Você está na página 1de 52

16

Electrically Detected Magnetic Resonance (EDMR) of defects in GaN Light Emitting Diodes
Martin W. Bayerl* , Martin S. Brandt, Martin Stutzmann Compared to standard Electron Spin Resonance (ESR), EDMR has proven to be a more sensitive method in detecting paramagnetic states in semiconductors. If certain transport processes like recombination or hopping are spin-dependent, they can be detected by EDMR via resonant changes in the conductivity of the sample under ESR conditions. The application of this technique to electronic devices is particularly interesting because performance limitations in electrical transport can be correlated with recombination processes due to defects. In addition, the microscopic structure of such defects can be determined from the resonance spectra observed. EDMR measurements in group-IV semiconductors, in particular silicon, are well established. We have recently shown that EDMR can also be applied to III-V compound semiconductors such as GaAs and GaN. Among these materials, GaN currently receives considerable attention for use in light emitting diodes. Despite the high brightness already achieved, little is known about defects in such devices and their influence on device properties and stability. In contrast to the older EDMR studies on double-heterostructure and quantum-well GaN diodes, we investigate here commercial light emitting diodes, fabricated by CREE, with a much simpler structure, which basically consists of a GaN p-n junction, grown on a N-doped SiC substrate. This allows an easier assignment of the resonances observed. The EDMR investigations were performed at two different frequencies using a standard X-band ESR spectrometer and a Q-band spectrometer with microwave frequencies of 9.4 and 34 GHz respectively. We will show that the measurements at the higher microwave frequency provide an enhanced resolution for the g-tensor, leading to a better separation of different donor resonances involved in recombination processes. Figure 1 shows the spectra obtained at liquid helium temperatures. Both are dominated by a resonance at g=2.003 with a peak-to-peak linewidth of HPP 160 G. The typical resonant change / was 10-7. At 9.4 GHz, a substructure at high magnetic fields is visible, which can be attributed to an additional resonance at g=1.97. This high field structure is resolved more clearly in the 34 GHz measurement, which shows two resonances at g=1.983 and g=1.973. At the low temperatures used for the EDMR measurements, hopping in the SiC:N donor band of the substrate will also be a rate-limiting transport process. Indeed, the 34 GHz spectrum shows an additional narrow line which, on a different magnetic field scale, exhibits the characteristic 12 G hyperfine splitting of N-donors in SiC. The two dominant spin resonances observed in GaN with conventional ESR are the deep defect (DD) at g2.00 and the effective mass donor at g=1.948. The deep defect, which is the intrinsic defect in GaN epilayers, has been linked by ODMR measurements to the yellow * phone: +49 89 289-12755, fax: -12737, email: martin.bayerl@wsi.tu-muenchen.de

17 luminescence. In addition, the intensity of the corresponding EDMR resonance increases under degradation of various GaN-LEDs. The fact that the linewidth of the deep defect resonance is very similar at 9 GHz and 34 GHz suggests unresolved hyperfine interactions as a probable broadening mechanism. The shallow EM donor, however, is not observed in the present study. Instead, two deep donor like states are found at g=1.983 and g=1.973. Similar resonance signatures have been observed with ODMR on n-type and semi-insulating GaN. Although it is likely that these resonances are due to impurities, the exact origin remains to be identified. However, it can be noted that low-temperature recombination in the CREE light emitting diodes studied include spin-dependent transitions between two different deep donors and the deep defect.

2x10

-7

Pmw = 2000 mW = 9.37 GHz T = 4.5 K

6x10

-8

deep donor 1 g = 1.97 H pp = 30 G

4x10

-8

Pmw = 250 mW = 34.0 GHz T = 4.5 K

deep donor 2 g = 1.983 H pp = 46 G deep donor 1 g = 1.973 H pp = 80 G

1x10

-7

2x10

-8

/
-7

-1x10

-2x10

-8

-2x10

-7

native defect g = 2.00 Hpp = 170 G

-4x10

-8

native defect g = 2.003 H pp = 160 G

-6x10

-8

3000

3200

3400

3600

11800

12000

12200

12400

Magnetic Field (G)

Magnetic Field (G)

Figure 1: EDMR spectra of the CREE GaN LED measured at microwave frequencies of 9.4 and 34 GHz. The dotted lines show simulations of the constituent resonances obtained by a least square fit. The dashed line represents the sum over all the constituent lines. *

In summary, we have shown that EDMR at 34 GHz can provide additional spectroscopic information compared to standard EDMR at 9 GHz. The excellent signal to noise ratio of the high-frequency EDMR is currently used to investigate recombination in other GaN-based green and blue LEDs including quantum-well based structures by Nichia. The EDMR measurements will be complemented by optically detected magnetic resonance (ODMR) investigations on GaN and AlGaN films at both frequencies 9 and 34 GHz, with the aim to identify the microscopic nature of some of the defects which are found in group-III nitride films and quantum-wells.

*Supported

by Deutsche Forschungsgemeinschaft (SFB 348 and Stu 139/3-1).

18

Recombination centers in GaAs investigated by electrically detected magnetic resonance


Thomas Wimbauer*,Martin W. Bayerl, Nils M. Reinacher, Martin S. Brandt and Martin Stutzmann Out of the large variety of experimental methods to characterize semiconductor materials, magnetic resonance techniques are probably the most powerful to obtain detailed information on point defects in the materials. In particular, knowledge about the chemical nature of the defects can be obtained from electron paramagnetic resonance (EPR) by the observation of fine-structure-, hyperfine-, or ligand-hyperfine-interactions. The major problem of the application of the EPR spectroscopy to epitaxial layers or quantum well structures is the insufficient sensitivity of the method. The reduced thickness of the sample pushes the number of defects well below the detection limit of 1011 spins/G halfwidth. Thus, most spin resonance investigations on such structures make use of the more sensitive technique of optically detected magnetic resonance (ODMR). However, when such structures are investigated in electronic devices, the detection of the effect of defect states on transport properties is important. This can be achieved by the related method of electrically detected magnetic resonance (EDMR). Most EDMR experiments reported so far deal with defects in bulk semiconductors, bulk like thin layers or simple p/n diode structures, mostly in silicon. Here we present an extension of this technique to III - V semiconductors in the form of an exploratory investigation on a GaAs/AlGaAs heterostructure. The structure of the MBE grown sample essentially consists of an 50 nm thick GaAs layer co-doped with Si (1 x 1016 cm3 ) and Be (3 x 1016 cm-3) grown on a semiinsulating GaAs substrate with an 800 nm Al0.4Ga0.6As buffer layer. The spin-dependent photoconductivity signals depend strongly on the electrical contacts. For the first type of EDMR investigations silver paste contacts on top of the GaAs cap layer were made. A single isotropic resonance line located at g = 2.001 with a peak - to - peak linewidth of Bpp = 18 G was observed, previously attributed to surface dangling bonds. This result indicates that the recombination current is localized in the GaAs cap layer. In a second step, In-contacts (alloyed at 480C for one minute) were used in order to contact all regions of the layered structure. Using these contacts the spectrum shown in Fig. 1 was detected. The dominating central signal is well described by a gaussian, has a g-value of g = 1.99 and a FWHM of 200 G. These resonance parameters are consistent with Cr4+ (3d2, S = 1) located in the semiinsulating GaAs substrate to which electrical contact was also made by the alloying of the contacts. In order to describe the low field and high field parts of the experimental spectrum the well known EPR parameters of the arsenic antisite defect AsGa were used (g = 2.047, hyperfine constant A = 0.089 cm-1, nuclear spin I = 3/2 of As). The simulation shows that the two outermost resonances of the experimental spectrum can be explained by the existence of AsGa, which is also located in the semiinsulating GaAs substrate. Further including the Ga-interstitial Gai (g = 2.009, nuclear spin I = 3/2 of 69Ga and 71Ga, natural abundance ratio of 61/39, hyperfine constants of both isotopes A69 = 0.053 cm-1 and A71 = 0.067 cm-1) observed in the same sample by ODMR in the co-doped GaAs layer, a better agreement with the experimental spectrum can be achieved with respect to the shoulders localized more closely to the central high intensity line. A clear separation of the signals from the substrate and the co-doped layer can be achieved using inhomogeneously absorbed light (h > 1.5 eV instead of white light) and phase shift analysis. Strong absorption of light in the co-doped layer and weak absorption in the substrate leads to different photoconductivities of the involved layers. Therefore each layer has a different RC time constant which results in different phase delays between modulation and detection. This fact enables the experimental separation of the substrate signals and the signals originating from the co-doped layer. In this *phone: +49 89 289-12755, fax: -12737, email: thomas.wimbauer@physik.tu-muenchen.de

19

Fig. 1: EDMR spectrum of the heterostructure using alloyed In-contacts and the decomposition of the experimental spectrum into Cr4+, AsGa and Gai.

Fig. 2: EDMR spectra of the separated Gai at different microwave frequencies.

way the spectra of the Gai shown in Fig. 2 were recorded, which shows comparative X-band and Q-band measurements. Included in the figure are the simulations of the Gai resonances for the corresponding frequencies using the above mentioned defect parameters. After this demonstration that EDMR can be successfully employed to identify recombination processes in GaAs - based heterostructures, studies are under way to investigate the details of spin dependent recombination in III - V materials, in particular aimed at a possible determination of defect concentrations using this method.*

*Supported

by SFB 348, Teilprojekt A3

20 Raman characterization of AlxGa1-xN alloys

Ana Cros* , Helmut Angerer, Frank Freudenberg, Oliver Ambacher** and Martin Stutzmann We have performed a Raman study of the optical phonons of AlxGa1-xN layers grown by plasma-induced molecular beam epitaxy on sapphire substrates. The goal of this work is to determine the behavior (one-mode vs. two-mode) of the optical modes in the entire compositional range (0 x 1). The optical phonons were studied at room temperature using a micro-Raman system, which allowed us to focus the laser beam to a spot size of about 2m, with a power density of 105W/cm2. The micro-Raman system has the advantage, over the standard macroscopic setup, of restricting the focal depth to about 2m, which efficiently diminishes the signal arising from the sapphire substrate. This is particularly important in our case, where the samples to be characterized are around 1m thick. The scattered light was analyzed by a DILOR triple spectrograph with a cooled CCD detector array.

Fig. 1: Room temperature Raman spectra obtained for different AlxGa1-xN films. Typical Raman spectra obtained in backscattering polarization along the wurtzite c axis are shown in Fig. 1 for various alloy compositions. The position of the different phonon modes is indicated with circles (E2), empty triangles (A1(TO)) and full triangles (A1(LO)). Both A1(TO) and E2 phonon modes shift to higher energy as the aluminum content of the alloy is increased, but the A1(TO) mode moves faster than the E2 line. An interesting feature arises for an alloy composition of 36% aluminum, where a new mode is observed as a broad peak centered at 629 cm-1. The intensity of this line slowly increases with increasing aluminum content, and it shifts slowly to higher energy towards the position of the E2 line of AlN. The line is indicated by arrows in the different spectra of Fig. 1a. At the same time, the intensity of the E2 mode
* Permanent ** Tel:

address: Dep. Fsica Aplicada, Valencia, Spain. e-mail: Ana.Cros@uv.es +49(89)289-12735, Fax:+49(89)289-12737, e-mail : Ambacher@physik.tu-muenchen.de

21
decreases, and for an aluminum content greater than 89% the peak cannot be detected anymore. This behavior indicates the existence of one E2 mode characteristic of GaN and one characteristic of AlN for an intermediate alloy composition (two-mode behavior). Figure 1b shows the Raman spectra corresponding to the A1(LO) phonon mode. This mode shifts smoothly from the GaN to the AlN energy position with increasing aluminum content. No AlN localized mode has been found for this spectral range, indicating a one-mode behavior for this mode.

Fig. 2: Dependence of the frequencies of the A1(TO), E2 and A1(LO) phonon modes on alloy composition.

Fig. 3: Dependence of the linewidth of the A1(TO) and E2 phonon modes on aluminum content.

A better insight of the behavior of the different phonon modes can be obtained from Fig. 2, where the position of the phonon peaks is shown as a function of alloy composition. The A1(LO) mode clearly shows a one-mode behavior. However, it does not follow the straight line joining GaN and AlN (full line), but bends up towards higher energy. At lower energy in the figure, the two-mode behavior of the E2 mode (dots) can be clearly identified. The full lines represent a linear fit to the points. Figure 3 shows the compositional dependence of the linewidth of the A1(TO), E2(GaN) and E2(AlN) phonon modes on aluminum content. The A1(TO) linewidth remains initially almost constant as the amount of aluminum is enhanced, but increases quickly for an aluminum content greater than 70%. The curve shows a clear maximum at x = 0.8, wich correlates with the structural quality of the samples, measured by means of X-ray diffraction. On the other hand, the linewidth of the GaN E2 mode increases with aluminum content, while the AlN E2 line clearly becomes narrower for x 1, as expected in a two-mode behavior. In conclusion, the behavior of the optical phonons in AlGaN has been studied by means of micro-Raman spectroscopy. Both one and two-mode behavior are simultaneously observed in the III-V semiconductor for different phonon modes. While the A1 phonon modes show a onemode behavior, the E2 phonon mode shows a two-mode type behavior. From our measurements, the frequency of the Al local mode in GaN is estimated to be 616 cm-1, while the Ga gap mode in AlN would arise at 605.5cm-1.
Supported by DFG 138/3-1 and Ministerio de Educacin y Ciencia, Spain.

22

Vibrational Spectroscopy of Siloxene


Nikta Zamanzadeh*, Martin S. Brandt and Martin Stutzmann Silicon with its indirect bandgap cannot be used for active optoelectronic devices. On the other hand there exist a variety of silicon-based materials which exhibit strong luminescence, such as amorphous Si:O:H, porous silicon and siloxene. Among these materials, siloxene is a particular interesting substance: its layered structure with two dimensional character leads to a direct bandstructure with a bandgap of 2.5 eV. Siloxene is prepared by a topochemical reaction of concentrated HCl and metallic CaSi2. During the last years, our investigations have concentrated on the preparation of crystalline siloxene and the study of the anisotropy of its physical properties such as conductivity. Vibrational spectroscopy of this two-dimensional modification of silicon has been extensively used to study ligands as well as degradation of the siloxene layers. However, the current assignment of dominant vibrational modes is still under discussion, in particular the exact origin of the IR-active mode at 515 cm-1. We have therefore studied the effects of isotope substitution (1H 2D and 16O 18O) on the FTIR- and Raman-spectra of siloxene.

Transmittance (a. u.)

hydrogenated

Figure 1 :FTIRSpectra of siloxene: prepared with HCl (straight line) and DCl (dashed).

deuterated

Siloxene 300 K 0 500 1000 1500 2000 Wavenumber (cm-1) 3000 4000

The FTIR-spectra of hydrogenated and deuterated siloxene are shown in Fig. 1. The FTIR spectra change as expected under H-D substitution. In particular, the 2100 cm-1 SiH stretching vibration shifts to 1530 cm-1. More interesting is the strong FTIR line at 515 cm-1, which was previously attributed to silicon phonons which become IR-active due to the different H- and OH-ligands at the silicon planes. However, this line also shifts under H-D substitution to 370 cm-1, as expected for a hydrogen-related mode.

: +49 89 289 12768, fax: +49 89 289 12737 e-mail: nikta@wsi.tu-muenchen.de

23 Similar substitution experiments on a-Si:O:H have shown that the 880 cm-1 spectral line, which has been linked to the red photoluminescence exhibited by this material, is in fact a combination of two degenerate local vibrational modes, one involving hydrogen atoms. Under the substitution of siloxene with 18O no major changes in the spectrum are observed because of the small relative mass difference of 16O and 18O. However a clear shift of the Si-O-Sistretching mode (from 1034 cm-1 to 1013 cm-1) and of the O-H-stretching vibration (from 3595 cm-1 to 3574 cm-1) is found. The corresponding Raman spectra have been obtained with a He-Ne Laser (632.8 nm, 1 mW). Figure 2 shows the room temprature Raman spectra. The top spectrum shows siloxene prepared with HCl only. The strongest Raman line in hydrogenated siloxene is observed at 496 cm-1. Similar to the 515 cm-1 mode in IR, it has been attributed to silicon phonons. Here, an anomalous and rather surprising behaviour is observed upon H-D substitution: The Raman mode appears to harden and shifts upwards in energy to 575 cm-1. The hardening is a continuous function of the degree of deuterium substitution as seen in Fig. 2.

Whler-Siloxene
HCl DCl
1 0

HCl
0.84

DCl
0.16

Raman Intensity (a. u.)

HCl
0.67

DCl
0.33

HCl
0.45

DCl
0.55

Figure 2: RamanSpectra of siloxene made with different mixtures of HCl and DCl. The top curve corresponds to purely hydrogenated, the bottom curve to purely deuterated siloxene.

HCl
0.25

DCl
0.75

HCl DCl
0 1

450

500

550 600 650 Raman Shift (cm -1)

700

In fact a closer inspection of the Raman spectra suggests that, the 496 cm-1 softens and decreases in intensity upon gradual deuteration, while at medium D concentration at least one additional mode occurs. This behaviour, which in this clarity has not been observed in any SI-H system before, is probably due to a Fermi resonance between the 496 cm-1 mode and a hydrogen-related local vibrational mode (such as the bending mode at 730 cm-1) shifted downward to this energy range by the H-D substitution. Theoretical work is under way to further understand this anomalous behaviour.

DFG-Schwerpunkt Silizium-Chemie (Stu 139/4-1)

24

Inelastic Light Scattering by Electronic Excitations and Phonons


Edilson Silveira1, Werner Dondl, Michael K. Kelly, Gerhard Abstreiter In a perfect isotopically pure crystal lattice, the introduction of another type of isotope leads, in a first approximation, to an ideal example of a mass defect. The isotope mass fluctuations in bulk natural Ge are not large enough to produce phonon localization near such mass defects. However, in contrast to 3D systems, 2D and 1D structures with isotope mass fluctuations will always have localized states. The detailed theoretical understanding and the controlled preparation of compounds with defined mass changes has been developed recently. Raman scattering is a non-destructive method to investigate structural properties of isotopic superlattices in which layers of two isotopically enriched materials alternate periodically. Such superlattices are not accessible for example to x-ray analysis. Fig. 1(a) shows experimental Raman spectra of a (70Ge)8(74Ge)16 superlattice after various annealing steps at 500 C. The evolution of the positions and intensities of these modes, as a function of the annealing time at a fixed temperature, gives information about the interface intermixing and consequently about the selfdiffusion of Ge. In Fig. 1(b) theoretical results are presented. The Raman spectra were calculated using the planar force constant model and the bond polarizability approach. The average masses of the atoms in a crystal plane perpendicular to the growth direction were interpolated from the concentration profiles obtained from Ficks diffusion equation. The best agreement between the calculated and the experimental results is achieved using a selfdiffusion constant Dsd = 5.50x10-24 m2s-1.

Fig. 1. (a) Experimental Raman spectra of a (70Ge)8(74Ge)16 superlattice for different annealing steps at 500 C. (b) Calculated Raman spectra for the same superlattice.
1

tel.: ++49/89/28912775; fax: ++49/89/3206620; email:silveira@wsi.tu-muenchen.de

25 In another project we used inelastic light scattering to study electronic excitations in microstructured GaAs/AlGaAs systems. Due to the lack of inversion symmetry in the zincblende crystalline structure, GaAs has a spin-orbit splitting in the conduction band, even at zero magnetic field. In heterostructures the effect can be inequivalent along the [1 1] and PW, depolarized EL=1695.2 meV the [11] directions. Two samples were laterally q // (105 cm-1) wires // [11] structured along these directions, Sample A 1.49 [11], and Sample B [1 1], in order to obtain spatially separated wire systems. The samples 1.32 were structured using two polarized pulsed 1.10 laser beams forming an angle of 90 between them. The resulting interference pattern 0.86 produces a permanent, spatially sharp potential 0.59 modulation, due to the transient thermal grating at the surface of the semiconductor 0.30 heterostructure. Fig 2 shows the wave vector dependence of the depolarized spectra of 0 Sample A in plane wave scattering geometry. In this scattering geometry intrasubband SPE 2 4 6 8 10 are expected to dominate the spectra. For Energy (meV) vanishing momentum transfer, only one peak Fig. 2. Depolarized Raman spectra of the wire at about 3 meV can be detected. It is attributed sample in plane wave geometry for different wave to the only parity-allowed, n=2 intersubband vector q//. The wires were structured along the [11] transitions, also observed in standing wave direction. geometry. For large wave vectors, a strong broadening and a decrease in its intensity is observed. For non-zero q//, a peak at smaller PW, depolarized energies appeas. It shows a linear dispersion EL=1695.4 meV q // (105 cm-1) with in-plane wave vector. This peak is attriwires // [11] 1.49 buted to intrasubband SPE. In Fig. 3 depolarized spectra of sample B are displayed 1.32 in plane wave scattering geometry for different in-plane wave vector transfer. The 1.10 wires in this sample are structured along the 0.86 [1 1] direction. Similar as in sample A, the spectrum for zero wave vector shows one peak 0.59 at about 3 meV, due to n=2 intersubband excitations. This confirms that the structuring 0.30 is identical for the two samples, except for wire orientation. With increasing q// one 0 observes two peaks with almost the same 2 4 6 8 10 dispersion relation. The doublet structure of Energy (meV) the intrasubband SPE for non-vanishing inplane momentum transfer is attributed to tranFig. 3. Depolarized Raman spectra of sample B in sitions between the two spin-split conduction plane wave geometry for different wave vector q//. bands. The dispersions of the two peaks are The wires are oriented along the direction. linear and parallel, which supports their assignment to inter-spin-subband excitations.
supported by: DFG (SFB 348) and DAAD
Intensity (a. u.) Intenstiy (a. u.)

26

Micro-photoluminescence studies of quantum wells, wires and dots


Wolfgang Heller#, Ulrich Bockelmann*, Arianna Filoramo+ and Philippe Roussignol+ The application of a magnetic field to a single quantum dot (QD) allows to measure the Zeeman spin splitting and the diamagnetic shift. The former gives information on the exciton gfactor, the latter is determined by the interplay of lateral confinement, magnetic confinement and Coulomb interaction. We perform photoluminescence (PL) and photoluminescence excitation (PLE) spectroscopy with a spatial resolution of 1.5 m at a temperature of 4 K. The laser beam is focused through a microscope objective onto the sample, which is mounted in a continuous flow magnet cryostat (5 T). The luminescence is collected by the same objective and dispersed by a triple grating Raman spectrometer. A pinhole in an image plane defines the detection area. We investigate two different types of QDs, a 3.5 nm wide GaAs quantum well where monolayer fluctuations give rise to a localization of excitons and a series of QDs fabricated by laser induced thermal interdiffusion of a 3 nm wide GaAs quantum well. Fig. 1a shows PL and PLE spectra of a QD formed by monolayer fluctuations. The luminescence line of the QD appears about 13 meV below the dominant quantum well luminescence at 1671 meV. The two sharp peaks in the PLE spectrum are attributed to excited

# * +

tel.: ++49-89-3209 2736, fax: ++49-89-3206 620, e-mail: heller@wsi.tu-muenchen.de currently at LPMC, Ecole Normale Suprieure, Paris, email: ulrich@physique.ens.fr LPMC, Ecole Normale Suprieure, 24 rue Lhomond, 75005 Paris

27 states of the dot. We find linewidths as narrow as 0.1 meV. In a magnetic field the three lines split into doublets (shown in the inset), the splittings at 5T are 0.4 meV for the ground state and about 0.7 meV for the excited states. The sign of the g-factor of the second excited state is reversed. This is attributed to valence band mixing effects. Another interesting feature is that the spin relaxation depends on the magnetic field B. This is shown in the inset. When the dot is pumped at energy Ea the lower energy line of the ground state doublet is dominating, which means that at nonzero B the spin is conserved to a large extent during relaxation. Fig. 1b shows data from an experiment where the dot is pumped resonantly into one of the excited states with defined circular polarization and the polarization P = ( I + I ) / ( I + + I ) of the ground state luminescence is measured. We find that the spin relaxation is increasingly quenched when the field is raised. While at zero magnetic field an elastic scattering process is possible between the spin states, an energy exchange with the lattice is required when the spin states are split by a magnetic field. In another experiment we apply lateral electric fields to these dots. We find that the ground state as well as the excited states shift to lower energy when the field is raised which is explained by the Stark effect. The line width of the ground state broadens with increasing field due to lateral tunneling out of the dot.
1.0

B=5T
0.8

0.6

0.4

0.2 1000 nm 500 nm 450 nm 400 nm 1700 1710 1720 1730

0.0

PEAK ENERGY (meV)

We also investigate a series of QDs produced by laser induced thermal interdiffusion of a quantum well. The geometrical width varies from 400 nm to 1 m. In a magnetic field we find that the diamagnetic shifts and spin splittings of these dots are different. The Zeeman spin splitting increases with the peak energy as shown in Fig 2. In quantum wells it has been shown that the exciton g-factor increases with decreasing well width. This was attributed to the increase of the quantization energy. Therefore an increase of g with the PL energy might also be expected for a QD. The diamagnetic shift decreases with the dot size because the magnetic confinement becomes less important if there is already a strong lateral confinement.

In collaboration with the Ecole Normale Suprieure Paris, the spatial resolution is Fig. 2: Spin splitting at 5T as a function of combined with a time resolution system in the the peak energy for various quantum dots. picosecond range. This setup is employed for time resolved spectroscopy of the single QDs produced by thermal interdiffusion. The time evolution of the PL lines gives detailed information on the relaxation mechanisms in quantum dots. For example, the measured short rise time (about 10 ps) of the PL indicates that Coulomb scattering plays an important role in the energy relaxation process. The same setup is used for exciton diffusion experiments in quantum wells and wires. In this case, a displacement of the pinhole with respect to the exciting laser spot allows us to monitor the lateral change of the PL time-dependence.

Supported by: DFG (SFB 348) and a PROCOPE Project

SPLITTING (meV)

28

Spin-polarized excitons in GaAs/AlAs coupled quantum well structures


Artur Zrenner1 , Andreas Schaller and Marcus Hagn2 In GaAs/AlAs coupled quantum well (CQW) structures various interesting excitonic properties of low-dimensional systems can be studied. In the last issues of our WSI report we have already presented selected topics, such as the properties of quantum dots formed by interface fluctuations and the electric field-induced exciton transport. In the first case the electric-field tunable CQW structures are operated as internal energy spectrometers to investigate the properties of natural quantum dots in magnetooptic studies, in the second case advantage is taken from the extended lifetime and the huge Stark-shift of the indirect excitons. In the present issue of our report we address spin-related properties of CQW-structures. In our studies we investigate the magnetooptic properties of excitons in single natural quantum dots, in particular under the condition of resonant charge injection via the AlAs X-point. Due to the absence of inhomogeneous broadening single quantum dots exhibit atom-like emission spectra and the Zeeman-splittings can be fully resolved in optical experiments (see fig. 1). Whereas the effective gfactor of the ground state exciton 0D-X (g=-1.49) is basically given by the dominant confinement of the 30 GaAs QW, the intensities of the Zeeman components show an anormal behaviour in the regime of high magnetic fields (B10T). The observed increase of the high-energy component suggests an inverted occupancy of the Zeeman levels. As shown in figure 2a and b, this behaviour originates from the path of relaxation and recombination in the GaAs/AlAs CQW structure. Roughly equal occupancy of the Zeeman-levels is found for the case of excitation, relaxation and recombination in the GaAs QW (see figure 2a and corresponding inset). The previously described inverted occuFig 1: Magnetic field dependence of the ground state pancy of the Zeeman levels is only exciton line (0D-X) and the corresponding biexciton observed if the photoexcited electron line (0D-XX) in a natural quantum dot. is scattered to the AlAs-layer, forming an indirect exciton with a
1

tel: +49/89/289-12772, fax: +49/89/3206620, e-mail: zrenner@wsi.tu-muenchen.de tel: +49/89/289-12756, fax: +49/89/3206620, e-mail: hagn@wsi.tu-muenchen.de

29

Fig. 2: (a) Exciton (0D-X) and biexciton (0D-XX) line for various magnetic fields. Relaxation and recombination is via -point states only. (b) The AlAs X-point subband is introduced as intermediate state. The associated long-lived indirect exciton allows for spinthermalization prior to backinjection to the -point and subsequent recombination. lifetime in the range of 100 nsec (see fig. 2b). In contrast to direct excitons complete spin thermalization (T=700mK) becomes now possible within the exciton lifetime. With the positive effective g-factor of the GaAs/AlAs indirect exciton only the |3/2,-1/2> Zeeman state will remain occupied (see fig. 3). By spin-conserving -X tunneling, this state can be converted to a direct quantum dot exciton with negative g-factor, which is represented now however by the high-energy Zeeman branch. Subsequent recombination from the quantum dot state is under such conditions therefore highly spin-polarized and population inverted. Fig. 3: Zeeman levels for the direct and indirect exciton in a GaAs/AlAs CQW structure. Spin-conserving -X tunneling results in population inversion for the direct exciton.

Supported by: DFG (ZR 2/1-1) and DFG (SFB 348)

30

Self-assembled InAs/GaAs quantum dots


Markus W. Arzberger1, Markus Hauser, Gerhard Bhm, Artur Zrenner, Gerhard Abstreiter The electronic and optical properties of semiconductors can be changed drastically by reducing the dimensionality. Whereas two-dimensional systems can be achieved quite easily by molecular beam epitaxy (MBE) or metal-organic chemical vapour deposition (MOCVD), it is much more difficult to obtain high quality one (quantum wires) or zero-dimensional structures (quantum dots, QDs). During the past few years the Stranski-Krastanow growth mode, which appears in heteroepitaxy of strained systems, e.g. InAs/GaAs or Ge/Si, was studied by many groups in order to realize self-organized QDs. Here we are interested in the optical properties of InAs/GaAs-QDs grown by MBE with the aim to use them as active medium in a vertical cavity surface emitting laser (VCSEL). These lasers should have superior properties compared to conventional semiconductor quantum well lasers. Due to the three dimensional confinement the QDs have a discrete energy spectrum, similar to atoms. However, the photoluminescence (PL) measured from an ensemble of about 107 QDs is inhomogeneously broadened because of size fluctuations of the QDs. We have studied the excitation power and temperature dependence of MBE-grown selfassembled InAs dots embedded in GaAs. The growth temperature and the InAs layer thickness were varied. Typical PL spectra of a nominally 2.25 ML thick InAs layer Fig. 1: PL-intensity at different excitation densities. grown at T=530C are shown in Fig. 1. When the excitation density is increased, higher states in the QDs are populated and PL peaks at higher energies are observed. The ground state PL (1.085 eV) is already saturated at 10 Wcm-2. The wetting-layer luminescence, 350 meV apart from the ground level PL, is only seen at fairly high excitation densities. There is no hint for suppressed relaxation from the 2Dwetting layer into the 0D-QD states. Based on the PL studies we have realized a Bragg resonator, where three layers of InAs QDs are embedded (Fig. 2). They are separated by 30 nm of GaAs, so that the amount of strain is too low to generate dislocations and QDs of different layers do not electrically couple. The cavity consists of two distributed Bragg reflectors (DBRs) composed of 28.5 respectively 18 AlAs/GaAs mirror pairs and a 3-spacer. The current is injected through an n-doped back mirror and a lateral p-contact, in order to avoid p-doping of the upper mirror. The energy of the cavity mode is designed to coincide with the QD ground state luminescence at low tempera1

tel: ++49-89-28912759, fax: ++49-89-3209620, email: arzberger@wsi.tu-muenchen.de

31 tures. The design and processing technology was adopted from the InGaAs-quantum well-VCSEL project also described in this report. The PL obtained from the cleavage plane of this structure is plotted in Fig. 3 together with the normal- incidence reflectivity and the electroluminescence. The cleaved edge PL of the QDs in the cavity (Fig 3a) is very similar to the PL of QDs without mirrors. In addition there appear two sharp peaks due to the resonator cavity. The reflectivity of the unprocFig. 2: Schematic structure of the processed Bragg-resonator. essed structure at T=7.5 K is shown in Fig. 3b. The cavity dip in the middle of the stop band is clearly observed at about 1.08 eV, which is close to the QD ground state luminescence. Since the luminescence of the QDs (3a) is modulated by the properties of the resonator (3b), the electroluminescence (EL) (Fig. 3c, intensity plotted on a logarithmic scale) is dominated by a narrow (FWHM 0.38 meV) and intensive line at the energy position of the cavity resonance. The intensity of this line already saturates at an injection current of about 40 Acm-2, which is very similar to the value we obtained for the saturation of the ground state EL-intensity from a sample without DBRs. Therefore, laser threshold is not yet reached in the cavity sample. Using the QD density of 1.41010 cm-2 determined by AFM measurements at a sample with QDs grown under similar conditions but without any cap layer, we can roughly estimate that only about 200 QDs contribute to the sharp line. The maximum gain is not sufficient to compensate the absorption and mirror losses, most probably because of this small active volume.
supported by: DFG (SFB 348) Fig. 3: Cleaved edge PL, reflectivity (unprocessed structure) and EL of the diode shown in Fig. 2.

32

Quantum dots fabricated by twofold cleaved edge overgrowth


Gert Schedelbeck, Werner Wegscheider*, Martin Rother, Max Bichler, Gerhard Abstreiter Cleaved Edge Overgrowth (CEO) has proven to be a powerful technique for the fabrication of atomic-scale T-shaped quantum wires (QWRs) which form at the intersection of two quantum wells (QWs). Here we report on the first experimental demonstration of quantum dots (QDs) which result when three QWs intersect each other at a right angle. Optical emission from zerodimensional (0D) states in these QDs, which were fabricated by a twofold CEO technique, is clearly identified by means of micro-photoluminescence (PL) and PL excitation (PLE) spectroscopy. In contrast to the inhomogeneously broadened QW and QWR signals originating from the complex sample structure, the QD response, which is characterised by sharp lines (FWHM < 70 eV), is strongly spatially localised at a position where the QWs meet. Figure 1 illustrates the layer sequence of the QD structure obained after three molecular beam growth steps each separated from the following one by an in situ cleave. It consists of 22 (001) oriented 7 nm wide GaAs QWs, labeled MQW0, which were overgrown two times with slightly narrower single QWs (labeled QW1/QW2), oriented normal to the two <110> directions. At the intersection of each two QWs, T-shaped QWRs form as schematically depicted by the white contour plots (||2 = const.) of the one-dimensional (1D) wavefunction. The intersection of the three orthogonal QWs (MQW0, QW1 and QW2) and QWRs, labeled QWR0,1/QWR0,2/QWR1,2, each leads to the formation of a QD as visualized in the magnified portion of Fig. 1. Fig. 1: Schematic illustration of QD formation at 3 intersecting QWs and QWRs, respectively. The approx. 7 nm wide GaAs QWs are embedded in Al0.35Ga0.65As barrier material.

Typical low-temperature PL spectra (5K) and grey scale intensity plots from a 5 m 5 m area scan (pixel size 200 nm) recorded from the third growth surface are shown in Fig. 2. Due to the high spatial resolution (FWHM of the detection window 800 nm) and the high degree of perfection in the structure, distinct luminescence features from the individual types of lowdimensional structures, i.e. from the QWs, QWRs and QDs, are well resolved both spatially as well as spectrally.
*

tel.: ++49/89/2891-2776; fax: ++49/89/3206620; e-mail: wegscheider@wsi.tu-muenchen.de

33 Fig. 2: PL spectra and grey scale area scans taken from the third growth surface. The positions on the sample surface where the spectra were recorded are given by the black square in the inset. The arrows indicate the spectral position of the corresponding grey scale intensity plot.

At an energy of 1589 meV, emission from MQW0 is observed. Due to exciton diffusion into the QWRs and QDs, the PL intensity decreases over a length of more than 1m [Fig. 2(a)]. The PL energy of the (110) oriented quantum wells QW1/QW2 is about 1575meV, well below the corresponding energy of MQW0. This is a result of the large heavy hole mass for <110> type confinement directions. Again, bleaching of the QW PL signal in the vicinity of QWRs and QDs, indicating considerable exciton diffusion, is observed [Fig. 2(b)]. Emission from the QWRs, centered at around 1565 meV, can be detected at all positions shown in Fig. 2, i.e. ontop of QWR0,2, QWR1,2 and QWR0,1. The wire emission is strongest at the position studied in Fig. 2(c) (not shown in the grey scale plot). This is caused by the contribution from the wires QWR0,1, which are oriented along the detection direction. Precisely and exclusively at the intersection of the wells and wires, several sharp and intense lines appear at about 1562 meV, indicating that the QD state is about 3 meV deep with respect to the QWR state. Their intensity decrease in all directions is only limited by the experimental resolution. From this strong spatial localisation, the energetic position and a pronounced absorption under resonant PLE conditions, these lines are clearly identified as PL originating from quasi-0D excitons bound to the QDs.

supported by: DFG (SFB 348), BMBF (01 BM 630)

34

Transport in Si/SiGe-nanostructures* i
Martin Holzmann**, Markus Riedinger, Gerhard Abstreiter Modulation doped Si/SiGe-heterostructures are of interest for quantum devices such as quantum wire or single electron transistors because of their superior transport properties. The electron mobilities exceed by far those of Si-MOSFETs both at room and at helium temperatures. The top electron mobilities achieved in samples grown with our MBE system are above 150000 cm2/Vs. Up to now the main problem for nanostructures is the patterning process. The great challenge is the fabrication of structures in the order of 100nm and below without loosing the high quality of the two-dimensional electron gas (2 DEG). We focused mainly on two nanopatterning processes which were developed in collaboration with the LMU Mnchen (Prof. Kotthaus) and the RWTH Aachen (Prof. Kurz). The first method is a combination of a sophisticated reactive ion etching (RIE) process with an in-situ SiO2-passivation of the etched trenches (Fig. 1b) which form the barriers. We used in-plane-gate transistors to investigate the quality of the nanopatterning process. The 2 DEG in the strained Si-channel has a mobility of = 71400cm2/Vs and a corresponding carrier density of ns = 7.71011cm-2 at T = 4.2K. The desired gate geometry (Fig. 1a) was defined by electron beam lithography (EBL). Two V-shaped gates in a distance wgeo form the point contact transistor. A width of 100nm was chosen for the etch trenches throughout the whole series of transistors. The etching of the Si/SiGeheterostructure and the SiO2-passivation were performed in a two chamber clustertool. The dry chemical etching and the oxide deposition take place in the same machine but in different chambers. Immediately after the RIE process, which cut through the 2 DEG the samples are transferred into the deposition chamber within the same ultra-high-vacuum system. This process is designed to avoid contamination of the etched surface and guarantees a high quality interface between the different layers of the heterostructure and the SiO2.

The transistor behaviour has been studied in dependence on wgeo. Transistors with a geometrical gate width of only 0.3m could still be operated in the normally-on mode (Fig. 2). From the saturation currents of a series of transistors with different gate widths between 0.3m and 2.1m we deduce lateral depletion lengths below 50nm. Therefore the resulting barrier widths
*

Fig. 1: (a): Schematic drawing of an IPGtransistor. (b): Atomic force micrograph of IPGs. The tip distance is 0.5m. The black lines are the etched and SiO2-passivated barriers.

**

in collaboration LMU Mnchen and RWTH Aachen tel.:+49-89-28912756, fax.+49-89-3206620, email: holzmann@wsi.tu.muenchen.de

35

in the plane of the 2 DEG are below 200nm including the 100nm broad etch trenches. This is far below the values recently obtained by focused laser and ion beam patterning of Si/SiGe-hetero-structures. The developed fabrication method turned out to be a very promising tool for nanostructuring Si/SiGe. In a second approach we realized singleelectron-transistors by a split-gate-technique. Therefore, Pd-Schottky-gates with a minimum width of 100nm were defined by electron beam lithography. The 2 DEG also embedded in a strained Si-channel exhibits a mobility of = 167000cm2/Vs and a corresponding carrier density of ns = 4.451011cm-2 at T = 0.37K. The top-gate geometry is shown as inset of Fig. 3. The first observation of coulomb blockade oscillations in this system is plotted in Fig. 3. The measurements were performed in a dilution refrigerator at T = 30mK. The gate fingers marked by arrows were tuned between 0V and -0.4V. The lower part displays the region between -0.2V and -0.4V. The conductance G clearly oscillates in this voltage regime. From the oscillation period ( V ) one can deduce the gate capacitances ( CG ) and thereby the dot capacitance. The corresponding dot radius ( R ) amounts to about 60nm. Altogether, the observed phenomena are very similar to those known from GaAs/AlGaAs structures. However, the stability of the Si/SiGe devices is the main problem. Charging effects lead to a continous change of the electrostatic potential of the island. Supported by BMBF under Grant No DLR 400101 M 2413

Fig. 2: Transistor characteristics of an IPG-transistor at T=4.2K. The geometrical gate width is 0.3m.

36

Photoconductivity of polycristalline CVD diamond films


Eberhard Rohrer, Christoph E. Nebel* and Martin Stutzmann After the discovery that large-area diamond films can be grown by the chemical vapour deposition (CVD) technique, diamond attracts increased interest due to its unique thermal, optical and electronic properties. However, diamond thin films still contain a lot of defects, grain boundaries and non-diamond phases like graphite and amorphous carbon. In addition, up to now, efficient n-type doping has not been achieved. The knowledge of the nature of dopants and defects is required to improve the quality of the layers. Sub-bandgap absorption and photoconductivity is a common technique to probe the density of states of wide bandgap semiconductors where measurements like deep level transient spectroscopy (DLTS) are difficult to perform. The diamond films investigated at the Walter Schottky Institute were grown at the Daimler-Benz Research Center in Ulm. Nitrogen and boron are the main dopants incorporated during the growth of CVD diamond. Boron gives rise to an acceptor level at 0.37 eV above EV and nitrogen introduces various energy levels in the bandgap of diamond, known from measurements on synthetic crystals. Fig. 1 shows a typical absorption spectrum of nitrogen-rich CVD- and synthetic diamond, measured by the constant photocurrent method (CPM).

The broad absorption band between 2 eV and 4 eV measured in the synthetic crystal is due to excitations of electrons from nitrogen donors into the conduction band. The absorption features from 4 eV up to the fundamental bandgap at 5.5 eV are also due to defects introduced by nitrogen atoms. The dip around 4.5 eV correlates with a maximum in optical absorption and can therefore be explained by generation of trapped carriers. The main difference to spectra of CVD grown diamond is found in the lower energy range. Photoconductivity starts around 1 eV with some structure around 2.4 eV and 3.3 eV. Electrons transfered to - states of sp2-bound carbon at grain boundaries or a broad energy distribution of 3-fold coordinated nitrogen are probable origins of this absorption band which is not found in single crystal diamond. Fig. 1: CPM-spectra of nitrogen rich diamond. ( CVD film, synthetic diamond type Ib).
*

phone: +49 (089) 289-12766, FAX: 289-12737, email: cnebel@physik.tu-muenchen.de

37

The excited states of the boron acceptors can be studied by photothermal ionization spectroscopy at energies below the photoionization threshold. Transitions of holes from the ground states of the acceptors to excited states, followed by thermal excitation into the valence band, are observed as a photocurrent. Above the photoionization threshold, the photocurrent shows periodic fluctuations where the minima are separated by the energy of the diamond LOphonons (Fig. 2).

Fig. 2: Extrinsic oscillatory photoconductivity in synthetic (type IIb) and CVD diamond. The modulations are due to free holes, rapidly captured by excited states of boron acceptors. If the energy difference between photogenerated holes deep in the valence band and the capturing states can be dissipated by emission of LO-phonons, the relaxation and capturing process by cascade emission is very fast compared to the relaxation to the top of the valence band by emission of acoustic phonons. Thus, the lifetime of those holes is short, giving rise to a minimum in the photocurrent. Up to 15 phonon replica can be observed in synthetic diamond (left side of Fig. 2). The energy separation of the periodic minima decreases slightly from 165 meV to 155 meV with photon energy because a photoexcited hole deep in the valence band has to lose more momentum during the relaxation than a hole located around the -point. The lifetime of holes, which cannot dissipate the excess energy by resonant LO-phonon emission, is larger, giving rise to maxima in the photocurrent response. On the right side of Fig. 2, the spectrum of the synthetic diamond is compared with a CVD grown film containing a large boron concentration. The synthetic diamond (dashed lines) shows two sets of minima due to two different capturing states. The minima are even more pronounced in the CVD diamond, indicating the higher quality of the film consisting of a thin (0.6 m) layer of boron doped (46 ppm) diamond grown on a highly oriented insulating CVD film (25 m). The positions of some lines are shifted in energy and a new set of minima, separated by approximately 182 meV, was found in the CVD film. ___________________________ supported by: Bundesministerium fr Bildung, Wissenschaft, Forschung und Technologie (03M2727 C7)

40

Spatially resolved techniques at low temperatures and high magnetic fields


Artur Zrenner1 , Thomas Huber and Markus Markmann Basic research in the field of modern semiconductor nanostructures and devices requires powerful experimental methods to get access to local optical and electrical properties on m and nm length scales. Furthermore it is desirable to have those methods available in the regime of low temperatures and high magnetic fields to avoid excess phonon scattering and to make use of magnetic quantisation phenomena and spin-related effects for advanced analysis. To meet those requirements we have developed a low-temperature microscope for both optical and electrical analysis, which is constructed for use in an Oxford He3-system (T350mK) with superconducting magnet (B15T). As shown in Fig. 1, optical access is via free space optics from an integrated optical minibench, which contains beamsplitters, polarizers, single-mode fiber couplers and a CCD-camera, to a remote controllable lowtemperature microscope, which is immersed into liquid He3. The low-temperature microscope has altogether six translation stages for in-situ manipulations. 3 translation stages are used to implement x-y-adjustment Fig 1: He3-System for spatially resolved magnetooptic of the specimen under experiments. The 6-axis low-temperature microscope can be investigation and z-adoperated with different scanning probes as shown above. justment of the objective
1

tel: +49/89/289-12772, fax: +49/89/3206620, e-mail: zrenner@wsi.tu-muenchen.de

41 lens. The remaining 3 translation stages are used to manipulate the position of an additional scanning probe in close proximity to both the species and the laser focus. Each of the translation stages is driven by precise piezoelectric stepper motors, which allow for an overall walking distance in the range of 2mm. With an NA=0.6 objective lens the spatial resolution of the optical system is below 1m. An image of the specimen under investigation is obtained via a CCD camera. Advanced spatial resolution in the nm-regime is obtained by introducing different scanning probes. As indicated in Fig. 1, there are conducting cantilevers or etched tungsten tips to perform STM-cathodoluminescence or to make local electric or mechanic contact to nanostructures. With cantilever-based near-field probes the microscope can be operated as SNOM. A photograph of the 6axis low-temperature microscope is shown in Fig. 2. The instrument is currently used for research in the field of single quantum dots, diffusive and electric-field-induced exciton transport, electric-field-induced exciton traps, surface emitting quantum dot lasers and transport in mesocopic systems. Part of this work can be found in the previous and current issue of the WSI report.

Fig. 2: Photograph of a 6-axis low temperature microscope with integrated tungsten STM-tip.

Supported by: DFG (ZR 2/1-1) and DFG (SFB 348)

42

Low Temperature Scanning Near-Field Optical Spectroscopy


Christian Obermller1 , Andreas Deisenrieder, Khaled Karra2 , Gerhard Abstreiter In scanning near-field optical microscopy (SNOM), a subwavelength light source is scanned over the sample surface while the relevant optical information is mapped as a function of the probe position. In this project we use subwavelength sized light probes consisting of tapered optical fiber tips. Such a tip is metal coated in a way that only an aperture is left open at its apex (Fig. 1). A small portion of light from the single mode fiber is transmitted through the circular aperture. As a typical result the transmission of tips with 60 nm to 250 nm aperture ranges between 1 nW and 1 W. Such high levels of photon throughput allow for imaging as well as for spectroscopic measurements with subwavelength resolution.

Fig.1 ( ): SEM of an aluminum coated tapered optical fiber tip. The aperture is about /10 of the 633 nm HeNe laser line. Fig. 2 ( ): Setup of the scanning near-field optical microscope for PL-spectroscopy on opaque samples at cryogenic temperatures. The near-field region away from the tips is of the order of the aperture diameter and its intensity strongly decreases with distance as discussed already in earlier WSI reports. Consequently it is necessary to control the distance between the tip and the sample, and maintain it constant to nanometer levels. Therefore we have developed a non-optical feedback, based on shear force detection by using piezoelectric tuning forks. With this basis we have constructed a room temperature SNOM, capable of measuring simultaneously topography, reflection and transmission as well as spectroscopy. The following room temperature experiments have been performed: Transmission microscopy of microcrystalline/amorphous silicon films. PL-spectroscopy of siloxene microcrystallites. Fluorescence microscopy of human chromosomes.

1 2

Tel: ++49-89-289 12778, FAX ++49-89-3206620, e-mail: Christian.Obermueller@wsi.tu-muenchen.de now at: Ludwig-Maximilians-Universitt Mnchen.

43

In order to perform PL spectroscopy of semiconductor quantum structures we have built up a SNOM, which allows work at cryogenic temperatures (4.2 K). The optical setup (Fig.2) is optimized for high efficiency light collection in reflection mode, which allows spectroscopy even of opaque samples. Therefore the light response of the local excited sample is collected with an ellipsoidal mirror. An optical fiber bundle guides the emitted light to a spectrometer which is equipped with a LN-cooled CCD multichannel detector. For coarse positioning and scanning of the sample, two different piezo-actuator based xyz-units with the range of 2 mm and 1.4 m in each direction are used. Furthermore an image control of the sample position is possible by free space optics. Also the nonoptical feedback and the coating of the tips was improved to work reliably at low temperatures.
energy [eV] 1,75 1,7 1,65 1,6 1,55 1,5 30 QW 6 nm QW 10 nm

PL-intensity [a.u.]

20

tip position relative to double QW: right left

10 Al.15Ga.85As

0 700 750 800 wavelength [nm] 850

Fig. 3: Spectra of the test sample at two different tip positions: left and right of the double quantum well.

Fig.4: Position dependent PL intensity of the QWs and the AlGaAs barrier (top). Band structure of the test sample (bottom).

An MBE-grown GaAs double quantum well ( 6 nm and 10 nm) separated by a 20 nm wide barrier served as a test sample (Fig. 4 bottom). The near-field probe illuminated the cleaved plane of the sample with laser light (458nm, 10nW). PL spectra were collected for two different tip positions (Fig. 3). The spectral lines of the two QWs and the low AlGaAs barrier were identified. Their intensities change drastically with probe position. In Fig. 4 (top) the integrated intensities of these three PL-lines are plotted versus tip position in the growth direction relative to the sample edge. The intensities of the individual peaks depend on the position of the tip (excitation) and the diffusion of excitons. As the tip moves from the Al.15Ga.85As to the GaAs range, which are separeted by a thin high barrier, a strong decrease of the Al.15Ga.85As line can be observed within a length scale of only 90 nm (for 10%-90%). This drop-off length determines the spatial optical resolution of the probe. It demonstrates that spectroscopy at 4.2 K on opaque samples with high signal to noise ratio and spatial resolution far beyond the theoretical diffraction limit is achieved.

supported by DFG (SFB 348)

44

Molecular beam epitaxy of ultrapure GaAs for cleaved edge overgrowth


Werner Wegscheider*, Max Bichler and Robert Neumann Molecular beam epitaxial (MBE) growth on the cleavage face of a previously prepared GaAs/AlGaAs multilayer has recently attracted increasing attention because of its potential for the fabrication of nearly perfect 1D and 0D quantum confined structures suitable for optical and transport investigations. While quantum wells (QWs) and two-dimensional electron systems (2DESs) exhibiting narrow linewidths and low-temperature mobilities well above 106 cm2/Vs, respectively, are routinely achieved on (001) oriented substrates, the realisation of corresponding structures on the natural GaAs(110) cleavage surface is hampered by poor surface morphology of the layers, caused by insufficient sticking of the As atoms on this nonpolar surface under conventional growth conditions. One possibility to circumvent this problem is a drastic reduction of the growth temperature to below 500oC. This leads in combination with an increase of the As4 partial pressure by a factor of 4 compared to (001) crystal growth to smooth and planar interfaces and surfaces. These unusual growth conditions, however, demand extreme precautions with respect to the cleanliness of the MBE system. This is visualised in Fig. 1, where the evolution of the low-temperature electron mobilities in a 2DES at the GaAs/Al0.3Ga0.7As interface for (110) and (001) growth is compared. After roughly one Fig. 1: Evolution of the electron mobility in modultion-doped single interface GaAs/AlGaAs structures oriented along the (001) and (110) directions with time after the MBE growth chamber has been exposed to atmospheric pressure.

year of growth without breaking the vacuum in the growth chamber, the peak mobility in (001) oriented 2DESs reaches 11x106 cm2/Vs while the corresponding value for growth on the (110) cleavage face is 1.7x106 cm2/Vs. Both values are among the highest ever reported in modulation-doped high-mobility structures. From this comparison and a detailed analysis of the involved scattering mechanisms it can be concluded that the electron mobility for (110) oriented 2DESs is very sensitive to the incorporation of impurities during crystal growth. It
*

tel.: ++49/89/2891-2776; fax: ++49/89/3206620; e-mail: wegscheider@wsi.tu-muenchen.de

45 should be noted at this point that even (001) electron mobilities as high as 2-3x106 cm2/Vs, which represent the upper limit for most MBE systems, can be achieved at background impurity levels which are still inadequate for high-quality (110) growth and lead therefore to mobilites as low as 105 cm2/Vs. The limiting scattering mechanism for the highest (110) electron mobility samples was, however, identified as remote ionised impurity scattering. In contrast to (001) crystal growth, Si is incorporated on As as well as on Ga sites as a result of the amphoteric nature of this dopant material. Consequently the amount of Si to achieve a given electron density at the (110) GaAs/Al0.3Ga0.7As interface exceeds that for the (001) interface by more than a factor of 5. For the preparation of modulation-doped quantum wires the charge carriers can, however, be completely electrostatically induced by gates. In this way the electron mobilities for transport along these 1D structures should be well above 106 cm2/Vs, values which correspond to elastic mean free paths of more than 10 m and which make ballistic transport experiments feasible. In addition to the cleaved edge overgrowth projects which rely mainly on high-quality growth on the (110) GaAs crystal face, several experimental groups within and outside the WSI are supported with conventional 2DESs. One example represents the investigation of composite fermions. For this purpose ultrahigh-mobility 2DESs with relatively low sheet carrier densities are required. The quantum Hall and Shubnikov-de Haas spectra of such a sample are depicted in Fig. 2. Around filling factor =1/2 a large number of fractional quantum Hall states can be identified. The occurence of these states up to =7/15 and 8/15 at magnetic fields below 15 T directly demonstrates the high-quality of these samples.

Fig. 2: Shubnikov-de Haas and quantum Hall effect measured in a ultra high-mobility 2DES. The sample was provided to Dr. J. Smet at the MPI fr Festkrperforschung (Stuttgart).
*

supported by: DFG (SFB 348), BMBF (01 BM 630)

46

MBE Growth of Metamorphic In(Ga)AlAs Buffers


Markus Sexl* eimann, Gerhard Abstreiter

Metamorphic buffer layers allow the arbitrary combination of active semiconductor structures with substrates having different lattice constants. These buffers are essential to accommodate the lattice mismatch between the active layers and the substrate in a controllable way, to generate a new substrate with desired lattice constant and moderate dislocation densities. Modulationdoped InAlAs/InGaAs-heterostructures, with their inherently high electron densities and mobilities, are the ideal basis for ultrafast, low noise transistors. Their superior device performance can be combined with the advantages of lar ge, less expensive substrates by growth on GaAs-substrates. The properties of the active layers, e.g. the morphology and the electron mobility in modulation doped structures depend on the composition and growth conditions of the buffer layer. For a preliminary characterization of the buffer quality, we grew single quantum wells with different well widths metamorphically with two dif ferent types of buf fers and for comparison lattice-matched on InP-substrates. The photoluminescence spectra 4 3 1.5 nm 32 15 7 of these structures are shown in Fig. 1. A very simple metamorphic buffer is the growth of one micron directly on the GaAs-substrate (one step buffer). The PL yield of the quantum wells on top of this buf fer is very low compared to the latticematched (LM) structure on InP. A much better result is achieved by using a quarternary InGaAlAs Fig. 1: Single quantum well structures on metamorphic buffer with linear grading of the buffers, photoluminescence at 4.2 K. In-content (linear graded buffer).

As linear grading gives the best results, such buffers are studied in more detail. A linear grading of InAlAs up to x In = 0.52, grown at 420 C, results in a moderate cross hatching with 3 m spacing and mean surface roughness of 5 nm. The surface morphology was improved by using quarternary InGaAlAs buffers. The In-content is increased linearly, the ratio of Ga and Al concentration is adjusted to a constant bandgap of 1.48 eV. Since Ga-adatoms are more mobile than Al-adatoms on the MBE growth surface, lower growth temperatures can be used. W e obtained smoothest morphology using a linear temperature ramp from 350 C at the beginning of the buffer to 390 C at the end. The relaxation behavior of the metamorphic buf fers are studied by high resolution X-ray diffraction (HRXRD). Active relaxed. The degree of relaxation is 86% for the ternary buffer and 90% for the quarternary buffer, respectively. Furthermore, the ternary buffer is tilted by 0.56 in (011)-direction. No tilt is detected in the quarternary buffer. Grading the metamorphic buffer to an In-content of 0.52 (the value of the active layers) resulted in a residual strain.

phone: +49 89 2891 2788, fax: +49 89 3206 620, email: mas@e26.physik.tu-muenchen.de

47 X-ray in (011)-direction (224)-reection (004)-reection X-ray in (01-1)-direction (004)-reection

qx [ -1] qx [ -1] qx [ -1] Fig. 2: Reciprocal space mapping of quarternary InGaAlAs buffer with xIn up to 0.63. Increasing the nal In-composition of the buf fer up to a value of 0.63, the In 0.52Al0.48As layers on top of the metamorphic buffer are unstrained. In Fig. 2 the reciprocal space map of this structure is shown. T wo symmetric and one asymmetric reections indicate no macroscopic tilt of the layers. The metamorphic buffer is fully relaxed up to an In-content of 0.52, that is the lattice Fig. 3: High mobility structures on metamorphic buffconstant of InP; the part with the ers, mobility and mean surface roughness. overshoot in the In-content is pseudomorph. A reduction of the In-content to the desired value of 0.52 for the active layers results in complete relaxation. Interface roughness and its reduction by using dif ferent buffer layer structures were studied with respect to transport properties in 2DEG-structures with low carrier density, where interface scattering dominates. Heterostructures with single-sided homogeneous Si-doping and thick spacers of 15 nm were used; the channel width was 32 nm. The obtained mobilities of otherwise identical structures on the metamorphic buf fers discussed above are compared at three different temperatures in Fig. 3. The metamorphic buf fer with the overshoot in In-content has the best mobility, approaching the lattice matched reference on InP-substrate. The carrier density remains unchanged. Only a small increase in surface roughness is observed by growing the additional overshoot in In-content. The observed dif ferences in mobility between metamorphic and lattice-matched structures show the ef fect of scattering by interface roughness and scattering at mist dislocations in incompletely relaxed active layers. Supported by: Siemens AG via SFE and the BMBF

qz [

-1]

48

Self-assembled growth of Ge on Si
Peter Schittenhelm*, Frank Findeis, Cornelia Engel, Edilson Silveira, Hubert Riedl, Gerhard Abstreiter The strain-driven self-assembled formation of Ge-nanostructures on Si during MBE growth in the Stranski-Krastanow mode has recently attracted a lot of interest. One promising aspect is the in-situ formation of quasi-zero-dimensional structures without the need of sophisticated technology. The second important aspect is the rather low critical thickness of pseudomorphic SiGe heterostructures, that has turned out to be one of the most severe restrictions on the road to many of the proposed device applications. Several concepts, like e.g. relaxed graded buffers, had been tried in the past to overcome this limitation. The Stranski-Krastanow growth mode, which leads to the self-assembled formation of islands in strained systems, allows to exceed the critical thickness without introducing dislocations. We are able to control the size and thus the optical and electrical properties of the islands by an appropriate choice of the parameters growth temperature, flux, composition and thickness of the deposited layer. This enables us to achieve Si/Ge heterostructures with effective valence-band offsets as large as 420meV.

Fig. 1: I-V characteristic of an Esaki tunneling diode with embedded Ge dots and a Si reference structure. First attempts have been made to realize a number of device concepts based on this large valence band offset and the related small bandgap, such as pin-diodes and npn-transistors for IRdetection, or carrier storage. Besides these optoelectronic applications, we have also tried to fabricate silicon-based tunneling diodes. For this purpose, Ge dots have been embedded in the depletion zone of a heavily doped pn-Esaki diode. Due to the reduced bandgap in the dots, the tunneling probability is expected to increase strongly as compared to a pure silicon structure. Fig. 1 shows the I-V-charateristics of such a diode, together with a silicon reference structure. As expected, the tunneling current in the diode containing the Ge dots is increased by several
*

tel.: ++49-89-2891 2776, fax: ++49-89-3206 620, e-mail: schittenhelm@wsi.tu-muenchen.de

49 orders of magnitude: However, also the excess current is increased stronger then expected, probably due to the creation of additional defects. Therefore, no clear negative differential resistance is observed.

Fig. 2: TEM image of a 19 period Si/Ge-dots superlattice. A second problem connected with the self-assembled dots is their rather broad size distribution, which results in a large variation in the energy of the electronic states in an ensemble of dots. To increase the size homogeneity, Si/Ge-dot/Si superlattices have been grown. In these structures, the spatially modulated strain field of the buried dots leads to a more regular arrangement of the dots in the subsequent layer, resulting in a short-range lateral correlation and a significantly increased size homogeneity of the dots. Fig. 2 shows a TEM image, revealing the strong vertical correlation of the stacked dots. Reciprocal space maps received from XRD show not only the satellites of the vertical Si/Ge/Si superlattice, but also lateral satellites next to the superlattice peaks. The latter have been attributed to a regular arrangement of the dots in the topmost layers. The TEM and X-ray analysis has been performed in collaboration with A.O. Kosogov and P. Werner (Max-Planck-Institut fr Mikrostrukturphysik, Halle) and the group of Prof. Bauer (Univ. Linz), respectively.

Supported by: SIEMENS AG, BMBF

50

Focused Laser Beam Induced Patterning of Heterostructures


Peter Baumgartner* , Matthias Herbst, Cornelia Engel, Gerhard Abstreiter Novel device concepts are often based on ultrasmall lateral dimensions to exploit Coulomb-blockade or quantum mechanical effects. Our recently developed fabrication technique, the local doping by heating with a focused laser beam (FLB), combines a high spatial resolution with large potential modulations and low damage of the starting material. The achievable resolution of this direct optical writing method is not limited by the optical diffraction, due to nonlinear effects during the diffusion of the dopants. The technique is therefore ideally suited to fabricate such advanced nanostructures. The fabrication starts from n-modulation doped GaAs/AlGaAs heterostructures containing a high mobility two-dimensional electron gas (2DEG). Insulating regions are fabricated by depleting the 2DEG by local p-doping. Regions of the 2DEG, which are insulated from the active area of the device by p-doped regions, can be used as in-plane gates to control the nanostructures. It is also possible to fabricate lateral bipolar-structures such as npn-transistors or pn-diodes. The high potential modulation of lateral npn-structures, combined with a small width of the p-region, results in high lateral electric fields. These structures may be used as optical modulators, since the lateral fields are tunable by a bias voltage. We have performed some wavelength dependent photocurrent measurements with excitation energies below the bandgap of the GaAs quantum-well. A high electric field makes the absorption of a photon with an energy below the band gap possible, due to the Franz-Keldysh effect. We have estimated the lateral electric field in these structures, by analyzing the wavelength-dependent photocurrent measurements. The inset of Fig. 1 shows the wavelength-dependent photocurrent data for different bias voltages (solid lines) together with theoretical curves resulting from the FranzKeldysh theory (dashed lines). The only free parameter in the model is the electric field. It is therefore possible to estimate the electric field as a function of bias voltage by fitting the theoretical curves to the experimental data. The resulting field as a function of the bias voltage is shown in Fig. 1. There is a built-in electric field of about 120 kV/cm. The field ascends with increasing voltage and reaches about 260 kV/cm at a bias voltage of 2V. It is possible to estimate a width of the depleted region from the slope of the field-voltage dependency, which results in a value below 200nm. The electric field saturates for higher voltages. This saturation is caused by leakage currents in the surface layers of the structure, which cause an accumulation of surface charges. Fig. 1: Electric field as a function of bias These surface charges screen the applied voltvoltage for a lateral npn-structure. The inset age and the field saturates. The effect also limits shows wavelength-dependent photocurrent the useful maximum gate-voltage of in-plane measurements. gated devices. The achievable fields are rather high, com*

tel.: ++49-89-289 12777, fax: ++49-89-3206 620, e-mail:baumgartner@wsi.tu-muenchen.de

51 parable to vertical modulator structures. This shows, that it is possible to fabricate structures with a very high potential modulation and a small structure size with the FLB-technique. The inset of Fig. 2 shows the devicegeometry of a single electron transistor, fabricated by the FLB-technique. A small dot, a source and drain reservoir, and in-plane-gates are all built from the two-dimensional electron gas of the heterostructure. Laser-written p-doped lines are used to define a dot of about 70nm diameter and to insulate it from the in-plane-gates. Tunnel junctions connect the dot with source and drain. The in-plane-gates are used to tune the tunnel junctions and to change the electrostatic potential of the dot. Fig. 2 shows the conductance of the device as a function of the center-gate voltage for Fig. 2: Conductance as a function of gatetwo temperatures, T=1.5K and T=4.2K. Clear voltage of a single electron transistor. The oscillations are observed for conductances beinset shows the device geometry low about <e2/h, which result from the Coulomb-blockade effect in the small dot. The period of the oscillations of about U=150mV is consistent with a center-gate capacitance of C=1.1aF. This value is about one order of magnitude smaller than gate capacitances typically observed in split-gate devices. A high charging energy, about 5meV, of the dot is achieved due to the low coupling capacitance to the gate-regions in this IPG-device. The high charging energy results in a rather high maximum working temperature of the device, up to about 7K. First results on quantum interference devices, like Aharonov-Bohm rings, are very promising. It is possible to shift the phase-difference of two interfering wavefunctions by applying magnetic or electric fields. This results in an oscillating current through the device. The experimental results show, that a phase coherence length in the micrometer range is achieved in FLB-fabricated electron channels. The inset of Fig. 3 shows schematically the device geometry of a lateral pn Esaki tunnel-diode. A sharp transition between a modulation doped 2DEG and a two-dimensional hole gas is achieved by pdoping of a rectangular region on an n-type modulation doped InGaAs quantum well with the FLBtechnique. The n- and p- regions are contacted by GeAuNiAu and BeAu contacts. The current-voltage characteristics of the device is shown in Fig. 3. There is a strong negative differential resistance from about U=0.325V to U=0.4V. The peak to valley ratio of the Fig. 3: Current-voltage characteristics device reaches 2.9 at room temperature. of a lateral Esaki tunnel-diode. The inset shows schematically the device. Supported by: DFG (SFB348)

52

Transient-Thermal-Grating-Processing of III-V Semiconductors


Michael K. Kelly*, Christoph E. Nebel, Joseph Rogg, Werner Wegscheider, Gerhard Bhm, Oliver Ambacher, Martin Stutzmann We are investigating and developing the ability to do high-resolution structuring in III-V semiconductor systems using interference patterns from a high-power pulsed laser, as a project of the Sonderforschungsbereich 348 - "Nanometer Semiconductor Devices". The interference of two or three coherent beams at a material surface produces a sinusoidal intensity modulation in one or two dimensions, respectively. The energy absorbed from this intensity distribution produces a thermal grating in the semiconductor that can induce permanent material modifications with the same modulation period. Because the thermal energy is deposited in a very short time by the pulsed laser (less than 10 ns), higher spatial resolution than usual for thermal treatment is possible because of reduced heat diffusion in the solid. After the pulse the cooling is also rapid, typically on a scale of tens of nanoseconds. This can be useful for "freezing in" nonequilibrium structures, as well as minimizing some Fig. 1. AFM profile of an interference structured GaN types of damage, like arsenic loss in GaAs. film. We have produced interference gratings with periods as small as 130 nm by this technique. As an illustration of the patterning, an atomic force microscopic measurement is shown in Fig. 1 of a GaN surface with a 250 nm period. In the case of GaN, we have found that thermal decomposition with nitrogen effusion can be induced with a pulsed laser, which provides a flexible means of structuring this increasingly technologically relevant material. We have been able to structure two-dimensional electron gases in GaAs/AlGaAs heterostructures, introducing lateral potential barriers to the electrons that are already confined to the 2-D channel. This is shown schematically in Fig. 2,

Fig. 2. Schematic diagram of the laterally structured 2D-electron gas.

* Tel:

+49(89)289-12765, Fax: +49(89)289-12737, e-mail: mkelly@physik.tu-muenchen.de

53

where the shaded regions close to the surface of the MBE-grown system have been heated by a single pulse with the interference pattern. Our measurements based on photoluminescence and transport indicate that these processed regions then accumulate electrons from the Si-doped AlGaAs, and the resulting potential depletes the 2DEG at these locations and creates barriers. When these barriers have a striped pattern, resulting from the interference of two beams, the conductivity perpendicular to these lines is dramatically decreased, while the conductivity parallel to the wire-like channels remains comparable to the initial high values of the 2DEG. The results of a temperature-dependent measurement of this anisotropic conductivity is shown in Fig. 3 for a 380 nm-period grating. Magnetotransport measurements are being used to characterize these systems and investigate the onset of quantum confinement effects as the dimensions are made smaller. Three interfering beams can be similarly used in order to produce a laser intensity modulation in two dimensions at the sample surface. The surface of such a sample, characterized with an atomic force microscope, is shown in Fig. 4. In this way, lattices of dots and antidots have been created in the 2DEG. In the case of dots, the electrons are then confined in all three dimensions, which can result in desirable optical characteristics. An antidot lat- Fig. 3. Temperature-dependent anisotropic conductivity of a tice provides a periodic potential bar- 2DEG structured with a 380 nm-period line grating. rier for the electrons, but leaves the electron gas connected so transport remains possible. A magnetic field perpendicular to the 2DEG encourages curved electron trajectories during such transport, which can be frustrated by the antidot lattice. At field strengths where the curvature radius corresponds to stable orbits in this lattice, commensurability maxima in the magnetoresistance can then be observed. Fig. 5 illustrates such a result that was observed in this way.
1.0x10
-3

8.0x10

-4

B = 0.095 T 0.181 T 0.35 T

Resistivity (a.u.)

6.0x10

-4

4.0x10

-4

antidot array
2.0x10
-4

unpatterned
0.0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

B(T)

Fig. 5. Magnetoresistivity from an antidot lattice sample. Fig. 4. AFM image of the surface of a 2DEG sample structured with a 3-beam interference pattern to give an antidot lattice.

Supported by the Deutsche Forschungsgemeinschaft, SFB 348.

54

Nanopatterning of Silicon by Laser Processing


Gerhard Groos* , Martin Stutzmann Today, silicon structures with sizes down to about 100 nm are mainly produced by expensive processes like e-beam lithography and dry etching. We investigate a simpler method based on direct patterning of amorphous silicon (a-Si) films by pulsed laser annealing. With this method we can generate homogeneous arrays of stripes, grids or dots of microcrystalline silicon (cSi) in one or two parallel processing steps. Periods down to 210 nm and linewidths of less than 100 nm could be demonstrated. Because the process is also suited for transparent substrates such as fused silica or glass, the resulting arrays are easily accessible to optical characterization. The structures are defined by the interference pattern of a frequency-doubled or -tripled Nd:YAG laser (532 nm or 355 nm respectively, pulse duration 8 ns), which modifies the a-Sifilm at the regions of highest intensity. This kind of process is very fast and robust against perturbations like mechanical vibrations, because the whole pattern is generated during a single laser pulse of 8 ns duration.In contrast to a typical laser-recrystallization process we use higher intensities, which leads to film ablation at the hot regions and dynamically limited contraction at the colder regions of the interference pattern. This enables us to produce c-Si patterns in one step, without further etching. Widths down to 1/5 of the period can be realized for periods from 300 to 1700 nm (Fig. 1).

10

et ched pi led up

10

w = 0.2 0 p w = 0.1 5 p 10 period p [ nm]


3

10

Fig. 1: Structure widths after recrystallization and selective etch (circles) or pile-up by high intensity processing (diamonds).

Fig. 2: Atomic force micrograph of 477 nmperiod c-Si stripes after a multi-line annealing (MLA) step.

To increase the crystallinity in the stripes, we apply a second annealing step using a similar line pattern, oriented perpendicular to the stripes. With a pulse-repetition rate of 10 Hz, these lines are scanned along the c-Si stripes, so that consecutive pulses overlap and create moving melting zones. Fig. 2 shows a stripe pattern after this multiline annealing (MLA) step. The main result is a drastic increase in the aspect ratio (height/width), in some samples up to 0.85,
*

phone: (+49) 89 289-12769, fax: -12737, e-mail: Groos@WSI.TU-Muenchen.de

55 a value characteristic for almost hemispherical wire profiles. Detailed analysis revealed that not much material is lost during these processing steps, but that larger flat regions contract (due to surface tension) to form the piled-up structures observed after MLA. After MLA the Si wires are up to a factor of 6.5 higher than the initial film thickness. Fig. 3 shows the effect of the MLA step on the position of the first-order silicon Raman signal, which occurs at 520 cm-1 in crystalline silicon (c-Si) and is affected by mechanical stress. With increasing intensity of the MLA step, the Raman peak shifts towards the position in c-Si, indicating that the mechanical stress is reduced by this processing step. The increasing intensity of the c-Si Raman peak and absorption measurements (not displayed) show the increasing crystalline fraction compared to that of the initial structure.
52 0 . 8
518 5h 900C 5h 850C not oxidized c-Si wafer

517

1 .66 m peri od
516

5 19. 7

2 Int ensit y of MLA [ mJ/ cm ]

50

100

150

200

250

530

520

510

wavenumber [ cm ]
Fig. 4 The c-Si Raman peak position suggests tensile stress without and compressive stress with oxidation at different Temperatures.

-1

500

Fig. 3: Position of the c-Si Raman peak with increasing annealing intensity, indicating reduction of mechanical stress.

To reduce the lateral size of the c-Si stripes, an additional oxidation step can be used: The oxidation of c-Si-columns by annealing in an O2 atmosphere is a self-limiting process due to the strain induced by the larger volume of the oxide. Fig. 4 shows this effect in the Raman spectra of stripes annealed during 5 h in O2 at different temperatures. The conditions were chosen to reach the self-limiting regime and to oxidize the initial film totally. The Raman signal of the stripes prior to oxidation is shifted to lower wavenumbers compared to c-Si. This is due to tensile strain building up during the laser processing, because of the lower thermal expansion of the quartz substrate compared to the Si structures. In contrast, the oxidized samples are under compressive strain, as indicated by a Raman signal shifted to higher wavenumbers. This is caused by the above-mentioned larger oxide volume leading to the selflimitation of the process. Our preliminary results suggest, that laser-interference annealing is a fast and simple patterning method and suitable for silicon structures smaller than 100 nm. Future research efforts will include electrical characterization and will focus on further decreasing the structure size until effects like bandgap variation or coulomb blockade can be seen.

CBE of integrated 1.55 m lasers on exact and misoriented (100) - InP substrates
Andreas Nutsch* , Heinrich Kratzer, Bahram T #, Gerhard Abstreiter Weimann
+

Chemical beam epitaxy is a well suited technology for monolithic inte gration of dif ferent photonic devices. Area selecti ve gro wth occurs only on unmask ed InP surf aces. After removing the silicon nitride mask, grooves are etched into the substrate. The laser structure is lled in these groo ves. The use of a blocking layer with a quaternary etch stop layer allo ws chemical etching with the precision of the Chemical Beam Epitaxy . The blocking layer also avoids short circuits in the laser structure.

0,75

height [ m ]

1.3

(100)

<- [01-1]

0,00
(100) misoriented towards [110]

-0,75 0 2,5 5,0 7,5 10

11 m m

surface distance [ m ]
Fig. 1: left: AFM cross section indicating a tilted (100) growth surface of a laser on a misoriented substrate (2 off towards [110] = 1.3 in the (110) plane). right: SEM cross section of a selective laser on an exactly oriented (100) InP substrate. The laser structure shows a very good symmetry. The typical threshold current for a 200 m long and 3 m wide laser is 11 mA.

Our laser structure consists of six compressi vely strained GaInAsP QWs, quaternary barriers ( = 1.2 m) and connement layers ( = 1.1 m). Typical photoluminescence linewidths of the laser structures gro wn on (100) InP epi - ready substrates misoriented 2 to wards the nearest [110] direction are 23 meV at 300 K and 6.4 meV at 4.2 K. For 6 QW, the threshold current density is 410 Acm-2 and the internal losses are 11 cm-1 (broad area laser). F or selectively gro wn lasers on misoriented substrates, the PL line widths are 24.5 meV and 5.6 meV at 300 and 4.2 K respectively. The threshold current density is 440 Acm-2; the internal losses are 14 cm-1. Growing the same laser structure on e xactly oriented (100) InP substrates (max. deviation 0.2) yields a PL linewidth of 23 meV at 300 K and 3.6 meV at 4.2 K. The threshold current density is 420 Acm-2, the internal losses are somewhat higher (26 cm1 ). The higher losses are due to higher mist dislocation densities on e xactly oriented substrates. On the other hand we obtain a v ery good symmetry of the selecti vely grown laser
* Phone: 49 89 2891 2787, Fax.: 49 89 3206620, email: arn@e26.physik.tu-muenchen.de + now at: Ferdinand - Braun - Institut, Rudower Chaussee 5, D - 12489 Berlin #

now at:

structure on such exactly oriented substrates (see right part of Figure 1). The laser performance for our selectively grown structures is equal to that on epi-ready substrates. No change in composition of the quaternary core appeared for lasers in 7 m and 30 m wide grooves, when compared to lasers on planar substrates. Quantum wells grown in narrow 4 m wide grooves on misoriented substrates, on the other hand showed wavelength shifts depending on substrate misorientation. A 10 nm redshift is observed on (100) substrates oriented 2 off towards the next [110] direction, indicating a change in material composition, whereas we observe a slight blueshift on (100) exact oriented substrates.

[011] [011]

1 m 0

1 m

Fig. 2: InP: Be surface of selectively grown lasers. The left side shows the smooth surface of the laser in a 4 m wide groove. On the right side the surface of a large area is shown. The surface steps are perpendicular to the [011] direction and to the [010] direction, which is the misorientation of the substrate.

AFM measurements show that the growth surface is tilted by 1.3 with respect to the misoriented substrate in a (011) plane; the angle corresponds to the 2 misorientation of the substrate. Figure 1 shows this misorientation as measured by AFM. For comparison the right part of Figure 1 shows the SEM cross section ((011) plane) of a laser on an exactly oriented substrate. Step densities on InP: Be contact ridges of selectively grown lasers are shown in Figure 2. With decreasing misorientation of the substrate, i.e. lower surface step density, the gallium and the phosphorus incorporation of GaInAsP ( = 1.05 m) were found to decrease. A reduced surface step density in narrow grooves results in InAs rich layers. This is in good agreement with the energy shift for narrow grooves during selective growth on 2 misoriented substrates. Investigations on wide area selective growth on (100) oriented substrates support this result. The change in compositions occurs when the width of the groove is comparable to the migration length of the adsorbed metalorganic species, resulting in a reconstruction of the (100) plane in the narrow groove. The redshift is not observed on exactly oriented substrates. The presented results show the capability of Chemical Beam Epitaxy for monolithic integration of optoelectronic devices. Optimized growth parameters of planar grown laser structures allow the selective growth of high quality lasers. Supported by BMBF Photonik II

64 Monolithic Integration of Lasers by Selective Chemical Beam Epitaxy


* +

eimann#,

Gerhard Abstreiter

One of the rst devices using monolithic integration will be a monolithic integrated transceiver (transmitter plus receiver) for optical communication (suggested by Heinrich Hertz Institut, Berlin), integrating different optical components such as a laserdiode, waveguide and a photodetector onto one chip. Fabrication of such a device requires selective growth either of the laserstructure or the waveguides and detector. In contrast to growth by MOCVD, selective area epitaxy (SAE) with chemical beam epitaxy (CBE) allows the growth of very homogenous and at layers with a small (about 0.5 m) disturbed area at the edge of the dielectric mask, e.g. a Si3N4 mask. Details and characteristics of such structures can be seen in our previous reports and other sections of this report (B. Torabi, A. Nutsch). We investigated some 'key technologies' for fabrication of photonic devices by SAE. One is the growth of thermally stable layers to allow a regrowth at high temperatures with a high optical quality. The other is the control of the shape at the edge of the mask to enable the integration of laserdiodes with 'state of the art' performance and butt-joined laser and waveguides with a high coupling efciency as shown in the report by B.Torabi. To test the thermal stability we have grown quaternary laserstructures covered with a 100 nm InP:Be cap layer by CBE at a growth temperature of 540 oC. The stability of our quaternary layers was tested by a 2nd growth of 1.5 m InP:Zn with MOVPE at a substrate temperature of 600oC in cooperation with Siemens AG (Fig. 1, Test 1). No thermal degradation could be observed; compared to CBE-grown lasers there is no signicant dif ference in XRD- (Fig. 2) and photoluminescence spectra. Threshold current densities, optical losses and internal quanFig. 2: XRD spectra of

2nd Epitaxy

1,5 m InP:Be SiN - Mask 100 nm InP:Be SCH 1 SCH 2 + QW SCH 1 InP:Si

resp. active layers grown by CBE and pcontact regrown by MOVPE.

Intensity [a.u.]

CBE grown lasers

MOVPE

CBE

1st Epitaxy

2 [Grad] Fig. 3: CL spectra of lasers with selective grown pcontact. Luminescence was detected in the middle of each laser. Broad lasers show luminescence comparable to conventional lasers, the small one a degradation of luminescence. Energy [eV] CL - Intensity [a.u.]

Test 1

Test 2

Fig. 1: Structures used for probing the thermal stability of our quaternary laserstructure. One structure was regrown by MOVPE at 600oC, the other one was masked with SiN and regrown with CBE at 540oC.

* phone: + #

+49 89 289 12786, fax: +49 89 3206 620, email: hjk@e26.physik.tu-muenchen.de

65 tum efciency of laserdiodes remain unchanged compared to all CBE grown lasers. In a second run before 2nd growth, we determined stripe apertures of 100 nm thick SiN (Fig. 1, Test 2). 1.5 m InP:Be and a thin GaInAs:Be contact layer were selective grown by CBE at a temperature of 540 oC and a low growth rate of about 0.4 m/h. Spatially resolved CL measurements on structures below 10 m show a degradation through thermal stress: a redshift of about 10 meV and a weak luminescence (Fig. 3). Laserdiodes show no or only a poor performance compared to conventional lasers. The SiN mask seems to induce some thermal degradation of the quaternary layers.
4a 4b <---- [01-1] 4c

Fig. 4:

Crossection of selectively lled-in lasers. Growth was performed on 2" InP wafers oriented (100) 2o off towards next (110)). The grooves were 1.5 m deep. Fig. 4a: Fig. 4b: Fig. 4c: no lateral undercut, groove dened by ECR/RIE with CH4/H2/N2 at high bias. lateral undercut of about 200 nm, groove dened by ECR/RIE with CH4/H2/N2 at low bias. lateral undercut of about 700 nm, groove etched selective by HCl etchant (wet chemical).

Due to the off-axis conguration of sources in our CBE equipment, we observe a high lateral growth rate at sidewalls of grooves during selective growth. In grooves with perfect rectangular shape, this lateral growth induces a highly disturbed region at the edge of the mask forming an electrical short-circuit (Fig. 4a). Etching a lateral undercut under the SiN mask of about 200 nm, we observe lateral growth but no electrical short at the edge of the mask (Fig.4b). The layers were almost at; the small disturbance at the left edge is induced by using an of f-oriented substrate. A very large undercut of about 700 nm (Fig. 4c) results in growth of an independent ridge in the middle of the groove without any contact to sidewalls. This new mode we call 'shadow mask' growth. For both undercuts, laser diodes show an extremely low threshold current, e.g 10 mA for a 4 m wide laser, and excellent performance. For denition of this undercut we used two different methods: ECR/RIE with CH4/H2/N2 at low bias and wet chemical by HCl-based etchant. By etching with ECR/RIE, the shape of the grooves is almost independent of orientation and material composition. The uniformity of the depth of grooves is about 15% over a 2" wafer . Addition of N 2 to the CH 4/H2 plasma blocks the formation of polymers at the sidewalls of grooves. Selective grown laserstructures show intensive and narrow photoluminescence and 'state of the art' device performance. Wet chemical etching requires a base structure with an etch stop layer of 50 nm (GaIn)(AsP) and 1.5 m InP. By selective etching with HCl-based etchants, we can dene grooves in [011] orientation with a perfect rectangular shape. The uniformity of etch depth is better than 3% over a complete 2" wafer , limited by uniformity of the base structure grown by CBE. Selectively grown layers show intensive and extremely narrow (5 meV) photoluminescence. The drawback of this method is the dependence of shape on the orientation of the groove; grooves in [01-1] direction show {1 11}-facets at the sidewall. Due to the high selectivity to InP , this method is not suitable to dene grooves in waveguides for butt-joining lasers and waveguides. Supported by BMBF (Photonik II)

66

Vertical cavity surface emitting lasers with current connement

Vertical cavity surface emitting lasers (VCSELs) are no longer just subjects of research and development: In 1996 they became commercially available as discrete devices. As the light emission is perpendicular to the wafer surface with small divergence and longitudinally single moded, VCSELs are well suited for integration in photonic circuits, e.g. two-dimensional arrays for parallel optical communications. Small arrays of 16 devices, each of them mounted to an optical ber are already used for short-range, high-speed optical data transfer. But there is still a lot of basic research to be performed on VCSELs. Due to their small dimensions, these devices can be seen as photonic quantum dots. Low threshold currents and series resistances are required for high integration densities. In our VCSEL-project we are using a conventional vertical n-contact via the n-doped substrate and the n-doped bottom distributed Bragg reector, providing large area n-contacts, while the p-contact is made laterally by a p-doped GaAs layer in the cavity , as the p-contact makes the major contribution to the series resistance. We restricted the current distribution to the center of the cavity, by inserting a current conning n-doped GaAs barrier into the p-cladding. The incorporation of this lateral current connement requires two epitaxial steps. In the initial MBE growth run the bottom n-doped DBR, the 1 10 nm thick n-AlGaAs cladding, the active MQW region with 3 strained InGaAs wells tuned for an emission wavelength of 980 nm, and a 100 nm thick fraction of the p-AlGaAs cladding are grown, followed by the 50 nm thick nGaAs blocking layer. The lasing region is dened laterally by opening apertures in this GaAs layer with selective reactive ion etching using CCl2F2. In the subsequent second MBE run, the p-GaAs contact layer and the top DBR were grown to complete the diode. The central regions of the VCSELs are, despite the low growth temperature of 540 oC, planar and extremely

(a)

(b)

Fig. 1: (a) CW light against current characteristics at room temperature for blocking layer apertures from 10 m to 5 m diameter. (b) Resonance frequency vs. square root of output power (8 m aperture diameter). *phone: +49 89 2891-2767, fax: +49 89 320 66 20, email: mjh@e26.physik.tu-muenchen.de

67 smooth. The optical power versus current characteristics for aperture diameters from 5 to 10 m is shown in Fig. 1(a) for VCSELs with lateral p-contact. The lowest threshold currents of 470 A are found for 6 m aperture diameter. VCSELs with 9 m blocking layer aperture diameter have threshold currents of 0.65 mA and a maximum output power of 2.7 mW. They show the best overall performance with an external quantum ef ciency of 34% and a power conversion efciency of 13 %. Operation voltage and series resistance are 1.6 V and 380 at threshold and 3.5 V and 1 10 at peak power , respectively. The lasers with 6 m aperture diameter operate in single-mode emission and stable linear polarization at all current levels. For larger apertures higher order transversal modes can be observed. Last year the rst linewidth and RIN measurements of our VCSELs were performed in collaboration with the group of Prof. Harth from the department of electrical engineering. In Fig. 1(b), the resonance frequency is plotted as a function of light power . The curve shows the expected linear behavior with a slope of 7.7 GHz/ mW, one of the best values reported for VCSEL so far. The resonance frequency increases at a rate of 4.4 GHz/ mA above threshold. The measured linewidth of 30 MHz is extremely small, and comparable to the best result reported so far, and shows the excellent quality of the grown VCSEL structures. VCSELs having two lateral intracavity contacts allow easier integration since both n - and p contact are on the same side of the wafer . Such laser diodes with current blocking layer are being investigated, with present threshold currents of 0.58 mA, a maximum output power of 2.3mW and series resistances of about 80 at peak power. With a newly designed set of masks, 8x8 VCSEL arrays have been fabricated. This was possible as the yield of the device fabrication process is very high (better than 95%). Fig. 2(a) shows a photograph. Spreading of threshold current and output power is very small, as shown in Fig. 2(b). Recently we fabricated VCSEL with equal diameters of the top DBR mesa and aperture diameter. The 14 m diameter VCSELs show a maximum output power of 5 mW with a very high external quantum efciency of 45% and a power conversion efciency of 16.5 %. The threshold for these devices is 1.06mA.

Pmax

0.05 mW

Ith
(a) (b)

50 A

Fig. 2: (a) Photograph of a 8x8 VCSEL array with a separate p-contact for each laser. (b) cw light vs. current for all VCSELs of that array (6 m blocking layer aperture). The homogeneity of threshold current and output power is extremely good. Supported by: DFG (SFB 348)

68

Amorphous Silicon Suboxide Light Emitting Diodes


Rainer Janssen* , Uwe Karrer, Doriana Dimova-Malinovska and Martin Stutzmann Amorphous substoichiometric silicon oxide (a-SiOx:H) is deposited by plasma-enhanced chemical vapour deposition using SiH4 and CO2 as source gases. The oxygen content and the optical gap E04 of the samples can be controlled by varying the CO2- partial pressure and the deposition power. Doped samples are realized by adding PH3 and B2H6 to the source gases. P-i-n diodes are deposited on transparent conducting oxide using p- and n- doped layers with lower oxygen content (25 at.% , E04=2.0 eV) and thicknesses of 30 nm and 100 nm respectively for an efficient injection of carriers into an i- layer with approx. 45 at.% oxygen (E04=2.3 eV) and variable thickness as the active luminescent layer. Fig. 1 shows the I-V characteristics of a series of suboxide p-i-n diodes at 300 K.

1 0.1

30 nm i - layer thickness

150 nm

2)

0.01 1E-3 1E-4

200 nm 0 nm

Current Density (A/cm

400 nm
1E-5 1E-6

800 nm
1E-7 1E-8 1E-9 -20

-15

-10

-5

10

15

20

Voltage (V)

Fig. 1: I-V characteristics at of p-i-n diodes with different i-layer thicknesses. Modelling these data with a basic circuit diagram for nonideal diodes gives values of 109 for the parallel resistance influencing the current density at small voltages. The ideality factor describing the exponential increase of the current density at intermediate voltages rises from 3 for an i-layer thickness of 30 nm to about 20 for thicker i-layers. At higher voltages, the I-V curves of thin diodes are dominated by the diodes series resistance of approx. 800 which is due to the low conductivity of the p-doped layer. The rectification ratio of thin suboxide diodes exceeds 5 orders of magnitude, and a significant current density under reasonable forward bias is reached for diodes with an i-layer thickness

* phone:

+49 89 289-12768, fax: -^12737, email: janssen@physik.tu-muenchen.de

69 less than 200 nm. Applying high voltage stress for several hours demonstrates stability of the devices for i-layer thicknesses larger than 100 nm. The intensity of electroluminescence (EL) at room temperature was measured with a silicon detector (Fig. 2). A significant EL signal appears at current densities exceeding 10-3 A/cm2 and increases linearly with the current density for more than two orders of magnitude. This indicates that recombination of a constant fraction of the injected carriers leads to photon emission. The best EL performance was detected for an i-layer thickness of 150 nm making this diode the starting point for further improvements mentioned below.

10 200

i - layer thickness

300 K

Electroluminescence (a.u.)

150 nm
1

200 nm 60 nm

Current density = 0.5 A/cm2

Electroluminescence (a.u.)

150

30 nm
0.1

i-layer thickness 150 nm


100

60 nm

0.01

0 nm

50

0 nm

E
04

(i-layer) = 2,33 eV 300 K


0.01 0.1 1

1E-3 1E-3

0 0.7

0.8

0.9

1.0

1.1

1.2

1.3

1.4

1.5

1.6

Current Density (A/cm2)

Energy (eV)

Fig. 2: EL intensity versus injection current measured by a Si-photodetector.

Fig. 3: Spectral shape of EL as a function of i-layer thickness.

Sufficiently high current densities can be realized for diodes with i-layer thicknesses smaller than 200 nm under forward bias only. Thus no EL is detected under reverse bias. Evidence for the i-layer being the origin of this EL comes from a comparison with the much lower signal of a p-n junction shown in Fig. 2. Spectrally resolved EL measurements at 300 K using a cooled germanium detector (Fig. 3) confirm this conclusion showing a defect-related luminescence peak at 0.8 eV for the p-n diode which was reported from EL studies of amorphous silicon p-i-n structures and attributed to defect recombination at the p-i interface. In addition to the luminescence peak at 0.8 eV a shoulder appears at higher energies and becomes more prominent with increasing i-layer thickness. A possible origin is recombination at defects being related to the incorporated oxygen. Obviously so far no bandtail-to-bandtail recombination is detected, as the expected luminescence for an i-layer bandgap of 2.3 eV should occur at higher photon energies. This fact points towards problems with the injection of holes into the i-layer because of the band discontinuity at the p-i interface. Therefore our major concern in the near future will be to improve the hole injection by modifying the p-layer and the p-i interface.

70

Erbium-Oxygen doped Silicon for Light Emitting Diodes


Eduard Neufeld*, Jrg Stimmer, Antonius Reittinger, Gerhard Abstreiter Excited erbium ions, when incorporated into a suitable host, are known to emit photons at about 1.54 m originating from an intra-4f transition. This wavelength coincides with the absorption minimum of silica-based fibers. Therefore erbium-doped materials are of high interest for optoelectronic applications. In particular, erbium-doping of Si and SiGe seems to be a promising apO 1020 proach towards silicon-based optoelectronics. Er 1019 In the present work we have invesP B tigated the optical properties of sili1018 con doped with erbium and oxygen, SiH -Background the latter dopant being a necessary 17 10 requirement for growth without erbium segregation and for optical 1016 activation of the Er3+-ions. To avoid 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 Depth (m) crystal damage, as in the case of doping by ion implantation, the Fig. 1: SIMS analysis of an as-grown Si:Er:O LED. samples have been grown completely using molecular beam epitaxy (MBE). This allows the incorporation of optically active Er3+-ions even without subsequent annealing. Samples with erbium and oxygen PL at 5 K concentrations of up to 21020 cm-3 T Substrate = 5 0 0 C , and 1021 cm-3, respectively, have [E r ] = 1 0 19 c m -3 been investigated. The concentra[O ] = 1 0 19 c m -3 tion profiles as studied by secondary ion mass spectroscopy (SIMS) (Fig. 1) show no noticeable segregation even at the highest erbium concentrations. Fig. 2 shows typical photoluminesa n n e a le d fo r 3 0 m i n a t 9 0 0 C cence spectra of an as grown and an annealed Si:Er:O-sample with the characteristic erbium signal at about as grown 1.54 m. The strongest photoluminescence emission was obtained 1 .1 1 .2 1 .3 1 .4 1 .5 1 .6 from samples doped with W avelength [ m ] 41019 cm-3 Er and 1020 cm-3 O. These concentrations correspond to Fig. 2: Photoluminescence spectra of an as grown and an anthe limit up to where two dimen- nealed Si:Er:O-sample sional growth could be observed by
-3 Concentration (cm )
X

tel.: ++49/89/289-12775; fax: ++49/89/3206620; e-mail: neufeld@wsi.tu-muenchen.de

PL-Intensity [a.u.]

71 reflection high energy electron diffraction (RHEED). At higher concentrations, three dimensional growth occurred. All samples investigated by photoluminescence show a characteristic temperature quenching of about three orders of magnitude when going from 4K to room temperature. In a further step, Si:Er:O pndiodes have been processed for electroluminescence measure[Er] = 2x10 cm ments. In order to restrict the [O] = 10 cm current flow through the active additional annealing (900C, 30 min) J = 150 A/cm (forward bias) layer, mesas of 300400m2 were defined photolithographically. To finish the diode structure, metal top and back contact 5K layers (Ti/Au) were evaporated on both sides of the samples. A 180 K 200200m2 window was de300 K, x3 fined on the top contact to pro1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7 vide optical access. W avelength [m] The processed diodes show typiFig. 3: Forward bias electroluminescence spectra at three different cal current voltage characteristemperatures of a diode grown with high erbium and oxygen concen- tics in forward bias. In reverse tration. The spectra are shifted vertically for clarity, the roombias, however, a smooth onset of temperature spectrum being magnified by a factor of three.
20 -3 21 -3

EL Intensity [a.u.]

the current occurs at a rather low voltage (approximately -3 V) which is most likely due to trap[Er] = 5x10 cm [O] = 10 cm mediated conduction. No degraadditional annealing (900C, 30 min) dation has been observed so far J = -42 A/cm (reverse bias) when operating the diodes under reverse bias with current densi300 K ties up to 85 A/cm2. The characteristic erbium electroluminescence emission has been measured for both forward 5K and reverse bias conditions 1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7 (Fig. 3 and 4, respectively). UnW avelength [m] der forward bias the erbium sigFig. 4: Reverse bias electroluminescence spectra at 5 K and room nal shows a similar temperature temperature of a diode with lower erbium and oxygen concentration. quenching as in the case of phoThe spectra are again shifted vertically. toluminescence. Under reverse bias, however, the emission at 1.54 m is almost independent of temperature between 5 K and room-temperature for low erbium and oxygen concentrations and is considerably stronger than under forward bias operation. Diodes with the highest erbium and oxygen concentrations, however, show a similar temperature quenching for both polarities, the reverse bias electroluminescence intensity still being considerably higher. The strongest electroluminescence was measured for erbium concentrations between 51019 cm-3 and 1.51020 cm-3 with typical oxygen concentrations being a factor of six higher.
19 -3 20 -3

EL Intensity [a.u.]

supported by: Bayerische Forschungsstiftung (FOROPTO) and Volkswagenstiftung (Photonik)

72

Opto-electronic devices based on Si/SiGe


Cornelia Engel*, Peter Schittenhelm, Christoph Schelling, Karoline Bernhard-Hfer, Markus Gail, Ralf Brederlow, Joachim F. Ntzel, Gerhard Abstreiter In spite of the indirect fundamental band gap, the Si/Ge material systems offers the possibility to realize various opto-electronic devices operating in the visible, the near and far infrared regions. Such devices could be integrated in Si-based electronic circuits, merging the advantages of VLSI technology and fiber optics. In strained Si/Ge heterostructures, the fundamental band gap can be tuned to the wavelength regime of 1.3 m and 1.55 m, which makes the system interesting for detector devices used for fiber communication. With increasing Ge content, the fundamental band gap shifts to smaller energies; concomitantly the strain in the SiGe layer reduces the critical thickness of a pure Ge layer in Si to a few monolayers. The small layer thickness leads to quantization effects enlarging the effective band gap. The Stranski-Krastanow growth mode, which results in the self-assembled formation of islands in strained systems, allows exceeding the critical thickness without introducing dislocations. Embedding self-assembled Ge-rich dots in an active detector structure promises a smaller effective band gap and consequently a higher responsivity in the near-infrared region. The small absorption due to the indirect fundamental bandgap can be compensated by a device concept with a high internal gain like in the case of a phototransistor. Photogenerated electron hole pairs are separated in the internal field and accumulated holes in the p-doped region tune the thermionic electron current across the base, as schematically depicted in the inset of Fig.1. Fig.1 shows the energy dependence of the photocurrent of an npn-phototransistor structure with 3 layers of nominal 12 monolayers pure Ge in the p-doped base (solid line). The dashed line shows the photocurrent of a sample with 3 Si0.6Ge0.4 quantum wells instead of the Ge-dot layers. The photocurrent of the sample with dots is more than one order of magnitude higher than the signal of the reference sample without dots. Together with the onset of the photocurrent at lower energies this demonstrates that the effective band gap in SiGe/Si heterostructures can be reduced by the Stranski-Krastanov growth mode. The responsivity of the phototransistor structure depends on intrinsic parameters like the n- and p-doping concentration as well as on extrinsic measuring conditions. With decreasing doping concentration in the base, the potential modulation is lowered. Consequently a certain density of photogenerated holes

Fig. 1:Photocurrent versus photon energy of a npnphototransistor with (---) Ge-dots and Si0.6Ge0.4quantum wells (- - -) in the base.
*

tel.: ++49-89-289 12776, fax: ++49-89-3206 620, e-mail: engel@wsi.physik.tu-muenchen.de

73 compensate a larger percentage of the barrier, resulting in a higher responsivity. As the sample temperature, the applied bias voltage and the incident illumination also influence the potential modulation, the internal gain depends on the measurement conditions. Optimizing the intrinsic and extrinsic parameters, a responsivity of 0.04 A/W (70A/W) can be achieved at =1.3m (=800nm). Another application of the npn-phototransistor structure may be deduced from the time-resolved photocurrent measurements (s. Fig. 2). Due to the separation of electrons and holes in k-space as well as in real space, the lifetime of confined holes is rather long. In a logarithmic plot two decay constants can be determined, 1=673ms and 2=72ms. These time constants are attributed to the lifetime of the confined holes in the Ge-layer. Fig. 2:Time dependence of the photocur- Such long lifetimes may allow the realization of memory devices based on the self-assembled dots. rent of a npn-phototransistor structure. Since the effective band gap is smaller than in the surrounding Si layer, the storage could be written optically with infrared radiation. The inhomogeneity of the dot sizes even makes it possible to charge specific dots selected by the incident photon energy (spectral hole burning). The large valence band offset of the Si/SiGe system is ideal for the realization of efficient intersubband detectors in the far infrared. Fig. 3 shows the transmission spectra of different modulation-doped multiquantum well samples. By varying the width of the quantum wells, the narrow absorption lines can be tuned in wavelength. With increasing well width, the absorption shifts from 10 to 20m, due to the decreasing energetic separation of the subbands. The observed absorption lines are in good agreement with theoFig. 3: Transmission spectra of p-Si/SiGe multi quantum well retical calculations (solid lines) taking into account the comsamples with different well width. plex valence band structure. In samples with higher Ge content the absorption could also be tuned to shorter wavelengths down to 4m. First experimental results show a responsivity of up to 0.1 A/W, a value which is comparable to commercially available III-V or II-VI detectors in this wavelength regime. The intersubband detector work has been performed in collaboration with the group of Prof. Bauer (Universitt Linz). supported by: Bayerische Forschungstiftung via FOROPTO and Siemens AG.

80

InGaAs/InAlAs HEMTs with extremely low source and drain resistances

S. Kraus*, H. Hei, D. Xu, M. Sexl, G. T

+,

G. Weimann#

InP based InAlAs/InGaAs high electron mobility transistors (HEMTs) have demonstrated excellent high frequency and low noise performance. But high contact and sheet resistances in the access regions between alloyed contact and gate electrode can cause lar ge parasitic source and drain resistances. The reduction of the parasitic resistances by parallel conduction in a thick cap layer is a well known concept from pseudomorphic AlGaAs/InGaAs HEMTs. At rst glance this recipe is not suitable for InAlAs/InGaAs HEMTs, because of the large conduction band discontinuity between InAlAs and InGaAs, which does not allow a current ow between these layers. Here we demonstrate that a tunneling current through the high barrier is possible by inserting a high planar Si doping layer into the top of the InAlAs layer . Further we show that a nondepleted thick cap layer can signicantly improve the parasitic resistances without deteriorating the diode characteristics. Fig. 1 :
source Rc1 Rcap Rcap alloyed contact Rc2 Rt Rt InAlAs channel Rch Rch nondepleted cap gate

Distributed equivalent circuit for the source resistance (R c1, R c2: contact resistances; Rt: tunneling resitance; Rch: channel sheet resistance; R cap: cap sheet resistance).

Standard lattice-matched HEMT structures normally use a nonconductive depleted cap layer . As a result, one only gets a single path for the current ow from the alloyed contact to the channel and the total resistance is xed by the contact and the channel sheet resistance (Fig. 1 : R tot = R c1 + R c2 + n Rch). There Rc1 is the contact resistance between metallization and InGaAs cap and Rc2 is the tunneling resistance through the InAlAs barrier to the channel.
600 InGaAs nSi = 1.0 1019cm-3 nSi = 2.0 180 In0.4Al0.6As nSi = 5.0 120 InGaAs nSi = 2.5 1012cm-2 260 2300 In0.4Al0.6As InAlAs InP - substrate 1013cm-2 1012cm-2

Fig. 2 : HEMT layer structure for the nondepleted cap concept.

But the sheet resistances of the channel are quite high; a typical value is 200 / , depending on the doping concentration. The concept of the nondepleted cap is based on the electrical connection between cap layer and channel (Fig. 1). If a current ow through the InAlAs barrier (repesented by the tunneling resistance Rt) can be achieved, then a cap layer with low sheet resistance can function as an extension of the alloyed contact and the total resistance is reduced by a distributed circuit of resistances. * phone : +49 89 3209 2789, fax : +49 89 3206 620, email : kraus@e26.physik.tu-muenchen.de + now at : Ferdinand Braun Institut, Berlin # now at : IAF, Freiburg

81

A calculation of the distributed circuit shows, that it is necessary to reduce both the tunneling as well as the cap resistance, to get a very low total resistance (Rtot = Rc1 + Rdistributed). The schematic cross section of the epitaxial structure is illustrated in Fig. 2. For device fabrication we used a double sided -doping heterostructure with 5.0 1012cm-2 in the upper and 2.5 1012cm-2 in the lower InAlAs layer and a 120 thick channel. The cap layer consists of 600 1.0 1019cm-3 homogeneously doped InGaAs for high parallel conductivity with only 46 / sheet resistance. The contact and sheet resistance were determined by standard TLM measurement for the whole layer structure. The parallel conduction in the cap layer leads to a very low contact resistance Rc1 below 0.04 mm and to a total sheet resistance of only Rsh = 38 / . A tunneling current through the InAlAs barrier is possible by inserting a high Si -doping into the top of the InAlAs layer . To avoid a deterioration of the diode characteristic, we used a strained Schottky barrier with 60% aluminium. T-shaped gates of 0.15 m length were dened by electron beam lithography in a two layer photoresist process. Gate recess was performed in two steps: rst by a selective succinic acid etch for the lateral recess denition, second by a nonselective phosphoric acid etch for removing the top Si -doping.
Ugs = 0.25 V Ugs = 0 V extr. transconductance [S/mm] drain current density [A/mm]

drain source voltage [V]

drain current density [A/mm]

Fig. 3 : I-V characteristics of nondepleted InAlAs/InGaAs HEMT.

Fig. 4 : Extrinsic transconductance versus drain current density at Uds = 1 V.

cap

Fig. 3 shows excellent output characteristic without any kink and an output conductance of 84 mS/mm at Uds = 1V and a maximum drain current of 1150 mA/mm. A very high extrinsic transconduction of 1150 mS/mm (Fig. 4) results from the extremely low source and drain resistance of Rs+Rd = 0.34 mm. These are the lowest source and drain resistances for 2 m source drain spacing ever reported for HEMTs on InP. Maximum extrinsic transconduction was obtained at a drain current density of 466 mA/mm. Regardless of the highly doped nondepleted cap, a breakdown voltage of -5.5 V has been measured at I g = 1mA/mm. This is due to the higher band gap of the strained In0.4Al0.6As which enhances the Schottky barrier. Fig. 5 :
|h21|2, mug [dB]

mug

h21 -20 dB

Extrapolated cut off frequencies fT = 206 GHz, fmax = 232 GHz for 150 nm gate length.

frequency [Hz]

The S-parameters of 150 m wide HEMTs were measured from 2-50 GHz using on-wafer probing. The current gain cut of f frequency f T and the maximum oscillation frequency f max were calculated by extrapolating |h 21|2 and mug with a -20dB/dec slope. W e get values of 206 GHz for f T and 232 GHz for fmax, respectively (Fig. 5). This leads to a very good fT-LG product of 31 GHz m. supported by : Siemens AG via SFE

Você também pode gostar