Anda di halaman 1dari 8

Supplemental Proceedings: Volume 3: General Paper Selections

TMS (The Minerals, Metals & Materials Society), 2010

ELECTROCHEMISTRY OF TANTALUM PENTACHLORIDE IN THE


ROOM TEMPERATURE IONIC LIQUID 1-BUTYL-3-METHY
IMIDAZOLIUM HEXAFLUOROPHOSPHATE

Xiaoxiang Zhang1, Huimin Lu1∗, Tao Zhang1


1
Beihang University, School of Materials Science & Engineering;
37 Xueyuan Road, Haidian District, Beijing 100191, China
* Corresponding author: lhm0862002@yahoo.com.cn

Keywords: tantalum; electrochemical deposition; ionic liquid; [Bmim]PF6

Abstract

Electrochemical behavior of 0.25M tantalum pentachloride on platinum substrate was


investigated in the 1-butyl-3-methy imidazolium hexafluorophosphate ([Bmim]PF6) by cyclic
voltammetry, chronoamperometry. Cyclic voltammetrys of tantalum pentachloride in the
employed ionic liquid at 100ºC exhibit four reduction peaks, the first reduction peak contributes
to the reduction of trichloride anion (Cl3-) at -0.4V vs. Pt, the second corresponds to the
reduction of the tantalum (V) to tantalum (III) at -0.9V vs. Pt, the third is attributed to the
reduction of the tantalum (III) to tantalum metal at -1.25V vs. Pt and the fourth is the formation
of various tantalum subchlorides at -1.6V vs. Pt. The results have shown that the
electrodeposition of tantalum from tantalum pentachloride in [Bmim]PF6 is quite complicated
process, but tantalum can be electrodeposited in thin layers.

Introduction

Tantalum has been widely used in various fields, from electronics to mechanical and aeronautics
and astronautics industry, due to its excellent properties such as high melting point, sufficient
ductility, toughness and excellent corrosion resistance [1]. As the raw material of tantalum
capacitor, the quantity demanded of tantalum increased gradually. Owing to low diffusion
coefficient, high thermal conductivity and good adhesion in copper and silicon, as well as good
thermal stability, tantalum becomes the best diffusion interlayer materials for the application in
ultra large-scale integrated circuits [2, 3]. In addition, its resistance to body fluids and
biocompatibility make tantalum a perfect medical material.

Because of the narrow electrochemical window of aqueous solution and the negative deposition
potential of tantalum, it is impossible to electrodeposit tantalum from aqueous solution. High
temperature molten salts were found to the suitable electrolyte for the electrodeposition of
refractory metals. There are some reports about the electrodeposition of tantalum in high
temperature molten salts [4 ~ 7]. For example, Senderoff and Mellors in first reported that the
eletrodeposition of coherent tantalum coatings in ternary eutectic molten salts LiF-NaF-KF used
K2TaF7 as tantalum source between 650 ºC and 850ºC [4]. Balikhin has shown that the optimum
conditions for tantalum electroplating in the electrolyte LiF-NaF-K2TaF7 under an excess
pressure of argon in a temperature of 800ºC [5]. At the same time, Wu and Chen performed that
a new method of producing tantalum by electrochemical reduction of solid Ta2O5 directly in high
temperature salts CaCl2 at 850ºC [6]. However, the above tantalum electrochemical reduction
methods are based on high temperature molten salts, which bring many problems, such as high
temperature corrosion, low current efficiency and large energy consumption.

867
In recent years, the electrodeposition of refractory metals has attracted the interest of many
researchers by use of room temperature ionic liquids as electrolyte instead of high temperature
molten salts, due to its wide electrochemical window, negligible volatility, high thermal stability,
and high conductivity. Some studies on electrochemical deposition of tantalum (V) from ionic
liquids have been reported [8 ~ 13]. The first report that a tantalum film can be electrodeposited
in a mixture of 30mol% TaCl5, 10 mol% LiF and 60 mol% [EMIM]Cl at 100ºC by Masatsugu
and Morihisa [8]. Abedin and Endres have showed that thin layers of Ta can be electrodeposited
at 200ºC from TaF5 in the air and water-stable ionic liquid [Py1,4]TFSA [9], and the quality and
the thickness of the deposit were improved significantly by addition of LiF, and they also
obtained about 1μm thick crystalline Ta layers from TaF5-LiF- [Py1,4]TFSA mixtures on Ni–Ti
alloy substrate at 200ºC [10]. Babushkina have investigated the complex formation of
tantalum(V) in the mixtures of (x)1-Butyl-1-methyl- pyrrolidinium chloride-(1-x)TaCl5 in solid
and molten states by Raman spectroscopy over the temperature range 20ºC ~ 160ºC [11].
Recently, Borisenko and Endres studied the tantalum electrodeposition from TaF5 in the air and
1-butul-1- methylpyrrolidinium bis (trifluoromethylsulfonyl) amide ionic liquid in situ STM and
EQCM, the results indicated that the formation of micrometer thick tantalum layers was
restricted at room temperature because of some kinetic problems [12]. In this paper, the
electrochemical behavior of 0.25M tantalum pentachloride on platinum substrate was
investigated in the 1-butyl-3-methy imidazolium hexafluorophosphate, because the tantalum
pentachloride is a common and cheap tantalum material and the [Bmim]PF6 is also an often-used,
inexpensive and air and water-stable ionic liquid.

Experimental

Chemicals

1-butyl-3-methy imidazolium hexafluorophosphate ([Bmim]PF6) ionic liquid (≥99%, from Lihua


Pharmaceutical Ltd., Henan Province, China) was further dried under vacuum at a temperature of
105ºC for 12h to remove any trace of water. Tantalum pentachloride (≥99%, Orient Tantalum
Industry Ltd., Ningxia Province, China) was used without further handling. The mixtures of
[Bmim]PF6 with 0.25M TaCl5 were prepared by adding appropriate amount of TaCl5 to
[Bmim]PF6 in Bunsen beaker, and stirring the solutions inside a glove box at 60ºC for several
hours.

Electrochemical experiments and apparatus

All electrochemical experiments were conducted under a purified nitrogen atmosphere inside a
glove box with the moisture and oxygen contents below 1 ppm. The electrochemical experiments
were carried out in a three-electrode glass cell on a CHI6108C electrochemical workstation (CH
Instrument Co., USA), and electrodepositing experiments were accomplished with a DH1719
potentiostat /galvanostat (Beijing, China). For cyclic voltammetry and chronoamperometry,
platinum wires (1mm diameter) were used as working electrode and counter electrode. Platinum
substrates were polished with a corundum suspension, and cleaned in an ultrasonic in acetone for
10 mim before use. A platinum wire (0.5mm diameter) was used as reference electrode. The
electrodepositing experiments were conducted on platinum plates (area = 0.5cm2) rinsed by
acetone and dilute nitric acid. The deposits were cleaned in acetone to remove residual ionic
liquid, and washed with deionized water repeatedly. The scanning electron microscope (SEM)
with an energy dispersive spectroscopy (EDS) was employed to examine the surface morphology
and the element analysis of the elecrodeposited film.

868
Results and Discussion

Typical cyclic voltammograms that were recorded on the Pt electrode in [Bmim]PF6 ionic liquid
without and with 0.25 M TaCl5 with a sweep rate of 50mv/s at 100ºC are showed in Fig.1. The
ionic liquid itself exhibits an electrochemical window of above 4V on Pt electrode (inset in
Fig.1). Scans were started from open circuit potential, the cathodic limit is mainly due to the
irreversible reduction of the [Bmim]+, and the anodic limit is due to the irreversible anion
oxidation. As shown in Fig.1, there are four reduction peaks (C1-C4) on negative scan, and three
oxidation peaks (A1-A3) on positive scan. The first reduction peak C1 was very obvious after the
appearance of the oxidation peak A1 in the following circle, the oxidation peak A1 appears at
0.86V due to the anodic conversion of Cl- to Cl3- [14], associated with a reduction peak C1 at -
0.4V. The second reduction peak C2 located at -0.9V corresponds to the reduction of tantalum (V)
to tantalum (III), and the third reduction peak C3 at -1.25V is attributed to the formation of
tantalum metal. The anodic peaks A2 at -0.86V and A3 at -1.2V are accorded with the cathodic
peaks C2 and C3 respectively. It could be inferred that the fourth reduction peak C4 at -1.6V
perhaps originated from the formation of subvalent non-stochiometric tantalum chlorides and
other ones, and the nucleation of tantalum metal was bound to be chloride [12].

Fig.2 shows the effect of different temperature on the cyclic voltammograms of 0.25 M TaCl5 in
[Bmim]PF6 on platinum electrode at a scanning rate of 50mV/s. The cathodic peak potentials
slightly move to less negative values and the potential of oxidation peaks slightly shift to the
positive direction with temperature increasing. The peak currents of the reduction and oxidation
peaks remarkably increase with temperature rising. The conductivity of such types of ionic
liquids increases and the viscosity of them decreases with temperature increasing, which is
beneficial for the migration of electroactive species, as well as the transfer of charge.

Fig.1 Typical cyclic voltammograms of [Bmim]PF6 without and with 0.25M TaCl5 on Pt
electrode at 50mV/s

869
Fig.2 Cyclic voltammograms of 0.25M TaCl5 in [Bmim]PF6 on platinum electrode at different
temperature at 50mV/s

Fig.3 Cyclic voltammograms of 0.25M TaCl5 in [Bmim]PF6 on platinum electrode at various


scanning rates at 100ºC

Fig.3 shows the effect of various scanning rates on the cyclic voltammograms of 0.25M TaCl5 in
[Bmim]PF6 on platinum electrode at 100ºC. We can see that an increase in the scanning rate
shifts the reduction peak to a more negative potential, and moves slightly the oxidation peak to
positive potential, which indicate the irreversibility of the reduction process. In the irreversible
reduction electrode process, the cathodic transfer coefficient can be calculated by following
equation [15]:

|Ep-Ep/2|=1.857RT/( αnF) (1)

where Ep and Ep/2 are the peak potential and half peak potential, respectively, and α is the
cathodic transfer coefficient. The value of cathodic transfer coefficients obtained by applying
this equation and the data of cyclic voltammograms are listed in the table I and table II. The
results show that the cathodic transfer coefficients of Ta (V) /Ta (III) and Ta ( III) /Ta are 0.156

870
and 0.392 in ionic liquid [Bmim]PF6 in table I and table II. The value of |Ep-Ep/2| is larger than
0.011V theoretical value, which expects to be a two-electron electrode process at this
temperature [16].

Table I Cyclic voltammogram data for Ta (V) /Ta (III) with various scanning rates
v/(Vs-1) Ipc/mA v1/2(V-1/2 s-1) Epc/V Epc/2/V Ep/2- Ep/V α
0.01 0.5296 0.1 -0.884 -0.695 0.189 0.158
0.02 0.5954 0.141 -0.865 -0.680 0.185 0.161
0.05 0.7781 0.224 -0.900 -0.711 0.189 0.158
0.06 0.8583 0.245 -0.905 -0.714 0.191 0.156
0.08 0.9511 0.283 -0.925 -0.731 0.194 0.154
0.1 0.997 0.316 -0.935 -0.737 0.198 0.151
0.2 1.318 0.447 -0.941 -0.750 0.191 0.156
Average 0.156

Table II Cyclic voltammogram data for Ta ( III) /Ta with various scanning rates
v/(Vs-1) Ipc/mA v1/2(V-1/2 s-1) Epc/V Epc/2/V Ep/2- Ep/V α
0.01 0.3765 0.1 -1.220 -1.130 0.090 0.332
0.02 0.4086 0.141 -1.210 -1.132 0.078 0.383
0.05 0.5651 0.224 -1.250 -1.181 0.069 0.433
0.06 0.6147 0.245 -1.264 -1.188 0.076 0.392
0.08 0.6627 0.283 -1.283 -1.212 0.071 0.421
0.1 0.7245 0.316 -1.297 -1.223 0.074 0.403
0.2 0.9240 0.447 -1.313 -1.235 0.078 0.383
Average 0.392

The dependence plot of the peak currents of cathodic peak C2 and C3 is given in Fig.4. The
diffusion coefficient of electroactive species could be obtained by applying the Berzins-Delahay
equation for a diffusion controlled process [17]:

Ip = 0.61(nF)3/2(RT)-1/2AC0D1/2v1/2 (2)

here Ip is the peak current, C0 the bulk concentration of the electroactive species, D the diffusion
coefficient, and A is the electrode area. As shown in Fig.4, the peak currents of cathodic peak C2
and C3 are both linear relationship against the square root of sweep rate, which indicated that the
electrode reaction process is diffusion controlled process. But the fitting line does not pass
through the origin, which shows that it is not a simple linear diffusion controlled process, some
other phenomena such as adsorption phenomena, surface chemical reaction process could be
considered. The value of Ta (V) diffusion coefficient calculated is 9.348 × 10-10 cm2s-1/2.

871
Fig.4 A dependence plot of the cathodic peak current against square root of sweep rate on
platinum electrode at 50mV/s, constructed from peak C2 (Fig.4a) and peak C3(Fig.4b), of curves
shown in Fig.1

Chronoamperometry experiments were conducted by stepping the potential from a value where
no reduction process would occur to those potentials negative to induce reduction process. The
current-time transients on platinum electrode in 0.25M TaCl5 in [Bmim]PF6 at 100ºC are shown
in Fig.5. The initial current increased sharply is due to double-layer charging, and with more
negative step potential, the greater the maximum current value. The liner I-t-1/2 relationship
derived from the data in Fig.5 indicates that the reduction processes are diffusion control ones
(Fig.6). The diffusion coefficient can be calculated by applying Cottrell equation [18]:

I = nFAC0D1/2(πt) -1/2 (3)

and the average D value of Ta (V) estimated from chronoamperometry experiments is in the
range of 10-10 cm2s-1/2, which is almost agreement with the result by cyclic voltammogram. The
extremely slow diffusion coefficient is very unfavorable for the electrodeposition of tantalum in
this ionic liquid studied.

Fig.5 Current-time transients resulting from chronoamperometry experiments of 0.25M TaCl5 in


[Bmim]PF6 on platinum electrode at 100ºC at various overpotentials: (a)-0.5V, (b)-0.8V, (c)
-1.2V, (d)-1.4V, (e)-1.8 V

872
Fig.6 The linear I-t-1/2 relationship obtained from chronoamperometry experiments. (a)-0.5V, (b)
-0.8V, (c)-1.2V, (d)-1.4V, (e)-1.8V

Conclusion

The investigation of cyclic voltammograms shows that the electrodeposition of tantalum from
tantalum pentachloride on platilum substrate in [Bmim]PF6 is quite complicated process at 100ºC.
First the tantalum (V) is reduced to tantalum (III), later tantalum (III) is reduced to tantalum
metal and tantalum subchlorides. The reduction peaks shifted to negative direction with the
increase of scanning rate indicates that the electrode reaction process is irreversible reduction
electrode process by diffusion control. The cathodic transfer coefficients of Ta (V) /Ta (III) and
Ta (III) /Ta are 0.156 and 0.392 respectively, and the diffusion coefficient of Ta (V) is 9.348 ×
10-10cm2s-1/2. The extremely slow diffusion coefficient is very unfavorable for the
electrodeposition of tantalum in this ionic liquid studied.

References

[1] C. E. Mosheim, “Niobium and Tantalum: a Review of Industry Statistics,” TIC, 112 (2002),
2 ~ 9.
[2] T. Laurila et al., “Chemical Stability of Tantalum Diffusion Barriers between Cu and Si,”
Thin Solid Films, 373 (2000), 64 ~ 67.
[3] J. W. Lim et al., “Improvement of Ta Barrier Film Properties in Cu Interconnection by
Using a Non-Mass Separated Iron Beam Deposition Method,” Materials Transactions, 43
(3) (2002), 478 ~ 481.
[4] S. Senderoff, G. W. Mellors, and W.J. Reinhardt, “The Electrodeposition of Coherent
Deposits of Refractory Metal,” J. Electrochem. Soc, 112 (8) (1965), 840 ~ 845.
[5] V.S. Balikhin, “Electroplating of protective tantalum coating,” Zasch. Met, 10 (1974), 459 ~
460.
[6] Tian Wu et al., “Thin Pellets: Fast Electrochemical Preparation of Capacitor Tantalum
Powders,” Chem. Mater, 19 (2) (2007), 153 ~ 160.
[7] F. Lantelme et al., “Electrodeposition of Tantalum in NaCl-KCl-K2TaF7 Melts,” J.
Electrochem. Soc., 139(5) (1992), 1249 ~ 1225.
[8] M. Masatsugu and M. Morihida. JPn. Kokai Tokyo Koho; Coden: JKXXAF JP2001279486
A200111010, 2001.

873
[9] S. Zein El Abedin et al., “Electroreduction of Tantalum Fluoride in a Room Temperature
Ionic Liquid at Variable Temperatures,” Phys. Chem. Chem. Phys, 7 (2005), 2333 ~ 1339.
[10] S. Zein El Abedin, U. Welz-Biermann, and F.Endres, “A Study on the Electrodeposition of
Tantalum on NiTi Alloy in an Ionic Liquid and Corrosion Behavior of the Coated Alloy,”
Electrochemistry Communications, 7 (2005), 941-946.
[11] Olga B. Babushkina, Silvia Ekres and Gerhard E. Nauer, “Raman Spectroscopy of the
Mixtures (x)1-Butyl-1-Methylpyrrolidinium Chloride-(1- x)TaCl5 in Solid and Molten
States,” J. Phys. Chem. A, 112 (2008), 8288 ~ 8294.
[12] N. Borisenko et al., “In Situ STM and EQCM Studies of Tantalum Electrodeposition from
TaF5 in the Air-and Water-Stable Ionic Liquid 1-Butyl-1-Methylpyrrolidinium Bis
(Trifluoromethylsulfonyl) Amide,” Electrochimica Acta, 54 (2009), 1519 ~ 1528.
[13] O. B. Babushkina, S. Ekres, G. E. Z. Nauer, “Spectroscopy and Electrochemistry of
Tantalum(V) in 1-Butyl-1-Methylpyrrolidinium Trifluoromethanesulfonate,” Naturforsch,
63(A) (2008),73-80.
[14] Huijiao Sun et al., “Unusual Anodic Behavior of Chloride Ion in 1-Butyl-3-
Methylimidazolium Hexafluorophosphate,” Electrochemistry Communications, 7(2005 ),
685 ~ 691.
[15] E. G. S. Jeng and I. W. Sun, “Electrochemistry of Thallium in the Basic Aluminum
Chloride-1-Methyl-3-Ethylimidazolium Chloride Room Temperature Molten Salt,” J.
Electrochem. Soc, 145 (4) (1998), 1196 ~ 1201.
[16] P. Y. Chen and I. W. Sun, “Electrochemistry of Cd(II) in the Basic 1-Ethyl-3-
Methylimidazolium Chloride/Tetrafluoroborate Room Temperature Molten Salt,”
Electrochimica Acta, 45 (2000), 3163 ~ 3170.
[17] A. M. Martínez et al., “Electrodeposition of Magnesium from CaCl2-NaCl-KCl-MgCl2
Melts,” J. Electrochem. Soc., 151(7) (2004), C508 ~ C513.
[18] B. R. Scharifker and G. Hill, “Theoretical and Experimental Studies of Multiple
Nucleation,” Electrochimica Acta, 28 (7) (1983), 879 ~ 889.

874

Anda mungkin juga menyukai