Sie sind auf Seite 1von 11

Russian Journal of Electrochemistry, Vol. 38, No. 7, 2002, pp. 732742. Translated from Elektrokhimiya, Vol. 38, No.

7, 2002, pp. 825835.


Original Russian Text Copyright 2002 by Danilov, Molodkina, Polukarov.

Initial Stages of Copper Electrocrystallization


on a Glassy-Carbon RingDisk Electrode from Sulfate
Electrolytes of Various Acidity: A Cyclic Voltammetry Study
A. I. Danilov, E. B. Molodkina, and Yu. M. Polukarov
Institute of Physical Chemistry, Russian Academy of Sciences, Leninskii pr. 31, Moscow, 117915 Russia
Received November 5, 2001

AbstractInitial stages of copper electrocrystallization on glassy carbon from sulfuric acid electrolytes of pH
0.3 and 3.7 are studied by the cyclic voltammetry method on rotating and stationary ringdisk electrode. The
rate of nucleation and growth of a metallic phase of copper in a 0.5 M Na2SO4 + 0.01 CuSO4 (pH 3.7) solution
is marginally higher than in a 0.5 M H2SO4 + 0.01 M CuSO4 acid electrolyte (pH 0.3). Regularities governing
the multistage discharge of copper ions, the formation of the new phase nuclei, and the deposit dissolution are
analyzed. No copper adlayers form on glassy carbon at potentials more positive than the equilibrium potential
of a reversible copper electrode, the copper nucleation occurs via the VolmerWeber mechanism. The oxygencontaining surface groups of glassy carbon (quinonehydroquinone, carbonyl, etc.) are probably active centers
for the discharge of copper ions and the nucleation of the new phase. The results of the study are compared with
the data on the kinetics of copper electrocrystallization on a platinum electrode.

INTRODUCTION
Processes of electrochemical nucleation of metals
on foreign substrates are of fundamental and applied
importance, studying them is of current interest for
modern electrochemistry [13], as the modification of
electrode surfaces by nanosize objects (monolayers,
islet films, quantum dots, and conductors) has become
a major direction of the development of nanotechnology. Carbonaceous materials are frequently employed
as supports for catalysts, electrodes for extracting ions
of heavy metals from dilute solutions, and for analytical
purposes [4]. That is why the researchers devote meticulous attention to studying regularities of metal deposition on carbonaceous electrodes. The process of copper
electrocrystallization involving a multistage discharge
of ions is widely used in industry and is of interest for
studying complex electrochemical processes.
Initial stages of copper electrodeposition from acid
electrolytes have been studied quite comprehensively
on metallic [2, 3, 520, and references therein] and carbonaceous substrates [2124 and references therein].
Some works have been devoted to studying these processes in moderately acid solutions (pH > 2) [1317,
21, 25, 26]. It was shown that the discharge of copper
ions on a platinum electrode accelerates in conditions
of low acidity, and the deposit dissolution yields a much
smaller amount of intermediate species (Cu+ ions) [14
18, 25] than in acid solutions. The authors of [21] discovered no perceptible influence of the electrolyte acidity on the kinetics of copper electrodeposition on glassy
carbon at high cathodic overvoltages. These researchers observed a lower copper crystallization rate in the
presence of hydrochloric acid, i.e. in conditions where

complexes of univalent copper formed. It is noteworthy


that the role played by solution pH and intermediate
species in the phase formation of copper has been analyzed in the literature wholly inadequately. Our studies
of the mechanism of formation of adlayers and nucleation of copper on polycrystalline platinum electrodes
and single crystals of Pt(111) [1520, 27] have shown
that the specific adsorption of oxygen-containing compounds (copper oxides, hydroxide ions, sulfate ions,
hydrogen peroxide reduction products) at the positively-charged surface of platinum electrodes exerts a
catalytic effect on the discharge of copper ions, accelerates the adsorption of atoms and three-dimensional
nucleation, and affects the concentration of Cu+ ions in
processes of discharge and ionization of copper.
The aim of this work is to examine the effect the
solution acidity and the concentration of intermediate
species have on the kinetics and mechanism of copper
electrodeposition at glassy carbon (GC) and perform a
comparative analysis of data on the Pt and GC electrodes.
EXPERIMENTAL
The experiments were carried out on rotating ring
disk electrodes (Tacussel, France). Diameter of GC or
Pt disks was 4 mm; the inside and outside diameters of
the Pt ring were 4.4 and 4.8 mm, respectively; and the
collection coefficient of the ring, N = 25 5%. The auxiliary electrode was made of platinum, and the reference electrode was a mercury/mercury sulfate electrode
in a 0.5 M H2SO4 solution. Values of potentials in the
paper are given in the NHE scale.

1023-1935/02/3807-0732$27.00 2002 I Nauka /Interperiodica

INITIAL STAGES OF COPPER ELECTROCRYSTALLIZATION

Solutions 0.5 M H2SO4 + 0.01 CuSO4 (pH 0.3)


and 0.5 M Na2SO + 0.01 CuSO4 (pH 3.7) were prepared from reactants of p.a. grade (Merck, Germany)
and Milli-Q water (Millipore, the United States) with a
resistivity of 18 Mohm cm and an organic-impurity
content of less than 0.01 ppm. The dissolved oxygen
was removed by argon of extra purity grade. In the
course of experiments, a low excess pressure of argon
was maintained over the solution in the cell. The electrode was polarized and the measurements were carried
out with the aid of a computer-controlled bipotentiostat
designed at the Institute of Physical Chemistry of the
Russian Academy of Sciences.
To ensure a standard state of the GC surface, the
electrode potential was cycled in the range 0.250.80 V
at the rate v = 100 mV s1. To prevent the electrode surface from corrosion, no high anodic potentials (E >
0.8 V) were used. The platinum pretreatment was performed when cycling potential in the range 0.271.25 V
at v = 100 mV s1. The potential 0.8 V imposed on the
ring electrode ensured the occurrence of the reaction
Cu+ e = Cu2+

(1)

Cu+

for monitoring the


concentration in the near-electrode layer of solution.
The equilibrium potential Eeq of the Cu2+/Cu0 pair in
the acid electrolyte of pH 0.3 is equal to 0.25 V. In the
weakly-acid solution (pH 3.7), Eeq = 0.26 V, as determined from steady-state polarization curves and measurements of the open-circuit potential. The 10-mV
shift might have been due to a diffusion drop of potential at the interface between the working solution and
0.5 M H2SO4 in the mercury/mercury sulfate reference
electrode [16].
Details of the experimental procedure are described
in [1618]. The explanations will be provided in text
whenever necessary.
RESULTS AND DISCUSSION
The kinetics of formation and growth of threedimensional nuclei of a metallic phase is traditionally
studied by recording chronoamperograms while varying the electrode potential in a stepwise manner (potentiostatic current transients). The establishment of a
specified value of potential and the EDL charging occur
within a few microseconds. Then there begins the accumulation of metal adatoms (if copper electrocrystallized on platinum at cathodic overvoltages, the amount
of adatoms exceeds one monolayer [5, 15, 16, 28]) and
the accumulation of univalent copper ions in the nearelectrode layer of solution [1418, 27, 29] via the reaction
Cu2+ + e

Cu+

(2)

and quasi-equilibrium occurs between univalent copper


ions and free copper adatoms Cuad (free adatoms are
not bound into a rigid lattice with coadsorbed anions,
RUSSIAN JOURNAL OF ELECTROCHEMISTRY

Vol. 38

733

on platinum these emerge after the formation of a complete monolayer of copper)


Cu+ + e
Cuad,
(3)
because univalent copper ions move from the electrode
surface into the bulk solution by means of natural diffusion or forced convection. These processes give rise to
a sharp maximum of a potentiostatic current transient
for time periods shorter than 1 s.
Depending on the cathodic overvoltage = Eeq E,
copper nuclei form at the electrode surface at one rate
or another; the current of their growth is dependent on
the cathodic overvoltage, the active surface area of the
deposit, and the magnitude of the flow of electroactive
ions to the electrode/solution and nucleus/solution
interfaces. The appearance of a metallic phase of copper and incorporation sites for the products of reduction
of divalent copper ions leads to a decrease in the concentration of free adatoms of copper on the electrode
surface and univalent copper ions in the near-electrode
layer of solution [16, 17]. With the copper crystallites
growing under diffusion or mixed control, a chronoamperogram for a stationary electrode usually displays
a second, less steep maximum, which corresponds to
the instant when hemispherical diffusion zones
(regions of diminished concentration of discharging
ions) start overlapping. Processing experimental current transients in accordance with some model or other
allows one to estimate kinetic and thermodynamic
parameters of the formation of a new phase [2].
Potentiostatic deposition and dissolution of copper
on GC and Pt electrodes will be discussed exhaustively
in the second part of the paper [30]. Here we will contemplate results of exploration of the kinetics and
mechanism of copper electrocrystallization in conditions of linear variations of the electrode potential,
where varying the cathodic limit of the potential
cycling, the potential scan rate, and the electrode rotation rate allows one to analyze processes of accumulation of adatoms and intermediate species in the stage
preceding the phase formation.
On platinum, the copper nucleation occurs via the
StranskyKrastanov mechanism on either a monolayer
or a supermonolayer adatomic coating [1520, 27].
According to the authors of [31, 32], copper adatoms
form in discernible amounts also on carbonaceous
materials at E > Eeq, but interpretation of data presented
by these authors is not indisputable. Our studies have
produced no evidence as to the formation of adatomic
copper layers on GC electrodes. Figure 1 shows cyclic
voltammograms (CVAs) for stationary and rotating GC
disk electrode in a 0.5 M Na2SO4 + 0.01 CuSO4
(pH 3.7) solution (curves 1 and 2, respectively). At E >
0.4 V, the curves virtually coincide with one another
and the CVA recorded in the supporting solution free of
copper ions. Once the direction of the scan of the disk
potential is changed at the beginning of an anodic scan
(increasing positive values of ED) there is a maximum
of current ID in curve 1 (ED 0.3 V), which could have
No. 7

2002

734

DANILOV et al.
ID, A
(a)

1
0

2
I, A

3
2
1

20

50

( rpm )

2
0.4

IR, A

0.6

12

0.8

(b)

2
2
1
0

0.22

0.26

0.30
ED, V

Fig. 1. CVAs for (a) GC-disk and (b) Pt-ring electrodes in 0.5 M Na2SO4 + 0.01 CuSO4 (pH 3.7), recorded at v = 10 mV s1.
Arrows mark the direction of variation of the disk potential. (a) (1) stationary electrode and (2) electrode rotating at a frequency of
500 rpm, in the range 0.250.80 V. Insert: dependences of cathodic current of accumulation of Cu+ at the disk on the electrodes
rotation frequency (5003000 rpm) at potentials of (1) 0.245, (2) 0.240, and (3) 0.235 V. (b) Variations in the current of oxidation
of univalent copper ions on the ring electrode at potential ER = 0.8 V and cycling the potential of the GC disk in the ranges of (1)
0.25 to 0.80 and (2) 0.21 to 0.80 V. The frequency of rotation of the electrode is 500 rpm.

been ascribed to the oxidation of nearly 0.1 monolayers


of copper adatoms accumulated over the cathodic scan
of the potential. However, no such maximum is
observed on a rotating electrode, although considerable
cathodic currents are recorded at E < 0.4 V (curve 2).

currents of the ring and disk electrodes, IR /ID, in this


region of potential is close to 0.2, which corresponds to
the collection coefficient of a disk electrode with a ring.
A similar effect is observed on a GC electrode in the
0.5 M H2SO4 + 0.01 CuSO4 solution.

An analysis of experimental data shows that the


cathodic current at ED < 0.3 V corresponds chiefly to
the occurrence of one-electron reaction (2) of the accumulation of univalent copper ions under the conditions
of diffusion removal of univalent copper ions into the
bulk solution (Tafel slope is equal to 6070 mV per a
decade of change in the current on the disk and ring
electrodes), the current is linearly dependent on the
square root of the frequency of the electrode rotation
(Fig. 1a, insert), as is the case with a platinum electrode
at potentials close to Eeq [17]. The ratio between the

Thus, the maximum of the anodic current in CVA 1


in Fig. 1a is due to the oxidation of univalent copper
ions accumulated in the near-electrode layer of solution
at the end of a cathodic scan of the potential of a stationary electrode, rather than to the oxidation of an
adatomic layer of copper. As the quasi-equilibrium concentrations of free adatoms and univalent copper ions
decrease with the potential displaced in the direction of
more positive values, the excess of univalent copper
ions undergoes oxidation on the surface of the GC disk
via reaction (1). Agitating the electrolyte makes the

RUSSIAN JOURNAL OF ELECTROCHEMISTRY

Vol. 38

No. 7

2002

INITIAL STAGES OF COPPER ELECTROCRYSTALLIZATION


ID, mA

735

ID, mA
(a)

(c)

1
2

Pt

GC

1
0

2
1
IR, A

IR, A
(d)

(b)

1
4

4
2
2

2
0

0.2

1
0.4

0
0
0.6
ED, V

0.2

0.4

0.6

Fig. 2. (a, c) CVAs in (1) 0.5 M H2SO4 + 0.01 CuSO4, pH 0.3 and (2) 0.5 M Na2SO4 + 0.01 CuSO4, pH 3.7, solutions during
a scan of the disk potential, ED, in the range (a) 0.80.00.8 V on GC and (c) 1.250.01.25 V on Pt electrode at a rate of 10 mV s1.
(b, d) Corresponding variations in the current of oxidation of univalent copper ions on the ring electrode at potential ER = 0.8 V. The
frequency of rotation of the electrode is 500 rpm. Arrows mark the direction of variation of the disk potential.

excess of univalent copper ions rapidly leave for the


electrolyte bulk, and the maximum of the current disappears.
However, it would have been wrong to completely
rule out the presence of copper adatoms at the GC surface at the very least at cathodic overvoltages if for no
other reason than that they represent a building material
for the formation of nuclei of a metallic phase. Besides,
it is worthy to note the presence of a small hysteresis of
the cathodic current in CVA following a decrease and
an increase of potential in the range 0.20.4 V. On the
disk, the current of an anodic scan is smaller than that
of a cathodic one (Fig. 1a), and on the ring, vice versa
(Fig. 1b). This effect may be explained by assuming
that in an anodic scan there occurs a two-stage oxidation of the excess of free copper adatoms on the disk
electrode, leading to the emergence of an additional
amount of univalent copper ions in the near-electrode
layer and their oxidation on the ring.
RUSSIAN JOURNAL OF ELECTROCHEMISTRY

Vol. 38

Figures 2a and 2b show CVAs for a GC disk electrode (Fig. 2a) and a Pt ring electrode (Fig. 2b) in solutions of different acidity. With the disk potential ED
decreased from 0.8 to 0.2 V, the dependence of the current ID is similar to the cathodic branch of curve 2 in
Fig. 1a, the current of the ring electrode at ED < 0.35 V
starts increasing and reaches a maximum at ED 0.2 V
(Fig. 2b). A further decrease in ED leads to the formation and growth of metal nuclei; the disk current
increases and reaches a plateau; and the ring current IR,
which is proportional to the near-electrode concentration of univalent copper ions, dramatically drops,
almost to zero (Fig. 2b). Cathodic scans in CVAs
recorded in solutions of different acidity barely differ
(curves 1, 2; Figs. 2a, 2b), we can note only an earlier
nucleation in the weakly acidic electrolyteat more
positive potentials there occurs a sharp increase in the
cathodic current on the disk and begins a decrease in
the anodic current on the ring.
No. 7

2002

736

DANILOV et al.

After changing the direction of the scan and increasing potential up to ED 0.2 V the currents vary insignificantlythe deposit continues growing. However, after
a further increase in the potential the system behavior
depends on the solution acidity. In the acid solution the
deposit dissolution begins at a potential (ED = 0.24 V)
less positive than that in the weakly acidic solution
(ED = 0.26 V) and occurs more rapidly: the potential of
the anodic peak of dissolution of the Cu phase is
observed at ED = 0.33 V (0.37 V for curve 2 in Fig. 2a),
albeit the amount of the copper deposit, calculated by
integrating anodic scans of CVAs, is approximately the
same, specifically, 236 effective monolayers (EM) in
the acid solution and 248 EM at pH 3.7. The difference
between cathodic and anodic charges in the course of
depositiondissolution of copper on GC does not
exceed 1%, whereas the anodic charge on platinum in
the acid solution is smaller than the cathodic charge by
approximately 5% at the expense of intensive formation of univalent copper ions in the course of dissolution of the deposit on the rotating electrode (curve 1,
Fig. 2d). Despite a considerable amount of the deposit
(236 and 248 EM), a part of the surface of GC remains
in contact with electrolyte, copper crystals did not coalesce to form a continuous coating, because the nucleation occurred at moderate cathodic overvoltages (even
a 700-EM deposit formed on GC at 0.18 V for 12 min
in an acid solution comprises individual copper crystallites [30]). Very substantial differences are observed on
the ring electrode: the amount of intermediate species
during the deposit dissolution in the weakly acidic solution is considerably smaller than at pH 0.3 (Fig. 2b). A
similar but more clearly pronounced effect of acidity of
the electrolyte on the number of forming univalent copper ions during the deposit dissolution is observed on a
platinum electrode (Figs. 2c, 2d), this effect was studied in detail in [1518, 25]. Apparently, specific adsorption of sulfate and hydroxide ions on dissolving copper
crystallites hampers the anodic process and extends the
+
lifetime of intermediate species ( Cu ad ) on the surface
of the Cu phase in a weakly acidic electrolyte so that a
large part of univalent copper ions have enough time to
undergo oxidation to Cu2+ before being desorbed into
solution [1618].
It is worth noting that the number of univalent copper ions formed in the stage of copper nucleation on
platinum is smaller than on glassy carbon (Figs. 2b, 2d;
region of maximum in the IR,ED dependence at ED
0.2 V). This fact is due to a delayed nucleation of a
phase of Cu on the GC electrode (at higher overvoltages in these conditions one should expect an increase
in the concentration of free adatoms and univalent copper ions at equilibrium (3) with them) and the difference between spatial distributions of copper crystallites
on the surface of electrodes. The gaps between copper
crystallites growing on glassy carbon are as a rule much
greater than on platinum. In the vicinity of copper
islets, Cu2+ ions undergo a stage-by-stage discharge to

Cu0 and insert themselves into the deposit, while


between copper islets on glassy carbon and platinum
only reaction (2) occurs, yielding Cu+. It is obvious that
the fraction of these ions diffusing into the bulk solution depends on the spatial position of centers of the
deposit crystallization (density of nuclei at the electrode surface)the closer the growth centers to one
another, the lower the concentration of univalent copper
ions in the near-electrode space.
One can obtain important information for an analysis of a multistage mechanism of discharge and ionization of copper by comparing dependences of the ratio
IR /ID on the disk potential ED during the deposition and
dissolution of metallic phase on different substrates in
solutions of different acidity (Fig. 3). Such dependences for steady-state polarization curves of a copperplated platinum electrode are presented in [16] for
cathodic and anodic overvoltages reaching 100 mV
values of the ratio IR /ID in a weakly acidic solution are
considerably smaller than in an electrolyte of pH 0.3,
the amount of univalent copper ions that formed during
the copper deposition is smaller than during the deposit
dissolution.
With the potential scanned from 0.8 V to zero, the
ratio IR /ID passes through a maximum at small cathodic
overvoltages where reaction (2) predominates. Considerable current oscillations are observed on the ring
electrode at these potentials, especially on glassy carbon (Fig. 3a). This phenomenon is of individual interest
but its discussion is beyond the scope of this paper.
With the formation and growth of metallic phase following a further decrease in potential values of the ratio
IR /ID steadily decrease. After a change in the potential
scan direction the copper deposition process continues
and IR /ID barely alters (Fig. 3b, dashed lines), but absolute values of IR /ID in an anodic scan are smaller than
after shifting ED into the cathodic region throughout the
entire range of the copper deposition potentials. The
authors of [8] analyzed the process of formation of univalent copper ions during deposition of copper on
Pt(111) from the viewpoint of thermodynamics and
drew a conclusion that the number of univalent copper
ions in the near-electrode layer of solution is determined mainly by the electrode potential (determining
region of stability of univalent copper ions). The data
presented in the foregoing testify that a thermodynamic
analysis is clearly inadequate for describing this process: besides the electrode potential and the volume
concentration of divalent copper ions, one should take
into account the amount and spatial distribution of
metallic copper on the electrode surface, the solution
acidity, and the substrate origin.
The disk current decreases and alters sign (metal
dissolution commences) near the equilibrium potential
of copper in an anodic scan. In the process, a sharp
increase in negative values of IR /ID is observed (Fig. 3b,
dashed lines), disruption of continuity of function
IR /ID = f(ED) (ID passes through zero, this portion of

RUSSIAN JOURNAL OF ELECTROCHEMISTRY

Vol. 38

No. 7

2002

INITIAL STAGES OF COPPER ELECTROCRYSTALLIZATION

737

IR /ID
0

0.1

0.2

(a)

0.2

0.4

2
1

0.01

4 1

0.02
(b)
1
0

0.1

3 2
0.2

1
ED, V

Fig. 3. (a) Dependences of the ratio between currents of the ring electrode and the disk electrode, IR /ID, on the disk potential ED in
cathodic scans of CVAs presented in Fig. 2 for (1, 2) glassy carbon and (3, 4) platinum in solutions (1, 3) 0.5 M H2SO4 + 0.01
CuSO4, pH 0.3 and (2, 4) 0.5 M Na2SO4 + 0.01 CuSO4, pH 3.7; (b) an enlarged fragment of Fig. 3a (solid lines), dashed lines
represent initial portions of anodic scans after changing the direction of the potential scan.

curves is not shown in the figure), and a subsequent


decay of positive values of the ratio IR /ID with increasing anodic overvoltage and the current of the deposit
dissolution (Fig. 4a). At more positive potentials (ED >
0.35 V) on platinum after the dissolution of the major
bulk of the metallic phase of copper there occurs the
removal of a monolayer of adatoms (curves 3, 4,
Fig. 4a), and the dependence of IR /ID on ED is practically a mirror image of the anodic branch of CVA of the
disk, as the current of the ring electrode in these conditions varies rather slowly (curve 1', Fig. 4b).
An analysis of curve 3 in Fig. 4a may give an
impression that considerably more univalent copper
ions form during the desorption of an adatomic layer
from the platinum surface (ED > 0.35 V) than during the
RUSSIAN JOURNAL OF ELECTROCHEMISTRY

Vol. 38

dissolution of the metallic phase of copper. However,


during the cycling of the electrode potential in the range
0.251.25 V (Fig. 4b, insert) univalent copper forms in
perceptible amounts only at ED < 0.4 V, while during
the formation of a coadsorption lattice of copper adatoms and anions (cathodic scan, ED = 0.40.8 V) or during the desorption of copper adatoms (anodic scan,
ED = 0.40.8 V) the current of the ring electrode is
equal to practically zero. Hence, the higher values of
the ratio IR /ID at ED > 0.35 V for curve 3 in Fig. 4a are
due not to the desorption of adatoms but to the process
of slow dissolution of residues of phase copper or a surface alloy of copper and platinum [19]the current of
the ring electrode remains relatively high practically
until the end of a cathodic scan during the dissolution
No. 7

2002

738

DANILOV et al.
IR/ID

(a)

0.06
3
0.04

0.02

1 4

1
0
ID, A

0.3

0.4

0.5

ED, V
IR, A

(b)
IR, A

12

12

0.2
8

8
0
1'

1.0 ED, V

0.5

4
2
0
0.2

0.4

0.6

ED, V

Fig. 4. (a) Dependences of the ratio between currents of the ring electrode and the disk electrode, IR /ID, on the disk potential ED
for the process of dissolution of copper in anodic scans of CVAs presented in Fig. 2 for (1, 2) glassy carbon and (3, 4) platinum in
solutions (1, 3) 0.5 M H2SO4 + 0.01 CuSO4, pH 0.3 and (2, 4) 0.5 M Na2SO4 + 0.01 CuSO4, pH 3.7; (b) fragments of anodic
scans of CVAs for (1, 2) platinum disk and (1', dashed line) ring electrodes in 0.5 M H2SO4 + 0.01 CuSO4 solution of pH 0.3;
curves 1 and 1' are identical to curves 1 in Figs. 2c and 2d; curve 2 is obtained in similar conditions with a cathodic limit of the
potential scan of 0.08 V. Insert: the current of oxidation of Cu+ on the ring during the cycling of the potential of the Pt disk in the
range 0.251.25 V at a rate of 10 mV s1 and the electrode rotation rate of 500 rpm.

of considerable amounts of copper deposit (for example, at ED = 0.4 V, the current of oxidation of univalent
copper ions for curve 1' in Fig. 4b is ten times that for
the curve in the insert). Values of the ratio IR /ID at some
potentials in the course of deposition and dissolution of
copper on platinum and glassy carbon, determined
from the data of Figs. 3 and 4, appear in the table.
We see that, for all studied conditions of experiments, the amount of univalent copper ions formed during the copper deposition is dependent not on potential
only (see data for 0.1 and 0.2 V in cathodic and anodic
scans): of importance is the deposit mass and, probably,
the spatial distribution of copper on the surface of the
substrate and the magnitude of adsorption of anions on
the substrate and the growing crystallites.

The maximum amount of univalent copper ions in


the cathodic process
Cu2+ + 2e

Cu

(4)

at ED < 0.2 V forms on glassy carbon at pH 3.7, and the


minimum amount, on platinum in the same solution. It
seems possible that the almost-twofold decrease in the
quantity IR /ID (from 11 104 to 6 104) during a
potential scan from zero to 0.1 V for GC in a weakly
acidic solution is caused by an increase in the deposit
mass by approximately 1.5 times (from 114 to 175 EM)
and a corresponding decrease in the fraction of the electrode surface free of copper crystallites (as we have
already mentioned, reaction (2) proceeds at these portions). A similar increase in the amount of copper on
GC at pH 0.3 by a factor of 1.5 leads to a corresponding
decrease in the ratio IR /ID by a factor of 1.5 (from

RUSSIAN JOURNAL OF ELECTROCHEMISTRY

Vol. 38

No. 7

2002

INITIAL STAGES OF COPPER ELECTROCRYSTALLIZATION

739

Values of (104 IR/ID) in the process of deposition--dissolution of copper on platinum and glassy carbon at a potential scan
rate of 10 mV s1
Acidity
pH 0.3
pH 3.7

Electrode
GC
Pt
GC
Pt

Potential of disk electrode, V


0.2*

0.1*

0.0

0.1**

0.2**

0.3**

2000 (0.7)
130 (1.1)
160 (1.5)
12 (15)

22 (45)
4 (13)
31 (52)
1 (78)

3 (112)
1 (38)
11 (114)
1 (145)

2 (179)
3 (50)
6 (175)
1 (210)

13 (230)
50 (67)
10 (232)
3 (273)

+21 (178)
+150 (58)
+7 (111)
+16 (280)

Note: In the parentheses we show amounts of copper deposit (EM) at the instant of realization of one or another potential.
* Cathodic scan from 1.25 V (Pt) or 0.8 V (GC) to 0.0 V.
** Anodic scan from 0.0 V to a corresponding anodic limit.

3 104 at 0.0 V to 2 104 at 0.1 V). Microscopy of the


deposits that were deposited on the GC electrode at
cathodic overvoltages of up to 0.1 V showed that crystals formed in the acid solution (the height was 25
times as small as the diameter of a conventional disk)
were flatter than those formed at pH 3.7 (the height and
the side size were approximately identical), consequently, at the same amount of deposit in solutions with
different acidity a different fraction of free surface of
GC was in contact with electrolyte. This fact may come
useful for explaining a lower magnitude of the ratio
IR /ID in the acid solution at ED = 0.0 V as compared
with the weakly acidic solution. The number of copper
nuclei (crystallization centers) on platinum is considerably larger, and a practically continuous coating forms
after the deposition of about 50 EMthe ratios IR /ID =
1 104 are observed on this electrode at ED = 0.0 V
irrespective of the solution acidity. It should be noted
that a further increase in the amount of the deposit during a potential scan from 0.1 to 0.2 V leads to an
increase in the absolute value of the ratio IR /IDin this
case a predominant role is played by the electrode
potential [8].
Other reasons responsible for an elevated concentration of univalent copper ions at 0.0 V on the GC electrode in the weakly acidic solution (IR /ID = 11 104)
may be a decrease in the adsorption of sulfate anions
and especially hydroxide ions on the substrate and the
copper crystals at negative charges of the surface (for
copper, the potential of zero charge is equal to 0.09 V
[33], for carbonaceous materials in acidified solutions,
about 0.2 V [4, 34]) and a corresponding decrease in the
probability of formation of hydroxy complexes of copper [25].
During the deposit dissolution (ED = 0.3 V, see table)
the minimum amount of univalent copper ions forms on
GC at pH 3.7 at the expense of an increase in the life+
time of species Cu ad in the presence of specifically
adsorbed anions and surface oxygen-containing groups
of glassy carbon [4], and the maximum, in the acid
solution on platinum, as the weak adsorption of bisulfate anions on the substrate and the deposit fails to
RUSSIAN JOURNAL OF ELECTROCHEMISTRY

Vol. 38

ensure effective retention of intermediate species on the


surface.
Figure 5 shows CVAs for the stationary and rotating
GC electrode with variable cathodic limits in solutions
of pH 0.3 and 3.7. The cathodic current of the copper
deposition at ED < 0.2 V passes through a maximum on
the stationary electrode (Figs. 5a, 5c) as a result of
heavy depletion of the near-electrode layer of solution
by discharging ions. Rotating the electrode at a frequency of 500 rpm leads to partial removal of diffusion
limitations, and the current of the deposit growth in this
case rises considerably. Nevertheless, in the weakly
acidic solution there is observed a limiting current at
ED < 0.15 V (Fig. 5d), which points to a diffusion control of copper electrocrystallization. At pH 0.3, no limiting current is reached on the rotating electrode up to a
potential of 0.0 V (Fig. 5b), in this case there is realized
a mixed control diffusion + discharge. Such a situation can emerge when the discharge of copper ions in
the solution of pH 3.7 occurs faster than in the acid
electrolyte, and the limiting stage of the deposition is
the mass transfer. We had come to similar conclusions
when analyzing the copper deposition kinetics on platinum [17].
Figure 6 presents summary data on the copper deposition kinetics at GC and Pt electrodes in sulfate solutions of different acidity in the mode of a linear potential scan at a rate of 10 mV s1 (similar dependences
were obtained at 1 and 100 mV s1). Curves 1 and 2
were obtained by integrating anodic scans of CVAs presented in Fig. 5, while curves 3 and 4 were taken from
[17]. As seen, a minimum copper deposition rate (the
copper amounts corresponding to an identical EC were
deposited within practically identical time periods at
v = 10 mV s1; therefore, quantities Q may be viewed
as estimates of an integral deposition rate) is realized in
the acid electrolyte on the platinum electrode (curves 3,
Figs. 6a, 6b). On glassy carbon in either solution and on
platinum in the electrolyte of pH 3.7 the copper deposition rates are rather close.
As mentioned earlier, the high rate of the copper
crystallization on platinum in the weakly acidic solution is due to accelerated discharge of ions and
No. 7

2002

740

DANILOV et al.
ID, mA

(a)
0.6

0.4

0.2
1
0

0.2

6 5
0.2

0
6 5

32

(d)

(b)

1
0
6
0

6
5
3

0.4

0.2

6
5
4
3
2

(c)

ID, mA

1
1
0.2

2
1
6

0.4

0.6
0
ED, V

1
0.2

0.4

0.6

Fig. 5. CVAs for GC electrode in solutions (a, b) 0.5 M H2SO4 + 0.01 CuSO4, pH 0.3 and (c, d) 0.5 M Na2SO4 + 0.01 CuSO4,
pH 3.7, in the ranges 0.8EC0.8 V, at a rate of 10 mV s1. Cathodic limits EC are as follows: (1) 0.19, (2) 0.17, (3) 0.15, (4) 0.10,
(5) 0.05, and (6) 0 V. The frequency of rotation of the electrode is (a, c) 0 and (b, d) 500 rpm. Arrows mark the direction of variation
of the disk potential.

increased number of active centers on the substrate.


Local electrostatic effects in EDL during specific
adsorption of sulfate and hydroxide ions [1520, 27] on
a positive charged surface of platinum ensure a
decrease in the activation energy for discharge and an
elevated surface concentration of electroactive species
[33]. Surface oxides of copper, which act as active centers for the formation of an adatomic layer and threedimensional nucleation [1618, 27], form on platinum
even in the acid solution. In the solution of pH 3.7
oxides form more rapidly and the nucleation on a supermonolayer adatomic coating commences at more positive potentials and proceeds at a higher rate. That is why
the relatively high rate of the copper electrodeposition
observed on GC in the weakly acidic solution is not
unexpected.
However, the question of the reason for a considerable acceleration of the copper deposition process on
GC in the sulfuric acid solution of pH 0.3, as compared
with a platinum substrate, warrants special analysis. At
the nucleation stage, the increase in the process rate
could have been caused by a higher concentration of
free copper adatoms. As we discovered no ordered
adatomic layer on the GC electrode at E > Eeq (Fig. 1),

all adatoms at E < Eeq may be viewed as free to participate in the nucleation process. On the grounds of
equilibrium (3) and the large current of oxidation of
univalent copper on the ring electrode (Figs. 2b, 2d), at
cathodic overvoltages one should expect that the concentration of free adatoms of copper on GC would be
higher than that on platinum. Oxygen-containing
groups (quinonehydroquinone, carbonyl, etc.), which
are always present at the surface of carbonaceous materials [4], may act as active centers of the substrate. An
elevated electron density on the electronegative oxygen
of such groups must provide for a local increase in the
surface concentration of electroactive species and an
accelerated charge transfer (Fig. 7).
However, it should be noted that the factors we enumerated do not ensure a larger number of copper nuclei
on the surface of the GC electrode as compared with the
platinum electrodeaccording to the electron and optical microscopy data, their number on platinum is
greater by a few orders of magnitude. Apparently, the
high copper deposition rate observed on the GC electrode results from relatively large distances between
crystallization centers (the degree of depletion of the
near-electrode layer of solution by copper ions is

RUSSIAN JOURNAL OF ELECTROCHEMISTRY

Vol. 38

No. 7

2002

INITIAL STAGES OF COPPER ELECTROCRYSTALLIZATION

The adsorption of sulfate and hydroxide ions on GC


in the weakly acidic solution accelerates the charge
transfer and the process of copper deposition as a whole
as compared with the acid electrolyte (Figs. 2, 5, 6).
However, the acceleration is less pronounced than on
platinum, where there occurs the most intensive deposition for the systems we consider here. The adsorption
of these anion on platinum is probably stronger than on
the GC electrode. Another factor of importance at the
stage of phase nucleation is the energy of interaction
between copper adatoms and the substrate. The presence of a monolayer of adatoms on Pt and its absence
on GC at E > Eeq definitely testifies in favor of a higher
bonding energy of Cu atoms with platinum than with
glassy carbon. As a result, all other conditions being the
same, the energy of the formation of copper nuclei on
platinum is expected to be smaller than on the GC electrode. The combination of a high nucleation rate and a
fast discharge of ions during the growth of phase is the
most probable factor accelerating the copper deposition
on platinum in the weakly acidic solution (pH 3.7).

Q, EM
(a)

40

20
3

0
(b)

200

100
1

0.1

0.2
0.3
Cathodic limit, EC, V

Fig. 6. Dependences of the amount of the copper deposit on


(1, 2) GC and (3, 4) platinum electrodes (effective monolayers) in solutions (1, 3) 0.5 M H2SO4 + 0.01 CuSO4,
pH 0.3 and (2, 4) 0.5 M Na2SO4 + 0.01 CuSO4, pH 3.7,
on the cathodic limit of potential scan at a rate of 10 mV s1.
The frequency of rotation of the electrode is (a) 0 and (b)
500 rpm. Data for the platinum electrode are taken
from [17].

Cu2+ + e = Cu+ + e = Cu0

OH
C

O
C

741

CONCLUSIONS
The data presented in the foregoing and their analysis allow us to draw the following conclusions. The
three-dimensional nucleation of copper on glassy carbon occurs via the VolmerWeber mechanismnuclei
of the new phase form on the GC surface from a small
amount of copper adatoms that are at equilibrium with
univalent copper ions. The oxygen-containing surface
groups of glassy carbon (quinonehydroquinone, carbonyl, etc.) are probably the active centers for processes of discharge of copper ions and nucleation of the
metallic phase. The rate of copper deposition in the
weakly acidic solution of pH 3.7 is marginally higher
and the dissolution rate is lower than in the acid electrolyte of pH 0.3 at the expense of adsorption of sulfate
and hydroxide ions on the substrate and the deposit
crystallites. No formation of an adatomic layer of copper has been discovered on glassy carbon at potentials
more positive than a reversible copper electrode.

OH

OHad

ACKNOWLEDGMENTS
This work was supported by the Russian Foundation
for Basic Research, projects nos. 98-03-32 117 and 0015-97381.

Surface of GC-electrode
Fig. 7. Schematic depiction of the glassy carbon/solution
interface. See text for explanations.

smaller than on platinum) and an elevated concentration of univalent copper ions in the near-electrode layer
(the authors of [21] observed deceleration of copper
nucleation in the presence of a chloride complexing
with univalent copper).
RUSSIAN JOURNAL OF ELECTROCHEMISTRY

Vol. 38

REFERENCES
1. Petrii, O.A. and Tsirlina, G.A., Usp. Khim., 2001,
vol. 70, p. 330.
2. Danilov, A.I. and Polukarov, Yu.M., Usp. Khim., 1987,
vol. 56, p. 1082.
3. Danilov, A.I., Usp. Khim., 1995, vol. 64, p. 818.
4. Tarasevich, M.R., Elektrokhimiya uglerodnykh materialov (The Electrochemistry of Carbon Materials), Moscow: Nauka, 1984.
No. 7

2002

742

DANILOV et al.

5. Tindall, G.W. and Bruckenstein, S., Anal. Chem., 1968,


vol. 40, p. 1051; p. 1637.
6. Michailova, E., Vitanova, I., Stoychev, D., and Milchev, A.,
Electrochim. Acta, 1993, vol. 38, p. 2455.
7. Rigano, P.M., Mayer, C., and Chierchie, T., Electrochim.
Acta, 1990, vol. 35, p. 1189.
8. Markovic, N.M., Gasteiger, H.A., and Ross, P.N., Langmuir, 1995, vol. 11, p. 4098.
9. Ikemiya, N., Miyoka, S., and Hara, S., Surf. Sci., 1995,
vol. 327, p. 261.
10. Nichols, R.J., Bunge, E., Meyer, H., and Baumgartel, H.,
Surf. Sci., 1995, vol. 335, p. 110.
11. Eliadis, E.D., Nuzzo, R.G., Gewirth, A.A., and Alkire, R.C., J. Electrochem. Soc., 1997, vol. 144, p. 96.
12. Dietterle, M., Will, T., and Kolb, D.M., Surf. Sci., 1995,
vol. 342, p. 29.
13. Holzle, M.H., Zwing, V., and Kolb, D.M., Electrochim.
Acta, 1995, vol. 40, p. 1237.
14. De Agostini, A., Schmidt, E., and Lorenz, W.J., Electrochim. Acta, 1989, vol. 34, p. 1243.
15. Danilov, A.I., Molodkina, E.B., and Polukarov, Yu.M.,
Elektrokhimiya, 1997, vol. 33, pp. 313, 320.
16. Danilov, A.I., Molodkina, E.B., and Polukarov, Yu.M.,
Elektrokhimiya, 2000, vol. 36, p. 1106.
17. Danilov, A.I., Molodkina, E.B., and Polukarov, Yu.M.,
Elektrokhimiya, 2000, vol. 36, pp. 1118, 1236.
18. Danilov, A.I., Molodkina, E.B., and Polukarov, Yu.M.,
Elektrokhimiya, 2000, vol. 36, p. 1130.
19. Danilov, A.I., Andersen, J.E.T., Molodkina, E.B., et al.,
Electrochim. Acta, 1998, vol. 43, p. 733.
20. Danilov, A.I., Molodkina, E.B., Polukarov, Yu.M., et al.,
Electrochim. Acta, 2001, vol. 46, p. 3137.

21. Gunawardena, G., Hills, G., and Montenegro, I., J. Electroanal. Chem., 1985, vol. 184, p. 357.
22. Kovarskii, N.Ya., Avramenko, V.A., Voit, A.V., et al.,
Elektrokhimiya, 1990, vol. 26, p. 521.
23. Trofimenko, V.V., Kosenko, I.N., Zhitnik, V.P., et al.,
Elektrokhimiya, 1990, vol. 26, p. 387.
24. Danilov, A.I., Molodkina, E.B., and Polukarov, Yu.M.,
Elektrokhimiya, 1994, vol. 30, p. 748.
25. Lezhava, T.I. and Meladze, K.G., Elektrokhimiya, 1978,
vol. 14, p. 1651.
26. Zhang, J., Sung, Y.-E., Rikvold, P.A., and Wieckowski, A.,
J. Chem. Phys., 1996, vol. 104, p. 5699.
27. Danilov, A.I., Molodkina, E.B., and Polukarov, Yu.M.,
Elektrokhimiya, 1997, vol. 34, pp. 1387, 1395.
28. Schultze, J.W., Ber. Bunsen-Ges. Phys. Chem., 1970,
vol. 74, p. 705.
29. Fabricius, G., Kontturi, K., and Sundholm, G., J. Appl.
Electrochem., 1996, vol. 26, p. 1179.
30. Danilov, A.I., Molodkina, E.B., Baitov, A.A., et al., Elektrokhimiya, 2002, vol. 38, p. 743.
31. Jaya, S., Rao, T.P., and Rao, G.P., Electrochim. Acta,
1986, vol. 31, p. 343.
32. Brainina, Kh.Z., Roitman, L.I., Kalnishevskaya, L.N.,
and Podkorhitov, E.M., J. Electroanal. Chem., 1980,
vol. 106, p. 235.
33. Damaskin, B.B. and Petrii, O.A., Vvedenie v elektrokhimicheskuyu kinetiku (Electrochemical Kinetics:
An Introduction), Moscow: Vysshaya Shkola, 1983.
34. Frumkin, A.N., Korobanov, A.A., Vilinskaya, V.S., and
Burshtein, R.Kh., Dokl. Akad. Nauk SSSR, 1976,
vol. 229, p. 153.

RUSSIAN JOURNAL OF ELECTROCHEMISTRY

Vol. 38

No. 7

2002

Das könnte Ihnen auch gefallen