study of threading dislocation density reduction in algan ...2of8s. lazarev et al. dislocation...

8
research papers J. Appl. Cryst. (2013). 46 doi:10.1107/S0021889812043051 1 of 8 Journal of Applied Crystallography ISSN 0021-8898 Received 26 July 2012 Accepted 15 October 2012 # 2013 International Union of Crystallography Printed in Singapore – all rights reserved Study of threading dislocation density reduction in AlGaN epilayers by Monte Carlo simulation of high- resolution reciprocal-space maps of a two-layer system S. Lazarev, a M. Barchuk, b,a S. Bauer, a * K. Forghani, c V. Holy ´, b F. Scholz c and T. Baumbach a a Karlsruhe Institute of Technology (KIT)/Synchrotron Facility ANKA, 76344 Eggenstein-Leopold- shafen, Germany, b Charles University in Prague, Faculty of Mathematics and Physics, Ke Karlovu 5, 12116 Prague 2, Czech Republic, and c Institute of Optoelectronics, University of Ulm, Albert- Einstein-Allee 45, 89081 Ulm, Germany. Correspondence e-mail: [email protected] High-resolution X-ray diffraction in coplanar and noncoplanar geometries has been used to investigate the influence of an SiN x nano-mask in the reduction of the threading dislocation (TD) density of high-quality AlGaN epitaxial layers grown on sapphire substrates. Our developed model, based on a Monte Carlo method, was applied to the simulation of the reciprocal-space maps of a two- layer system. Good agreement was found between the simulation and the experimental data, leading to an accurate determination of the dislocation densities as a function of the overgrowth layer thickness. The efficiency of the SiN x nano-mask was defined as the ratio of the TD densities in the AlGaN layers below and above the mask. A significant improvement in the AlGaN layer quality was achieved by increasing the overgrowth layer thickness, and a TD density reduction scaling law was established. 1. Introduction Nitride-based ultraviolet light-emitting diodes have been investigated for their applications in high-density optical storage media, biotechnology, air/water purification and curing resins. By changing the Al composition, the cutoff wavelength of AlGaN photoactive devices can cover the spectrum from 200 to 360 nm at room temperature. However, in comparison with GaN, AlGaN layers grown directly on sapphire typically exhibit a large number of threading dislo- cations (TDs), which mainly occur in the form of edge-type TDs (Walker et al., 1996). The reduction of the number of threading dislocations, which act as nonradiative recombina- tion centres, is essential (Sugahara, Hao et al. , 1998; Sugahara, Sato et al. , 1998) for the improvement of the performance of ultraviolet light-emitting diodes. Recently the optimization of Al 0.2 Ga 0.8 N layers directly grown on sapphire by metal organic vapour phase epitaxy (MOVPE) has been investigated (Forghani et al., 2011). The measurement of the crystal quality of the AlGaN epilayers was determined qualitatively from transmission electron microscopy (TEM) micrographs. An improvement was found by applying an in situ nano-masking technology with ultra-thin SiN x interlayers. The edge-type dislocation density could be significantly reduced by optimizing the growth parameters and depositing the SiN x interlayer on a 150 nm Al 0.2 Ga 0.8 N inter- layer grown on an AlN nucleation layer. The width of the rocking curve of the asymmetric 10.2 reflection was used as a good indicator for the reduced defect density in the Al 0.2 Ga 0.8 N layer grown on top of the SiN x nano-mask (Forghani et al., 2011). A deeper insight into the growth mechanism and the defect- reduction process has been obtained by exploiting weak-beam dark-field micrographs in high-resolution TEM investigations (Klein et al., 2011). These micrographs show a significant difference of the dislocation densities below and above the SiN x mask. However, this method does not enable the deter- mination of the TD densities. To overcome the limitations of the TEM investigation we have applied several approaches to determine the dislocation densities from X-ray diffraction data. In our previous work, the TD densities were determined from simple equations based on the full width at half- maximum of the rocking curve (Gay et al. , 1953; Lazarev et al. , 2012). Recently, Monte Carlo simulation has been used for the modelling of reciprocal-space maps (RSMs) in coplanar and grazing-incidence diffraction (GID) of GaN layers grown on sapphire. By comparing the simulated and measured maps in the diffuse part generated by defects, the TD densities have been accurately determined (Barchuk et al., 2010, 2011). In this paper, we demonstrate the possibility of using a Monte Carlo method to compute more sophisticated Al 0.2 Ga 0.8 N epilayer structures and to establish the scaling law of the TD density reduction with overgrowth thickness. The TD densities below and above the SiN x nano-mask have been

Upload: others

Post on 21-Oct-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

  • research papers

    J. Appl. Cryst. (2013). 46 doi:10.1107/S0021889812043051 1 of 8

    Journal of

    AppliedCrystallography

    ISSN 0021-8898

    Received 26 July 2012

    Accepted 15 October 2012

    # 2013 International Union of Crystallography

    Printed in Singapore – all rights reserved

    Study of threading dislocation density reduction inAlGaN epilayers by Monte Carlo simulation of high-resolution reciprocal-space maps of a two-layersystem

    S. Lazarev,a M. Barchuk,b,a S. Bauer,a* K. Forghani,c V. Holý,b F. Scholzc and T.

    Baumbacha

    aKarlsruhe Institute of Technology (KIT)/Synchrotron Facility ANKA, 76344 Eggenstein-Leopold-

    shafen, Germany, bCharles University in Prague, Faculty of Mathematics and Physics, Ke Karlovu 5,

    12116 Prague 2, Czech Republic, and cInstitute of Optoelectronics, University of Ulm, Albert-

    Einstein-Allee 45, 89081 Ulm, Germany. Correspondence e-mail: [email protected]

    High-resolution X-ray diffraction in coplanar and noncoplanar geometries has

    been used to investigate the influence of an SiNx nano-mask in the reduction of

    the threading dislocation (TD) density of high-quality AlGaN epitaxial layers

    grown on sapphire substrates. Our developed model, based on a Monte Carlo

    method, was applied to the simulation of the reciprocal-space maps of a two-

    layer system. Good agreement was found between the simulation and the

    experimental data, leading to an accurate determination of the dislocation

    densities as a function of the overgrowth layer thickness. The efficiency of the

    SiNx nano-mask was defined as the ratio of the TD densities in the AlGaN layers

    below and above the mask. A significant improvement in the AlGaN layer

    quality was achieved by increasing the overgrowth layer thickness, and a TD

    density reduction scaling law was established.

    1. IntroductionNitride-based ultraviolet light-emitting diodes have been

    investigated for their applications in high-density optical

    storage media, biotechnology, air/water purification and

    curing resins. By changing the Al composition, the cutoff

    wavelength of AlGaN photoactive devices can cover the

    spectrum from 200 to 360 nm at room temperature. However,

    in comparison with GaN, AlGaN layers grown directly on

    sapphire typically exhibit a large number of threading dislo-

    cations (TDs), which mainly occur in the form of edge-type

    TDs (Walker et al., 1996). The reduction of the number of

    threading dislocations, which act as nonradiative recombina-

    tion centres, is essential (Sugahara, Hao et al., 1998; Sugahara,

    Sato et al., 1998) for the improvement of the performance of

    ultraviolet light-emitting diodes.

    Recently the optimization of Al0.2Ga0.8N layers directly

    grown on sapphire by metal organic vapour phase epitaxy

    (MOVPE) has been investigated (Forghani et al., 2011). The

    measurement of the crystal quality of the AlGaN epilayers

    was determined qualitatively from transmission electron

    microscopy (TEM) micrographs. An improvement was found

    by applying an in situ nano-masking technology with ultra-thin

    SiNx interlayers. The edge-type dislocation density could be

    significantly reduced by optimizing the growth parameters and

    depositing the SiNx interlayer on a 150 nm Al0.2Ga0.8N inter-

    layer grown on an AlN nucleation layer. The width of the

    rocking curve of the asymmetric 10.2 reflection was used as a

    good indicator for the reduced defect density in the

    Al0.2Ga0.8N layer grown on top of the SiNx nano-mask

    (Forghani et al., 2011).

    A deeper insight into the growth mechanism and the defect-

    reduction process has been obtained by exploiting weak-beam

    dark-field micrographs in high-resolution TEM investigations

    (Klein et al., 2011). These micrographs show a significant

    difference of the dislocation densities below and above the

    SiNx mask. However, this method does not enable the deter-

    mination of the TD densities. To overcome the limitations of

    the TEM investigation we have applied several approaches to

    determine the dislocation densities from X-ray diffraction

    data. In our previous work, the TD densities were determined

    from simple equations based on the full width at half-

    maximum of the rocking curve (Gay et al., 1953; Lazarev et al.,

    2012).

    Recently, Monte Carlo simulation has been used for the

    modelling of reciprocal-space maps (RSMs) in coplanar and

    grazing-incidence diffraction (GID) of GaN layers grown on

    sapphire. By comparing the simulated and measured maps in

    the diffuse part generated by defects, the TD densities have

    been accurately determined (Barchuk et al., 2010, 2011).

    In this paper, we demonstrate the possibility of using a

    Monte Carlo method to compute more sophisticated

    Al0.2Ga0.8N epilayer structures and to establish the scaling law

    of the TD density reduction with overgrowth thickness. The

    TD densities below and above the SiNx nano-mask have been

    http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB16

  • derived from the Monte Carlo simulation with the goal of

    estimating the SiNx nano-mask efficiency in defect reduction

    in AlGaN epilayers.

    2. Experiment

    2.1. Sample description

    In the present work we investigated seven samples of

    c-plane epitaxial AlGaN structures grown on sapphire (00.1)

    substrates with a miscut of 0.3� towards the a plane. Growth

    was carried out in a low-pressure horizontal MOVPE reactor

    (Aixtron AIX-200/4 RF-S) using the standard precursors

    trimethylgallium (C3H9Ga), trimethylaluminium (C3H9Al)

    and ammonia (NH3); the growth temperature was set to

    1393 K. An Al content of 20% for the AlGaN samples was

    derived from photoluminescence measurements (Neuschl et

    al., 2010). The growth was initiated by an oxygen-doped AlN

    nucleation layer (about 25 nm). After about 150 nm of AlGaN

    growth, an SiNx nano-mask layer was deposited in situ by

    flowing SiH4 and ammonia into the reactor for approximately

    5 min. On the top of this layer, a

    second AlGaN layer was deposited,

    with thicknesses of 90, 290, 500, 1000,

    1850, 2500 and 3500 nm corresponding

    to samples A1–A7, respectively. The

    layout of the AlGaN samples is shown

    in the inset of Fig. 1(a) (left), where the

    layer above the SiNx mask is termed

    the overgrowth layer while the layer

    below the mask is denoted the 150 nm

    AlGaN interlayer.

    2.2. Experimental setup and X-raymeasurements

    High-resolution X-ray coplanar and

    noncoplanar diffraction measurements

    have been performed at the bending-

    magnet single-crystal diffraction

    (SCD) beamline at the synchrotron

    facility ANKA in Karlsruhe in

    Germany. Fig. 1(a) (left) shows the six-

    circle diffractometer used to bring the

    sample into the Bragg condition and to

    measure symmetric and GID reflec-

    tions. The diffractometer has four

    degrees of freedom for the sample and

    two degrees of freedom for the detec-

    tor. All X-ray data were recorded with

    a microstrip solid-state detector

    (MYTHEN 1K, manufactured by

    DECTRIS Ltd), having 1280 channels

    with a channel size of 50 mm and apoint-spread function of one channel

    (Bergamaschi et al., 2010).

    Fig. 1(a) (middle) gives a schematic

    representation of the coplanar diffrac-

    tion geometry used to record RSMs of the symmetric 00.2,

    00.4, 00.6 and 00.8 reflections, at an energy of 12 keV for all

    samples. In this geometry the incident and the outgoing beams

    lie in the plane perpendicular to the sample surface. In this

    case the distance between the crystalline planes (hkl) parallel

    to the surface, in this paper later referred to as ‘d spacing’, has

    been determined.

    As is illustrated in Fig. 1(a) (middle), �i and �f are definedas the angles between the incident and the outgoing beams

    and the sample surface, respectively. The total angular range

    of the scattering angle �f and the angular resolution are givenby the effective detector length (64 mm), the sample–detector

    distance and the number of detector channels, as is shown in

    Fig. 1(a) (middle).

    The components of the scattering wavevector Q in the

    radial and angular directions, qrad and qang, respectively, are

    given by the following equations:

    qrad ¼ ð2�=�Þ sin �f þ sin �ið Þ; ð1Þ

    qang ¼ ð2�=�Þ cos�f � cos�ið Þ; ð2Þ

    research papers

    2 of 8 S. Lazarev et al. � Dislocation density reduction in AlGaN epilayers J. Appl. Cryst. (2013). 46

    Figure 1(a) The six-circle diffractometer at the SCD bending magnet beamline at ANKA (left), showing thelinear detector, the crystal analyser and a sample mounted on the Euler cradle stage. The inset showsthe layout of the Al0.2Ga0.8N samples. Schematic representations of the coplanar geometry used forsymmetric high-resolution X-ray diffraction (middle) and for the noncoplanar GID geometry (right).(b) Reciprocal-space map of the 00.2 reflection with two distinguishable peaks corresponding to AlNand AlGaN (left). The 00.8 reflection (middle) shows two AlGaN peaks, corresponding to theovergrowth layer and to the 150 nm AlGaN interlayer. The GID reflection 10.0 (right). All mapswere obtained using sample A4.

  • where � is the wavelength, qrad corresponds to the componentof Q along the surface normal and qang represents the in-plane

    component of Q.

    The variation of �i at the Bragg reflection was achieved bytilting the sample with respect to the incoming beam, which

    leads to the acquisition of RSMs of the 00.2, 00.4, 00.6 and 00.8

    symmetric reflections. In the RSM of the 00.2 reflection of

    sample A4 (Fig. 1b, left), two main peaks are distinguishable:

    an AlN peak coming from the nucleation layer and an AlGaN

    peak, which shows diffuse scattering in the transverse direc-

    tion (at qrad = 24.4 nm�1) due to the presence of dislocations.

    In the RSM the intensity along the qrad direction at qang = 0

    corresponds to the crystal truncation rod (CTR) where �i = �f.An additional streak (M) due to the crystal monochromator is

    visible in the RSM of the 00.2 reflection.

    Fig. 1(b) (middle) shows the highest measured Bragg

    reflection, 00.8, of sample A4, where the two AlGaN peaks

    designated as the AlGaN overgrowth layer and the 150 nm

    AlGaN interlayer could be resolved. Such high-order reflec-

    tions, which could readily be measured at the synchrotron

    facility, are advantageous for resolving diffraction peaks

    originating from only small differences in d spacing in the

    growth direction. Besides the two AlGaN peaks, additional

    peaks of intensity originating from AlN- and GaN-rich areas

    have been detected.

    The resolution function (�qrad, �qang) depends on theangular divergence of the incident and diffracted beams as

    well as on the monochromaticity of the X-ray beam. The energy

    resolution of the Si(111) crystal monochromator used in our

    investigation is �E/E ’ 0.0001, where the energy E is 12 keV.Since the scattering plane corresponds to the vertical plane,

    the resolution function in reciprocal space depends on the

    vertical divergence of the incident beam, which was defined by

    the vertical opening of the slit (0.3 mm) situated upstream

    from the sample. The angular resolution (0.004�) of the

    diffracted beam was predefined by the channel size of 50 mmand the distance between sample and detector of 700 mm.

    Fig. 1(a) (right) is a schematic representation of the non-

    coplanar diffraction geometry used for GID measurements,

    which were performed at an energy of 8 kev using the

    MYTHEN 1K line detector in combination with an Si(111)

    crystal analyser. In this case the resolution function (�qrad,�qang) was defined by the crystal analyser resolution and theenergy resolution of the Si(111) crystal monochromator,

    which in our investigation is �E/E ’ 0.0001 at an energy E of8 keV.

    In the GID geometry the diffracted intensity originates

    from the crystalline planes (hkl) perpendicular to the sample

    surface. In noncoplanar geometry the incident and the

    outgoing wavevectors are each defined by two angles: (�i, �i)and (�f, �f), respectively (see Fig. 1a, right). However theincident angle �i and the outgoing angle �f are relatively smalland close to the critical angle of 0.336�, and will be neglected

    in the calculation of the components of the scattering wave-

    vector. The angles �i and �f are the angles between the incidentbeam and the outgoing beam, respectively, and the diffracting

    planes.

    For GID reflections the components of the scattering

    wavevector Q in the radial and angular directions are given by,

    respectively,

    qrad ¼ ð2�=�Þ sin �f þ sin �ið Þ; ð3Þ

    qang ¼ ð2�=�Þ cos �f � cos �ið Þ: ð4Þ

    For each sample, RSMs of the 10.0 and 11.0 GID reflections

    were recorded. Fig. 1(b) (right) gives a typical RSM of the 10.0

    reflection. The broadening of the reflection in the angular

    direction qang is related to the defect density, while from the

    radial direction qrad one can determine the in-plane d spacing.

    3. Symmetric X-ray diffraction from the AlGaN epilayersystem

    In order to characterize the structural parameters (i.e. crys-

    talline parameters, defect densities etc.) of Al0.2Ga0.8N

    epilayers, we compared in our previous study (Lazarev et al.,

    2012) the RSMs of a GaN sample and an Al0.2Ga0.8N sample

    having similar overgrowth layer thicknesses of about 2.5 mm.Differences in peak positions have been detected in the RSMs

    of symmetric reflections. One of the aims of this paper is to

    identify the origin of the Bragg reflections detected in the

    RSMs of the AlGaN structure with respect to the individual

    layers of the sample.

    The evolution of the X-ray pattern as a function of the

    overgrowth thickness was examined by recording the RSMs of

    the 00.8 reflection shown in Fig. 1(b) (middle) for the different

    samples. The intensity distribution along the CTR was derived

    from the RSMs and is displayed in arbitrary units in Fig. 2(a).

    In order to clearly differentiate between the curves, each

    profile in Fig. 2(a) has been multiplied by a factor of five

    relative to the previous one.

    In order to identify the origin of the different peaks and to

    demonstrate accurately the resulting changes in peak position

    and broadening as function of the overgrowth thickness, a

    decomposition procedure was systematically applied to all

    samples.

    Figs. 2(b) and 2(c) contain the intensity distribution along

    the CTR as well as the contributing profiles for samples A1

    and A7, respectively. It should be emphasized that A1 and A7

    correspond to the lowest and highest overgrowth layer

    thicknesses, respectively. The choice of Lorentzians as a fitting

    function for the individual peaks enables us to achieve a good

    fit for the CTR profiles of the samples.

    We propose to follow the change in the peak intensities

    when the thickness of the overgrowth layer increases. For this

    purpose we define h1 = 150 nm as the thickness of the AlGaN

    layer below the SiNx interlayer and h as the thickness of the

    overgrowth layer, varying from 90 up to 3500 nm (see x2.1). Asis shown in Fig. 2(b) for sample A1, where h = 90 nm is less

    than h1, the intensity of peak 3 is higher than that of peak 4,

    while in the case of sample A7 (see Fig. 2c), where h = 3500 nm

    is greater than h1, the intensity of peak 4 is dominant, because

    of the increase of the diffracting volume coming from the

    overgrowth layer. As a consequence, the increase of the

    research papers

    J. Appl. Cryst. (2013). 46 S. Lazarev et al. � Dislocation density reduction in AlGaN epilayers 3 of 8

  • overgrowth layer thickness h leads to a continuous increase in

    the peak 4 intensity (see Fig. 2a). In conclusion, peak 3 could

    be attributed to the 150 nm AlGaN interlayer and peak 4 to

    the overgrowth layer. The decomposition procedure applied

    systematically to all the samples does not reveal any change in

    the position of peaks 3 and 4. Since peaks 3 and 4 have

    different positions in the CTR cut, one can conclude that the

    corresponding vertical lattice parameters of the structure

    below and above the SiNx mask are different.

    A previous study using spatially resolved scanning electron

    microscope cathodoluminescence (Neuschl et al., 2010)

    revealed the growth of GaN-rich islands in the openings of the

    SiNx interlayer, surrounded by AlN-rich areas. These islands

    coalesce to a fairly uniform AlGaN layer, corresponding to the

    overgrowth layer in this manuscript. The observed diffraction

    peaks 1 and 2 occur close to the values of a pure GaN crys-

    talline structure [at qrad(GaN) = 97.87 nm�1], while peak 5 is

    closer to the peak of pure AlN [at qrad(AlN) = 100.93 nm�1];

    these peaks are therefore presumably attributable to the GaN-

    and AlN-rich areas, respectively.

    4. Application of Monte Carlo simulation to a two-layersystem

    Numerical Monte Carlo simulation was applied successfully

    for the determination of the edge (De) and screw (Ds) dislo-

    cation densities from the diffuse scattering of GaN samples

    having nearly the same growth structure as our AlGaN

    samples (Barchuk et al., 2010, 2011). For these samples it has

    been demonstrated that the TD densities determined from the

    Monte Carlo simulation are comparable to those measured by

    the etch pit density method. The simulation of RSMs of GaN

    samples was carried out with the assumption that the

    diffracted intensity distribution originates only from the GaN

    layer above the SiNx layer. The comparison of RSMs of

    symmetrical and asymmetrical reflections of AlGaN and GaN

    by Lazarev et al. (2012) has revealed different behaviour:

    contrary to the case of AlGaN, GaN shows only one peak

    along the CTR in the RSM of the 00.8 reflection measured

    with the same instrumental function. An attempt to apply the

    Monte Carlo method to simulate the RSM by considering only

    the AlGaN overgrowth layer was not successful in achieving a

    good fit with the measured RSM. For this reason a further

    development was implemented into the model to calculate the

    RSM of a more complex system consisting of two AlGaN

    layers, in order to derive independently and accurately the

    density of edge and screw threading dislocations of each

    individual AlGaN layer.

    Because of their relatively low density we do not take into

    account the presence of screw dislocations in the calculation of

    the RSMs in GID for either the 150 nm interlayer or the

    overgrowth layer. Nevertheless, the simulations of the RSMs

    of the symmetric reflection of the overgrowth layer and the

    150 nm interlayer are slightly different because of the

    different types of interfaces: the 150 nm AlGaN interlayer

    does not have a free surface and hence does not reveal any

    surface stress relaxation since it forms a top interface with the

    AlGaN overgrowth layer and a lower interface with the AlN

    nucleation layer. In contrast, the overgrowth layer possesses a

    discontinuity at the surface, which gives rise to an additional

    nonzero vertical component parallel to the diffraction vector

    research papers

    4 of 8 S. Lazarev et al. � Dislocation density reduction in AlGaN epilayers J. Appl. Cryst. (2013). 46

    Figure 2(a) CTR profiles for samples A1–A7 derived from RSMs of the 00.8reflections. The profiles were shifted vertically in order to make thechanges in peak positions and intensities visible as a function of theovergrowth thicknesses. It is observed that the intensity of peak 3decreases while that of peak 4 increases. (b) The CTR profile of the 00.8reflection and the corresponding decomposed peaks for sample A1,where the overgrowth layer is 90 nm, i.e. lower than the thickness of the150 nm AlGaN interlayer. The intensity of peak 3 is higher than that ofpeak 4. (c) The CTR profile of the 00.8 reflection and the correspondingdecomposed peaks for sample A7, where the overgrowth layer is3500 nm, i.e. higher than the thickness of the 150 nm AlGaN interlayer.Peak 4 is dominant because of the increase of the diffracting volumecoming from the overgrowth layer.

  • of the symmetric reflections, resulting from the displacement

    field created by edge dislocations (Shaibani & Hazzledine,

    1981).

    Since in the case of the overgrowth layer the edge dislo-

    cations play an important role in the general intensity distri-

    bution in the RSMs of symmetric reflections, De serves as an

    input parameter in the simulations of these RSMs. In contrast,

    these symmetric reflections are not affected by De in the case

    of the 150 nm AlGaN interlayer.

    In our experiment the penetration depth in GID was esti-

    mated to be less than 150 nm, which enables us to probe solely

    the upper part of the overgrown AlGaN layer for samples A2

    through A7 with h > h1. Hence, we could determine the

    densities of edge TDs for this thickness. This statement does

    not apply to sample A1, where the overgrowth layer was only

    90 nm, i.e. smaller than the penetration depth. The study of

    this sample will be presented in the next section. A detailed

    description of the different steps involved in the application of

    Monte Carlo simulation to RSMs was previously introduced

    by Barchuk et al. (2010).

    To derive the TD density the RSMs of the GID reflections

    10.0 and 11.0 were simulated. The simulated RSMs were then

    convoluted with a resolution function to achieve a finite

    experimental resolution. The resolution function was chosen

    as a two-dimensional Gaussian with a width corresponding to

    the coherently irradiated area of the sample surface estimated

    for our experimental setup. A similar procedure for GaN films

    is described by Barchuk et al. (2011). The comparison of

    experimental and simulated RSMs for sample A4 is shown in

    Figs. 3(b) and 3(c). In the experimental 11.0 RSM (Fig. 3c) one

    can see some asymmetry in the radial direction, which indi-

    cates the presence of a local difference in the in-plane lattice

    parameters. However, for the TD density determination it is

    more important to fit the tails of the peaks, and therefore this

    asymmetry does not affect the accuracy or reliability of the

    values determined by the simulation.

    Subsequently, Ds was determined by

    considering a system with two AlGaN

    layers as described above. Contrary to the

    GID method, the penetration depth in the

    coplanar diffraction geometry was about

    5000 nm for the 00.8 reflection, so all layers

    deposited on the sapphire substrate

    contribute to the RSM.

    In the previous section it was shown in

    Fig. 1(b) (middle) that two peaks

    appearing in the RSM of the 00.8

    symmetric reflection are attributable to the

    AlGaN overgrowth layer and to the

    150 nm AlGaN interlayer. Consequently,

    the intensity distribution of the RSMs

    represents the overlap of these two peaks.

    The RSMs from these two layers were

    simulated independently and then the

    intensity distributions were superposed.

    The densities of screw dislocations in both

    layers were derived after achievement of

    the best agreement between experimental

    and simulated RSMs (see Fig. 3a). The qangaxis is chosen as the (10.0) direction and

    the qrad axis as the (00.1) direction. One

    can observe some disagreement between

    the experimental and simulated RSMs of

    the 00.8 reflection in the lower qrad values

    due to the fact that the GaN-rich peak was

    not taken into consideration in our simu-

    lation model.

    In addition to comparing the shapes of

    intensity distributions in the RSMs we

    analysed their cuts along the angular

    directions, following the approach intro-

    duced by Kaganer et al. (2005). For the

    symmetric 00.8 reflection we compared the

    cuts through both peaks. The experimental

    and simulated profiles in log–log repre-

    research papers

    J. Appl. Cryst. (2013). 46 S. Lazarev et al. � Dislocation density reduction in AlGaN epilayers 5 of 8

    Figure 3Measured (red, dots) and simulated (blue, solid) reciprocal-space maps for the reflections 00.8(a), 10.0 (b) and 11.0 (c) for sample A4. The intensity steps are in logarithmic scale: 100.5 (a) and100.25 (b), (c).

    Figure 4The cuts through the main peak of symmetric reflection 00.8 along the angular direction qang; theasymptotic behaviour of cuts obeys q�2ang � q�3ang (a). The cuts through the maximum of intensity ofthe 10.0 GID reflection (b) and the 11.0 reflection (c). For all the samples, the experimentalcurves are depicted by dots and the theoretical curves by solid lines.

  • sentation are in good agreement, especially in the diffuse part

    of the intensity distributions, which confirms that our model

    was applied successfully and could be used for accurate

    determination of TD densities. The cuts through the main

    peaks of symmetric reflections are depicted in Fig. 4(a). The

    asymptotic behaviour of the tails seems to follow a q�nang law,

    with n between 2 and 3. This indicates the presence of screw

    TDs as the prevailing type of dislocation, as is proposed by

    Holý et al. (2008). In a similar way, the cuts through the

    maximum of intensity of the GID reflections are displayed in

    Figs. 4(b) and 4(c). In contrast to the symmetric reflections, the

    asymptotic behaviour of GID reflection cuts is mainly influ-

    enced by the edge type of TDs. The tails of the curves lie

    between q�1ang and q�2ang rather than q

    �2ang and q

    �3ang, as might be

    expected from the universal law q�3ang (Kaganer et al., 2005).

    One possible explanation is the excess of the edge TD density

    with respect to the screw TD density.

    5. Determination of the edge dislocation density belowand above the SiNx nano-mask

    Our TEM investigation has led to a comprehensive under-

    standing of growth and defect reduction mechanisms occur-

    ring by the formation of dislocation half-loops at the SiNxinterface. A qualitative estimation of the dislocation densities

    was based on the observation of TEM images (Klein et al.,

    2011, 2010). In fact, a homogenous edge-type dislocation

    density has been observed in the layer below SiNx, while

    defect-free and defect-rich areas have been detected side by

    side on a scale of several hundreds of nanometres above the

    SiNx nano-mask. In order to determine quantitatively the

    efficiency of the SiNx nano-mask in the reduction of threading

    dislocation densities, it is very important to probe the AlGaN

    layers below and above the SiNx nano-mask simultaneously

    with X-rays. For all samples, the GID 11.0 reflection was

    measured with a maximum achievable penetration depth of

    150 nm, calculated using an approach described by Dosch et al.

    (1986). For samples A2–A7 the overgrowth thickness h was

    greater than the penetration depth of 150 nm, and only for

    sample A1 was it possible to measure simultaneously the

    signals from the overgrowth layer completely and those from

    the 150 nm AlGaN interlayer partially (60 nm). Consequently

    the RSM of the 11.0 GID reflection of sample A1 shows

    different features in comparison to the remaining samples. The

    RSM of the 11.0 reflection presented in Fig. 5 (left) shows the

    presence of two overlapped signals shifted along the qraddirection and originating from the overgrowth layer and the

    150 nm interlayer. In order to derive the TD densities from the

    overlapped signals, Fig. 5 (middle) illustrates the decomposi-

    tion procedure applied to the RSM, involving decomposition

    into two-dimensional Lorentzians. The parameters of the

    Lorentzians were determined from the decomposition of the

    radial and angular scans through the maxima of the signals

    indicated in the RSM in Fig. 5 (left). The two derived RSMs

    were simulated separately by considering two separate layers

    having 90 and 150 nm thicknesses. The RSMs simulated by the

    Monte Carlo method for the two layers are compared with the

    decomposed RSMs in Fig. 5 (right). The edge dislocation

    densities were consequently determined from the best fit and

    were found to be 6.4 � 1010 cm�2 for the 90 nm AlGaNovergrowth layer and 9.8 � 1010 cm�2 for the 150 nm AlGaNinterlayer. This method will allow us to quantify the efficiency

    of the SiNx nano-mask, as discussed in the next section.

    6. Scaling law of the threading dislocation densityreduction

    The De and Ds data derived from the Monte Carlo simulation

    are plotted in log–log scale as a function of the overgrowth

    research papers

    6 of 8 S. Lazarev et al. � Dislocation density reduction in AlGaN epilayers J. Appl. Cryst. (2013). 46

    Figure 5The 11.0 GID reflection of sample A1 (left), decomposition of the overlapped signals from the 150 nm AlGaN interlayer and the 90 nm overgrowth layer(middle), and Monte Carlo simulations of the peaks (right).

  • thickness h in Fig. 6(a). The fitting of the De variation with a

    scaling law h�n gives an exponent n = 0.95.

    We define here the efficiency factor of the SiNx nano-mask

    as the ratio between De = 9.8 � 1010 cm�2 of the 150 nmAlGaN interlayer and De of the overgrowth layer thickness h

    (Fig. 6b). The increase of the overgrowth layer thickness

    conspicuously improves the efficiency of the SiNx nano-mask.

    In fact, this factor increases up to 52 for the overgrowth

    thickness of 3500 nm.

    As described by Klein et al. (2011), the dislocation bending

    leads to the annihilation of TDs by the formation of disloca-

    tion loops. A growth model was proposed by these authors,

    illustrated in Figs. 10 and 11 of Klein et al. (2011), where it is

    demonstrated that the overgrowth will favour the coalescence

    of two nearby islands leading to small areas with TDs at the

    surface. This dislocation reduction mechanism proposed by

    the model leads to the enhancement of the efficiency factor

    with increasing overgrowth thickness h.

    The value of Ds is found to be lower than De for all over-

    growth thicknesses h, and follows a scaling law h�n, with an

    exponent n = 0.18 for the overgrowth layer and n = 0.33 for the

    150 nm AlGaN interlayer. A similar result was derived from

    high-resolution TEM investigation (Klein et al., 2010, 2011).

    Furthermore, Ds of the overgrowth layer is about a factor of 2

    lower than Ds of the 150 nm AlGaN interlayer. This demon-

    strates that the SiNx nano-mask plays a role in the reduction of

    Ds as well as of De.

    7. Conclusions

    In this work we found a local difference in the d spacing along

    the growth and the lateral directions as determined by the

    analysis of symmetric and GID reflections, respectively.

    The crystalline lattice parameters of the 150 nm AlGaN

    interlayer and of the overgrowth layer are different. Our

    investigation of the AlGaN epilayer shows evidence of the

    presence of several phases, including two AlGaN phases as

    well as GaN- and AlN-rich phases. These latter two have not

    been the main focus of this manuscript.

    By a systematic study of the intensity distribution along the

    CTR of the 00.8 symmetric reflection we were able to attribute

    the different peaks to the different individual layers

    composing the AlGaN epilayer system. We conclude that the

    lower and upper main peaks of the symmetric reflection

    originate from the 150 nm AlGaN interlayer and from the

    overgrowth layer, respectively. The decomposition of the CTR

    profile for all the samples into several peaks shows a change in

    the peak intensities but no change in the peak positions, which

    means that the local difference in the d spacing is not influ-

    enced by the overgrowth thickness.

    Our simulation based on the Monte Carlo method has

    proven to be reliable in modelling the RSMs in coplanar and

    noncoplanar geometries of a two-layer system. The experi-

    mental and the Monte Carlo-simulated RSMs are in good

    agreement. This allows us to derive screw and edge-type

    threading dislocation densities with an error of less than 15%.

    Furthermore, the reduction of De follows a scaling law h�n

    with the overgrowth thickness, where n = 0.95. The asymptotic

    behaviour of the tails of the symmetric reflections obeys a q�nanglaw, with 2 < n < 3, while the tails of the diffraction peaks in the

    GID follow a q�nang law, with 1 < n < 2. The deviation of

    asymptotic decay from the universal law q�3ang, particularly in

    GID, is due to the large edge-type TD densities.

    We have demonstrated that the analysis of diffuse scattering

    by using the simulation of decomposed RSMs of GID reflec-

    tions is a powerful method to determine separately the TDs in

    a two-layer system, e.g. below and above an SiNx nano-mask.

    We would like to thank Dr Gernot Buth, beamline scientist

    at the SCD beamline at ANKA (Karlsruhe Institute of

    Technology), for his support during beamtime and Dr Stephen

    Doyle for his help with the manuscript. The epitaxial growth

    of the AlGaN layers under investigation was financially

    supported by the BMBF.

    References

    Barchuk, M., Holý, V., Miljevic, B., Krause, B. & Baumbach, T. (2011).Appl. Phys. Lett. 98, 021912.

    Barchuk, M., Holý, V., Miljevic, B., Krause, B., Baumbach, T.,Hertkorn, J. & Scholz, F. (2010). J. Appl. Phys. 108, 043521.

    Bergamaschi, A., Cervellino, A., Dinapoli, R., Gozzo, F., Henrich, B.,Johnson, I., Kraft, P., Mozzanica, A., Schmitt, B. & Shi, X. (2010). J.Synchrotron Rad. 17, 653–668.

    research papers

    J. Appl. Cryst. (2013). 46 S. Lazarev et al. � Dislocation density reduction in AlGaN epilayers 7 of 8

    Figure 6(a) The variation of Ds determined from coplanar high-resolution X-raydiffraction and De determined from GID measurement as a function ofthe overgrowth thickness h. (b) The variation of the efficiency of the SiNxnano-mask in dislocation reduction with overgrowth thickness h.

    http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB1http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB1http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB2http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB2http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB3http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB3http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB3

  • Dosch, H., Batterman, W. & Dack, D. C. (1986). Phys. Rev. Lett. 56,1144–1147.

    Forghani, K. et al. (2011). J. Cryst. Growth, 315, 216–219.Gay, P., Hirsch, P. B. & Kelly, A. (1953). Acta Metall. 1, 315–319.Holý, V., Baumbach, T., Lübbert, D., Helfen, L., Ellyan, M., Mikulı́k,

    P., Keller, S., DenBaars, S. P. & Speck, J. (2008). Phys. Rev. B, 77,094102.

    Kaganer, V. M., Brandt, O., Trampert, A. & Ploog, K. H. (2005). Phys.Rev. B, 72, 045423.

    Klein, O., Biskupek, J., Forghani, K., Scholz, F. & Kaiser, U. (2011). J.Cryst. Growth, 324, 63–72.

    Klein, O., Biskupek, J., Kaiser, U., Forghani, K., Thapa, S. B. & Scholz,F. (2010). J. Phys. Conf. Ser. 209, 0120018.

    Lazarev, S., Bauer, S., Forghani, K., Barchuk, M., Scholz, F. &Baumbach, T. (2012). J. Cryst. Growth. In the press.

    Neuschl, B., Fujan, K. J. & Feneberg, M. (2010). Appl. Phys. Lett. 97,192108.

    Shaibani, S. J. & Hazzledine, P. M. (1981). Philos. Mag. A, 44, 657–665.

    Sugahara, T., Hao, M., Wang, T., Nakagawa, D., Naoi, Y., Nishino, K.& Sakai, S. (1998). Jpn. J. Appl. Phys. 37, L1195–L1198.

    Sugahara, T., Sato, H., Hao, M., Naoi, Y., Kurai, S., Tottori, S.,Yamashita, K., Nishino, K., Romano, L. T. & Sakai, S. (1998). Jpn.J. Appl. Phys. 37, L398–L400.

    Walker, D., Zhang, X., Kung, P., Saxler, A., Javadpour, S., Xu, J. &Razeghi, M. (1996). Appl. Phys. Lett. 68, 2100–2101.

    research papers

    8 of 8 S. Lazarev et al. � Dislocation density reduction in AlGaN epilayers J. Appl. Cryst. (2013). 46

    http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB15http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB15http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB5http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB6http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB7http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB7http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB7http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB8http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB8http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB9http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB9http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB10http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB10http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB11http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB11http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB12http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB12http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB13http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB13http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB14http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB14http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB15http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB15http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB15http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB16http://scripts.iucr.org/cgi-bin/cr.cgi?rm=pdfbb&cnor=he5567&bbid=BB16