Vous êtes sur la page 1sur 14

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/262142961

The Kinetics of the Electrochemical Oxidation of


Cyanide on PbO2-Coated Stainless Steel

Article in Zeitschrift fur Metallkunde January 2000

CITATIONS READS

0 17

1 author:

G. Cifuentes
University of Santiago, Chile
83 PUBLICATIONS 153 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Uranio View project

Fluid Dynamic Modelling of the Operational System of Capture of Gases for the Copper Mattes
Conversion Process View project

All content following this page was uploaded by G. Cifuentes on 22 December 2016.

The user has requested enhancement of the downloaded file.


THE KINETICS OF THE ELECTROCHEMICAL OXIDATION OF CYANIDE ON
PbO2-COATED STAINLESS STEEL.

G. Cifuentes*, J. Torrealba* , L. Cifuentes** and R. Kammel***

* Departamento de Ingeniera Metalrgica, Universidad de Santiago de Chile, Av. Lib. B.


OHiggins 3363, Santiago, Chile.
** Departamento de Ingeniera de Minas, Universidad de Chile, Tupper 2069, Santiago, Chile.
***Institut fr Metallurgie, Technische Universitt Berlin, Strae des 17. Juni 135, 10623
Berlin, Germany.

Abstract

The kinetics of the electrolytic oxidation of cyanide to cyanate on PbO2-coated stainless


steel anodes have been studied by both experimental and theoretical means. The anodes
were based on 316 stainless steel and performed in a satisfactory fashion. The studied KCN
solutions exhibited a constant pH = 11 value. Values for all the kinetic parameters have
been determined and a rate law for the studied reaction is proposed.

1 Introduction

The toxicity of cyanide is a serious problem for both the metal finishing and metallurgical.
Industries. Chlorination [1] is normally used as a way of oxidizing residual cyanide to
cyanate, which later reacts to form ammonium carbonate according to:

NaCN + NaOCl + H2O ClCN + 2NaOH (1)

ClCN + 2NaOH NaOCN + NaCl + H2O (2)

2NaOCN + 5H2O (NH4)2CO3 + 2NaOH + CO2 (3)

However, among the disadvantages presented by this method is the formation of the ClCN
intermediate, which is as poisonous as cyanide and exhibits low solubility in water. An
alternative to this procedure is the electrochemical path, i.e., the electrolytic oxidation of
cyanide to cyanate, which does not produce toxic intermediate compounds. This sort of
treatment has been extensively studied since the 40s [2]. Graphite and stainless steel anodes
have been used; current efficiency was low and the consumption of anodic material was
high. Lead (IV) oxide layers were tested, but they showed to be brittle and weakly bonded to
the substrate. Recent work [3] has shown that new methods of producing Lead (IV) oxide
layers on stainless steel afford a much better anode performance. In the present work, the
kinetics of cyanide oxidation on such anodes is studied and the kinetic parameters and
reaction mechanism are established.

2 Experimental

The experimental work was carried out with PbO2-coated stainless steel anodes whose
dimensions and preparation method have been given elsewhere [3]. The solutions were aqueous
KCN of pH=11. Cyanide concentrations were 0.5, 1, 2 and 4 gl-1. The behaviour of the anodic
current density was monitored at constant potential until a constant c.d. was achieved. The
potential was then increased by 30 mV and the anodic c.d. behaviour was again determined. This
procedure was repeated to cover the range between the rest potential and the oxygen evolution
potential (200 and 600 mV against SCE). A Radiometer Copenhagen PGP 201 potentiostat-
galvanostat with Voltamaster 1.0 software was used. The anodes were pre-conditioned by
immersing them in diluted hydrogen peroxide during two hours. During the experiments the
anodes rotated at 100 rpm and the solution temperature was 30 C.

3 Results

Fig. 1 shows the constant potential curves for cyanide concentrations of 4, 2, 1 y 0.5 gl-1
(grams per liter). The oxidation rate increased with increasing cyanide concentration. Tables
I and II give the results of the regressions calculated by the Voltamaster program.

Table I Quadratic regressions for different cyanide concentrations.


[CN-] Quadratic interpolation Correlation
[gl-1] i = aV2+bV+c coefficients, R2
V in [mV/SCE]
i in [mA/cm2]
4 I = 7.40E-6 V2 + 1.10E-2 V 2.4895 0.99
2 I = 6.96E-6 V2 + 7.40E-3 V 1.8559 0.99
1 I = 3.31E-6 V2 + 4.96E-3 V 1.1615 0.99
0.5 I = 2.97E-6 V2 + 2.89E-3 V 7.70E-1 0.99

4
i (mA/cm2)

0
160 200 240 280 320 360 400 440 480 520 560 600

mV / SCE

0.5 gl-1 1gl-1 2 gl-1 4 gl-1

Figure 1 i vs. curves for the 200-600 mV/E SCE range.

Current density vs. Concentration data were obtained at 300, 350, 400, 450, 500, 550 and
600 mV/SCE. Results are shown in Fig. 2. The slopes allow the calculation of the reaction
order with respect to cyanide. Tables II and III show the regression results.
1

m600 = 0.5958 m450 = 0.6164 m300 = 0.6641


m550 = 0.6017 m400 = 0.6264 m250 = 0.7299
0.8 m500 = 0.6084 m350 = 0.6404 m230 = 0.8284

0.6

0.4

0.2
230
250

Log (ia [mA/cm2])


300
0
350
400
-0.2 450
500
550
-0.4
600

-0.6

-0.8

-1

-1.2
-1.8 -1.6 -1.4 -1.2 -1 -0.8

Log ([CN-] [M])

Figure 2: Determination of slopes in Fig. 1, 300-600 mV/SCE range.

Table II Regression results for log i (mA/cm2) v/s log [CN-] curves for a range of potential
values.
mV Linear interpolation Correlation
[SCE] Log(ia) = m Log [CN-]+ n coefficient, R2
[CN-] is given as molarity
ia in [ mA/cm2 ]
600 Log (ia) = 0.5983 Log [CN-] + 1.3383 0.98
550 Log (ia) = 0.6032 Log [CN-] + 1.2746 0.98
500 Log (ia) = 0.6093 Log [CN-] + 1.2028 0.98
450 Log (ia) = 0.6172 Log [CN-] + 1.1200 0.99
400 Log (ia) = 0.6282 Log [CN-] + 1.0220 0.99
350 Log (ia) = 0.6455 Log [CN-] + 0.9013 0.99
300 Log (ia) = 0.6793 Log [CN-] + 0.7441 0.99
Figures 3 to 6 show the Tafel curves obtained experimentally for 0.5, 1, 2 and 4 gl-1 CN-. .
Tables III and IV show results for Tafel slopes (Ba), polarization resistance (Rp) and
simmetry factor (), obtained from Figs. 3 to 6.

-1.5
Log ( ia [ mA/cm2 ] )

-2

-2.5 0.5 gpl


Tafel

-3

-3.5

-4
0 4 8 12 16
[ mV ]

Figure 3: Tafel curve for [CN-] = 0.5 gl-1.

-0.3

-0.8

-1.3
Log ( ia [mA/cm2] )

-1.8
1 gpl
Tafel
-2.3

-2.8

-3.3

-3.8
0 8 16 24 32
[ mV ]

Figure 4: Tafel curve for [CN-] = 1 gl-1.


-0.3

-0.8

-1.3
Log ( ia [ mA/cm2 ] )

-1.8
2 gpl
Tafel
-2.3

-2.8

-3.3

-3.8
0 5 10 15 20 25 30
[ mV ]

Figure 5: Tafel curve for [CN-] = 2 gl-1.

-0.3

-0.8
Log ( ia [ mA/cm2 ] )

-1.3

4 gpl
-1.8
Tafel

-2.3

-2.8

-3.3
0 4 8 12 16 20 24
[ mV ]

Figure 6: Tafel curve for [CN-] = 4 gl-1.

From the results presented in Table II it can be concluded that the activation (charge
transfer) control range is 0 < < 30 mV and at higher overpotentials there is mixed control.
The average value for the reaction order is 0.6 (Table II).
Table III: Regressions for Tafel curves for a range of cyanide concentrations.
[CN-] Linear interpolation Correlation
gl-1 Log(ia) = m + Log io coefficient, R2
in [mV], ia and io in [ mA/cm2 ]
0.5 Log (ia) = 0.0371 1.7575 0.99

1 Log (ia) = 0.0153 1.1759 0.99

2 Log (ia) = 0.0178 - 1.0338 0.99

4 Log (ia) = 0.0225 1.0033 0.99

Table IV: Experimental values for Tafel slopes Ba, exchange current density io, polarisation
resistance Rp and symmetry factor for a range of cyanide concentrations.
[CN-] Ba io Rp
gl-1 [mV / dec] [ mA/cm2 ] [Ohmcm2]
0,5 26.9 0.017 1493.8 0.45
1 65.4 0.066 391.5 1.08
2 56.2 0.092 282.2 0.93
4 44.4 0.099 263.1 0.74

The mean value for the Tafel slopes is 50mV/dec or 2RT/F mV/dec.

4 Modelling the electrochemical kinetics

In agreement with the experimental results, a mixed-control type equation was used to fit the
data.
i o nF (4)
[
i = t CN ]
Ord RT
i
where 1 o
i lox
: overpotential (V)
: symmetry factor
t : fitting parameter
[CN-]: cyanide concentration, molar
Ord: reaction order for cyanide oxidation
io: exchange current density [mA/cm2].
n: number of electrons in the reaction
ilox: limiting c.d. for cyanide oxidation [mA/cm2].

The exchange current density (io) and the limiting current density for the cyanide oxidation
reaction (ilox) are calculated by the following iteration algorithm, where ireal is an
experimentally determined c.d. value:

i predicted i real = where 0

with
[ ]
i predicted = f ( CN , T , i 0 , i Lox , , i real )

ireal, [CN-], T and are known; io and iLox are calculated by the model.

Data used for model calculations were as follows: T = 303 K, R = 1.987 Cal/mol/K, F =
23,060 Cal/V/eq. The starting value for the reaction order was 0.6, which was obtained from
Table II. The calculation was iterated until the difference between the measured and
predicted current density values was minimised. The modelling results are shown in Tables
V and VI.
Table V: Kinetic parameters obtained from the model.
[CN-] Order t n io ilox
gl-1 [mA/cm2] [mA/cm2]
0.5 0.2 10.138 1 0.018 2.00 0.5
1 0.2 10.015 1 0.070 5.00 0.5
2 0.2 10.022 1 0.091 7.20 0.5
4 0.2 10.178 1 0.099 12.00 0.5
Table VI: Tafel slopes Ba and polarization resistances obtained from the model for a range of
concentrations.
[CN-] Ba Rp
gl-1 [mV/dec] [Ohmcm2]
0.5 120.3 1457.2
1 120.3 373.1
2 120.3 288.6
4 120.3 263.8
Figures 7, 8, 9 and 10 show a comparison of experimental and model predicted values.

2.50

1.40

1.20 2.00

1.00

1.50

i (mA/cm2)
i (mA/cm2)

0.80

0.60 1.00

0.40

0.50

0.20

0.00
0.00
0 100 200 300
0 100 200 300
(mV) (mV)
2
i experimental (mA/cm ) i simulated (mA/cm 2)
i experimental (mA/cm2) i simulated (mA/cm2)

Figure 7: i vs. curve, [CN-]=0.5gl-1. Figure 8: i vs. curve, [CN-]=1gl-1.

4.00 3 . 50

3.50
3 . 00

3.00
2 . 50
(m A/ cm 2)

2.50
2 . 00
i (mA/cm2)

2.00
i

1 . 50

1.50

1 . 00

1.00

0 . 50

0.50

0 . 00

0.00 0 100 200 300


0 100 200 300
( mV )
(mV) 2
i ex pe ri me nt al (m A/ cm ) i s i m u l a t e d ( m A / c m 2)
i experimental (mA/cm2) i simulated (mA/cm2)

Figure 9: i vs. curve, [CN-]=2 gl-1 Figure 10: i vs. curve, [CN-]= 4 gl-1.
5 Discussion

Figures 7 to 11 clearly show that the model predictions for the reaction rate as a function of
the anodic overpotential give a very good fit. Arikado et al. (1) have produced a model
based on results obtained on graphite electrodes, where the cyanide oxidation rate law is:

i = 2 K a a CN exp( FE / RT )

where Ka is the kinetic constant for activation control, aCN- is the cyanide activity in
solution, is the symmetry factor (a value of 0.5 is assumed), F is Faraday's constant, E is
the electrode potential, R is the universal gas constant and T is the temperature. Arikado's
model predicts reaction orders of 1 and 0 for CN- and OH- respectively.
The present model is different from Arikado's a) in the form of the rate law (see
above); b) in the fact the the parameters are easier to determine; c) being fundamental
kinetic parameters, io and iLox give a direct view of the behaviour of the system. In addition,
the anodic limiting current density is also very significant from an operational point of view;
d) the predicted reaction orders for CN- (expressed as concentration rather than activity) and
for OH- are 0.2 and 0 respectively; this means that the reaction rate is barely affected by
variations in CN- concentration. This result confirms previously published results (2), but it
differs from the first order rate law given by Arikado (1).
It is also interesting to compare the model results with those obtained directly from
the experimental curves. In the case of the determination of the reaction order for cyanide, it
is evident from figures 3 to 6 that the slope value is quite difficult to determine, especially
as the cyanide concentration increases. The value obtained from the model (0.2) is different
from the one obtained from the Tafel slopes (0.6), but it is the value given by the model
which produces the best agreement with experimental reaction rate data. Comparing Tables
II and V it becomes clear that the Tafel slopes vary from point to point as the potential
varies, while the model predicts a constant 0.2 reaction order in the 0.5 to 4 gl-1 cyanide
concentration range.
It is also useful to compare the values for the Tafel slopes in Tables IV and VI. The
experimentally obtained values (Fig. IV) range from 26.9 mV/dec to 65.4 mV/dec, while the
model produced value (Table VI) is constant at 120.3 mV/dec. As a) the latter agrees with
published results (2) and b) model calculated rates closely agree with i vs. plots, it can be
concluded that the model predictions are more reliable than those values obtained from the
experimental curves via regression techniques.
It is worth noting that the proposed equation represents a mathematical procedure to
fit the experimental data in a way which allows better process control than previously
published expressions. It is not proof of any particular reaction mechanism.

6 Conclusions

1. Two distinct rate control zones were established: an activation (charge-transfer) control
zone and a mixed control one.

2. A technique to obtain kinetic parameters via numerical modelling in the mixed control
zone was developed and applied.

3. The rate law for the electrolytic oxidation of cyanide is:

[
i = 10 CN ]
0.2 F
2RT
io
i

1 o
i l ox

4. The cyanide reaction order predicted by the model is 0.2. The predicted cyanide oxidation
rates agree very closely with experimental i vs. curves.
5. The kinetic parameter values calculated by the model for the cyanide oxidation reaction
are more reliable than those values obtained from the experimental curves via regression
techniques.
6. The proposed equation allows better process control than previously published
expressions, but it is not proof of any particular reaction mechanism.

Acknowledgements

The authors would like to thank the National Committee for Scientific and
Technological Research (CONICYT, Chile) for supporting this work by means of
FONDECYT project No. 1950532.

Literature

1. T. Arikado, C. Iwakura, H. Yoneyama and H. Tamura, Anodic oxidation of potassium cyanide


on the graphite electrode, Electrochimica Acta, 21 (1976), 1021-1027.

2. F. Hine, M. Yasuda, T. Iida and Y. Ogata, On the oxidation of cyanide solutions with lead
dioxide coated anode, Electrochimica Acta, 31 (1986), 1389-1395.

3. G. Cifuentes, L. Cifuentes, R. Kammel, J. Torrealba and A. Campi, New methods to produce


electrolytic Lead (IV) dioxide coatings on titanium and stainless steel, Z. Metallkunde , 89
(1998), 363-367.

View publication stats

Vous aimerez peut-être aussi