You are on page 1of 4

Journal of The Electrochemical Society, 152 共11兲 A2151-A2154 共2005兲 A2151

0013-4651/2005/152共11兲/A2151/4/$7.00 © The Electrochemical Society, Inc.

Lithium-Ion Transfer at the Interface Between Lithium-Ion


Conductive Ceramic Electrolyte and Liquid Electrolyte-A
Key to Enhancing the Rate Capability of Lithium-Ion Batteries
Takeshi Abe,*,z Fumihiro Sagane, Masahiro Ohtsuka, Yasutoshi Iriyama,* and
Zempachi Ogumi**
Department of Energy and Hydrocarbon Chemistry, Graduate School of Engineering, Kyoto University,
Nishikyo-ku, Kyoto 615-8510, Japan

In this study, lithium-ion transfer through the electrode/electrolyte interface was examined using a model interface composed of a
lithium-ion-conductive ceramic and liquid electrolytes to focus on lithium-ion transfer. Lithium-ion transfer resistances at the
interface and their activation energies were evaluated by impedance spectroscopy. The activation energies were quite large and
consistent with the interaction between lithium-ion and solvents in an electrolyte as determined by a theoretical calculation.
© 2005 The Electrochemical Society. 关DOI: 10.1149/1.2042907兴 All rights reserved.

Manuscript submitted February 11, 2005; revised manuscript received June 21, 2005. Available electronically September 21, 2005.

Rechargeable lithium-ion batteries have been extensively studied 共DMSO兲, and fluoroethylene carbonate 共FEC, Kanto Denka Inc.兲
to improve their performance even after they were commercialized containing 1 mol dm−3 LiCF3SO3 were used as liquid electrolytes.
in 1991. Since lithium-ion batteries have high energy densities, they Four-probe ac impedance measurements were performed using
are useful in portable electronic devices, such as cellular phones, the cell shown in Fig. 2. Lithium metal was used for the four elec-
notebook computers, etc. Recently, lithium-ion batteries have been trodes. The contact area of solid and liquid electrolytes is kept
considered as power sources for hybrid electric vehicles 共HEV兲 and constant to be 0.20 cm2 by O-ring. Resistances at the Li/liquid elec-
electric vehicles 共EV兲, although there are still some safety issues to trolyte interface can be eliminated by the four-probe system. Detect-
be taken into account. For the use of lithium-ion batteries in HEV able resistances are the bulk and grain boundary resistances of solid
and EV applications, a high rate performance, i.e., fast charge and electrolytes, bulk resistance of liquid electrolytes, and interfacial
discharge reactions, is needed.1 lithium-ion transfer resistance between electrolyte/electrolyte. Mea-
The reaction in lithium-ion batteries is very uncomplicated and surements were conducted over a frequency region of 100 kHz–
can be described as the lithium-ion transfer between the positive and 10 mHz, with applied voltage amplitude of 30 mV at various
negative electrodes through an ionically conducting phase. There- temperatures.
fore, ion transfer at the interface between the positive 共and negative兲 To elucidate a quantitative correlation between the obtained ac-
electrode and electrolyte is an essential reaction during the charging tivation energies and the interaction of lithium ion with solvents, a
and discharging of lithium-ion batteries. Although the diffusivity of theoretical calculation with the density functional theory using
lithium ion in the active materials for lithium-ion batteries has been GAUSSIAN 98 共Ref. 9兲 was used. Reaction enthalpies of solvation
studied extensively,2,3 little information is available regarding ion of the first molecule of solvent with lithium ion were calculated at
transfer at the interfaces in lithium-ion batteries. 298.15 K. Molecular structures were fully optimized with B3LYP/
In our previous study, lithium-ion transfer and solvated lithium- 6-31G共d兲 in advance. Vibrational frequency calculation was per-
ion transfer at graphite electrodes were compared.4 The results are formed with the corresponding basis sets to confirm that the geom-
shown schematically in Fig. 1. The activation energies for solvated etries are at the minimum of the potential energy surface and to
lithium-ion transfer are around 25 kJ mol−1, and those for lithium- make zero-point energy corrections. Finally, single-point energies
ion transfer are twice as large 共53–59 kJ mol−1兲. Consequently, we were calculated at the B3LYP/6-311 + g共3dp,3df兲 level by using
found that solvated lithium-ion transfer at a graphite electrode pro- the optimized geometries obtained. In this calculation, reaction en-
ceeds very fast, and desolvation plays an important role in lithium- thalpies 共Li+ + solvent = Li+ − solvent兲 at 298.15 K were obtained
ion kinetics at an electrode/electrolyte interface. by ⌬H = H共Li+ − solvent兲 − 兵H共solvent兲 + H共Li+兲其.
In the above studies, the lithium-ion transfer through the
electrode/electrolyte interface was studied. Therefore, a redox reac-
tion also proceeds in addition to lithium-ion transfer at the graphite Results and Discussion
electrode. To shed light on the interfacial lithium-ion transfer, stud-
ies on ion transfer at the interface between a lithium-ion-conductive A typical Nyquist plot is given in Fig. 3a. Two semicircles were
solid electrolyte and liquid electrolyte should be interesting, since no observed in the higher and lower frequency regions in Fig. 3a. The
redox reaction takes place in this system. higher-frequency semicircle is identified as a grain-boundary resis-
tance of LLT based on a comparison with the results of ac imped-
ance measurement in Au/LLT/Au.10 Since bulk resistances of LLT
Experimental and PC-based electrolyte give no semicircle in the frequency region
of 100 kHz to 10 mHz, and interfacial resistance between Li metal
Crystalline solid electrolyte of La0.55Li0.35TiO3 共LLT兲 共Ref. 5 and and PC-based electrolyte is eliminated by the four-probe measure-
6兲 10 mm␾, thickness: 1 mm兲 and Li–Al–Ti-phosphate-based glass ment, the latter semicircle should be ascribed to lithium-ion transfer
electrolyte donated by Ohara Inc.7,8 共OHARA glass兲 共20 ⫻ 20 resistance at the LLT/PC-based electrolyte interface. The fact that
⫻ 0.33 mm兲 were used as ceramic electrolytes. Both electrolytes the lower-frequency semicircle is dependent on the salt concentra-
showed high lithium-ion conductivities with a high transference tion of LiCF3SO3 in PC-based electrolyte strongly supports the
number for Li+ 共tLi+ was determined to be almost unity by use of above assignment. Further, when the electrolyte salt was changed
concentration cell兲. Propylene carbonate 共PC兲, dimethylsulfoxide from LiCF3SO3 to LiClO4, the same concentration-dependent be-
havior was observed as given by the Nyquist plots in Fig. 3b. Unless
otherwise mentioned, lithium-ion transfer resistances were clarified
* Electrochemical Society Active Member. by the dependency of Nyquist plots on the salt concentration in
** Electrochemical Society Fellow. electrolytes. Note that lithium-ion transfer occurred in the lower
z
E-mail: abe@elech.kuic.kyoto-u.ac.jp frequency region, indicating that ion transfer across the interface has

Downloaded on 2014-10-18 to IP 141.214.17.222 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
A2152 Journal of The Electrochemical Society, 152 共11兲 A2151-A2154 共2005兲

Figure 1. Comparison of solvated lithium-ion transfer and pure lithium-


ion transfer at graphite electrodes. The activation energies for solvated
lithium-ion transfer are much lower than those for pure lithium-ion transfer
共Ref. 4兲.

a large time constant. Based on the assignment mentioned above, the


lithium-ion transfer resistances were evaluated by the equivalent cir-
cuit as given in Fig. 3c.
Figure 4 shows the temperature dependence of the lithium-ion
transfer resistances at LLT/PC-based electrolyte and OHARA glass
electrolyte/PC-based electrolyte interfaces. The activation energies
共57.3 and 56.2 kJ mol−1兲 obtained from the slope in Fig. 4 were very
similar, regardless of the ceramic electrolytes. The activation ener-
gies for lithium-ion conduction in LLT and glass electrolytes were
within the range of 28–36 kJ mol−1, which are in the range of values
reported.5-8 In addition, the “activation energies” of lithium ion con-
duction in PC-based electrolyte are much smaller. Hence, ion trans-
fer across the interface is much more difficult than lithium ion con-
duction in the present electrolytes. These results suggest that a high
energy barrier should exist at the interface. Because the activation

Figure 3. 共a兲 Typical Nyquist plot for a LLT/PC-based electrolyte interface


using the cell shown in Fig. 2. Numbers on solid symbols refer to the loga-
rithms of frequencies. The bulk resistances of the LLT and PC-based elec-
trolyte are very small and give no semicircles in these frequency regions. The
contact area of LLT and PC-based electrolyte is kept constant at 0.20 cm2.
共b兲 Dependence of salt concentration in a Nyquist plots for the LLT/PC-
based electrolyte interface using the cell described in Fig. 2. The electrolytes
used are 0.25, 0.50, and 1.0 mol dm−3 LiClO4 dissolved in PC. Numbers in
Figure 2. Four-probe cell for ac impedance spectroscopy. C. E. and R. E. solid squares refer to the logarithms of the frequencies. The measurements
denote counter- and reference electrodes, respectively. Li metal was used for are performed at ambient temperature. 共c兲 Equivalent circuit used in this
the four electrodes. The four-probe system eliminates resistances at the Li/ study. Rbulk: resistances of solid and liquid electrolytes; RLLT gb: grain bound-
liquid electrolyte interface. Detectable resistances are the bulk- and grain- ary resistance of LLT; CLLT gb: capacitance of LLT grain boundary; Rint:
boundary resistances of ceramic electrolytes, the bulk resistance of liquid resistance of lithium-ion transfer at interface between solid and liquid elec-
electrolytes, and the resistance at the electrolyte/electrolyte interface. Mea- trolytes; Cint: capacitance for lithium-ion transfer at interface between solid
surements are performed over the frequency region of 100 kHz to 10 mHz. and liquid electrolytes.

Downloaded on 2014-10-18 to IP 141.214.17.222 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 152 共11兲 A2151-A2154 共2005兲 A2153

Table I. Reaction enthalpies determined by the density functional


theory using GAUSSIAN 98W (Ref. 9).

Heats of reaction 共⌬H兲


Solvent Donor no. 共kJ mol−1兲a
DMSO 29.8 共Ref. 11兲 −235.5
PC 15.1 共Ref. 11兲 −217.1
FEC —b −190.3
a
Reaction enthalpies 共Li+ + solvent = Li+ − solvent兲 at 298.15 K ac-
cording to ⌬H = H共Li+ − solvent兲 − 兵H共solvent兲 + H共Li+兲其.
b
Not applicable.

Figure 4. Temperature dependence of interfacial resistance for 共solid


circles兲 La0.55Li0.35TiO3 /PC-LiCF3SO3 and 共open circles兲 OHARA glass/PC-
LiCF3SO3. Activation energies were evaluated from the slope. The difference
in interfacial resistances is ascribed to the difference in wettability between
ceramic electrolytes and PC-based electrolyte.

energies are not influenced by ceramic electrolytes, the barrier


should be due to the liquid electrolyte. We focused on the solvation
abilities in electrolytes.
Three electrolytes, PC, DMSO, and FEC, containing LiCF3SO3
were selected. OHARA glass was used as a ceramic electrolyte for
experimental simplicity. The cell shown in Fig. 2 was used and the
activation energies for lithium ion transfer at the interface between
glass and liquid electrolytes were obtained in the same manner.
Figure 5 shows the activation energies for the three
liquid electrolytes. The activation energies are in the order

Figure 6. 共a兲 Typical Nyquist plot for a LLT/EMIBF4–LiBF4 interface at


323 K using the cell shown in Fig. 2. To avoid the reaction between elec-
Figure 5. Temperature dependence of lithium-ion transfer resistances at the trode and ionic liquid, Pt wires were used for the four electrodes. Measure-
interface between glass electrolyte and liquid electrolytes of PC, DMSO, and ments are performed over the frequency region of 100 kHz to 10 mHz. Num-
FEC containing 1 mol dm−3 LiCF3SO3. Activation energies were obtained bers in solid squares refer to the logarithms of frequencies. The contact area
from the slope. The difference in charge transfer resistances should be due to of LLT and EMIBF4–LiBF4 is kept constant at 0.20 cm2. 共b兲 Temperature
the difference in wettability between the ceramic electrolyte and the liquid dependence of lithium-ion transfer resistance at the interface between LLT
electrolytes of PC, DMSO, and FEC. and EMIBF4–LiBF4. The activation energy was obtained from the slope.

Downloaded on 2014-10-18 to IP 141.214.17.222 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
A2154 Journal of The Electrochemical Society, 152 共11兲 A2151-A2154 共2005兲

DMSO 共69.9 kJ mol−1兲 ⬎ PC共57.3 kJ mol−1兲 ⬎ FEC共31.5 kJ Conclusions


mol−1兲. Since DMSO gives a higher Gutmann donor number than Lithium-ion transfer through the electrode/electrolyte interface
PC,11 the interaction between lithium ion and DMSO might be was examined using a model interface composed of a lithium ion-
stronger than that between lithium ion and PC, which is consistent conductive ceramic and liquid electrolytes 共or ionic liquid兲 to focus
with the above order. Although no donor number has been reported only on lithium ion transfer without a redox reaction. As a result, the
for FEC, the electron-donating ability of FEC is expected to be following conclusions can be drawn:
lower than that of PC because fluorine atom is strongly electrone-
gative. Therefore, the interaction between lithium ion and FEC 1. The activation energies were quite large and consistent with
should be weaker than that between lithium ion and PC. These facts the interaction between lithium ion and solvents in an electrolyte as
qualitatively support the above order for activation energies. determined by a theoretical calculation.
Table I gives the reaction enthalpies, and the values are 2. When 1-ethyl-3-methylimidazolium tetrafluoroborate contain-
in the order DMSO 共−235.5 kJ mol−1兲 ⬎ PC共−217.1 kJ mol−1兲 ing 1 mol dm−3 LiBF4 was used as an ionic liquid, the activation
⬎ FEC共−190.3 kJ mol−1兲. The difference in activation energies is energy for the lithium ion transfer at the interface between LLT and
consistent with the difference in the reaction enthalpies among sol- the ionic liquid was as small as 25 kJ mol−1. An ionic liquid may be
vents: the difference between ⌬Ea and ⌬H for the DMSO and PC useful as an electrolyte in lithium ion batteries to achieve high-
systems = 12.6 and 18.4 kJ mol−1, respectively, and the difference power use.
between ⌬Ea and ⌬H for the PC and FEC systems = 25.8 and
26.8 kJ mol−1, respectively. Hence, the solvating ability of solvents Practical lithium ion batteries are more complicated and contain
in electrolytes, i.e., ion-dipole interaction, decisively affects the ac- other factors such as a solid electrolyte interphase, potential, current
tivation barrier at the interface. distribution in both phases of the electrolyte, and active materials
The above results show that the activation energies correlate with inside composite electrodes. Nevertheless, lithium ion transfer at the
lithium ion and solvent interaction 共under the assumption that the interface between the positive 共negative兲 active material and electro-
transition states are at the same energy level in these systems兲. Thus, lyte plays an important role in battery reactions. Based on the
a study on ion transfer at the interface between an ionic liquid and present results, reduction of the interaction between lithium ion and
solvent in electrolyte appears to be a promising method for enhanc-
ceramic electrolyte should be very interesting. In this system, no
ing the rate performance of lithium ion batteries.
ion-dipole interactions should exist, but repulsive and attractive
force should be present among constituent ions in ionic liquids. Kyoto University assisted in meeting the publication costs of this article.
Ionic liquid 1-ethyl-3-methylimidazolium tetrafluoroborate
共EMIBF4兲 containing 1 mol dm−3 LiBF4 was used. Hereafter, this References
ionic liquid is called EMBIF4-LiBF4. Lithium-ion transfer at the 1. T. Iwahori, Y. Ozaki, A. Funahashi, H. Momose, I. Mitsuishi, S. Shiraga, S. Yoshi-
take, and H. Awata, J. Power Sources, 81-82, 872 共1999兲.
LLT/EMIBF4-LiBF4 interface was studied by ac impedance spec- 2. For example, A. Van der Ven and G. Ceder, Electrochem. Solid-State Lett., 3, 301
troscopy using a four-electrode-electrochemical cell as mentioned 共2000兲.
above. In this case, Li electrodes cannot be applied due to the high 3. A. Funabiki, M. Inaba, T. Abe, and Z. Ogumi, J. Electrochem. Soc., 146, 2443
共1999兲.
reactivity of EMIBF4-LiBF4 toward Li metal, and therefore Pt wires 4. T. Abe, H. Fukuda, Y. Iriyama, and Z. Ogumi, J. Electrochem. Soc., 151, A1120
are used as four electrodes. Measurements were conducted over a 共2004兲.
frequency region of 100 kHz to 100 mHz with the applied voltage 5. Y. Inaguma, C. Liquan, M. Itoh, and T. Nakamura, Solid State Ionics, 70/71, 196
amplitude of 30 mV. 共1994兲.
6. Y. Inaguma, C. Liquan, M. Itoh, and T. Nakamura, Solid State Commun., 86, 689
Figure 6a shows a typical Nyquist plot for the above system with 共1993兲.
two semicircles. The semicircle in the higher frequency region is 7. J. Fu, J. Am. Ceram. Soc., 80, 1901 共1997兲.
ascribed to the grain-boundary resistance of LLT as before 共Fig. 3a兲. 8. J. Fu, Solid State Ionics, 96, 195 共1997兲.
Figure 6b shows the temperature dependence of the lithium ion 9. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R.
Cheeseman, V. G. Zakrzewski, J. A. Montgomery Jr., R. E. Stratmann, J. C. Burant,
transfer resistances at the interface between LLT and EMIBF4- S. Dapprich, J. M. Millam, A. D. Daniels, K. N. Kudin, O. F. M. C. Strain, J.
LiBF4. The activation energy is evaluated to be ca. 25 kJ mol−1. Tomasi, B. Barone, M. Cossi, R. Cammi, B. Mennucci, C. Pomelli, C. Adamo, S.
This small value is almost identical to that for solvated lithium ion Clifford, J. Ochterski, G. A. Petersson, P. Y. Ayala, Q. Cui, K. Morokuma, D. K.
Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. Ciolovski, J. V. Ortiz,
transfer at the graphite electrode.4 We used other ionic liquids, and V. V. Stefanov, G. Liu, A. Liashensko, P. Piskorz, I. Komaroni, R. Gomperts, R. L.
found that the activation energy for lithium ion transfer at the inter- Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, C.
face between LLT and ionic liquids depends on the ionic liquids. Gonzalez, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, J. L.
The details are under investigation. Based on these results, ionic Andres, M. Head-Gordon, E. S. Replogle, and J. A. Pople, GAUSSIAN 98, Gauss-
ian Inc., Pittsburgh, PA 共1998兲.
liquids may be promising electrolytes, which can reduce the high 10. T. Abe, M. Ohtsuka, F. Sagane, Y. Iriyama, and Z. Ogumi, J. Electrochem. Soc.,
activation barrier at the electrode/electrolyte interface in lithium ion 151, A1950 共2004兲.
batteries. 11. V. Gutmann, Electrochim. Acta, 26, 661 共1976兲.

Downloaded on 2014-10-18 to IP 141.214.17.222 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

You might also like