jp810279h

7

Click here to load reader

Transcript of jp810279h

Page 1: jp810279h

8/13/2019 jp810279h

http://slidepdf.com/reader/full/jp810279h 1/7

Theoretical Insight into Oxidative Decomposition of Propylene Carbonate in the Lithium

Ion Battery

Lidan Xing,† Chaoyang Wang,† Weishan Li,*,†,‡,§ Mengqing Xu,†,‡ Xuliang Meng,† andShaofei Zhao†

School of Chemistry and EnVironment, South China Normal UniVersity, Guangzhou 510006, China, College of  Materials Science and Engineering, South China UniVersity of Technology, Guangzhou 510641, China, and 

Key Lab of Electrochemical Technology on Energy Storage and Power Generation in Guangdong UniVersities,Guangzhou 510006, China

 ReceiVed: NoVember 22, 2008; ReVised Manuscript ReceiVed: February 6, 2009

The detailed oxidative decomposition mechanism of propylene carbonate (PC) in the lithium ion battery isinvestigated using density functional theory (DFT) at the level of B3LYP/6-311++G(d), both in the gasphase and in solvent. The calculated results indicate that PC is initially oxidized on the cathode to a radicalcation intermediate, PC•+, and then decomposes through three pathways, generating carbon dioxide CO 2 andradical cations. These radical cations prefer to be reduced on the anode or by gaining one electron from PC,forming propanal, acetone, or relevant radicals. The radicals terminate by forming final products, includingtrans-2-ethyl-4-methyl-1,3-dioxolane, cis-2-ethyl-4-methyl-1,3-dioxolane, and 2,5-dimethyl-1,4-dioxane. Among

all the products, acetone is most easily formed. The calculations in this paper give detailed explanations of the experimental findings that have been reported in the literature and clarify the role of intermediate propyleneoxide in PC decomposition. Propylene oxide is one of the important intermediates. As propylene oxide isformed, it isomerizes forming a more stabile product, acetone.

1. Introduction

The lithium ion battery has been an irreplaceable powersource for electronic equipment required by today’s high-techsociety and is believed to be the most competitive power sourcefor electric vehicles needed in future, thus attracting extensiveinterest.1-6 A typical lithium ion battery system usually com-prises a transition metal oxide cathode, a graphite anode, divided

by a separator, and an electrolyte solution, which enables iontransfer between the two electrodes.7,8 The performance of alithium ion battery depends to a great extent on the stability of electrolyte solution, because the high voltage of the battery maycause the decomposition of lithium salt or organic solvents.9-12

Especially, gaseous products are formed from the decompositionof electrolyte solution, which increase the inside pressure of the battery and lead to a safety risk.13 To improve the stabilityof an electrolyte solution, it is necessary to understand thepossible mechanisms on the decomposition of the componentsin the electrolyte solution.

Currently, the most suitable solvents for lithium ion batteryremain the mixtures of cyclic alkyl carbonates, such as propylene

carbonate (PC) and ethylene carbonate (EC), and linear alkylcarbonates, such as dimethyl carbonate (DMC) and diethylcarbonate (DEC).6 PC is a most valuable solvent component of electrolytes for lithium ion battery because its high permittivity(64.4) provides the electrolytes with high ionic conductivity andits low melting point (-49   °C) provides a battery with betterperformance under low temperature.14-16 Therefore, the stabilityof PC has attracted much attention.

The reduction of PC on the anode in the lithium ion batteryhas been well understood.16-19 PC is easy to reduce on the anodeand co-intercalated into the graphite with lithium ions in theelectrolyte, resulting in the exfoliation of the graphite ac-companied by the formation of gaseous products such as propeneand hydrogen.20-24 Fortunately, the reduction and the co-intercalation of PC can be suppressed by an SEI film formedby additives such as vinylene carbonate and butyl sultone.25

The oxidation of PC on the cathode in the lithium ion batteryis also well investigated experimentally.26-29 Carbon dioxide,ethane, acetone, propanal, and both   cis- and   trans-2-ethyl-4-methyl-1,3-dioxolane have been identified as the products fromPC oxidation. However, the mechanisms that have beenproposed by different researchers to explain the experimentalphenomena are inconsistent. For example, Krtil et al.30 identifiedacetone, propanal, and carbon dioxide as the products for PCoxidation and proposed a mechanism involving the intermediatescationic propyl and isopropyl carbonates, as shown in eq 1.Arakawa et al.31 identified carbon dioxide, propanal, and bothcis- and   trans-2-ethyl-4-methyl-1,3-dioxolane as the productsfor PC oxidation and proposed a mechanism involving the

intermediate propylene oxide. Ufheil et al.28

believed thatpropylene oxide was one of the final products of PC oxidation.

With the aim of understanding the oxidative decompositionmechanism, high-level density functional theory (DFT) calcula-tions were carried out to identify the initial, transition state, andfinal products for PC oxidation in this paper. The calculations

* Corresponding author. Tel: +86 20 39310256. Fax: +86 20 39310256.E-mail address: [email protected].

† South China Normal University.‡ South China University of Technology.§ Key Lab in Guangdong Universities.

 J. Phys. Chem. B  2009,   113, 5181–5187   5181

10.1021/jp810279h CCC: $40.75   © 2009 American Chemical SocietyPublished on Web 03/23/2009

Page 2: jp810279h

8/13/2019 jp810279h

http://slidepdf.com/reader/full/jp810279h 2/7

give detailed explanations on the experimental findings that havebeen reported in the literature and clarify the role of intermediatepropylene oxide in PC decomposition.

2. Computational Details

All the calculations have been performed using the Gaussian03 package.32 The equilibrium and transition state structures arefully optimized by the B3LYP method33 in conjunction withthe 6-311++G(d) basis set.34 To confirm each optimizedstationary point and make zero-point energy (ZPE) corrections,frequency analyses are done with the same basis set. For each

transition state, intrinsic reaction coordinate (IRC) calculationis also performed in both directions to connect these corre-

Figure 1.   The optimized geometries of the various reactants, intermediates, transition states, and initial oxidative decomposition products of PC

at the B3LYP/6-311++G(d) level. Bond lengths are in angstroms.

TABLE 1: Calculated Structural Parameters and RelativeEnergies (in kJ/mol) for PC and PC•+ a)

PC PC•+

C1-O5 1.44   1.44   1.49   1.50

C1-C2 1.54   1.54   1.53   1.54C2-O3 1.45   1.45   1.54   1.54

C2-C7 1.52   1.52   1.50   1.51C4-O3 1.36   1.36    1.28   1.28

C4-O5 1.36   1.36    1.28   1.29C4dO6 1.19   1.20   1.29   1.29

∆ E    0.0 1023.4   1003.3

a Bond lengths are in angstroms. The italic data is taken from ref 36, which was calculated at the B3LYP/6-31+G(d) level.

5182   J. Phys. Chem. B, Vol. 113, No. 15, 2009   Xing et al.

Page 3: jp810279h

8/13/2019 jp810279h

http://slidepdf.com/reader/full/jp810279h 3/7

sponding intermediates at the above level. Enthalpies and Gibbsfree energies are calculated at 298.2 K. Charge distribution isanalyzed by the natural bond orbital (NBO) theory.

To investigate the role of solvent effects, the bulk solventeffect is estimated by single point calculations using thepolarized continuum models (PCM).35 A dielectric constant of 64.4 for PC is used for all PCM calculations.

3. Results and Discussion

3.1. Initial Oxidation of PC.  The initial oxidation of PCinvolves a one-electron transfer from one PC molecule to thecathode of the lithium ion battery, resulting in radical cationPC•+, which has been detected by in situ FT-IR.36 The optimizedgeometries and the calculated structural parameters of PC andPC•+ are presented in Figure 1. The structural parameters of PC and PC•+ are also presented in Table 1 with a comparison

of the values obtained from the calculation at the B3LYP/6-31+G(d) level.36 The optimized geometry indicates that the

initial oxidation is accompanied by the cleavage of the C2-O3

bond. With the cleavage of the C2-O3 bond in PC, the C4

-O3

and C4-O5 bonds are shortened from 1.36 Å in PC to 1.28 Å

in PC•+, and the C4-O6 bond is elongated from 1.19 Å in PC

to 1.29 Å in PC•+, which is close to C4-O3 or C4

-O5 in PC•+.Table 2 presents the charge distribution on atoms in PC and

PC•+ obtained by natural population analysis (NPA). Chargedistributes mainly on C and O atoms in PC and PC•+. It can beseen from Table 2 that the charge on O atoms in PC •+ (O3

)

-0.44, O5) -0.42, O

6) -0.61) is more positive than that onO atoms in PC (O3

) -0.56, O5) -0.56, O6

) -0.63), whilethe charge on the C atoms in PC and PC •+ are almost the same.This indicates that the electron is taken from O atoms duringthe initial oxidation of PC.

3.2. Decomposition Mechanism of PC•+.   Based on thegeometry of PC•+, it can be known that there are four possiblepathways for its decomposition, involving four initial products(PC-1, PC-2, PC-3, and PC-4), three intermediates (M1, M2,and M3), and seven transition states (TS1, TS2, TS3, TS4, TS5,TS6, and TS7), as shown in Scheme 1.

The geometry of all the intermediates, transition states andinitial products involved in the decomposition of PC•+ is

presented in Figure 1. The calculated relative energies (∆ E ),enthalpies (∆ H ), free energies (∆G), and the imaginary frequen-cies (cm-1) for transition states (TS) are presented in Table 3.Figure 2 presents the potential energy profile for the decomposi-tion process of PC•+ in gas phase and solvent.

In order to confirm the geometry of transition states,frequency analyses and IRC calculations are carried out. Eachtransition state (TS1, TS2, TS3, TS4, TS5, TS6, and TS7)corresponds to one imaginary frequency (392, 571, 308, 166,587, 441, and 733i cm-1 in gas phase, respectively), and IRCresults indicate that all the transition states are connected withthe relevant reactants and products.

As shown in Table 3, in the gas phase, the  ∆ E  of the initial

products, PC-1, PC-2, PC-3, and PC-4 is 23.1, 5.1, 0.4, and302.0 kJ/mol, respectively, indicating that the order of stabilityis PC-3 ≈ PC-2  >  PC-1  >  PC-4. The data of ∆ E + ∆ZPE, ∆ H ,and   ∆G  show the same results. Accordingly, in solvent, the∆ E  of the initial products, PC-1, PC-2, PC-3, and PC-4 is 4.9,4.1, 3.6, and 310.1 kJ/mol, respectively, indicating that the orderof stability is PC-3   ≈  PC-2   ≈  PC-1   >  PC-4. Obviously, theinitial products PC-1, PC-2, and PC-3 are far more stable thanPC-4. Therefore, PC-4 is unstable and less probable to be theinitial product from the oxidation of PC•+. Carbon dioxide isincluded in PC-1, PC-2, and PC-3 and has been observed bymany researchers.3-5 The calculated standard electrode potentialfor paths 1, 2, and 3 in the gas phase is 4.51, 4.43, and 4.41 V(vs Li/Li+), respectively.

On the other hand, the possibility for a reaction to take placedepends on its ∆ E  or the activation energy of its transition state.In path 1, PC•+ converts into PC-1 via only one transition stateTS1, with the cleavage of C2

-O3 and C4-O5 bonds, the

corresponding activation energy is 44.7 and 19.0 kJ/mol in thegas phase and in solvent, respectively. In path 2, PC •+ convertsfirst into intermediate M1 via transition state TS2 with thecleavage of C2

-C7 and C4-O3 bonds; the corresponding

activation energy is 34.3 and 13.5 kJ/mol in the gas phase andin solvent, respectively. Then M1 dissociates to PC-2 via thetransition state TS3 with the direct cleavage of the C1

-O5 bondand the formation of the C1

-C7 bond, the correspondingactivation energy is 105.0 and 94.3 kJ/mol in the gas phase

and in solvent, respectively. In path 3, PC•+ converts first intointermediate M2 via transition state TS4 with the cleavage of 

TABLE 2: Atomic Charges Based on NPA for PC and PC•+

C1 C2 O3 C4 O5 O6 C7

PC   -0.05 0.09   -0.56 1.03   -0.56   -0.63   -0.60PC•+

-0.05 0.11   -0.44 1.02   -0.42   -0.05   -0.61

TABLE 3: Relative Energies, Enthalpies, and Free Energies(in kJ/mol) of the Stationary Points and ImaginaryFrequency (ω , cm-1) of the Transition States for theOxidative Decomposition of PC•+ a

structure   ∆ E    ∆ E +∆ ZPE    ∆ H    ∆G   ω

PC•+ 0.0 (0.0) 0.0 0.0 0.0TS1 44.7 (19.0) 29.0 29.3 27.4 392iPC-1 23.1 (4.9)   -1.7 6.4   -19.4TS2 34.3 (13.5) 20.9 20.9 20.4 571iM1 30.4 (10.0) 21.7 23.1 20.1TS3 105.0 (94.3) 83.7 88.9 71.9 308iPC-2 5.1 (4.1)   -11.8   -5.0   -26.6TS4 128.0 (105.6) 111.7 113.7 108.3 166iM2 120.0 (107.5) 106.0 109.7 99.5TS5 145.8 (142.5) 129.3 132.6 123.1 587iPC-3 0.4 (3.6)   -15.9   -9.4   -29.2TS6 191.1 (180.8) 176.4 177.8 174.9 441iM3 120.0 (107.4) 106.0 109.7 99.5TS7 342.8 (332.3) 321.1 326.9 308.3 733i

PC-4 302.0 (310.1) 278.0 286.4 258.3a The data in parentheses refer to those in solvent.

TABLE 4: Atomic Charges Based on NPA for PC-1, PC-3,PC-2, and M1

C1 C2 O3 C4 O5 O6 C7

PC-1   -0.26 0.39   -0.56 1.06   -0.25   -0.47   -0.72PC-2   -0.08 0.09 0.00 1.04   -0.53   -0.48   -0.62PC-3   -0.08 0.09   -0.53 1.04   -0.02   -0.47   -0.62M1   -0.06 0.11   -0.31 1.04   -0.46   -0.47   -0.44

TABLE 5: Relative Energies, Enthalpies, and Free Energies(in kJ/mol) of the Stationary Points and Imaginary

Frequency (ω 

, cm

-1

) of the Transition States for theConversion of the Radical Cations a

structure   ∆ E    ∆ E  +  ∆ZPE   ∆ H    ∆G   ω

P1•+ 23.9 (4.4) 16.0 17.5 12.1TS-1 30.7 (10.6) 22.3 21.9 22.3 659i

P3•+-125.6 (-137.9)   -126.3   -125.6   -128.3

TS-2 30.0 (19.0) 24.3 23.3 24.3 141i

P2•+ 0.0 (0.0) 0.0 0.0 0.0TS-3 100.2 (94.3) 97.0 96.4 97.1 435i

a The data in parentheses refer to those in solvent.

Decomposition of PC in Li Ion Battery   J. Phys. Chem. B, Vol. 113, No. 15, 2009   5183

Page 4: jp810279h

8/13/2019 jp810279h

http://slidepdf.com/reader/full/jp810279h 4/7

C2-O3 bond; the corresponding activation energy is 128.0 and

105.6 kJ/mol in the gas phase and in solvent, respectively. ThenM2 dissociates to PC-3 via the transition state TS5 with thedirect cleavage of the C4

-O5 bond; the corresponding activationenergy is 145.8 and 142.5 kJ/mol in the gas phase and in solvent,respectively. In path 4, PC•+ converts first into intermediate M3

via transition state TS6 with the cleavage of C2-

O3

and C1-

O5

bonds and the formation of C1-O3 bond; the corresponding

activation energy is 191.1 and 180.8 kJ/mol in gas phase andsolvent, respectively. Then M3 dissociates to PC-4 via thetransition state TS7 with the direct cleavage of the C1

-O3 bond;the corresponding activation energy is 342.8 and 332.3 kJ/molin the gas phase and in solvent, respectively. Obviously, path 1is the most favorable pathway, followed by path 2 and path 3(Figure 2). Path 4 is infeasible because of the high activationenergy for transition state TS7 and the instability of PC-4.

Since the relative energy of M1 (10 kJ/mol in solvent) isclose to that of PC-1, PC-2, or PC-3 and the activation energyfor its formation (13.5 kJ/mol in solvent) is quite approximateto that for the formation of PC-1 and lower than that the

formation of PC-2 or PC-3, M1 is one of the possible initialproducts from the oxidation of PC•+. Based on the activation

energy of the formation of the possible initial products, it canbe inferred that the content of the initial products should be inthe order PC-1  > M1  > PC-2  > PC-3.

3.3. Charge Distribution of the Initial Products.   Thecharge distributions on atoms in PC-1, M1, PC-2, and PC-3 insolvent, obtained by using natural population analysis (NPA),

are shown in Table 4. The sum of the charge on O3, C4, and O6of the formed CO2 in PC-1, PC-2, and PC-3 is  -0.03,  -0.03,and 0.04, respectively. It is almost zero, indicating that afterthe decomposition of PC•+, the formed CO2 maintains electricalneutrality and the positive charge mainly concentrates on theresidual structure in PC-1, PC-2, and PC-3. The residualstructure is radical cation. The radical cation in PC-1 is denotedas P1•+ and the radical cation in PC-2 is similar to that in PC-3, denoted as P2•+. For M1, the sum of the charge on C7, H11,H12, and H13 of the formed methyl CH3

• radical in M1 is 0.29.Thus the positive charge mainly concentrates on the residualstructure in M1, which is a cation and denoted as P4+. Onecation similar to P4+ has been detected during the galvanostatic

oxidation of PC.3 The CH3• terminates forming ethane, whichhas been observed experimentally.8,37

Figure 2.   Potential energy profile for the decomposition process of PC•+ in gas phase and in solvent calculated with B3LYP/6-311++(d) andPCM-B3LYP/6-311++(d).

SCHEME 1: Possible Pathways for the Decomposition of PC•+

5184   J. Phys. Chem. B, Vol. 113, No. 15, 2009   Xing et al.

Page 5: jp810279h

8/13/2019 jp810279h

http://slidepdf.com/reader/full/jp810279h 5/7

3.4. Termination Reactions of Radical Cations. Figure 3presents the optimized geometry of radical cations P1•+ and P2•+

and the possible conversion pathways of them are shown inFigure 4, together with the potential energy profile. The

geometry of all the intermediates, transition states, and initialproducts involved in the conversion of P1•+ and P2•+ is also

presented in Figure 3. The calculated relative energies, enthal-pies, free energies, and the imaginary frequencies (cm-1) fortransition states are presented in Table 5.

As shown in Figure 4, P1•+ converts into P3•+ via transitionstate TS-1 with a activation energy of 30.7 and 10.6 kJ/mol inthe gas phase and in solvent, respectively, and into P2 •+ via

transition state TS-2 with a activation energy of 44.7 and 19.0kJ/mol in the gas phase and in solvent, respectively. This

Figure 3.   The optimized geometries of the various reactants, intermediates, transition states, and final oxidative decomposition products of PC atthe B3LYP/6-311++G(d) level. Bond lengths are in angstroms. The underlined data in DD/DD+ are the parameters of DD+.

Figure 4.   Potential energy profile for conversion of P1•+, P2•+, andP3•+ in the gas phase and in solvent calculated with B3LYP/6-311++(d) and PCM-B3LYP/6-311++(d).

Figure 5.   Singlet and triple potential energy surface for the reductionreaction of P1•+, P2•+, and P3•+ in solvent calculated with PCM-B3LYP/ 6-311++(d).

Decomposition of PC in Li Ion Battery   J. Phys. Chem. B, Vol. 113, No. 15, 2009   5185

Page 6: jp810279h

8/13/2019 jp810279h

http://slidepdf.com/reader/full/jp810279h 6/7

suggests that P1•+ converts into P3•+ more easily than into P2•+.It seems that P2•+ cannot be converted into P3•+ via transitionstate TS-3 due to its high activation energy, 100.2 and 94.3kJ/mol in the gas phase and in solvent, respectively. Thus, there

are three radical cations available during the PC oxidation, P1•+

,P2•+, and P3•+.

The radical cation may polymerize by itself or with anotherradical cation forming a divalent cation or gain one electronfrom the anode or a solvent molecule forming a radical. Theoptimization in this work fails to locate the divalent cationsformed from P1•+, P2•+, and P3•+. Therefore these radicalcations cannot polymerize but prefer to gain one electron fromthe anode forming a radical or from PC forming a radical anda new radical cation, PC•+. The final products are the same forthe radical cations to gain electron on the anode and from PC.

There are two possible electronic configurations for thereduction products of P1•+, P2•+, and P3•+, singlet and triplet.

The prefix “t” is used to denote the structure in the tripletelectronic state. The optimized structures of various species onthe singlet and the triplet state potential energy profiles arepresented in Figure 3, and the relative energies of various speciesand energy diagram along singlet and triplet reaction pathwaysin solvent are shown in Figure 5.

As shown in Figure 5, for the single potential energy surface,P2•+ is reduced to propylene oxide, and the formation of propylene oxide is exothermic with 748.2 kJ/mol. This suggeststhat propylene oxide is highly active and easy to isomerizeforming a more stable product, acetone, via transition state TS-4, the corresponding activation energy is 239.6 kJ/mol. Thisgives an explanation why some researchers believed thatpropylene oxide was one of the products of PC oxidation but

could not provide any evidence for the existence of propyleneoxide during the oxidation of PC.3,4 Thus, the product from P2•+

reduction is acetone, which has been observed as the oxidativedecomposition product of PC,   8 and the product from P1•+ orP3•+ reduction is propanal, which has also be observed.5,8

For the triplet potential energy surface, P1•+ and P2•+ are

reduced to t-P1• and P3•+ is reduced to t-P2•. Subsequently,radical t-P1• or t-P2• terminates with each other or with theunreduced radical cations, forming the products shown inTable 6 with the optimized structures presented in Figure 3.The products of termination between radicals t-P1• and t-P2•

are trans-2-ethyl-4-methyl-1,3-dioxolane (TEMD), cis-2-ethyl-4-methyl- 1,3-dioxolane (CEMD), and 2,5-dimethyl-1,4-dioxane(DD). Among them, TEMD and CEMD have been observedby Arakawa and Yanmaki.5 DD has not yet been reported asthe product of PC oxidation. This may be due to the loweramount of DD in the products compared with the amount of TEMD and CEMD. The products of the termination betweenradical t-P1• or t-P2• and radical cation P1•+ and P2•+ are cations

TEMD+

, CEMD+

, and DD+

. These cations are reduced on theanode forming TEMD, CEMD, and DD or gain one electronfrom PC forming TEMD, CEMD, DD, and PC•+.

To further confirm the products of PC oxidation, the relativeenergies, enthalpies, and free energies of PC and its oxidativedecomposition products are summarized in Table 7. From Table7, it can be noticed that the pathway for PC oxidation to formCO2   and propylene oxide has the lowest thermodynamicpossibility. Among the possible products of PC, the mostthermodynamically favorable product is acetone, followed bypropanal, CEMD, TEMD, and DD.

4. Conclusions

The detailed mechanism for the oxidative decomposition

of PC in the lithium ion battery has been investigated in thepresent work by DFT calculation at the B3LYP/6-311++G(d)level. Based on the calculated results, PC•+ is generated afterPC transfers one electron to the cathode. Subsequently, PC•+

converts into the initial products PC-1, M1, PC-2, and PC-3. Radical cations P1•+ and P2•+ are generated after CO2

released from the initial products, and P1•+ is easily convertedinto a more stable radical cation, P3•+. The radical cationsP1•+, P2•+, and P3•+ are reduced on the anode or by gainingone electron from PC, forming propanal, acetone, or relevantradicals. The radicals terminate by forming final products,including  trans-2-ethyl-4-methyl-1,3-dioxolane,  ci s-2-ethyl-4-methyl-1,3-dioxolane, and 2,5-dimethyl-1,4-dioxane.

Acknowledgment. This work is supported by the NationalNatural Science Foundation of China (Grant No. NS-

TABLE 6: Relative Energies, Enthalpies, and Free Energies (in kJ/mol) of the Stationary Points and Imaginary Frequency (ω ,cm-1) of the Transition States for the Termination of the Radical Cations  a

structure   ∆ E    ∆ E  +  ∆ZPE   ∆ H    ∆G   ω

P1•+ 149.5 (142.4) 142.3 143.1 140.4t-P1•

-621.9 (-378.1)   -626.8   -626.1   -629.6P3•+ 0.0 (0.0) 0.0 0.0 0.0t-P2•

-646.6 (-401.6)   -647.3   -647.4   -648.0P2•+ 125.6 (137.9) 126.3 125.6 128.3TS-4   -577.3 (-370.7)   -583.8   -585.6   -579.5 888i

propylene oxide  -

850.9 (-

610.3)  -

840.8  -

843.1  -

835.8propanal   -951.4 (-713.2)   -943.9   -944.5   -942.2(t-P1•

+ t-P2•) CEMD)/2   -959.1 (-708.4)   -940.8   -944.4   -910.9

[(t-P1•+ P3•+) or  (t-P2•

+ P1•+)  ) CEMD+]/2   -529.7 (-374.8)   -514.8   -517.9   -486.1t-P1•

+ t-P2•) TEMD   -958.5 (-707.9)   -940.2   -943.9   -910.5

[(t-P1•+ P3•+) or  (t-P2•

+ P1•+)  ) TEMD+]/2   -530.3 (-375.3)   -515.5   -518.4   -486.9(t-P1•

+ t-P1•) DD)/2   -953.7 (-703.3)   -934.4   -938.8   -902.4

(t-P1•+ P1•+

) DD+)/2   -544.1 (-389.1)   -526.4   -530.5   -495.2

a The data in parentheses refer to those in solvent.

TABLE 7: Relative Energies, Enthalpies, and Free Energies(in kJ/mol) of the Stationary Points of PC and the OxidativeDecomposition Products a

structure   ∆ E    ∆ E  +  ∆ZPE   ∆ H    ∆G

PC 0.0 (0.0) 0.0 0.0 0.0CO2  +  propyleneoxide

73.3 (62.3) 38.3 42.8   -5.3

CO2  +  acetone   -57.8 (-71.5)   -97.4   -90.4   -145.5CO2  +  propanal   -27.2 (-40.6)   -64.9   -58.6   -111.8CO2  +  TEMD/2   -34.3 (-35.3)   -61.2   -58.0   -80.1CO2  +  CEMD/2   -34.9 (-35.8)   -61.7   -58.5   -80.5CO2  +  DD/2   -29.5 (-30.8)   -55.4   -52.9   -71.9

a The data in parentheses refer to those in solvent.

5186   J. Phys. Chem. B, Vol. 113, No. 15, 2009   Xing et al.

Page 7: jp810279h

8/13/2019 jp810279h

http://slidepdf.com/reader/full/jp810279h 7/7

FC20873046) and Specialized Research Fund for the DoctoralProgram of Higher Education (Grant No. 200805740004).

References and Notes

(1) Tarascon, J. M.; Armand, M  Nature  2001,  414, 359.(2) Aurbach, D.; Zinigrad, E.; Cohen, Y.; Teller, H.  Solid State Ionics

2002,  148, 405.(3) Tasaki, K.  J. Phys. Chem. B  2005,  109, 2920.(4) Xu, M. Q.; Xing, L. D.; Li, W. S.; Zuo, X. X.; Shu, D.; Li, G. L.

 J. Power Sources   2008,  184, 427.

(5) Gireaud, L.; Grugeon, S.; Pilard, S.; Guenot, P.; Tarascon, J. M.;Laruelle, S.   Anal. Chem.   2006,  78, 3688.(6) Zhang, H. L.; Sun, C. H.; Li, F.; Liu, C.; Tan, J.; Cheng, H. M. J.

Phys. Chem. C   2007,  111, 4740.(7) Lu, D. S.; Li, W. S.; Zuo, X. X.; Yuan, Z. Z.; Huang, Q. M.   J.

Phys. Chem. C   2007,  111, 12067.(8) Wang, Y. X.; Nakamura, S.; Tasaki, K.; Balbuena, P. B.  J. Am.

Chem. Soc.  2002,  124, 4408.(9) Ue, M.; Murakami, A.; Nakamura, S.   J. Electrochem. Soc.  2002,

149, A1572.(10) Arakawa, M.; Yamaki, J. I.  J. Power Sources  1995,  54, 250.(11) Ue, M.; Takeda, M.; Takehara, M.; Mori, S.   J. Electrochem. Soc.

1997,  144, 2684.(12) Koch, V. R.; Dominey, L. A.; Nanjundiah, C.; Ondrechen, M. J.  J.

 Electrochem. Soc.  1996,  143, 798.(13) Joho, F.; Novak, P.   Electrochim. Acta  2000,  45, 3589.(14) Kameda, Y.; Umebayashi, Y.; Takeuchi, M.; Wahab, M. A.;

Fukuda, S.; Ishiguro, S. I.; Sasaki, M.; Amo, Y.; Usuki, T.  J. Phys. Chem. B  2007,  111, 6104.(15) Xu, M. Q.; Li, W. S.; Zuo, X. X.; Liu, J. S.; Xu, X.   J. Power 

Sources  2007,  174, 705.(16) Wang, Y. X.; Balbuena, P. B.  J. Phys. Chem. B  2002,  106 , 4486.(17) Buqa, H.; Wursig, A.; Goers, D.; Hardwick, L. J.; Holzapfel, M.;

Novak, P.; Krumeich, F.; Spahr, M. E.  J. Power Sources  2005,  146 , 136.(18) Herstedt, M.; Andersson, A. M.; Rehsmo, H.; Siegbahn, H.;

Edstrom, K.   Electrochim. Acta  2004,  49, 4939.(19) Aurbach, D.; Levi, M. D.; Levi, E.; Schechter, A.  J. Phys. Chem.

 B  1997,  101, 2195.(20) Herstedt, M.; Andersson, A. M.; Rehsmo, H.; Siegbahn, H.;

Edstrom, K.   Electrochim. Acta  2004,  49, 4939.(21) Winter, M.; Besenhard, J. O.; Spahr, M. E.; Novak, P.  Ad V. Mater.

1998,  10, 725.

(22) Xing, L. D.; Wang, C. Y.; Xu, M. Q.; Li, W. S.; Cai, Z. P.   J.Power Sources  2008, doi: 10.1016/j.jpowersour.2008.08.076.

(23) Wang, Y.; Balbuena, P. B.  J Phys Chem A  2002,  106 , 9585.(24) Hahn, M.; Wursig, Gallay, A. R.; Nov; ak, P.; Kotz, R.  Electrochem.

Commun.  2005,  7 , 925.(25) Zhang, X. R.; Kostecki, R.; Richardson, T. J.; Pugh, J. K.; Ross,

P. N.  J. Electrochem. Soc.  2001,  148, A1341.(26) Matsuta, S.; Kato, Y.; Ota, T.; Kurokawa, H.; Yoshimura, S.;

Fujitani, S.  J. Electrochem. Soc.  2001,  148, A7.(27) Cattaneo, E.; Rash, B.; Vielstich, W.  J. Appl. Electrochem.   1991,

21, 885.(28) Ufheil, J.; Wursig, A.; Schneider, O. D.; Novak, P.  Electrochem.

Commun.  2005,  7 , 1380.(29) Xu, K.; Ding, S.; Jow, T. R. J. Electrochem. Soc. 2001, 148, A267.(30) Krtil, P. Ph.D. Thesis, The J. Heyrovsky Institute of Physical

Chemistry and Electrochemistry, Prague, Czech Republic, 1993.(31) Arakawa, M.; Yamaki, J.  J. Power Sources   1995,  54, 250.(32) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb,

M. A.; Cheeseman, J. R.; Montgomery, J. A.; Vreven, T., Jr.; Kudin, K. N.;Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.;Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.;Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.;Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li,X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.;Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.;Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.;Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich,S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.;Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.;

Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz,P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.;Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson,B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A.,  Gaussian 03,revision B.05; Gaussian Inc., Pittsburgh, PA, 2003.

(33) Abbotto, A.; Streitwieser, A.; Schleyer, P. R.   J. Am. Chem. Soc.1997,  119, 11255.

(34) Wang, Y.; Balbuena, P. B.  J. Phys. Chem. A  2001,  105, 9972.(35) Zhang, X. R.; Pugh, J. K.; Ross, P. N.   J. Electrochem. Soc.  2001,

148, E183.(36) Xu, K.  Chem. ReV.  2004,  104, 4303.(37) Novak, P.; Goers, D.; Hardwick, L.; Holzapfel, M.; Scheifele, W.;

Ufheil, J.; Wursig, A.  J. Power Sources  2005,  146 , 15.

JP810279H

Decomposition of PC in Li Ion Battery   J. Phys. Chem. B, Vol. 113, No. 15, 2009   5187