Sunteți pe pagina 1din 6

Materials Research Bulletin 43 (2008) 36013606

www.elsevier.com/locate/matresbu

The synthesis of composite SnSnSb films for lithium-ion


batteries by electrochemical deposition
S.Q. Zhang a,b, C.H. Chen a,*
a

Department of Materials Science and Engineering, University of Science and Technology of China,
Hefei, Anhui 230026, PR China
b
Department of Chemistry and Life Science, Chuzhou University, Chuzhou, Anhui 239012, PR China
Received 24 August 2007; received in revised form 6 December 2007; accepted 9 January 2008
Available online 18 January 2008

Abstract
Composite SnSnSb nano-crystalline films were fabricated on Cu substrate by an electrochemical deposition process. X-ray
diffraction, scanning electron microscopy, and galvanostatic cell cycling were used to characterize the structures and electrochemical properties of the films. The as-deposited films consist of only Sn/SnSb composite nanocrystals with a rather dense
morphology. The SnSnSb composite electrode gives rise to a very small (6%) initial capacity loss and a rather high specific
capacity of about 650 mAh/g, which is significantly higher than the values reported in the literature. The coulombic efficiency
during chargedischarge cycles is close to 100%.
# 2008 Elsevier Ltd. All rights reserved.
Keywords: A. Alloys; B. Chemical synthesis; C. X-ray diffraction; C. Impedance spectroscopy; D. Electrochemical properties

1. Introduction
Lithium storage metals and alloys (intermetallic compounds), such as Sn, Sb and SnSb, have been the subject of
considerable investigation as potential new negative electrodes in lithium-ion batteries [16]. They can deliver a
substantially higher capacity of over 1000 mAh/g, compared to carbonaceous materials such as graphite which has a
theoretical capacity of only 372 mAh/g. What has hindered their practical applications so far are the enormous volume
changes which occur during cycling. In the case of compact and coarse electrode morphology this volume change
immediately results in cracking and crumbling of the active material. The too large volume-change issue can be to
some extent resolved by using nano-crystalline or porous materials, and by introducing multi-component instead of
single-component materials. The pores or second component may serve as buffer to digest the volume change during
lithium insertion and extraction so that the mechanical integrity of the electrode is secured. Successful examples are
binary systems like SnySb [79], SnyAg [10], CuxSn [11] and FexSn [12]. In this study, we prepared a composite Sn
SnSb thin film by electrochemical deposition technique. It has exhibited excellent electrochemical performance as a
negative electrode material for lithium-ion batteries.

* Corresponding author. Fax: +86 551 3601592.


E-mail address: cchchen@ustc.edu.cn (C.H. Chen).
0025-5408/$ see front matter # 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.materresbull.2008.01.004

3602

S.Q. Zhang, C.H. Chen / Materials Research Bulletin 43 (2008) 36013606

2. Experimental
An aqueous solution containing 18 g/L SbCl3, 30 g/L SnCl2, 70 g/L citric acid and 50 g/L sodium citric acid was
used as a precursor solution for electrochemical deposition. At a current density of 2 mA/cm2 and room temperature, a
film was electroplated onto a copper substrate without stirring the solution during deposition [13]. The
electrodeposited layer covered a geometric area of 3.0 cm2 and had a thickness of around 2.5 mm after 1.5 h of
deposition. It was dried in a vacuum oven at 70 8C. For comparison, a Sn film was also deposited with only the SnCl2
solution as the precursor.
The electrochemical measurements were conducted by using a three-electrode test cell. Lithium metal was used as
both the counter and reference electrodes. The Sn and SnSb films were used to fabricate CR2032 coin cells with 1 M
LiPF6 in an ethylene carbonatediethyl carbonate (ECDMC) = 1/1 (v/v) mixture as the electrolyte. The cell assembly
was carried out in an argon-filled dry glove box.
The electrochemical performance of these cells was cycled with a Newware 610 multi-channel battery cycler
between 0.1 and 1.8 V at several different current densities at room temperature. The structure of these films before
cycling and after 40 cycles was investigated by scanning electron microscopy (SEM, Hitachi S-4500) and X-ray
diffractometry (XRD, Philips XPerPro Super, Cu Ka radiation).
The impedance of these cells after different cycles was also measured with an electrochemical workstation
(CHI604A) in the frequency range from 0.01 Hz to 100 kHz. The applied ac voltage amplitude was 5 mV. All of the
cells had been discharged to 0.1 V at 0.2 mA/cm2 and rested for 0.5 h before each impedance measurement.
3. Results and discussion
3.1. Structural analyses of SnSb film
Fig. 1 shows the XRD pattern of the SnSb film made by electrodeposition. It can be seen that besides the substrate
copper the film is composed of Sn and SnSb metallic compound. Both phases have rather high degree of crystallinity.
Based on the peak intensity of these two phases detected, their amount is comparable. Therefore, the film composition is a
SnSnSb composite. It should be mentioned that this sample is an as-deposited film that has been only dried at 70 8C.
Apparently, no annealing treatment is needed to form crystalline film. Thus, the particle size can be controlled fine.
3.2. Electrochemical performance of SnSnSb and Sn electrodes
Theoretically, in the potential range of 0.11.8 V investigated in this study, each Sn atom can accommodate
maximum 2.6Li (i.e. Li13Sn5) whereas each Sb atom can accommodate 3Li (i.e. Li3Sb), the theoretical capacity of Sn
and SnSb is 586 and 622 mAh/g, respectively. If the molar ratio in the SnSnSb composite is 1:1, its theoretical
capacity can be calculated as 611 mAh/g. Fig. 2 shows the electrochemical cycling results of the SnSnSb composite
and pure Sn films. The first discharge capacity of the SnSnSb film is about 697 mAh/g and its reversible capacity is
about 650 mAh/g (Fig. 2a). This means that its composition is around SnSnSb (1:1), which is consistent with above

Fig. 1. XRD patterns of a SnSnSb composite film.

S.Q. Zhang, C.H. Chen / Materials Research Bulletin 43 (2008) 36013606

3603

Fig. 2. Electrochemical cycling performance of Sn/Li and SnSnSb/Li cells. (a) Capacity vs. cycle number and (b) coulombic efficiency vs. cycle
number. The cells were cycled in the voltage range of 1.80.1 V at a current density of 0.2 mA/cm2.

XRD result (Fig. 1). Note that the measured capacity here is slightly higher than its theoretical capacity, which might
be attributed to the contribution of lithium stored at the phase boundaries between Sn and SnSb. However, the first
discharge capacity of the Sn film is about 500 mAh/g and its reversible capacity is only 200 mAh/g, which is much less
than its theoretical value. The coulombic efficiency of the two half cells (Fig. 2b) also indicates that the SnSnSb film
is better than the pure Sn film. The difference in the initial coulombic efficiency is remarkable because it is 94.3% for
the SnSnSb and only 43% for the Sn. The coulombic efficiency of 94.3% is also significantly higher than that reported
in literature [14].
The advantages of the SnSnSb composite in electrochemical performance lie in its microstructure. During the
charge/discharge, the lithiation or delithiation in the component Sn, which is in both Sn and SnSb, cause expansion or
contraction of the electrode volume. The component antimony (Sb) is not only active electrochemically but also can
perform as a matrix buffering the expansion or contraction of the tin [15]. Additional preconditions for above are (1)
that the reactant Sn is finely dispersed in the matrix, (2) that the matrix allows the electrons and lithium ions move between
the reactive domains or is a mixed (electron and lithium cation) conductor and (3) that the reactive domains are
sufficiently small, so that the benefits of small particle size are effective. Hence, the cycling performance of lithium
alloys can be significantly improved if intermetallic and/or composite hosts are employed instead of pure metals.
3.3. Electrochemical alloying of lithium with SnSnSb
Fig. 3 shows the initial discharge curve of the SnSnSb sample cycled at constant current of 0.2 mA/cm2 over the
voltage range of 0.11.8 V. The discharge profiles were similar to those of Sn as reported in the literature [16]. The
electrochemical reduction of a SnSnSb electrode in a Li+-conducting electrolyte leads to the subsequent formation of
Li3Sb and a number of intermetallic phases LixSny at room temperature. The cyclic voltammetry measurement (Fig. 4)
shows the similar dischargecharge characteristic as obtained from the voltage profiles (Fig. 3). After removal of
lithium from LixSnSbn, the original two-phase matrix (Sn + SnSb) is restored.

3604

S.Q. Zhang, C.H. Chen / Materials Research Bulletin 43 (2008) 36013606

Fig. 3. Characteristic of the first dischargecharge curve of the electrochemical deposited SnSnSb electrode in 1 M LiPF6.

Fig. 4. The cyclic voltammogram of a SnSnSb/Li cell measured at the voltage scan rate of 5 mV s 1.

3.4. Electrochemical impedance spectroscopy


A recovery of the SnSb phase can still take place after complete lithium extraction from the separate Li3Sb and
LixSn phases during the second cycle. A gradual deterioration of the SnSb crystal structure is also unavoidable. It
seems that insertion of lithium into, or extraction from, the SnSb host matrix is not a completely topotatic reaction.
Excess lithium insertion will break down the host crystal lattice and essentially change the basic structural
configuration. It is found from Fig. 5 that the electrochemical impedance of the LiSnSb alloy composite electrode
becomes progressively larger during cycling. The high impedance increases the measured electrode overpotential,
resulting in a decline in the capacity for a fixed cutoff potential [7].

Fig. 5. Change of impedance spectra of a SnSnSb composite electrode during cycling.

S.Q. Zhang, C.H. Chen / Materials Research Bulletin 43 (2008) 36013606

3605

Fig. 6. SEM micrographs of electrochemical deposited a SnSnSb film on Cu substrate, before cycling (a) and after 40 cycles (b).

3.5. Scanning electron microscopy


To our knowledge, pure Sn is severely cracked and delaminated from the Cu substrate after a few cycles, which has
been reported in the literature [10]. Judging from micrographs (see Fig. 6) the small differences in the antimonytin
intermetallic film before and after the process of chargedischarge, the morphology of the SnSnSb matrix remains
fairly intact even after 40 cycles. The surface becomes even smoother during cycling. Obviously, after the considerable
expansion due to the first insertion of lithium, the composite intermetallic SnSnSb material is almost dimensionally
stable in the following cycles. This structural stability can explain the good cyclability of the SnSnSb composite film.
Furthermore, it can be noticed that the particle size of the cycled film is smaller that of the as-deposited composite film.
This pulverization effect can explain the initial rise in the capacity during the first few cycles (Fig. 2a) because more
lithium can be stored at the phase boundaries.
4. Conclusions
Nano-crystalline SnSnSb composite has been successfully fabricated using electrochemical deposition onto the
Cu substrate. The antimonytin intermetallic materials have been investigated as anode materials to replaced carbon in
lithium-ion batteries. This preliminary study has demonstrated the strong effect that an active matrix element can have
on the utilization of an active Sn component than pure metal tin. It is anticipated that with improved processing
techniques, further enhancement of practical capacities and cycling behavior of these antimonytin intermetallic
materials will be achieved.
Acknowledgements
This study was supported by National Science Foundation of China (grant Nos. 50372064 and 20471057).
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]

J. Yang, M. Winter, J.O. Besenhard, Solid State Ionics 90 (1996) 281.


I. Rom, M. Wachtler, I. Papst, M. Schmied, J.O. Besenhard, F. Hofer, M. Winter, Solid State Ionics 143 (2001) 329.
O. Mao, R.A. Dunlap, J.R. Dahn, Solid State Ionics 118 (1999) 99.
T. Zhang, S.C. Zhang, X.P. Qiu, W.T. Zhu, L.Q. Chen, J. Power Sources 166 (2007) 503.
H.P. Zhao, C.Y. Jiang, X.M. He, J.G. Ren, C.R. Wan, Electrochim. Acta 52 (2007) 7820.
S. Matsuno, M. Noji, T. Kashiwagi, M. Nakayama, M. Wakihara, J. Phys. Chem. C 111 (2007) 7548.
J. Yang, Y. Takeda, N. Imanishi, J. Electrochem. Soc. 146 (1999) 4009.
Y. Jun, W. Mario, W. Martin, J.O. Besenhard, Electrochem. Solid State Lett. 2 (1999) 161.
M. Wachtler, M. Winter, J.O. Besenhard, J. Power Sources 105 (2002) 151.

3606
[10]
[11]
[12]
[13]
[14]
[15]
[16]

S.Q. Zhang, C.H. Chen / Materials Research Bulletin 43 (2008) 36013606


J. Yang, M. Winter, J.O. Besenhard, Solid State Ionics 90 (1999) 383.
K.D. Kepler, J.T. Vaughey, M.M. Thackeray, J. Power Sources 8182 (1999) 383.
O. Mao, J.R. Dahn, J. Electrochem. Soc. 146 (1999) 414.
J.O. Besenhard, J. Yang, M. Winter, J. Power Sources 68 (1997) 87.
M. Wachtler, J.O. Besenhard, M. Winter, J. Power Sources 94 (2001) 189.
M. Winter, J.O. Besenhard, M.E. Spahr, P. Novak, Adv. Mater. 10 (1998) 725.
M. Winter, J.O. Besenhard, J.H. Albering, J. Yang, M. Wachtler, Prog. Batteries Battery Mater. 17 (1998) 208.

S-ar putea să vă placă și