Vous êtes sur la page 1sur 4

G Model EA-17421; No.

of Pages 4

ARTICLE IN PRESS
Electrochimica Acta xxx (2011) xxxxxx

Contents lists available at ScienceDirect

Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta

Direct electrochemical reduction of titanium dioxide in Lewis basic AlCl3 1-butyl-3-methylimidizolium ionic liquid
Xiao-Ying Zhang, Yi-Xin Hua , Cun-Ying Xu, Qi-Bo Zhang, Xiao-Bo Cong, Nan Xu
Faculty of Metallurgical and Energy Engineering, Kunming University of Science and Technology, Kunming 650093, PR China

a r t i c l e

i n f o

a b s t r a c t
The direct electrochemical reduction of titanium dioxide (TiO2 ) to metallic titanium at room temperature is rstly studied in Lewis basic AlCl3 1-butyl-3-methylimidizolium (AlCl3 BMIC) ionic liquid. In this study, cyclic voltammetry, potentiodynamic polarization, sampled current voltammetry and X-ray photoelectron spectroscopy (XPS) techniques were utilized. Analysis of the cyclic voltammetry suggested that TiO2 lm can be reduced to metallic Ti. The sampled current voltammetry was applied to elaborate the reduction mechanism and the results showed that this reduction process may include two steps. When the output potential difference of 2.8 V was applied, a TiO2 cylindrical pellet was partly reduced to metallic Ti. However, due to the very slow reaction rate, there was only about 12 wt% of TiO2 was reduced during the electrolysis time of 48 h. It was predicted that the process for the direct reduction of solid TiO2 would be explained as follows: given enough cathode potential, the reduction happened at the cathode/ionic liquids interface, where the oxygen was ionized, then dissolved in the ionic liquid and discharged at the anode, with the metallic Ti left at the cathode. 2011 Published by Elsevier Ltd.

Article history: Received 10 February 2011 Received in revised form 10 July 2011 Accepted 11 July 2011 Available online xxx Keywords: Titanium dioxide Electrochemical reduction Ionic liquids Cyclic voltammetry Sampled current voltammetry XPS

1. Introduction Titanium has been acknowledged as a favorable structural material and used in many elds due to its many desirable properties such as high strength, low density and excellent corrosion resistance [1]. The Kroll process is a main method used for the industrial production of sponge titanium. However, its use has been restricted because of the high processing cost compared with other common metals, which is resulted from its complicated and discontinuous procedures [2]. To overcome these shortcomings, a lot of efforts have been done on producing metallic titanium, such as FFC Cambridge method [3], OS process [4], EMR process [5] and the perform reduction process (PRP) [6]. The FFC method, proposed by Fray et al., has inspired not only academic research activity on titanium but also related industries. It is most likely to be applied in larger scale industrial production, though there are still several technical problems, such as separation of Ti product from the bath constituents, purication to very low chlorine level and its high reaction temperature, needed to be solved. In our opinion, a new electrolyte needs to be found for this electrochemical reduction process. The discovery of the new class of substance with low melting points, good conductivity and wide electrochemical window, named as room temperature ionic liquids

Corresponding author. Tel.: +86 871 5162008; fax: +86 871 5161278. E-mail address: huayixin@gmail.com (Y.-X. Hua). 0013-4686/$ see front matter 2011 Published by Elsevier Ltd. doi:10.1016/j.electacta.2011.07.037

(RTILs), has attracted much attention of many researchers in the last few years. Although much admirable progress has been made in exploring their application, as electrolyte in electro-deposition of metals, there are only few reports on the application of RTILs in electro-deposition of titanium. More recently, much more attempts have been made by a number of researchers to electro-deposit titanium in this excellent electrolyte. Mukhopadhyay et al. [7,8] have obtained a ultrathin titanium lm of 12 nm from TiCl4 in ionic liquid [BMIm]BTA at room temperature on highly oriented pyrolytic graphite and Au(1 1 1) substrate. They concluded that TiCl4 was rstly reduced to TiCl2 , which was subsequently reduced to metallic Ti. During this process, a brown lm, probably the TiCl3 , was deposited on the electrode, possibly inhibiting the reduction. Endres et al. [9] have investigated the electro-deposition of Ti in the air- and water-stable TiCl4 -containing ionic liquids [EMIm]Tf2 N, [BMP]Tf2 N and [P14,6,6,6 ]Tf2 N. Their research implied that only nonstoichiometric halides were formed instead of Ti. Alternatively, the reduction of TiCl4 to elemental Ti was at least extremely difcult if not even impossible. Andriyko et al. [10,11] studied the electrochemistry of TiCl4 and TiF4 in BMMImN3 and BMIMBF4 , respectively. The results of the former experiment denoted that the reduction was kinetically hindered due to the formation of strong azide complex of Ti (IV). In the latter one, some poorly soluble low valence intermediates formed, which was reduced to thin Ti coatings on a Pt substrate. The research by Tsuda et al. [12] implied that electro-deposition of AlTi alloy in the Lewis acidic AlCl3 EMIC ionic liquid was feasible. So far to our knowledge, only can Ti alloy

Please cite this article in press as: X.-Y. Zhang, et al., Direct electrochemical reduction of titanium dioxide in Lewis basic AlCl3 1-butyl-3methylimidizolium ionic liquid. Electrochim. Acta (2011), doi:10.1016/j.electacta.2011.07.037

G Model EA-17421; No. of Pages 4 2

ARTICLE IN PRESS
X.-Y. Zhang et al. / Electrochimica Acta xxx (2011) xxxxxx

be obtained in ionic liquids but pure metallic Ti. Thus, new kinds of ionic liquids or Ti precursors need to be found to obtain Ti at low temperature. As well known, AlCl3 -based ionic liquids are easily synthesized by simple mix addition of Lewis acidic AlCl3 to a 1-butyl-3methylimidizolium chloride under an inert atmosphere. If the mole fraction of AlCl3 is less than 0.5 in the nal product, the ionic liquids are basic. In contrast, they are acidic. In the case of 0.5, the salts are neutral. The electrochemical window and viscosity of Lewis basic AlCl3 -based IL are suitable to act as a supporting electrolyte for the reduction of TiO2 . Nevertheless the main disadvantages of this IL are their corrosiveness and their instability against air and moisture [13]. The major purpose of our present work is to investigate if TiO2 can be electrochemically reduced in Lewis basic AlCl3 BMIC at low temperature. Some studies will be carried out to preliminarily analyze the mechanism of this reduction process. 2. Experimental 2.1. Chemicals Anhydrous AlCl3 (99 wt%) was used without further treatment. 1-Butyl-3-methylimidizolium chloride (BMIC) was synthesized based on reported procedures [14]. The electrolyte of Lewis basic ionic liquid AlCl3 BMIC used in this paper was prepared by the slow addition of a certain weight of AlCl3 into BMIC in a mole ratio of 0.8:1.0, respectively, giving a mole fraction of AlCl3 of 0.44. 2.2. Preparation of Ti electrodes A titanium foil (commercial purity, 30 mm 10 mm 0.1 mm) was polished with a sequence of emery papers of different grades (400, 600 and 800), degreased with acetone in ultrasonic bath for 5 min, washed three times with distilled water, and dried at room temperature. Subsequently, the titanium foil was oxidized in a furnace at 823 K in air for different time (48, 96 and 144 h). Obviously, the more oxidation time, the thicker oxide lm (0.28, 0.56 and 0.67 m). In our precious experiment, this electrochemical reduction cannot be achieved with commercial purity TiO2 . Therefore a cylindrical pellet (10 mm i.d., weight 0.8 g) was made of TiO2 powders synthesized by solgel method [15] to improve the activity of TiO2 particles, which was completely mixed and ground with 6 wt% polyvinyl alcohol (PVA). After drying at room temperature, the TiO2 pellet was then heated in air by programmed heating in four stages: increasing from 298 to 398 K at 5 K/min, holding at 398 K for 1 h, increasing from 398 to 623 K at 100 K/h, increasing from 623 to 723 K at 40 K/h, holding at 723 K for 3 h, increasing from 723 to 823 K at 100 K/h, holding at 823 K for 4 h. The heating was aimed at removing water and pore-forming. 2.3. Electrochemical experiments Electrochemical tests were conducted in a standard threeelectrode electrochemical cell using the oxidized Ti foil mentioned above as the working electrode, a platinum wire as a counter electrode, and a silver wire as a reference electrode. Cyclic voltammetry, cathode polarization and sampled current voltammetry were carried out to investigate the electrochemistry of oxidized Ti foil in AlCl3 BMIC. The scan rate in cyclic voltammetry was 50 mV/s and the rst scan direction was negative. The cathode polarization curves were measured by linear sweep voltammetry with a scan rate of 5 mV/s. In the sampled current voltammetry test, the voltage range was from 0.5 to 2.4 V vs. Ag with the sample period of 10 s and step voltage of 0.1 V.

Fig. 1. Cyclic voltammograms of titanium foil with oxide lm (0.67 m) and glassy carbon electrode in Lewis basic AlCl3 BMIC vs. Ag obtained at 50 mV/s and 343 K.

During all experiments the cell was purged by an argon ow to avoid the ingress of air. The surface area of working electrode was estimated from the depth of the electrode immersed in the electrolyte. In each test, a new oxidized Ti foil was used and the temperature was hold at 343 K. All the electrochemical experiments were performed on a Gamry PCI4/300 electrochemical work station for the experimental control and data acquisition. 2.4. Electrochemical reduction Electrolysis experiment with a two-electrode arrangement was performed in the same cell that was used for electrochemical experiments. The TiO2 pellet was placed into a Ti basket made of commercially pure Ti sheet with little holes, which acted as the cathode. The same platinum wire in the electrochemistry test served as the anode. During all the electrochemical reduction process, the cell was also purged by an argon ow and the output potential difference of 2.8 V was controlled by a potentiostat. Temperature was maintained at 373 K to increase the reaction rate. 3. Results and discussion Three typical cyclic voltammograms of titanium electrode (solid and dash lines) with the oxide lm and glassy carbon electrode (dot line) in AlCl3 BMIC melt at 343 K are shown in Fig. 1. In the scan voltage range of 0.5 to 0.9 V, two reduction waves A and B at 0.33 and 0.60 V vs. Ag and oxidation wave C at a potential of

Fig. 2. Cyclic voltammograms of bare Ti foil and oxidized titanium foil (0.67 m) after electrolysis in Lewis basic AlCl3 BMIC vs. Ag obtained at 50 mV/s and 343 K.

Please cite this article in press as: X.-Y. Zhang, et al., Direct electrochemical reduction of titanium dioxide in Lewis basic AlCl3 1-butyl-3methylimidizolium ionic liquid. Electrochim. Acta (2011), doi:10.1016/j.electacta.2011.07.037

G Model EA-17421; No. of Pages 4

ARTICLE IN PRESS
X.-Y. Zhang et al. / Electrochimica Acta xxx (2011) xxxxxx 3

Fig. 3. Cathode polarization curves in Lewis basic AlCl3 BMIC at 343 K for oxidized Ti foils with different oxide lm thickness (0.28, 0.56 and 0.67 m) at a sweep rate of 5 mV/s.

Fig. 4. Sampled current voltammogram of oxidized Ti foil (0.67 m) in Lewis basic AlCl3 BMIC at 343 K. The potential range was from 0.5 to 2.4 V vs. Ag with the sample period of 10 s and step voltage of 0.1 V.

0.18 V were observed in the cyclic voltammetry curve of oxidized Ti electrode at negative direction (solid line). The appearance of reduction waves A and B were most probably owing to reduction of the TiO2 lm. After 5 min electrolysis with the cathode potential set as 1.6 V (dash line), the reduction waves A, B still appeared (A , B ) with the oxidation peak C becoming greater, which can be possibly assigned to the reoxidation of already reduced Ti. In Fig. 2, two cyclic voltammetry curves of bare Ti foil (solid line) and oxidized Ti foil after long time electrolysis (dash line) were illustrated. As a result of the electrolysis of 8 h with the cathode potential set as 1.6 V, the reduction waves A, B and their corresponding anodic waves C of oxidized Ti foil were absent, the appearance of which was very similar to the one of the bare Ti foil. Additionally, we found that the overall slope of current and potential decreased in some degree compared with the cyclic voltammetry of oxidized Ti foil in Fig. 1. The reduction process of TiO2 in the IL is an interfacial reaction. In our opinion, after the outer oxide lm was reduced, the contact area between un-reduced oxide and IL was decreased. Alternatively, when the IL diffused thorough the reduced Ti, the effective reduction area of un-reduced oxide lm became less than beginning. The reduction process was consisted of reduction of high valence Ti ion to low one and reduction of low valence Ti ion to elemental Ti. As well known, the composition of oxide lm thermally formed on the Ti foil is nonstoichiometric. Moreover, the reduction potential of each valence Ti ion is very close to each other. Thus the reduction current is possibly responsible for the united contribution of some reduction processes of respective valence Ti ions. The cyclic voltammetry experiments denoted a dependence of the reduction potential on the thickness of TiO2 lm formed on the oxidized Ti foil. As clearly can be seen from Fig. 3, the thicker was oxide lm, the more negative was cathode polariza-

tion potential. In another word, the increasing oxide lm thickness (resistance) makes a greater cathode overpotential on the reduction. This result was due to the cathode inhibition effect of the oxide lm on the reduction process. We believe that direct electrochemical reduction of bulk TiO2 should need a greater driving force to be completed. In the cyclic voltammetry experiment, two reduction waves A, B appeared because of the oxide lm. However, it is not easy to determine directly how many steps are in the reduction process of oxidized Ti foil. Thus sampled current voltammetry was performed in AlCl3 BMIC melt at 343 K. When the sample period was set as 10 s, it showed evidently two charge transfer reactions in a voltage range of 0.5 to 2.4 V (Fig. 4). Each of them can account, respectively, for one of the two steps of this reduction process. At the beginning of each step of the reduction process, there was an increasing current as the voltage increased, when the reaction rate was controlled by charge transfer. Then both the currents reached a limit value, which means that each reaction got to its steady state. The mass transfer or the diffusion control step may be assigned to the reason that the current did not increase with the voltage increased [16]. From the viewpoint of solid physics, TiO2 lm thermally formed on titanium, well known as an n-type semiconductor, has a wide bandgap (Eg 3.2 eV), which makes its charge transfer more difcult than other narrow ones [17]. For the aim of reduction of TiO2 lm, increasing the cathodic polarization can make the electronic bands bend downward at the TiO2 lm/ionic liquid interface, where the oxygen in the TiO2 cathode was ionized, and as a result, the TiO2 lm was reduced [18]. We suppose that the reduction of TiO2 may occur as following steps: Given an enough great cathode potential, the charge transfer

Fig. 5. Photographs of TiO2 pellet before (a) and after (b, c) electrolysis (2.8 V, 373 K, 48 h) in Lewis basic AlCl3 BMIC.

Please cite this article in press as: X.-Y. Zhang, et al., Direct electrochemical reduction of titanium dioxide in Lewis basic AlCl3 1-butyl-3methylimidizolium ionic liquid. Electrochim. Acta (2011), doi:10.1016/j.electacta.2011.07.037

G Model EA-17421; No. of Pages 4 4

ARTICLE IN PRESS
X.-Y. Zhang et al. / Electrochimica Acta xxx (2011) xxxxxx

and c). Before the XPS test, the TiO2 pellet was rinsed by acetonitrile in a dry box to avoid re-oxidation of already reduced Ti. If not, the pellet turned to be light grey or even nearly white. As can be seen from Fig. 6a, the X-ray photoelectron spectroscopy of the TiO2 pellet after electrolysis contains two sets of peaks, respectively, for Ti and TiO2 , which were obviously different from the one set of peaks just for TiO2 (Fig. 6b). Due to the very slow reaction rate, there was only about more than 12 wt% of TiO2 reduced to metallic Ti in this experiment without the re-oxidation of reduced Ti. 4. Conclusions (1) Veried by cyclic voltammetry experiment, TiO2 lm can be reduced to metallic Ti in AlCl3 BMIC melt at room temperature. (2) Bulk TiO2 synthesized by solgel method also can be reduced to metallic Ti, but it needs greater overpotential. (3) Cyclic voltammetry and sampled current voltammetry conrm that the reduction process of TiO2 possibly includes two steps. (4) A further study ought to be carried out to obtain more information of the mechanism of this reduction process. In addition, some technical problems need to be improved, such as methods of making TiO2 electrode, selection of ionic liquids, optimization of experiment condition and so on. Acknowledgements The authors gratefully appreciate for the support of the National Natural Science Foundation of China (Project No. 50864009) for this work. References
Fig. 6. XPS of TiO2 pellet after electrolysis (2.8 V, 373 K, 48 h) in Lewis basic AlCl3 BMIC (a) and pure TiO2 (b). [1] [2] [3] [4] [5] [6] [7] [8] [9] H.M. Flower, Nature 407 (2000) 305. O. Taiji, T.H. Okabe, Titanium Jpn. 56 (2008) 18. G.Z. Chen, D.J. Fray, T.W. Farthing, Nature 407 (2000) 361. K. Ono, R.O. Suzuki, J. Met. 54 (2002) 59. T. Abiko, I. Park, T.H. Okabe, Proceedings 10th World Conference on Titanium, Hamburg, Germany, 14 July, vol. 7, 2003, p. 13. T.H. Okabe, T. Oda, Y. Mitsuda, J. Alloys Compd. 364 (2004) 156. I. Mukhopadhyay, W. Freyland, Langmuir 19 (2003) 1951. I. Mukhopadhyay, C.L. Aravinda, D. Borissov, W. Freyland, Electrochim. Acta 50 (2005) 1275. F. Endres, S. Zein, E.I. Abedin, A.Y. Saad, E.M. Moustafa, N. Borissenko, W.E. Price, G.G. Wallace, D.R. MacFarlane, P.J. Newman, A. Bund, Phys. Chem. Chem. Phys. 10 (2008) 2189. Y. Andriyko, G.E. Nauer, Electrochim. Acta 53 (2007) 957. Y. Andriyko, A. Andriiko, O.B. Babushkina, G.E. Nauer, Electrochim. Acta 55 (2010) 1081. T. Tsuda, C.L. Hussey, G.R. Stafford, J.E. Bonevich, J. Electrochem. Soc. 150 (2003) C234. F. Endres, D. MacFarlane, A. Abbott (Eds.), Electrodeposition from Ionic Liquids, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, 2008. X.H. Xu, C.L. Hussey, J. Electrochem. Soc. 140 (1993) 619. S. Eiden-Assmann, J. Widoniak, G. Maret, Chem. Mater. 16 (2004) 6. S.L. Wang, Y.J. Li, J. Electroanal. Chem. 571 (2004) 37. J.W. Schultze, M.M. Lohrengel, Electrochim. Acta 45 (2000) 2499. D.S. Kong, J.X. Wu, J. Electrochem. Soc. 155 (2008) C32.

occurs at the interface between oxide lm and ionic liquid, with diffusion of O2 dissolved in the melt through the pores of the cathode and a reduced porous Ti lm formed. Through the partly reduced porous Ti lm, ionic liquid diffuses and reaches the unreduced oxide lm. As mentioned above, the reduction still occurred at the interface between oxide lm and IL. In addition, the reduction process may be summarized as follows: Cathode : TiO2 + 4e Ti + 2O2 Anode : 2O2 O2 + 4e All the experiments above implied that electrochemical reduction of TiO2 lm in AlCl3 BMIC melt was feasible. To further conrm this, electrolysis of a small TiO2 pellet placed into a homemade Ti basket was performed. After electrolysis of 48 h, the Ti basket with TiO2 pellet was removed from the cell. Compared with the initial one (Fig. 5a), the TiO2 pellet visually turned to be dark grey (Fig. 5b

[10] [11] [12] [13] [14] [15] [16] [17] [18]

Please cite this article in press as: X.-Y. Zhang, et al., Direct electrochemical reduction of titanium dioxide in Lewis basic AlCl3 1-butyl-3methylimidizolium ionic liquid. Electrochim. Acta (2011), doi:10.1016/j.electacta.2011.07.037

Vous aimerez peut-être aussi