+ All Categories
Home > Documents > Toward First Principles Prediction of Voltage Dependences of Electrolyte/Electrolyte Interfacial...

Toward First Principles Prediction of Voltage Dependences of Electrolyte/Electrolyte Interfacial...

Date post: 21-Dec-2016
Category:
Upload: craig-m
View: 212 times
Download: 0 times
Share this document with a friend
12
Toward First Principles Prediction of Voltage Dependences of Electrolyte/Electrolyte Interfacial Processes in Lithium Ion Batteries Kevin Leung* and Craig M. Tenney Sandia National Laboratories, MS 1415, Albuquerque, New Mexico 87185, United States * S Supporting Information ABSTRACT: In lithium ion batteries, Li + intercalation into electrodes is induced by applied voltages, which are in turn associated with free energy changes of Li + transfer (ΔG t ) between the solid and liquid phases. Using ab initio molecular dynamics (AIMD) and thermodynamic integration techniques, we compute ΔG t for the virtual transfer of a Li + from a LiC 6 anode slab, with pristine basal planes exposed, to liquid ethylene carbonate conned in a nanogap. The onset of delithiation, at ΔG t = 0, is found to occur on LiC 6 anodes with negatively charged basal surfaces. These negative surface charges are evidently needed to retain Li + inside the electrode and should aect passivation (SEI) lm formation processes. Fast electrolyte decomposition is observed at even larger electron surface densities. By assigning the experimentally known voltage (0.1 V vs Li + /Li metal) to the predicted delithiation onset, an absolute potential scale is obtained. This enables voltage calibrations in simulation cells used in AIMD studies and paves the way for future prediction of voltage dependences in interfacial processes in batteries. I. INTRODUCTION Imposing a potential dierence between electrodes is the among most basic operations in electrochemical experiments. In lithium ion batteries (LIB), critical processes such as Li + intercalation into anodes and cathodes, and electrochemical reductive/oxidative decomposition of organic-solvent based electrolytes, are governed by half-cell voltages. For example, widely used electrolytes based on a mixture of ethylene carbonate (EC), dimethyl carbonate (DMC, or similar linear carbonates), lithiun ions, and hexauorophosphide counterions (PF 6 ) start to decompose on the anode at 0.70.8 V relative to lithium metal foil reference (Li + /Li(s)), while Li + insertion into commerical graphite anodes occurs at much lower, 0.10.2 V, potentials. To prevent continuous loss of electrolyte and exfoliation of graphite during charging, anode passivation by self-limiting lms (called solidelectrolyte interphase or SEI lms) formed via electron-injection-induced electrolyte degra- dation is critical for LIB operations. 14 Proposed high-voltage cathode materials like LiMn 1.5 Ni 0.5 O 4 intercalate Li + above the experimentally observed stability voltage limit of EC/DMC/ LiPF 6 . Either new electrolytes need to be discovered, or the liquidsolid interfaces must be articially passivated to avoid electrolyte oxidation on these cathode surfaces. To understand and control degradation processes at atomic/electronic length scales, there is arguably an urgent need to use electronic structure computational tools (e.g., density functional theory, DFT) to calculate the voltage dependence of interfacial processes. However, DFT and quantum chemistry techniques deal with xed numbers of electrons (N e ), not xed voltages. The two properties are conjugate, like pressure and volume; specifying N e means that potentials are implicitly dened. In cluster calculations, where periodic boundary conditions are not used, 58 intrinsic redox potentials can be readily calculated at the expense of excluding electrodes in the models. Ab initio molecular dynamics-based redox calculations with explicit treatment of pure liquid environments have also been an area of fruitful study. 9 However, calculating potentials on electrodes in a DFT context and condensed-phase settings has long been recognized as a challenge in computational electrochemis- try. 1025 It is further complicated by subtle issues of whether Galvaniand Voltavoltages are well-dened, can be measured, or even computed. 12,26,27 Part of the challenge arises because, unlike classical force elds depiction of electrodes, e.g., using the polarizable SiepmannSprik model, 2834 one-electrode simulation cells are universally applied in DFT electrochemistry calculations (Figure 1). 1022 This is partly because accounting for both a cathode and an anode in a DFT simulation cell (Figure 1a) can easily lead to unphysical e migration between the electrodes. In lithium ion batteries, the salt concentration is typically 1.0 M, consistent with small a Debye screening length of 3 Å, so anodes and cathodes are indeed well-screened from each other and independent. But with only one electrode present, imposing a xed voltage would mean allowing N e to uctuate on the electrode. This implies permitting the net charge of periodically replicated simulation cells to change, which makes the total energy undened 3537 except in special cases or with specialized boundary conditions. 15,38 These voltage-specic diculties are compounded by other challenges associated with electronic structure modeling of solidliquid interfaces, 39 Received: September 7, 2013 Published: October 22, 2013 Article pubs.acs.org/JPCC © 2013 American Chemical Society 24224 dx.doi.org/10.1021/jp408974k | J. Phys. Chem. C 2013, 117, 2422424235
Transcript
Page 1: Toward First Principles Prediction of Voltage Dependences of Electrolyte/Electrolyte Interfacial Processes in Lithium Ion Batteries

Toward First Principles Prediction of Voltage Dependences ofElectrolyte/Electrolyte Interfacial Processes in Lithium Ion BatteriesKevin Leung* and Craig M. Tenney

Sandia National Laboratories, MS 1415, Albuquerque, New Mexico 87185, United States

*S Supporting Information

ABSTRACT: In lithium ion batteries, Li+ intercalation into electrodes is induced byapplied voltages, which are in turn associated with free energy changes of Li+ transfer (ΔGt)between the solid and liquid phases. Using ab initio molecular dynamics (AIMD) andthermodynamic integration techniques, we compute ΔGt for the virtual transfer of a Li+

from a LiC6 anode slab, with pristine basal planes exposed, to liquid ethylene carbonateconfined in a nanogap. The onset of delithiation, at ΔGt = 0, is found to occur on LiC6anodes with negatively charged basal surfaces. These negative surface charges are evidentlyneeded to retain Li+ inside the electrode and should affect passivation (“SEI”) filmformation processes. Fast electrolyte decomposition is observed at even larger electronsurface densities. By assigning the experimentally known voltage (0.1 V vs Li+/Li metal) tothe predicted delithiation onset, an absolute potential scale is obtained. This enables voltagecalibrations in simulation cells used in AIMD studies and paves the way for futureprediction of voltage dependences in interfacial processes in batteries.

I. INTRODUCTION

Imposing a potential difference between electrodes is theamong most basic operations in electrochemical experiments.In lithium ion batteries (LIB), critical processes such as Li+

intercalation into anodes and cathodes, and electrochemicalreductive/oxidative decomposition of organic-solvent basedelectrolytes, are governed by half-cell voltages. For example,widely used electrolytes based on a mixture of ethylenecarbonate (EC), dimethyl carbonate (DMC, or similar linearcarbonates), lithiun ions, and hexafluorophosphide counterions(PF6

−) start to decompose on the anode at 0.7−0.8 V relativeto lithium metal foil reference (Li+/Li(s)), while Li+ insertioninto commerical graphite anodes occurs at much lower, 0.1−0.2V, potentials. To prevent continuous loss of electrolyte andexfoliation of graphite during charging, anode passivation byself-limiting films (called solid−electrolyte interphase or SEIfilms) formed via electron-injection-induced electrolyte degra-dation is critical for LIB operations.1−4 Proposed high-voltagecathode materials like LiMn1.5Ni0.5O4 intercalate Li

+ above theexperimentally observed stability voltage limit of EC/DMC/LiPF6. Either new electrolytes need to be discovered, or theliquid−solid interfaces must be artificially passivated to avoidelectrolyte oxidation on these cathode surfaces. To understandand control degradation processes at atomic/electronic lengthscales, there is arguably an urgent need to use electronicstructure computational tools (e.g., density functional theory,DFT) to calculate the voltage dependence of interfacialprocesses.However, DFT and quantum chemistry techniques deal with

fixed numbers of electrons (Ne), not fixed voltages. The twoproperties are conjugate, like pressure and volume; specifyingNe means that potentials are implicitly defined. In cluster

calculations, where periodic boundary conditions are notused,5−8 intrinsic redox potentials can be readily calculated atthe expense of excluding electrodes in the models. Ab initiomolecular dynamics-based redox calculations with explicittreatment of pure liquid environments have also been an areaof fruitful study.9 However, calculating potentials on electrodesin a DFT context and condensed-phase settings has long beenrecognized as a challenge in computational electrochemis-try.10−25 It is further complicated by subtle issues of whether“Galvani” and “Volta” voltages are well-defined, can bemeasured, or even computed.12,26,27

Part of the challenge arises because, unlike classical forcefields depiction of electrodes, e.g., using the polarizableSiepmann−Sprik model,28−34 one-electrode simulation cellsare universally applied in DFT electrochemistry calculations(Figure 1).10−22 This is partly because accounting for both acathode and an anode in a DFT simulation cell (Figure 1a) caneasily lead to unphysical e− migration between the electrodes.In lithium ion batteries, the salt concentration is typically 1.0 M,consistent with small a Debye screening length of ∼3 Å, soanodes and cathodes are indeed well-screened from each otherand independent. But with only one electrode present,imposing a fixed voltage would mean allowing Ne to fluctuateon the electrode. This implies permitting the net charge ofperiodically replicated simulation cells to change, which makesthe total energy undefined35−37 except in special cases or withspecialized boundary conditions.15,38 These voltage-specificdifficulties are compounded by other challenges associatedwith electronic structure modeling of solid−liquid interfaces,39

Received: September 7, 2013Published: October 22, 2013

Article

pubs.acs.org/JPCC

© 2013 American Chemical Society 24224 dx.doi.org/10.1021/jp408974k | J. Phys. Chem. C 2013, 117, 24224−24235

Page 2: Toward First Principles Prediction of Voltage Dependences of Electrolyte/Electrolyte Interfacial Processes in Lithium Ion Batteries

such as the inherent difficulty of finding reasonable surfaceatomic, electronic, and magnetic structures40−42 and theincreased computational cost of modeling liquids at finitetemperature.39,43

In this work, we focus on the interface between liquid ECand fully lithiated graphite (stoichiometry LiC6)

44 in lithiumion batteries. When the coulomb efficiency of LIB is close to100% (e.g., with well-chosen voltage windows or passivatedelectrodes, so that the electrolyte is no longer beingdecomposed), the applied voltage should be governed by Li+

transfer between electrodes and the electrolyte, and e− shouldonly move along the external electrical circuit:

→ +−− +Li C Li C Li (solv)n n n n6 1 6 (1)

The free energy associated with Li+ transfer, denoted ΔGthenceforth, is relatively straightforward, if costly, to compute.When ΔGt is tuned to zero by adjusting the net surfaceelectronic density, LiC6 is at the onset of delithiationexperimentally known to occur at 0.1 V vs Li+/Li(s). In otherwords, the e− left behind by delithiation is at a Fermi level (EF)0.1 eV below that of Li(s) used as reference in batteries. Thisfixed point permits concrete comparison of predictions withmeasurements. Further discussions along these lines are givenin section II.ΔGt associated with Li+ transfer at liquid EC/solid LiC6

interfaces are conducted using thermodynamic integration (TI)techniques which closely follow AIMD solvation ΔGsolvcalculations.45 Our ΔGt calculations are operationally similarto some AIMD pKa simulations at water−oxide interfaces.46−48These methods fall under the umbrella of “chemical space” or

alchemical transformations.49 The Li+ transfer is virtual; no lowenergy, physical pathway exists for Li+ to diffuse from insideLiC6 solid, through the graphite (0001) plane, to the liquidregion.This paper is organized as follows. Section II provides further

justifications for our approach. Section III describes thethermodynamic integration method used in bulk liquid ECand at interfaces. Section IV discusses the computed voltages assurface charge densities on electrodes vary and compares thepredictions with changes in electrostatic potentials andinstantaneous Kohn−Sham band alignments. A discussion ofmethodology improvement is given in section V, and section VIconcludes the paper.

II. JUSTIFICATION OF APPROACHA more detailed rationale for our voltage assignment can bemade as follows. The casual reader is encouraged to skip tosection III.The implicit reference electrode is Li+/Li(s) (Figure 2a) at

its equilibrium potential:

→ ++−

−Li (s) Li (solv) Li (s)n n 1 (2)

When an Li+ is removed to the electrolyte, an e− is left behindat the metallic Li(s) Fermi level, just as an e− is left at EF of theworking LiC6 electrode (eq 1). If we had used a two-electrodesimulation cell like Figure 2a or 2b, the potential of the workingelectrode relative to the Li(s) reference can be obtainedwithout experimental input. The lithiated graphite potentialshould be +0.1 V vs Li(s), modulo DFT errors. However, wedo not actually model Li(s) electrodes and their complex, SEI-covered surfaces which exhibit unknown surface structures andcharge densities. Instead, when both Li(s) and LiC6 are at theirrespective ΔGt = 0.0 eV, EF in the two electrodes differs by 0.10V according to experiments, and eq 1 must be at 0.10 V relativeto eq 2. In the language of ref 12, eq 18, Li+ has the sameGalvani potentials inside the electrolyte, LiC6, and Li(s) underthese conditions, and the full-cell voltage is just the difference inμe of the electrodes divided by |e|.The schematic in Figure 2b further clarifies that the “Fermi

levels” of Figure 2a can be formally related to measurable work

Figure 1. (a) Schematic of a cathode, an anode, and a liquid electrolytein the same simulation cell. Classical force field electrode models canpredict the voltage drop between two polarizable electrodes, but so farwith no DFT methods. (b) Periodically replicated simulation cellsapplied in this work. They consist of alternating slabs of LiC6 andliquid EC, in most cases with Li+, “Liλ+,” and/or PFPF6

− ions solvatedwithin the liquid region. The solid and liquid regions are roughly 15and 19 Å thick, respectively. Gray, red, white, and blue spheres/linesdenote C, O, H, and Li atoms.

Figure 2. (a, b) Schematic of working (“test”) and reference electrodes(eqs 1 and 2) hypothetically connected by electrolytes or separated byvacuum, respectively. Li(s) is never explicitly included in calculations;only its potential difference with LiC6 is used (see text). (c) Charge-neutral LiC6/vacuum interfaces used to illustrate the computedlithiation potential of zero charge.

The Journal of Physical Chemistry C Article

dx.doi.org/10.1021/jp408974k | J. Phys. Chem. C 2013, 117, 24224−2423524225

Page 3: Toward First Principles Prediction of Voltage Dependences of Electrolyte/Electrolyte Interfacial Processes in Lithium Ion Batteries

function differences referenced to the vacuum.12 Here avacuum layer is heuristically inserted in middle of theelectrolyte. The surface potentials of the two vacuum/liquidinterfaces are clearly equal for the two electrolyte layerscovering the electrodes. They cancel each other and yield thesame potential difference between the electrodes as Figure 2a.In other words, Li(s) and LiC6 are in equilibrium with a liquidEC/Li+ electrolyte at the same absolute potential relative tovacuum, regardless of whether the vacuum layer exists. Asdiscussed below, it is computationally advantageous to avoidvacuum regions.By incorporating the experimentally known 0.1 V voltage

difference, an external “electrical circuit” (Figure 2a) isheuristically completed without directly computing e− orbitallevels in or the cohesive energy of Li(s). Ambiguities about themeasurability of half-cell potentials12,26,27 should be avoided.The assumption that EF on Li metal is at 1.37 V versus vacuumunder eq 2 conditions, which is analogous to the 4.44 Vstandard hydrogen electrode reference in aqueous systems, isnot used; this information is subsumed into the 0.1 V potentialdifference and LiC6 ΔGt calculations which incorporate Li+

solvation effects.By construction, the hidden Li(s) reference is at equilibrium

(ΔGt = 0). But it is a requirement in our scheme to search forΔGt = 0 with LiC6. ΔGt is a function of the surface electrondensity (σ) on LiC6. For the purpose of this computationalpaper, at each σ, we assign −ΔGt/|e| as the voltage of the anodeon short time scales, as if LiC6 were a capacitor or an “ideallypolarizable” electrode. This assumes that σ is adjusted on timescales fast compared to Li+ lithiation/delithiation which mayalter the voltage. Indeed, we have frozen all atoms in the LiC6electrodes in our simulations. By calculating ΔGt over a rangeof σ, the condition under which eq 1 is at equilibrium (ΔGt =0) is obtained. This aspect appears related to Rossmeisl et al.’scalculations of H+ at Pt(111)/H2O interfaces, with thedifference that there is no vacuum region in our work. Wealso define the “lithiation potential of zero charge” (LPZC) ofLiC6 as the −ΔGt/|e| value when σ = 0. LPZC is not necessarilymeasurable, but it will allow a check against an alternate,thermodynamic estimate that assumes liquid−solid interfaceeffects are minimal on charge-neutral electrodes (Figure 2c).In reality, at low voltages, anodes should be covered with SEI

films regardless of its material composition.1−4,50 Ourcalculations on pristine basal planes are meant to reflect thetransient period before SEI starts to form and covers the anodesurface; the predicted voltage dependence is precisely what isneeded to understand initial SEI formation processes. The LiC6stochiometry is chosen over unintercalated graphite (which ismore appropriate at higher potentials) because the onset ofLiC6 delithiation (0.1 V vs Li+/Li(s)) is much betterestablished44 than the onset of lithiation into graphite; thelatter depends on edge site chemistry. The relatively unreactivegraphite basal terminating surfaces are chosen to slow downparasitic reactions that may occur while ΔGt statistics are beingcollected. In the future, LiC6 edge planes, more pertinent tolithium intercalation, will be considered. At the same appliedvoltage, the surface charge densities of different crystal facetsdiffer, as do those of pristine electrodes and electrodes coveredwith SEI.A brief comparison between our approach/philosophy and

methods used in the aqueous electrochemistry (especially fuelcell) literature is given next. Many computational electro-chemistry applications are based on surface science methods.

Structural optimization is applied at zero temperature; thisextends to the configuration of electrolyte molecules, if present.A dielectric approximation is sometimes used to mimic thefinite-temperature liquid environment. For catalytic metalelectrode surfaces, the theoretical voltage dependence ofmolecular reactions is calibrated by a combination of Fermilevel and thermodynamic cycle estimates. While successful forfuel cell applications, zero temperature-based methods appearto lack the versatility to deal with ionic processes, such as Li+

intercalation from solvent into anodes and the parasiticreactions of its counterion, PF6

−,2,51 in batteries. Indeed, netcharges on electrode and ions in the electrolyte are seldomexplicitly included in T = 0 K calculations because ions areinsoluble (i.e., they “salt out”) in crystallized solvents andbecause simulation cells containing material/vacuum interfacemust be charge-neutral for their energies to be meaningful.24,52

Most previous assignments of Li potential on LIB electrodesurfaces, including our own, apply this imperfect surface scienceroute.43,53 We will show that LiC6 basal plane/liquid electrolyteinterfaces should in fact be negatively charged. Note thatcombining dielectric continuum approaches with condensed-phase applications of DFT has shown promise.16,20,58,59

However, in LIB, both solvents and salts providing thedielectric environment can participate in electrochemicalreactions. Using an all-atom DFT treatment permits uncon-strained simulations of their reaction mechanisms and ispreferred.Recently, metal surfaces covered by thin layers of water have

been simulated at finite temperature using AIMD meth-ods.13−16 Even though statistical uncertainties can besignificant,13 average work functions can be unambiguouslypredicted. The work function of an electrode covered with asufficiently thick liquid layer should yield electronic bandalignment needed to elucidate its potential relative to vacuum,defined as absolute zero energy.12 However, adding a vacuumlayer in LIB models significantly increases the simulation cellsize, particularly because symmetric slabs with liquid layers onboth electrode surfaces are desirable to mimic battery systems.Here, we are most interested in calculating the voltagedependence of condensed-phase simulation cells used tostudy electrolyte decomposition43,51,54,55 and Li+ intercalationdynamcis.56,57 Model systems with vacuum gaps are no longerthe same simulation cells used for studying the phenomenamentioned above. In battery experiments, vacuum or vaporregions do not exist unless an ionic liquid is used aselectrolyte.60 It appears ideal to avoid artificially openingvacuum gaps in simulations if an alternative connection toexperimental fixed points can be established.Finally, we compare our approach with rigorous, seminal

theoretical studies of band alignment at TiO2/H2O inter-faces.11,12 Sprik et al. reference their TiO2 electronic bands tothe standard hydrogen electrode (SHE) potential at the pH ofzero charge of TiO2.

+ → +− + •TiOH H (aq) TiOH H2

+ + → ++ − •Ti OH OH (aq) (1/2)H Ti OH H O(aq)2 2 2 2(3)

All processes associated with eq 3 are computed in water-filledsimulation cell containing the oxide slab. The pH of zero chargeof TiO2 has been measured, and properties computed at thePZC can be compared to experiments. In LIB, potentials ofzero charge, determined not by pH but by the net electron

The Journal of Physical Chemistry C Article

dx.doi.org/10.1021/jp408974k | J. Phys. Chem. C 2013, 117, 24224−2423524226

Page 4: Toward First Principles Prediction of Voltage Dependences of Electrolyte/Electrolyte Interfacial Processes in Lithium Ion Batteries

surface density (σ), are unknown. Furthermore, the lack of aliquid H2O environment inside LIB renders calculating eq 3 inLIB simulation cells meaningless. In LIB half-cell measure-ments, lithium metal foils are used as counter electrodes.Unfortunately, it is impractical to place a Li metal slab in ourLiC6 simulation cell and compute reference propertiesanalogous to eq 3. Not only is there substantial latticemismatch, but pristine Li(s) reacts violently with organicsolvents.54 Our approach is agnostic to the vacuum level or theSHE but uses the experimental LiC6 voltage at the onset of Li

+

delithiation as reference. Unlike TiO2, LiC6 is metallic;removing a Li+ from LiC6 leaves an excess e− on the Fermilevel, not in a localized state. For LIB cathode oxides that areelectronic insulators, our approach needs to be significantlymodified.

III. METHODA. Li+ Solvation Free Energy. The basic thermodynamic

integration formula used in this work is

∫ λ λΔ = ⟨ ⟩λ

λG Hd ( )/d(4)

where λ parametrizes the continuous creation or deletion of aLi+ ion and ⟨ ⟩λ denotes averaging with intermediateHamiltonian, 0 < λ < 1, sampled using molecular dynamicstrajectories.First we apply it to compute Li+ absolute solvation free

energies (ΔGsolv) in bulk EC liquid. Liquid EC simulation cellsare of dimensions (15.2442 Å)3 and are filled with 32 ECmolecules and a Liλ+ ion (Figure 3a). Liλ+ is represented by aVASP Li+ pseudopotential without core 1s electrons and with

all r-dependent parts scaled uniformally by λ. The success ofsuch a scaling has been previously demonstrated in anothersolvent.45 The integrands ⟨dH(λ)/dλ⟩λ are evaluated at either 2or 6 discrete λ points, each sampled every 0.1 ps interval of anAIMD trajectory with a Liλ+ pseudopotential. λ-derivatives arecomputed via finite difference with δλ = 0.025. Integrand valuesat λ ≠ 0 or 1 are not physically significant. They only serve toevaluate the integral in eq 4, which should be path independent.Other functional forms for λ can be used to scale thepseudopotentials and should give the same result.Since a periodically replicated simulation cell is used in

conjunction with Ewald summation of electrostatics, and cellsthat contain Liλ+ ions have a net +λ charge, several electrostaticcorrections are needed.35−38,61 The well-known monopolecorrection is αλ2/(2Lε0), where L is the box length and α is theMadelung constant. ε0 = 1 is imposed for gas phase calculationswhen evaluating the energies of bare Liλ+ which needs to besubtracted while ε0 = ∞ is assumed for the high dielectric ECliquid. The quadrupole correction27,35,45

∫ηπ ρΔ = − −E r r r2 /3 ( )( )r

quad 02

(5)

is computed using optimized gas phase EC geometry. The ECcarbonyl carbon, positioned at r0, is chosen as molecular center;η and ρ are the EC density and the total (electronic plusnuclear) charge density, respectively. Finally, the dipolarcontribution to the surface potential, computed using thecarbonyl carbon as molecular center for consistency,27 isestimated using classical force fields; see the SupportingInformation for details. The corrections arise from the long-range nature of electrostatics; all higher multiple contributionsvanish.27,35,36

B. Free Energy of Li+ Transfer from LiC6 to Liquid EC.There is neither sufficient static/dynamic symmetry norreasonable physical boundaries to evaluate the quadrupolemoment correction (eq 5) in simulation cells containing anelectrode (Figure 1b).36 Consequently, the energies of chargedsimulation cells require corrections that cannot be readilyevaluated. In this work, only interfacial cells with constant Neand overall charge neutrality are considered. We freeze allatoms of a “LiC6” slab with a C288Li36 stoichiometry (3 layers ofLi intercalated between 4 graphite sheet with basal planesexposed, Figure 1b), select one Li in the middle Li slab, andscale all r-dependent parts of its pseudopotential by (1 − λ). Atthe same time, a Liλ+ pseudopotential is created at a fixedposition in the liquid electrolyte, halfway between electrodesurfaces. Bare Liλ+ contributions cancel, and their energies dono need to computed. Equation 4 is approximated with a 2-point Gaussian quadrature formula, which is justified andcorrected in section IVA. The simulation cells have dimensions34.00 × 12.96 × 14.96 Å3. 32 EC molecules, 0 to 4 mobile Li+,and in some cases a Liλ+ and/or 2 Li+PF6

− ion pairs fill the gapbetween electrode surfaces. The presence of mobile Li+ mimicsexperimental conditions but creates hurdles for AIMDtrajectories which are typically short compared to Li+ diffusiontime scale. Despite this, we have found reasonable agreementamong predicted ΔGt when initial Li+ configurations are varied.We have only considered removing Li+ from the middle of

the LiC6 slab because these are proof-of-principle “virtual”calculations aimed at elucidating the thermodynamics of Li+

transfer. In reality, Li+ escapes LiC6 through graphite edge, andLi+ intercalation potentials at those edge sites often function-alized with oxygen groups will differ from those inside LiC6.

Figure 3. (a) A snapshot of a Liλ+ ion in a bulk-like simulation cell. (b)Dipolar contribution to the surface potential, computed using classicalforce fields at T = 350 K (red line) and T = 450 K (green).

The Journal of Physical Chemistry C Article

dx.doi.org/10.1021/jp408974k | J. Phys. Chem. C 2013, 117, 24224−2423524227

Page 5: Toward First Principles Prediction of Voltage Dependences of Electrolyte/Electrolyte Interfacial Processes in Lithium Ion Batteries

This will be addressed in future publications. ΔGt variations asa function of Li+ position inside the electrolyte is expected to besmall and is discussed at the end of section IVE.The Li(1−λ)+ and Liλ+ ions being “transferred” are frozen in

space. Hence, the TI procedure omits vibrational andtranslational entropies of these ions. Assuming each Li+ isindependent inside LiC6, the Hessian matrix elements Kij =d2Etotal(x)/dxi dxj = 3.04, 0.90, and 0.92 eV/Å2 for i = j and arefound to be zero otherwise, where Etotal(x) is the total energy ofthe LiC6 solid and xi are the Cartesian coordinates of the taggedLi+. The Li+ translational/volumetric entropy corresponding toa 1.0 M salt solution (∝ [−kBT ln(1660 Å3)]) and itsvibrational free energy (∝ [−kBT∑i ln(2πkBT/Kii)

1/2]) areadded to and subtracted from ΔGt, respectively. The correctionamounts to −0.22 eV.C. AIMD Details. AIMD calculations are conducted using

the Vienna Atomic Simulation Package (VASP) version 4.663,64

and the PBE functional.65 A 400 eV planewave energy cutoff, Γ-point Brillouin zone sampling, and a 10−6 eV convergencecriterion are applied at each Born−Oppenheimer time step, 1 fsin duration. The k-space sampling is spot-checked using adenser 1 × 2 × 2 grid. The trajectories are kept at an averagetemperature of T = 450 K using Nose thermostats. Theelevated temperature reflects the need to “melt” EC, which hasan experimental freezing point above room temperature, and toimprove sampling efficiency.62 In real batteries DMC cosolventmolecules reduce the viscosity, but DMC is not includedherein. Minor differences in ΔGt that may arise from the use ofa mixed solvent in real batteries are neglected in this work.Tritium masses on EC are substituted for protons. The first 1ps of each AIMD trajectory is discarded, and the rest is used forsampling eq 4. The different AIMD simulations and trajectorylengths are described in Table 2.AIMD trajectories are initialized from configurations pre-

equilibrated using simple, rigid-body classical molecular forcefields,43,66 Monte Carlo (MC) simulations, and the Towheecode.67 At least 40 000 MC passes at T = 1000, 700, and then400 K are successively conducted, and the final configuration isused for AIMD simulations. In that sense, the electrical doublelayer should be well-equilibrated to the extent that the simpleclassical force field used is accurate. When PF6

− and excessmobile Li+ are both present in the electrolyte, the MCsimulation procedure yields Li+/PF6

−, but not Li(λ+)/PFPF6−,

contact ion pairs.We have also applied flexible classical molecular force fields

to perform molecular dynamics so as to estimate the dipolecontribution to the surface potential of pure liquid EC.68 Thesecalculations enable the prediction of absolute Li+ ΔGsolv,defined as the free energy change of moving an Li+ fromvacuum through the liquid−vacuum interface into EC liquid(see the Supporting Information).

IV. RESULTSA. Li+ Solvation in EC Liquid. Although not the main

purpose of this work, Li+ ΔGsolv calculations illustrate thenontrivial effect of interfaces and highlight important electro-static considerations. Table 1 lists ΔGsolv predictions for Li

+ inEC liquid at T = 450 K. Our predicted ΔGsolv is larger inmagnitude than that computed using a bare Li+ plus a dielectriccontinuum approximation.69 It is smaller than ΔGsolv predictedin acetonitrile solvent, reported without surface potentialcorrections.57,70 The 2- and 6-point formulas for Li+ solvationdiffer by 0.15 eV (see the Supporting Information), which is

within twice the standard deviation but larger than thediscrepancy obtained in H2O solvent.45 In the rest of thiswork, a global −0.15 eV correction is applied to all ΔGtcomputed using the 2-point formula because ΔGt also involvesLi+ solvation effects. This correction does not affect the relativeΔGt as the voltage varies.We have also considered the energy (EC;λ) of a Liλ+

embedded in bulk LiC6 solid while all atoms are frozen andNe is held fixed. The Supporting Information shows that a low-order integration suffices for EC;λ. These findings are used tojustify the 2-point formula for eq 4 when we simultaneouslyannihilate a Li+ inside LiC6 and create a Li+ inside EC liquid.The dipole contribution to the EC liquid−vapor surface

potential is depicted in Figure 3b. This quantity depends on thechoice of the molecular center; only the sum of the dipolar andquadrupolar contribution is physical.27 Nevertheless, Figure 3bserves to illustrate that the surface potential between a pureliquid and the vacuum is in general on the order of a fraction ofa volt.72 When salt is present, some ions may be repelled fromthe liquid surface while others may be attracted there, settingup further, nontrivial electric fields.73 Thus, if we had openedup an artificial gap in the electrolyte region to estimate bandalignment relative to absolute zero energy, we would haveintroduced two addition interfaces where distributions of ionsare additional sources of concern. In our interfacial simulations,vacuum gaps are avoided, and no classical force-field-generatedcontributions are included in ΔGt.

B. ΔGt of Li+ Transfer from LiC6 to EC Liquid. Figure 4

depicts ΔGt associated with Li+ transfer from LiC6 to the ECliquid with 0, 1, and 2 excess e− in the LiC6 anode compensatedwith the same number of mobile Li+ ions in the electrolyte. It iscompiled using trajectories A−J (Table 2) and constitutes themain result of the paper.The x-axis denotes surface e− density. In the Supporting

Information, we show, as expected from classical electrostatics,that the excess e− density on the electrode is fairly evenlydistributed on the two surfaces, although instantaneously thetwo surfaces can exhibit variations in charges; the x-axis reflectsan average over them. Henceforth, we report a uniform e−

surface density LiC6 surfaces, σ = −QLi/(2A), where A = 194 Å2

is the lateral surface area of the simulation cell. After a Li+ istransferred from LiC6 to liquid EC, an extra e− resides on theanode compared to before Li+ transfer. In our finite simulationcell, the resulting change in σ is not negligable. Hence, we havemarked the three ΔGt’s at halfway points, at σ values consistentwith −0.5|e|, −1.5|e|, and −2.5|e| excess e− in LiC6. Onmacroscopic electrodes, adding one extra e− does not affect σ.By placing our data point at halfway marks, we are effectivelyextrapolating toward this infinite size limit. The y-axis depicts−ΔGt/|e| in units of volts. y = 0.0 V represents the point where

Table 1. Monopole, Dipole (Dipolar Surface Potential), andQuadrupole Correction to ΔGsolv, and the Final Results, forLi+ Solvation in EC Liquid (in eV)a

Emono Edip Equad 2-pt 6-pt

2.042 +0.300 +5.389 −5.054 ± 0.04 −5.201 ± 0.04aThe 6 λ integration points are at 1.0, 0.789. 0.6, 0.4, 0.211, and 0.05(the last approximates λ = 0, which cannot be easily computed).Equation 4 is evaluated using a trapezoidal rule with these points. Forthe 2-point (Gaussian) quadrature, a 0.5 weight is applied for λ =0.211 and 0.789. Statistics for each λ-point are compiled over AIMDtrajectories of at least 24 ps.

The Journal of Physical Chemistry C Article

dx.doi.org/10.1021/jp408974k | J. Phys. Chem. C 2013, 117, 24224−2423524228

Page 6: Toward First Principles Prediction of Voltage Dependences of Electrolyte/Electrolyte Interfacial Processes in Lithium Ion Batteries

Li+ is equally favored inside or outside LiC6 anode. Thus, ΔGt =0 denotes the onset of delithiation. Experimentally, this isknown to occur at 0.1 V versus Li+/Li(s). This reference pointallows us to assign an absolute voltage scale. Above the greenline, LiC6 is thermodynamically unstable. This emphasizes thatanode surfaces, or at least pristine LiC6 basal planes, need to benegatively charged to retain Li+.

Despite the statistical noise, the three points approximatelylie on a straight line, with δV/δQe ≈ 0.44 V/|e|. Linearity isreasonable over such a small voltage window but is not essentialto our analysis. If we treat LiC6 as a capacitor and factor in theelectrode surface area, δσ/δ(potential) = 8.3 μC/(cm2 V). Thisvalue is larger than the predicted capacitances of unlithiatedcarbon nanotube arrays in propylene carbonate31 and porouscarbon electrode models in ionic liquids30 and smaller than thecapacitances of metallic electrodes.20

C. Statistical Uncertainties and Other Spot Checks.⟨dH(λ)/dλ⟩λ values are tabulated in Table 2, which also listsstatistical uncertainties and illustrates the dependence on initialconfigurations. We have run two trajectories each for λ = 0.211and λ = 0.789 with 0 and 2 excess e− on the anode. In all cases,⟨dH(λ)/dλ⟩λ with λ = 0.789 exhibits smaller numericaluncertainties and dependence on initial conditions comparedto integrands evaluated at λ = 0.211, even when the AIMDsampling trajectory is longer for the smaller λ. We have furtherconsidered a third set with two excess e− on the anode plus fourLi+ and two PF6

− mobile ions in the electrolyte region(trajectories K and L). During equilibration with force fields,each PF6

− forms a contact ion pair (CIP) with a Li+. Such ion-pairing has been predicted to occur with considerableprobability using polarizable classical force fields.71 As mightbe expected, the charge-neutral CIPs do not appreciably affect⟨dH(λ)/dλ⟩λ (trajectories K and L).In the Supporting Information, snapshots of Li+ config-

uration as well as their distributions as functions of x aredepicted. They clearly show that Li+ spatial configurations arenot completely converged within 15−35 ps AIMD trajectories.Despite this, the configurations of large-dipole-moment ECmolecules appear to compensate for differences in Li+ positionsand make λ = 0.789 integrand values relatively consistent (e.g.,trajectories H and J). On the other hand, even λ = 0.211integrands in simulation cells without mobile Li+ (trajectories Aand C) exhibit a considerable dependence on initialconfigurations. In contrast, Li+ ΔGsolv calculations conductedin the absence of the electrode do not exhibit larger uncertaintyat small λ. This enhanced sensitivity at small λ is as yet notcompletely understood. In Figure 4, we have averaged theresults from the two sets of initial conditions and have reportedthe error bar as the difference between these two runs. Theexception is the one mobile Li+ simulation (trajectories E andF); with only one set of data, we have reported the uncertaintyby assuming Gaussian distributions of noise in λ-windows. As σbecomes more negative, EC orients themselves so that theirpartially positively charged C2H4 termini start to align and pointtoward the LiC6 surface (not shown). Detailed studies ofcharge-dependent orientational effects are more suited toclassical force field methods71 than AIMD.The effect of k-point sampling in the lateral (y and z)

directions has been spot-checked as follows. Trajectories G andH are sampled every 1 ps. These configurations are used tocalculate ΔGt with both 1 × 2 × 2 and Γ-point Mohhorst−Packgrids. The resulting ΔGt computed differ by only −0.04 eV. Analmost identical small discrepancy of −0.04 eV is found fortrajectories A and B when using the two k-point grid sizes. Wehave not added this small correction to the final results (Figure4 and Table 2).

D. Lithiation Potential of Zero Charge. Extrapolating theresults in Figure 4 to σ = 0 yields the lithiation potential of zerocharge (LPZC) for LiC6 basal planes. It is predicted to occur at1.14 V vs Li+/LiC6 (1.24 V vs Li+/Li(s)). At this potential, Li+

Figure 4. Predicted potential (−ΔGt/|e|) for virtual Li+ transfer from

the LiC6 slab to the middle of the liquid EC region as the surfacecharge (σ) varies. Crosses denote the three data points computed, with0, 1, and 2 mobile Li+, respectively. AIMD simulations with 4 mobileLi+ and no counterions lead to EC decomposition.

Table 2. Details of AIMD Trajectoriesa

N(Li) N(PF6−) λ T ⟨dH(λ)/dλ⟩λ ΔGt

A 0 0 0.211 35.7 +5.32 ± 0.19B 0 0 0.789 16.7 −6.57 ± 0.11 −1.00C 0 0 0.211 30.1 +5.52 ± 0.10D 0 0 0.789 17.9 −6.54 ± 0.10 −0.88E 1 0 0.211 18.0 +6.06 ± 0.14F 1 0 0.789 17.9 −6.21 ± 0.05 −0.45G 2 0 0.211 15.1 +6.23 ± 0.11H 2 0 0.789 18.4 −5.65 ± 0.11 −0.08I 2 0 0.211 18.0 +6.45 ± 0.07J 2 0 0.789 16.8 −5.73 ± 0.11 −0.04K 4 2 0.211 15.6 +6.58 ± 0.12L 4 2 0.789 16.0 −5.82 ± 0.11 +0.01M 4 0 0.211 3.9 NAN 4 0 0.789 3.9 NAO 0 0 NA 16.4 NAP 2 0 NA 15.0 NAQ 4 0 NA 9.2 NA

aN(Li) and N(PF6−) are respectively the number of mobile Li+ and

PFPF6− ions in the liquid region of the simulation, and λ is the net

charge of the frozen Liλ+ ion if one exists. The reported trajectorydurations (T in picoseconds) include the first 1 ps equilibration timediscarded when collecting statistics. Integrands and ΔGt are in eV; thelatter includes a −0.22 eV entropic correction and a −0.15 eVcorrection for using a 2-point treatment of Li+ solvation (see text).

The Journal of Physical Chemistry C Article

dx.doi.org/10.1021/jp408974k | J. Phys. Chem. C 2013, 117, 24224−2423524229

Page 7: Toward First Principles Prediction of Voltage Dependences of Electrolyte/Electrolyte Interfacial Processes in Lithium Ion Batteries

should diffuse out of LiC6 into the electrolyte. Thus, LPZC forLiC6 basal planes cannot be measured. Delithiated graphiteexhibits a potential of zero charge which has no relation to theLiC6 LPZC.

74

A back-of-the-envelope calculation75 supports the existenceof a positive LPZC. If interfacial effects and excess negativecharge at basal plane surfaces were absent, the contributions toΔGt can be estimated via the following thermodynamicpathway. (1) Remove an Li atom from bulk LiC6; (2) Li(g)→ Li+(g) + e−; (3) put the ionized e− back into LiC6 (reverseof the work function, see the Supporting Information); (4)Li+(g) → Li+(EC). The energies of these processes arecomputed using DFT/PBE and listed in Table 3. They add

to a −1.84 eV exothermicity for removing Li+ to the EC liquidor about +1.94 V versus Li+/Li(s). An also identical value of2.00 V is obtained by focusing only on the energy of e− at theFermi level, i.e., the LiC6 work function, and subtracting thestandard 1.37 V. This is because the experimental equivalents ofTable 3 are present when deriving the 1.37 V Li(s) reference.In Figure 4, −ΔGt/|e| is 1.24 V vs Li+/Li(s) at σ = 0. This

estimate, and the 1.94 V discussed above, are both large andpositive. Their difference must be due to the neglect of solid−liquid interface effects in the latter, known to reduce the workfunction of charge-neutral water-covered metal surfaces by ∼1eV.12,77 Other factors and systematic/statistical errors may alsocontribute to the difference. Battery techologists focus onvoltage variations and arguably do not have a pressing need tomeasure σ. However, it is crucial for theorists to impose thecorrect explicit surface charge in DFT calculations to representrealistic, experimental potentials. This consideration hasarguably been neglected in most AIMD interfacial calculations(see the discussions in ref 39), although the previous worksreveal important electrolyte decomposition mechanisms whichshould be relevant over large voltage windows.E. Electrostatic/Exchange-Correlation Potentials as σ

Varies. Next, we analyze potential differences in cases where 0,2, and 4 excess e− reside on the LiC6 slab (trajectories O, P, andQ) using a electrostatic analysis complementary to calculatingion transfer free energies. The simulation cells considered (O−P in Table 2) do not contain λ-scaled lithium ions and areindependent of Li+-transfer ΔGt simulations. Figure 5 depictsthe average potential (V(x)) which includes electrostatic andDFT exchange-correlation contributions, sampled every 1.0 ps.The differences between V(x) in the electrolyte and LiC6regions, bracketed by the green lines in the figure, are ΔV =5.80, 4.75, and 4.15 V respectively for 0, 2, and 4 excess e−.More excess e− on the anode translates into a higher V(x) there(less favorable for electrons to reside in). The 1.05 V differencein ΔV between the first two runs is reasonably similar to the0.88 V difference observed in Figure 4. The two values are not

expected to be identical because Figure 5 averages almost theentire electrolyte region, not at one value of x. However,absolute voltages cannot be estimated from electrostaticpotential differences.From the similarity in the σ-dependence of ΔGt and ΔV, it

may be argued that ΔGt only needs to be computed at one σ;the σ−potential relation can then be determined using δΔV/δσ. This intriguing alternate strategy comes with the followingcaveat. The ΔV prediction for 4 excess e− on the anode (Figure5) deviates from the linear relation of Figure 4. It should betaken with a grain of salt because Li+ ions adsorbed directly onthe basal planes (Figure 7) are excluded from the regionarbitarily chosen for ΔV sampling (green lines in Figure 5).Thus, ΔV may depend on how many mobile Li+ ions areincluded in the averaging procedure. The electrolyte is alsoexperiencing decomposition, and this trajectory has to beprematurely terminated (see section IVG).The statistical uncertainties depicted in Figure 4 can in

principle be measured as time-dependent properties. Thenumerical uncertainties associated with Figure 5, which areabout 0.15 V, are however unphysical unless a nanosize probe isused. Assuming the system size is well-converged andfluctuations in each copy of the simulation cell are statisticallyindependent, doubling the lateral dimensions of the cell (i.e.,increasing A by a factor of 4) should yield a standard deviationhalf as large. Such cell size-dependence fluctuations should betrue of previous calculations of work functions of water-coveredmetal surfaces as well.13,14

V(x) is rather structured in the nanoconfined electrolyteregion, reflecting solvent layering which is observed even inclassical force field simulations (Figure 2b). We have conductedshort AIMD test runs to show that such layering has littleimpact on Li+ distributions. First, a Li+ is frozen at either x =23.0 Å or x = 24.7 Å, respectively, a valley and peak in the V(x)curve. The rest of the electrolyte around the fixed Li+ is pre-equilibrated with classical force fields. Then AIMD is initiated,with the tagged Li+ and all electrolyte molecules allowed to

Table 3. Contributions to ΔGt for Li+ Transfer from LiC6 or

Li(s) to Liquid EC If Liquid−Solid Interfacial Effects WereAbsent, in eVa

anode Li in anode IP (Li gas) −Φ (anode) Li+ ΔGsolv net

LiC6 +1.65 +5.30 −3.37 −5.20 −1.84Li(s) +1.56 +5.30 −3.05 −5.20 −1.56

aThey are respectively the binding energy of an Li atom in LiC6 orLi(s), gas phase Li ionization potential, the negative of LiC6 or Li(s)work function, and Li+ solvation free energy in EC liquid, all computedusing DFT/PBE. A −0.22 eV entropy correction is added to “net”.

Figure 5. Average electrostatic-plus-exchange/correlation potential(V(x)) with 0 (black line), 2 (red), and 4 (blue) excess e− on LiC6.High V(x) regions repel e−.

The Journal of Physical Chemistry C Article

dx.doi.org/10.1021/jp408974k | J. Phys. Chem. C 2013, 117, 24224−2423524230

Page 8: Toward First Principles Prediction of Voltage Dependences of Electrolyte/Electrolyte Interfacial Processes in Lithium Ion Batteries

move. In each case, the tagged Li+ is found to fluctuate in spaceover the course of a few picoseconds but does not exhibitsignificant net displacement in the x direction. The one initiallyat x = 23 Å does not fall into potential wells in V(x) computedin the absence of the tagged Li+. This underscores the fact thatLi+−solvent interactions are far stronger than the solvent−solvent interactions which determine V(x). This observationdovetails with classical force field predictions that the freeenergy profile (i.e., “potential of mean force”) of Li+

displacement toward the electrode is much less structuredthan the mean electrostatic potential; it suggests that using Li+

transfer to calculate electrode voltages may lead to fasterconvergence with respect to the thickness of the liquidelectrolyte layer.F. Kohn−Sham Band Structure as σ Varies. Next we

correlate ΔGt with shifts in the bottom of the conduction bandin the electrolyte region as σ varies. Figure 6 depicts Kohn−

Sham band structures in a single snapshot toward the end oftrajectories O, P, and Q, with 0, 2, and 4 excess e− in LiC6,respectively. Here the charge density on each Kohn−Shamorbital is split among atoms i at positions xi = x; if ρi ≥ 0.005,the histogram at x is incremented. The arbitrary cutoff meansthat some orbitals in the electrode are omitted; the density ofstate is somewhat higher than shown in Figure 6. Referenced toEF, always set at 0.0 eV, the electrolyte conduction bandminimum is shifted downward by roughly 1 eV with eachsuccessive injection of 2 excess e− into the LiC6 slab. Themagnitude of the shift is consistent with both ΔGt predictions(Figure 4, 0.88 V for every two e− added) and electrostaticpotential estimates (Figure 5, 1.05 V per two e−), althoughthese shifts may vary somewhat from one snapshot to the next.Figure 6c is particularly interesting because the snapshot is

taken less than 1 ps prior to an EC absorbing two e− and

decomposing. The bottom of the electrolyte conduction bandis almost at the Fermi level, enabling rapid e− transfer to anddecomposition of the electrolyte.

G. Electrolyte Decomposition at Low Voltages. AIMDsimulations with >1 excess e− in the model anode aremetastable. According to Figure 4, their potentials are belowthe experimentally known EC reduction potential (∼0.7 V vsLi+/Li(s)). With 2 excess e−, −ΔGt/|e| is close to the Li(s)plating potential (∼0.0 V vs Li+/Li(s)), which is another sourceof battery safety concern. Despite this, electrolyte decom-position is not yet observed in those short AIMD trajectoriesbecause the basal plane is relatively inert. However, with at least4 excess e− in the LiC6 slab, a EC molecule absorbs two e− anddecomposes within picoseconds. Figure 7 shows snapshots atthe end of trajectories Q and N. In each case, a EC moleculebreaks an oxygen−ethylene carbon (O−CE) bond, not anoxygen−carbonyl carbon (O−CC) bond. Quantum chemistrycalculations on EC2− predict that the latter barrier is smaller.76

However, on a material surface, we have found that the (O−CE) cleavage barrier can be reduced.54 It is also possible thatthe extremely low effective potential associated with thesetrajectories has changed the dominant decomposition mecha-nism. Given the limited bond-breaking statistics available inthese trajectories, the potential dependence of bond-breakingmechanism cannot be completely resolved but is an importantconsideration for future studies.The very fast, adiabatic e− transfer to EC molecules

apparently occurs via fluctuation-induced band-crossing. Asdiscussed in section IVF, the bottom of the conduction band inthe electrolyte region almost coincides with the Fermi levelresiding in the electron-conducting anode (Figure 6c). e− canthus pour into the liquid region. This seems to explain why inFigure 7a even an EC in the middle of the electrolyte regioncan accept two e− and decompose. Electron motion of thisnature may be sensitive to details of DFT functionals (e.g.,accuracy of predicted conduction band energy levels).

Figure 6. Instantaneous Kohn−Sham orbital alignment. The orbitaldensity is decomposed on to atoms at their x-coordinates for (a) 0, (b)2, and (c) 4 excess e− on LiC6. x < 15 Å denotes the LiC6 region;outside that range resides the electrolyte. Fermi levels are at E = 0.00eV.

Figure 7. Two snapshot of trajectories with 4 mobile Li+: (a)trajectory Q, after 9.2 ps; (b) trajectory N, after 4 ps (with an extrafrozen Li0.8+ ion present). Decomposed/intact EC molecules aredepicted as ball-and-stick/stick figures, respectively.

The Journal of Physical Chemistry C Article

dx.doi.org/10.1021/jp408974k | J. Phys. Chem. C 2013, 117, 24224−2423524231

Page 9: Toward First Principles Prediction of Voltage Dependences of Electrolyte/Electrolyte Interfacial Processes in Lithium Ion Batteries

Fortunately, such negative potentials vs Li+/Li(s) are notrelevant to battery operations, where voltage control isexercised to prevent overcharging. Under normal conditions,the electrolyte conduction band mininum resides above theanode Fermi level (Figure 6a or 6b). Fluctuations in ECgeometries are then required to lower EC lowest unoccupiedmolecular orbital levels and permit e− transfer to EC. Suchgeometric fluctuations help surmount e− transfer barriersassociated with reorganization free energies enunciated inMarcus theory.54

We have not observed Li(s) plating on LiC6 surfaces even atthe lowest voltages because Li+ diffusion and nucleation to formLi(s) clusters occur on long time scales.H. EC/Li(s) Interface: Preliminary Investigations. It

must be stressed that other interfaces may exhibit very differentbehavior. For example, EC molecules directly coordinated tomore reactive electrode surfaces may more readily participate ine− transfer due to strong electronic coupling. We havereanalyzed the initial configuration of an AIMD simulation ofEC/Li(s) (100) interfaces reported in ref 54, in the absence ofexcess electrons (i.e., at instantaneous lithiation potential ofzero charge, Figure 8). The electrolyte conduction band

minimum lies at least 2.0 eV above the Li metal Fermi level(not shown in Figure 8a). However, there is significanthybridization between Fermi level metallic states on theelectrodes and EC orbitals spatially located near metal surfaces.When an AIMD trajectory is launched from this configuration,EC molecules on the surface rapidly absorb e− and decomposeto form mostly CO gas.54 Using a hybrid DFT functional doesnot change this picture. In contrast, on LiC6 basal planes, theKohn−Sham conduction bands in the electrolyte region exhibitnegligable coupling with the anode near the Fermi level unlessσ is large and negative (Figure 7).Explicit ΔGt calculations are unlikely to succeed for pristine

Li(s) anodes because their extreme reactivity causes rapid

solvent decomposition and precludes sufficient AIMD sam-pling. Instead, in Table 3, we have estimated the LPZC bycalculating contributions to the free energy of Li+ transfer fromLi(s) to liquid EC if interfacial effects were absent and thesurface were uncharged. Analogous to LiC6, LPZC is large andpositive under these assumptions. Yet, as mentioned above, ECrapidly decomposes on the charge-neutral Li metal surface,contrary to EC on LiC6 basal planes where reductivedecomposition is observed only at large negative potentials.This comparison emphasizes that rates of e− transfer andelectrolyte decomposition strongly depend on surface details.

V. DISCUSSIONSo far we have performed voltage calculations on one type ofelectrode surface. At the same applied voltage, different anodecrystal facets should exhibit variations in surface chargedensities. On carbon surfaces containing edge sitese.g.,pockets framed by CO functional groups where Li+ can bestrongly trapped43the present formalism can determine theinstantaneous voltage with a fixed number of Li+ trapped atedge sites if we virtually transfer a Li+ from the anode interiorto the solvent. At longer time scales such that Li+ can desorbfrom surface sites, the potential becomes a function of bothexcess e− surface density and the amount of bound Li+ on theelectrode surface in equilibrium with mobile Li+ in the liquidregion. The resulting surface concentration of Li+ ions (whichare likely coordinated to some solvent molecules) may besubstantially different from zero temperature DFT estimatesperformed in the absence of the liquid electrolyte.39,53 In asimilar vein, on SEI-covered anodes, Li+ (and perhaps lesslikely, PF6

−) may adsorb on the SEI surface, and the total netcharge of a SEI-covered anode model may have a surfacecharge−potential relationship quite different from that inFigure 4. The goal in this work is to use free energy methodsto elucidate initial SEI formation prior to full SEI formation, butthe ability to compute surface charge/voltage profiles for SEI-coated electrodes will be essential for future studies of voltage-dependent Li+ transport through the SEI film.The spatial distribution of mobile Li+ ions is not fully

converged within AIMD time scales currently available. Hence,it is difficult to quantify the structure or function of the doublelayer in the simulations reported in this work. It is possible thatthe up to ∼0.12 eV statistical uncertainties and the small Debyescreening lengths have obscured double-layer effects. Advancedsimulation techniques to accelerate salt sampling will be anextremely valuable improvement. However, we stress that theinitial AIMD configuration is pre-equilibrated by Monte Carlosimulations of classical force field models, where the electrodesare not polarizable but exhibit the expected constant surfacecharge density. Therefore, the electric double layer should berepresented to the extent that the force fields for electrodes andelectrolytes are accurate.Previous studies have shown that the M06-L functional yields

the most accurate Li+ solvation free energy in acetonitrile.70

The M05-2X and PBE-derived functionals have also beenshown to predict Li+/EC binding energies that differ by up to afew kcal/mol (∼0.12 eV). This can be a source of smallsystematic error in ΔGt, which involves Li+ solvation in ECliquid as an end point.We have not considered the different stages of lithium-

intercalated graphite. At higher voltages, LiC12, LiC18, and LiC24stoichiometries, and even graphite free of Li content,dominate.79,80 AIMD simulations of interfaces are not ideally

Figure 8. (a) Instantaneous Kohn−Sham orbital decomposed on toatoms at their x-coordinates. Two < x < 14 Å denotes the Li(s) region;outside that range resides the electrolyte. The Fermi level is at E = 0.0eV. This simulation cell has a net +|e| charge due to the solvated Li+.

The Journal of Physical Chemistry C Article

dx.doi.org/10.1021/jp408974k | J. Phys. Chem. C 2013, 117, 24224−2423524232

Page 10: Toward First Principles Prediction of Voltage Dependences of Electrolyte/Electrolyte Interfacial Processes in Lithium Ion Batteries

suited to accommodating the changes in Li content and latticeconstant variations as these stages transform into each other.Instead, we have focused on the LiC6 stoichiometry, consistentwith Li intercalation at the lowest potential, to facilitate thestudy of liquid−solid interfacial effects. We have also frozen allLi ions inside the slab so far. Finally, dispersion-corrected DFTcan be used in the future.81 This is not expected to change theresults substantially because AIMD simulations start fromclassical force-field-initiated configurations which are partlydetermined by dispersion forces.

VI. CONCLUSIONS

In this work, we have applied AIMD simulations to calculatethe free energy of Li+ transfer (ΔGt) from an electronicallyconducting LiC6 anode to liquid EC electrolyte, in condensed-phase settings appropriate to lithium ion batteries. We havecorrelated ΔGt with the voltage on the anode, which in turndepends on the net surface charge. Negative surface chargedensities are compensated by mobile, solvated Li+ in theelectrolyte in charge-neutral simulation cells. The approach,which does not require a vacuum gap, should be rigorous formodeling electrochemical reactions on macroscopic metallicelectrodes in lithium ion batteries in the limit of largesimulation cell sizes and long trajectories. Even in the presentapplications to nanoscale simulation cells, the results are usefulfor calibrating voltages that cause low-barrier electrolytedecomposition reactions in the same simulation cells.43,51,54,55

These calculations can also potentially be corroborated withstate-of-the-art nanoelectrochemical measurements.As a proof-of-principle example, we have considered LiC6

basal planes and their interfaces with ethylene carbonate (EC).These electrochemically inert graphite surfaces slow down ECdecomposition and permit sufficient sampling of ΔGt overpicosecond time scales. Only at large negative potentialsrelative to Li+/Li(s) are electrolyte decomposition eventsobserved in picosecond time scales. We predict that the basalplanes need to be negatively charged to retain Li+. A “lithiationpotential of zero charge” (LPZC) of 1.24 V vs Li+/Li(s) ispredicted for LiC6 basal planes. This quantity is notmeasurable; Li deintercalation would have occurred at thisvoltage if edge planes were present to permit it.At present, the statistical uncertainty in these computation-

ally intensive voltage calculations is on the order of 0.12 V. Thespatial distributions of mobile Li+ may not be extremely well-converged within AIMD time scales used. However, ourempirical finding is that voltages predicted with salt present donot exhibit significantly larger statistical uncertainties thansimulations conducted without mobile ions. Our qualitativeconclusion about average surface charges at electrode/electro-lyte interfaces is robust, well within the margins of estimateduncertainties. Interesting future applications include anodesurfaces which are more reactive, e.g., graphite edge planes andelectrode surfaces containing spatial/chemical inhomogeneities.

■ ASSOCIATED CONTENT

*S Supporting InformationFurther documentation regarding classical force field calcu-lations, snapshots and Li+ distributions along AIMDtrajectories, and DFT work function predictions. This materialis available free of charge via the Internet at http://pubs.acs.org.

■ AUTHOR INFORMATIONCorresponding Author*E-mail [email protected]; Tel (505) 844-1588 (K.L.).NotesThe authors declare no competing financial interest.

■ ACKNOWLEDGMENTSWe thank John Sullivan, Michiel Sprik, Andrew Leenheer,Marie-Pierre Gaigeot, Marialore Sulpizi, Oleg Borodin, KevinZavadil, and Peter Feibelman for useful discussions. SandiaNational Laboratories is a multiprogram laboratory managedand operated by Sandia Corporation, a wholly ownedsubsidiary of Lockheed Martin Corporation, for the U.S.Deparment of Energy’s National Nuclear Security Admin-istration under Contract DE-AC04-94AL85000. AIMD simu-lations were supported by Nanostructures for Electrical EnergyStorage (NEES), an Energy Frontier Research Center fundedby the U.S. Department of Energy, Office of Science, Office ofBasic Energy Sciences, under Award DESC0001160. Classicalforce field simulations were funded by Sandia’s Laboratory-Directed Research and Development program.

■ REFERENCES(1) Advances in Lithium-Ion Batteries; van Schalkwijk, W. A., Scrosati,B., Eds.; Kluwer: New York, 2002.(2) Lithium-Ion Batteries: Solid-Electrolyte Interphase; Wang, Y.,Balbuena, P. B., Eds.; Imperial College: London, 2004.(3) Lithium Batteries; Peled, E., Ed.; Academic Press: London, 1983.(4) Xu, K. Nonaqueous Liquid Electrolytes for Lithium-BasedRechargeable Batteries. Chem. Rev. 2004, 104, 4303−4417.(5) Winget, P.; Cramer, C. J.; Truhlar, D. G. Computation ofEquilibrium Oxidation and Reduction Potentials for Reversible andDissociative Electron-Transfer Reactions in Solution. Theor. Chem. Acc.2004, 112, 217−227.(6) Chen, S. T.; Raugei, S.; Rousseau, R.; Dupuis, M.; Bullock, R. M.Homogeneous Ni Catalysts for H2 Oxidation and Production: AnAssessment of Theoretical Methods, from Density Functional Theoryto Post Hartree-Fock Correlated Wave-Function Theory. J. Phys.Chem. A 2010, 114, 12716−12724.(7) Hughes, T. F.; Friesner, R. A. Development of Accurate DFTMethods for Computing Redox Potentials of Transition MetalComplexes: Results for Model Complexes and Application toCytochrome P450. J. Chem. Theory Comput. 2012, 8, 442−459.(8) Konezny, S. J.; Doherty, M. D.; Luca, O. R.; Crabtree, R. H.;Soloveichik, G. L.; Batista, V. S. Reduction of Systematic Uncertaintyin DFT Redox Potentials of Transition-Metal Complexes. J. Phys.Chem. C 2012, 116, 6349−6356.(9) Adriaanse, C.; Cheng, J.; Chau, V.; Sulpizi, M.; VandeVondele, J.;Sprik, M. Aqueous Redox Chemistry and the Electronic BandStructure of Liquid Water. J. Phys. Chem. Lett. 2012, 3, 3411−3415.(10) Halley, J. W.; Mazzolo, A.; Zhou, Y.; Price, D. First-PrinciplesSimulations of the Electrolde/Electrolyte Interface. J. Electroanal.Chem. 1998, 450, 273−280.(11) Cheng, J.; Sprik, M. Aligning Electronic Energy Levels at theTiO2/H2O Interface. Phys. Rev. B 2010, 82, 081406.(12) Cheng, J.; Sprik, M. Alignment of Electronic Energy Levels atElectrochemical Interfaces. Phys. Chem. Chem. Phys. 2012, 14, 11245−11267.(13) Schnur, S.; Gross, A. Properties of Metal-Water InterfacesStudied from First Principles. New J. Phys. 2009, 11, 125003.(14) Schnur, S.; Gross, A. Challenges in the First-PrinciplesDescription of Reactions in Electrocatalysis. Catal. Today 2011, 165,129−137.(15) Otani, M.; Hamada, I.; Sugino, O.; Morikawa, Y.; Okamoto, Y.;Ikeshoji, T. Electrode Dynamics from First Principles. J. Phys. Soc. Jpn.2008, 77, 024802.

The Journal of Physical Chemistry C Article

dx.doi.org/10.1021/jp408974k | J. Phys. Chem. C 2013, 117, 24224−2423524233

Page 11: Toward First Principles Prediction of Voltage Dependences of Electrolyte/Electrolyte Interfacial Processes in Lithium Ion Batteries

(16) Otani, M.; Sugino, O. First-principles Calculations of ChargedSurfaces and Interfaces: A Plane-Wave Non-repeated Slab Approach.Phys. Rev. B 2006, 73, 115407.(17) Zipoli, F.; Car, R.; Cohen, M. H.; Selloni, A. Simulation ofElectrocatalytic Hydrogen Production by a Bioinspired CatalystAnchored to a Pyrite Electrode. J. Am. Soc. Chem. 2010, 132, 8593−8601.(18) Nazmutdinov, R. R.; Bronshtein, M. D.; Berezin, A. S.; Soldano,G.; Schmickler, W. Bond Breaking Electron Transfer Across aConducting Nanowire(nanotube)/Electrolyte Solution Interface:The Role of Electrical Double Layer Effects. J. Electroanal. Chem.2011, 660, 309−313.(19) Santos, E.; Quaino, P.; Schmickler, W. Theory of Electro-catalysis: Hydrogen Evolution and More. Phys. Chem. Chem. Phys.2012, 14, 11224−11233.(20) Letchworth-Weaver, K.; Arias, T. A. Joint Density FunctionalTheory of the Electrode-Electrolyte Interface: Application to FixedElectrode Potentials, Interfacial Capacitances, and Potentials of ZeroCharge. Phys. Rev. B 2012, 96, 075140.(21) Viswanathan, V.; Hansen, H. A.; Rossmeisl, J.; Jaramillo, T. F.;Pitsch, H.; Norskov, J. K. Simulating Linear Sweep Voltammetry fromFirst-Principles: Application to Electrochemical Oxidation of Water onPt(111) and Pt3Ni(111). J. Phys. Chem. C 2012, 116, 4698−4704.(22) Taylor, C. D.; Wasileski, S. A.; Filhol, J.-S.; Neurock, M. FirstPrinciples Reaction Modeling of the Electrochemical Interface:Consideration and Calculation of a Tunable Surface Potential FromAtomic and Electronic Structure. Phys. Rev. B 2006, 73, 165402.(23) Bjorketun, M. E.; Zeng, Z. H.; Ahmed, R.; Tripkovic, V.;Thygesen, K. S.; Rossmeisl, J. Avoiding Pitfalls in the Modeling ofElectrochemical Interfaces. Chem. Phys. Lett. 2013, 555, 145−148.(24) Rossmeisl, J.; Skulason, E.; Bjorketun, M. E.; Tripkovic, V.;Norskov, J. K. Modeling the Electrified Solid-Liquid Interface. Chem.Phys. Lett. 2008, 466, 68−71.(25) Tripkovic, V.; Bjorketun, M. E.; Skulason, E.; Rossmeisl, J.Standard Hydrogen Electrode and Potential of Zero Charge in DensityFunctional Calculations. Phys. Rev. B 2011, 84, 115452.(26) Guggenheim, E. A. The Conceptions of Electrical PotentialDifference Between Two Phases and the Individual Activities of Ions.J. Phys. Chem. 1929, 33, 842−849.(27) Pratt, L. R. Contact Potentials of Solution Interfaces: PhaseEquilibrium and Interfacial Electric Fields. J. Phys. Chem. 1992, 96,25−33.(28) Siepmann, J. I.; Sprik, M. Influence of Surface Topology andElectrostatic Potential on Water/Electrode Systems. J. Chem. Phys.1995, 102, 511−524.(29) Vatamanu, J.; Borodin, O.; Smith, G. D. Molecular DynamicsSimulation Studies of the Structure of a Mixed Carbonate/LiPF6Electrolyte near Graphite Surface as a Function of Electrode Potential.J. Phys. Chem. C 2012, 116, 1114−1121.(30) Xing, L.; Vatamanu, J.; Borodin, O.; Bedrov, D. On theAtomistic Nature of Capacitance Enhancement Generated by IonicLiquid Electrolyte Confined in Subnanometer Pores. J. Phys. Chem.Lett. 2013, 4, 132−140.(31) Yang, L.; Fishbine, B. H.; Migliori, A.; Pratt, L. R. MolecularSimulation of Electric Double-Layer Capacitors Based on CarbonNanotube Forests. J. Am. Chem. Soc. 2009, 131, 12373−12376.(32) Merlet, C.; Salanne, M.; Rotenberg, B.; Madden, P. A.Imidazolium Ionic Liquid Interfaces with Vapor and Graphite:Interfacial Tension and Capacitance from Coarse-Grained MolecularSimulations. J. Phys. Chem. C 2011, 115, 16613−16618.(33) Merlet, C.; Rotenberg, B.; Madden, P. A.; Taberna, P.-L.;Simon, P.; Gogotsi, Y.; Salanne, M. On the Molecular Origin ofSupercapacitance in Nanoporous Carbon Electrodes. Nat. Mater.2012, 11, 306−310.(34) Willard, A. P.; Reed, S. K.; Madden, P. A.; Chandler, D. Water atan Electrochemical Interface−a Simulation Study. Faraday Discuss.2009, 141, 423−441.(35) Saunders, V. R.; Freyria-Fava, C.; Dovesi, R.; Salasco, L.; Roetti,C. On the Electrostatic Potential in Crystalline Systems Where the

Charge-Density Is Expanded in Gaussian Functions. Mol. Phys. 1992,77, 629−665.(36) Leung, K.; Marsman, M. Energies of Ions in Water andNanopores within Density Functional Theory. J. Chem. Phys. 2007,127, 154722.(37) Lozovoi, A. Y.; Alavi, A.; Kohanoff, J.; Lynden-Bell, R. M. AbInitio Simulation of Charged Slabs at Constant Chemical Potential. J.Chem. Phys. 2001, 115, 1661−1669.(38) Hummer, G.; Pratt, L. R.; Garcia, A. E. Free Energy of IonHydration. J. Phys. Chem. 1996, 100, 1206−1215.(39) Leung, K. Electronic Structure Modeling of ElectrochemicalReactions at Electrode/Electrolyte Interfaces in Lithium Ion Batteries.J. Phys. Chem. C 2013, 117, 1539−1547.(40) Qian, D.; Hinuma, Y.; Chen, H.; Du, L.-S.; Carroll, K. J.; Ceder,G.; Grey, C. P.; Meng, Y. S. Electronic Spin Transition in NanosizeStochimetric Lithium Cobalt Oxide. J. Am. Chem. Soc. 2012, 134,6096−6099.(41) Chen, M. K. Y.; Wolverton, C.; Greeley, J. P. First PrinciplesSimulations of the Electrochemical Lithiation and Delithiation ofFaceted Crystalline Silicon. J. Am. Chem. Soc. 2012, 134, 14362−14374.(42) Karim, A.; Fosse, S.; Persson, K. A. Surface Structure andEquilibrium Particle Shape of the LiMn2O4 Spinel from First-Principles Calculations. Phys. Rev. B 2013, 87, 075322.(43) Leung, K.; Budzien, J. L. Ab Initio Molecular DynamicsSimulations of the Initial Stages of Solid-Electrolyte InterphaseFormation on Lithium Ion Battery Graphitic Anodes. Phys. Chem.Chem. Phys. 2010, 12, 6583−6586.(44) Ohzuku, T.; Iwakoshi, Y.; Sawai, K. Formation of Lithium-Graphite Intercalation Compounds in Nonaqueous Electrolytes andTheir Application as a Negative Electrode for a Lithium Ion(Shuttlecock) Cell. J. Electrochem. Soc. 1993, 140, 2490−2498.(45) Leung, K.; Rempe, S. B.; von Lilienfeld, O. A. Ab InitioMolecular Dynamics Calculations of Ion Hydration Free Energies. J.Chem. Phys. 2009, 130, 204507.(46) Tazi, S.; Rotenberg, B.; Salanne, M.; Sprik, M.; Sulpizi, M.Absolute Acidity of Clay Edge Sites from Ab-Initio Simulations.Geochim. Cosmochim. Acta 2012, 94, 1−11.(47) Sulpizi, M.; Gaigeot, M.-P.; Sprik, M. The S−Water Interface:How the Silanols Determine the Surface Acidity and Modulate theWater Properties. J. Chem. Theory Comput. 2012, 8, 1037−1047.(48) Motta, A.; Gaigeot, M.-P.; Costa, D. Ab initio MolecularDynamics Study of the AlOOH Boehmite/Water Interface: Role ofSteps in Interfacial Grotthus Proton Transfers. J. Phys. Chem. C 2012,116, 12514−12524.(49) von Lilienfeld, O. A. First Principles View on ChemicalCompound Space: Gaining Rigorous Atomistic Control of MolecularProperties. Int. J. Quantum Chem. 2013, 113, 1676−1689.(50) See, e.g. Shi, S.; Lu, Peng; Liu, Z.; Qi, Y.; Hector, L. G.; Li, H.;Harris, S. J. Direct Calculation of Li-Ion Transport in the SolidElectrolyte Interphase. J. Am. Soc. Chem. 2012, 134, 15476−15487 andreferences therein.(51) Ganesh, P.; Kent, P. R. C.; Jiang, D. Solid-Electrolyte InterphaseFormation and Electrolyte Reduction at Li-Ion Battery GraphiteAnodes: Insights from First-Principles Molecular Dynamics. J. Phys.Chem. C 2012, 116, 24476−24481.(52) Exceptions are calculations that include H+ and compensatinge− on metal electrodes. See ref 23.(53) For a recent review of modeling of solid state cathode materials,see: Meng, Y. S.; Arroyo-de Dompablo, M. E. First PrinciplesComputatoinal Materials Design for Energy Storage Materials inLithium Ion Batteries. Energy Environ. Sci. 2009, 2, 589−609.(54) Leung, K.; Qi, Y.; Zavadil, K. R.; Jung, Y. S.; Dillon, A. C.;Cavanagh, A. S.; Lee, S. H.; George, S. M. Using Atomic LayerDeposition to Hinder Solvent Decomposition in Lithium IonBatteries: First-Principles Modeling and Experimental Studies. J. Am.Chem. Soc. 2011, 133, 14741−14754.

The Journal of Physical Chemistry C Article

dx.doi.org/10.1021/jp408974k | J. Phys. Chem. C 2013, 117, 24224−2423524234

Page 12: Toward First Principles Prediction of Voltage Dependences of Electrolyte/Electrolyte Interfacial Processes in Lithium Ion Batteries

(55) Leung, K. First-Principles Modeling of the Initial Stages ofOrganic Solvent Decomposition on LixMn2O4(100) Surfaces. J. Phys.Chem. C 2012, 116, 9852−9861.(56) Abe, T.; Fukuda, H.; Iriyama, Y.; Ogumi, Z. Solvated Li-IonTransfer at Interface between Graphite and Electrolyte. J. Electrochem.Soc. 2004, 151, A1120−A1123.(57) Cresce, A.v.W.; Borodin, O.; Xu, K. Correlating Li+ SolvationShealth Structure with Interphasial Chemistry on Graphite. J. Phys.Chem. C 2012, 116, 26111−26117.(58) Fatteberg, J. L.; Gygi, F. Density Functional Theory for EfficientAb Initio Molecular Dynamics Simulations in Solution. J. Comput.Chem. 2002, 23, 662−666.(59) Andreussi, O.; Dabo, I.; Marzari, N. Revised Self-ConsistentContinuum Solvation in Electronic-Structure Calculations. J. Chem.Phys. 2012, 136, 064102.(60) Zhang, L. Q.; Liu, X. H.; Liu, Y.; Huang, S.; Zhu, T.; Gui, L.;Mao, S. X.; Ye, Z. Z.; Wang, C. M.; Sullivan, J. P.; Huang, J. Y.Controlling the Lithiation-Induced Strain and Charging Rate inNanowire Electrodes by Coating. ACS Nano 2011, 5, 4800−4809.(61) Hummer, G.; Pratt, L. R.; Garcia, A. E. Ion Sizes and Finite-SizeCorrections for Ionic-Solvation Free Energies. J. Chem. Phys. 1997,107, 9275−9277.(62) In DFT/PBE simulations of water, it is well known that elevatedtemperature is needed to reproduce experimental pair correlation at T= 300 K. See, e.g.: Schwegler, E.; Grossman, J. C.; Gygi, F.; Galli, G.Towards an Assessment of the Accuracy of Density Functional Theoryfor First Principles Simulations of Water. II. J. Chem. Phys. 2004, 121,5400−5409.(63) Kresse, G.; Furthmuller, J. Efficienct Iterative Schemes for AbInitio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys.Rev. B 1996, 54, 11169.(64) Kresse, G.; Joubert, J. From Ultrasoft Pseudopotentials to theProjector Augmented-Wave Method. Phys. Rev. B 1999, 59, 1758−1775.(65) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized GradientApproximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865−3868.(66) Li, T.; Balbuena, P. B. Theoretical Studies of LithiumPerchlorate in Ethylene Carbonate, Propylene Carbonate, and TheirMixtures. J. Electrochem. Soc. 1999, 146, 3613−3622.(67) Martin, M. G.; Thompson, A. P. Industrial Property Predictionusing Towhee and LAMMPS. Fluid Phase Equilib. 2004, 217, 105−110.(68) Tenney, C. M.; Cygan, R. T. Analysis of Molecular Clusters inSimulations of Lithium-Ion Battery Electrolytes. J. Phys. Chem. C,submitted.(69) Johansson, P.; Jacobsson, P. Lithium Salt Dissociation in Non-aqueous Electrolytes Modeled by Ab Initio Calculations. Solid StateIonics 2006, 177, 2691−2697.(70) Bryantsev, V. S. Calculation of Solvation Free Energies of Li+

and O2− and Neutral Lithium-Oxygen Compounds in Acetonitrile

Using Mixed Cluster/Continuum Models. Theor. Chem. Acc. 2012,131, 1250−1257.(71) Borodin, O.; Smith, G. D. Quantum Chemistry and Moleculardynamics Simulation Study of Dimethyl Carbonate:Ethylene Carbo-nate Electrolytes Doped with LiPF6. J. Phys. Chem. B 2009, 113, 1763−1776.(72) Wilson, M. A.; Pohorille, A.; Pratt, L. R. Surface Potential of theWater Liquid-Vapor Interface. J. Chem. Phys. 1987, 88, 3281−3285.(73) Jungwirth, P.; Winter, B. Ions at Aqueous Interfaces: FromWater Surface to Hydrated Proteins. Annu. Rev. Phys. Chem. 2008, 59,1457−1463.(74) Zhu, L.; Hommet, F.; Salace, G.; Claude-Montigny, B.; Melrot,A. STM Observation of the Electro-adsorption of Lithium Ions ontoGraphite and of the Ensuing Solid Electrolyte Interphase Formation.Surf. Sci. 2002, 512, 84−96.(75) See, e.g.: Trasatti, S. Surface Science and Electrochemistry:Concepts and Problems. Surf. Sci. 1995, 335, 1−9.

(76) Leung, K. Two-electron Reduction of Ethylene Carbonate: aQuantum Chemistry Re-examination of Mechanisms. Chem. Phys. Lett.2013, 568−569, 1−8.(77) The “quantum mechanical push back” effect, operative at water-metal interfaces, is however likely less significant on graphite basalplanes. See: Bagus, P. S.; Staemmler, V.; Woll, C. Exchangelike Effectsfor Closed-Shell Adsorbates: Interface Dipole and Work Function.Phys. Rev. Lett. 2002, 89, 096104.(78) Smith, G. D.; Borodin, O.; Russo, S. P.; Rees, R. J.; Hollenkamp,A. F. A Molecular Dynamics Simulation Study of LiFePO4/ElectrolyteInterfaces: Structure and Li+ Transport in Carbonate and Ionic LiquidElectrolytes. Phys. Chem. Chem. Phys. 2008, 11, 9884−9897.(79) Persson, K.; Hinuma, Y.; Meng, Y. S.; van der Ven, A.; Ceder, G.Thermodynamic and Kinetic Properties of the Li-Graphite Systemfrom First Principles Calculations. Phys. Rev. B 2010, 82, 125416.(80) Holzwarth, N. A. W.; Louie, S. G.; Rabii, S. Lithium-IntercalatedGraphite: Self-Consistent Electronic Structure for Stages One, Two,and Three. Phys. Rev. B 1983, 28, 1013−1025.(81) Grimme, S. Semi-empirical GGA-Type Density FunctionalConstructed with a Long-Range Dispersion Correction. J. Comput.Chem. 2006, 27, 1787−1799.

The Journal of Physical Chemistry C Article

dx.doi.org/10.1021/jp408974k | J. Phys. Chem. C 2013, 117, 24224−2423524235


Recommended