20.04.2014 Views

Electrolyte Solvation and Ionic Association II. Acetonitrile-Lithium ...

Electrolyte Solvation and Ionic Association II. Acetonitrile-Lithium ...

Electrolyte Solvation and Ionic Association II. Acetonitrile-Lithium ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Journal of The Electrochemical Society, 159 (9) A1489-A1500 (2012)<br />

0013-4651/2012/159(9)/A1489/12/$28.00 © The Electrochemical Society<br />

<strong>Electrolyte</strong> <strong>Solvation</strong> <strong>and</strong> <strong>Ionic</strong> <strong>Association</strong><br />

<strong>II</strong>. <strong>Acetonitrile</strong>-<strong>Lithium</strong> Salt Mixtures: Highly Dissociated Salts<br />

Daniel M. Seo, a,∗ Oleg Borodin, b Sang-Don Han, a,∗ Paul D. Boyle, c<br />

<strong>and</strong> Wesley A. Henderson a,∗∗,z<br />

a <strong>Ionic</strong> Liquids & <strong>Electrolyte</strong>s for Energy Technologies (ILEET) Laboratory, Department of Chemical & Biomolecular<br />

Engineering, North Carolina State University, Raleigh, North Carolina 27695, USA<br />

b Electrochemistry Branch, Sensor & Electron Devices Directorate, U.S. Army Research Laboratory, Adelphi,<br />

Maryl<strong>and</strong> 20783, USA<br />

c X-ray Structural Facility, Department of Chemistry, North Carolina State University, Raleigh,<br />

North Carolina 27695, USA<br />

A1489<br />

The electrolyte solution structure for acetonitrile (AN)-lithium salt mixtures has been examined for highly dissociated salts. Phase<br />

diagrams are reported for (AN) n -LiN(SO 2 CF 3 ) 2 (LiTFSI) <strong>and</strong> -LiPF 6 electrolytes. Single crystal structures <strong>and</strong> Raman spectroscopy<br />

have been utilized to provide information regarding the solvate species present in the solid-state <strong>and</strong> liquid phases, as well as<br />

the average solvation number variation with salt concentration. Molecular dynamics (MD) simulations of the mixtures have been<br />

correlated with the experimental data to provide additional insight into the molecular-level interactions. Quantum chemistry (QC)<br />

calculations were performed on (AN) n -Li-(anion) m clusters to validate the ability of the developed many-body polarizable force field<br />

(used for the simulations) to accurately describe cluster stability (ionic association). The combination of these techniques provides<br />

tremendous insight into the solution structure within these electrolyte mixtures.<br />

© 2012 The Electrochemical Society. [DOI: 10.1149/2.035209jes] All rights reserved.<br />

Manuscript submitted February 23, 2012; revised manuscript received April 6, 2012. Published August 14, 2012.<br />

The key properties of electrolytes are determined by the solvate<br />

species present in solution, but only limited information is available<br />

regarding the structure of liquid electrolytes. A preliminary study of<br />

(glyme) n -LiX mixtures indicated the following order for increasing<br />

ionic association with varying anions: 1<br />

LiTFSI, LiAsF 6 < LiClO 4 , LiI < LiBF 4<br />

< LiCF 3 SO 3 < LiNO 3 , LiBr < LiCF 3 CO 2<br />

based upon the phase behavior <strong>and</strong> crystalline solvates that form in<br />

relatively dilute mixtures. This resulted in the following classification:<br />

LiN(SO 2 CF 3 ) 2 (LiTFSI) <strong>and</strong> LiAsF 6 are highly dissociated; LiClO 4 ,<br />

LiI <strong>and</strong> LiBF 4 are intermediately associated; <strong>and</strong> LiCF 3 SO 3 , LiNO 3 ,<br />

LiBr <strong>and</strong> LiCF 3 CO 2 are associated or highly associated. Upon dissolution<br />

in aprotic solvents, Li + cations are coordinated by the solvent<br />

molecules. The anions remain uncoordinated by the solvent as anion<br />

solvation occurs principally through hydrogen bonding (which does<br />

not occur in aprotic solvents such as AN). A competition therefore<br />

exists between the solvent molecules <strong>and</strong> anions for coordination to<br />

the Li + cations in solution. This results in the formation of a variety of<br />

solvate species which may be designated as solvent-separated ion pair<br />

(SSIP), contact ion pair (CIP) or aggregate (AGG) solvates in which<br />

the anions are coordinated to zero, one <strong>and</strong> two or more Li + cations,<br />

respectively. The solvates present in solution are a strong function<br />

of salt concentration <strong>and</strong> the structure of the solvent <strong>and</strong> anions, <strong>and</strong><br />

to a lesser extent the temperature—with ionic association (desolvation)<br />

tending to increase with increasing temperature. Thus, for dilute<br />

solutions with dissociated (e.g., AsF − 6 ) <strong>and</strong> highly associated (e.g.,<br />

CF 3 CO − 2 ) anions, SSIP <strong>and</strong> AGG solvates, respectively, are expected<br />

to predominate.<br />

This paper continues a previous exploration of electrolyte solution<br />

structure of acetonitrile (AN) mixtures with intermediate (LiClO 4 <strong>and</strong><br />

LiBF 4 ) <strong>and</strong> associated (LiCF 3 SO 3 , LiNO 3 <strong>and</strong> LiCF 3 CO 2 ) salts. 2 AN<br />

has been used in this study as this solvent has relatively simple interactions<br />

with Li + cations due to the single electron lone-pair on<br />

the nitrogen atom of the solvent molecule. Nitrile <strong>and</strong> dinitrile solvents<br />

may also be of interest as practical lithium battery electrolyte<br />

components. 3–5 The present manuscript reports the phase behavior<br />

<strong>and</strong> solution structure of (AN) n -LiTFSI <strong>and</strong> -LiPF 6 mixtures (highly<br />

dissociated salts). This work is part of a larger study linking electrolyte<br />

∗ Electrochemical Society Student Member.<br />

∗∗ Electrochemical Society Active Member.<br />

z E-mail: whender@ncsu.edu<br />

solution structure to the transport properties of aprotic solvent-lithium<br />

salt mixtures.<br />

Experimental <strong>and</strong> Computational Methods Section<br />

Materials.— Anhydrous AN (Sigma Aldrich, 99.8%) was used<br />

as-received. LiTFSI <strong>and</strong> LiPF 6 (battery grade) were purchased from<br />

3M Company <strong>and</strong> Novolyte, respectively. LiTFSI was dried at 120 ◦ C<br />

for 24 hr. Samples were prepared in a Vacuum Atmospheres inert<br />

atmosphere (N 2 ) glove box (< 5 ppm H 2 O) by combining the appropriate<br />

amounts of salt <strong>and</strong> solvent in hermetically sealed vials <strong>and</strong><br />

heating/stirring on a hot plate to form homogeneous solutions. The<br />

water content in the samples was confirmed to be negligible using Karl<br />

Fischer titration (Mettler Toledo DL39X Coulometer). The compositions<br />

are described generally using three notations: (1-x) AN-(x) LiX,<br />

(AN) n -LiX <strong>and</strong> (AN) n :LiX for discussions focusing on mole fraction,<br />

ratio of AN/Li <strong>and</strong> specific crystalline solvates, respectively.<br />

Thermal characterization.— DSC measurements were performed<br />

with a TA Instruments Q2000 DSC with liquid N 2 cooling. The instrument<br />

was calibrated with cyclohexane (solid-solid phase transition<br />

at −87.06 ◦ C, melt transition (T m ) at 6.54 ◦ C) <strong>and</strong> indium (T m at<br />

156.60 ◦ C). Hermetically sealed Al sample pans were prepared in the<br />

glove box. Sample pans were cycled (5 ◦ Cmin −1 ) <strong>and</strong> annealed repeatedly<br />

at subambient temperature to fully crystallize the samples when<br />

possible. Once the samples were crystallized, the pans were cooled to<br />

−150 ◦ C <strong>and</strong> then heated (5 ◦ Cmin −1 ) to fully melt the samples. Only<br />

the final heating traces are reported. Peak temperatures from this data<br />

were then used to construct the phase diagrams.<br />

Raman measurements.— Raman vibrational spectra were collected<br />

with a Horiba-Jobin Yvon LabRAM HR VIS high-resolution<br />

confocal Raman microscope using a 632 nm −1 He-Ne laser as the<br />

exciting source <strong>and</strong> a Linkam stage for temperature control with a<br />

long distance 50X objective. Spectra were typically collected using a<br />

10 s measurement time <strong>and</strong> five accumulations. Raman spectra were<br />

processed using LabSpec software.<br />

Quantum chemistry (QC) calculations.— QC calculations were<br />

performed on (AN) n -Li-(anion) m clusters (where anion = PF − 6 or<br />

TFSI − ;n= 0, 2 or 3; m = 1 or 2) using Gaussian 09 Revision B.01<br />

software. 6 Geometries were optimized at the M05-2X/6-31+G(d)<br />

level, while the binding energies were calculated at the M05-2X/6-<br />

31+G(d), MP2/aug-cc-pvDz <strong>and</strong> MP2/aug-cc-pvTz levels. Basis set


A1490 Journal of The Electrochemical Society, 159 (9) A1489-A1500 (2012)<br />

Figure 1. DSC heating traces (5 ◦ Cmin −1 ) <strong>and</strong> the corresponding phase diagrams of (a) (1-x) AN-(x) LiTFSI <strong>and</strong> (b) (1-x) AN-(x) LiPF 6 mixtures.<br />

superposition error (BSSE) was calculated using the counterpoise<br />

correction methodology.<br />

Molecular dynamics (MD) simulations.— MD simulations were<br />

performed on AN doped with LiPF 6 or LiTFSI employing a recently<br />

developed, many-body polarizable APPLE&P (version 1e19) force<br />

field for AN <strong>and</strong> the PF 6 − <strong>and</strong> TFSI − anions. 7–9 Extensive details regarding<br />

the MD simulation methodology are provided in the previous<br />

related manuscript. 2<br />

Results <strong>and</strong> Discussion<br />

Solvent-lithium salt phase behavior.— Pure AN.—AN undergoes a<br />

solid-solid phase transition at −56 ◦ C prior to the T m at −46 ◦ C<br />

(Fig. 1). 10–17 The low temperature (β or <strong>II</strong>) <strong>and</strong> high temperature (α or<br />

I) phases both consist of ordered AN molecules with the phase transition<br />

consisting of a 90 ◦ rotation of double slabs of the AN molecules<br />

packed in the crystal layers. 14<br />

AN-LiTFSI.—DSC measurements <strong>and</strong> the corresponding phase diagram<br />

for (AN) n -LiTFSI mixtures (Fig. 1) agree well with a previously<br />

reported partial phase diagram. 18 Three different crystalline solvate<br />

phases are found, consisting of 6/1, 4/1 <strong>and</strong> 1/1 AN/LiTFSI solvates,<br />

respectively. The structures of the 6/1 <strong>and</strong> 4/1 solvates are not yet<br />

known. For the 6/1 solvate, it is possible that the Li + cations are fully<br />

solvated by six AN molecules (octahedral coordination), but several<br />

studies have suggested that it is not energetically favorable (in the gas<br />

phase) to coordinate more than four AN molecules to a Li + cation. 19–23


Journal of The Electrochemical Society, 159 (9) A1489-A1500 (2012)<br />

A1491<br />

The spectroscopic data (see below) confirms that this is a SSIP phase<br />

with uncoordinated TFSI − anions, as expected for a solvate containing<br />

six AN molecules. The 4/1 solvate is also a SSIP phase which may<br />

have coordination similar to that found in the structures of the SSIP<br />

(AN) 4 :LiClO 4 24 <strong>and</strong> (AN) 4 :LiI 25 phases in which the Li + cations are<br />

fully solvated by four AN molecules (tetrahedral coordination) <strong>and</strong><br />

the anions are uncoordinated. For compositions between the 4/1 <strong>and</strong><br />

1/1 phases, it was not possible to fully crystallize the samples despite<br />

subjecting the samples to extensive heating-cooling cycles at<br />

subambient temperature. Thus, a “crystallinity gap” occurs for these<br />

compositions due to either the slow nucleation of crystalline solvates<br />

(beyond the time frame of the crystallization procedure utilized) or<br />

the inhibition of ordered crystalline solvate formation due to unfavorable<br />

packing. Single crystals of the 1/1 phase, i.e., (AN) 1 :LiTFSI<br />

solvate, were obtained <strong>and</strong> the structure was determined <strong>and</strong> reported<br />

elsewhere (see Supporting Information). 26 In this phase, there are two<br />

different Li + cations. One Li + cation has six-fold coordinated by six<br />

oxygen atoms from four TFSI − anions. The second Li + cation has<br />

four-fold coordination by two oxygen atoms from two TFSI − anions<br />

<strong>and</strong> two nitrogen atoms from two AN solvent molecules. Each TFSI −<br />

anion is coordinated to three Li + cations (AGG solvate).<br />

AN-LiPF 6 .—(AN) n -LiPF 6 mixtures form both 6/1 <strong>and</strong> 5/1 AN/LiPF 6<br />

crystalline solvate phases. In dilute mixture (n ≥ 10), the 6/1 SSIP<br />

solvate crystallizes with the peak at 18 ◦ C corresponding to the T m for<br />

this phase (Fig. 1 <strong>and</strong> Supporting Information). It is difficult to characterize<br />

the thermal behavior of this phase more definitively as more<br />

concentrated mixtures (i.e., 10 > n ≥ 6) form both the 6/1 <strong>and</strong> 5/1<br />

phases. For these samples, in each case, some of the sample crystallized<br />

into the 5/1 phase upon cooling from the melt, even when a rapid<br />

cooling procedure was used. The remainder of the sample then crystallized<br />

at low temperature into the 6/1 phase resulting in the complicated<br />

thermal behavior noted (Fig. 1). The crystal structure of the 6/1 phase,<br />

i.e., (AN) 6 :LiPF 6 solvate, has recently been reported. 27 The structure<br />

consists of Li + cations coordinated by four AN molecules with uncoordinated<br />

PF 6 − anions <strong>and</strong> uncoordinated AN molecules (two per Li +<br />

cation) located between the solvated cations (see Supporting Information).<br />

The crystal structure of the 5/1 phase, i.e., (AN) 5 :LiPF 6 solvate,<br />

has also been determined. 28 This resembles the 6/1 structure with Li +<br />

cations coordinated by four AN molecules, uncoordinated PF 6 − anions<br />

<strong>and</strong> uncoordinated AN molecules (one per Li + cation) located<br />

between the solvated cations (see Supporting Information). The same<br />

solvate structure is found for (AN) 5 :CuPF 6 . 29,30 The 5/1 solvate with<br />

LiPF 6 has a T m at 67 ◦ C with a solid-solid phase transition at 38 ◦ C.<br />

It is noteworthy that CuClO 4 <strong>and</strong> CuBF 4 form 4/1 phases 31–36 rather<br />

than a 5/1 phase, as is also found for LiClO 4 <strong>and</strong> LiBF 4 2 given that<br />

the Li + <strong>and</strong> Cu(I) + cations are nearly identical in size. 27,37 A eutectic<br />

point is observed for more concentrated mixtures (n < 5) between the<br />

5/1 solvate <strong>and</strong> a more aggregated solvate (composition unknown).<br />

Despite the fact that the phase diagrams for both LiPF 6 <strong>and</strong> LiTFSI<br />

suggests that these are highly dissociated salts (i.e., both are able to<br />

form a 6/1 phase), significant differences are also evident. In particular,<br />

LiPF 6 forms crystalline phases with a (relatively) high T m ,in<br />

contrast with LiTFSI which forms phases which melt at low temperature.<br />

Thus, (AN) n -LiTFSI mixtures remain liquid at −30 ◦ C over a<br />

large concentration range, whereas all of the (AN) n -LiPF 6 mixtures<br />

(except for the most dilute) crystallize readily at ambient temperature.<br />

These differences are attributed to the differences in size, shape <strong>and</strong><br />

flexibility of the anions. The PF 6 − anions are essentially spherical <strong>and</strong><br />

easily pack together symmetrically with the solvated Li + cations. At<br />

higher temperature, the uncoordinated PF 6 − anions may become disordered<br />

by spinning about one axis or tumbling. The solid-solid phase<br />

transition for the 5/1 solvate is a structural change in the crystal structure<br />

which accomodates this disorder at elevated temperature. The<br />

long-range order, however, is retained. The non-spherical shape of the<br />

TFSI − anions, in contrast, likely results in a less symmetrical packing<br />

of the solvated Li + cations amongst the uncoordinated anions. Further,<br />

the anions are flexible 38,39 <strong>and</strong> may become conformationally disordered<br />

at elevated temperature thus disrupting the long-range packing<br />

of the solvate structure leading to melting instead of a disordered solid<br />

crystalline (plastic crystalline) phase.<br />

Raman characterization of AN-Li + cation solvation.— Upon coordination<br />

with Li + cations, the vibrational b<strong>and</strong>s associated with the<br />

solvent C–C <strong>and</strong> C≡N stretching modes shift. Thus, these b<strong>and</strong>s may<br />

be utilized to determine the average solvation number (N) forAN<br />

coordinated to the Li + cations. Uncoordinated AN has a ν 4 b<strong>and</strong> at<br />

918 cm −1 for the C–C stretching vibration <strong>and</strong> a ν 2 b<strong>and</strong> at 2254 cm −1<br />

for the C≡N stretching vibration (Fig. 2), with the 922 <strong>and</strong> 2251 cm −1<br />

shoulders attributed to hot b<strong>and</strong>s. 40,41 When the electron lone-pair<br />

on the nitrogen is coordinated to a Li + cation, these b<strong>and</strong>s shift to<br />

930 <strong>and</strong> 2277 cm −1 , respectively (Fig. 2). 18,20,42,43 The following calculation<br />

may be used to determine the fraction of uncoordinated <strong>and</strong><br />

coordinated AN molecules:<br />

A AN-C<br />

= N c LiX<br />

[1]<br />

A AN-C + A AN−UC c AN<br />

where A AN-C <strong>and</strong> A AN-UC are the integrated area intensities of the b<strong>and</strong>s<br />

for the coordinated <strong>and</strong> uncoordinated AN, respectively, c LiX <strong>and</strong> c AN<br />

are the concentrations of the salt <strong>and</strong> AN, respectively, <strong>and</strong> N is the<br />

average solvation number. 44,45 From this information, values for N<br />

may be determined by multiplying the fraction of coordinated solvent<br />

by the total number of AN molecules present at a specified concentration<br />

(Fig. 3). The analysis in Fig. 3 is based upon the assumption that<br />

the relative activities of the b<strong>and</strong>s associated with the uncoordinated<br />

<strong>and</strong> coordinated AN have equivalent Raman activity (no scaling is required),<br />

as was done in the previous manuscript. 2 Data were collected<br />

for mixtures of (AN) n -LiTFSI <strong>and</strong> (AN) n -LiPF 6 at 60 ◦ C (note that<br />

similar data has previously been reported for (AN) n -LiTFSI mixtures<br />

at ambient temperature 18 ). For comparison, data for (AN) n -LiClO 4<br />

mixtures is included in Fig. 3. 2 The data for the different salts may<br />

then be compared for a specified composition. For example, for the<br />

composition x = 0.20 (or four AN molecules per Li + cation), the average<br />

solvation number is approximately 3.2 for LiPF 6 , 2.8 for LiTFSI,<br />

<strong>and</strong> 2.7 for LiClO 4 (Fig. 3). This, along with a previous study for more<br />

associated salts (for which the values were 2.1 for LiBF 4 <strong>and</strong> 1.0 for<br />

LiCF 3 CO 2 for x = 0.20), 2 indicates the following order for the Li +<br />

cation solvation number in liquid (AN) n -LiX mixtures:<br />

LiPF 6 > LiTFSI ≥ LiClO 4 > LiBF 4 >> LiCF 3 CO 2<br />

Based upon the suggested order for ionic association tendency indicated<br />

previously, however, the results for LiTFSI are surprising as the<br />

TFSI − anion was proposed to be a highly dissociated anion. 1 A comparison<br />

of the crystalline phases that form for (AN) n -LiTFSI, -LiPF 6<br />

<strong>and</strong> -LiClO 4 mixtures 2 does indeed suggest that, in the crystalline<br />

phases, the TFSI − −<br />

anions are highly dissociated (comparable to PF 6<br />

anions). But in the liquid phase, as will be shown below, the TFSI −<br />

anions may have an ionic association tendency closer to ClO − 4 than<br />

PF − 6 anions. Note, however, that TFSI − also tends to form bidentate<br />

coordination to a single Li + cation to a much greater extent than anions<br />

such as PF − 6 ,ClO − 4 <strong>and</strong> BF − 4 . This is evident from the QC studies<br />

of (AN) n -Li-(TFSI) m complexes <strong>and</strong> MD simulation results (see<br />

below), as well as the known crystal structures of numerous LiTFSI<br />

solvates. 18,26,46–49 Most of the solvated Li + cations have tetrahedral<br />

coordination. Thus, if a given TFSI − anion contributes two rather<br />

than one donor oxygens to the cation coordination, this may displace<br />

an additional AN molecule. . . thereby lowering the value of N to some<br />

extent from what would otherwise be predicted. This could therefore<br />

be one possible explanation for the lower than expected N values noted<br />

for the TFSI − anion.<br />

Raman characterization of ionic association.— Fig. 4 shows the<br />

Raman b<strong>and</strong> vibration from the expansion <strong>and</strong> contraction of the entire<br />

TFSI − anion 50,51 with changing concentration at −80 <strong>and</strong> 60 ◦ C. At<br />

−80 ◦ C, all of the samples are crystalline solids, except for those in the<br />

crystallinity gap which either remain liquid or are amorphous solids,<br />

depending upon the sample T g (Fig. 1). Two b<strong>and</strong>s are observed in the<br />

dilute mixtures (n ≥ 3.6) which do not vary in position with varying


A1492 Journal of The Electrochemical Society, 159 (9) A1489-A1500 (2012)<br />

Figure 2. Raman spectra at 60 ◦ C of AN C–C stretching mode (920 cm −1 )<strong>and</strong>C≡N stretching mode (2250 cm −1 ) b<strong>and</strong>s for (a) (AN) n -LiTFSI <strong>and</strong> (b) (AN) n -LiPF 6<br />

mixtures (AN/LiX ratio indicated).<br />

Figure 3. Raman spectroscopic analysis at 60 ◦ C of solvent b<strong>and</strong>s for uncoordinated AN <strong>and</strong> Li + cation coordinated AN in (AN) n -LiX mixtures with (a) LiPF 6 ,<br />

(b) LiTFSI <strong>and</strong> (c) LiClO 4 (the latter shown for comparison). The calculated Li + cation average solvation number (N) is shown at the top. The dark solid line<br />

corresponds to the average of the two sets of data from Fig. 2.


Journal of The Electrochemical Society, 159 (9) A1489-A1500 (2012)<br />

A1493<br />

Figure 5. TFSI − anion b<strong>and</strong> variation with temperature in (AN) n -LiTFSI<br />

mixtures with (a) n = 6.0 (crystalline in the −80 to −45 ◦ C range) <strong>and</strong> (b)<br />

n = 4.0 (crystalline in the −100 to −35 ◦ C range).<br />

Figure 4. TFSI − anion b<strong>and</strong> variation with concentration in (AN) n -LiTFSI<br />

mixtures at (a) −80 ◦ C<strong>and</strong>(b)60 ◦ C. The mixtures with n = 3.0, 2.5, 2.0 <strong>and</strong><br />

1.55 are in the crystallinity gap <strong>and</strong> remain either fully amorphous liquids or<br />

glassy solids at −80 ◦ C.<br />

concentration at −80 ◦ C. Although two different crystalline solvates<br />

(6/1 <strong>and</strong> 4/1 phases) are present for this concentration range (n ≥ 4)<br />

(Fig. 1), the anion coordination to the Li + cations does not change<br />

because both phases appear to consist of SSIP solvates in which the<br />

anions are uncoordinated. The TFSI − anion is known to have two<br />

different low-energy conformational states: a cisoid form (C 1 ) with<br />

the CF 3 groups on the same side of the S–N–S plane <strong>and</strong> a transoid<br />

form (C 2 ) with the CF 3 groups on opposite sides of the plane. 51 This<br />

difference in the b<strong>and</strong> position was initially thought to originate from<br />

different conformations of the anions in the solvate crystal structures.<br />

The data in Fig. 4a thus suggested that the 6/1 phase consists of<br />

uncoordinated TFSI − anions with the C 1 conformation, whereas the<br />

4/1 phase consists of uncoordinated TFSI − anions with both the C 1 <strong>and</strong><br />

C 2 conformations. To confirm this, the 250–450 cm −1 region of the<br />

spectra from (AN) n -LiTFSI mixtures with n = 6 <strong>and</strong> 4 was examined<br />

as the anion vibrational b<strong>and</strong>s in this region provide a fingerprint for<br />

the uncoordinated anion conformations (Fig. 5a). 51 The data, however,<br />

clearly indicate that both phases consist only of anions with the C 2<br />

conformation. Note, however, that new b<strong>and</strong>s are evident at 382 cm −1<br />

for the n = 6 sample (Fig. 5a) <strong>and</strong> at 315 cm −1 for the n = 4 sample<br />

(Fig. 5b) which do not correspond to the b<strong>and</strong>s typically noted for<br />

either the C 1 or C 2 conformations, so it may be that one or more<br />

different conformations are present which would also account for the<br />

b<strong>and</strong> at 741 cm −1 . Perhaps these b<strong>and</strong>s are related to a third lowenergy<br />

conformation for the TFSI − anion which has been reported, 52<br />

but unfortunately the Raman spectrum for this conformation is not<br />

yet known. In the crystallinity gap for more concentrated mixtures,<br />

the samples are liquid or amorphous solids at −80 ◦ C <strong>and</strong> broad b<strong>and</strong>s<br />

are evident in the Raman spectra. As the concentration of LiTFSI<br />

increases, the Raman b<strong>and</strong> variation indicates that the amount of SSIP<br />

solvates decreases <strong>and</strong> more aggregated solvates, i.e., CIP <strong>and</strong> AGG<br />

solvates, increase. As the composition approaches n = 1, a crystalline<br />

phase is formed again (along with an amorphous phase if 1 < n<br />

< 1.5). The Raman spectra show sharper b<strong>and</strong>s for the crystalline<br />

phases <strong>and</strong> broader b<strong>and</strong>s for the amorphous phase. For the n = 1<br />

mixture, the entire sample is crystalline with one sharp Raman b<strong>and</strong><br />

at 751 cm −1 , corresponding to the AGG-<strong>II</strong>b (AN) 1 :LiTFSI crystalline<br />

solvate (with the TFSI − anion coordinated to three Li + cations through<br />

four oxygens).<br />

For the liquid mixtures at 60 ◦ C (Fig. 4b), the shift of the<br />

Raman b<strong>and</strong>s for the TFSI − anion varies with concentration in a similar<br />

manner to that noted for the solid phase. However, unlike the solid


A1494 Journal of The Electrochemical Society, 159 (9) A1489-A1500 (2012)<br />

Figure 6. Solvate species distribution in (AN) n -LiTFSI mixtures at 60 ◦ C.<br />

phase, the b<strong>and</strong>s shift smoothly with changing concentration. Due to<br />

the structural flexibility of the TFSI − anion, different conformations<br />

are possible <strong>and</strong> each of these may be coordinated in varying ways<br />

to one or more Li + cation(s). Therefore, numerous types of solvates<br />

are formed in the liquid mixtures with changing concentration, giving<br />

overlapping b<strong>and</strong>s. This makes it challenging to deconvolute the<br />

spectra conclusively to identify specific forms of anion coordination<br />

to the Li + cations. This has been done, however, using preliminary<br />

assignments from a study underway in which the Raman spectra of<br />

crystalline LiTFSI solvates have been correlated with known crystal<br />

structures. From this analysis, the peak positions at −80 ◦ C/60 ◦ Care<br />

C 2 -SSIP (741/740 cm −1 – uncoordinated TFSI − anions), C 2 -CIP-<strong>II</strong><br />

(747/746 cm −1 – TFSI − anion coordinated to a single Li + cation<br />

through two oxygen atoms), C 2 -AGG-Ib (749/748 cm −1 – TFSI − anion<br />

coordinated to two Li + cations through three oxygen atoms) <strong>and</strong><br />

C 1 -AGG-<strong>II</strong>b (750-752/749-750 cm −1 – TFSI − anion coordinated to<br />

three Li + cations through four oxygen atoms). Note that in the liquid<br />

phase, these b<strong>and</strong>s may shift somewhat due to variability in the anion<br />

conformations <strong>and</strong> coordination bond lengths to the Li + cations.<br />

The close proximity of the b<strong>and</strong>s for AGG coordination (along with<br />

other forms of AGG solvates which have not yet been characterized)<br />

suggests that it is fruitless to attempt the identification of different<br />

forms of the AGG solvates in solution. Thus, Fig. 6 simply notes<br />

the deconvolution of the peaks in Fig. 4 in terms of uncoordinated<br />

TFSI − anions (SSIP), anions coordinated to a single Li + cation (CIP)<br />

<strong>and</strong> those coordinated to more than one Li + cation (AGG). Such an<br />

analysis suggests that dilute mixtures contain > 80% SSIP solvates,<br />

with the remainder CIP solvates. With increasing salt concentration,<br />

the fraction of both CIP <strong>and</strong> AGG solvates increases at the expense<br />

of the SSIP solvates, but the latter persist even for very concentrated<br />

mixtures.<br />

To examine why the crystallinity gap occurs for the (AN) n -LiTFSI<br />

mixtures, variable-temperature Raman spectra have been measured<br />

for compositions within the crystallinity gap (Fig. 7 <strong>and</strong> Supporting<br />

Information). For the n = 4.0 mixture, the Raman b<strong>and</strong>s at 741<br />

<strong>and</strong> 743 cm −1 (from uncoordinated TFSI − anions) increase as the<br />

temperature decreases, <strong>and</strong> the b<strong>and</strong> at 747–749 cm −1 (attributed<br />

to CIP <strong>and</strong> AGG-I solvates) decreases, indicating that the amount<br />

of SSIP solvates increases at lower temperature. This results in the<br />

nucleation <strong>and</strong> growth of the SSIP 4/1 crystalline phase. The n = 3.0<br />

sample, however, did not crystallize <strong>and</strong> the Raman spectra show that<br />

even though the amount of SSIP solvates increased as the temperature<br />

decreased, there remains a sizeable amount of coordinated anions<br />

(corresponds to Raman b<strong>and</strong> on 748 cm −1 ) thus hindering/preventing<br />

the formation of the 4/1 crystalline phase. Similarly, although the n<br />

Figure 7. Variable-temperature TFSI − anion b<strong>and</strong> variation with concentration<br />

in (AN) n -LiTFSI mixtures with (a) n = 4.0, (b) n = 3.0, (c) n = 2.0 <strong>and</strong><br />

(d) n = 1.0.<br />

= 2.0 sample has a dominant peak at 750 cm −1 (corresponding to an<br />

AGG-<strong>II</strong> solvate as found in the 1/1 phase), a significant fraction of<br />

the anions persist as SSIP, CIP <strong>and</strong> perhaps AGG-I solvates which<br />

hinder/prevent the nucleation <strong>and</strong> growth of the 1/1 crystalline phase.<br />

In contrast, for the n = 1.0 sample, most of the anions have AGG-<strong>II</strong><br />

(or perhaps even higher) aggregation, thus facilitating the nucleation<br />

of the 1/1 crystalline phase. Note that in the phase diagram for the<br />

(AN) n -LiTFSI mixtures in Fig. 1, the“×” symbols indicate the T g


Journal of The Electrochemical Society, 159 (9) A1489-A1500 (2012)<br />

A1495<br />

Figure 8. PF 6 − anion b<strong>and</strong> variation with concentration in (AN) n -LiPF 6 mixtures<br />

at (a) −80 ◦ C<strong>and</strong>(b)60 ◦ C.<br />

of the fully amorphous samples, whereas the “triangles” are the T g<br />

for the amorphous phase which remains after some portion of a sample<br />

has crystallized as the 1/1 phase. The composition of the latter<br />

amorphous phase will thus be more dilute than the composition of<br />

the sample itself. These samples have nearly the same T g value of<br />

−82 ◦ C <strong>and</strong> this value can be used to estimate the composition of the<br />

stabilized amorphous phase (i.e., x ∼ 0.25). This value corresponds<br />

to an average of 3 AN molecules per Li + cation.<br />

Fig. 8a shows the b<strong>and</strong> variation of the PF 6 − anion with varying<br />

concentration at −80 <strong>and</strong> 60 ◦ C. Based upon O h symmetry, the b<strong>and</strong><br />

assignments have been determined. 53–58 The most intense Raman b<strong>and</strong><br />

for the PF 6 − anion is observed at 740–750 cm −1 .At−80 ◦ C, all of the<br />

samples are crystalline solids. For dilute mixtures (n ≥ 5), only a single<br />

b<strong>and</strong> is evident at 744 cm −1 . This b<strong>and</strong> corresponds to SSIP solvates<br />

in which the PF 6 − anion is uncoordinated. There appears to be a slight<br />

shift to higher wavenumber between the 6/1 <strong>and</strong> 5/1 samples (perhaps<br />

corresponding to differences in structure (i.e., lattice packing) between<br />

the (AN) 6 :LiPF 6 <strong>and</strong> (AN) 5 :LiPF 6 crystalline solvates). A new b<strong>and</strong><br />

appears at 748 cm −1 as the concentration increases (n ≤ 5). This<br />

b<strong>and</strong> corresponds to an as yet undetermined more aggregated phase<br />

(perhaps a 1/1 phase).<br />

For the liquid mixtures at 60 ◦ C(Fig.8b), the Raman b<strong>and</strong> for the<br />

PF 6 − anion shifts to lower wavenumber <strong>and</strong> broadens. In dilute mixtures,<br />

one Raman b<strong>and</strong> is observed at 741 cm −1 . This b<strong>and</strong> corresponds<br />

to SSIP solvates. As the concentration increases, an asymmetric shoulder<br />

at higher wavenumbers grows with increasing concentration. This<br />

Figure 9. Geometries of the (a) (AN) 3 -Li-TFSI (A: C 1 -CIP-I), (b) (AN) 3 -<br />

Li-TFSI (B: C 2 -CIP-I), (c) (AN) 3 -Li-TFSI (C: C 1 -CIP-<strong>II</strong>), (d) (AN) 3 -Li-TFSI<br />

(D: C 1 -CIP-<strong>II</strong>), (e) (AN) 2 -Li-(TFSI) 2 (A: C 1 -CIP-I, C 2 -CIP-<strong>II</strong>), (f) (AN) 2 -Li-<br />

(TFSI) 2 (B: C 2 -CIP-<strong>II</strong>, C 2 -CIP-<strong>II</strong>), (g) Li-(TFSI) 2 (C 2 -CIP-<strong>II</strong>, C 2 -CIP-<strong>II</strong>), (h)<br />

(AN) 3 -Li-PF 6 <strong>and</strong> (i) (AN) 2 -Li-(PF 6 ) 2 clusters optimized using the M05-2X<br />

functional with the 6-31+G(d) basis set (optimized in the gas phase) (Li-purple,<br />

N-blue, O-red, S-yellow, F-light green).<br />

is attributed to one or more b<strong>and</strong>s due to CIP <strong>and</strong> AGG solvates. It is<br />

clear that SSIP solvates dominate the mixtures for dilute concentrations<br />

<strong>and</strong> such solvates are found even in the concentrated mixtures.<br />

This confirms that LiPF 6 is a highly dissociated salt.<br />

QC studies <strong>and</strong> force field validation.— QC studies of (AN) n -Li-<br />

(anion) m complexes focused on the calculation of the binding energies,<br />

which were then compared with the corresponding energies obtained<br />

from molecular mechanics (MM) geometry optimizations employing<br />

a polarizable force field (FF). QC studies were performed on the<br />

clusters shown in Fig. 9. Binding energies relative to the isolated<br />

species in the gas-phase are shown in Table I together with the previously<br />

reported binding energies for the analogous complexes with<br />

BF 4 − <strong>and</strong> ClO 4 − anions. 2 The binding energies, after BSSE correction<br />

from the (computationally) inexpensive DFT M05-2X/6-31+G*<br />

method, are found to be generally in fair agreement (< 3 kcal mol −1 )<br />

with the significantly more expensive results from MP2/aug-cc-pvTz<br />

calculations. The MP2/aug-cc-pvDz binding energies, on the other<br />

h<strong>and</strong>, showed larger deviations from the MP2/aug-cc-pvTz energies<br />

(Table I). Thus, the inexpensive M05-2X/6-31+G* binding energies<br />

were deemed acceptable for the solvent-Li-anion force field validation,<br />

instead of the MP2/aug-cc-pvDz energies, for the cases when<br />

the MP2/aug-cc-pvTz energies are too expensive to calculate, such as<br />

for the (AN) n -Li-(TFSI) m clusters shown in Fig. 9. An examination of


A1496 Journal of The Electrochemical Society, 159 (9) A1489-A1500 (2012)<br />

Table I. (AN) n -Li-(anion) m complex binding energies (in kcal mol −1 ) from QC calculations of the geometry using the M05-2X/6-31+G(d) level<br />

<strong>and</strong> molecular mechanics MM (FF).<br />

M052x/6-31+G(d) MP2/Dz MP2/Tz MM (FF) MM(FF)-MP2/Tz MM(FF)-M052x/6-31+G*<br />

(AN) 3 -Li-BF 4 −191.4 −185.9 −189.1 −188.9 0.2 2.5<br />

(AN) 3 -Li-ClO 4 −188.5 −184.3 −188.0 −188.0 0.1 0.5<br />

(AN) 3 -Li-TFSI(A) (C 1 -CIP-I) −184.0 −180.4 −183.6 0.5<br />

(AN) 3 -Li-TFSI(B) (C 2 -CIP-I) −180.8 −177.3 −182.0 −1.2<br />

(AN) 3 -Li-TFSI(C) (C 1 -CIP-<strong>II</strong>) −181.5 −178.2 −181.4 0.1<br />

(AN) 3 -Li-TFSI(D) (C 1 -CIP-<strong>II</strong>) −180.9 −177.4 −180.8 0.0<br />

(AN) 3 -Li-PF 6 −184.5 −179.0 −182.8 −183.4 −0.6 1.1<br />

(AN) 2 -Li-(BF 4 ) 2 −205.8 −199.6 −202.9 −203.8 −0.9 2.0<br />

(AN) 2 -Li-(ClO 4 ) 2 −201.1 −196.4 −200.6 −201.6 −1.0 −0.5<br />

(AN) 2 -Li-(TFSI) 2 (A) −199.7 −198.8 0.9<br />

(AN) 2 -Li-(TFSI) 2 (B) −198.5 −194.5 −197.8 0.7<br />

(AN) 2 -Li-(PF 6 ) 2 −193.6 −188.9 −192.1 −191.8 0.3 1.8<br />

Li-(TFSI) 2 −185.0 −178.8 −183.3 1.6<br />

the solvate structures in Fig. 9 suggests that these do not necessarily<br />

represent the lowest energy structures in solution as the AN molecules<br />

‘wrap’ around the anions to optimize the solvate energetics as there<br />

is no surrounding shell of solvent molecules <strong>and</strong> ions with which to<br />

interact. Nevertheless, this is not expected to impact the results significantly<br />

<strong>and</strong> a comparison of the calculated solvate binding energies<br />

<strong>and</strong> geometries is very informative.<br />

The DFT binding energies shown in Table I indicate that the magnitude<br />

of the complex stability follows the order PF − 6 ∼ TFSI −<br />

< ClO − 4 < BF − 4 with the BF − 4 anion being the most stable. This<br />

order is different from the order of the Li + . . . anion binding energies<br />

without any solvent present (i.e., TFSI − < ClO − 4 < PF − 6 < BF − 4 )<br />

reported by Johansson from MP2/6-31G(d) level calculations. 59 Thus,<br />

inclusion of the solvent in the Li + . . . anion complex calculations<br />

changes the order of the binding energy stabilization. Fig. 9a, 9b<br />

<strong>and</strong> 9h also shows that the Li + cation binds to one TFSI − oxygen<br />

or one PF − 6 fluorine atom in the most stable configurations for the<br />

(AN) 3 -Li-anion clusters (Table I), which is different from the binding<br />

arrangement in the most stable Li-PF 6 <strong>and</strong> Li-TFSI complexes (without<br />

AN), where the Li + cation is bound to three fluorine atoms of<br />

the PF − 6 anion (CIP-<strong>II</strong>I) or to two oxygen atoms of the TFSI − anion<br />

(CIP-<strong>II</strong>). 60–64 It was found, however, that a stable configuration of the<br />

(AN) 3 -Li-TFSI complex with two TFSI − oxygen atoms bound to a<br />

Li + cation (geometry C - Fig. 9c) is only 2-3 kcal mol −1 less stable<br />

than the configuration with one TFSI − anion oxygen bound to a Li +<br />

cation (geometry A - Fig. 9a). The (AN) 3 -Li-TFSI complexes C <strong>and</strong> D<br />

have the TFSI − anion in the same conformation with φ C-S..S-C = 105 ◦<br />

<strong>and</strong> bidentate binding to a Li + cation, but different arrangements of the<br />

AN molecules (Fig. 9c <strong>and</strong> 9d) resulting in a slight (< 1 kcal mol −1 )<br />

energy difference between the solvates. Amongst the (AN) 3 -Li-TFSI<br />

complexes A <strong>and</strong> B (Fig. 9a <strong>and</strong> 9b) with monodentate binding of a<br />

TFSI − anion to a Li + cation, complex A with the C 1 TFSI − anion<br />

conformer is more stable than complex B with the C 2 TFSI − anion<br />

conformer.<br />

In order to further investigate the tendency of the TFSI − anion<br />

to contribute one or two oxygen atoms to the anion. . . Li + cation<br />

coordination, two (AN) 2 -Li-(TFSI) 2 complexes were examined: in<br />

complex A, one TFSI − anion is bound to a Li + cation with one oxygen,<br />

while the other TFSI − anion is bound to the Li + cation with two<br />

oxygen atoms (Fig. 9e); in complex B, both TFSI − anions contribute<br />

two oxygen atoms each to the anion. . . Li + cation coordination (Fig.<br />

9f). The DFT geometry optimization of these complexes started from<br />

a configuration with two AN molecules bound to a Li + cation, but<br />

the optimized geometry instead converging to a structure with only<br />

one of the AN molecules bound to the Li + cation while the other<br />

AN is located in close proximity, but directed away from the Li +<br />

cation (Fig. 9e <strong>and</strong> 9f). Complex A was found to be more stable<br />

by about 1 kcal mol −1 than complex B (Table I) indicating that, for<br />

the (AN) 2 -Li-(TFSI) 2 complexes, the configuration with two TFSI −<br />

anions contributing both oxygen atoms to the anion. . . Li + cation<br />

coordination is not the most energetically favorable (although the<br />

difference is not large). In contrast to the (AN) 2 -Li-(TFSI) 2 complexes,<br />

which do not have two AN molecules bound to the Li + cationinthe<br />

most stable configurations, the most stable (AN) 2 -Li-(PF 6 ) 2 complex<br />

does have two AN molecules, in addition to the two PF 6 − anions,<br />

bound to the Li + cation (Fig. 9i).<br />

Next, the ability of the many-body polarizable force field to predict<br />

the binding energies <strong>and</strong> geometries (relative to the QC calculations)<br />

was examined for the (AN) n -Li-(anion) m clusters. The MM<br />

optimization using the developed FF predicted cluster geometries is in<br />

acceptable agreement with the geometries from the DFT calculations.<br />

For example, for the (AN) 3 -Li-PF 6 cluster, the distances are R(Li-N)<br />

= 2.0–2.1 Å from both DFT <strong>and</strong> MM(FF) <strong>and</strong> R(Li-F) = 1.8 Å from<br />

MM(FF) <strong>and</strong> 1.9 Å from DFT. For the (AN) 3 -Li-(PF 6 ) 2 complex,<br />

R(Li-N) = 2.1 Å from both DFT <strong>and</strong> MM(FF), while R(Li-F) was in<br />

the range of 1.7–1.9 Å from MM(FF) <strong>and</strong> 1.9–2.1 Å from the DFT calculations.<br />

Excellent agreement (less than 0.05 Å deviations) between<br />

R(Li-O) <strong>and</strong> Li-N(AN) distances from the DFT <strong>and</strong> MM(FF) calculations<br />

were also observed for both the A <strong>and</strong> B (AN) 3 -Li-TFSI geometries.<br />

The complex binding energies from the MM(FF) calculations are<br />

compared with the QC results in Table I. Excellent agreement is observed<br />

between the DFT M05-2X/6-31+G* <strong>and</strong> MM(FF) predictions<br />

with deviations of 1.1 kcal mol −1 or less for the (AN) 3 -Li-anion complexes<br />

<strong>and</strong> 1.8 kcal mol −1 or less for the (AN) 2 -Li-(anion) 2 complexes<br />

with the PF 6 − <strong>and</strong> TFSI − anions. Similar quality MM(FF) predictions<br />

were observed when compared to a more accurate MP2/aug-cc-pvTz<br />

level. Note that MM(FF) correctly predicts that the A geometries<br />

of (AN) 3 -Li-TFSI <strong>and</strong> (AN) 2 -Li-(TFSI) 2 are more stable than the B<br />

geometries.<br />

DFT calculations were also employed to examine the shift of the<br />

TFSI − anion Raman b<strong>and</strong>s <strong>and</strong> intensities upon Li + cation complexation.<br />

While previous DFT calculations 62–64 have investigated the influence<br />

of Li + cation coordination on the TFSI − anion Raman <strong>and</strong><br />

IR b<strong>and</strong> vibrations, no solvent (i.e., AN) was included in the studies.<br />

The solvent is included, however, explicitly <strong>and</strong> implicitly through a<br />

PCM model in the present work. For this, an additional set of DFT<br />

calculations was performed on the (AN) n -Li-(TFSI) m solvates at a<br />

M05-2X/6-31+G* level using a self-consistent reaction field with the<br />

polarized continuum PCM model with AN parameters from the Gaussian<br />

g09 package. Initial geometries were taken from the gas-phase<br />

DFT calculations without PCM (Fig. 9). The resulting optimized geometries<br />

with PCM(AN) are shown in the Supporting Information.<br />

The b<strong>and</strong> frequencies <strong>and</strong> Raman activities are given in Table <strong>II</strong> for<br />

the spectral range around 740–750 cm −1 that is used for the experimental<br />

characterization of solvates. In this region, similar frequencies<br />

are noted for the C 1 <strong>and</strong> C 2 TFSI − anion conformations. Bidentate


Journal of The Electrochemical Society, 159 (9) A1489-A1500 (2012)<br />

A1497<br />

Table <strong>II</strong>. Raman frequency (ν) <strong>and</strong> activity (I) for (AN) n -Li-(TFSI) m solvates from M05-2X/6-31+G* calculations with PCM(AN) (solvates shown<br />

in Supporting Information Fig. S10).<br />

complex ν (cm −1 ) scaled ν a (cm −1 ) ν complex −ν SSIP I complex /I SSIP φ C-S...S-C ( ◦ )<br />

TFSI (C 1 -SSIP) 749.7 742.2 −39<br />

TFSI (C 2 -SSIP) 751.4 743.9 −174<br />

(AN) 3 -Li-TFSI(A) (C 1 -CIP-I) 753.9 746.3 3.3 0.93 −45<br />

(AN) 3 -Li-TFSI(B) (C 2 -CIP-I) 750.5 743.0 0.0 0.95 177<br />

(AN) 3 -Li-TFSI(C) (C 1 -CIP-<strong>II</strong>) 755.5 747.9 5.0 0.98 −127<br />

(AN) 3 -Li-TFSI(D) (C 1 -CIP-<strong>II</strong>) 756.6 749.0 6.1 0.98 −112<br />

(AN) 2 -Li-TFSI 758.3 750.7 7.7 1.01 −169<br />

(AN) 2 -Li-(TFSI) 2 758.9 751.3 8.4 1.09 179,179<br />

Li-(TFSI) 2 756.4 748.9 5.9 1.01 178,178<br />

a M05-2X/6-31+G* frequency multiplied by scaling factor of 0.99.<br />

binding of a Li + cation to the TFSI − anion (i.e., CIP-<strong>II</strong>) results in a<br />

frequency shift of 5–8 cm −1 (from the uncoordinated anion), while the<br />

monodentate binding of the TFSI − anion to a Li + cation (i.e., CIP-I)<br />

in (AN) 3 -Li-TFSI clusters results in a much smaller frequency shifts<br />

of 0 <strong>and</strong> 3 cm −1 , respectively, for C 2 <strong>and</strong> C 1 TFSI − anion conformers.<br />

The observed frequency shift upon bidentate binding of 5–8 cm −1<br />

is in good agreement with previous DFT calculations 64 <strong>and</strong> experimental<br />

data for solvates with known crystal structures. The calculated<br />

much smaller shift of 0-3 cm −1 for monodentate coordination is also<br />

qualitatively in accord with previously reported calculated results for<br />

Li + (TFSI(C 2 )) 4 complexes. 64 The small shift for TFSI − anion CIP-I<br />

coordination (relative to the uncoordinated anion) is an important<br />

point to note as this is potentially a source of error, perhaps a significant<br />

one, in the analysis performed in Figs. 4 <strong>and</strong> 6. No experimental<br />

data is available at present, however, to validate the DFT prediction<br />

that Li + cation complexation to a single oxygen atom of the TFSI −<br />

anion results in a frequency shift of less than < 3cm −1 for the TFSI −<br />

anion vibrational b<strong>and</strong> located near 740 cm −1 . Finally, the total energy<br />

of the (AN) 3 -Li-TFSI complexes A-D was calculated to estimate<br />

their stability with the inclusion of the implicit solvent PCM(AN).<br />

It was found that the energy of the (AN) 3 -Li-TFSI (A-D) complexes<br />

varied by less than 1 kcal mol −1 with the relative binding energies (to<br />

A) following the order: A (0 kcal mol −1 ) > B (0.4 kcal mol −1 ) > D<br />

(0.7 kcal mol −1 ) > C (1.0 kcal mol −1 ) (in contrast to the results reportedinTableI<br />

which did not included the polarized continuum PCM<br />

model for the surrounding solvent).<br />

MD simulations of (AN) n -LiTFSI <strong>and</strong> -LiPF 6 mixtures.— MD simulations<br />

have been used to examine the calculated solution solvate<br />

structures <strong>and</strong> to explore the ionic association tendence of the salts in<br />

AN. Three-dimensional snapshots of the simulations for the (AN) n -<br />

LiTFSI <strong>and</strong> -LiPF 6 mixtures with n = 30, 20, 10, 5 <strong>and</strong> 2 may be<br />

viewed using the .xyz files provided in the Supporting Information <strong>and</strong><br />

a freeware structure viewing program such as Mercury. The snapshots<br />

were created by modifying the simulation box results as follows—<br />

unwrapping all of the molecules/ions so that they were not divided,<br />

applying periodic boundary conditions so as not to divide the solvate<br />

structures across the boundaries of the simulation box (thereby retaining<br />

entire solvates) <strong>and</strong> removing solvent molecules if they were ><br />

2.90 Å from the Li + cations (uncoordinated <strong>and</strong> not in close proximity)<br />

to facilitate the viewing of the solvates.<br />

Table <strong>II</strong>I shows the fraction of the uncomplexed Li + cations <strong>and</strong><br />

anions (TFSI − or PF 6 − ) <strong>and</strong> the coordination of the anions as determined<br />

by the MD simulations. Li + cation coordination to the anions<br />

was defined as Li + cations within 4.74 or 3.70 Å, respectively, of<br />

the TFSI − or PF 6 − anions (N or P atoms, respectively). These values<br />

were chosen because the radial distribution function (RDF) g Li–N (r)<br />

for Li–N(anion) <strong>and</strong> g Li–P (r) for Li–P equals one after the first peak<br />

for these distances (Fig. 10). In the previous manuscript associated<br />

with this study of solution structure, 2 it was found to be advantageous<br />

to also scrutinize the coordination using a distance defined by g Li–X (r)<br />

= max(1st peak)/2, where for the present study X = P or N(anion)<br />

(i.e., 4.56 or 3.50 Å, respectively for TFSI − or PF 6 − (Table <strong>II</strong>I)),<br />

which may be interpreted as representing the strongly bound Li +<br />

cations, whereas the former values (4.74 or 3.70 Å) include both the<br />

strongly <strong>and</strong> loosely bound Li + cations.<br />

A comparison of the MD simulation results (Table <strong>II</strong>I) with the<br />

experimental data (Fig. 8) for the (AN) n -LiPF 6 mixtures suggests that<br />

there is reasonable agreement. The dilute mixtures are dominated<br />

by uncoordinated anions <strong>and</strong> fully solvated Li + cations, i.e., SSIP<br />

solvates (see Supporting Information). Examples of the solvate complexes<br />

found in the simulation dilute mixtures are shown in Fig. 11.<br />

Note that the bidentate coordination of the PF 6 − aniontoaLi + cation<br />

(i.e., Fig. 11f) is only rarely observed, as reflected by the similarity<br />

of the values for the number of fluorine <strong>and</strong> phosphorus atoms in<br />

close proximity (within 2.40 <strong>and</strong> 3.70 Å, respectively) to a Li + cation<br />

(Table <strong>II</strong>I). The average solvation numbers (for AN coordinated to Li +<br />

cations) from the MD simulations (Table <strong>II</strong>I) are also in reasonable<br />

accord with the experimental values (Fig. 3) for the most concentrated<br />

mixtures, but the values deviate for the n = 10 composition<br />

Figure 10. Radial distribution function (RDF) from the MD simulations for<br />

the (a) (AN) n -LiTFSI <strong>and</strong> (b) AN-LiPF 6 (n = 10) mixtures.


A1498 Journal of The Electrochemical Society, 159 (9) A1489-A1500 (2012)<br />

Table <strong>II</strong>I. Composition of the ion coordination shell from MD simulations of (AN) n -LiTFSI <strong>and</strong> (AN) n -LiPF 6 mixtures at 60 ◦ C (Note: some of the<br />

percentages do not sum to 100% due to the rounding off of the values).<br />

LiTFSI<br />

Property (AN:Li ratio): 30 20 10 5 2<br />

No. solvent in MD box 480 640 640 640 512<br />

No. LiTFSI in MD box 16 32 64 128 256<br />

Concentration (M) 0.56 0.81 1.46 2.43 3.98<br />

Molality (mol kg −1 ) 0.72 1.03 1.78 2.80 4.27<br />

Simulation run length a (ns) 17.3 (6) 12.2 (4) 7.0 (9) 15.3 (16) 15.0 (16)<br />

Simulation box length (Å) 36.30 40.35 41.74 44.38 43.11<br />

MD density (g cm −3 ) 0.844 0.896 1.019 1.197 1.469<br />

Expt density (g cm −3 ) 0.861 0.887 1.005 1.176 (1.385) b<br />

Fraction of free Li (r Li–N > 4.74 Å) (SSIP) 0.37 0.31 0.20 0.10 0.02<br />

Fraction of free N* (r Li–N > 4.74 Å) (SSIP) 0.36 0.27 0.16 0.06 0.01<br />

Li + coordination numbers<br />

# N** (within 2.40 Å of Li + ) 3.00 2.88 2.63 2.23 1.45<br />

# O (within 2.40 Å of Li + ) 0.83 0.97 1.24 1.67 2.49<br />

# N* (within 4.74 Å of Li + ) 0.75 0.89 1.14 1.57 2.42<br />

Probability of finding the following number of Li + cations within the given distance from the N of TFSI −<br />

0Li + within 4.56/4.74 Å of N (SSIP) 0.41/0.36 0.32/0.27 0.25/0.16 0.10/0.06 0.02/0.01<br />

1Li + within 4.56/4.74 Å of N (CIP) 0.52/0.55 0.57/0.58 0.50/0.56 0.47/0.41 0.20/0.12<br />

2Li + within 4.56/4.74 Å of N (AGG-I) 0.07/0.10 0.11/0.14 0.22/0.25 0.36/0.42 0.45/0.41<br />

3Li + within 4.56/4.74 Å of N (AGG-<strong>II</strong>) 0.00/0.00 0.01/0.01 0.03/0.03 0.07/0.10 0.28/0.37<br />

4Li + within 4.56/4.74 Å of N (AGG-<strong>II</strong>I) 0.00/0.00 0.00/0.00 0.00/0.00 0.00/0.01 0.05/0.09<br />

* N from TFSI − , ** N from AN<br />

LiPF 6<br />

Property (AN:Li ratio): 30 20 10 5 2<br />

No. solvent in MD box 480 640 640 640 512<br />

No. LiPF 6 in MD box 16 32 64 128 256<br />

Concentration (M) 0.58 0.87 1.67 3.06 5.89<br />

Molality (mol kg −1 ) 0.72 1.03 1.78 2.80 4.27<br />

Simulation run length a (ns) 12.0 (4) 11.0 (4) 7.7 (2) 18.0 (4) 14.0 (4)<br />

Simulation box length (Å) 35.78 39.42 39.71 40.51 40.02<br />

MD density (g cm −3 ) 0.802 0.844 0.938 1.091 1.379<br />

Expt density (g cm −3 ) 0.814 0.842 0.936<br />

Fraction of free Li (r Li–P > 3.70 Å) (SSIP) 0.66 0.61 0.47 0.30 0.06<br />

Fraction of free P (r Li–P > 3.70 Å) (SSIP) 0.66 0.60 0.44 0.24 0.02<br />

Li + coordination numbers<br />

# N (within 2.40 Å of Li + ) 3.42 3.37 3.15 2.84 1.81<br />

# F (within 2.40 Å of Li + ) 0.38 0.44 0.68 1.02 2.13<br />

# P (within 3.70 Å of Li + ) 0.36 0.42 0.62 0.96 2.00<br />

Probability of finding the following number of Li + cations within the given distance from the P of PF − 6<br />

0Li + within 3.50/3.70 Å of P (SSIP) 0.70/0.66 0.65/0.60 0.50/0.44 0.31/0.24 0.11/0.02<br />

1Li + within 3.50/3.70 Å of P (CIP) 0.29/0.33 0.34/0.38 0.46/0.50 0.56/0.58 0.34/0.24<br />

2Li + within 3.50/3.70 Å of P (AGG-I) 0.01/0.01 0.01/0.02 0.04/0.06 0.12/0.18 0.36/0.50<br />

3Li + within 3.50/3.70 Å of P (AGG-<strong>II</strong>) 0.00/0.00 0.00/0.00 0.00/0.00 0.00/0.01 0.16/0.22<br />

4Li + within 3.50/3.70 Å of P (AGG-<strong>II</strong>I) 0.00/0.00 0.00/0.00 0.00/0.00 0.00/0.00 0.02/0.02<br />

a equilibration run lengths are given in parentheses,<br />

b experimental density value in parentheses was extrapolated from experimental data.<br />

(N = 3.15 from the MD simulation, but ∼4 from the experimental<br />

results). Note that for the n = 10 composition at 60 ◦ C, there<br />

is no significant evidence of the presence of CIP or AGG solvates<br />

(Fig. 8b), but CIP <strong>and</strong> AGG solvates are found in the MD simulations<br />

for the most dilute mixtures studied with n = 10, 20 <strong>and</strong> 30<br />

(see Supporting Information). Thus, the simulations (i.e., force fields)<br />

appear to overpredict the degree of ionic association of the electrolyte<br />

mixtures. This was also true for the mixtures with LiBF 4 (<strong>and</strong> perhaps<br />

LiClO 4 ). 2<br />

For the highly concentrated (AN) n -LiTFSI mixtures, reasonable<br />

agreement is noted between the MD simulations <strong>and</strong> experimental<br />

data. For example, for the n = 5 concentration, simulations predict<br />

that the Li + cations have (on average) 0.6 less AN in the first coordination<br />

shell for the (AN) n -LiTFSI mixture as compared to the analogous<br />

(AN) n -LiPF 6 mixture, in reasonable agreement with the experimental<br />

observation for x = 0.20 (i.e., n = 4) where coordination of the Li +<br />

cations in the (AN) n -LiTFSI mixture contains 0.4 less AN (on average)<br />

than in the (AN) n -LiPF 6 mixture (Fig. 3). For the dilute mixtures,<br />

however, poor agreement is found between the MD simulations<br />

(Table <strong>II</strong>I) <strong>and</strong> experimental data (Fig. 6) for the SSIP, CIP <strong>and</strong> AGG<br />

distributionin(AN) n -LiTFSI mixtures. The experimental spectroscopic<br />

data for the anion (Figs. 4 <strong>and</strong> 6) indicates that the LiTFSI salt<br />

is highly dissociated, although to a lesser extent than LiPF 6 .Thisis<br />

also evident in the phase behavior (Fig. 1), as both LiPF 6 <strong>and</strong> LiTFSI<br />

form a SSIP 6/1 crystalline phase which is not found for the LiClO 4<br />

<strong>and</strong> LiBF 4 (intermediately associated) salts. 2 The results from the MD<br />

simulations, however, suggest that, for the dilute mixtures, the majority<br />

of the TFSI − anions are coordinated to the Li + cations as CIP


Journal of The Electrochemical Society, 159 (9) A1489-A1500 (2012)<br />

A1499<br />

Figure 11. Representative Li + cation solvate species (i.e., coordination shells)<br />

extracted from the MD simulations for the (AN) n -LiPF 6 mixtures (n = 30, 20<br />

<strong>and</strong> 10) at 60 ◦ C: (a) SSIP, (b) CIP-I, (c) AGG-I, (d) CIP-I (×2), (e) CIP-I,<br />

AGG-I <strong>and</strong> (f) CIP-I, CIP-<strong>II</strong> (Li-purple, N-blue, P-orange, F-light green).<br />

solvates. Further, from the MD simulations, the solvation number for<br />

LiTFSI (Table <strong>II</strong>I) is lower than for LiClO 4 2 which also conflicts with<br />

expectations from the experimental data—note that most of the Li +<br />

cations have 4-fold coordination. . . thus, a lower solvation number<br />

equates to an increase in anion coordination. There is no experimental<br />

evidence which supports the MD simulation conclusion that LiTFSI<br />

is more associated than LiClO 4 . One potential source of error in the<br />

estimated experimental solvate distribution (Fig. 6) is the use of a<br />

b<strong>and</strong> at 747 cm −1 to estimate the fraction of CIP solvates during the<br />

b<strong>and</strong> deconvolution. The QC calculation results (Table <strong>II</strong>), however,<br />

suggest that this may only account for the CIP-<strong>II</strong> (bidentate) coordinated<br />

TFSI − anions, with the CIP-I (monodentate) coordinated anions<br />

included, in part,with the SSIP fraction. In reality, a new b<strong>and</strong> would<br />

need to be introduced during the deconvolution (for CIP-I solvates)<br />

which would change both the SSIP <strong>and</strong> CIP (i.e., CIP-<strong>II</strong>) b<strong>and</strong> areas.<br />

Thus, the fraction of SSIP <strong>and</strong> CIP solvates should perhaps be<br />

smaller <strong>and</strong> larger, respectively. There is no experimental proof yet<br />

available regarding the b<strong>and</strong> position for the monodentate anion coordination,<br />

but this is discussed to bring this possible source of error<br />

in the analysis to the reader’s attention. Even so, it is still evident that<br />

the MD simulations predict for the dilute mixtures increased ionic<br />

association (fraction of CIP <strong>and</strong> AGG solvates) relative to the experimental<br />

data—this is true for all of the salts studied thus far (i.e., LiPF 6 ,<br />

LiTFSI, LiClO 4 <strong>and</strong> LiBF 4 ).<br />

Despite the discrepancies between the simulation <strong>and</strong> experimental<br />

results, it is quite informative to examine the manner in which the<br />

TFSI − anions coordinate the Li + cations in the solvates found in the<br />

MD simulations (Fig. 12). In contrast with semi-spherical anions such<br />

as PF 6 − ,ClO 4 − <strong>and</strong> BF 4 − , bidentate coordination of the TFSI − anion<br />

Figure 12. Representative Li + cation solvate species (i.e., coordination shells) extracted from the MD simulations for the (AN) n -LiTFSI mixtures (n = 30, 20 <strong>and</strong><br />

10) at 60 ◦ C: (a) C 1 -SSIP, (b) C 1 -CIP-I, (c) C 2 -AGG-I, (d) C 2 -CIP-I, C 1 -CIP-<strong>II</strong> (e) C 2 -CIP-I (×2), C 1 -AGG-I, (f) C 2 -CIP-<strong>II</strong> (×2), (g) C 1 -CIP-<strong>II</strong>, (h) C 1 -AGG-I<br />

(×2), (i) C 1 -AGG-I, C 2 -AGG-I <strong>and</strong> (j) C 1 -CIP-I (×2), C 1 -AGG-I, C 2 -AGG-I, C 2 -AGG-<strong>II</strong> (Li-purple, N-blue, O, red, S-yellow, F-light green).


A1500 Journal of The Electrochemical Society, 159 (9) A1489-A1500 (2012)<br />

to a given Li + cation in the MD simulations is not uncommon<br />

(Fig. 12d, 12f <strong>and</strong> 12g). For example, for the n = 10 concentration,<br />

12% of the TFSI − anions have two oxygen atoms from the same anion<br />

bound to the same Li + cation, which is similar to the observations from<br />

previous simulations of (EC) n -LiTFSI 65 <strong>and</strong> (ionic liquid) n -LiTFSI 66<br />

mixtures. As noted above, this, in part, may explain the lower than<br />

expected average solvent coordination numbers noted experimentally<br />

for LiTFSI (as comparable to LiClO 4 ) despite the additional experimental<br />

evidence (i.e., phase diagram <strong>and</strong> Raman spectroscopy of the<br />

TFSI − anion coordination) which suggest that LiTFSI is more dissociated<br />

than LiClO 4 (but less dissociated than LiPF 6 ) in AN. Thus,<br />

the designation of the TFSI − anion as “highly dissociated” may be<br />

questionable <strong>and</strong> it may be more appropriate to categorize this anion<br />

instead as “dissociated” with the PF − 6 anion retaining the “highly<br />

dissociated” appellation.<br />

Conclusions<br />

The solid-state <strong>and</strong> solution structure of electrolyte mixtures consisting<br />

of AN <strong>and</strong> either LiTFSI or LiPF 6 have been examined in<br />

detail using phase diagrams <strong>and</strong> Raman spectroscopy. Both salts are<br />

able to form a 6/1 AN/LiX crystalline phase which is not found for<br />

more associated salts. LiPF 6 forms crystalline phases with a high T m ,<br />

in contrast with LiTFSI which forms phases which melt at low temperature.<br />

Thus, (AN) n -LiTFSI mixtures remain liquid at −30 ◦ C over<br />

a large concentration range, whereas the (AN) n -LiPF 6 mixtures (except<br />

for the most dilute) crystallize readily at ambient temperature.<br />

LiPF 6 <strong>and</strong> LiTFSI are found to be highly dissociated <strong>and</strong> dissociated,<br />

respectively, in dilute mixtures. When CIP or AGG solvates do form,<br />

bidentate coordination of Li + cations is more prominent for TFSI −<br />

than for PF − 6 anions. QC calculations <strong>and</strong> MD simulations have been<br />

used to extensively complement this study. Taken together, the thermal<br />

phase behavior <strong>and</strong> spectroscopic analysis, when paired with the<br />

structural information provided by the crystalline solvates <strong>and</strong> simulations,<br />

give tremendous insight into the solution structure of these<br />

electrolytes.<br />

Supporting Information<br />

Supporting Information files include exp<strong>and</strong>ed views of the DSC<br />

data for the (AN) n -LiTFSI <strong>and</strong> LiPF 6 mixtures. In addition, Li +<br />

cation coordination <strong>and</strong> ion packing diagrams are provided for the<br />

(AN) 1 :LiTFSI, (AN) 6 :LiPF 6 <strong>and</strong> (AN) 5 :LiPF 6 solvate crystal structures.<br />

Files are also provided for the MD simulations in .xyz format<br />

(text files). These files are linked to the electronic version of this<br />

manuscript. 67<br />

Acknowledgments<br />

The authors wish to express their gratitude to the U.S. Department<br />

of Energy, Office of Basic Energy Sciences, Division of Materials<br />

Sciences <strong>and</strong> Engineering which fully supported the experimental research<br />

under Award DE-SC0002169. The computational work was<br />

partially supported by an Interagency Agreement between the U.S.<br />

Department of Energy <strong>and</strong> the U. S. Army Research Laboratory under<br />

DE-IA01-11EE003413 for the Office of Vehicle Technologies<br />

Programs including the Batteries for Advanced Transportation Technologies<br />

(BATT) Program.<br />

References<br />

1. W. A. Henderson, J. Phys. Chem. B, 110, 13177 (2006).<br />

2. D. M. Seo, O. Borodin, S.-D. Han, Q. Ly, P. D. Boyle, <strong>and</strong> W. A. Henderson, J.<br />

Electrochem. Soc., 159, A553 (2012).<br />

3. Y. Abu-Lebdeh <strong>and</strong> I. Davidson, J. Electrochem. Soc., 156, A60 (2009).<br />

4. M. Nagahama, N. Hasegawa, <strong>and</strong> S. Okada, J. Electrochem. Soc., 157, A748 (2010).<br />

5. Y. Abu-Lebdeh <strong>and</strong> I. Davidson, J. Power Sources, 189, 576 (2009).<br />

6. M. J. Frisch et al., Gaussian, Inc., Wallingford CT, 2010.<br />

7. J. B. Hooper <strong>and</strong> O. Borodin, Phys. Chem. Chem. Phys., 12, 4635 (2010).<br />

8. O. Borodin, J. Phys. Chem. B, 113, 11463 (2009).<br />

9. O. Borodin <strong>and</strong> G. D. Smith, J. Phys. Chem. B, 113, 1763 (2009).<br />

10. M. J. Barrow, Acta Crystallogr. B, 37, 2239 (1981).<br />

11. O. K. Antson, K. J. Tilli, <strong>and</strong> N. H. Andersen, Acta Crystallogr. B, 43, 296<br />

(1987).<br />

12. B. H. Torrie <strong>and</strong> B. M. Powell, Mol. Phys., 75, 613 (1992).<br />

13. H. Abramczyk <strong>and</strong> K. Paradowska-Moszkowska, Chem. Phys., 265, 177 (2001).<br />

14. R. Enjalbert <strong>and</strong> J. Galy, Acta Crystallogr. B, 58, 1005 (2002).<br />

15. Y. Suzuki, M. Sato, K. Takanohashi, T. Ida, <strong>and</strong> M. J. Mizuno, J. Phys. Chem. A, 112,<br />

13481 (2008).<br />

16. A. Olejniczak <strong>and</strong> A. Katrusiak, J. Phys. Chem. B, 112, 7183 (2008).<br />

17. S. Hore, R. Dinnebier, W. Wen, J. Hanson, <strong>and</strong> J. Maier, Z. Anorg. Allg. Chem., 635,<br />

88 (2009).<br />

18. D. Brouillette, D. E. Irish, N. J. Taylor, G. Perron, M. Odziemkowski, <strong>and</strong><br />

J. E. Desnoyers, Phys. Chem. Chem. Phys., 4, 6063 (2002).<br />

19. W. Kunz, J. Barthel, L. Klein, T. Cartailler, P. Turq, <strong>and</strong> B. Reindl, J. Soln. Chem.,<br />

20, 875 (1991).<br />

20. E. M. Cabaleiro-Lago <strong>and</strong> M. A. Ríos, Chem. Phys., 254, 11 (2000).<br />

21. X. Xuan, H. Zhang, J. Wang, <strong>and</strong> H. Wang, J. Phys. Chem. A, 108, 7513 (2004).<br />

22. D. Spångberg <strong>and</strong> K. Hermansson, Chem. Phys., 300, 165 (2004).<br />

23. E. Pasgreta, R. Puchta, A. Zahl, <strong>and</strong> R. van Eldik, Eur. J. Inorg. Chem., 1815 (2007).<br />

24. Y. Yokota, V. G. Young Jr., <strong>and</strong> J. G. Verkade, Acta Crystallogr. C, 55, 196 (1999).<br />

25. C. L. Raston, C. R. Whitaker, <strong>and</strong> A. H. White, Aust. J. Chem., 42, 201 (1989).<br />

26. D. M. Seo, P. D. Boyle, <strong>and</strong> W. A. Henderson, Acta Crystallogr. E, 67, m534 (2011).<br />

27. D. M. Seo, P. D. Boyle, O. Borodin, <strong>and</strong> W. A. Henderson, RSC Adv., in-press (2012).<br />

28. D. M. Seo, P. D. Boyle, <strong>and</strong> W. A. Henderson, Acta Crystallogr. E, 67, m1148 (2011).<br />

29. J. R. Black, W. Levason, <strong>and</strong> M. Webster. Acta Crystallogr. C, 51, 623 (1995).<br />

30. L. A. Dakin, P. C. Ong, J. S. Panek, R. J. Staples, <strong>and</strong> P. Stavrospoulos,<br />

Organometallics, 19, 2896 (2000).<br />

31. I. Csöregh, P. Kierkegaard, <strong>and</strong> K. Norrestam, Acta Crystallogr. B, 31, 314 (1975).<br />

32. G. A. Bowmaker, D. S. Gill, B. W. Skelton, N. Somers, <strong>and</strong> A. H. White, Z. Naturforsch.,<br />

B59, 1307 (2004).<br />

33. E. K. Beloglazkina, A. V. Shimorsky, A. G. Mazhuga, O. V. Shilova, V. A. Tafeenko,<br />

<strong>and</strong> N. V. Zyk, Russ. J. Gen. Chem., 79, 1504 (2009).<br />

34. H.-G. Hao, X.-D. Zheng, <strong>and</strong> T.-B. Lu, Angew. Chem. Int. Ed., 49, 8148 (2010).<br />

35. P. G. Jones <strong>and</strong> O. Crespo, Acta Crystallogr. C, 54, 18 (1998).<br />

36. J. W. Bats, T. Kretz, <strong>and</strong> H.-W. Lerner, Acta Crystallogr. C, 65, m94 (2009).<br />

37. R. D. Shannon, Acta Crystallogr. A, 32, 751 (1976).<br />

38. P. Johansson, S. P. Gejji, J. Tegenfeldt, <strong>and</strong> J. Lindgren, Electrochim. Acta, 43, 1375<br />

(1998).<br />

39. M. Herstedt, W. A. Henderson, M. Smirnov, L. Ducasse, L. Servant, D. Talaga, <strong>and</strong><br />

J.-C. Lassègues, J. Mol. Struct., 783, 145 (2006).<br />

40. P. Neelakantan, P. Indian Acad. Sci. A, 60, 422 (1964).<br />

[http://www.springerlink.com/content/t428037518568053/<br />

41. G. Fini <strong>and</strong> P. Mirone, Spectrochim. Acta A, 32, 439 (1976).<br />

42. J. Barthel, R. Buchner, <strong>and</strong> E. Wismeth, J. Soln. Chem., 29, 937 (2004).<br />

43. J. M. Alia, H. G. M. Edwards, <strong>and</strong> J. Moore, Spectrochim. Acta A, 51, 2039 (1995).<br />

44. J.-S. Seo, B.-S. Cheong, <strong>and</strong> H.-G. Cho, Spectrochim. Acta A, 58, 1747 (2002).<br />

45. B. G. Oliver <strong>and</strong> G. J. Jantz, J. Phys. Chem., 74, 3819 (1970).<br />

46. M. G. Davidson, P. R. Raithby, A. L. Johnson, <strong>and</strong> P. D. Bolton, Eur. J. Inorg. Chem.,<br />

18, 3445 (2003).<br />

47. Q. Zhou, P. D. Boyle, L. Malpezzi, A. Mele, J.-H. Shin, S. Passerini, <strong>and</strong><br />

W. A. Henderson, Chem. Mater., 23, 4331 (2011).<br />

48. Q. Zhou, K. Fitzgerald, P. D. Boyle, <strong>and</strong> W. A. Henderson, Chem. Mater., 22, 1203<br />

(2010).<br />

49. W. A. Henderson, F. McKenna, M. A. Khan, N. R. Brooks, V. G. Young Jr., <strong>and</strong><br />

R. Frech, Chem. Mater., 17, 2284 (2005).<br />

50. I. Rey, P. Johansson, J. C. Lassègues, J. Grondin, <strong>and</strong> L. Servant, J. Phys. Chem. A,<br />

102, 3249 (1998).<br />

51. M. Herstedt, M. Smirnov, P. Johansson, M. Chami, <strong>and</strong> J. Grondin, J. Raman Spectrosc.,<br />

36, 762 (2005).<br />

52. L. Xue, C. W. Padgett, D. D. DesMarteau, <strong>and</strong> W. T. Pennington, Solid State Sciences,<br />

4, 1535 (2002).<br />

53. A. M. Heyns <strong>and</strong> D. de Waal, J. Chem. Phys., 97, 8086 (1992).<br />

54. X. Xuan, J. Wang, <strong>and</strong> H. Wang, Electrochim. Acta, 50, 4196 (2005).<br />

55. R. Aroca, M. Nazri, G. A. Nazri, A. J. Camargo, <strong>and</strong> M. Trsic, J. Soln. Struct., 29,<br />

1047 (2000).<br />

56. A. M. Heyns, Spectrochim. Acta A, 33, 315 (1977).<br />

57. C. M. Burba <strong>and</strong> R. Frech, J. Phys. Chem. B, 109, 15161 (2005).<br />

58. J. Grondin, L. Ducasse, J.-L. Bruneel, L. Servant, <strong>and</strong> J.-C. Lassègues, Solid State<br />

<strong>Ionic</strong>s, 166, 441 (2004).<br />

59. P. Johansson, Phys. Chem. Chem. Phys., 9, 1493 (2007).<br />

60. O. Borodin, G. D. Smith, <strong>and</strong> R. L. Jaffe, J. Comput. Chem., 22, 641 (2001).<br />

61. S. E. Popov, A. E. Nikiforov, O. V. Bushkova, <strong>and</strong> V. M. Zhukovskii, Russ. J. Electrochem.,<br />

41, 546 (2005).<br />

62. O. Borodin <strong>and</strong> G. D. Smith, J. Phys. Chem. B, 110, 6293 (2006).<br />

63. S. P. Gejji, C. H. Suresh, K. Babu, <strong>and</strong> S. R. Gadre, J. Phys. Chem. A, 103, 7474<br />

(1999).<br />

64. J. C. Lassegues, J. Grondin, C. Aupetit, <strong>and</strong> P. Johansson, J. Phys. Chem. A, 113, 305<br />

(2009).<br />

65. O. Borodin <strong>and</strong> G. D. Smith, J. Phys. Chem. B, 110, 4971 (2006).<br />

66. O. Borodin, G. D. Smith, <strong>and</strong> W. Henderson, J. Phys. Chem. B, 110, 16879 (2006).<br />

67. See Supporting Information at http://dx.doi.org/10.1149/2.035209jes.html.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!