24.08.2013 Views

Diploma Thesis - Erich Schmid Institute

Diploma Thesis - Erich Schmid Institute

Diploma Thesis - Erich Schmid Institute

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Thermal Behaviour of Al/Si(0 0 1) and GaN/BGaN/Al2O3(0 0 0 1)<br />

structures characterized using X-ray diffraction<br />

<strong>Diploma</strong> <strong>Thesis</strong><br />

submitted by<br />

Hafok Martin<br />

<strong>Erich</strong> <strong>Schmid</strong> Institut für Materialwissenschaft<br />

Leoben, March 2004


Danksagung<br />

Diese Arbeit entstand im Zeitraum September 2003 bis März 2004, wobei sämtliche<br />

Messungen am <strong>Erich</strong> <strong>Schmid</strong> Institut für Materialwissenschaft in Leoben durchgeführt<br />

wurden.<br />

Das Schreiben dieser Arbeit wäre ohne die professionelle Hilfe und Unterstützung von<br />

Herrn Dr. DI. Jozef Keckes nicht möglich gewesen. Ich verdanke ihm, dass er mir die<br />

Methoden der Spannungsmessung mittels Röntgendiffraktion sowie deren<br />

Auswertung und Interpretation näher gebracht hat. Ebenso möchte ich seinen Einsatz<br />

und guten Rat, der mir bei der Lösung vieler Probleme im Rahmen dieser Arbeit<br />

geholfen hat, hervorheben sowie dass er für mich und meine Anliegen immer Zeit<br />

gefunden hat. Aus all diesen Gründen will ich mich herzlich bedanken.<br />

Weiters möchte ich auch Herrn Ao. Univ. Prof. Balder Ortner meinen Dank für seine<br />

Ratschläge und Bemühungen aussprechen.<br />

Ein weiterer großer Dank gilt Herrn DI. Ernst Eiper, der mir bei der schwierigen<br />

Messung der einkristallinen GaN/BGaN/Saphire Probe behilflich war und mir bei der<br />

Korrektur der Arbeit eine wertvolle Hilfestellung gegeben hat. Ebenso möchte ich mich<br />

auch bei Stefan Massl bedanken, der mir bei der Überarbeitung des Textes geholfen<br />

hat.<br />

Zu guter letzt möchte ich meiner Familie meinen Dank aussprechen, die mich in<br />

meinem Bestreben stets unterstützt hat und mir immer einen wertvollen Rückhalt<br />

geboten hat.


Content:<br />

i<br />

Content<br />

1. Motivation.............................................................................................................1<br />

2. Introduction ..........................................................................................................2<br />

3. Basic X-ray diffraction expressions ......................................................................3<br />

3.1. Reciprocal lattice ......................................................................................3<br />

3.2. Bragg equation .........................................................................................4<br />

4. Mechanical properties of crystals .........................................................................6<br />

4.1. Stress, strain and displacement ...............................................................6<br />

4.2. Anisotropic Elasticity.................................................................................7<br />

4.3. Intrinsic and extrinsic stress .....................................................................9<br />

5. Orthogonal tensors and rotation.........................................................................11<br />

5.1. Basic relations ........................................................................................11<br />

5.2. Rotation in a plane..................................................................................12<br />

5.3. Euler angles ...........................................................................................13<br />

6. Basic Physical Properties of layer and substrate materials ................................15<br />

6.1. Polycrystalline aluminium on monocrystalline silicon .............................15<br />

6.1.1. Properties of aluminium layer .................................................................15<br />

6.1.2. Properties of silicon substrate ................................................................16<br />

6.2. Monocrystalline GaN on monocrystalline sapphire.................................17<br />

6.2.1. GaN........................................................................................................17<br />

6.2.2. Boron nitride ...........................................................................................21<br />

6.2.3. Sapphire substrate .................................................................................22<br />

7. Polycrystalline and monocrystalline Model.........................................................24<br />

7.1. Calculating isotropic elastic constants....................................................24<br />

7.1.1. Isotropic material ....................................................................................24<br />

7.1.2. Voigt model ............................................................................................24<br />

7.1.3. Reuss model: .........................................................................................27<br />

7.1.4. Hill model................................................................................................28<br />

7.2. Calculating stresses for a single crystalline material ..............................29<br />

7.3. Strain evaluation.....................................................................................31


ii<br />

Content<br />

8. Deposition of Thin Films.....................................................................................33<br />

8.1. Magnetron sputtering of polycrystalline Al thin films on Si(1 0 0) ...........33<br />

8.2. Molecular beam epitaxy of GaN/BGaN on Al2O3(0 0 0 1).......................34<br />

9. X-ray Diffraction – Measurements and Alignment ..............................................36<br />

9.1. Four Circle Goniometer ..........................................................................36<br />

9.2. DHS 900 Domed Hot Stage ...................................................................37<br />

9.3. The High-Resolution Monochromator.....................................................38<br />

9.4. Alignment of the diffractometer – point focus .........................................41<br />

9.5. Alignment of the diffractometer – line focus............................................42<br />

10. Aluminium on silicon measurement....................................................................45<br />

10.1. Experiment .............................................................................................45<br />

10.2. Shift of diffraction peaks .........................................................................46<br />

10.3. Peak broadening with increasing ψ tilt....................................................48<br />

11. GaN and GaBN on sapphire measurement........................................................50<br />

11.1. Measuring with the high resolution monochromator ...............................50<br />

11.2. Stereographic projections and crystal orientation...................................51<br />

11.3. Phi adjustment........................................................................................55<br />

11.4. Omega adjustment .................................................................................57<br />

11.5. Theta scans:...........................................................................................58<br />

12. Results of aluminium on silicon ..........................................................................61<br />

12.1. Sin(ψ)² vs. a plot.....................................................................................61<br />

12.2. Lattice spacing of aluminium layer .........................................................64<br />

12.3. Thermal expansion coefficient of aluminium...........................................65<br />

12.4. Lattice spacing of silicon substrate.........................................................66<br />

12.5. Thermal expansion coefficient of silicon substrate .................................66<br />

12.6. Stress curve ...........................................................................................67<br />

12.7. Strain curves ..........................................................................................69<br />

12.8. Discussion ..............................................................................................70<br />

13. Results of GaN/GaBN/Al2O3(0 0 0 1) .................................................................74<br />

13.1. Sapphire lattice parameters....................................................................74<br />

13.2. Thermal expansion coefficients of sapphire ...........................................75<br />

13.3. In-plane stress in GaN and GaBN layer: ................................................75<br />

13.4. Discussion: .............................................................................................81


iii<br />

Content<br />

14. Conclusion and Outlook .....................................................................................82<br />

15. Literature ............................................................................................................83<br />

16. Appendix ............................................................................................................87<br />

16.1. Serial port communication ......................................................................87<br />

16.2. Stereographic projection of Silicon (0 0 1)..............................................88<br />

16.3. Stereographic projection of gallium nitride (0 0 1) ..................................89<br />

16.4. Stereographic projection of sapphire (0 0 1)...........................................90<br />

16.5. GaN stress evaluation written in Mathematica: ......................................91


1. Motivation<br />

Motivation<br />

Virtually all types of thin films are expected to contain some amount of residual strain<br />

decisively influencing their mechanical behaviour and, secondary, modifying band-<br />

gaps in semiconductors, transition temperatures in superconductors, magnetic<br />

anisotropy, wear resistance or other important physical parameters. The strains can<br />

be formed unintentionally as an unavoidable consequence of the deposition process<br />

or intentionally in order to control required parameters of devices.<br />

The residual stresses represent very important issue especially in the case of modern<br />

electronics packages representing nowadays complicated composites of<br />

semiconductors, metals, dielectrics and plastics with specific thermal expansion<br />

coefficients, manufacturing temperatures and geometry. For the fabrication as well as<br />

for the practical application of such structures, the control of residual stresses has<br />

turned out to be of utmost importance.<br />

For the production of interconnects in microelectronic chips, aluminium and copper<br />

have been used. When the interconnects are thermally cycled during operation,<br />

various micro-structural effects can occur including grain growth, diffusion, plastic<br />

flow, electromigration etc. All these phenomena are influenced and partly also<br />

controlled by the magnitude of residual stresses in the metals.<br />

On the other hand, residual stresses in semiconductors directly influence important<br />

optical and electronic properties through the deformation of crystal lattice and<br />

subsequently the modification of the band gap parameters. The nitride<br />

semiconductors including the family of refractory materials like indium nitride,<br />

aluminium nitride and especially gallium nitride possess a significant potential for<br />

optoelectronic and piezoelectric applications. The presence of high compressive<br />

residual stresses in nitride-based thin films, however, influence not only the optical<br />

properties but is responsible also for crack formation – a very serious problem in<br />

nitride technology.<br />

The characterization of residual stresses in thin films represent thus a very important<br />

issue for nowadays technology. The main aim of this thesis to perform elevated-<br />

temperature X-ray diffraction characterization of residual stresses in aluminium thin<br />

films and in BGaN/GaN structure focusing micro-structural changes and phenomena<br />

related to intrinsic and extrinsic stresses.<br />

1


2. Introduction<br />

Introduction<br />

Already in ancient times, materials were tested for their reliability. For example the<br />

bending of swords to proof their elasticity and the tapping on ceramic vessels, like<br />

amphora, to detect defects were wide spread testing methods at that time. The first<br />

systematic examinations of materials properties and their quantification have been<br />

reported since the middle age, where this knowledge served for the constructions of<br />

buildings as well as to improve shipbuilding. With the appearance of the first<br />

industrial steam engines in the 19 th century, the demands not only on the material but<br />

also on adequate testing methods raised. Since that time, many important testing<br />

methods have been developed and utilized till nowadays especially for bulk<br />

materials. With the development of microelectronics and with the application of<br />

integrated circuits, new testing methods have been introduced allowing a significant<br />

progress in the miniaturization of the thin film – based structures.<br />

Today’s increasing performance of electronic devices is accompanied by a higher<br />

energy consumption stimulating requirements for the cooling. The heat dissipation<br />

influences not only the electrical, optical and magnetic properties of the devices but<br />

also the magnitude of internal stresses. The presence of residual stresses in thin<br />

films and in sublayers of sandwich structures can not be underestimated due to their<br />

direct influence on all basic physical as well as on mechanical properties of the<br />

devices. Especially in the context of the electromigration in copper or in aluminium<br />

interconnects, the residual stresses represent a very important issue.<br />

Up to now, stresses in thin films have been analyzed predominantly ex situ using<br />

X-ray diffraction, Raman and photoluminescence spectroscopy, the wafer curvature<br />

method, and high resolution transmission electron microscopy. The main practical<br />

advantage of the curvature technique resides in the application for new materials with<br />

unknown elastic constants. The diffraction techniques, on the other hand, are<br />

capable of resolving anisotropic deformation of crystal lattice, thin film behaviour on<br />

anisotropic substrates and stresses even in multilayered structures. Recently, a<br />

significant attention was devoted to the studying of residual stress origins in thin<br />

films. In this case, the elevated-temperature X-ray diffraction provided in important<br />

results leading to the understanding the role of intrinsic-stresses and their formation<br />

in thin films.<br />

Within this thesis structural properties of polycrystalline Al thin film deposited on<br />

Si(1 0 0) substrate and the properties of GaN/BGaN multilayers deposited on c plane<br />

sapphire are studied using elevated-temperature X-ray diffraction technique<br />

implemented recently at <strong>Erich</strong> <strong>Schmid</strong> <strong>Institute</strong> for Materials Science in Leoben. For<br />

the studies of GaN/BGaN/Al2O3(0 0 0 1) structures, a high-resolution monochromator<br />

was used.<br />

2


3. Basic X-ray diffraction expressions<br />

3.1. Reciprocal lattice<br />

Basic X-ray diffraction<br />

A mathematical formalism will be defined to describe scattering phenomena on<br />

crystal structures with translation symmetry [1]. Supposing the crystal periodicity, the<br />

selection of available functions is reduced. A periodical function alone, like sinus or<br />

cosines, is unable to describe the electron density distribution and phase<br />

phenomenon, the basic attributes of X-ray diffraction effect. Fourier series can be<br />

applied to describe the scattering effect.<br />

n<br />

G<br />

= ∫ N(<br />

r)<br />

exp( −i<br />

G . r)<br />

dV<br />

(equ. 3.1)<br />

cell<br />

∑<br />

n ( r ) = nG<br />

exp( i G . r)<br />

(equ. 3.2)<br />

G<br />

The imaginary unit, square root of minus one, is expressed by i. N(r) defines the<br />

electron density within one unit cell while, on the other hand, n(r) denotes the<br />

electron density of the whole crystal. The two vectors r and G in equation 3.1 and 3.2<br />

that appear in the exponential term can be expressed as:<br />

r = x a + y a + z a<br />

(equ. 3.3)<br />

1<br />

1<br />

2<br />

2<br />

3<br />

G = h b + k b + l b<br />

(equ. 3.4)<br />

The summation over G in equation 3.2 should be understood as a summations over<br />

h, k and l from minus infinity to plus infinity. The relation between ai and bi vectors<br />

can be found by calculating the scalar product of G and r. According to the Fourier<br />

series the scalar product must have the following form.<br />

3<br />

G . r = 2π<br />

(h x + k y + l z)<br />

(equ. 3.5)<br />

Considering the equation 3.5, the scalar product of a1 and b2 or a1 and b3 is zero, so<br />

a system of linear equations can be written as:<br />

3


⎛ a1<br />

. b1<br />

⎜<br />

⎜a<br />

2 . b1<br />

⎜<br />

⎝a<br />

3 . b1<br />

The solution of this system is:<br />

a<br />

a<br />

a<br />

1<br />

2<br />

3<br />

b<br />

b<br />

b<br />

. b<br />

. b<br />

2<br />

. b<br />

1<br />

2<br />

3<br />

2<br />

2<br />

a1<br />

. b3<br />

⎞ ⎛1<br />

⎟ ⎜<br />

a2<br />

. b3<br />

⎟ = 2π<br />

⎜0<br />

a ⎟ ⎜<br />

3 . b3<br />

⎠ ⎝0<br />

a2<br />

× a3<br />

= 2π<br />

a . ( a × a )<br />

1<br />

a3<br />

× a1<br />

= 2π<br />

a . ( a × a )<br />

1<br />

a1<br />

× a2<br />

= 2π<br />

a . ( a × a )<br />

1<br />

2<br />

2<br />

2<br />

3<br />

3<br />

3<br />

0<br />

1<br />

0<br />

0⎞<br />

⎟<br />

0⎟<br />

1⎟<br />

⎠<br />

Basic X-ray diffraction<br />

The vectors b1, b2 and b3 of equations 3.7 are called reciprocal lattice vectors.<br />

3.2. Bragg equation<br />

(equ. 3.6)<br />

(equ. 3.7)<br />

The Bragg equation represents a relatively simple way to describe the scattering<br />

phenomenon on a crystal lattice. Consider an incident beam that is reflected by a<br />

family of net planes. It’s worth to mention that the path difference between<br />

neighbouring net planes causes a phase difference between the diffracted beams, so<br />

the reflected radiation shows constructive and destructive interference. The Bragg<br />

equation, formula 3.8, predicts that constructive interference only occurs if the ratio<br />

between the phase difference and wavelength is an integer value of n.<br />

2 d sin( θ)<br />

n =<br />

(equ. 3.8)<br />

λ<br />

The variable θ is the angle between the incident or the reflected beam and the net<br />

plane. Because elastic scattering is assumed, the incident and reflected beam must<br />

have same wavelength λ. The last parameter is the lattice spacing d, the shortest<br />

distance between two neighbouring net planes with same Miller indices (h k l) and the<br />

surface normal G that is inversely proportional to the plane spacing d.<br />

π<br />

=<br />

G<br />

2<br />

d (equ. 3.9)<br />

4


Basic X-ray diffraction<br />

Equation 3.9 is valid for all crystal systems, and the components of G can be<br />

replaced by the points of intersection u1, u2 and u3 of the net plane with the elongated<br />

translation vectors a1, a2 and a3.<br />

Figure 3.1: Crystallographic plane and surface normal<br />

The components of G are for that reason:<br />

1<br />

h =<br />

u1<br />

1<br />

k =<br />

u2<br />

1<br />

l = (equ. 3.10)<br />

u<br />

3<br />

5


4. Mechanical properties of crystals<br />

4.1. Stress, strain and displacement<br />

Mechanical properties of crystals<br />

Internal stresses are produced by external forces acting on a body [2]. These<br />

external forces can be separated into two categories, like the distribution of forces<br />

over the surface, such as hydrostatic pressure, and distributed forces over the<br />

volume, for example gravitational forces or magnetic forces. A motion of the body is<br />

also able to influence the internal stresses. The most important expression to<br />

describe stresses can be derived by assuming the conservation of impulse.<br />

ρ a = div(<br />

σ)<br />

+ f<br />

(equ. 4.1)<br />

Where ρ is the density of the body, a is the acceleration, f denotes the distributed<br />

forces over the volume and σ, a second ranked tensor, represents the internal<br />

stresses.<br />

σ<br />

σ n<br />

n = (equ. 4.2)<br />

The distributed forces or stresses over a surface are written in equation 4.2 that is the<br />

product between the stress tensor and the surface normal vector n, and can be<br />

understood as a boundary condition for equation 4.1. In the case of stress<br />

measurement on thin films, which is a static problem, the acceleration is zero and the<br />

volume forces are neglected.<br />

div ( σ ) = 0<br />

(equ. 4.3)<br />

A body under stresses will deform, this means that a point P of the unstressed body<br />

moves to the position P’. The vector connecting P and P’ is called displacement u.<br />

We consider a short and a long rod with same cross section and made of the same<br />

material. Let us say that the change of length is the displacement, than one can see<br />

that the displacement of the longer rod is greater than the displacement of the small<br />

rod, if the same stress is applied at the ends of both rods. An assessment based on<br />

strains to characterize the deformation will show same results for both rods. This<br />

means that the influence of geometry does not play a role.<br />

6


ε<br />

ij<br />

=<br />

1 ⎛<br />

⎜<br />

∂ ui<br />

2 ⎜<br />

⎝ ∂x<br />

j<br />

∂ u j ⎞<br />

+ ⎟<br />

∂x<br />

⎟<br />

i ⎠<br />

Mechanical properties of crystals<br />

(equ. 4.4)<br />

Equation 4.4 defines the strain tensor for small displacements. A closer look on the<br />

strain tensor reveals the symmetric property. It was not mentioned above but the<br />

stress tensor is also symmetric.<br />

4.2. Anisotropic Elasticity<br />

ε = ε<br />

(equ. 4.5)<br />

ij<br />

ij<br />

ji<br />

σ = σ (equ. 4.6)<br />

Further equations are needed to solve a general mechanical problem, because if all<br />

unknown quantities of the previous chapter are count together the total sum will be<br />

15, in detail there are six stresses, just as much strains and additionally three<br />

displacements. On the other side we have three expressions to describe stresses<br />

(equation 4.1 or 4.3) and six formulas of strain/displacement relations (equation 4.4),<br />

so remains a lack of six equations.<br />

The missing equations are based on the mechanical behaviour of a material that is<br />

assumed to be elastic. In literature a material is often treated in an isotropic way. It<br />

means that a property like the electrical resistant or thermal expansion coefficient is<br />

not depending on the materials direction. However, a crystal is anisotropic [3], and<br />

therefore a general expression will define the elastic stresses/strains relations.<br />

ij<br />

ijkl<br />

ji<br />

σ = c ε<br />

(equ. 4.7)<br />

ij<br />

ijkl<br />

kl<br />

ε = s σ<br />

(equ. 4.8)<br />

The stiffness tensor cijkl of equation 4.7 is the relation between stress and strain<br />

tensor, both of second rank including nine components, thus the stiffness tensor must<br />

consist of 81 elements. Very similar to the stiffness is the compliance tensor sijkl in<br />

equation 4.8. It is simple to find a relation between both expressions.<br />

−1<br />

c = s<br />

kl<br />

−1<br />

s = c<br />

(equ. 4.9)<br />

7


Mechanical properties of crystals<br />

Because of the anisotropic material behaviour, with unequal properties in different<br />

directions, the position of the crystal system with reference to the sample is of<br />

interest. Generally the crystal and sample system do not coincide due to angular<br />

differences. In order to express the anisotropic behaviour in a sample system, the<br />

compliance or stiffness tensors must be rotated.<br />

A ' = O O A<br />

(equ. 4.10)<br />

For a fourth ranked tensor the expression has the following form:<br />

ijkl<br />

ij<br />

im<br />

ik<br />

jn<br />

jl<br />

ko<br />

kl<br />

A ' = O O O O A<br />

(equ. 4.11)<br />

In literature a second preferred way to express the relation between stresses and<br />

strains according to W. Voigt [4] is often used. By applying the matrix notation, the<br />

stress and strain tensors are reduced to vectors with six components, thus the<br />

compliance and stiffness matrices have 36 components. A comparison between<br />

stresses and strains written in matrix notation and tensor notation shows:<br />

⎛ ε<br />

⎜<br />

⎜ε<br />

⎜<br />

⎝ ε<br />

⎛ σ<br />

⎜<br />

⎜σ<br />

⎜<br />

⎝ σ<br />

11<br />

12<br />

13<br />

11<br />

12<br />

13<br />

ε<br />

ε<br />

ε<br />

12<br />

22<br />

23<br />

σ<br />

σ<br />

σ<br />

12<br />

22<br />

23<br />

ε<br />

ε<br />

ε<br />

13<br />

23<br />

33<br />

σ<br />

σ<br />

σ<br />

13<br />

23<br />

33<br />

⎞ ⎛ σ<br />

⎟ ⎜<br />

⎟ = ⎜σ<br />

⎟ ⎜<br />

⎠ ⎝ σ<br />

⎛<br />

⎜ ε1<br />

⎞<br />

⎟<br />

⎜<br />

⎜ 1<br />

⎟ = ε<br />

⎟<br />

⎜ 2<br />

⎠ ⎜ 1<br />

⎜ ε<br />

⎝ 2<br />

6<br />

5<br />

1<br />

6<br />

5<br />

lp<br />

1<br />

2<br />

σ<br />

σ<br />

σ<br />

ε<br />

1<br />

2<br />

ε<br />

2<br />

ε<br />

6<br />

2<br />

4<br />

mnop<br />

6<br />

4<br />

σ5<br />

⎞<br />

⎟<br />

σ4<br />

⎟<br />

σ ⎟<br />

3 ⎠<br />

1<br />

ε<br />

2<br />

1<br />

ε<br />

2<br />

ε<br />

3<br />

5<br />

4<br />

⎞<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎠<br />

(equ. 4.12)<br />

(equ. 4.13)<br />

The first two and last two suffixes of the tensor notation are merged to a single<br />

number running from one to six, according to the scheme:<br />

Tensor<br />

notation<br />

Matrix<br />

notation<br />

11 22 33 23, 32 31, 13 12, 21<br />

1 2 3 4 5 6<br />

8


The relations between the compliance tensor and matrix are:<br />

sijkl = smn when m and n are 1, 2 or 3.<br />

2 sijkl = smn when either m or n are 4, 5 or 6.<br />

4 sijkl = smn when both m and n are 4, 5 or 6.<br />

Mechanical properties of crystals<br />

The advantage of working with matrix notation is the compacter formalism, though<br />

the rotation of the compliance and the stiffness matrix is more complicated than the<br />

rotation in the tensor notation. For example an isotropic material can be<br />

characterized by the following stiffness matrix:<br />

⎛ c11<br />

c12<br />

c12<br />

0 0 0 ⎞<br />

⎜<br />

⎟<br />

⎜c12<br />

c11<br />

c12<br />

0 0 0 ⎟<br />

⎜c<br />

⎟<br />

12 c12<br />

c11<br />

0 0 0<br />

c = ⎜<br />

⎟<br />

(equ. 4.14)<br />

⎜ 0 0 0 c44<br />

0 0 ⎟<br />

⎜ 0 0 0 0 c ⎟<br />

44 0<br />

⎜<br />

⎟<br />

⎜<br />

⎟<br />

⎝ 0 0 0 0 0 c44<br />

⎠<br />

Isotropic materials have only two independent components, so the third one can be<br />

calculated:<br />

4.3. Intrinsic and extrinsic stress<br />

1<br />

c44 = ( c11<br />

− c12<br />

)<br />

2<br />

The residual stresses in thin films results from a growth procedure and from a cooling<br />

down to the operation temperature after the deposition with specific intrinsic and<br />

extrinsic stress contributions, respectively. The intrinsic stresses originate from a<br />

specific microstructure development and the film densification during the growth [5].<br />

An atomic disorder caused by foreign atoms integrated into the lattice can serve as<br />

an example of the intrinsic stress origin. An impurity atom can substitute a lattice<br />

atom or it can be found in lattice gaps as an interstitial atom. The compressive/tensile<br />

stress is getting higher if the atomic radius of the impurity atoms is larger/smaller than<br />

the atomic radius of the original lattice atoms. Intrinsic stresses can be reduced by<br />

applying high substrate temperatures during deposition, because of the high mobility<br />

of the atoms reduces the disorder. An additional part of intrinsic stresses occur in<br />

sputtered layers. The incoming sputtered atoms are hitting the layer and this is like<br />

9


Mechanical properties of crystals<br />

shot peening, setting the layer under compressive stress. This effect is known as<br />

atomic peening.<br />

Extrinsic stresses originate from the cooling down procedure and basically depend on<br />

the different constants of thermal expansion of layer and substrate. After cooling or<br />

heating, starting from the deposition temperature TS, the layer material is elastically<br />

or plastically deformed. In the elastic region the extrinsic stress is:<br />

hkl<br />

σ = ( α − α ) ( T − T )<br />

(equ. 4.15)<br />

ex<br />

M l s<br />

S<br />

The thermal expansion coefficient αl belongs to the layer, the other one αs to the<br />

substrate and the quantity M hkl is the biaxial modulus of the layer [6] that is<br />

depending on the crystallographic orientation. The final stresses are the sum of<br />

intrinsic and extrinsic stresses.<br />

Residual stresses are reduced through crack formation under tensile stress,<br />

delamination of the layer or plastic deformation forming hills under compressive<br />

stress.<br />

10


5. Orthogonal tensors and rotation<br />

5.1. Basic relations<br />

Orthogonal tensors and rotation<br />

Crystals are anisotropic bodies, which means that the mechanical behaviour depends<br />

on the crystal orientation. For this reason, it is important to exactly define the<br />

orientation of the crystal in the sample coordinate system. Generally the sample and<br />

the crystal coordinate systems can be related by a rotation (Chapter 4.2) expressed<br />

by an orthogonal tensor O.<br />

We consider two right angled coordinate systems with the unit vectors ei and ei’<br />

(figure 5.1). A unit vector belonging to one coordinate system, for example ei, is<br />

projected on the axes of the other system so that the position of this vector can be<br />

obtained in the new coordinate system ei’ [7]. The length of the projected unit vector<br />

on one axis is:<br />

Figure 5.1: Relation between two tilted coordinate systems<br />

e i ' . e j = ei<br />

' e j cos( ei<br />

',<br />

e j)<br />

e i ' = 1<br />

j 1 = e<br />

e . e = cos( e ',<br />

e )<br />

(equ. 5.1)<br />

i ' j<br />

i j<br />

By applying equation 5.1 on all unit vectors of one coordinate system, an orthogonal<br />

tensor is received that contains only cosine functions. Such a expression is also<br />

known as direction cosine. For example let us use e1 to express e1‘, where the<br />

components of e1‘ and e1 are x1’, y1’, z1’ and x1, y1, z1.<br />

11


⎛x<br />

1'⎞<br />

⎛ cos( e1'<br />

, e1)<br />

⎜ ⎟ ⎜<br />

⎜ y1'⎟<br />

= ⎜cos(<br />

e 2'<br />

, e1)<br />

⎜ ⎟ ⎜<br />

⎝ z1'<br />

⎠ ⎝cos(<br />

e3'<br />

, e1)<br />

5.2. Rotation in a plane<br />

cos( e ',<br />

e )<br />

1<br />

2<br />

3<br />

2<br />

cos( e ',<br />

e )<br />

2<br />

cos( e ',<br />

e )<br />

1<br />

1<br />

2<br />

Orthogonal tensors and rotation<br />

cos( e1'<br />

, e3<br />

) ⎞ ⎛ x1<br />

⎞<br />

⎟ ⎜ ⎟<br />

cos( e 2'<br />

, e3<br />

) ⎟ ⎜ y1<br />

⎟<br />

cos( e ⎟ ⎜ ⎟<br />

3'<br />

, e3<br />

) ⎠ ⎝ z1<br />

⎠<br />

(equ. 5.2)<br />

e ' = O e<br />

(equ. 5.3)<br />

To learn more about how a rotation around a certain axis will affect the position of the<br />

sample, it is necessary to consider a body on which points the vector r, like in figure<br />

5.2. A positive rotation of the body is synonymous with the rotation of the vector r<br />

around the e3 axis and at the end of the operation the vector r coincides with r’.<br />

The new point r’ with the components x’, y’ and z’ can be expressed by the starting<br />

point r:<br />

r ' = O<br />

3<br />

r<br />

⎛cos(<br />

ϕ)<br />

− sin( ϕ)<br />

0⎞<br />

⎜<br />

⎟<br />

3<br />

O = ⎜ sin( ϕ)<br />

cos( ϕ)<br />

0⎟<br />

(equ. 5.4)<br />

⎜<br />

⎟<br />

⎝ 0 0 1⎠<br />

The orthogonal tensors, that are describing a rotation around the two other axes, are<br />

shown in equation 5.5 and 5.6.<br />

Figure 5.2: Rotation of an object<br />

⎛1<br />

0 0 ⎞<br />

⎜<br />

⎟<br />

1<br />

O = ⎜0<br />

cos( ψ)<br />

− sin( ψ)<br />

⎟<br />

(equ. 5.5)<br />

⎜<br />

⎟<br />

⎝0<br />

sin( ψ)<br />

cos( ψ)<br />

⎠<br />

12


Orthogonal tensors and rotation<br />

⎛ cos( ω)<br />

0 sin( ω)<br />

⎞<br />

⎜<br />

⎟<br />

2<br />

O = ⎜ 0 1 0 ⎟<br />

(equ. 5.6)<br />

⎜<br />

⎟<br />

⎝−<br />

sin( ω)<br />

0 cos( ω)<br />

⎠<br />

The body in position r’ can be turned back by a rotation around e3 axis in the<br />

negative direction. Thus the angle’s sign is negative and the body’s starting position<br />

is supposed to be r’.<br />

⎛x<br />

⎞ ⎛ cos( ϕ)<br />

⎜ ⎟ ⎜<br />

⎜ y⎟<br />

= ⎜ − sin( ϕ)<br />

⎜ ⎟ ⎜<br />

⎝ z ⎠ ⎝ 0<br />

sin( ϕ)<br />

cos( ϕ)<br />

0<br />

0⎞<br />

⎛ x'⎞<br />

⎟ ⎜ ⎟<br />

0⎟<br />

⎜ y'⎟<br />

1⎟<br />

⎜ ⎟<br />

⎠ ⎝ z'<br />

⎠<br />

A closer look shows that the orthogonal tensor is the transposed or the inverse<br />

version of O 3 .<br />

5.3. Euler angles<br />

3 T<br />

( O ) r'<br />

r = (equ. 5.7)<br />

In the previous chapter we only thought about a rotation in a plane. In practice more<br />

possibilities are needed to express the tilt between two coordinate systems. One<br />

possibility is the concept of Euler angles [8], that can be understood as three<br />

separated rotations around certain axes, similar like before. The first step is the<br />

rotation around the e3 axis so e1 and e2 will change their position. After that the new<br />

e1 is supposed to be the next rotation axis, which will lead to two new e2 and e3 axes,<br />

and the last step is like the first one.<br />

After the first rotation the new axis is e1.<br />

⎛cos(<br />

ϕ1)<br />

− sin( ϕ1)<br />

0⎞<br />

⎜<br />

⎟<br />

ei<br />

' = ⎜ sin( ϕ1)<br />

cos( ϕ1)<br />

0⎟<br />

e<br />

⎜ 0 0 1⎟<br />

⎝<br />

⎠<br />

⎛1<br />

0 0 ⎞<br />

⎜<br />

⎟<br />

ei<br />

'' =<br />

⎜0<br />

cos( φ)<br />

sin( φ)<br />

⎟ ei<br />

'<br />

⎜0<br />

sin( ) cos( ) ⎟<br />

⎝ − φ φ ⎠<br />

i<br />

13


The last step is the rotation around the new e3 axis.<br />

All equations written in full are:<br />

⎛cos(<br />

ϕ2<br />

) − sin( ϕ2<br />

) 0⎞<br />

⎜<br />

⎟<br />

ei<br />

'' ' = ⎜ sin( ϕ2<br />

) cos( ϕ2<br />

) 0⎟<br />

ei<br />

''<br />

⎜ 0 0 1⎟<br />

⎝<br />

⎠<br />

Orthogonal tensors and rotation<br />

⎛cos(<br />

ϕ2<br />

) − sin( ϕ2<br />

) 0⎞<br />

⎛1<br />

0 0 ⎞ ⎛cos(<br />

ϕ1)<br />

− sin( ϕ1)<br />

0⎞<br />

⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟<br />

ei<br />

' ' ' = ⎜ sin( ϕ2<br />

) cos( ϕ2<br />

) 0⎟<br />

⎜0<br />

cos( φ)<br />

sin( φ)<br />

⎟ ⎜ sin( ϕ1)<br />

cos( ϕ1)<br />

0⎟<br />

e<br />

⎜ 0 0 1⎟<br />

⎜0<br />

sin( ) cos( ) ⎟ ⎜ 0 0 1⎟<br />

⎝<br />

⎠ ⎝ − φ φ ⎠ ⎝<br />

⎠<br />

After the multiplications the orthogonal Euler tensor has the following formula:<br />

e<br />

O<br />

⎛ cos( ϕ1)<br />

cos( ϕ2<br />

) − cos( φ)<br />

sin( ϕ1)<br />

sin( ϕ2<br />

)<br />

⎜<br />

= ⎜−<br />

cos( ϕ1)<br />

sin( ϕ2<br />

) − cos( φ)<br />

sin( ϕ1)<br />

cos( ϕ2<br />

)<br />

⎜<br />

⎝<br />

sin( φ)<br />

sin( ϕ1)<br />

sin( ϕ ) cos( ϕ ) + cos( φ)<br />

cos( ϕ ) sin( ϕ )<br />

1<br />

− sin( ϕ ) sin( ϕ ) + cos( φ)<br />

cos( ϕ ) cos( ϕ )<br />

1<br />

2<br />

2<br />

− sin( φ)<br />

cos( ϕ )<br />

1<br />

1<br />

1<br />

2<br />

2<br />

sin( φ)<br />

sin( ϕ2<br />

) ⎞<br />

⎟<br />

sin( φ)<br />

cos( ϕ2<br />

) ⎟<br />

cos( φ)<br />

⎟<br />

⎠<br />

i<br />

(equ. 5.8)<br />

14


Basic Physical Properties<br />

6. Basic Physical Properties of layer and substrate materials<br />

6.1. Polycrystalline aluminium on monocrystalline silicon<br />

6.1.1. Properties of aluminium layer<br />

Aluminium ore, most commonly bauxite, occurs mainly in tropical and sub-tropical<br />

areas. The raw material bauxite is converted into alumina in the Bayer process, which<br />

is reduced to aluminium metal in electrolytic cells known as pots by adding cryolite [9].<br />

Pure aluminium shows no phase transformation in the solid state and will crystallise at<br />

an equilibrium temperature of about 660°C in a closed packed face centred lattice [10]<br />

with a mono-atomar basis (figure 6.1).<br />

As basis for the stress evaluation the anisotropic elastic behaviour of a cubic material<br />

is represented in equation 6.1, where the low number of independent elastic constants<br />

corresponds to the high symmetry of the cubic system [3, 4].<br />

⎛s11<br />

⎜<br />

⎜s12<br />

⎜s12<br />

s = ⎜<br />

⎜ 0<br />

⎜ 0<br />

⎜<br />

⎝ 0<br />

s<br />

s<br />

s<br />

12<br />

11<br />

12<br />

0<br />

0<br />

0<br />

s<br />

s<br />

s<br />

12<br />

12<br />

11<br />

0<br />

0<br />

0<br />

s<br />

0<br />

0<br />

0<br />

44<br />

0<br />

0<br />

Figure 6.1: Cubic face centred unit cell<br />

s<br />

0<br />

0<br />

0<br />

0<br />

44<br />

0<br />

0 ⎞<br />

⎟<br />

0 ⎟<br />

0 ⎟<br />

⎟<br />

0 ⎟<br />

0 ⎟<br />

⎟<br />

s ⎟<br />

44 ⎠<br />

⎛ c<br />

⎜<br />

⎜c<br />

⎜c<br />

c = ⎜<br />

⎜ 0<br />

⎜ 0<br />

⎜<br />

⎝ 0<br />

11<br />

12<br />

12<br />

c<br />

c<br />

c<br />

12<br />

11<br />

12<br />

0<br />

0<br />

0<br />

c<br />

c<br />

c<br />

12<br />

12<br />

11<br />

0<br />

0<br />

0<br />

c<br />

0<br />

0<br />

0<br />

44<br />

0<br />

0<br />

c<br />

0<br />

0<br />

0<br />

0<br />

44<br />

0<br />

0 ⎞<br />

⎟<br />

0 ⎟<br />

0 ⎟<br />

⎟<br />

0 ⎟<br />

0 ⎟<br />

⎟<br />

c ⎟<br />

44 ⎠<br />

(equ 6.1)<br />

15


Basic Physical Properties<br />

The components of the compliance and stiffness matrix or tensor are not real constant<br />

values due to their temperature dependence. In figure 6.2 the three independent<br />

compliance matrix components of aluminium were plotted [11]. Another characteristic<br />

parameter associated with the temperature influence is the coefficient of thermal<br />

expansion that has a value of 23,8 10 -6 K -1 for aluminium [12].<br />

s 12 / [10 -3 GPa -1 ]<br />

-5<br />

-6<br />

-7<br />

-8<br />

-9<br />

-10<br />

s 11 / [10 -3 GPa -1 ]<br />

24<br />

22<br />

20<br />

18<br />

16<br />

14<br />

34<br />

0 100 200 300 400 500<br />

6.1.2. Properties of silicon substrate<br />

The Czochralski technique is a widespread production method capable of providing<br />

silicon single crystals with a relatively large size. After cutting the silicon single crystals<br />

into thin wafers, followed by surface treatments like lapping, etching and polishing,<br />

they are ready to serve as substrate for thin film deposition.<br />

Silicon has a diamond-like structure [1, 10], which is based on a face centred cubic<br />

lattice with the primitive basis consisting of two atoms at position [0 0 0] and [¼¼¼].<br />

This basis reduces therefore the symmetry but it is not changing the general elastic<br />

mechanical behaviour [13] of the cubic lattice (figure 6.3) described by the compliance<br />

or stiffness matrix (at room-temperature).<br />

T / °C<br />

Figure 6.2: Temperature dependent compliance constants of aluminium<br />

c11 = 165,64 GPa ; c12 = 63,94 GPa ; c44 = 79,51 GPa;<br />

s 11<br />

s 12<br />

s 44<br />

50<br />

48<br />

46<br />

44<br />

42<br />

40<br />

38<br />

36<br />

s 44 / [10 -3 GPa -1 ]<br />

16


Basic Physical Properties<br />

During cooling, starting from the melting point at 1410°C, silicon crystallises in the<br />

diamond structure without showing allotropic transformation in the solid state. The<br />

cooling process is associated with the thermal contraction of the silicon material that<br />

has a thermal expansion coefficient of 2,616 10 -6 K -1 at room temperature [13].<br />

6.2. Monocrystalline GaN on monocrystalline sapphire<br />

6.2.1. GaN<br />

Unfortunately, bulk crystals of nitrides cannot be obtained by conventional methods of<br />

liquid phase epitaxy, because of extremely high melting temperatures and very high<br />

decomposition pressures at the melting point.<br />

Several production techniques are available for growing thin gallium nitride (GaN)<br />

films. These methods can be divided in two groups, one where chemical reactions play<br />

an important role and the other one where only physical process are responsible for<br />

the layer formation. For example a representative of the first group would be<br />

metalorganic vapour deposition (MOCVD) technique or reactive sputtering and for<br />

instance a production method belonging to the second group would be molecular<br />

beam epitaxy (MBE).<br />

Figure 6.3: Diamond structure<br />

17


Basic Physical Properties<br />

Group III nitrides like AlN, GaN and InN can crystallize in wurtzite, zinc-blende and<br />

rock-salt crystal structures. At ambient conditions the thermodynamically stable phase<br />

is the wurtzite structure and only at higher pressures an allotropic transformation<br />

changes the crystal to rock salt structure. The zinc-blende structure is metastable and<br />

may be stabilized by heteroepitaxial growth on substrates reflecting topological<br />

compatibility.<br />

In the preparation phase of the experiment several Bragg reflections of the thin layer<br />

were measured, leading to the result that the thin GaN film is based on wurtzite<br />

structure (figure 6.4), where the anions form a closed packed hexagonal structure and<br />

the cations with smaller atomic radius will occupy the tetrahedral gaps [10]. Owing to<br />

the stoichiometry only the half of all available tetrahedral positions can be filled.<br />

GaN crystallized in wurtzite structure and therefore it has a lower symmetry than<br />

native hexagonal structures. All hexagonal based lattices have same compliance and<br />

stiffness matrix [3] that are expressed in equation 6.2.<br />

⎛s11<br />

⎜<br />

⎜s12<br />

⎜s13<br />

s = ⎜<br />

⎜ 0<br />

⎜ 0<br />

⎜<br />

⎝ 0<br />

s<br />

s<br />

s<br />

12<br />

11<br />

13<br />

0<br />

0<br />

0<br />

s<br />

s<br />

s<br />

13<br />

13<br />

33<br />

0<br />

0<br />

0<br />

s<br />

0<br />

0<br />

0<br />

44<br />

0<br />

0<br />

s<br />

0<br />

0<br />

0<br />

0<br />

44<br />

0<br />

2 ( s<br />

Figure 6.4: Wurtzite structure<br />

11<br />

0<br />

0<br />

0<br />

0<br />

0<br />

− s<br />

11<br />

⎞<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

) ⎟<br />

⎠<br />

⎛ c11<br />

⎜<br />

⎜c12<br />

⎜c13<br />

⎜<br />

c =<br />

⎜ 0<br />

⎜<br />

0<br />

⎜<br />

⎜ 0<br />

⎝<br />

c<br />

c<br />

c<br />

12<br />

11<br />

13<br />

0<br />

0<br />

0<br />

c<br />

c<br />

c<br />

13<br />

13<br />

33<br />

0<br />

0<br />

0<br />

c<br />

0<br />

0<br />

0<br />

44<br />

0<br />

0<br />

c<br />

0<br />

0<br />

0<br />

0<br />

44<br />

0<br />

1<br />

( c<br />

2<br />

11<br />

0<br />

0<br />

0<br />

0<br />

0<br />

− c<br />

11<br />

⎞<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

) ⎟<br />

⎠<br />

(equ 6.2)<br />

18


Basic Physical Properties<br />

During the measurement the temperature is increased or decreased, and it must be<br />

taken into account that the elastic stiffness tensor has no longer constant values. In<br />

the evaluation of the data, the temperature dependence of the elastic constants,<br />

measured by R. R. Reeber and K. Wang [14], will be taken into consideration.<br />

c 13 / GPa<br />

100<br />

98<br />

96<br />

94<br />

92<br />

c11 / GPa<br />

c33 / GPa<br />

390<br />

385<br />

380<br />

375<br />

370<br />

365<br />

360<br />

134<br />

200 300 400 500 600 700 800 900 1000<br />

T / °C<br />

Figure 6.5: Temperature dependent stiffness constants of wurtzite GaN<br />

c 11<br />

c 33<br />

c 12<br />

c 13<br />

c 44<br />

146<br />

144<br />

142<br />

140<br />

138<br />

136<br />

c 12 / GPa<br />

99,0<br />

98,5<br />

98,0<br />

97,5<br />

97,0<br />

96,5<br />

c 44 / GPa<br />

19


Basic Physical Properties<br />

The plane spacing of hexagonal lattices is expressed through h, k, l, a0 and c0. Both<br />

unknown unstressed lattice parameters a0 and c0 must be considered when refining<br />

the structural parameters. R.R. Reeber and K. Wang [15] examined the lattice<br />

parameters of 99,99% pure and annealed GaN powder using neutron scattering. The<br />

results of their experiment can be seen in figure 6.6. With this data a relation between<br />

a0 and c0 is found by expressing c0 through the lattice parameter a0 and the ratio c0/ a0<br />

given by the neutron scattering measurement.<br />

-1<br />

c 0 a 0<br />

1,6263<br />

1,6262<br />

1,6261<br />

1,6260<br />

1,6259<br />

1,6258<br />

1,6257<br />

1,6256<br />

a 0 / A<br />

3,200<br />

3,198<br />

3,196<br />

3,194<br />

3,192<br />

3,190<br />

3,188<br />

3,186<br />

200 300 400 500 600 700 800 900 1000<br />

T / K<br />

a 0 spacing<br />

c 0 spacing<br />

-1<br />

c0 a0 Figure 6.6: Lattice parameters of wurtzite GaN<br />

5,195<br />

5,190<br />

5,185<br />

5,180<br />

c 0 / A<br />

20


6.2.2. Boron nitride<br />

Basic Physical Properties<br />

Boron nitride (BN) has different properties than other group III nitride members [16].<br />

The crystal structures and related physical properties are analogous to modifications<br />

of carbon, thus boron nitride exists in graphite like hexagonal structure. The former<br />

includes the stable zinc-blende and metastable wurtzite crystal structure. In contrast to<br />

the other group III nitrides no transition to the rock-salt structure at high pressures has<br />

been observed. The hexagonal boron nitride has outstanding mechanical properties,<br />

but it is less interesting for electronic applications.<br />

Figure 6.7: Modifications of boron nitride [17]<br />

gBN … hexagonal BN, zBN … zinc-blende BN, wBN … wurtzite BN<br />

The buffer layer contains only a small fraction of boron nitride, thus it is supposed that<br />

the boron will substitute gallium, so the buffer layer will also crystallise in wurtzite<br />

structure. The different lattice constants between boron nitride (a0=2,553A and<br />

c0=4,228A) and gallium nitride (a0=3,188A and c0=5,185A) [16] will cause tension in<br />

the buffer layer because of the smaller atomic radius of boron.<br />

21


Basic Physical Properties<br />

The elastic stiffness constants for wurtzite boron nitride at 300K, according to K. Kim,<br />

W.R.L. Lambrecht and B. Segall [18], are:<br />

c11 = 987 GPa; c12 = 143 GPa; c13= 70 GPa; c33 = 1020GPa; c44= 369GPa;<br />

6.2.3. Sapphire substrate<br />

Sapphire is the most used substrate for the growth of group III nitrides, which can be<br />

produced by the Czrochalski technique with high crystal quality and at low cost.<br />

Based on the 2:3 stoichiometry aluminium cations that take an octahedral position<br />

must fill two third of available sites [19]. To see how this occur a cation layer between<br />

two layers of close packed oxygen ions is drawn in figure 6.8.<br />

Figure 6.8: Shifting of aluminium cation layer<br />

These octahedral sites occupied by aluminium ions form a hexagonal array with the<br />

same spacing as the oxygen layer. The next cation layer has the same honeycomb<br />

configuration but is shifted by one atomic spacing in the direction of the vector 1. After<br />

another close packed oxygen layer, a third cation layer is placed, that is shifted by the<br />

vector 2. If a vertical slice as indicated by the dashed line is taken than the<br />

arrangement of cathions is drawn like in figure 6.9. The columns of octahedral sited<br />

perpendicular to the (0 0 0 1) plane alternate in having every two sides occupied and<br />

one empty. Only by considering the stacking of the closed packed oxygen ion layers<br />

must follow that the sapphire crystal is based on a hexagonal lattice. The anisotropic<br />

22


Basic Physical Properties<br />

elastic behaviour is for that reason expressed by equation 6.2. The following stiffness<br />

constants for sapphire are valid for room temperature [20].<br />

c11 = 496 GPa; c12 = 164 GPa; c13= 115 GPa; c33 = 498GPa; c44= 148GPa;<br />

A thermal property influencing the origin of extrinsic stress is the thermal expansion<br />

coefficient that is distinguished between expansion parallel and perpendicular to<br />

c-axis. According to Landolt and Börnstein [21] the thermal expansion coefficients for<br />

sapphire are 7,5 10 -6 K -1 perpendicular to c-axis and 8,5 10 -6 K -1 parallel to c-axis.<br />

Figure 6.9: (1 0 1 0) vertical slice according to dashed line in figure 6.8<br />

23


7. Polycrystalline and monocrystalline Model<br />

7.1. Calculating isotropic elastic constants<br />

7.1.1. Isotropic material<br />

Polycrystalline and monocrystalline Model<br />

An untextured material will behave in an isotropic way due to the randomly oriented<br />

grains, though a grain can be seen as single crystal with anisotropic elastic<br />

properties. For calculating stresses in isotropic materials, the Hill model [22] can be<br />

used which is based on Voigt and Reuss approach.<br />

7.1.2. Voigt model<br />

W. Voigt [4] assumed an untextured polycrystalline material where all grains are<br />

under the same strain. Starting with the matrix notation of the stiffness matrix for a<br />

cubic material (equation 6.1) and converting the expression into tensor notation<br />

(chapter 4.2) a rotation can be performed using Euler angles (equation 5.8) to rotate<br />

the crystal in any possible position. A mean value of all random orientated crystals is<br />

taken to express an isotropic behaviour (equation 4.14) of an untextured cubic<br />

material.<br />

2π<br />

π 2π<br />

V 1<br />

e e e e<br />

c ijkl = ∫ ∫ ∫ Oim<br />

O jn Oko<br />

Olp<br />

cmnop<br />

sin( φ)<br />

dϕ1<br />

dφ<br />

dϕ2<br />

(equ 7.1)<br />

8π<br />

0 0<br />

c<br />

c<br />

c<br />

0<br />

V<br />

1111<br />

V<br />

1122<br />

V<br />

2323<br />

= c<br />

= c<br />

= c<br />

V<br />

11<br />

V<br />

12<br />

V<br />

44<br />

1<br />

= ( 3c<br />

5<br />

1<br />

= ( c11<br />

5<br />

1<br />

= ( c11<br />

5<br />

11<br />

+ 2 c<br />

+ 4 c<br />

− c<br />

12<br />

12<br />

12<br />

+<br />

−<br />

+<br />

3 c<br />

4 c<br />

2 c<br />

44<br />

44<br />

)<br />

44<br />

)<br />

)<br />

(equ 7.2)<br />

The stiffness form is unusable for calculating strains, thus the elastic constants must<br />

be converted into compliance form by applying equation 4.9. First the stiffness<br />

components of the cubic system must be replaced on the right side, and afterwards<br />

the isotropic stiffness constants of the Voigt average must be expressed by<br />

compliance components of an isotropic material, that has only two independent<br />

parameters.<br />

s<br />

V<br />

44<br />

=<br />

2 ( s<br />

V<br />

11<br />

− s<br />

V<br />

12<br />

)<br />

24


Polycrystalline and monocrystalline Model<br />

Therefore the compliance components of the Voigt average defined by compliance<br />

constants of a cubic material are:<br />

s<br />

V<br />

12<br />

s<br />

V<br />

11<br />

=<br />

= 2 s<br />

11<br />

1 ⎛<br />

⎜<br />

⎜−<br />

s<br />

2 ⎝<br />

s<br />

11<br />

− s<br />

12<br />

+ 3s<br />

2<br />

5 ( s11<br />

− s12<br />

)<br />

−<br />

3s<br />

− 3s<br />

+ s<br />

12<br />

5 ( s<br />

11<br />

12<br />

44<br />

2<br />

5 ( s11<br />

− s12<br />

)<br />

+<br />

3s<br />

− 3s<br />

+ s<br />

− s<br />

11<br />

) s<br />

12<br />

44<br />

⎟ ⎞<br />

⎠<br />

V<br />

11 12 44<br />

44 = (equ 7.3)<br />

3s11<br />

− 3s12<br />

+ s44<br />

The sample can be rotated around three different angles ω, ψ and ϕ, but the rotation<br />

cannot be performed in any order of these angles. In figure 7.1 the sample and two<br />

coordinate systems are shown. The first one is the laboratory coordinate system eL<br />

which will be fixed and independent of the sample rotation. In this system the plane<br />

spacing is measured, because G will always be parallel to eL3. The second one is the<br />

sample coordinate system eS that is associated with the sample and will change<br />

position during rotation. The problem is, that the strains are measured in the<br />

laboratory system, but the stresses must be expressed in the sample system, so a<br />

rotation order must be performed to describe the sample’s tilt during the<br />

measurement. Let us start at a position where both systems coincide. The first<br />

rotation performed by the diffractometer is around eS1 direction by an angle ω,<br />

marked with the double arrow. In this position the new rotation is done around the<br />

new eS2 axis and the last one around the new eS3 axis. This is similar to the Euler<br />

angles procedure.<br />

a.) starting position b.) ψ-rotation c.) ϕ-rotation d.) end position<br />

and ω-rotation<br />

Figure 7.1: Rotation of the sample with respect to the laboratory system<br />

According to chapter 4.2 the laboratory system can be transformed by rotation into<br />

the sample system.<br />

25


e =<br />

The transformation matrix written in full is:<br />

Polycrystalline and monocrystalline Model<br />

3 1 2<br />

D<br />

S = O O O e L O e L<br />

(equ 7.4)<br />

⎛cos(<br />

ϕ)<br />

cos( ω)<br />

− sin( ϕ)<br />

sin( ψ)<br />

sin( ω)<br />

− sin( ϕ)<br />

cos( ψ)<br />

cos( ϕ)<br />

sin( ω)<br />

+ sin( ϕ)<br />

sin( ψ)<br />

cos( ω)<br />

⎞<br />

⎜<br />

⎟<br />

D<br />

O = ⎜sin(<br />

ϕ)<br />

cos( ω)<br />

+ cos( ϕ)<br />

sin( ψ)<br />

sin( ω)<br />

cos( ϕ)<br />

cos( ψ)<br />

sin( ϕ)<br />

sin( ω)<br />

− cos( ϕ)<br />

sin( ψ)<br />

cos( ω)<br />

(equ 7.5)<br />

⎟<br />

⎜<br />

⎟<br />

⎝ − cos( ψ)<br />

sin( ω)<br />

sin( ψ)<br />

cos( ψ)<br />

cos( ω)<br />

⎠<br />

Equation 7.4 will transform the laboratory system into the sample system, but for the<br />

evaluation of stresses the opposite way is demanded. For that reason the inverse<br />

orthogonal tensor O D is used.<br />

D D<br />

σ = O O σ<br />

ϕψ (equ 7.6)<br />

ij<br />

ki<br />

The change of the orthogonal tensor’s suffix, in equation 7.6, represents the<br />

transposed or inverted version of O D . The thin film is considered to be under plane<br />

stress, which means that every component of the stress tensor on the right side that<br />

has at least one three as suffix is zero.<br />

ε<br />

ϕψ,<br />

V<br />

33<br />

=<br />

lj<br />

kl<br />

V ϕψ<br />

s33kl σkl<br />

In the experiment the ω−angle has no significance and it is set to zero. After<br />

simplifying the expression for the Voigt model is:<br />

= ( σ<br />

) V<br />

ϕψ,<br />

V<br />

ε33 11 + σ22<br />

1 + σϕ<br />

2 ( s<br />

V =<br />

1<br />

σ ϕ<br />

11<br />

− s<br />

V<br />

2<br />

12<br />

V<br />

) ( s11<br />

+ 2 s12<br />

) − ( s<br />

2 (3s<br />

− 3s<br />

+ s<br />

11<br />

12<br />

2<br />

11<br />

44<br />

5 ( s11<br />

− s12<br />

) s44<br />

=<br />

2 (3s<br />

− 3s<br />

+ s<br />

11<br />

12<br />

44<br />

sin<br />

2<br />

− 3s<br />

)<br />

)<br />

( ψ)<br />

2<br />

2<br />

= σ11<br />

sin ( ϕ)<br />

− σ12<br />

sin( 2 ϕ)<br />

+ σ22<br />

cos ( ϕ)<br />

(equ 7.7)<br />

12<br />

) s<br />

44<br />

26


7.1.3. Reuss model:<br />

Polycrystalline and monocrystalline Model<br />

In contrast to Voigt supposes Reuss that all grains of an untextured material are<br />

under same stresses, so only grains with a net plane normal parallel to eL3 will diffract<br />

[23]. Therefore the crystal must be rotated in a position where the net plane is<br />

perpendicular to the measuring direction eL3.<br />

v<br />

3<br />

h b1<br />

+ k b2<br />

+ l b<br />

=<br />

h b + k b + l b<br />

v<br />

2<br />

1<br />

l a<br />

=<br />

l a<br />

1<br />

2<br />

2<br />

2<br />

− k a<br />

− k a<br />

v = v × v<br />

The wanted net plane normal can be rotated into a parallel position to eL3 by using<br />

the direction cosine.<br />

⎛ ec1<br />

. v1<br />

ec2<br />

. v1<br />

ec3<br />

. v1<br />

⎞<br />

⎜<br />

⎟<br />

e L = ⎜e<br />

c1 . v 2 ec2<br />

. v2<br />

ec3<br />

. v2<br />

⎟ eC<br />

= O<br />

⎜e<br />

c1 . v3<br />

ec2<br />

. v3<br />

ec3<br />

. v ⎟<br />

⎝<br />

3 ⎠<br />

The following othogonal tensor O hkl is only valid for cubic systems.<br />

O<br />

hkl<br />

⎛<br />

⎜<br />

⎜<br />

⎜<br />

= ⎜<br />

⎜<br />

⎜<br />

⎜<br />

⎝<br />

h<br />

h<br />

2<br />

2<br />

k<br />

2<br />

+ k<br />

+ l<br />

0<br />

h<br />

2<br />

+ k<br />

2<br />

2<br />

+ l<br />

2<br />

+ l<br />

2<br />

−<br />

k<br />

2<br />

+ l<br />

h<br />

+ k<br />

2<br />

2<br />

k + l<br />

k<br />

2<br />

2<br />

h k<br />

h<br />

l<br />

2<br />

2<br />

+ k<br />

2<br />

+ l<br />

2<br />

3<br />

2<br />

3<br />

3<br />

+ l<br />

2<br />

3<br />

3<br />

−<br />

k<br />

hkl<br />

2<br />

e<br />

−<br />

C<br />

+ l<br />

h<br />

2<br />

2<br />

h l<br />

h<br />

k<br />

+ k<br />

2<br />

2<br />

k + l<br />

l<br />

An untextured material will have lots of grains with G parallel to eL3 so in every<br />

position around the net plane normal the same amount of crystals is present.<br />

O<br />

λ<br />

⎛cos(<br />

λ)<br />

⎜<br />

= ⎜ sin( λ)<br />

⎜<br />

⎝ 0<br />

− sin( λ)<br />

cos( λ)<br />

0<br />

0⎞<br />

⎟<br />

0⎟<br />

1⎟<br />

⎠<br />

2<br />

+ k<br />

2<br />

+ l<br />

2<br />

2<br />

+ l<br />

2<br />

⎞<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎟<br />

⎠<br />

27


2π<br />

hkl,<br />

R 1 λ λ λ λ λ λ λ λ<br />

sijkl =<br />

2π<br />

∫ Oim<br />

O jn Oko<br />

Olp<br />

Omq<br />

Onr<br />

Oos<br />

Opt<br />

sqrst<br />

0<br />

Polycrystalline and monocrystalline Model<br />

After calculating the Reuss average the mean elastic constants are dependent of the<br />

miller indices h, k and l. The strain is written in the following way, by using the<br />

stresses in the laboratory system from before:<br />

After simplifying:<br />

ϕψ,<br />

hkl,<br />

R hkl,<br />

R ϕψ<br />

ε33 = s33 kl σkl<br />

ϕψ , hkl,<br />

R<br />

hkl<br />

ε33 = ( σ11<br />

+ σ22<br />

) R1<br />

+ σϕ<br />

R<br />

R<br />

7.1.4. Hill model<br />

R<br />

hkl<br />

2<br />

hkl<br />

1<br />

= s<br />

= s<br />

11<br />

12<br />

− s<br />

+ Γ<br />

12<br />

⎛<br />

⎜s<br />

⎝<br />

11<br />

− s<br />

⎛<br />

+ 3 Γ ⎜s<br />

⎝<br />

11<br />

12<br />

hkl<br />

2<br />

s<br />

−<br />

2<br />

− s<br />

12<br />

44<br />

2<br />

sin ( ψ)<br />

⎞<br />

⎟<br />

⎠<br />

s<br />

−<br />

2<br />

44<br />

⎞<br />

⎟<br />

⎠<br />

dλ<br />

2 2 2 2 2 2<br />

h k + k l + h l<br />

Γ =<br />

(equ 7.8)<br />

2 2 2 2<br />

( h + k + l )<br />

The Reuss and Voigt model are both limits of elastic stresses. In the hill model, that<br />

shows good correspondence with practical results, simply the average value of both<br />

limits is taken by allocating x in equation 7.9 to be a half.<br />

ε = x ε + ( 1 − x)<br />

ε<br />

(equ 7.9)<br />

hill<br />

33<br />

ϕψ,<br />

hkl,<br />

R<br />

33<br />

The strain in direction of the net plane normal is:<br />

After rewriting equation 7.10 follows:<br />

ε<br />

hill<br />

33<br />

ϕψ<br />

d − d<br />

=<br />

d<br />

0<br />

0<br />

ϕψ,<br />

V<br />

33<br />

(equ 7.10)<br />

28


d<br />

ϕψ<br />

ϕψ,<br />

hkl,<br />

R<br />

= d (1+<br />

( x ε + ( 1−<br />

x)<br />

ε<br />

0<br />

33<br />

Polycrystalline and monocrystalline Model<br />

ϕψ,<br />

V<br />

33<br />

Equation 7.11 will be used for nonlinear regression and d0 is substituted by:<br />

d<br />

ϕψ<br />

a 0 ϕψ,<br />

hkl,<br />

R<br />

ϕψ,<br />

V<br />

= (1+<br />

( x ε33<br />

+ ( 1−<br />

x)<br />

ε33<br />

)) (equ 7.11)<br />

2 2 2<br />

h + k + l<br />

7.2. Calculating stresses for a single crystalline material<br />

The examination of isotropic polycrystalline material differs from anisotropic single<br />

crystalline material in such a way that an untextured polycrystalline material fulfils the<br />

Bragg condition at every ψ-ϕ tilt because there are always crystals with net plane<br />

normals parallel to the measuring direction in the laboratory system. However Bragg<br />

reflections of single crystals can only be detected at strict defined tilts [24].<br />

ε = O O O O s σ<br />

(equ 7.12)<br />

ij<br />

e<br />

mi<br />

e<br />

nj<br />

e<br />

ok<br />

Only strains can be measured, hence the stresses must be expressed with a linear<br />

elastic anisotropic material model. The strains in the sample system are depending<br />

on the crystal position therefore the orthogonal tensor O e , which is described by Euler<br />

angles, is introduced to express the relation between the sample and the crystal<br />

system.<br />

ij<br />

D<br />

ki<br />

e<br />

pl<br />

D<br />

lj<br />

kl<br />

mnop<br />

kl<br />

))<br />

ε = O O ε<br />

ϕω (equ 7.13)<br />

The strain measurement is done in the laboratory system, so the sample’s tilt is<br />

defined by equation 7.13 which is similar to equation 7.6 except that the strain tensor<br />

is rotated and ω−angle is used instead of ψ.<br />

In the case of gallium nitride and gallium boron nitride on sapphire the basal plane of<br />

the GaN crystal and BGaN crystal is perpendicular to the surface normal, so further<br />

calculations are simplified in such a way that no orthogonal tensor O e is needed,<br />

because of the transversal isotropy of the hexagonal lattice.<br />

ε = O O s σ<br />

ϕω (equ 7.14)<br />

33<br />

D<br />

i3<br />

D<br />

j3<br />

ijkl<br />

kl<br />

29


Polycrystalline and monocrystalline Model<br />

If we are assuming a biaxial stress state than all stresses that have at least one three<br />

as suffix will be zero.<br />

ε ϕ<br />

= ( s<br />

11<br />

σ<br />

11<br />

+ s<br />

12<br />

σ<br />

22<br />

2<br />

) cos ( ϕ)<br />

+ ( s<br />

12<br />

ε<br />

= ε<br />

+ ε<br />

sin ( ω)<br />

ϕω<br />

33<br />

C ϕ 2<br />

ε<br />

σ<br />

c<br />

11<br />

= s13 ( σ11<br />

+ σ22<br />

)<br />

+ s<br />

11<br />

σ<br />

22<br />

2<br />

) sin ( ϕ)<br />

+ ( s<br />

11<br />

− s<br />

12<br />

) σ<br />

12<br />

sin(2 ϕ)<br />

− s13<br />

( σ11<br />

+ σ<br />

(equ 7.15)<br />

The plane spacing is calculated by applying the Bragg equation and can be<br />

expressed by the two lattice parameters a and c of the hexagonal lattice.<br />

The strain parallel to G is:<br />

2<br />

2 2<br />

1 4 (h + h k + k ) l<br />

= +<br />

(equ 7.16)<br />

2<br />

2<br />

2<br />

d 3a<br />

c<br />

ε<br />

ϕω<br />

33<br />

ϕω<br />

d − d<br />

=<br />

d<br />

The final regression formula has the following form:<br />

d<br />

0<br />

0<br />

ϕω<br />

C ϕ<br />

= d0<br />

(1+<br />

ε ) = d0<br />

(1+<br />

ε + ε sin ( ω)<br />

)<br />

(equ 7.17)<br />

ϕω 2<br />

33<br />

It is possible to replace one lattice parameter by using the ratio of cN0 and aN0<br />

measured by R. R. Reeber and K. Wang [15].<br />

c = a<br />

0<br />

The unstressed lattice parameter a0 is found using equation 7.16 and by replacing d0<br />

in the regression formula.<br />

0<br />

c<br />

a<br />

N0<br />

N0<br />

30<br />

22<br />

)


d<br />

2<br />

2<br />

N0<br />

Polycrystalline and monocrystalline Model<br />

2<br />

N0<br />

0 = a 0<br />

(equ 7.18)<br />

2<br />

l<br />

4<br />

+ (h<br />

3<br />

⎛ c<br />

⎜<br />

⎝ a<br />

⎞<br />

⎟<br />

⎠<br />

2 ⎛ c<br />

+ h k + k ) ⎜<br />

⎝ a<br />

The elastic behaviour of the buffer layer that is containing a small fraction of boron<br />

nitride, about 3%, can be estimated by supposing a linear change of elastic constants<br />

with increasing boron nitride content.<br />

s<br />

BGaN<br />

ij<br />

= s<br />

GaN<br />

ij<br />

+ ( s<br />

It is assumed that the compliance constants of the BGaN show the same thermal<br />

behaviour like the GaN layer constants.<br />

s ( T)<br />

=<br />

BGaN<br />

ij<br />

s(<br />

T<br />

)<br />

GaN<br />

ij<br />

BN<br />

ij<br />

− s<br />

s(<br />

300<br />

GaN<br />

ij<br />

s(<br />

300 K)<br />

The approximation of the temperature dependence was done by assuming that the<br />

compliance constants will increase in the same way like the compliance constants of<br />

gallium nitride. To estimate the new lattice parameters of BGaN the same procedure<br />

) x<br />

K)<br />

N0<br />

N0<br />

BGaN<br />

ij<br />

GaN<br />

ij<br />

was applied like before, hence the value of x is again 3%.<br />

And the new c0/a0 ratio is:<br />

c<br />

a<br />

0<br />

0<br />

a<br />

c<br />

( T)<br />

( T)<br />

7.3. Strain evaluation<br />

BGaN<br />

0<br />

BGaN<br />

0<br />

BGaN<br />

BGaN<br />

= a<br />

= c<br />

GaN<br />

0<br />

GaN<br />

0<br />

c0<br />

=<br />

a<br />

0<br />

+<br />

+<br />

( T)<br />

( T)<br />

( a<br />

( c<br />

GaN<br />

GaN<br />

BN<br />

0<br />

BN<br />

0<br />

− a<br />

− c<br />

0<br />

GaN<br />

0<br />

GaN<br />

0<br />

) x<br />

) x<br />

c0<br />

( 300 K)<br />

a 0(<br />

300 K)<br />

c0<br />

( 300 K)<br />

a ( 300 K)<br />

The strains are related to the sample system like in the stress evaluation. The only<br />

strain that can be measured is expressed in the laboratory system. This strain is<br />

parallel to G and so to eL3 vector of the laboratory system, thus the sample<br />

coordinates must be expressed by the laboratory coordinates. The opposite way is<br />

GaN<br />

⎞<br />

⎟<br />

⎠<br />

BGaN<br />

BGaN<br />

GaN<br />

31


Polycrystalline and monocrystalline Model<br />

expressed in equation 7.7, where the laboratory system is transformed into sample<br />

system. For that reason the inverse tensor O D will be used to transform the strains<br />

from the sample into the laboratory system.<br />

ε = O O ε<br />

ϕψ (equ 7.19)<br />

ij<br />

D<br />

ki<br />

The transposed matrix of O D is obtained by changing the suffixes. Like before the<br />

measured strain in the laboratory system is written as:<br />

ε<br />

ϕψ<br />

33<br />

D<br />

lj<br />

ϕψ<br />

d − d<br />

=<br />

d<br />

The equation used in the nonlinear regression has the following form:<br />

d<br />

ϕψ<br />

= d ( 1+<br />

( ε<br />

0<br />

33<br />

+ ( ε<br />

22<br />

+ ( ε<br />

13<br />

2<br />

cos ( ϕ)<br />

− ε<br />

sin( ϕ)<br />

− ε<br />

12<br />

23<br />

0<br />

0<br />

kl<br />

sin( 2 ϕ)<br />

+ ε<br />

11<br />

cos( ϕ))<br />

sin( 2 ψ)))<br />

2<br />

sin ( ϕ)<br />

− ε<br />

If the ω−axis is involved in the sample rotation the plane spacing is:<br />

d<br />

ϕω<br />

= d ( 1+<br />

( ε<br />

0<br />

33<br />

+ ( ε<br />

11<br />

+ ( ε<br />

2<br />

cos ( ϕ)<br />

+ ε<br />

13<br />

cos( ϕ)<br />

+ ε<br />

12<br />

23<br />

sin( 2 ϕ)<br />

+ ε<br />

22<br />

sin( ϕ))<br />

sin( 2 ω)))<br />

33<br />

2<br />

sin ( ϕ)<br />

− ε<br />

33<br />

2<br />

) sin ( ψ)<br />

2<br />

) sin ( ω)<br />

(equ 7.20)<br />

(equ 7.21)<br />

32


8. Deposition of Thin Films<br />

Deposition of Thin Films<br />

8.1. Magnetron sputtering of polycrystalline Al thin films on Si(1 0 0)<br />

Magnetron sputtering of polycrystalline thin films is a widely used deposition<br />

technique. Thin layers of chromium, gold, titanium, aluminium and many other<br />

materials can be produced using the technique in a deposition chamber<br />

schematically depicted in Fig. 8.1. In the evacuated chamber, ionized gas atoms<br />

(usually argon ions) are accelerated under influence of an electrical field towards a<br />

target. The high kinetic energy enables the ions to strike out surface bound atoms<br />

which are leaving, with rather low kinetic energy (some keV), the target’s surface in<br />

all possible directions. A substrate, fixed usually on the opposited side of the<br />

chamber, is covered by a thin film of the sputtered atoms. The advantages of the<br />

magnetron sputtering process are a good control of process parameters, high purity<br />

of the films and a possibility to use low process temperatures.<br />

For the production of polycrystalline Al thin films on monocrystalline Si(1 0 0) wafers,<br />

magnetron sputtering device installed at the “<strong>Institute</strong> of Physics, Technical University<br />

Wien” was used. As a substrate for the specimen preparation, a 1 mm thick Si(1 0 0)<br />

wafer cleaned using isopropanol and acetone was used. On this native-oxidized<br />

silicon wafer, a polycrystalline aluminium thin film with a thickness of 2 µm was<br />

deposited using magnetron sputtering. The deposition was performed at 150 °C<br />

applying a pressure of 4 x 10 -3 mbar.<br />

+<br />

Figure 8.1: A schematic view of a sputtering reactor<br />

1 Vacuum chamber, 2 upper electrode, 3 plasma,<br />

4 lower electrode, 5 gas inlet, 6 vacuum pump connection<br />

7 substrate, 8 high frequency voltage, 9 target<br />

10 plasma ion, 11 target atom<br />

33


8.2. Molecular beam epitaxy of GaN/BGaN on Al2O3(0 0 0 1)<br />

Deposition of Thin Films<br />

The molecular beam epitaxy (MBE) is a thin film deposition process in which thermal<br />

beams of atoms or molecules react on a single crystalline substrate surface that is<br />

held under ultra high vacuum at elevated temperatures [25, 26].<br />

Such devices usually consist of at least three chambers (Figure 8.2). The first<br />

chamber serves as a sample container at medium high vacuum. The second<br />

chamber is used for sample preparations such as outgassing or sputter etching and<br />

for an accommodation of surface analytical facilities under ultra high vacuum. The<br />

last chamber or growth chamber is also under ultrahigh vacuum and equipped with<br />

facilities capable of forming and monitoring the vacuum level, heating and monitoring<br />

the temperature of the substrate, generating and determining the molecular or atomic<br />

beam’s intensity and controlling of the composition profiles.<br />

The composition of the grown film and its doping level depend on the relative<br />

deposition rates of the constituent elements, which is controlled by the evaporation<br />

rate and the geometrical configuration between source and substrate. The<br />

evaporation rate of particular elements is set by temperature of the effusion cells,<br />

containing the element in the liquid phase.<br />

MBE has in comparison with other epitaxial growth methods some unique<br />

advantages. The growth rate is generally low allowing a compositional and doping<br />

profile changes within atomic dimensions leading also to an atomic smooth surface.<br />

The growth temperature is relatively low and thus diffusion between layers of different<br />

composition is negligible. A disadvantage is due to low growth rates the small<br />

throughput having an effect on long depositions times, about one hour for a one<br />

micron high layer.<br />

Figure 8.2: A schematic drawing of the growth chamber<br />

For the purposes of this diploma thesis, the GaN/B0.03Ga0.97N multilayer structure<br />

(with the individual thickness of 1 µm in the case of both sublayers) was prepared in<br />

34


Deposition of Thin Films<br />

the development laboratories of “Infineon AG, München”. In the preparation phase<br />

the substrate is degreased and etched for the removal of surface contaminates and<br />

mechanical damage due to polishing, and finally rinsed in deionised water. The<br />

deposition of both sublayers on monocrystalline Al2O3(0 0 0 1) was performed at<br />

700 °C in ultra-high vacuum gradually - at first BGaN sublayer was prepared and<br />

then the deposition process was concluded with the deposition of GaN sublayer. The<br />

original goal was to understand an influence of the BGaN buffer layer on the stress<br />

state in the GaN top layer.<br />

35


X-ray Diffraction – Measurements and Alignment<br />

9. X-ray Diffraction – Measurements and Alignment<br />

9.1. Four Circle Goniometer<br />

The basic principle of characterising residual stress in thin layers with the help of a<br />

diffractometer is to measure lattice spacing by using X-ray diffraction at different<br />

sample tilts. The deviation from the unstressed lattice parameters in combination with<br />

a material model enables the evaluation of the residual stresses. All present<br />

measurements were carried out on a Seifert 3000 PTS four circle diffractometer<br />

equipped with the DHS 900 heating stage.<br />

The X-ray radiation is produced by rapid deceleration of electrons and can be<br />

separated in continuous and characteristic radiation. In a diffraction experiment it is<br />

desirable to have monochromatic Kα radiation. This is achieved by placing a filter<br />

material, which has its K absorption edge between the Kα and Kβ wavelengths, in<br />

the beam path. A common filter material for copper radiation is nickel [27].<br />

Figure 9.1: Diffraction geometry<br />

36


X-ray Diffraction – Measurements and Alignment<br />

In order to obtain a convergent and parallel beam, a collimator is mounted on the X-<br />

ray tube housing. The collimated beam irradiates the sample surface, where the<br />

biggest part is absorbed and only a small part is diffracted, which can be detected by<br />

a scintillation counter at certain θ-angles.<br />

The collimator is replaced if the high resolution monchromator will be used. To<br />

achieve a sufficient high intensity the disturbing filter material must be removed, so<br />

the restriction of wavelengths is now guaranteed by the high resolution<br />

monochromator, because only beams of a certain wavelength undergo two times<br />

double diffraction.<br />

The geometry of a diffraction experiment is sketched in figure 9.1, where the θ-angle<br />

concerns the detector rotation and the three other angles ψ, ϕ and ω characterise the<br />

sample tilt.<br />

9.2. DHS 900 Domed Hot Stage<br />

A stress measurement at elevated temperatures requires a devise capable of a<br />

controlled heating of various thin film samples. In cooperation with the company<br />

“Anton Paar”, the “<strong>Erich</strong> <strong>Schmid</strong> Institut für Materialwissenschaften“ and the<br />

“Technische Universität Graz” the heating chamber DHS 900, Domed Hot stage, was<br />

developed to fulfil important demands on elevated-temperature stress analysis.<br />

The DHS 900 operates in a temperature range from 25°C up to 900°C controlled by<br />

the TCU TEX temperature controller, where temperatures can be entered manually<br />

or by a simple C-program. The advantage of setting the temperature via the serial<br />

port is an automatic controlled temperature dependent stress measurements. The C-<br />

Program can be seen in the appendix.<br />

During the heating the reactivity of a sample increases with temperature. To protect<br />

the layer from oxidation an inert gas or nitrogen is kept under a small dome. This<br />

circumstance means that the primary beam and the scattered beams must go<br />

through the dome, and no additional diffraction should be produced by the dome<br />

itself. Thus organic material like PEEK is an excellent option, because of its small<br />

absorption and scattering length, but the disadvantage of plastic is the low melting<br />

temperature and therefore it is necessary to cool the dome with air. So the supply by<br />

gas, air and electricity causes a limitation in rotation, thus it is recommended to stay<br />

in a ϕ range from -100° to 50°.<br />

37


9.3. The High-Resolution Monochromator<br />

X-ray Diffraction – Measurements and Alignment<br />

To measure residual stresses in a multilayered structure with very small differences<br />

in lattice spacing, it is necessary to use a monochromatic radiation with a very narrow<br />

wavelength (or energy) distribution. In this case, it is usual to use a high-resolution<br />

monochomator<br />

Figure 9.2: Monolithic monochromator with housing<br />

Figure 9.2 shows a picture of the high resolution monochromator, produced by the<br />

“Instituto dei Materiali per l’Elettronica ed il Magnetismo, Parma, Italy”. The design is<br />

based on the Bartel’s monochromator, shown in figure 9.3, which is including two<br />

channel-cut crystals cut out of a piece single crystalline germanium. In each crystal<br />

two separate symmetric {2 2 0} diffractions occur, so the final beam is in the same<br />

direction as the incident beam from the X-ray source.<br />

Figure 9.3: Bartels monochromator with possible rotations for alignment<br />

38


X-ray Diffraction – Measurements and Alignment<br />

In the present case, the high resolution monochromator was manufactured out of a<br />

piece of germanium single crystal, where the two channel cut crystals are joined by a<br />

spring. This significantly simplifies the usually complicated alignment procedure.<br />

The cut through the two blocks was done by taking into account two different<br />

scattering planes. In the first block (1 1 1) and ( 1 1 1 ) and in the second block<br />

( 2 2 0) and (2 2 0) planes are involved. The basic idea is that two couples of<br />

reflection with non parallel scattering vectors are being obtained in the same crystal<br />

for some specific geometry and by inducing a small bending δ of the crystal, as can<br />

be seen in figure 9.4. The monolithic design makes it not necessary to tilt the<br />

alignment like it is done with Bartels monochromator, since the crystal cut is designed<br />

to guarantee that the scattering plane in the first and the second block coincide within<br />

a good accuracy.<br />

Figure 9.4: Monolithic monochromator under small bending<br />

The final wavelength, dispersion and divergence are of interest for the measurement.<br />

A way to qualitatively evaluate these parameters can be found by using DuMond<br />

diagrams for multiple diffractions. According to the dynamical theory of X-ray<br />

diffraction, the reflectivity of a perfect crystal for a fixed wavelength is close to 100%<br />

in a finite ∆α range of several seconds of arc. Thus the DuMond diagram of a perfect<br />

crystal is a band of width ∆α in the λ−α space. In the DuMond diagram, shown in<br />

figure 9.5, two bands of wavelengths for several diffractions are plotted, one for the<br />

first block with asymmetric (1 1 1) reflection and one for the second block with<br />

symmetric (2 2 0) reflection. The cross section of the DuMond diagram for two<br />

crystals gives the wavelength dispersion and the divergence of the final X-ray beam.<br />

For the high resolution monochromator a wavelength dispersion ∆λ / λ = 2·10 -4 and a<br />

divergence of ∆α = 15,4 arcsec is obtained, by using CuKα X-ray radiation.<br />

39


X-ray Diffraction – Measurements and Alignment<br />

Figure 9.5: DuMond diagram for of the the Ge(1 1 1, 1 1 1 )( 2 2 0, 2 2 0) setting<br />

This device produces a diffracted beam with the minimum divergence in the<br />

scattering plane but with the maximum divergence of several degrees in the<br />

perpendicular direction, which does not affect the width of the diffraction profiles to be<br />

measured if the scattering plane of the sample is aligned to the scattering plane of<br />

the monochromator.<br />

The final intensity of the high resolution monochromator is more than 50% higher<br />

than that of the Bartels setting, thus the high resolution monochromator is a powerful<br />

device for high resolution X-Ray diffraction.<br />

40


9.4. Alignment of the diffractometer – point focus<br />

X-ray Diffraction – Measurements and Alignment<br />

To obtain reliable and reproducible results from X-ray diffraction stress<br />

characterization in ψ-geometry, the beam must be directed to the intersection of<br />

three rotation axes ω, ϕ and ψ, so that the beam spot is not allowed to move on the<br />

sample surface during rotation or tilting. The alignment procedure is carried out for<br />

the collimator with a point focus.<br />

At the beginning the detector is set to 90° position an d the ω, ψ and ϕ angles are set<br />

to 0°. Afterwards a laser is mounted on the shade hol der. If the laser is turned on, a<br />

little red spot will appear on the diffractometer plate. Now a small fluorescent plate is<br />

fixed on the diffractometer plate and the grid’s origin printed on the fluorescent plate<br />

must be in the middle of the laser spot.<br />

A dial gauge is fixed in front of the diffractometer plate to adjust the height until the<br />

distance measured from the small fluorescent plate surface is 5,75 mm. After that the<br />

dial gauge is removed.<br />

The ϕ angle is set to 180° and the laser spot must stay in the origin of the grid. If not<br />

the shade holder must be oriented, in such a way that the laser spot is back in the<br />

middle. After turning back to ϕ = 0° the spot should stay in the origin.<br />

The laser spot is now parallel to the ϕ-axis and, as next, the ψ angle is set to 85°.<br />

The spot is distorted into a line, but the middle of the line must stay in the origin of the<br />

grid. One more verification is made by setting ψ back to 0° and going with θ to 5°.<br />

Again a line appears while its midpoint must coincide with the grid’s origin on the<br />

fluorescent plate. After this procedure, the laser beam is directed to the intersection<br />

of ω, ϕ and ψ axes (table 9.1).<br />

θ ψ ω φ Note<br />

90° 0° 0° 0° / 180° The laser spot must stay in the origin.<br />

90° ± 85° 0° 0° The midpoint must stay in the origin.<br />

5° 0° 0° 0° The midpoint must stay in the origin.<br />

Table 9.1: Detector alignment<br />

The next alignment procedure will concern the collimator (table 9.2). First the ψ angle<br />

is set to 0°, the detector is driven to 155° and after wards the ω−angle must be in<br />

90° position. If the shutter is opened a weak spot will appear on the fluorescent plate<br />

surface. Like in the detector system alignment the middle of the spot must coincide<br />

with the grid’s origin. If not, the shutter must be closed at first and then the collimator<br />

will be adjusted. For that reason the whole procedure must be repeated several times<br />

until the middle of the X-ray spot coincides with the grid’s origin.<br />

41


X-ray Diffraction – Measurements and Alignment<br />

The previous step was only a rough alignment. To be sure that the spot is in the<br />

origin, ψ angle is set to 85°. The beam spot will appear as a li ne, and if the midpoint<br />

does not match with the origin, the collimator has to be adjusted in vertical direction.<br />

The alignment is nearly finished only the adjustment in the horizontal direction must<br />

be done, thus ψ is changed to 0° and ω is set to 3°. Afterwards the procedure is done<br />

in a similar way like in the vertical alignment.<br />

At the end the small fluorescent plate is replaced by an annealed gold film. A θ-scan,<br />

to measure the (4 2 0) peak with Cu-Kα radiation at a ψ-tilt of 0° and 60°, is<br />

performed after the height adjustment, where the dial gauge must show a value of<br />

5,75mm. The difference of the peak’s maxima at different ψ-tilts must not be more<br />

than 0,015° and the small angular misalignment betwe en measurement value and<br />

(4 2 0) peak position will be electronically corrected.<br />

θ ψ ω φ Note<br />

155° 0° 90° 0° The x-ray spot must be in the origin.<br />

155° 85° 90° 0° The midpoint must be in the origin.<br />

155° 0° 3° 0° The midpoint must be in the origin.<br />

Table 9.2 Collimator alignment<br />

9.5. Alignment of the diffractometer – line focus<br />

In the case of the point focus alligment, the main aim was to find the intersection<br />

point of all three rotation axes and focus the centre of the X-ray beam into this point.<br />

Performing measurements with the high resolution monochromator in line focus<br />

mode, the beam must just intersect the ω−rotation axis.<br />

Before the X-ray beam alignment is done it is important to ensure that the Bragg<br />

condition is fulfilled by the high resolution monochromator. Therefore all angles are<br />

set to 0° and the secondary slit is removed so the whole detector area is used to<br />

collect incoming photons. The tube’s voltage and current must be less than 20kV and<br />

20mA to avoid a damage of the detector. By inducing a small bending on the<br />

germanium crystal, the channel cut parts of the crystal are tilted a little bit, turning the<br />

net planes into diffraction position. A maximum intensity of 80.000 to 90.000 counts<br />

per second is a good standard value.<br />

Then the shade holder with the laser is mounted and the θ-angle is set to 90°. A<br />

cuboid shape like metallic block with parallel faces is fixed on the diffractometer plate,<br />

42


X-ray Diffraction – Measurements and Alignment<br />

so that the centre of the top face is in the middle of the laser spot. Afterwards the<br />

shade holder together with the laser is removed and the detector is turned back to<br />

0° position. Afterwards a dial gauge is mounted in fro nt of the diffractometer plate to<br />

adjust the height to 5,75 mm.<br />

The top surface of the metallic block intersects with the ω−axis. If no radiation is<br />

detected by opened shutter the height of the monochromator is too low and all<br />

radiation will be adsorbed by the metallic block (figure 9.6a). Otherwise if it is too high<br />

the detector will measure the full intensity (figure 9.6b). To find the correct height the<br />

monochromator is moved in such a way that half of the maximum intensity is<br />

measured, so it is guaranteed that the X-ray beam intersects with the ω−axis<br />

(figure 9.6c).<br />

a.) Monochromator is too low b.) Monochromator is too high c.) X-ray beam intersects with<br />

rotation axe<br />

The fine adjustment is done by rotating the metal block around ω−axis. A scan shows<br />

that the intensity has a maximum at a certain ω−angle (figure 9.7). It is supposed that<br />

at the maximum intensity the top surface of the metallic block is parallel to the<br />

incident beam, so the deviation from the zero ω−angle can be electronically<br />

corrected. The last step is to measure gold standard like it is done in the point focus<br />

alignment.<br />

Figure 9.6: Height adjustment of the monochromator<br />

43


I / counts s -1<br />

83000<br />

82000<br />

81000<br />

80000<br />

79000<br />

78000<br />

77000<br />

76000<br />

75000<br />

X-ray Diffraction – Measurements and Alignment<br />

-0,4 -0,2 0,0 0,2 0,4<br />

ω / °<br />

Figure 9.7: Fine adjustment<br />

44


10. Aluminium on silicon measurement<br />

10.1. Experiment<br />

Aluminium on silicon measurement<br />

The aluminium layer, deposited by magnetron sputtering on silicon substrate, was<br />

examined in a temperature controlled stress analysis performed on a four circle<br />

diffractometer operating in point focus mode. The thermal load of the sample, which<br />

included two thermal subcycles at elevated temperatures, was in a range from 25°C<br />

to 450°C.<br />

The temperature remained unaltered during the X-ray diffraction analysis and the<br />

production of new measurement files by copying and by rewriting the θ-range. All<br />

these processes took about 27 minutes together with the setting of the new<br />

temperature followed by the beginning of the next X-ray diffraction analysis. The<br />

order of temperature steps that were set in the experiment can be seen in figure 10.1.<br />

At the beginning the specimen was heated up from room temperature to 275°C by a<br />

step width of 25°C. The last temperature coincided wit h the beginning of the first<br />

thermal subcycle, where the temperature was decreased step by step to a lower limit<br />

of 150°C followed by an increase up to 375°C. During this heating the first thermal<br />

subcycle ended at a temperature of 275°C.<br />

T / °C<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

First thermal<br />

subcycle<br />

0 200 400 600 800 1000 1200<br />

t / min<br />

Second thermal<br />

subcycle<br />

Figure 10.1: Time table of the cycle experiment<br />

45


Aluminium on silicon measurement<br />

The last temperature was also the beginning of the second thermal subcycle that<br />

caused the sample’s temperature to decrease by a step width of 25°C to the lowest<br />

value of 250°C. Like at the lowest temperature in th e first subcycle, the sample was<br />

heated up again to the maximum temperature of 450°C . At the end, the cooling of the<br />

specimen was shortened by an increase of the step width to a value of 100°C.<br />

10.2. Shift of diffraction peaks<br />

The peak positions were changing during heating, caused by three effects, without<br />

considering delamination, crack formation and precipitation or solution of phases at<br />

elevated temperatures. The first effect is based on thermal or extrinsic stresses that<br />

are strongly depending on the difference of thermal expansion coefficients between<br />

film and substrate and/or between matrix and precipitations. The second effect is the<br />

thermal expansion of the sample and the last effect is due to the changes in thin film<br />

architecture by diffuse processes, like grain growth and recrystallisation. In figure<br />

10.2 the change of peak position is shown for the (6 2 0) substrate reflection at a<br />

ψ-angle of 0°.<br />

I / counts s -1<br />

60000<br />

50000<br />

40000<br />

30000<br />

20000<br />

10000<br />

0<br />

127,0 127,2 127,4 127,6 127,8 128,0 128,2 128,4 128,6 128,8<br />

2 θ / °<br />

Figure 10.2: Peak shift of (6 2 0) substrate peak at ψ = 0°<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

T / °C<br />

46


Aluminium on silicon measurement<br />

The two peaks originated from identical Bragg reflection and were traced back to<br />

characteristical copper Kα1 and Kα2 radiation. The peak located at lower θ-angles<br />

was referred to copper Kα1 and the other one to copper Kα2 radiation.<br />

The shown silicon peaks in figure 10.2 were measured in the first heating phase from<br />

room temperature to 275°C. With increasing temperatur e the sample expanded, so<br />

that the size of the lattice enlarged and therefore the reciprocal lattice spacing was<br />

reduced.<br />

I / counts s -1<br />

In figure 10.3 the peak shift of the (3 3 1) aluminium reflection at ψ = 0° is shown in<br />

the same temperature range as mentioned before, where the peaks shifted to lower<br />

θ-angles due to the thermal expansion of the lattice in combination with residual<br />

stresses.<br />

300<br />

250<br />

200<br />

150<br />

100<br />

110<br />

111<br />

112<br />

2 θ / °<br />

It is obvious, that the intensity of the layer peaks in comparison with the substrate<br />

peaks was much lower because of the structural differences. The substrate is single<br />

113<br />

300<br />

Figure 10.3: Peak shift of (3 3 1) aluminium peak at ψ = 0°<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

T / °C<br />

47


Aluminium on silicon measurement<br />

crystalline, so at a certain θ, ϕ and ψ position all net planes within the irradiated<br />

volume were involved in the scattering process, in contrast to the polycrystalline<br />

layer, where only grains in particular orientations were diffracting.<br />

Another difference between layer and substrate is the peak shift which is reflected by<br />

the thermal expansion coefficients of the silicon substrate, 2,6 10 -6 K -1 , and the thin<br />

aluminium layer, 23,8 10 -6 K -1 . The mean peak shift obtained from the measurement<br />

data was for the aluminium layer 0,159°K -1 and that of the substrate had a value of<br />

0,0187°K -1 , so the ratio of thermal expansion coefficients and the ratio of peak shifts<br />

are similar.<br />

10.3. Peak broadening with increasing ψ tilt<br />

During the stress analysis it was essential to tilt the sample by combination of two<br />

angles, like it was done in the experiment by using ψ and ϕ-angle.<br />

I / counts s -1<br />

150<br />

100<br />

50<br />

0<br />

(3 3 1); ψ = 0°<br />

(3 3 1); ψ = 30°<br />

(3 3 1); ψ = 45°<br />

(3 3 1); ψ = 60°<br />

111 112 113 114<br />

2 θ / °<br />

Figure 10.4: Aluminium peaks at 25°C<br />

Figure 10.4 shows (3 3 1) aluminium reflections of the first X-ray diffraction<br />

measurement at room temperature for ϕ = 0° and at different ψ-tilts. Consider the<br />

black curve, where a separation between Kα1 and Kα2 peak can be easily done. At a<br />

48


Aluminium on silicon measurement<br />

tilt of ψ = 30° the peak shape is becoming broader (red), theref ore both copper Kα<br />

peaks start to merge, but one can still separate between Kα1 and Kα2 peak. The<br />

peak profile at ψ = 45° (purple) is asymmetric and broader than the last two curves.<br />

At the highest ψ-tilt, the peak profile (blue) seems to be symmetric, although it<br />

consists of Kα1 and Kα2 radiation.<br />

One reason for the peak broadening was the elongation of the beam spot caused by<br />

sample tilt. Figure 10.5 shows the change of an originally circular beam spot at<br />

different ψ and θ-angles. The region from where X-rays were diffracted enlarged with<br />

increasing sample and decreasing θ-tilt. A rotation around the ψ-axis had the effect<br />

that half of the sample surface was behind and half of it was in front of its original<br />

plane. In these regions the beam was not longer exactly focused on the sample<br />

surfaces, which led to a decreasing intensity simultaneously with a broadening of the<br />

reflection profile.<br />

Usually the integral intensity of the peaks remains constant, but the detector<br />

equipped with a system of receiving slits, recorded only radiation diffracted from a<br />

small, constant area of the sample surface, therefore the intensity decreased with<br />

increasing sample tilt. It is worth to mention that such a beam spot distortion has no<br />

effect on the θ-position of the peaks.<br />

Figure 10.5: Distortion of beam spot<br />

49


11. GaN and GaBN on sapphire measurement<br />

GaN and GaBN on sapphire measurement<br />

11.1. Measuring with the high resolution monochromator<br />

For this measurement it should be noted that the GaN-layer, the BGaN buffer-layer<br />

as well as the substrate are single crystalline, so a Bragg reflection can only be<br />

obtained by rotating the sample in a certain position depending on the<br />

crystallographic orientation. Therefore a couple of certain angles has to be used to<br />

measure lattice spacing for each (h k l) reflection.<br />

To differentiate between BGaN and GaN reflections, a high resolution X-ray<br />

diffraction experiment was performed to resolve the small difference in-plane spacing<br />

between GaN layer and BGaN buffer layer. For that reason the collimator was<br />

replaced by the high resolution monochromator that operates in line focus mode.<br />

In figure 11.1a the combination of ψ and ϕ-axes would distort the primary beam’s<br />

projection and would cause a radiation of different specimen regions by varying the<br />

sample tilt. On the other hand the ω - ϕ couple is more favourable, because there is<br />

no influence on the spot shape, as can be seen in figure 11.1b.<br />

a.) Psi – ϕ tilting b.) Omega - ϕ tilting<br />

Figure 11.1: Choice of rotation axes and distortion of the primary beam’s projection<br />

50


GaN and GaBN on sapphire measurement<br />

It has to be noted that the ω−axis will always be parallel to the θ-axis if the<br />

ω - ϕ couple is chosen. Thus the measured ω value is not the true tilt of the crystal<br />

plane with respect to the sample system.<br />

Figure 11.2: Definition of omega tilt<br />

Figure 11.2 illustrates a small volume cut out of the sample. The thick black line<br />

symbolizes a net plane where only a small part of incident beam is diffracted, by<br />

showing an angular difference of 2θ between primary and scattered beam.<br />

The measured ω−angle ωm is defined as angle between the sample surface and the<br />

primary beam. However, a true ω value, which is demanded in the evaluation, is<br />

represented as angle between net plane and surface.<br />

2 θ<br />

ω = − ωm<br />

(equ 11.1)<br />

2<br />

In figure 11.2 it is obvious that the true ω can be calculated in equation 11.1 by<br />

calculating the difference between θ and the measured ω−angle.<br />

11.2. Stereographic projections and crystal orientation<br />

Because there exists no simple procedure to estimate the find reflections in the<br />

orientation space, the whole ϕ - ω space must be scanned to detect a certain Bragg<br />

reflection. After two peaks were found, a calculation can be performed by using<br />

numerical solvers to estimate the reflection that is perpendicular to the layer surface.<br />

In the case of GaN and BGaN the (0 0 0 1) direction is parallel to the surface normal.<br />

This information was used to generate a stereographic projection by the program<br />

JWulf [28], that helped to find additional Bragg reflections by converting the true<br />

51


GaN and GaBN on sapphire measurement<br />

ω−angle, that is obtained by the stereographic projection, into the ω−value, that was<br />

set in the measurement (equation 11.1). Due to the monochromatic wavelength and<br />

the loss of intensity by the monochromator, it was necessary to perform scans<br />

without shade, because the expected peaks had a low intensity and a low peak<br />

width.<br />

ω / °<br />

22,0<br />

21,5<br />

21,0<br />

20,5<br />

20,0<br />

19,5<br />

19,0<br />

18,5<br />

The peak detection of a single crystalline material in a high resolution X-ray<br />

diffraction measurement is a very sensitive process, where a small change of the tilt<br />

angle ω can cause a peak to vanish as can be seen in figure 11.3. In the scan the<br />

GaN peaks in ϕ−direction had a peak broadening of about 3,5° and the broadening in<br />

ω−direction was about 0,5°.<br />

-80 -60 -40 -20 0 20 40<br />

A scan of the sapphire substrate can be seen in figure 11.4. The X-ray beam that<br />

penetrates into the substrate was diffracted in a bigger volume than in the layer, so a<br />

higher amount of atoms were taking part in the scattering process, which led to a<br />

ϕ / °<br />

Figure 11.3: {2 1 3 4} GaN Peaks<br />

smaller peak width in comparison to the layer peaks.<br />

50 counts s -1<br />

100 counts s -1<br />

150 counts s -1<br />

200 counts s -1<br />

250 counts s -1<br />

52


GaN and GaBN on sapphire measurement<br />

The substrate peak was of about 1° in ϕ-direction, and in ω−direction the broadening<br />

was only about 0,2°. Since the peaks shifted with increasi ng temperature, their<br />

position could sometimes be only detected by performing a scan without slit a small<br />

ϕ - ω range.<br />

ω / °<br />

All Bragg peaks that were used to examine stresses in the thin film are plotted in the<br />

stereographic projection, shown in figure 11.5. The surface normal of the sample<br />

coincides with the (0 0 0 1) direction and the ( 1 0 1 0) direction is parallel to eS2<br />

vector of the sample system if the misalignment of the sample in the DHS 900 is<br />

neglected.<br />

25,5<br />

25,0<br />

24,5<br />

24,0<br />

23,5<br />

23,0<br />

22,5<br />

22,0<br />

-80 -60 -40 -20 0 20 40<br />

ϕ /°<br />

0 counts s -1<br />

20 counts s -1<br />

40 counts s -1<br />

60 counts s -1<br />

80 counts s -1<br />

Figure 11.4: {3 1 4 11} Sapphire peak<br />

100 counts s -1<br />

120 counts s -1<br />

140 counts s -1<br />

53


GaN and GaBN on sapphire measurement<br />

eS2<br />

Figure 11.5: Stereographic projection of GaN peaks<br />

54<br />

eS1


11.3. Phi adjustment<br />

GaN and GaBN on sapphire measurement<br />

Figure 11.3 and 11.4 show that the reflections were stretched in the ω - ϕ space, so<br />

the main task was to find the peak maximum by varying ϕ and ω−angle. Before the<br />

first ϕ-scan was carried out by setting an estimated ω and θ-angle, the shade holder<br />

was removed to ensure that the whole detector area could be used to detect a peak<br />

maxima. The ϕ-scan at room temperature is plotted in figure 11.6 where three peaks<br />

belonging to the same family of net planes are shown.<br />

I / counts s -1<br />

300<br />

200<br />

100<br />

0<br />

ϕ (2 2 0 5) / °<br />

18 20 22 24 26 28 30 32 34<br />

400<br />

-105 -100 -95 -90<br />

ϕ (2 0 2 5) / °<br />

-42 -40 -38 -36 -34 -32 -30 -28<br />

ϕ (0 2 2 5) / °<br />

Figure 11.6: Phi scans of {2 0 2 5} peaks<br />

ϕ (2 0 2 5)<br />

ϕ (0 2 2 5)<br />

ϕ (2 2 0 5)<br />

55


GaN and GaBN on sapphire measurement<br />

The peak profiles in figure 11.6 were not fitted by a Gaussian peak profile. An<br />

explanation of such a peak shape is given by the wide detector range measuring<br />

without a shade. Figure 11.7 shows a peak that is going to be scanned where the<br />

detector is assumed to be the reference system. For that reason the environment and<br />

the peaks are moving. A peak that will come into the measuring range of the detector<br />

increases the measured intensity until the whole peak is in the detector range, thus<br />

the peak profile will show a plateau region.<br />

So the peaks were not fitted by a simple Gaussian peak function that is unable to<br />

describe the plateau region of the measured ϕ-peaks, but the evaluation can be done<br />

by introducing the regression parameter c instead of the quadratic function in the<br />

exponential term.<br />

Figure 11.7: Peak shape distortion<br />

I ( θ)<br />

= I<br />

0<br />

⎛<br />

⎜ ⎛ θ - θ<br />

+ a exp - ⎜<br />

⎜ ⎜<br />

⎝ ⎝ b<br />

0<br />

c<br />

⎞ ⎞<br />

⎟ ⎟<br />

⎟<br />

⎠<br />

⎟<br />

⎠<br />

56


11.4. Omega adjustment<br />

GaN and GaBN on sapphire measurement<br />

Like in chapter 11.3 all scans were carried out without a slit. In the ω−scans the new<br />

ϕ-angle at maximum intensity was set and the former estimated θ-angle was kept.<br />

Such scans, performed at room temperature, of {2 0 2 5} net planes are shown in<br />

figure 11.8. The second peak cannot be traced back to copper Kα2 radiation<br />

because of the restricted wavelength spectrum. Therefore it is supposed that the<br />

peak at higher ω−angles refers to BGaN due to lower lattice constant caused by<br />

substitution of gallium by boron atoms.<br />

The ω−position of GaN and BGaN is assumed to be hardly influenced by the<br />

ϕ-position because of the large peak width in ϕ-direction. Therefore ω−scans for both<br />

layers were performed at same ϕ-tilt.<br />

I / counts s -1<br />

800<br />

600<br />

400<br />

200<br />

0<br />

31,0 31,2 31,4 31,6 31,8 32,0<br />

ω / °<br />

Figure 11.8: Omega scans of {2 0 2 5} peaks<br />

ω (2 0 2 5)<br />

ω (0 2 2 5)<br />

ω (2 2 0 5)<br />

57


11.5. Theta scans:<br />

GaN and GaBN on sapphire measurement<br />

The θ-angle is required to calculate the plane spacing, the crucial parameter for<br />

stress and strain evaluation. At first, the sample was tilted into the ω and ϕ-orientation<br />

determined by the procedures from Sec. 11.3 and 11.4. It was necessary to<br />

distinguish between GaN and BGaN θ-peaks with respect to their ω−positions. These<br />

scans were performed using a 1mm slit shade. The results are presented in figure<br />

11.9 for GaN and in figure 11.10 for BGaN.<br />

I / counts s -1<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

135,8 136,0 136,2 136,4 136,6 136,8 137,0 137,2<br />

2 θ / °<br />

2θ (2 0 2 5)<br />

2θ (0 2 2 5)<br />

2θ (2 2 0 5)<br />

Figure 11.9: Theta scans of {2 0 2 5} GaN peaks<br />

58


I / counts s -1<br />

800<br />

600<br />

400<br />

200<br />

GaN and GaBN on sapphire measurement<br />

0<br />

136,4 136,6 136,8 137,0 137,2 137,4 137,6 137,8<br />

The same procedure was used to gain information about sapphire peaks as well. The<br />

substrate peaks, like (0 0 0 12), showed much higher intensity but smaller peak width<br />

than the layer peaks, so the detection was very difficult even the shade holder was<br />

removed. The substrate peak shape had no Gaussian profile, as can be seen in<br />

figure 11.11, because of the same reasons mentioned in chapter 11.3 or maybe<br />

because of the dynamic theory of X-ray diffraction, where this peak could be<br />

interpreted as Darwin curve.<br />

2 θ / °<br />

2θ (2 0 2 5)<br />

2θ (0 2 2 5)<br />

2θ (2 2 0 5)<br />

Figure 11.10: Omega scans of {2 0 2 5} GaBN peaks<br />

59


I (2 2 0 5) / counts s -1<br />

500<br />

400<br />

300<br />

200<br />

100<br />

2 θ (0 0 0 12) / °<br />

2 θ (2 2 0 5) / °<br />

GaN and GaBN on sapphire measurement<br />

90,3 90,4 90,5 90,6 90,7 90,8 90,9<br />

600<br />

7000<br />

2θ (0 0 0 12)<br />

2θ (2 2 0 5)<br />

0<br />

0<br />

135,6 135,8 136,0 136,2 136,4 136,6 136,8 137,0 137,2 137,4<br />

6000<br />

5000<br />

4000<br />

3000<br />

2000<br />

1000<br />

Figure 11.11: Comparison between GaN layer and sapphire substrate peak<br />

I (0 0 0 12) / counts s -1<br />

60


12. Results of aluminium on silicon<br />

12.1. Sin(ψ)² vs. a plot<br />

Results of aluminium on silicon<br />

The principle behind the stress evaluation is the projection of stresses defined in the<br />

sample coordinate system on the measuring direction that is usually the eL3 axis of<br />

the laboratory system. If the sample is not tilted around the ψ-axis the eS3 direction<br />

will be parallel to the eL3 direction. For that reason no stresses can be evaluated if the<br />

sample is in a biaxial stress state that has no components in eS3 direction.<br />

σ33 ϕψ<br />

= ( σ<br />

11<br />

2<br />

sin ( ϕ)<br />

− σ<br />

12<br />

sin( 2<br />

ϕ)<br />

+ σ<br />

22<br />

2<br />

2<br />

cos ( ϕ))<br />

sin ( ψ)<br />

Because of the isotropic behaviour of the thin Aluminium film, it is supposed that the<br />

in-plane stresses (σ11, σ22) are equal, which leads to the non-diagonal elements<br />

σ12 = 0.<br />

a / A<br />

4,080<br />

4,075<br />

4,070<br />

4,065<br />

4,060<br />

4,055<br />

4,050<br />

4,045<br />

4,040<br />

275°C<br />

250°C<br />

225°C<br />

200°C<br />

175°C<br />

150°C<br />

125°C<br />

100°C<br />

75°C<br />

25°C<br />

0,0 0,2 0,4 0,6 0,8<br />

sin 2 (ψ)<br />

Figure 12.1: Lattice spacing dependence on sin² (ψ)<br />

61


Results of aluminium on silicon<br />

Such a simplified regression is mainly depending on the ψ-tilt and implies a linear<br />

behaviour of a vs sin²(ψ). The measurement points, in figure 12.1, show the lattice<br />

spacing at ψ = 0°, 30°, 45° and 60° evaluated from the 2- θ positions of the peaks,<br />

starting from room temperature to 275°C.<br />

If the sample is tilted the peak position will be influenced by residual stresses that are<br />

responsible for the change of lattice spacing and the shift of peaks at constant<br />

temperature. The slope value of the linear smoothing function provides information<br />

about the stress state, so a positive sign indicates that the layer is under tensile<br />

stress and if a negative sign is obtained then the opposite stress state will be present.<br />

Furthermore the absolute value of the slope is proportional to the magnitude of<br />

stress.<br />

In figure 12.1 can be seen that the layer was in a tensile stress state at room<br />

temperature. With raising temperature the slope is slightly decreasing and is nearly<br />

zero at 125°C. A further heating was increasing the com pressive stress until 225°C<br />

where the layer started to relax to lower compressive stress values.<br />

a / A<br />

4,080<br />

4,075<br />

4,070<br />

4,065<br />

4,060<br />

4,055<br />

4,050<br />

275°C<br />

250°C<br />

225°C<br />

200°C<br />

175°C<br />

275°C<br />

250°C<br />

225°C<br />

200°C<br />

175°C<br />

150°C<br />

0,0 0,2 0,4 0,6 0,8<br />

sin 2 (ψ)<br />

Figure 12.2: First thermal subcycle<br />

Heating<br />

Cooling<br />

62


Results of aluminium on silicon<br />

In the first thermal subcycle the temperature was changed in steps of 25°C from<br />

275°C to 150°C, followed by an increase of temperatur e up to 275°C. In figure 12.2<br />

the stresses increased from compression to tension during the cooling, which is<br />

expressed by the blue lines. The approximate border between the two stress states<br />

was found at 225°C.<br />

The heating, symbolised by the black lines, reduced the tension and turned the layer<br />

back into a compressive stress state at a temperature of about 225°C. It can be seen<br />

that there is nearly no difference in stress behaviour between heating and cooling<br />

operation because the linear smoothing functions are nearly parallel.<br />

The dependence of the lattice spacing on the sample tilt in the second thermal<br />

subcycle, which is shown in figure 12.3, was measured in a temperature range from<br />

250°C to 375°C with a step width of again 25°C. In t he first X-ray analysis of the<br />

second subcycle at a temperature of 375°C the specimen was under compressive<br />

stresses. The followed cooling increased the stress value until 325°C, where a further<br />

reduction of heat did not seem to affect the stress of the following temperature steps.<br />

a / A<br />

4,090<br />

4,085<br />

4,080<br />

4,075<br />

4,070<br />

4,065<br />

375°C<br />

350°C<br />

325°C<br />

300°C<br />

275°C<br />

250°C<br />

375°C<br />

350°C<br />

325°C<br />

300°C<br />

275°C<br />

0,0 0,2 0,4 0,6 0,8<br />

sin 2 (ψ)<br />

Figure 12.3: Second thermal subcycle<br />

Heating<br />

Cooling<br />

63


Results of aluminium on silicon<br />

Tension was dominating at the lowest temperature of 250°C. A followed heating<br />

enlarged the lattice spacing and decreases the slope of the smoothing function, so<br />

the stress changed back to compression at 300°C.<br />

In the cooling operation the border between tension and compression was detected<br />

at about 350°C and in the heating operation it was f ound about 50°C lower than<br />

before. At the end of the second thermal subcycle the black and blue linear<br />

smoothing functions are nearly at same position.<br />

12.2. Lattice spacing of aluminium layer<br />

In equation 7.11 not only the in-plane stress but also the mean unstressed lattice<br />

parameter, of all cubic face centred aluminium crystals that are present in the<br />

irradiated sample volume, were evaluated.<br />

a 0 / A<br />

4,10<br />

4,09<br />

4,08<br />

4,07<br />

4,06<br />

4,05<br />

4,04<br />

0 100 200 300 400 500<br />

T / °C<br />

a 0 = a + b T + c T 2<br />

a = 4,0488<br />

b = 6,8827 10 -5<br />

c = 7,4653 10 -8<br />

Figure 12.4: Thermal expansion of aluminium lattice spacing<br />

The thermal expansion of the lattice spacing a0 of the thin aluminium layer is plotted<br />

in figure 12.4 and shows only weak scattering of the data. An exponential function<br />

proved to be favourable to fit the temperature dependent lattice spacing.<br />

64


12.3. Thermal expansion coefficient of aluminium<br />

Results of aluminium on silicon<br />

The thermal expansion coefficient of the aluminium layer can be calculated [29] by<br />

entering the temperature dependent fit function, which expresses the unstressed<br />

lattice spacing of the previous paragraph, in equation 12.1.<br />

1 ∂ a 0(<br />

T)<br />

α ( T)<br />

=<br />

(equ. 12.1)<br />

a ( T)<br />

∂T<br />

0<br />

The result of this application is plotted in figure 12.5. A comparison between values<br />

from literature (chapter 6.1.1) and the calculated expansion shows a difference of<br />

20% at room temperature.<br />

α Al / 10 -6 K -1<br />

34<br />

32<br />

30<br />

28<br />

26<br />

24<br />

22<br />

20<br />

18<br />

16<br />

0 100 200 300 400 500<br />

T / °C<br />

Figure 12.5: Thermal expansion coefficient of aluminium<br />

65


12.4. Lattice spacing of silicon substrate<br />

Results of aluminium on silicon<br />

The lattice spacing of silicon substrate was evaluated by making use of the (6 2 0)<br />

reflection. It is assumed that the thermal stresses of the substrate are negligible with<br />

respect to relatively deep penetration depth of the X-ray beam, so the measurements<br />

can provide an information about the thermal expansion of the substrate.<br />

a 0 / A<br />

5,440<br />

5,438<br />

5,436<br />

5,434<br />

5,432<br />

5,430<br />

12.5. Thermal expansion coefficient of silicon substrate<br />

The procedure of calculating the thermal expansion coefficient of the silicon substrate<br />

is done in the same manner as it was mentioned in chapter 12.3. The result can be<br />

seen in figure 12.7, where the thermal expansion coefficient at room temperature is<br />

2,43 10 -6 K -1 .<br />

0 100 200 300 400 500<br />

T / °C<br />

a 0 = a + b T + c T 2<br />

a = 5,4308<br />

b = 1,2380 10 -5<br />

c = 1,2058 10 -8<br />

Figure 12.6: Thermal expansion of silicon lattice spacing<br />

66


α Si / 10 -6 K -1<br />

4,5<br />

4,0<br />

3,5<br />

3,0<br />

2,5<br />

2,0<br />

12.6. Stress curve<br />

Results of aluminium on silicon<br />

0 100 200 300 400 500<br />

Owing to the heating and cooling operations, it is necessary to differ between these<br />

two temperature gradients. Consequently the stress dependence of the heating is<br />

drawn using black lines and the blue lines symbolize the stress dependence during<br />

the cooling in figure 12.8.<br />

T / °C<br />

Figure 12.7: Thermal expansion coefficient of silicon substrate<br />

At 25°C the layer was in a tensile stress state of 185 M Pa, but with raising<br />

temperature the stress decreased linearly by 1,702 MPa·K -1 (12.8-1) until the<br />

temperature reached 200°C where the linearity was lost (12.8-2) and the curve<br />

shows a minimum at 225°C. At a temperature of 275°C the first thermal subcycle<br />

was commenced. The linear stress-temperature dependences of the heating and<br />

cooling are very similar which is reflected by the slope values of 1,957 MPa·K -1 of the<br />

cooling operation and 1,995 MPa·K -1 of the heating operation (12.8-3 and 12.8-4).<br />

67


Results of aluminium on silicon<br />

After the first subcycle, the residual stress was in range between -70MPa and -<br />

40MPa that was measured in a temperature interval from 275°C to 375°C.<br />

The specimen temperature of 375°C was also the beginnin g of the second thermal<br />

subcycle that started with a cooling operation (12.8-5). From 375°C to 325°C the<br />

stress increased linearly with a slope of 1,581 MPa·K -1 , afterwards the slope changed<br />

to 0,663 MPa·K -1 thus the stress increased less than before, but the dependence was<br />

still linear. At the lower limit of the second thermal subcycle at 250°C, the sample<br />

was heated again. During the heating the stress decreased with 1,876 MPa·K -1 until a<br />

temperature of 325°C was reached (12.8-6). Afterwards the stress remained in a<br />

range between -55MPa and -40MPa. At the end the sample was cooled to room<br />

temperature with a step width of 100°C between the X -ray diffraction measurements.<br />

It is supposed that the temperature dependent stress of the last cooling operation<br />

(12.8-8) can be described by a linear behaviour with a slope value of 0,815 MPa·K -1 .<br />

σ 11 / MPa<br />

400<br />

300<br />

200<br />

100<br />

0<br />

-100<br />

-200<br />

1<br />

4<br />

3<br />

0 100 200 300 400 500<br />

T / °C<br />

Heating<br />

Cooling<br />

Figure 12.8: In-plane stress in aluminium layer<br />

1 First heating phase, 2 deviation from linear behaviour,<br />

3 cooling operation of first thermal subcycle, 4 heating operation of first thermal subcycle,<br />

5 cooling operation of second thermal subcycle, 6 heating operation of second thermal subcycle,<br />

7 decrease of compressive yield stress, 8 last cooling operation<br />

2<br />

6<br />

8<br />

5<br />

7<br />

68


12.7. Strain curves<br />

Results of aluminium on silicon<br />

The evaluated stress from chapter 12.6 is based on the strain measurement, since<br />

the shift of Bragg reflections is referred to the change of plane spacing, which is<br />

proportional to the change of strain ε33 in laboratory system.<br />

In the chapter 12.1, the in-plane stresses σ11 and σ22 were assumed to be equal, and<br />

the shear stress was neglected, so the strain evaluation must be based on this<br />

assumptions. In the evaluation of the strain data use has been made of the<br />

unstressed lattice spacing a0 received from the previous stress evaluation. The<br />

reason why not calculating a0 in the strain evaluation is that it was done before, and<br />

therefore more measurement points per regression parameter can be used in the<br />

nonlinear regression. The results are shown in figure 12.9 and 12.10 for the in-plane<br />

and out of plane strain-temperature dependence.<br />

ε 11<br />

0,004<br />

0,003<br />

0,002<br />

0,001<br />

0,000<br />

-0,001<br />

-0,002<br />

0 100 200 300 400 500<br />

T / °C<br />

Figure 12.9: In-plane strain of aluminium layer<br />

Heating<br />

Cooling<br />

69


ε 33<br />

0,002<br />

0,000<br />

-0,002<br />

-0,004<br />

12.8. Discussion<br />

Results of aluminium on silicon<br />

0 100 200 300 400 500<br />

T / °C<br />

Figure 12.10: Out of plane strain of aluminium layer<br />

During the deposition of the thin film, target atoms were adsorbed on the layer<br />

surface transferring a part of their kinetic energy to the layer and the substrate, where<br />

this additional energy is converted into heat and plastic deformation of the layer. The<br />

rather low deposition temperature of 150°C, that was kept constant by a temperature<br />

control devise during the film formation, led to a high supercooling accompanied by a<br />

high density of nuclei. Owing to the low temperature diffusion was limited, and so<br />

grain growth was negligible, resulting in a fine-grained film structure.<br />

The deposition process was followed by a cooling operation that caused a<br />

contraction of the layer as well as the substrate. The prevention of free contraction by<br />

the substrate during cooling down was responsible for extrinsic in-plane tensile<br />

stresses in the aluminium layer, because of the ten times higher thermal expansion<br />

coefficient of aluminium. For that reason tensile in-plane stresses were obtained in<br />

the first X-ray diffraction experiment at room temperature.<br />

Heating<br />

Cooling<br />

70


Results of aluminium on silicon<br />

The measured temperature dependent stress behaviour is shown in figure 12.8,<br />

where the stress/temperature dependence of the first heating operation from room<br />

temperature to 200°C is elastic. Only the extrinsic in-p lane stresses, but no additional<br />

plastic deformation, are superimposed on the intrinsic in-plane stress state (12.8-1).<br />

Therefore, the intrinsic in-plane stress value of -39MPa [30, 31] was derived at<br />

150°C, which corresponds to the deposition temperature where no extrinsic stresses<br />

were assumed to be present. The compressive intrinsic in-plane stress state was<br />

traced back to atomic peening caused by the fabrication technique.<br />

In the present case diffuse processes were activated at elevated temperatures<br />

[6, 32], where at 200°C a deviation from the linear elastic behaviour was observed<br />

(12.8-2). This nonlinearity is related to plastic deformation, grain growth and<br />

recovery. According to F. Vollertsen and S. Vogler [32] dynamic recrystallisation of<br />

aluminium can be excluded, because a high dislocation density is rapidly reduced by<br />

recovery processes at elevated temperatures.<br />

As mentioned before, the deposited layer formed as fine-grained structure, so the<br />

reduction of interface boundary that is synonymous with a reduction of the integral<br />

surface energy is the driving force behind the coarsening process [32]. The grain<br />

growth will be inhibited if the grains reach a dimension which is in the order of the<br />

layer thickness, and so it is assumed that the grain growth process ended at a<br />

temperature of 275°C [6]. Such a coarsened structure m ay be regarded as bamboo<br />

structure which represents an obstacle for electro migration.<br />

The material responded in an elastic manner during the first thermal subcycle (12.8-3<br />

and 12.8-4). A completely different behaviour was measured in the second thermal<br />

subcycle which showed a hysteresis. The stress elastically increased starting at<br />

375°C and -40,3MPa until 325°C, where the sample sta rted to yield again at a tensile<br />

stress value of 38,5MPa (12.8-5). At 250°C the sample w as heated again and an<br />

elastic behaviour was obtained up to 325°C (12.8-6). It can be seen that the<br />

compressive yield stress was decreasing in the temperature range from 275°C to<br />

450°C (12.8-7). This is a well known phenomenon. Plasti c deformation is described<br />

through dislocation movement and diffusion driven processes. Those processes are<br />

temperature dependent and cause a material to yield at lower stress values than at<br />

room temperature.<br />

Regarding to figure 12.8, it is obvious that the elastic behaviour of the cooling<br />

operation in the second thermal subcycle (12.8-5) was measured within a<br />

temperature interval of 325°C to 375°C, which is small er than the elastic response of<br />

the heating operation (12.8-6) that lay in a temperature interval of 250°C to 325°C. If<br />

the plastic yield stress of the cooling operation in the second thermal subcycle is<br />

extrapolated to higher temperatures values, like it is done in figure 12.11, than the<br />

extrapolated yield stress will intersect with the compressive yield stress of the cycle<br />

(12.8-7) at a temperature lower than 450°C. Conseque ntly it is questionable if an<br />

71


Results of aluminium on silicon<br />

elastic behaviour is obtained during the last cooling operation (12.8-8) which cannot<br />

be shown by the experimental data because of the rough step width of 100°C.<br />

However, an elastic behaviour would have only little influence on the in-plane<br />

stresses of the last cooling operation starting from 450° to room temperature that is<br />

dominated by plastic deformation (12.8-8).<br />

σ 1 / MPa<br />

200<br />

100<br />

0<br />

-100<br />

-200<br />

In figure 12.12 the decrease of intensity as a function of temperature is shown. The<br />

data is scattered but an exponential decay is obvious. Of interest are the points<br />

belonging to the last cooling steps (350°C, 250°C, 150 °C and 25°C), which lie close<br />

to the dotted line.<br />

250 300 350 400 450<br />

T / °C<br />

Figure 12.11: Extrapolation of plastic yield stress<br />

Heating<br />

Cooling<br />

An explanation of the deviation of the exponential decay could be given by<br />

considering the elongation of grains, caused by plastic deformation accompanied by<br />

a preferred orientation of crystallographic planes. For that reason the lower intensity<br />

values may be traced back to a change of texture during the last cooling operation.<br />

72


I / counts s -1<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

T / °C<br />

Results of aluminium on silicon<br />

0 100 200 300 400 500<br />

Figure 12.12: Decrease of intensity<br />

(3 3 1); ψ = 0°<br />

73


13. Results of GaN/GaBN/Al2O3(0 0 0 1)<br />

13.1. Sapphire lattice parameters<br />

Results of GaN/GaBN/Al2O3(0 0 0 1)<br />

Temperature dependencies of sapphire lattice parameters were evaluated by<br />

measuring (0 0 0 12) and (3 1 4 11) reflections of the substrate. Since the<br />

penetration depth of the X-ray beam is much higher than the GaN/BGaN film<br />

thickness, it can be supposed that the lattice spacings evaluated from the measured<br />

data represent unstressed parameters.<br />

The lattice parameters can be calculated using equation 7.16 and by entering the<br />

calculated spacing from the Bragg equation 3.8. The temperature dependent lattice<br />

parameters are plotted in Figure 13.1 and were fitted by a quadratic function.<br />

c 0 / A<br />

13,06<br />

13,04<br />

13,02<br />

13,00<br />

12,98<br />

12,96<br />

c 0<br />

a 0<br />

a 0 = a + b T + c T 2<br />

a = 4,7581<br />

b = 2,9193 10 -5<br />

c = 7,5746 10 -9<br />

100 200 300 400 500 600<br />

T / °C<br />

c 0 = a + b T + c T 2<br />

a = 12,9849<br />

b = 9,6315 10 -5<br />

c = 1,5527 10 -8<br />

Figure 13.1: Temperature dependencies of sapphire lattice parameters.<br />

4,790<br />

4,785<br />

4,780<br />

4,775<br />

4,770<br />

4,765<br />

4,760<br />

4,755<br />

74<br />

a 0 / A


13.2. Thermal expansion coefficients of sapphire<br />

Results of GaN/GaBN/Al2O3(0 0 0 1)<br />

The thermal expansion coefficients of the sapphire were evaluated from the functions<br />

describing temperature dependencies of the lattice spacings using equation 12.1.<br />

The results are plotted in figure 13.2, where the evaluated thermal expansion<br />

coefficients show good correlation to literature values.<br />

α / 10 -6 K -1<br />

9,0<br />

8,5<br />

8,0<br />

7,5<br />

7,0<br />

6,5<br />

6,0<br />

100 200 300 400 500 600<br />

T / °C<br />

Figure 13.2: Thermal expansion coefficient of sapphire<br />

13.3. In-plane stress in GaN and GaBN layer:<br />

The stress evaluation in GaB and in BGaN is based on the assumption that single<br />

crystalline hexagonal thin films are oriented with (0 0 0 1) parallel to the sample’s<br />

surface normal as described in chapter 7.2. Considering these specific geometrical<br />

conditions, it can be supposed that the in-plane stress is isotropic (σ11 = σ22) while<br />

surface normal stress components σi3 = 0. As a consequence of the assumed stress<br />

state follows that the non-diagonal components σ12 = 0.<br />

75


Results of GaN/GaBN/Al2O3(0 0 0 1)<br />

The unstressed lattice spacing a0 and c0 of the hexagonal crystals were evaluated<br />

using the equation 7.18. It is assumed that the ratio of the unstressed lattice<br />

parameters a0/c0 is corresponding to the literature values measured by K. Wang and<br />

R.R. Reeber [15]. The evaluated unstressed lattice parameter a0 is plotted in<br />

figure 13.3.<br />

a 0 / A<br />

3,200<br />

3,198<br />

3,196<br />

3,194<br />

3,192<br />

3,190<br />

3,188<br />

3,186<br />

The in-plane stresses were measured in a temperature range from 50°C to 600°C<br />

with a step width of 50°C. They behave in a nearly l inear way in the lower<br />

temperature region, as can be seen in figure 13.4. In the beginning the in-plane<br />

stress was in a compressive state of 440 MPa that is further reduced during the<br />

heating cycle. By approaching the end of the heating procedure the thin film was in<br />

an unstressed state.<br />

0 100 200 300 400 500 600<br />

T / °C<br />

Figure 13.3: Measured lattice parameter a0 of GaN<br />

76


σ 11 / GPa<br />

0,0<br />

-0,1<br />

-0,2<br />

-0,3<br />

-0,4<br />

-0,5<br />

T / °C<br />

Results of GaN/GaBN/Al2O3(0 0 0 1)<br />

0 100 200 300 400 500 600<br />

Figure 13.4: Inplane stress of GaN layer<br />

The properties of BGaN sublayer were evaluated using the same approach as those<br />

of the GaN top layer. Both have wurtzite structure and same orientation with respect<br />

to the substrate. For the BGaN sublayer, the unstressed lattice parameter as well as<br />

the in-plane stresses were evaluated. The unstressed lattice spacing ao is slightly<br />

smaller than the ao spacing of the GaN layer, as can be seen in figure 13.5.<br />

77


a 0 / A<br />

3,194<br />

3,192<br />

3,190<br />

3,188<br />

3,186<br />

3,184<br />

3,182<br />

3,180<br />

Results of GaN/GaBN/Al2O3(0 0 0 1)<br />

0 100 200 300 400 500 600<br />

Figure 13.6 shows the in-plane stress of the buffer layer containing about 3% boron<br />

nitride that is at 50°C under tensile stress of 350MPa. An increase of temperature<br />

was attended by an increase of in-plane stress, for that reason the in-plane stress at<br />

600°C has a value of 710MPa.<br />

T / °C<br />

Figure 13.5: Unstressed lattice parameter a0 of the GaBN buffer layer<br />

78


σ 11 / GPa<br />

0,8<br />

0,7<br />

0,6<br />

0,5<br />

0,4<br />

0,3<br />

Results of GaN/GaBN/Al2O3(0 0 0 1)<br />

0 100 200 300 400 500 600<br />

In figure 13.7 both in-plane and out of plane strain of the GaN top layer are<br />

approaching an unstrained state with raising sample temperature. The in-plane strain<br />

of the buffer layer in figure 13.8 is in a tensile state forcing the out of plane stress into<br />

compressive strain regions.<br />

T / °C<br />

Figure 13.6: Inplane stress of GaBN buffer layer<br />

79


ε 11<br />

ε 11<br />

0,0000<br />

-0,0002<br />

-0,0004<br />

-0,0006<br />

-0,0008<br />

-0,0010<br />

0,0016<br />

0,0014<br />

0,0012<br />

0,0010<br />

0,0008<br />

0,0006<br />

0,0004<br />

T / °C<br />

Figure 13.7: In and out of plane strains of GaN layer<br />

Results of GaN/GaBN/Al2O3(0 0 0 1)<br />

0 100 200 300 400 500 600<br />

0 100 200 300 400 500 600<br />

T / °C<br />

Figure 13.8: In and out of plane strains of BGaN buffer layer<br />

ε 11<br />

ε 33<br />

ε 11<br />

ε 33<br />

5e-4<br />

4e-4<br />

3e-4<br />

2e-4<br />

1e-4<br />

0<br />

-0,0002<br />

-0,0003<br />

-0,0004<br />

-0,0005<br />

-0,0006<br />

-0,0007<br />

ε 33<br />

ε 33<br />

80


13.4. Discussion:<br />

Results of GaN/GaBN/Al2O3(0 0 0 1)<br />

Sapphire is no perfectly suitable substrate for the growth of GaN layer due to the<br />

enormous lattice mismatch. For that reason, attempts are made to reduce the<br />

mismatch by introducing a buffer layer to minimize the defect density which would<br />

degrade the GaN layer’s quality.<br />

In the present case a GaN buffer layer containing about 3% boron nitride was<br />

deposited on the sapphire substrate. If the anion sublattices of the sapphire and the<br />

BGaN are continuous than the deposited buffer layer should be under compressive<br />

stress [33], which cannot be shown by the experimental data.<br />

The simple approach, by only considering geometrical aspects, is not sufficient to<br />

explain the transition region between sapphire and GaN buffer layer that is<br />

influencing the intrinsic stresses. Not only the structural difference between GaN and<br />

sapphire that seems to be responsible for a tensile stress state, caused by island<br />

coalescence [34] of the growing GaN layer and surface tension, but also the<br />

substitution of GaN by BN, which has a smaller unit cell than GaN, contributes to<br />

tensile stress in the buffer layer. An extrapolation of buffer layer’s in-plane<br />

stresses/temperature dependence shows a value of 766MPa at the deposition<br />

temperature at 700°C. The tensile stresses of the buffe r layer affects the GaN layer<br />

that is, if the smoothing function is extrapolated, in a tensile stress state of 128MPa at<br />

the deposition temperature. Therefore the intrinsic stresses of the buffer and GaN<br />

layer are supposed to have a value of 766MPa and 128MPa, respectively.<br />

Extrinsic stresses do not exist at a temperature of 700°C, because they can only<br />

originate from a heating or cooling after the thin film formation. A cooling procedure<br />

starting from the deposition temperature will cause a higher contraction of the<br />

substrate than of the layer materials. For that reason the substrate will induce<br />

compressive in-plane stresses in the BGaN and GaN layer, as is observed in the<br />

experiment where the stresses at room temperature were much lower than at<br />

elevated temperatures.<br />

High resolution X-ray diffraction stress measurements were performed during the<br />

cooling to provide information about an eventually plastic flow of the material.<br />

However, the results showed a good correspondence between heating and cooling of<br />

the sample, so it is assumed that the sample is only elastically deformed.<br />

81


14. Conclusion and Outlook<br />

Conclusion and Outlook<br />

Elevated-temperature X-ray diffraction has been applied to evaluate temperature<br />

dependencies of residual stresses in aluminium thin film on silicon (0 0 1) and in<br />

sublayers of BGaN/GaN structure on sapphire. The results indicate that the method<br />

can provide useful information with respect to the elastic and plastic response of the<br />

(sub)layers, resolve residual stresses in sublayers with very small differences in<br />

composition, extrapolate intrinsic and extrinsic stresses, observe phenomena related<br />

to the changes of thin film architecture etc. The part of the data has been published<br />

and submitted [31, 35]<br />

Further studies on this topic should focus the stress relaxation in transition region<br />

between plastic and elastic deformation that is attributed to grain growth of the<br />

polycrystalline aluminium. A peak profile analysis coupled with scanning electron<br />

microscopy measurements should provide an information about the coarsening<br />

process of the aluminium layer in future examinations.<br />

The weak textured layer was treated in an isotropic way during the evaluation. This<br />

assumption cannot be applied on all polycrystalline thin films due to the circumstance<br />

that most layers show a strong texture. The isotropic Hill model in chapter 7.1 can be<br />

seen as groundwork for a future based ODF dependent formula capable of serving<br />

as basis for a nonlinear stress evaluation of textured layer materials.<br />

The sign and magnitude of the GaN layer in-plane stresses is an important issue in<br />

the manufacturing process. For that reason a buffer layer is usually deposited on the<br />

substrate, in order to provide a suitable lattice for the GaN film. The qualitiy of the<br />

buffer lattice is influenced by residual stresses that are supporting defect formation<br />

and finally leading to a degradation of the BGaN and GaN layer.<br />

It was demonstrated that a separation of Bragg reflections belonging to different<br />

materials with nearly same lattice parameters can be done by a restriction of the<br />

wavelength spectrum using the high resolution monochromator. Further examinations<br />

of GaN/BGaN/Sapphire multilayered structures should be performed to characterise<br />

the in-plane stresses/temperature behaviour depending on the boron nitride content<br />

of the buffer layer, so that the boron nitride influence on the temperature dependent<br />

stresses can be resolved. Such results can serve in a future development of GaN<br />

based applications.<br />

82


15. Literature<br />

[1] C. Kittel<br />

Einführung in die Festkörperphysik<br />

R. Oldenbourg Verlag, 1999<br />

[2] P. Chadwick<br />

Continuum mechanics, concise theory and problems<br />

Dover publications, 1999<br />

[3] J. F. Nye<br />

Properties of crystals, their representation by tensor and matrices<br />

Oxford university press, 2000<br />

[4] W. Voigt<br />

Lehrbuch der Kristallphysik<br />

B. G. Teubners Lehrbücher, 1910<br />

[5] K. O’Donell, J. Kostetsky, R. S. Post<br />

Controlling stress in thin films<br />

IMAPS Flipchips, Wilmington, 2002<br />

[6] W. D. Nix<br />

Mechanical properties of thin films<br />

Metallurgical transactions A, Vol 20A., pp 2217, November 1989<br />

[7] Bronstein, Semendjajew, Musiol, Mühlig<br />

Taschenbuch der Mathematik<br />

Verlag Harri Deutsch, 1995<br />

[8] V. Randle, O. Engler<br />

Introduction to texture analysis, macrotexture, microtexture & Orientation<br />

Mapping<br />

Gordon and Breach Science Publishers, 2000<br />

[9] Hermann Schumann<br />

Metallographie, 13. neu bearbeitete Auflage<br />

Deutscher Verlag für Grundstoffindustrie Stuttgart, 1991<br />

Literature<br />

83


[10] G. Harsch, R. <strong>Schmid</strong>t<br />

Kristallgeometrie, Packungen und Symmetrie in Stereodarstellungen<br />

Diesterweg/Salle, 1993<br />

[11] Landolt Börnstein<br />

Literature<br />

Elastische, piezoelektrische, piezooptische und elektrooptische Konstanten<br />

von Kristallen<br />

Springer Verlag Berlin – Heidelberg – New York, 1966<br />

[12] Horst Kuchling<br />

Taschenbuch der Physik<br />

Fachbuchverlag Leipzig, 1996<br />

[13] Micheal H. Jones, Stephan H. Jones<br />

Basic Mechanical and Thermal Properties of Silicon<br />

www.virginiasemi.com<br />

[14] R. R. Reeber, K. Wang<br />

High Temperature Elastic Constant Prediction of Some Group III-Nitrides<br />

MRS Internet Journal Nitride Semiconductor Research Vol. 6, 2001<br />

[15] K. Wang, R.R. Reeber<br />

Thermal Expansion of GaN and AlN<br />

Nitride Semiconductor Symposium 863-8, 1998<br />

[16] J. I. Pankove, T. D. Moustakas<br />

Gallium nitride (GaN) I<br />

Semiconductors and semimetals, volume 50, 1998<br />

[17] Anna E. McHale<br />

Phase Equilibria Diagrams / Phase diagrams for ceramicists<br />

Volume X, Figures 8665-9114<br />

The American Ceramic Society, Inc. 1994<br />

[18] K. Kim, W.R.L. Lambrecht, B. Segall<br />

Elastic constants and related properties of tetrahedrally bonded BN, AlN, GaN,<br />

and InN.<br />

Physical Review B 53, S. 16310, 1996<br />

84


[19] Y. M. Chiang, W. D. Kingery, D. Birnie III<br />

Physical ceramics, principles for ceramic science and engineering<br />

Wiley, 1997<br />

[20] MarkeTech International<br />

Sapphire Table of General Properties<br />

www.mkt-intl.com<br />

[21] Landolt-Börnstein<br />

Numerical Data and Functional Relationship in Science and Technology<br />

Crystal and solid Physics, Vol. 17, Semiconductors, Springer New York.<br />

[22] P. van Houtte, L. de Buyser<br />

The influence of crystallographic texture on diffraction measurement of<br />

residual stress<br />

Acta Metallurgica et Materialia, Vol. 41, No. 2, pp 323, 1991<br />

[23] T. Uchida, N. Funamori, T. Yagi<br />

Lattice strains in crystals under uniaxial stress field<br />

Journal of applied physics, Vol 80, No. 2, pp 739, 1996<br />

[24] J. Keckes, J. W. Gerlach, B. Rauschenbach<br />

Literature<br />

Residual stresses in cubic and hexagonal GaN grown on sapphire using ion<br />

beam-assisted deposition<br />

Journal of crystal growth 219 1-9, 2000<br />

[25] J. I. Pankove, T. D. Moustakas<br />

Gallium nitride (GaN) II<br />

Semiconductors and semimetals, volume 57, 1999<br />

[26] G. Koblmüller<br />

Molecular beam epitaxy of group III-nitrides on silicon carbide<br />

<strong>Diploma</strong> thesis TU Wien, 2001<br />

[27] I. C. Noyan, J. B. Cohen<br />

Residual stress, measurement by diffraction and interpretation<br />

Springer Verlag, 1986<br />

85


[28] Steffen Weber<br />

JWulf<br />

http://jcrystal.com/steffenweber/java.html<br />

[29] P. A. Tipler<br />

Physik<br />

Spektrum Akademischer Verlag, Heidelberg – Berlin – Oxford, 1995<br />

[30] E. Eiper<br />

Literature<br />

Einfluss der Temperatur auf mechanische Spannungen in dünnen, auf Silizium<br />

gewachsenen Al-Filmen<br />

<strong>Diploma</strong>rbeit TU Graz, 2003<br />

[31] E. Eiper, R. Resel, et al.<br />

Thermally-induced stresses in thin aluminium layers grown on silicon<br />

Powder Diffraction 19(1), March 2004<br />

[32] F. Vollertsen, S. Vogler<br />

Werkstoffeigenschaft und Mikrostruktur<br />

Hanser Verlag, 1989<br />

[33] E. V. Etzkorn, D. R. Clarke<br />

Cracking of GaN films<br />

Journal of Applied Physics, Vol. 89, Nr. 2, pp 1025, 2000<br />

[34] K. A. Dunn, S.E Babcock, D. S. Stone, et al.<br />

Dislocation Arrangement in a thick LEO GaN Film on Sapphire<br />

MRS Internet Journal Nitride Semiconductor Research 5S1, W2.11, 2000<br />

[35] J. Keckes, M. Hafok, E. Eiper, et al.<br />

Evaluation of experimental stress-strain dependence from thermally cycled Al<br />

thin film on Si<br />

Acta Materialica<br />

86


16. Appendix<br />

16.1. Serial port communication<br />

#include <br />

#include <br />

#include <br />

#include <br />

#include <br />

void main(int argc, char *argv[])<br />

{<br />

FILE *fschr;<br />

unsigned int i, bcc, zeit;<br />

char temp[6],zeich[0];<br />

}<br />

strcpy(temp,argv[1]);<br />

bcc= 'S'^'L';<br />

for (i=0;i


16.2. Stereographic projection of Silicon (0 0 1)<br />

Appendix<br />

88


16.3. Stereographic projection of gallium nitride (0 0 1)<br />

Appendix<br />

89


16.4. Stereographic projection of sapphire (0 0 1)<br />

Appendix<br />

90

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

Saved successfully!

Ooh no, something went wrong!