You are on page 1of 6

Electrochemical Synthesis of Silver Nanoparticles

L. Rodr guez-Sanchez,* M. C. Blanco, and M. A. Lopez-Quintela


Department of Physical Chemistry, UniVersity of Santiago de Compostela,
E-15706 Santiago de Compostela, Spain
ReceiVed: May 11, 2000; In Final Form: July 21, 2000
An electrochemical procedure, based on the dissolution of a metallic anode in an aprotic solvent, has been
used to obtain silver nanoparticles ranging from 2 to 7 nm. By changing the current density, it is possible to
obtain different silver particle sizes. The influence of the different electrochemical parameters on the final
size was studied by using different kinds of counter electrodes. The effect of oxygen presence in the reaction
medium as well as the type of particle stabilizer employed have also been investigated. In some conditions
an oscillatory behavior is observed. Characterization of particles was carried out by TEM and UV-vis
spectroscopy. The maximum and the bandwidth of the plasmon band are both strongly dependent on the size
and interactions with the surrounding medium. The presence of different silver clusters was detected by UV-
vis spectroscopy. By using this technique, the existence of an autocatalytic step in the synthesis mechanism
is proposed.
1. Introduction
Metal small particles have received increasing attention in
the last years.
1,2
These particles are usually prepared by chemical
reduction of metal salts
3
. Reetz et al.
4
have described an
electrochemical procedure to obtain particles in which a metal
sheet is anodically dissolved and the intermediate metal salt
formed is reduced at the cathode, giving rise to metallic particles
stabilized by tetraalkylammonium salts. Some of the advantages
of this method are the high purity of the particles and the
possibility of a precise particle-size control achieved by adjusting
the current density. The primary aim of this work is to adjust
the general scheme to the synthesis of silver nanoparticles. For
the optimization, it is necessary to take into account the
following parameters: the choice of the right solvent, supporting
electrolyte, type of electrode, and the current density. It is well-
known that colloidal dispersions of metals exhibit absorption
bands in the UV-vis region, due to collective excitations of
the free electrons (surface plasmon band).
5
Optical properties
of clusters have been investigated for many years.
6
For a single
cluster, the width and position of the plasmon modes are
influenced by the shape, volume, and surface/interface effects.
In many cluster samples, interactions between clusters and the
surrounding media, due to the presence of stabilizers or to the
existence of cluster size or shape distributions, made the
understanding of the optical behavior difficult.
7
Charge transfer
from/toward the ligands may change the electron density, and
lead to blue or red shifts of the plasmon resonance band.
Henglein and co-workers
8
have demonstrated the influence of
redox reactions upon the position when reduction or oxidation
take place, showing blue and red shift, respectively.
In the present work, the optical properties of the obtained
silver colloids are correlated with the silver particle size. The
evolution of the optical properties has been used for the
understanding of some kinetic aspects of the synthesis mech-
anism.
2. Experimental Section
All chemicals were of analytical grade and used without
further purification: H
2
SO
4
from Merck, tetrabutylammonium
bromide (TBABr) and acetate (TBAAcO) from Aldrich, alu-
minum oxide, alpha; 99.99%, 1.0 m from Alfa, were employed.
An Autolab PGSTAT 20 potentiostat was used both in the
synthesis and the electrochemical study. Temperature was kept
at 25 ( 0.1 C using a Grant thermostatic bath. All potentials
were measured against Ag/AgCl reference electrode. The
experiments were carried out in a standard Metrohm electrolysis
beaker containing a sacrificial silver sheet as anode (counter
electrode), and the same size platinum sheet was used as cathode
(working electrode). These two electrodes were vertically placed
face-to-face inside the cell. Platinum electrode was hand-
polished with 1 m alumina powder to a mirror-like finish.
Then, the electrode was activated by triangular potential cycling
between 1.35 and -0.15 V, at a scan rate of 500 mV s
-1
for 5
min in 1.0 M H
2
SO
4
. Before each experiment, the silver
electrode was hand polished by fine grade emery paper and
washed with bidistillated water and a small amount of acetone.
The electrolyte solution consisting of tetrabutylammonium
bromide 0.1 M dissolved in acetonitrile was deaerated by
bubbling nitrogen for about 15 min, keeping an inert atmosphere
during the whole process. A freshly prepared dissolution was
used in each experiment. Strong stirring was kept during the
galvanostatic electrolysis.
TEM measurements were performed on a transmission
electron microscope JEOL 2000, working at 200 kV and
equipped with an EDXA detector. Samples were prepared by
adding acetonitrile to a fraction of the obtained sol, and a droplet
of it was placed on a carbon-coated copper grid covered with
an acetatecellulose polymer. Silver particle formation was
confirmed by dispersive X-ray energy analysis. Size data were
averaged over 100-200 particles from different TEM micro-
graphs
To obtain the absorption spectra, ultracentrifugation of the
sol at 10000 rpm, for 10 min in a Sigma 2-15 ultra-centrifuge,
was carried out to make faster the sedimentation process of the
* Author to whom correspondence should be addressed. Fax: 00 34 81
595 012. E-mail: qflrs@usc.es.
9683 J. Phys. Chem. B 2000, 104, 9683-9688
10.1021/jp001761r CCC: $19.00 2000 American Chemical Society
Published on Web 09/23/2000
particle aggregates, which are always present during the particle
synthesis. Spectra from the supernatant were recorded on a
diode-array Hewlett-Packard HP8452 spectrophotometer in 1
cm light path length cuvettes.
3. Results and Discussion
Optimization of Synthesis Variables. (1) SolVent and
Supporting Electrolyte. Although the solvent proposed in the
method described by Reetz et al.
4
was a mixture of acetonitrile
and tetrahydrofurane, this mixture was found unsuitable for the
silver synthesis, because the presence of tetrahydrofurane
induces the aggregation of the metal particles obtained, which
can be detected by the blue color of the solution. The same
color was found when silver colloidal particles are agglomerated
in the presence of hydrogen peroxide.
9,10
For this reason, pure
acetonitrile was used as solvent. Its aprotic character is necessary
because when protons are in the medium, passivation of the
silver anode occurs, and the synthesis cannot take place.
It is necessary to have a supporting electrolyte as well as a
stabilizer of the particles in the medium. When synthesis is
carried out in 0.1 M NaClO
4
(a supporting electrolyte recom-
mended in electrochemical literature when acetonitrile is the
solvent), no particles are obtained; the electroreduction produces
silver deposition on the cathode. By using tetrabutylammonium
bromide or acetate, which act as supporting electrolytes and as
stabilizers, silver nanoparticles are obtained. Cyclic voltamme-
tries of these two salts were recorded. As it can be seen in Figure
1a, both salts are electrochemically inert in the range 0.8-1.8
V. The processes that occur at higher and lower potentials are
the following:
The solvent must be oxygen free, to avoid the oxidation of the
small metal particles produced; but this is not the only reason:
in silver reduction, oxygen interferes with the electrochemical
process. Figure 1b shows the potential evolution when a
supporting electrolyte 0.1 M tetrabutylammonium bromide is
used, the solution initially deoxygenated without keeping the
inert atmosphere during the electrolysis. In the beginning, the
synthesis occurs at -2.75 V but as the time goes by (about
1000 s), oscillations in the potential can be observed, and finally
(about 2200 s) the potential is not high enough to let the silver
reduction take place and no particles are obtained. Figure 1c
displays linear voltammetries of the system, using two platinum
electrodes, in the presence of different amounts of O
2.
The
presence of a reduction peak about -1.5 V, shows clearly the
oxygen reduction at this potential, therefore the oscillations in
the potential can be attributed to the diffusion of small amounts
of O
2
from the atmosphere into the solution. The oscillatory
behavior disappears when tetrabutylammonium acetate is em-
ployed, which clearly shows a participation of bromide in this
oscillatory phenomenon.
It is also observed that the solution becomes slightly yellow
as the synthesis takes place at any of the current densities
employed. This also happens when silver is not in the medium.
Figure 2 shows the spectrum of the bromide salt solution after
electrolysis using platinum electrodes, under the same experi-
mental synthesis conditions. A wide absorption band, attributed
to the anodic formation of Br
2
, at 364 nm is evident. Taking
into account that the silver plasmon band appears at about 400
nm, this secondary reaction is undesirable for any spectroscopic
study of silver particles. For this reason, acetate salt was
employed in this work.
A typical synthesis of silver nanoparticles is carried out in
acetonitrile using 0.1 M tetrabutylammonium acetate. Under
these conditions, upon applying the current, the electrolyte
Anodic potentials: Cathodic potentials:
(1) 2 CH
3
CO
2
-
f2 CH
3
CO
2

+ 2 e
-
NBu
4
+
+ 1e
-
fNBu
3
+ Bu

(2) Br
-
fBr

+ 1 e
-
2 Br

fBr
2
Figure 1. A) Linear voltammetries of supporting electrolytes using
Pt electrodes. B) Potential variation with time in TBA bromide 0.1 M,
without inert atmosphere. Working electrode: Pt, Counter electrode:
Ag. Current density: -1.25 mA cm
-2
. C) Linear voltammetries using
Pt electrodes in TBA bromide 0.1 M. (A) O2 free. Bubbling O2: (B)
10 s. (C) 15 s.
Figure 2. Spectrum of tetrabutylammonium bromide after Br2
electrolysis at current density of -3 mA cm
-2
.
9684 J. Phys. Chem. B, Vol. 104, No. 41, 2000 Rodr guez-Sanchez et al.
becomes dark yellow and as the reduction process goes on, a
black precipitate is formed. This precipitate can be redispersed
by dilution with acetonitrile and the dark yellow color reappears.
Therefore, this color change can be explained because, when
the concentration of silver colloidal particles is too high,
interactions between the stabilizer chains increase and floccula-
tion takes place. These flocculated particles can be redispersed
again when enough solvent is added.
(2) Current Density and Type of Electrode. Table 1 shows
the reduction products obtained in different conditions, employ-
ing platinum or aluminum as cathode. The morphology of the
cathodic deposits obtained using platinum as cathode at different
current densities is shown in Figure 3. These results are in
accordance with other results already reported in the literature:
11,12,13
when the growth of a new phase takes place close to the
thermodynamic equilibrium (low current densities), a rather
compact phase is formed. Conversely, when the growth condi-
tions are far from equilibrium (high current densities), irregular
aggregates may be formed. The study of these parameters shows
the existence of a competition between two different cathode
surface processes that are summarized in Figure 4: the particle
formation, by reduction and stabilization of silver ions by the
tetrabutylammonium salt, and the film deposition at the cathode
surface (see Table 1). This second process limits the yield of
the particle synthesis, and must be minimized, because when
the electrode surface is totally covered by silver deposition, the
only process that occurs is the silver deposition. Table 2 shows
the crystallographic characteristics and atomic radius of silver
and the different materials employed. The similarity between
silver and aluminum may explain the high tendency to the silver
deposition on aluminum and the fact that, in this case, no
particles are obtained. On the contrary, the difference in radius
and lattice parameters for platinum and silver leads mainly to
particle formation.
Characterization. Figure 5 shows TEM photographs as well
as the corresponding size distributions of the biggest and the
smallest particles obtained. To confirm the composition, EDXA
was carried out, showing that only the black images come from
silver (particles), whereas the gray areas come from the
stabilizer. Table 3 shows the average size of the particles
obtained for different current densities. The observed size
decreases as the current density is increased. This is an expected
result, previously found and explained by Reetz
4
for other
metals.
Spectroscopic Study. The optical behavior of silver colloid
was studied by UV-vis spectroscopy. This shows a clearly
resolved surface plasmon resonance, well separated from the
interband transition.
14
In Figure 6 a typical absorption spectrum from the silver sol
is shown. A broad plasmon absorption band centered at 444
nm can be observed. This peak appears in the range 420-444
nm, depending on the average size of the colloidal particle. Table
4 shows the maximum of the plasmon band position as a
function of the current density employed. A red shift can be
observed as the current density is increased and therefore as
the particle size decreases. This red shift, usually found in small
metal particles, can be explained by the spilling out of the
Figure 3. SEM photographs of the platinum surface after silver reduction at (A) low current density, (-1.4 mA cm
-2
); (B) high current density,
(-7 mA cm
-2
); and (C) magnification of B).
TABLE 1: Summary of the Reduction Products Obtained at
the Working Electrode Depending on the Experimental
Conditions
cathode
nature conditions products
platinum low current density
(-1.4 mA cm
-2
)
(Figure 3.A).
silver nanoparticles and thin
white film
platinum high current density
(-7 mA cm
-2
)
(Figure 3.B)
silver nanoparticles and dendritic
deposition of black film
aluminum -2 mA cm
-2
no particles, only thin white film
Figure 4. Schematic picture showing the competition of two pro-
cesses: (1) silver particle formation, (2) silver deposition.
TABLE 2: Some Relevant Characteristics of Silver and the
Employed Cathodic Materials
species
crystallographic structure/
lattice parameter () atomic radius ()
Ag fcc; a ) 4.08626 1.444
Al fcc; a ) 4.04959 1.431
Pt fcc; a ) 3.9240 1.380
Electrochemical Synthesis of Silver Nanoparticles J. Phys. Chem. B, Vol. 104, No. 41, 2000 9685
conduction electrons, which is more important as the particle
size decreases.
7
To obtain the bandwidths, spectra were fitted to Lorentzians
according to the simple free-electron theory.
15,16
Figure 7 shows
the bandwidth linear increase with the particle size. This linear
dependence is in agreement with a modified version of Drudes
theory, introduced by Doyle
17
and Kreibig et al.
18
For small
particles, these authors assume that the mean free path of the
electrons is limited by the particle size. It should be noted that
other theories (quantum confinement,
15,16
quantum box model,
19
etc.) also predict the 1/R law of the plasmon bandwidth.
We have also studied the evolution with time of the spectra
during the synthesis carried out at a current density of -5.83
mA cm
-2
. More than thirty spectra were recorded, some of them
are shown in Figure 8. In the beginning, a band at about 315
nm, a shoulder at 375 nm, and a weak absorption band at 437
Figure 5. TEM images and size distribution of silver nanoparticles synthesized in TBA acetate 0.1 M in acetonitrile, at various current densities.
TABLE 3: Mean Particle Size Obtained from Different
Current Densities
j (mA cm
-2
) L (nm)
-1.35 6 ( 0.7
-2.85 4.5 ( 0.8
-4.14 3.2 ( 0.6
-6.90 1.7 ( 0.4
Figure 6. Absorption spectrum, taken from the electroreduced sample
at -5.8 mA cm
-2
.
9686 J. Phys. Chem. B, Vol. 104, No. 41, 2000 Rodr guez-Sanchez et al.
nm are observed. The intensity of this last band increases as
time goes by. The band at 315 nm could be attributed
20
to the
existence of Ag
2
+
, and the band at 375 nm to long-lived
clusters.
21
The band about 420-440 nm is commonly attributed
to metallic silver particles. All spectra were fitted to three
Lorentzians. Table 5 summarizes some of the results. In the
beginning no particles are observed (band at 400 nm), and the
main contribution to the absorption is due to the small silver
clusters. At long times the relative contribution due to the
absorption of this kind of clusters is small, but even at longer
times three Lorentzians are needed for the obtaining of good
fits. This fact shows the existence of the initial clusters at any
time, what is in agreement with the suggestion that in the
synthesis method employed, clusters are continuously generated.
The evolution of the optical density at 430 nm means a
sigmoidal shape. This suggests that an autocatalytic growth
could be involved.
22
A possible explanation can be given in
the following terms. The first silver ions are reduced at the
platinum surface, forming the initial clusters. Some of these
clusters diffuse to the bulk and can be detected by UV-vis
spectroscopy (see Figure 8). As reaction proceeds, the number
of clusters increases and the probability that some clusters be
located near the electrode surface is higher. These clusters can
be charged at the same potential (by tunneling). The reduction
on those charged clusters will be easier than on the platinum
surface, giving rise to an autocatalytic effect. This explanation
is in agreement with the observed shift of the cathodic potential
toward less negative values as the reaction proceeds. The inset
of Figure 8 shows a typical autocatalytic behavior. From the
fit, a rate constant of k
obs
) (2.6 ( 0.2) 10
-3
s
-1
is obtained.
It is interesting to note that similar kinetic results were obtained
9
for silver chemical reduction in aqueous media: (k ) (1-5)
10
-3
s
-1
), as well as for silver reduction in microemulsions.
23,24
It has been pointed out that silver electrodeposition on the
cathode takes place simultaneously, diminishing the effective
surface for particle production. This is responsible for the
leveling off observed in the absorbance at long time. When the
surface is totally covered by the silver electrodeposit, the
production of particles finally stops.
4. Conclusions
Silver nanoparticles obtained by electroreduction of anodically
solved silver ions in acetonitrile containing tetrabutylammonium
salts (TBA bromide or TBA acetate) have been studied. An
interesting oscillatory behavior is observed in some conditions.
The optimization of the general synthesis method has been
carried out. The current density plays an important role, not
only on particle size but also on the efficiency of the process.
The cathode nature seems to be also a decisive parameter,
because only particles have been obtained when platinum instead
of aluminum was employed. The spectra of silver sols show
the presence of two different silver clusters. The smallest clusters
are present from the beginning to the end of the process, and
this suggests that clusters are continuously generated. From the
plot of the optical density of the plasmon band versus time an
autocatalytic effect has been found, which is explained assuming
that the reduction can proceed easier on charged clusters located
near the surface. A linear variation of the bandwidth with the
inverse of the particle radius (predicted by Drudes model) as
well as a red shift of the plasmon band (due to the electron
spill out) with decreasing particle size are observed.
Figure 7. Linear variation of the half-width (HW) with the inverse of
the particle radius.
TABLE 4: Variation of
max
with Current Density
j (mA cm
-2
) max
-1.35 422
-1.43 426
-2.85 430
-3.00 430
-3.29 432
-5.83 438
-6.90 444
TABLE 5: Fits of the UV-Vis Spectra Obtained during the Synthesis Performed at - 5.83 mA cm
-2
t (s) max1 (nm)
a
HW1 (nm)
d
max2 (nm)
b
HW2 (nm)
d
max3 (nm)
c
HW3 (nm)
d
250 316 ( 3 46 ( 8
500 314 ( 1 48 ( 4
860 441.7 ( 0.8 58 ( 5 372 ( 1 62 ( 6 306 ( 12 66 ( 17
1100 442.5 ( 0.5 47 ( 3 372.4 ( 0.7 40 ( 3 314 ( 2 41 ( 6
1500 438.4 ( 0.9 79 ( 6 372 ( 1 74 ( 7 307 ( 9 59 ( 16
2300 437.2 ( 0.7 83 ( 3 375 ( 1 88 ( 9 321 ( 30 30 ( 18
2500 435.4 ( 0.8 91 ( 3 373 ( 1 93 ( 9 324 ( 21 30 ( 15
2900 437 ( 1 89 ( 7 383 ( 9 118 ( 50 316 ( 38 42 ( 27
3200 434.7 ( 0.9 100 ( 3 373 ( 9 106 ( 27 320 ( 40 43 ( 29
3600 433.1 ( 0.4 104 ( 13 381 ( 25 231 ( 95 317 ( 50 40 ( 30
a

max1
) 437.5 nm.
b

max2
) 375.2 nm.
c

max3
) 315.5 nm.
d
HW) half-width.
Figure 8. Spectra obtained during the nanoparticle synthesis at -5.83
mA cm
-2
. Inset: Plot of ln(a/(1 - a)) vs time (a ) Abst/Abs).
Electrochemical Synthesis of Silver Nanoparticles J. Phys. Chem. B, Vol. 104, No. 41, 2000 9687
Acknowledgment. Partial financial support by the Xunta
de Galicia is gratefully acknowledged (PGIDT99PXI20905B).
References and Notes
(1) Ozin, G. A. AdV. Mater. 1992, 4, 612.
(2) Henglein, A. Chem. ReV. 1989, 89, 1861.
(3) Lopez-Quintela, M. A.; Rivas, J. Curr. Opin. Colloid Interface Sci.
1996, 1, 806.
(4) Reetz, M. T.; Helbig, W. J. Am. Chem. Soc. 1994, 116, 7401.
(5) Wilcoxon, J. P.; Williamson, R. L.; Baughman, R. J. Chem. Phys.
1993, 12, 9933.
(6) a) Hughes, A. E.; Jain, S. C. AdV. Phys. 1979, 20, 717, and
references therein. (b) Kreibig, U.; Fragstein, C. Z. Phys. 1962, 224, 307.
(7) Kreibig, U.; Vollmer, M. Optical Properties of Metal Clusters;
Springer-Verlag: Berlin, Heidelberg, 1995.
(8) Henglein, A.; Mulvaney, P.; Linnert, T. Faraday Disc. 1991, 92,
31.
(9) Huang, Z.-Y.; Mills, G.; Hajek, B. J. Phys. Chem. 1993, 97, 11542.
(10) Kreibig, U. Z. Phys. B: Condens. Matter Quanta 1978, 31, 39.
(11) Stranski, J.; Krastanov, L. Akad. Wiss. Math. Nat. K111b 1938,
797.
(12) Franck, F. Van der Merwe, J. Proc. R. Soc. London, Ser. A 198
1949, 205.
(13) Volmer, M.; Weber, A. Z. Phys. Chem. 1926, 119, 277.
(14) Henglein, A. J. Phys. Chem. 1993, 97, 5457.
(15) Kawata, A.; Kubo, R. J. Phys. Soc. Jpn. 1966, 21, 1765.
(16) Kubo, R. J. Phys. Soc. Jpn. 1962, 17, 1765.
(17) Doyle, W. T. Phys. ReV. 1958, 111, 1067.
(18) Kreibig, U.; Fragstein, C. V. Z. Phys. 1969, 224, 307.
(19) Genzel, L.; Martin, T. P.; Kreibig, U. Z. Phys. B 1975, 21, 339.
(20) Ershov, B. G.; Janata, E.; Henglein, A.; Fojtik, A. J. Phys. Chem.
1993, 97, 4589.
(21) Henglein, A.; Linnert, T.; Mulvaney, P. Ber. Bunsen-Ges. Phys.
Chem. 1990, 94, 1449.
(22) Boudart, M. Kinetics of Chemical Processes; Prentice Hall:
Englewood Cliffs, NJ, 1968; Chapter 6.
(23) Rivadulla, J. F.; Vergara, M. C.; Blanco, M. C.; Lopez-Quintela,
M. A.; Rivas, J. J. Phys. Chem. 1997, 101, 8997.
(24) Tojo, C.; Blanco, M. C.; Rivadulla, J. F.; Lopez-Quintela, M. A.
Langmuir 1997, 13, 1970.
9688 J. Phys. Chem. B, Vol. 104, No. 41, 2000 Rodr guez-Sanchez et al.

You might also like