anisotropic surface reaction limited phase 74 Aziz, M. J.; Sabin, P. C.; Lu, G. Q. The activation strain tensor: Nonhydrostatic stress effects
transformation dynamics in LiFePO4. ArXiv 2007, No. arXiv:0707.1858v1. on crystal growth kinetics. Phys. Rev. B 1991, 41, 9812–9816.
44 Beard, D. A.; Qian, H. Relationship between thermodynamic driving force and one-way 75 Raistrick, I. D.; Mark, A. J.; Huggins, R. A. Thermodynamics and kinetics of the
fluxes in reversible processes. PLoS ONE 2007, 2, No. e144. electrochemical insertion of lithium into tungsten bronzes. Solid State Ionics 1981, 5,
45 Bard, A. J.; Faulkner, L. R. Electrochemical Methods; J. Wiley & Sons, Inc.: New York, 2001. 351–354.
46 Kulikovsky, A. A. Analytical Modelling of Fuel Cells; Elsevier: New York, 2010. 76 Smith, K. C.; Mukherjee, P. P.; Fisher, T. S. Columnar order in jammed LiFePO4 cathodes:
47 Eikerling, M.; Kornyshev, A. A. Modelling the performance of the cathode catalyst layer of Ion transport catastrophe and its mitigation. Phys. Chem. Chem. Phys. 2012, 14,
polymer electrolyte fuel cells. J. Electroanal. Chem. 1998, 453, 89–106. 7040–7050.
48 Ioselevich, A. S.; Kornyshev, A. A. Phenomenological theory of solid oxide fuel cell anode. 77 Bai, P.; Tian, G. Statistical kinetics of phase-transforming nanoparticles in LiFePO4 porous
Fuel Cells 2001, 1, 40–65. electrodes. Electrochim. Acta 2013, 89, 644–651.
49 Lai, W.; Ciucci, F. Thermodynamics and kinetics of phase transformation in intercalation 78 Laughlin, R. B.; Pine, D.; Schmalian, J.; Stojkovic, B. P.; Wolynes, P. The middle way.
battery electrodes - phenomenological modeling. Electrochim. Acta 2010, 56, 531–542. Proc. Natl. Acad. Sci. U.S.A. 2000, 97, 32–37.

1160 ’ ACCOUNTS OF CHEMICAL RESEARCH ’ 1144–1160 ’ 2013 ’ Vol. 46, No. 5

Das könnte Ihnen auch gefallen