Sie sind auf Seite 1von 8

JOURNAL OF APPLIED PHYSICS 101, 103112 2007

Silicon nanoparticles: Absorption, emission, and the nature of the electronic bandgap
Cedrik Meier,a Andreas Gondorf, Stephan Lttjohann, and Axel Lorke
Experimental Physics and CeNIDE, University of DuisburgEssen, Lotharstrae 1, D-47048 Duisburg, Germany

Hartmut Wiggers
Institute for Combustion and Gas Dynamics, University of DuisburgEssen, Lotharstrae 1, D-47048 Duisburg, Germany

Received 18 January 2007; accepted 21 February 2007; published online 30 May 2007 Silicon nanoparticles synthesized in the gas phase are studied. From time-resolved photoluminescence measurements we determine, quantitatively, the size-dependence of the oscillator strength of the nanoparticles. We investigate experimentally the absorption and photoluminescence emission of nanoparticle ensembles with a broad size distribution. Using a model which accounts for size-effects in both oscillator strength and quantum-connement, we are able to calculate absorption and emission spectra of ensemble samples. From these results we have determined, whether silicon nanoparticles should be regarded as indirect or direct semiconductors. Moreover, we systematically study the inuence of the particle size-distribution on the optical spectra. 2007 American Institute of Physics. DOI: 10.1063/1.2720095
I. INTRODUCTION

Today, silicon Si is the most important material for electronic applications, e.g., in the microelectronics industry. This is especially due to its high availability and low cost, as well as its specic electronic and material properties, which allow the fabrication of planar ultralarge scale integrated electronic devices. For optoelectronic applications, silicon has widely been used in photodetectors or photovoltaic cells. For the use as an active material in light emitting devices, however, silicon has not been considered to be attractive. The reason for this is the indirect nature of the electronic bandgap, which requires the involvement of phonons for an optical recombination process. Therefore, the optical emission of bulk silicon is very weak in comparison to semiconductors with direct bandgaps, such as GaAs. With the advent of nanostructured silicon, this changed. The discovery of photoluminescence from porous silicon by Canham1 triggered a strong effort of research in this eld. Major milestones were the observation of amplied stimulated emission ASE2 and the demonstration of electroluminescence using ac-eld effect injection in silicon nanoparticles.3 Especially the latter experiment strongly supports that silicon nanoparticles have the potential to be used in optoelectronic devices. However, to successfully realize such optoelectronic devices, it is mandatory to better understand both emission and absorption in these nanoparticles. Here, it is of particular importance to take the specic properties of nanoparticle ensembles into account, such as their size distribution and quantum size effects. In this paper, we will discuss the oscillator strength of silicon nanoparticles and its implication on the absorption
a

and photoluminescence spectra thereof. The size-dependence of the oscillator strength is directly derived from timeresolved photoluminescence measurements. We will then address the question if silicon nanoparticles should be considered as direct or indirect semiconductors. We will introduce a simple model which allows us to calculate both absorption and PL emission with a few simple assumptions. Finally, we will use this model to t our photoluminescence data and compare the results for particle size and geometric standard deviation with results from transmission electron microscopy TEM and atomic force microscopy AFM measurements.

II. EXPERIMENTAL DETAILS

URL: http://aglorke.uni-duisburg.de; electronic mail: cedrik.meier@unidue.de

We investigate silicon nanoparticles fabricated by rfstimulated dissociation of silane SiH4. The fabrication process has been described by Knipping et al.4 The nanoparticles are mostly spherical and single crystalline. The particle diameter has been determined by the Brunauer-Emmet-Teller BET method5 and by analyzing atomic force micrographs AFM. The particles studied in this work have diameters between d = 4 nm and d = 20 nm. As the particles are not sizeselected, we nd a log-normal size distribution for the nanoparticle ensembles6 with a geometric mean deviation of typically = 1.3. A more detailed discussion on the impact of mean deviation will be presented later in this paper. For the absorption spectroscopy, the nanoparticles have been dispersed in water and deposited on quartz substrates. The absorption spectroscopy was carried out at room temperature in an evacuated Fourier-transform spectrometer, using a mercury-cadmium-telluride MCT photoconductive detector. The photoluminescence measurements have been carried out in a confocal microphotoluminescence setup. As an excitation source, a frequency-doubled Nd: YVO4 laser at 532
2007 American Institute of Physics

0021-8979/2007/10110/103112/8/$23.00

101, 103112-1

Downloaded 30 May 2007 to 134.91.45.15. Redistribution subject to AIP license or copyright, see http://jap.aip.org/jap/copyright.jsp

103112-2

Meier et al.

J. Appl. Phys. 101, 103112 2007

nm was used. Both excitation and collection were performed using a 10 microscope objective with a numerical aperture of NA= 0.25.
III. OSCILLATOR STRENGTH

As it is essential to our discussion of both absorption and photoluminescence experiments, we will start with our results on the size dependence of the oscillator strength. The radiative recombination rate RR = 1 / R is linked to 2 and the rethe optical transition matrix element 0p 1 combination frequency via Fermis Golden Rule7
1 R = RR =

e 2n 2 . 0p 1 0m 2 c 3

FIG. 1. Color online Photoluminescence decay of silicon nanoparticles for an emission energy of E = 1.67 eV.

We use the common denition for the oscillator strength f f osc = 2 2 . 0p 1 m 2 dence of the oscillator strength can be tted over the entire energy range using an exponential function. It should be made clear that this type of function is only chosen to t the energy dependence well and that there is no physical mechanism connecting oscillator strength and energy using such an analytical expression. The t is shown as the solid line in Fig. 2. From these results, the energy dependence of the oscillator strength can be described using the following empirical formula:

Combining the above equations, the radiative recombination time R can be directly linked to the oscillator strength7 f osc = 20mc3 1 . e 2n 2 R 3

In the above equations, we assume that the nanoparticles are in the weak connement regime, i.e., we use for the exciton mass m = me + mh, the sum of the electron and hole mass. For the latter, we use me = 0.19m0 and mh = 0.286m0 after Xia and Ioffe.9,10 It is, therefore, possible to determine the oscillator strength directly using time-resolved spectroscopy. However, one has to take into account that the photoluminescence decay is usually a combination of radiative and nonradiative decay processes. Thus, the PL decay time PL is linked to the radiative and nonradiative decay times via
1 1 1 PL = R + NR .

f oscEeV = 1.4 105 + 5.80 108 exp

EeV . 0.332 5

Delerue et al.19 have suggested the following size dependence for the electronic bandgap of silicon nanoparticles based on photoluminescence results and simulations using a linear combination of atomic orbitals LCAO: 3.73 . dnm1.39

4 EgeVd = E0eV +

In the case of silicon nanoparticles we have shown in a separate experiment11 that the nonradiative lifetime is 200 s. Therefore, we can easily link the radiative lifetime R to the experimentally measured photoluminescence decay time PL. Such a PL decay trace recorded using time-resolved spectroscopy is shown in Fig. 1. It can be seen that the silicon nanoparticle photoluminescence exhibits a clear single exponential decay, unlike other silicon nanocrystal systems.2,12 In these systems, a stretched exponential decay is found which obscures an accurate determination of the PL decay and, thus, of the radiative recombination time R. Using the above formalism, we can now determine the oscillator strength f osc directly from the measured PL decay times. The results are shown in Fig. 2. In this gure, our ndings for the radiative recombination times are plotted on the left axis as a function of the PL emission energy E = . On the right axis, the results for the oscillator strength are shown. It can be seen that the oscillator strength increases with increasing PL emission energy, as is expected due to the stronger connement in smaller particles. The energy depen-

Using this equation, we can directly obtain the size dependence of the oscillator strength

FIG. 2. Color online Radiative recombination lifetime R and oscillator strength f as a function of emission energy .

Downloaded 30 May 2007 to 134.91.45.15. Redistribution subject to AIP license or copyright, see http://jap.aip.org/jap/copyright.jsp

103112-3

Meier et al.

J. Appl. Phys. 101, 103112 2007

est particles. Similar high quantum efciencies have also been reported by other groups for silicon nanoparticles produced in a very similar fashion.14,15 This result clearly demonstrates that a high quantum efciency does not automatically indicate a high emission rate.
IV. ABSORPTION A. Experimental results

FIG. 3. Color online Size dependence of the oscillator strength f for silicon nanoparticles obtained from Eq. 7.

f oscdnm = 1.4 106 + 1.7 106 exp

11.24 . dnm1.39 7

The size dependence according to Eq. 7 is plotted in Fig. 3. It can be seen that for particle diameters between 2 and 10 nm the oscillator strength is in the order of 105. For particles below 2 nm, the oscillator strength increases over several orders of magnitude and reaches unity for particle diameters of d 1 nm. However, one should note that the formula by Delerue et al. was used for the derivation of Eq. 7, which leads to a singularity for d 0. Also, values for the radiative recombination times have only been sampled for energies up to E = 2.1 eV, corresponding to a particle size of about d = 2.6 nm. Therefore, one should be cautious to use Eq. 7 for particle sizes d 2.5 nm. As mentioned before, for nanoparticle sizes d 2.5 nm we obtain values for the oscillator strength in the order of 105. For a quantum mechanically forbidden transition, the oscillator strength is typically f osc 108, while an allowed transition should yield an oscillator strength of at least 102. The recombination process in silicon nanoparticles is, therefore, in an intermediate regime. The excitons in silicon nanoparticles behave like bound excitions which show, in contrast to free excitions, a large oscillator strength. After Rashba and Gurgenishvili,13 the oscillator strength of a bound exciton f BE can be almost four or ve orders of magnitude larger than the oscillator strength f osc of free excitons7 f BE =

In this part of the paper we will address the question, if silicon nanoparticles should be regarded as indirect or direct semiconductors. In fact, this is not such a clear decision. While bulk silicon shows only an extremely weak emission, the photoluminescence from silicon nanoparticles is orders of magnitude stronger and shifted to higher energies. The reason for the weak photoluminescnce in bulk silicon is the fact that the involvement of phonons makes radiative recombinations a second-order process. Therefore, one might guess that a more efcient radiative recombination is due to a transition to a direct semiconductor caused by the quantum connement in the nanoparticles. In the cluster regime, Smith et al. explain the nding of fast decay rates by the transition to a direct semiconductor.16 Another remarkable fact is that amorphous silicon, which is a direct semiconductor, also has a blue-shifted bandgap energy in comparison to bulk silicon. Additionally, for amorphous silicon a red-shift of the phonon modes in Raman spectroscopy is observed, similar to the red-shift of the TO-phonon in silicon nanoparticles usually explained by the phonon connement effect.17 Therefore, it is necessary to assess this question spectroscopically. The experimental results from absorption spectroscopy are shown in Fig. 4. From the raw data a linear background has been subtracted to remove the inuence of Rayleigh scattering from the data. More details on the light scattering properties of the particles can be found in Ref. 18. The data in Fig. 4 have been plotted as 1/2 as a function of the photon energy . In such a plot, a linear behavior is expected for semiconductors with indirect bandgaps, for which one expects an energy dependence of the fundamental interband absorption as given by

1 Eg2 .

10

aex 3 f osc , a

where a is the unit-cell length and aex is the excitonic Bohr radius. This shows the strong localization in the silicon nanoparticles. Apart from the oscillator strength, we can also estimate the quantum efciency of the nanoparticles using

1 R

1 R 1 . + NR

As can be seen from Fig. 4, for all particle sizes between d = 4.3 nm and d = 19.2 nm, a clear absorption edge corresponding to the electronic bandgap can be found. Moreover, one can see that indeed above the bandgap, there is a soft onset, followed by an almost linear increase as expected for indirect bandgap semiconductors. The observed soft absorption edge can either be due to phonon emission/absorption or to effects from the particle size distribution. In case of phonon effects, absorption/emission of phonons already leads to a signicant increase in the absorption below the bandgap, as parts of the energy required for electronic excitations is provided from the phonon bath, i.e., the nal state energy is given by E f = Ei + , 11 where is the photon energy and is the phonon energy. However, apart from phonon effects, the particle size distri-

As mentioned before, we have concluded that NR = 200 s for the entire studied energy range. From this we yield quantum efciencies of up to = 34% and = 86% for the small-

Downloaded 30 May 2007 to 134.91.45.15. Redistribution subject to AIP license or copyright, see http://jap.aip.org/jap/copyright.jsp

103112-4

Meier et al.

J. Appl. Phys. 101, 103112 2007

FIG. 5. Color online Absorption data for silicon nanoparticles plotted as d2, corresponding to direct semiconductors.

FIG. 4. Color online Absorption data for silicon nanoparticles with diameters between d = 4.3 nm and 19.2 nm, plotted as 1/2, corresponding to an indirect semiconductor. Dashed lines Extrapolations for determination of the electronic bandgap.

tal data and the linear t. It can be seen that for the upper plot indirect semiconductor behavior, the agreement is very good for an energy interval as large as 100 meV, while for the lower plot direct semiconductor behavior the same quality of agreement is only reached in a range of about 25 meV, and the t location is completely random. This clearly shows that the smallest particles investigated in this study still exhibit indirect semiconductor behavior. From the data plotted in Fig. 4 we have determined the bandgap using linear extrapolation. The used linear approximations are plotted in Fig. 4 as dashed lines. To validate the accuracy of our method, we have determined the electronic bandgap of bulk silicon commercial wafer and of microcrystalline silicon powder. For bulk silicon, we have obtained a bandgap of Eg = 1.121 eV, and for the microcrystalline powder a bandgap of Eg = 1.081 eV data not shown here. Therefore, we assume that the method has an experimental inaccuracy of Eg = 50 meV. The results for the nanoparticles are summarized in Table I and plotted together with results from theoretical considerations discussed later in this paper in Fig. 7. It can be seen that the experimentally obtained bandgaps increase with decreasing particle sizes down to d 5 nm. For

bution can also play a role in the soft onset in the absorption measurements. We will discuss this later in more detail, together with the results from simple theoretical calculations. To be sure that indeed the absorption behavior supports the assumption that silicon nanoparticles are indirect semiconductors, we have also plotted the experimental data in Fig. 5 as 2. In such a plot, a linear behavior is expected for direct bandgap semiconductors. As can be seen from Fig. 5, the experimental data certainly does not show a linear behavior. This is highlighted for the smallest investigated particle size of d = 4.3 nm in Fig. 6. In the upper plot, the experimental absorption data are plotted as 1/2, corresponding to indirect semiconductor behavior. In the lower part, we have plotted the data as 2, corresponding to direct semiconductor behavior. Both plots are normalized for unity at E = 1.5 eV in order to be able to quantify the quality of the linear ts. In both plots, the linear ts used are plotted together with the difference between the experimen-

FIG. 6. Color online Experimental absorption for the d = 4.3 nm sample plotted together with linear ts and the differences between the experimental data and the t for upper plot indirect semiconductor behavior and lower plot direct semiconductor behavior.

Downloaded 30 May 2007 to 134.91.45.15. Redistribution subject to AIP license or copyright, see http://jap.aip.org/jap/copyright.jsp

103112-5

Meier et al.

J. Appl. Phys. 101, 103112 2007

TABLE I. Electronic bandgaps for silicon nanoparticles used in this study. Experimental values determined from absorption, monodisperse bandgaps obtained after Delrue et al. Ref. 20, and ensemble data obtained from the model described in the text assuming a log-normal particle size distribution with a geometrical standard deviation of = 1.3. Bandgap exp. eV 1.310 1.233 1.310 1.326 1.286 1.245 Bandgap monodisperse eV 1.611 1.581 1.518 1.276 1.247 1.181 Bandgap ensemble eV 1.405 1.392 1.382 1.258 1.239 1.179 FIG. 8. Color online Calculated absorption spectra.

Particle diameter nm 4.3 4.5 5.0 9.8 11.4 19.2

the smallest particles used in this study, the experimentally determined bandgap uctuates between 1.23 and 1.30 eV. In the equation suggested by Delerue et al.,19 the particles are assumed to be monodisperse. The bandgap obtained from Eq. 6 is plotted in Fig. 7 as the solid line together with experimental values squares. It can be seen, that for larger particles the bandgap is underestimated by Eq. 6. For nanoparticles with diameters below 5 nm, however, the experimentally determined bandgap is signicantly lower, even when the scatter of the data points due to uncertainties in determination of the particle size and the absorption edge are taken into account.
B. Theoretical description

the paper Eq. 7 we have derived an expression for the size dependence of the oscillator strength. From the calculated spectra, we will then determine the electronic bandgap in a similar fashion as we have used for the experimental data. In our approach, we assume that an individual nanoparticle behaves as an indirect semiconductor and has an absorption spectrum as described by Eq. 10. This assumption is justied by the experimental results from absorption and time-resolved PL measurements in the previous sections. For the particle size distribution, we assume a lognormal size distribution6 1 1 1 ln2 dn d , d 0, d/d0 = exp . 2 2 2 ln d ln dd

12

As can be seen from the experimental data discussed above, there is a signicant difference between Eq. 6, which is applicable to mondodisperse particles, and the experimental results obtained from a broad particle ensemble. In order to ll this gap and to study the inuence of particle size distribution, we develop a simple model to calculate the absorption properties of silicon nanoparticle ensembles. Here, we can benet from the fact that in the rst section of

Here, d0 is the mean particle size and is the geometrical standard deviation. From TEM measurements see, e.g., Knipping et al.4, we know that our particle ensembles have a standard deviation typically of 1.3. To determine the electronic bandgap for a given particle size, we use Eq. 6. Then we can obtain the spectral absorbance AE of a particle ensemble with a mean diameter of d0 and a geometrical standard deviation by integrating over the particle size distribution, the oscillator strength Eq. 7, and the spectral absorbance of an individual particle A = 1

f osc d

dn d , d 0, . 13 Eg d2dd dd

FIG. 7. Color online Size dependence of the bandgap of silicon nanoparticles, experimentally determined from absorption squares, for monodisperse nanoparticles following Delerue et al. solid line, and for particle ensembles with a size distribution with a geometrical standard deviation of = 1.3 dotted diamond.

The resulting spectra obtained using this method are shown in Fig. 8 for particle diameters in the range between d = 4.3 nm and d = 14.0 nm. In this plot, the data are again plotted as 1/2 as a function of the photon energy . The simulated spectra represent the measured absorption spectra fairly well. A systematic blue-shift of the absorption edge is resembled, as it is expected due to the use of the Eq. 6 for the size-dependence of the bandgap of an individual nanoparticle. While on the high-energy side of the absorption edge the linear behavior expected in this plot for an indirect semiconductor is very visible, for lower energies a soft tail in the calculated absorption spectra is found. This soft tail becomes more pronounced when the particle size is decreased. As we have not included any phonon processes in the depen-

Downloaded 30 May 2007 to 134.91.45.15. Redistribution subject to AIP license or copyright, see http://jap.aip.org/jap/copyright.jsp

103112-6

Meier et al.

J. Appl. Phys. 101, 103112 2007

FIG. 9. Color online PL spectra of silicon nanoparticles at room temperature.

dence in Eq. 10, it is clear that the particle size distribution is sufcient to explain this soft onset. As the quantum connement increases with decreasing particle size, it is evident that this effect must become more pronounced for particle distributions with a smaller mean particle diameter when the geometric standard deviation is kept constant. It is unclear why the increase in the tail below the absorption edge in the experimental data is not as pronounced as in the theoretically calculated spectra. On the other hand, it supports the assumption that in case of absorption spectra on nanoparticle ensemble spectra, the inuence of phonons can be neglected. In the same way as for the experimental data, from the theoretically calculated spectra the absorption edge has been determined by linear extrapolation. The results for the obtained bandgap energies are summarized in Table I and plotted in Fig. 7, together with the experimental values and the energies obtained from Eq. 6 for monodisperse particles. It can be seen that for the smaller particles, the values from the ensemble calculations describe the experimentally observed behavior much better than Eq. 6. However, it should be noted that the above equation is implicitly used in the ensemble bandgap values plotted here. The difference in energies is obviously mostly caused by the particle size distribution, in which particles always contribute to the absorption that are substantially larger than the mean particle diameter, thus reducing the observed bandedge energy.
V. PHOTOLUMINESCENCE

FIG. 10. Color online Comparison between experimental PL data room temperature and spectra obtained by a simple numerical model.

the power law dependence of the electronic bandgap, to a larger effect for the smaller particles. It should also be noted that a decrease in the mean particle diameter of only 1.1 nm from d = 5.2 nm to d = 4.1 nm already leads to a peak shift of nearly 400 meV. Starting from the size dependence of the oscillator strength discussed above and the power law dependence for the electronic bandgap, we can also develop a simple model for calculating the PL emission spectra of a particle ensemble. In this case, we make the assumption that a single nanoparticle at room temperature acts as an inhomogeneously broadened emitter with a spectral width of E. The PL intensity for a single particle with a peak emission energy of Eg is then given by INP, Eg, E E2 1 exp

In this section of the paper, we will discuss ensemble photoluminescence measurements at room temperature and extend the above discussed simple model to the PL emission spectra. We will also discuss the validity of the experimental data obtained from the t to the model. PL measurements on nanoparticle ensembles with sizes between d = 4.1 nm and d = 5.2 nm are shown in Fig. 9. Although the PL emission from a single particle is expected to be narrow, the measured ensemble spectra exhibit a broad single peak. Depending on the mean particle diameter the peak width varies between 250 and 400 meV. For smaller particles, the energetic width of the peak is broader. This observation strongly suggests that the peak broadening is caused by the particle size distribution which leads, due to

E g 2 . 2E2

14

We can then compute the PL emission of a nanoparticle ensemble with mean particle diameter of d0 using I =

f osc d

dn d , d 0, . INP, Eg,E ddd dd

15

Using the above equation we can compute the PL emission spectra of the silicon nanoparticle ensemble, taking fully into account, the inuence of the size dependence of the oscillator strength. We have used this equation to t the experimental data shown in Fig. 9. Representative results for d = 4.5 nm and d = 5.2 nm are shown in Fig. 10. It can be seen

Downloaded 30 May 2007 to 134.91.45.15. Redistribution subject to AIP license or copyright, see http://jap.aip.org/jap/copyright.jsp

103112-7

Meier et al.

J. Appl. Phys. 101, 103112 2007

ticle ensemble sample does not decrease signicantly when cooling the sample to liquid Helium temperatures not shown here. From these results it is obvious that even using such a model it is impossible to determine from the ensemble data a meaningful value for the spectral width E of a single nanoparticle oscillator. This value can only be determined from either single nanoparticle spectroscopy or samples with a very low size dispersion. Although the signicance of the determination of the spectral width is only limited, the calculated spectra are very sensitive to slight changes in the geometrical standard deviation of the particle size distribution. In the lower part of Fig. 11 one can see calculated spectra with the mean particle diameter d and the spectral width E kept xed. The geometrical standard deviation is varied between 1.1 and 1.3, which represents a realistic range of values depending on the particle synthesis process. It can be seen that a change of only 0.05 already leads to a signicant change in the spectral shape, especially in the width of the calculated PL peak. As the inuence of the tting parameter E, as discussed above, is not so pronounced, it can be concluded that the values for obtained from the t should be very reliable.
VI. CONCLUSIONS
FIG. 11. Color online Comparison between the inuence of the spectral width of a single nanoparticle oscillator E and the geometrical standard deviation on the ensemble PL spectra.

that the agreement between the experimental emission curve and the calculated curve is quite convincing. Especially, the asymmetry of the peaks toward the high-energy wing is well reproduced. This asymmetry is at least partly induced by the higher oscillator strength of smaller nanoparticles in the particle ensemble. For tting the data, we have kept the spectral width of a single silicon nanoparticle xed at E = 70 meV and varied the particle diameter d and the geometrical standard deviation . It can be seen that the difference between the particle diameter obtained by BET measurements from the powder and the t to the above discussed model is quite good, in the worst case a difference of 0.7 nm is found. Taking the intrinsic uncertainties of the BET method into account density variations, etc., this is to be considered an extremely good agreement. The geometric standard deviations obtained from the t vary between = 1.25 and 1.3. In fact, these results agree very well with results obtained from TEM measurements on these particles discussed elsewhere.4 However, the question still arises on the sensitivity of the model to the different t parameters. Therefore, we have kept the particle diameter xed at d = 4.7 nm and varied either the spectral width of a single nanoparticle E or the geometric standard deviation within experimentally reasonable limits. The results for these model calculations are shown in Fig. 11. It can be seen that a change in the spectral width of a single nanoparticle in the range between 10 and 100 meV is not expected to have a great impact on the emission spectra. We believe that this is also the reason for the experimental observation that the spectral width of an silicon nanopar-

From the above discussed experiments and calculations, we draw the following conclusions. First, the oscillator strength of silicon nanoparticles is size dependent. We have given an empirical formula to determine the oscillator strength quantitatively, which should deliver reliable results for particle sizes down to d = 2.5 nm. For particle sizes above 2.5 nm the oscillator strength is about 105, which is in an intermediate range between allowed and forbidden transitions, compared, i.e., to atomic physics. Second, silicon nanoparticles are indirect semiconductors. This could be shown in time-resolved PL and stationary absorption experiments. Including the results for the oscillator strength, we have developed a simple model to calculate both, absorption and emission spectra numerically, using straightforward assumptions. The results support the assumption that Si nanoparticles are indirect semiconductors. The inclusion of the size dependence of the bandgap rst introduced by Delerue et al.19 does lead to absorption bandedges and PL peak energies which agree very well with the experiment.
ACKNOWLEDGMENTS

Financial support by the Deutsche Forschungsgemeinschaft DFG under Grant No. SFB 445 Nanoparticles from the gas phase and Grant No. GRK 1240 Nanotronics is gratefully acknowledged.
L. T. Canham, Appl. Phys. Lett. 57, 1046 1990. L. Pavesi, L. DalNegro, C. Mazzoleni, G. Franzo, and F. Priolo, Nature 408, 440 2000. 3 R. J. Walters, G. I. Bourianoff, and H. A. Atwater, Nat. Mater. 4, 143 2005. 4 J. Knipping, H. Wiggers, B. Rellinghaus, P. Roth, D. Konjhodzic, and
1 2

Downloaded 30 May 2007 to 134.91.45.15. Redistribution subject to AIP license or copyright, see http://jap.aip.org/jap/copyright.jsp

103112-8

Meier et al.
14

J. Appl. Phys. 101, 103112 2007 D. Jurbergs, E. Rogojina, L. Mangolini, and U. Kortshagen, Appl. Phys. Lett. 88, 233116 2006. 15 L. Mangolini, E. Thimsen, and U. Kortshagen, Nano Lett. 5, 655 2005. 16 A. Smith, Z. H. Yamani, N. Roberts, J. Turner, S. R. Habbal, S. Granick, and M. H. Nayfeh, Phys. Rev. B 72, 205307 2005. 17 C. Meier, S. Lttjohann, V. G. Kravets, H. Nienhaus, A. Lorke, and H. Wiggers, Physica E Amsterdam 32, 155 2006. 18 V. G. Kravets, C. Meier, D. Konjhodzic, A. Lorke, and H. Wiggers, J. Appl. Phys. 97, 084306 2005. 19 C. Delerue, G. Allan, and M. Lannoo, Phys. Rev. B 48, 11024 1993.

C. Meier, J. Nanosci. Nanotechnol. 4, 1039 2004. S. Brunauer, P. Emmet, and E. Teller, J. Am. Chem. Soc. 60, 309 1938. 6 C. G. Granqvist and R. Buhrmann, J. Appl. Phys. 47, 2200 1976. 7 G. W. t Hooft et al., Phys. Rev. B 35, 8281 1987. 8 C. Henry, Phys. Rev. B 1, 1628 1970. 9 J. B. Xia, Phys. Rev. B 40, 8500 1989. 10 A. D. Yoffe, Adv. Phys. 42, 173 1993. 11 S. Lttjohann, C. Meier, M. Offer, A. Lorke, and H. Wiggers Europhys. Lettsubmitted. 12 M. L. Brongersma et al., Appl. Phys. Lett. 76, 351 2000. 13 E. I. Rashba and G. E. Gurgenishvili, Sov. Phys. Solid State 4, 759 1962.
5

Downloaded 30 May 2007 to 134.91.45.15. Redistribution subject to AIP license or copyright, see http://jap.aip.org/jap/copyright.jsp

Das könnte Ihnen auch gefallen