Vous êtes sur la page 1sur 12

J. Electroanal Chew 198 (1986) 319-330 Elsevter &quota S.A.

, Lausanne - Prmted

319 m The Netherlands

THE REDUCTION OF CHLORATE ACTIVE TITANIUM ELECTRODE

AND PERCHLORATE

IONS AT AN

GILBERT Chemutry (Recetved

M. BROWN Diorsron, Oak Ridge Natzonal Laboratory Oak Ridge. TN 37831 (U.S.A.) 1985)

3rd July 1985; in revised form 2nd September

ABSTRACT

Perchlorate ion is electrochemically reduced to chloride ion at an active tttanium electrode in aqueous 1.0 M HClO,. The reductton occurs by direct reaction at the surface rather than a pathway involving catalysis by soluble titanium corrosion products. The reaction occurs by oxygen atom transfer to the tttanium surface, and the implicattons of this mechanism for the surface composition of active tttanium electrodes are discussed. Chlorate ion is also reduced at titanium, and the rate constant for chlorate reduction is at least lo5 times greater than that for perchlorate reduction.

INTRODUCTION

Although perchloric acid is thermodynamically a powerful oxidizing agent, the direct reduction of perchlorate ion at an electrode surface is unusual. In fact, the perchlorate salts of tetraalkylammonium ions or alkali metal ions are frequently employed as supporting electrolytes for non-aqueous electrochemistry. The cathodic process limiting the usefulness of alkali metal perchlorates as supporting electrolyte in acetonitrile is the reduction of the alkali metal ion rather than ClO; [l]. Nevertheless there are a few reports [2-111 in the literature of the direct reduction of perchlorate at an electrode, although most of these are poorly characterized. Additionally, only a few examples of the direct reduction of chlorate ion have been published [7,12,13]. A number of relatively well characterized examples of the metal ion catalyzed electrochemical reduction of ClO, and ClO; have been reported [14-171. The reduction of perchlorate ion has been reported at platinum [2,3], tungsten carbide [4], ruthenium [5], carbon impregnated with Cr,O, or Al,O, [6], aluminum [7,8], and titanium [9-111 electrodes. The reactions at Pt, WC, and Ru are the reduction of adsorbed ClO; by a surface hydride species. The reduction of ClO; was observed during oxide film breakdown at Al or Ti with a high anodic potential applied to the interface. There are also reports of the oxidation of Ti in non-aqueous media in which perchlorate is the source of oxygen for film growth, The direct reduction of chlorate was observed at platinized titanium [12], stainless steel [12],

0022-0728/86/$03.50

0 1986 Elsevier Sequoia

SA

320

Armco Iron [13], and aluminum [7]. The reduction of chlorate is catalyzed by numerous metal ions including those of MO, W, V, and Ti. The catalysis by molybdenum-catechol complexes has been discussed recently [17]. There are also a few reports of the catalyzed reduction of perchlorate ion by metal ions [14]. The metal ion catalyzed pathway should be available any time a metal ion can be reduced at an electrode surface to generate a species which will reduce perchlorate or chlorate. The reduction of perchlorate ions by metal ions in solution has been discussed recently by Taube [18]. The results reported in this paper are an outgrowth of a study of the influence of anions on the anodic dissolution of titanium [19]. Perchlorate ions are generally considered to be weakly complexing in solution and to be weakly adsorbed at electrode surfaces. Perchlorate ions do not appear to assist the dissolution of titanium; however, at low concentrations of activating anion (chloride, sulfate, or oxalate) in a perchlorate electrolyte the stationary-state current-potential behavior of titanium changes. This change in behavior is attributed to the reduction of perchlorate at the titanium surface.
EXPERIMENTAL

Aqueous solutions of perchloric and hydrochloric acids were prepared from Baker Ultrex concentrated acids. The 72% perchloric acid has a stated chloride impurity of < 0.5 ppm so that 1.0 M HClO, solutions should have < 2 X low6 M Cl-. House line distilled water was further purified by triple distillation: first from acid chromate, then from alkaline permanganate, and finally from an all quartz still and collected in quartz containers. Sodium chlorate was Fisher certified grade, and solutions were prepared by adding weighed amounts to the desired acid solutions. Chlorate ion reacts with chloride ion to produce intermediate oxidation states of chlorine. On the basis of literature data [20], dilute solutions of chlorate in 1 M HCl should decay with a half life of about 4 x lo5 s at 30C. Accordingly, solutions of NaClO, in HCl were prepared immediately before use and kept in an ice bath while being degassed. Solutions were analyzed for chloride ion by Volhards method as outlined by Vogel [21]. Titanium was estimated by converting it to the peroxy complex and analyzing spectrophotometrically [22]. The electrochemical cell was constructed of glass and teflon and had three compartments (reference, test, auxiliary) separated by ungreased precision-ground glass stopcocks. The test and reference compartments were jacketed, and the temperature was maintained at 3O.OC with a circulating water bath. The test compartment had a flat bottom, and solution agitation was achieved with a magnetic stirrer and a teflon-coated spin bar. Dissolved oxygen was removed and protection from atmospheric contamination was maintained by continuously purging the cell solutions with hydrogen gas. The source of this gas was a GE model 15EHG hydrogen generator. A PAR Model 173 or PAR Model 371 potentiostat was used for potential control and current measurements. The current output was monitored with a Keithley Model 171 or Model 177 digital multimeter and simultaneously recorded

321

on a Hewlett-Packard Model 7100B strip chart recorder. In solutions containing perchlorate ion, commercial SCE reference electrodes could not be used because KClO, precipitated in the ceramic or fiber junction, and potential measurements became erratic. Therefore the reference electrode used in this work was a saturated sodium chloride calomel electrode (NaSCE). The saturated KC1 fill solution of a commercial ceramic junction SCE was removed and replaced with saturated NaCl. The potential of this electrode is 9 mV positive of the SCE at 30C. Titanium electrodes were fabricated from polycrystalline 6.4 mm diameter rods. The source of titanium was MRC (zone refined, MARZ grade) or Alfa Inorganics (99.98% pure); the stationary-state current-voltage behavior of these materials was identical. Internally threaded titanium samples were held on the end of a glass or Kel-F electrode holder with a threaded stainless steel rod, and the titanium to holder seal was made leak tight with a teflon washer. A Kel-F nut and teflon washer sealed the other planar surface of 12.7 mm long electrodes such that only the cylindrical surface was exposed to solution. Larger samples (63.5 mm long) had the cylindrical surface and the bottom planar surface in contact with solution. Titanium electrodes were pretreated to insure reproducible current-voltage behavior. Samples were mechanically and/or chemically polished and then held at constant potential in 0.5 M H,SO, until the current reached a stationary state and the parameters were in reasonable agreement with those reported by Kelly: potential maximum, E,,, = -530 mV vs. SCE; current at potential maximum, i,,, = 54 PA cmp2; and open circuit corrosion potential, E,, = -740 mV vs. SCE [23,24]. Several days were usually required for a new or freshly polished electrode to come to a stationary state. The electrodes were then removed from 0.5 M H,SO, solution, rinsed well with triply distilled water, and put in the medium of interest or stored in an air filled bottle. With subsequent use, electrodes would reach a stationary state current at an active potential in a few hours.
RESULTS

Stationary-state current-voltage data for titanium in 1.0 M HCl, 1.0 M HClO,, and mixtures of HCl and HClO, at a total acidity of 1.0 M are shown in Fig. 1. The current-voltage behavior for the hydrogen evolution reaction (HER) on titanium is included in this figure, using the data of Kelly and Bronstein [24] and corrected to the acidity of this experiment (1.0 M H+). The potential range shown in Fig. 1 encompasses the hydrogen evolution reaction (HER), the active dissolution state of titanium, and the active-passive transition region [23]. In 1.0 M HCl, as the potential is made more positive from open circuit, the current increases with potential until a critical potential is reached. The current then decreases with potential. This critical potential (E,,,) divides the active and the active-passive transition potential regions. The anodic dissolution current density in the active and active-passive transition regions decreases as the concentration of chloride ion decreases at constant acidity. At a chloride ion concentration of less than approximately 0.1 M, the current in this entire potential region is net cathodic. The

322

-400

-500

- 600
E /mV vs. NaSCE

-700

-800

Frg. 1. Plot of the logarithm of the stationary-state current vs. potential for tttanium at an acidity of 1.0 M m HCl. HClO,. and mixtures of the two acids. (0) 1.0 M HCl; (0) 0.5 A4 Cl-; (A) 0.2 M Cl-: (A) 0.1 M Cl-; (0) 1.0 M HClO,. In solutions containing Cl-. the solid lines are the calculated net current assuming a partial cathodic process with a 120 mV per decade current-potential dependence and a partial anodic process given by Kellys equatton [23] for the anodtc dissolution of titanium (fitted by a non-hnear least squares routine). (- - - ) Calculated current for the HER [24].

influence of chloride and other anions on the dissolution of titanium will be discussed in a future publication [19b]. The cathodic current density in 1.0 M HClO, is nearly a factor of 7 larger than the current for the HER in 1.0 M HCl. Kelly and Bronstein found that the current density for the HER reaction is dependent on solution pH and independent of the electrolyte medium (1.0 M chloride or 0.5 M sulfate) [24]. Therefore it does not seem reasonable to expect the HER kinetics to change by such a large factor in going from chloride or sulfate to perchlorate. At potentials greater than -700 mV vs. NaSCE, where the contribution to the net current by the anodic dissolution reaction is minimized, the observed cathodic current density (or that extrapolated from more positive values) increases with increasing concentration of perchlorate. Although the dependence is non-linear, it seems reasonable to attribute this increase in cathodic current density over that of the HER alone to perchlorate reduction. Controlled potential electrolysis experiments in 1.0 M HClO, with analysis of the solution for chloride ion indicate that perchlorate is reduced. Although these

323 TABLE Current solution E/(mV - 625 - 600 - 525 1 den&es volume. and yields of chlonde V = 100 ml 103[Cl_ J/M 1.27 1.61 i 0.1 ion m 1.0 M HCIO, (30C). Electrode area, A =12.73 cm;

vs. NaSCE)

10-%/s 3.47 7.7 8.53

r/pAcrn-* 22 13 _

I,,,,./~A 21 17 -3

cm-

a Determmed from the coulometnc b Difference between the observed

relationship. I = (nFV[CII)/( At ). current density and that measured for the HER m ref. 24.

experiments used a relatively large electrode, the electrolysis required a few days to build up sufficient product. The yield of chloride ion as a function of potential in 1.0 M HClO, is reported in Table 1. The current was recorded continuously and did not vary more than f4%. The difference between the observed current density and that calculated for the HER alone is also given in Table 1. The current density for perchlorate reduction was calculated from a Coulombs law expression, assuming the stoichiometry is given by eqn. (1). ClOi + 8 e--t 8 H+-t Cl-+ 4 H,O (1)

The current density, determined from chloride analysis, is in good agreement with the difference in the observed current density and that calculated for the HER alone. The polarization curves in Fig. 1 can be accounted for as the summation of three partial reactions: anodic dissolution of titanium, hydrogen evolution, and reduction of perchlorate to chloride. Several groups have reported [23,25,26] that the current density for titanium corrosion in H,SO, or HCl solutions is independent of stirring rate. It was therefore somewhat surprising to find a slight stirring dependence of the current in HClO, solutions. At potentials between - 500 and -600 mV vs. NaSCE, a net cathodic current increased with decreasing stirring rate. This stirring dependence becomes insignificant, however, in the presence of a sufficient concentration of activating anion such as chloride, sulfate, or oxalate. Preliminary experiments [27] with a titanium rotating disk electrode indicate the dependence of the current on rotational frequency saturates at several hundred rpm. The mass transport provided by a rotating spin-bar at cylindrical electrodes was sufficient to obtain a limiting current. As will be discussed later, this influence of mass transport on the current density seems to be due to the mass transport limited reaction of an intermediate in the reduction of perchlorate [27a]. The influence of perchloric acid on the passivation of titanium electrodes has been reported by Romanushkina et al. [15], in HClO, + HCl mixtures. Their study indicated that perchlorate was not reduced directly at titanium, but rather the influence of increasing the concentration of perchloric acid is to increase the anodic dissolution of titanium. The corrosion product is aquo Ti(II1) ions which are oxidized to Ti(IV) by perchlorate.

324 8

Ti(II1) + ClOT + 8 H +-,8Ti(IV)+Cl-+4H,O

(2)

The product of reaction (2) is Ti(IV) ions which are reduced at these potentials to Ti(III) which can again react with perchlorate ions. Passivation of titanium occurs as Ti(IV) ions build up in solution, and the current density for the Ti(IV) reduction reaction exceeds the critical current density for passivation. The reduction of perchlorate to chloride is a Ti(III)-Ti(IV) catalyzed process given by the reaction sequence of equations (3), (2) and (4). Ti-z-Ti(II1) Ti(IV)+$-Ti(II1) (3) (4)

Reaction (3) is the electrochemical dissolution of titanium metal to form Ti(II1) ions in solution, and reaction 4 is the electrochemical reduction of Ti(IV) ions to form Ti(II1) ions. The rates of reactions (2), (3), and (4) are either known or an upper limit can be estimated for the conditions of this work. The time required to generate a concentration of chloride ion, with a relatively large volume of solution, should allow a distinction to be made between the catalyzed pathway and direct reduction of perchlorate at the electrode surface. The kinetics for reaction (2) have been reported [28], and the indicated rate expression is given by eqn. (5). - 2 [Ti(III)] = 8-$[Cll] = k,[Ti(III)] [ClO;]

(5)

The rate of formation of Ti(II1) in solution by reaction (3) is i,A/3FV where i, is the anodic current density for the dissolution of titanium (reaction 3), V is the solution volume, and A is the electrode area. The rate of disappearance of Ti(II1) by reaction (2) is k,,,[Ti(III)] where k,, is the pseudo-first-order rate constant for reaction (2) (kobs = k,[ClO;]). The rate of appearance of Ti(II1) in solution by reaction (4) is (k,A/V)[Ti(IV)] where k, is the electrochemical rate constant for reduction of Ti(IV) at the electrode surface. The rate of appearance of Ti(IV) by reaction 2 is k,,,[Ti(III)]; the rate of disappearance of Ti(IV) by reaction 4 is (k,A/V)[Ti(IV)]. These considerations lead to the coupled differential eqns. (6) and (7). z 2 [Ti(III)] [Ti(IV)] = (i,A/3FV) = k,,,[Ti(III)] - k,,,[Ti(III)] + (k,A/V)[Ti(IV)]

- (k,A/V)[Ti(IV)]

These coupled differential equations were solved by the method of Laplace transformation, with the boundary condition [Ti(III)] = [Ti(IV)] = 0 at time zero. S[=E(III>] = (i,A/3FV)l/s s[Z(IV)] = kobs[~(III)] - kobs[~(III)] -(k4A/V)[n(IV)] +(k,A/V)[E(IV)] (8) (9)

325

where s is the Laplace variable and a superscript bar indicates the concentration the Laplace plane. Equation (9) is solved for [z(IV)] as a function of [E(III)]. [n(Iv)] Equation = { kobs/(S + k,A/lr)} (8) with [z(IV)] [Ti(III)]

in

(10)

given by eqn. (lo), is solved for [Ti(III)].

Inverse Laplace form [32]. LP[(s+a)/s2(.s+~)] Therefore (13).

transformation

of eqn. (11) was obtained

from a tabulated

standard

=(1/a-a/cy)(l-e-*)+(a/cy)t of Ti(II1) ions as a function

(12) of time is given by eqn.

the concentration

k obs tkobs + k,W)

[1+exp[-(k,,,+k,A/V)1]]

(b%V) t + (kobs + M/V)

(13)

The exponential term in equation 13 becomes insignificant at the time of observation in Table 1, and it can be neglected. This approximate expression for [Ti(III)] as a function of time is substituted in eqn. (14) and then integrated using the boundary condition [Cl-] = 0 at time zero to give an expression for [Cl-] which can be compared with the experimental values. $[Cll] = ik,,,[Ti(III)] (14)

Lc1-

i,k,tdt
24FV

koba i
( kobs + k,A/V)*

+1

(k,A/V)t

2 (kotx + k,A/V)

(15)

The current density for the anodic dissolution of Ti is less than 1 x 10-h A cm- at -600 mV vs. NaSCE (see Fig. 1). The rate constant for reduction of Ti(IV) to Ti(II1) is dependent on the concentration of complexing anions. as noted in a following section, and a conservative estimate is 1 X 1O-6 cm SK. The pseudo-firstorder rate constant for the reaction of Ti(II1) with perchlorate in 1.0 M HCIO, is 3 x lo- s- [18,28]. Using the other parameters appropriate for the experiments in Table 1, the calculated concentration of chloride ion after 8 X lo5 s is - 5 X lo- M, a value far lower than the observed concentration. Thus, perchlorate reduction in these experiments occurs by direct reduction at the electrode rather than by a catalyzed path involving Ti solution species. It should be noted that the catalyzed pathway is still present, but it is too slow to account for the product yields in these

326

experiments. If the electrode area to solution volume ratio (A/V) were increased by a factor of 100, the catalyzed pathway would be significant. The total cathodic currents in Fig. 1 are the sums of the partial currents for the HER and perchlorate reduction. The acid concentration is constant at 1.0 M, and the partial current for the HER is assumed to be constant. As previously noted. the partial current for perchlorate reduction is not linearly dependent on perchlorate concentration. This current seems to be inversely dependent on the concentration of chloride ion. When the concentration of chloride exceeds about 0.2 M, the partial cathodic current decreases to the value for the HER alone. Adsorption of perchlorate on the surface seems to be a necessary condition for its reduction to occur. A sufficient concentration of chloride ion will displace the adsorbed perchlorate from the electrode surface. The current-voltage data in Fig. 1 also show that the decrease in the partial cathodic process for perchlorate reduction is paralleled by an increase in the partial anodic current for titanium dissolution with an increase in chloride concentration. Reduction of Ti(IV) at a titanium electrode

An estimate of the rate at which Ti(IV) in solution is electrochemically reduced to Ti(II1) at an active titanium electrode is needed to gange the effect of the catalyzed pathway for ClOi reduction. This rate constant has been reported in 1.0 M HCl and 0.5 M H,SO, at various potentials by Kelly [29,30]. The rate constant was determined by introducing a relatively large concentration of Ti(IV) to the test compartment of the electrochemical cell, and measuring the current before and after the addition of Ti(IV). The difference in current density as a function of concentration of Ti(IV) is linear, demonstrating the reaction is first-order in Ti(IV). In this work, solutions of Ti(IV) in 1.0 M HCIO, were too unstable to precipitation of hydrous oxides of titanium for this method to be successful. Small amounts of sulfuric acid stabilize dilute solutions of Ti(IV), probably by complexation. and the rate constant for Ti(IV) reduction was measured at several concentrations of sulfuric acid. The value reported at 30C in 0.5 M H,SO, is 7.9 x 10K5 cm ss [29] at - 600 mV vs. SCE: rate constants measured in this work were 2.4 X 10m5 cm ss and 2.8 X lop6 at -600 mV vs. NaSCE in 1.0 M HClO, containing 1 X lo- M H,SO, and 1 X 1O-4 M H,SO, respectively. An estimate of the rate constant, extrapolated to 10e6 M H,SO, by a log-log plot of the rate constant versus concentration of H,SO,, is in the range 3 X 10-cm s-l to 1 X 10e6 cm s-. Reduction of chlorate and the The at a of

The good agreement between the current density for perchlorate reduction the appearance of chloride in solution indicates that the concentration of intermediate species ClO, does not build up to a significant level in solution. reaction scheme for perchlorate reduction requires that chlorate be reduced faster rate. This was confirmed by independent experiments in which solutions

321 TABLE 2 for the reaction of perchlorate k/cm s -la c10; (1 M HClO,) 1 7x10m3 1.2x10-3 o.9x1o-3 c10; (1 M HCl) 1.6x 10-j 1.1 x10-3 1.4x 10-7 and chlorate Ions at titamum (30C)

Rate constants E/mV

vs. NaSCE

ClO,(1 M HClO,) - 600 - 550 - 500 A The reactions are assumed 2.2x10-s 7.2X10- 1.3x1o-9 to be first order in [ClO;

] or (ClO; 1, I = nFAkc.

NaClO, in 1.0 A4 H+ were added to the test electrode compartment, and the difference in current was attributed to chlorate reduction. The reductions of chlorate and perchlorate are assumed to be first order. and the stoichiometries are given by eqns. (1) and (16). ClO; + 6 e-+ 6 H++ Cl-+ 3 H,O (16)

The rate constants are summarized in Table 2. The current for reduction of ClO; is highly stirring dependent and only weakly dependent on potential or ionic medium. The dependence of the current on mass transport was probably not completely removed in these experiments. The stirring conditions of these experiments produce a diffusion layer thickness, 6, of 1O-2 to 1O-3 cm. The diffusion controlled rate constant is D/S, 10e3 to 1O-2 cm-. Thus the apparent rate constants for ClO; reductions in Table 2 represent a lower limit only. It can be concluded, however, that the rate constant for ClO, reduction is at least 5 orders of magnitude greater than the rate constant for ClO,- reduction. The reductions of ClO, and ClO,- are presumed to occur stepwise with the lower oxidation state 0x0 species (ClO_, CIO- ) occurring as intermediates.
DISCUSSION

In a 1966 review article, Pearson stated that the perchlorate ion shows no oxidizing properties in solution [31]. This statement is still correct in the sense that perchlorate ion must be bound to another species or a surface before it can be reduced to a lower valence of chlorine. Taube [18] has recently reviewed the evidence for reduction of ClOc in solution. Perchlorate has no low lying unfilled electronic orbitals, and thus it cannot easily accommodate an additional electron to form the C10i2 ion. Reduction necessarily occurs via transfer of an oxygen atom. The process in reactions (17) and (18) can occur if some species is available to stabilize 02Cloy ClO; + e-j + 2 e-+ ClO, + 02ClO; + 02(17) (18)

328

Protons can stabilize O-. and this may account for the activity of concentrated solutions of HClO, as oxidizing agents. Taube [IS] notes that metal ions are the only species known to reduce Cloy at reasonable rates at ambient temperatures. The most active ions are those forming yl ions such as V(IV). Thus the reduction of Cloy by V(II1) and Ti(II1) are understandable if oxygen atom transfer occurs. ClO,- + M( H?O);f + + ClO, + MO( H,O);+ Reduction reactant. ClO, can also occur by a two-electron +UO; +2H+ + H,O mechanism M = Ti.V (19) as

such as that for U(II1)

+ U(II1) + Hz0 -ClO,

(20)

The reduction of ClO; on an active titanium electrode can be taken as evidence that surface bound species are responsible for this rate enhancement. This enhanced rate of perchlorate reduction is consistent with the kinetic model proposed by Kelly [23] for the dissolution of titanium. The surface of Ti in the active and active-passive transition potential regions is covered by a monolayer of adsorbed oxy or hydroxy Ti(II), Ti(III), and Ti(IV) species. At potentials negative to E,,,, Ti(I1) is the predominant species. The surface has an oxygen vacancy at least, and the potential dependence (120 mV per decade Tafel slope) of perchlorate reduction is consistent with a rate limiting one-electron transfer reaction, assuming a transfer coefficient of 0.5. The reaction of perchlorate with the adsorbed Ti(I1) species by transfer of an oxygen atom is indicated as ClO; + (Ti(II)),+ + Hi+ H++ (Ti(III)OH),+ (Ti(II))I+ + ClO, + H,O species. The net reaction is reduction (21) (22) of

(Ti(III)OH),:+

+ e-+

where subscript a indicates an adsorbed Cloy to ClO, catalyzed by the surface. Cloy + 2 H++ em--, ClO, + H,O

(23) pathway mentioned above or a two-electron

Either the one-electron reduction reduction pathway is possible. ClOL + (TiOH): (Ti(OH),)z+ + H+-, (Ti(OH),)t+

+ ClO, + H,O

(24) (25)

+ 2 e-+

H+-+ (TiOH):

The relative stability of the intermediate ClO, will determine whether the reaction will proceed via the one-electron pathway or directly to ClO; via the two-electron pathway. The same mechanistic arguments that have been made for ClO< reduction apply to ClO, reduction. For the one-electron pathway, ClO, is the indicated product, and it is a relatively stable species. This may account for the greatly enhanced reaction rate of ClO; over ClOL. The stirring dependence of the current mentioned earlier is probably due to the mass transport-dependent reduction of either ClO, or ClO, at the surface. It may be possible to distinguish between the

329

one-electron and two-electron reduction pathways by a rotating disk electrode investigation of the ClOL and ClO, reduction reactions [27b]. The rate of perchlorate reduction is retarded by a significant concentration of other anions such as Clby a competitive adsorption process. Reactions (21) and (24) may involve the preliminary adsorption of ClO; at a hydroxy Ti(II) surface species in a reaction such as eqn. (26). (Ti(OH),) u + ClO; + Ht + (Ti(OH)(ClO,)). + H,O (26)

The dependence of the current on stirring disappears with an increase in the concentration of activating anions. This observation is consistent with an adsorption process in which these anions (chloride. sulfate, or oxalate) compete with perchlorate for active sites.
ACKNOWLEDGEMENT

Research sponsored by the Division of Materials Sciences. Office of Basic Energy Sciences, U.S. Department of Energy under contract DE-AC05840R21400 with the Martin Marietta Energy Systems, Inc.
REFERENCES 1 (a) RN. Adams, Electrochemistry at Solid Electrodes. Marcel Dekker. New York. 1979. p 29: (b) C K. Mann m A.J. Bard (Ed.). Electroanalytical Chemistry. Vol 3. Marcel Dekker. New York. 1969, p. 57. 2 S.Ya. Vasma and O.A. Petrlt. Sov. Electrochem , 6 (1969) 231. 3 G. Horanyl and G. V&es. J. Electroanal. Chem.. 64 (1975) 252. 4 (a) G HorBnyl and G. Vertes, Inorg. Nucl. Chem Lett . IO (1974) 767; (b) G Vertes and G. HorBnyl. J. Electroanal. Chem.. 54 (1974) 445. (c) G Horinyl and G. Vtrtes, J Electroanal Chem.. 63 (1975) 359. 5 M.J. Gonzales TeJera and F. Colom Polo. An. Quim.. 80 (1984) 219. 6 M.T. Mouhandeas, F. Chassagneux and 0. Vtttorl. C. R. Acad. SCI.. Ser. C. 290 (1980) 417. 7 L. KISS. M. L-VarsBnyl and E. Dud&s. Acta Chum. (Budapest). 79 (1973) 73 8 J. Pamot and J. Augustynski, Electrochlm. Acta. 20 (1975) 747. 9 A.I. Tsinman, L.M. Pischlk and G.L. Makovel, Sov. Electrochem.. 11 (1975) 1598 10 (a) J.B. Mathleu. H.J. Mathleu and D. Landolt. J. Electrochem. Sot, 125 (1978) 1039. (b) J B. Mathieu and D. Landolt, J. Electrochem. Sot.. 125 (1978) 1044. 11 L E. Tsygankova. V.I. Vlgdorovich, T.V. Korneeva and E.K. Oahe. Sov Electrochem.. 19 (1983) 99. I2 J.S. Booth, H. Hamzah and A.T Kuhn. Electrochlm Acta.. 25 (1980) 1347. 13 I.E. Veselovskaya. E.M. Kuchmsku and L.V. Morochko, J. Appl. Chem USSR, 37 (1964) 85. 14 T. Muwm and G. Falta in A.J. Bard (Ed.), Encyclopedla of Electrochemistry of the Elements. Vol. 1. Marcel Dekker, New York, 1973. p. 1. and references cited therein. 15 A.E. Romanushkina. M.V. Mamyhkhma and S.Ya. Maer. Sov. Prot. Met., 12 (1974) 13. 16 (a) V.A. Vekslina, V.V. Vashchenko and V F. Toropova, Sov. Electrochem., 8 (1972) 1445; (b) Yu.S. Mdyavsku. Sov. Electrochem.. 10 (1974) 435. 17 SK. Lahr. H.O. Fmklea and F.A Schultz, J Electroanal. Chem.. 163 (1984) 237. and references cited therem 18 H. Taube m D.B. Rorabacher and J.F Endxott (Eds ). Mechanwc Aspects of Inorgamc ReactIons. ACS Symposium Series No. 198. 1982, p. 151. 19 (a) G.M. Brown. Paper 60 presented at the I63rd Electrochemical Society Meetmg, San Francisco. CA, Vol. 83-l. The Electrochemical Society. Prmceton. NJ, 1983. p 93; (b) G.M. Brown, manuscript

330 m preparation. 20 I.R. Wilson in C H. Bamford and C.F.H. Tlpper (Eds.). Comprehenslve Chermcal Kinetics, Vol. 6, Elsevier. Amsterdam, 1972, p. 375. 21 A.I. Vogel. Quantitative Inorgamc Analysis, 3rd ed., Longmans, London. 1961, p. 266. 22 Z. Marczenko. Spectrophotometric Determmation of Elements, Ellis Horwood, ChIchester, 1976, p. 556. 23 E.J Kelly m J.OM. Bockns, B.E. Conway and R.E. White (Eds.). Modern Aspects of ElectrochemIstry. Vol. 14, Plenum Press. New York, 1982. p. 319. 24 E.J. Kelly and H.R. Bronstem, J. Electrochem. Sot., 131 (1984) 2232 25 R D. Armstrong and R.E. Firman, J. Electroanal. Chem.. 34 (1972) 391. 26 A. Caprani and J.P. Frayret. Electrochim. Acta. 24 (1979) 835. 27 (a) F.A. Posey and G.M Brown. Chem. DIV. Annu. Prog. Rep. January 31. 1985, ORNL-6152. p. 125: (b) F.A. Posey and G.M. Brown. work in progress. 28 V.W Cope, R.G. Miller and R.T.M. Fraser, J. Chem. Sot., A. (1967) 301. 29 E.J. Kelly, J. Electrochem. Sot.. 123 (1976) 162. 30 E.J Kelly, J Electrochem. Sot.. 126 (1979) 2064. 31 G.S. Pearson, Adv. Inorg. Chem. Radlochem.. 8 (1966) 177. 32 G.E. Roberts and H. Kaufman, Table of Laplace Transforms. W.B. Saunders. Pluladelphia. 1966. p. 185.

Vous aimerez peut-être aussi