Highly Concentrated Electrolytes Containing a Phosphoric Acid Ester

1. Highly Concentrated Electrolytes Containing a. Phosphoric Acid Ester Amide with Self-. Extinguishing Properties for Use in Lithium. Batteries. Tohr...
0 downloads 6 Views 1MB Size
Subscriber access provided by UNIV OF NEW ENGLAND ARMIDALE

C: Energy Conversion and Storage; Energy and Charge Transport

Highly Concentrated Electrolytes Containing a Phosphoric Acid Ester Amide with Self-Extinguishing Properties for Use in Lithium Batteries Tohru Shiga, Chika-aki Okuda, Yuichi Kato, and Hiroki Kondo J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.7b12461 • Publication Date (Web): 20 Apr 2018 Downloaded from http://pubs.acs.org on April 20, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Highly Concentrated Electrolytes Containing a Phosphoric Acid Ester Amide with SelfExtinguishing Properties for Use in Lithium Batteries Tohru Shiga*, Chika-aki Okuda, Yuichi Kato, and Hiroki Kondo Toyota Central Research & Development Laboratories Inc. Yokomichi, Nagakute-city, Aichi-ken, 480-1192 Japan

1 Environment ACS Paragon Plus

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ABSTRACT. Phosphoric acid ester amides were examined as a new self-extinguishing solvent for Li-ion batteries. The phosphoric acid ester amides used in this study contained two fluorinated alkyl groups and one amino group, (CF3CH2O)2(NR1R2)P=O (PNR1R2). The thermal stability of the highly concentrated electrolyte of lithium bis(fluorosulfonyl) amide (LiFSA) and PNR1R2 with a molar ratio of [LiFSA]/[PNR1R2] = 0.5 under overcharge depended on the modification of the amino substituent. Introduction of a phenyl group (R1=CH3, R2=C6H5) was effective for improving thermal stability. The release of gases and heat that typically accompanied reaction of the solvent with the charged graphite anode was greatly suppressed. Density functional theory calculations indicated that PNR1R2 decomposed reductively near 0.5 V vs. Li+/Li, suggesting poor Li ion insertion into the graphite. However, the highly concentrated electrolyte using LiFSA and PNR1R2 reduced the reductive potential of PNR1R2, and enabled not only the insertion of Li ions into the graphite but also reversible Li plating/stripping.

2 Environment ACS Paragon Plus

Page 2 of 23

Page 3 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

1. Introduction Lithium batteries using volatile and flammable electrolytes are excellent power sources for vehicles of the future but pose serious safety issues.1,2 For example, fire or explosion upon overcharge may occur in some cases.3-5 The use of incombustible compounds, such as phosphazenes with a -N=P- group, and trialkyl phosphates, (CnH2n+1O)3P=O, can eliminate this risk and have been investigated in this application.6-8 However, Li salts are not very soluble in phosphazenes.9-12 Phosphates can attack the graphite anode and break down its laminar structure due to co-intercalation of Li ions and solvent. Moreover, under overcharged conditions, phosphates generate large amounts of gases at a temperature lower than that of organic electrolytes.13 One strategy for the suppression of graphite destruction is the use of a high concentration

of

lithium

bis(fluorosulfonyl)amide

salt.14-16

[LiFSA;

Highly

concentrated

Li(FSO2)2N)]

and

electrolytes

trimethyl

using

phosphate

lithium

(TMP)

or

tris(trifluoroethyl) phosphate [(CF3CH2O)3P=O, TFEP] have been investigated.17 The TFEP is a fluorinated alkyl phosphate with self-extinguishing properties. A highly-concentrated electrolyte with a molar ratio of [LiFSA]/[solvent] = 0.5 enabled the insertion of Li ions into the graphite anode. In addition, an electrolyte with [LiFSA]/[TFEP] = 0.5 showed good thermal stability in the presence of charged graphite and the gas generation was reduced dramatically. The present study describes non-aqueous electrolytes of phosphoric acid ester amide as a new class of selfextinguishing solvent. The phosphoric acid ester amides in this study have a chemical formula of (CF3CH2O)2(NR1R2)P=O (R1, R2: alkyl or phenyl group, abbreviation: PNR1R2), in which one trifluoroethyl group in TFEP is exchanged by one amino substituent. Therefore, the phosphoric acid ester amide involves a P-N single bond similar to phosphazene compounds, but were expected to provide better thermal stability and incombustibility. Highly concentrated

3 Environment ACS Paragon Plus

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

electrolytes composed of LiFSA and PNR1R2 were prepared up to a molar ratio of [LiFSA]/[PNR1R2]=0.5, and the Li-ion solvation and electrochemical properties were examined. The results indicated that a highly concentrated electrolyte employing PNR1R2 is an excellent candidate as a safe electrolyte for lithium batteries. 2. Experimental Materials: The two types of phosphoric acid ester amides used were dimethylaminodi(trifluoroethyl) phosphate (PNMeMe), and methylphenylamino-di(trifluoroethyl) phosphate (PNMePh) (Tohso Finechem Corporation) (Fig.1). The physical properties of PNMeMe and PNMePh were listed in Table S1. The water contents in the liquids were 12 ppm (PNMeMe) and 37 ppm (PNMePh), respectively. Lithium bis(fluorosulfonyl) amide (LiFSA), and propylene carbonate (PC) were obtained from Kishida Chemicals. Vinylene carbonate (VC), an electrolyte additive, was available from Aldrich.

PNMeMe

PNMePh

Figure 1. Chemical structures of the phosphoric acid ester amides.

Electrolytes: Highly concentrated electrolytes were prepared by mixing LiFSA and the phosphoric acid ester amide with molar ratios of [Li salt]/[solvent] of 0.125 to 0.5. For reference, a 1 mol/L LiPF6/EC+DMC (1:1 v/v) electrolyte [battery grade, ethylene carbonate (EC), dimethyl carbonate (DMC), Kishida Chemicals] was used.

4 Environment ACS Paragon Plus

Page 4 of 23

Page 5 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Computational details: The Gaussian program 09E01 was used to calculate one-electron oxidation and one-electron reduction potentials of the two phosphoric acid ester amides. The geometries of the solvents were fully optimized at the B3LYP/6-311++G(d,p) level. DSC measurements: The charged anode material was prepared as a pellet by compressing graphite (Osaka Gas, average particle diameter D50: 6 µm). A coin cell composed of the anode pellet, Li metal, and an electrolyte of 1 mol/L LiPF6/EC+DMC was fabricated. The electrolyte was added to the cell in an argon-filled glove box, and the cell was connected to a charge/discharge measurement apparatus (Hokuto Denko) and charged up to 0.01 V to obtain a sample with a 100% SOC using constant-current mode. The temperature was maintained at 25 °C. The charged anode pellet was removed from the cell and washed several times with dimethyl carbonate. The charged anode pellet (2.5 mg) was wetted with 1.35 µL highly concentrated electrolyte at a molar ratio of [Li salt]/[solvent] = 0.5, and DSC measurements (Thermo Plus TG8120, Rigaku) were obtained under an argon flow of 10 mL/min and a heating rate of 5 °C/s. The thermal stability of 1 mol/L LiPF6/EC+DMC was also determined. The charged cathode material was prepared by compressing LiNiO2 powder (Sakai Chemical Industry, LiNi0.8Co0.15Al0.05O2, D50: 7 µm) to obtain a disk electrode. A coin cell composed of the cathode disk, Li metal, and an electrolyte of 1 mol/L LiPF6/EC+EMC then was fabricated. The cell was connected to a charge/discharge measurement apparatus (Hokuto Denko) and charged up to 4.2 V under constant-current mode and continuous constant-voltage mode at 4.2 V to obtain a sample with 100% SOC. The composition of the charged cathode was Li0.52Ni0.8Co0.15Al0.05O2 (145 mAh/g). After washing the charged cathode disk, 2.5 mg of the cathode disk was wetted with 1.35 µL of the electrolyte for DSC analysis.

5 Environment ACS Paragon Plus

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Ignition test: The ignition tests were conducted using glass fiber filter paper strips (Whatman, 8×50 mm2) saturated with 0.15 mL electrolyte at a molar ratio of [Li salt]/[solvent] = 0.5; a gas burner with a flame at ca. 700 °C was used to attempt ignition. Raman analysis: Some electrolytes were prepared by mixing LiFSA and PNMePh at [LiFSA]/[PNMePh] ratios of 0.125, 0.25, and 0.5. The electrolytes and pure PNMePh solvent were poured into glass capillaries. The LiFSA powder was placed in a quartz cell filled with argon (Fig. S1). Raman measurements were made through the capillaries or quartz cell using a Raman spectrophotometer (NR-3300, Jasco). The wavelength of the excitation laser was 532 nm and the beam diameter in the excitation region was 1 mm. FTIR analysis: The FTIR measurements were made under attenuated total reflection mode to study the composition of the SEI on graphite using a Nicolet FTIR spectrophotometer (Avatar360). Electrochemical measurements: The anode was prepared by mixing graphite powder (90% w/w, Osaka Gas) with polyvinylidene difluoride (#1120 Kureha) and N-methylpyrrolidone (NMP; Wako Chemicals) in a kneading machine (ARE-310, Thinky Co. Ltd.) at 2200 rpm for 5 min. The anode slurry was spread onto a copper current collector (20 µm thick) using a doctorblade technique and dried at 150 °C under vacuum for 5 h. The anode sheet was pounded to obtain a 14-mm diameter disk electrode with a thickness of 35 µm. The active material loading was 2.79 mg/cm2. Half coin cells (Fig. S2) were fabricated using the 14-mm diameter disk electrodes, one filter paper (25-µm thick, Kiriyama 5C), a Li metal disk 18 mm in diameter and 0.4 mm thick (Honjo Metal), and the highly concentrated electrolyte. The coin cell was connected with a charge-discharge unit (Hokuto Denko). The charge-discharge current was 0.01–0.1 mA/cm2 without using constant-voltage mode. The cutoff for the graphite disk was set

6 Environment ACS Paragon Plus

Page 6 of 23

Page 7 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

at 0.03–2.2 V. Charge-discharge tests were performed at temperatures up to 60 °C. Electrochemical impedance spectroscopy (EIS) was conducted with an LF impedance analyzer (Hewlett Packard, 4192A) with an amplitude of 100 mV over a frequency range from 5 Hz to 12 MHz to investigate the formation of passive layer on graphite.

3. Results and discussion 3.1 Electrochemical stability of phosphoric acid ester amide One-electron oxidation and reduction potentials of phosphoric acid ester amides were calculated using DFT calculations with the (B3LYP/6-311++G(d,p)) basis set.18-20 The optimized geometries of PNMeMe and PNMePh were determined to minimize their free energy, and then the potentials were calculated. The one-electron oxidation potential of PNMeMe was calculated to 5.130 V vs. Li+/Li, whereas PNMePh had a one-oxidation potential of 4.744 V vs. Li+/Li (Fig. 2). In contrast, one-electron reduction potentials of the two phosphoric acid ester amides were 0.436 V and 0.547 V, suggesting decomposition of the phosphoric acid ester amides at the surface of graphite anode, leading to passivation layer formation. Each potential window of PNMeMe or PNMePh, which was the difference between the oxidative and reductive potentials, was 4.694 V or 4.197 V. In addition, Mulliken electric charge distribution on each element in PNMeMe and PNMePh were provided by the DFT calculations (Fig.S3). The electric charges on some elements are summarized in Table 1. The results showed a characteristic difference for nitrogen, i.e., the nitrogen in PNMeMe was negatively charged (-0.263), whereas PNMePh had a positive-charge of +0.124. This indicated the effect of the electron-donating nature of the phenyl substituent, leading to reduction in the oxidative potential for PNMePh. We calculated an optimized geometry of LiFSA solvated by two PNMePh molecules, [LiFSA]/[PNMePh] = 0.5, to estimate Mulliken electric charge distribution. The solvation structure is shown in Fig.3. The

7 Environment ACS Paragon Plus

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

coordination of the solvent to Li ion occurs through the oxygen atom in OCH2CF3 group of PNMePh. The electric charges on phosphorus and nitrogen in one PNMePh of the solvated structure were +0.002 and +0.330, respectively (Fig.S4). The other PNMePh molecule had a negatively charged phosphorus (-0.434) and a positively charged nitrogen (+0.252). The results denoted the similar tendency as pure PNMePh solvent.

Figure 2. One-electron oxidation and reduction potentials of phosphoric acid ester amides

Figure 3. Optimized geometry of Li+ solvated structure for [LiFSA]/[PNMePh] =0.5. The symbols represent phosphorus (orange), nitrogen (blue), oxygen (red), fluorine (light blue), carbon (dark grey), sulfur (yellow) and hydrogen (light grey)

8 Environment ACS Paragon Plus

Page 8 of 23

Page 9 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Table 1. Mulliken electric charge distribution of phosphoric ester amides

Cyclic voltammetry for the phosphoric acid ester amide electrolytes ([LiFSA]/[PNMePh] = 0.125 and 0.5) was performed at 25 °C using a three-electrode cell with Ag+/Ag reference electrode [BAS, RE-7 electrolyte: acetonitrile solution containing 0.01M AgNO3 and 0.1M (C4H9)4NClO4]. The working electrode for potential greater than 3.0 V vs. Li+/Li was a platinum plate, whereas a glassy carbon electrode was used as the working electrode to avoid Li-Pt alloys. As shown in Figs. 4a and 4b, large cathodic currents due to lithium plating were observed below 0.5 V (vs. Li+/Li) in the PNMePh electrolytes. In dilute electrolyte, the current corresponding to Li stripping was not observed, and no stripping was caused by the reductive decomposition of PNMePh. The super-concentrated electrolyte of PNMePh had a peak at 1.95 V, reflecting lithium stripping. The efficiency of lithium plating/stripping increased with cycle number (Fig. S5). Therefore, the high concentration enabled elongation of the electrochemical window and improved electrochemical stability.21,22 The CV curves of the dilute and super-concentrated electrolytes in the high potential region are displayed in Fig. S6. The rise in the oxidative current, which began close to 4.2 V at 25 °C, in the two electrolytes was unchanged. The CV test was also done for the PNMeMe electrolytes, and the results were the same as those using the PNMePh electrolytes. As shown in Fig. S7, partially reversible Li plating/stripping was observed in the electrolyte with [LiFSA]/[PNMeMe] = 0.5. The oxidative current was observed near 5.1 V vs. Li+/Li.

9 Environment ACS Paragon Plus

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 4. CV profiles for the PNMePh electrolytes: (a) [LiFSA]/[PNMePh] = 0.125, and (b) [LiFSA]/[PNMePh] = 0.5. Cycle number was 1 (black), 2 (pink), 3 (orange), 4 (blue), and 5 (green).

3.2. Thermal stability of highly concentrated electrolyte Thermal stability of the electrolyte in the presence of charged active materials was determined using differential scanning calorimetry (DSC) under an argon flow. Graphite was examined with a 100% state of charge (SOC). Previous DSC measurements indicated an exothermic reaction of the charged graphite with a typical non-aqueous electrolyte that occurred between 120 and 140 °C.23,24 Figure 5a shows DSC profiles for charged graphite with an SOC of 100% immersed in various electrolytes. In the 1 mol/L LiPF6-EC+DMC electrolyte, three peaks were observed at 140, 265, and 298 °C. The first small signal was due to decomposition of the passivation layer on the charged graphite, and the two latter peaks reflected reactions between LiPF6 and the alkyl carbonate solvents.25,26 The heat flow of the main peak observed at 298 °C was 10.4 W/g (watts per g graphite). The amount of heat generated in the total reaction ∆Hf, was 2800 J/g. The superconcentrated electrolyte with PNMeMe ([LiFSA]/[PNMeMe] = 0.5) exhibited three exothermic peaks between 200 and 300 °C: 220 °C, 257 °C, and 288 °C. The heat flow of the main peak observed at 288 °C was 7.78 W/g (watts per g graphite). The amount of heat generated in the total reaction ∆Hf, was 3100 J/g. Thus, a reduction in the thermal stability of the anode is a

10 Environment ACS Paragon Plus

Page 10 of 23

Page 11 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

serious disadvantage when using PNMeMe against a standard electrolyte. In contrast, the PNMePh system had smaller exothermic peaks and a much lower ∆Hf of 2350 J/g. The first two signals for the PNMePh system were smaller than those for the two phosphoric acid ester amides, and the main peak was shifted to a higher temperature (5.39 W/g, at 339 °C). To investigate reactivity of the charged cathode material with the electrolyte, DSC profiles were recorded for LiNi0.8Co0.15Al0.05O2 powders with 100% SOC in the presence of several electrolytes (Fig. 5b). In this study, the LiPF6/EC+DMC electrolyte also had two exothermic peaks with calorific values of 6.12 W/g (watts per g of Li0.52Ni0.8Co0.15Al0.05O2) at 240 °C and of 1.74 W/g at 285 °C, with a ∆Hf of 2050 J/g. The former signal was due to the main reaction with the electrolyte, attributed to a structural change in the de-lithiated cathode accompanied by oxygen liberation and combustion of the electrolyte with the liberated oxygen. The subsequent peak observed at 290 °C was caused by reaction of the remaining electrolyte with the cathode.27 The DSC profiles for the highly concentrated electrolytes with PNMeMe or PNMePh are shown as green and red lines, respectively, in Fig. 5b. A small signal observed near 180 °C for both electrolytes was identified as decomposition of the passivation film on the charged cathode. Compared to the standard electrolyte, the two DSC peaks for the PNMeMe electrolyte shifted to a higher temperature by 22 °C. The PNMeMe system first showed a peak at 262 °C, with a heat flow of 6.47 W/g, and a second peak with a heat flow of 2.21 W/g at 329 °C (∆Hf = 2350 J/g). In contrast, the [LiFSA]/[PNMePh]= 0.5 electrolyte had two large signals between 250 and 360 °C with respective ∆Hf values of 3.11 and 1.49 J/g. The total amount of heat in the reactions was ∆Hf = 1600 J/g. The separation between the exothermic peaks of the PNMePh and standard systems was 45 °C (285 °C vs. 240 °C). Therefore, the highly concentrated electrolyte using

11 Environment ACS Paragon Plus

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

PNMePh had good thermal stability against the charged LiNi0.8Co0.15Al0.05O2 cathode material. Thus, introduction of a phenyl group into the amide improved thermal stability.

Figure 5. DSC profiles for various electrolytes with immersed (a) graphite, and (b) LiNi0.8Co0.15Al0.05O2 with 100% SOC; highly concentrated electrolytes of [LiFSA]/[PNMeMe] = 0.5 (green), [LiFSA]/ [PNMePh] = 0.5 (red), and 1 mol/L LiPF6/EC+DMC (black).

3.3. Ignition test Ignition behavior of the highly concentrated electrolytes was investigated with a [LiFSA]/[solvent] of 0.5. For comparison, highly concentrated electrolyte with PC as the solvent ([LiFSA]/[solvent] = 0.5), as well as the typical LiPF6-EC+DMC electrolyte, were also prepared. Ignition tests were performed using glass fiber filter papers saturated with the electrolytes. After the electrolyte was applied to the paper, a flame at 700 °C was introduced using a gas burner. Figure 6 shows photographs of the glass fiber filter papers during the ignition test. The typical LiPF6/EC+DMC electrolyte and the super-concentrated electrolyte with PC were flammable (Time until the ignition: within 0.2 seconds), whereas the highly concentrated electrolytes with the phosphoric acid ester amides (PNMeMe, PNMePh) did not burn in the test over 15 seconds, indicating that they were self-extinguishing.

12 Environment ACS Paragon Plus

Page 12 of 23

Page 13 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 6. Photographs of ignition tests using (a) highly concentrated electrolytes of [LiFSA]/ [PNMeMe] =0.5, (b) [LiFSA]/[PNMePh] = 0.5, (c) [LiFSA]/[PC] = 0.5, and (d) 1 mol/L LiPF6-EC+DMC. Glass fiber filter papers were saturated with electrolyte and ignition was attempted using a gas burner.

3.4 Raman studies of electrolytes employing phosphoric acid ester amide The electrochemical and thermal stabilities and ignition test led to a focus on PNMePh in subsequent experiments. Raman spectroscopic measurements for phosphoric acid ester amide electrolytes with various molar ratios of LiFSA to PNMePh were conducted first to elucidate the coordination of Li ions with the solvent molecules. Phosphoric acid ester amides have characteristic Raman signals for P=O and P-N stretching. The spectra span the regions of 680– 860 cm-1 and 1200–1350 cm-1 (Figs. 7a and 7b). Raman profiles of LiFSA powder and pure PNMePh solvent are also shown in Figs. 7a and 7b. The signal near 710 cm-1 was due to the stretching vibration of P-O-C, adjacent to P=O bond, in PNMePh, and the peak located at 840 cm-1 was assigned to the stretching vibration of P-N.28-30 As the concentration of LiFSA increased, the former signal shifted to higher Raman wavenumber by 6 cm-1, which suggests coordination of PNMePh molecules with Li ions. However, the signal for the P-N stretching vibration was impervious to high concentrations. The waveform separation for the 710 cm-1 peak represented free and solvated PNMePh molecules. The peak was fitted by two components with peaks of 710 cm-1 and 716 cm-1. The proportion of the former decreased with

13 Environment ACS Paragon Plus

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

[LiFSA]/[PNMePh] molar ratio, whereas the proportion of latter increased (Table I), which reflect free and solvated PNMePh molecules, respectively.31-33 Upon addition of LiFSA salt, a new Raman band appeared near 730 cm-1, assigned to the S-N vibration in LiFSA. Upon an increase in Li+ ion concentration, this band was intensified and red-shifted. This spectral behavior was reported by Umebayashi et al.34 The phosphoric acid ester amide electrolytes produced three peaks between 1200 and 1350 cm-1. The peak located near 1260 cm-1 was assigned to the P=O stretching vibration in PNMePh. As the concentration of LiFSA was increased, the peak shift at 1260 cm-1 became more significant, up to 6 cm-1 in the superconcentrated solution. According to the wave-form separation, two components with peaks of 1260 cm-1 and 1267 cm-1 were obtained, which corresponded to free and solvated PNMePh molecules, respectively (Table 2). The signals near 1220 cm-1 and 1290 cm-1 were due to the S=O stretching vibration in LiFSA35 and deformation mode of P-O-CH2CF3 in PNMePh. The former signal was intensified and red-shifted with increasing LiFSA concentration, which reflected bulk (free) FSA- and that bound to Li+ ions. The signal on the deformation mode was uninfluenced. The results shown in Figs. 7a and 7b support the solvation of Li ions by the PNMePh solvent, and the structures and conformations of solvates such as contact ion pairs and aggregates.36-38

14 Environment ACS Paragon Plus

Page 14 of 23

Page 15 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 7. Raman spectra of the LiFSA- PNMePh electrolytes. Ratios indicate the molar ratio of LiFSA to PNMePh; [LiFSA]/[PNMePh] = 0.125 (pink), [LiFSA]/[PNMePh] = 0.25 (green), and [LiFSA]/[PNMePh] = 0.5 (red). Blue and black lines represent Raman profiles of LiFSA powder and pure PNMePh solvent, respectively.

Table 2. Ratio of free and solvated solvents

3.5. Intercalation of Li ions into active materials Although Li ion insertion into graphite is unlikely to occur in 1mol/L LiPF6-PC and LiFSAtrimethyl phosphate solutions, a high concentration of Li salt in the electrolyte improves Li ioninsertion behavior.14-17 Non-aqueous electrolytes were prepared with various concentrations of LiFSA in PNMePh solvent, followed by application to the graphite/Li cell. Figure 8 shows chargedischarge curves for the graphite/Li half cells at 60 °C. The test started at charging (i.e., insertion of Li ions into graphite). The cell using the electrolyte with a molar ratio of [LiFSA]/[PNMePh] = 0.125 charged at high potentials; however, it could not discharge sufficiently (charge capacity = 315

15 Environment ACS Paragon Plus

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

mAh/g, discharge capacity = 8 mAh/g). The charge plateau near 1 V was due to the co-intercalation of PNMePh molecules with Li ions and subsequent destruction of the graphite structure.39,40 The electrolyte with [LiFSA]/[PNMePh] = 0.25 enabled Li-ion intercalation into graphite. The charge and discharge capacities were 352 mAh/g and 92 mAh/g, respectively. In contrast, the electrolyte with [LiFSA]/[PNMePh] = 0.5, i.e., highly concentrated electrolyte, had a profile similar to that of the conventional electrolyte system (1 mol/L LiPF6/EC+DMC). The three potential plateaus, characterized as Li+ insertion into graphite, appeared clearly at 0.175, 0.091, and 0.056 V for the [LiFSA]/[PNMePh] = 0.5 electrolyte,41 charging up to 360 mAh/g was possible. When the polarity of the current was changed, three discharge plateaus corresponding to the charge behavior were fuzzy. Total discharge capacity was 124 mAh/g. Thus, the highly concentrated LiFSA in PNMePh enabled repeated Li-ion insertion and extraction, and enhanced the reversible capacity. The mechanism of Li-ion insertion in the highly concentrated electrolyte currently is not well understood. Insertion may be affected by the passivation layer (solid electrolyte interphase, SEI) at the surface of the graphite anode. The SEI formation is a factor that made high irreversible capacity (236 mAh/g). Low ionic conductivity of [LiFSA]/[PNMePh ] = 0.5 electrolyte (0.07 mS/cm at 25 °C) is an additional factor of the poor discharge behavior. It is incomparable with that of the conventional electrolyte (10.46 mS/cm for 1mol/L LiPF6-EC+DMC). We described temperature dependence of ionic conductivity for the both electrolytes in the supporting information. The DFT calculations and CV measurements for the PNMePh electrolyte revealed reductive decomposition near 0.5 V vs. Li+/Li, indicating SEI formation. To enhance discharge capacity for the cell using [LiFSA]/[PNMePh ] = 0.5 electrolyte, vinylene carbonate (VC) was added to the electrolyte.42,43 Since decomposition of VC occurs at 1.0 V, film formation occurs before decomposition of PNMePh. The effect of VC addition affected discharge behavior of the

16 Environment ACS Paragon Plus

Page 16 of 23

Page 17 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

cell (Fig. 8). Discharge capacity increased two-fold (from 124 mAh/g to 245 mAh/g). The cycle performance up to the fiftieth cycle is shown in Fig. S11.

Figure 8. Charge/discharge curves of graphite/Li half cells with electrolytes composed of LiFSA and PNMePh at 60 °C. Ratios indicate the molar ratio of LiFSA to PNMePh. The charge/discharge current was 0.03 mA/cm2.The light green line represents the charge/discharge curve for cell using [LiFSA]/[PNMePh] = 0.5 electrolyte with 2%

dissolved vinylene carbonate.

Figure 9 shows the Cole-Cole plots for the graphite/Li cells before and after the 1st charge/discharge cycle at 60 °C. Two semicircles were observed in all samples. The semicircle between 5 kHz and 12 MHz reflects interface resistance at Li metal. The semicircle appearing at the low frequency region should be corresponding to interface between electrolyte and graphite. The two semicircles became clearly larger after the charge/discharge process. Therefore, the large polarization of graphite/Li cells at the 1st cycle was caused by resistance increase at both electrodes. When comparing the samples with and without VC after the charge/discharge test, the diameter of the semicircle at the low frequency region significantly increased in no additive system. Thus, the addition of VC into the electrolyte is effective to suppress the reductive decomposition of the PNMePh solvent at graphite.

17 Environment ACS Paragon Plus

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 9. Cole-Cole plots for graphite/Li cells using electrolytes of [LiFSA]/[PNMePh] =0.5 with no additive (red) and 2% VC (green). The open and solid symbols represent the cells before and after the first charge/discharge cycle at 60 °C, respectively.

FTIR spectra were obtained to analyze the SEI on graphite in graphite/Li cells using the [LiFSA]/[PNMePh] = 0.5 electrolytes (Fig.10). The FTIR profiles of LiFSA powder and PNMePh in the region of 600–1600 cm-1 are shown in the supporting information (Fig.S12). In the FTIR spectrum for LiFSA powder, signals at 759 cm-1, 854 cm-1, 1176 cm-1, and 1365 cm-1 were assigned to symmetric S-N-S vibrations, asymmetric S-N-S vibrations, symmetric SO2 vibrations, and SO2 asymmetric vibrations, respectively.44-46 For PNMePh, the P-O symmetric vibration was located at 763 cm-1. The symmetric and asymmetric vibrations of CF3 appeared at 960 cm-1 and 1162 cm-1, respectively. The peak at 1060 cm-1 reflected deformation of P-OCH2CF3. Two signals at 1265 cm-1and 1288 cm-1 were assigned to stretching vibrations of P=O and P-N, respectively. Two strong signals between 1400 and 1500 cm-1 were due to amino groups. Several FTIR signals were observed for SEI on graphite without VC additive and were assigned to PNMePh: 862 cm-1, 961 cm-1, 1087 cm-1, 1278 cm-1, 1417 cm-1, and 1483 cm-1. The peak at 1151 cm-1 assigned to the symmetric stretching of SO2 was not clear. No stretching modes of CH2CF3 groups were observed at 2846 cm-1, 2958 cm-1, or 2917 cm-1. These signals

18 Environment ACS Paragon Plus

Page 18 of 23

Page 19 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

were also detected in spectra of for SEI on graphite with VC additives. In addition, a peak shift near 848 cm-1 was observed, which suggests the inclusion of FSA ions (S-N-S unit). The absorbances between 1400 and1500 cm-1, due to a decreased number of amino units in cells with VC addition, indicate a small amount of PNMePh degradation.

Figure 10. FITR spectra of graphite anode after the first charge/discharge cycle of the graphite/Li cell using electrolytes of [LiFSA]/[PNMePh] =0.5 with no additive (red) and 2% VC (green).

Cell performance for LiNi0.8Co0.15Al0.05O2/Li cells using the electrolytes [LiFSA]/[PNMePh] = 0.125 and 0.5 are shown in the supporting information (Figs.S13-S14). A large capacity drop-off in dilute electrolyte ([LiFSA]/[PNMePh] = 0.125) was detected at 60 °C. When the cell was dismantled after the test, the transparent electrolyte was found to have changed to a purple liquid (Fig.S15). A similar phenomenon was observed in a storage test of LiNi0.8Co0.15Al0.05O2 cathode dipped into a dilute solution of the [LiFSA]/[PNMePh] = 0.125 electrolyte at 60 °C. Therefore, the coloring of the electrolyte may be caused by release of Co or Ni from LiNi0.8Co0.15Al0.05O2. The Cole-Cole plots for LiNi0.8Co0.15Al0.05O2/Li cells using the [LiFSA]/[PNMePh] = 0.5 electrolyte was shown in Fig.S16. A large semicircle and other small one were obtained in the pristine sample. The former semicircle at the high frequency region increased its diameter during

19 Environment ACS Paragon Plus

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the cycle. It indicates that the reductive decomposition of PNMePh at the surface of Li metal proceeded fairly. The latter semicircle at the low frequency region was unchanged before and after the 1st charge/discharge cycle at 60 °C, suggesting that a passive layer at cathode was not thicker.

4. Conclusions

Alkyl phosphates have been investigated because of their non-flammability characteristics. However, their poor Li-ion insertion capability and their instability under overcharge have limited their application to Li-ion battery technologies. A previous study demonstrated that a super-concentrated fluorinated alkyl phosphate electrolyte possessed excellent Li-ion insertion capability and good thermal stability under overcharge. The present study examined phosphoric acid ester amides having two fluorinated alkyl groups and one amino group as a new class of self-extinguishing solvent. Results showed that introduction of a phenyl substituent into the amino group (PNMePh) provided greater thermal stability than that of the fluorinated phosphate. Although the reductive potential of PNMePh was 0.61 V vs. Li+/Li, it was lowered by a superconcentrated solution of PNMePh. Therefore, Li-ion insertion was observed in the PNMePh electrolyte. Although the Li-ion intercalation was largely irreversible, it could be reversed by addition of vinylene carbonate. Partially reversible Li plating-stripping was also observed in the super-concentrated electrolyte. Thus, two major issues with alkyl phosphates were resolved by highly concentrated electrolyte of LiFSA and PNMePh.

■ Supporting Information This material is available free of charge via the Internet at http://pubs.acs.org.

20 Environment ACS Paragon Plus

Page 20 of 23

Page 21 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Test apparatus, DFT calculation data, Typical CV profiles, DSC curves, Raman and FTIR profiles, Cell performance of graphite/Li and LiNiO2/Li cells. ■ AUTHOR INFORMATION Corresponding Author E-mail: [email protected] Tel:+81-561-71-7607 ■ NOTES The authors declare no competing financial interest. ■ Author Contributions T.S. conceived and conducted the experiments, analyzed the data, and wrote the paper. C.O. analyzed the DSC data, H. K. provided the DFT calculations, and Y.K performed the Raman analysis. ■ ACKNOWLEDGMENTS The authors thank Shigeharu Kawauchi of Toyota CRDL for ionic conductivity measurement. ■ REFERENCES (1) Xu, K. Electrolytes and Interphases in Li-Ion Batteries and Beyond. Chem. Rev. 2014, 114, 1150311618. (2) Watanabe, M.; Thomas, M.L.; Zhang, S.; Ueno, K.; Yasuda, T.; Dokko, K. Application of Ionic Liquids to Energy Storage and Conversion Materials and Devices. Chem. Rev. 2017, 117, 7190-7239. (3) Hess, S.; Wohlfahrt-Mehrens, M.; Wachtler, M. Flammability of Li-Ion Battery Electrolytes: Flash Point and Self-Extinguishing Time Measurements. J. Electrochem. Soc. 2015, 162, A3084-A3097. (4) Nowak, S.; Winter M. Review-Chemical Analysis for a Better Understanding of Aging and Degradation Mechanisms of Non-Aqueous Electrolytes for Lithium Ion Batteries: Method Development, Application and Lessons Learned. J. Electrochem. Soc, 2015, 162, A2500-A2508. (5) Hess, A.; Barber, G.; Chen, C.; Mallouk, T.E.; Allcock, H.R. Organophosphates as Solvents for Electrolytes in Electrochemical Devices. ACS Appl. Mater. Interfaces 2013, 5, 13029-13034. (6) Wang, X.; Yasukawa, E.; Kasuya, S. Nonflammable Trimethyl Phosphate Solvent –Containing Electrolyte for Lithium-Ion Batteries. J. Electrochem. Soc. 2001, 148, A1058-A1065. (7) Xu, K.; Ding, M.S.; Zhang, S.; Allen, J.L.; Jow, T.R. Evaluation of Fluorinated Alkyl Phosphates as Flame Retardants in Electrolytes for Li-Ion Batteries. J. Electrochem. Soc. 2003, 150, A161-A169. (8) Wang, X.; Yamada, C.; Naito, H.; Segami, G.; Kibe, K. High-Concentration Trimethyl PhosphateBased Nonflammable Electrolytes with Improved Charge-Discharge Performance of a Graphite Anode for Lithium-Ion Cells. J. Electrochem. Soc. 2006, 153, A135-A139. (9) Peearse, A.J.; Schmitt, T.E.; Fuller, E.J.; EI-Gabaly, F.; Lin, C-F.; Gerasopoulos, K.; Kozen, A.C.;

21 Environment ACS Paragon Plus

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Talin, A.A.; Rubloff, G.; Gregorczyk, K.E. Nanoscale Solid State Batteries Enabled by Thermal Atomic Layer Deposition of a Lithium Polyphosphazene Solid State Electrolyte. Chem. Mater. 2017, 29, 37403753. (10) Sun, J.; Wang, X.; Wu, D. Novel Spirocyclic Phosphazene-Based Epoxy Resign for HalogenFree Fire Resistance: Synthesis, Curing, Behaviors, and Flammability Characteristics. ACS Appl. Mater. Interfaces 2012, 4, 4047-4061. (11) King, A.K.; Plant, D.F.; Golding, P. Using Density Functional Theory to Interpret the Infrared Spectra of Flexible Cyclic Phosphazenes. J. Phys. Chem. A 2012, 116, 2080-2088. (12) Ganapathiappan, S.; Chen, K.; Shriver, D.F. Synthesis, Characterization, and Electrical Response of Phosphazene Polyelectrolyte. J. Am. Chem. Soc. 1989, 111, 4091-4095. (13) Shigematsu, Y.; Ue, M.; Yamaki, J. Thermal Behavior of Charged Graphite and LiCoO2 in Electrolytes Containing Alkyl Phosphate for Lithium-Ion Batteries. J. Electrochem. Soc. 2009, 156, A176-A180. (14) Kameda, Y.; Umebayashi, Y.; Takeuchi, M.; Wahab, M.A.; Fukuda, S.; Ishiguro, S.; Sasaki, M.; Amo, Y.; Usuki, T. Solvation Structure of Li+ in Concentrated LiPF6-Propylene Carbonate Solutions. J. Phys. Chem. Lett. 2007, 111, 6104-6109. (15) Moon, H.; Tatara, R.; Mandai, T.; Ueno, K.; Yoshida, K.; Tachikawa, N.; Yasuda, T.; Dokko, K.; Watanabe, M. Mechanism of Li Ion Desolvation at the Interface of Graphite Electrode and Glyme-Li Salt Solvate Ionic Liquids. J. Phys. Chem. C 2014, 118, 20246-20256. (16) Yamad, Y.; Yamada, A. Review-Super Concentrated Electrolytes for Lithium Batteries. J. Elechrochem. Soc. 2015, 162, A2406-A2423. (17) Shiga, T.; Kato, Y.; Kondo, H.; Okuda, C. Self-Extinguishing Electrolytes Using Fluorinated Phosphates for Lithium Batteries. J. Mater. Chem. A 2017, 5, 5156-5162. (18) Frisch, M.J.; Pople, J.A.; Binkley, J. S. Self-Consistent Molecular Orbital Methods Supplementary Functions for Gaussian Basis Sets. J. Chem. Phys. 1984, 80, 3265-3269. (19) Becke, A.D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648-5652. (20) Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Review B 1988, 37, 785-789. (21) Chen, Z.; Ma, Q.; Liu, P.; Hu, Y.-S.; Zhou, Z.; Li, H.; Huang, X.; Chen, L. Novel Concentrated Li[(FSO2)(n-C4H9SO2)N]-Based Ether Electrolyte for Superior Stability of Metallic Lithium Anode, ACS Appl. Mater. Interfaces 2017, 9, 4282-4289. (22) Sodeyama, K.; Yamada, Y.; Aikawa, K.; Yamada, A.; Tateyama, Y. Sacrificial Anion Reduction Mechanism for Electrochemical Stability Improvement in Highly Concentrated Li-Salt Electrolyte. J. Phys. Chem. C 2014, 118, 14091-14097. (23) Du Pasquier, A,; Disma, F.; Bowmer, T.; Gozdz, A.S.; Amatucci, G.; Tarascon, J-M. Differential Scanning Calorimetry Study of the Reactivity of Carbon Anodes in Plastic Li-Ion Batteries. J. Electrochem. Soc. 1998, 145, 472-477. (24) Maleki, H.; Deng, G.; Anani, A.; Howard, J. Thermal Stability Studies of Li-Ion Cells and Components. J. Electrochem. Soc. 1999, 146, 3224-3229. (25) MacNeil, D.D.; Larcher, D.; Dahn, J.R. Comparison of the Reactivity of Various Carbon Electrode Materials with Electrolyte at Elevated Temperature. J. Electrochem. Soc. 1999, 146, 3596-3602. (26) Botte, G.G.; White, R.E.; Zhang, Z. Thermal Stability of LiPF6-EC:EMC Electrolyte for Lithium Ion Batteries. J. Power Sources 2001, 97-98, 570-575. (27) Bang, H.J.; Joachin, H.; Yang, H.; Amine, K.; Prakash, Contribution of the Structure Changes of LiNi0.8Co0.15Al0.05O2 Cathodes on the exothermic Reactions in Li-Ion Cells. J. J. Electrochem. Soc. 2006, 153, A731-A737. (28) Gejji, S.P.; Hermansson, K.; Lindgren, J. Ab Initio Vibrational Frequencies of the Triflate Ion. J. Phys. Chem. 1993, 97, 3712-3715. (29) Rey, I.; Johansson, P.; Lindgren, J.; Lassegues, J.C.; Grondin, J.; Servant, L. Spectroscopic and Theoretical Study of (CF3SO2)2N- and (CF3SO2)2NH. J. Phys. Chem. A 1998, 102, 3249-3258. (30) Umebayashi, Y.; Mitsugi, T.; Fukuda, S.; Fujimori, T.; Fujii, K.; Kanzaki, R.; Takeuchi, M.; Ishiguro, S. Lithium Ion Solvation in Room-Temperature Ionic Liquids Involving

22 Environment ACS Paragon Plus

Page 22 of 23

Page 23 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Bis(trifluoromethanesulfonyl)imide Anion Studied by Raman Spectroscopy and DFT Calculations. J. Phys. Chem. B 2007, 111, 13028-13032. (31) Ramaathan, N.; Sundararajan, K.; Prasad Kar, B.; Viswanathan, K.S. Conformations of Trimethyl Phosphates: A Matrix Isolation Infrared and ab Initio Study. J. Phys. Chem. A 2011, 115, 10059-10068. (32) Reva, I.; Simao, A.; Fausto, R. Conformational Properties of Trimethyl phosphate Monomer. Chem. Phys. Let. 2005, 406, 126-136. (33) Nakagawa, H.; Ochida, M.; Domi, Y.; Doi, T.; Tsubouchi, S.; Yamanaka, T.; Abe, T.; Ogumi, Z. Electrochemical Raman Study of Edge Plane Graphite Negative-Electrodes in Electrolytes Containing Trialkyl Phosphoric Ester. J. Power Sources 2012, 212, 148-153. (34) Fujii, K.; Hamano, H.; Doi, h.; Song, X.; Tsuzuki, S.; Hayamizu, K.; Seki, S.; Kameda, Y.; Dokko, K.; Watanabe, M, et al. Unusual Li+ Ion Solvation Structure in Bis(Fluorosulfonyl)amide Based Ionic Liquid. J. Phys. Chem. C 2013, 117, 19314-19324. (35) Gejji, S.P.; Hermansson, K.; Lindgren, J. Ab Initio Vibrational Frequencies of the Triflate Ion, J. Phys. Chem. 1993, 97, 3712-3715. (36) Yoshida, K.; Nakamura, M.; Kazue, Y.; Tachikawa, N.; Tsuzuki, S.; Seki, S.; Dokko, K.; Watanabe, M. Oxidative Stability Enhancement and Charge Transport Mechanism in Glyme-Lithium Salt Equimolar Complexes. J. Am. Chem. Soc. 2011, 133, 13121-13129. (37) Lu, D.; Tao, J.; Yan, P.; Henderson, W.A.; Li, Q.; Shao, Y.; Helm, Lu.; Borodin, O.; Graff, G.L.; Polzin, B, et al. Formation of Reversible Solid Electrolyte Interface on Graphite Surface from Concentrated Electrolytes. Nano Lett. 2017, 17, 1602-1609. (38) Takenaka, N.; Suzuki, Y.; Sakai, H.; Nagaoka, M. On Electrolyte-Dependent Formation of Solid Electrolyte Interphase Film in Lithium-Ion Batteries: Strong Sensitivity to Small Structural Difference of Electrolyte Molecules. J. Phys. Chem. C 2014, 118, 10874-10882. (39) Zhang, H.-L.; Sun, C.-H.; Li, F.; Liu, C.; Tan J.; Cheng, H.-M. New Insight into the Interaction between Propylene Carbonate-Based Electrolytes and Graphite Anode Material for Li-Ion Batteries. J. Phys. Chem. C 2007, 111, 4740-4748. (40) Chung, G.-C.; Kim, H.-J.; Yu, S-I.; Jun, S.-H.; Choi, J.-W. Origin of Graphite Exfoliation: AN Investigation of the Important Role of Solvent Co-intercalation. J. Electrochem. Soc. 2000. 147, 43914398. (41) Guyomard, D.; Tarascon, J.M. Rechargeable LiMn2O4/Carbon Cells with a New Electrolyte Composition: Potentiostatic Studies and Application to Practical Cells. J. Electrochem. Soc. 1993, 140, 3071-3081. (42) Ouatani, L.EI.; Dedryvere, R.; Siret, C.; Biensan, P.; Reynaud, S.; Iratcabal, P.; Gonbeau, D. The Effect of Vinylene Carbonate Additive on Surface Film Formation on Both Electrodes in Li-Ion Batteries. J. Electrochem. Soc. 2009, 156, A103-A113. (43) Jung, H.M.; Park, S-H.; Jeon, J.; Choi, Y.; Yoon, S.; Cho, J-J.; Oh, S.; Kang, S.; Han, Y-K.; Lee, H. Fluoropropane Sultone as an SEI-Forming Additive That Outperforms Vinylene Carbonate. J. Mater. Chem. A 2013, 1, 11975-11981. (44) Huang, W.; Frech, R.; Wheeler, R.A. Molecular Structure and Normal Vibrations of Trifluoromethane and Its Lithium Ion Pair and Aggregates. J. Phys. Chem. 1994, 98, 100-110. (45) Brouillette, D.; Irish, D.; Taylor, N.J.; Perron, G.; Odziemkowski, M.; Desnoyers, J. Stable Solvates in Solution of Lithium Bis(trifluoromethylsulfone)imide in Glymes and Other Aprotic Solvents: Phase Diagrams, Crystallography and Raman Spectroscopy. Phys. Chem. Chem. Phys. 2002, 4, 60636071. (46) Nie, M.; Abraham, D.P.; Seo, D.M.; Chen, Y.; Bose, A.; Lucht, B.L. Role of Solution Structure in Solid Electrolyte Interphase Formation on Graphite with LiPF6 in Propylene Carbonate. J. Phys. Chem. C 2013, 117, 25381-2589.

23 Environment ACS Paragon Plus