18
Theory of Chemical Kinetics and Charge Transfer based on Nonequilibrium Thermodynamics The MIT Faculty has made this article openly available. Please share how this access benefits you. Your story matters. Citation Bazant, Martin Z. "Theory of Chemical Kinetics and Charge Transfer based on Nonequilibrium Thermodynamics." Acc. Chem. Res., 2013, 46 (5), pp 1144–1160. As Published http://dx.doi.org/10.1021/ar300145c Publisher American Chemical Society (ACS) Version Author's final manuscript Citable link http://hdl.handle.net/1721.1/90946 Terms of Use Article is made available in accordance with the publisher's policy and may be subject to US copyright law. Please refer to the publisher's site for terms of use.

Theor y of Chemical Kinetics and Charge Tr ansfer based on

  • Upload
    others

  • View
    8

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Theor y of Chemical Kinetics and Charge Tr ansfer based on

Theory of Chemical Kinetics and Charge Transferbased on Nonequilibrium Thermodynamics

The MIT Faculty has made this article openly available. Please share how this access benefits you. Your story matters.

Citation Bazant, Martin Z. "Theory of Chemical Kinetics and Charge Transferbased on Nonequilibrium Thermodynamics." Acc. Chem. Res., 2013,46 (5), pp 1144–1160.

As Published http://dx.doi.org/10.1021/ar300145c

Publisher American Chemical Society (ACS)

Version Author's final manuscript

Citable link http://hdl.handle.net/1721.1/90946

Terms of Use Article is made available in accordance with the publisher'spolicy and may be subject to US copyright law. Please refer to thepublisher's site for terms of use.

Page 2: Theor y of Chemical Kinetics and Charge Tr ansfer based on

Theory of Chemical Kinetics and Charge Transferbased on Nonequilibrium Thermodynamics

Martin Z. Bazant∗

Departments of Chemical Engineering and Mathematics,Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA

CONSPECTUSAdvances in the fields of catalysis and electrochemical energy conversion often involve nanopar-

ticles, which can have kinetics surprisingly different from the bulk material. Classical theories ofchemical kinetics assume independent reactions in dilute solutions, whose rates are determined bymean concentrations. In condensed matter, strong interactions alter chemical activities and createvariations that can dramatically affect the reaction rate. The extreme case is that of a reactioncoupled to a phase transformation, whose kinetics must depend not only on the order parameter,but also its gradients at phase boundaries. Reaction-driven phase transformations are common inelectrochemistry, when charge transfer is accompanied by ion intercalation or deposition in a solidphase. Examples abound in Li-ion, metal-air, and lead-acid batteries, as well as metal electrodepo-sition/dissolution. In spite of complex thermodynamics, however, the standard kinetic model is theButler-Volmer equation, based on a dilute solution approximation. The Marcus theory of chargetransfer likewise considers isolated reactants and neglects elastic stress, configurational entropy, andother non-idealities 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, whichscientists believe limits its performance. Chemists first modeled phase separation in LFP as anisotropic “shrinking core” within each particle, but experiments later revealed striped phase bound-aries on the active crystal facet. This raised the question: What is the reaction rate at a surfaceundergoing a phase transformation? Meanwhile, dramatic rate enhancement was attained withLFP nanoparticles, and classical battery models could not predict the roles of phase separation andsurface modication.

In this Account, I present a general theory of chemical kinetics, developed over the past sevenyears, which is capable of answering these questions. The reaction rate is a nonlinear function ofthe thermodynamic driving force – the free energy of reaction – expressed in terms of variationalchemical potentials. The theory unifies and extends the Cahn-Hilliard and Allen-Cahn equationsthrough a master equation for non-equilibrium chemical thermodynamics. For electrochemistry, Ihave also generalized both Marcus and Butler-Volmer kinetics for concentrated solutions and ionicsolids.

This new theory provides a quantitative description of LFP phase behavior. Concentration gradi-ents and elastic coherency strain enhance the intercalation rate. At low currents, the charge-transferrate is focused on exposed phase boundaries, which propagate as “intercalation waves”, nucleatedby surface wetting. Unexpectedly, homogeneous reactions are favored above a critical current andbelow a critical size, which helps to explain the rate capability of LFP nanoparticles. Contraryto other mechanisms, elevated temperatures and currents may enhance battery performance andlifetime by suppressing phase separation. The theory has also been extended to porous electrodesand could be used for battery engineering with multiphase active materials.

More broadly, the theory describes non-equilibrium chemical systems at mesoscopic length andtime scales, beyond the reach of molecular simulations and bulk continuum models. The reactionrate is consistently defined for inhomogeneous, non-equilibrium states; for example, with phaseseparation, large electric fields, or mechanical stresses. This research is also potentially applicableto fluid extraction from nanoporous solids, pattern formation in electrophoretic deposition, andelectrochemical dynamics in biological cells.

I. INTRODUCTION

Breakthroughs in catalysis and electrochemical energyconversion often involve nanoparticles, whose kineticscan differ unexpectedly from the bulk material. Per-haps the most remarkable case is lithium iron phosphate,

∗Electronic address: [email protected]

LixFePO4 (LFP). In the seminal study of micron-sizedLFP particles, Padhi et al. [5] concluded that “the ma-terial is very good for low-power applications” but “athigher current densities there is a reversible decrease incapacity that... is associated with the movement of atwo-phase interface” between LiFePO4 and FePO4. Iron-ically, over the next decade – in nanoparticle form – LFPbecame the most popular high-power cathode materialfor Li-ion batteries [6–8]. Explaining this reversal of for-tune turned out to be a major scientific challenge, with

arX

iv:1

208.

1587

v3 [

cond

-mat

.mtr

l-sc

i] 2

4 Fe

b 20

13

Page 3: Theor y of Chemical Kinetics and Charge Tr ansfer based on

2

(a) (b) (c)

FIG. 1: Motivation to generalize charge-transfer theory. Observations by (a) Chen et al. [1] and (b) Ramana et al.[2] of separated FePO4 and LiFePO4 phases on the active 010 facet, which suggest (c) focusing of lithiumintercalation reactions on the phase boundary, so it propagates as an “intercalation wave” [3] (or “domino

cascade” [4]). [From Refs. [1–3]]

important technological implications.

It is now understood that phase separation is stronglysuppressed in LFP nanoparticles, to some extent in equi-librium [9–12], but especially under applied current [11,13–15], since reaction limitation [3], anisotropic lithiumtransport [4, 16–18], elastic coherency strain [11, 19–21],and interfacial energies [12, 13, 22, 23] are all enhanced.At low currents, anisotropic nucleation and growth canalso occur [3, 11–13, 24], as well as multi-particle mo-saic instabilities [25–28]. These complex phenomena can-not be described by traditional battery models [29, 30],which assume a spherical “shrinking core” phase bound-ary [31, 32].

This Account summarizes my struggle to develop aphase-field theory of electrochemical kinetics [3, 10–13, 21, 28, 33–35] by combining charge-transfer the-ory [36] with concepts from statistical physics [37] andnon-equilibrium thermodynamics [38–40]. It all beganin 2006 when my postdoc, Gogi Singh, found the paperof Chen et al. [1] revealing striped phase boundaries inLFP, looking nothing like a shrinking core and suggestingphase boundary motion perpendicular to the lithium flux(Fig. 1). It occurred to me that, at such a surface, inter-calation reactions must be favored on the phase boundaryin order to preserve the stable phases, but this could notbe described by classical kinetics proportional to concen-trations. Somehow the reaction rate had to be sensitiveto concentration gradients.

As luck would have it, I was working on models ofcharge relaxation in concentrated electrolytes using non-equilibrium thermodynamics [35, 41], and this seemedlike a natural starting point. Gerbrand Ceder suggestedadapting the Cahn-Hilliard (CH) model for LFP [42], butit took several years to achieve a consistent theory. Ourinitial manuscript [43] was rejected in 2007, just afterGogi left MIT and I went on sabbatical leave to ESPCI,faced with rewriting the paper [3].

The rejection was a blessing in disguise, since it mademe think harder about the foundations of chemical kinet-ics. The paper contained some new ideas – phase-fieldchemical kinetics and intercalation waves – that, the re-

viewers felt, contradicted the laws of electrochemistry.It turns out the basic concepts were correct, but KenSekimoto and David Lacoste at ESPCI helped me real-ize that my initial Cahn-Hilliard reaction (CHR) modeldid not uphold the De Donder relation [37]. In 2008 inParis, I completed the theory, prepared lecture notes [33],published generalized Butler-Volmer kinetics [35] (Sec.5.4.2) and formulated non-equilibrium thermodynamicsfor porous electrodes [28]. (See also Sekimoto [37].)

Phase-field kinetics represents a paradigm shift inchemical physics, which my group has successfully ap-plied to Li-ion batteries. Damian Burch [10] used theCHR model to study intercalation in nanoparticles, andhis thesis [27] included early simulations of “mosaic insta-bility” in collections of bistable particles [25, 26]. Simula-tions of galvanostatic discharge by Peng Bai and DanielCogswell led to the unexpected prediction of a criticalcurrent for the suppression of phase separation [13]. LiamStanton modeled anisotropic coherency strain [21], whichDan added to our LFP model [11], along with surfacewetting [12]. Using material properties from ab initiocalculations, Dan predicted phase behavior in LFP [11]and the critical voltage for nucleation [12] in excellentagreement with experiments. Meanwhile, Todd Fergu-son [28] did the first simulations of phase separation inporous electrodes, paving the way for engineering appli-cations.

What follows is a general synthesis of the theory and asummary its key predictions. A thermodynamic frame-work is developed for chemical kinetics, whose applica-tion to charge transfer generalizes the classical Butler-Volmer and Marcus equations. The theory is then unifiedwith phase-field models and applied to Li-ion batteries.

Page 4: Theor y of Chemical Kinetics and Charge Tr ansfer based on

3

II. REACTIONS IN CONCENTRATEDSOLUTIONS

A. Generalized Kinetics

The theory is based on chemical thermodynamics [40].In an open system, the chemical potential of species i(per particle),

µi = kBT ln ai + µΘi = kBT ln ci + µexi (1)

is defined relative to a standard state (Θ) of unit activity(ai = 1) and concentration ci = cΘi , where ci = ci/c

Θi is

the dimensionless concentration. The activity coefficient,

γi = e(µexi −µΘi )/kBT (2)

is a measure of non-ideality (ai = γici). In a dilute so-lution, µexi = 0 and γi = 1. For the general chemicalreaction,

S1 =∑r

srAr →∑p

spBp = S2, (3)

the equilibrium constant is

KΘ =

(a2

a1

)eq= e(µΘ

1 −µΘ2 )/kBT (4)

where a1 =∏r a

srr , a2 =

∏p a

spp , µΘ

1 =∑i srµ

Θr and

µΘ2 =

∑p spµ

Θp .

The theory assumes that departures from equilibriumobey linear irreversible thermodynamics (LIT) [38, 39].The flux of species i is proportional to the thermody-namic driving force −∇µi:

Fi = −Mici∇µi = −Di

(∇ci + ci∇

µexikBT

)= −Dchem

i ∇ci(5)

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

diffusivity, and Dchemi = Di

(1 + ∂ ln γi

∂ ln ci

)is the chemical

diffusivity [30]. In Eq. 5, the first term represents ran-dom fluctuations and the second, drift in response to thethermodynamic bias, −∇µexi .

In a consistent formulation of reaction kinetics [33, 37],therefore, the reaction complex explores a landscape ofexcess chemical potential µex(x) between local minimaµex1 and µex2 with transitions over an activation barrierµex (Fig. 2(a)). For rare transitions (µex‡ −µex1,2 kBT ),

the reaction rate (per site) is

R = k→c1e−(µex‡ −µ

ex1 )/kBT − k←c2e−(µex‡ −µ

ex2 )/kBT (6)

Enforcing detailed balance (R = 0) in equilibrium(µ1 = µ2) yields the reaction rate consistent with non-equilibrium thermodynamics:

R = k0

(e−(µex‡ −µ1)/kBT − e−(µex‡ −µ2)/kBT

)(7)

FIG. 2: (a) Landscape of excess chemical potentialexplored by the reaction S1 → S2. (b) Adsorption from

a liquid, where the transition state (TS) excludesmultiple surface sites (s > 1) while shedding the

first-neighbor shell. (c) Solid diffusion on a lattice,where the transition state excludes two sites.

where k0 = k→ = k← (for properly defined µ). Eq. 7upholds the De Donder relation [37],

R→R←

=KΘa1

a2= e(µ1−µ2)/kBT (8)

which describes the steady state of chemical reactions inopen systems [44].

The thermodynamic driving force is

∆µ = µ2 − µ1 = kBT lna2

KΘa1= ∆G (9)

also denoted as ∆G, the free energy of reaction. The re-action rate Eq. 7 can be expressed as a nonlinear functionof ∆µ:

R = R0

(e−α∆µ/kBT − e(1−α)∆µ/kBT

)(10)

where α, the symmetry factor or generalized Brønstedcoefficient [36], is approximately constant with 0 < α < 1for many reactions. Defining the activity coefficient ofthe transition state γ‡ by

µex‡ = kBT ln γ‡ + (1− α)µΘ1 + αµΘ

2 , (11)

the exchange rate R0 takes the form,

R0 =k0a

1−α1 aα2γ‡

= k0c1−α1 cα2

(γ1−α

1 γα2γ‡

)(12)

where the term in parentheses is the thermodynamic cor-rection for a concentrated solution.

Page 5: Theor y of Chemical Kinetics and Charge Tr ansfer based on

4

B. Example: Surface Adsorption

Let us apply the formalism to Langmuir adsorptionfrom a liquid mixture with µ1 = kT ln a (Fig. 2(b)). Thesurface is an ideal solution of adatoms and vacancies,

µ2 = kBT lnc

1− c+ Ea (13)

with coverage c = c/cs, site density cs, and adsorptionenergy Ea = µΘ

2 − µΘ1 . Equilibrium yields the Langmuir

isotherm,

ceq =KΘa

1 +KΘa, KΘ = e−Ea/kBT (14)

If the transition state excludes s surface sites,

µex‡ = −skBT ln(1− c) + E‡ (15)

then Eq. 7 yields,

R = k1(1− c)s[KΘa(1− c)− c

](16)

where k1 = k0e(Ea−E‡)/kBT . With only configura-

tional entropy, we recover standard kinetics of adsorp-tion, Asol+sV → Asurf +(s−1)V , involving s vacancies.With attractive forces, however, Eq. 7 predicts novel ki-netics for inhomogeneous surfaces undergoing condensa-tion (below).

C. Example: Solid diffusion

We can also derive the LIT flux Eq. 5 for macroscopictransport in a solid by activated hopping between adja-cent minima of µex having slowly varying chemical poten-tial, |∆µ| kBT and concentration ∆c 1. Linearizingthe hopping rate,

R ∼ −R0∆µ

kBT, R0 ∼

k0cγ

γ‡(17)

over a distance ∆x through an area ∆y∆z with ∂µ∂x ∼

∆µ∆x ,

we obtain Eq. 5 with

D

D0=

γ

γ‡(18)

where D0 = k0∆xcΘ∆y∆z . Eq. 18 can be used to derive the

tracer diffusivity in a concentrated solid solution by esti-mating γ‡, consistent with γ. For example, for diffusionon a lattice (Fig. 2(c)) with γ = (1 − c)−1, the transi-tion state excludes two sites, γ‡ = (1 − c)−2; the tracerdiffusivity, D = D0(1− c), scales with the mean numberof empty neighboring sites, but the chemical diffusivityis constant, Dchem = D0 = D(0) (particle/hole duality).

x

µex

O+ ne−

RΔµ = neη

E = −∇φ eq

e−x1x2 x†

electrode electrolyte

FIG. 3: Landscape of excess chemical potentialexplored by the Faradaic reaction O + ne− → R, in

Nernst equilibrium (blue) and after a negativeoverpotential η = (µ2 − µ1)/ne is applied (red) to favor

reduction, as illustrated below.

III. ELECTROCHEMISTRY INCONCENTRATED SOLUTIONS

A. Electrochemical Thermodynamics

Next we apply Eq. 7 to the general Faradaic reaction,

S1 =∑i

si,OOzi,Oi + ne− →

∑j

sj,RRzj,Rj = S2 (19)

converting the oxidized state OzO =∑i si,OO

zi,Oi to the

reduced state RzR =∑j sj,RR

zj,Rj while consuming n

electrons. Let µ1 = µO + nµe =∑i si,Oµi,O + nµe and

µ2 = µR =∑j sj,rµj,r. Charge conservation implies zO−

n = zR where zO =∑i si,Ozi,O and zR =

∑j sj,Rzj,R.

The electrostatic energy zieφi is added to µexi to definethe electrochemical potential,

µi = kBT ln ai + µΘi + zieφi = kBT ln ci + µexi (20)

where zie is the charge and φi is the Coulomb potentialof mean force.

The electrostatic potential is φe in the electrode and φin the electrolyte. The difference is the interfacial volt-age, ∆φ = φe−φ. The mean electric field −∇φ at a pointis unique, so φi = φe for ions in the electrode and φi = φfor those in the electrolyte solution. In the most generalcase of a mixed ion-electron conductor, the reduced andoxidized states are split across the interface (Fig. 4(a)).Charge conservation implies zOe + zOs − n = zRe + zRs,and the net charge nce transferred from the solution tothe electrode is given by nc = zOs− zRs = zRe− zOe+n.

Let us assume that ions only exist in the electrolyte(zRe = zOe = 0, nc = n) since the extension to mixedion-electron conductors is straightforward. For redox re-

Page 6: Theor y of Chemical Kinetics and Charge Tr ansfer based on

5

(a)$ OzOs(s )

RzRs(s)

ne−

φφe electrode$$|$$solu/on$

OzOe(e)

RzRe(e)

(b)$ OzO

RzR

(c)$ OzO

RzR

ne−

ne−

FIG. 4: Types of Faradaic reactions O + ne− → R. (a)General mixed ion-electron conductor

electrode/electrolyte interface. (b) Redox in solution.(c) Ion intercalation or electrodeposition.

actions (Fig. 4(b)), e.g. Fe3+ + e− → Fe2+, the re-duced state is in the solution at the same potential,φR = φO = φ. For electrodeposition (Fig. 4(c)), e.g.Cu2+ + 2e− → Cu, or ion intercalation as a neutral po-laron, e.g. CoO2+Li+ +e− → LiCoO2, the reduced stateis uncharged, zR = 0, so we can also set φR = φ, eventhough it is in the electrode. For this broad class ofFaradaic reactions, we have

µO = kBT ln aO + µΘO + zOeφ (21)

µR = kBT ln aR + µΘR + zReφ (22)

µe = kBT ln ae + µΘe − eφe (23)

(aO =∏i asji , µΘ

O =∑i siµi,...) where µe is the Fermi

level, which depends on φe and the electron activity ae =γece.

In equilibrium (µ1 = µ2), the interfacial voltage isgiven by the Nernst equation

∆φeq = EΘ +kBT

ncelnaOa

ne

aR(24)

where nc = n and

EΘ =µΘO + nµΘ

e − µΘR

ne(25)

is the standard half-cell potential. Out of equilibrium,the current I = neR (per active site) is controlled by theactivation over-potential,

η = ∆φ−∆φeq =∆µ

ne=

∆G

ne(26)

Specific models of charge transfer correspond to differentchoices of µex‡ .

B. Generalized Butler-Volmer Kinetics

The standard phenomenological model of electrode ki-netics is the Butler-Volmer equation[30, 45],

I = I0

(e−αcneη/kBT − eαaneη/kBT

)(27)

where I0 is the exchange current I0. For a single-stepcharge-transfer reaction, the anodic and cathodic charge-transfer coefficients αa and αc satisfy αa = 1−α and αc =α with a symmetry factor, 0 < α < 1. The exchangecurrent is typically modeled as I0 ∝ cαaO cαcR , but this is adilute solution approximation.

In concentrated solutions, the exchange current is af-fected by configurational entropy and enthalpy, elec-trostatic correlations, coherency strain, and other non-idealities. For Li-ion batteries, only excluded volumehas been considered, using[29, 30], I0(c) ∝ (cs− c)αccαa .For fuel cells, many phenomenological models have beendeveloped for electrocatalytic reactions with surface ad-sorption steps [46–48]. Electrocatalysis can also betreated by our formalism [33], but here we focus on theelementary charge-transfer step and its coupling to phasetransformations, which has no prior literature.

In order to generalize BV kinetics (Fig. 3), we modelthe transition state

µex‡ = kBT ln γ‡+(1−α)(zOeφ−neφe+µΘO+nµΘ

e )+α(zReφ+µΘR)

(28)by averaging the standard chemical potential and electro-static energy of the initial and final states, which assumesa constant electric field across the reaction coordinate xwith α =

x‡−xRxO−xR . Substituting Eq. 28 into Eq. 7 using

Eq. 24, we obtain Eq. 27 with

I0 =k0ne(aOa

ne )1−αaαR

γ‡= k0ne(cOc

ne )1−αcαR

[(γOγ

ne )1−αγαRγ‡

](29)

The factor in brackets is the thermodynamic correctionfor the exchange current.

Generalized BV kinetics (Eq. 27 and Eq. 29) consis-tently applies chemical kinetics in concentrated solutions(Eq. 10 and Eq. 12, respectively) to Faradaic reactions.In Li-ion battery models, ∆φeq(c) is fitted to the opencircuit voltage, and I0(c) and Dchem(c) are fitted to dis-charge curves [29, 31, 32], but these quantities are relatedby non-equilibrium thermodynamics [13, 28, 35]. Lai and

Page 7: Theor y of Chemical Kinetics and Charge Tr ansfer based on

6

Oxidized State

Transition State

Reduced State

Solid Host Liquid Electrolyte

e−

e−

xOxR x†

µexO + nµe

µexR

µex†

O+ ne−

R

x

µex

ΔGex

FIG. 5: Above: The Faradaic reaction O + ne− → R inconcentrated solutions. Each state explores a landscapeof excess chemical potential µex. Charge transfer occurs

where the curves overlap, or just below, by quantumtunneling (dashed curves). Below: Example of ionintercalation into a solid electrode from a liquid

electrolyte.

Ciucci [49–51] also recognized this inconsistency and usedEq. 5 and Eq. 24 in battery models, but they postulateda barrier of total (not excess) chemical potential, in con-trast to Eq. 7, Eq. 29 and charge-transfer theory.

C. Generalized Marcus Kinetics

The microscopic theory of charge transfer, initiated byMarcus [52, 53] and honored by the Nobel Prize in Chem-istry [54], provides justification for the BV equation and

a means to estimate its parameters based on solvent re-organization [45]. Quantum mechanical formulations pi-oneered by Levich, Dogonadze, Marcus, Kuznetsov, andUlstrup further account for Fermi statistics, band struc-ture, and electron tunneling [36]. Most theories, however,make the dilute solution approximation by considering anisolated reaction complex.

In order to extend Marcus theory for concentrated so-lutions, our basic postulate (Fig. 5) is that the Faradaicreaction Eq. 19 occurs when the excess chemical poten-tial of the reduced state, deformed along the reaction co-ordinate by statistical fluctuations, equals that of the oxi-dized state (plus n electrons in the electrode) at the samepoint. (More precisely, charge transfer occurs at slightlylower energies due to quantum tunneling [36, 45].) Fol-lowing Marcus, we assume harmonic restoring forces forstructural relaxation (e.g. shedding of the solvation shellfrom a liquid, or ion extraction from a solid) along thereaction coordinate x from the oxidized state at xO tothe reduced state at xR:

µex1 (x) = µΘO+nµΘ

e +kBT ln(γOγne )+zOeφ−neφe+

kO2

(x−xO)2

(30)

µex2 (x) = µΘR + kBT ln γR + zReφ+

kR2

(x− xR)2 (31)

The Nernst equation Eq. 24 follows by equating the to-tal chemical potentials at the local minima, µ1(xO) =µ2(xR) in equilibrium. The free energy barrier is set bythe intersection of the excess chemical potential curves,µex‡ = µex1 (x‡) = µex2 (x‡), which determines the barrierposition, x = x‡ and implies

∆Gex = µex2 (xR)−µex1 (xO) =kO2

(x‡−xO)2−kR2

(x‡−xR)2

(32)where ∆Gex is the excess free energy change per reaction.

From Eq. 26, the overpotential is the total free energychange per charge transferred,

neη = ∆G = ∆Gex + kBT lncRcO cne

(33)

In classical Marcus theory [45, 54], the overpotential isdefined by neη = ∆Gex without the concentration fac-tors required by non-equilibrium thermodynamics, whichis valid for charge-transfer reactions in bulk phases (A−+B → A+B−) because the initial and final concentrationsare the same, and thus ∆G = ∆Gex = ∆G0 (standardfree energy of reaction). For Faradaic reactions at inter-faces, however, the concentrations of reactions and prod-ucts are different, and Eq. 33 must be used. The missing“Nernst concentration term” in Eq. 33 has also beennoted by Kuznetsov and Ulstrup [36] (p. 219).

In order to relate µex‡ to ∆Gex, we solve Eq. 32 forx‡. In the simplest approximation, kO = kR = k, thebarriers for the cathodic and anodic reactions,

∆Gexc = µex‡ − µex1 (xO) =λ

4

(1 +

∆Gex

λ

)2

(34)

Page 8: Theor y of Chemical Kinetics and Charge Tr ansfer based on

7

∆Gexa = µex‡ − µex2 (xR) =λ

4

(1− ∆Gex

λ

)2

(35)

are related to the reorganization energy, λ = k2 (xO −

xR)2. These formulae contain the famous “inverted re-gion” predicted by Marcus for isotopic exchange [54],where (say) the cathodic rate, kc ∝ e−∆Gexc /kBT reachesa minimum and increases again with decreasing drivingforce ∆Gex, for x‡ < xR in Fig. 5(a). This effect re-mains for charge transfer in concentrated bulk solutions,e.g. A− + B→ A + B−. For Fardaic reactions, however,it is suppressed at metal electrodes, since electrons cantunnel through unoccupied conduction-band states, butcan arise in narrow-band semiconductors [36, 53, 54].

Substituting µex‡ into Eq. 7, we obtain

R = k0e−λ/4kBT e−(∆Gex)2/4kBTλ

×(coc

ne e−∆Gex/2kBT − cRe∆Gex/2kBT

)(36)

Using Eq. 33, we can relate the current to the overpo-tential,

I = I0 e−(neη)2/4kBTλ

(e−αneη/kBT − e(1−α)neη/kBT

)(37)

via the exchange current,

I0 = nek0e−λ/4kBT (cO c

ne )

3−2α4 c

1+2α4

R , (38)

and symmetry factor,

α =1

2

(1 +

kBT

λlncO c

ne

cR

). (39)

In the typical case λ kBT , the current Eq. 37 is wellapproximated by the BV equation with α = 1

2 at mod-

erate overpotentials, |η| > kBTne

√λ

kBTand non-depleted

concentrations, | ln c| λkBT

.

Comparing Eq. 38 with Eq. 29 for α ≈ 12 , we can re-

late the reorganization energy to the activity coefficientsdefined above,

λ ≈ 4kBT lnγ‡

(γOγne γR)1/2(40)

For a dilute solution, the reorganization energy λ0 canbe estimated by the classical Marcus approximation,λ0 = λi + λo, where λi is the “inner” or short-rangecontribution from structural relaxation (sum over nor-mal modes) and λo is the “outer”, long-range contri-bution from the Born energy of solvent dielectric relax-ation [45, 54]. For polar solvents at room temperature,the large Born energy, λo > 0.5n2eV ≈ 20n2kBT (atroom temperature), implies that single-electron (n = 1),symmetric (α ≈ 1

2 ) charge transfer is favored. Quantummechanical approximations of λ0 are also available [36].For a concentrated solution, we can estimate the thermo-dynamic correction, γc‡ , for the entropy and enthalpy ofthe transition state and write

γ‡ = γc‡eλ0/4kBT . (41)

which can be used in either Marcus (Eqs. 37-40) or BV(Eqs. 27-29) kinetics. An example for ion intercalationis given below, Eq. 80, but first we need to develop amodeling framework for chemical potentials.

IV. NONEQUILIBRIUM CHEMICALTHERMODYNAMICS

A. General theory

In homogeneous bulk phases, activity coefficients de-pend on concentrations, but for reactions at an interface,concentration gradients must also play a role (Fig. 1).The main contribution of this work has been to formulatechemical kinetics for inhomogeneous, non-equilibriumsystems. The most general theory appears here for thefirst time, building on my lectures notes [33].

The theory is based the Gibbs free energy functional

G[ci] =

∫V

g dV +

∮A

γs dA = Gbulk +Gsurf (42)

with integrals over the bulk volume V and surface areaA. The variational derivative [55],

δG

δci(x) = lim

ε→0

G[ci(x) + εδε(x)]−G[ci(x)]

ε(43)

is the change in G to add a “continuum particle” δ(x)(delta function) of species i at point x, where δε(x) →δ(x) is a finite-size approximation, e.g. δε(x) = e−x

2/2ε√

2πε.

This is the consistent definition of diffusional chemicalpotential [39, 56],

µi =δG

δci(44)

If g depends on ci and ∇ci, then

µi =∂g

∂ci−∇ · ∂g

∂∇ci(45)

The continuity of µi at the surface yields the “naturalboundary condition”,

n · ∂g

∂∇ci=∂γs∂ci

. (46)

We can also express the activity variationally,

ai = exp

(1

kBT

δGmixδci

)(47)

in terms of the free energy of mixing

Gmix = Gbulk −∑i

µΘi

∫V

ci dV (48)

Page 9: Theor y of Chemical Kinetics and Charge Tr ansfer based on

8

locally conserved order parameter

non-conserved order parameter

R

R

R

R

R

R

R

R R R R

R R

Heterogeneous Chemistry (Cahn-Hilliard reaction)

(a)

Homogeneous Chemistry

(b)

Phase Transformation (Allen-Cahn reaction)

(c)

FIG. 6: Types of reactions (R) in non-equilibriumchemical thermodynamics. (a) Heterogeneous chemistryat a surface Eq. 52. (b) Homogeneous chemistry Eq. 54

with diffusing species. (c) Phase transformations, orhomogeneous reactions with immobile species Eq. 56.

which we define relative to the standard states of eachspecies.

The simplest approximation for an inhomogeneous sys-tem is the Cahn-Hilliard [56] (or Landau-Ginzburg, orVan der Waals [57]) gradient expansion,

g = g(ci) +∑i

µΘi ci +

1

2

∑j

∇ci · κij∇cj

(49)

for which

µi − µΘi = kBT ln ai =

∂g

∂ci+∑j

∇ · κij∇cjcΘj

(50)

where g is the homogeneous free energy of mixing and κis the gradient penalty tensor. (Higher-order derivativeterms can also be added [58, 59].)

With these definitions, Eq. 7 takes the variationalform,

R = k0eµex‡kBT

[exp

(∑r

srkBT

δG

δcr

)− exp

(∑p

spkBT

δG

δcp

)](51)

for the general reaction, Eq. 3, in a concentrated solu-tion.

B. Phase-Field Chemical Kinetics

The rate expression Eq. 51 can be applied to any typeof reaction (Fig. 6):

1. Heterogeneous chemistry

At an interface, Eq. 51 provides a new reaction bound-ary condition [3, 10, 13, 35]

siAr n ·(~u ci −

DicikBT

∇δGδci

)= ±R

(δG

δcj

)(52)

(+ for reactants, − for products; Ar = reaction site area)for the Cahn-Hilliard (CH) equation [39],

∂ci∂t

+ ~u · ∇ci = ∇ ·(DicikBT

∇δGδci

), (53)

expressing mass conservation for the LIT flux Eq. 5 withconvection in a mean flow ~u. For thermodynamic consis-tency, Di is given by Eq. 18, which reduces Eq. 53 tothe “modified” CH equation [58] in an ideal mixture [28].This is the “Cahn-Hilliard reaction (CHR) model”.

2. Homogeneous chemistry

For bulk reactions, Eq. 51 provides a new source termfor the CH equation,

∂ci∂t

+ ~u · ∇ci = ∇ ·(DicikBT

∇δGδci

)∓ cssiR

(δG

δcj

)(54)

(cs = reaction sites/volume). This also generalizes theAllen-Cahn equation [39] (AC), which corresponds toDi = 0 and linearization of R for |µj | kBT . Eq.54 is the fundamental equation of non-equilibrium chem-ical thermodynamics. It unifies and extends the CH andAC equations via a consistent set of reaction-diffusionequations based on variational principles. Eq. 52 is itsintegrated form for a reaction localized on a boundary.

3. Phase transformations

As a special case, Eq. 54 also describes phase transfor-mations with an immobile, non-conserved order parame-ter. For example, if f(c) has two local equilibrium states,cA and cB , then

ξ =c− cAcA − cB

(55)

is a phase field with minima at ξ = 0 and ξ = 1 satisfying

∂ξ

∂t= R

(δG

δξ

)(56)

This is the “Allen-Cahn reaction (ACR) model”, which isa nonlinear generalization of the AC equation for chemi-cal kinetics [3, 11, 13, 33].

Page 10: Theor y of Chemical Kinetics and Charge Tr ansfer based on

9

C. Example: Adsorption with Condensation

To illustrate the theory, we revisit surface adsorp-tion with attractive forces, strong enough to driveadatom condensation (separation into high- and low-density phases) on the surface [33]. Applications mayinclude water adsorption in concrete [60] or colloidal de-position in electrophoretic displays [61]. Following Cahnand Hilliard [56], the simplest model is a regular solutionof adatoms and vacancies with pair interaction energy Ω,

g = cs kBT [c ln c+ (1− c) ln(1− c)]

+Ωc(1− c) + Eac+κ

2|∇c|2 (57)

µ = kBT lnc

1− c+ Ω(1− 2c) + Ea +

κ

cs∇2c (58)

Below the critical point, T < Tc = Ω2kB

, the enthalpy

of adatom attraction (third term, favoring phase sepa-ration c = 0, 1) dominates the configurational entropyof adatoms and vacancies (first two terms, favoring mix-ing c = 1

2 ). The gradient term controls spinodal de-composition and stabilizes phase boundaries of thickness

λb =√

κcsΩ

and interphasial tension γb =√κcsΩ. Using

Eq. 15 to model the transition state with

µ = kBT lnc

1− c+ Ω(1− 2c) + Ea +

κ

cs∇2c (59)

the ACR model Eq. 56 takes the dimensionless form

∂c

∂t= KΘa(1− c)− c exp

(Ω(1− 2c) + κ∇2c

)(60)

where t = k1t, Ω = ΩkBT

= 2TcT , κ = κ

L2cskBTand

∇ = L∇ (with length scale L). This nonlinear PDEdescribes phase separation coupled to adsorption at aninterface (Fig. 7), controlled by the reservoir activity a.It resembles a reaction-diffusion equation, but there is nodiffusion; instead, −κ∇2c is a gradient correction to thechemical potential, which nonlinearly affects the adsorp-tion reaction rate. With modifications for charge transferand coherency strain, a similar PDE describes ion inter-calation in a solid host, driven by an applied voltage.

V. NONEQUILIBRIUM ELECTROCHEMICALTHERMODYNAMICS

A. Background

We thus return to our original motivation – phase sep-aration in Li-ion batteries (Fig. 1). Three important pa-pers in 2004 set the stage: Garcia et al. [62] formulatedvariational principles for electromagnetically active sys-tems, which unify the CH equation with Maxwell’s equa-tions; Guyer et al. [63] represented the metal/electrolyte

interface with a continuous phase field ξ evolving by ACkinetics [64]; Han et al. [42] used the CH equation tomodel diffusion in LFP, leading directly to this work.

When the time is ripe for a new idea, a number of sci-entists naturally think along similar lines. As describedin the Introduction, my group first reported phase-fieldkinetics (CHR and ACR) [3, 43] and modified Poisson-Nernst-Planck (PNP) equations [41] in 2007, the gen-eralized BV equation [35] in 2009, and the completetheory [13, 33] in 2011. Independently, Lai and Ciuccialso applied non-equilibrium thermodynamics to electro-chemical transport [49], but did not develop a variationalformulation. They proceeded to generalize BV kinet-ics [50, 51] (citing Singh et al. [3]) but used µ in placeof µex and neglected γex‡ . Tang et al. [65] were the firstto apply CHR to ion intercalation with coherency strain,but, like Guyer et al. [64], they assumed linear AC ki-netics. Recently, Liang et al. [66] published the BV-ACRequation, claiming that “in contrast to all existing phase-field models, the rate of temporal phase-field evolution...is considered nonlinear with respect to the thermody-namic driving force”. They cited my work [3, 10, 11, 13]as a “boundary condition for a fixed electrode-electrolyteinterface” (CHR) but overlooked the same BV-ACRequation for the depth-averaged ion concentration [3, 13],identified as a phase field for an open system [11, 13].They also set I0 =constant, which contradicts chemicalkinetics (see below).

B. Phase-Field Electrochemical Kinetics

We now apply phase-field kinetics to charged species.The Gibbs free energy of ionic materials can be modeledas [3, 10, 11, 13, 59, 62, 63, 67]:

G = Gmix +Gelec +Gsurf +∑i

µΘi

∫V

ci dV (61)

Gmix =

∫V

f(~c)dV +Ggrad (62)

Ggrad =1

2

∫V

(∇~c · κ∇~c−∇φ · εp∇φ+ σ : ε

)dV (63)

Gelec =

∫V

ρeφdV +

∮A

qsφdA (64)

where Ggrad is the free energy associated with all gradi-ents; Gelec is the energy of charges in the electrostaticpotential of mean force, φ; ~c is the set of concentrations(including electrons for mixed ion/electron conductors);f is the homogeneous Helmholtz free energy density, ρeand qs are the bulk and surface charge densities; εp is thepermittivity tensor; and σ and ε are the stress and straintensors. The potential φ acts as a Lagrange multiplier

Page 11: Theor y of Chemical Kinetics and Charge Tr ansfer based on

10

FIG. 7: Surface adsorption with condensation when an empty surface is brought into contact with a reservoir(µres = µ1 = kBT ln a > Ea = −kBT lnKΘ). 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.

constraining the total ion densities [11, 62] while enforc-ing Maxwell’s equations for a linear dielectric material( δGδφ = 0),

−∇ · εp∇φ = ρe =∑i

zieci (65)

− n · εp∇φ = qs (66)

The permittivity can be a linear operator, e.g. εp =ε0(1 − `2c∇2), to account for electrostatic correlations inionic liquids [59] and concentrated electrolytes [35, 68] (asfirst derived for counterion plasmas [69, 70]). ModifiedPNP equations [35, 41, 49, 50] correspond to Eq. 53 andEq. 65.

For elastic solids, the stress is given by Hooke’s law,σij = Cijklεkl, where C is the elastic constant tensor.The coherency strain,

εij =1

2

(∂ui∂xj

+∂uj∂xi

)−∑m

ε0ijmcm (67)

is the total strain due to compositional inhomogeneity(first term) relative to the stress-free inelastic strain (sec-ond term), which contributes to Gmix. In a mean-fieldapproximation (Vegard’s law), each molecule of speciesm exerts an independent strain ε0m (lattice misfit betweencm = 0, 1 with cΘm = cs). Since elastic relaxation (sound)is faster than diffusion and kinetics, we assume mechan-ical equilibrium, δG

δ~u = ∇ · σ = 0.For Faradaic reactions Eq. 19, the overpotential is the

thermodynamic driving force for charge transfer,

neη =∑j

sj,RδG

δcj,R−∑i

si,OδG

δci,O− n δG

δce, (68)

determined by the electrochemical potentials µi = δGδci

.For thermodynamic consistency, the diffusivities Eq. 18,Nernst voltage Eq. 24 and exchange current Eq. 29 must

depend on ~c, ∇~c, and σ via the variational activities Eq.47, given by

kBT ln ai =∂f

∂ci− ∇ · κ∇ci + σ : ε0i

cs−∇φ · ∂εp

∂ci∇φ (69)

for the ionic model above. The Faradaic current densityis

I = I0 F

(neη

kBT

)(70)

where

F (η) =

e−αη − e(1−α)η Butler-Volmer

e−η2/4λ

(e−αη − e(1−α)η

)Marcus

(71)and I0 is given by either Eq. 29 or Eqs. 38-41, respec-tively (λ = λ

kBT) . The charge-transfer rate, R = I

ne ,defines the CHR and ACR models, Eqs. 52-56, for elec-trochemical systems.

C. Example: Metal Electrodeposition

In models of electrodeposition [63, 64] and electroki-netics [71], the solid/electrolyte interface is representedby a continuous phase field ξ for numerical convenience(to avoid tracking a sharp interface). If the phase fieldevolves by reactions, however, it has physical significance,as a chemical concentration. For example, consider elec-trodeposition, Mn+ +ne− → M, of solid metal M from abinary electrolyte M+A− with dimensionless concentra-tions, ξ = c = c/cs and c±/c0, respectively. In order toseparate the metal from the electrolyte, we postulate

f = W [h(c) + c(c+ + c−)] + fion(c+, c−) (72)

with W kBT , where h = c2(1 − c)2 is an arbi-trary double-welled potential. For a dilute electrolyte,fion = kBT (c+ ln c+ + c− ln c−), without phase separa-tion [67], we include gradient energy only for the metal.

Page 12: Theor y of Chemical Kinetics and Charge Tr ansfer based on

11

The activities Eq. 69 for reduced metal

cskBT ln a = W [h′(c) + c+ + c−]− κ∇2c− ∂εp∂c|∇φ|2

(73)and metal cations

c0kBT ln a+ = Wc+ kBT ln c+ −∂εp∂c+|∇φ|2 (74)

define the current density Eq. 70 via

I0 = K0aαa1−α

+ , K0 =nek0a

ne cs

γ‡(75)

η =kBT

neln

a

a+ae− EΘ (76)

Note that the local potential for electrons and ions isunique (φ = φe, ∆φ = 0), but integration across thediffuse interface yields the appropriate interfacial voltage.

The ACR equation Eq. 56 for ξ = c with Eqs. 70- 76differs from prior phase-field models [64, 66]. Eq. 75 hasthe thermodynamically consistent dependence on reac-tant activities (rather than I0 =constant). Coupled withEq. 54 for c±, our theory also describes Frumkin correc-tions to BV kinetics [72, 73] and electro-osmotic flows [71]associated with diffuse charge in the electrolyte.

D. Example: Ion Intercalation

Hereafter, we neglect double layers and focus onsolid thermodynamics. Consider cation intercala-tion, An++B+ne− →AB, from an electrolyte reservoir(aO =constant) into a conducting solid B (ae =constant)as a neutral polaron (cR = c(x, t), zR = 0). The overpo-tential Eq. 68 takes the simple form

neη =δG

δc− (µO + µe) =

δGmixδc

+ ne∆Φ (77)

where

∆Φ = ∆φ− EΘ − kBT

neln aOae (78)

is the interfacial voltage relative to the ionic standardstate. The equilibrium voltage is

ne∆Φeq = −kBT ln a = −δGmixδc

. (79)

Note that potentials can be shifted for convenience: Baiet al. [13] and Ferguson and Bazant [28] set µΘ = 0 forions, so µ = kBT ln a = δGmix

δc ; Cogswell and Bazant [11]

defined “∆φ”= ∆Φ and shifted g by −c∆Φ, so eη = δGδc .

Our surface adsorption model Eq. 57 can be adaptedfor ion intercalation by setting Ea = e∆Φ. If the tran-sition state excludes s sites (where s > 1 could account

for the An+ solvation shell) and has strain −ε‡, then itsactivity coefficient Eq. 41 is

γ‡ = (1− c)−se−σ:ε‡+λ0/4 (80)

where λ0 = λ0

kBTand σ = σ

cskBT. The exchange current

Eq. 29 is

I0 = nek(c)cα(1− c)s−α eσ:∆ε+αΩ(1−2c)−α∇·κ∇c (81)

k(c) = k0css(a+ae(c))

1−α e−λ0/4 (82)

where a+ is the ionic activity in the electrolyte and ∆ε =ε‡−αε0 is the activation strain [74]. For semiconductors,

the electron activity ae = e∆Ef/kBT depends on c, ifthe intercalated ion shifts the Fermi level by donating anelectron to the conduction band, e.g. ∆Ef ∝ (1 + βc)2/d

for free electrons in d dimensions (as in LiWO3 with d = 3[75]).

VI. APPLICATION TO LI-ION BATTERYELECTRODES

A. Allen-Cahn-Reaction Model

The three-dimensional CHR model Eqs. 52-53 withcurrent density I = neR given by Eq. 70 and Eq. 81describes ion intercalation in a solid particle from anelectrolyte reservoir. In nanoparticles, solid diffusiontimes (ms-s) are much shorter than discharge times, soa reaction-limited ACR model is often appropriate. Inthe case of LFP nanoparticles, strong crystal anisotropyleads to a two-dimensional ACR model over the active(010) facet by depth averaging over Ns sites in the [010]direction [3, 13]. For particle sizes below 100nm, the con-centration tends to be uniform in [010] due to the fast dif-fusion [16] (uninhibited by Fe anti-site defects [18]) andelastically unfavorable phase separation [11].

Using Eq. 70 and Eq. 81 with ae =constant, ε‡ = αε0,α = 1

2 and s = 1, the ACR equation Eq. 56 takes thesimple dimensionless form [11, 13],

∂c

∂t= I0 F (µ+ ∆Φ) (83)

µ = lnc

1− c+ Ω(1− 2c)− κ∇2c+ σ : ε (84)

I0 =√c(1− c) e(Ω(1−2c)−κ∇2c)/2 (85)

where ∆Φ = ne∆ΦkBT

, t = Nskt. The total current inte-grated over the active facet

I(t) =

∫A

∂c

∂tdxdy (86)

is either controlled while solving for ∆Φ(t) (as in Fig. 8),or vice versa.

Page 13: Theor y of Chemical Kinetics and Charge Tr ansfer based on

12

x

c

c

c

I = 0.01

I = 0.25

I = 2

(a) (c) (b)

1 wave 2 waves

Spinodal decomposition à intercalation waves

Quasi-solid solution (partial phase separation)

Solid solution (suppressed phase separation)

FIG. 8: Suppression of phase separation at constant current in an Li-ion battery nanoparticle (ACR model withoutcoherency strain or surface wetting) [13]. (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 (Fig. 1(c)); quasi-solid

solution at I = 0.25; homogeneous filling at I = 2.

B. Intercalation Waves and Quasi-Solid Solutions

The theory predicts a rich variety of new intercala-tion mechanisms. A special case of the CHR model [3]is isotropic diffusion-limited intercalation [29, 30] with ashrinking-core phase boundary [31, 32], but the reaction-limited ACR model also predicts intercalation waves(or “domino cascades” [4]), sweeping across the ac-tive facet, filling the crystal layer by layer (Fig. 1(c))[3, 11, 13, 34, 65]. Intercalation waves result from spin-odal decomposition or nucleation at surfaces [13] andtrace out the voltage plateau at low current (Fig. 8).

The theory makes other surprising predictions aboutelectrochemically driven phase transformations. Singhet al. [3] showed that intercalation wave solutions of theACR equation only exist over a finite range of thermo-dynamic driving force. Based on bulk free energy cal-culations, Malik et al. [14] argued for a “solid solutionpathway” without phase separation under applied cur-rent, but Bai et al. [13] used the BV ACR model to showthat phase separation is suppressed by activation over-potential at high current (Fig. 8), due to the reducedarea for intercalation on the phase boundary (Fig. 1(c)).Linear stability analysis of homogeneous filling predictsa critical current, of order the exchange current, abovewhich phase separation by spinodal decomposition is im-possible. Below this current, the homogeneous state isunstable over a range of concentrations (smaller than thezero-current spinodal gap), but for large currents, thetime spent in this region is too small for complete phaseseparation. Instead, the particle passes through a tran-sient “quasi-solid solution” state, where its voltage andconcentration profile resemble those of a homogeneoussolid solution. When nucleation is possible (see below),a similar current dependence is also observed.

For quantitative interpretation of experiments, it is es-

sential to account for the elastic energy [11]. Coherencystrain is a barrier to phase separation (Fig. 9), whichtilts the voltage plateau (compared to Fig. 8) and re-duces the critical current, far below the exchange current.An unexpected prediction is that phase separation rarelyoccurs in situ during battery operation in LFP nanopar-ticles, which helps to explain their high-rate capabilityand extended lifetime [11, 13].

Phase separation occurs at low currents and can be ob-served ex situ in partially filled particles (Fig. 10). Crys-tal anisotropy leads to striped phase patterns in equi-librium [19–21], whose spacing is set by the balance ofelastic energy (favoring short wavelengths at a stress-freeboundary) and interfacial energy (favoring long wave-lengths to minimize interfacial area) [11]. Stanton andBazant [21] predicted that simultaneous positive and neg-ative eigenvalues of ε0 make phase boundaries tilt withrespect to the crystal axes. In LFP, lithiation causescontraction in the [001] direction and expansion in the[100] and [010] directions [1]. Depending on the degreeof coherency, Cogswell and Bazant [11] predicted phasemorphologies in excellent agreement with experiments(Fig. 10) and inferred the gradient penalty κ and theLiFePO4/FePO4 interfacial tension (beyond the reach ofmolecular simulations) from the observed stripe spacing.

C. Driven Nucleation and Growth

The theory can also quantitatively predict nucleationdynamics driven by chemical reactions. Nucleation isperhaps the least understood phenomenon of thermody-namics. In thermal phase transitions, such as boiling orfreezing, the critical nucleus is controlled by random het-erogeneities, and its energy is over-estimated by classicalspherical-droplet nucleation theory. Phase-field models

Page 14: Theor y of Chemical Kinetics and Charge Tr ansfer based on

13

3.2

3.25

3.3

3.35

3.4

3.45

3.5

0 0.2 0.4 0.6 0.8 1

Batte

ry v

olta

ge (V

)

x in LixFePO4

I/I0=.001I/I0=.25I/I0=.5I/I0=1I/I0=2

I = 0.001 I = 0.01 I = 0.033 I = 0.05

phase separation

solid solution

quasi-solid solutions at X=0.6 with increasing current

Li+ + FePO4 + e− → LiFePO4

+

FePO4 LiFePO4

FIG. 9: ACR simulations of galvanostatic discharge in a 100nm LiXFePO4 nanoparticle [11]. As the current isincreased, transient quasi-solid solutions (images from the shaded region) transition to homogeneous filling for

I > 0.1, as phase separation is suppressed.

(a)

(b)

Simulations Experiments

Spinodal decomposition at zero current (X=0.5)

FePO4

LiFePO4

FIG. 10: Phase separation of a 500nm particle of Li0.5FePO4 into Li-rich (black) and Li-poor phases (white) at zerocurrent in ACR simulations [11], compared with ex situ experimental images [1, 2]. (a) Coherent phase separation

with [101] interfaces. (b) Semi-coherent phase separation, consistent with observed 100 microcracks [1].

address this problem, but often lack sufficient details tobe predictive.

For battery nanoparticles, nucleation turns out to bemore tractable, in part because the current and voltagecan be more precisely controlled than heat flux and tem-perature. More importantly, the critical nucleus has awell-defined form, set by the geometry, due to strongsurface “wetting” of crystal facets by different phases.Cogswell and Bazant [12] showed that nucleation in bi-nary solids occurs at the coherent miscibility limit, as asurface layer becomes unstable and propagates into the

bulk. The nucleation barrier, Eb = −e∆Φ is set by co-herency strain energy (scaling with volume) in large par-ticles and reduced by surface energy (scaling with area)in nanoparticles. The barrier thus decays with the wet-ted area-to-volume ratio A/V and vanishes at a criticalsize, below which nanoparticles remain homogeneous inthe phase of lowest surface energy.

The agreement between theory and experiment – with-out fitting any parameters – is impressive (Fig. 11). Us-ing our prior ACR model [11] augmented only by ab initiocalculated surface energies (in Eq. 46), the theory is able

Page 15: Theor y of Chemical Kinetics and Charge Tr ansfer based on

14

FePO4 LiFePO4

Li “wetted” facets

metastable

stable

+ (a)

(b)

0

20

40

60

80

100

120

140

0 0.2 0.4 0.6 0.8 1

size

(n

m)

x in LixFePO4(d)

Theory

Experiments Wagemaker et al. (2011)

Miscibility Gap -Δφ∗∞

0 0.05 0.1 0.15 0.2 0.25 0

5

10

15

20

25

30

35

401/Lc

-Δφ∗

(mV)

A/V (nm-1)

Simulated limitFarkhondeh [20]

Meethong [11,12]Safrononv [18,19]

Yu [13]Come [16]

Dreyer [15]Zhu [17]

Theory N

ucle

atio

n B

arrie

r

dc = 22nm

-Δφ∗∞

0 0.05 0.1 0.15 0.2 0.25 0

5

10

15

20

25

30

35

401/Lc

-Δφ∗

(mV)

A/V (nm-1)

Simulated limitFarkhondeh [20]

Meethong [11,12]Safrononv [18,19]

Yu [13]Come [16]

Dreyer [15]Zhu [17]

Wetted Area/Volume (nm-1)

Par

ticle

Dia

met

er (

nm)

(c)

Li+

Farkhondeh & Delacourt (2012) Meethong et al. (2007)

Safronov et al. (2011-12) Yu et al. (2007)

Come et al. (2011) Dreyer et al. (2010) Zhu & Wang (2011)

Experiments

FIG. 11: (a) ACR simulation of galvanostatic nucleation in a realistic LFP nanoparticle shape (C3) [76] with a 150nm × 76 nm top (010) active facet [12]. 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 coherentmiscibility limit. (b) Discharge plot indicating nucleation by fluctuations in voltage or composition [12]. (c) Collapse

of experimental data for the nucleation voltage by the theory, without any fitting parameters [12]. (d) Sizedependence of the miscibility gap, fitted by the theory [11].

to collapse Eb data for LFP versus A/V , which lie eitheron the predicted line or below (e.g. from heterogeneities,lowering Eb, or missing the tiniest nanoparticles, lower-ing A/V ) [12]. This resolves a major controversy, sincethe data had seemed inconsistent (Eb = 2.0 − 37 mV),and some had argued for [3, 24, 77] and others againstthe possibility of nucleation (using classical droplet the-ory) [14]. The new theory also predicts that the nucle-ation barrier (Fig. 11(c)) and miscibility gap (Fig. 11(d))vanish at the same critical size, dc ≈ 22 nm, consistentwith separate Li-solubility experiments [15].

D. Mosaic Instability and Porous Electrodes

These findings have important implications for porousbattery electrodes, consisting of many phase separatingnanoparticles. The prediction that small particles trans-form before larger ones is counter-intuitive (since largerparticles have more nucleation sites) and opposite to clas-sical nucleation theory. The new theory could be usedto predict mean nucleation and growth rates in a sim-ple statistical model [77] that fits current transients inLFP [24] and guide extensions to account for the particlesize distribution.

Discrete, random transformations also affect voltagetransients. Using the CHR model [10] for a collection ofparticles in a reservoir, Burch [27] discovered the “mo-saic instability”, whereby particles switch from uniform

to sequential filling after entering the miscibility gap.Around the same time, Dreyer et al. [25] published a sim-ple theory of the same effect (neglecting phase separationwithin particles) supported by experimental observationsof voltage gap between charge/discharge cycles in LFPbatteries (Fig. 12(c)), as well as pressure hysteresis inballon array [26].

The key ingredient missing in these models is the trans-port of ions (in the electrolyte) and electrons (in theconducting matrix), which mediates interactions betweennanoparticles and becomes rate limiting at high current.Conversely, the classical description of porous electrodes,pioneered by Newman [29, 30], focuses on transport,but mostly neglects the thermodynamics of the activematerials [28, 50], e.g. fitting [31], rather than deriv-ing [13, 25, 49, 51], the voltage plateau in LFP. Theseapproaches are unified by non-equilibrium chemical ther-modynamics [28]. Generalized porous electrode theoryis constructed by formally volume averaging over themicrostructure to obtain macroscopic reaction-diffusionequations of the form Eq. 54 for three overlapping con-tinua – the electrolyte, conducting matrix, and activematerial – each containing a source/sink for Faradaic re-actions, integrated over the internal surface of the activeparticles, described by the CHR or ACR model.

The simplest case is the “pseudo-capacitor approxi-mation” of fast solid relaxation (compared to reactionsand macroscopic transport), where the active particlesremain homogeneous. Using our model for LFP nanopar-

Page 16: Theor y of Chemical Kinetics and Charge Tr ansfer based on

15

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 13.1

3.15

3.2

3.25

3.3

3.35

3.4

3.45

3.5

Filling Fraction

Volta

ge

C/50 DischargeC/50 Charge10C Discharge

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.5

1

x / L x

y / L

y

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.5

1

x / L x

y / L

y

0 0.2 0.4 0.6 0.8 10

0.5

1

x / L x

y / L

y

−0.5 0 0.5 10

0.5

1

x / L x

y / L

y

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 13.1

3.15

3.2

3.25

3.3

3.35

3.4

3.45

3.5

Filling Fraction

Volta

ge

C/50 DischargeC/50 Charge10C Discharge

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.5

1

x / L x

y / L

y

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.5

1

x / L x

y / L

y

0 0.2 0.4 0.6 0.8 10

0.5

1

x / L x

y / L

y

−0.5 0 0.5 10

0.5

1

x / L x

y / L

y

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 13.1

3.15

3.2

3.25

3.3

3.35

3.4

3.45

3.5

Filling Fraction

Volta

ge

C/50 DischargeC/50 Charge10C Discharge

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.5

1

x / L x

y / L

y

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.5

1

x / L x

y / L

y

0 0.2 0.4 0.6 0.8 10

0.5

1

x / L x

y / L

y

−0.5 0 0.5 10

0.5

1

x / L x

y / L

y

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 13.1

3.15

3.2

3.25

3.3

3.35

3.4

3.45

3.5

Filling Fraction

Voltage

C/50 DischargeC/50 Charge10C Discharge

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.5

1

x / L x

y / Ly

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.5

1

x / L x

y / Ly

0 0.2 0.4 0.6 0.8 10

0.5

1

x / L x

y / Ly

−0.5 0 0.5 10

0.5

1

x / L x

y / Ly

A

C

B

A

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 13.1

3.15

3.2

3.25

3.3

3.35

3.4

3.45

3.5

Filling Fraction

Volta

ge

C/50 DischargeC/50 Charge10C Discharge

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.5

1

x / L x

y / L

y

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.5

1

x / L x

y / L

y

0 0.2 0.4 0.6 0.8 10

0.5

1

x / L x

y / L

y

−0.5 0 0.5 10

0.5

1

x / L x

y / L

y

B

C

(a) (b)

(c) discharge

charge

Dreyer et al. (2010) Experiments

LiFePO4 FePO4

FePO4

FePO4

LiFePO4

LiFePO4

FIG. 12: Finite-volume simulations of a porous LFP cathode (T. Ferguson [28]). (a) Voltage versus state of chargeat different rates with profiles of the mean solid Li concentration (A-C), separator on the left, current collector on

the right. (b) SEM image of LFP nanoparticles represented by three finite volumes (P. Bai). (c) Experimentsrevealing a zero-current gap between noisy charge and discharge voltage plateaus (From Dreyer et al. [25]).

ticles [11], the porous electrode theory predicts the zero-current voltage gap, without any fitting (Fig. 12). (Usingthe mean particle size, the gap is somewhat too large,but this can be corrected by size-dependent nucleation(Fig. 11), implying that smaller particles were preferen-tially cycled in the experiments.) Voltage fluctuations atlow current correspond to discrete sets of transformingparticles. For a narrow particle size distribution, mosaicinstability sweeps across the electrode from the separatoras a narrow reaction front (Fig. 12(a) inset). As the cur-rent is increased, the front width grows, and the activematerial transforms more uniformly across the porouselectrode, limited by electrolyte diffusion. A wide parti-cle size distribution also broadens the reaction front, asparticles transform in order of increasing size. These ex-amples illustrate the complexity of phase transformationsin porous media driven by chemical reactions.

VII. CONCLUSION

This Account describes a journey along the “middleway” [78], searching for organizing principles of the meso-scopic domain between individual atoms and bulk mate-rials. The motivation to understand phase behavior inLi-ion battery nanoparticles gradually led to a theory ofcollective kinetics at length and time scales in the “mid-dle”, beyond the reach of both molecular simulations andmacroscopic continuum models. The work leveraged ad-vances in ab initio quantum-mechanical calculations andnanoscale imaging, but also required some new theoreti-cal ideas.

Besides telling the story, this Account synthesizes my

work as a general theory of chemical physics, which tran-scends its origins in electrochemistry. The main result,Eq. 54, generalizes the Cahn-Hilliard and Allen-Cahnequations for reaction-diffusion phenomena. The reac-tion rate is a nonlinear function of the species activitiesand the free energy of reaction (Eq. 7) via variationalderivatives of the Gibbs free energy functional (Eq. 51),which are consistently defined for non-equilibrium states,e.g. during a phase separation. For charged species, thetheory generalizes the Poisson-Nernst-Planck equationsof ion transport, the Butler-Volmer equation of electro-chemical kinetics (Eq. 29), and the Marcus theory ofcharge transfer (Eq. 37) for concentrated electrolytes andionic solids.

As its first application, the theory has predicted newintercalation mechanisms in phase-separating batterymaterials, exemplified by LFP:

• intercalation waves in anisotropic nanoparticles atlow currents (Fig. 8);

• quasi-solid solutions and suppressed phase separa-tion at high currents (Fig. 9);

• relaxation to striped phases in partially filled par-ticles (Fig. 10);

• size-dependent nucleation by surface wetting (Fig.11);

• mosaic instabilities and reaction fronts in porouselectrodes (Fig. 12);

These results have some unexpected implications, e.g.that battery performance may be improved with elevated

Page 17: Theor y of Chemical Kinetics and Charge Tr ansfer based on

16

currents and temperatures, wider particle size distribu-tions, and coatings to alter surface energies. The modelsuccessfully describes phase behavior of LFP cathodes,and my group is extending it to graphite anodes (“stag-ing” of Li intercalation with ≥ 3 stable phases) and aircathodes (electrochemical growth of Li2O2).

The general theory may find many other applicationsin chemistry and biology. For example, the adsorp-tion model (Fig. 7) could be adapted for the deposi-tion of charged colloids on transparent electrodes in elec-trophoretic displays. The porous electrode model (Fig.12) could be adapted for sorption/desorption kineticsin nanoporous solids, e.g. for drying cycles of cemen-titious materials, release of shale gas by hydraulic frac-turing, carbon sequestration in zeolites, or ion adsorp-tion and impulse propagation in biological cells. Thecommon theme is the coupling of chemical kinetics withnon-equilibrium thermodynamics.

Acknolwedgements

This work was supported by the National ScienceFoundation under Contracts DMS-0842504 and DMS-

0948071 and by the MIT Energy Initiative and would nothave been possible without my postdocs (D. A. Cogswell,G. Singh) and students (P. Bai, D. Burch, T. R. Fergu-son). P. Bai noted the Nernst factor in Eq. 39.

Bibliographical Information

Martin Z. Bazant received his B.S. (Physics, Math-ematics, 1992) and M.S. (Applied Mathematics, 1993)from the University of Arizona and Ph.D. (Physics, 1997)from Harvard University. He joined the faculty at MIT inMathematics in 1998 and Chemical Engineering in 2008.His honors include an Early Career Award from the De-partment of Energy (2003), Brilliant Ten from PopularScience (2007), and Paris Sciences Chair (2002,2007) andJoliot Chair (2008,2012) from ESPCI (Paris, France).

[1] G. Chen, X. Song, and T. Richardson, Electrochemicaland Solid State Letters 9, A295 (2006).

[2] C. V. Ramana, A. Mauger, F. Gendron, C. M. Julien,and K. Zaghib, J. Power Sources 187, 555 (2009).

[3] G. Singh, D. Burch, and M. Z. Bazant, Electrochim-ica Acta 53, 7599 (2008), arXiv:0707.1858v1 [cond-mat.mtrl-sci] (2007).

[4] C. Delmas, M. Maccario, L. Croguennec, F. L. Cras, andF. Weill, Nature Materials 7, 665 (2008).

[5] A. Padhi, K. Nanjundaswamy, and J. Goodenough, Jour-nal of the Electrochemical Society 144, 1188 (1997).

[6] J. Tarascon and M. Armand, Nature 414, 359 (2001).[7] B. Kang and G. Ceder, Nature 458, 190 (2009).[8] M. Tang, W. C. Carter, and Y.-M. Chiang, Annual Re-

view of Materials Research 40, 501 (2010).[9] N. Meethong, H.-Y. S. Huang, W. C. Carter, and Y.-M.

Chiang, Electrochem. Solid-State Lett. 10, A134 (2007).[10] D. Burch and M. Z. Bazant, Nano Letters 9, 3795 (2009).[11] D. A. Cogswell and M. Z. Bazant, ACS Nano 6, 2215

(2012).[12] D. A. Cogswell and M. Z. Bazant, submitted (????).[13] P. Bai, D. Cogswell, and M. Z. Bazant, Nano Letters 11,

4890 (2011).[14] R. Malik, F. Zhou, and G. Ceder, Nature Materials 10,

587 (2011).[15] M. Wagemaker, D. P. Singh, W. J. Borghols, U. Lafont,

L. Haverkate, V. K. Peterson, and F. M. Mulder, J. Am.Chem. Soc. 133, 10222 (2011).

[16] D. Morgan, A. V. der Ven, and G. Ceder, Electrochemicaland Solid State Letters 7, A30 (2004).

[17] L. Laffont, C. Delacourt, P. Gibot, M. Y. Wu, P. Kooy-man, C. Masquelier, and J. M. Tarascon, Chem. Mater.18, 5520 (2006).

[18] R. Malik, D. Burch, M. Bazant, and G. Ceder, NanoLetters 10, 4123 (2010).

[19] N. Meethong, H. Y. S. Huang, S. A. Speakman, W. C.Carter, and Y. M. Chiang, Adv. Funct. Mater. 17, 1115(2007).

[20] A. V. der Ven, K. Garikipati, S. Kim, and M. Wage-maker, J. Electrochem. Soc. 156, A949 (2009).

[21] L. G. Stanton and M. Z. Bazant, phase separation withanisotropic coherency strain arXiv:1202.1626v1 [cond-mat.mtrl-sci].

[22] A. V. der Ven and M. Wagemaker, ElectrochemistryCommunications 11, 881 (2009).

[23] M. Wagemaker, F. M. Mulder, and A. V. der Ven, Ad-vanced Materials 21, 2703 (2009).

[24] G. Oyama, Y. Yamada, R. Natsui, S. Nishimura, andA. Yamada, J. Phys. Chem. C 116, 7306 (2012).

[25] W. Dreyer, J. Jamnik, C. Guhlke, R. Huth, J. Moskon,and M. Gaberscek, Nat. Mater. 9, 448 (2010).

[26] D. Dreyer, C. Guhlke, and R. Huth, Physica D 240, 1008(2011).

[27] D. Burch, Intercalation Dynamics in Lithium-Ion Bat-teries (Ph.D. Thesis in Mathematics, Massachusetts In-stitute of Technology, 2009).

[28] T. R. Ferguson and M. Z. Bazant, J. Electrochem. Soc.159, A1967 (2012).

[29] M. Doyle, T. F. Fuller, and J. Newman, Journal of theElectrochemical Society 140, 1526 (1993).

[30] J. Newman, Electrochemical Systems (Prentice-Hall,Inc., Englewood Cliffs, NJ, 1991), 2nd ed.

[31] V. Srinivasan and J. Newman, Journal of the Electro-chemical Society 151, A1517 (2004).

[32] S. Dargaville and T. Farrell, Journal of the Electrochem-ical Society 157, A830 (2010).

Page 18: Theor y of Chemical Kinetics and Charge Tr ansfer based on

17

[33] M. Z. Bazant, 10.626 Electrochemical Energy Systems(Massachusetts Institute of Technology: MIT Open-CourseWare, http://ocw.mit.edu, License: CreativeCommons BY-NC-SA, 2011).

[34] D. Burch, G. Singh, G. Ceder, and M. Z. Bazant, SolidState Phenomena 139, 95 (2008).

[35] M. Z. Bazant, M. S. Kilic, B. Storey, and A. Ajdari,Advances in Colloid and Interface Science 152, 48 (2009).

[36] A. M. Kuznetsov and J. Ulstrup, Electron Transfer inChemistry and Biology: An Introduction to the Theory(Wiley, 1999).

[37] K. Sekimoto, Stochastic Energetics (Springer, 2010).[38] S. R. D. Groot and P. Mazur, Non-equilibrium Thermo-

dynamics (Interscience Publishers, Inc., New York, NY,1962).

[39] R. W. Balluffi, S. M. Allen, and W. C. Carter, Kineticsof materials (Wiley, 2005).

[40] I. Prigogine and R. Defay, Chemical Thermodynamics(John Wiley and Sons, 1954).

[41] M. S. Kilic, M. Z. Bazant, and A. Ajdari, Phys. Rev. E75, 021503 (2007).

[42] B. Han, A. V. der Ven, D. Morgan, and G. Ceder, Elec-trochimica Acta 49, 4691 (2004).

[43] G. K. Singh, M. Z. Bazant, and G. Ceder (2007),arXiv:0707.1858v1 [cond-mat.mtrl-sci].

[44] D. A. Beard and H. Qian, PLoS ONE 2, e144 (2007).[45] A. J. Bard and L. R. Faulkner, Electrochemical Methods

(J. Wiley & Sons, Inc., New York, NY, 2001).[46] A. A. Kulikovsky, Analytical Modelling of Fuel cells (El-

sevier, New York, 2010).[47] M. Eikerling and A. A. Kornyshev, J. Electroanal. Chem.

453, 89 (1998).[48] A. S. Ioselevich and A. A. Kornyshev, Fuel Cells 1, 40

(2001).[49] W. Lai and F. Ciucci, Electrochim. Acta 56, 531 (2010).[50] W. Lai and F. Ciucci, Electrochimica Acta 56, 4369

(2011).[51] W. Lai, Journal of Power Sources 196, 6534 (2011).[52] R. A. Marcus, J. Chem .Phys. 24, 966 (1956).[53] R. A. . Marcus, J. Chem .Phys. 43, 679 (1965).[54] R. A. Marcus, Rev. Mod. Phys. 65, 599 (1993).[55] I. M. Gelfand and S. V. Fomin, Calculus of Variations

(Dover, New York, 2000).[56] J. W. Cahn and J. W. Hilliard, J. Chem Phys. 28, 258

(1958).[57] J. D. van der Waals, Verhandel. Konink. Akad. Weten.

Amsterdam (Sect. 1) 1, 8 (1893), (Translation by J. S.Rowlinson, J. Stat. Phys. 1979, 20, 197-244).

[58] E. B. Nauman and D. Q. Heb, Chemical EngineeringScience 56, 1999 (2001).

[59] M. Z. Bazant, B. D. Storey, and A. A. Kornyshev, Phys.Rev. Lett. 106, 046102 (2011).

[60] M. Z. Bazant and Z. P. Bazant, J. Mech. Phys. Solids60, 1660 (2012).

[61] P. Murau and B. Singer, J. Appl. Phys. 49, 48204829(1978).

[62] R. E. Garcia, C. M. Bishop, and W. C. Carter, ActaMater. 52, 11 (2004).

[63] J. E. Guyer, W. J. Boettinger, J. A. Warren, and G. B.McFadden, Phys. Rev. E 69, 021603 (2004).

[64] J. E. Guyer, W. J. Boettinger, J. A. Warren, and G. B.McFadden, Phys. Rev. E 69, 021604 (2004).

[65] M. Tang, J. F. Belak, and M. R. Dorr, The Journal ofPhysical Chemistry C 115, 4922 (2011).

[66] L. Liang, Y. Qi, F. Xue, S. Bhattacharya, S. J. Harris,and L.-Q. Chen, Phys. Rev. E 86, 051609 (2012).

[67] S. Samin and Y. Tsori, J. Phys. Chem. B 115, 75 (2011).[68] B. D. Storey and M. Z. Bazant, Phys. Rev. E 86, 056303

(2012).[69] C. D. Santangelo, Physical Review E 73, 041512 (2006).[70] M. M. Hatlo and L. Lue, Europhysics Letters 89, 25002

(2010).[71] M. M. Gregersen, F. Okkels, M. Z. Bazant, and H. Bruus,

New Journal of Physics 11, 075016 (2009).[72] P. M. Biesheuvel, M. van Soestbergen, and M. Z. Bazant,

Electrochimica Acta 54, 4857 (2009).[73] P. Biesheuvel, Y. Fu, and M. Bazant, Physical Review E

83 (2011).[74] M. J. Aziz, P. C. Sabin, and G. Q. Lu, Phys. Rev. B 41,

9812 (1991).[75] I. D. Raistrick, A. J. Mark, and R. A. Huggins, Solid

State Ionics 5, 351 (1981).[76] K. C. Smith, P. P. Mukherjee, and T. S. Fisher, Phys.

Chem. Chem. Phys. 14, 7040 (2012).[77] P. Bai and G. Tian, Electrochimica Acta 89, 644 (2013).[78] R. B. Laughlin, D. Pine, J. Schmalian, B. P. Stojkovic,

and P. Wolynes, PNAS 97, 32 (2000).