Sie sind auf Seite 1von 6

Appl. Phys.

A 82, 125130 (2006)

Applied Physics A

DOI: 10.1007/s00339-005-3372-4

Materials Science & Processing

t. gleitsmann
t.m. bernhardtu

l. woste

Luminescence properties
of femtosecond-laser-activated silver oxide
nanoparticles embedded in a biopolymer matrix
Institut fr Experimentalphysik, Freie Universitt Berlin, Arnimallee 14, 14195 Berlin, Germany

Received: 8 February 2005/Accepted: 6 July 2005


Published online: 30 September 2005 Springer-Verlag 2005
ABSTRACT Strong visible luminescence is observed from silver clusters generated by femtosecond-laser-induced reduction
of silver oxide nanoparticles embedded in a polymeric gelatin
matrix. Light emission from the femtosecond-laser-activated
matrix areas considerably exceeds the luminescence intensity of
similarly activated bare silver oxide nanoparticle films. Optical
spectroscopy of the activated polymer films supports the assignment of the emissive properties to the formation of small silver
clusters under focused femtosecond-laser irradiation. The size
of the photogenerated clusters is found to sensitively depend
on the laser exposure time, eventually leading to the formation of areas of metallic silver in the biopolymer matrix. In this
case, luminescence can still be observed in the periphery of the
metallic silver structures, emphasizing the importance of the organic matrix for the stabilization of the luminescent nanocluster
structures at the metalmatrix interface.
PACS 78.66.Qn;

78.20.-e; 78.40.-q; 78.67.Bf; 78.68.+m

Introduction

In the quest for efficiently fluorescent nanoscale


materials, atomic silver clusters have attracted considerable
interest in recent years. Approaching the ultimate scale of
miniaturization at the atomic dimensions, the envisaged applications of clusters of a few atoms lie in nanoscale all-optical
logic devices as well as in advanced optical data-storage media [1]. Embedded in rare-gas matrices, silver clusters with
two, three, four, eight, or nine atoms were found to exhibit
light-emitting properties under UV- or visible-light excitation [24]. In contrast, free silver clusters in the gas phase that
are exposed to light excitation are not known to show indications of luminescence [3]. Experimental clues to the origin
of such unexpected optical properties of silver clusters in the
size range below about 20 atoms per particle come from twophoton excitation and electron spectroscopic experiments on
these clusters in ultra-cold rare-gas droplets [3, 5, 6]. The
prevalence of long-lived discrete electronic levels obviously
determines the observed fluorescence, indicating that these
clusters cannot yet be considered metallic. Theoretical calculations confirm this picture [4, 7]. The apparent essential role
u Fax: +49-30-838-55567, E-mail: tbernhar@physik.fu-berlin.de

of the matrix in this respect might be assigned to a stabilizing effect on the electronic as well as on the geometric cluster
structure (caging effect) [3, 6].
Readily prepared through UV-light-induced reduction of
silver oxide thin films, atomic silver clusters have also been
identified as the origin of strong visible luminescence observed in this new photosensitive material [8, 9]. In similar
silver oxide thin films, electroluminescence of atomic silver clusters at nanoscale break junctions has been employed
for single-nanocluster quantum optoelectronic logic operations [10]. With respect to advanced optical data-storage applications, the UV-light-induced formation of luminescent
clusters on silver oxide was proposed to enable local multilevel writing followed by read out via fluorescence excitation
in the visible [8]. However, very fast data writing and read out
are necessary for the utilization of this system as an efficient
storage medium. Even more important, the written data have
to be long-time stable and inert against the read-out process
itself. These prerequisites have been recently met by employing a focused femtosecond (fs) near-infrared laser for spatially
resolved and fast photoactivation. In contrast, for read out
low-power Ar+ /Kr+ laser excitation was used, which does
not generate fluorescent structures itself [11]. With this approach it was possible to write diffraction-limited structures
which could be read out through fluorescence excitation without degradation even after months.
In particular, with respect to the activation sensitivity and
the luminescence intensity, the interaction of the photoactive
silver oxide material with the surrounding environment has
been found to be very crucial in previous experiments [9].
Silver oxide in vacuum, for example, showed an order of
magnitude lower activation rate compared to that in air at
atmospheric pressure, whereas the fluorescence intensity increased in nitrogen or argon atmospheres [9]. This interface
influence is also well known from the photoactivation of photographic silver halide materials, in which a similar process of
light-induced silver cluster generation gives rise to the latent
image [12], which is in turn stabilized by a surrounding transparent matrix. The biopolymer gelatin is the classical matrix
for photosensitive materials. It consists of a polymeric collagen structure. Collagen itself is not a homogeneous substance,
but consists of a whole family of proteins. Similar in all collagen types is the triple-helical structure. The macromolecular
network undergoes coagulation under partial hydrolysis, de-

126

Applied Physics A Materials Science & Processing

termining the particular mechanical stability and flexibility of


the gelatin material that make it ideally suited for a wide range
of applications [13].
The most important functions of the gelatin with respect to
photographic products consists in the prevention of grain coagulation and aggregation, the regulation of particle growth,
and the optimization of the light sensitivity of the photographic layers. In addition to its importance in silver halide
photography, gelatin is widely applied in pharmacology as
biocompatible encapsulation material for medication and in
drug-delivery systems.
In the present contribution we investigate the influence
of a polymeric gelatin matrix on the photoactivated luminescence properties of silver oxide nanoparticles. The reductive
transformation of the silver oxide within the matrix upon fslaser exposure is evidenced and emphasis is put on the defined
generation (writing) of fluorescent microstructures, which is
found to be much more efficient in the presence of the matrix.
2

Experimental

The polymer matrix with embedded silver oxide


nanoparticles was prepared by grinding of solid Ag2 O
(99.9%, Aldrich Chemical) together with gelatin powder at
a mass ratio of 1 : 10. 500 mg of the mixture was subsequently
suspended in 10 ml water and about 0.5 ml of this suspension
were placed on microscope glass slides. After evaporation of
the liquid, a few m thick transparent polymer films were obtained that exhibit a homogeneous light-brownish color. For
comparison, pure silver oxide nanoparticle films were also
prepared according to previous descriptions [8, 11]. In this
case, ground solid Ag2 O was suspended in chloroform and
sonicated. After settling, the supernatant was sonicated again
and drops from this suspension were placed on microscope
glass slides. After evaporation of the chloroform, thin films
consisting of nanoscale silver oxide particles with diameters
of a few 100 nm were obtained.
Both kinds of films initially exhibited no luminescence.
Spatially resolved activation of luminescence was achieved
by scanning of a tightly focused ultra-fast fs-laser beam
over the sample. The experimental setup for photoactivation and luminescence detection is depicted schematically
in Fig. 1. For activation (Fig. 1a: Write) 60-fs laser pulses
with a wavelength of act = 800 nm and a repetition rate of
87 MHz from a home-built Ti:sapphire laser oscillator were
used. The fs-laser beam was coupled into a confocal scanning laser microscope (Leica TCS) and focused by a 10
objective (numerical aperture 0.4) onto the sample. Using
a biaxial pivoted scanning mirror, it was possible to scan over
defined areas on the sample surface with a frequency of up to
1 kHz and an activation power of up to Pact = 12.7 MW/cm2
in the diffraction-limited focus. Luminescence from the activated areas in the silver oxide nanoparticle films was subsequently excited by scanning over the activated region with
blue (488 nm) or green (568 nm) light from an Ar+ /Kr+
laser (Fig. 1b; Readout). For the luminescence excitation
an Ar+ /Kr+ -laser intensity of more than four orders of magnitude lower than the activating fs-laser power was used, in
order to avoid further activation throughout the read-out process [11]. The emitted fluorescence was collected by the same

Schematic illustration of the experimental arrangement. (a) Activation (writing) by scanning with focused fs-laser pulses (800 nm) over
a square on the sample surface. (b) Excitation (read out) of fluorescence
by subsequent scanning of the activated region with blue (488 nm) or green
(568 nm) light from an Ar+ /Kr+ laser. The luminescence light passes
through long-pass filters (LP1) and is detected by a photomultiplier (PM).
The lower part shows a perspective view of an image of a pure Ag2 O
nanoparticle film acquired in this way and digitized to an 8-bit false-color
intensity scale. The nine squares were individually written by 10-min activation at powers ranging from 0.6 to 6.5 MW/cm2 (from upper left to lower
right corners). Alternatively, filtered light of a mercury lamp is used to record
real-color fluorescence images with a color digital camera (D). A dichroic
mirror (DM) in combination with long-pass filters (LP2) is used to separate
excitation light from fluorescence light in this case
FIGURE 1

objective. Finally, the photons were detected by a photomultiplier to yield an 8-bit gray-scale image.
The micrograph shown in the lower part of Fig. 1 was obtained in this way from a pure silver oxide nanoparticle film.
Nine squares (40 40 m2 each) were individually activated
over 10 min with different fs-laser powers, resulting in different luminescence intensities. To block scattered excitation
light and to bracket the spectral range of the fluorescence,
long-pass filters were used in front of the photomultiplier.
In contrast to the pure silver oxide particle films, luminescence from previously activated gelatin matrix embedded
Ag2 O nanoparticles could already be excited with a weak
mercury (Hg) lamp. Real-color fluorescence images of the
activated structures under filtered Hg-lamp excitation were
recorded in this case directly with a Samsung UCA-3 color
digital camera (detector D in Fig. 1). This detector was also
used to obtain microscopic images of the samples in reflection
of unfiltered, unfocused radiation from an externally mounted
tungsten lamp (not shown in Fig. 1).
3

Results and discussion

In the following, first, the microscopic structure of


a silver oxide thin film will be presented in conjunction with
the spatial distribution of the light emission. Second, the fslaser writing of luminescent line scans in a gelatin embedded
Ag2 O nanoparticle film will be investigated together with the
optical absorbance and emission of the activated thin films.
Finally, the fs-laser exposure-time-dependent reductive transformation of the silver oxide nanoparticles will be presented

GLEITSMANN et al.

Luminescence properties of fs-laser-activated silver oxide nanoparticles . . .

127

and the long-time stability of luminescent structures in the


gelatin embedded silver oxide nanoparticle film will be discussed.

no apparent formation of areas of metallic silver could be detected [11]. This is in contrast to the matrix embedded Ag2 O
nanoparticle films, as will be described below.

3.1

3.2

Microscopic structure
of luminescent silver oxide nanoparticle films

Figure 2a shows a 5 5 m2 topographic image of


a pure silver oxide nanoparticle film obtained with an atomic
force microscope (AFM). Several aggregates of silver oxide
particles can be distinguished. These agglomerations consist
of individual nanoparticles of sizes of a few hundred nanometers. If a silver oxide nanoparticle film has been previously
activated by irradiation with fs-laser light, fluorescence is observed to appear around these nanoparticle agglomerations:
Fig. 2b to d show fluorescence images recorded on a similar
film as in Fig. 2a. An agglomeration of silver oxide nanoparticles is visible. A 32 32 m2 area of this film has been
previously activated and the three micrographs show the local
evolution of the fluorescence on the same surface area with
increasing photoactivation time. Bright luminescence grows
in at the periphery of this agglomeration. With increasing activation time the fluorescent spots become more pronounced
and broader until they cover the whole agglomerate perimeter [11]. Apparently, the luminescent centers generated during
fs-laser activation are confined to the surface of the silver
oxide nanoparticles and the number of luminescent centers
increases with activation time. Although gradual saturation
of fluorescence intensity with increasing activation time and
power has been observed for the pure silver oxide thin films,

FIGURE 2 (a) Topographic AFM image of a pure silver oxide thin film
showing several agglomerations of nanoparticles with a size of a few 100 nm.
(b)(d) Evolution of the fluorescence at the periphery of a similar aggregate
of Ag2 O nanoparticles as in (a). Activation times were 12, 22, and 60 min for
(b)(d), respectively (Pact = 6.5 MW/cm2 , 32 32 m2 total activated area,
ecx = 568 nm of Ar+ /Kr+ laser, photomultiplier detection yielding a falsecolor image). The fluorescence images were recorded with the same low
signal amplification to avoid photomultiplier saturation at high fluorescence
intensity. The length scale is the same in all images

Femtosecond-laser-induced activation
of biopolymer matrix embedded
silver oxide nanoparticles

Single, 4-m-thick lines were activated in a gelatin


embedded Ag2 O nanoparticle film by repeated x -direction
scanning of the fs laser, while leaving the y position of the
scanning mirror fixed. Figure 3a shows a fluorescence image
of such a line activated by scanning for 0.1 s with a frequency
of 1 kHz and an activation power of 12.7 MW/cm2 . Under
these conditions luminescence can be excited predominantly
from the laser turning points and not in the center region of
the activated line (dashed). Obviously, due to the non-linear
motion of the scanning mirror, the activating fs-laser exposure time is longer at the laser turning points than at the scan
line center. This leads to a discontinuous distribution of activated fluorescent centers with a maximum at the line end,
where the laser focus remains for a comparably long time. Luminescence is observed at these short activation times directly
in the area that has been exposed to the focal point of the fs
laser (Fig. 3a).
Most interestingly, the spatial distribution of the fluorescence changes drastically if the activation time is increased
to several seconds under otherwise identical conditions. Figure 3b shows that under these conditions luminescence no
longer emerges from the area illuminated by the fs-laser focus, but from the periphery of the fs-laser scan line. Obviously, the areas of the matrix embedded silver oxide nanoparticle film that have been directly exposed to the fs-laser focus
are already saturated and have been transformed into non-

FIGURE 3 Single lines (total length: 500 m) written with the activating fs laser into a gelatin matrix embedded silver oxide nanoparticle film.
Line width: 4 m. Activation power Pact = 12.7 MW/cm2 . Scan frequency:
1 kHz. False-color images: 568-nm Ar+ /Kr+ -laser excitation, recorded with
the photomultiplier. (a) Activation time: 0.1 s. Note that fluorescence is only
observed along the scan line (dashed) at the laser turning point region.
(b) Activation times: A: 1 s; B:2 s; C: 4 s

128

Applied Physics A Materials Science & Processing

fluorescent material, whereas the perimeter of the scan line


which has been illuminated only weakly due to the Gaussian
beam profile or by scattered light of the fs laser now exhibits
strong light emission. Thus, the generation of fluorescent centers in the silver oxide film strongly depends on the activating
fs-laser exposure time at constant laser intensity.
A reductive transformation of the matrix embedded silver oxide particles with exposure time as postulated previously [8, 9, 11, 14] is observed, leading first to the formation
of very small silver clusters that show luminescent properties at the surface of the Ag2 O nanoparticles (cf. Fig. 2). With
increasing fs-laser exposure time apparently larger silver particles are generated that eventually exhibit metallic properties,
resulting in a quenching of the luminescence. This fs-laserinduced transformation will be evidenced in the following by
measurements of the optical properties of the matrix embedded silver oxide nanoparticles and by a more detailed analysis
of the structure of the activated areas.
3.3

Optical properties of activated matrix embedded


silver oxide particles

Figure 4 shows the absorbance of a gelatin matrix embedded sliver oxide nanoparticle film in the wavelength range between 250 and 900 nm. Trace A was obtained
from a luminescent film. The sample was activated to yield
maximum visible fluorescence from a 1 mm2 activated area.
Trace B was obtained from a non-activated area of the same
sample. The increase in absorbance around 300 nm can be
attributed to the absorption of the gelatin matrix itself, as evidenced from spectra taken of a pure gelatin film (not shown).
The maximum at 265 nm in spectra A and B most likely originates from silver oxide nanoparticles, because a pure silver
oxide thin film also exhibits an absorbance maximum at this
wavelength as can be seen from trace D in the inset of Fig. 4.
Trace C results from the difference of traces A and B. This
spectrum exhibits maxima at 290 nm and 425 nm, as well as
a shoulder around 610 nm.

FIGURE 4 Absorbance spectra of a gelatin film with embedded silver


oxide nanoparticles. Trace A (- - -): activated film. Trace B (-.-.-): nonactivated film. Trace C (): difference spectrum of A and B. Trace D (inset):
absorbance of a pure Ag2 O nanoparticle thin film

Silver clusters with a diameter between 5 and 50 nm display a plasmon absorption resonance at about 410 nm [15, 16].
This plasmon peak shifts to longer wavelength under the influence of the dielectric gelatin matrix and also as a function
of the matrix filling factor [1618]. A new peak in the optical
spectrum around 400 nm has also been observed after thermal
annealing of silver oxide thin films and it has been assigned to
the formation of nanometer-sized silver clusters [19].
Absorption spectra of silver clusters in photosensitive
glass display, in addition to this plasmon absorption, a peak
around 300-nm wavelength which has been attributed to a molecular absorption of very small, not yet metallic, silver clusters [16]. Also, atomic silver clusters with one to four atoms
that are stabilized by encapsulation in an organic matrix material exhibit distinct absorption bands at 424 and 520 nm [20].
These bands are found to coincide with the electronic transitions calculated and observed for small silver clusters, in
particular Ag2 and Ag3 , in this spectral region [21, 22].
Thus, the absorbance spectrum (trace C in Fig. 4) clearly
identifies the fs-laser-activated generation of nanometer-sized
silver particles exhibiting the plasmon absorption around
400 nm. However, it furthermore provides indications of the
initial formation of atomic silver clusters consisting of only
a few atoms which are expected to be responsible for the observed luminescence properties of the activated silver oxide
nanoparticle films.
3.4

Luminescence characterization

The spectral distribution of the luminescence from


the activated areas is found to be excitation-wavelength dependent, as demonstrated earlier [9, 11]. Real-color fluorescence images of activated samples are shown in Fig. 5c and

FIGURE 5 Series of lines written with the focused fs-laser into a gelatin
embedded Ag2 O nanoparticle film. (a) Image recorded in reflection of an
unfocused tungsten lamp with the color digital camera showing the microscopic structure of the film after activation. Note the apparent formation of
metallic silver at the laser turning point on the left-hand side. (b) False-color
fluorescence image: Ar+ /Kr+ -laser (ecx = 568 nm)-excited fluorescence,
recorded with the photomultiplier after long-pass filtering (LP 590). (c)
Real-color fluorescence image: Hg-lamp (ecx = 546 nm)-excited fluorescence, recorded with the color digital camera after long-pass filtering (LP
590, transmitted wavelength trans > 590 nm). (d) Real-color fluorescence
image: Hg-lamp (ecx = 363366 nm)-excited fluorescence, recorded with
the color digital camera after long-pass filtering (LP 430, transmitted wavelength trans > 430 nm)

GLEITSMANN et al.

Luminescence properties of fs-laser-activated silver oxide nanoparticles . . .

129

d. A series of lines written in a gelatin embedded silver


oxide film exhibit bright-red fluorescence under 546-nm Hglamp excitation (Fig. 5c), whereas excitation with the Hglamp lines at around 365 nm causes fluorescence in the green
and yellow range (Fig. 5d). No polarization dependence of
the fluorescence intensity is observed. The emission spectra
of a similar activated sample at 488- and 568-nm Ar+ /Kr+ laser excitation exhibit broad fluorescence bands of about
200 nm width centered at 520 and 620 nm, respectively [23].
These spectra are comparable to luminescence spectra obtained from UV-light-activated silver oxide thin films [9]. Interestingly, very similar luminescence spectra have also been
observed from photoactivated colloidal or fractal silver nanostructures in air [24].
3.5

Exposure-time-dependent
spatial luminescence distribution

A micrograph of an activated structure in a gelatin


embedded silver oxide film consisting of several activated horizontal lines is depicted in Fig. 5a. The image was taken using
the reflected light of an unfocused external tungsten lamp and
observing the sample morphology through the microscope
with the color digital camera. Most strikingly, under these observation conditions the turning point areas of the activating
fs-laser on the left-hand side can be identified to consist of
metallic silver. As detailed in Sect. 3.2, the non-linear motion of the scanning mirror causes an enhanced laser exposure
time around the turning points. This resulted in the case of
the activated structure in Fig. 5 in the complete light-induced
reduction of the silver oxide and thus to the formation of
metallic silver.
In the Ar+ /Kr+ -laser-excited fluorescence image shown
in Fig. 5b the area identified to consist of metallic silver exhibits no luminescence at all. This is the same in the realcolor fluorescence images obtained under Hg-lamp excitation
displayed in Fig. 5c and d. However, at the interface of the
metallic silver and the silver oxide matrix, fluorescence can
still be observed, emphasizing the importance of the matrix
interface for the stabilization of the luminescent nanocluster
structures. This observation is similar to the scan line perimeter fluorescence displayed for long exposure times in the line
scan of Fig. 3b. Apparently, the silver cluster size increases
with increasing fs-laser exposure time, leading to the formation of larger silver particles (exhibiting plasmon absorption,
cf. Fig. 4) and eventually to the generation of microscopic
silver islands as observed in Fig. 5a. Luminescent blinking
has also been observed from the above-mentioned photoactivated colloidal or fractal silver nanostructures [24]. However,
no clear connection between the luminescence and the silver
structure perimeter was seen in this contribution. Figure 6a
displays an enlarged view of the fluorescence image in Fig. 5d
and a schematic model of the observed spatial luminescence
distribution, summarizing the exposure-time-dependent silver cluster formation. The dashed lines in Fig. 6 mark the
focus center position of the scanning fs-laser. Starting from
the right-hand side, the scanning mirror speed is largest in
region A, resulting in short exposure times and no luminescence activation, i.e. no silver cluster formation. Propagating
to the left, in region B, the exposure time increases, because

FIGURE 6 (a) Enlarged view of the real-color fluorescence image


in Fig. 5d showing the spatial distribution of fluorescence emission along
the scan lines of the activating fs laser (dashed lines). Laser exposure time
is increasing from right to left and maximal at the laser turning point on the
left-hand side. (b) Schematic model of the evolution of fluorescence activation along the activating fs-laser scan lines (dashed lines). Different spatial
regions (AD) can be distinguished. Cross-hatched areas exhibit luminescence, gray-shaded areas consist of large silver particles or are fully reduced
to metallic silver. See text for details

the scanning mirror slows down approaching its turning point.


Consequently, as also already seen in Fig. 3a, luminescence
along the laser focus line is observed (cross-hatched area in
Fig. 6b), indicating the fs-laser-induced formation of atomic
silver clusters on the silver oxide particles in the matrix (cf.
also Fig. 2). Careful inspection of the spatial distribution of
the observed luminescence shows that at the transition from
region B to region C, the luminescence intensity along the
laser focus scan line disappears and is displaced to the area
between two scan lines. According to the above, this leads to
the conclusion that the even more elongated exposure time in
region C resulted in the generation of larger silver particles
along the laser focus line (light-gray-shaded area in Fig. 6b)
and fluorescent atomic silver clusters at the focus perimeter in between the laser focus lines (cross-hatched area in
Fig. 6b). Finally, the scanning laser turning point is reached
in region D. Here, the laser exposure time is so long that the
activation leads to a complete reductive transformation of the
silver oxide and the formation of microscopic metallic silver
islands, clearly visible in the micrograph of Fig. 5a (grayshaded area in Fig. 6b). Thus, it can be concluded that the size
of the photogenerated luminescent silver clusters sensitively
depends on the activating fs-laser exposure time at constant
laser power, suggesting the possibility of a sophisticated micropatterning of gelatin embedded silver oxide nanoparticle
films.
3.6

Long-time stability of laser-activated luminescence

As distinct from the previously investigated fslaser-activated pure silver oxide nanoparticle films [11], luminescence from the fs-laser-activated areas of gelatin polymer
matrix embedded silver oxide particle films can already be
excited with light from a weak mercury lamp. Activated sam-

130

Applied Physics A Materials Science & Processing

ples were stored in air under ambient conditions, however


avoiding light exposure, over a period of 10 weeks. Repeated
investigation of the samples with respect to the luminescent
properties of the activated structures yielded no fluorescenceintensity degradation during this time period. Also, no change
in the morphological structure of the activated areas could be
detected.
4

Conclusions and perspectives

Femtosecond-laser-activated luminescence from


a biopolymer matrix embedded silver oxide nanoparticle film
has been observed. The luminescence intensity from the matrix embedded particles considerably exceeds the observed
light emission of pure silver oxide thin films. A strong dependence of the degree of photoinduced reductive transformation
of silver oxide nanoparticles on the laser exposure time is detected, eventually leading to the formation of metallic silver
microstructures in the matrix. Light emission is also observed
from the perimeter of these silver structures.
The favorable mechanical properties of the matrix material, as well as the measured long-time stability of the luminescent structures in the matrix, suggest possible applications
as versatile light-emitting materials for all-optical logic devices or data-storage media. In addition, the stabilization of
the formed luminescent clusters by interaction with the organic matrix might determine gelatin encapsulated fluorescent silver clusters as potential biocompatible luminescent
pharmacological probes.
It is however not yet clear whether the efficient fs-laserinduced reduction of silver oxide and the formation of luminescent cluster structures are due to a local rapid heating of
the matrix or to multi-photon excitation effects. Our own future efforts, in collaboration with theoretical modeling of the
cluster support system [22], will therefore aim at a detailed
microscopic understanding of the molecular origin of the observed light emission. Experimental insight into the electronic
structure of the cluster support system is expected from the
investigation of mass-selectively deposited silver clusters and
time-resolved spectroscopic probes. Furthermore, the modification of the matrix and support material will open new routes
to tailor the optical properties of these promising luminescent
materials with respect to spectral sensitization and sensitivity
control.
ACKNOWLEDGEMENTS
The authors are grateful to
Dr. B. Stegemann for help in the initial phase of the experiments, to C. Ritter
and Prof. K. Rademann for the AFM analysis of the thin films, to Prof. R.
Menzel for providing the scanning microscope equipment, and to I. Wallat
and Prof. M.P. Heyn for support concerning the optical absorption measurements. Fruitful discussions with Prof. V. Bonacic-Koutecky are gratefully
acknowledged. This work was supported through priority program SPP 1153
of the Deutsche Forschungsgemeinschaft.

REFERENCES
1 J. Tominaga, D.P. Tsai (eds.), Optical Nanotechnologies (Springer,
Berlin, 2003)
2 W. Krasser, U. Kettler, P.S. Bechthold, Chem. Phys. Lett. 86, 223 (1982);
S. Fedrigo, W. Harbich, J. Buttet, J. Chem. Phys. 99, 5712 (1993);
L. Knig, I. Rabin, W. Schulze, G. Ertl, Science 274, 1353 (1996);
I. Rabin, W. Schulze, G. Ertl, J. Chem. Phys. 108, 5137 (1998); C. Felix,
C. Sieber, W. Harbich, J. Buttet, I. Rabin, W. Schulze, G. Ertl, Chem. Phys.
Lett. 313, 105 (1999); I. Rabin, W. Schulze, G. Ertl, C. Felix, C. Sieber, W.
Harbich, J. Buttet, Chem. Phys. Lett. 320, 59 (2000); D. Ievlev, I. Rabin,
W. Schulze, G. Ertl, Chem. Phys. Lett. 328, 142 (2000); D. Ievlev, I. Rabin,
W. Schulze, G. Ertl, Eur. Phys. J. D 16, 157 (2001); W. Schulze, I. Rabin,
G. Ertl, Chem. Phys. Chem. 5, 403 (2004)
3 C. Felix, C. Sieber, W. Harbich, J. Buttet, I. Rabin, W. Schulze, G. Ertl,
Phys. Rev. Lett. 86, 2992 (2001)
4 C. Sieber, J. Buttet, W. Harbich, C. Felix, R. Mitric, V. BonacicKoutecky, Phys. Rev. A 70, 041 201 (2004)
5 F. Federmann, K. Hoffmann, N. Quaas, J.P. Toennies, Eur. Phys. J. D
9, 11 (1999); T. Diederich, J. Tiggesbumker, K.H. Meiwes-Broer,
J. Chem. Phys. 116, 3263 (2002)
6 P. Radcliffe, A. Przystawik, T. Diederich, T. Dppner, J. Tiggesbumker,
K.H. Meiwes-Broer, Phys. Rev. Lett. 92, 173 403 (2004)
7 V. Bonacic-Koutecky, V. Veyret, R. Mitric, J. Chem. Phys. 115, 10 450
(2001)
8 L.A. Peyser, A.E. Vinson, A.P. Bartko, R.M. Dickson, Science 291, 103
(2001)
9 L.A. Peyser, T.-H. Lee, R.M. Dickson, J. Phys. Chem. B 106, 7725
(2002)
10 T.-H. Lee, J.I. Gonzalez, R.M. Dickson, Proc. Natl. Acad. Sci. USA 99,
10 272 (2002); T.-H. Lee, R.M. Dickson, Proc. Natl. Acad. Sci. USA
100, 3043 (2003); T.-H. Lee, R.M. Dickson, J. Phys. Chem. B 107,
7387 (2003); T.-H. Lee, C.R. Hladik, R.M. Dickson, Nano Lett. 3, 1561
(2003); T.H. Lee, C.R. Hladik, R.M. Dickson, Appl. Phys. Lett. 84, 118
(2004)
11 T. Gleitsmann, B. Stegemann, T.M. Bernhardt, Appl. Phys. Lett. 84,
4050 (2004)
12 P. Fayet, F. Granzer, G. Hegenbart, E. Moisar, B. Pischel, L. Wste,
Phys. Rev. Lett. 55, 3002 (1985); J.F. Hamilton, Adv. Phys. 37, 359
(1988); R.S. Eachus, A.P. Marchetti, A.A. Muenter, Annu. Rev. Phys.
Chem. 50, 117 (1999)
13 P. Ball, Designing the Molecular World (Princeton University Press,
Princeton, 1994); W. Babel, Chem. unserer Zeit 30, 86 (1996)
14 X.-Y. Pan, H.-B. Jiang, C.-L. Liu, Q.-H. Gong, X.-Y. Zhang, Q.-F. Zhang,
B.-X. Xu, J.-L. Wu, Chin. Phys. Lett. 20, 133 (2003); A. Maali,
T. Cardinal, T. Treguer-Delapierre, Physica E 17, 559 (2003)
15 U. Kreibig, Z. Phys. D 3, 239 (1986)
16 U. Kreibig, M. Vollmer, Optical Properties of Metal Clusters (Springer,
Berlin, 1995)
17 U. Kreibig, A. Althoff, H. Pressmann, Surf. Sci. 106, 308 (1981)
18 V. Hornebebecq, M. Antonietti, T. Cardinal, M. Treguer-Delapierre,
Chem. Mater. 15, 1993 (2003)
19 T. Shima, J. Tominaga, J. Vac. Sci. Technol. A 21, 634 (2003)
20 J. Zheng, R.M. Dickson, J. Am. Chem. Soc. 124, 13 982 (2002);
J.T. Petty, J. Zheng, N.V. Hud, R.M. Dickson, J. Am. Chem. Soc. 126,
5207 (2004)
21 W. Harbich, S. Fedrigo, F. Meyer, D.M. Lindsay, J. Lignieres,
J.C. Rivoal, D. Kreisle, J. Chem. Phys. 93, 8535 (1990); V. BonacicKoutecky, J. Pittner, M. Boiron, P. Fantucci, J. Chem. Phys. 110, 3876
(1999)
22 V. Bonacic-Koutecky, private communication
23 T. Gleitsmann, T.M. Bernhardt, to be published
24 C.D. Geddes, A. Parfenov, I. Gryczynski, J.R. Lakowicz, J. Phys.
Chem. B 107, 9989 (2003)

Das könnte Ihnen auch gefallen