06.01.2015 Views

Photoluminescence studies of isotopically enriched - Max Planck ...

Photoluminescence studies of isotopically enriched - Max Planck ...

Photoluminescence studies of isotopically enriched - Max Planck ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

phys. stat. sol. (b) 235, No. 1, 63–74 (2003) / DOI 10.1002/pssb.200301540<br />

Editor’s Choice<br />

<strong>Photoluminescence</strong> <strong>studies</strong> <strong>of</strong> <strong>isotopically</strong> <strong>enriched</strong> silicon<br />

D. Karaiskaj *; 1 , M. L. W. Thewalt 1 ,T.Ruf 2 , and M. Cardona 2<br />

1 Department <strong>of</strong> Physics, Simon Fraser University, Burnaby, British Columbia, V5A 1S6, Canada<br />

2 <strong>Max</strong>-<strong>Planck</strong>-Institut fçr Festkærperforschung, 70569 Stuttgart, Germany<br />

Received 24 July 2002, revised 19 August 2002, accepted 19 August 2002<br />

Published online 30 December 2002<br />

PACS 63.20.Dj, 63.20.Ry, 78.55.Ap<br />

We report the first high resolution photoluminescence <strong>studies</strong> <strong>of</strong> <strong>isotopically</strong> pure silicon. New information<br />

is obtained on isotopic effects on the indirect band gap energy, phonon energies, and phonon<br />

broadenings, which is in good agreement with previous results obtained in germanium and diamond.<br />

Remarkably, the line widths <strong>of</strong> the no-phonon boron and phosphorus bound exciton transitions in the<br />

28 Si sample (99:896% 28 Si) are much sharper than in natural Si, revealing new fine structure in the<br />

boron bound exciton luminescence. Most surprisingly, the small splittings <strong>of</strong> the neutral acceptor<br />

ground state in natural Si are absent in the photoluminescence spectra <strong>of</strong> acceptor bound excitons in<br />

<strong>isotopically</strong> purified 28 Si, demonstrating conclusively that they result from the randomness <strong>of</strong> the Si<br />

isotopic composition.<br />

1. Introduction The effects <strong>of</strong> isotopic composition and disorder on the properties <strong>of</strong> semiconductors<br />

have been the subject <strong>of</strong> numerous <strong>studies</strong>, as detailed in several reviews [1–4]. Isotopically<br />

purified germanium (Ge) and diamond (C) were the first such semiconductors to become available in<br />

the form <strong>of</strong> bulk single crystals with reasonable purity, allowing for the investigation <strong>of</strong> isotope effects<br />

on thermal conductivity, phonon energies and broadenings, and direct and indirect electronic band<br />

gaps [1, 5–16]. Previous <strong>studies</strong> on isotopic effects in silicon (Si) have been limited to Raman scattering<br />

[17, 18], thermal conductivity [19, 20], and the direct gap due to a lack <strong>of</strong> bulk samples <strong>of</strong><br />

sufficient purity for a study <strong>of</strong> the indirect band gap via photoluminescence (PL) until the recent<br />

growth <strong>of</strong> the samples for the aforementioned thermal conductivity <strong>studies</strong>. In this article, we report<br />

on the differences between the PL spectra <strong>of</strong> natural Si (92:23% 28 Si þ 4:67% 29 Si þ 3:10% 30 Si) and<br />

<strong>isotopically</strong> <strong>enriched</strong> 28 Si, 29 Si, and 30 Si, obtaining new experimental data on the isotope effect on the<br />

indirect band gap <strong>of</strong> Si, together with new information on changes in phonon energies and lifetimes [21].<br />

Remarkably, the no-phonon bound exciton (BE) PL lines, which have already been shown to be extremely<br />

sharp in natural Si, are much narrower in the 28 Si sample. Thus, the limiting broadening mechanism<br />

for the no-phonon BE PL in high purity Si is found to result from the isotopic randomness. The<br />

removal <strong>of</strong> this broadening in fact reveals new fine structure in the boron (B) BE [22]. Furthermore, the<br />

small but sample-independent splittings <strong>of</strong> the neutral acceptor (A 0 ) ground state in natural Si, observed<br />

by different experimental techniques [23–39], was absent in the <strong>isotopically</strong> pure 28 Si sample. This<br />

surprising result suggests that the A 0 splittings in natural Si result from the isotopic disorder [40].<br />

* Corresponding author: e-mail: dkaraisa@sfu.ca<br />

# 2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 0370-1972/03/23501-0063 $ 17.50þ.50/0


64 D. Karaiskaj et al.: <strong>Photoluminescence</strong> <strong>studies</strong> <strong>of</strong> <strong>isotopically</strong> <strong>enriched</strong> silicon<br />

2. Experimental methods The PL spectra were collected from samples mounted in a strain-free<br />

manner and immersed in liquid helium at a bath temperature <strong>of</strong> either 4.2 or 1.3 K, with excitation<br />

provided either by the 514.5 nm line <strong>of</strong> the Ar þ laser or, in order to ensure bulk excitation, 1020 nm<br />

from a Ti : sapphire laser. The natural Si sample used for comparison was a float-zone crystal selected<br />

to have approximately the same B and phosphorus (P) concentration as the 28 Si sample, which was<br />

cut from the SI283 crystal. This crystal was used previously for thermal conductivity <strong>studies</strong> and Hall<br />

measurements [19, 20, 41]. Spectra were collected with the Bomem DA8 Fourier transform interferometer<br />

using a cooled Ge detector (North Coast) as described in previous <strong>studies</strong> <strong>of</strong> ultrahigh resolution<br />

PL spectroscopy on natural Si [36–39, 42, 43]. The maximum instrumental resolution achieved in<br />

this study is estimated to be 0.014 cm 1 from the observed FWHM <strong>of</strong> the line-narrowed Ti-sapphire<br />

laser line, using an intracavity etalon, and collected under the same conditions as the PL spectra.<br />

The samples used to collect the aluminum (Al), gallium (Ga), and indium (In) A 0 X spectra were<br />

cut from the same crystal (SI283) as used in the study <strong>of</strong> the B and P A 0 X luminescence, but Al, Ga,<br />

and In, were introduced via variable-energy ion implantation, providing a roughly uniform concentration<br />

<strong>of</strong> the aforementioned impurities <strong>of</strong> 2 10 15 cm 3 to a depth <strong>of</strong> 3 mm followed by an anneal at<br />

1000 C for 20 min in order to activate the implant.<br />

3. Results and discussion The most remarkable features <strong>of</strong> the PL data presented in this work are<br />

related to the impurity BE no-phonon PL transitions in the <strong>isotopically</strong> <strong>enriched</strong> silicon [21, 22].<br />

While phonon assisted transitions are broadened through the wave-vector conserving phonons, these<br />

PL transitions are very sharp in natural Si and can only be resolved by Fourier transform spectroscopy<br />

[44]. Our investigations in <strong>isotopically</strong> purified 28 Si (SI283, <strong>enriched</strong> to 99.896% 28 Si) revealed that<br />

these transitions become even sharper in this material. We start our discussion with the P donor and B<br />

acceptor BE transitions in <strong>isotopically</strong> <strong>enriched</strong> 28 Si. These impurities were unintentionally incorporated<br />

during the growth process, but their concentrations were sufficient to collect the no-phonon<br />

transitions.<br />

The PL spectra <strong>of</strong> natural Si, <strong>isotopically</strong> <strong>enriched</strong> 28 Si, 29 Si, and 30 Si in the TO replica region <strong>of</strong><br />

the BE and bound multi-exciton complexes (BMEC) are shown in Fig. 1. The spectra <strong>of</strong> the natural<br />

<strong>Photoluminescence</strong> Intensity<br />

T= 4.2 K<br />

(a)<br />

(b)<br />

(c)<br />

(d)<br />

α 1 TO<br />

B 1 TO<br />

α 1 TO<br />

B 1 TO<br />

(a) 28 Si<br />

(b) Natural Si<br />

(c) 29 Si<br />

(d) 30 Si<br />

B 1 TO<br />

8800 8850<br />

Photon Energy (cm -1 )<br />

Fig. 1 <strong>Photoluminescence</strong> spectra <strong>of</strong> natural Si, <strong>isotopically</strong> <strong>enriched</strong> 28 Si, 29 Si, and 30 Si in the TO<br />

phonon replica region are superimposed. The natural Si, 28 Si, and 30 Si spectra are dominated by the B BE<br />

(B 1 TO ) and BMEC lines, while the 29 Si spectrum shows stronger P BE (a 1 TO ) line. The energy shift between<br />

the spectra is due to the indirect electronic band gap and wave-vector conserving phonon shifts [40].


phys. stat. sol. (b) 235, No. 1 (2003) 65<br />

Si, 28 Si and 30 Si samples are dominated by the B impurity showing strong B BE (B 1 TO ) and BMEC<br />

transition lines with a weaker P BE (a 1 TO ) line [52]. The natural Si sample was intentionally lightly<br />

doped with 1 10 15 cm 3 B impurities and, using the method introduced by Tajima [45], we estimate<br />

the impurity concentration in the 28 Si sample to be [B] ¼ 7 10 14 cm 3 and [P] ¼ 7 10 13 cm 3 ,in<br />

good agreement with the previous Hall measurements [19, 41], and in the 30 Si sample, the B concentration<br />

is approximately [B] ¼ 2 10 15 cm 3 . The uncertainty in estimating the impurity concentration<br />

using the method <strong>of</strong> Tajima is larger for the 30 Si sample due to the much weaker TO phonon<br />

assisted transition <strong>of</strong> the free exciton. The 29 Si crystal is dominated by the P impurity showing a<br />

strong P BE (a 1 TO ) (Fig. 1c) transition and the P concentration is estimated to be approximately<br />

[P] ¼ 1:5 10 15 cm 3 .<br />

The BE and BMEC TO replicas are energetically shifted due to the different isotopic composition<br />

<strong>of</strong> the samples. This energy shift is composed <strong>of</strong> two contributions: the shift <strong>of</strong> the electronic band<br />

gap and the shift <strong>of</strong> the TO phonon energy. The band gap shift can be measured directly from the nophonon<br />

spectra, which are shown in Fig. 2 at a spectral resolution <strong>of</strong> 0.014 cm 1 , with the 28 Si spectrum<br />

shifted up in energy by 0.92 cm 1 so as to align the features in both spectra. The most remarkable<br />

differences between the PL spectra <strong>of</strong> the 28 Si and natural Si samples are the much sharper nophonon<br />

BE lines seen in the <strong>isotopically</strong> pure sample and the resulting extra fine structure revealed in<br />

the B BE transition. It must be remembered that the no-phonon BE and BMEC transitions are already<br />

very narrow in natural Si, and the detailed study <strong>of</strong> their linewidths and fine structure only became<br />

practical with the advent <strong>of</strong> ultra high resolution Fourier transform PL spectroscopy using Michelson<br />

interferometers [44]. We discuss first the P BE (a 1 ) transition, which has no electronic substructure,<br />

and in the natural Si sample, was found to have a FWHM <strong>of</strong> 0.041 cm 1 , in agreement with previous<br />

measurements [42, 43]. The observed FWHM in the 28 Si sample is only 0.014 cm 1 , much less than<br />

<strong>Photoluminescence</strong> Intensity<br />

α 1 NP<br />

natural Si<br />

28<br />

Si<br />

B 1 NP<br />

T=1.3 K<br />

9275.0 9280 9281<br />

Photon Energy for natural Si (cm -1 )<br />

Fig. 2 A high resolution spectrum in the no-phonon region is shown for the natural Si sample, together<br />

with the same spectrum obtained from the 28 Si sample shifted up in energy by 0.92 cm 1 (the length <strong>of</strong><br />

the horizontal arrow) so as to align the transitions. The P BE (a 1 ) line is seen to consist <strong>of</strong> a single<br />

component, while the B BE (B 1 ) line has considerable fine structure. The vertical tick marks below the B<br />

BE show the predicted locations <strong>of</strong> the fine structure components obtained from a theoretical fit to the<br />

behavior <strong>of</strong> the spectrum under uniaxial stress [42]. An identical instrumental resolution <strong>of</strong> 0:014 cm 1<br />

FWHM was used for both spectra. The oscillations on either side <strong>of</strong> the P BE line are not noise, but rather<br />

the sincðxÞ lineshape resulting from transforming an interferogram using boxcar apodization when there is<br />

still strong modulation at the maximum optical path difference <strong>of</strong> 50 cm [21].


66 D. Karaiskaj et al.: <strong>Photoluminescence</strong> <strong>studies</strong> <strong>of</strong> <strong>isotopically</strong> <strong>enriched</strong> silicon<br />

that observed in the best natural Si samples, even though the 28 Si sample is <strong>of</strong> only moderate chemical<br />

purity. It must also be stressed that the observed FWHM <strong>of</strong> 0.014 cm 1 is dominated by the<br />

instrumental resolution, which at our maximum optical path length difference <strong>of</strong> 50 cm, and using<br />

boxcar apodization and ignoring resolution degradation from the finite size <strong>of</strong> the collimating aperture,<br />

would give a FWHM <strong>of</strong> 0.012 cm 1 . At present, we can thus only set an upper limit <strong>of</strong><br />

0.005 cm 1 on the true linewidth <strong>of</strong> the P bound exciton no-phonon transition in 28 Si. The broadening<br />

seen in Si is consistent with a simple calculation combining the isotopic dependence <strong>of</strong> the band gap<br />

energy determined here with statistical fluctuations <strong>of</strong> the isotopic composition in an ‘‘effective” excitonic<br />

volume having a radius <strong>of</strong> approximately 3.5 nm [53].<br />

A similar reduction in the no-phonon linewidths is seen in Fig. 2 for the components <strong>of</strong> the B BE<br />

line, which had been shown to have 9 resolved components in natural Si, explained in terms <strong>of</strong> a<br />

threefold splitting <strong>of</strong> the electron levels <strong>of</strong> the BE due to the valley-orbit effect, coupled with a threefold<br />

splitting <strong>of</strong> the hole states due to hole–hole coupling. More detailed <strong>studies</strong> <strong>of</strong> the B BE in<br />

natural Si followed, involving both uniaxial stress [42] and Zeeman [43] perturbations revealing that a<br />

complete description required the inclusion <strong>of</strong> additional terms, such as the electron–hole coupling.<br />

The linewidths <strong>of</strong> the components <strong>of</strong> the B BE are again much narrower in the 28 Si sample than in<br />

natural Si, with the narrowest components having an observed FWHM <strong>of</strong> 0.017 cm 1 , slightly larger<br />

than that <strong>of</strong> the P BE. As a result, the B BE spectrum in the 28 Si sample now reveals 13 resolved<br />

components and one partially resolved doublet [21]. This new fine structure is well accounted for by<br />

the earlier investigations <strong>of</strong> the effects <strong>of</strong> uniaxial stress and magnetic fields on the B BE [42, 43].<br />

The vertical tick marks along the bottom <strong>of</strong> the B BE spectrum, which are seen to be in quite good<br />

agreement with the newly observed fine structure, are the predicted zero-stress energies <strong>of</strong> the various<br />

components obtained from the theoretical fit to the observed energies <strong>of</strong> the many stress-split components<br />

as a function <strong>of</strong> the applied uniaxial stress [42].<br />

The evidence that the residual broadening <strong>of</strong> the no-phonon BE transition lines in natural Si is<br />

mostly due to local band gap fluctuations resulting from isotopic randomness also raised further ques-<br />

<strong>Photoluminescence</strong> Intensity<br />

T=1.8 K<br />

Si (Al)<br />

natural Si<br />

28<br />

Si<br />

Γ 5<br />

Γ 1<br />

A O X<br />

Γ<br />

Γ 5<br />

3<br />

Γ 3<br />

A O<br />

Γ 1<br />

9271 9272<br />

Photon Energy for natural Si (cm -1 )<br />

Fig. 3 A high resolution PL spectrum <strong>of</strong> the no-phonon region <strong>of</strong> the aluminum BE in natural Si is compared<br />

with the aluminum BE spectrum in a 28 Si sample, shifted up in energy (for ease <strong>of</strong> comparison) by the<br />

0.92 cm 1 difference in band gap energy between natural Si and 28 Si (the actual energies <strong>of</strong> the 28 Si transitions<br />

can be obtained by shifting the spectrum down by the 0.92 cm 1 length <strong>of</strong> the arrow). The inset shows<br />

a level scheme explaining the transitions in the natural Si sample. The A 0 X initial state has three populated<br />

levels corresponding to the three valley-orbit electron states having G 1 , G 3 and G 5 symmetry, and each <strong>of</strong><br />

these initial states produces a doublet in the PL spectrum due to the splitting <strong>of</strong> the A 0 final state, which is<br />

clearly absent in the 28 Si sample [22].


phys. stat. sol. (b) 235, No. 1 (2003) 67<br />

tions that needed to be pursued. One <strong>of</strong> the unexpected discoveries resulting from the previous applications<br />

<strong>of</strong> ultrahigh resolution spectroscopic techniques to the PL spectra <strong>of</strong> natural Si was the observation<br />

<strong>of</strong> a small, but reproducible splitting <strong>of</strong> the ground state <strong>of</strong> the Al, Ga, and In acceptors in<br />

natural Si [36, 37]. It was a natural extension <strong>of</strong> the work on the P and B no-phonon BE transitions to<br />

investigate the acceptor ground state splitting in <strong>isotopically</strong> <strong>enriched</strong> Si, since the broadening <strong>of</strong> the<br />

lines should be reduced substantially.<br />

The acceptor states and acceptor BE transitions have been studied by a variety <strong>of</strong> different experimental<br />

techniques such as electron spin resonance (ESR) [23, 24], thermal conductivity <strong>studies</strong> [32–34],<br />

phonon absorption spectroscopy [25–31, 46] and, most recently, by photoluminescence [35–39].<br />

While electron paramagnetic resonance (EPR) was detected for electrons bound to neutral donor impurities<br />

(D 0 ) in Si very early in the development <strong>of</strong> the semiconductor physics [47], careful searches<br />

for EPR from holes bound to neutral acceptors (A 0 ) in Si were initially unsuccessful, as summarized<br />

by Kohn [48], who was also the first to point out that the problem likely resulted from the four-fold<br />

degeneracy <strong>of</strong> the A 0 ground state. Unlike the D 0 ground state, whose two-fold spin degeneracy can<br />

only be lifted by a magnetic field, the A 0 ground state in the absence <strong>of</strong> perturbations is a four-fold<br />

degenerate state, which can be split into doublets by perturbations such as axial strain or an electric<br />

field. Thus, a random distribution <strong>of</strong> perturbations, such as strain fields from dislocations and/or impurities,<br />

acting on the A 0 ground state produces a distribution <strong>of</strong> doublet splittings which in turn can<br />

broaden the EPR transitions to the point where they are unobservable. This conjecture was soon verified<br />

by the observation <strong>of</strong> EPR from A 0 in silicon in the presence <strong>of</strong> externally applied uniaxial<br />

strain, which produces a large but uniform doublet splitting <strong>of</strong> the ground state, and thus diminishes<br />

the effects <strong>of</strong> the internal random strain fields [49]. While there were several predictions that the<br />

rapidly improving perfection <strong>of</strong> single crystal silicon would soon permit the observation <strong>of</strong> EPR without<br />

necessity <strong>of</strong> external strain, the first success in this regard did not occur until considerably later<br />

[50, 51].<br />

However, Neubrand’s pioneering observation <strong>of</strong> ESR from A 0 in B doped silicon in the absence <strong>of</strong><br />

external strain, with doping concentrations ranging from 10 15 to 10 16 cm 3 , introduced a new problem<br />

which has remained unresolved until this work [23, 24]. After accounting for all <strong>of</strong> the possible random<br />

strain splittings from the known contaminants in the purest, dislocation-free samples, there remained<br />

a sample-independent distribution <strong>of</strong> A 0 splittings whose origin could not be explained.<br />

A similar small (0:10 0:01 cm 1 ) but sample-independent splitting <strong>of</strong> the Al A 0 ground state in<br />

silicon was later observed as a final-state splitting <strong>of</strong> the components <strong>of</strong> the Al acceptor bound exciton<br />

(A 0 X) photoluminescence transitions, and it was suggested that this splitting arose from a spontaneous<br />

lowering <strong>of</strong> the A 0 symmetry via the Jahn–Teller effect [35, 36]. A subsequent PL study confirmed<br />

that the doublet splittings arose from the A 0 final state <strong>of</strong> the A 0 X PL transitions and revealed a<br />

similar 0:10 0:01 cm 1 doublet splitting for the Ga A 0 ground state, and a 0:15 0:03 cm 1 A 0<br />

splitting for the deeper In acceptor [37–39]. The 0:15 0:03 cm 1 splitting found for the indium A 0<br />

state in the PL study [37] agrees well with the phonon spectroscopy [31] result <strong>of</strong> 0:11 to 0:16 cm 1 ,<br />

and the failure to detect the B splitting (0:043 cm 1 as determined by phonon spectroscopy [28, 29,<br />

31]) in the PL <strong>studies</strong> was consistent with the PL observed linewidth <strong>of</strong> the boron BE components <strong>of</strong><br />

0:08 cm 1 resulting from a combination <strong>of</strong> insufficient instrumental resolution and PL broadening<br />

effects.<br />

We return now to a more detailed explanation <strong>of</strong> the PL spectra shown in Fig. 3. A review and<br />

further references to the details <strong>of</strong> PL related to shallow donors and acceptors in Si can be found in<br />

Ref. [52]. Both spectra are from what is known as the no-phonon region <strong>of</strong> the Si PL spectrum, in<br />

which the energy <strong>of</strong> the emitted photon is exactly equal to the energy difference between the electronic<br />

initial and final states <strong>of</strong> the transition. The Al BE no-phonon recombination spectrum, shown in<br />

Fig. 3 (natural Si), consists <strong>of</strong> three identical doublets with the same splitting <strong>of</strong> 0:1 cm 1 . The results<br />

shown in Fig. 3 are explained by three initial states <strong>of</strong> the A 0 X in both the natural Si and the 28 Si<br />

samples, which are labelled by their symmetries <strong>of</strong> G 1 , G 3 , and G 5 in the tetrahedral point group (T d )<br />

appropriate for substitutional impurities in Si. These three A 0 X initial states arise from a splitting <strong>of</strong><br />

the electron states in the A 0 X via the valley-orbit splitting, as also found on a much larger energy


68 D. Karaiskaj et al.: <strong>Photoluminescence</strong> <strong>studies</strong> <strong>of</strong> <strong>isotopically</strong> <strong>enriched</strong> silicon<br />

Photoluminscence Intensity<br />

T=1.8 K<br />

Si (Ga)<br />

natural Si<br />

28<br />

Si<br />

Γ 5<br />

Γ 3<br />

Γ 1<br />

9266 9268<br />

Photon Energy for natural Si (cm -1 )<br />

Fig. 4 High resolution no-phonon PL spectra <strong>of</strong> the gallium A 0 X transitions are compared for natural<br />

Si and 28 Si samples. For ease <strong>of</strong> comparison, the 28 Si spectra have been shifted up in energy by the<br />

0.92 cm 1 difference in band gap energy between 28 Si and natural Si. The inset shows the doublet<br />

resulting from a singlet A 0 X initial state due to the A 0 final state splitting which is used to fit the<br />

observed spectra in the natural Si samples [22].<br />

scale for the D 0 ground state in indirect band-gap semiconductors such as Si [38, 39]. It should also<br />

be noted that there is a much larger splitting <strong>of</strong> the A 0 X initial state due the hole–hole coupling <strong>of</strong> its<br />

two holes. At the temperatures used in this study, only the lowest in energy G 1 spin singlet state <strong>of</strong><br />

A 0 X is populated resulting from the reduction <strong>of</strong> the (G 8 G 8 ¼ G 1 þ G 3 þ G 5 ) two-hole state, and all<br />

transitions in the Al, Ga, and In A 0 X spectra originate from an initial state <strong>of</strong> G 1 symmetry. Thus,<br />

only a triplet due to the valley-orbit splitting <strong>of</strong> the electron states is expected. The obvious explanation<br />

is that the extra, identical doublet splittings <strong>of</strong> the three valley-orbit components are due to a final<br />

state splitting <strong>of</strong> the A 0 X (or a ground state splitting <strong>of</strong> the A 0 ) confirming the already strong evidence<br />

previously given by ESR, phonon spectroscopy and thermal conductivity <strong>studies</strong>. A diagram <strong>of</strong><br />

these transitions and the effect <strong>of</strong> the additional splitting <strong>of</strong> the A 0 ground state is shown in Fig. 3<br />

(inset).<br />

Further investigations on the Ga BE no-phonon recombination spectrum shown in Fig. 4 (natural<br />

Si), indicated that this spectrum also consists <strong>of</strong> three doublets. The first doublet originating from the<br />

G 3 electron state is clearly visible. The second component is thought to result from the overlap <strong>of</strong> two<br />

doublets as the one shown in the inset <strong>of</strong> Fig. 5. This was also confirmed by PL <strong>studies</strong> in magnetic<br />

fields [39] and under uniaxial stress [38]. The second component most likely consists <strong>of</strong> two doublets<br />

originating from the G 1 and G 5 electron states, which are too close in energy to be fully resolved in<br />

the natural Si spectrum. Furthermore, the In bound exciton no-phonon spectrum is shown in Fig. 5<br />

and consists <strong>of</strong> four components also resulting from the G 1 , G 3 , and G 5 transitions. The fourth component<br />

is thought to originate from the G 5 component that is split into a doublet by the spin-orbit interaction.<br />

This has also been confirmed by further investigations in magnetic fields and under uniaxial<br />

stress [38, 39]. The line shape <strong>of</strong> the components indicates the presence <strong>of</strong> underlying structure, which<br />

is not resolved. These extra unresolved components do appear in the no-phonon absorption spectra<br />

when the temperature is lowered from 4.2 to 0.4 K (see Ref. [37] for further details).<br />

The Jahn–Teller theorem [54] states that in any nonlinear configuration <strong>of</strong> atoms, coupling between<br />

a degenerate electronic state and asymmetric vibrations <strong>of</strong> nuclei, must distort this configuration and<br />

lift the electronic degeneracy by reducing its symmetry. The fact that the maximum <strong>of</strong> the valence<br />

band in Si belongs to the G 8 symmetry and is a fourfold degenerate state that could be unstable<br />

towards vibrations, makes the Jahn–Teller effect the most suitable explanation for the acceptor ground


phys. stat. sol. (b) 235, No. 1 (2003) 69<br />

T=1.8 K<br />

Si (In)<br />

<strong>Photoluminescence</strong> Intensity<br />

natural Si<br />

28<br />

Si<br />

Γ 1<br />

Γ 3<br />

Γ 5<br />

9201 9203<br />

Photon Energy for natural Si (cm -1 )<br />

Fig. 5 High resolution no-phonon PL spectra <strong>of</strong> the indium A 0 X transitions are compared for natural<br />

Si and 28 Si samples. For ease <strong>of</strong> comparison, the 28 Si spectra have been shifted up in energy by the<br />

0.92 cm 1 difference in band gap energy between 28 Si and natural Si. The inset shows the doublet<br />

resulting from a singlet A 0 X initial state due to the A 0 final state splitting which is used to fit the<br />

observed spectra in the natural Si samples [22].<br />

state splitting. Vibrations and distortions <strong>of</strong> the crystal cannot break the time reversal symmetry, and<br />

as a result, can cause only a splitting into Kramers doublet.<br />

At this point we are left with two possibilities for explaining the small A 0 splittings: either a Jahn–<br />

Teller effect as postulated by the optical <strong>studies</strong>, which would be in conflict with the much higher<br />

energy Jahn–Teller resonance invoked in the phonon <strong>studies</strong> [30, 33], or, as postulated in the phonon<br />

<strong>studies</strong>, a distribution <strong>of</strong> splittings from random fields <strong>of</strong> unknown origin, but essentially identical in<br />

all high-quality samples. Clearly, the first explanation had some difficulties, at least in Si, where the<br />

phonon measurements applied, and the second was incomplete in that the origin <strong>of</strong> the sample-independent<br />

random fields was left unexplained.<br />

The Al A 0 X spectrum from the <strong>isotopically</strong> pure 28 Si sample shown in Fig. 3 ( 28 Si) is shifted up in<br />

energy by the 0:92 cm 1 reduction in band gap energy <strong>of</strong> 28 Si relative to natural Si. It is clear that not<br />

only has the isotopic broadening been reduced in the 28 Si sample, but the A 0 splitting has also vanished.<br />

The doublets resulting from the splitting <strong>of</strong> the A 0 ground state, as shown schematically in the<br />

inset, are now single sharp transitions. The same change is also seen in the Ga A 0 X spectrum collected<br />

from the 28 Si sample. The G 3 line, which is clearly a doublet in natural Si, is a single line in<br />

28 Si. The more complicated component, thought to be composed <strong>of</strong> two doublets, is now reduced to<br />

two transitions most likely originating from the G 5 and G 1 electron states. The doublet final state<br />

splittings are less resolved in the natural Si(In) spectrum, although they have been confirmed by careful<br />

thermalization <strong>studies</strong> in both PL and absorption spectroscopy, and as mentioned earlier, the In<br />

A 0 X spectrum is further complicated by a splitting <strong>of</strong> the A 0 X initial state <strong>of</strong> G 5 symmetry by spinorbit<br />

coupling [38, 39]. Nevertheless, it is again clear that the doublet splitting resulting from the A 0 X<br />

final state, as modelled in the inset, is also absent for the In A 0 Xin 28 Si.<br />

In conclusion, the linewidths <strong>of</strong> the no-phonon BE transitions in 28 Si were found to be much narrower<br />

than those <strong>of</strong> natural Si, revealing new fine structure in the spectrum <strong>of</strong> the B bound exciton<br />

predicted in earlier <strong>studies</strong>. The residual, sample-independent A 0 ground state splittings in Si are<br />

absent in <strong>isotopically</strong> pure samples, and as a result, these splittings must also arise from the effects <strong>of</strong><br />

isotopic randomness in all <strong>of</strong> the previously studied natural Si samples. This makes good qualitative<br />

sense – any perturbation which lowers the crystal symmetry from T d at the impurity site can split the


70 D. Karaiskaj et al.: <strong>Photoluminescence</strong> <strong>studies</strong> <strong>of</strong> <strong>isotopically</strong> <strong>enriched</strong> silicon<br />

A 0 ground state into a doublet, and the random fluctuations <strong>of</strong> isotopic composition around individual<br />

acceptors will provide such a symmetry lowering perturbation, which has the added necessary property<br />

<strong>of</strong> having an identical distribution for all samples having the same (natural) isotopic abundances.<br />

One might therefore ask why such an obvious explanation has never been postulated in the extensive<br />

literature on these splittings. We believe that the reason is that the effect <strong>of</strong> isotopic randomness was<br />

assumed to be too small to explain the observed splitting – this was certainly our assumption before<br />

obtaining the present results. Previous calculations <strong>of</strong> the A 0 splittings due to the known perturbations<br />

<strong>of</strong> impurities such as oxygen and carbon were based on the long-range strains produced by these<br />

relatively dilute and highly perturbing impurities, where long-range refers to strains which were essentially<br />

uniform across the wave function <strong>of</strong> a given A 0 state. From the point <strong>of</strong> view <strong>of</strong> such long-range<br />

fields, the perturbations arising from isotopic randomness in natural silicon would indeed be much too<br />

small to explain the observed A 0 splittings. Instead, we believe a new approach is necessary, namely,<br />

a detailed analysis <strong>of</strong> how isotopic fluctuations on a scale smaller than the A 0 Bohr radius couple to<br />

the different A 0 states, and we are currently investigating such effects using realistic A 0 wave functions.<br />

Isotopic effects on the indirect electronic band gap and phonon energies The electronic states <strong>of</strong><br />

crystals, such as valence and conduction band states, are renormalized by their interaction with phonons.<br />

Crystal vibrations also exist at zero Kelvin (zero-point vibrations) and interact with electrons<br />

renormalizing their energy. Changes in the average isotopic mass lead to changes in the lattice vibrations,<br />

and as a consequence <strong>of</strong> the electron–phonon interaction, to an energy shift <strong>of</strong> the electronic<br />

states. To a lesser extent this energy shift also results from the change in the lattice constant with the<br />

isotopic composition. The zero-point renormalization energy <strong>of</strong> a crystal is proportional to the inverse<br />

square root <strong>of</strong> the average isotopic mass, and by using this relation and measuring the energy shift <strong>of</strong><br />

the electronic band gap between two crystals <strong>of</strong> different isotopic composition, the zero-point renormalization<br />

energy can be estimated.<br />

From the relative energy shift <strong>of</strong> the impurity bound exciton no-phonon transitions, between two Si<br />

crystals <strong>of</strong> different isotopic composition, the relative shift <strong>of</strong> the indirect electronic band gap can be<br />

determined. Due to the much simpler electronic structure <strong>of</strong> the donor P BE, these transitions were<br />

used to obtain the band gap shifts. From this quantity, the zero-point renormalization energies <strong>of</strong> the<br />

indirect electronic band gap in Si were estimated using the method <strong>of</strong> Ref. [3]. The samples used to<br />

determine the relative band gap shift are the single crystals 28 Si, 29 Si, 30 Si and the solution grown<br />

isotopic mixture <strong>of</strong> 28 Si and 30 Si. Their isotopic compositions and estimated average masses are given<br />

in Table 1.<br />

The four renormalization energies obtained are listed in Table 2 and are similar to the value obtained<br />

for germanium ( 51:9 2:1 meV) using the same experimental method [6]. The energy shift<br />

<strong>of</strong> the direct electronic band gap with isotopic composition has been measured recently for Si using<br />

ellipsometry, leading to renormalization energies <strong>of</strong> 90 meV [55].<br />

The simplest approximation, the harmonic virtual crystal approximation (VCA), predicts that the<br />

phonon energy decreases proportionally to the inverse square root <strong>of</strong> the average isotopic mass. In<br />

reality, additional interactions such as anharmonic and disorder effects influence the phonon energies<br />

and lifetimes. The anharmonic effects to the frequency shift result from the phonon–phonon interac-<br />

Table 1<br />

The isotopic composition and average isotopic mass <strong>of</strong> the Si samples investigated.<br />

sample isotopic composition average mass (amu)<br />

natural Si 92.23% ( 28 Si) þ 4.67% ( 29 Si) þ 3.10% ( 30 Si) 28.086<br />

28 Si 99.896% ( 28 Si) þ 0.09% ( 29 Si) þ 0.014% ( 30 Si) 27.978<br />

29 Si 2.15% ( 28 Si) þ 97.58% ( 29 Si) þ 0.27% ( 30 Si) 28.958<br />

30 Si 0.70% ( 28 Si) þ 0.62% ( 29 Si) þ 98.68% ( 30 Si) 29.954<br />

28 Si þ 30 Si 26% ( 28 Si) þ 74% ( 30 Si) 29.45


phys. stat. sol. (b) 235, No. 1 (2003) 71<br />

Table 2 The indirect electronic band gap shifts relative to the <strong>isotopically</strong> <strong>enriched</strong> 28 Si sample. The<br />

renormalization energies are obtained from these band gap shifts.<br />

natural Si <strong>enriched</strong> 29 Si <strong>enriched</strong> 30 Si mixture <strong>of</strong> 28Si þ 30 Si<br />

band gap 0.92 cm 1 8.4 cm 1 16.9 cm 1 12 cm 1<br />

shift (114 meV) (1.04 meV) (2.09 meV) (1.48 meV)<br />

band gap –59.2 meV –60.4 meV –61.3 meV –57.7 meV<br />

renormalization (1 meV) (2 meV) (1 meV) (2 meV)<br />

tion processes. The coherent potential approximation (CPA) takes into account the random distribution<br />

<strong>of</strong> the different isotopes, and as a result, enables us to estimate the contributions <strong>of</strong> the isotopic<br />

disorder to the energy shift and broadening <strong>of</strong> the phonons.<br />

The frequency shifts for the transverse optical (TO) and transverse acoustical (TA) wavevector-conserving<br />

phonons were obtained from the phonon replicas <strong>of</strong> the P and B BE transitions and are listed<br />

in Table 3. All phonon shifts are given relative to the 28 Si sample, because <strong>of</strong> its high isotopic purity,<br />

together with the phonon shifts theoretically predicted by the VCA. The TO(X) and TA(X) phonon<br />

shifts <strong>of</strong> natural Si relative to 28 Si are in reasonable agreement with the VCA predicted values. Small<br />

deviations are mainly due to the isotopic disorder, caused by the random distribution <strong>of</strong> the 29 Si and<br />

30 Si isotopes in natural Si, and to a lesser extent, due to anharmonic effects. The TO(X) phonon<br />

frequency shift <strong>of</strong> the 29 Si sample agrees also with the VCA. There might be some disorder-induced<br />

frequency shift due to 2.15% 28 Si still contained in the sample, but it is much smaller than our experimental<br />

certainty. The TO(X) and TA(X) phonon shifts <strong>of</strong> the high <strong>isotopically</strong> <strong>enriched</strong> 30 Si crystal<br />

are in good agreement with the value predicted by the VCA within the experimental uncertainty. This<br />

reconfirms the high isotopic purity <strong>of</strong> the crystal showing a very weak isotopic disorder induced<br />

frequency shift [40].<br />

Table 3 The experimentally obtained energy shifts <strong>of</strong> the TO(X), TA(X), and Raman phonons are<br />

summarized together with the energetic shifts predicted by the VCA for natural Si, 29 Si, 30 Si,<br />

28 Si þ 30 Si, compared to <strong>isotopically</strong> <strong>enriched</strong> 28 Si.<br />

TO(X) phonon TA(X) phonon Raman phonon<br />

phonon shift <strong>of</strong> natural Si measured ( 1.14 0.1) cm 1 ( 0.28 0.1) cm 1 ( 0.81 0.05) cm 1<br />

with respect to 28 Si<br />

corresponding shifts predicted<br />

from the VCA compared to 28 Si<br />

0.91 cm 1 0.29 cm 1 1.01 cm 1<br />

phonon shift <strong>of</strong> 29 Si measured ( 7.8 0.5) cm 1<br />

with respect to 28 Si<br />

corresponding shifts predicted 8cm 1<br />

from the VCA compared to 28 Si<br />

phonon shift <strong>of</strong> 30 Si measured ( 15.7 0.5) cm 1 ( 5.1 0.5) cm 1<br />

with respect to 28 Si<br />

corresponding shifts predicted 15.7 cm 1 5.1 cm 1<br />

from the VCA compared to 28 Si<br />

phonon shift <strong>of</strong> isotopic mixture ( 12.6 0.5) cm 1<br />

( 28 Si þ 30 Si) measured with<br />

respect to 28 Si<br />

corresponding shifts predicted 11.9 cm 1<br />

from the VCA compared to 28 Si


72 D. Karaiskaj et al.: <strong>Photoluminescence</strong> <strong>studies</strong> <strong>of</strong> <strong>isotopically</strong> <strong>enriched</strong> silicon<br />

We were able to obtain accurate values <strong>of</strong> the zone-center longitudinal optical Raman phonon energies<br />

in natural Si and 28 Si, since the Stokes–Raman line <strong>of</strong> the 1020 nm pump laser falls within the<br />

transparency region <strong>of</strong> Si, and is in the same spectral region over which the PL spectra were collected.<br />

The energy shift <strong>of</strong> the Raman phonon is also given in Table 3. The additional anharmonic<br />

contributions to this frequency shift have been estimated. These contributions to the phonon frequency<br />

shift result from phonon–phonon interaction processes and can be estimated using perturbation theory.<br />

As a result, we obtain an additional correction term, which is proportional to the inverse average<br />

vibrating mass. By including these contributions, the expected frequency shift for the Raman phonon<br />

is 0.83 cm 1 , which is very close to the measured one <strong>of</strong> 0.81 cm 1 [18, 40].<br />

A closer examination <strong>of</strong> Fig. 1 reveals that the TO wavevector-conserving phonon replicas <strong>of</strong><br />

natural Si are noticeably broader than those <strong>of</strong> 28 Si, indicating that in natural Si the TO wavevectorconserving<br />

phonon has a substantial lifetime broadening arising from isotopic randomness. Extracting<br />

this additional broadening parameter is complicated by the fact that the electronic bound exciton<br />

transitions are themselves not delta functions (see for example the structure <strong>of</strong> the B bound exciton<br />

revealed in Fig. 2), and the impurity-related wavevector-conserving phonon replicas have an additional<br />

broadening which arises from the uncertainty in the center-<strong>of</strong>-mass electron and hole wavevector<br />

due to the localization on the impurity, which causes a broadening via the energy dispersion<br />

<strong>of</strong> the particular phonon involved. While the detailed analysis <strong>of</strong> this broadening will be the subject<br />

<strong>of</strong> a future study, one can estimate for the surprisingly large difference between Si and 28 Si by<br />

noting that for the TO phonon the FWHM <strong>of</strong> the B 3 line in the former is 2.6 cm 1 and in the latter<br />

only 2.1 cm 1 [21].<br />

In addition, we investigated an <strong>isotopically</strong> disordered alloy with an isotopic mixture <strong>of</strong> 28 Si and<br />

30 Si grown by liquid-phase epitaxy from indium solution. The isotopic composition and average atomic<br />

mass are given in Table 1. The PL spectrum showed a high concentration <strong>of</strong> indium and phosphorus<br />

that could not be determined exactly, but it is estimated to be <strong>of</strong> order 1 10 16 cm 3 or greater.<br />

In such highly <strong>isotopically</strong> disordered crystals, the disorder induced phonon frequency shift and<br />

linewidth broadening become significant. Therefore, the discrepancy between the VCA predicted frequency<br />

shift and the measured value, given in Table 3, is believed to be primarily due to the isotopic<br />

disorder induced shift. In fact, recent calculations [18] using the CPA predict for the TO(X) phonon in<br />

this isotopic composition a disorder-induced frequency shift <strong>of</strong> 0:5 cm 1 , which is in good agreement<br />

with our experimental value [18, 40].<br />

In conclusion, we have determined the indirect band gap shifts <strong>of</strong> natural Si, 29 Si, 30 Si, and isotopic<br />

mixture, relative to the almost single isotope 28 Si crystal. Using these energy shifts <strong>of</strong> the indirect<br />

electronic band gap, the zero-point renormalization energies were estimated. We measured the phonon<br />

frequency shifts <strong>of</strong> natural Si, 29 Si, 30 Si and an isotopic mixture <strong>of</strong> 28 Si and 30 Si, relative to the<br />

<strong>isotopically</strong> pure 28 Si sample. These shifts were compared with the values predicted by the VCA, and<br />

the eventual discrepancies due to anharmonic and disorder effects were discussed.<br />

Acknowledgements D. Karaiskaj and M. L. W. Thewalt acknowledge Natural Sciences and Engineering Council<br />

<strong>of</strong> Canada for financial support.<br />

References<br />

[1] E. E. Haller, J. Appl. Phys. 77, 2857 (1995).<br />

[2] M. Cardona, P. Etchegoin, H. D. Fuchs, and P. Molins-Mata, J. Phys.: Condens. Matter 5, A61 (1993).<br />

[3] M. Cardona, phys. stat. sol. (b) 220, 5 (2000).<br />

[4] M. Cardona and T. Ruf, Solid State Commun. 117, 201 (2001).<br />

[5] K. C. Hass, M. A. Tamor, T. R. Anthony, and W. F. Bahnholzer, Phys. Rev. B 45, 7171 (1992).<br />

[6] P. Etchegoin, J. Weber, M. Cardona, W. L. Hansen, K. Itoh, and E. E. Haller, Solid State Commun. 83, 843 (1992).<br />

[7] P. Etchegoin, H. D. Fuchs, J. Weber, M. Cardona, L. Pintschovius, N. Pyka, K. Itoh, and E. E. Haller, Phys.<br />

Rev. B 48, 12 661 (1993).<br />

[8] G. Davies, E. C. Lightowler, T. S. Hui, V. Ozhogin, K. M. Itoh, W. L. Hansen, and E. E. Haller, Semicond.<br />

Sci. Technol. 8, 2201 (1993).


phys. stat. sol. (b) 235, No. 1 (2003) 73<br />

[9] G. Davies, E. C. Lightowler, K. Itoh, W. L. Hansen, E. E. Haller, and V. Ozhogin, Semicond. Sci. Technol.<br />

7, 1271 (1992).<br />

[10] C. Parks, A. K. Ramdas, S. Rodriguez, K. M. Itoh, and E. E. Haller, Phys. Rev. B 49, 14 244 (1994).<br />

[11] H. D. Fuchs, C. H. Grein, C. Thomsen, M. Cardona, W. L. Hansen, E. E. Haller, and K. Itoh, Phys. Rev. B<br />

43, 4835 (1991).<br />

[12] H. D. Fuchs, C. H. Grein, M. Cardona, W. L. Hansen, K. Itoh, and E. E. Haller, Solid State Commun. 82,<br />

225 (1992).<br />

[13] H. D. Fuchs, C. H. Grein, M. Bauer, and M. Cardona, Phys. Rev. B 45, 4065 (1992).<br />

[14] J. M. Zhang, M. Giehler, A. Gæbel, T. Ruf, M. Cardona, E. E. Haller, and K. Itoh, Phys. Rev. B 57, 1348<br />

(1998).<br />

[15] A. Gæbel, D. T. Wang, M. Cardona, L. Pintschovius and W. Reichardt, J. Kulda, N. M Pyka, K. Itoh, and<br />

E. E. Haller, Phys. Rev. B 58, 10 510 (1998).<br />

[16] S. Zollner, M. Cardona, and S. Gopalan, Phys. Rev. B 45, 3376 (1992).<br />

[17] F. Widulle, T. Ruf, A. Gæbel, I. Silier, E. Schænherr, M. Cardona, J. Camacho, A. Cantarero, W. Kriegseis,<br />

and V. I. Ozhogin, Physica (b) 263/264, 381 (1999).<br />

[18] F. Widulle, T. Ruf, M. Konuma, I. Silier, M. Cardona, W. Kriegseis, and V. I. Ozhogin, Solid State Commun.<br />

118, 1 (2001).<br />

[19] T. Ruf, R. W. Henn, M. Asen-Palmer, E. Gmelin, M. Cardona, H.-J. Pohl, G. G. Devyatych, and P. G. Sennikov,<br />

Solid State Commun. 115, 243 (2000).<br />

[20] T. Ruf, R. W. Henn, M. Asen-Palmer, E. Gmelin, M. Cardona, H.-J. Pohl, G. G. Devyatych, P. G. Sennikov,<br />

and J. Bollmann, Proc. 25th Int. Conf. Phys. Semicond., Springer-Verlag, Berlin, to be published.<br />

[21] D. Karaiskaj, M. L. W. Thewalt, T. Ruf, M. Cardona, H.-J. Pohl, G. G. Deviatych, P. G. Sennikov, and<br />

H. Riemann, Phys. Rev. Lett. 86, 6010 (2001).<br />

[22] D. Karaiskaj, M. L. W. Thewalt, T. Ruf, M. Cardona, and M. Konuma, Phys. Rev. Lett. 89, 16401 (2002).<br />

[23] H. Neubrandt, phys. stat. sol. (b) 86, 269 (1978).<br />

[24] H. Neubrandt, phys. stat. sol. (b) 90, 301 (1978).<br />

[25] K. Suzuki and N. Mikoshiba, Phys. Rev. Lett. 28, 94 (1972).<br />

[26] T. Ishiguro, Phys. Rev. B 8, 629 (1973).<br />

[27] K. Lassmann and H. P. Schad, Solid State Commun. 18, 449 (1976).<br />

[28] H. Zeile, U. Harten, and K. Lassmann, phys. stat. sol. (b) 111, 213 (1982).<br />

[29] H. Zeile and K. Lassmann, phys. stat. sol. (b) 111, 555 (1982).<br />

[30] J. Maier and E. Sigmund, in: Phonon Scattering in Condensed Matter, Springer Series in Solid State<br />

Science, Vol. 51, edited by W. Eisenmenger (Springer-Verlag, Berlin 1984), p. 364.<br />

[31] A. Ambbrosy, K. Lassmann, A. M. de GoŸr, B. Salce and H. Zeile, in: Phonon Scattering in Condensed<br />

Matter, Springer Series in Solid State Science, Vol. 51, edited by W. Eisenmenger (Springer-Verlag, Berlin<br />

1984), p. 361.<br />

[32] L. J. Challis, A. M. de GoŸr, and S. C. Haseler, Phys. Rev. Lett. 39, 558 (1977).<br />

[33] A. M. de GoŸr, M. Locatelli, and K. Lassmann, J. Phys. C 42, 235 (1981).<br />

[34] L. J. Challis and A. P. Heraud, in: Phonon Scattering in Condensed Matter, Springer Series in Solid State<br />

Science, Vol. 51, edited by W. Eisenmenger (Springer-Verlag, Berlin 1984), p. 368.<br />

[35] A. S. Kaminskii and A. N. Safonov, Sov. Phys. JETP Lett. 55, 242 (1992).<br />

[36] V. A. Karasyuk, M. L. W. Thewalt, S. An, and E. C. Lightowlers, Phys. Rev. Lett. 73, 2340 (1994).<br />

[37] V. A. Karasyuk, S. An, M. L. W. Thewalt, A. C. Lightowlers, and A. S. Kaminskii, Solid State Commun.<br />

93, 379 (1995).<br />

[38] V. A. Karasyuk, M. L. W. Thewalt, S. An, and E. C. Lightowlers, Phys. Rev. B 56, 15 672 (1997).<br />

[39] V. A. Karasyuk, M. L. W. Thewalt, S. An, E. C. Lightowlers, and A. S. Kaminskii, Phys. Rev. B 54, 10 543<br />

(1996).<br />

[40] D. Karaiskaj, M. L. W. Thewalt, T. Ruf, M. Cardona, and M. Konuma, Solid State Commun., to be published.<br />

[41] A. D. Bulanov, G. G. Devyatych, A. V. Gusev, P. G. Sennikov, H.-J. Pohl, H. Riemann, H. Schilling, and<br />

P. Becker, Cryst. Res. Technol. 35, 1023 (2000).<br />

[42] V. A. Karasyuk, A. G. Steele, A. Mainwood, E. C. Lightowler, G. Davies, D. Brake, and M. L. W. Thewalt,<br />

Phys. Rev. B 45, 11 736 (1992).<br />

[43] V. A. Karasyuk, D. M. Brake, and M. L. W. Thewalt, Phys. Rev. B 47, 9354 (1993).<br />

[44] M. L. W. Thewalt, and D. M. Brake, Mater. Sci. Forum 65/66, 187 (1990).<br />

[45] M. Tajima, Appl. Phys. Lett. 32, 719 (1978).<br />

[46] E. Ortlieb, H. P. Schad, and K. Lassmann, Solid State Commun. 19, 599 (1976).


74 D. Karaiskaj et al.: <strong>Photoluminescence</strong> <strong>studies</strong> <strong>of</strong> <strong>isotopically</strong> <strong>enriched</strong> silicon<br />

[47] R. C. Fletcher, W. A. Yager, G. L. Pearson, A. N. Holden, W. T. Read, and F. R. Merritt, Phys. Rev. 94,<br />

1392 (1954).<br />

[48] W. Kohn, in: Solid State Physics, Vol. 5, edited by F. Seitz and D. Turnbull (Academic Press, New York<br />

1957), p. 257.<br />

[49] G. Feher, J. C. Hensel, and E. A. Gere, Phys. Rev. Lett. 5, 309 (1960).<br />

[50] E. O. Kane, Phys. Rev. 119, 40 (1960).<br />

[51] Y. Yafet, J. Phys. Chem. Solids 26, 647 (1965).<br />

[52] M. L. W. Thewalt, in: Excitons, edited by E. I. Rashba and M. D. Sturge (North-Holland Press, Amsterdam<br />

1982), p. 393–458.<br />

[53] V. G. Plekhanov, Opt. Spectrosc. (USSR) 79, 715 (1995).<br />

[54] H. A. Jahn and E. Teller, Proc. Roy. Soc. 1, 220 (1937).<br />

[55] L. F. Lastras-MartÌnez, T. Ruf, M. Konuma, M. Cardona, and D. E. Aspnes, Phys. Rev. B 61, 12 946 (2000).

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!