You are on page 1of 12

Article

pubs.acs.org/cm

Fundamental Mechanisms of Solvent Decomposition Involved in


Solid-Electrolyte Interphase Formation in Sodium Ion Batteries
Hemant Kumar,† Eric Detsi,† Daniel P. Abraham,‡ and Vivek B. Shenoy*,†

Department of Materials Science and Engineering, University of Pennsylvania, Philadelphia, Pennsylvania 19104, United States

Chemical Sciences and Engineering Division, Argonne National Laboratory, Lemont, Illinois 60439, United States
*
S Supporting Information

ABSTRACT: Prolonged decomposition of electrolytes form-


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ing a thick and unstable solid-electrolyte interphase (SEI)


continues to be a major bottleneck in designing sodium-ion
batteries (SIBs). We have carried out quantum chemistry
Downloaded via ESS INFLIBNET PCA 1 on July 25, 2022 at 13:23:47 (UTC).

simulations to investigate the fundamental mechanisms of


reduction-induced decomposition of electrolyte solvents in the
vicinity of a sodium ion. Kinetics and thermodynamics of
several reaction pathways for one- and two-electron reduction
of ethylene carbonate (EC) have been examined. Our
calculations indicate that the high reduction potential and
low barrier for the ring opening of EC is the main cause for the
continuous growth of SEI observed in SIBs. The impact of two well-known electrolyte additives, vinyl carbonate (VC) and
fluoroethylene carbonate (FEC), on SEI composition was evaluated by studying decomposition pathways of (1) VC and FEC
molecules in the bulk EC solvent and (2) an EC molecule in a supermolecular cluster comprising an EC and the additive
molecule. The additive molecules have significantly low barriers for decomposition and therefore decompose first. Additionally,
the presence of an additive molecule was also shown to increase the barrier for decomposition of EC. Another observation
suggests that the preferred reduction state of an EC molecule changes when it forms a dimer with additive molecules, and these
reduction states have different decomposition pathways which leads to formation of different SEI compounds. On the basis of
these observations, we predict that not only do the additive molecules protect solvent molecules from reductive decomposition
but also they can promote alternate pathways for the decomposition, leading to qualitatively different and potentially stable SEI
products.

1. INTRODUCTION Another leading cause of the short cycle life is the formation of
Solid electrolyte interphase (SEI) is a crucial determinant of unstable SEI near the anode surface.6−8 Despite the similar
battery life.1 During the first few charging cycles of a battery, ionic nature of sodium and lithium ions, the SEI in SIBs has
organic solvents in the electrolyte are reduced near the anode. been shown to continuously evolve upon cycling, while it
The reduced solvent molecules decompose into chemically becomes stable after the first few cycles in LIBs.9 An
active radicals that form new chemical entities. Mosaic structure experimental study comparing the SEI structure between SIB
of these entities generates a film composed of inorganic and and LIB for a given electrolyte composition has shown that the
organic decomposition products called the SEI.2 With the right SEI in SIBs is thicker and unstable compared to the SEI in
composition, the SEI can impart kinetic stability to the LIBs.6 Furthermore, recent XPS experiments suggest that the
electrolyte against further reduction in successive cycles while SEI in SIBs has a higher percentage of inorganic salts, alkyl
allowing ionic mobility through it.3,4 The SEI acts to extend carbonates, and CO rich compounds compared to LIBs,
battery life by preventing further decomposition of the solvent which predominantly contain hydrocarbon chains
molecules and by hindering solvent intercalation into the (−CH2−).7,10 In addition, it has also been shown6 that the
battery anode.2 Although the SEI extends the battery life, the SEI in LIBs has a well-defined layered structure in contrast to a
formation of SEI also results in irreversible capacity loss due to relatively homogeneous structure in SIBs.
consumption of metal ions, a factor that plays an important role The chemical composition of the battery electrolyte
in determining the cycle life of batteries. determines the resulting structure and the passivating
Sodium-ion batteries (SIBs) are cost-effective alternatives to capabilities of the SEI. However, while the chemical
lithium-ion batteries (LIBs) but often suffer from short cycle composition of the electrolytes for LIBs has been greatly
lifetime,5 partly attributed to the relatively large size of the
sodium atom (Na radius 190 pm vs Li radius 167 pm), which Received: August 15, 2016
results in deleterious sodiation-induced volume changes, Revised: November 29, 2016
compared to the volume changes occurring during lithiation.1 Published: November 30, 2016

© 2016 American Chemical Society 8930 DOI: 10.1021/acs.chemmater.6b03403


Chem. Mater. 2016, 28, 8930−8941
Chemistry of Materials Article

optimized in recent years, the optimum electrolyte composition (Figure 1). The effect of the bulk solvent was accounted for by using a
for SIBs is still not known. Previous research has shown that continuum solvent model (C-PCM) (eps = 89.78, rsolv = 2.476) for
using the electrolyte compositions optimized for LIBs in SIBs
leads to the formation of unstable SEIs.11 Komaba et al. has
shown that an electrolyte containing ethylene carbonate (EC)
and diethyl carbonate (DEC) solvents retains a stable capacity
of 240 mAh g−1 after 100 cycles with hard carbon electrodes in
SIBs.12 Ponrouch et al. systematically investigated various
solvents (EC, propylene carbonate (PC), DEC, dimethyl
carbonate (DMC), tetrahydrofuran (THF), triglyme) and
sodium salts (NaClO4, NaPF6) for the optimum performance
of SIBs and concluded that a binary solvent consisting of EC
and PC with NaClO4 salt results in the best electrochemical
performance.13 Film-forming materials such as vinylene
carbonate (VC) and fluoroethylene carbonate (FEC) are
well-known additives in LIB and Li−Si batteries.14−18 Initial
studies have shown that the addition of FEC improves the cycle
life of SIBs as well.19 However, an electrolyte mixture for SIB Figure 1. Optimized geometries for different organic solvent and
that can provide a cycle life comparable to LIBs is still not additive molecules in the unreduced form.
known. Furthermore, the underlying reason for such character-
istically different electrolyte decomposition behaviors of these EC.28,29 Both the geometry optimization and frequency calculations
two battery systems (in spite of the similar electronic shell were carried out using the same level of DFT in conjugation with the
structure of their cations) is not understood. The final SEI continuum solvent model. This direct approach has been shown to
composition can be affected by a multitude of different factors, reproduce free energy values in solution phase comparable to the
such as the order of the bond breaking events, the relative thermodynamic cycle method.30 The role of the additives on EC
stability of resulting radical ions, and the thermodynamic decomposition was explored by considering the decomposition of the
feasibility of different decomposition reactions.20 Understand- supermolecular clusters composed of EC and other additive molecules
in the gas phase simulations. These simple theoretical calculations may
ing the decomposition mechanism of the electrolyte not incorporate the complex details of SEI growth conditions in
components in a reducing environment is central to the experiments, but accurate estimates of the thermodynamics and
rational design of SEI for optimum performance of the battery. barriers for various decomposition reactions help to better understand
The recent surge in experimental activities related to the SEI in the feasibility of different reactions and infer the nature of dominant
SIBs has brought important insights on the formation SEI products.
mechanism,21−25 but a mechanistic understanding is missing
due to a limited number of theoretical studies. 3. RESULTS
Therefore, we investigated the mechanism of reduction 3.1. Reduction Pathway for the Decomposition of EC.
induced decomposition of different alkyl carbonate molecules The reaction pathways for the reduction induced dissociation of
and additives in the vicinity of a sodium ion using the density the EC−Na+ complex are shown in Figure 2. The changes in
functional theory (DFT) method. Thermodynamics and the Gibbs free energy with respect to the unreduced structure
kinetics of reductive decomposition of the most commonly (E0) are also shown for all stationary points along the pathway.
used solvent molecule EC and additive molecules VC and FEC Two distinct reduction states with different values of reduction
were quantified using a hybrid functional DFT scheme. Using a free energies were found. The first reduction state corresponds
molecular cluster approach, we have further studied the effect to a structure where the excess electron is completely localized
of additive molecules (PC, VC, and FEC) on the decom- on the sodium ion. In the second reduction state, the excess
position pathways of the EC. In this study, we elucidate the electron is shared by the carbon and oxygen atoms of the
thermodynamics of SEI formation which enables us to make carbonyl group of the EC molecule. These two reduction states
predictions to improve the performance of SEIs in SIBs. are subsequently designated as A and B, respectively. Our
calculations show that the free energy of reduction for the state
2. COMPUTATIONAL METHODS A (−52.65 kcal/mol) is lower than the corresponding free
Structural optimization of cyclic carbonate molecules was performed
energy for the state B (−51.61 kcal/mol). This indicates that
using DFT with the hybrid B3PW91 exchange-correlation function as EA1 is the thermodynamically preferred reduction state. A
implemented in GAMESS.26 Gaussian basis set 6-311++G(d,p) was comparison with the corresponding free energy values in LIBs
employed to expand the Kohn−Sham orbitals in all the optimization (−45.8 kcal/mol) implies that the reduction of an EC molecule
calculations as well as for the single point energy calculations. This is easier in SIBs.31 The reduction follows the homolytic
hybrid functional DFT has been shown to accurately reproduce the dissociation of either the CCOC bond or the CEOC bond,
thermochemistry data for open-shell organic molecules.27 Free energy opening the five-membered ring of the EC molecule. The
of the decomposition reactions was quantified at T = 298.15 K and P = relative free energies of the transition states govern which of the
1 atm after taking into account entropy corrections obtained from two possible bond dissociation reactions dominates. The
frequency analysis. Zero-point energy (ZPE) corrections were
dissociation of the CEOC bond has two transition states
included in all energy values unless noted otherwise. Transition states
were verified by identifying the imaginary modes along the reaction indicated as EA2 and EB2 in Figure 2 for both of the reduction
coordinates and then performing intrinsic reaction coordinate (IRC) states A and B. The state A has a lower value of the bond
simulations. breaking barrier for reduction state A (4.50 kcal/mol (EA2)
First, reduction induced decompositions of individual molecules of compared to 9.31 kcal/mol in state B (EB2)), which suggests
EC, VC, and FEC were investigated in the vicinity of the Na+ ion that the CEOC bond can break easily in this reduction state.
8931 DOI: 10.1021/acs.chemmater.6b03403
Chem. Mater. 2016, 28, 8930−8941
Chemistry of Materials Article

Figure 2. Reaction pathways for the reduction induced decomposition of a EC−Na+ complex studied using the continuum solvent model (PCM).
Gibbs free energies at T = 298.15 K and P = 1 atm are given with respect to the unreduced structure (E0).

Since the reduction potential for the state A is higher and the On the basis of the above observations, we concluded that an
bond breaking barrier is lower compared to state B, CEOC bond EC molecule connected to a sodium cation is easily
breaking is likely to occur through the reduction state A. This decomposable compared to an EC molecule connected to a
behavior is in contrast to an EC−Li+ complex for which lithium cation. This result is consistent with experimental
reduction state B is the preferred state and the dissociation observations6 which show that the thickness of the SEI in SIBs
barrier is ∼9.5 kcal/mol.31 After the dissociation of the CEOC is larger compared to LIBs, indicating the prevalent
bond, both states A and B yield the same radical anion (E3) decomposition of solvent molecules near anodes. Experimental
coordinated with a sodium cation. Bond dissociation free studies have also found dissimilar proportions of organic and
energies corresponding to these exothermic reactions are inorganic compounds in these two battery systems. The SEIs in
−29.59 and −30.63 kcal/mol, respectively. In addition to the LIBs are mainly composed of carbon rich organic compounds
CEOC bond dissociation discussed above, another probable while, in SIBs, the SEIs mainly contain oxygen rich inorganic
pathway for the ring-opening reaction involves homolytic compounds.11,33 In order to determine what factors contribute
dissociation of a CCOC bond. However, we did not find any to this difference in the composition, we further examine the
stable equilibrium structures corresponding to the CCOC bond termination reactions of these radicals in the following section.
dissociation. This suggests that the CCOC bond dissociation is 3.2. SEI Components from the Decomposition of EC.
not feasible in a single electron reduction state of EC. This The anion radicals generated from the bond dissociation
reactions described above (E3 from CEOC bond) are highly
behavior is similar to the EC−Li+ complex studied by Leung.32
reactive and therefore are very likely to undergo secondary
2e Reduction of EC. The CCOC bond of an EC molecule
reactions. Various reactions leading to the termination of the
spontaneously dissociates when an extra electron is added to
anion radical (E3) are shown in Figure 3. Sodium butylene
the singly reduced state B. The anion radical (EB3) obtained
carbonate ([CH2CH2OCO2Na]2, Na2BDC) is one of the major
from this dissociation can further decompose in two different
components of the SEI and is generated as a result of the direct
ways, depending upon whether the remaining CCOC bond recombination of two E3 anion radicals (Figure 3a). Barrierless
(EB4) or CEOC bond (EB5) breaks. Thermodynamically, recombination of these radicals also results in a free energy gain
dissociation of the CCOC bond (ΔG= −0.78 kcal/mol), which of −54.89 kcal/mol. Another possible pathway for radical
results in the release of a CO gas molecule and formation of an termination is the nucleophilic attack of anion radical E3

OCH2CH2O− anion (EB4), is favorable compared to the (Figure 3b) on the oxygen of the neighboring E3 radical. This
dissociation of the CEOC (ΔG= 6.17 kcal/mol) bond, which r e a c t io n p r o d u c e s s o d i u m et h y l e n e d i c a r b o n a t e
results in the release of a CO2 gas molecule and formation of an ([CH2OCO2Na]2, Na2EDC) and ethylene (C2H4) gas.
·CH2CH2O− anion (EB5). These 2e reduction-induced Na2EDC is the most abundant chemical compound in a
decompositions are widely accepted as the primary reaction SEI,8,25 and its formation reduces the free energy by −43.81
mechanisms to explain the release of CO and CO2 gases in kcal/mol. In addition to these two direct recombination
LIBs.32 In our calculations, CCOC bond dissociation is reactions, the next pathway involves further reduction of the
spontaneous upon reduction and no energy barrier is observed, anion radical E3. After reduction, the spontaneous dissociation
while in the case of LIBs, a small barrier of ∼2.3 kcal/mol was of the remaining CEOC bond follows (Figure 3c). This reaction
observed by Leung.32 is also exothermic, and −35.46 kcal/mol of free energy is
8932 DOI: 10.1021/acs.chemmater.6b03403
Chem. Mater. 2016, 28, 8930−8941
Chemistry of Materials Article

the termination of this radical (E4) are explored here. In one


pathway, the radical E4 combines with a sodium ion (Figure
3e) to form Na2CO3, an inorganic salt experimentally observed
in SEIs. The second pathway involves a nucleophilic attack on
the oxygen of another EC−Na+ complex, which results in the
release of an ethylene gas molecule along with the formation of
a Na2EDC molecule (Figure 3f). Both of these termination
reactions are thermodynamically favorable and are equally
plausible as they produce similar free energy gains of −121.79
and −122.96 kcal/mol, respectively.
3.3. Reaction Pathway for the Decomposition of VC.
Reduction chemistry of VC in the vicinity of a sodium ion is
shown in Figure 4. Close inspection of the intermediate
reaction products shows that most of the reaction pathways
remain qualitatively similar to the case of EC. Like the EC−Na+
complex, VC−Na+ has two distinct reduction states A and B
(VA1 and VB1). However, in this case, B has a lower free
energy and is therefore the preferred reduction state. Various
decomposition pathways for this reduction state, i.e., the
dissociation of the CCOC bond (VB1−VB2−VB3) and the
dissociation of the CEOC bond (VB1−VB6−VB7), are
discussed below.
The free energy of reduction for a VC−Na+ complex
(−55.96 kcal/mol) is lower than that of an EC−Na+ complex
(−51.61 kcal/mol) in the equivalent reduction state B.
Figure 3. EC anion radical termination reactions and their Gibbs free However, there is a difference between the reductive
energy at T = 298.15 and P = 1 atm. decomposition of the VC and EC molecules. The homolytic
dissociation of a CCOC bond has an unfavorable reaction free
released. As a result, an ethylene gas molecule and NaCO−3 energy in the single electron reduction states of an EC−Na+
anion (E4) is produced. The resulting anion radical is highly complex but may occur readily for the VC−Na+ complex. The
reactive due to an unpaired electron and goes through energy barrier for this bond dissociation (VB2) is very low
subsequent reaction pathways. Two plausible pathways for (1.13 kcal/mol) implying a rapid decomposition of the VC

Figure 4. Reaction pathways for the reduction induced decomposition of VC−Na+ complex studied using a continuum solvent model (PCM). Gibbs
free energy at T = 298.15 K and P = 1 atm are given with respect to the unreduced structure (V0).

8933 DOI: 10.1021/acs.chemmater.6b03403


Chem. Mater. 2016, 28, 8930−8941
Chemistry of Materials Article

molecule after reduction. Although this bond dissociation is an energy release (−55.96 kcal/mol) due to the reduction of the
exothermic reaction, the thermodynamic gain from this intact VC−Na+ complex. Thus, once the radical is formed, it
decomposition (VB3) is quite small (ΔG = −1.53 kcal/mol) will get preferentially reduced. This radical remains stable after
compared to the ring opening (through CE O C bond the excess electron is reduced, as opposed to the EC radical
dissociation) in an EC−Na+ complex (ΔG = −30.63 kcal/ anion (E3) which spontaneously decomposes upon reduction
mol). The anion radical VB3 produced by this bond by the second electron. This metastable reduction state (VB9)
dissociation initiates further chemical reactions. Dissociation decomposes exothermically into CO32− and C2H2, and a
of the remaining CCOC bond of the VB3 radical yields an thermodynamic gain in free energy of −12.66 kcal/mol is

OCHCHO− anion (VB4) and a CO gas molecule. This observed.
exothermic decomposition reaction reduces the system free To summarize, the VC−Na+ complex decomposes readily
energy by −38.55 kcal/mol. Another route for the decom- after reduction. For both the 1e and 2e reduction states, the
position is the release of CO2 and the anion VB5 as a result of reaction mechanism involving the release of CO (VB1−VB3−
the CEOC bond dissociation of VB3. While this decomposition VB4, VB1−VB11−VB12) is the most feasible reaction from a
is also thermodynamically possible, the free energy gain is thermodynamic point of view. This leads to the formation of
comparatively small (ΔG = −12.32 kcal/mol). the −OCHCHO− anion which is known as an oligomerization
The next plausible decomposition in the reduction state B agent due to the presence of CHCH moieties. Another
involves a CEOC bond dissociation (VB7) of the VC molecule. decomposition pathway results in the formation of CO32− and
The gain in Gibbs free energy as a result of this reaction is C2H2 (or other organic dicarbonate, e.g., NaVDC). This
considerable (ΔG = −15.85 kcal/mol) compared to the CCOC pathway has large thermodynamic gain but is kinetically
bond dissociation discussed above. However, the energy barrier restricted due to a large free energy barrier and is therefore
for this reaction is very high (VB6, ΔGbarrier = 17.84 kcal/mol), less likely to contribute to SEI formation.
and therefore, the CCOC bond dissociation will kinetically 3.4. Reaction Pathway for the Decomposition of FEC.
dominate over the CEOC bond dissociation in reduction state B. The presence of a fluorine atom creates asymmetry in the five-
Because the anion radicals generated from these dissociations membered ring which in-turn leads to various possibilities for
have different reaction centers, they follow different reaction reduction states. In addition to the reduction states A and B
pathways and generate different decomposition products. discussed above, two additional states can be created by placing
Hence, the relative thermodynamic viability of these bond the sodium and fluorine atoms opposite the ring bisector.
dissociation reactions govern SEI composition. Furthermore, Na and F can be placed either on the same or on
We also investigated the pathways originating from the the opposite sides of the ring plane. Six distinct stable reduction
reduction state A (reduction free energy = −50.55 kcal/mol, structures arising from such permutations are shown in Figure
not shown in Figure 4). Homolytic dissociation of the CEOC 5. The relative free energy values with respect to the unreduced
bond in the reduction state A is also exothermic and reduces
the free energy by −21.25 kcal/mol. The energy barrier for this
reaction is 8.34 kcal/mol which is much smaller than 17.84
kcal/mol in reduction state B for the same bond dissociation. A
smaller energy barrier suggests that the dissociation of the
CEOC bond is easier in the reduction state A. However, the
dissociation of the CCOC bond is not feasible in this reduction
state. We discuss this issue in Section 3.5.
2e Electron Reduction of VC. In the initial phase of SEI
formation, electron tunneling rates are high and a VC molecule
may receive two electrons. The transfer of two electrons Figure 5. Different reduction states of the FEC−Na+ complex along
initiates spontaneous dissociation of the CCOC bond (VB11) of with the reduction free energy (in kcal/mol) at T = 298.15 K and P =
a VC molecule. This decomposition reaction has no energy 1 atm obtained using the PCM method.
barrier as opposed to the 1e electron reduction with an energy
barrier of 1.13 kcal/mol. The anion generated from this bond structure are also shown in Figure 5. The free energies of
dissociation decomposes further to produce CO and −OCH- reduction for four out of six FEC−Na+ structures are lower
CHO− (VB12) and lowers free energy by −58.56 kcal/mol. than that of an EC−Na+ complex. All structures corresponding
Another possibility of this anion decomposition via dissociation to the reduction state B have lower free energies compared to
of the CEOC bond produces CO2 and CHCHO− (VB13) anion. those of reduction state A, irrespective of the placement of the
However, this reaction is less exothermic, and the free energy is fluorine atom. However, for a given reduction state (A or B),
only lowered by −41.77 kcal/mol. A similar reaction route structures with sodium and fluorine placed on the same side of
releasing CO was also observed by Ushirogata et al.,34 while the ring bisector (e.g., Figure 5d,e) have a free energy lower
investigating two-electron pathways of VC decomposition in than those structures where sodium and fluorine are placed
LIBs using ab-initio molecular dynamics (AIMD) simulations. opposite the bisector by ∼1.5 kcal/mol. Among all the
However, they did not observe a pathway that releases CO2 as a structures, Figure 5d has the lowest reduction in free energy
result of the direct decomposition of VC. This might possibly and is therefore the most probable reduction state.
be related to the large kinetic barrier for the dissociation of the Decomposition pathways of FEC−Na+ in the 1e reduction
CEOC bond that was not overcome during the short AIMD states are shown in Figure 6. Unlike an EC−Na+ complex, the
trajectories they studied. CCOC bond of a FEC can easily break, even in the singly
Another two-electron pathway involves reduction of the reduced state with a small free energy barrier of 0.96 kcal/mol.
radical anion (VB7). Reduction of this radical lowers the free The radical anion generated from this bond dissociation
energy by −86.88 kcal/mol which is much lower than the free undergoes further decomposition, and two subsequent bond
8934 DOI: 10.1021/acs.chemmater.6b03403
Chem. Mater. 2016, 28, 8930−8941
Chemistry of Materials Article

Figure 6. Reaction pathways for the reduction induced decomposition of FEC−Na+ complex studied using continuum solvent model (PCM). Gibbs
free energy at T = 298.15 K and P = 1 atm is given with respect to the unreduced structure (F0). Numbers in blue correspond to the reduction state,
Figure 5f.

dissociation reactions are possible: a CEOC bond can break that leads to the formation of HF. It should also be noted that, for
releases a CO2 gas molecule and generates CHCHO− anion an EC molecule, these secondary reactions are thermodynami-
(FB4) or a CCOC bond can break that releases a CO gas cally feasible only in the 2e reduction state and do not occur in
molecule and generates a CH2CHO2− (FB5) anion. Thermo- the 1e reduction states.
dynamically, the reaction involving CO2 is more favorable (ΔG Like VC molecule decomposition, the alternative ring-
= −43.99 kcal/mol) than the reaction in which CO is released opening reaction through dissociation of a CEOC bond is
(ΔG = −29.76 kcal/mol). Dissociation of the second CCOC kinetically unfavorable. However, while this bond dissociation
bond in the FEC molecule is accompanied by simultaneous has a high free energy barrier (FB7) of 16.83 kcal/mol, making
breaking of the C−F bond leading to the release of the fluorine it kinetically unfavorable, free energy gain is favorable compared
atom. This sequential dissociation of multiple bonds as a result to the CCOC bond dissociation (−78.16 vs −65.20 kcal/mol).
of the single electron reduction of FEC was experimentally The radical anion (FB8) generated from this reaction
observed15 for LIBs as well. The fluorine atom generally decomposes further, thereby releasing the CH2CHF and
combines with a nearby sodium ion, creating NaF, a commonly NaCO3− anion (FB9). It is important to note that, in a single
observed inorganic salt when FEC is introduced as an additive. electron reduction state, this reaction is thermodynamically
The presence of LiF in LIBs was recently shown to act as a link uphill (ΔG = 3.14 kcal/mol) and thus unlikely to happen. On
between multiple decomposition products which in-turn the other hand, subsequent attack of another electron on the
promotes a network structure in SEI.35 radical makes this decomposition thermodynamically favorable
In addition to these two dissociation reactions, there is a and the free energy of reaction would be −106.65 kcal/mol
third possibility when fluorine and sodium do not form NaF (FB10).
after dissociation of CCOC and CEOC bonds. In this case, both 2e Electron Reduction of FEC. Like EC and VC, the CCOC
H and F atoms are released simultaneously and can combine to bond spontaneously dissociates after 2e reduction of FEC
form a HF molecule. The vinyl anion (FB6) generated as a (FB11). Subsequent reactions generate CO or CO2 gases along
result of this reaction is the same as the previous reaction with anions (FB12 and FB13) similar to the case of the 1e
involving the release of CO. However, this reaction is less reduction described above. Two significant differences were
exothermic (ΔG = −24.36 kcal/mol) compared to the other observed between 1e and 2e decomposition pathways. Free
two reactions discussed above. It should be noted that this energy gain for these reactions is very large for 2e reduction
decomposition mechanism of FEC releasing HF was also (ΔGFB12 = −98.8 kcal/mol, ΔGFB13 = −69.52 kcal/mol)
proposed for LIBs.36 However, Leung et al. did not observe this compared to 1e reduction (ΔGFB4 = −43.99 kcal/mol, ΔGFB5 =
reaction for LIB in their AIMD simulations.14 We believe the −29.76 kcal/mol). Additionally, the fluorine atom drifts away as
different outcomes stem from the different sequences of bond fluorine ion instead of combining with the sodium atom as in
breaking considered in the two studies. As we observed here, the case of 1e reduction.
sequential bond decompositions (of CCOC and CEOC bonds) 3.5. CEOC Bond Dissociation Is Preferred in the
do not lead to the release of HF, but simultaneous dissociation Reduction State A while CCOC Bond Dissociation Is
8935 DOI: 10.1021/acs.chemmater.6b03403
Chem. Mater. 2016, 28, 8930−8941
Chemistry of Materials Article

Preferred in the Reduction State B. For all the solvent sodium ion solvation shell (Table 1). The addition of the fourth
molecules studied above, energy barriers for the CEOC bond EC molecule to an EC3−Na+ cluster leads to a positive
dissociations are higher in reduction state B than reduction
state A. On the other hand, transition states corresponding to Table 1. Formation Free Energy with an Increasing Number
the CCOC bond dissociation exist only for reduction state B. All of EC Molecules in the Solvation Shell of a Sodium Iona
attempts to find a transition state for the CCOC bond
formation free reduction free
dissociation in state A led to a transition state for reduction cluster energy energy
state B, suggesting that the CCOC bond dissociation can be
EC + (EC)Na+ → (EC)2Na+ −18.70 −64.06
initiated only in the reduction state B. Surprisingly, the CCOC EC + (EC)2Na+ → (EC)3Na+ −8.22 −56.75
bond dissociation reactions in LIBs systems have been EC + (EC)3Na+ → (EC)4Na+ 10.23 −54.12
discussed only in reduction state B without any reference to a
state A.32 Observations made here suggest that the two distinct Reduction free energy is also given for different cluster sizes. All
energy values are in kcal/mol.
reduction states, A and B, have a preference over which bond is
likely to dissociate after reduction, and hence, qualitatively
different SEI products might be observed depending on the formation free energy (ΔG = GEC4−Na+ − GEC3−Na+ − GEC) of
preferred reduction state of the decomposing molecule. The
10.23 kcal/mol. The inability of the sodium ions to solvate four
CEOC bond dissociation is preferred in reduction state A, which
EC molecules in an electrolyte is in sharp contrast to the case of
eventually leads to the formation of the carbonate salts (e.g.,
LIBs which are known to form a stable tetrahedral structure
Na2CO3, Na2EDC, Na2BDC) and hydrocarbon gas molecules.
with four EC molecules.31,41 Due to the smaller size of a lithium
On the other hand, the CCOC bond dissociation is favored in
ion, the charge density at the surface of the ion is higher than a
reduction state B, which leads to the formation of oxygen rich
sodium ion, and therefore, a lithium ion can solvate more
SEI products (e.g., aloxide, glycol group) along with CO and
molecules than a sodium ion, accounting for the larger
CO2 gas molecules.37 These differing outcomes of the two
solvation shell of lithium ions. The smaller solvation shell of
reduction states are likely to be valid for LIBs as well but, to the
the sodium ion is beneficial for battery operation because a
best of our knowledge, have not been identified yet. A clear small cluster of molecules can be transported easily in the
distinction between the two reduction states has important electrolyte solution resulting in faster charge/discharge rates.
implications for controlling the SEI composition. As we will Reduction free energies of the EC with increasing number of
show later in this paper, preference for reduction states A or B molecules in the solvation shell are also given in Table 1. With
can be adjusted by introducing suitable additive molecules in increasing cluster size, free energy of reduction increases and
the electrolyte. approaches the value obtained from the continuum solvent
3.6. Solvation Behavior of Different Solvent Mole- model. This gives us confidence that the PCM model employed
cules: Structure of a Sodium Ion Solvation Shell. Before above can accurately reproduce the reduction characteristics of
investigating the effects of additives on reductive decom- explicit solvent molecules.
position, it is important to understand the solvation behavior of Next, we studied the effects of additive molecules on the
commonly used molecules in the electrolyte. The structure of structure of the solvation shell. Inclusion of an additive
an ion solvation shell plays a major role in battery operation. A molecule (PC, VC, FEC) in an EC3−Na+ cluster has a positive
large solvation shell not only limits ion-mobility during the formation free energy, illustrating that the three-coordinated
charge/discharge cycle but also changes ion-transfer kinetics solvation shell is still the most stable configuration even in the
near the electrodes.38 From the point of view of SEI formation, presence of additives. To investigate the possibility of additive
the structure of an ion solvation shell can have significant molecules replacing an EC from the three-coordinated
effects as well.39,40 Preferential solvation of an additive molecule solvation shell, we computed the formation free energies of
over the solvation of the solvent molecules changes the clusters where an EC molecule is replaced by an additive
electronic environment of the decomposing molecules and, as a molecule. For each of the additive molecules, we considered the
result, changes the reduction chemistry.40 We examine the free energy of formation of the EC2−X−Na+ complex (X = EC,
solvation characteristics of a sodium ion in the EC solvent and PC, VC, FEC) defined as ΔGformation= ΔGEC2−X − ΔGEC2 −
investigate potential consequences of additive molecules (PC,
ΔGX. Values of the free energies of formation for all the
VC, and FEC) on solvation thermodynamics (Figure 7).
additive molecules have been shown in Table 2. For all the
Formation free energy of ECn−Na+ (n = 1···4) clusters
trimers studied, EC3−Na+ is the most stable state, and
suggest that a maximum of three EC molecules can stay in the
substituting an EC molecule with any other molecule results
in a less stable solvation shell. The formation energies of Na+−

Table 2. Formation Free Energy of Adding Different


Molecules to EC-Dimer Completing the Solvation Shell of a
Sodium Iona
formation free reduction free
cluster energy energy
PC + (EC)2Na+ → (EC)2Na+PC −8.08 −55.76
VC + (EC)2Na+ → (EC)2Na+VC −5.13 −58.88
Figure 7. Solvation shell structure of a sodium ion in the EC solvent FEC + (EC)2Na+ → (EC)2Na+FEC −5.15 −59.86
(a) before reduction and (b) after reduction. Structures were
a
optimized in the gas phase simulation using B3PW91/6-311++G(d,p) Reduction free energies of the EC molecule in these solvation shells
DFT simulations. are also given for all clusters. All energy values are in kcal/mol.

8936 DOI: 10.1021/acs.chemmater.6b03403


Chem. Mater. 2016, 28, 8930−8941
Chemistry of Materials Article

Figure 8. (a) Three distinct reduction states for a heterodimer (EC−VC, EC−PC, or EC−FEC), optimized using B3PW91/6-311++G(d,p) DFT
simulations. (b) Change in the reduction free energy for both reduction states of EC in the vicinity of different additive molecules.

Figure 9. Reaction pathways for the reduction induced decomposition of dimer complexes studied in the gas phase. Gibbs free energies at T =
298.15 K and P = 1 atm are given with respect to an unreduced structure. Molecules marked with A represent additive molecules (A = EC, PC, VC,
FEC). The upper half of the figure represents pathways involving the ring opening of the EC while the lower half represents pathways involving
decomposition of the additive molecules.

EC3 and Na+−EC2−PC are almost identical (ΔGformation = been proven to extend battery life by improving the SEI
−8.21 kcal/mol vs ΔGformation = −8.08 kcal/mol), while composition.29,42,43 Additive molecules generally have higher
solvation shells incorporating a VC (ΔGformation = −5.13 kcal/ reduction potentials compared to solvent molecules and are
mol) or a FEC (ΔGformation = −5.15 kcal/mol) molecule are therefore preferentially reduced at the anode. This sacrificial
comparatively less stable. These results suggest that the most reduction is considered to be the main protection mechanism
stable solvation shell is formed with EC molecules, and additive for the additives. However, it was recently shown that the
molecules are less likely to replace an EC molecule from the sacrificial reduction of additives can only be partially accounted
solvation shell, although a solvation shell with an additive for in the enhancement of the battery life.34 Ushirogata et al.
molecule is thermodynamically stable. showed that a VC molecule preferentially reacts with an EC
3.7. Role of Additive Molecules on the Decomposition anion radical.34 This suppresses the two-electron reduction of
of EC. In LIB electrolytes, the addition of VC and FEC has EC molecules and thus enhances the SEI composition.
8937 DOI: 10.1021/acs.chemmater.6b03403
Chem. Mater. 2016, 28, 8930−8941
Chemistry of Materials Article

Figure 10. Change in (a) the free energy barrier and (b) the free energy released after dissociation of the CEOC bond of EC and additive molecule in
the reduction state A of different dimers.

Additionally, reaction pathways for decomposition of the 10). In the presence of a VC molecule, the value of the energy
solvent molecules are also altered when additive molecules barrier is 4.58 kcal/mol, and in the presence of a FEC molecule,
are present in the vicinity.35,42 In order to investigate the role of it is 4.78 kcal/mol. These results suggest that, even though the
additives on EC decomposition, we simulated decomposition reduction of the EC molecule becomes thermodynamically
pathways for four dimers, namely, EC−EC, EC−PC, EC−VC, favorable in the presence of additive molecules, the ring
and EC−FEC. VC and FEC are commonly used additives; PC opening as a result of CEOC bond dissociation becomes
is often used as a cosolvent of EC, and the EC−EC cluster kinetically unfavorable.
representing the pure EC solvent is considered as a reference The heat of bond dissociation, defined as the change in the
structure. In the case of a heterodimer, any of the two dimer free energy due to bond dissociation, also varies in the
molecules can get reduced in the A or B reduction states. Since presence of another molecule in the dimer. While homolytic
the reduction state A involves localization of the excess electron dissociation of the CEOC bond remains thermodynamically
on the sodium ion, this state is identical for the reduction of favorable in the presence of all molecules studied here, the heat
both molecules, leaving three distinct reduction states for a of bond dissociation varies with the additive molecule (see
heterodimer as shown in Figure 8a. Figures 9 and 10). The presence of VC or FEC not only
Variation in the reduction free energy of an EC molecule in increases the energy barrier but also lowers the free energy (by
conjugation with other molecules is depicted in Figure 8b for A ∼2.5 kcal/mol) released upon bond dissociation, making this
and B reduction states. The reduction free energy of the bond dissociation thermodynamically even more unfavorable.
homodimer molecule in the presence of another EC molecule On the other hand, in the presence of a PC molecule, the heat
is −64.0 kcal/mol, which is marginally lower than the reduction of the bond dissociation increases by ∼1 kcal/mol making the
free energy (−52.65 kcal/mol) in the continuum EC solvent bond dissociation slightly more favorable.
discussed in Section 3.1. To quantify the impact of different Reaction pathways discussed above lead to the formation of
additive molecules on the decomposition pathways, we anion radicals which undergo secondary termination reactions
consider the EC−Na+−EC complex as a reference and similar to the monomer reactions discussed in the previous
systematically replace the nonreducing EC molecule with section. Among many different reaction possibilities, the attack
additive molecules. of an excess electron on an anion radical to generate a
When the EC molecule is replaced with a PC molecule, carbonate anion and to release an ethylene molecule is the most
reduction of EC becomes favorable by −1.2 kcal/mol. Similarly, prominent one. Thermodynamic data corresponding to this
replacing the EC molecule with a VC or FEC molecule also reaction in the presence of the additive molecules are also
lowers reduction free energy of the EC molecule (Figure 8b). shown in Figure 9. These data suggest that the dissociation of
However, the addition of these two additive molecules changes the second CEOC bond of an EC anion radical is mostly
reduction free energy more significantly (∼4 kcal/mol) for exothermic in the presence of a PC molecule, followed by EC,
both of the reduction states. This demonstrates that the VC, and finally FEC.
reduction of an EC molecule forms a more stable structure in 3.8. Preferential Decomposition of the Additive
the presence of an intact additive molecule compared to the Molecules. In addition to the decomposition of EC discussed
pure EC solvent. EC decomposition pathways that follow these above, reductive decompositions of other molecules (i.e.,
reductions are shown in Figure 9. As elucidated in the previous additive) in a dimer are also plausible. In fact, the
section, because the decomposition of an EC molecule after 1e decomposition of additive molecules is considered to be the
reduction is most likely to occur through the reduction state A primary mechanism of solvent protection by the additives.18
followed by the dissociation of a CEOC bond, we consider only When a dimer is reduced, an excess electron is more likely to
those pathways that initiate from reduction state A.The energy localize on the molecule with a higher reduction potential, and
barrier for the dissociation of a CEOC bond in conjugation with by this virtue, the additive molecules, instead of the solvent EC
a PC molecule is 4.94 kcal/mol which is slightly higher molecule in the dimer, undergo reduction-induced decom-
compared to the EC dimer. The same trend of a higher energy position. The thermochemistry of such decomposition
barrier holds for the other two molecules as well (Figures 9 and reactions is also shown in Figure 9. From the single molecule
8938 DOI: 10.1021/acs.chemmater.6b03403
Chem. Mater. 2016, 28, 8930−8941
Chemistry of Materials Article

decomposition studies, it is evident that the homolytic dissociation. Several experiments have demonstrated reduction
dissociations of both CEOC and CCOC bonds are thermody- potentials of EC-based electrolytes in the range of 0.4 to 0.7 eV
namically feasible in the 1e reduction state of VC and FEC but for different anode materials.19,44,45 Komaba et al. observed
not of EC. We have also shown that CEOC bond dissociation is electrochemical decompositions at 0.7 eV vs Na+/Na (2.41 eV
likely to occur in reduction state A, while CCOC bond physical scale) for the PC−FEC-based electrolyte.44 The
dissociation is favored in reduction state B. Keeping this in reduction potential of an EC molecule (52.65 kcal/mol or
mind, we have investigated both bond dissociations starting 2.28 eV), computed in our calculations, is in good agreement
from their preferred reduction states. with these experiments. Using DFT calculations with the same
For the EC−VC dimer reduction, localization of the excess exchange functional, the reduction potential of EC in LIBs was
electron over the carbonyl group of the VC (as seen in the calculated to be 45.8 kcal/mol. In addition to a higher tendency
reduction of VC in state B) gives the most stable structure for reduction, the barrier for the ring opening is smaller in both
(shown in [iii] of Figure 8a) compared to the other two the 1e and 2e reduction states of EC in SIBs as opposed to
possibilities ([i] and [ii] in Figure 8a). The free energy barrier LIBs. In the 2e reduction state, EC decomposes without any
for the CCOC bond dissociation of VC in this reduction state is barriers in SIB while there is a small barrier of ∼2.0 kcal/mol in
0.97 kcal/mol, indicating that bond dissociation will be almost LIBs. Similarly, the decomposition barrier in the preferred 1e
spontaneous after reduction. However, thermodynamic gain as reduction state of EC−Na+ is smaller (4.50 kcal/mol) than for
a result of this bond dissociation is much smaller (−2.82 kcal/ the preferred reduction state of EC−Li+ (9.5 kcal/mol). These
mol). In the reduction state A, dissociation of a CEOC bond is results suggest that EC molecules decompose rapidly during
preferred compared to the CCOC bond. However, this reaction the initial phase of SEI formation through a CCOE dissociation
has a free energy barrier of ∼9.04 kcal/mol. For the same after the 2e electron reduction. Such decomposition leads to an
reduction state, the free energy barrier for CEOC bond initial deposition of polymeric compounds on the anode surface
dissociation of EC is 4.58 kcal/mol. The thermodynamic gain which then slows electron transfer to the electrolytes.46Under
as a result of the CEOC bond dissociation is −17.46 and −32.91
these circumstances, the 1e electron reduction mechanism
kcal/mol for VC and EC, respectively.
dominates SEI formation. As shown earlier, the 1e induced
On the basis of these free energy values, it is evident that the
decomposition leads to the formation of sodium carbonates
EC−VC dimer is most likely reduced in the reduction state B of
(−CO3) or sodium alkyl carbonates. A recent study has
VC. In this reduction state, dissociation of the CCOC bond of a
suggested that the chemical stability of the decomposition
VC molecule will occur almost spontaneously and the EC
molecule will remain undecomposed. This sacrificial decom- products can further change the characteristics of the SEI.37
position of the VC molecules to protect the EC molecules is in Keeping this in mind, we have investigated the chemical
contrast to the findings from Wang et al.17 which showed that stability of two major decomposition products Na2BDC and
the VC molecules are more likely to remain intact due to the Na2EDC against radical attack. Our calculations suggest (see
large free energy barrier for their decomposition. Due to this the Supporting Information) that the interaction of the OH
higher free energy barrier, if the dimer is reduced in the radical with one of the sodium ions of Na2BDC considerably
reduction state A, the CEOC bond dissociation of EC is weakens the C1−O1 bond compared to the bare Na2BDC
kinetically preferred. We suggest this discrepancy is because molecule, and only 4.82 kcal/mol energy is needed to break this
Wang et al. only considered the dissociation of the CEOC bond. This homolytic dissociation results in the formation of a
bond.17 Our calculations also show that C E O C bond new radical OC2H4CO3 and CO2. In the case of Na2EDC, the
dissociation has a higher free energy barrier but an alternative same bond-dissociation reaction is even exothermic and −9.62
decomposition pathway (dissociation of CCOC bond) is more kcal/mol is released upon homolytic dissociation of the C1−
likely to occur due to the smaller barrier. O1 bond. These new radicals will interact with other solvent
Like the EC−VC dimer, reduction of a EC−FEC dimer gives molecules initiating additional decomposition reactions and
the most stable structure ([iii] in Figure 8a) if localization of contribute to additional SEI growth. In comparison, lithium
the excess electron occurs over the carbonyl group of FEC (as counterparts of the same salts are more stable and bond
seen in the reduction of FEC in state B). Due to the small free dissociation energies for the homolytic dissociation of the C1−
energy barrier for CCOC bond dissociation in this reduction O1 bond is 7.43 kcal/mol for Li2EDC and 3.92 kcal/mol for
state, spontaneous dissociation of the CCOC bond occurs in Li2BDC. These results suggest that alkyl carbonate salts in SIBs
FEC if the excess electron is localized to form reduction state B. will decompose easily in the vicinity of radicals and have higher
In comparison, the free energy barrier for CEOC bond propensity for decomposition than in LIBs. This is consistent
dissociation of FEC is 7.76 kcal/mol, while it is 4.78 kcal/ with the experimental observations of thicker SEI in SIBs. In
mol for the CEOC bond dissociation of EC in the reduction addition, two distinct reduction states (A and B) were identified
state A of this EC−FEC dimer. This suggests that, in the for all monomers and these reduction states have preferences
reduction state A, EC can also undergo bond dissociation, over which bond is likely to dissociate after reduction.
further evidence opposing the conventional belief that only Dissociation of the CEOC bond (preferred in state A) is more
additive molecules decompose first. However, since the likely to produce carbonate salts while dissociation of the CCOC
reduction potential of reduction state A is lower compared to bond (preferred in state B) is more like to produce molecules
reduction state B, dissociation of the FEC remains the more containing the −O−R−O− group. We have also shown that
thermodynamically favorable pathway of dimer decomposition. the preferred reduction state of EC changes from A to B in the
presence of additives (Figure 8b); this suggests that additives
4. DISCUSSION AND CONCLUSIONS may also change the decomposition products by changing the
We have characterized the reductive decomposition of EC, VC, preferred reduction state. Additional studies involving explicit
and FEC by computing the reduction potentials, the free solvent molecules with different proportions of additive
energy barriers for decomposition, and heats of bond molecules are needed to further investigate this aspect.
8939 DOI: 10.1021/acs.chemmater.6b03403
Chem. Mater. 2016, 28, 8930−8941
Chemistry of Materials Article

Free energy of solvation of sodium ions suggests that additive Electrolyte Interphase in Li-Ion Batteries. J. Phys. Chem. C 2013, 117,
molecules are more likely to be present as individual molecules 8579−8593.
rather than in a solvation shell with EC molecules. The (5) Slater, M. D.; Kim, D.; Lee, E.; Johnson, C. S. Sodium-Ion
reductive decomposition of individual additive molecules such Batteries. Adv. Funct. Mater. 2013, 23, 947−958.
as VC and FEC in bulk EC solvent show that not only are the (6) Philippe, B.; Valvo, M.; Lindgren, F.; Rensmo, H.; Edström, K.
Investigation of the Electrode/Electrolyte Interface of Fe 2 O 3
reduction potentials of both VC and FEC molecules higher
Composite Electrodes: Li vs Na Batteries. Chem. Mater. 2014, 26,
than that of the EC (55.96 kcal/mol for VC and 58.47 kcal/mol 5028−5041.
for FEC), but also the barrier for the ring opening through the (7) Muñoz-Márquez, M. A.; Zarrabeitia, M.; Castillo-Martínez, E.;
CCOC bond dissociation along its most probable pathway is Eguía-Barrio, A.; Rojo, T.; Casas-Cabanas, M. Composition and
small compared to the barrier for the EC ring opening. This Evolution of the Solid-Electrolyte Interphase in Na 2 Ti 3 O 7
leads to early decomposition of both molecules. Intermediate Electrodes for Na-Ion Batteries: XPS and Auger Parameter Analysis.
products generated from the dissociation of CCOC bonds are ACS Appl. Mater. Interfaces 2015, 7, 7801−7808.
very susceptible to oligomerization and are more likely to form (8) Weadock, N.; Varongchayakul, N.; Wan, J.; Lee, S.; Seog, J.; Hu,
oxygen-rich polymeric compounds.32 In addition, the decom- L. Determination of Mechanical Properties of the SEI in Sodium Ion
position of FEC generates NaF which is a very good passivating Batteries via Colloidal Probe Microscopy. Nano Energy 2013, 2, 713−
agent. Not only do the additive molecules decompose first, but 719.
also the decomposition pathways of EC molecules are affected. (9) Cheng, X.-B.; Zhang, R.; Zhao, C.-Z.; Wei, F.; Zhang, J.-G.;
The reduction potential of the EC increases in the presence of Zhang, Q. A Review of Solid Electrolyte Interphases on Lithium Metal
Anode. Adv. Sci. 2016, 3, 1500213.
both VC and FEC, indicating EC molecules will be easily
(10) Baggetto, L.; Ganesh, P.; Meisner, R. P.; Unocic, R. R.; Jumas, J.
reduced in the vicinity of these additive molecules. However, C.; Bridges, C. a.; Veith, G. M. Characterization of Sodium Ion
the free energy barrier to the ring opening also increases and Electrochemical Reaction with Tin Anodes: Experiment and Theory. J.
makes reductive decomposition kinetically unfavorable. Power Sources 2013, 234, 48−59.
In summary, several aspects of the reductive decomposition (11) Iermakova, D. I.; Dugas, R.; Palacín, M. R.; Ponrouch, A. On the
of cyclic carbonates in SIB have been studied to explain Comparative Stability of Li and Na Metal Anode Interfaces in
experimentally observed characteristics of SEI in SIBs. In Conventional Alkyl Carbonate Electrolytes. J. Electrochem. Soc. 2015,
addition, a new pathway to tailor SEI composition by changing 162, A7060−A7066.
the preferred reduction states of electrolyte molecules has been (12) Komaba, S.; Murata, W.; Ishikawa, T.; Yabuuchi, N.; Ozeki, T.;
elucidated. More detailed studies involving electrolyte salt Nakayama, T.; Ogata, A.; Gotoh, K.; Fujiwara, K. Electrochemical Na
molecules and the anode surface will shed additional light and Insertion and Solid Electrolyte Interphase for Hard-Carbon Electrodes
will be the subject of future research. and Application to Na-Ion Batteries. Adv. Funct. Mater. 2011, 21,


3859−3867.
ASSOCIATED CONTENT (13) Ponrouch, A.; Dedryvere, R.; Monti, D.; Demet, A. E.; Ateba
Mba, J. M.; Croguennec, L.; Masquelier, C.; Johansson, P.; Palacin, M.
*
S Supporting Information R. Towards High Energy Density Sodium Ion Batteries through
The Supporting Information is available free of charge on the Electrolyte Optimization. Energy Environ. Sci. 2013, 6, 2361−2369.
ACS Publications website at DOI: 10.1021/acs.chemma- (14) Leung, K.; Rempe, S. B.; Foster, M. E.; Ma, Y.; Martinez del la
ter.6b03403. Hoz, J. M.; Sai, N.; Balbuena, P. B. Modeling Electrochemical
Decomposition of Fluoroethylene Carbonate on Silicon Anode
Energetics for the decomposition reactions of Na2EDC
Surfaces in Lithium Ion Batteries. J. Electrochem. Soc. 2014, 161,
and Na2BDC with and without OH radical (PDF) A213−A221.

■ AUTHOR INFORMATION
ORCID
(15) Shkrob, I. A.; Wishart, J. F.; Abraham, D. P. What Makes
Fluoroethylene Carbonate Different? J. Phys. Chem. C 2015, 119,
14954−14964.
(16) El Ouatani, L.; Dedryvère, R.; Siret, C.; Biensan, P.; Reynaud, S.;
Hemant Kumar: 0000-0003-4339-5711 Iratçabal, P.; Gonbeau, D. The Effect of Vinylene Carbonate Additive
Notes on Surface Film Formation on Both Electrodes in Li-Ion Batteries. J.
The authors declare no competing financial interest. Electrochem. Soc. 2009, 156, A103.


(17) Wang, Y.; Nakamura, S.; Tasaki, K.; Balbuena, P. B. Theoretical
ACKNOWLEDGMENTS Studies to Understand Surface Chemistry on Carbon Anodes for
Lithium-Ion Batteries: How Does Vinylene Carbonate Play Its Role as
This work is supported by the Grants EFMA-542879, CMMI- an Electrolyte Additive? J. Am. Chem. Soc. 2002, 124, 4408−4421.
1363203, and CBET-1235870 from the US National Science (18) Kim, S.-P.; van Duin, A. C. T.; Shenoy, V. B. Effect of
Foundation.


Electrolytes on the Structure and Evolution of the Solid Electrolyte
Interphase (SEI) in Li-Ion Batteries: A Molecular Dynamics Study. J.
REFERENCES Power Sources 2011, 196, 8590−8597.
(1) Yabuuchi, N.; Kubota, K.; Dahbi, M.; Komaba, S. Research (19) Ji, L.; Gu, M.; Shao, Y.; Li, X.; Engelhard, M. H.; Arey, B. W.;
Development on Sodium-Ion Batteries. Chem. Rev. 2014, 114, 11636− Wang, W.; Nie, Z.; Xiao, J.; Wang, C.; Zhang, J. G.; Liu, J. Controlling
11682. SEI Formation on SnSb-Porous Carbon Nanofibers for Improved Na
(2) Xu, K. Electrolytes and Interphases in Li-Ion Batteries and Ion Storage. Adv. Mater. 2014, 26, 2901−2908.
Beyond. Chem. Rev. 2014, 114, 11503−11618. (20) Shkrob, I. a.; Zhu, Y.; Marin, T. W.; Abraham, D. Reduction of
(3) Zhang, Q.; Pan, J.; Lu, P.; Liu, Z.; Verbrugge, M. W.; Sheldon, B. Carbonate Electrolytes and the Formation of Solid-Electrolyte
W.; Cheng, Y. T.; Qi, Y.; Xiao, X. Synergetic Effects of Inorganic Interface (SEI) in Lithium-Ion Batteries. 1. Spectroscopic Observa-
Components in Solid Electrolyte Interphase on High Cycle Efficiency tions of Radical Intermediates Generated in One-Electron Reduction
of Lithium Ion Batteries. Nano Lett. 2016, 16, 2011−2016. of Carbonates. J. Phys. Chem. C 2013, 117, 19255−19269.
(4) Shi, S.; Qi, Y.; Li, H.; Hector, L. G. Defect Thermodynamics and (21) Bhide, A.; Hofmann, J.; Katharina Dürr, A.; Janek, J.; Adelhelm,
Diffusion Mechanisms in Li2CO 3 and Implications for the Solid P. Electrochemical Stability of Non-Aqueous Electrolytes for Sodium-

8940 DOI: 10.1021/acs.chemmater.6b03403


Chem. Mater. 2016, 28, 8930−8941
Chemistry of Materials Article

Ion Batteries and Their Compatibility with Na 0.7 CoO 2. Phys. Chem. (39) Borodin, O.; Bedrov, D. Interfacial Structure and Dynamics of
Chem. Phys. 2014, 16, 1987−1998. the Lithium Alkyl Dicarbonate SEI Components in Contact with the
(22) Singh, G.; Aguesse, F.; Otaegui, L.; Goikolea, E.; Gonzalo, E.; Lithium Battery Electrolyte. J. Phys. Chem. C 2014, 118, 18362−18371.
Segalini, J.; Rojo, T. Electrochemical Performance of NaFex- (40) von Cresce, A.; Xu, K. Preferential Solvation of Li+ Directs
(Ni0.5Ti0.5)1−xO2 (X = 0.2 and X = 0.4) Cathode for Sodium-Ion Formation of Interphase on Graphitic Anode. Electrochem. Solid-State
Battery. J. Power Sources 2015, 273, 333−339. Lett. 2011, 14, A154.
(23) Ming, J.; Ming, H.; Yang, W.; Kwak, W.-J.; Park, J.-B.; Zheng, J.; (41) Von Wald Cresce, A.; Gobet, M.; Borodin, O.; Peng, J.; Russell,
Sun, Y.-K. A Sustainable Iron-Based Sodium Ion Battery of Porous S. M.; Wikner, E.; Fu, A.; Hu, L.; Lee, H. S.; Zhang, Z.; Yang, X. Q.;
carbon−Fe3O4/Na2FeP2O7 with High Performance. RSC Adv. 2015, Greenbaum, S.; Amine, K.; Xu, K. Anion Solvation in Carbonate-Based
5, 8793−8800. Electrolytes. J. Phys. Chem. C 2015, 119, 27255−27264.
(24) Mortazavi, M.; Wang, C.; Deng, J.; Shenoy, V. B.; Medhekar, N. (42) Takamatsu, D.; Orikasa, Y.; Mori, S.; Nakatsutsumi, T.;
V. Ab Initio Characterization of Layered MoS2 as Anode for Sodium- Yamamoto, K.; Koyama, Y.; Minato, T.; Hirano, T.; Tanida, H.;
Arai, H.; Uchimoto, Y.; Ogumi, Z. Effect of an Electrolyte Additive of
Ion Batteries. J. Power Sources 2014, 268, 279−286.
Vinylene Carbonate on the Electronic Structure at the Surface of a
(25) Lee, J.; Chen, Y.-M.; Zhu, Y.; Vogt, B. D. Tuning SEI Formation
Lithium Cobalt Oxide Electrode under Battery Operating Conditions.
on Nanoporous Carbon−titania Composite Sodium Ion Batteries
J. Phys. Chem. C 2015, 119, 9791−9797.
Anodes and Performance with Subtle Processing Changes. RSC Adv. (43) Chen, L.; Wang, K.; Xie, X.; Xie, J. Effect of Vinylene Carbonate
2015, 5, 99329−99338. (VC) as Electrolyte Additive on Electrochemical Performance of Si
(26) Schmidt, M. W.; Baldridge, K. K.; Boatz, J. A.; Elbert, S. T.; Film Anode for Lithium Ion Batteries. J. Power Sources 2007, 174,
Gordon, M. S.; Jensen, J. H.; Koseki, S.; Matsunaga, N.; Nguyen, K. A.; 538−543.
Su, S.; Windus, T. L.; Dupuis, M.; Montgomery, J. A. General Atomic (44) Komaba, S.; Ishikawa, T.; Yabuuchi, N.; Murata, W.; Ito, A.;
and Molecular Electronic Structure System. J. Comput. Chem. 1993, 14, Ohsawa, Y. Fluorinated Ethylene Carbonate as Electrolyte Additive for
1347−1363. Rechargeable Na Batteries. ACS Appl. Mater. Interfaces 2011, 3, 4165−
(27) Han, Y. K.; Lee, S. U. Performance of Density Functional for 4168.
Calculation of Reductive Ring-Opening Reaction Energies of Li+-EC (45) Han, M. H.; Gonzalo, E.; Singh, G.; Rojo, T. A Comprehensive
and Li+-VC. Theor. Chem. Acc. 2004, 112, 106−112. Review of Sodium Layered Oxides: Powerful Cathodes for Na-Ion
(28) Li, H. Quantum Mechanical/molecular Mechanical/continuum Batteries. Energy Environ. Sci. 2015, 8, 81−102.
Style Solvation Model: Linear Response Theory, Variational Treat- (46) Ponrouch, A.; Dedryvère, R.; Monti, D.; Demet, A. E.; Ateba
ment, and Nuclear Gradients. J. Chem. Phys. 2009, 131, 184103. Mba, J. M.; Croguennec, L.; Masquelier, C.; Johansson, P.; Palacín, M.
(29) Tasaki, K. Solvent Decompositions and Physical Properties of R. Towards High Energy Density Sodium Ion Batteries through
Decomposition Compounds in Li-Ion Battery Electrolytes Studied by Electrolyte Optimization. Energy Environ. Sci. 2013, 6, 2361.
DFT Calculations and Molecular Dynamics Simulations. J. Phys. Chem.
B 2005, 109, 2920−2933.
(30) Ho, J.; Ertem, M. Z. Calculating Free Energy Changes in
Continuum Solvation Models. J. Phys. Chem. B 2016, 120, 1319−1329.
(31) Wang, Y.; Nakamura, S.; Ue, M.; Balbuena, P. B. Theoretical
Studies to Understand Surface Chemistry on Carbon Anodes for
Lithium-Ion Batteries: Reduction Mechanisms of Ethylene Carbonate.
J. Am. Chem. Soc. 2001, 123, 11708−11718.
(32) Leung, K. Two-Electron Reduction of Ethylene Carbonate: A
Quantum Chemistry Re-Examination of Mechanisms. Chem. Phys. Lett.
2013, 568−569, 1−8.
(33) Shkrob, I. A.; Zhu, Y.; Marin, T. W.; Abraham, D. Reduction of
Carbonate Electrolytes and the Formation of Solid-Electrolyte
Interface (SEI) in Lithium-Ion Batteries. 2. Radiolytically Induced
Polymerization of Ethylene Carbonate. J. Phys. Chem. C 2013, 117,
19270−19279.
(34) Ushirogata, K.; Sodeyama, K.; Okuno, Y.; Tateyama, Y. Additive
Effect on Reductive Decomposition and Binding of Carbonate-Based
Solvent toward Solid Electrolyte Interphase Formation in Lithium-Ion
Battery. J. Am. Chem. Soc. 2013, 135, 11967−11974.
(35) Okuno, Y.; Ushirogata, K.; Sodeyama, K.; Tateyama, Y.
Decomposition of the Fluoroethylene Carbonate Additive and the
Glue Effect of Lithium Fluoride Products for the Solid Electrolyte
Interphase: An Ab Initio Study. Phys. Chem. Chem. Phys. 2016, 18,
8643−8653.
(36) Mogi, R.; Inaba, M.; Iriyama, Y.; Abe, T.; Ogumi, Z. Study of the
Decomposition of Propylene Carbonate on Lithium Metal Surface by
Pyrolysis-Gas Chromatography-Mass Spectroscopy. Langmuir 2003,
19, 814−821.
(37) Soto, F. A.; Ma, Y.; Martinez De La Hoz, J. M.; Seminario, J. M.;
Balbuena, P. B. Formation and Growth Mechanisms of Solid-
Electrolyte Interphase Layers in Rechargeable Batteries. Chem.
Mater. 2015, 27, 7990−8000.
(38) Yamada, Y.; Iriyama, Y.; Abe, T.; Ogumi, Z. Kinetics of Lithium
Ion Transfer at the Interface between Graphite and Liquid
Electrolytes: Effects of Solvent and Surface Film. Langmuir 2009, 25,
12766−12770.

8941 DOI: 10.1021/acs.chemmater.6b03403


Chem. Mater. 2016, 28, 8930−8941

You might also like