Sei sulla pagina 1di 39

0

e
0 0
0
PROCESSES AND APPLICATIONS FOR TIN AND TIN-BASED ALLOY SURFACE COATING TECHNOLOGIES

0 0

0
0

A Technical Review and Assessment of Recent Developments


,

Compiled for Tin Technology

0 0

STAGE 1 ELECTROLYTIC DEPOSITION PART 3 ELECTRODEPOSITION AND ELECTROPLATING OF TIN ALLOYS ELECTROCHEMICAL POWER SOURCES

e
0 0 0
0 0

0 0
0
BY
L. M. Baugh, BSc, MSc, PhD, Chem, FRSC Consultant

0 0
0 0

0
0

0 0
0
0 0

March 2005

0
0

CONTENTS

1. 2

Introduction

...........................................................................

Batteries 2.1

............................................................................... Lithium Ion Batteries .........................................................

4 4

2.1.1 Commercial and Historical Perspective Relating ...................4 to Lithium Batteries 2.1.2 Background on Alkali Metal Electrochemistry ..................... Relevant to the Development of Rechargeable Lithium-Ion B att eries 2.1.3 Applications of Tin in Lithium-Ion Batteries ........................
5

2.1.4 Pure Tin .................................................................. 2.1.5 Tin-Copper Alloys ...................................................... 2.1.6 Tin-Nickel Alloys ...................................................... 2.1.7 Tin-Zinc Alloys ......................................................... 2.1.8 2.19
2.2

6
9 19 23 27 27 28 28 28 29

Tin-Iron Alloys ......................................................... Tin-Antimony Alloys ..................................................

Lead-Acid Batteries

.........................................................

2.2.1 Background ............................................................. 2.2.2 Electrodeposited Pb-Sn Alloys .......................................


2.3 Alkaline-Manganese Batteries 3.

...........................................

Fuel Cells 3.1

............................................................................. Direct Methanol Fuel Cells ................................................

29 29 29

3.1.1 Background .............................................................

e
0 0
0 0
3.1.2 Research on Tin Modified Platinum Electrodes.. ... . . ......... ...30 3.1.3 Research on Tin Modified Carbon Supported.. .... ..... .... ...... 33 Platinum Electrodes

0
0
4.

Conclusions and Recommendations.. References.. . .. .

..........................................33

0
0 0
5 .
6.

. . ...................................................................36 ...............................................................39

a
0
0 0
0 0 0

Acknowledgements..

0 0 0 0

0 0
0

0
0

0 0
0

0
0

0 0

0 0
1.
INTRODUCTION

About 20 different electrodeposited tin alloys are known. These can be roughly divided according to whether their primary application is in the field of corrosion protection or as solders and coatings in the field of electronics. However, the current review (Part 3) is concerned with recent developments in a lesser known application of electrodeposited tin alloys, that is, in the field of electrochemical power sources. A subsequent review will deal with the topic of the application of tin alloys in solders and electronics (Part 4). Earlier reviews were concerned with the electrodeposition and electroplating of pure tin [ 11 and applications in the field of corrosion resistant coatings [2]. It is convenient to discuss the current topic in terrns of applications in either batteries or fuel cells.

2.

Batteries

This topic is hrther sub-divided according to whether lithium-ion batteries or lead-acid batteries are considered.

2.1 2.1.1

Lithium-Ion Batteries
Commercial and Historical Perspective Relating, to Lithium Batteries

0 0 0 0 0 0 0 0 cl 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

'

The enormous and rapid development of digital electronics that has taken place in recent years has enabled the creation of a large number of portable devices, such as computers, cell-phones, video cameras, etc. These devices are almost exclusively equipped with storage batteries built according to a single technology; namely nickelcadmium batteries. However, the need to produce ever smaller and lighter portable devices, together with the need for long operating autonomy, has forced researchers to consider new technologies. In the late eighties and early nineties, two new batteries appeared on the market : nickel-metal hydride and lithium-ion batteries. In just a few years, the performance of batteries produced using these technologies has improved considerably, to the extent that today it is superior to that of nickel cadmium batteries.
An important advantage of the above newer technologies from the environmental

standpoint is the absence of cadmium. But whilst it is known that Li-ion batteries are the most promising in energy terms, their production cost is relatively high. Nevertheless, if the cost per cycle of these systems is considered, it is noted that they become competitive as compared to the more economical Ni-Cd batteries and on the basis of the technical characteristics, environmental impact and cost per cycle, a considerable growth is expected for the Li-ion system on the world market. By contrast, Ni-metal hydride batteries are likely to have a more modest growth. Li-ion batteries utilise a so-called 'rocking-chair' technology in which the metallic lithium anode of conventional cells is replaced by a composite material capable of intercalating Li' ions in their structure via insertion reactions. This overcomes the

e
0
0
0 0 0 0 0
problem of dendrite formation and loss of reversibility during charge/discharge cycling. 2.1.2 Background on Alkali Metal Electrochemistry Relevant to the Development of Rechargeable Lithium-Ion Batteries Lithium batteries require the use of non-aqueous electrolyte systems that may be either solid or liquid. The former include lithium ion conductingpolymers and the latter are polar aprotic solvents, eg. organic ethers, esters and alkyl carbonates [31, containing dissolved salts having anions of the type: C104-, A&-, PF6- and BF4 . It should be noted that on a thermodynamic basis, lithium metal reacts readily with any relevant polar aprotic solvent, many salt anions and unavoidable atmospheric contaminants eg. H20, 02, CO2 and N2. Nevertheless, lithium metal is apparently stable in many polar aprotic systems due to its coverage by electronically insulating, passivating surface films. These films are formed due to the reduction of solution components by the active metal, which forms insoluble Li salts that precipitate on the lithium surface. It appears that many Li salts such as LiX (X = F, C1, Br); Li2CO3; Li20; LiOH; LiOR; and LiOC02R (R = alkyl group) when precipitated on lithium as thin surface films, are Li ion conductors. Hence, the lithium metal in the relevant electroyte solution is electrochemically active in spite of its passivation by the surface films [4]. This operational situation for Li metal anodes in battery systems limits their reversibility and may lead to severe safety problems in abuse cases such as short circuit, overcharge and exposure to air.
,

0
0
0 0

0 0

e
0
0
0

e
0 0

0
0

0 0

As a result of these considerations, the major efforts in this field were shifted towards R&D on Li-ion batteries in which Li metal anodes are replaced-bycarbonaceous materials that can insert lithium reversibly at potentials close to that of the Li/Li+ couple. It should be noted that the electrochemical behaviour of Li-C electrodes is also controlled by surface films of similar structure to those formed on lithium electrodes in the same solutions [ 5 ] . However, changing the anodic reaction from Li dissolutioddeposition to Li intercalatioddeintercalation (dicharginglcharging processes) leads to a considerable gain in stability, reversibility and enhanced safety, albeit at the expense of some loss in energy density. Li-ion batteries, in which the anodes are Li-C insertion compounds, the cathodes are Li,MO, insertion materials (M = transition metal oxide such as Mn, Ni, CO,V), and the electrolyte systems are organic polar aprotic solutions, have become a significant commercial reality, and as discussed previously, their market share continues to increase. It is claimed that a gravimetric energy density of >1 50Wh/kg, wide temperature range of operation (-3O0CCT<6O0C) and the possibility of many hundreds of full capacity chargedischarge cycles can be obtained [6].
2.1.3 Applications of Tin in Lithium Ion Batteries

0 0
0 0 0

The background has been summarised by Tamura et al. [7]. Many kinds of carbon materials are now widely used for the negative electrode of lithium-ion batteries. As a result of competitive development efforts, improvements have been made and graphite with high crystallinity, offering a specific capacity of 370 mAh g- and providing a good flatness in its voltage-capacity profile, is now

0
0
5

available. However, the demand for lithium ion batteries as a power supply for portable electric devices has steadily increased and their capacity requirements have become larger. The capacity of graphite, on the other hand, has already approached the theoretical limit of 372 mAh g-*for the compound CsLi. In order to enhance the energy density of lithium batteries, materials having a larger specific capacity than graphite must therefore be found. Tin-based materials have received much attention because of their much greater specific capacities. The theoretical capacity is 994 mAh g" based on the compound Li4.4Sn7far exceeding that of graphite. Tin oxides offer large capacity and good cycle performance so that the energy density of lithium-ion batteries is expected to rise as a result of their use. However, one key issue remains unresolved before commercialisation. This is the irreversible capacity decay experienced during the first cycle as a result of the formation of lithium oxide. In order to reduce this irreversible capacity decline, research has been focused on the use of tin and tin alloys as substrate materials for the negative electrode (anode) in lithium-ion batteries. Recent progress achieved in this field is reviewed below.

0 0 0

0 0 0 0 0 -0

2.1.4 Pure Tin


Although most of the succeeding discussion is concerned with lithium-ion battery developments involving the use of conventional non-aqueous liquid electrolytes containing a salt dissolved in a non-aqueous solvent, it is interesting to note that room temperature molten salts (RTMS) can also be used as electrolytes for these batteries. The background to this subject has been outlined by Fung and Zhu [8]. Materials intercalated with lithium ions such as LiCoO2, LiMn204 and LiNi02 have been used successfully as positive (cathode) materials in RTMS batteries. However, a suitable negative (anode) material for the battery is more difficult to find. Recently, Fung and Zhu [81 have developed a new RTMS system containing 1-methyl-3-ethylimidazolinium chloride/AlC13/SnC12 (3 :2:0.5) for the electrodeposition of tin on a copper substrate electrode. The 1-methyl-3ethylimidazolinium chloride component can be abbreviated as MEIC1. The tin film deposited at a potential more positive than -0.72V (reference electrode AlRTMS, where AlC13:MEICl= 66.7:33.3) under different temperatures produced tin crystallites with different surface morphologies and particle sizes. These exhibited widely different performances in a lithium cell having the composition: (Cu)Sn I LiCl buffered RTMS I Li(foi1) assembled using MEICl-AlC13 melts as electrolyte. It was found that the tin film crystallites deposited at a temperature of 50C or higher gave a desirable crystal structure and offered the best electrochemical performance as compared to those obtained at lower temperature. The effect of temperature on the deposition of tin on a copper electrode in the MEICl/AlC13/SnC12 melt was investigated using a constant current density of 3.3 mA cm-2at temperatures ranging from 22-70C. Scanning electron microscopy was used to study the particle size and surface morphologies of the deposited tin film and the results are shown in Fig. 1.

0 0 0 0 0 0' 0 0 0 0 0 0 0 0 0 0 0 0 0

0 0

0
0

0
0 0 0 0 0
0

a
0

a
0 0 0
0

Fig. 1. SEM micrographs of a tin film deposited at different temperatures at constant current density of 3.3 mA cm-2[8].

At low temperatures (22 and 35"C), dendritic and large crystallites were found. As the temperature increased, the deposited tin crystals became more regular. The grain size of the crystallites was found to decrease at higher temperatures, with a grain size of 4 pm at 50C and 2 pm at 70C. The results of cyclic voltammetry and galvanostatic cycling reflected the deposition of three major Li/Sn alloy phases corresponding to the two phase transition for LiSdLi7Sn3, Lil3Sn~Li7Snz Li7Snz/Li22Sn~ and among the six intermediate phases identified in the Li/Sn system. Both the charging and discharging capacities were found to increase with the deposition of Li-Sn alloys with higher lithium activities. However, the irreversible capacity at the first cycle was also increased, indicating that more lithium was needed for the initial development of the tin film electrode when cycling at higher voltages. To study the effect of current density, the electrode was cycled at constant current densities from 0.4 to 1.0 mA cm-2.The change in the discharging capacity on cycling at different current densities is shown in Fig. 2. The use of a current density higher than 0.4 mA cm-2led to a rapid degradation of the discharging capacity upon cycling. Moreover, the tin crystallites were found to expand and become irregular and loose when charging at a high current density of 1.O mA cm-2for 20 cycles as shown by the SEM micrographs of the spent electrode in Fig. 3. However, when cycling at a low current density of 0.4 mA cm-2,no obvious change of the morphology of the tin crystallites were obtained after 20 cycles, giving a more intact and regular structure. It was postulated that these differences in crystal structures produced the large differences observed in the rate of capacity loss when cycling at high current densities (Fig. 2). Thus the charging current density should be kept below 0.4 mA cm-2.

a
0 0 0
0

e
a
0
0

0
0

a
0 0
7

0
250

rn

0
0

0
0

rn rn
H

0
I

0 0 0

50

100
Cycle number

150

200

Fig. 2. Discharging capacity for tin film electrode cycling at different current densities [8].

a
0

0
0
0

0
0

e
0

0
0 0
0 0 0

0
0 0
0
Fig. 3. SEM micrographs showing the morphology of tin films after 20 cycles of charging-discharging at different current densities: top, 0.4 and bottom, 1.O mA cm-2 [8].
,

e e
0

0
0

0 0 0
0

In conclusion, it was claimed that the tin film electrode offered a better cycling performance at low current densities compared with other metal alloy electrodes such as the lithiudaluminium electrode. This improvement was attributed to the special texture of the tin deposited film that allowed sufficient space between the tin crystals for the incorporation of lithium without a significant volume change during cycling. 2.1.5 2.1.5.1 Tin-Copper Alloys Pulse Plating
,

0
0
0

0
0

0 0
0

0
0

Beattie et al. [9] have discussed the anomalous, high-voltage irreversible capacity in tin electrodes for lithium batteries. This is observed in electrodeposited and sputtered Sn films when used as working electrodes in electrochemical lithium cells. Tin electrodes can fbnction as expected in Li/Sn cells provided the upper charging cut-off voltage is kept below about 1.4V. However, when this cut-off voltage is increased to about 1.5V and above, an anomalous irreversible capacity loss is experienced. Beattie et al. showed that the amount of irreversible anomalous high voltage irreversible capacity depends on the discharge rate. A model involving catalytic decomposition of the electrolyte (1M LiPF6 dissolved in ethylene carbonate and diethyl carbonate at a ratio of 3357 by volume) by pure tin crystallites was proposed. The authors suggested that there were two ways of avoiding the capacity loss. Firstly, to rapidly discharge the cell to 0.8V, then limit the recharge voltage to below -1.W. Secondly, to alloy some other metal with the tin so that pure tin crystallites were not present. Alloying the tin with copper was suggested as a good remedy. The background to he use of Cu-Sn alloys in lithium-ion cells has been discussed by Beattie and Dahn [ 101. Cu-Sn alloys have been proposed as possible anode materials. Kepler et al. [ 111 showed that Cu6Sn5 can react reversibly with lithium for a few tens of cycles. Larcher et al. [ 121 studied the Li-Cu& reaction using in situ X-ray diffraction. Tamura et al. [ 131 showed that Cu-Sn alloys could be prepared by electrodepositing pure tin on copper foil, followed by an annealing step. Beattie et al. [ 141 showed that Cu6Sn5 could be directly prepared from a single bath containing both dissolved Cu2+and Sn2+ ions. These films were shown to have approximately the same electrochemical behaviour as the powdered samples prepared by Larcher et al. The electrodeposited films have the advantage that they can be directly utilised as electrodes in lithium-ion batteries without further processing. There are a number of binary intennetallic Cu-Sn phases [ 151 and the electrochemical properties of each should be considered. Beattie and Dahn have recently prepared a range of Cu-Sn alloys by electroplating on nickel foil [ 101. Based on previous work [ 141 it was believed that a method of electrodeposition fi-om a single bath usingpulsed deposition in a Hull cell could be used. This method was similar to those discussed previously [ 16,171. However, Beattie and Dahn claimed that their method was unique in that a low frequency odoff pulsed waveform and the geometry of the Hull cell were used to exploit the exchange of deposited Sn by aqueous Cu to achieve a composition and structure spread as a function of position on the substrate. Cu-Sn alloys were deposited from a pyrophosphate solution having the composition: Sn2P207 (36 gh); K2P207 (135 dl); Cu2P207.3H20 (1 gA). Without the presence of

a
0

0
0

0
0

0
0

0
0
0
0

0 0 0
9

additives, a pyrophosphate bath has poor throwing power [ 181. A solution with good throwing power deposits a constant thickness film regardless of macroscopic cathode irregularities. A solution with poor throwing power was needed in this work to obtain a varying thickness along the cathode in the Hull cell. The deposition was performed in a cell 7.5 cm wide made of PVC. The cathode was inclined at an angle of 5 1So with respect to the anode, which was 7.5 cm wide. The closest approach of anode to cathode was lcm and the furthest separation distance was 9 cm. The cathode was 1lcm in length and the cell was 6.9 cm tall. The cell had a volume of -25Oml. In order to prepare the composition-spread library of Cu-Sn alloys, several requirements needed to be satisfied. Firstly, copper deposits at more positive potentials than tin, so the tin concentration in solution should be much larger than copper. Secondly, the bath should have a poor throwing power so that the current density varies with the anode andor cathode distance in the Hull cell. This ensures that the thickness of the deposit varies. Finally, pulsed deposition was used to create the compositional spread [ 141. A series of on-off pulses were used, primarily because Sn is deposited during the on pulse and Cu from the plating solution by the process of immersion plating (or ion exchange) onto the plated tin during the off pulse according to the reaction:

0 0
0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0

This produced a deposit like that shown in Fig. 4. Fig. 4(a) shows the layers of tin rich material deposited during the on pulses, whose thickness varied along the length of the substrate electrode, and the layers of immersion plated copper deposited during the off pulses, whose thickness was constant along the length of the substrate electrode. In the depositions, the on pulse/off pulse sequence was repeated hundreds of times to deposit a film several microns thick. During the on pulse, the layer thickness deposited was estimated to range between 4.7 and 1.3 nm along the length of the substrate. Examinations of the film using small angle X-ray scattering showed that a composition modulated superlattice structure had not been formed. It was assumed this was because the copper-rich and tin-rich layers had been eliminated by interdiffusion. Therefore, Fig.4(b) shows a schematic of the deposited film after the Cu-Sn interdifhsion process.

0
Fig. 4. (a) A schematic of film structure expected using odoff pulsed deposition in a hull cell with a Sn pyrophosphate solution containing a small amount of Cu pyrophosphate in the absence of solid-state interdifhsion, (b) the actual film stoichiometry due to interdiffusion
[101.

10

0 0 0 0 0 0

0 0 0 0
0
Fig. 5 shows that the procedure deposits a film where the average stoichiometry vanes along the length of the substrate. Each data point represents a 40 x 40 pm area.
4
0 *= 3 2 .c)

20

10

0 0 0 0
0 0

25 c
v)

t - ,

E 2
C

33 .o
50

5'
C l

v)

5 2

.o .U

I
'

v ) I

3 1

1
0 0

4 6 8 Position (crn)

a
0

Fig. 5 . Cu:Sn atomic ratio and atomic % Sn as determined by electron probe microanalysis versusposition on the electrodeposited film. The various symbols correspond to six separate scan lines measured along the film. An image of the deposited film is shown on the right [ 101.

0
0 0
0

In that area, many crystallites were sampled and did not necessarily have the same composition. Fig. 6 shows an electron micrograph of the surface of the film with squares indicating the approximate size of the probe beam.

a
0
0
0

0
0

a a
0 0 0 0

Fig. 6 . An SEM image of the deposited film. The size of the 40 pm electron probe beam is indicated by the squares. Different numbers of large crystallites are sampled fiom place to place along the film, causing scatter in the electron microprobe data of Fig. 5 [ 101. For the three indicated positions of the squares there were different numbers of crystallites present. Large crystallites were rich in tin, so if the beam was analysing an area with a large number of crystallites, a large tin concentration would be expected. Alternatively, if the beam was analysing an area with few crystallites, the tin concentration would be expected to be low. Between adjacent points, the number of crystallites being analysed could vary. This variation resulted in a significant difference in the Cu:Sn ratio between adjacent points, causing scatter in the data as

0
0 0
11

seen in Fig. 5. Nevertheless, it was clear that the average stoichiometry inferred from Fig. 5 varied in an approximately linear manner. Using the electron microprobe results shown in Fig. 5 , the atomic percent Sn could be determined along the length of the film, (right hand ordinate). The position of the film in the phase space could then be mapped onto the Cu-Sn phase diagram and this is shown in Fig. 7. The arrow in this figure corresponds with the range of compositions I sampled by the deposited film.
1100

loo0 900 800

700 600
500

400

300 200

I,

40

A 1

I i A

Atomic Percent Sn
Fig. 7. The Cu:Sn phase diagram. The arrow indicates the region of phase space sampled by the film (Fig. 5) deposited in a Hull cell using odoff pulsed deposition [ 101. Fig. 8 shows XRD patterns along the length of the film starting from the bottom (that is, the part of the film closest to the anode during electrodeposition and therefore rich in tin) up to the top of the film (that is, the part of the filmfurthest from the anode and therefore rich in copper).
h)

VI 0

t
9

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

VI

z
w

Fig. 8. 20 XRD patterns measured at equal spacings along the length of the film. The Cu content increases from bottom to top. Sn rich phases are observed at the bottom of the film, while tin rich phases are observed at the top. Peaks fiom known phases are indicated [101.

0 0 0 0

12

As the patterns progress up the film towards the Cu rich end, the XRD pattern from Sn steadily decreased and disappeared approximately a third of the way up. Approximately two thirds of the way up the film the Cu6Sn~ phase also disappeared. Correspondingly, the XRD patterns for Cu-rich alloys began to emerge. However, it can be seen that a smooth transition in the relative abundance of the alloys was observed as a function of distance up the film. These results agreed with those in the phase diagram of Fig. 7 except that the high temperature Cuq1Snl1 phase was observed in place of Cu3Sn in the phase diagram. Cu-Sn fl samples, punched fi-om circular sections of the film at various locations, im were tested in Li/Cu-Sn cells for their specific capacity retention. Standard 23 mm diameter 'coin' cell hardware was employed. The cells used a polypropylene microporous separator; 1M LiPF6 dissolved in ethylene carbonate:diethyl carbonate electrolyte (EC:DEC, 33:67 by volume); and lithium metal as the negative electrode. All cells were assembled in an argon-filled glove box and tested using constant charge and discharge current densities of 30 mA g-'. Fig. 9 illustrates the speclJic capacity of the cells vs. cycle number and the voltage vs. cumulative capacity for five electrodes taken along the length of the film.

a
Q

.E
0

0.8

>
Q)

n Q 0
E 0
0 .-

0.4
0

= 8
Q

cn

0.8
0.4
0 0.8

0.4
n

"0 5 10 15 20 0

Cycle #

4000 6000 8060 Cumulative Capacity (mAh/g)


2000

Fig.9. Voltage vs. cumulative capacity and capacity vs. cycle number for five Li/Cu-Sn cells with electrodes taken along the length of the film. The copper content increases from bottom to top. The specific capacity decreases with increasing Cu content with corresponding increase in capacity retention. (Solid symbols correspond to discharge and open symbols to recharge capacities) [ lo].
13

Figs. 9(a) and (b) correspond to an electrode taken from the Sn-rich end of the film and Figs. 9(i) and 0') correspond to an electrode taken from the Cu-rich end of the film. The results demonstrated that the Sn-rich regions exhibited a high capacity (600 mAh g-'), but poor capacity retention. Conversely, the Cu-rich regions exhibited a lower capacity (- 300 mA g- l), but with a dramatically improved capacity retention. These results were in agreement with earlier work [ 11,12,14]. 2.1.5.2 Electrodeposition followed by Annealing

In two recent papers, Tamura et al. [7, 191 have described a technique for fabricating Sn and Sn-Cu thin film electrodes utilising electrodeposition followed by annealing.
The electrodeposition process enhances the interface strength and electronic conductivity between the active metals and the copper current collector (substrate). It also leads to a stable reaction between the active material and lithium. In the h l l charge-discharge condition, charging and discharging between 0 and 2.0V (vs. Li/Li+), the first discharge capacity was 940 mAh g-', which was 2.5 times higher than that of graphite. The coulombic efficiency in the first cycle was 93%, but the cycling performance was poor. It was considered that this resulted from a lack of interface strength between the entire part of the active material and the current substrate, in spite of the formation of a small amount of Cu&S due to the electrodeposition.

0 0 0 0 0 0 0 0

e
0 0 0 0 0 6 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

In order to enhance the interface strength between the active material and copper, Tamura et al. investigated an anode fabricated by annealing the as-deposited anode. In the full charge-discharge condition, the first charge-discharge characteristics were almost equivalent to those of the as-deposited anode and the retention capacity ratio after 10 cycles was improved from 10 to 94%. This annealing induced the formation of two different Sn-Cu compound layers between the tin layer and the copper current collector. It was concluded that introducing copper in the tin phase and its concentration gradient enhanced the interface strength between the active material and the current collector and thus the cycling performance was improved.
Fig. 10 shows the preparation sequence for the thin film electrode. An electrodeposited tin layer on electrolytic copper foil was fabricated. After acid treatment, the tin layer was electrodeposited at a current density of 1A cm'* from a bath consisting of: SnS04 (40 g/l); H2S04, 98% (150 g/l; formalin (5 ml/l); and an addition agent (40 ml/l). The thickness of the tin layer was 2 pm. The size of the electrodeposited tin anode for the electrochemical charge discharge test was 20mm x 20mm. The colour of the active material was 'grey'.

In order to investigate the effects of heat treatment, an electrodeposited tin anode was annealed for 24h in vacuum. The colour of the active material became 'blackish' from this treatment.
Fig. 11 shows the X-ray diffraction patterns of the as-deposited and annealed samples. A CusSnS phase was observed in addition to tin and copper in the as-deposited anode. This result indicated that that the electrodeposition process without heat treatment formed a slight amount of Sn-Cu interrnetallic compound at the interface c o p p e r h .

14

Cu foil

(18pm)

U
a

Acid treatment

on the surface

Electrodeposition of Sn from SnS04 aqueous

solution

Fig. 10. Schematic flow for the preparation of electrodeposited tin anode with and without heat treatment [7].

ACu OSn . 0 Cu6Sn,

b,
3 .E

0 Sn

A Cu
0

Cu,Sn,

A Cu3Sn

9 3.U

2
C

10

20

30

40

50

60

70

80

90

2theta (degree)

Fig. 11. XRD patterns of (a) as-deposited and (b) annealed anodes [7].

15

Fig. 11(b) shows that there was a Sn-Cu alloy phase like Cu3Sn in addition to tin, Cu6Sn5 and copper in the annealed anode. Fig. 12 shows a schematic of the structures of both types of anode.
I

Sn

.-.
\

'.,

/i

..

0 0 0 0 0 0 0 0
0 0 0 0
3

Fig. 12. Schematic models of the microstructures of (a) as-deposited and (b) annealed anodes. The thickness of both Cu6Sn5 the as-deposited anode and surface Sn on the annealed anode in are < 0.5 pm [7].

In order to study the charge-discharge characteristics and cycling performance of the electrodeposited heat treated tin anode, a small lithium cell was constructed using this electrode as the anode (thickness 2 pm) The cell consisted of one stack of LiCoOz/separator/annealed anode. The size of the positive electrode was 20 x 20 mm and that of the negative 25 x 25 mm. The electrolyte was ethylene carbonate + diethyl carbonate (1 :1) containing 1M LiPF6. These were set in an aluminium laminated package, whose size was 35 x 50 x 0.4 mm. This small cell was charged for 6 mAh at a constant current density of 1.2 mA and then discharged to 2.25V at the same current density. The charge-discharge curve of this cell in the first cycle is shown in Fig. 13. The average discharge voltage was 3.27V, which was slightly lower than that of a cell using graphite as the negative electrode.

0 0
(3

3.0 -

2.5

0 0 0 0 0

Capacity (mAh)

Fig. 13. Initial charge-discharge curves for a small cell using the annealed anode as a negative electrode. The cell was charged for 6 mAh at 1.2 mA and discharged to 2.25V at 1.2 mA [7].' The cell exhibited a good 'flatness' in its voltage capacity profile and the cycling performance was good in the first 20 cycles but then gradually deteriorated, as shown in Fig. 14.

16

0 0 0 0 0 0 0

0
0 0 0 0
1 00

80

.d c1

.2
.Y

60
40

g
c

$
20
0

0 0
0 0 0

10

15

20

25

30

Cycle number

Fig. 14. Cycling performance for a small cell using the annealed anode as a negative electrode. The cell was charged for 6 mAh at 1.2 mA and discharged to 2.25V at 1.2 mA [ 7 ] . In later work [ 191, Tamura et al. have continued to perfect their technique for preparing lithium anodes by examining the role of the substrate roughness in determining electrode performance. The active material was deposited on copper foil that was either described as flat or rough. Annealed and non-annealed electrodes were also considered, yielding a total of four different experimental electrode systems that could be compared for their electrochemical performance. The effect of the surface roughness on the structure of the active material layer was studied after the initial cycle in a test half-cell where lithium was used both as the counter and reference electrode. The test cell was fully charged to O vs. Li/Li+ in V three successive steps commencing at 0.25 mA cm-2,followed by 0.13 and 0.05 mA cm-2.The cell was then fully discharged to 2.0V vs. Li/Li in the same steps. Fig. 15 shows a plan and side view of the annealed electrode produced with the rough surface copper foil after the first cycle. From the plan view, the active material was observed to divide into islands by cracks. From the side view, each island appeared like a column. The cross-section dimensions of these microcolumns/islands were about 10 pm2.

0
0

0
0

e
0 0 0

0 0
0

0
0

0 0 0
0
0
10pm Fig. 15. SEM images of (a) plan view and (b) side view of the active material of an annealed electrode after the initial discharge to 2.0V vs. Li/Li (~2000) [19]. Fig. 16 shows backscattered electron images of cross-sections of the annealed electrodes with rough and flat surface foils after the first cycle. For the rough surface foil, it was observed that the active material cracked in accordance with the surface profile of the foil and formed the microcolumnar structure. For the flat surface

H 1Opm

0
0

0 0 0

17

profile, by contrast, the active material did not form this structure and was likely to be delaminated from the foil.

0 0 0 0 0 0 0

Fig. 16. SEM cross-section images (backscattered electron images) of the annealed electrodes after the initial discharge discharge to 2.0V vs. Li/Li+), (a) with rough surface copper foil ) substrate and (b) with flat surface copper foil substrate ( ~ 2 5 0 0 [19]. Fig. 17 illustrates the cycle performance differences for the four experimental situations studied. The annealed electrodes exhibited a better cycle performance than those non-annealed as demonstrated in the previous work [7]. In particular, the annealed electrode with the rough surface profile retained 9 1% of its initial capacity for 10 cycles under the full charge and discharge conditions and its average coulombic efficiency was 98% between the second and tenth cycle.

1000 800

.
.C

0
0 0

600

400

0 0 0 0 0 0 0 0 0 0

Fig. 17. Cycling performance of the electrodeposited electrodes at 25C. Electrode potential range 0-2.OV vs. Li/Li. (a) non-annealed with flat surface foil, (b) non-annealed with rough surface foil, (c) annealed with flat surface foil, and (d) annealed with rough surface foil [ 191. Fig. 18 shows schematic models of the structures of an annealed electrode with flat and rough surface foils. It was postulated that the effect of the surface roughness of the foil on improving cyclability resulted from the formation of the microcolumnar structure. The active material on the rough surface foil cracks periodically in accordance with the surface profile of the foil, in this case every 10 pm, forming the stable microcolumnar structure. This provides a moderate size for the columns and moderate spaces for the active material to swell in all directions during charge (lithiation) so that it can adhere more strongly to the foil under the much reduced stress and strain caused by its volume change.
18

0 0 6 0 0 0 0 0 0 0

0
0

0
0

- Cu foil + t '

0
0
0

Rough surface foil Cracked in accordance with surface profile


Enough space to swell in either upper or side directions

Flat surface foil

Cracked randomly
Too little space to swell sidewards

-c

Strongly stuck to the foil

Easily delaminated from the foil

0
0 0

Fig. 18. Schematic model of the annealed electrode structure with 'rough' and 'flat' surface copper foil [ 191. 2.1.6 Tin-Nickel Alloys

0 0 0 0
0

a
0 0 0

The selection of an adequate metallic matrix is the key to the formation of a high performance lithium anode material based on tin. A minimisation of the volume changes, rather than their accommodation, as discussed above, has also been assumed to be important. Elements that are inactive against lithium have been assumed to suppress these volume changes effectively without much irreversible capacity loss. The use of tin compounds with elements such as Cu, Mn and CO are therefore a possibility. Ni is a typical element that does not react with lithium. Therefore, this element can be expected to serve as an appropriate matrix to improve the cyclability of the electrode without a high initial irreversible capacity. Mukaibo et al. [20] has studied Sn-Ni alloy thin-films with various Ni/Sn ratios using an electroplating technique. The electrolyte consisted of: NiC12.6H20; SnC12.2HzO; potassium pyrophosphate (GP207; glycine (H2NCH2COOH); and NH40H. The current density was 5 mA cm-2and the bath temperature was 50 "C. Thin films containing different Ni/Sn ratios were prepared on Cu sheets from the plating baths. The Sn-Ni alloy samples contained 54, 62, 84 and 92 atom % Sn. The deposition time was 5 min for most samples and the various atom ratios were achieved by changing the molar ratios of metal ions (and glycine) in the plating solutions. Electrochemical measurements were performed in conventional glass cells with two pieces of lithium foil as counter and reference electrodes. The working electrolyte was 1M LiC104 in an ethylene carbonate/propylene carbonate mixture (1 :1 by volume). The cyclability of half-cells was evaluated by galvanostatic charge-discharge tests in the potential range 0-3V vs. Li/Li+. Both charge and discharge were camed out at the same current density of 50 mA g-' . Field emission electron microscopy (FESEM) was used to observe the morphology of the samples. Fig. 19 shows images for the 62 atom % Sn sample: as-deposited; after the first charge; and after the first discharge. It can be seen that the as-deposited film consisted of aggregates of small grains with an average size of 95 nm. After the first charge at a constant current, the grain size of the film increased to eight times that of the as-deposited sample (average size -760 nm). With further cycling the grain size

0
0

0 0

0
0 0
0 0

0
0 0

1 9

decreased to an average of 150 nm and the grains were aggregated to form a . complicated three-dimensional porous structure. These changes in the electrode structure in different states of the initial cycle were assumed to be related to the insertion and de-insertion reactions of lithium.

0 0 0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

I urn

lbm
Fig. 19. FESEM images of 62 atom % Sn electrode (a) as-deposited, (b) after the first charge until OV vs. Li/Li', and (c) after the first discharge to 3V vs. LdLi' [20]. The electrochemical cycling tests are shown in Fig. 20. They reveal that the initial drop in the discharge capacity varied with the Sn content of the sample. For samples with 54 and 62 atom % Sn, the drop was less than 100 mAh g-', whereas for those

20

0
0

0
0

0
0 0 0 0 0 0 0
0

with 84 and 92 atom % Sn the drop exceeded 500 mAh g-'. Among these films, the 62 atom % Sn film exhibited the highest capacity of ca. 650 mAh g-' at about the 70th cycle, whereas the other samples exhibited lower values in the region of 300 rnAh g-'.

10

20

30 40 Cycle Number

50

60

70

Fig.20. Cycle performance of electrodes with various NUSn ratios at a constant current density of 50 MA g-' in the potential range 0-3V vs. Li/Li' [20]. Crosnier et al. [21] have investigated small particle size electroplated Ni-Sn alloys as anodes for Li-ion batteries. The electrochemical conditions were modified to change the morphology of the material. Samples of Sn-Ni were synthesised by deposition on a copper foil using a standard electrolytic bath consisting of: NiC12.H20 (250 g/l); SnC12.2H20 (50 gA); NH4F (40 g/l); and N H 3 35% (35 mlll). The mixture was stirred at 65C. The time of the electrodeposition was 1 min with 100 mA cm-2current density, leading. to a 3 pm thick layer. The films were dried at 70 "C. (Films were also deposited at 20 A cm-2,but the experimental conditions were not specified). The electrodes were tested 'as-deposited' . Galvanostatic cycling measurements were performed on electrodes 12 mm in diameter in simple cells using metallic lithium foil as both counter and reference electrodes. The electrolyte was a 1M solution of LiPF6 dissolved in ethylene carbonate/diethyl carbonate (2: 1).
l

0 0 0

0
0
0

0
0 0

0
0

0
0 0

0 0
0 0

The concentrations of Ni observed in the Ni-Sn alloy, as well as the morphology of the electrochemically synthesised material, were strongly influenced by the current density used in the preparation of the films. At low values of the current density (100 mA cm-2),in order to obtain an alloy with a 1:1 ratio of Sn-Ni it was necessary to use an electrolytic solution containing a molar ratio of Sn:Ni of 1:0.2, whereas when a higher fabrication rate was used (20 A cm-2)it was possible to obtain the desired 1:1 Sn-Ni alloy from a solution containing a 1:1 molar ratio. An XRD analysis revealed that only in the case of the alloy prepared at the lower current density was a single phase system developed. This is a rather metastable composition outside the equilibrium range of Ni3Sn2. An EDX analysis on large areas indicated a 52%Sn-48%Ni composition for the low rate deposit. The use of the high current density lead to a multiphase compound. The scanning electron micrographs shown in Fig. 21 indicate that the rate of deposition changed the morphology of the alloy deposit. The material prepared at the higher rate exhibited a much smaller particle size (less than 1 pm) and higher porosity than the material prepared at the lower rate (particle size 5 pm).

0
0
21

0 0

0
0

0
0

0 0 0 0 0 0
0 0
I

0
0 0
Fig. 21. SEM micrographs of Ni-Sn alloy deposited at current densities of (a) 100 mA cm-2 and (b) 20 A cmm2 [21]. The results from the electrochemical cycling tests are depicted in Fig. 22. The specific capacity of the very dense thin film obtained with the low density current was much smaller than the capacity obtained at the high rate, However, this capacity was still smaller than the theoretical one of 682 mAh g-' for Ni 0.48 Sn 0.52. It was postulated that the better cycling behaviour of the high current density material may not only be due to the smaller particle size, but also to the higher porosity of the deposit, which allows for expansion due to the muhi-phase character of the deposit.

0 0 0

0
0

0
0

0
0

0 0
0
Fig. 22. Specific capacity vs. cycle number for Ni-Sn alloys [21].

0 0 0

22

a
0

a
0

2.1.7

Tin-Zinc Alloys

a a
0

In recent work, Wang et al. [22] have used an electroplating method to prepare Sn-Zn alloys on Cu foil and have then carried out a heat-treatment procedure to improve the performance of the alloys as anode materials in lithium ion negative electrodes. The plating bath consisted of: Sn2+(0.18M); Zn2+(0.04M); sulphosuccinic acid (0.5M); pH 3-6; temperature 25C. The current density was 10 mA cm-2.The thickness of the plated layer was 5pm. The concentration of Zn in the plated layer was 10 wt. %. The anode size for charge-discharge testing was 1 cm2. Heat treatment was applied in the range 100-200C under vacuum. XRD patterns of the as-plated and heat treated Sn-1O%Zn alloy are shown in Fig. 23. In the as-plated sample, a slight Cu6Sn~ intermetallic compound phase could be discerned beside the Sn-Zn plated phase and the Cu foil. This result indicated that a slight Cu6Sn5 intermetallic compound was formed and existed on the interface between the Sn-Zn plated film and the Cu foil during electroplating. After heat treatment, the Cu6Sn5 intermetallic compound became a major phase in the plated film of the electrode. However, the Zn diffraction peaks disappeared and no Zn-based compounds were observed in the heat-treated electrode.

a a
0

a
0
0

0
0

0 0 0 0
3 m

a
.-

0
\

a c l c

20000

a a
0 0

loo00

0 20

30

40

50

60

70

80

20/O

Fig. 23. XRD patterns of Sn-lOwt.%Zn alloy samples, (a) as-plated and (b) heat-treated [22]

a
0

a a
0

Fig. 24 gives the SEM-EDX images of a cross-section of as-plated and heat-treated electrodes. In the as-plated electrode, Sn and Zn were evenly mixed together. During heat treatment Cu diffused from the Cu foil to the surface of the electrode and the compound Cu6SnS was formed. In this process, Zn segregated and formed two Zn layers. Based on this result, Fig. 25 gives a schematic model of a cross-section of electrode before and after heat treatment. Test half-cells were constructed of the Sn-Zn alloys on the Cu foil as working electrodes and lithium sheet as counter electrode. The latter was also employed as reference electrode. The electrolyte was 1M LiPF6 in ethylene carbonate/dimethyl carbonate (1:2). The charge (lithium insertion) and discharge (lithium extraction) current density was 0.2 mA cm-2for the initial three cycles of activation, followed by 0.4 mA cm-2 (charge) and 1.0 mA cm-2(discharge) in subsequent cycles. The potential range was 0-1 .OVvs. Li/Li+.
23

a
0
0

0 0 0

I
\

(a)

As-plated

(b)

Heat-treated

Fig. 24. SEM-EDXS spectra of cross-section of electrodes (a) as-plated and (b) heat-treated 1221-

0 0 0 0 0 0 0 0 0 0 0

,'

.
Heat-treated

As-pla ted

Fig. 25. Schematic models of the cross-sections of as-plated and heat-treated Sn-Zn electrodes [221* The charge-discharge curves of an as-plated electrode are shown in Fig. 26. In the lithium insertion process, there was a long plateau around 0.4V and then the potential decreased gradually until OV. In contrast, there was a reversible plateau at about 0.6V in the lithium extraction process. It was considered that the two plateau regions resulted from the reaction Li,Sn (x = 1-2.5).
2.0

r-----l
-1 cycle
-.-.-.

- - - - - 5 cycles
...... .. 10 cycles

1 .o

I\ 50th

200
300
400

- 20 cycles -..-.-. 30 cycles . . _.___-.-_. 50 cycles

1st

0.5

nn -.-

100

500

600

700

800

Capacity / mAh.g"

Fig. 26. Charge-discharge curves for an as-plated electrode Sn-1O%Zn electrode [22].

24

0 0 0 0 0 0 0 0 0 0 0 0 0 0 -0 0 0 0 0

0 0

0
0 0
0

Fig. 27 illustrates the cycle behaviour of an as-plated electrode without heat treatment. At the first 10 cycles, the discharge capacity was above 600 mAh g-', about twice as large as that of graphite. The charge-discharge efficiency was high to 95% at the first 10 cycles.
800
1100

0 0
0
,

0 0

0
0
0 0 0 0

10

20

30

40

50

Cycle number

Fig. 27. Cycle behaviour of as-plated Sn-1O%Zn electrode [22]. Compared to electroplated pure Sn [ 131, the discharge capacity and cyclability of the Sn-Zn alloy was increased considerably. It was postulated that the presence of Zn 'relaxes' the mechanical stress resulting from phase transition and volume expansion and contraction during lithium insertion and extraction. During the subsequent cycles, the discharge capacity decreased significantly to 25 mAh g-' after 50 cycles. This suggested that the plated electrode gradually decayed due to the loss of interface strength between the plated film and the Cu foil. In order to improve the performance, the heat treatment procedure was applied.

0 0
0

Fig. 28 gives the charge-discharge curves for a heat treated Sn-Zn electrode.
20 .

0 0 0 0 0
0 0
00 . 0 200
400

1.5

1 .o

0.5

600

800

1000

1200

Capacity / mAh.g-'

e
0
0

Fig. 28. Charge-discharge curves for a heat-treated electrode Sn-1O%Zn electrode [22].

0 0

25

0
Compared with the as-plated electrode in Fig. 26, the plateau around 0.4V disappeared and emerged at a more negative potential at around 0.2V in the electrode after heat-treatment. This result indicated that the heat-treated electrode had a different charge-discharge reaction mechanism to that of the as-plated electrode. The first charge capacity of 1013 mAh g-' was higher than that of the theoretical capacity for Sn-1O%Zn of 9 10 mAh g-'. This was attributed to both the conversion of a small quantity of Sn02 on the surface of the electrode to metallic Sn and the formation of a 'solid-electrolyte interface' region at the electrolyte/solution interface.
,

Fig. 29 shows the cycle behaviour of the heat-treated electrode. This exhibited a high discharge capacity and enhanced cycle performance. The initial discharge capacity of 725 mAh g-' was higher than that of the as-plated electrode. At the same time the cycle stability increased significantly. The discharge capacity maintained a capacity of 450 mAh g-' for over 50 cycles.
1200

0 0 0 0 0 0 0 0 0

loo0
' ) 800 s

0
0 0 0 0 0 0 0 0 0 0

2 E
0

.c

+ 600

400
-0-

0 Li-extraction
I

200m
I

10

20

30

40

50

Cycle number

Fig. 29. Cycle behaviour of heat-treated Sn-1O%Zn electrode [22].

In conclusion, Wang et al. considered that the good cycle performance of the heattreated electrode could be attributed to three factors:-

(0

The CusSnS phase in the film offered a host matrix that was capable of relaxing the mechananical stress resulting from the phase transition structure and volume expansion during Li insertion and extraction. The segregated Zn layers strengthened the interface between the plated film and the Cu current collector. Voids, present in the plated film to around 30% of the total volume, also contributed to a reduction in the detrimental effects pf volume expansion during the cycling of the heat-treated electrode.

(ii) (iii)

0 0 0

26

0 0

0 0
0 0
2.1 8 Tin-Iron Alloys

0
0 0

Sonoda et al. [23] have very recently described the charge discharge behaviour of electroplated Sn- 12%Fe alloy films on electrolytic copper foils. These exhibited a capacity of 374 mAh g-' after 50 cycles in simple electrochemical tests involving halfcells. A coin type battery with a Sn-l2%Fe alloy plated anode and a LiCoO2 cathode had a discharge capacity of 272 mAh g-' and a coulombic efficiency of 90.4% after 95 cycles in charge-discharge tests at a constant charge capacity of 302 mAh g-'. The authors claimed that Sn-!O%Fe alloy plated anodes might be expected to substitute for carbon anodes for advanced lithium batteries. 2.19 Tin-Antimony Alloys

0
0

0
0 0 0 0
0

With the use of a high speed plating method at currents above the limiting current density, Peled and Ulus [24,25] were able to produce the following nanosize tin-antimony alloys: (i) alloys with low antimony content exhibited a high reversible capacity (up to 700 mAh g-'), a lower irreversible capacity, a better rate capability, and a lower average working potential vs. Li/Lif; (ii) antimony-rich alloys that had longer cycle life, but low rate capability and a high average working potential vs. Li/Li+. The addition of a little Cu to the alloy improved the cycle life and with no penalty in capacity . More recently, Ulus et al. [26] have described the preparation of a composite anode containing nanosize (<loo nm) particles of tin alloy Sn65Sbl8Cu17 and Sn62Sb21Cu17. The alloys were electroplated at high current densities (above the limiting current) from aqueous solutions, directly onto copper current collectors and were coated by a polyvinylidine fluoride-graphite matrix at a ratio of al1oy:graphite matrix 70:30 and 80:20 w/w. Over 40 (100% depth of discharge) cycles were demonstrated in a halfcell and over 30 were demonstrated with a LiCoO;! battery containing 1M LiPF6 ethylene carbonate-diethyl carbonate electrolyte. The faradaic efficiency (Q de-ins/Qins) was less than 100%. Lithium was fully de-inserted fkom the host lattice only when the anode was cycled at low current densities. The kinetics of lithium insertion to and de-insertion from the composite anode material gradually slowed as the cycle number increased. X-ray diffraction patterns of the anode material showed that the alloy became amorphous during cycling, while the graphite did n0t.X-ray photoelectron spectroscopy measurements revealed that the solid electrolyte interphase consisted of mainly LiF, small amounts of Li20 and possibly polymeric substances. The electrochemical behaviour of the alloy changed with cycle number, while that of the graphite did not. The fall of the discharge capacity of the graphite from the first cycle to the 34thby more than 50% proved that the active material in the anode suffered from particle-to-particle 'break off.

e
0 0
0

0
0

e e
a

e
0
0

e
0
27

2.2 Lead-Acid Batteries

2.2.1 Background The lead-acid battery is unique amongst battery systems in that both the positive and negative active materials, together with their respective alloy current collectors or grids, are both derived from the same element, viz. lead. Modem maintenance-free and sealed (or valve regulated) lead-acid batteries have eliminated the use of Pb-Sb alloys in the grid materials because of the need to minimise gassing and <waterloss resulting from the use of Sb (a low hydrogen-overvoltage metal) in the negative alloy composition. The preferred alloy is now based upon the Pb-Ca system, with additions of Sn, for both positive and negative grids. The addition of tin provides for improved corrosion resistance of lead-acid battery grids; enhanced conductivity of the corrosion film generated at the interface between the active materials and the grids; and increased mechanical properties. This requires a tin content in the region of 1-2 wt. % for modem positive grids that are subject to the most corroding conditions. Lower tin contents do not provide adequate mechanical strength, improved corrosion resistance or enhanced conductivity in the corrosion film. Higher tin contents may provide improved corrosion resistance but may result in decreased mechanical properties. A low calcium content in the region of 0.08 wt.%, combined with an adequate tin concentration (and possibly a small amount of silver in the region of 0.05 wt. %), appears to be the best-performing composition for positive grids.
2.2.2

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Electrodeposited Pb-Sn Alloys

In addition to the use of Pb-Ca-Sn alloys in lead-acid battery systems, so called pure lead tin technology is receiving increased interest. This involves the use of positive grid materials composed of Pb-Sn alloys containing very low levels of impurities. Using this technology it is possible to produce thin grids and thereby increase the energy density of the battery.
Petersson and Ahlberg [27] have electroplated Pb-Sn alloys on a glassy carbon substrate with the objective of forming light-weight electrodes for lead-acid batteries. This goal is particularly important for the lead-acid system if it is to compete with other contenders in battery powered vehicles. The deposits were electroplated from a 1M HBF4 solution containing various concentrations of the metals to be plated. The electrochemical properties of the deposits were investigated by cyclic voltammetry in solutions of 5M H2S04. The experimental results showed that the electrodeposited layer contained a supersaturated solid solution phase up to 12 at. % Sn. Oxidation of this phase lead to the formation of a PbO phase with increased conductivity compared to pure PbO. In addition, the amounts of PbO and Pb02 decreased with increasing amounts of tin in the alloy and for high tin alloys, where a bulk tin phase was present, no PbO was observed.

c
0

c
0 0
0 0 0

28

0
0

0
0 0 0

2.3

Rechargeable Alkaline Manganese Batteries

0 0 0

Ghaemi et al. [28] have described the use of ternary Cu-Sn-Zn (55:25:20) alloys in the development of improved rechargeable alkaline manganese batteries. The improvements were obtained through minimising the passivation problems associated with the zinc electrode (anode). A relationship was observed between the tin alloying elements and gas evolution during cycling. The Cu-Sn-Zn alloy (trademark Optalloy) displayed better corrosion resistance and a higher hydrogen overvoltage when used as the anode current collector. This alloy is well known for its hardness, good conductivity and good electrochemical properties. It acts like a noble metal with an open-circuit potential of about 700 mV more positive than some binary tin alloys. A very thin sheet of this alloy, in the order of 2-4 pm, can meet all the usual corrosion test requirements. A typical bath formulation for the alloy consisted of CuCN (3 g l ) ; Na2Sn03 (2 gl); Zn(CN)2 (1.4 gl); KOH (19.8 gl); and &CO3 (23.03 gl). Pure Zn foil substrates were employed. The current density was 0.5 A dm-2.Manufactured foils (9 cm2) were used as anode current collectors in C-size (LR- 14) rechargeable alkaline manganese (RAM) cells. To increase the electrochemical reversibility and electronic conductivity of the anode mass,porous zinc was treated with Optalloy. This was to obtain a modified zinc electrode, which was found to be effective in terms of raising the cycle performance of the bipolar RAM batteries. A porous zinc deposit was first prepared with a current density of around 12 A dm-2.This was obtained using a saturated alkaline zincate solution. The chemically modified zinc was prepared through direct dipping of the porous zinc electrode in Optalloy electrolyte, using a dual-bathtechnique.

e
0 0

0 0 0 0
0 0

a
0
0

3.

Fuel Cells

3.1 .3.1.1

Direct Methanol Fuel Cells Background

0
0

The background to this subject has been discussed by Rahim and Hassan [29]. Considerable interest has been directed to the use of fuel cells in electric vehicle technology. One promising system is the direct methanol fuel cell, since the fuel is best transported and converted into energy from the Ziquid state [30]. Unfortunately, efficient ezectrocatalysts for the reactions of the fuel are still far from satisfactory. To improve both the oxidation rate and electrode stability, considerable efforts have been directed towards the study of binary metallic electrodes for the direct electrochemical oxidation of methanol. Recent advances have shown that platinum based bimetallic or multimetallic catalysts exhibit enhanced efficiency for methanol oxidation [3 1-33]. Alloy electrodes such as Pt-Ru, Pt-Rh, Pt-Rb, Pt-Re and Pt-Sn have been studied [34]. However, Pt-Ru and Pt-Sn are the most active of the binary alloys [35]. By contrast, purely surface adsorbed tin metal was found by Campbell and Parsons [36] to inhibit methanol oxidation.

0 0
0

0
0
0

a
0 0

29

The exact role of the alloying metal with Pt as promoters in the methanol electrooxidation is still not understood. Contradictory assumptions about its role were found in the literature. One of these is that the alloying metal facilitates the adsorption of oxygen containing species such as OH,d, [37,38]. Subsequently, the adsorbed PtOH reacts with organic poison. Another is that the enhancement occurs by preventing formation of a strongly adsorbed poisoning species such as CO either by blocking the sites necessary for its adsorption [35], or by completely oxidising CO in solution to CO2 [39]. It was also assumed that a hybrid homogeneous mechanism accounts for the enhancement [40]. Sn spontaneously adsorbed on Pt as Sn(n) might catalyse the methanol oxidation in acid medium [41]. A pronounced enhancement of:the catalytic activity towards methanol electrooxidation on Pt-Sn oxide was observed [42]. 3.1.2 Research on Tin Modified Platinum Electrodes

. -

Rahim and Hassan [29] have recently shown how Pt electrodes, modified by partial electrodeposited tin, can be used as anodes for the catalytic electrooxidation of methanol in an acid medium. Sn was electrodeposited both galvanostatically and potentiostatically. The technique of cyclic voltammetry was used to study the methanol electrooxidation. The electrochemical measurements were performed on a platinum sheet working electrode (area 1 cm2). The Sn modified Pt electrodes were prepared by partial electrodeposition of tin at a constant current of 1 mA cm-2 from a solution of O.3M SnC12 in 0.5M H2SO4. In the potentiostatic method of electrodeposition, a constant potential of -0.W (vs. Hg/HgS04/1 M H2SO4) was applied. After Sn deposition by either method, the electrode was strongly heated in an 02/natural gas environment with temperatures around 1000C. To activate the surface, the electrode was polarised in a cyclic mode from the H2 evolution potential (-0.8V) to the 0 2 evolution potential (+1.4V) at a scan rate of 100 mV s-l for 50 cycles before each experiment. Cyclic voltammograms for the clean Pt electrodes were traced repeatedly in pure 0.5M H2S04 between H2 and 0 2 evolution potentials to ensure surface cleanliness. The voltammogram was then repeated in the presence of methanol. Fig. 30 shows the results for platinum in the absence of tin.

0.8

0.4

0.0

-0.2

0.0

0.2

0.4

E I V vs MMS

Fig. 30. Electrooxidation of 0.5M methanol on a pure Pt electrode in 0.5M H2S04at a scan rate of 50 mV s-*. Cycle 2 (fdl line) and cycle 50 (dotted line) [29].

30

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

0 0

0
0 0

0
0
0

The anodic oxidation of methanol is represented by two peaks, one in the anodic direction and one with slightly lower current density in the cathodic direction. Both the peak current densities decreased as a result of repetitive cycling. In fact, the peak current density for the 50thcycle was only 38% of that of the 2ndcycle. It is well established that the loss of methanol oxidation efJiciency is due to surface poisoning either by some intermediate products of the methanol oxidation process [43] or by the formation of platinum oxide [36]. The strong adsorption of -CHO type intermediates can lead to the formation of CO which blocks the surface to methanol oxidation.

a
0 0 0 0 0 0 0 0
0

A radical increase in the peak current density was observed on the Sn-rnodzfied Pt electrodes. The results are shown in Fig. 31. Sn was galvanostatically deposited on the Pt surface for 7 min at 1 mA cm-2. It must also be noted that the deposition of Sn was onlypartial. In order to reduce the m o u n t of tin deposited on the electrode surface, only about 2% of the available area was utilised, the remainder being blocked by a paraffin wax treatment prior to the tin plating process. After electrodeposition of Sn, the wax layer was removed by heating.

-0.2

0.0

0.2

0.4

E I V vs MMS

0 0

Fig. 3 1. Electrooxidation of 0.5M methanol on a Sn-modified Pt electrode in 0.5M H2S04at a scan rate of 50 mV s-'. Cycle 2 (full line) and cycle 50 (dotted line) [29]. From the results shown in Figs. 30 and 3 1,the following conclusions were reached:(i) The peak current density for the oxidation of methanol was much higher in the presence of Sn compared to a pure Pt surface The Sn-modified Pt electrodes exhibited a slight loss of efficiency after many repeated cycles (93% of the peak current density was retained after 50 cycles) compared to the pure Pt electrode. The Sn-modified electrode exhibited a slight decrease in overpotential compared with pure Pt.

0
0

0
0 0
(ii)

e
0 0

(iii)

a
0
0

The results suggested that Sn has to be alloyed with Pt for the electrocatalysis of methanol oxidation in acid solution. Electrooxidation of methanol was not observed on pure Sn surfaces. The treatment camed out after the deposition of Sn (removing the wax layer by heating) suggested that Sn probably dissolved to an appreciable extent in the Pt. Sn is known to form active oxygen compounds that assist the fast

31

oxidation of intermediates, e.g. -CHO compounds and CO, to CO2 [35].Fig. 32 shows the relationship between the amount of tin deposited and the methanol oxidation peak current density in the forward scan. It was observed that in both the galvanostatic and potentiostatic methods of Sn deposition, a certain criticaZ ratio of Sn:Pt produced the maximum electrocatalytic effect.

0 0 0

0 0 0 0 0

40

80 Time I s

300

400

2
7
0

0' 0

100

200

300

Time I s

Fig. 32. Variation of the peak current density (forward scan) for methanol electooxidation with the time of Sn deposition on Pt, (a) galvanostatic deposition at 1mA cm-2. (b) potentiostatic deposition at -0.W vs. Hg/Hg2S04[29]. It was shown that the mass of Sn deposited in the galvanostatic experiments was 9.22 x 10" mg cm-2and that in the potentiostatic experiments 26.6 x 10-3mg cm-2.It was concluded that the platinum 'active centres' are 'blocked' if this ratio is exceeded for each of the methods of deposition, respectively. These results agreed qualitatively with those of Wrighton et al. [34], at electrodes modified with polyaniline and Pt/Sn particles, which displayed a variation of the catalytic activity as a fbnction of Pt to Sn ratio. It was shown that the most active catalysts had surface ratios of Pt:Sn between 8:l and 2:l. Sn dissolves in Pt forming Pt$n [3 11. The surface composition of Pt-Sn alloys are almost independent of the bulk concentration. A bulk concentration of 1-5% will produce a surface concentration of about 33% Sn.

0 0 0 0 0' 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 0

32

0 0
0 0
3.1.3
Research on Tin Modified Carbon Supported Platinum Electrodes

0 0
0

Carbon-supported platinum (Pt/C) with an adsorbed layer of underpotential deposited (upd) Sn is a much better catalyst for methanol oxidation than a carbon-supported platinum-tin (PtSdC) aZZoy [44]. Mukerjee and McBreen have used in-situ absortion spectroscopy to determine differences in the effects that the two methods of Sn addition have on the electronic properties and the structural properties of the catalyst. Alloying of Pt with Sn in a 3: 1 ratio resulted in the formation of a Pt3Sn L12 phase, a decrease in the Pt d-band vacancies, an increase in the lattice parameters and hence larger Pt-Pt bond distances. (These changes were opposite to the effect of alloying Pt with the first row transition metal elements or with Ru, where there are increases in the d-band vacancies and contractions in the lattice parameters). The deposition of Sn as upd adatoms on Pt/C had minimal effects on the electronic or structural aspects of PVC. Both upd Sn on Pt/C and Sn surface atoms in the PtSn/C alloy were found to be similar in that both exhibit Sn-oxygen interactions that varied with potential. It was concluded that the Sn atoms in both upd Sn on Pt/C and in PtSdC catalysts are associated with oxygen species and as a result can initiate the oxidation of adsorbed CO and CHO at lower potentials as compared to Pt/C. The main differences in the catalytic activity of PtSn/C and upd Sn on Pt/C for methanol oxidation is due to the differences in the electronic properties of the Pt in these catalysts. The upd Sn does not interfere with the ability of the platinum to adsorb methanol and dissociate C-H bonds, whereas alloying with Sn inhibits the ability of Pt to carry out these functions. This provides important pointers for explaining the different activities exhibited by alloyed PtSn/C and upd Sn on Pt/C for methanol oxidation. Santos and Bulhoes [45] have very recently discussed the underpotential deposition of Sn on Pt in acid media and have provided a detailed bibliography.

0 0

a
0
0 0 0

a
0
0

e
0

0
0
4.

Conclusions and Recommendations

0
0

0
0 0 0 0 0
0

The application of electrodeposited tin in the fabrication of improved electrodes for both batteries and fuel cells has to received considerable interest during the past few years. These electrochemical power sources will continue to be very important and the developments have been described in some detail in the current review. The following is a list of general observations considered by the reviewer to have particular merit :-

A)

a
0 0

In the field of lithium-ion batteries, the use of electrodeposited tin, in place of the more conventional carbon substrates in the anode structure, has allowed the energy density of these batteries to be increased as a direct result of an increase in the capacity density of the negative electrode. Tin deposited on a copper or nickel substrate has been found to be particularly effective and further improvements have accrued from the electrodeposition of tin alloys. These include Sn-Cu, Sn-Ni, Sn-Zn, Sn-Fe and Sn-Sb. Additional improvements in the performance of these electrodes can be accomplished in some cases by post-deposition annealing processes. These

33

0
fabrication routes can generate preferred alloy phases, which in turn produce an active anode material better equipped to accommodate the volume changes that occur during continuous charge/discharge (Li insertiodde-insertion) cycling.

0 0
0 0

B)

Despite an appreciable literature dealing with electrode fabrication techniques using different alloy systems, there have been no comparative studies that would help to identify any specific alloy having a superior performance. Electrode performance has simply been optimised for a particular alloy, anode design and processing route. Furthermore, the plating methods and plating solutions are extremely diverse, leading to the belief that much of the research work continues to be somewhat empirical in origin. Similar conclusions to these were reached in the earlier reviews concerning the electrodeposition of pure tin and the electrodeposition of corrosion resistant tin alloys. However, despite these conclusions, a substantial amount of progress has been made, particularly when the large number of experimental variables are recognised.

0 0

0
0 0 0

a
0 0 0

In the field of lead-acid batteries, the addition of tin provides for improved corrosion resistance of lead-acid battery grids; enhanced conductivity of the corrosion film generated at the interface between the active materials and the grids; and increased mechanical properties. In addition to the use of Pb-Ca-Sn alloys in lead-acid battery systems, so called pure lead tin technology is receiving increased interest. This involves the use of positive grid materials composed of Pb-Sn alloys containing very low levels of impurities. Using this technology it is possible to produce thin grids and thereby increase the energy density of the battery. Electroplated Pb-Sn alloys have recently been electrodeposited on a glassy carbon substrate with the objective of forming light-weight electrodes for lead-acid batteries. This goal is particularly important for the lead-acid system if it is to compete with other contenders in battery powered vehicles. The experimental results showed that the electrodeposited layer containing a supersaturated solid solution phase up to 12 at. % Sn could act as an effective positive electrode. (D)
Ternary Cu-Sn-Zn (55:25:20) alloys have been employed in the development of improved rechargeable alkaline manganese batteries. The improvements were obtained through minimising the passivation problems associated with the zinc electrode (anode). A relationship has been observed between the tin alloying elements and gas evolution during cycling. The Cu-Sn-Zn alloy (trademark Optalloy) displayed better corrosion resistance and a higher hydrogen overvoltage when used as the anode current collector. This alloy is well known for its hardness, good conductivity and good electrochemical properties. It acts like a noble metal with an open-circuit potential of about 700 mV more positive than some binary tin alloys. A very thin sheet of this alloy, in the order of 2-4 pm, can meet all the usual corrosion test requirements. To increase the electrochemical reversibility and electronic conductivity of the anode mass, porous zinc was treated with electroplated Optalloy

0
0

e
0
0

0
0

0
0

a
0
0

e
34

a
0

0 0
0
(E)

0
0

0 0 0 0 0 0
0 0 0
(F)

Considerable interest has been directed to the use offuel cells in electric vehicle technology. One promising system is the direct methanol fuel cell, since the fuel is best transported and converted into energy from the liquid state. Unfortunately, efficient electrocatalysts for the reactions of the fuel are still far from satisfactory. To improve both the oxidation rate and electrode stability, considerable efforts have been directed towards the study of binary metallic electrodes for the direct electrochemical oxidation of methanol. The use of Pt electrodes, modified by partially electrodeposited tin, have recently been shown to be very effective anodes for the catalytic electrooxidation of methanol in an acid medium. It was shown that the most active catalysts had surface ratios of Pt:Sn between 8:l and 2: 1. Sn is known to form active oxygen compounds that assist the fast oxidation of intermediates, e.g. -CHO compounds and CO, to CO*. Carbon-supported platinum (Pt/C) with an adsorbed layer of underpotential deposited (upd) Sn is a much better catalyst for methanol oxidation than a carbon-support ed plat hum-tin (Pt Sn/C) alloy . Fundamenta1 studies have recently shown that the main differences in the catalytic activity of PtSn/C and upd Sn on Pt/C for methanol oxidation is due to the differences in the electronic properties of the Pt in these catalysts. The upd Sn does not interfere with the ability of the platinum to adsorb methanol and dissociate C-H bonds, whereas alloying with Sn inhibits the ability of Pt to carry out these functions. This provides important pointers for explaining the different activities exhibited by alloyed PtSn/C and upd Sn on Pt/C for methanol oxidation.

e
0 0 0
0 0 0

0
0 0 0 0
0 0

0
0

0
0
0
35

5 REFERENCES .
1. Processes and Applications for Tin and Tin-Based Alloy Surface Coating Technologies, Stage 1,Part1 ,Electrodeposition and Electroplating of Tin. Compiled for Tin Technology by L. M. Baugh, January, 2005. 2. Processes and Applications for Tin and Tin-Based Alloy Surface Coating Technologies, Stage 1,Part 2, Electrodeposition and Electroplating of Tin Alloys (Corrosion Resistant Coatings) . Compiled for Tin Technology by L M. Baugh, 2005. 3. L. Dominey, in: D. Aurbach (Ed.), Non-aqueous Electrochemistry, Marcel Dekker, New York, 1999, (Chapter 8). 4. E. Peled, in: J.P. Gabano (Ed.), Lithium Batteries, Academic Press, New York, 1983, (Chapter 3).

5. D. Aurbach, B. Makovsky, I. Weissman, et al., Electrochim Acta, 45 (1999) 67.


6 Y. Nishi, in: M. Wakihara and 0. Yamamoto (Eds.), Li-ion Batteries, WileyNCH, Weinheim, 1998, (Chapter 18). 7. Sudy of the Anode Behaviour of Sn and Sn-Cu Alloy Tin-Film Electrodes. N. Tamura, R. Ohshita, M. Fujimoto, S. Fujitani, M. Kamino and I. Yonezu, J. Power Sources, 107, (202) pp 48-55.

8. Electrodeposited Tin Coating as Negative Electrode Material f o r Lithium-Ion Battery in Room Temperature Molten Salt. Y. S. Fung and D. R. Zhu, J. Electrochem. Soc. 149, no. 3, (2002) pp A319-324.
9. Anomalous, High Voltage Irreversible Capacity in Tin Electrodes f o r Lithium Batteries. S. D. Beattie, T. Hatchard, A. Bonakdarpour, K. C. Hewitt and J. R. Dahn, J. Electrochem. Soc. 150, no. 6, (2003) pp A701-705. 10. Single Bath Electrodeposition of a Combinatorial Library of Binary CUI,Sn,Alloys. S. D. Beattie and J. R. Dahn, J. Electrochem. Soc. 150, no. 7, (2003) pp C457-460. 11. K. D. Kepler, J. T. Vaughey and M. M. Thackeray, J. Power Sources, 81-82 (1999) 383. 12. D. Larcher, L. Y. Beaulieu, D. D. MacNeil and J. R. Dahn, J. Electrochem. Soc. 147 (2000) 1658. 13. Study of the Anode Behaviour of Sn and Sn-Cu Thin Film Electrodes. N. Tamura, R. Ohshita, M. Fujimoto, S. Fujitani, M. Kamino and I. Yonezu, J. Power Sources 107 (2002) pp 48-55.

36

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

a
0

0 0 0 0

14. Single Bath Pulsed Electrodeposition o Copper-Tin Alloy Negative Electrodes for f Lithium-Ion Batteries. S. D. Beattie and J. R. Dahn, J. Electrochem. Soc. 150, no. 7, (2003) pp A894-898. 15. M. Hansen, Constitution o Binary Alloys, McGraw-Hill Book Company, New f York, 1958. 16. D. Landolt, Plat. Surf. Finish., 88 (2001) 70. 17. S. Roy, M. Matlosz and D. Landolt, J. Electrochem. Soc., 141 (1994) 1509. 18. M. Paunovic and M. Schlesinger, Fundamentals of Electrochemical Deposition, John Wiley and Sons, New York, 1998. 19. Advanced Structures in Electrodeposited Tin Base Negative Eletrodes for Lithium Secondary Batteries. N. Tamura, R. Ohshita, M. Fujimoto, M. Kamino and S. Fujitani, J. Electrochem. Soc. 150, no. 6, (2003) pp A679-683. 20. Electrodeposited Sn-Ni Alloy Film as a High Capacity Anode Material for Lithium-Ion Batteries. H. Mukaibo, T . Sumi, T. Yokoshima, T. Momma and T. Osaka, Electrochemical and Solid-state Letters 6, no. 10, (2003) pp A2 18-220. 21. New Anode System for Lithium-lon Cells. 0. Crosnier, T. Brousse, X. Devaux, P. Fragnaud and D. M. Schleich, J. Power Sources, 94 (2001) pp 169-174. 22. Electroplated Sn-Zn Alloy Electrode for Li Secondary Batteries. L. Wang, S. Kitamura, T. Sonuda, K. Obata, S. Tanase and T. Sakai, J. Electrochem. Soc. 150, no. 10, (2003) pp A1346-1350. 23. Preparation of Advanced Lithium Secondary Batteries with Tin-Iron Alloy Plating Anodes and their Charge-Discharge Behaviour. T. Sonoda, H. Kobayashi, K. Komoto, H. Sakaebe and K. Tatsumi, J. Surf. Finish. Soc. Japan 54, no. 8, (2003) pp 528-532. 24. E. Peled and A. Ulus, Abstract99, The Electrochemical Society Abstracts , Vol. 98-16, p123, Boston, MA, Nov 1-6, 1998. 25. E. Peled and A. Ulus, in New Materials for Batteries and Fuel Cells, Materials Research Society Proceedings, Vol. 575 (1999). 26. Tin Alloy-Graphite Composite Anode for Lithium-Ion Batteries. A. Ulus, Yu. Rosenberg, L. Burnstein and E. Peled, J. Electrochem. Soc. 149, no. 5, (2002) pp A635-643. 27. Oxidation of Electrodeposited Lead-Tin Alloys in %MH2SO4. I. Peterson and E. Ahlberg, J. Power Sources, 91 (2000) 143-149. 28. New Advances on Bipolar Rechargeable Alkaline Manganese Dioxide-Zinc Batteries. M. Ghaemi, R. Amrollahi, F. Ataherian and M. 2. Kassaee, ibid, 117 (2003) 233-241.

e e
0

a a
0 0

a
0 0

a a
0

0 0
0

0 0 0 0 0 0
0

0 0 0

37

29. Platinum-Tin Electrodes for Direct Methanol Fuel Cells. M. A. Abdel Rahim, M. W. Khalil and H. B. Hassan, J. Appl. Electrochem. 30 (2000) pp 1151-1 155. 30. A. Hamnett and G. L. Throughton, Chem. And Ind. 6 (1992) 480. 31. P. N. Ross, Electrochim. Acta 36 (1991) 2053. 32. M.M.P. Janssen and J. Moolhuysen, J. Catal. 46 (1977) 289. 33. J. Clavilier, A. Femandez-Vega, J. M. Feliu and A. Aldaz, J. Electroanal. Chem. 258 (1989) 89,101; 261 (1989) 113. 34. T. C. Hable and M. S. Wrighton, Langmuir 7 (1991) 1305. 35. A. Hamnettand B. J. Kennedy, Electrochim. Acta 33 (1988) 1613. 36. S. A. Campbell and R. Parsons, J. Chem. Soc. Faraday Trans. 88 (1992) 833. 37. Y. B. Vassiliev, V. S. Bagotzky, N. Osetrova and A. A. Mikhailova, J. V. Electroanal. Chem. 209 (1986) 151. 38. M. Shibata and S. Motoo, ibid., 209 (1986) 151. 39. H. A. Gasteiger, N. Markovic and P. N. M. Ross,Jr., Catalysis Letters 36 (1996) 1. 40. K. Wang, H. A. Gasteiger, N.M. Markovic and P. N. Ross, 'Jr., Electrochim. Acta 41 (1996) 2587. Haner andP. N. Ross, Phs. Chern. 95 (1991) 3740. 41. A. N. 42. A. Aramata, I. To-yoshimaand M. Enyo, Electrochim. Acta 37 (1992) 1317. 43. G. Meli, J. M. Leger, G. Lamy and R. Durand, J. Appl. Electrochem. 23 (1993) 197. 44. An In-Situ X-Ray Absorption Spectroscopy Investigation of the Effect of Sn Additions to Carbon-Supported Pt Catalysts. S. Mukerjee and J. McBreen, J. Electrochem. Soc. 146, no. 2, (1999) pp 600-606. 45. The Underpotential Deposition o Sn on Pt in Acid Media. M. C. Santos and L. 0. f S. Bulhoes, Electrochim. Acta 48, (2003) pp 2607-2614.

0 0 0 0 0 0

0 0.

e
0 0
0 0 0

0 0 0 0 0 0 0 0 0 0

38

0 0

6.

Acknowledgements

The following have kindly given their consent to reproduce the figures in this report:Figs. 1-9. Reproduced by permission of The Electrochemical Society, Inc. Figs. 10-14. Reproduced by permission of Elsevier. Figs. 15-20. Reproduced by permission of The Electrochemical Society, Inc.
\

Figs. 2 1 and 22. Reproduced by permission of Elsevier. Figs. 23-29. Reproduced by permission of The Electrochemical Society, Inc. Figs. 30-32. Reproduced by permission of Springer Publications.

39

Potrebbero piacerti anche