Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Journal of The Electrochemical

Society

OPEN ACCESS

Ion Diffusivity through the Solid Electrolyte Interphase in Lithium-Ion


Batteries
To cite this article: Laura Benitez and Jorge M. Seminario 2017 J. Electrochem. Soc. 164 E3159

View the article online for updates and enhancements.

This content was downloaded from IP address 110.93.234.10 on 13/08/2021 at 10:42


Journal of The Electrochemical Society, 164 (11) E3159-E3170 (2017) E3159

JES FOCUS ISSUE ON MATHEMATICAL MODELING OF ELECTROCHEMICAL SYSTEMS AT MULTIPLE SCALES IN HONOR OF JOHN NEWMAN
Ion Diffusivity through the Solid Electrolyte Interphase in
Lithium-Ion Batteries
Laura Beniteza,b and Jorge M. Seminarioa,b,c,z
a Department of Chemical Engineering, Texas A&M University, College Station, Texas 77843, USA
b Department of Electrical and Computer Engineering, Texas A&M University, College Station, Texas 77843, USA
c Department of Materials Science and Engineering, Texas A&M University, College Station, Texas 77843, USA

Understanding the transport properties of the solid electrolyte interphase (SEI) is a critical piece in the development of lithium ion
batteries (LIB) with better performance. We studied the lithium ion diffusivity in the main components of the SEI found in LIB
with silicon anodes and performed classical molecular dynamics (MD) simulations on lithium fluoride (LiF), lithium oxide (Li2 O)
and lithium carbonate (Li2 CO3 ) in order to provide insights and to calculate the diffusion coefficients of Li-ions at temperatures in
the range of 250 K to 400 K, which is within the LIB operating temperature range. We find a slight increase in the diffusivity as
the temperature increases and since diffusion is noticeable at high temperatures, Li-ion diffusion in the range of 1300 K to 1800 K
was also studied and the diffusion mechanisms involved in each SEI compound were analyzed. We observed that the predominant
mechanisms of Li-ion diffusion included vacancy assisted and knock-off diffusion in LiF, direct exchange in Li2 O, and vacancy and
knock-off in Li2 CO3 . Moreover, we also evaluated the effect of applied electric fields in the diffusion of Li-ions at room temperature.
© The Author(s) 2017. Published by ECS. This is an open access article distributed under the terms of the Creative Commons
Attribution 4.0 License (CC BY, http://creativecommons.org/licenses/by/4.0/), which permits unrestricted reuse of the work in any
medium, provided the original work is properly cited. [DOI: 10.1149/2.0181711jes] All rights reserved.

Manuscript submitted March 9, 2017; revised manuscript received May 5, 2017. Published May 17, 2017. This paper is part of the
JES Focus Issue on Mathematical Modeling of Electrochemical Systems at Multiple Scales in Honor of John Newman.

The continuous demand for smart phones, tablets and other mo- detach it from the active material surface.28,29 Moreover, most of the
bile devices has enormously promoted the improvement of Li-ion SEI components are highly reactive when exposed to contaminants,
batteries. Moreover, volatile oil prices, pollution and climate change air or humidity.16 For these reasons, ex situ characterization of the SEI
concerns have further increased the need for energy storage de- becomes very difficult whereas in situ experiments require specially
vices for renewable energies and electric vehicles. The latter ap- designed tools and measurement set ups.30,31 Given the difficulties
plications require batteries with even higher energy density, bet- of experimental techniques, computational simulations, increasingly
ter cycle life, better power performance, lower cost and improved become a valuable tool to study properties of the SEI.
safety.1 Therefore, improving and increasing the fundamental scien- Many studies try to understand parameters such as composition,
tific knowledge of lithium ion batteries is essential for their continual morphology and thickness, as well as to clarify formation and growth
advancement. mechanisms of SEI films.13,23,32–37 Its structure and composition have
A key component in lithium ion batteries is the solid electrolyte been investigated using, both in situ and ex situ, experimental tech-
interphase (SEI) film. The SEI film is a thin passivating layer that is niques such as X-ray photoelectron spectroscopy (XPS), infrared
initially formed, on both anode and cathode surfaces, from the reduc- spectroscopy (IR), nuclear magnetic resonance (NMR) spectroscopy,
tion of the electrolyte during the first charging/discharging cycles.2–9 Raman spectroscopy, X-ray diffraction (XRD), transmission electron
It consists of a mixture of inorganic and organic products, such as LiF, microscopy (TEM), scanning electron microscopy (SEM), and time-
Li2 O and LiCO3 , and (CH2 OCO2 Li)2 , ROCO2 Li and ROLi, where R is of-flight secondary ion mass spectroscopy (TOF-SIMS); the most
an organic group such as CH2 , CH3 , CH2 CH2 , CH2 CH3 , CH2 CH2 CH3 generally recognized inorganic compounds in literature are LiF, Li2 O
that depends on the electrolyte solvent.10–24 Proposed structure mod- and Li2 CO3 .16,18,21,31,38–43 Moreover, there is an increasing number of
els suggest that a dense layer of inorganic products is found near the studies, both experimental and theoretical, on Li-ion transport within
electrode (inner layer) followed by a porous organic layer near the the electrolyte and in the electrodes.44–65 Other investigations focus on
electrolyte/SEI interface (outer layer).19–21 In carbon anodes, the SEI the boundary between the liquid electrolyte and the electrode, even the
layer prevents further decomposition of the electrolyte by hindering SEI as a whole,21,25,42,66–71 yet little attention has been put toward ex-
electron transport from the electrode to the electrolyte, and by block- plaining transport mechanisms and predicting diffusion coefficients in
ing the travel of solvent molecules to the active surface of the anode, the individual interphase components. Lithium ion diffusion in the SEI
where they can react with lithium ions and electrons.25 In silicon an- is thought to occur through grain boundaries, through porous regions,
odes, i.e., the huge volume expansion of the anode creates cracks in or through interstitials and vacancies.9,72–74 Each SEI component may
the SEI layer and generates new surfaces which are freshly exposed exhibit one or more of these ion transport mechanisms.
to the electrolyte.26 In both carbon and silicon anodes, the SEI film One of the most often-reported to be found in the SEI is LiF; this in-
still allows the transport of Li+ from the electrolyte to the negative organic compound has been observed in both carbon and silicon based
electrode; lithium ions can move through the SEI by the exchange anodes. One study reported that its cation diffusivity is much lower
of ions between the electrolyte, the SEI compounds, and the lithium than in other SEI inorganic compounds, and suggested that diffusion
intercalated in the electrodes.25 in LiF cause rate limitations in Li-ion batteries;73 they used periodic
It has been noted that battery performance, cycle life, and safety density functional theory (DFT) calculations to determine dominant
are highly dependent on the characteristics of the SEI, yet it is the diffusion carriers and diffusion pathways, as well as used the nudge
“least understood component in lithium ion batteries”.27 This is due, elastic band (NEB) method to calculate energy barriers of diffusion;
in part, to the fact that the experimental analysis is very challenging. they found that vacancies were “energetically more favorable” than
The SEI film thickness is very small (few Å to tens or hundreds of interstitials, and reported energy barriers of 0.73 and 1.09 eV for neu-
nanometers) and it is formed on the electrode surface thus making it tral vacancies and neutral Schottky vacancies, respectively; moreover
almost impossible to define the boundaries of the SEI, and to precisely the associated diffusion coefficients they found were in the range of
10−26 to 10−20 m2 /s. Similar results for the energy barrier were found
in other DFT studies where the lithium dynamics were investigated.8,75
z
E-mail: seminario@tamu.edu Also, earlier experimental investigations using NMR obtained energy
E3160 Journal of The Electrochemical Society, 164 (11) E3159-E3170 (2017)

barriers in the range of 0.65 to 0.73 eV.76 Two recent studies, one us- This study provides insight on lithium ion transport through indi-
ing molecular dynamics,77 and the other applying phase-field model vidual SEI compounds: 1) At both low and high temperatures; and 2)
together with Fick’s law,21 reported diffusion coefficients of Li in LiF At room temperature under an applied external electric field. This fun-
at room temperatures (298 K to 318 K) in the range of 3.7 × 10−16 damental knowledge is useful in multi-scale computational methods
m2 /s. Moreover, diffusion in LiF has been more extensively studied at developed to simulate SEI nucleation, growth and evolution.21,105,106
high temperatures since this compound is an important component of The long-term goal is to eventually have more control over interface
the molten salt mixtures used as primary coolant and fuel in nuclear parameters such as composition, structure, porosity and thickness, and
fission reactors. Investigations using molecular dynamics and experi- thus accurately design SEI films and therefore better Li-ion batteries.
mental techniques obtained diffusion coefficients in high temperature This work is a step toward this ultimate goal.
range (700 K to 1400 K), and in temperatures close to the melting In this paper, the main components of the inorganic SEI layer
point of LiF, and reported values in the range from 0.25 to 1.86 eV are studied: lithium fluoride (LiF), lithium oxide (Li2 O) and lithium
for the diffusion energy barrier.77–83 carbonate (Li2 CO3 ).15,17 First, the force fields needed for accurate
Only a couple of theoretical studies reported diffusion barriers molecular dynamics are evaluated. MD has the advantage that no
in Li2 O at low temperature.8,75 Chen et al.8 used DFT to investi- previous knowledge of diffusion pathways is required and it can con-
gate the electronic structure and the vacancy-assisted Li diffusion sider many-particle effects. Moreover, MD simulations allow studying
using NEB method; they showed that Li2 O electronic structure had phenomena that is not easily accessible with experimental techniques.
insulating character and obtained a diffusion barrier of 0.15 eV. Ad- Diffusion coefficients at several temperatures are calculated in order
ditionally, Guan et. al21 and Tasaki et. al,84 in addition to studying to provide fundamental understanding on the transport properties of
LiF, also studied diffusion in Li2 O and found transport coefficients the solid electrolyte interphase found in lithium ion batteries. Further-
in the range of 1.7 × 10−16 m2 /s. Lithium diffusion in Li2 O has more, the room temperature diffusivity of Li-ions as function of an
been widely studied at high temperatures due to its superionic (fast applied electric field is obtained.
ion mobility) behavior at elevated temperatures.49,85–96 Most of these In Methodology section, details of the computational methods used
works focused on developing and evaluating force field parameters are given. In Results and Discussion section, the results and discus-
for molecular dynamics simulations, which were then used to an- sions are presented. The results from the defect free SEI structures
alyze structure, physical properties, and diffusion.85–94 Experimen- are first discussed; our findings on diffusion of Li-ions over low and
tally, diffusion of Li+ in Li2 O was studied by Oishi et. al95 using high temperatures are then examined; and finally, the details of the
mass spectrometry and the 6 Li radioisotope as a tracer. In a recent diffusion coefficients obtained after applying an electric field are re-
study,96 non-equilibrium molecular dynamics (NEMD) were used to ported. In Conclusions section, the findings and conclusions are briefly
obtain diffusion coefficients of Li at temperatures from ∼870 K to summarized.
∼1600 K. In NEMD, a fictitious electric field is applied in order Methodology
to increase the occurrence of diffusion-related hopping events and
thus accurately calculate coefficients. They found that diffusion be- The SEI products in this study consist of lithium fluoride (LiF),
low 1200 K is dominated by synchronous nearest-neighbor hoping lithium oxide (Li2 O) and lithium carbonate (Li2 CO3 ). Their structures
also called ring diffusion,49 where two or more adjacent atoms move and crystal lattice parameters are obtained from the Inorganic Crystal
at the same time; at higher temperatures, interstitial-assisted diffusion Structure Database (ICSD).107 A 5 × 5 × 5 supercell is used for LiF,
mechanism governs. They also report, energy diffusion barriers of a 4 × 4 × 4 supercell for Li2 O and a 3 × 3 × 4 supercell for L2 CO3 ;
0.26 and 1.11 eV for high temperatures and for superionic regime, simulation box sizes are 20 × 20 × 20 Å3 for LiF, 22 × 22 × 22 Å3
respectively. for Li2 O, and 24 × 21 × 25 Å3 for Li2 CO3 . Figure 1 shows the SEI
Lithium carbonate (Li2 CO3 ) has been shown to be the main compo- structures studied.
nent of the dense inner SEI layer and it is claimed to be responsible for In addition, all studied SEI structures are set up with three-
stabilizing SEI films.74 Recent studies have mainly focused on finding dimensional periodic boundary conditions. Both lithium fluoride and
the dominant diffusion carriers (interstitials and/or vacancies), their lithium oxide are modeled by the standard 6–12 Lennard-Jones (LJ)
diffusion pathways and diffusion mechanisms, as well as the associ- potential energy, which for a pair of atoms i and j at a distance r is
ated diffusion barriers. Shi and co-workers9,74 in two separate works given by
  
σ 12  σ 6
determined the dominant diffusion carriers, among interstitials, va-
cancies and Frenkel pairs, over a voltage range from 0 to 4.4V; they E6−12 LJ = 4ε − [1]
found that below 0.98 V interstitial Li+ ions are the main carriers,74 r r
that above ∼4 V vacancies dominate diffusion, and that between 0.98 ε is the depth of the LJ potential well, σ is the distance at which the
and ∼4 V interstitials and vacancies have the same influence.9 More- potential is zero and σ = 2(-1/6) rm ; rm is the distance at the minimum
over, using the climbing image NEB (CI-NEB) method they observed (equilibrium) energy.
that interstitial Li+ moves via knock off mechanism and vacancy- In addition, a Coulombic pairwise interaction is also used given
assisted Li+ diffusion is via direct hopping, and calculated energy by
barriers in the range of 0.31 for interstitials and 0.24 eV for vacancies. qi qj
In addition, they report diffusion coefficient values in the range of ECoul = [2]
10−11 m2 /s over the voltage range they studied. Other investigations r
of Li-ion diffusion in bulk monoclinic Li2 CO3 used DFT to find mi- qi and qj are the charges of the pair of atoms i and j, and ε is the di-
gration barriers and reported values of 0.23 to 0.49 eV.6,8 While there electric constant. In considering the charges, we take into account that
are studies focused on investigating the structure and thermodynamics fluorine presents the possibility of forming a σ-hole,108–110 a noncova-
of Li2 CO3 ,97–102 to the best of our knowledge, there are no reports on lent interaction between a covalently-bonded halogen and a negative
diffusion of Li-ion at high temperatures. site; however, the fraction of covalent bonding in LiF is very little
While the previously mentioned studies calculate diffusion coef- ∼0.1 due to the large differences in electronegativitives of the two
ficients only at room temperature, or only at very high temperatures atoms, which actually yield a strong ionic bond.
(near melting), our work provides coefficients over a wider temper- For both potentials, LJ and Coulombic, there is an rc cutoff distance
ature range from 250 K to 1800 K, which includes LIB extreme after which the energy is not calculated; the cutoff distance (rc ) in all
charging temperatures (−20 to 45◦ C).103 Since SEI components are calculations is 10 Å. In order to compensate for the abrupt change in
highly sensitive to temperature,104 characterization under different energy at the cutoff distance, long range Coulombic interactions in LiF
temperatures is critical. Moreover, we investigate what effect an ap- and Li2 O are evaluated by the Ewald summation technique outside
plied electric field has on the diffusion coefficient and on the diffusion this cutoff distance.111 Lithium carbonate uses the standard 6–12 LJ
mechanisms. and coulombic interaction potentials given by 1 and 2 respectively,
Journal of The Electrochemical Society, 164 (11) E3159-E3170 (2017) E3161

LiF Li2O Li2CO3


5x5x5 4x4x4 3x3x4

21 Å
(a) (b) (c)

Figure 1. SEI structures studied. (a) LiF: 5 × 5 × 5 supercell, 20 × 20 × 20 Å box size, 500 Li atoms and 500 F atoms. (b) Li2 O: 4 × 4 × 4 supercell, 22 × 22
× 22 Å box size, 512 Li atoms and 256 O atoms. (c) Li2 CO3: 3 × 3 × 4 supercell, 24 × 21 × 25 Å box size, 288 Li atoms and 144 C atoms and 432 O atoms.
Li (purple), F (cyan), O (red), C (brown).

together with bond, angle and improper dihedral interactions given by software, Visual Molecular Dynamics (VMD).116 In this investiga-
tion, the defect-free samples are first studied, and then the point de-
Ebond = kb (r−r0 ) 2
[3]
fects are added in each SEI sample and the diffusion is evaluated.
We start with an energy minimization of each SEI product in order
to find a stable point; we then equilibrate the samples at 5 K for 1 ns
Eangle = kθ (θ−θ0 )2 [4] to allow the relaxation of the structure; next we follow a simple MD
simulated annealing with the following steps: first, a gradual tem-
perature ramp up to 800 K is performed at a rate of 0.795 K/ps;
Eimproper = kχ (χ−χ0 )2 [5] subsequently, the sample is cooled down to 250 K at a rate of 0.55
where kb , kθ and kχ are force constants, r is the distance between K/ps, and finally it is stabilized at 250 K for 1 ns. Annealing is done
the atoms and r0 is the equilibrium bond distance; θ is the angle in order to find a structure closer to a global minimum in the energy
between atoms and θ0 is the equilibrium angle; χ is the improper and to optimize the structure of the samples. Then temperature equi-
dihedral angle and χ0 is the equilibrium value of the improper dihedral librations lasting 1 ns each are done in 25 K increments from 250 K
angle. The 6–12 LJ potential parameters and ionic charges used for to 400 K, in 50 K increments from 400 K to 600 K, and finally in 100
LiF are tabulated in Table I.84,112 ε and σ are mixed using Lorentz- K increments from 600 K to 1800 K; increase rates are 1 K/ps for
Berthelot rules.112 The parameters and charges used to describe Li2 O all the aforementioned temperature ranges. All operations are done at
and Li2 CO3 are presented in Table I.113,114 ε and σ are mixed using zero pressure since experimental diffusion studies and, in general, SEI
geometric combination rules.114 Bond, angle and improper interaction studies must be carried out in vacuum. The NPT ensemble is used for
parameters utilized are tabulated in Table II. all MD simulations and then some features are compared with those
All molecular dynamics simulations are done using Large-scale under the NVT ensemble. The time step used for LiF and Li2 O is 1 fs,
Atomic/Molecular Massively Parallel Simulator (LAMMPS),115 and a typical value for MD simulations, and for Li2 CO3 , a 0.1 fs time step
all visualizations of the structures are performed using the graphics is used since values smaller than 1 fs are used in samples modeled
with bonded interactions.59,117
First, we evaluate the quality of the force fields by checking the
crystal structure and bond distances of each structure after the energy
Table I. 6–12 Lennard-Jones Parameters and Atom Charges for minimization, and after equilibrium at 300 K. In addition, snapshots
LiF, Li2 O and Li2 CO3. 84, 112–114

SEI Product Atom ε (kJ /mol) σ (Å) Q (e)


Table II. Bond, Angle and Improper Dihedral Interaction
LiF Li 0.24125 1.715 0.78 Parameters for Li2 CO3 .
F 0.02707 3.954 −0.78
Li2 O Li 0.1046 2.183 0.75
Bond C-O Angle O-C-O Improper O-C-O-O
O 0.25104 3.118 −1.5
Li2 CO3 Li 0.1046 2.183 1.0 kb (kJ/mol-Å2 ) r0 kθ (kJ/mol-rad2 ) θ0 kχ (kJ/mol-rad2 ) χ0
C 0.43932 3.431 0.9853
O 0.25104 3.118 −0.9951 3222 1.3 460.2 120 1050 180
E3162 Journal of The Electrochemical Society, 164 (11) E3159-E3170 (2017)

Figure 2. Energy versus time curves and Temperature


versus time curves under NPT and NVT ensembles for
(a-b) LiF structures (vacancy defect ratio of 2/1000), (c-d)
Li2 O structures (vacancy defect ratio of 3/798), and (e-f)
Li2 CO3 structures (interstitial defect ratio of 1/864).

of the structures are taken during each temperature equilibration every Lastly, an electric field is applied in the range of 0.1 V/Å to 0.85
2 ps, and the coordinates are saved. Then, the radial distribution func- V/Å to each of the studied structures and the diffusion coefficients
tions (RDF’s) for all the atom pairs in each compound are calculated are calculated. In different simulation runs for each SEI product and
with VMD at each temperature using the saved atom coordinates from for each field applied, energy minimization, equilibration at 5 K for
the 500 snapshots taken. Afterwards, the diffusion of Li+ at various 100 ps, thermalization up to 800 K at a rate of 7.95 K/ps, cool down
temperatures is evaluated. First, defects are created in each compound to 300 K at a rate of 7.95 K/ps, and finally equilibration at 300 K
(LiF, Li2 O, and Li2 CO3 ) in order to allow Li+ ion diffusion. For that, for 100 ps are done. The electric field is applied to the sample at the
one Li-ion and one F-ion are removed in LiF sample; two Li-ions and 300 K equilibration and the MSD of the Li-ions is recorded during
one O-ion are removed in Li2 O sample; and for Li2 CO3 , one Li-ion is a 100 ps timeframe and snapshots of the structures are taken every
added as interstitial. Again, energy minimization, equilibration at 5 K, 0.1 ps. As the electric field is increased from zero, only results in the
thermalization, and equilibration at 250 K are done. Then temperature linear regime of the MSD can be directly compared to results with no
equilibrations are done from 250 K to 600 K and at high temperatures field applied.96 We increase the field up to 0.85 V/Å where non-linear
in the range of 700 K to 1800 K; temperature increments and the rates effects are clearly observed.
for the temperature ramps are the same as above. LiF and Li2 O are
compared with available experimental and MD results. Once the struc-
ture is equilibrated, the mean square displacement (MSD) of the Li+
ion is recorded at each temperature using the appropriate “compute” Results and Discussion
command in LAMMPS. Diffusion coefficients (D) are calculated from Energy and temperature profiles.—We first analyze the total en-
the mean square displacement (MSD), since D is proportional to the ergy obtained at each stage of the simulations performed for pristine
MSD71 as shown by and impure samples. Figures 2a–2f shows total energy and the tem-
1  2 perature versus time for each case studied.
 
D =  r (t) − r (t0 )  [6] For the case of LiF, shown in Figures 2a–2b, we can observe that
6 the energy of the pristine crystal structure (blue curve) starts at about
where r is the position of the particle at each time step, t is the time, t0 −76 Gcal/mol, then increases to −70 Gcal/mol as the temperature
  is increased to 800 K (∼2 ns), then it goes down to −74 Gcal/mol
is the initial time, r (t) − r (t0 ) is the distance traveled by the particle when the temperature decreases to 250 K (∼3 ns). Afterwards, the
 
over the time interval (t - t0 ), and | r (t) − r (t0 )|2  is the MSD. The energy increases in a stepwise manner to an average energy of −67.5
diffusion coefficient is found from the slope of the MSD vs time. Gcal/mol at a temperature of 1200K (∼20 ns). At this temperature
Journal of The Electrochemical Society, 164 (11) E3159-E3170 (2017) E3163

Figure 3. Structure after 300 K equilibrium for (a) LiF, (b) Li2 O and (c) Li2 CO3 . Li (purple), F (cyan), O (red), C (brown).

(1200 K), the energy significantly jumps to −63.8 Gcal/mol. This restarted using the last saved state from the NPT test run. The calcula-
suggests a change of phase in the structure, namely indicating the tions were restarted using the simulation box size and shape, boundary
melting of LiF. There is a significant volume increase that produces settings, atom positions and velocities, as well as atom attributes and
a corresponding energy increase. After that, the energy continues force field styles and coefficients. Therefore, for the case of Li2 CO3 ,
increasing to −58.4 Gcal/mol as the temperature goes to 1800 K shown in Figures 2e–2f, the energy of the defect-free structure (blue
(∼28 ns). Energy variations around average values are approximately curve) starts around the average value of −62 Gcal/mol and slowly
0.1 Gcal/mol at low temperatures and 0.7 Gcal/mol at high tempera- goes down to an average value of −74 Gcal/mol as the temperature
tures. The energy profile does not change when the vacancy defects goes up to 1800 K. When the interstitial Li+ is introduced to the
(black curve) are introduced to the structure. In both LiF cases, pure structure (black curve), no changes are observed in the energy profile.
and impure crystal structures, the major contributions to the energy The energy variations in the Li2 CO3 case are approximately 11 and
come mainly from the short range and long range coulombic interac- 17 Gcal/mol at low and high temperatures, respectively. In both cases,
tions. We then compare with simulations performed under the NVT the pristine crystal and impure structure, the major energy contribu-
ensemble. The energy of the pristine crystal structure (red curve) starts tion to the total energy comes from very strong short range coulombic
at about −71 Gcal/mol, then increases to −66 Gcal/mol as the tem- interactions. Comparing with the NVT ensemble, the energy of the
perature is increased to 800 K (∼2 ns), then it goes down to −70 pristine crystal structure (red curve) first starts at an average value of
Gcal/mol when the temperature decreases to 250 K (∼3 ns). After −73 Gcal/mol, then goes slightly up to −69 Gcal/mol and then slowly
that, the energy increases in a stepwise manner to an average energy goes down to −71 Gcal/mol as the temperature increases to 1800 K.
of −61 Gcal/mol as the temperature increases to 1800 K (∼20 ns). Energy variations are approximately the same in both NVT and NPT
In the NVT case, the volume is not able to change and thus the jump cases. Again no changes are observed in the energy profile when the
in energy at 1200 K is not observed. Energy variations are similar to interstitial defect (green curve) is introduced to the structure. Com-
the NPT case. Again, the energy profile does not change significantly paring the energy profiles of the different SEI components studied, we
when the vacancy defects (green curve) are introduced to the sample. observe that the lowest energies are obtained in the Li2 O case which
For the Li2 O defect-free structure (blue curve) shown in Figures yields energies in the range from −115 to −100 Gcal/mol (excluding
2c–2d, a large variation in the energy is observed at the beginning initial energy variations). Then LiF case yields energies from −76
of the simulation which stabilizes when the sample is cooled down to −58 Gcal/mol. Finally, Li2 CO3 , produces energies in the range of
to 250 K (∼3 ns). At 250 K the average energy value corresponds −50 to −85 Gcal/mol.
to −114 Gcal/mol. Then the energy increases in a stepwise style as
the temperature of the sample is increased all the way up to 1800 K
Bond distances and structure.—The defect free structures are then
(∼28 ns). Energy variations around average values are approximately
studied to evaluate the force fields. First, we verify that the crystal
0.1 Gcal/mol and 1 Gcal/mol at low and high temperatures, respec-
structure and bond distances of each SEI compound are maintained
tively. Again, the energy profile does not change when the vacancy
after the energy minimization and equilibrium at 300 K. Figure 3
defects (black curve) are introduced in the Li2 O crystal. In both cases,
shows the structures obtained after equilibration at 300 K for LiF,
pure crystal and impure structure, the major energy contributions to
Li2 O, and Li2 CO3 . It can be observed, in all cases, that the crystal
the total energy come from the short range and long range coulombic
structure was not changed significantly. For LiF, the crystal order is
interactions. In the NVT case, for the pristine crystal structure (red
well maintained both after minimization and equilibration at 300 K.
curve), the large variation in the energy at the beginning of the simula-
For Li2 O case, there is minor disorder after the energy minimization;
tion run is not observed; the energy of the defect-free structure starts at
however, the crystal order and structure are reestablished after ther-
−111 Gcal/mol, then it increases to −108 Gcal/mol as the temperature
malization up to 800 K and then, equilibration at 300 K. For Li2 CO3 ,
is increased to 800 K (∼2 ns); it then goes down to −111 Gcal/mol
the crystal structure is conserved during minimization; nonetheless,
when the temperature decreases to 250 K (∼3 ns). Then the energy
in the equilibration at 300 K, slight crystal distortion is seen. Table III
increases in a stepwise style as the temperature is increased all the
presents a comparison of all bond distances obtained in this work and
way up to 1800 K (∼28 ns). Energy variations around average values
the data available in the ICSD.107 For all compounds, bond distances
are approximately the same in both NVT and NPT cases. Moreover,
remain quite similar after minimization; at 300 K equilibrium, even
the energy profile does not change when the vacancy defects (green
though there is more variation than that of the minimization, the bond
curve) are introduced in the Li2 O structure.
In the case of Li2 CO3 , in order to find the best the volume for distances are still within ±0.2 Å.
this structure, a test simulation was performed under the NPT en-
semble first. In this test run, the structure was first minimized, then RDF curves and melting temperature.—RDF curves for temper-
heated to 800 K (0.795 K/ps), subsequently cooled down to 250 K atures ranging from 250 K to 1800 K are analyzed and only selected
(0.55 K/ps), then slowly heated back up to 1800 K (∼0.060 K/ps). representative RDF curves are presented in Figure 4. For LiF, the
Then, production runs under the NPT and NVT ensembles were RDF curves from 250 K to 800 K are similar. Characteristic peaks
E3164 Journal of The Electrochemical Society, 164 (11) E3159-E3170 (2017)

Table III. Comparison of Calculated and Experimental Bond lower temperatures, peaks can still be observed yet they are shifted
Distances. and broader, indicating that some disorder in the structure is starting
to appear. At 1200 K, the characteristic peaks disappear and g(r) ap-
Bond distance (Å) proximates to 1 at long range, indicating that the simulated structure
becomes liquid at this temperature; this suggests that the result for
SEI Bergerhoff After After melting temperature of 1200 K is very close to the experimental melt-
Product Bond et al.107 Minimization Equilibration at 300 K ing point of LiF (1122 K).118 At temperatures greater than 1200 K, all
RDF curves are equal to the 1200 K curve.
LiF Li-F 2.01 2.00 2.20
Li2 O RDF curves from 250 K to 600 K are alike; their peaks
Li-O 2.00 1.93 1.95
Li2 O Li-Li 2.31 2.32 2.26
decrease and broaden as temperature increases. From 800 K to 1400 K
C-O1 1.30 1.29 1.30 peaks continue to decrease and widen until individual peaks combine.
C-O2 1.30 1.29 1.30 At 1800 K, a considerable difference is observed, peaks start vanishing
Li2 CO3 C-O3 1.27 1.30 1.31 and g(r) begins approaching a value of 1 at long range; however,
Li-O1 1.97 1.96 1.95 the simulated structure does not become liquid at this temperature,
Li-O2 1.93 1.94 1.94 implying that a value greater than 1800 K for the melting temperature
Li-O3 1.89 1.89 1.96 is obtained, which is higher than the experimental result of 1700 K.119
For Li2 CO3 compound, the RDF curves in the temperature range
from 250 K to 1100 K are almost the same. A small decrease in the first
representing the crystalline structure are observed, and the intensity peak intensity is observed as the temperature increases above 1200 K,
of the peaks decreases and their widths broadens as the temperature this suggests that melting is starting to occur around this temperature;
increases. At 1000 K, the RDF is significantly different from those at this melting temperature obtained is higher than the experimental

Figure 4. Selected radial distribution functions of LiF, Li2 O and Li2 CO3 at various temperatures. For LiF at 300 K, 1000 K and 1200 K: (a) Li-Li, (b) F-F, and
(c) Li-F. For Li2 O at 300 K, 1400 K and 1800 K: (d) Li-Li, (e) O-O, and (f) Li-O. For Li2 CO3 at 300 K, 1200 K and 1800 K: (g) Li-Li, (h) C-C, (i) O-O, (j) Li-C,
(k) Li-O, and (l) C-O. Snapshots used to calculate RDF’s are taken every 2 ps during 1000 ps temperature equilibrations using NPT ensemble.
Journal of The Electrochemical Society, 164 (11) E3159-E3170 (2017) E3165

Table IV. Diffusion Coefficients for Li-ion in LiF, Li2 O and


Li2 CO3 .

Diffusion Coefficient (D) (m2 /s)

Temperature (K) LiF Li2 O Li2 CO3


250 6.83 × 10−17 1.71 × 10−18 1.00 × 10−16
275 1.12 × 10−15 1.74 × 10−16 7.11 × 10−16
300 3.93 × 10−16 4.01 × 10−16 3.30 × 10−16
325 1.22 × 10−15 5.30 × 10−16 9.14 × 10−15
350 7.10 × 10−16 6.92 × 10−16 6.73 × 10−14
375 4.64 × 10−17 1.10 × 10−16 1.08 × 10−15
400 1.07 × 10−15 3.63 × 10−16 1.68 × 10−14

value of 996 K.120,121 At 1800 K the RDF curves are clearly different
than the curves at lower temperatures.

Diffusion Coefficients and Diffusion Mechanism at Low and


High Temperatures
Diffusion coefficients at low temperatures.—After vacancies and
interstitial are introduced in each SEI product, the lithium ion dif-
fusion coefficients for each case are obtained from the slope of the
MSD curves. Coefficient values for each temperature below 400 K are
tabulated in Table IV. At 300 K, for Li+ diffusion in LiF it is D = 3.93
× 10−16 m2 /s which is in the range of values (10−26 m2 /s and 10−16
m2 /s) that have been reported for diffusion of Li in LiF found using
phase-field, molecular dynamics and NEB methods.21,73,84 Similarly,
the Li-ion diffusion in Li2 O is D = 4.01 × 10−16 m2 /s at 300 K, which
is also within the range (10−20 m2 /s and 10−16 m2 /s) of published the-
oretical values of diffusivity of Li in Li2 O obtained also by phase-field
method and molecular dynamics.21,84 For Li2 CO3 at 300 K, D = 3.3 ×
10−16 m2 /s, which is one order of magnitude smaller than the reported
values (10−15 and 10−12 m2 /s) of diffusion of Li in Li2 CO3 found by
NEB and molecular dynamics studies.9,84 Moreover, all of the results
at 300 K are within the range of values (10−18 and 10−14 m2 /s) of
diffusion of Li in amorphous silicon that have been experimentally
Figure 5. Li-ion diffusion in (a) LiF, (b) Li2 O and (c) Li2 CO3 at high tem-
found.44,122,123
peratures. Initial (left) and final (right) position of Li-ions. Li (purple, blue,
yellow), F (cyan), O (red), C (brown).
Diffusion mechanisms.—At low temperatures, mainly the atom
vibrations due to the increased temperature are observed and no dif-
fusion hopping events during the simulation time are seen. On the where D0 is a pre-exponential factor for the diffusion coefficient, Ea is
contrary, at high temperatures, Li-ion diffusion can be clearly ob- the diffusion activation energy, T is the temperature and R is the ideal
served. Sample snapshots of Li-ion diffusion are presented in Figure 5. gas constant. The activation energy Ea can be found from the nega-
Predominant mechanisms in the simulations include both vacancy tive slope of the Arrhenius plot, and the intercept of the line can be
assisted and knock-off diffusion in LiF. This agrees with previous used to determine the temperature-independent pre-factor D0 . Figure 6
theoretical results that found vacancy diffusion more favorable than shows the ln(D) as function of 1/T obtained from the diffusion coeffi-
interstitial diffusion.73 The diffusion mechanism observed in Li2 O is cient results. First, at temperatures from 250 K to 650 K, the diffusion
direct exchange, where two neighboring ions move simultaneously. A coefficients are very low; a dependency on temperature is not ob-
similar mechanism, ring diffusion, was found in a recent study.96 Va- served. As mentioned before, only atom vibrations are observed in
cancy and knock-off diffusion mechanisms are observed in Li2 CO3 ; the simulations.
in contrast, a previous theoretical study found that interstitial Li+ ions Second, at high temperatures, it can be observed in all cases, LiF,
are the main carriers,74 but also that this interstitial Li+ moves via Li2 O and Li2 CO3, that the diffusion coefficients increase as the temper-
knock off mechanism. In LiF case, shown in Figure 5a, a Li-ion (blue) ature increases. Moreover, the LiF and Li2 O results are comparable to
displaces a Li-ion (yellow) that is able to move to a neighbor va- the experimental values and other MD results.77,88,92–95 Furthermore,
cant site. In Li2 O, presented in Figure 5b, direct exchange, where two the diffusion coefficients of Li+ at high temperatures ranged from
neighboring ions (blue and yellow) move simultaneously, is observed. 700 K to 1800 K are higher than the low temperature results by at
Figure 5c shows the knock-off mechanism observed in Li2 CO3 , where least five orders of magnitude; this is expected since the movement of
a Li interstitial ion (blue) removes a Li-ion (yellow) from its original atoms is usually higher in liquid phase than in solid phase. The dif-
site. fusion pre-exponential factor (D0 ) and activation energy Ea obtained
from the linear fit of the Arrhenius plot for all SE products studied
Diffusion coefficient in Arrhenius plot and fitting parameters.— are summarized in Table V. Results at low temperature show that the
The temperature dependence of diffusion coefficients follows the Ar- activation energy in LiF is the smallest of the three samples studied,
rhenius equation124 given by followed by Li2 O then Li2 CO3 . The results for the LiF and Li2 CO3
activation energy, Ea = 0.04 eV and Ea = 0.12 eV, are below the val-
D = D0 e−Ea /RT [7] ues obtained experimentally and by other MD studies.6,8,9,75,76 On the
E3166 Journal of The Electrochemical Society, 164 (11) E3159-E3170 (2017)

Figure 6. The ln(D) versus 1/T for (a) LiF, (b)Li2 O, and (c) Li2 CO3 . (d) Diffusion coefficient versus applied electric field for all SEI products. Results from
molecular dynamics (solid circles) and experimental studies (solid lines) previously reported by others are shown for comparison in (a) LiF,77, 84 (b) Li2 O,84,95 and
(c) Li2 CO3. 84

contrary, the Li2 O energy result, Ea = 0.22 eV, is within the already re- due to the fact that less diffusion events occur and it becomes more
ported values.8,75 At high temperatures, LiF still has activation energy, difficult to calculate the actual diffusion coefficients since the error in
Ea = 0.32 eV, lower than Li2 O, Ea = 2.3 eV, even above 1300 K where the calculation increases as the temperature decreases.
Li2 O presents superionic conductivity. In addition, the LiF and Li2 O
results for the diffusion pre-factor (D0 ) are within one order of magni-
tude of the published values, and the results for the activation energy Diffusion Coefficients and Diffusion Mechanisms at Room
are close to values obtained by other studies.77,95 For Li2 CO3 case, in Temperature under an Electric Field
the temperature range of 1300 K to 1800 K the activation energy is Ea Diffusion coefficients.—We then apply an electric field to the
= 1.34 eV. Moreover, the slope changes at approximately 650 K and sample and calculate the diffusion coefficients at 300 K. As the electric
1000 K which corresponds to a phase transformation and melting of field is increased, Li-ion diffusion exponentially increases as observed
Li2 CO3 .120,121 To the best of our knowledge, no studies have reported in Figure 6d. Li-ion diffusion is clearly observed in the simulations at
on diffusion of Li+ in Li2 CO3 at high temperatures. In summary, our
0.4V/Å in Li2 CO3 and at 0.7 V/Å in both LiF and Li2 O. Coefficients
results at high temperatures are closer to those from previous works.
at these fields are at least five orders of magnitude greater than those
However, in the case of low temperatures, the discrepancies can be
where no field is applied (Table IV). Shi et al.9 reported coefficients

Table V. Diffusion pre-factor (D0 ) and diffusion activation energy (Ea ) for LiF, Li2 O and Li2 CO3 at low temperatures and high temperatures.

Temperature Range SEI Product Reference D0 (m2 /s) Ea (eV)


Low Temperature 250 K-400 K LiF This work 1.8 × 10−15 0.04
Other8,75,76 0.65-0.80
Li2 O This work 5.4 × 10−13 0.22
Other8,75 0.15-0.34
Li2 CO3 This work 7.6 × 10−13 0.12
Other6,8,9,74,101 0.23-0.80
High Temperature 1300 K–1800 K LiF This work 3.5 × 10−7 0.32
Other77,83 1.6 × 10−8 0.25-1.86
Li2 O This work 3.1 × 10−2 2.30
Exp.95 4.1 × 10−1 2.5
Li2 CO3 This work 7.3 × 10−6 1.34
Journal of The Electrochemical Society, 164 (11) E3159-E3170 (2017) E3167

mechanism in Li2 CO3 fully agrees with Shi et al.9 who found that
above ∼4 V, vacancy diffusion dominates Li-ion diffusion but diffu-
sion through interstitials is also energetically favorable; vacancy and
interstitial formation energies are ∼0.6 eV and 1 eV, respectively.
Notice that 4 V corresponds to an electric field of ∼0.17 V/Å in our
Li2 CO3 box as its length in the field direction is 24 Å.

Energy and temperature profiles.—We also analyze the total en-


ergy and temperature profiles for each of the SEI samples at different
applied electric fields. In the case of LiF (Figure 8a, 8b), we observe
that the energy slightly increases as the electric field is increased to 0.6
V/Å, then at 0.7 V/Å, 0.8 V/Å and 0.85 V/Å a significant increase in
the energy occurs, which goes from −73 Gcal/mol to −68 Gcal/mol.
An increase in the temperature of the sample is also observed when 0.7
V/Å, 0.8 V/Å and 0.85 V/Å electric fields are applied to the sample.
In the case of the Li2 O (Figure 8c, 8d), a large variation in the
energy at the beginning of our simulation is also observed, as explained
previously. Then at 300 K, when the electric field is applied the energy
remains at −113 Gcal/mol up to 0.7 V/Å, then at 0.8 V/Å and 0.85 the
energy jumps to −110 Gcal/mol. An increase in the temperature of
the sample is also observed when 0.8 V/Å and 0.85 V/Å are applied
to the sample.
In the case of Li2 CO3 (Figure 8e, 8f), the energy of the samples
remain at −60 Gcal/mol up to 0.2 V/Å, then from 0.4 V/Å to 0.85 V/Å
the energy decreases to and stays at −80 Gcal/mol. The temperature
of the sample stays at 300 K up to 0.4 V/Å then from 0.5 V/Å to 0.85
V/Å, it slowly increases from 300 K to 500 K. When comparing the
results of the SEI products analyzed, it can be observed that in the LiF
and Li2 O cases the energy of the sample increases as the electric field
applied is increased. However, in the Li2 CO3 case, the contrary occurs
with the energy of the sample decreasing when an electric field is ap-
plied. This may be due to the defect types created in the sample. In the
LiF and Li2 O cases vacancies are introduced whereas in the Li2 CO3
sample an interstitial Li atom with a positive charge is introduced.

Effect of electric field in atom pairs distribution curves.—Se-


Figure 7. Snapshots of initial (left) and final (right) positions of Li+ ions lected representative RDF curves for the different atom pairs of
to show diffusion displacement in (a) LiF, (b) Li2 O and (c) Li2 CO3 when
an electric field is applied. Li+ ions (multiple colors), F (cyan), O (red), C
Li2 CO3 at various applied electric fields are presented in Figure 9.
(brown). These distribution plots are taken during a 100 ps temperature equi-
libration at 300 K in different simulation runs for each applied field.
Each atom pair distribution plot includes curves for 0, 0.4, and 0.85
close to 10−12 m2 /s for Li2 CO3 at a voltage range from 0 to 4.4 V. For V/Å in black, red and blue respectively. In all atom pair cases, RDF
the same voltage range, which corresponds to electric fields from 0 curves for 0.1 and 0.2 V/Å are equal to the 0 V/Å curve. Charac-
to 0.18 V/Å, the diffusion coefficient values are found to be between teristic peaks representing the crystalline structure are observed, and
10−14 and 10−13 m2 /s. the intensity of the peaks and their widths do not change as the field
is increased. At 0.4 V/Å, the RDF curves are clearly different from
Diffusion mechanisms.—Diffusion events observed in our simu- those at lower electric fields. In the case of Li-Li, O-O and Li-O, the
lations consist of Li+ ions moving mainly via knock-off diffusion in peaks can still be observed yet they are broader and their intensity
LiF, where multiple Li+ ions knock-off their nearest neighbors in a is decreased. In the rest of the atom pairs, C-C, Li-C, and C-O, the
cascade manner in the direction of the electric field. In the LiF case, changes are more significant. In C-C distribution plot the characteris-
shown in Figure 7a, first a Li+ ion (green) starts a chain movement tic peaks at 3.5, 4.3 and 5 Å combine and make a broad peak at 4.3 Å,
of Li+ ions (purple, orange and yellow) as indicated by the top set of similarly the peak at 7.6 Å combines with smaller peaks at 6.9 and 8.6
arrows; then in a subsequent simulation step, another Li+ ion (red) Å and form a very broad peak at 7.9 Å. In Li-C, the intensity of the first
starts a second chain of displacements as indicated by the bottom set
peak at 2.3 Å increases and the intensity of the second peak at 2.8 Å
of arrows. Li+ ions move in Li2 O from one lattice site to an empty
decreases; the rest of the peaks combine and make broad peaks. In the
one in sequence, as well as multiple Li+ ions move simultaneously
in the direction of the applied field. Shown in Figure 7b, first a Li+ C-O distribution plot the characteristic peak at 1.29 Å remains very
ion (top orange and bottom yellow) moves to a vacant site, conse- similar, only a small decrease in intensity is observed; individual peaks
quently the other Li+ ions move, both individually and in pairs, in at 3 Å, 3.75 Å and 4.4 Å combine a form a broad peak at 3.83; sim-
the direction of the electric field. No studies were found in the litera- ilarly, the rest of the peaks in the distribution plot combine and form
ture that study LiF and Li2 O diffusion mechanisms under an applied broader peaks. As the electric field is increased from 0.4 to 0.85 V/Å,
electric field. In Li2 CO3 , the predominant mechanism observed is a no significant changes occur; only a very small decrease in the first
combined vacancy-interstitial diffusion, shown in Figure 7c. First a peak intensity is observed in Li-Li, Li-C and Li-O distribution plots.
Li+ ion (yellow) moves slightly from its lattice position thus allowing
another Li+ ion (blue) to easily take its place and further displace it
Conclusions
to an interstitial position; then in a subsequent step another Li+ ion
(green) moves to a vacant site, and the Li-ion in the interstitial position Classical molecular dynamics simulations were used to obtain the
(yellow) moves accordingly as indicated in the arrows. The observed diffusion coefficients of Li-ion in the three main components of the
E3168 Journal of The Electrochemical Society, 164 (11) E3159-E3170 (2017)

Figure 8. Energy versus time curves (left) and Temperature versus time curves (right) for samples with applied electric fields in the range from 0 V/Å to 0.85
V/Å. (a-b) LiF with a vacancy defect ratio of 2/1000, (c-d) Li2 O with a vacancy defect ratio of 3/798, and (e-f) Li2 CO3 with interstitial defect ratio of 1/864.

solid electrolyte interphase found in Li-ion batteries. Suitable force energies (Ea ) and pre-exponential factors (D0 ) were obtained. Activa-
fields were successfully evaluated by comparing the resulting bond tion energies found at low temperatures are close to those previously
distances of each SEI structure at 300 K. In all SEI compounds the obtained by others experimentally and theoretically. At high tempera-
resulting bond distances were within ±0.2 Å of experimental results. tures the pre-factor values and activation energies obtained were 0.32,
Then the radial distribution functions of atom pairs in each structure 2.3, and 1.34 eV for LiF, Li2 O and Li2 CO3 , respectively. These re-
were calculated in a wide range of temperatures from 250 K to 1800 sults are very close to the available literature results. Moreover, the
K. Melting temperatures obtained for both LiF and Li2 CO3 were very predominant diffusion mechanisms observed were vacancy assisted
close to the experimental results; the result for Li2 O was higher than and knock-off diffusion in LiF, direct exchange in Li2 O, and vacancy
the experimental value. The mean square displacements of Li-ions in and knock-off in Li2 CO3 . These findings generally agree with previous
LiF, Li2 O and Li2 CO3 were obtained and the diffusion coefficients theoretical results. The effect of an applied electric field in the dif-
over the operating temperature range of 250 K to 400 K for LIB were fusion coefficients and the diffusion mechanisms was also analyzed.
found. Since most of the experimental diffusion measurements are Diffusion coefficients increased exponentially with the increase of
done at high temperature, diffusion coefficients at temperatures rang- the electric field, and diffusion of Li-ions observed under the applied
ing from 600 K to 1800 K were also found. Simulation results showed electric field occurred via knock-off in LiF, and via vacancies in Li2 O
that Li-ion diffusion coefficients at 300 K are 3.93 × 10−16 m2 /s, and in Li2 CO3 . Whereas no studies were found of Li-ion diffusion in
4.01 × 10−16 m2 /s, and 4.90 × 10−17 m2 /s for LiF, Li2 O and Li2 CO3 , LiF or Li2 O under an applied electric field, the diffusion coefficient
respectively. These results are comparable with available experimen- and mechanism obtained in Li2 CO3 case fully agrees with previous
tal data and other computational results. In addition, the diffusion theoretical findings. Previously mentioned studies calculate diffusion
coefficients were fitted to the Arrhenius equation, and the activation coefficients only at room temperature, or only at very high tempera-
Journal of The Electrochemical Society, 164 (11) E3159-E3170 (2017) E3169

Figure 9. Selected radial distribution functions of Li2 CO3 at various applied electric fields. (a) Li-Li, (b) C-C, (c) O-O, (d) Li-C, (e) Li-O, and f) C-O. Snapshots
used to calculate RDF’s are taken every 0.1 ps during a 100 ps temperature equilibration at 300 K using NPT ensemble.

tures, this work provides coefficients over a wider temperature range 18. M. Nie, D. Chalasani, D. P. Abraham, Y. Chen, A. Bose, and B. L. Lucht, J. Phys.
from 250 K to 1800 K. Furthermore, the evaluated force field param- Chem. C, 117, 1257 (2013).
19. J. Zhang, R. Wang, X. Yang, W. Lu, X. Wu, X. Wang, H. Li, and L. Chen, Nano
eters can be used in this range of temperatures to further study lithium Lett., 12, 2153 (2012).
ion transport in structures combining two or more SEI products using 20. J. Zheng, H. Zheng, R. Wang, L. Ben, W. Lu, L. Chen, L. Chen, and H. Li, Phys.
classical molecular dynamics simulations. Chem. Chem. Phys., 16, 13229 (2014).
21. P. Guan, L. Liu, and X. Lin, J. Electrochem. Soc., 162, A1798 (2015).
22. Y. Wang, S. Nakamura, M. Ue, and P. B. Balbuena, J. Am. Chem. Soc., 123, 11708
Acknowledgments (2001).
23. P. Arora, R. E. White, and M. Doyle, J. Electrochem. Soc., 145, 3647 (1998).
We acknowledge the support of the Texas A&M Supercomputer 24. D. Aurbach, A. Zaban, A. Schechter, Y. Ein-Eli, E. Zinigrad, and B. Markovsky, J.
Electrochem. Soc., 142, 2873 (1995).
Facility and the Texas Advanced Computing Center, as well as the 25. M. B. Pinson and M. Z. Bazant, J. Electrochem. Soc., 160, A243 (2013).
financial support from the Assistant Secretary for Energy Efficiency 26. H. Wu, G. Chan, J. W. Choi, I. Ryu, Y. Yao, M. T. McDowell, S. W. Lee, A. Jackson,
and Renewable Energy, Office of Vehicle Technologies of the U.S. Y. Yang, L. Hu, and Y. Cui, Nat. Nanotechnol., 7, 310 (2012).
Department of Energy under Contracts No. DE-EE0007766 and No. 27. M. Winter, Z. Phys. Chem., 223, 1395 (2009).
28. T. Yoshida, M. Takahashi, S. Morikawa, C. Ihara, H. Katsukawa, T. Shiratsuchi, and
DE-AC02-05CH11231, Subcontract 7060634, under the Advanced J. I. Yamaki, J. Electrochem. Soc., 153, A576 (2006).
Battery Materials Research (BMR) Program. Stimulating discussions 29. K. Edström, M. Herstedt, and D. P. Abraham, J. Power Sources, 153, 380 (2006).
on molecular simulations and research in general with Narendra Ku- 30. J. M. Tarascon and M. Armand, Nature, 414, 359 (2001).
mar and Karim Salazar as well as a reading of this manuscript by 31. P. P. R. M. L. Harks, F. M. Mulder, and P. H. L. Notten, J. Power Sources, 288, 92
(2015).
Tillman Austin are also acknowledged. 32. D. Aurbach, Y. Ein-Eli, O. Chusid, Y. Carmeli, M. Babai, and H. Yamin, J. Elec-
trochem. Soc., 141, 603 (1994).
33. L. Benitez, D. Cristancho, J. M. Seminario, J. M. Martinez de la Hoz, and
References P. B. Balbuena, Electrochim. Acta, 250 (2014).
34. F. A. Soto, Y. Ma, J. M. Martinez De La Hoz, J. M. Seminario, and P. B. Balbuena,
1. M. T. McDowell, S. W. Lee, W. D. Nix, and Y. Cui, Adv. Mater., 25, 4966 (2013). Chem. Mater., 27, 7990 (2015).
2. D. Aurbach, K. Gamolsky, B. Markovsky, G. Salitra, Y. Gofer, U. Heider, R. Oesten, 35. A. L. Michan, M. Leskes, and C. P. Grey, Chem. Mater., 28, 385 (2016).
and M. Schmidt, J. Electrochem. Soc., 147, 1322 (2000). 36. G. M. Veith, M. Doucet, J. K. Baldwin, R. L. Sacci, T. M. Fears, Y. Wang, and
3. S. Matsuta, Y. Kato, T. Ota, H. Kurokawa, S. Yoshimura, and S. Fujitani, J. Elec- J. F. Browning, J. Phys. Chem. C, 119, 20339 (2015).
trochem. Soc., 148, A7 (2001). 37. E. Rejovitzky, C. V. Di Leo, and L. Anand, J. Mech. Phys. Solids, 78, 210
4. M. Moshkovich, M. Cojocaru, H. E. Gottlieb, and D. Aurbach, J. Electroanal. (2015).
Chem., 497, 84 (2001). 38. A. M. Andersson and K. Edström, J. Electrochem. Soc., 148, A1100 (2001).
5. J. Ufheil, A. Würsig, O. D. Schneider, and P. Novák, Electrochem. Commun., 7, 39. F. Kong, R. Kostecki, G. Nadeau, X. Song, K. Zaghib, K. Kinoshita, and
1380 (2005). F. McLarnon, J. Power Sources, 97–98, 58 (2001).
6. H. Iddir and L. A. Curtiss, J. Phys. Chem. C, 114, 20903 (2010). 40. A. M. Andersson, D. P. Abraham, R. Haasch, S. MacLaren, J. Liu, and K. Amine,
7. L. Yang, B. Ravdel, and B. L. Lucht, Electrochem. Solid-State Lett., 13, A95 (2010). J. Electrochem. Soc., 149, A1358 (2002).
8. Y. C. Chen, C. Y. Ouyang, L. J. Song, and Z. L. Sun, J. Phys. Chem. C, 115, 7044 41. X. H. Liu and J. Y. Huang, Energy Environ. Sci., 4, 3844 (2011).
(2011). 42. P. Lu and S. J. Harris, Electrochem. Commun., 13, 1035 (2011).
9. S. Shi, Y. Qi, H. Li, and L. G. Hector, J. Phys. Chem. C, 117, 8579 (2013). 43. I. V. Veryovkin, C. E. Tripa, A. V. Zinovev, S. V. Baryshev, Y. Li, and D. P. Abraham,
10. D. Aurbach and A. Zaban, J. Electroanal. Chem., 348, 155 (1993). Nucl. Instrum. Methods Phys. Res., Sect. B, 332, 368 (2014).
11. D. Aurbach, B. Markovsky, A. Shechter, Y. Ein-Eli, and H. Cohen, J. Electrochem. 44. J. Xie, N. Imanishi, T. Zhang, A. Hirano, Y. Takeda, and O. Yamamoto, Mater.
Soc., 143, 3809 (1996). Chem. Phys., 120, 421 (2010).
12. D. Aurbach, M. D. Levi, E. Levi, and A. Schechter, J. Phys. Chem. B, 101, 2195 45. G. A. Tritsaris, K. Zhao, O. U. Okeke, and E. Kaxiras, J. Phys. Chem. C, 116, 22212
(1997). (2012).
13. P. B. Balbuena and Y. Wang, Lithium-Ion Batteries: Solid-Electrolyte Interphase, 46. H. Wang, X. Ji, C. Chen, K. Xu, and L. Miao, AIP Advances, 3, 112102 (2013).
London (2004). 47. P. M. Panchmatia, A. R. Armstrong, P. G. Bruce, and M. S. Islam, Phys. Chem.
14. H. L. Zhang, F. Li, C. Liu, J. Tan, and H. M. Cheng, J. Phys. Chem. B, 109, 22205 Chem. Phys., 16, 21114 (2014).
(2005). 48. Z. Wang, Q. Su, H. Deng, W. He, J. Lin, and Y. Q. Fu, J. Mater. Chem. A, 2, 13976
15. Y. M. Lee, J. Y. Lee, H. T. Shim, J. K. Lee, and J. K. Park, J. Electrochem. Soc., (2014).
154, A515 (2007). 49. M. Wu, B. Xu, and C. Ouyang, Chin. Phys. B, 25, 018206 (2015).
16. P. Verma, P. Maire, and P. Novák, Electrochim. Acta, 55, 6332 (2010). 50. M. D. Bhatt, M. Cho, and K. Cho, Modell. Simul. Mater. Sci. Eng., 20, 065004
17. X. Xiao, P. Lu, and D. Ahn, Adv. Mater., 23, 3911 (2011). (2012).
E3170 Journal of The Electrochemical Society, 164 (11) E3159-E3170 (2017)

51. M. Park, X. Zhang, M. Chung, G. B. Less, and A. M. Sastry, J. Power Sources, 195, 87. T. X. T. Sayle, P. E. Ngoepe, and D. C. Sayle, J. Mater. Chem., 20, 10452 (2010).
7904 (2010). 88. T. Oda and S. Tanaka, J. Nucl. Mater., 386–388, 1087 (2009).
52. J. Yu, P. B. Balbuena, J. Budzien, and K. Leung, J. Electrochem. Soc., 158, A400 89. T. Oda, Y. Oya, S. Tanaka, and W. J. Weber, J. Nucl. Mater., 367–370A, 263 (2007).
(2011). 90. M. Hayoun, M. Meyer, and A. Denieport, Acta Mater., 53, 2867 (2005).
53. P. Ganesh, D. E. Jiang, and P. R. C. Kent, J. Phys. Chem. B, 115, 3085 (2011). 91. M. Wilson, S. Jahn, and P. A. Madden, J. Phys.: Condens. Matter, 16, S2795 (2004).
54. O. Borodin and G. D. Smith, J. Phys. Chem. B, 110, 4971 (2006). 92. P. Goel, N. Choudhury, and S. L. Chaplot, Phys. Rev. B, 70, 174307 (2004).
55. Y. Wang and P. B. Balbuena, Int. J. Quantum Chem, 102, 724 (2005). 93. J. G. Rodeja, M. Meyer, and M. Hayoun, Modell. Simul. Mater. Sci. Eng., 9, 81
56. Y. Wang and P. B. Balbuena, J. Phys. Chem. B, 108, 15694 (2004). (2001).
57. A. Van Der Ven and G. Ceder, Electrochem. Solid-State Lett., 3, 301 (2000). 94. R. M. Fracchia, G. D. Barrera, N. L. Allan, T. H. K. Barron, and W. C. Mackrodt,
58. X. Bogle, R. Vazquez, S. Greenbaum, A. V. W. Cresce, and K. Xu, J. Phys. Chem. J. Phys. Chem. Solids, 59, 435 (1998).
Lett., 4, 1664 (2013). 95. Y. Oishi, Y. Kamei, M. Akiyama, and T. Yanagi, J. Nucl. Mater., 87, 341 (1979).
59. O. Borodin and G. D. Smith, J. Phys. Chem. B, 113, 1763 (2009). 96. A. D. Mulliner, P. C. Aeberhard, P. D. Battle, W. I. F. David, and K. Refson, Phys.
60. M. T. Ong, O. Verners, E. W. Draeger, A. C. T. Van Duin, V. Lordi, and J. E. Pask, Chem. Chem. Phys., 17, 21470 (2015).
J. Phys. Chem. B, 119, 1535 (2015). 97. M. Bruno and M. Prencipe, Surf. Sci., 601, 3012 (2007).
61. K. Persson, V. A. Sethuraman, L. J. Hardwick, Y. Hinuma, Y. S. Meng, 98. Y. Duan and D. C. Sorescu, Phys. Rev. B: Condens. Matter Mater. Phys., 79, 014301
A. Van Der Ven, V. Srinivasan, R. Kostecki, and G. Ceder, J. Phys. Chem. Lett., (2009).
1, 1176 (2010). 99. Y. N. Zhuravlev and I. A. Fedorov, J. Struct. Chem., 47, 206 (2006).
62. A. V. Churikov, A. V. Ivanishchev, A. V. Ushakov, and V. O. Romanova, J. Solid 100. I. A. Fedorov, Y. N. Zhuravlev, and D. V. Korabel’nikov, Russ. Phys. J., 49, 1106
State Electrochem., 18, 1425 (2014). (2006).
63. A. Van der Ven, J. Bhattacharya, and A. A. Belak, Acc. Chem. Res., 46, 1216 (2013). 101. M. T. Dunstan, J. M. Griffin, F. Blanc, M. Leskes, and C. P. Grey, J. Phys. Chem.
64. M. Schroeder, C. Eames, D. A. Tompsett, G. Lieser, and M. S. Islam, Phys. Chem. C, 119, 24255 (2015).
Chem. Phys., 15, 20473 (2013). 102. S. -L. Shang, L. G. Hector Jr, S. Shi, Y. Qi, Y. Wang, and Z. -K. Liu, Acta Mater.,
65. J. -C. Soetens, C. Millot, and B. Maigret, J. Phys. Chem. A, 102, 1055 (1998). 60, 5204 (2012).
66. O. Borodin and D. Bedrov, J. Phys. Chem. C, 118, 18362 (2014). 103. Y. Ji, Y. Zhang, and C. Y. Wang, J. Electrochem. Soc., 160, A636 (2013).
67. S. Ogata, N. Ohba, and T. Kouno, J. Phys. Chem. C, 117, 17960 (2013). 104. A. V. Churikov, Electrochim. Acta, 46, 2415 (2001).
68. K. Leung, J. Phys. Chem. C, 117, 1539 (2013). 105. J. Christensen and J. Newman, J. Electrochem. Soc., 151, A1977 (2004).
69. O. Borodin, G. V. Zhuang, P. N. Ross, and K. Xu, J. Phys. Chem. C, 117, 7433 106. S. Shi, J. Gao, Y. Liu, Y. Zhao, Q. Wu, W. Ju, C. Ouyang, and R. Xiao, Chin. Phys.
(2013). B, 25, 018212 (2015).
70. A. M. Colclasure, K. A. Smith, and R. J. Kee, Electrochim. Acta, 58, 33 (2011). 107. G. Bergerhoff and I. D. Brown, in Crystallographic Databases, F. H. Allen,
71. A. Márquez and P. B. Balbuena, J. Electrochem. Soc., 148, A624 (2001). G. Bergerhoff, and R. Sievers Editors, Chester, International Union of Crystal-
72. E. Peled, D. Golodnitsky, and G. Ardel, J. Electrochem. Soc., 144, L208 (1997). lography (1987).
73. H. Yildirim, A. Kinaci, M. K. Y. Chan, and J. P. Greeley, ACS Appl. Mater. Inter- 108. P. Politzer, J. S. Murray, and T. Clark, Phys. Chem. Chem. Phys., 15, 11178 (2013).
faces, 7, 18985 (2015). 109. T. Clark, M. Hennemann, J. S. Murray, and P. Politzer, J. Mol. Model., 13, 291
74. S. Shi, P. Lu, Z. Liu, Y. Qi, L. G. Hector, H. Li, and S. J. Harris, J. Am. Chem. Soc., (2007).
134, 15476 (2012). 110. J. S. Murray, P. Lane, and P. Politzer, J. Mol. Model., 15, 723 (2009).
75. Y. Koyama, Y. Yamada, I. Tanaka, S. R. Nishitani, H. Adachi, M. Murayama, and 111. S. Kale and J. Herzfeld, J. Chem. Theory Comput., 7, 3620 (2011).
R. Kanno, Mater. Trans., 43, 1460 (2002). 112. A. H. Mao and R. V. Pappu, J. Chem. Phys., 137, 064104 (2012).
76. J. Maier, Angew. Chem. Int. Ed., 32, 313 (1993). 113. K. Tasaki and S. J. Harris, J. Phys. Chem. C, 114, 8076 (2010).
77. H. Luo, S. Xiao, S. Wang, P. Huai, H. Deng, and W. Hu, Comp. Mater. Sci., 111, 114. A. K. Rappe, C. J. Casewit, K. S. Colwell, W. A. Goddard, and W. M. Skiff, J. Am.
203 (2016). Chem. Soc., 114, 10024 (1992).
78. M. Kobelev, Mol. Simul., 39, 868 (2013). 115. S. Plimpton, J. Comput. Phys, 117, 1 (1995).
79. C. Merlet, P. A. Madden, and M. Salanne, Phys. Chem. Chem. Phys., 12, 14109 116. W. Humphrey, A. Dalke, and K. Schulten, J. Mol. Graphics, 14, 33 (1996).
(2010). 117. O. Borodin, G. D. Smith, and P. Fan, J. Phys. Chem. B, 110, 22773 (2006).
80. O. L. Andreev, A. A. Raskovalov, and A. V. Larin, Russ. J. Phys. Chem. A, 84, 48 118. L. V. Gurvich, I. V. Veyts, and C. B. Alcock, Thermodynamic Properties of Individ-
(2010). ual Substances (1989).
81. V. Sarou-Kanian, A. L. Rollet, M. Salanne, C. Simon, C. Bessada, and P. A. Madden, 119. M. S. Ortman and E. M. Larsen, J. Am. Ceram. Soc., 66, 645 (1983).
Phys. Chem. Chem. Phys., 11, 11501 (2009). 120. M. A. K. L. Dissanayake and B. E. Mellander, Solid State Ionics, 21, 279 (1986).
82. G. Ciccotti, G. Jacucci, and I. R. McDonald, Phys. Rev. A, 13, 426 (1976). 121. C. W. Bale and A. D. Pelton, Calphad, 6, 255 (1982).
83. T. Stoebe and R. Huggins, J. Mater. Sci., 1, 117 (1966). 122. N. Ding, J. Xu, Y. X. Yao, G. Wegner, X. Fang, C. H. Chen, and I. Lieberwirth,
84. K. Tasaki, A. Goldberg, J. J. Lian, M. Walker, A. Timmons, and S. J. Harris, J. Solid State Ionics, 180, 222 (2009).
Electrochem. Soc., 156, A1019 (2009). 123. R. Ruffo, S. S. Hong, C. K. Chan, R. A. Huggins, and Y. Cui, J. Phys. Chem. C,
85. R. Asahi, C. M. Freeman, P. Saxe, and E. Wimmer, Modell. Simul. Mater. Sci. Eng., 113, 11390 (2009).
22, 075009 (2014). 124. H. Mehrer, Diffusion in Solids: Fundamentals, Methods, Materials, Diffusion-
86. J. Long, L. Yang, and D. Li, Solid State Sci., 19, 12 (2013). Controlled Processes (2007).

You might also like