Vous êtes sur la page 1sur 5

Solid State Ionics 285 (2016) 96100

Contents lists available at ScienceDirect

Solid State Ionics


journal homepage: www.elsevier.com/locate/ssi

Complex hydride for composite negative electrodeapplicable to


bulk-type all-solid-state Li-ion battery with wide temperature operation
Koji Yoshida a,, Shohei Suzuki b, Jun Kawaji b, Atsushi Unemoto a, Shin-ichi Orimo a,c
a
b
c

WPI-Advanced Institute for Materials Research (WPI-AIMR), Tohoku University, 2-1-1 Katahira, Aoba-ku, Sendai, 980-8577, Japan
Research and Development Group, Center for Exploratory Research, Hitachi. Ltd, 7-1-1, Omika, Hitachi, Ibaraki, 319-1292, Japan
Institute for Material Research, Tohoku University, 2-1-1 Katahira, Aoba-ku, Sendai, 980-8577, Japan

a r t i c l e

i n f o

Article history:
Received 19 April 2015
Received in revised form 18 June 2015
Accepted 17 July 2015
Available online 11 August 2015
Keywords:
Complex hydride
Lithium-ion battery
All-solid-state battery
Solid electrolyte
Li-ion conduction

a b s t r a c t
A composite negative electrode for use in bulk-type all-solid-state batteries was developed by combining the solid
electrolyte Li4(BH4)3I (LiBH4 doped with LiI), acetylene black (AB) (as a conductive additive), and lithium titanate
(Li4Ti5O12, LTO). Because of the highly deformable nature of the complex hydride, hand milling and subsequent uniaxial pressing were sufcient to ensure tight interfacial contact. As a result, the composite negative electrode allowed
successful operation of a bulk-type all-solid-state battery from room temperature (23 C) to 150 C. The composite
negative electrode was characterized via microstructural observation and using electrochemical techniques. In the
prepared composite negative electrode, LTO, Li4(BH4)3I, and AB were homogeneously dispersed and formed tightly
contacted interfaces. At 150 C, the battery exhibited high discharging capacities of 170 and 158 mAh g1 for the 1st
and 2nd cycles, respectively, which corresponded to LTO utilization ratios of 97% and 90%. Moreover, it retained
140 mAh g1 at the 100th cycle, which equates to a 90% capacity retention ratio from the 2nd cycle. Furthermore,
the battery was successfully operated at room temperature (23 C), exhibiting discharging capacities of
122 mAh g1 for the 1st cycle and 111 mAh g1 for the 5th cycle (capacity retention ratio of 91%). At both temperatures, the coulombic efciency was nearly 100%, indicating that the charge/discharge reactions proceeded
without any signicant side-reactions. This good performance is due to the small charge-transfer resistance at
the LTO interface resulting from the tight contact between the constituents of the composite electrode, and suggests that LiI-doped LiBH4 is a promising solid electrolyte for practical composite electrodes.
2015 Elsevier B.V. All rights reserved.

1. Introduction
Li-ion batteries are both promising energy storage devices due to
their high energy and power densities compared to other devices and
a key technology for the development of electronic devices, including
electric vehicles. To increase the scale of Li-ion batteries, however,
their safety proles must be improved; ammable organic solvents
are commonly used and dendrite formation can cause short circuits
[1]. In addition, temperature increases during battery operation often
lead to remarkable capacity fading [2,3]. All-solid-state batteries,
which use solid electrolytes instead of organic solvents, are one solution
for avoiding the above-mentioned issues.
Fig. 1 shows a schematic illustration of a typical all-solid-state battery, in which a solid electrolyte layer is sandwiched between composite negative and positive electrode layers that comprise active materials,
the solid electrolyte, and conductive additives when needed. To develop
an all-solid-state battery, it is important to choose an appropriate solid
electrolyte for each battery component; specically, the solid electrolyte
should have sufcient reducing and oxidizing abilities for the negative
Corresponding author. Tel.: +81 22 215 2094; fax: +81 22 215 2091.
E-mail address: k-yoshida@imr.tohoku.ac.jp (K. Yoshida).

http://dx.doi.org/10.1016/j.ssi.2015.07.013
0167-2738/ 2015 Elsevier B.V. All rights reserved.

and positive electrode layers, respectively. In addition to high Li-ion


conductivity, it should also be highly deformable in order to increase
the contact area between the constituents, and thereby minimize interfacial resistance. Over the past several decades, various inorganic solid
electrolytes suitable for use in all-solid-state batteries have been developed, including perovskite-type [4,5], NASICON [6,7], LISICON [8,9],
thio-LISICON [10,11], garnet-type [12,13], and glass-ceramic [14]
materials.
The typical complex hydride solid electrolyte LiBH4 undergoes a phase
transition at 110 C, and its high-temperature (HT) phase (space group:
P63mc) [15] exhibits an ionic conductivity as high as 2.0 103 S cm1
at this temperature [1618]. In addition, the HT phase of LiBH4 can be stabilized at lower temperatures by doping with a lithium halide (LiCl, LiBr,
or LiI) [1719]. In fact, when LiBH4 is doped with LiI, i.e., Li4(BH4)3I, the HT
phase is stable even at room temperature, which results in a higher Li-ion
conductivity (2.0 105 S cm1) than is possible without LiI
(2.0 108 S cm1) [19]. In addition, the doped material is highly deformable, thermally stable up to 150 C, highly stable toward lithium
metal electrodes due to its high reducing ability, and electrochemically
stable over the range of 0 to 5 V vs. Li/Li+. In recent years, a number of
studies on the development of all-solid-state batteries employing complex hydride electrolytes have been reported. Takahashi et al. fabricated

K. Yoshida et al. / Solid State Ionics 285 (2016) 96100

Fig. 1. Schematic illustration of a typical bulk-type all-solid-state battery conguration.

a bulk-type all-solid-state battery with a LiBH4 solid electrolyte and a


LiCoO2 positive electrode [20,21]. In their study, the LiCoO2 surface was
coated with Li3PO4 in order to avoid the reductive decomposition of the
LiCoO2 and enable repeated charge/discharge cycling at 120 C. The initial
discharge capacity was 89 mAh g1, and the capacity retention ratio at
the 30th cycle was 97%. Unemoto et al. incorporated a LiBH4 solid electrolyte into a bulk-type all-solid-state lithium-sulfur battery [22]. They realized discharge capacities as high as 1140 and 730 mAh g1 for the initial
and 45th discharges, respectively. Sveinbjrnsson et al. assembled an allsolid-state battery using a thin-lm lithium titanate (Li4Ti5O12, LTO) electrode and LiBH4LiI solid electrolyte [23]. The use of the LiBH4LiI solid
electrolyte allowed low temperature operation (60 C). Because LTO has
a low redox potential of 1.55 V (vs. Li/Li+), it forms a stable interface
with the LiBH4LiI electrolyte; therefore, in contrast to the LiCoO2 positive
electrode, no surface coating was necessary. This all-solid-state battery
exhibited a high initial discharge capacity of 142 mAh g1, corresponding
to an LTO utilization ratio of 81%. As demonstrated by these studies, complex hydride electrolytes should be considered as appropriate components for incorporation into composite negative electrodes intended for
use in bulk-type all-solid-state batteries.
Herein, we describe the fabrication of a composite electrode that
uses a Li4(BH4)3I solid electrolyte and an LTO electrode, and examine
its performance over a wide temperature range (23 to 150 C). LTO undergoes extremely small volume changes during battery operation (less
than 0.2%), and allows good cycle life for practical Li-ion batteries [24].
In general, LTO is used as a negative electrode active material, as described in Fig. 1. In the present study, however, we examined the performance of an LTO composite electrode by combining it with lithium
metal electrode prior to the assembly of a practical rechargeable battery
with LiCoO2 or other high-voltage positive electrodes. Because lithium
metal has lower redox potential, LTO functions as a positive electrode.
The system performance is discussed based on the microstructure and
the electrochemical analysis results.
2. Experimental
The Li4(BH4)3I solid electrolyte was synthesized from LiBH4 (N 95%,
Sigma-Aldrich) and LiI (99.999%, Sigma-Aldrich) using a high-energy
mechanical milling technique. All of the operations described below

97

were performed under argon. First, the powders were mixed in a 3:1
molar ratio for 30 min in an agate mortar, and then the mixture was
transferred into a stainless-steel vessel with 20 stainless-steel balls
(diameter: 7 mm). Mechanical ball milling was performed in a planetary ball mill (P-7, Fritsch) at a rotation rate of 400 rpm for 5 h under
Ar. LTO (N99%, Sigma-Aldrich) was used as the active material.
Acetylene black (AB), used as the conductive additive, was dried in a
vacuum oven at 120 C for more than 1 h before use. The LTO composite
electrode was prepared by mixing LTO, Li4(BH4)3I, and AB in a
32:61:7 wt.% ratio in an agate mortar for 30 min. To form the
electrode-electrolyte composite pellet, the powders of each (9.5 and
53 mg, respectively) were transferred to a die (8 mm in diameter) and
uniaxially pressed at 240 MPa and room temperature. The thicknesses
of the electrode and electrolyte layers were approximately 0.1 and
0.7 mm, respectively. A lithium foil was then used as the negative electrode. The experimental conguration for the battery tests is described
elsewhere [22]. The battery tests were performed using a Solartron1470
CellTest system (Solartron Analytical) at C-rates of 0.2 C (current density: 200 A cm2) at 150 C and 0.02 C (current density: 20 A cm2) at
23 C (room temperature) in the potential range of 1.02.0 V.
Electrochemical impedance measurements were obtained using a
chemical impedance meter (Model 3532-80, Hioki E. E. Corp.) at room
temperature in the frequency range of 10 MHz to 4 Hz. Prior to each impedance analysis, the cell was charged to a 50% state of charge (SOC).
The microstructure and elemental distribution of the composite
electrode were evaluated via scanning electron microscopy (SEM,
Hitachi S-4800) and energy dispersive X-ray spectroscopy (EDS, Horiba
EX-350), respectively. For these analyses, the composite was prepared
in the ratio LTO:Li4(BH4)3I:AB = 74:22:4 wt.% to avoid possible damage
to the sample during electron beam irradiation. SEM analysis revealed
that the constituents were lled successfully at this mixing ratio (see
Fig. 3(a)), and no remarkable differences were observed in the microstructure compared to that of the electrode used for the electrochemical
analyses (see Fig. S-1 in the Supplementary Information). An LTO composite pellet was also prepared using the foregoing procedure at a thickness of 0.1 mm and attached to a piece of indium foil (0.2 mm thick) for
handling. The cross-sectional surface was obtained via ion milling
(Hitachi E-3500).
3. Results and discussion
3.1. Microstructure of the composite electrode
Successful, stable battery operation and enhanced electrode utilization ratios require a homogeneous distribution of the LTO, Li4(BH4)3I,
and AB constituents in the composite electrode layer, including the formation of tight interfaces. As expected, homogeneous distributions of
the LTO, Li4(BH4)3I, and AB were observed in the cross-sectional SEM
image of the composite electrode, as displayed in Fig. 2(a). Although a
few pores were observed, which most likely formed during the pellet
fabrication process, the composite electrode pellet showed good lling
of the constituents. The elemental distributions from the same area as
used for the SEM observation revealed homogeneous distributions of
iodine and titanium over the entire composite electrode (Fig. 2(b) and
(c)). These results demonstrate that the solid electrolyte and the LTO active material were uniformly mixed.
A higher resolution image of the surface (Fig. 3(a)) shows particles
of various sizes observed in rather bright contrast. Both the large particles in the left and lower right portions of the image and the small particles distributed between them are considered to be LTO. The deviation
of the particle size is attributed to the original size distribution of the
commercially available LTO powder. The spaces between the small
LTO particles observed in darker contrast are assumed to be lled by
the solid electrolyte, because no charging of the edge effect was observed, indicating that the spaces are not pores. The elemental analysis
for the indicated area in Fig. 3(a) revealed the presence of titanium,

98

K. Yoshida et al. / Solid State Ionics 285 (2016) 96100

solid electrolyte should lead to low interfacial resistance and high LTO
utilization ratios.

3.2. Battery discharge/charge properties

Fig. 2. (a) Cross-sectional SEM image of the LTO/Li4(BH4)3I/AB composite electrode and
distributions of (b) titanium and (c) iodine evaluated using EDS. The red square shown
in (a) indicates the observation area for Fig. 3. The lower edges of the images show the
In foil that was attached for handling.

iodine, and carbon, corresponding to the LTO, Li4(BH4)3I, and AB,


respectively. The uctuation of the contrast between the LTO particles
may be due to the local distribution of the AB close to the crosssectional surface. The observed tight binding between the LTO and

Fig. 3. (a) Cross-sectional, high-resolution SEM image of the LTO/Li4(BH4)3I/AB composite


electrode and (b) a schematic illustration of the surface area indicated in Fig. 2(a). (c) EDS
spectrum obtained for the rectangular area shown in (a).

The stable operation of a bulk-type all-solid-state battery using the


LTO-Li4(BH4)3I-AB composite electrode was successfully demonstrated
over a wide temperature range (i.e., at 150 C and room temperature).
It should be noted that LTO is typically used as a negative electrode active material; however, the LTO-Li4(BH4)3I-AB composite electrode
functioned as a positive electrode because lithium metal was used as
the negative electrode in order to examine the performance of the composite electrode. Thus, the Li insertion reaction into LTO is written as a
discharge reaction, while the Li extraction reaction from LTO as charging
reaction. Fig. 4 shows the discharge/charge proles of the bulk-type allsolid-state battery for the rst ten cycles operated at (a) 150 C and
0.2 C, and (b) the discharge capacity of the rst 100 cycles. As expected
for the reaction of LTO, a voltage plateau appeared at 1.55 V [24].
The rst discharge capacity was 170 mAh g1, which shifted to
158 mAh g1 in the second cycle. This second discharge capacity
corresponded to 89% of the LTO utilization ratio. For the second through
to the tenth cycle, the discharge/charge curves did not show any obvious shift, except for a small deviation of the voltage response at the beginning of the discharging process. The 100th discharge capacity was
140 mAh g 1, corresponding to a 90% capacity retention ratio from
the second discharging capacity. In addition, the nearly 100% coulombic
efciency throughout the dischargecharge cycles suggests that no remarkable side reactions occurred during the battery tests. This highly
stable cycling property is due to both the small volumetric change of
LTO during the dischargecharge process and the high deformability
of the Li4(BH4)3I, which minimizes the stress at the LTO-electrolyte interface and prevents possible detachment.

Fig. 4. (a) Typical discharge/charge proles for the Li|Li4(BH4)3I|LTO/Li4(BH4)3I/AB cell


operated at 150 C and 0.2 C, and (b) Coulombic efciency and discharge capacity as a
function of the number of cycles.

K. Yoshida et al. / Solid State Ionics 285 (2016) 96100

99

The dischargecharge cycle properties of the bulk-type all-solidstate battery at 23 C and 0.02 C were also examined (Fig. 5). Prior to
battery operation at room temperature, several dischargecharge cycles
were performed at 120 C in order to determine whether the battery
was operable. The temperature was then decreased to 23 C. The battery
exhibited a specic capacity as high as 122 mAh g1 at the initial discharge, and it remained as high as 111 mAh g1 for the 5th discharge,
which corresponds to a discharge capacity retention ratio of 91%. The
specic capacity during these ve cycles decreased linearly as the number of cycles increased, and the coulombic efciency was nearly 100%,
indicating that the ideal discharge/charge reaction took place, as was
the case at 150 C. The fading of the capacity and the shift of the voltage
plateaus in the charge and discharge traces may be attributed to an increase in the internal resistance due to unavoidable gas leakage into
the cell during the long-term measurements or changes in the local iodine uctuation in the complex hydride, which can inuence the Li-ion
conductivity. The existence of this gap between the charge and discharge plateaus suggests that the overpotential due to an internal resistance is signicant, in contrast to the results obtained for the battery test
at 150 C.
3.3. Electrochemical impedance of the battery
To examine the factors determining the internal resistance of the
battery observed in the discharge/charge proles at 23 C, electrochemical impedance measurements were performed for the all-solid-state
battery with the LTO/Li4(BH4)3I/AB composite electrode at a 50% SOC
and an Li-symmetric cell using the Li4(BH4)3I electrolyte, and the results
are shown in Fig. 6. For this analysis, the Warburg impedance, which is
included in the overpotential observed in Fig. 5, was not considered. The
impedance factor attributed to Warburg impedance appeared at lower
frequencies (see Supplementary Information Fig. S-2). A typical Nyquist
plot for the battery showed three half circles that are ascribed to

Fig. 6. Typical AC impedance spectra for the (a) Li|Li4(BH4)3I|LTO/Li4(BH4)3I/AB cell at a


50% SOC and (b) an Li|Li4(BH4)3I|Li cell, both at 23 C. The lled circles indicate data points
obtained at frequencies of 10n Hz (n = 16).

different resistive factors at high (N104 Hz), middle (N 250 Hz), and
low (b 250 Hz) frequencies. For the symmetric cell (Fig. 6(b)), only
two resistive factors, high- and middle-frequency resistance, were observed. Based on these results, the high- and middle-frequency resistance can be attributed to the solid electrolyte, i.e., due to the solid
electrolyte bulk and grain boundaries, respectively. The low frequency
resistance, which appeared only for the battery, is attributed to charge
transfer at the LTO-electrolyte interface. Note that no resistive factor
was observed that could be ascribed to the Li-electrolyte interface, likely
because it was too small to be distinguished. As can be seen in Fig. 6(a),
the interfacial and grain boundary resistances were relatively small
compared to that of the solid electrolyte bulk, and in good agreement,
as expected based on the microstructure analysis, which suggested
tight interfaces between the LTO, Li4(BH4)3I, and AB (cf. Fig. 3). To
achieve better battery performance through the reduction of internal
resistance, the development of a solid-state electrolyte that has a higher
Li-ion conductivity is necessary.
4. Conclusions

Fig. 5. (a) Typical discharge/charge proles for the Li|Li4(BH4)3I|LTO/Li4(BH4)3I/AB cell operated at 23 C, and 0.02 C and (b) Coulombic efciency and discharge capacity as a function of the number of cycles.

A composite negative electrode for all-solid-state batteries that can


operate over the wide temperature range from room temperature to
150 C was developed using the complex hydride Li4(BH4)3I as the
solid electrolyte. In the composite negative electrode, all of the constituents were dispersed homogeneously and in close contact with one
another due to the highly deformable nature of Li4(BH4)3I. A bulk-type
all-solid-state battery fabricated with this composite electrode exhibited high discharge capacities of 170 and 158 mAh g1 in the 1st and 2nd
cycles, respectively, and high capacity stability, with a 90% capacity retention ratio from the 2nd to the 100th cycle at 150 C. At 23 C, the battery exhibited discharge capacities of 122 and 111 mAh g1 for the 1st
and 5th cycles, respectively, corresponding to a capacity retention ratio
of 91%. At both temperatures, the coulombic efciency was nearly 100%,
indicating that side reactions were negligible. Such highly stable battery

100

K. Yoshida et al. / Solid State Ionics 285 (2016) 96100

operation over a wide temperature range demonstrates the applicability of this complex hydride solid electrolyte for practical composite
electrodes designed for use in bulk-type all-solid-state batteries, even
though the ionic conductivity still requires improvement. We are currently working on the assembly of practical batteries like that illustrated
in Fig. 1 using the composite negative electrode developed in this study
in combination with high-voltage positive electrodes.
Supplementary data to this article can be found online at http://dx.
doi.org/10.1016/j.ssi.2015.07.013.
Acknowledgement
This work was performed through the Industrial-Academic Partnership Project, Next Generation Innovative Battery Laboratory, between
Hitachi Ltd. and WPIAIMR, Tohoku University.
References
[1] J.M. Tarascon, M. Armand, Nature 414 (2001) 359.
[2] D. Aurbach, J. Power Sources 119121 (2003) 497.
[3] J. Shim, R. Kostecki, T. Richardson, X. Song, K.a. Striebel, J. Power Sources 112 (2002)
222.
[4] S. Stramare, V. Thangadurai, W. Weppner, Chem. Mater. 15 (2003) 3974.
[5] Y. Inaguma, C. Liquan, M. Itoh, T. Nakamura, T. Uchida, H. Ikuta, M. Wakihara, Solid
State Commun. 86 (1993) 689.

[6] H. Aono, E. Sugimoto, Y. Sadaoka, N. Imanaka, G. Adachi, Solid State Ionics 47 (1991)
257.
[7] V. Thangadurai, A.K. Shukla, J. Gopalakrishnan, J. Mater. Chem. 9 (1999) 739.
[8] H.Y.-P. Hong, Mater. Res. Bull. 13 (1978) 117.
[9] P.G. Bruce, J. Electrochem. Soc. 130 (1983) 662.
[10] R. Kanno, M. Murayama, J. Electrochem. Soc. 148 (2001) A742.
[11] N. Kamaya, K. Homma, Y. Yamakawa, M. Hirayama, R. Kanno, M. Yonemura, T.
Kamiyama, Y. Kato, S. Hama, K. Kawamoto, A. Mitsui, Nat. Mater. 10 (2011) 682.
[12] V. Thangadurai, H. Kaack, W.J.F. Weppner, J. Am. Ceram. Soc. 86 (2003) 437.
[13] R. Murugan, V. Thangadurai, W. Weppner, Angew. Chem. Int. Ed. Engl. 46 (2007)
7778.
[14] M. Tatsumisago, S. Hama, A. Hayashi, H. Morimoto, T. Minami, Solid State Ionics
154155 (2002) 635.
[15] S. Orimo, Y. Nakamori, J.R. Eliseo, A. Zttel, C.M. Jensen, Chem. Rev. 107 (2007) 4111.
[16] M. Matsuo, Y. Nakamori, S. Orimo, H. Maekawa, H. Takamura, Appl. Phys. Lett. 91
(2007) 3.
[17] M. Matsuo, S. Orimo, Adv. Energy Mater. 1 (2011) 161.
[18] A. Unemoto, M. Matsuo, S. Orimo, Adv. Funct. Mater. 24 (2014) 2267.
[19] H. Maekawa, M. Matsuo, H. Takamura, M. Ando, Y. Noda, T. Karahashi, S. Orimo, J.
Am. Chem. Soc. 131 (2009) 894.
[20] K. Takahashi, K. Hattori, T. Yamazaki, K. Takada, M. Matsuo, S. Orimo, H. Maekawa,
H. Takamura, J. Power Sources 226 (2013) 61.
[21] K. Takahashi, H. Maekawa, H. Takamura, Solid State Ionics 262 (2014) 179.
[22] A. Unemoto, S. Yasaku, G. Nogami, M. Tazawa, M. Taniguchi, M. Matsuo, T. Ikeshoji,
S. Orimo, Appl. Phys. Lett. 105 (2014) 083901.
[23] D. Sveinbjrnsson, A.S. Christiansen, R. Viskinde, P. Norby, T. Vegge, J. Electrochem.
Soc. 161 (2014) A1432.
[24] T. Ohzuku, J. Electrochem. Soc. 142 (1995) 1431.

Vous aimerez peut-être aussi