Vous êtes sur la page 1sur 6

Home

Search

Collections

Journals

About

Contact us

My IOPscience

Photoacoustic Studies of Annealed CdSxSe1-x (x = 0.26) Nanocrystals in a Glass Matrix

This content has been downloaded from IOPscience. Please scroll down to see the full text.
1999 Jpn. J. Appl. Phys. 38 3163
(http://iopscience.iop.org/1347-4065/38/5S/3163)
View the table of contents for this issue, or go to the journal homepage for more

Download details:
IP Address: 130.153.73.137
This content was downloaded on 03/03/2015 at 00:18

Please note that terms and conditions apply.

Jpn. J. Appl. Phys. Vol. 38 (1999) pp. 31633167


Part 1, No. 5B, May 1999
c
1999
Publication Board, Japanese Journal of Applied Physics

Photoacoustic Studies of Annealed CdS x Se1x (x = 0.26) Nanocrystals


in a Glass Matrix
Qing S HEN and Taro TOYODA
Department of Applied Physics and Chemistry, The University of Electro-Communications, 1-5-1 Chofugaoka, Chofu, Tokyo 182-8585, Japan
(Received November 26, 1998; accepted for publication January 29, 1999)

Photoacoustic (PA) spectroscopy is applied to study an optical absorption of CdSx Se1x (x = 0.26) nanocrystals in a glass
matrix annealed at 700 C for different time durations. As the annealing time increases, a shift of the absorption peak (due to the
quantum confinement effect) to lower energy is observed in the PA spectra at room temperatures. The observed absorption peak
shifts are used to evaluate the average size of CdSx Se1x nanocrystals in terms of a simple model. The average radius increases
from 2.3 to 4.7 nm as the annealing time increases from 0 to 120 min. The PA signal intensities plotted in the semilogarithmic
scale vary linearly below the fundamental absorption edges in accordance with the Urbach rule for the optical absorption
coefficient. The steepness factor (slope of exponential optical absorption) of the PA spectra increases with increasing annealing
time. The change of the electronic states by annealing is explained by considering that the contribution of the surface states
decreases with the increase of the nanocrystal size.
KEYWORDS: CdSx Se1x nanocrystals, semiconductor doped glass, photothermal phenomena, photoacoustic spectroscopy,
annealing process, quantum confinement effect, Urbach rule, exponential optical absorption, surface states

1. Introduction
Semiconductor particles with sizes of the order of a few
nanometers are called semiconductor nanocrystals. The sizes
are larger than the lattice constants but comparable to the
spatial extension of the wave functions of excitons, electrons
or holes on the corresponding bulk semiconductor materials.
So quantum confinement effect occurs and the semiconductor nanocrystals have new interesting properties completely
different from those of bulk materials. For example, they
show size-dependent optical absorption and photoluminescence (PL) properties, and large optical nonlinearity. Thus,
they have potentials and promises both for the basic study
of the three-dimensional confinement in semiconductors and
for applications in the field of optoelectronic devices, such
as for fast switching, optical signal processing and solar energy conversion.1) The semiconductor nanocrystals can be
grown in different matrices, like glasses, solutions and polymers by different manufacturing methods. Semiconductordoped glass (SDG) is one typical example of the semiconductor nanocrystal materials. It typically consists of CdSx Se1x
(0 5 x 5 1) nanocrystals embedded in a glass matrix. It
has been widely used as a sharp-cut color filter glass in optics.2, 3) The study of optical properties, especially optical absorption property, is very important to understand the behavior of semiconductor nanocrystals. In general, the absorption spectra can be measured by the conventional transmitted absorption measurement method. However, the samples
should be sufficiently thin and have good quality surfaces by
pretreatment. Photoacoustic (PA) method is one of the photothermal detection techniques and it is proved to be useful
for investigating optical absorption and thermal properties of
various materials by measuring the nonradiative de-excitation
processes.410) Besides the important aspect of measuring optically opaque or scattering solid materials, the PA method
can also be used to measure low absorption coefficients with
high sensitivity.10, 11) In recent years, quantum confinement
effects on the excitonic transitions in CdS quantum dots embedded in a polymer matrix have been studied using PA spectroscopy and it was observed that excitonic transitions were
well resolved in PA spectra as compared to the correspond-

ing optical absorption spectra.12) The purpose of this paper


is to show the dependence of the PA spectra of a commercial
HOYA sharp-cut color filter (O56) with a thickness of 2.5 mm
(thick for transmitted absorption measurements) on annealing
time at heat treatment temperatures of 700 C. The mole fraction of sulfur of the CdSx Se1x nanocrystals in the O56 was
reported to be approximately x = 0.26 by the Raman scattering measurements.13)
2. Experimental
The filter samples were cut into suitable dimensions and
were annealed for 10, 20, 30, 60, 90 and 120 min in air after
the temperature was elevated to 700 C at a rate of 47 C/min.
Then they were quenched to room temperature. It has been
reported that the nanocrystal grows and the mole fraction is
almost unchanged during the annealing processes.2)
The PA cell was composed of an aluminum cylinder with
a small channel at its periphery in which a microphone was
inserted. The cell was suspended by four rubber bands to
avoid outside vibrations. In the PA measurements, a light
beam from a 300 W xenon arc lamp was passed through a
0.25 m monochromator with f /4.5 optics and a bandwidth of
12 nm to obtain monochromatic light. It was then modulated
by a mechanical chopper with a frequency of 33 Hz and focused on a sample placed inside a PA cell. The light power
reaching the sample was approximately 1 mW at 500 nm. The
inside volume of the cell was 0.5 cm3 . The light absorbed by
the sample is converted into heat by nonradiative processes,
which results in a pressure fluctuation of the air inside the
cell. The pressure fluctuation was detected by a sensitive microphone enclosed in the PA cell and the signal was amplified
by a preamplifier and a two-phase lock-in amplifier. PA measurements were carried out at room temperature in a wavelength range of 320 nm to 800 nm with 5 nm step scan. The
PA spectra were normalized using the PA spectrum obtained
from carbon black sample. X-ray diffraction (XRD) apparatus was used for the sample characterization using CuK radiation with a 0.3 receiving slit, a 1.0 divergence slit and a
tube power of 3.2 kW. No smoothing was used in the XRD
measurements. A Jeol JEM-2000FX transmission electron
microscope (TEM) operating at 200 kV was used to observe

3163

3164

Q. S HEN and T. TOYODA

Jpn. J. Appl. Phys. Vol. 38 (1999) Pt. 1, No. 5B

images of the nanocrystals in the sample.


3. Results and Discussion
Figure 1 shows the XRD patterns of the CdSx Se1x
(x = 0.26) nanocrystal-doped glasses annealed for different
times over a range of 2 from 40 to 55 . No peaks are seen
for the sample annealed for 20 min, while three peaks can be
observed for the samples annealed for more than 30 min. The
peaks become clear and their half-widths decrease slightly
upon increasing the annealing time. This indicates an increase
in the sizes of the nanocrystallites due to annealing. These
peaks are identified as reflections from the (110), (103), and
(200) planes of the hexagonal wurtzite structure for the sample annealed for a period of more than 60 min. Although the
position of the peaks of the sample annealed for a period of
less than 60 min is difficult to determine, no shift between
those patterns is observed. This implies no change in the selenium content with crystal growth within experimental accuracy.
Figure 2 shows a TEM micrograph of the unannealed sample. CdSx Se1x (x = 0.26) nanocrystals appear as dark particles with sizes of the order of a few nanometers. They are
randomly distributed in the glass matrix with a large particle
density.
All the samples were 2.5 mm thick. They were too thick to
obtain absorption peaks by quantum confinement effect exactly in the strong absorption region (above the absorption
edge) by the normal transmitted method, because the transmitted light intensities were so weak that the changes in absorption with the wavelength could not be distinguished.10)
Figure 3 shows the PA intensity spectra of several types of

Fig. 1. X-ray diffraction pattern of CdSx Se1x (x = 0.26) nanocrystal in


a glass matrix annealed at 700 C for 20, 60 and 120 min.

samples (annealed for 0, 10 and 90 min) measured under the


same experimental conditions (sample dimensions, optical
alignment and so on). The PA intensity spectra gradually in-

Fig. 2.

A TEM micrograph of unannealed sample.

Fig. 3. Photoacoustic intensity spectra of CdSx Se1x (x = 0.26)


nanocrystal in a glass matrix annealed at 700 C for 0, 10 and 90 min. The
arrows show the maximum peaks of the spectra just above the absorption
edges.

Jpn. J. Appl. Phys. Vol. 38 (1999) Pt. 1, No. 5B

Q. S HEN and T. TOYODA

crease from the absorption edges and the absorption peaks,


which may be due to the quantum confinement effects, can be
observed as indicated in Fig. 3 by arrows. The shapes of the
PA spectra change with the increase of annealing time. The
absorption edges and peaks in the PA spectra shift to lower energy with the increase of annealing time. A splitting of both
the valence and conduction bands into a series of subbands
occurs when the confinement potential dominates the overall interaction. The energy band in the nanocrystal becomes
discrete and the energy gap should increase with decreasing
nanocrystal size.14) The experimental results include the size
distribution effect and the obtained spectra are the summation
of those of the individual nanocrystals.15) We assume that the
lowest excited energies for each annealing time correspond
to the absorption peaks indicated by arrows in Fig. 3. The
lowest transition energy is between the n = 1, l = 0 hole
state and the n = 1, l = 0 electron state, while the next transition involves n = 1, l = 0 and n = 1, l = 1 states,2)
where n is the radial quantum number and l is the orbital angular momentum quantum number. The results can be qualitatively understood as follows. In the annealing processes, the
volume of the semiconductor nanocrystal grows linearly with
annealing time (t) corresponding to R (t)1/3 at a constant
temperature, where R is the average radius of the nanocrystals.1) The quantum confinement effect becomes weaker with
the increase of nanocrystal radius as shown in Fig. 3, namely
the absorption peaks in PA spectra shift to lower energy. Neglecting the correlation term coming from Coulombic interaction, the nanocrystal size can be approximately calculated
by
1E = E E g = h 2 /8R 2 ,

3165

Fig. 4. Dependence of the change of the average volume of the CdSx Se1x
(x = 0.26) nanocrystals in the annealed samples relative to that of the
unannealed sample on the annealing time.

(1)

where is the reduced electron-hole mass, E g is the bulk


crystal band gap, R is the average radius of nanocrystals, E is
the lowest energy electronic transition and h is the Plancks
constant (the effective mass approximation).2, 16) Using the
band gap energies of bulk CdS (2.42 eV) and bulk CdSe
(1.74 eV) as well as the fraction x(= 0.26), the band gap energy of bulk material (E g ) corresponding to the stoichiometry
of the CdSx Se1x mixed anion system in O56 is calculated
to be about 1.92 eV. E is the photon energy corresponding to
the absorption peaks in the PA spectra. The reduced electronhole mass of CdSx Se1x (x = 0.26) was determined to be
0.126 m 0 calculated from m e = 0.2 m 0 and m h = 1.34 m 0 for
CdS, and m e = 0.13 m 0 and m h = 0.74 m 0 for CdSe, where
me and m h are effective masses of the electrons and holes in
bulk crystals and m 0 is the free electron mass.17) Using equation (1), we estimated the average radius of the nanocrystals
in a glass matrix annealed at different times. The average radius R increases from 2.3 to 4.7 nm when the annealing time
increased from 0 to 120 min. Figure 4 shows the average
volume change 1V of nanocrystals in the annealed samples
relative to that in the unannealed sample as a function of annealing time, where 1V = V V0 . V0 and V are the averages
of volumes of the nanocrystals in the unannealed sample and
annealed one, respectively. The change of the average volume
of the nanocrystals in the glass matrix linearly increases versus the annealing time. This result suggests that the nanocrystals in a glass matrix grow by a diffusion process during the
annealing processes.17)
Below the absorption edge, our preliminary optical absorp-

Fig. 5. Logarithmic plots of the photoacoustic intensity spectra for


CdSx Se1x (x = 0.26) nanocrystals in a glass matrix below the absorption
edges.

tion measurements showed that this was the case for an optically opaque sample with a thermally thick condition (s < L
and s < where s , and L are the thermal diffusion
length, optical absorption length and sample thickness, respectively).4) This result shows that the PA signal intensity
below the absorption edge is proportional to the optical absorption coefficient.4) The intensities of the PA signals plotted
semilogarithmically vary linearly below the absorption edges
for different annealing times as shown in Fig. 5 in accordance
with the Urbach rule for the optical absorption coefficient (exponential tail).18) Investigations of these exponential tails can
provide information on the band structure, disorders, defects
and characteristics of electron-phonon interactions. An empirical relationship for the dependence of PA signal intensity P on the measuring temperature (T ) and photon energy
(h; : frequency) was fitted by
P = P0 exp[ (h h0 )/kT ],

(2)

where k is Boltzmans constant, and P0 , and 0 are fitting


parameters.19) is the steepness factor and is the slope of the
exponential optical absorption. To explain Urbach rule many
theoretical studies have been performed, but no explanation
has been universally accepted.1922) Figure 6 shows the de-

3166

Jpn. J. Appl. Phys. Vol. 38 (1999) Pt. 1, No. 5B

Fig. 6. Dependence of the steepness factor on the annealing time for the
CdSx Se1x (x = 0.26) nanocrystal in a glass matrix.

pendence of the steepness factor on the annealing time. It


is found that increases with the increase of annealing time.
Yamasaki demonstrated that the exponential tail of undoped
hydrogenated amorphous Si was correlated through thermal
as well as structural disorder and the weak absorption band
below the optical gap was tentatively attributed to singleoccupied dangling bonds.11) Zammit and coworkers showed
that the divacancies in implanted Si on sapphire strongly affect the population of band tail states (exponential optical absorption) and the annealing studies revealed that the progressive quenching of the divacancy band is followed.23, 24) In our
case, is the parameter of the surface states on nanocrystals. Two explanations can be considered for the increase of
value with the increase of annealing time. One is the decrease
of the surface state density relative to the electronic state density, because of the decrease of the ratio of the surface area
to the entire volume of the nanocrystals as the sizes increase.
Consequently, the capture cross section of the electrons to the
surface states in the small nanocrystal is larger than that in the
large one.15, 25) The other explanation is the increase of uniformity of the distribution of the nanocrystal sizes due to the
annealing effect. Figure 7 shows a linear relationship between
the steepness factor and the average radius. Therefore, the
former possibility stated above may be the dominant process
for the PA signal generation.
4. Conclusions
CdSx Se1x (x = 0.26)-doped glasses (HOYA O56) annealed at 700 C for various times are studied using PA spectroscopy. It was observed that the absorption edges and peaks
in the PA spectra shifted to lower energies with the increase
of annealing time. Unlike the conventional transmitted optical absorption measurements, the PA measurements give true
optical absorption spectra with high sensitivity even though
the samples are relatively thick. PA spectroscopy provides a
novel technique to measure quantum confinement effects of
CdSx Se1x nanocrystals in a glass matrix with large optical
density. It demonstrates that the quantum confinement effects
become weaker as the nanocrystal radii increase. From the
lowest transition energy in the spectra, we can estimate the average radius of nanocrystals using the physical parameters of
bulk crystals. The average radius changes from 2.3 to 4.7 nm
as the annealing time increases from 0 to 120 minutes. The
steepness factor (slope of exponential optical absorption) of

Q. S HEN and T. TOYODA

Fig. 7. Dependence of the steepness factor on the average radius for the
CdSx Se1x (x = 0.26) nanocrystal in a glass matrix.

the PA spectra increases with the increase of annealing time,


showing the changes of electronic states of the CdSx Se1x
(x = 0.26) nanocrystals during the annealing process. These
results demonstrate the usefulness of the PA method to study
the optical properties of nanocrystal-doped glasses.
PL measurements are necessary to provide detailed information regarding the defect states and for comparative studies
between the radiative and nonradiative processes of nanocrystals in a glass matrix. The temperature dependence of the PA
and PL spectra is also needed to clarify the nature of the electronic states.
Acknowledgements
Part of this work was supported by a Grant-in Aid for Scientific Research (No. 09650887) from the Ministry of Education, Science, Sports and Culture. We would like to thank
Professor N. Yamada for permitting us to use his XRD apparatus and Dr. J. Shi for his help in the TEM experiment. We
also thank S. Fukushima and K. Shinoyama for their help in
the experiments.

1) U. Woggon: Optical Properties of Semiconductor Quantum Dots, ed.


G. Hohler (Springer-Verlag, Berlin, 1996) p. 2.
2) N. F. Borrelli, D. W. Hall, H. J. Holland and D. W. Smith: J. Appl. Phys.
61 (1987) 5399.
3) N. R. Kulish, V. P. Kunets and M. P. Lisitsa: Opt. Eng. 34 (1995) 1054.
4) A. Rosencwaig and A. Gersho: J. Appl. Phys. 47 (1976) 64.
5) A. Mandelis and E. K. M. Siu: Phys. Rev. B 34 (1986) 7209.
6) A. Pinto Neto, H. Vargas, N. F. Leite and C. M. Miranda: Phys. Rev. B
41 (1990) 9971.
7) A. Yarai, K. Sakamoto and T. Nakanishi: Jpn. J. Appl. Phys. 33 (1994)
3255.
8) A. Zagadi, I. S. Al-Saffer, M. V. Yakushev and R. D. Tomlinson: Rev.
Sci. Instrum. 66 (1995) 4095.
9) T. Toyoda and K. Kato: Jpn. J. Appl. Phys. 35 (1996) 2912.
10) Q. Shen, Y. Kato and T. Toyoda: Jpn. J. Appl. Phys. 36 (1997) 3297.
11) S. Yamasaki: Philos. Mag. B 56 (1987) 79.
12) P. Nandakumar, A. R. Dhobale, Y. Babu, M. D. Sastry, C. Vijayan,
Y. V. G. S. Murti, K. Dhanalakshmi and G. Sundararajan: Solid State
Commun. 106 (1998) 193.
13) T. Miyoshi, T. Nakatsuka and N. Natsuo: Jpn. J. Appl. Phys. 34 (1995)
1835.
14) L. E. Brus: IEEE J. Quantum. Electron. 22 (1986) 1909.
15) T. Arai, H. Fujimura, I. Umezu, T. Ogawa and A. Fujii: Jpn. J. Appl.
Phys. 28 (1989) 484.
16) Y.-S. Yuang, Y.-F. Chen, Y.-Y. Lee and L.-C. Liu: J. Appl. Phys. 76

Jpn. J. Appl. Phys. Vol. 38 (1999) Pt. 1, No. 5B


(1994) 3041.
17) A. I. Ekimov, Al. L. Efros and A. A. Onuschenko: Solid State Commun.
56 (1985) 921.
18) F. Urbach: Phys. Rev. 92 (1953) 1324.
19) J. D. Dow and D. Redfield: Phys. Rev. B 5 (1972) 594.
20) H. Sumi and Y. Toyozawa: J. Phys. Soc. Jpn. 31 (1971) 342.
21) H. Schreiber and Y. Toyozawa: J. Phys. Soc. Jpn. 51 (1982) 152.
22) T. Skettrup: Phys. Rev. B 18 (1978) 2622.

Q. S HEN and T. TOYODA

3167

23) U. Zammit, F. Gasparrini, M. Marinelli, R. Pizzoferrato, A. Agostini, F.


Mercuri, F. Scudieri and S. Martellucci: Photoacoustic and Photothermal Phenomena III, ed. D. Bicanic, (Springer-Verlag, Berlin, 1992)
Springer Series of Optical Sciences, p. 393.
24) U. Zammit, K. N. Madhusoodanan, M. Marinelli, F. Scudieri, R.
Pizzoferrato, F. Mercuri, E. Wendler and W. Wesch: J. de Phys. IV 4
(1994) Suppl. C7, C7-113.
25) T. Arai, T. Yoshida and T. Ogawa: Jpn. J. Appl. Phys. 26 (1987) 396.

Vous aimerez peut-être aussi