Next Article in Journal
Cohesive Subgraph Identification in Weighted Bipartite Graphs
Next Article in Special Issue
Influence of Contacts and Applied Voltage on a Structure of a Single GaN Nanowire
Previous Article in Journal
Imitation Learning with Graph Neural Networks for Improving Swarm Robustness under Restricted Communications
Previous Article in Special Issue
Solid-State Transformation of an Additive Manufactured Inconel 625 Alloy at 700 °C
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rocking Curve Imaging Investigation of the Long-Range Distortion Field between Parallel Dislocations with Opposite Burgers Vectors

1
European Synchrotron Radiation Facility, BP 220, CEDEX 9, F-38043 Grenoble, France
2
SOITEC, Parc Technologique des Fontaines, F-38190 Bernin, France
3
UGA-CEA, CEDEX 9, F-38043 Grenoble, France
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(19), 9054; https://doi.org/10.3390/app11199054
Submission received: 18 August 2021 / Revised: 16 September 2021 / Accepted: 22 September 2021 / Published: 28 September 2021

Abstract

:
We observe a long-range distortion field between parallel dislocations with opposite Burgers vectors in a platelet-shaped single crystal of 4H-SiC with a low dislocation density (~103 cm/cm3). This distortion field is in the µradian range when the distance D between dislocations is in the ~50–250 µm range. We were able to characterise this weak distortion field through Rocking Curve Imaging (RCI), a highly sensitive Bragg diffraction imaging technique using monochromatic synchrotron radiation. From the experimental images, we generate maps of the angle of maximum reflectance (“peak position”) that provide a measurement of the local lattice orientation. Deviations from the crystal matrix orientation are associated with the long-range distortion field around dislocations. Between parallel dislocations with opposite Burgers vectors, this distortion does not decay to zero but towards a constant value α. We propose a simple model considering the angular parameter α characterising the distortion. This model indicates that α should roughly vary as 1/D. This appears to be in fair agreement with our experimental data.

1. Introduction

The interaction between dislocations is a well-known topic, discussed in scientific literature since the middle of last century. High-quality overviews on this topic can be found in reference books concerning dislocations (see for instance [1,2,3]). The aim of the present work is to investigate a particular aspect observed for non-mobile, isolated dislocations, in a platelet-shaped single crystal of 4H-SiC with low dislocation density (~103 cm/cm3).
The dislocations in 4H-SiC cannot move easily at room temperature, and we can consider their distance D as constant. When this distance D is in the range 50–250 µm, we observe a weak (~µradian) distortion of the region lying between parallel dislocations with opposite Burgers vectors. Dislocations in 4H-SiC crystals have been extensively studied through X-ray diffraction Rocking Curve Imaging (see for instance [4,5,6,7,8,9]). The faint distortion we observe had not been investigated previously because there was no available experimental technique allowing its visualisation. We took advantage of the capabilities of Rocking Curve Imaging (RCI) [10] to perform this investigation. RCI is a Bragg diffraction imaging technique that has been developed and pushed to the limits at the European Synchrotron Radiation Facility (ESRF) over the last twenty years (see, as examples, [11,12,13]). Nowadays, it is able to provide maps of a crystal with an angular resolution in the µradian range.
Our experiment shows that the region lying between parallel dislocations with opposite Burgers vectors exhibits an angular misorientation with respect to the crystal matrix that differs from the simple sum of the misorientation of the individual dislocations. The distortion field is modified and appears to extend the distortion associated with the neighbourhood of the core of the dislocations. In the present work, we characterise this misorientation and discuss its origin.

2. Materials and Methods

Single crystals of silicon carbide are grown industrially for high-power electronic applications. The experiment was carried out at the beamline ID06-HXM of the ESRF [14] on a commercially produced 0.35 mm thick single crystal wafer of 4H-SiC. The surface orientation of our sample was nearly (0001), with a 4° offset towards the [ 10 1 ¯ 0 ] direction. We studied the ( 40 4 ¯ 0 ) reflection in transmission mode (Laue geometry) using an X-ray beam energy of 10 keV with θB = 14.6°. The beam size was 1 × 1 mm2. An indirect two-dimensional detector with an effective pixel size of 0.6 × 0.6 µm2 (leading to a field of view of 1.2 × 1.2 mm2) was positioned approximately 40 mm downstream of the sample.
The experimental technique we used is Rocking Curve Imaging (RCI), which gives access to quantitative data associated with crystal distortion with a spatial resolution in the µm range. RCI is an extension of monochromatic Bragg diffraction topography. Experimentally, it is performed by recording a series of images (200 in the present work) at slightly different consecutive angular positions on the Bragg diffraction curve (“rocking curve”) using a two-dimensional pixelated detector (Figure 1). The data recorded along the rocking curve are processed and analysed pixel by pixel using purpose-built software [15]. Maps of the “local” integrated intensity, Full Width at Half-Maximum (FWHM) and peak position are then generated from the rocking curves recorded in each pixel. The last two maps result from a Gaussian fit of the rocking curve recorded on each individual pixel, as schematically indicated in Figure 1.
RCI was developed using synchrotron radiation because it is difficult to implement with conventional laboratory X-ray equipment. In particular, the number of photons recorded in a given pixel of the detector (with a surface typically in the 1 µm2 range) should be statistically meaningful, which implies that a high photon flux and a bijective (reversible) relationship between one pixel of the detector and one voxel in the crystal should exist. This implies a low divergence beam. This technique was initially implemented in the reflection geometry [10] and was used for the characterisation of crystals and deposited layers with thicknesses in the µm range [12]. In order not to be limited to the near-surface distortion, we use RCI in transmission geometry.
Most of our RCI work relies on two assumptions that simplify the interpretation of the various maps. We first assume that interferences, such as Kato fringes, associated with dynamical diffraction effects [16], are averaged out and that the diffracted beam recorded in a given pixel of the detector originates from a well-defined volume (“voxel”) in the crystal that corresponds to the intersection of the incident beam with the projection of the considered pixel along the Bragg diffracted direction. We assume, in addition, that all of the beam diffracted by the considered voxel goes to one pixel of the detector; this assumption is valid when the “mosaic spread” in this voxel, measured by the FWHM obtained from the Gaussian fit, is small enough (smaller than ~2 × 102 µradian for our experimental conditions). Both assumptions appear to be well-satisfied in the present work.
Crystal defects produce a contrast in the integrated intensity map that is similar to that obtained by classical Lang or white beam topography [16]. The FWHM and integrated intensity maps are rather similar in the X-ray low-absorption case. However, the FWHM map provides a quantitative measurement of the distortion associated with a given defect. The peak position map measures the effective misorientation [16] averaged over each of the voxels. In the case of a dislocation, this effective misorientation exhibits a change of sign when passing from one side of the dislocation line to the other. Let us consider, as an example, the case of an isolated screw dislocation perpendicular to the diffracting planes: the variation of peak position Δω measured when going from a distance (−r0) from the core of the dislocation to a distance (+r0) is given by Δω = br0, b being the modulus of the Burgers vector of the dislocation. A variation of the peak width associated with the dislocation results from the change of Δω within a pixel. This contribution to the FWHM can be written as ~p (δ(Δω)/δr), where p is the pixel size. We use this characteristic variation to identify dislocations in the peak position and FWHM maps [11]. In the present work, we pay close attention to the quantitative dependence of both Δω and FWHM when dislocations are not isolated but are parallel and exhibit opposite Burgers vectors.

3. Results

From the ensemble of images recorded along the rocking curve of our 4H-SiC sample, we produced maps of the local FWHM and peak position. These maps result, as described above, from a Gaussian fit of the rocking curve recorded in each individual pixel.
The FWHM map (Figure 2) shows for most pixels a local FWHM that is around 60 µradians. The width is rather uniform, with a few exceptions that will be discussed later. This value mainly results from the convolution between an “instrumental” divergence δζ and the local mosaic spread of the crystal. δζ includes the divergence of the faintly convergent beam (~40 µradians) and the width corresponding to the energy bandwidth of the Si (111) monochromator (~30 µradians). It appears that δζ is the dominant term of this convolution and that the mosaic spread corresponding to most of the voxels is smaller (<30 µradians), indicating a rather high crystalline quality of the crystal matrix.
A few regions exhibit, in Figure 2, a higher FWHM than the matrix. They are associated with the distortion field of crystalline defects. The most visible ones are red-yellow lines located in the upper right corner of the figure, which are probably associated with surface scratches. We also observe a number of red dots surrounded by yellow pixels. These dot-like images are also visible on the peak position map shown in Figure 3, where they display a double-dot shape, the dot on one side exhibiting a higher peak position than the matrix and the one on the other side a lower peak position (see, e.g., the image in the black circle on Figure 3). These FWHM and peak position images are characteristic of dislocations [11], which show up in RCI images as regions of higher FWHM (red-yellow dots in the FWHM map) and exhibit an abrupt change of peak position from one side of the dislocation core to the other, in the peak position map. In both Figure 2 and Figure 3, these dislocation-associated images are “dot” shaped, indicating that the dislocation lines are nearly perpendicular to the surface, i.e., nearly parallel to the c-axis. We estimated the dislocation density (~1300/cm2) by counting these dot-like images.
All the observed dot-like images are very similar, suggesting that the crystal displays one type of threading dislocation, having the same direction along the c-axis and the same Burgers vector modulus. The peak position map shows that both signs of the Burgers vector are present because the sign of the angular departure with respect to the matrix average on one (e.g., the right) side of a dislocation is not the same for all dislocations.
Let us now point out the behaviour of the region lying between two dislocations with opposite Burgers vectors, which constitutes the main topic of the present work. This behaviour is particularly visible for the pair of dislocations indicated by the red arrow on both the FWHM (Figure 2) and the peak position map (Figure 3). The two dislocations, located at a distance D~50 µm, exhibit Burgers vectors of opposite signs as indicated, in Figure 3, by reversed yellow and blue sides. The region between the dislocations displays a higher FWHM than the matrix (Figure 2) and is also misoriented with respect to the matrix (Figure 3).
Let the angle Δω be the difference in the peak position relative to the crystal matrix. Figure 4 shows how Δθ varies in the region between two dislocations of opposite sign, indicated by the red arrow in Figure 2 and Figure 3. The angular position of the ( 40 4 ¯ 0 ) Bragg diffracting planes does not relax to the matrix orientation in the region lying between the dislocations. Instead, it decays towards an angle α corresponding to a region of the distortion field of the dislocations located near the core. This angle α is not constant over the region between the two dislocations: it is largest along the line joining the cores of the dislocations and decreases when going away from this line, implying a gradient of distortion that is evidenced by a higher FWHM than that of the matrix in Figure 2. The different curves shown in Figure 4 correspond to averaging α over distances, perpendicular to the line joining the core of the dislocations, going from ~3 to 20 µm. The average angle α corresponds to the 20 µm-wide region that displays a darker blue color in the peak position map (Figure 3), corresponding to an orientation difference of ~10 µradians.
The described feature is clearly visible for the pair of dislocations indicated by the red arrow. It is also observable for several other, similar, pairs of dislocations in Figure 3. The sign of α reverses when the sign of both Burgers vectors is inverted in the region between the dislocations (blue-yellow-blue vs. yellow-blue-yellow in Figure 3). However, for the two cases, the modulus of α decreases when the distance D between the dislocations increases.

4. Discussion

The dislocations we observe display a similar contrast and are nearly parallel to the c-axis. The threading dislocations parallel to the c-axis in 4H-SiC that have been reported [5,6] are either screw dislocations or edge dislocations displaying a Burgers vector along 11 2 ¯ 0 . We discard the screw dislocation case because it is such that h.b = 0, h being the diffraction vector (perpendicular to the ( 40 4 ¯ 0 ) planes) and b the Burgers vector (parallel to c), and therefore, these dislocations should not be observed because they produce nearly no contrast in the image [16]. The Burgers vectors of the dislocations we observe display an edge component. These dislocations are such that they produce, on the ( 40 4 ¯ 0 ) lattice planes, an angular misorientation towards the c-axis, as indicated in Figure 4.
Figure 5 schematically shows the dislocation lines (violet) and the tilt of the ( 40 4 ¯ 0 ) lattice planes in the region between dislocation lines of opposite sign. The intersection of the scattering plane (the ( 1 2 ¯ 10 ) lattice planes of the crystal matrix) with the distorted ( 40 4 ¯ 0 ) lattice planes is shown as green lines. In the region between the pair of dislocations, these lines are rotated by an angle α with respect to the matrix. For simplicity, for the present discussion, we assume that the weak variation of the angle α within this region can be neglected and that α can be approximated by a constant, average angle, as indicated on the right side of Figure 5. Subgrain boundary-like regions (red rectangles in Figure 5) are generated at the boundaries between the regions where the ( 40 4 ¯ 0 ) lattice planes exhibit a rotation and the matrix.
Let Ed be the energy per unit length (comprising core and elastic energy) of the dislocation. When the distance D between the dislocations is large enough, the interaction between dislocations can be neglected, and the misorientation angle α in the region between the dislocations decays back to zero. In this case, the energy of the dislocation pair is simply the sum of the energies of the two dislocations 2 · t · E d , where t is the thickness of the crystal, i.e., also the length of the dislocation in the crystal. In the case of the pair of dislocations indicated by the red arrow in Figure 2 and Figure 3, the energy of the dislocations is smaller than 2 · t · E d because there is no return to the matrix orientation in between them. This is shown schematically in Figure 6, which shows a truncated distortion field for the dislocations. A zone in between the two dislocations retains a misorientation α instead of going to zero. We therefore suggest that the energy of the dislocations reduces to k · t · E d instead of 2 · t · E d , with k < 2.
The energy of the dislocations in the case indicated by the red arrow (Figure 2 and Figure 3) can therefore be expressed as k · t · E d , with 1 < k < 2. However, another energy, the one of the two subgrain boundary-like regions, indicated in Figure 5 by the red rectangular volumes, has to be added, in this simple model approximation, to determine the total energy of this configuration.
This total energy can therefore be written as E = k · t · E d + 2 · t · ( D 2 r 0 ) × γ , where γ is the energy per unit surface of the subgrain boundary-like volume. The expressions of γ and Ed are known [2]: γ = G · w · α 2 , and E d = G · b 2 , where G is the elastic shear modulus, and w is the width of the region where the plane orientation varies.
The parameter k is a function of the angle α: we know that k = 2 for α = 0 and that k diminishes if the modulus of α increases. The experiment shows that α is small, i.e., <~10−5 radians for all the cases that can be observed in Figure 3. This suggests a phenomenological approach that assumes a linear variation of k = k(α) as a function of the modulus of α, thus writing k = 2 − z |α|, where z is a proportionality factor for the linearization. We will consider, for simplicity in this discussion, that α is positive.
The shortest distance D observed in Figure 3 is 50 µm, and it corresponds to α = 10 µradians. Let us assume that this pair of dislocations and its (D, α) couple can be considered an “experimental minimum”. By writing dE/dα = 0 for this configuration, we can deduce that z = 4 (D − 2 r0).w.α/b2. The effective misorientation Δωexp of an edge dislocation lying perpendicular to the diffracting planes is expected to vary as a function of the angle between the diffraction and Burgers vectors and to display a maximum given by Δω = 1.75 br0 [16]. Using this formula for the isolated dislocation indicated by the black circle in Figure 3, we can estimate, with Δω~18 µradian and r0~5 µm, that the modulus of the Burgers vector b is ~0.2 nm. This, with w~1.5 µm, allows a rough estimation of z and k that leads to 1 < k < 2, which is consistent with the model.
We will consider, for the D larger distances case, (i) that the value of z remains constant when D varies and (ii) that D >> r0. With these approximations, we can write E = ( 2 z · α ) G · b 2 · t + 2 D · G · w · t · α 2 and, from dE/dα = 0, deduce that the α that minimises the energy for a given D is α~(1/D) (z b2/4w). This means that α should be nearly proportional to 1/D. This statement allows estimation of the expected α for the different D observed in Figure 3 and comparison between them with the experiment. The value of α = 10 µradians, corresponding to D = 50 µm, is the biggest we measured. It is therefore the one where the error is the smallest. We will use this point as a reference for all the other measured (α, D) pairs. The experimental and calculated (α, D) pairs are given in Table 1.
Figure 7 graphically visualises the comparison between the calculated curve and the experiment. It shows that the experimental points (measured in the same way as for the first point on Figure 4) are in fair agreement with the curve resulting from our very simple model.
Let us mention again that this phenomenological approach only applies when dislocations are such that they cannot move easily, i.e., are sessile. Otherwise, it is well-known [2] that dislocations lying in the same glide plane and exhibiting opposite Burgers vectors attract with a force F = G b 2 / 2 π D in such a way that they can annihilate upon meeting. We can notice that as F = dE/dD, within the frame of this phenomenological model that implies E~1/D, the force decays faster than in the standard model.

5. Conclusions

The present capabilities of the RCI experimental setup we developed, which is located at the ESRF, allowed us to show that a long-range distortion and interaction takes place between non-mobile parallel dislocations located a distance D (nearly horizontal in the image) and exhibits opposite Burgers vectors in a 4H-SiC crystal. This distortion is observable for a range of macroscopic distances (D~50–250 µm).
Our phenomenological model rests on a distortion coupling the non-mobile dislocations and allowing a reduction in the total energy. The main physical idea is that the region displaying an average angle α with respect to the matrix and lying between the dislocations with opposite Burgers vectors
(1)
permits the reduction in the extension of the distortion field associated with the dislocations, and consequently the energy of the dislocation pair,
(2)
introduces an extra energy associated with subgrain boundary-like areas.
For each distance D between the dislocations, there is an angle α that minimises the energy of the whole configuration.
The simple model resulting from the above considerations has only one parameter: the average angle α. It leads to a dependence of this average angle α~1/D, which is in fair agreement with the experimental results. This mechanism should modify the attractive force and thus potentially the dislocation dynamics.
This original experiment mainly provides the angular distortion of the reflecting planes along the diffraction vector (vertical for all the presented images) and, as such, opens the way to further diffraction experiments aimed at measuring the complete two-dimensional distortion and at visualising in this way the long-range interactions between dislocations that are not located in nearly the same horizontal plane of the image. This should provide a more complete view of the distortion and lead to further comparisons with more sophisticated theoretical calculations.

Author Contributions

Conceptualization: A.D., T.N.T.C. and J.B.; methodology: T.N.T.C. and C.D.; software: T.N.T.C.; validation: T.N.T.C., C.D. and A.D.; formal analysis: D.C., J.B. and C.D.; investigation: T.N.T.C.; resources: T.N.T.C.; data curation: T.N.T.C.; writing—original draft preparation: J.B.; writing—review and editing: C.D. and J.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the French National Research Agency (ANR) under the “Investissements d’avenir” programs: ANR-10-AIRT-0005 (IRT-NANOELEC).

Institutional Review Board

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data supporting the reported results are stored at the ESRF.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Read, W.T. Dislocations in Crystals; McGraw-Hill Inc.: New York, NY, USA, 1953. [Google Scholar]
  2. Hull, D.; Bacon, D.J. Introduction to Dislocations, 5th ed.; Butterworth-Heinemann: Oxford, UK, 2011. [Google Scholar]
  3. Anderson, P.M.; Hirth, J.P.; Lothe, J. Theory of Dislocations, 3rd ed.; Cambridge University Press: New York, NY, USA, 2017. [Google Scholar]
  4. Dudley, M.; Wang, H.; Wu, F.; Byrappa, S.; Shun, S.; Raghothamachar, B.; Sanchez, E.K.; Chung, G.; Hansen, D.; Mueller, S.G.; et al. Synchrotron Topography Studies of Growth and Deformation-Induced Dislocations in 4H-SiC. In MRS Online Proceedings Library (OPL), Volume 1433: Symposium H–Silicon Carbide 2012—Materials, Processing and Devices; mrss12-1433-h03-03; Materials Research Society: Warrendale, PA, USA, 2012. [Google Scholar] [CrossRef]
  5. Kallinger, B.; Polster, S.; Berwian, P.; Friedrich, J.; Danilewsky, A.N. Experimental verification of the model by Klapper for 4H-SiC homoepitaxy on vicinal substrates. J. Appl. Phys. 2013, 114, 183507. [Google Scholar] [CrossRef] [Green Version]
  6. Harada, S.; Yamamoto, Y.; Seki, K.; Horio, A.; Mitsuhashi, T.; Tagawa, M.; Ujihara, T. Evolution of threading screw dislocation conversion during solution growth of 4H-SiC. APL Mater. 2013, 1, 022109. [Google Scholar] [CrossRef] [Green Version]
  7. Guo, J.; Yang, Y.; Raghothamachar, B.; Dudley, M.; Stoupin, S. Mapping of Lattice Strain in 4H-SiC Crystals by Synchrotron Double-Crystal X-ray Topography. J. Electron. Mater. 2017, 47, 903–909. [Google Scholar] [CrossRef]
  8. Mahadik, N.A.; Das, H.; Stoupin, S.; Stahlbush, R.E.; Bonanno, P.L.; Xu, X.; Rengarajan, V.; Ruland, G.E. Evolution of lattice distortions in 4H-SiC wafers with varying doping. Sci. Rep. Nat. Res. 2020, 10, 10845. [Google Scholar] [CrossRef] [PubMed]
  9. Fujie, F.; Peng, H.; Ailihumaer, T.; Raghothamachar, B.; Dudley, M.; Harada, S.; Tagawa, M.; Ujihara, T. Synchrotron X-ray topographic image contrast variation of screw-type basal plane dislocations located at different depths below the crystal surface in 4H-SiC. Acta Mater. 2021, 208, 116746. [Google Scholar] [CrossRef]
  10. Lübbert, D.; Baumbach, T.; Härtwig, J.; Boller, E.; Pernot, E. μm-resolved high-resolution X-ray diffraction imaging for semiconductor quality control. Nucl. Inst. Meth. Phys. Res. B 2000, 160, 521–527. [Google Scholar] [CrossRef]
  11. Tsoutsouva, M.G.; Amaral De Oliveira, V.; Baruchel, J.; Camel, D.; Marie, B.; Lafford, T. Characterisation of defects in mono-like silicon for photovoltaic applications using X-ray Bragg diffraction imaging. J. Appl. Cryst. 2015, 48, 645–654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Yildirim, C.; Gout, E.; Pagot, A.; Tran Thi Caliste, T.N.; Baruchel, J.; Brellier, D.; Ballet, P. Bragg Diffraction Imaging of Cd Zn Te Single Crystals. J. Electron. Mater. 2020, 49, 4550–4556. [Google Scholar] [CrossRef]
  13. Simons, H.; Jakobsen, A.C.; Rosenlund Ahl, S.; Poulsen, H.F.; Pantleon, W.; Chu, Y.-H.; Detlefs, C.; Valanoor, N. Non-destructive mapping of long-range dislocation strain fields in an epitaxial complex metal oxide. Nano Lett. 2019, 19, 1445–1450. [Google Scholar] [CrossRef] [PubMed]
  14. Kutsal, M.; Bernard, P.; Berruyer, G.; Cook, P.K.; Hino, R.; Jakobsen, A.C.; Ludwig, W.; Ormstrup, J.; Roth, T.; Simons, H.; et al. The ESRF dark-field x-ray microscope at ID06. IOP Conf. Ser. Mater. Sci. Eng. 2019, 580, 012007. [Google Scholar] [CrossRef]
  15. Thi, T.N.T. RCIA Software. Available online: https://gitlab.com/l_sim/scripts/rcia (accessed on 1 September 2021).
  16. Authier, A. Dynamical Theory of X-ray Diffraction; Oxford University Press: Oxford, UK, 2001. [Google Scholar] [CrossRef]
Figure 1. RCI technique: (a) the sample is in Bragg position for a chosen reflection; the diffracted intensity is recorded by a pixelated detector. (b) A series of images (~200) is recorded for different angular positions (rotation ω) along the rocking curve corresponding to the whole sample; a few representative images along the rocking curve of the 4H-SiC crystal are displayed. (c) RCI maps generated from the rocking curves recorded on each of the pixels of the detector show the variation across the crystal of the integrated intensity, peak width and peak position (in ω).
Figure 1. RCI technique: (a) the sample is in Bragg position for a chosen reflection; the diffracted intensity is recorded by a pixelated detector. (b) A series of images (~200) is recorded for different angular positions (rotation ω) along the rocking curve corresponding to the whole sample; a few representative images along the rocking curve of the 4H-SiC crystal are displayed. (c) RCI maps generated from the rocking curves recorded on each of the pixels of the detector show the variation across the crystal of the integrated intensity, peak width and peak position (in ω).
Applsci 11 09054 g001
Figure 2. RCI peak width (FWHM) map of the 4H-SiC crystal. The red dots are the images of dislocations perpendicular to the surface. The red arrow indicates a pair of dislocations connected by an area exhibiting a higher FWHM than the matrix. The projection of the diffraction vector h, the scale, and the vertical z and horizontal y directions are indicated on the left side.
Figure 2. RCI peak width (FWHM) map of the 4H-SiC crystal. The red dots are the images of dislocations perpendicular to the surface. The red arrow indicates a pair of dislocations connected by an area exhibiting a higher FWHM than the matrix. The projection of the diffraction vector h, the scale, and the vertical z and horizontal y directions are indicated on the left side.
Applsci 11 09054 g002
Figure 3. RCI peak position map. The black circle indicates an isolated dislocation where the peak shift Δω has opposite signs on the left and right of the dislocation core. A similar pattern is observed for other isolated dislocations, the sign of Δω being reversed (blue-yellow or yellow-blue) as a function of the sign of the Burgers vector of the dislocation. The red arrow indicates the pair of dislocations already pointed out in Figure 2. They display opposite Burgers vectors because the sign of their peak position variation with respect to the matrix is reversed: when going from left to right along the y direction, the change is yellow to blue for the first dislocation and blue to yellow for the second one. Note that the peak shift in the region between the dislocation cores does not decay to the value of the surrounding matrix.
Figure 3. RCI peak position map. The black circle indicates an isolated dislocation where the peak shift Δω has opposite signs on the left and right of the dislocation core. A similar pattern is observed for other isolated dislocations, the sign of Δω being reversed (blue-yellow or yellow-blue) as a function of the sign of the Burgers vector of the dislocation. The red arrow indicates the pair of dislocations already pointed out in Figure 2. They display opposite Burgers vectors because the sign of their peak position variation with respect to the matrix is reversed: when going from left to right along the y direction, the change is yellow to blue for the first dislocation and blue to yellow for the second one. Note that the peak shift in the region between the dislocation cores does not decay to the value of the surrounding matrix.
Applsci 11 09054 g003
Figure 4. Plot of the angular misorientation Δω between the region lying in between the dislocations and the matrix. The inset shows the left-down quarter of Figure 3, and the investigated area is shown as a blue rectangle. The different curves correspond to averaging over distances, perpendicular to line joining the core of the dislocations, going from ~3 to 20 µm. The region in dark blue is ~20 µm wide, and the corresponding angle α (see text) is ~10 µrad.
Figure 4. Plot of the angular misorientation Δω between the region lying in between the dislocations and the matrix. The inset shows the left-down quarter of Figure 3, and the investigated area is shown as a blue rectangle. The different curves correspond to averaging over distances, perpendicular to line joining the core of the dislocations, going from ~3 to 20 µm. The region in dark blue is ~20 µm wide, and the corresponding angle α (see text) is ~10 µrad.
Applsci 11 09054 g004
Figure 5. Schematic drawing indicating the angular deviation of the Bragg diffracting planes (schematised by green lines) in the region lying between the dislocations (violet lines) with opposite Burgers vectors. This region is misoriented by an average angle α with respect to the matrix. The direction of the diffraction vector h is indicated. On the right side of the figure we show the approximation we use for the discussion, where we consider an angle α constant over the volume lying between the subgrain boundary-like regions, represented by the rectangular volumes in red.
Figure 5. Schematic drawing indicating the angular deviation of the Bragg diffracting planes (schematised by green lines) in the region lying between the dislocations (violet lines) with opposite Burgers vectors. This region is misoriented by an average angle α with respect to the matrix. The direction of the diffraction vector h is indicated. On the right side of the figure we show the approximation we use for the discussion, where we consider an angle α constant over the volume lying between the subgrain boundary-like regions, represented by the rectangular volumes in red.
Applsci 11 09054 g005
Figure 6. Schematic representation of the distortion of the Bragg reflecting planes associated with two parallel dislocations with opposite Burgers vectors and distant from D. A zone in between the two dislocations retains a misorientation α instead of going down to zero (red straight line instead of continuation of the decreasing curve). The energy of the dislocations is associated with the zones indicated by the green rectangles and reduces to k · t · E d instead of 2 · t · E d , with k < 2.
Figure 6. Schematic representation of the distortion of the Bragg reflecting planes associated with two parallel dislocations with opposite Burgers vectors and distant from D. A zone in between the two dislocations retains a misorientation α instead of going down to zero (red straight line instead of continuation of the decreasing curve). The energy of the dislocations is associated with the zones indicated by the green rectangles and reduces to k · t · E d instead of 2 · t · E d , with k < 2.
Applsci 11 09054 g006
Figure 7. The dashed curve corresponds to |α| proportional to 1/D, with the point D = 50 µm, α = 10 µradians being, by construction, on the curve. The circles are the measured modulus of α for five pairs of dislocations with opposite Burgers vectors that we can observe in Figure 3.
Figure 7. The dashed curve corresponds to |α| proportional to 1/D, with the point D = 50 µm, α = 10 µradians being, by construction, on the curve. The circles are the measured modulus of α for five pairs of dislocations with opposite Burgers vectors that we can observe in Figure 3.
Applsci 11 09054 g007
Table 1. Reported α values for different distance D, as measured in Figure 3 and as calculated by the formula in the text.
Table 1. Reported α values for different distance D, as measured in Figure 3 and as calculated by the formula in the text.
Distance D (µm)Calculated α (µrad)Experimental |α| (µrad)
501010
1054.73.8
1403.64.4
2002.53.3
2402.12.0
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Caliste, T.N.T.; Drouin, A.; Caliste, D.; Detlefs, C.; Baruchel, J. Rocking Curve Imaging Investigation of the Long-Range Distortion Field between Parallel Dislocations with Opposite Burgers Vectors. Appl. Sci. 2021, 11, 9054. https://doi.org/10.3390/app11199054

AMA Style

Caliste TNT, Drouin A, Caliste D, Detlefs C, Baruchel J. Rocking Curve Imaging Investigation of the Long-Range Distortion Field between Parallel Dislocations with Opposite Burgers Vectors. Applied Sciences. 2021; 11(19):9054. https://doi.org/10.3390/app11199054

Chicago/Turabian Style

Caliste, Thu Nhi Tran, Alexis Drouin, Damien Caliste, Carsten Detlefs, and José Baruchel. 2021. "Rocking Curve Imaging Investigation of the Long-Range Distortion Field between Parallel Dislocations with Opposite Burgers Vectors" Applied Sciences 11, no. 19: 9054. https://doi.org/10.3390/app11199054

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop