Sunteți pe pagina 1din 6

Electrochemical Studies of Aluminum Deposition in Ionic Liquids at Ambient Temperatures

Venkat Kamavaram and Ramana G. Reddy Department of Metallurgical and Materials Engineering The University of Alabama, Tuscaloosa, AL 35487-0202,

Abstract Ionic liquids are low melting salts used as potential non-aqueous electrolytes for the electrodeposition of aluminum. In the present study, an ionic liquid based on a mixture of AlCl3 and 1-methyl3-butylimidazolium chloride was used as an electrolyte for the electrodeposition of aluminum at ambient temperatures. Electrochemical studies were carried out using cyclic voltammetry (CV) and chronoamperometry (CA) techniques. Electrodeposition of aluminum was carried out in the abovementioned ionic liquid on copper as cathode. The process of electrodeposition was found to be quasi-reversible and the kinetic mechanism was identified as instantaneous nucleation followed by diffusion controlled three-dimensional growth. Kinetic parameters such as charge transfer coefficient () and diffusion coefficient (D0) of the species Al2Cl7 were obtained from the CV and CA data, which compared well with those available in the literature for aluminum deposition. Introduction Aluminum electrodeposition in various electrolytes such as nonaqueous and organic solutions has been reported in literature [16]. In recent years, ambient-temperature chloroaluminate molten salts prepared from the combination of AlCl3 in Nbutlypyridinium chloride (BuPyCl) and AlCl3 in 1-methyl-3ethylimidazolium chloride (MEIC) have been used as electrolytes for industrial aluminum electroplating yielding promising results [7,8]. Dense aluminum deposits with high current efficiency have also been obtained by electrolysis in a melt of AlCl3 and 1-butyl-3-methyl imidazolium chloride (BMIC) [9]. 4Al2Cl7 + 3e Al + 7AlCl4 (1)

In this process aluminum deposition proceeds by the reduction of the Al2Cl7 ion given by Eq. (1). AlCl3 and MCl (where M is an organic cation) mixtures having AlCl3: MCl ratios greater than one, called as acidic melts, contain large concentrations of Al2Cl7 ions enabling the forward reaction in Eq. (1) to take place. In melts with AlCl3: MCl molar ratio less than one, termed basic melts, the concentration of Al2Cl7 is extremely small and the dominant chloroaluminate anion, AlCl4, is not reducible within the electrochemical window of the melt [10]. Electrodeposition of aluminum has been extensively studied in many electrolytes, but only a few of these studies are devoted towards the nucleation and growth mechanism of aluminum [10,15]. In most electrodeposition reactions the formation of new metallic phase is fast and uninhibited process [11]. The metal atoms may initially alloy with the substrate or they may be accommodated on a large number of low energy nucleation sites. The formation and growth of an electrodeposited phase is a complex process associated with two or three-dimensional nucleation. The types of growth mechanisms that are likely to occur on a substrate metal, (a) when the substrate is same as the depositing metal [12] and (b) when the substrate is inert metal [11] were reviewed. Thirsk and Harrison [13] have conspicuously related the potentiostatic current-time relationships for the growth of discrete areas before the onset of overlap, in terms of onedimensional needle like growth, two-dimensional cylindrical growth and three-dimensional hemispherical or spherical growth. All these morphologies may occur under two different schemes: (1) the rate controlling process is that of incorporating the metal into the growth points, (2) the rate controlling process is mass transfer mechanism by diffusion or migration of

Light Metals 2002 Edited by Wolfgang Schneider TMS (The Minerals, Metals & Materials Society), 2002

depositing ions to the surface of the growth points. Astley and Harrison [14] have obtained expression for the diffusion controlled three-dimensional growth of hemispherical nucleus. In the present study cyclic voltammetry and chronoamperometry techniques were used to study the nucleation and growth phenomena of aluminum deposition in an ionic liquid. Experimental Procedure The BMIC (1-butyl-3-methylimidazolium chloride) was synthesized and purified as described elsewhere [16]. AlCl3 (anhydrous 99.985%) was purchased from Alfa Aesar. The ionic liquid melt was prepared by adding slowly measured quantities of AlCl3 to BMIC in the ratio 1.5:1, while the mixture is stirred. Care was taken to ensure that the temperature does not rise due to the exothermic reaction. Temperatures above 150C were reached during the preparation. The melt was allowed to cool before it was used for electrochemical studies. Melts were prepared in a glove box under purified nitrogen. All the experiments were carried out at 110 5C. Electrochemical studies were performed in the glove box using an electrochemical cell, which consisted of a glass beaker fitted with Teflon cap having holes for the electrodes. Fig. 1 shows the schematic diagram of the cell. The typical electrochemical cell consisted of three electrodes, working electrode (WE), reference electrode (RE) and counter or auxiliary electrode (CE). The reference electrode and counter electrode were made of high purity aluminum wire obtained from Alfa Aesar with outer diameters ( = 1.5 mm) and ( = 2.0 mm) respectively. The working electrode was a copper wire ( = 1.25 mm) and the active surface area of the electrode was determined from the length of the electrode immersed into the electrolyte. Before use, the working electrode was polished with emery paper (of grit size 4) and then rinsed with distilled water and acetone. The aluminum electrodes were chemically polished by immersing in a solution made up of H2SO4 (98%wt): H3PO4 (85%wt): HNO3 (50%wt) (25:70:5 by volume) for about 10min. The reference electrode consisted of aluminum wire immersed in a Pyrex capillary glass tube filled with the same electrolyte. The reference electrode was kept close to the working electrode so that the solution resistance can be neglected. All electrochemical experiments were performed using an EG&G PARC model 273A Potentiostat / Galvanostat. The cyclic voltammetry and chronoamperometric plots were generated using the EG&G PARC Model 270 software package. Results and Discussion Cyclic Voltammetry The typical voltammogram for the deposition of aluminum on the copper electrode is shown in Fig. 2. Two important features can be seen, namely, the peak in the negative region corresponding to the cathodic reduction reaction and the peak in the positive region corresponding to the anodic oxidation reaction. The anodic peak current exceeds that of the cathodic process. The cathodic peak and anodic peak were attributed to the deposition and dissolution of aluminum, respectively. The asymmetry between the deposition and stripping peaks and their

large separation indicates that the process is quasi-reversible [17].

Figure 1: The experimental set up for the electrochemical studies showing the working, reference and counter electrodes and thermometer inserted into the solution. The value of |Ep-Ep/2| (where Ep and Ep/2 are cathodic peak and half-peak potentials respectively) obtained from the Fig. 2 was considerably larger than those expected for a reversible process involving multiple electron transfer.

E p E p / 2 = 1.857 RT nF

(2)

The charge transfer coefficient can be calculated from Eq. (2) given for an irreversible process [18]. The average value of the charge transfer coefficient is 0.4 indicating that the Al+3 /Al system is quasi-reversible in the melt.

30 20 Current (mA) 10 0 -10 -20 -2 -1 0 1 Voltage (V) vs. Al / Al(III)

Figure 2: Voltammogram obtained at copper electrode in 1.5:1 AlCl3: BMIC mixture. Scan rate: 0.08V/s, Electrode area: 0.68 cm2.

40 30 Current (mA) 20 10 0 -10 -20 -2 -1 0 1 0.35 0.25 0.18 0.13

An equation relating the peak current Ip to the scan rate for a quasi-reversible process [18] assuming linear diffusion is given in Eq. (3).

* 1 I p = 0.4958n 3 / 2 AC0 D0 / 2 1 / 2 F 3 RT

1/ 2

(3)

Voltage (V) vs. Al/Al(III)

As shown in Fig. 4 a linear relation between Ip and 1/2 is typical for a diffusion-controlled process. The diffusion coefficient of the electro-active species, Al2Cl7 calculated from Eq. (3) is D0 = 1.13 x 107 cm2/sec. Where n (=3/4) is the number of electrons involved in rate determining step, F faradays constant, D0 is diffusion coefficient of the electroactive species Al2Cl7, C0* is the bulk concentration in moles/cm3 and A is the area of electrode and other variables have their usual meaning. The positive intercept for the plot Ip vs. 1/2 indicates the deviation from simple linear diffusion controlled process [11]. The deviation from linear diffusion may be due to the adsorption of the products on the electrode or due to the geometry of the electrode. Linear diffusion is applicable to planar electrodes. If the electrode is spherical or cylindrical as, is the case in the present study semi-infinite spherical diffusion could be possible [18]. The voltammograms do not reasonably show the typical characteristics of diffusion-controlled process, i.e., steep increase in cathodic current due to deposition [19]. This may be due to the preceding nucleation process. The effect of nucleation process on the kinetics of deposition reaction is under investigation. Chronoamperometry When a potential step applied to an electrode is large enough to cause an electrochemical reaction, the current changes with time. The study of this current response as a function of time is called chronoamperometry. Chronoamperometry was performed in order to understand the deposition kinetics of aluminum in detail. Collections of current-time transients were obtained as a function of voltage. Typical chronoamperometric current-time transients for the deposition of aluminum in 1.5:1.0 AlCl3: BMIC at different overpotentials are shown in Fig. 5. According to Cottrells equation, Eq. (4) for a simple diffusion controlled process the current response to a potentiostatic pulse should decay with time for all overpotentials. Where I is the current in Ampere, nF is the molar charge, A is the area of the electrode, C0* is the bulk concentration, t is time and D0 is the diffusion coefficient of the electroactive species.

Figure 3: Voltammograms on a copper electrode in 1.5:1 AlCl3: BMIC mixture at 110C. The numbers indicated in figure are scan rates in V/s. Voltammograms were recorded at different scan rates ranging from 0.06 to 0.5 (V/s) using copper as the working electrode. The Figure. 3 shows the plots for low scan rates from 0.13 to 0.35 (V/s). It can be seen that the cathodic peak shifted in the negative direction with increasing scan rate. This negative shift in the cathodic peak potential indicates the irreversible nature of the process. The dependence of cathodic peak potentials (Ep) on scan rate is predicted for an irreversible process as evident from the Fig. 3. The reversibility of the process is quantitatively measured from the value of charge transfer coefficient (). The charge transfer coefficient ( = 0.5) for a completely reversible process and it varies from 0.3 -0.7 for a quasi-reversible

process [18].

0.025 0.020

Ip (mA)

0.015 0.010 0.005 0.2 0.3 0.4 0.5 0.6 0.7

1/2 (mV/sec)1/2
Figure 4: Peak current Ip vs. 1/2 ( is scan rate) for the reduction of aluminum in AlCl3BMIC at 110C. Area of working electrode: 0.68 cm2

* I = nFAD0 C 0 (t ) 1 / 2 1/ 2

(4)

20
-0.5V

zones overlap, the hemispherical mass transfer gives way to linear mass-transfer, perpendicular to an effectively planar surface [21,22]. Accordingly, the current passes through a maximum value and then decreases rapidly to a steady state current as seen in Fig. 5. Cottrells equation can be applied to the decreasing portion of the transients in Fig. 5 by plotting I vs. t1/2. The linear relationship I t1/2 is typical for a diffusion-controlled process under uniform conditions of linear diffusion [23]. The linear relationship seen in Fig. 6 indicates that the electrodeposition of aluminum is diffusion-controlled process. The diffusion coefficient calculated from the slope of straight line in Fig. 6 is D0 = 0.63x 10 7 cm2 /s.

15

Current (mA)

10 5 0

-0.4V -0.3V -0.2V -0.1V

-1

Time (sec)
Figure 5: Potentiostatic currenttime transients for the reduction of Al from the mixture of AlCl3BMIC at the potentials indicated in figure, V vs. Al / Al (III). However, from the initial rising portion of the current-time transients and their marked potential dependence observed in Fig. 5, it can be concluded that the deposition of aluminum from this melt is preceded by a process of nucleation. And the extent and rate of nucleation depend on the magnitude of the applied overpotential. This effect of nucleation on the current-time transients was observed at potentials above 0.4V.

According to Astley and Harrison [14] the current-time relationship for the growth of instantaneous three-dimensional nucleus is given by Eq. (5). The current for the growth of N0 nuclei already present or formed instantaneously is given by Eq. (5).
* 3 I = 8 N 0 zFM 2 C 0 D0 / 2 t 1 / 2 2 1 / 2

(5)

Where M and are the molecular weight and the density of the depositing metal, C0 and D0 are the concentration and diffusion of the electroactive species, t is time and other terms have their usual meaning. Similarly for the progressive nucleation of threedimensional nuclei the current time relation is given as, I t3/2. The rising portions of the current-time transients followed the relation I t1/2, indicating that the deposition process of aluminum under these conditions is one of instantaneous threedimensional nucleation followed by growth controlled due to diffusion of aluminum ions from the melt. The marked dependence of the growth current, in Fig. 5 on the value of overpotential implies that the number of available nucleation sites is potential dependent. A further analysis of the chronoamperometric current-time transients was done to establish the type of nucleation. Comparison of the experimental data with the appropriate dimensionless theoretical equations derived on the basis of theoretical models available in literature [21] indicated the type of nucleation. The expressions given in Eq. (6) and Eq. (7) are the dimensionless current-time relations for diffusion controlled three-dimensional progressive and instantaneous nucleation respectively.

25 20

Current (mA)

15 10 5 0 -5 0 1 2 3

t -1/2 (s -1/2)
Figure 6: Plot of I vs. t 1/2 taken from the decreasing portion of the current-time transients (0.50V) in Fig. 5. Chronoamperograms were collected by stepping the potential from 0V to a deposition potential ranging from 0.1V to 1.0V. The rising portion of the curve indicates a rise in current as the electroactive area increases, either as each independent nucleus grows in size or the number of nuclei also increases [20]. During this stage of growth of the deposit, the nuclei develop hemispherical diffusion zones around themselves and as these

(i im )2 = 1.2254(t tm )1 (1 exp[ 2.3367(t tm )2 ])

(6)

(i im )2 = 1.9542(t tm )1 (1 exp[ 1.2564(t tm )])2

(7)

Three-dimensional nucleation with diffusion controlled growth; a common phenomenon in the electrodeposition of metals was verified by plotting the dimensionless plots of (i/im)2 and t/tm.

Where i, im are current density, maximum current density (A/cm2) and t, tm are time, maximum time (sec) respectively, obtained from the current-time transients in Fig. 5. For confirmation of this nucleation process, the experimental data was compared with those obtained using Eq. (6) and Eq. (7). The Fig.7 shows representative experimental curves for potentials ranging from 0.5 to 0.8V and two theoretical curves corresponding to instantaneous and progressive type of nucleation. It is seen that the experimental curves closely followed the instantaneous type of nucleation. However, at low overpotentials below 0.5V, the dimensionless current-time plots are positioned in between the progressive and instantaneous type of nucleation. Clearly, the aluminum deposition process involves instantaneous nucleation with diffusion-controlled growth of the three-dimensional centers as reported in literature [17].

Conclusions The early stage of electrodeposition of aluminum in ionic liquids is found to be a quasi-reversible process, with the charge transfer coefficient = 0.40. The rate controlling mechanism of deposition process was found to be diffusion controlled process preceded by nucleation. The kinetics of the deposition process is found to be of instantaneous nucleation with diffusion controlled three-dimensional growth mechanism. The diffusion coefficient of the electroactive species Al2Cl7 calculated from cyclic voltammetry and chronoamperometry are comparable with that reported for aluminum deposition in the literature. Acknowledgement The author expresses sincere gratitude for financial support from National Science Foundation (NSF), Center for Green Manufacturing (CGM) and The University of Alabama (UA). References D. E. Couch and A. Brenner, A Hydride Bath for the Electrodeposition of Aluminum, J. Electrochem. Soc, 99 (1952), 234. 2. J. H. Conner and A. Brenner, Electrodeposition of Metals from Organic Solutions, J. Electrochem. Soc, 103 (1956), 657. 3. R. D. Blue and F. C. Mathers, Electrodeposition of Aluminum from Non-Aqueous Solutions, Sixty-fifth general meeting, Asheville, NC, 1934, 339. 4. F. H. Hurley and T. P. Wier, Jr, Electrodeposition of Metals from Fused Quaternary Ammonium Slats, J. Electrochem. Soc, 98 (1951), 203. 5. F. H. Hurley and T. P. Wier, Jr, The Electrodeposition of Aluminum from Nonaqueous solution at Room Temperature, J. Electrochem. Soc, 98 (1951), 207. 6. I. A. Menzies and D. B. Salt, The Electrodeposition of Aluminum, Transactions of Institute of Metal Finishing, 43 (1965), 186. 7. C. L. Hussey, Advances in Molten Salt Chemistry, ed. G. Mamantov (Amsterdam: Elsevier Science. 1983) Vol. 5, 185. 8. R. A. Osteryoung, Molten Salt Chemistry, ed. G. Mamantov and R. Marassi (Holland, D. Reidel, 1987) 329. 9. B. Wu, R. G. Reddy and R. G. Rogers, Aluminum Recycling via Room Temperature Electrolysis in Ionic Liquids, Proceedings of Fourth International Symposium on Recycling of Metals and Engineered Materials TMS, Pittsburgh, 2000, 845. 10. R. T. Carlin, W. Crawford and M. Bersch, Nucleation and Morphology Studies of Aluminum Deposited from an Ambient Temperature Chloroaluminate Molten Salt, J. Electrochem. Soc, 139 (1992), 2720. 11. G. J. Hill, D. J. Schiffrin and J. Thompson, Electrochemical Nucleation from Molten SaltsI. Diffusion Controlled Electrodeposition of Silver from Alkali Molten Nitrates, Electrochimica. Acta, 19(1974), 657-670. 12. M. Fleischmann and H. R. Thrisk, Advances in Electrochemistry and Electrochemical Engineering, ed. P. Delahay and C. W. Toibas (New York, NY: Interscience, 1963), 123. 1.

1.20 1.00 0.80

[ i / i m ]2

0.60 0.40 0.20 0.00 -0.20 0 1 2 3 4 5

[t / t m ]

Figure 7: Non-dimensional (i/im)2 vs. t/tm plot of the data from figure 5. Where () is instantaneous, (----) is progressive nucleation respectively and () 0.5V, () 0.6V, () 0.7V and ( ) 0.8V. The treatment of the chronoamperograms discussed above requires that the product im2 tm be independent of the nucleation and growth rates. Thus at a given bulk concentration of electrodepositing species im2 tm should not vary with the overpotential for sufficiently high step potentials, for which the surface concentration of aluminum is effectively zero.

2 i m t m = 0.1629 nFC 0 D0

(8)

In addition, it is possible to calculate the diffusion coefficient of Al2Cl7 ions from the maxima in the current-time transients using Eq. (8) developed for instantaneous type of nucleation [21]. From the mean value of the product im2 tm = 6.48 x 105 A2 s /cm4, resulting from experimental chronoamperograms exhibited in Fig. 5, the diffusion coefficient is obtained as D0 = 0.39 x 107 cm2 sec1. The diffusion coefficients obtained by this method are comparable to that available for aluminum deposition in the literature.

13. H. R. Thirsk and J. A. Harrison, A Guide to the Study of Electrode Kinetics, (New York, NY: Academic Press, 1972), 115. 14. D. J. Astley, J. A. Harrison and H. R. Thirsk, Trans Faraday. Soc, 64 (1968), 192. 15. L. Legrand, A. Tranchant and R. Messina, Electrodeposition Studies of Aluminum on Tungsten Electrode from DMSO2 Electrolyte, J. Electrochem. Soc., 141 (1994), 378. 16. B. Wu, R. G. Reddy and R. G. Rogers, Aluminum Reduction via Near Room Temperature Electrolysis in Ionic Liquids, Light Metals 2001, (TMS, 2001), 237-243. 17. Y. Zhao and T. J. Vandernoot, Electrodeposition of Aluminum from Room Temperature AlCl3TMPAC Molten Salts, Electrochimica. Acta, 42 (1997), 1639-1643. 18. A. J. Bard and L. R. Faulkner, Electrochemical Methods: Fundamentals and Applications, (New York: NY, J. Wiley & Sons, Inc. 2000). 19. J. Robinson and R. A. Osteryoung, The Electrochemical Behavior of Aluminum in the Low Temperature Molten Salt System n-Butyl Pyridinium Chloride: Aluminum Chloride and Mixture of this Molten Salt with Benzene, J. Electrochem. Soc., 127 (1980), 122. 20. G. A. Gunawardena, G. J. Hill and I. Montenegro, Potentiostatic Studies of Electrochemical Nucleation, Electrochimica. Acta, 23 (1978), 693-697. 21. B. Scharifker and G. Hill, Theoretical and Experimental Studies of Multiple Nucleation, Electrochimica. Acta, 28 (1983), 879-889. 22. G. A. Gunawardena, G. J. Hill, I. Montenegro and B. Scarifker, Electrochemical Nucleation, J. Electroanal. Chem., 138 (1982), 225-239. 23. P. Delahay, New Instrumental Methods in Electrochemistry, (New York: NY. Interscience, 1954),

chapter 3.

S-ar putea să vă placă și