Вы находитесь на странице: 1из 13

J. Electroanal. Chem.

, 177 (1984) 25-31 25


Elsevier Sequoia S.A., Lausanne - Printed in The Netherlands

THREE-DIMENSIONAL NUCLEATION WITH DIFFUSION CONTROLLED


GROWTH

PART II. THE NUCLEATION OF LEAD ON VITREOUS CARBON

J. MOSTANY, J. MOZOTA l and B.R. SCHARIFKER

Depurtumento de Quimiu, Uniwrsidud Simbn Boliwr, Apurtudo 806.59, Corucus 1080. A (Veneruelu)
(Received 2nd March 1984; in revised form 8th May 1984)

ABSTRACT

The nucleation of lead onto vitreous carbon electrodes has been investigated. The number density of
active sites for nucleation was found to vary with the overpotential but not with the concentration of lead
ions in solution. The overpotential and concentration dependence of the nucleation rate were also studied,
and the results interpreted according to both atomistic and classical theories. Due to the small size
obtained for the critical nuclei, the latter was refined to incorporate the effect of the line tension along the
contact line between the three phases involved. However. the atomistic theory seems to afford a better
representation of this electrochemical nucleation process.

INTRODUCTION

Heterogeneous nucleation takes place on a limited number of specific sites on the


surface. This is an assumption made in almost all present day studies of electrochem-
ical nucleation, although little is known about the nature of such active sites, their
number density, or its variation with the concentration of electrodepositing species,
overpotential or temperature. Moreover, there is no clear cut definition in terms of
experimentally measurable quantities for the activity of preferred sites for nuclea-
tion, apart from the intuitive notion that the activity of a preferred site is somehow
related to its specific rate of conversion into a growing nucleus for given experimen-
tal conditions. In spite of the vague definition of the active sites for nucleation and
the lack of direct experimental information about them, some theoretical insight has
been gained through conceptually sound considerations. Thus Markov and Kashchiev
[l] assumed that the number of active sites was dependent on the potential, while
Fletcher and Lwin [2,3] have recently proposed that the activities of preferred sites
are distributed in a spectrum, or activity density function, which is characteristic of
the electrode surface. The basic difference between both models is that for the first

l Present address: Nalco de Venezuela C.A., Apartado 62176, Caracas, Venezuela

0022-0728/84/$03.00 0 1984 Elsevier Sequoia S.A.


26

of them the activity of all preferred sites is uniform at a given overpotential, while
for the latter there exists a hierarchy of activities of preferred sites. Although the
validation of either one of these models through appropriate measurements, e.g. by
recording N(t) curves [2] has been proposed, most commonly the final number of
nuclei obtained in potentiostatic experiments is much lower than the number of
active sites on the surface [4]. Thus, in general, the N(t) behaviour does not reflect
the exhaustion rate of active sites, making it difficult to distinguish through it either
one of these models. For simplicity, we will assume here that at a given overpotential
all preferred sites show the same activity towards nucleation, i.e., we will follow the
uniform model.
The electrodeposition of lead on vitreous carbon has some distinguishing features
from other systems studied so far. On the one hand, the underpotential deposition of
lead on vitreous carbon is now firmly established [5] as well as its electrocatalytic
influence on the reduction of oxygen. On the other hand, three dimensional
nucleation of lead onto vitreous carbon electrodes takes place at considerable rates
even at very low overpotentials [6], i.e. at much lower supersaturations than those
required for the nucleation of other metals such as mercury, silver, cadmium or
copper. This behaviour might become useful in studying the nucleation process over
a wide range of overpotentials in order to (a) test the treatment for the current
transient described in Part I of this series [7] and (b) obtain reliable data to contrast
with the classical, thermodynamic and atom&tic theories of nucleation. In addition,
lead ions remain in solution in the presence of several anions, particularly the
halides, thus making it possible to study their effects on the process of nucleation.

EXPERIMENTAL

Lead deposition was studied from aqueous solutions of Pb(NO,), in the l-20
mM concentration range, using 1 M KNO, as supporting electrolyte. A two
compartment, three electrode cell was used throughout the work. The working
electrode was built from a vitreous carbon rod (VlO, Atomergic Chemetals) and
sealed to the end of a glass tube with epoxy resin (Araldite) in order to expose a flat
surface of 7.07 x lop2 cm2 in area to the solution. The exposed surface was polished
to a mirror finish with alumina down to 0.05 pm on Microcloth (Buehler), and
rinsed with dilute nitric acid and triply distilled water before the experiments. The
secondary electrode was a platinum wire, placed in the same compartment as the
working electrode. In order to polarise the working electrode directly with respect to
the equilibrium potential of metallic lead with its ions in solution, a lead rod,
immersed in the working solution but in a separate compartment, was used as a
reference electrode. Its rest potential was stable and within 1 mV of that of the
Pb/Pb*+ couple. The reference compartment was connected to the main cell through
a Luggin capillary, placed at about 2 mm from the working electrode surface.
All solutions were prepared with triply distilled water and analytical grade
potassium and lead nitrates, used as received. Before each series of potentiostatic
experiments, argon was bubbled through the cell for 30 min, keeping the working
27

electrode at a positive potential of 0.3 V. This potential was chosen after verifying
that the underpotential deposition phenomenon reported by Mayer and Juttner [S]
takes place in the region from 50 to 200 mV. In order to ensure complete dissolution
of the nuclei deposited during the pulses, these were applied at intervals of 1 min
within a series, mantaining a 0.3 V potential between pulses.
The potentiostat and pulse generator were built in our laboratory and provided
control of the potential with fast rise time (< 3.5 ps) and low noise (< 100 pV). The
current transients were fed to a Tektronix 5223 digital oscilloscope via a current
follower and then transferred to hard copy with a Hewlett-Packard 7004-B x-y
recorder. The electrochemical cell, potentiostat and pulse generator were all placed
in a Faraday cage, while the necessary power was provided from power supplies
outside the cage, using shielded cables.

RESULTS AND DISCUSSION

Figure 1 shows a family of current transients obtained with different potential


pulses. Here we observe the previously described characteristics of this type of
system [8], i.e. a rising current followed by a descending portion which corresponds
to linear diffusion to the planar electrode, described by the Cottrell equation. In
order to obtain the number density of active sites on the surface and the nucleation
rate per site, the diffusion coefficient of lead ions in each of the solutions used was
obtained from the slope of plots of i vs. t -I/* from high overpotential experiments,

Fig. 1. Potentiostatic current transients for the nucleation of lead on vitreous carbon from aqueous 0.01 M
Pb(NO,), in 1 M KNO, solutions at the overpotentials indicated.
28

for which the surface concentration of electroactive species is zero, Fig. 2. The
diffusion coefficients thus obtained appear in Table 1. These values are similar to
those reported by Palmisano et al. [6] for Pb *+ in 0.1 M HCl, determined at several
concentrations by potentiostatic and potentiodynamic techniques. Using the values
of D in Table 1, N,, and A were obtained from the current maxima of potentiostatic
experiments by means of the numerical method described by eqns. (36)-(39) of Part
I [7]. The results reported are the averages and standard deviations of three
independent series of experiments for each concentration.

The number of active sites

The dependence of the number density of active sites with overpotential at several
Pb*+ concentrations is shown in Fig. 3. In all cases the number of sites on the
surface that are active towards nucleation increases with overpotential. Also, the
density of sites is approximately constant at a given overpotential, independent of
concentration. For the highest overpotential studied of 300 mV, the number density

IO’1 /A cm-*

50-

40-

0 5 IO 15

Fig. 2. Linear dependence between current and t- II2 for the descending part of current transients
obtained at different concentrations.

TABLE 1
Diffusion coefficient of Pb*+ m aqueous solutions in the presence of 1 M KNO,

lo3 cp,,z+ /mol dm-’ 10’ O/cm’ s-’


1 2.0
5 1.14
10 1.06
20 1.06
29

of sites is of the order of lo9 cm-*, corresponding to a very small fraction of the
electrode surface if we consider 1015 cm-* as the density of geometric sites in a
compact array of lead atoms. The number density of active sites towards nucleation
is even lower than that of the sites for lead adsorption on vitreous carbon [5], which
is about 5 X lOI cm-*. Therefore only one geometric site out of a million is active
towards nucleation, justifying our starting assumption of a limited number of active
sites on the surface.

Non-dimensional plots

Non-dimensional plots of I’/12 vs. t/t,,, can be constructed from the experimen-
tal data. Some of the plots obtained are shown in Fig. 4 for different overpotentials.

I 1n (No/cm-*)
I=
=I
INn No/cm-*)

III
III
111r*
I 20-
I$
11 = II’ II1
II
15-
1I 6
I I

7 /mV q /mV
101 0 'OO
1
0 100 200 300 100 200 300

In ( No/cm-* 1 ) \n(N /cm-*)


20-

I
I01
q /mV
IO
q/mV
0 100 200 300 0 100 200 300

Fig. 3. Dependence of the number density of active sites with overpotential at different concentrations:
(1) 1 m M, (b) 5 mM, (c) 10 mM, (d) 20 m M Pb’+ solutions.
30

Data from each of the solutions studied is included in them and it can be observed
that the most dilute solution approaches the lower curve, corresponding to progres-
sive nucleation, whereas the data obtained from the other solutions are shifted in the
intermediate region of the plots towards the upper curve, which corresponds to
instantaneous nucleation. Within the framework of the treatment for the current
transient presented in Part I, this behaviour is interpreted as a limited availability of
active sites on the surface. The value of I,,,tz’/a, where a = ~FD’/~c/rn’/~, is a
measure of the position of a non-dimensional curve within the instantaneous-pro-
gressive range, since its limiting values are 0.7153 for instantaneous nucleation, and
0.9034 for progressive nucleation (cf. eqns. (31) and (34) in Part I). Table 2 compiles
the values of Z,t,!,(’ /a obtained from the experimental data presented in the
non-dimensional plots of Fig. 4. Although in some cases the values of I,,,t!,(*/a fell
outside the predicted range, possibly due to non-linearities of the nucleation process
or overlapping of double-layer charging and deposit growth currents at high over-
potentials, it is confirmed that the data shift from the limiting situation of progres-
sive nucleation to that of instantaneous nucleation as the concentration of Pb2+ is
increased.

Fig. 4. Non-dimensional plots of potentiostatic experiments at different concentrations: 1 mM (0), 5


mM(O),10mM(+),20mM(a),(a)q=-100mV;(b)~=-200mV;(c)q=-300mV.
31

Nucleation rates

Given that the number of active sites was found to be independent of concentra-
tion, the transition from a progressive nucleation regime in dilute solutions to
instantaneous at higher concentrations indicates that the nucleation rates increase
with concentration. In the classical expression for the steady state rate of heteroge-
neous nucleation [9],
A = N,TD* exp( - AG*/kT) (1)
where I is the Zeldovich factor [lo], D* is the flux of subcritical clusters to critical
nuclei along the size coordinate of the kinetic model and AG* is the reversible work
for nucleation, only the preexponential term is subject to concentration dependence,
through D * . The detailed mechanism of heterogeneous nucleation admits two
possibilities: either nucleation occurs directly by attachment of monomers from the
bulk of the solution to subcritical clusters or it occurs through the attachment of
previously adsorbed intermediates. Only the first of these two mechanisms will show
a concentration dependence of the steady state nucleation rates [ll], because in the
case of adsorbed intermediates, the Gibbs-Duhem relation requires their surface
concentration to be dependent only on supersaturation, expressed here as an
overpotential.
Figure 5 shows the nucleation rates per active site obtained in our analysis of
data, as a function of the overpotential for the solutions studied. Thus the addition
of monomers to form critical nuclei occurs directly from the bulk of the solution
[ll], and under these conditions we can equate D * to the rate of reduction of
metallic ions onto critical clusters on the surface, i.e.,

D* = K’d-“) exp( azFq/RT) (2)


where K’ is a constant, independent of c or n, which contains the area of the critical
clusters as well as their population. According to eqns. (1) and (2) the slope of
straight lines traced through the values of In A corresponding to a given value of n
for different concentrations gives an estimate of the electrochemical transfer coeffi-
cient. The value of 0.3 + 0.1 obtained is within the values reported for the deposition
of lead onto different substrates [12].
Electrochemical phase formation is a special case of heterogeneous nucleation. In

TABLE 2

Values of I,,,tz’/u for the current maxima of the data in Fig. 4.

- 9/mV lo3 c/mol dmm3

1 5 10 20

100 0.888 0.753 0.742 0.732


200 0.954 0.752 0.770 0.746
300 1.037 0.734 0.732 0.630
32

the classical description of heterogeneous nucleation [13] the work of formation of a


critical nucleus is expressed as

AG* = (16ay3/3AG;)+(8) (3)

where y is the interfacial tension of nucleus with its mother phase, AG, is the Gibbs
energy of formation of the new phase per unit volume and +(O) is a function of the
contact angle 19 between the nucleus and the substrate, equal to the ratio of the
volume u of the spherical cap resting on the surface to the volume V of the sphere
with the same radius of curvature, i.e.

&‘)=~/V=(2-3cos8+cos~~)/4 (4)

In the electrochemical case, the supersaturation c/co is related to the the overpoten-
tial by [14],

17= (RT/zF) In( c/co) (5)

where co is the equilibrium monomer concentration at the reversible potential of the


new phase with the solution and c its concentration in the supersaturated system.
The Gibbs energy of formation of the new phase is the product of the charge
transferred per unit volume of the new phase and the potential difference with
respect to the phase equilibrium situation. i.e.

AG, = ( PzF/M >17 (6)

2
t q/mV
I I I

0 100 200

ln( A/s-‘)
8-

6-

4-

2
7j /mV I
t
I I 1
0 100 200

Fig. 5. Dependence of the nucleation rates with overpotential at different concentrations: (a) 1 m M, (b) 5
mM, (c) 10 mM, (d) 20 mM Pb*+ solutions.
33

In these terms, the critical radius is

r* = 2My/pzFg (7)
and the Gibbs energy of formation of the critical nucleus is

AC* = 16ry3M2+( 8)/3( pzF~j)~ (8)


Thus a plot of the logarithm of the nucleation rate as a function of l/n2 should be a
straight line with slope

a In A/a(1/n2) = - 16.rry3M2+( 0)/3( pzF)*kT (9)


This is often not so in studies of electrochemical nucleation and the plots corre-
sponding to the solutions studied here are shown in Fig. 6, where it can be seen that
the slope is a function of overpotential. The dependence of eqn. (9) on overpotential
can be attributed only to variations of y or C#Iwith overpotential, and because the
range of overpotentials studied is only 250 mV, we do not expect the variation of y
to be large. The function C+(O), however, may show important variations with
supersaturation especially at high supersaturations, where the effects of the line
tension between the three phases involved are large due to the small size of the

,. In(A/s-‘1

8t b
%

6 OS&,
o “YQ-L
4

2 rl-9V-z
I_-
0 100 200

10
c Ln(A/s-'1

d
8
6

I
0 100 200

Fig. 6. Plot of In A vs. l/q2 for the solutions studied. Points: experimental data. Continuous lines: best
fits of the data to eqn. (11). (a) 1 mM, (b) 5 mM, (c) 10 mM, (d) 20 m A4 Pb*+ solutions.
34

critical nucleus [15]. Therefore we will assume here the change in ClIn A/a< q- 2,
with 9 is only due to the variation of the contact angle with overpotential.
The surface tension of lead nuclei was estimated from extrapolation to room
temperature of data for the surface tension of lead at temperatures higher than its
melting point, using the Guggenheim-Katayama equation [16],

y=yO(l- T/T,)”
where T, is the critical temperature, y ’ the surface tension at 0 K and n is a
parameter that for most metals is unit [17]. T, was estimated as 1.6 times the melting
temperature. At room temperature, y$” = 500 mJ mP2, assuming that the surface
tension can be equated to the interfacial tension in our system.
In order to obtain the dependence of + with overpotential from the experimental
data, these were empirically adjusted to a power law of the type

In A = a(l/$)’ (11)
from whose analytical derivative a In ,4/a(~-~) was obtained for each value of 17.
The fit of the data to eqn. (11) by linear regression is shown in Fig. 6. Then $I was
obtained for each overpotential and concentration and from it 8 was calculated
from eqn. (4) using Newton’s method. Figure 7 shows the variation of the contact
angle with overpotential. The contact angles of nuclei in equilibrium with their

0 100 200 300

Fig. 7. Contact angle of lead nuclei on vitreous carbon as a function of overpotential and in 1 m A4 (O), 5
mM (A), 10 m M (0) and 20 m M (0) Pb*+ solutions.
35

supersaturated mother phase have a common regression at the equilibrium potential


of the system, of about 8O. This angle is then taken as the macroscopic contact angle
of lead with vitreous carbon in aqueous solution. As the critical nucleus decreases its
size with supersaturation, the effects of line tension become increasingly large, giving
way to the variation of the contact angle observed. A detailed analysis of the effects
of line tension on the microscopic contact angle [15] results in that for small radii the
Young equation must be written
6/r = sin 8 (cos 6, - cos 0) (12)
where 6 is defined as a/y, u being the line tension of the three phase contact line, 8
is the microscopic contact angle of the nucleus and 0, is the macroscopic contact
angle, corresponding to r + co. From the corresponding values of r and 8 and using
eqn. (12) one finds that 6 is approximately 0.2 x 10P8 cm, and therefore that
u = 10-r’ J m-‘. This value of u is of the same order of magnitude as that estimated
for other systems [18,19] and that resulting from a statistical-mechanic calculation
for a simple system [20]. Therefore due to the small size of critical nuclei, the line
tension must be taken into account in the classical analysis of data. The Gibbs
energy of formation of nuclei is then given by [15]

AG* = 16my3+(0)/3AGi - 27ryu sin B/AG, (13)


and expressed in terms of the overpotential by

AG* = 16ry3M2+( 8)/3( @‘q)* - 2ayuM sin fl/pzFq (14)


Using equation (14), the Gibbs energy of nucleation is found to be AG* = lO-*l J
practically invariant with overpotential. Furthermore, the size of the critical nucleus,
which can be evaluated from the critical radius and the microscopic contact angle, is
in all cases less than one atom. This would indicate that an atom adsorbed on an
active site is already supercritical and will therefore grow irreversibly but, in any
case, casts severe doubts on the classical treatment followed. In other cases,
including the line tension in the thermodynamic theory produced a significant
increase in the volume of the critical nucleus [18]. In our case, however, the critical
volume is much too low for classical theory to apply.
The atornistic theory of nucleation, by avoiding macroscopic quantities such as
surface tension, is adequate for the analysis of heterogeneous nucleation at high
supersaturations. For the case of direct attachment of ions to critical clusters the rate
of nucleation is [21],

A = K, exp[ (n* + (~)ze~q/kr] (15)


where e, is the electron charge and K,, the preexponential term, also depends on TJ
and c, as in the classical treatment. From the slope of the plots of In A vs. 7, cf. Fig.
5, it is confirmed that q* = 0 also according to the atomistic theory. To calculate the
work of formation of the critical nucleus in the framework of the theory it is
necessary that plots such as those shown in Fig. 5 show at least three linear regions
of different slope, in order to calculate the supersaturation interval over which a
36

given cluster size plays the role of critical nucleus. As this cannot be accomplished
with our data, the only information available here from the atomistic theory is about
the size of the critical nucleus which is in agreement with that given by the classical
theory, improved by inclusion of line tension effects.

CONCLUSION

Using the treatment of the potentiostatic current transient for three dimensional
nucleation developed in Part I, the number of active sites and the nucleation rates
per site for the nucleation of lead on vitreous carbon were obtained, over a wide
range of overpotentials and at various concentrations of lead ions in solution. The
number of active sites was found to be independent of concentration and its surface
density is relatively low even at high overpotentials, thus justifying the main
assumption of our analysis, namely that the number of active sites on the surface is
limited. Furthermore, the number of active sites for nucleation is much lower than
the number of sites for lead adsorption on vitreous carbon. The concentration
dependence of the nucleation rates, on the other hand, indicates that incorporation
of atoms to critical clusters occurs directly from the bulk of the solution and not
through adsorbed intermediates.
In spite of the large scatter of the nucleation rates obtained, both classical and
atomistic approaches led to the same conclusion as far as the size of the critical
nucleus is concerned. Due to the extremely small size of the critical nuclei, the
classical theory of electrolytic nucleation had to be refined to take into account the
effects of line tension along the contact line between the three phases involved in the
process. From the variation of the microscopic contact angle with overpotential, the
line tension of the lead-carbon-water system was estimated as 10-i* J m-‘. Even
including the line tension effects the size of the critical nucleus remained too small
for classical concepts such as surface tension and contact angle to apply, and
therefore we conclude that the atom&tic theory provides a better representation of
this electrochemical nucleation process.

ACKNOWLEDGEMENTS

This work was financed by the Consejo National de Investigaciones Cientificas y


Tecnologicas (CONICIT) of Venezuela, through grant N, Sl-1227. One of us (J.M.)
also wishes to acknowledge CONICIT for a research studentship during which this
work was carried out.

REFERENCES

1 I. Markov and D. Kashchiev, J. Cryst. Growth, 16 (1972) 170.


2 S. Fletcher and T. Lwin, Electrochim. Acta, 28 (1983) 237.
3 S. Fletcher, Electrochim. Acta, 28 (1983) 917.
4 B.R. Scharifker and G.J. Hills, Electrochim. Acta, 28 (1983) 879.
31

5 C. Mayer and K. Juttner, Electrochim. Acta, 27 (1982) 1609.


6 F. Palmisano. E. Desimoni, L. Sabbatini and G. Torsi, J. Appl. Electrochem., 9 (1979) 517.
7 B. Scharifker and J. Mostany, J. Electroanal. Chem., 177 (1984) 13.
8 G.J. Hills, D.J. Schiffrin and J. Thompson, Electrochim. Acta, 19 (1974) 657.
9 C.M. Pound and H. Karge in R. Niedermayer and H. Mayer (Eds.), Basic Problems in Thin Film
Physics, Vandenhoeck and Ruprecht, Gottingen, 1966, pp. 19-28.
10 J.B. Zeldovich, Acta Physicochim. URSS., 18 (1943) 1.
11 A. Milchev, E. Vassileva and V. Kertov, J. Electroanal. Chem., 107 (1980) 323.
12 N. Tanaka and R. Tamamushi, Electrochim. Acta, 9 (1964) 963.
13 K.L. Moazed and J.P. Hirth, Surf. Sci., 3 (1964) 49.
14 T. Erdey-Gruz and M. Volmer, Z. Phys. Chem., 157 (1931) 1965.
15 G. NavascuC and P. Tarazona, J. Chem. Phys., 75 (1981) 2441.
16 E.A. Guggenheim, J. Chem. Phys., 13 (1945) 253.
17 A.W. Adamson, Physical Chemistry of Surfaces, John Wiley and Sons, New York, 1976, p. 48.
18 G. Navascues and L. Mederos, Surf. Technol., 17 (1982) 79.
19 J.S. Robinson, J. Chem. Sot. Faraday Trans. 2, 79 (1983) 77.
20 P. Tarazona and G. Navascues, J. Chem. Phys., 75 (1981) 3114.
21 A. Milchev and E. Vassileva, J. Electroanal. Chem., 107 (1980) 337.

Вам также может понравиться