+ All Categories
Home > Documents > Suppression of Phase Separation in LiFePO Nanoparticles ... · LiFePO 4 battery.

Suppression of Phase Separation in LiFePO Nanoparticles ... · LiFePO 4 battery.

Date post: 23-Jul-2020
Category:
Upload: others
View: 9 times
Download: 0 times
Share this document with a friend
7
Published: October 10, 2011 r2011 American Chemical Society 4890 dx.doi.org/10.1021/nl202764f | Nano Lett. 2011, 11, 48904896 LETTER pubs.acs.org/NanoLett Suppression of Phase Separation in LiFePO 4 Nanoparticles During Battery Discharge Peng Bai, ,§ Daniel A. Cogswell, and Martin Z. Bazant* ,,Department of Chemical Engineering and Department of Mathematics, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, Massachusetts 02139, United States § State Key Laboratory of Automotive Safety and Energy, Department of Automotive Engineering, Tsinghua University, Beijing 100084, People's Republic of China b S Supporting Information ABSTRACT: Using a novel electrochemical phase-eld model, we question the common belief that Li X FePO 4 nanoparticles always separate into Li-rich and Li-poor phases during battery discharge. For small currents, spinodal decomposition or nucleation leads to moving phase boundaries. Above a critical current density (in the Tafel regime), the spinodal disappears, and particles ll homogeneously, which may explain the superior rate capability and long cycle life of nano-LiFePO 4 cathodes. KEYWORDS: Li-ion battery, LiFePO 4 , phase-eld model, ButlerVolmer equation, spinodal decomposition, intercalation waves L iFePO 4 -based electrochemical energy storage 1 is one of the most promising developments for electric vehicle power sy- stems and other high-rate applications such as power tools and renewable energy storage. 2,3 Nanoparticles of LiFePO 4 have re- cently been used to demonstrate ultrafast battery discharge 4 and high power density carbon pseudocapacitors. 5 However, the design of these systems remains both challenging and controversial due to poor understanding of lithium intercalation dynamics in nanoparticles with highly anisotropic transport, 6,7 size-depen- dent diusivity, 8 and a strong tendency to separate into Li-rich and Li-poor phases. 911 The fact that reported values of the dif- fusivity and exchange current for LiFePO 4 vary by orders of mag- nitude 6,8,1216 indicates fundamental uncertainty in the kinetic processes governing battery operation and highlights the need for models that take the unique properties of LiFePO 4 into account. The prevailing belief is that phase separation always occurs in Li X FePO 4 during battery operation. This assumption is built into widely accepted porous-electrode models, 12,13,17 where each particle is modeled an isotropic sphere with a shrinking coreof one phase displaced by a shell of the other phase, as suggested in the rst experimental paper. 1 In contrast, phase-eld models provide a more general, thermodynamically consistent treatment of nucleation and growth without resorting to the articial placement of phase boundaries. 18,19 Considering the excitement surrounding LiFePO 4 as a phase- separating cathode material, phase-eld methods have received relatively little attention. Tang et al. modeled phase separation and crystalline-to-amorphous transformations in spherical iso- tropic LiFePO 4 particles with a phase-eld model, 20 prompting experiments to seek the predicted amorphous surface layers. 21 Kao et al. compared the model to X-ray diraction data and rst proposed the idea of overpotential-dependent phase transforma- tion pathways. 22 It is now becoming appreciated that models should account for the strongly anisotropic transport in crystal- line LiFePO 4 , 6,7 as well as surface reaction kinetics, which are likely to be rate limiting in nanoparticles. 2325 Singh, Ceder, and Bazant (SCB) 26 rst modeled reaction-limited intercalation in LiFePO 4 nanoparticles by coupling an anisotropic phase-eld model with Faradaic reactions on the active facet. They predicted Received: August 10, 2011 Revised: September 26, 2011
Transcript
Page 1: Suppression of Phase Separation in LiFePO Nanoparticles ... · LiFePO 4 battery.

Published: October 10, 2011

r 2011 American Chemical Society 4890 dx.doi.org/10.1021/nl202764f |Nano Lett. 2011, 11, 4890–4896

LETTER

pubs.acs.org/NanoLett

Suppression of Phase Separation in LiFePO4 Nanoparticles DuringBattery DischargePeng Bai,†,§ Daniel A. Cogswell,† and Martin Z. Bazant*,†,‡

†Department of Chemical Engineering and ‡Department of Mathematics, Massachusetts Institute of Technology,77 Massachusetts Avenue, Cambridge, Massachusetts 02139, United States§State Key Laboratory of Automotive Safety and Energy, Department of Automotive Engineering, Tsinghua University,Beijing 100084, People's Republic of China

bS Supporting Information

ABSTRACT:

Using a novel electrochemical phase-field model, we question the common belief that LiXFePO4 nanoparticles always separate intoLi-rich and Li-poor phases during battery discharge. For small currents, spinodal decomposition or nucleation leads tomoving phaseboundaries. Above a critical current density (in the Tafel regime), the spinodal disappears, and particles fill homogeneously, whichmay explain the superior rate capability and long cycle life of nano-LiFePO4 cathodes.

KEYWORDS: Li-ion battery, LiFePO4, phase-field model, Butler�Volmer equation, spinodal decomposition, intercalation waves

LiFePO4-based electrochemical energy storage1 is one of themost promising developments for electric vehicle power sy-

stems and other high-rate applications such as power tools andrenewable energy storage.2,3 Nanoparticles of LiFePO4 have re-cently been used to demonstrate ultrafast battery discharge4 andhigh power density carbon pseudocapacitors.5 However, the designof these systems remains both challenging and controversialdue to poor understanding of lithium intercalation dynamics innanoparticles with highly anisotropic transport,6,7 size-depen-dent diffusivity,8 and a strong tendency to separate into Li-richand Li-poor phases.9�11 The fact that reported values of the dif-fusivity and exchange current for LiFePO4 vary by orders of mag-nitude6,8,12�16 indicates fundamental uncertainty in the kineticprocesses governing battery operation and highlights the needfor models that take the unique properties of LiFePO4 intoaccount.

The prevailing belief is that phase separation always occurs inLiXFePO4 during battery operation. This assumption is built intowidely accepted porous-electrode models,12,13,17 where eachparticle is modeled an isotropic sphere with a “shrinking core”of one phase displaced by a “shell” of the other phase, as suggestedin the first experimental paper.1 In contrast, phase-field models

provide a more general, thermodynamically consistent treatmentof nucleation and growth without resorting to the artificialplacement of phase boundaries.18,19

Considering the excitement surrounding LiFePO4 as a phase-separating cathode material, phase-field methods have receivedrelatively little attention. Tang et al. modeled phase separationand crystalline-to-amorphous transformations in spherical iso-tropic LiFePO4 particles with a phase-field model,20 promptingexperiments to seek the predicted amorphous surface layers.21

Kao et al. compared the model to X-ray diffraction data and firstproposed the idea of overpotential-dependent phase transforma-tion pathways.22 It is now becoming appreciated that modelsshould account for the strongly anisotropic transport in crystal-line LiFePO4,

6,7 as well as surface reaction kinetics, which arelikely to be rate limiting in nanoparticles.23�25 Singh, Ceder, andBazant (SCB)26 first modeled reaction-limited intercalation inLiFePO4 nanoparticles by coupling an anisotropic phase-fieldmodel with Faradaic reactions on the active facet. They predicted

Received: August 10, 2011Revised: September 26, 2011

Page 2: Suppression of Phase Separation in LiFePO Nanoparticles ... · LiFePO 4 battery.

4891 dx.doi.org/10.1021/nl202764f |Nano Lett. 2011, 11, 4890–4896

Nano Letters LETTER

moving phase boundaries (“intercalation waves”) sweepingacross FePO4 planes, consistent with experimental inferencesof a “domino cascade,”27 as well as the suppression of equilibriumphase separation with decreasing particle size.28 Tang et al.recently modeled the effect of coherency strain on intercalationwave structure.29 No model, however, has yet addressed phase-separation dynamics under the experimentally relevant conditionof constant applied current.

In this article, we develop a theory of reaction-limitedintercalation in anisotropic nanoparticles and predict thatphase separation is suppressed above a critical current. Thepossibility of a solid-solution pathway for LiFePO4 intercala-tion has also been suggested by Malik et al.30 based onequilibrium bulk free energy calculations, but here we showthat phase separation out of equilibrium (i.e., during batterydischarge) is controlled by surface reactions, which lead tounstable “quasi-solid solutions” and ultimately homogeneousfilling as current increases.General Theory. We assume a regular solution model for

homogeneous free energy density31 and adopt a diffusionalchemical potential derived from the Cahn�Hilliard free energyfunctional32,33 for intercalated lithium

μ ¼ Ωð1� 2~cÞ þ 2kBTln~c

1�~c

� �� VsK∇2~c ¼ kBTln a

ð1Þ~c is the local filling (mole) fraction of lithium in FePO4; theregular solution parameterΩ is the enthalpy of mixing per site(see Supporting Information); kB is Boltzmann’s constant andT is the absolute temperature; Vs is the volume per intercala-tion site;K is the Cahn�Hilliard gradient energy coefficient;32

a = γ~c is the local activity of lithium, scaled to unity at thevoltage plateau, and γ is the activity coefficient. The factor2kBT accounts for the configurational entropy of both lithiumions and electrons, due to low electron mobility.34 While otherchoices of the free energy function are possible, this particularchoice fits measured solubility limits35 while using an appro-priate energy barrier height determined in other modelingwork.20,22,29,30,36

The diffusional chemical potential is appropriate for systemswith a fixed number of available lattice sites. It is the free energychange per site for adding one lithium ion while consuming onevacancy and defines the electrochemical potential of the reducedstate (left-hand side) of the Faradic reaction

LiFePO4 � FePO4 h Liþ þ e� ð2ÞThe oxidized state (right-hand side) consists of a lithium ionin the electrolyte and an electron that either diffuses throughthe solid or conducts through a metallic additive. The electro-chemical potential of the oxidized state, μO = (kBT ln a+ + eϕ)� eϕe, consists of the chemical and electrostatic energies of Li+

in the electrolyte plus the electrostatic energy of electrons.a+ = γ+~c+ is the local activity of the Li

+ in the electrolyte, andϕ and ϕe are the mean electrostatic potentials of ions and ele-ctrons, respectively. The local voltage drop across the inter-face is Δϕ = ϕe � ϕ.Next we use transition state theory for concentrated

solutions57 to relate voltage to the local current density J throughthe solid surface for the net cathodic reaction (see SupportingInformation). We assume that the mean electrostatic energy ofthe activated state is equal to α times that of the reduced state

(zero) plus (1 � α) times that of the oxidized state, which leadsto the Butler�Volmer (BV) equation37

J ¼ J0 exp � αeηkBT

� �� exp ð1� αÞ eη

kBT

� �� �ð3Þ

where α is the electron-transfer symmetry factor, and η = Δϕ �Δϕeq is the surface overpotential (the activation polarizationonly). The Nernst equilibrium voltage Δϕeq and the exchangecurrent density J0 are given by

Δϕeq ¼ kBTe

lnaþa

and J0 ¼ ea1 � αþ aα

Asτ0γAð4Þ

whereAs is the area of a reacting site, and τ0 is the mean time for asingle reaction step. γA is the chemical activity coefficient of theactivated state, which we take to be (1 � ~c)�1 to account forexcluded volume at the reacting site.Equations 3 and 4 represent the most general form of the BV

equation for concentrated solutions.57 BothΔϕeq and J0 depend onthe lithium concentration and its gradients via eq 1. This unusualrate dependence focuses reactions where the phase boundarymeets the active facet. Our modified BV model reduces to thestandard BV equation in the limit of a dilute solution (γ = γ+ =γA = 1), thus correcting the SCB rate expression for consistencywith statistical thermodynamics. It is straightforward to addelastic strain energy to μ or activation strain energy36,38 to γA,and our modified BV eqs 3 and 4 are a natural boundarycondition for the Cahn�Hilliard equation in the solid26,36 andmodified Poisson�Nernst�Planck equations39,40 or porouselectrode theory in the electrolyte.12,17,41

Reduced Model. In order to capture just the essential physicsof nano-LiFePO4, we take a simple limit of the general theorywhich is valid for a single nanoparticle. The active material inhigh-rate LiFePO4 cathodes is an ultrafine powderwith 30�100 nmparticle sizes.4,5 At such small length scales, the electrolyte con-centration ~c+ is nearly uniform and quasi-steady, since the diffu-sion time through the electrolyte (<10 μs) is much smaller thanthrough the solid (<10 ms). For simplicity we set a+ = 1 aroundthe particle so thatΔϕeq =�μ/e and η =Δϕ + μ/e. Note that η <0 for the cathodic reaction (insertion). Neglecting macroscopicvoltage losses (electrolyte concentration polarization, separatorand anode resistances, etc.), the operating voltage of our modelnanoparticle battery is V = VΘ + Δϕ, where VΘ is the standardpotential defined by the open circuit voltage plateau (VΘ= 3.42V vsLi metal).Next we take into account the strongly anisotropic transport

properties of the crystal, which are neglected in spherically sym-metric continuum models.12,17,20 Experiments and ab initiocalculations have shown that Li transport in a perfect LiXFePO4

crystal is confined to 1D channels in the y-direction (bpnma) withlittle possibility of transverse diffusion6,7 (see Figure 1a). In largermicroparticles, randomly distributed defects provide sites forchannel blocking and interchannel hopping, causing diffusivity todecrease with increasing particle size.8 Here we focus on defect-free nanoparticles and neglect the possibility of amorphizationincluded in other isotropic particle models.20

Following SCB, we neglect phase separation in the y-directionand assume that the bulk concentration in each y-channel quicklyequilibrates to the surface concentration due to fast diffusion, andremains uniform due to elastic coherency strain.29,34,36,42 Withdecreasing particle radius R, the diffusion time τd = R2/D de-creases as R2, while the reaction time decreases only as τr = τ0R/l

Page 3: Suppression of Phase Separation in LiFePO Nanoparticles ... · LiFePO 4 battery.

4892 dx.doi.org/10.1021/nl202764f |Nano Lett. 2011, 11, 4890–4896

Nano Letters LETTER

(where l is the length of a Li site), making intercalation reaction-limited in sufficiently small particles. For an ideal crystal with R =25 nm (a 50 nm particle with two active surfaces) and D =10�8 cm2/sec from ab initio calculations,6 the diffusion time isonly τd = 0.6 ms, which is much less than the shortest reporteddischarge time of 10 s.4 Defects can lead to reduced diffusivity,8

but theoretically only for particles larger than 100 nm. Experi-ments cannot observe diffusion in single nanoparticles, but evenwithD = 10�13�10�12 cm2/s from a recent study,13 the diffusiontime for a 50 nm particle is only 1 minute.We thus arrive at a reaction-limited model where the 3D

concentration profile is represented by a 2D pattern~c(~x,~z,~t) overthe active (010) facet, which satisfies the following dimensionlessnonlinear PDE

∂~c∂~t

¼ ~J0½expð � αη~Þ � expðð1� αÞη~Þ� ð5ÞWe scale length to the facet width L, voltage to the thermalvoltage kBT/e, and time to themean timeNHτ0 to fill a channel ofNH sites. The dimensionless exchange current density ~J0 =Asτ0J0/e and overpotential η~ = eη/kBT are given by

~J0 ¼ ~c exp12ðΩ~ð1� 2~cÞ � ~K∇~2~cÞ

� �ð6Þ

η~ ¼ μ~ þ Δϕ~

¼ Ω~ð1� 2~cÞ þ 2ln~c

1�~c

� �� ~K∇~2~c þ Δϕ~ ð7Þ

where ~K = VsK/kBTL2 is the dimensionless gradient energy

coefficient that penalizes sharp interfaces, introducing interfacialenergy and setting the diffuse interface width.Since LiFePO4/FePO4 phase boundaries are most likely to be

aligned with the yz-plane (bc-plane) of minimum strain,42 wefurther simplify eqs 5�7 to a one-dimensional PDE for thedepth-averaged filling fraction ~c(~x,~t) of a particle of length ~L = 1in the direction of phase separation. The total current is anintegral over the active facet area A, I =

RA J dA, and takes the

dimensionless form

~I ¼ τ0IeNA

¼Z 1

0

∂~c∂~t

d~x ð8Þ

The exchange current is I0 = NAe/τ0, and NA = A/As is thenumber of active surface sites (or lithium channels). Undergalvanostatic conditions, eq 8 is an integral constraint thatimplicitly determines the voltage Δϕ~(~t).Equations 5�8 describe a new class of electrochemical phase

transformation kinetics. While Cahn�Hilliard32 and Allen�Cahn43 kinetics describe phase separation with and without localconservation, respectively, our model describes phase separationunder a boundary integral constraint. It applies to systems incontact with a reservoir of mass at fixed chemical potential withan imposed total flux and assumes a nonlinear flux-potential(current�voltage) relation, suitable for electrochemical reactions.The model provides a simple paradigm to understand nonequili-brium pattern formation driven by an applied voltage or current.Phase Separation at Constant Voltage. The condition of

thermodynamic equilibrium in eq 7 is η~ = Δϕ~ + μ~ = 0, asillustrated in Figure 2. Applying aΔϕ~ has the effect of raising theenergy of one phase relative to the other phase, and the inter-sections of constant Δϕ~ with the homogeneous chemical poten-tial curve�μ~h correspond to equilibrium states. For large appliedvoltages (i.e., Δϕ~ > Δϕ~max = �μ~h

min/e ≈ 1.54 or Δϕ~ < Δϕ~min =�μ~h

max/e ≈ �1.54), there is only one solution A2 (or B2),

Figure 2. Plot of the diffusional chemical potential against the fillingfraction of the particle at equilibrium. The thick solid curves are theequilibrium chemical potential in a phase-separating system. If phase-separation is suppressed, the chemical potential deviates from the voltageplateau and instead follows the metastable and unstable pathway. Dash-dot (red) lines are different Δϕ, and small circles indicate equilibriumstates at different applied voltages.

Figure 1. Schematic model of a LiXFePO4 nanoparticle at low over-potential. (FeO6 and PO4 structures are not shown in order to highlightinserted Li.) (a) Lithium ions are inserted into the particle (blue arrows)from the active (010) facet with fast diffusion and no phase separation inthe depth (y) direction, forming a phase boundary of thickness λbetween full and empty channels. (b) The resulting 1D concentrationprofile (local filling fraction) transverse to the FePO4 planes for a particleof size L. The average compositions X = 0.4, 0.5, and 0.6 reflect mixturesof coexisting Li-poor and Li-rich phases, X(A1) and X(B1), respectively,which vary as the phase boundary moves during charging or discharging.Homogeneous profiles at high and low concentrations,X(A2) andX(B2)respectively (see Figure 2), are also shown.

Page 4: Suppression of Phase Separation in LiFePO Nanoparticles ... · LiFePO 4 battery.

4893 dx.doi.org/10.1021/nl202764f |Nano Lett. 2011, 11, 4890–4896

Nano Letters LETTER

corresponding to a stable Li-poor (or Li-rich) composition. Atthe critical voltage Δϕ~max or Δϕ~min, a second solution appears,corresponding to a spinodal composition that is linearly unstableto composition fluctuations. In a closed system at zero current,the instability leads to stable phase separation with constant meancomposition. In an open system at constant voltage, transientphase separation can occur but the only stable state is a homo-geneous solution. As shown below, the most unstable wavelengthfor spinodal decomposition is set by the particle size, similar toAllen�Cahn dynamics.43

Three solutions appear for small applied voltages whereΔϕ~min < Δϕ~ < Δϕ~max. The two extreme solutions A1 and B1are linearly stable. The middle solution is unstable, lying withinthe spinodal region. Because of the applied voltage, one outersolution (A1) has a higher energy than the other (B1), which ismetastable. Phase transformation can be triggered with an inter-face that transforms the metastable phase into the stable phase.Focusing the reaction at the phase boundary causes it topropagate across crystal planes as an “intercalation wave”,26

as illustrated in Figure 1, with a velocity, ~v ∼ �λ~Δϕ~/2, scalingas the applied voltage (see Supporting Information). Stablephase separation (v = 0) occurs at the plateau voltage whereΔϕ~ = 0.Phase Separation at Constant Current. To investigate the

feasibility of phase separation during battery operation, we con-sider a homogeneous system at concentration~c under an appliedcurrent ~I. A miscibility gap cannot be defined since the system isout of equilibrium, although there is still a linearly unstable spinodalregion. This instability at constant total flux is the nonequilibriumorigin of phase separation. From eqs 5�8, the current�voltagerelation is:

Δϕ~ ¼ Δϕ~eqð~cÞ þ η~ ¼ � μ~ð~cÞ � 2 sinh�1~I

2~J0ð~cÞ

!ð9Þ

The first term is the (homogeneous) open circuit voltage, and thesecond is the overpotential, which is derived from BV eq 3 withα = 1/2. The battery voltage is the change in total free energyG perelectron transferred from the anode to the cathode, which has thedimensionless form Δϕ~ = �(d~G)/(d~c). By applying a secantconstruction18 to ~G(~c) and neglecting finite-size effects, we findthat the system undergoes spinodal decomposition if (d2~G)/(d~c2) < 0 or (dΔϕ~)/(d~c) > 0. The spinodal region has a noveldependence on the applied current via eq 9 which is captured by thestability boundary in Figure 3.During battery discharge,~I > 0 and the spinodal range shrinks

as the concentration-dependent overpotential overcomes thesolid-solution voltage barrier, Δϕ~ss = Δϕ~max � Δϕ~min. Above acritical current~I >~Ic, the battery voltage strictly decreases, dΔϕ~/d~c < 0, and the homogeneous state is linearly stable for all com-positions. The critical current~Ic and corresponding spinodal con-centration~cc satisfy [(∂Δϕ~)/(∂~c)](~c,~I) = [(∂

2Δϕ~)/(∂~c2)](~c,~I) = 0.Using Ω = 0.183 eV for LiFePO4 (fitted to the room-tempera-ture miscibility gap), the upper bound for the critical current(scaled to the exchange current) is ~Ic ≈ 1.9, as illustrated inFigure 3. Far into the solid-solution regime, where ~I . ~J0, thevoltage has a simple Tafel form Δϕ~ ∼ 2 ln[(1 � ~c)/~I].Unlike equilibrium phase diagrams, the nonequilibrium stabi-

lity diagram (Figure 3) is traversed from left to right duringdischarge, since X = ~c + ~I~t. Whenever the homogeneously fillingstate is unstable, we must also determine whether phase

separation has enough time to occur before the particle becomesfull. We will use the term “quasi-solid solution” to describe anonequilibrium system that is unstable, but lacks the time to fullyphase-separate.Performing a linear stability analysis of eqs 5�8 around a

homogeneous base state~c +~I~t, the dimensionless growth rate~s =sτ0 for a perturbation of wavenumber ~k has the general form

~sð~k,~c,~IÞ ¼ � ðμ~ 0hð~cÞ þ ~K~k2Þ

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi~I2

!2

þ ðJ̅0ð~cÞÞ2vuut

þ J̅00 ð~cÞJ̅0ð~cÞ þ 1

2~K~k2

!~I ð10Þ

where μ~h(~c) and J0(~c) are for homogeneous states and primesdenote d/(d~c) (see Supporting Information for derivation).Themost unstablemode has the smallest allowable wavenumber,~k = 2π (if boundary effects and coherency strain are neglected36).For eqs 6 and 7 in one dimension, the marginal stability curve~smax(~c,~I) = ~s(2π,~c,~I) = 0 is plotted in Figure 3 (see SupportingInformation eq (S24)). Requiring that the instability growth timeτi(c,I) = τ0/~smax(~c,~I) be greater than the time to fill the particle τf =Ne/I (where N = NANH is the total number of sites) yields adimensionless relation ~smax(~c,~I) < ~I, which approximates the condi-tions for a quasi-solid solution (above the dashed curve in Figure 3).The preceding analysis neglects the nonlinear growth of the

instability, the statistical occurrence of fluctuations (which tendto excite shorter, less unstable wavelengths), and other effectssuch as geometry, elasticity, and the amount and configuration ofinterfaces, all of which can further suppress phase separationsignificantly.36 The threshold between quasi-solid solution andphase separation (dotted curve in Figure 3) thus lies well belowthe linear instability prediction (dashed curve). To better under-stand the transition current ~Is beyond which phase separation is

Figure 3. Dependence of spinodal region on the applied current. Thesolid curve gives the marginal stability of the system under appliedcurrent, and the dashed curve is the boundary between phase separationand quasi-solid solution according to linear stability analysis. In practice,the boundary where phase separation is observed is expected to besignificantly below the linear stability limits, for example, the dottedcurve is inferred from numerical simulations and scaling arguments.

Page 5: Suppression of Phase Separation in LiFePO Nanoparticles ... · LiFePO 4 battery.

4894 dx.doi.org/10.1021/nl202764f |Nano Lett. 2011, 11, 4890–4896

Nano Letters LETTER

suppressed, we make a simple scaling argument based on thecreation ofNw intercalation waves. If the increased overpotentialfor the reduced active area Nw λ~ is larger than the voltage gainfrom homogeneous reaction, |Δϕ~min| ∼Ω~ /4, then a quasi-solidsolution is preserved, which implies~Is∼Nwλ~Ω~ /4 (see Support-ing Information). For two waves, we find ~Is ≈ 0.2 withcorresponding overpotential ηs ≈ 40 mV, consistent with thesimulation results. Typical overpotentials are much larger, so weconclude that phase separation is suppressed in nano-LiFePO4

during normal battery operation.Simulations of Battery Discharge. To test our analytical

predictions, we simulate reaction-limited lithium insertion in aLiXFePO4 nanoparticle for a range of applied currents and relatephase-separation kinetics to transient battery voltage (Figures 4and 5). Movies are available in the Supporting Information.We solve eqs 5�8 numerically using an explicit finite-differencemethod, adjustingΔϕ~ at each time step to maintain the constantcurrent. Langevin noise with a variance scaled to ~J0 is added tothe concentration variable to account for thermal fluctuations.44

We simulate a particle of length L = 100 nm with a phaseboundary thickness9 of λ = 5 nm at room temperature (VsK =0.684 eV nm2, comparable to Tang et al.20).

First we simulate battery discharge for very small current, ~I =0.01 , ~Is, well below the critical current for phase separation(Figure 3). In the absence of nucleation, the system fills tothe spinodal point and noise triggers spontaneous phase separa-tion. Two intercalation waves then propagate through the crystal(Figure 4b). At the spinodal composition, the voltage (Figure 4a)suddenly jumps to the equilibrium plateau, with a slightoverpotential due to a kinetic “wave resistance”, inversely pro-portional to the active interfacial area. When one wave anni-hilates at the left side facet, there is a small bump in the voltage atconstant current due to interfacial energy released when thephase boundary is removed. The overpotential then doubles (sincetwo waves are replaced by one), until the second wave reachesthe right side facet, producing another voltage bump and leavingthe particle full of lithium. Although the bumps overshoot thetheoretical equilibrium plateau at VΘ, the overpotential relativeto the Nernst equilibrium potential VΘ � μ/e remains negative.Next we consider larger currents with suppressed phase sepa-

ration. For ~I = 0.25 > ~Is, the system behaves as a quasi-solidsolution. Linear instability of the concentration profile occurs,but it lacks the time to grow before the particle becomes full(Figure 4c). The predicted voltage rise during partial instability isunusual for a battery material, but closely resembles some pre-viously unexplained data for reaction-limited nano-LiFePO4 (see20C curve of Figure 3a in ref 4). For ~I = 2 > ~Ic, the homo-geneous state remains stable throughout discharge (Figure 4d).

Figure 4. Numerical simulation of phase transformation via spinodaldecomposition. (a) Voltage responses at different constant currents.Concentration evolution of (b) spinodal decomposition at ~I = 0.01, (c)quasi-solid solution at~I = 0.25, and (d) solid solution at~I = 2. Labels onthe concentration curves indicate the mean filling fraction of the particle.

Figure 5. Numerical simulation of phase transformation triggered bywetting of the particle boundary. (a) Voltage responses at differentconstant currents. Concentration evolution of (b) waves at~I = 0.01, and(c) quasi-solid solution at ~I = 0.5.

Page 6: Suppression of Phase Separation in LiFePO Nanoparticles ... · LiFePO 4 battery.

4895 dx.doi.org/10.1021/nl202764f |Nano Lett. 2011, 11, 4890–4896

Nano Letters LETTER

It is noteworthy that the voltage is relatively flat for this case(Figure 4a), which could be misinterpreted as a sign of phaseseparation. The model also predicts a large overpotential at highfilling due to the excluded volume effect, which could be misinter-preted as concentration polarization (diffusion limitation).Experimental observations suggest that the lithiated phase

perfectly wets the inactive side facets of the particle,10 and so werepeat the simulation with wetting boundary conditions45 thatpermit heterogeneous nucleation on the particle surface (Figure 5).At a small current, ~I = 0.01 < ~Is, the wetted boundaries triggerphase separation close to the miscibility gap, and the particle fillsat constant voltage by two waves nucleated at the edge. For alarger current, ~I = 0.5 > ~Is, although the lithiated phase still wetsthe boundaries, the waves hardly propagate before the bulk fillslike a solid solution (Figure 5c). Interestingly, the voltage curvesshow no difference between spinodal decomposition and wettingif ~I > 1.Discussion.The key fitting parameter in our model is the local

exchange current density of LiFePO4 on the active (010) facet,but it is difficult to measure and varies by orders of magnitude,perhaps due to particle size distribution, agglomeration, and surfacechemistry. Recent experiments estimate 0.1 and 1.5 mA/cm2

over the geometrical area of the electrode,14,15 not the activesurface area of LiFePO4 particles as defined in eq 8. If the react-ing area is 100 times the geometrical area, then ~I = 2 iscomparable to 0.3C-rate discharge for the electrode in ref 14and 4.5C-rate for ref 15 (Sample A). Regardless of the precisecurrent values, an important qualitative feature to note is thetransition from discharge curves with a wide, flat voltage plateauat low current to smooth decay of voltage above the criticalcurrent,14,15 which our model suggests is due to the suppressionof phase separation. Similar behavior is seen in ultrafine, coatednano-LiFePO4 powders,

4 where the exchange current density is oforder 10 mA/cm2 according to Tafel analysis at low state of charge,making the critical current ~I = 2 comparable to 30C or evenhigher. Another important observation in this nano-LiFePO4

system is the fact that the 20C discharge curve overshoots andthen overlaps the 30C curve after a surprising voltage increase(negative differential capacitance) near the theoretically pre-dicted critical current. Our model suggests that this is a sign ofunstable quasi-solid solution behavior. A quantitative fit of thedata, however, would require modeling additional effects, suchas many-particle interactions, electrolyte diffusion, Ohmic losses,and elastic coherency strain.34

It is important to emphasize that there are nomeasurements ofsingle nanoparticle discharge to test our theory, although akinetic difference between phase separation in bulk and nano-LiFePO4 has recently been observed.

24 Existing data for compo-site porous electrodes is complicated by many particle interac-tions, which have been shown tomask complex phase transformationdynamics.46 In particular, nearly flat voltage plateaus at very lowcurrents that have long been attributed to single-particle phaseseparation are likely signs of discrete, one-by-one filling of manyparticles. For the same reason, the sudden jumps and non-monotonic voltage signatures in our single-particle simulationswould be difficult to observe in a macroscopic compositeelectrode. Moreover, an apparent voltage plateau can maskin situ homogeneous intercalation in discharging batteries, eventhough ex situ two-phase coexistence has been detected inelectrochemically lithiated particles.10 As shown in SupportingInformation Figure S3, when a large discharging current is turnedoff at X = 0.6 in our simulations, a homogeneous Li0.6FePO4

particle will relax to an equilibrium two-phase system of Li-richand Li-poor stripes, presumably before the sample can beobserved ex situ.Conclusion. In their seminal paper introducing LiFePO4

cathodes, Padhi, Nanjundaswamy, and Goodenough1 concludedthat “thematerial is very good for low-power applications” but “athigher current densities there is a reversible decrease in capacitythat... is associated with the movement of a two-phase interface”.Ironically, subsequent advances in carbon coating,47,48 sizereduction,49,50 cation doping,51 and optimized synthesis methods52

have made LiFePO4 the most popular cathode material for high-power applications. This incredible reversal of fortune has beenattributed to changes in material properties below the 100 nmscale, such as enhanced diffusivity8 and shrinking of the equilib-rium miscibility gap,53 but until now this has been difficult toreconcile with the well documented phase-separating behaviornear open circuit conditions.1,9�11

Our theory suggests that one key to the high rate capability ofnano-LiFePO4 is that phase separation is dynamically suppressedduring normal battery operation. Because of reaction limitationin nanoparticles, the surface overpotential easily exceeds thesolid�solution voltage barrier and thus removes the thermo-dynamic driving force for phase separation, once the currentbecomes comparable to the exchange current. This contradictsexisting models for LiFePO4, which assume artificial phase boun-daries and neglect phase separation dynamics. Only at very smallcurrents in large particles should phase separation play a majorrole, which seems consistent with the improved interpretation ofgalvanostatic intermittent titration data using an empirical modelwith a moving phase boundary.13

Our results have surprising implications for battery design andutilization. The theory predicts that increasing the interfacial resi-stance at the active surface tends to further suppress phase tran-sformation, thereby increasing the available active area for inter-calation. This leads to the counterintuitive conclusion that slowingthe surface reaction could be beneficial for battery performance,which may explain the improved discharge rate capability ofLiFePO4 nanoparticles with thin (5 nm) phosphate glass coatings.

4

Since phase separation causes mechanical deformation due tolattice-mismatch strain, its suppression would also reduce stres-ses in the crystal that can contribute to capacity fade, for example,by point defect formation8 or iron leaching.54 In other words,very slow discharge could actually reduce the cycle life of thebattery.Our arguments are quite general and are reinforced by includ-

ing more physics in the model, as suggested by ongoing work inour group. Elastic coherency strain further suppresses instabilityof the homogeneously filling state. Including size-dependentdiffusion helps to understand the transition to phase-separatingbehavior in larger particles, but it does not affect nanoparticlebehavior. Finally, including electrolyte depletion in a compositeelectrode allows us to make connections with experimental data,by extending porous electrode theory based on quasi-solidsolution nanoparticles.Beyond Li-ion batteries, our theory involves new concepts in

nonequilibrium thermodynamics, which may have other applica-tions. Phase transformations are typically studied in closed orperiodic bulk systems, close to equilibrium.18 Our theory de-scribes the suppression of phase transformations in a driven,open system, away from equilibrium. This leads to the notion ofan unstable quasi-solid solution, which exists only for certaincurrents as a nonequilibrium extension of the phase diagram

Page 7: Suppression of Phase Separation in LiFePO Nanoparticles ... · LiFePO 4 battery.

4896 dx.doi.org/10.1021/nl202764f |Nano Lett. 2011, 11, 4890–4896

Nano Letters LETTER

(Figure 3). Other examples of this general phenomenon mayinclude transient clustering in electrophoretic displays55 ordelayed shock-induced phase transformations in solids.56

’ASSOCIATED CONTENT

bS Supporting Information. Regular solution model, modi-fied Butler�Volmer equation, instability growth rate, 1D wavevelocity, relaxation behavior, andfive simulationmovies. Thismaterialis available free of charge via the Internet at http://pubs.acs.org.

’AUTHOR INFORMATION

Corresponding Author*Tel: (617)258-7039. Fax: (617)258-5766. E-mail: [email protected].

’ACKNOWLEDGMENT

This work was supported by the National Science Foundationunder Contracts DMS-0842504 and DMS-0948071 and by aseed grant from the MIT Energy Initiative. The authors thankTodd Ferguson for Tafel analysis of discharge data in Ref. 4. P.B.acknowledges the support of a Chinese Government Scholar-ship, File No 2009621147.

’REFERENCES

(1) Padhi, A. K.; Nanjundaswamy, K. S.; Goodenough, J. B.J. Electrochem. Soc. 1997, 144 (4), 1188–1194.(2) Armand,M.; Tarascon, J. M.Nature 2008, 451 (7179), 652–657.(3) Tarascon, J. M.; Armand, M.Nature 2001, 414 (6861), 359–367.(4) Kang, B.; Ceder, G. Nature 2009, 458 (7235), 190–193.(5) Wu, X. L.; Jiang, L. Y.; Cao, F. F.; Guo, Y. G.; Wan, L. J. Adv.

Mater. 2009, 21 (25�26), 2710.(6) Morgan, D.; Van der Ven, A.; Ceder, G. Electrochem. Solid-State

Lett. 2004, 7 (2), A30–A32.(7) Islam, M. S.; Driscoll, D. J.; Fisher, C. A. J.; Slater, P. R. Chem.

Mater. 2005, 17 (20), 5085–5092.(8) Malik, R.; Burch, D.; Bazant, M.; Ceder, G. Nano Lett. 2010, 10

(10), 4123–4127.(9) Chen, G. Y.; Song, X. Y.; Richardson, T. J.Electrochem. Solid-State

Lett. 2006, 9 (6), A295–A298.(10) Laffont, L.; Delacourt, C.; Gibot, P.; Wu, M. Y.; Kooyman, P.;

Masquelier, C.; Tarascon, J. M.Chem. Mater. 2006, 18 (23), 5520–5529.(11) Ramana, C. V.; Mauger, A.; Gendron, F.; Julien, C. M.; Zaghib,

K. J. Power Sources 2009, 187 (2), 555–564.(12) Srinivasan, V.; Newman, J. J. Electrochem. Soc. 2004, 151 (10),

A1517–A1529.(13) Zhu, Y. J.; Wang, C. S. J. Phys. Chem. C 2010, 114 (6), 2830–2841.(14) Pasquali, M.; Dell’Era, A.; Prosini, P. P. J. Solid State Electro-

chem. 2009, 13 (12), 1859–1865.(15) Wang, C. S.; Kasavajjula, U. S.; Arce, P. E. J. Phys. Chem. C 2007,

111 (44), 16656–16663.(16) Shenouda, A. Y.; Liu, H. K. J. Alloys Compd. 2009, 477 (1�2),

498–503.(17) Dargaville, S.; Farrell, T. W. J. Electrochem. Soc. 2010, 157 (7),

A830–A840.(18) Balluffi, R. W.; Allen, S. M.; Carter, W. C. Kinetics of materials;

Wiley-Interscience: Hoboken, NJ, 2005.(19) Nauman, E. B.; He, D.Q.Chem. Eng. Sci. 2001, 56 (6), 1999–2018.(20) Tang, M.; Huang, H. Y.; Meethong, N.; Kao, Y. H.; Carter,

W. C.; Chiang, Y. M. Chem. Mater. 2009, 21 (8), 1557–1571.(21) Leriche, J. B.; Hamelet, S.; Shu, J.; Morcrette, M.; Masquelier,

C.; Ouvrard, G.; Zerrouki, M.; Soudan, P.; Belin, S.; Elkaim, E.; Baudelet,F. J. Electrochem. Soc. 2010, 157 (5), A606–A610.(22) Kao, Y. H.; Tang, M.; Meethong, N.; Bai, J. M.; Carter, W. C.;

Chiang, Y. M. Chem. Mater. 2010, 22 (21), 5845–5855.

(23) Dupre, N.; Oliveri, J.; Degryse, J.; Martin, J. F.; Guyomard, D.Ionics 2008, 14 (3), 203–207.

(24) Tan, H. J.; Dodd, J. L.; Fultz, B. J. Phys. Chem. C 2009, 113 (48),20527–20530.

(25) Kuwahara, A.; Suzuki, S.; Miyayama, M. J. Electroceram. 2010,24 (2), 69–75.

(26) Singh, G. K.; Ceder, G.; Bazant, M. Z. Electrochim. Acta 2008,53 (26), 7599–7613.

(27) Delmas, C.;Maccario, M.; Croguennec, L.; Le Cras, F.;Weill, F.Nat. Mater. 2008, 7 (8), 665–671.

(28) Burch, D.; Bazant, M. Z. Nano Lett. 2009, 9 (11), 3795–3800.(29) Tang, M.; Belak, J. F.; Dorr, M. R. J. Phys. Chem. C 2011, 115

(11), 4922–4926.(30) Malik, R.; Zhou, F.; Ceder, G.Nat.Mater. 2011, 10 (8), 587–590.(31) Huggins, R. A. Advanced batteries: materials science aspects;

Springer: New York, 2009.(32) Cahn, J. W.; Hilliard, J. E. J. Chem. Phys. 1958, 28 (2), 258–267.(33) Han, B. C.; Van der Ven, A.; Morgan, D.; Ceder, G. Electrochim.

Acta 2004, 49 (26), 4691–4699.(34) Van der Ven, A.; Garikipati, K.; Kim, S.; Wagemaker, M.

J. Electrochem. Soc. 2009, 156 (11), A949–A957.(35) Yamada, A.; Koizumi, H.; Sonoyama, N.; Kanno, R. Electro-

chem. Solid-State Lett. 2005, 8 (8), A409–A413.(36) Cogswell, D. A.; Bazant, M. Z., arXiv:1110.3877v2 [cond-mat.

mtrl-sci].(37) Bockris, J. O. M.; Reddy, A. K. N.; Gamboa-Aldeco, M. E.

Modern Electrochemistry 2A: Fundamentals of Electrodics, 2nd ed.; KluwerAcademic/Plenum Publishers: New York, 2000.

(38) Aziz, M. J.; Sabin, P. C.; Lu, G. Q. Phys. Rev. B 1991, 44 (18),9812–9816.

(39) Bazant, M. Z.; Kilic, M. S.; Storey, B. D.; Ajdari, A. Adv. ColloidInterface Sci. 2009, 152 (1�2), 48–88.

(40) Guyer, J. E.; Boettinger, W. J.; Warren, J. A.; McFadden, G. B.Phys. Rev. E 2004, 69, 021603.

(41) Newman, J. S.; Thomas-Alyea, K. E. Electrochemical systems, 3rded.; J. Wiley: Hoboken, NJ, 2004.

(42) Meethong, N.; Huang, H. Y. S.; Speakman, S. A.; Carter, W. C.;Chiang, Y. M. Adv. Funct. Mater. 2007, 17 (7), 1115–1123.

(43) Allen, S. M.; Cahn, J. W. Acta Metall. 1979, 27 (6), 1085–1095.(44) Karma, A.; Rappel, W. J. Phys. Rev. E 1999, 60 (4), 3614–3625.(45) Granasy, L.; Pusztai, T.; Saylor, D.; Warren, J. A. Phys. Rev. Lett.

2007, 98 (3), 035703.(46) Dreyer, W.; Jamnik, J.; Guhlke, C.; Huth, R.; Moskon, J.;

Gaberscek, M. Nat. Mater. 2010, 9 (5), 448–53.(47) Ravet, N.; Goodenough, J. B.; Besner, S.; Simoneau, M.;

Hovington, P.; Armand, M. Improved Iron Based Cathode Material,196th Meeting of the Electrochemical Society, Abstract No. 127,Honolulu, Hawai, Oct. 17�22, 1999; The Electrochemical SocietyInc: Pennington, New Jersey.

(48) Ravet, N.; Chouinard, Y.; Magnan, J. F.; Besner, S.; Gauthier,M.; Armand, M. J. Power Sources 2001, 97�8, 503–507.

(49) Yamada, A.; Chung, S. C.; Hinokuma, K. J. Electrochem. Soc.2001, 148 (3), A224–A229.

(50) Huang, H.; Yin, S. C.; Nazar, L. F. Electrochem. Solid-State Lett.2001, 4 (10), A170–A172.

(51) Chung, S. Y.; Bloking, J. T.; Chiang, Y. M. Nat. Mater. 2002,1 (2), 123–128.

(52) Whittingham, M. S. Chem. Rev. 2004, 104 (10), 4271–4301.(53) Meethong, N.; Huang, H. Y. S.; Carter, W. C.; Chiang, Y. M.

Electrochem. Solid-State Lett. 2007, 10 (5), A134–A138.(54) Amine, K.; Liu, J.; Belharouak, I. Electrochem. Commun. 2005,

7 (7), 669–673.(55) Murau, P.; Singer, B. J. Appl. Phys. 1978, 49 (9), 4820–4829.(56) Branicio, P. S.; Kalia, R. K.; Nakano, A.; Vashishta, P. Phys. Rev.

Lett. 2006, 96 (6), 065502.(57) Bazant, M. Z. 10.626 Electrochemical Energy Systems (Mas-

sachusetts Institute of Technology: MIT Open-CourseWare, http://ocw.mit.edu, License: Creative Commons BY-NC-SA, 2011).


Recommended