07.12.2012 Views

Stresses in Cu Thin Films and Ag/Ni Multilayers - Harvard School of ...

Stresses in Cu Thin Films and Ag/Ni Multilayers - Harvard School of ...

Stresses in Cu Thin Films and Ag/Ni Multilayers - Harvard School of ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

<strong>Stresses</strong> <strong>in</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> <strong>and</strong> <strong>Ag</strong>/<strong>Ni</strong> <strong>Multilayers</strong><br />

A thesis presented<br />

by<br />

Dillon Dodd Fong<br />

to<br />

The Division <strong>of</strong> Eng<strong>in</strong>eer<strong>in</strong>g & Applied Sciences<br />

<strong>in</strong> partial fulfillment <strong>of</strong> the requirements<br />

for the degree <strong>of</strong><br />

Doctor <strong>of</strong> Philosophy<br />

<strong>in</strong> the subject <strong>of</strong><br />

Applied Physics<br />

<strong>Harvard</strong> University<br />

Cambridge, Massachusetts<br />

September 2001


c○2001 - Dillon Dodd Fong<br />

All rights reserved.


Thesis advisor Author<br />

Frans Spaepen Dillon Dodd Fong<br />

<strong>Stresses</strong> <strong>in</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> <strong>and</strong> <strong>Ag</strong>/<strong>Ni</strong> <strong>Multilayers</strong><br />

Abstract<br />

The stresses <strong>in</strong> substrate-bonded <strong>Cu</strong> th<strong>in</strong> films <strong>and</strong> <strong>Ag</strong>/<strong>Ni</strong> multilayers were studied us<strong>in</strong>g<br />

wafer curvature <strong>and</strong> x-ray scatter<strong>in</strong>g techniques. The biaxial modulus <strong>of</strong> sputtered <strong>Cu</strong> films<br />

<strong>in</strong> the as-deposited state is less than that <strong>of</strong> the dynamic by approximately 15%. After<br />

anneal<strong>in</strong>g, the modulus <strong>in</strong>creases due to enhanced 〈111〉 texture but still rema<strong>in</strong>s compar-<br />

atively low. The deficit <strong>in</strong>creases at higher temperatures, suggest<strong>in</strong>g anelastic processes<br />

contribute to the modulus reduction. The annealed <strong>Cu</strong> films show high flow strengths due<br />

to the presence <strong>of</strong> the rigid substrate. Attempts were made to f<strong>in</strong>d a dislocation boundary<br />

layer <strong>in</strong> the plastically stra<strong>in</strong>ed film by graz<strong>in</strong>g <strong>in</strong>cidence diffraction studies. No confirma-<br />

tion <strong>of</strong> a stra<strong>in</strong> gradient near the substrate could be found, however, due to the <strong>in</strong>sensitivity<br />

<strong>of</strong> the measurement to stra<strong>in</strong>s at buried <strong>in</strong>terfaces. Significant stra<strong>in</strong> relaxation was seen at<br />

the surface, however. The effect <strong>of</strong> <strong>in</strong>terphase boundaries on <strong>in</strong>ternal stresses was also stud-<br />

ied <strong>in</strong> <strong>Ag</strong>/<strong>Ni</strong> multilayers. The {111} <strong>in</strong>terface stress was found to be −2.17 ± 0.15Nm −1 ,<br />

<strong>in</strong> agreement with previous studies. High <strong>and</strong> low-angle x-ray scatter<strong>in</strong>g experiments show<br />

that this value is unaffected by <strong>in</strong>termix<strong>in</strong>g. The widths <strong>of</strong> <strong>Ag</strong>/<strong>Ni</strong> <strong>and</strong> <strong>Ni</strong>/<strong>Ag</strong> <strong>in</strong>terfaces were<br />

determ<strong>in</strong>ed <strong>and</strong> correlated to an excess volume that derives from the volume expansion <strong>of</strong><br />

misfit dislocations. A positive excess volume is consistent with a negative <strong>in</strong>terface stress<br />

when account<strong>in</strong>g for anharmonic terms <strong>in</strong> the <strong>in</strong>teratomic potential.


Contents<br />

Title Page ...................................... i<br />

Abstract ....................................... iii<br />

Table <strong>of</strong> Contents .................................. iv<br />

List <strong>of</strong> Figures ....................................viii<br />

List <strong>of</strong> Tables ....................................xiii<br />

Acknowledgments .................................. xiv<br />

Dedication ......................................xvii<br />

1 Background 1<br />

1.1 Elasticity ................................... 2<br />

1.2 Dislocation Plasticity ............................. 6<br />

1.3 Th<strong>in</strong> Film Growth <strong>and</strong> Microstructure . ................... 10<br />

1.4 Two Common Stress Determ<strong>in</strong>ation Techniques ............... 10<br />

1.4.1 Wafer <strong>Cu</strong>rvature ........................... 11<br />

1.4.2 X-Ray Diffraction Measurements .................. 14<br />

1.5 Sputter Deposition .............................. 16<br />

2 The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 19<br />

2.1 The Variation <strong>of</strong> Stress with Temperature <strong>in</strong> a Th<strong>in</strong> Film on an Elastic Substrate<br />

...................................... 21<br />

2.2 The Biaxial Modulus Measurement . . . ................... 23<br />

2.2.1 The Effect <strong>of</strong> Texture on the Biaxial Modulus ............ 24<br />

2.3 Experiment .................................. 25<br />

2.3.1 Film Preparation ........................... 25<br />

2.3.2 Film Characterization ......................... 26<br />

2.3.3 X-Ray Diffraction .......................... 26<br />

2.3.4 <strong>Cu</strong>rvature Measurements ....................... 27<br />

2.4 Results ..................................... 28<br />

2.4.1 Microstructure ............................ 28<br />

2.4.2 Film Texture ............................. 28<br />

2.4.3 Biaxial Modulus ........................... 36<br />

iv


Contents v<br />

Measur<strong>in</strong>g Both the Biaxial Modulus <strong>and</strong> the CTE . ........ 36<br />

As-deposited <strong>and</strong> Annealed <strong>Films</strong> at Room Temperature ...... 37<br />

Biaxial Modulus as a Function <strong>of</strong> Temperature . . . ........ 39<br />

Biaxial Modulus as a Function <strong>of</strong> Oscillation Frequency ...... 40<br />

2.5 Discussion ................................... 43<br />

2.5.1 Texture . . .............................. 43<br />

2.5.2 Cause <strong>of</strong> the Modulus Deficit .................... 44<br />

2.6 Conclusions <strong>and</strong> Future Research ....................... 47<br />

3 The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 49<br />

3.1 The <strong>Cu</strong>rrent Underst<strong>and</strong><strong>in</strong>g <strong>of</strong> Plastic Properties <strong>of</strong> Th<strong>in</strong> <strong>Films</strong> ....... 49<br />

3.1.1 High Temperature Deformation ................... 50<br />

3.1.2 Low Temperature Deformation . ................... 51<br />

Discrete Dislocation Models . . ................... 54<br />

Stra<strong>in</strong> Gradient Plasticity <strong>in</strong> Th<strong>in</strong> <strong>Films</strong> ............... 59<br />

3.2 Determ<strong>in</strong><strong>in</strong>g the Stra<strong>in</strong> Pr<strong>of</strong>ile ........................ 63<br />

3.3 Experiment .................................. 64<br />

3.3.1 Film Preparation ........................... 64<br />

3.3.2 Film Characterization ......................... 65<br />

3.3.3 <strong>Cu</strong>rvature Measurements ....................... 66<br />

3.3.4 X-Ray Diffraction .......................... 66<br />

3.4 Results ..................................... 68<br />

3.4.1 Microstructure ............................ 68<br />

3.4.2 <strong>Cu</strong>rvature Results ........................... 70<br />

3.4.3 X-Ray Diffraction Results . . . ................... 72<br />

Measurements from the Surface <strong>of</strong> the Film ............. 72<br />

Measurements from the Film/Substrate Interface . . ........ 73<br />

Comparison with s<strong>in</strong>2ψ 2ψ 2ψ Measurements ............... 74<br />

3.5 Discussion ................................... 75<br />

3.5.1 The As-Deposited <strong>Films</strong> ....................... 77<br />

The Stra<strong>in</strong> Pr<strong>of</strong>ile........................... 78<br />

Anisotropy .............................. 78<br />

3.5.2 The Thermally Cycled <strong>Films</strong> . . ................... 79<br />

The Stra<strong>in</strong> Pr<strong>of</strong>ile........................... 79<br />

Anisotropy .............................. 80<br />

3.6 Conclusions <strong>and</strong> Future Research ....................... 81<br />

4 The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 82<br />

4.1 The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface .......................... 82<br />

4.1.1 The Structure <strong>and</strong> Energy <strong>of</strong> Semicoherent Interfaces ........ 83<br />

Structure <strong>of</strong> the <strong>Ag</strong>/<strong>Ni</strong> {111} Interface ................ 83<br />

Energy <strong>of</strong> the <strong>Ag</strong>/<strong>Ni</strong> {111} Interface ................. 88


Contents vi<br />

4.2 Interfacial Thermodynamics ......................... 89<br />

4.2.1 Fluid-Fluid Interfaces ......................... 89<br />

4.2.2 Solid-Fluid Interfaces ......................... 90<br />

4.2.3 Solid-Solid Interfaces ......................... 92<br />

4.3 Pr<strong>in</strong>ciple <strong>of</strong> the Interface Stress Measurement ................ 95<br />

4.4 Previous Measurements <strong>of</strong> the <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Energy <strong>and</strong> Stress . . 96<br />

4.5 Experiment .................................. 98<br />

4.5.1 Film Preparation ........................... 98<br />

4.5.2 Film Characterization ......................... 98<br />

4.5.3 <strong>Cu</strong>rvature Measurements ....................... 99<br />

4.5.4 X-Ray Diffraction .......................... 99<br />

4.6 Results .....................................100<br />

4.6.1 Microstructure ............................100<br />

4.6.2 Substrate <strong>Cu</strong>rvature Stress . . . ...................101<br />

4.6.3 Volume Stress .............................101<br />

4.6.4 The Interface Stress ..........................105<br />

4.7 Models for the Compressive Interface Stress .................107<br />

4.7.1 Intermix<strong>in</strong>g ..............................107<br />

A Systematic Method <strong>of</strong> Extract<strong>in</strong>g Stra<strong>in</strong> <strong>and</strong> Composition Pr<strong>of</strong>iles<br />

<strong>of</strong> <strong>Multilayers</strong> . . ...................110<br />

Results from Diffuse Scatter<strong>in</strong>g ...................113<br />

4.7.2 Excess Volume / Misfit Dislocations .................117<br />

Excess Volume ............................122<br />

Misfit Dislocations ..........................123<br />

4.8 Conclusions <strong>and</strong> Future Research .......................125<br />

Bibliography 126<br />

A X-Ray Stress Analysis 145<br />

A.1 Untextured Th<strong>in</strong> <strong>Films</strong> ............................147<br />

A.1.1 Biaxial Stress Case ..........................148<br />

Equibiaxial Stress Case ........................148<br />

A.2 Textured Th<strong>in</strong> <strong>Films</strong> <strong>in</strong> Equibiaxial Stress ..................149<br />

A.3 The Stra<strong>in</strong>-Free Lattice Parameter <strong>in</strong> Biaxially Stressed Samples ......150<br />

B X-Ray Errors 152<br />

B.1 M<strong>and</strong>atory Corrections ............................153<br />

B.1.1 Lorenz-Polarization-Absorption Corrections to the Intensity ....153<br />

B.1.2 Refraction Corrections ........................154<br />

B.2 Geometric Errors . ..............................155<br />

B.2.1 Receiv<strong>in</strong>g Position Errors . . . ...................155<br />

B.2.2 Horizontal Divergence Errors . ...................157


Contents vii<br />

B.2.3 Alignment Errors ...........................157<br />

Specimen <strong>and</strong> Beam Displacement ..................157<br />

Effect <strong>of</strong> the ω-Axis Not Correspond<strong>in</strong>g to the 2θ-Axis .......159<br />

B.3 Corrections for a Position Sensitive Detector .................159<br />

B.3.1 Parallax Corrections for a PSD . ...................159<br />

B.3.2 Small Beam Spot Corrections . ...................160<br />

B.3.3 PSD Calibration Constant Errors ...................160<br />

B.4 Count<strong>in</strong>g <strong>and</strong> Fitt<strong>in</strong>g Errors ..........................160<br />

C Alignment Procedure for the <strong>Harvard</strong> Four-Circle Diffractometer 161<br />

C.1 Coarse Alignment . ..............................161<br />

C.2 F<strong>in</strong>e Alignment . . ..............................162<br />

C.3 Align<strong>in</strong>g Samples . ..............................163<br />

C.4 Calibrat<strong>in</strong>g the Position Sensitive Detector ..................164<br />

D Elastic Constants 165<br />

D.1 Important Elastic Quantities .........................165<br />

D.2 The Rotation Matrix .............................166<br />

D.3 Average Elastic Quantities For Polycrystall<strong>in</strong>e <strong>Ag</strong>gregates . ........167<br />

D.4 The Vook-Witt Model <strong>of</strong> Gra<strong>in</strong> Interaction ..................167<br />

D.4.1 The Hornstra-Bartels Calculation Technique .............169<br />

D.4.2 Results with the Vook-Witt Model ..................171<br />

E The Stoney Equation 175<br />

F List <strong>of</strong> Symbols 179


List <strong>of</strong> Figures<br />

1.1 Interatomic potentials (a) <strong>and</strong> their first (b) <strong>and</strong> second (c) derivatives.<br />

Both the Lennard-Jones potential (solid) [140] <strong>and</strong> a harmonic potential<br />

(dashed) are shown. .............................. 4<br />

1.2 (a) A depiction <strong>of</strong> the vertical force balance between dislocation l<strong>in</strong>e tension<br />

<strong>and</strong> the Peach-Koehler force. (b) An illustration <strong>of</strong> a dislocation bow<strong>in</strong>g<br />

between two impenetrable obstacles with separation ℓ. ......... 8<br />

1.3 A schematic show<strong>in</strong>g how the film stress causes substrate curvature. (a)<br />

The film <strong>and</strong> substrate are both shown <strong>in</strong> their stra<strong>in</strong>-free state. (b) The<br />

film is stretched to fit onto the substrate. (c) The film is “glued” onto the<br />

substrate. (d) The film/substrate system bends <strong>in</strong> order to achieve force<br />

<strong>and</strong> moment balance. ............................. 12<br />

1.4 A schematic <strong>of</strong> the curvature measurement apparatus at <strong>Harvard</strong> University.<br />

Adapted from Mull<strong>in</strong> [158]. ......................... 13<br />

1.5 A schematic <strong>of</strong> the four-circle Huber diffractometer at <strong>Harvard</strong> University. . 16<br />

1.6 A schematic <strong>of</strong> different scatter<strong>in</strong>g geometries. (a) The symmetric θ −<br />

2θ geometry <strong>and</strong> χ rotation axis, (b) the geometry used for longitud<strong>in</strong>al<br />

diffuse scatter<strong>in</strong>g, <strong>and</strong> (c) the graz<strong>in</strong>g <strong>in</strong>cidence diffraction geometry. . . . 17<br />

1.7 A schematic <strong>of</strong> the sputter<strong>in</strong>g chamber at <strong>Harvard</strong> University. ........ 18<br />

2.1 Stress as a function <strong>of</strong> temperature for a 1.91 µm sputtered <strong>Cu</strong> film on<br />

Si(100). The dotted l<strong>in</strong>es show the hypothetical elastic stress <strong>in</strong> the film<br />

if no relaxation could take place (thermoelastic l<strong>in</strong>es), <strong>and</strong> the labels (A),<br />

(B), <strong>and</strong> (C) denote elastic thermal cycles as described <strong>in</strong> the text. The<br />

upper thermoelastic l<strong>in</strong>e was drawn us<strong>in</strong>g the elastic constant data found <strong>in</strong><br />

Bechmann et al. [12] <strong>and</strong> assum<strong>in</strong>g mild 〈111〉 texture. . . . ........ 21<br />

2.2 Plan-view TEM micrographs <strong>of</strong> 0.87 µm thick copper <strong>in</strong> the as-deposited<br />

state (a) <strong>and</strong> after anneal<strong>in</strong>g at 600 ◦ C for 0.5 hrs (b). ............ 29<br />

2.3 A cross-sectional TEM micrograph <strong>of</strong> 1.90 µm thick copper grown on<br />

Ge(111). .................................... 30<br />

viii


List <strong>of</strong> Figures ix<br />

2.4 (a) Powder diffraction x-ray spectra for sputter-deposited copper films on<br />

Ge(111) (dashed), α-Al2O3(0001) (dotted), <strong>and</strong> Si(100) (solid). (b) (111)<br />

fiber texture plot <strong>of</strong> copper films on Ge(111) (dashed) <strong>and</strong> Si(100) (solid). . 31<br />

2.5 (a) Powder diffraction spectra, <strong>and</strong> (b) a (111) fiber texture plot for asdeposited<br />

<strong>and</strong> annealed copper films on Si(100). .............. 32<br />

2.6 Powder diffraction spectra for annealed copper films on Si(100) (solid) <strong>and</strong><br />

α-Al2O3(0001) (dotted). ........................... 33<br />

2.7 (111) fiber texture plot <strong>of</strong> as-deposited <strong>and</strong> annealed 0.50 µm thick copper<br />

films. ..................................... 33<br />

2.8 (111) fiber texture plot <strong>of</strong> an electron beam evaporated copper film. .... 34<br />

2.9 The calculated biaxial modulus for 〈111〉 texture (a) <strong>and</strong> 〈100〉 texture (b)<br />

as a function <strong>of</strong> the texture pole st<strong>and</strong>ard deviation. A Gaussian texture<br />

pole is assumed. . . .............................. 35<br />

2.10 Plot show<strong>in</strong>g how the biaxial modulus <strong>and</strong> CTE can be determ<strong>in</strong>ed for a set<br />

<strong>of</strong> different substrates. ............................. 37<br />

2.11 Biaxial modulus as a function <strong>of</strong> the maximum anneal<strong>in</strong>g temperature exposed<br />

to a 1.88 µm thick copper film on Si(100). .............. 38<br />

2.12 Biaxial modulus as a function <strong>of</strong> the number <strong>of</strong> times thermally cycled to<br />

600 ◦ Cfora1.91 µm <strong>and</strong> 1.93 µm thick copper film. ............ 39<br />

2.13 Stress as a function <strong>of</strong> temperature for a 1.19 µm thick <strong>Cu</strong> film on Si(100).<br />

The elastic thermal cycles for the measurements <strong>of</strong> biaxial modulus at five<br />

different temperatures are shown for the as-deposited film (a) <strong>and</strong> the thermally<br />

cycled film (b). ............................. 40<br />

2.14 Biaxial modulus as a function <strong>of</strong> temperature for as-deposited copper films.<br />

The error bar denotes two times the st<strong>and</strong>ard deviation. The shaded area<br />

represents the calculated biaxial moduli after account<strong>in</strong>g for texture. It is<br />

based on a value <strong>of</strong> 226 ± 22 GPa at room temperature (cf. Table 2.4) <strong>and</strong><br />

has the temperature dependence found <strong>in</strong> Simmons <strong>and</strong> Wang [207]. .... 41<br />

2.15 Biaxial modulus as a function <strong>of</strong> temperature for a copper film previously<br />

annealed at 600 ◦ C. The error bar denotes two times the st<strong>and</strong>ard deviation.<br />

The shaded area represents the calculated biaxial moduli after account<strong>in</strong>g<br />

for texture. It is based on a value <strong>of</strong> 256 ± 22 GPa at room temperature<br />

(cf. Table 2.4) <strong>and</strong> has the temperature dependence found <strong>in</strong> Simmons <strong>and</strong><br />

Wang [207]. .................................. 42<br />

2.16 Biaxial modulus as a function <strong>of</strong> temperature for copper films previously<br />

thermally cycled to 600 ◦ C for different stra<strong>in</strong> oscillation frequencies. . . . 43<br />

2.17 Depiction <strong>of</strong> dislocation bow<strong>in</strong>g at high temperatures. Thermal fluctuations<br />

help the dislocation to bow past weak obstacles, but the radius <strong>of</strong> curvature<br />

rema<strong>in</strong>s below that required for slip. . . ................... 46


List <strong>of</strong> Figures x<br />

3.1 Diagram show<strong>in</strong>g the nucleation <strong>and</strong> propagation <strong>of</strong> dislocations along a<br />

slip plane from the surface to the gra<strong>in</strong> boundaries <strong>and</strong> film/substrate <strong>in</strong>terface.<br />

...................................... 55<br />

3.2 Schematic show<strong>in</strong>g the predicted dislocation behavior at different po<strong>in</strong>ts<br />

on the stress-temperature pr<strong>of</strong>ile. Based on the diagram by Weihnacht <strong>and</strong><br />

Brückner [245]. . . .............................. 57<br />

3.3 The scatter<strong>in</strong>g geometry for the GID measurements. As shown <strong>in</strong> (a), the<br />

2θ value is not the same as the motor position for the detector, 2θ ′ . The correction<br />

can be made us<strong>in</strong>g the well known formulas for the right-spherical<br />

triangle shown <strong>in</strong> (b). ............................. 67<br />

3.4 Cross-sectional views <strong>of</strong> an as-deposited (a) <strong>and</strong> annealed (b) 0.87 µm<br />

thick copper film................................ 69<br />

3.5 Deposition stress for ion beam sputtered <strong>Cu</strong> films as a function <strong>of</strong> total film<br />

thickness. ................................... 70<br />

3.6 Room temperature flow strength as a function <strong>of</strong> <strong>in</strong>verse thickness for <strong>Cu</strong><br />

samples thermally cycled to 660 ◦ C. ..................... 71<br />

3.7 The stress-temperature pr<strong>of</strong>iles for the <strong>Cu</strong> films used <strong>in</strong> the stra<strong>in</strong> gradient<br />

studies. .................................... 71<br />

3.8 Lattice parameter <strong>of</strong> CeO2 as a function <strong>of</strong> <strong>in</strong>cidence angle, α. The lattice<br />

parameters were measured with the (331) reflection. ............. 72<br />

3.9 Elastic stra<strong>in</strong> gradients <strong>in</strong> <strong>Cu</strong> films. The solid <strong>and</strong> dotted curves refer to the<br />

thermally cycled films <strong>and</strong> as-deposited films, respectively. The reflection<br />

is denoted by different markers, as shown <strong>in</strong> the legend. The plastic stra<strong>in</strong><br />

is given by 0.9% m<strong>in</strong>us the elastic stra<strong>in</strong>. ................... 73<br />

3.10 s<strong>in</strong> 2 ψ-plot for an as-deposited <strong>Cu</strong> film. ................... 74<br />

3.11 s<strong>in</strong> 2 ψ-plot for a thermally cycled <strong>Cu</strong> film................... 75<br />

3.12 Hypothetical stra<strong>in</strong> pr<strong>of</strong>iles <strong>and</strong> their normalized f<strong>in</strong>ite thickness Laplace<br />

transforms for a 1.8 µm thick <strong>Cu</strong> film. .................... 77<br />

4.1 The <strong>Ag</strong>/<strong>Ni</strong> phase diagram [149]. ....................... 85<br />

4.2 Top view <strong>of</strong> the <strong>Ag</strong>/<strong>Ni</strong> {111} <strong>in</strong>terface, with <strong>Ag</strong> <strong>and</strong> <strong>Ni</strong> denoted by closed<br />

<strong>and</strong> open circles, respectively. After relaxation, misfit dislocations are<br />

formed, as del<strong>in</strong>eated by the dark l<strong>in</strong>es. The coherent areas <strong>and</strong> <strong>in</strong>tr<strong>in</strong>sic<br />

stack<strong>in</strong>g fault areas lie <strong>in</strong> the centers <strong>of</strong> the white <strong>and</strong> shaded triangles,<br />

respectively. .................................. 86<br />

4.3 An edge-on view <strong>of</strong> the <strong>Ag</strong>/<strong>Ni</strong> {111} <strong>in</strong>terface along the [110] direction,<br />

with the <strong>Ag</strong> <strong>and</strong> <strong>Ni</strong> atoms denoted by closed <strong>and</strong> open circles, respectively.<br />

The large circles are atoms <strong>in</strong> the plane <strong>of</strong> the diagram; the small circles are<br />

atoms <strong>in</strong> the planes immediately above <strong>and</strong> below the plane <strong>of</strong> the figure.<br />

Two dist<strong>in</strong>ct dislocation nodes can be seen <strong>and</strong> associated with the different<br />

stack<strong>in</strong>g conditions. Adapted from Gao <strong>and</strong> Merkle [75]. . . ........ 87


List <strong>of</strong> Figures xi<br />

4.4 A schematic demonstrat<strong>in</strong>g how misfitt<strong>in</strong>g surface layers can lead to <strong>in</strong>ternal<br />

bulk stresses. . .............................. 93<br />

4.5 A schematic show<strong>in</strong>g how <strong>in</strong>terface stresses can lead to <strong>in</strong>ternal bulk stresses. 93<br />

4.6 A cross-sectional view <strong>of</strong> a <strong>Ag</strong>/<strong>Ni</strong> multilayer with � = 4.2nm........101<br />

4.7 Offset low-angle θ − 2θ diffraction spectra for the three smallest bilayer<br />

thicknesses. The spectra were taken with an x-ray energy <strong>of</strong> 8020 eV. . . . 102<br />

4.8 (a) Volume stresses <strong>in</strong> <strong>Ni</strong> (closed squares) <strong>and</strong> <strong>Ag</strong> (open squares) layers as<br />

a function <strong>of</strong> <strong>in</strong>verse bilayer thickness. (b) Substrate curvature stress (open<br />

circles) <strong>and</strong> average volume stress (closed circles) as a function <strong>of</strong> <strong>in</strong>verse<br />

bilayer thickness. The error bars for all the data are with<strong>in</strong> the size <strong>of</strong> the<br />

symbols. The arrow drawn from (a) to (b) illustrates how the layer stresses<br />

are averaged to give 〈σ 〉x−ray. .........................103<br />

4.9 Offset GID spectra for <strong>Ag</strong>/<strong>Ni</strong> multilayers hav<strong>in</strong>g the <strong>in</strong>dicated bilayer thickness<br />

�. The vertical l<strong>in</strong>es <strong>in</strong>dicate the {220} peak positions for stra<strong>in</strong>-free<br />

bulk <strong>Ag</strong> <strong>and</strong> <strong>Ni</strong>. The x-ray energy was 8020 eV. ...............104<br />

4.10 In-plane stra<strong>in</strong>s <strong>in</strong> nickel <strong>and</strong> silver as a function <strong>of</strong> bilayer thickness. . . . 105<br />

4.11 Difference between the substrate curvature stress <strong>and</strong> the average volume<br />

stress for <strong>Ag</strong>/<strong>Ni</strong> multilayers as a function <strong>of</strong> <strong>in</strong>verse bilayer thickness. The<br />

l<strong>in</strong>e is a weighted least-squares fit for the three smallest bilayer thicknesses. 106<br />

4.12 The average {111} d-spac<strong>in</strong>g for nickel. S<strong>in</strong>ce the nickel is <strong>in</strong> tension parallel<br />

to the {111} planes, d111 is expected to be less than the stra<strong>in</strong>-free<br />

d-spac<strong>in</strong>g <strong>of</strong> 0.20345 nm for a positive Poisson ratio. ............108<br />

4.13 Depiction <strong>of</strong> the k<strong>in</strong>etics <strong>of</strong> segregation. At cold temperatures, the <strong>Ag</strong> lacks<br />

the mobility to segregate. At high temperatures, the <strong>Ag</strong> “floats” to the<br />

surface. At <strong>in</strong>termediate temperatures, a compositional gradient exists. . . . 110<br />

4.14 Depiction <strong>of</strong> a superlattice unit cell us<strong>in</strong>g the parameters <strong>in</strong> equation (4.34). 111<br />

4.15 High-angle x-ray diffraction peaks for a <strong>Ag</strong>/<strong>Ni</strong> multilayer with � = 4.2nm.<br />

The order <strong>of</strong> the satellite is given above its peak. The data taken at E =<br />

8320 eV are <strong>of</strong>fset from the E = 8020 eV data. ...............112<br />

4.16 An illustration <strong>of</strong> a reciprocal space map along qx <strong>and</strong> qz. The specular<br />

ridge samples along qz, transverse scans sample along qx, <strong>and</strong> longitud<strong>in</strong>al<br />

scans sample both components. ........................114<br />

4.17 Specular reflectivity pr<strong>of</strong>iles (a) <strong>and</strong> longitud<strong>in</strong>al diffuse scatter<strong>in</strong>g pr<strong>of</strong>iles<br />

(b). The tilt <strong>of</strong> the sample from the specular condition is <strong>in</strong>dicated by �θ,<br />

<strong>and</strong> the x-ray spectra taken at E = 8020 eV <strong>and</strong> E = 8320 eV are <strong>of</strong>fset.<br />

The data <strong>and</strong> fitted/simulated pr<strong>of</strong>iles are given by the open circles <strong>and</strong><br />

solid (or dashed) curves, respectively. The solid curves show the results <strong>of</strong><br />

a simulation with a <strong>Ag</strong>/<strong>Ni</strong> (surface side / substrate side) chemical roughness<br />

<strong>of</strong> 0.24 nm <strong>and</strong> a <strong>Ni</strong>/<strong>Ag</strong> chemical roughness <strong>of</strong> 0.30 nm. The <strong>of</strong>fset dashed<br />

curves show the results <strong>of</strong> a simulation with a <strong>Ag</strong>/<strong>Ni</strong> chemical roughness<br />

<strong>of</strong> 0.30 nm <strong>and</strong> a <strong>Ni</strong>/<strong>Ag</strong> chemical roughness <strong>of</strong> 0.40nm............115


List <strong>of</strong> Figures xii<br />

4.18 (a) A schematic show<strong>in</strong>g the width between <strong>Ag</strong> <strong>and</strong> <strong>Ni</strong> (111) planes at<br />

the <strong>in</strong>terface. (b) The <strong>in</strong>terface width for <strong>Ag</strong>/<strong>Ni</strong> <strong>and</strong> <strong>Ni</strong>/<strong>Ag</strong> <strong>in</strong>terfaces.<br />

The mean (111) d-spac<strong>in</strong>g, 0.2197 nm, is <strong>in</strong>dicated by the horizontal l<strong>in</strong>e.<br />

(d111,<strong>Ag</strong> = 0.23592 nm, <strong>and</strong> d111,<strong>Ni</strong> = 0.20345 nm). .............118<br />

4.19 Schematic show<strong>in</strong>g the method by which SUPREX treats gradients <strong>in</strong> the<br />

out-<strong>of</strong>-plane d-spac<strong>in</strong>g. ............................119<br />

4.20 High-angle x-ray diffraction peaks for the � = 3.2nm(a) <strong>and</strong> � = 5.9nm<br />

(b) <strong>Ag</strong>/<strong>Ni</strong> multilayers. The <strong>in</strong>tensity axis is on a logarithmic scale, <strong>and</strong> the<br />

data taken at E = 8320 eV is <strong>of</strong>fset above the E = 8020 eV data. The<br />

circles are the measured <strong>in</strong>tensities, the solid curve is the fit for the stra<strong>in</strong><br />

gradient model, <strong>and</strong> the dotted curve is the best fit when stra<strong>in</strong> gradients<br />

are disallowed. The order <strong>of</strong> the reflection is <strong>in</strong>dicated above the peak. . . . 119<br />

4.21 Calculated {111} d-spac<strong>in</strong>g pr<strong>of</strong>iles for the � = 3.2nm(a) <strong>and</strong> 5.9nm(b)<br />

<strong>Ag</strong>/<strong>Ni</strong> multilayers. The results <strong>of</strong> Jaszczak et al. for their theoretical multilayer<br />

are provided <strong>in</strong> (c) [113]. The vertical dashed l<strong>in</strong>es are the stra<strong>in</strong>-free<br />

d-spac<strong>in</strong>gs for the bulk phases. ........................120<br />

A.1 Diagram show<strong>in</strong>g the relationship between the sample coord<strong>in</strong>ate system,<br />

S, <strong>and</strong> the laboratory coord<strong>in</strong>ate system, L. .................146<br />

B.1 The focus<strong>in</strong>g circle. ..............................156<br />

D.1 The ratio <strong>of</strong> ɛ ′ 33 /ɛ′ D.2<br />

11 <strong>in</strong> the Vook-Witt model. .................172<br />

The biaxial modulus <strong>in</strong> units <strong>of</strong> 100 GPa us<strong>in</strong>g the Vook-Witt model. ....173<br />

D.3 The maximum resolved shear stress on the (111) slip plane given <strong>in</strong> GPa<br />

us<strong>in</strong>g ɛ ′ 11 = 0.704%. .............................174


List <strong>of</strong> Tables<br />

1.1 Ability <strong>of</strong> wafer curvature <strong>and</strong> x-ray techniques to measure different types<br />

<strong>of</strong> stresses. Adapted from Ruud [195]. . ................... 11<br />

1.2 Parameters for ion beam sputter deposition <strong>and</strong> substrate clean<strong>in</strong>g. ..... 18<br />

2.1 Measured <strong>and</strong> calculated values <strong>of</strong> the Young modulus, E, <strong>and</strong> the biaxial<br />

modulus, Y , for both free-st<strong>and</strong><strong>in</strong>g <strong>and</strong> substrate-bonded films on the order<br />

<strong>of</strong> 1 µm thick. . . . .............................. 20<br />

2.2 Parameters for a-SiNx reactive sputter deposition. .............. 26<br />

2.3 Thermomechanical properties <strong>of</strong> the substrates used for the thermal cycl<strong>in</strong>g<br />

experiments. .................................. 27<br />

2.4 Calculated values <strong>of</strong> the biaxial modulus for selected films. The st<strong>and</strong>ard<br />

deviation is ±22GPa.............................. 35<br />

2.5 Measured values <strong>of</strong> the biaxial modulus for films <strong>of</strong> thickness 1 µm or greater. 38<br />

2.6 Comparison <strong>of</strong> the measured <strong>and</strong> calculated values <strong>of</strong> the biaxial modulus<br />

for the as-deposited films. ........................... 43<br />

2.7 Comparison <strong>of</strong> the measured <strong>and</strong> calculated values <strong>of</strong> the biaxial modulus<br />

for the annealed films. ............................. 44<br />

4.1 The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Energy . . . ................... 96<br />

4.2 {111} Interface Stress Results for FCC Metallic <strong>Multilayers</strong> . ........ 97<br />

xiii


Acknowledgments<br />

The acknowledgments is the only section anyone ever reads. That means I should have<br />

started writ<strong>in</strong>g it long ago, rather than wait<strong>in</strong>g until the eleventh hour. I would have loved<br />

to make it witty <strong>and</strong> thoughtful, but there’s no time. Plus, for some reason, it’s not very<br />

easy to write. I hope nobody reads the acknowledgments simply to see what I thought <strong>of</strong><br />

him or her. I love everybody here — it’s been great.<br />

First <strong>and</strong> foremost, thanks to Pr<strong>of</strong>. Frans Spaepen, my advisor. Anyone who meets him<br />

knows how <strong>in</strong>telligent he is. Hav<strong>in</strong>g been his student, I also know how great he is as a<br />

person. I’m honored to have been his advisee.<br />

Thanks to Pr<strong>of</strong>s. Mike Aziz <strong>and</strong> John Hutch<strong>in</strong>son. Along with Frans, these two demon-<br />

strate that it’s possible to be a brilliant scientist, a good teacher, <strong>and</strong> a nice person, prov<strong>in</strong>g<br />

that these traits are not necessarily mutually exclusive.<br />

Thanks to Dr. Denis McWhan. Without him, there would have been no <strong>Ag</strong>/<strong>Ni</strong> project.<br />

He taught me everyth<strong>in</strong>g I know about synchrotrons <strong>and</strong> gave me a glimpse <strong>of</strong> the way sci-<br />

ence is done outside academia; plus, he gave me that near-death experience people some-<br />

times talk about. Well okay, it wasn’t quite near-death, but it had potential. Thanks for<br />

keep<strong>in</strong>g your cool, Denis.<br />

Thanks to Pr<strong>of</strong>. David Wu. He’s another great guy I’m honored to know. Despite what<br />

I’ve told him, I hope he never changes. Science needs more people like him.<br />

Thanks to Pr<strong>of</strong>. David Turnbull. I don’t feel like I know him very well, but he’s my<br />

hero, nonetheless. He’s the epitome <strong>of</strong> a scientist, as well as an outst<strong>and</strong><strong>in</strong>g human be-<br />

<strong>in</strong>g. Everyone <strong>in</strong> the materials sciences, <strong>and</strong> especially those at <strong>Harvard</strong> obviously owe a<br />

tremendous amount to him.<br />

Thanks to the staff here at <strong>Harvard</strong>. There’s Yuan Lu, Warren MoberlyChan, John


Acknowledgments xv<br />

Cherv<strong>in</strong>sky, <strong>and</strong> others. They’re all very good at what they do, <strong>and</strong> they’re good people, to<br />

boot. And, <strong>of</strong> course, there’s Frank Molea: to me, Frank represents everyth<strong>in</strong>g I love about<br />

the fourth floor <strong>of</strong> McKay, <strong>and</strong> I’m proud to know him <strong>and</strong> call him my friend. I wish the<br />

best for him.<br />

Thanks to Denis Yu. Denis is the most selfless guy I’ve every known; do<strong>in</strong>g th<strong>in</strong>gs<br />

above <strong>and</strong> beyond the call <strong>of</strong> duty/friendship is normal for him. I’m embarrassed to th<strong>in</strong>k<br />

<strong>of</strong> the number <strong>of</strong> times he’s bailed me out. Good luck to you, Denis. Please let me know if<br />

there’s any way I can possibly pay you back.<br />

Thanks to Andrea Del Vecchio <strong>and</strong> Jen Sage; they’ve been great friends <strong>and</strong> <strong>of</strong>fice-<br />

mates. Andrea <strong>and</strong> V<strong>in</strong>ce have been my surrogate family this past year. Jen also watches<br />

out for me by br<strong>in</strong>g<strong>in</strong>g back leftovers from the Women <strong>in</strong> Physics meet<strong>in</strong>gs. Thanks also<br />

to Anita Bowles, Jason Draut, Jeff Warrender, Bola George, <strong>and</strong> Cheng-Yen Wen. I can’t<br />

th<strong>in</strong>k <strong>of</strong> a more car<strong>in</strong>g group <strong>of</strong> <strong>in</strong>dividuals. Special mention goes to Cheng-Yen. He took<br />

all the TEM images used <strong>in</strong> this thesis. He’s some sort <strong>of</strong> TEM prodigy, if such a th<strong>in</strong>g<br />

exists. My wish for him is that he doesn’t get stuck do<strong>in</strong>g TEM for everybody like he did<br />

for me.<br />

Thanks to the post-docs: John Leonard, Alex <strong>Cu</strong>enat, Manoj Pillai, Yuechao Zhao, <strong>and</strong><br />

Joris Proost. With the exception <strong>of</strong> John, this club <strong>of</strong> post-docs is like my guide to the<br />

world outside the states. Interest<strong>in</strong>gly, John is the guide to everywhere with<strong>in</strong> the states.<br />

I’ve enjoyed chatt<strong>in</strong>g with all <strong>of</strong> you.<br />

The dynamics <strong>of</strong> the lab change every year. I could go through all the names <strong>of</strong> past<br />

students <strong>and</strong> post-docs, but it’s really po<strong>in</strong>tless s<strong>in</strong>ce I know they won’t come back to read<br />

my silly acknowledgments. However, <strong>in</strong> case they do, I thank them, too.


Acknowledgments xvi<br />

I would be remiss, however, if I didn’t mention Tim Van Rompaey. He worked on <strong>Ni</strong>/Ti<br />

multilayers for a summer over here, <strong>and</strong> he’s a great guy. I hope he’s enjoy<strong>in</strong>g graduate<br />

school as much as I did.<br />

My mom <strong>and</strong> dad have been terrific. I might as well thank them here for shap<strong>in</strong>g me<br />

<strong>in</strong>to who I am. Everyone I meet <strong>and</strong> talk to affects me, but these are only higher order<br />

effects. Thank you, Mom <strong>and</strong> Dad.<br />

F<strong>in</strong>ally, thanks to Jessica for lov<strong>in</strong>g me unconditionally. I’m honestly not sure why she<br />

does, but I’m all the happier for it. I hope we’ll spend the rest <strong>of</strong> our lives together.


Dedicated to my family


Chapter 1<br />

Background<br />

This thesis deals with the measurement <strong>of</strong> stresses <strong>in</strong> th<strong>in</strong> films <strong>and</strong> multilayers us<strong>in</strong>g<br />

two different techniques: wafer curvature <strong>and</strong> x-ray diffraction. The techniques are com-<br />

plementary <strong>in</strong> that wafer curvature measures the average biaxial stress <strong>in</strong> the film, while<br />

x-ray diffraction measures the stra<strong>in</strong> with<strong>in</strong> the diffract<strong>in</strong>g gra<strong>in</strong>s. This stra<strong>in</strong> can be con-<br />

verted <strong>in</strong>to a stress if the reference lattice parameter <strong>and</strong> the elastic constants <strong>of</strong> the gra<strong>in</strong><br />

are known.<br />

In Chapter 2, I will discuss the elastic properties <strong>of</strong> copper th<strong>in</strong> films. Our <strong>in</strong>terest is not<br />

<strong>in</strong> the “true” modulus that depends on the curvature <strong>of</strong> the potential energy function <strong>and</strong> is<br />

usually measured by dynamic methods, but <strong>in</strong> the macroscopic <strong>and</strong> static modulus that is<br />

sensitive to material defects. We f<strong>in</strong>d the modulus <strong>of</strong> copper films is reduced relative to the<br />

dynamic value; it also falls <strong>of</strong>f more quickly than the dynamic value at higher temperatures<br />

(≈ 0.4 × the melt<strong>in</strong>g temperature). Anelastic mechanisms appear to contribute to the<br />

modulus reduction.<br />

In the next section, I describe our experiments on the plastic behavior <strong>of</strong> copper th<strong>in</strong><br />

1


Chapter 1: Background 2<br />

films on rigid substrates. While “smaller is stronger” is now an old saw <strong>in</strong> the th<strong>in</strong> film<br />

mechanics world, why it is so rema<strong>in</strong>s an issue <strong>of</strong> debate. A materials scientist looks at<br />

the problem from a dislocation st<strong>and</strong>po<strong>in</strong>t; a mechanician considers constitutive equations.<br />

Stra<strong>in</strong> gradient plasticity (SGP) merges these viewpo<strong>in</strong>ts <strong>and</strong> thus is a mathematically <strong>and</strong><br />

physically satisfy<strong>in</strong>g theory. It is still a theory under development, however, <strong>and</strong> requires<br />

additional experimental support. Although we were unable to verify one <strong>of</strong> the predictions<br />

<strong>of</strong> SGP <strong>in</strong> th<strong>in</strong> films, our experiment can be used as a start<strong>in</strong>g po<strong>in</strong>t for future attempts.<br />

F<strong>in</strong>ally, we will deal with <strong>in</strong>terface stresses <strong>in</strong> <strong>Ag</strong>/<strong>Ni</strong> multilayers. This field is also<br />

t<strong>in</strong>ged with controversy, although not for a lack <strong>of</strong> good reason. J. W. Gibbs died before<br />

formulat<strong>in</strong>g the thermodynamics <strong>of</strong> solid-solid <strong>in</strong>terfaces, <strong>and</strong> we’ve been struggl<strong>in</strong>g ever<br />

s<strong>in</strong>ce. We show that the <strong>in</strong>terface stress <strong>of</strong> <strong>Ag</strong>/<strong>Ni</strong> {111} <strong>in</strong>terfaces is compressive, i.e., the<br />

<strong>in</strong>terfaces push out. This is slightly counter<strong>in</strong>tuitive, but the compressive nature results<br />

from higher order effects <strong>in</strong> the <strong>in</strong>teratomic potential. (Our <strong>in</strong>tuition is usually only good to<br />

first order). I also show that our method <strong>of</strong> measur<strong>in</strong>g the <strong>in</strong>terface stress is <strong>in</strong>deed valid for<br />

<strong>Ag</strong>/<strong>Ni</strong>; for other multilayers, complications such as <strong>in</strong>termix<strong>in</strong>g can affect the measurement<br />

<strong>and</strong> should not be ignored.<br />

S<strong>in</strong>ce we will discuss stresses <strong>and</strong> stra<strong>in</strong>s <strong>in</strong> th<strong>in</strong> films, this chapter will review some<br />

<strong>of</strong> the essential mechanics <strong>and</strong> th<strong>in</strong> film basics (growth <strong>and</strong> microstructure). Subsequently,<br />

I will describe the wafer curvature <strong>and</strong> x-ray method <strong>of</strong> extract<strong>in</strong>g stresses <strong>and</strong> stra<strong>in</strong>s, as<br />

well as expla<strong>in</strong> how we grow our films <strong>and</strong> multilayers.<br />

1.1 Elasticity<br />

In what follows, I will only consider “stretch” deformations <strong>and</strong> no rotations.


Chapter 1: Background 3<br />

The elastic stra<strong>in</strong>, i.e., the time-<strong>in</strong>dependent reversible stra<strong>in</strong>, results from atoms resid-<br />

<strong>in</strong>g at high energy positions away from the equilibrium spac<strong>in</strong>g, a0. Figure 1.1 (a) shows<br />

two pair potentials; the solid l<strong>in</strong>e corresponds to a common nonconvex pair potential, the<br />

Lennard-Jones [140], <strong>and</strong> the dashed l<strong>in</strong>e corresponds to a harmonic potential. Their asso-<br />

ciated first <strong>and</strong> second derivatives are depicted <strong>in</strong> Figure 1.1 (b) <strong>and</strong> (c). Wefirst consider<br />

the nonconvex potential.<br />

By expansion around the equilibrium <strong>in</strong>teratomic spac<strong>in</strong>g, the energy per unit volume<br />

<strong>of</strong> a crystal can be expressed as<br />

w = w0 + 1<br />

2! cijklηijηkl + 1<br />

3! cijklmnηijηklηmn + ... (1.1)<br />

where the cs are nth order elastic constants correspond<strong>in</strong>g to the nth order derivatives <strong>of</strong><br />

energy <strong>and</strong> the ηs are Lagrangian stra<strong>in</strong>s. (Appendix F has a list <strong>of</strong> symbols used <strong>in</strong> the<br />

thesis). They are def<strong>in</strong>ed <strong>in</strong> terms <strong>of</strong> the displacement vector u by<br />

ηij = 1 � �<br />

ui, j + u j,i + uk,iuk, j . (1.2)<br />

2<br />

The Lagrangian stra<strong>in</strong>s are referenced to the undeformed state, as opposed to Eulerian<br />

stra<strong>in</strong>s, which are referenced to the deformed state. The thermodynamic tensions, Tij, are<br />

then def<strong>in</strong>ed by the differential <strong>of</strong> stra<strong>in</strong> energy with respect to the Lagrangian stra<strong>in</strong>s:<br />

dw = Tij dηij. (1.3)<br />

If all the components <strong>of</strong> the displacement gradient are small, i.e., |ui, j| ≪1, the La-<br />

grangian stra<strong>in</strong>s reduce to<br />

ɛij = 1 � �<br />

ui, j + u j,i , (1.4)<br />

2


Chapter 1: Background 4<br />

σ = F/a 0 2<br />

Figure 1.1: Interatomic potentials (a) <strong>and</strong> their first (b) <strong>and</strong> second (c) derivatives. Both<br />

the Lennard-Jones potential (solid) [140] <strong>and</strong> a harmonic potential (dashed) are shown.


Chapter 1: Background 5<br />

<strong>and</strong> the Hookean stresses are def<strong>in</strong>ed by<br />

dw = σij dɛij. (1.5)<br />

For the small stra<strong>in</strong> assumption, the dist<strong>in</strong>ction between the undeformed <strong>and</strong> deformed state<br />

is unnecessary.<br />

Mak<strong>in</strong>g the small stra<strong>in</strong> approximation is tantamount to choos<strong>in</strong>g a harmonic potential<br />

(the dashed curve <strong>in</strong> Figure 1.1 (a)). The stress then varies l<strong>in</strong>early with stra<strong>in</strong>, as shown<br />

<strong>in</strong> Figure 1.1 (b), <strong>and</strong> we may write<br />

σij = cijklɛkl (1.6a)<br />

or ɛij = sijklσkl (1.6b)<br />

depend<strong>in</strong>g on which is the <strong>in</strong>dependent variable. For the one-dimensional model depicted<br />

<strong>in</strong> Figure 1.1, c1111 is <strong>in</strong>dependent <strong>of</strong> stra<strong>in</strong> (Figure 1.1 (c)).<br />

In addition to writ<strong>in</strong>g the stress <strong>in</strong> terms <strong>of</strong> stra<strong>in</strong>, I can also write stress <strong>in</strong> terms <strong>of</strong> its<br />

def<strong>in</strong>ition <strong>of</strong> force per unit area:<br />

σij = Fi<br />

A j<br />

where the area vector po<strong>in</strong>ts along the plane normal.<br />

(1.7)<br />

A material’s measured elastic constants may differ substantially from those calculated<br />

from the curvature <strong>of</strong> the potential well. This may result from either po<strong>in</strong>t, l<strong>in</strong>e, areal, or<br />

volume defects <strong>in</strong> the material. If the properties are time-dependent <strong>and</strong> reversible, they are<br />

anelastic. The reversible motion <strong>of</strong> defects can cause an abnormally large stra<strong>in</strong>, thereby<br />

reduc<strong>in</strong>g the measured elastic constant. This phenomenon may or may not be measurable<br />

depend<strong>in</strong>g on the time scale <strong>of</strong> the experiment s<strong>in</strong>ce the defect k<strong>in</strong>etics have characteristic<br />

time constants.


Chapter 1: Background 6<br />

For more <strong>in</strong>formation on l<strong>in</strong>ear elasticity, the texts by L<strong>and</strong>au <strong>and</strong> Lifshitz [136], Timo-<br />

shenko <strong>and</strong> Goodier [224], <strong>and</strong> Sokolnik<strong>of</strong>f [210] are recommended. F<strong>in</strong>ite-stra<strong>in</strong> elasticity<br />

is discussed by Murgnahan [169] <strong>and</strong> Gurt<strong>in</strong> [91], <strong>and</strong> details on the higher order elastic<br />

constants can be found <strong>in</strong> the sem<strong>in</strong>al papers by Wallace [240] <strong>and</strong> Brugger [23]. The<br />

classic text on anelasticity is by Nowick <strong>and</strong> Berry [175].<br />

1.2 Dislocation Plasticity<br />

Plastic flow is time-dependent <strong>and</strong> irreversible. Here, I focus on the subject <strong>of</strong> dis-<br />

location plasticity <strong>and</strong> on one particularly important equation relevant to the strength <strong>of</strong><br />

materials.<br />

Arzt [6] emphasizes the strong coupl<strong>in</strong>g between characteristic lengths <strong>and</strong> size pa-<br />

rameters <strong>in</strong> determ<strong>in</strong><strong>in</strong>g the material strength. The characteristic length is def<strong>in</strong>ed as the<br />

dimension associated with the relevant physical process (such as the Burgers vector or the<br />

equilibrium diameter <strong>of</strong> a dislocation loop). The size parameter is related to the microstruc-<br />

ture (such as the gra<strong>in</strong> size or dislocation spac<strong>in</strong>g <strong>in</strong> a forest dislocation network). A simple<br />

example follow<strong>in</strong>g Orowan is given to show how a relation between the Burgers vector <strong>and</strong><br />

obstacle spac<strong>in</strong>g can determ<strong>in</strong>e material strength.<br />

tion:<br />

An applied shear stress, σ , imparts a Peach-Koehler force per unit length on a disloca-<br />

F = σ b (1.8)<br />

where b is the Burgers vector. The dislocation l<strong>in</strong>e tension, T , resists bow<strong>in</strong>g as shown <strong>in</strong>


Chapter 1: Background 7<br />

Figure 1.2 (a). The l<strong>in</strong>e tension is given by<br />

T ≈ 1<br />

2 Gb2<br />

where G is the shear modulus. Balanc<strong>in</strong>g the vertical forces shown <strong>in</strong> (a) gives<br />

(1.9)<br />

σ b (2�R) = 2T � (1.10)<br />

where the length <strong>of</strong> the l<strong>in</strong>e segment is 2�R, <strong>and</strong> we assume � is small. Thus, if a mov<strong>in</strong>g<br />

dislocation encounters the two obstacles shown <strong>in</strong> Figure 1.2 (b), the shear stress needed to<br />

bow the dislocation to the maximum curvature <strong>of</strong> κ is determ<strong>in</strong>ed by the obstacle spac<strong>in</strong>g,<br />

ℓ = 2/κ:<br />

σOr ≈ Gb<br />

. (1.11)<br />

ℓ<br />

Bow<strong>in</strong>g the dislocation beyond this critical radius requires no additional stress; thus, σOr is<br />

the maximum resistance to the applied shear stress for the case shown <strong>in</strong> Figure 1.2 (b).<br />

Assum<strong>in</strong>g that these obstacles, whether they are other dislocations or precipitates, rep-<br />

resent the overall maximum resistance to dislocation motion <strong>in</strong> a material, the Orowan<br />

stress, σOr, represents the strength <strong>of</strong> a material. This ignores the problem <strong>of</strong> thermal acti-<br />

vation as described by Kocks et al. [130].<br />

Stra<strong>in</strong> gradient plasticity merges dislocation mechanics with cont<strong>in</strong>uum mechanics.<br />

The importance <strong>of</strong> plastic stra<strong>in</strong> gradients had been recognized by Nye, who formulated<br />

the idea <strong>of</strong> geometrically necessary dislocations <strong>in</strong> association with lattice curvature [178],<br />

prompt<strong>in</strong>g mechanicians to develop a cont<strong>in</strong>uum description <strong>of</strong> lattice curvature with a non-<br />

Riemannian space formulation [15, 132, 133]. Every dislocation present <strong>in</strong> a material can<br />

be categorized as either a geometrically necessary dislocation (GND) required for defor-<br />

mation compatibility or a statistically stored dislocation (SSD) [43] that is trapped by other


Chapter 1: Background 8<br />

T<br />

T s<strong>in</strong>Θ<br />

σb (2ΘR)<br />

R<br />

Θ Θ<br />

T<br />

T s<strong>in</strong>Θ<br />

(a) (b)<br />

Figure 1.2: (a) A depiction <strong>of</strong> the vertical force balance between dislocation l<strong>in</strong>e tension<br />

<strong>and</strong> the Peach-Koehler force. (b) An illustration <strong>of</strong> a dislocation bow<strong>in</strong>g between two<br />

impenetrable obstacles with separation ℓ.<br />

dislocations <strong>in</strong> a r<strong>and</strong>om way. Ashby [7, 8] formulated a more general theory <strong>of</strong> GNDs <strong>and</strong><br />

emphasized their importance <strong>in</strong> determ<strong>in</strong><strong>in</strong>g a material’s plastic properties.<br />

Accord<strong>in</strong>g to Ashby, the density <strong>of</strong> the GNDs, ρGND, depends on the geometrical ar-<br />

rangement <strong>of</strong> gra<strong>in</strong>s <strong>and</strong> phases <strong>and</strong> is therefore microstructure-dependent. On the other<br />

h<strong>and</strong>, the density <strong>of</strong> SSDs, ρSSD, depends on the material properties <strong>and</strong> the total plastic<br />

stra<strong>in</strong>. The total dislocation density is approximately (ρGND + ρSSD), although ρGND can<br />

<strong>in</strong>crease ρSSD rapidly at large stra<strong>in</strong>s [8].<br />

<strong>and</strong><br />

The densities <strong>of</strong> GNDs due to edge <strong>and</strong> screw dislocations are given by [5]<br />

ρ edge<br />

GND = −∇ɛsh · s<br />

b<br />

ρ screw<br />

GND = ∇ɛsh · m<br />

b<br />

l<br />

(1.12)<br />

(1.13)


Chapter 1: Background 9<br />

where ɛsh is the plastic shear stra<strong>in</strong>, s is a unit vector along the slip direction, <strong>and</strong> m is given<br />

by m = s × n (where n is the slip plane unit normal). Ashby [7, 8] makes use <strong>of</strong> a simpler<br />

relation,<br />

ρGND = 4ɛsh<br />

ℓGNDb<br />

where ℓGND is called the geometrical slip distance characteristic <strong>of</strong> the microstructure.<br />

(1.14)<br />

If we assume a Taylor model <strong>of</strong> stra<strong>in</strong> harden<strong>in</strong>g [216, 217] (see Chapter 3 for details),<br />

the flow stress is given by<br />

σflow = kTGb √ ρSSD + ρGND<br />

(1.15)<br />

where kT is a constant on the order <strong>of</strong> 0.1 to 1. By comb<strong>in</strong><strong>in</strong>g (1.14) <strong>and</strong> (1.15) <strong>and</strong><br />

assum<strong>in</strong>g ρGND ≫ ρSSD, we see that for cases <strong>in</strong> which the local shear stra<strong>in</strong> is proportional<br />

to b/ℓGND, the Orowan stress equation is recovered.<br />

By <strong>in</strong>sert<strong>in</strong>g a gradient <strong>of</strong> plastic shear <strong>in</strong>to the strength problem, Ashby opened up the<br />

door to mechanicians who had the tools to develop cont<strong>in</strong>uum theories <strong>of</strong> plasticity that<br />

can <strong>in</strong>corporate a characteristic length scale. Probably the most well developed <strong>of</strong> these is<br />

by Fleck <strong>and</strong> Hutch<strong>in</strong>son [61]. They extended the J2 flow theory to <strong>in</strong>clude both stretch<br />

gradients <strong>and</strong> rotation gradients, each <strong>of</strong> which can be represented by a length parameter.<br />

For more <strong>in</strong>formation, Cottrell’s books, The Mechanical Properties <strong>of</strong> Matter [43] <strong>and</strong><br />

Dislocations <strong>and</strong> Plastic Flow <strong>in</strong> Crystals [42] serve as excellent <strong>in</strong>troductions to plasticity<br />

from a materials science viewpo<strong>in</strong>t. Argon’s recent reviews <strong>in</strong> [3, 4] summarize the current<br />

view <strong>of</strong> the field. Hill’s Mathematical Theory <strong>of</strong> Plasticity is the usual start<strong>in</strong>g po<strong>in</strong>t for a<br />

mechanician [99].


Chapter 1: Background 10<br />

1.3 Th<strong>in</strong> Film Growth <strong>and</strong> Microstructure<br />

When a deposit<strong>in</strong>g atom arrives at a substrate <strong>of</strong> a different phase, its <strong>in</strong>corporation<br />

depends on its <strong>in</strong>com<strong>in</strong>g energy, the overall deposition rate, the temperature <strong>of</strong> the substrate,<br />

the defect structure <strong>of</strong> the surface, etc. [25, 230]. If the mismatch with the layer below it<br />

is small, nucleation <strong>and</strong> growth may proceed pseudomorphically, <strong>and</strong> the <strong>in</strong>terface may<br />

be described as coherent if the lattice registry is complete [104]. The stra<strong>in</strong> energy <strong>of</strong><br />

the film grows l<strong>in</strong>early with the film thickness; consequently, it may become energetically<br />

favorable for a periodic array <strong>of</strong> misfit dislocations to form at the <strong>in</strong>terface which ma<strong>in</strong>ta<strong>in</strong>s<br />

the crystallographic registry between them. Such an <strong>in</strong>terface is termed semicoherent. If<br />

the mismatch is very large, the registry <strong>of</strong> the atomic planes is lost, <strong>and</strong> the <strong>in</strong>terface is<br />

<strong>in</strong>coherent.<br />

Non-epitaxial films can develop texture <strong>and</strong> undergo gra<strong>in</strong> growth dur<strong>in</strong>g deposition<br />

[127, 220]. The substrate temperature is <strong>of</strong>ten the most important parameter <strong>in</strong> determ<strong>in</strong><strong>in</strong>g<br />

the film microstructure. This has led to the development <strong>of</strong> structural zone models that<br />

predict gra<strong>in</strong> sizes <strong>and</strong> shapes as a function <strong>of</strong> substrate temperature [156, 145, 144].<br />

1.4 Two Common Stress Determ<strong>in</strong>ation Techniques<br />

Equations (1.6) <strong>and</strong> (1.7) show that stresses may be determ<strong>in</strong>ed by both stra<strong>in</strong> <strong>and</strong> force<br />

measurements. Ideally, both measurements are taken simultaneously dur<strong>in</strong>g a deformation;<br />

this allows an evaluation <strong>of</strong> the elastic constants [107]. Unfortunately such techniques<br />

are not <strong>of</strong>ten performed on th<strong>in</strong> film samples. The most common force measurement for<br />

th<strong>in</strong> films on substrates is the wafer curvature measurement. The equation used to extract


Chapter 1: Background 11<br />

Table 1.1: Ability <strong>of</strong> wafer curvature <strong>and</strong> x-ray techniques to measure different types <strong>of</strong><br />

stresses. Adapted from Ruud [195].<br />

Stress Type Wafer <strong>Cu</strong>rvature X-Ray<br />

Growth Yes Yes<br />

Thermal Yes Yes<br />

Coherency No Yes<br />

Interface Phase Formation Yes No<br />

Interface Stress Yes No<br />

the stress from the substrate curvature is derived from force <strong>and</strong> moment balance. X-ray<br />

diffraction is the most common way <strong>of</strong> mak<strong>in</strong>g stra<strong>in</strong> measurements.<br />

Both <strong>of</strong> these techniques benefit from their ease <strong>in</strong> experimental setup <strong>and</strong> analysis.<br />

They may not give the same results, however. Follow<strong>in</strong>g Table 1.1, the wafer curvature<br />

method can be used to measure the follow<strong>in</strong>g stresses: growth stress, thermal stress, <strong>in</strong>ter-<br />

face stress, <strong>and</strong> stresses associated with phase formation at the <strong>in</strong>terface. X-ray diffraction<br />

can be used to measure growth stress, thermal stress, <strong>and</strong> coherency stress. Thus, we can<br />

isolate certa<strong>in</strong> stresses by us<strong>in</strong>g both techniques on the same sample.<br />

1.4.1 Wafer <strong>Cu</strong>rvature<br />

Figure 1.3 (a) shows a stra<strong>in</strong>-free free-st<strong>and</strong><strong>in</strong>g film <strong>and</strong> substrate. Forc<strong>in</strong>g the film to<br />

fit onto the substrate requires an applied force as shown <strong>in</strong> (b), <strong>and</strong> the film is subsequently<br />

“glued” onto the substrate (c). Force <strong>and</strong> moment balance require that the substrate bends<br />

to accommodate the film stress (d).<br />

The relationship between the substrate curvature change, �κ, <strong>and</strong> the film stress, σf, is<br />

given by Stoney’s equation:<br />

σf = Yst 2 s �κ<br />

6tf<br />

(1.16)


Chapter 1: Background 12<br />

(a) (b)<br />

(c) (d)<br />

Figure 1.3: A schematic show<strong>in</strong>g how the film stress causes substrate curvature. (a) The<br />

film <strong>and</strong> substrate are both shown <strong>in</strong> their stra<strong>in</strong>-free state. (b) The film is stretched to fit<br />

onto the substrate. (c) The film is “glued” onto the substrate. (d) The film/substrate system<br />

bends <strong>in</strong> order to achieve force <strong>and</strong> moment balance.


Chapter 1: Background 13<br />

Figure 1.4: A schematic <strong>of</strong> the curvature measurement apparatus at <strong>Harvard</strong> University.<br />

Adapted from Mull<strong>in</strong> [158].<br />

where Ys is the biaxial modulus <strong>of</strong> the substrate, ts is the substrate thickness, <strong>and</strong> tf is the<br />

film thickness. This equation assumes tf ≪ ts; <strong>in</strong> this limit, determ<strong>in</strong>ation <strong>of</strong> the film stress<br />

does not rely on knowledge <strong>of</strong> the biaxial modulus <strong>of</strong> the film. Appendix E provides a<br />

derivation <strong>of</strong> Stoney’s equation.<br />

The wafer curvature measurement / anneal<strong>in</strong>g apparatus at <strong>Harvard</strong> University is called<br />

ROC <strong>and</strong> is shown <strong>in</strong> Figure 1.4. The technical details <strong>of</strong> how it measures the curvature can<br />

be found <strong>in</strong> Witvrouw’s or Mull<strong>in</strong>’s thesis [256, 158], but the pr<strong>in</strong>ciple is well known [62].<br />

The 1 <strong>in</strong>ch by 1/4 <strong>in</strong>ch sample rests <strong>in</strong> a furnace on two sapphire prisms <strong>and</strong> is abutted on<br />

two adjacent sides by steel p<strong>in</strong>s to allow repeatable sample placement. A HeNe laser scans<br />

across the long axis <strong>of</strong> the sample <strong>in</strong> 1 mm steps, <strong>and</strong> the reflected beam is collected <strong>in</strong> a<br />

null detector. The position <strong>of</strong> the reflected spot is recorded as a function <strong>of</strong> position along<br />

the sample, thereby giv<strong>in</strong>g the sample curvature.


Chapter 1: Background 14<br />

Prior to thermal cycl<strong>in</strong>g, the chamber is evacuated to a pressure <strong>of</strong> 3 × 10 −4 Pa by a<br />

diffusion pump <strong>and</strong> then backfilled to a slight overpressure <strong>of</strong> 3 psi with a flow<strong>in</strong>g gas <strong>of</strong><br />

95 wt.% He <strong>and</strong> 5 wt.% H2.<br />

1.4.2 X-Ray Diffraction Measurements<br />

As mentioned, stra<strong>in</strong> measurements are <strong>of</strong>ten performed by x-ray diffraction, with the<br />

stra<strong>in</strong>-free lattice parameter, a0, <strong>and</strong> the elastic constants known a priori. Although this<br />

measurement is straightforward <strong>in</strong> pr<strong>in</strong>ciple, it still has several problems. First, not every-<br />

one agrees on the value <strong>of</strong> a0. Its value is <strong>of</strong>ten determ<strong>in</strong>ed by perform<strong>in</strong>g x-ray diffraction<br />

experiments on a stra<strong>in</strong>-free sample, e.g., an annealed powder. Extremely precise tech-<br />

niques have been developed to m<strong>in</strong>imize the errors <strong>of</strong> these measurements [60]. In practice,<br />

however, a completely stra<strong>in</strong>-free sample is difficult to produce. Furthermore, the mean<strong>in</strong>g<br />

<strong>of</strong> “stra<strong>in</strong>-free” can be different depend<strong>in</strong>g on what the researcher is try<strong>in</strong>g to measure. It<br />

may refer to the constra<strong>in</strong>ed zero po<strong>in</strong>t between an applied tension <strong>and</strong> compression, or, it<br />

may refer to equilibrium lattice parameter. These may differ depend<strong>in</strong>g on the thermody-<br />

namic state <strong>of</strong> the material.<br />

The determ<strong>in</strong>ation <strong>of</strong> elastic constants <strong>in</strong> polycrystall<strong>in</strong>e materials is also problematic.<br />

There exists a substantial body <strong>of</strong> literature devoted to the proper ways <strong>of</strong> averag<strong>in</strong>g the<br />

s<strong>in</strong>gle crystal elastic constants for textured <strong>and</strong> untextured samples [24]. In pr<strong>in</strong>ciple, one<br />

can determ<strong>in</strong>e a sample’s orientation distribution function (ODF) <strong>and</strong> use a gra<strong>in</strong> <strong>in</strong>teraction<br />

model (e.g., Voigt or Reuss) to determ<strong>in</strong>e the stresses from the measured stra<strong>in</strong>s. ODF<br />

determ<strong>in</strong>ation is a tedious process, however, <strong>and</strong> researchers typically either assume the<br />

sample has fiber texture or no texture at all. The gra<strong>in</strong> <strong>in</strong>teraction model is also <strong>of</strong>ten


Chapter 1: Background 15<br />

assumed.<br />

Appendix A provides some details on how x-ray stra<strong>in</strong> measurements work. For more<br />

<strong>in</strong>formation, the classic text for stress determ<strong>in</strong>ation by x-rays is the book by Noyan <strong>and</strong><br />

Cohen [176]; there are several errors <strong>in</strong> this book that are corrected <strong>in</strong> Appendix B. Noyan<br />

et al. [177] have also written a review on x-ray measurements <strong>of</strong> th<strong>in</strong> film stresses, <strong>and</strong><br />

Hauk et al. [95] have recently completed a comprehensive review <strong>of</strong> techniques used to<br />

measure residual stresses.<br />

A schematic <strong>of</strong> the diffractometer used at <strong>Harvard</strong> is shown <strong>in</strong> Figure 1.5. It consists<br />

<strong>of</strong> a four-circle Huber diffractometer (a Huber 424 two-circle with a 511.1 Eulerian cra-<br />

dle) equipped with a Model 20/20 l<strong>in</strong>ear position sensitive detector (PSD) from Reflection<br />

Imag<strong>in</strong>g (www.reflectionimag<strong>in</strong>g.com). The system was put together with the<br />

help <strong>of</strong> Molecular Metrology (www.molmet.com). S<strong>in</strong>ce the p<strong>in</strong>hole collimator has a<br />

diameter <strong>of</strong> 0.3 mm, the samples are rocked ±2 ◦ <strong>in</strong> the diffractometer plane to achieve<br />

adequate sampl<strong>in</strong>g statistics [111] (Appendix B).<br />

Beaml<strong>in</strong>e X22C at the National Synchrotron Light Source at Brookhaven National Lab-<br />

oratory was used for reflectivity [194], θ − 2θ, <strong>and</strong> graz<strong>in</strong>g <strong>in</strong>cidence diffraction (GID)<br />

[148] measurements. The diffractometer is a six-circle Frank <strong>and</strong> Heydrich, <strong>and</strong> the <strong>in</strong>ci-<br />

dent beam is conditioned with a Pt focus<strong>in</strong>g mirror <strong>and</strong> a two-crystal Si(111) monochro-<br />

mator.<br />

A schematic <strong>of</strong> the different scatter<strong>in</strong>g geometries is presented <strong>in</strong> Figure 1.6. Figure<br />

1.6 (a) shows the symmetric θ − 2θ geometry used for x-ray reflectivity or powder scans.<br />

As <strong>in</strong>dicated by the arrow on the left, χ-scans are performed by rotat<strong>in</strong>g the sample normal<br />

<strong>in</strong>to or out <strong>of</strong> the plane <strong>of</strong> the page while ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g the diffraction condition. Figure


Chapter 1: Background 16<br />

Figure 1.5: A schematic <strong>of</strong> the four-circle Huber diffractometer at <strong>Harvard</strong> University.<br />

1.6 (b) shows the longitud<strong>in</strong>al scan geometry used <strong>in</strong> diffuse scatter<strong>in</strong>g (Chapter 4). The<br />

sample is tilted slightly away from the specular condition by an amount �θ. The geometry<br />

for graz<strong>in</strong>g <strong>in</strong>cidence diffraction is depicted <strong>in</strong> Figure 1.6 (c); here, the scatter<strong>in</strong>g vector is<br />

tilted with respect to the sample surface by a small angle α. The magnitude <strong>of</strong> the scatter<strong>in</strong>g<br />

vector, q,isgivenby<br />

q =<br />

4π s<strong>in</strong> θ<br />

λ<br />

where λ is the x-ray wavelength <strong>and</strong> 2θ is the diffraction angle.<br />

1.5 Sputter Deposition<br />

(1.17)<br />

The majority <strong>of</strong> the films used <strong>in</strong> this thesis were grown us<strong>in</strong>g the sputter<strong>in</strong>g chamber<br />

built by Spaepen et al. [213] <strong>and</strong> shown <strong>in</strong> Figure 1.7. Prior to deposition, 1 <strong>in</strong>ch by 1/4 <strong>in</strong>ch


Chapter 1: Background 17<br />

χ<br />

θ−∆θ<br />

2θ<br />

θ<br />

α<br />

q<br />

∆θ<br />

n<br />

Figure 1.6: A schematic <strong>of</strong> different scatter<strong>in</strong>g geometries. (a) The symmetric θ − 2θ<br />

geometry <strong>and</strong> χ rotation axis, (b) the geometry used for longitud<strong>in</strong>al diffuse scatter<strong>in</strong>g,<br />

<strong>and</strong> (c) the graz<strong>in</strong>g <strong>in</strong>cidence diffraction geometry.<br />

θ<br />

θ+∆θ<br />

α<br />

(a)<br />

(b)<br />

(c)


Chapter 1: Background 18<br />

Figure 1.7: A schematic <strong>of</strong> the sputter<strong>in</strong>g chamber at <strong>Harvard</strong> University.<br />

substrates were cleaved from the wafer <strong>and</strong> subsequently cleaned <strong>in</strong> acetone, methanol, <strong>and</strong><br />

isopropanol.<br />

The base pressure <strong>of</strong> the deposition chamber is 1.3 × 10 −5 Pa. The sputter<strong>in</strong>g targets<br />

are bolted onto a rotat<strong>in</strong>g target block to facilitate multilayer growth, <strong>and</strong> both the targets<br />

<strong>and</strong> samples are water-cooled. One ion gun is used for target sputter<strong>in</strong>g, <strong>and</strong> another gun<br />

can be used for either substrate clean<strong>in</strong>g or reactive sputter<strong>in</strong>g.<br />

The parameters for sputter deposition <strong>and</strong> substrate clean<strong>in</strong>g are listed <strong>in</strong> Table 1.2.<br />

Table 1.2: Parameters for ion beam sputter deposition <strong>and</strong> substrate clean<strong>in</strong>g.<br />

Deposition Substrate Clean<strong>in</strong>g<br />

Beam 1400 V 45 mA 500 V 35 mA<br />

Accelerator 150 V 0 mA 500 V 0 mA<br />

Discharge 25 V 2 mA 30 V 2 mA<br />

Cathode 8 V - 8 V -<br />

Ar Pressure 1.1 × 10 −2 Pa 1.1 × 10 −2 Pa


Chapter 2<br />

The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong><br />

Us<strong>in</strong>g a tensile tester especially designed for free-st<strong>and</strong><strong>in</strong>g films, Huang [106] found<br />

that the Young moduli <strong>of</strong> electron beam evaporated films were about 20% lower than those<br />

calculated from the elastic constants found <strong>in</strong> the literature (such as those from L<strong>and</strong>olt-<br />

Börnste<strong>in</strong> [12], Simmons <strong>and</strong> Wang [207], or Hunt<strong>in</strong>gton [108]). Table 2.1 shows that this<br />

“deficit” is well documented <strong>and</strong> only observed <strong>in</strong> polycrystall<strong>in</strong>e films.<br />

A variety <strong>of</strong> reasons for this modulus deficit have been hypothesized. They <strong>in</strong>clude<br />

elastic anisotropy (from preferred orientation), porosity, gra<strong>in</strong> boundary compliance, gra<strong>in</strong><br />

boundary crack<strong>in</strong>g, reversible microplasticity, dislocation anelasticity, <strong>and</strong> gra<strong>in</strong> bound-<br />

ary slid<strong>in</strong>g [107]. We studied the orig<strong>in</strong> <strong>of</strong> the modulus deficit by measur<strong>in</strong>g the biaxial<br />

modulus <strong>of</strong> sputtered copper films on substrates us<strong>in</strong>g the wafer curvature method. Before<br />

discuss<strong>in</strong>g how the measurement is made, it is important to first underst<strong>and</strong> the stress-<br />

temperature pr<strong>of</strong>ile.<br />

19


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 20<br />

Table 2.1: Measured <strong>and</strong> calculated values <strong>of</strong> the Young modulus, E, <strong>and</strong> the biaxial modulus,<br />

Y , for both free-st<strong>and</strong><strong>in</strong>g <strong>and</strong> substrate-bonded films on the order <strong>of</strong> 1 µm thick.<br />

Material Emeas E 〈111〉<br />

calc ∗ Ymeas Y 〈111〉<br />

calc ∗ (GPa) (GPa) (GPa) (GPa)<br />

Preparation<br />

Method<br />

Ref.<br />

<strong>Ag</strong>/<strong>Cu</strong> 87.5 107 - - e-beam [106]<br />

<strong>Cu</strong> 102 130 - - e-beam [106]<br />

<strong>Cu</strong> 104 130 - - e-beam [187]<br />

<strong>Cu</strong> 110 130 - - sputtered [187]<br />

<strong>Cu</strong> 66 130 - - electrodep. [187]<br />

<strong>Cu</strong> 104 130 161 261 sputtered [262]<br />

<strong>Cu</strong> 116 130 - - electrodep. [192]<br />

<strong>Ag</strong> 63 84 - - e-beam [106]<br />

<strong>Ag</strong> - - 50 174 thermal evap. [46]<br />

Al 57 72 - - e-beam [106]<br />

Al 49 72 - - thermal evap. [97]<br />

Al 64 72 - - sputtered [192]<br />

Al - - 62 115 thermal evap. [46]<br />

W 340 389 - - CVD [192]<br />

Si (110) † 166 166 - - etched from bulk [215]<br />

∗ Calculated from the dynamic elastic constants found <strong>in</strong> Bechmann et al. [12]<br />

† S<strong>in</strong>gle crystal


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 21<br />

Figure 2.1: Stress as a function <strong>of</strong> temperature for a 1.91 µm sputtered <strong>Cu</strong> film on Si(100).<br />

The dotted l<strong>in</strong>es show the hypothetical elastic stress <strong>in</strong> the film if no relaxation could take<br />

place (thermoelastic l<strong>in</strong>es), <strong>and</strong> the labels (A), (B), <strong>and</strong> (C) denote elastic thermal cycles<br />

as described <strong>in</strong> the text. The upper thermoelastic l<strong>in</strong>e was drawn us<strong>in</strong>g the elastic constant<br />

data found <strong>in</strong> Bechmann et al. [12] <strong>and</strong> assum<strong>in</strong>g mild 〈111〉 texture.<br />

2.1 The Variation <strong>of</strong> Stress with Temperature <strong>in</strong> a Th<strong>in</strong><br />

Film on an Elastic Substrate<br />

As shown <strong>in</strong> Figure 2.1, the <strong>in</strong>itial stress <strong>in</strong> the film is compressive due to atomic peen-<br />

<strong>in</strong>g [254]. (Tensile stresses are always plotted as positive values <strong>in</strong> this thesis). Due to the<br />

thermal expansion mismatch with the substrate, the copper film is compressed upon heat-<br />

<strong>in</strong>g <strong>and</strong> stretched upon cool<strong>in</strong>g (cf. Figure 1.3). For small temperature changes �T , the<br />

deformation is elastic, <strong>and</strong> cycl<strong>in</strong>g back <strong>and</strong> forth with<strong>in</strong> a small temperature range allows<br />

a measure <strong>of</strong> the biaxial modulus, Yf, if the thermal expansion coefficients <strong>of</strong> the substrate


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 22<br />

<strong>and</strong> film (αs <strong>and</strong> αf) are known s<strong>in</strong>ce<br />

�σf = Yf�ɛf (2.1)<br />

= Yf<br />

� T1<br />

T0<br />

dT (αs − αf)<br />

≈ Yf(αs − αf)�T.<br />

These small temperature cycles are called elastic thermal cycles from now on. Figure 2.1<br />

shows two elastic thermal cycles from 10 ◦ Cto70 ◦ C for the as-deposited <strong>and</strong> the annealed<br />

film (labeled (A) <strong>and</strong> (B), respectively).<br />

Compress<strong>in</strong>g or stretch<strong>in</strong>g beyond the elastic limit <strong>in</strong>duces plastic deformation by a<br />

mechanism that depends on the temperature. If the thermal cycle is started higher than<br />

some critical temperature (typically about 300 ◦ C for most <strong>of</strong> the sputtered films), nearly all<br />

the stra<strong>in</strong> is plastic. An example <strong>of</strong> this is shown <strong>in</strong> Figure 2.1, where an “elastic” thermal<br />

cycle (C) from 500 ◦ C to 550 ◦ C demonstrates hysteresis <strong>and</strong> is therefore not elastic. Below<br />

the critical temperature, it is possible to obta<strong>in</strong> elastic deformation <strong>and</strong> hence to determ<strong>in</strong>e<br />

the biaxial modulus as a function <strong>of</strong> temperature. This so-called critical temperature is<br />

dependent on the heat<strong>in</strong>g/cool<strong>in</strong>g rate, stress state, <strong>and</strong> the mechanical properties <strong>of</strong> the<br />

film.<br />

Several researchers have attempted to expla<strong>in</strong> the shape <strong>of</strong> the σf vs. T pr<strong>of</strong>ile [54, 124,<br />

125, 172]. The <strong>in</strong>itial ramp-up <strong>in</strong> temperature from the as-deposited state causes a large<br />

tensile <strong>in</strong>crease at approximately 200 ◦ C. Chaudhari [35] has shown that this is due to gra<strong>in</strong><br />

growth. (A ref<strong>in</strong>ement <strong>of</strong> his model has recently been published [58]). A further <strong>in</strong>crease<br />

<strong>in</strong> temperature results <strong>in</strong> relaxation <strong>of</strong> the stress by dislocation flow <strong>and</strong> atomic motion.<br />

Hold<strong>in</strong>g at any <strong>of</strong> these temperatures produces stress relaxation that can be described by<br />

various creep models [3, 4, 6, 69].


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 23<br />

<strong>Ag</strong>a<strong>in</strong> follow<strong>in</strong>g Figure 2.1, atomic motion fully relaxes the film at 550 ◦ C,<strong>and</strong>anew<br />

stra<strong>in</strong>-free lattice parameter is established. Cool<strong>in</strong>g from 550 ◦ C puts the film <strong>in</strong> tension.<br />

As the film cools, it becomes <strong>in</strong>creas<strong>in</strong>gly difficult to relax the stress: thus, the residual<br />

stress builds up. The result<strong>in</strong>g stress at room temperature is dependent on the temperature<br />

at which the stra<strong>in</strong>-free lattice parameter is established <strong>and</strong> how difficult it is for the film to<br />

undergo plastic relaxation. The dotted l<strong>in</strong>es <strong>in</strong> Figure 2.1 illustrate the stress-temperature<br />

behavior the film would take if no plastic flow were possible. At any given temperature, the<br />

difference between the dotted l<strong>in</strong>e <strong>and</strong> the measured residual stress <strong>in</strong>dicates the amount <strong>of</strong><br />

plastic stra<strong>in</strong> <strong>in</strong> the film. After the first thermal cycle takes place, the shape <strong>of</strong> the stress<br />

vs. temperature pr<strong>of</strong>ile rema<strong>in</strong>s constant. Indeed, it is now commonly accepted that the<br />

microstructure is stable after heat<strong>in</strong>g through the gra<strong>in</strong> growth phase <strong>of</strong> the first thermal<br />

cycle [10].<br />

It should be noted that this stress-temperature pr<strong>of</strong>ile changes dramatically when the<br />

films are prepared <strong>and</strong> annealed <strong>in</strong> ultra-high vacuum conditions rather than <strong>in</strong> high vacuum<br />

[246].<br />

2.2 The Biaxial Modulus Measurement<br />

As equation (2.1) shows, the biaxial modulus <strong>of</strong> the film can be determ<strong>in</strong>ed if the coef-<br />

ficients <strong>of</strong> thermal expansion (CTE) <strong>of</strong> the film <strong>and</strong> substrate are known.<br />

Alternatively, if elastic thermal cycles are performed on different substrates with known<br />

thermomechanical constants, it is possible to determ<strong>in</strong>e both the biaxial modulus <strong>and</strong> the<br />

CTE <strong>of</strong> the film. For <strong>in</strong>stance, if we deposit the same film on two different substrates 1 <strong>and</strong>


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 24<br />

2,<br />

<strong>and</strong><br />

Yf =<br />

1<br />

(α1 − α2)<br />

� �σ1<br />

�T<br />

�<br />

�σ2<br />

−<br />

�T<br />

(2.2)<br />

αf = α2�σ1 − α1�σ2<br />

. (2.3)<br />

�σ1 − �σ2<br />

If more than two substrates are used, we can perform a l<strong>in</strong>ear fit <strong>of</strong>�σi/�T vs. αi where<br />

�σi is the change <strong>in</strong> film stress over the elastic thermal cycle for a substrate i. The slope<br />

<strong>of</strong> this l<strong>in</strong>e is equal to −Yf, <strong>and</strong> the <strong>in</strong>tercept is αfYf.<br />

Note that pair<strong>in</strong>g a substrate with a s<strong>in</strong>gle CTE with a substrate with anisotropic CTEs,<br />

such as Y -cut quartz, adds yet another equation <strong>and</strong> can lead to an <strong>in</strong>dependent determi-<br />

nation <strong>of</strong> Young’s modulus, Poisson’s ratio, <strong>and</strong> the CTE <strong>of</strong> the film (assum<strong>in</strong>g the biaxial<br />

modulus <strong>of</strong> the film is isotropic <strong>in</strong> the plane) [262].<br />

2.2.1 The Effect <strong>of</strong> Texture on the Biaxial Modulus<br />

Copper exhibits strong elastic anisotropy; this can manifest itself if the polycrystall<strong>in</strong>e<br />

film has preferred orientation, i.e., texture. The calculated value <strong>of</strong> the biaxial modulus<br />

is 261 GPa for 〈111〉 textured copper, 196 GPa for a r<strong>and</strong>omly oriented polycrystal, <strong>and</strong><br />

115 GPa for 〈100〉 texture (cf. Figure D.2 <strong>in</strong> Appendix D). Without effects from defects<br />

<strong>in</strong> the microstructure, the biaxial modulus <strong>of</strong> polycrystall<strong>in</strong>e films is approximately the<br />

volume average over all the gra<strong>in</strong> orientations:<br />

〈X〉 = 1<br />

8π 2<br />

� ϕ1=2π � �=π � ϕ2=2π<br />

X (ϕ1,�,ϕ2) f (ϕ1,�,ϕ2) s<strong>in</strong> � dϕ1 d� dϕ2, (2.4)<br />

ϕ1=0 �=0 ϕ2=0<br />

where X is either the biaxial modulus or the <strong>in</strong>verse biaxial modulus, f (ϕ1,�,ϕ2) is the<br />

orientation distribution function (ODF) <strong>and</strong> ϕ1, �, <strong>and</strong> ϕ2 are the Euler angles [24, 131].


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 25<br />

See Appendix D for more on averag<strong>in</strong>g techniques. This calculation does not take <strong>in</strong>to<br />

account stress cont<strong>in</strong>uity at the gra<strong>in</strong> boundaries [10].<br />

2.3 Experiment<br />

2.3.1 Film Preparation<br />

A number <strong>of</strong> substrates were used: Si(100) with a 1 µm thermal oxide, bare Si(100),<br />

Ge(111), α-Al2O3(0001), <strong>and</strong> fused silica; Si(100) with the thermal oxide was used for<br />

most experiments. Before deposition, the basel<strong>in</strong>e curvatures <strong>of</strong> the substrates were mea-<br />

sured.<br />

Copper films (from a 99.99 at.% <strong>Cu</strong> target) <strong>of</strong> thicknesses rang<strong>in</strong>g from 0.2 µm to<br />

2.3 µm were deposited by ion beam sputter<strong>in</strong>g with an Ar pressure <strong>of</strong> 1.1 × 10 −2 Pa after<br />

reach<strong>in</strong>g a base pressure <strong>of</strong> 1.3 × 10 −5 Pa. The sputter<strong>in</strong>g chamber described by Spaepen<br />

et al. [213] (Figure 1.7) was used with the sputter<strong>in</strong>g parameters given <strong>in</strong> Table 1.2.<br />

For some substrates, a th<strong>in</strong> amorphous silicon nitride (a-SiNx) layer (nom<strong>in</strong>ally 40 nm<br />

thick) was put down by dual ion gun sputter<strong>in</strong>g before deposit<strong>in</strong>g the copper to act as a<br />

diffusion barrier between the film <strong>and</strong> substrate. A nitride layer <strong>of</strong> the same thickness was<br />

also deposited onto the back <strong>of</strong> the substrate so as not to affect the curvature measure-<br />

ment. These deposition parameters are <strong>in</strong> Table 2.2. Submicron copper films were grown<br />

on th<strong>in</strong> (97 µm thick) Si(100) substrates. Prior to deposition, the substrates were sputter<br />

cleaned for 2-4 m<strong>in</strong>utes. The substrates were also water cooled from the backside, lead<strong>in</strong>g<br />

to temperature rise <strong>of</strong> 10 ◦ C after three hours <strong>of</strong> deposition. A typical deposition rate was<br />

0.15 nm/s.


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 26<br />

Table 2.2: Parameters for a-SiNx reactive sputter deposition.<br />

Argon <strong>Ni</strong>trogen<br />

Beam 1200 V 35 mA 600 V 11 mA<br />

Accelerator 200 V 0 mA 200 V 0 mA<br />

Discharge 30 V 1.2 mA 40 V 2 mA<br />

Cathode 8 V - 8 V -<br />

Pressure 8 × 10 −3 Pa 5.3 × 10 −3 Pa<br />

Additionally, some films were electron beam (e-beam) evaporated from 99.99 at.% <strong>Cu</strong><br />

shot (see [106]) at rates <strong>of</strong> 1 nm/s. In this chamber, the substrate temperature rises to about<br />

90 ◦ C after two hours <strong>of</strong> deposition. These films were grown primarily for comparative<br />

purposes. In what follows, nearly all <strong>of</strong> copper films discussed were sputter deposited.<br />

2.3.2 Film Characterization<br />

After deposition, the thickness <strong>of</strong> the films was measured us<strong>in</strong>g a Tencor Alpha-Step<br />

200 Pr<strong>of</strong>ilometer. For very th<strong>in</strong> films (less than one micrometer thick), the thickness was<br />

also measured with Rutherford Backscatter<strong>in</strong>g Spectrometry (RBS).<br />

Microstructural characterization was performed by both plan-view <strong>and</strong> cross-sectional<br />

transmission electron microscopy (TEM). Film purity was checked by RBS <strong>and</strong> energy<br />

dispersive spectroscopy (EDS) <strong>in</strong> the TEM.<br />

2.3.3 X-Ray Diffraction<br />

Symmetric θ − 2θ scans (also known as powder scans) <strong>and</strong> (111) fiber texture plots<br />

(the χ-scans described <strong>in</strong> Chapter 1) were performed <strong>in</strong> order to estimate the film texture<br />

[44, 126]. The powder scans sample only a small portion <strong>of</strong> reciprocal space but can show<br />

the relative strengths <strong>of</strong> the 〈111〉 <strong>and</strong> 〈100〉 textures. The (111) fiber texture plots are cuts


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 27<br />

Table 2.3: Thermomechanical properties <strong>of</strong> the substrates used for the thermal cycl<strong>in</strong>g<br />

experiments.<br />

Material Ys (GPa) ∗ CTE (10 −6 K −1 ) † Thickness (µm) Ref.<br />

Si(001) 180.5 2.708 395 & 97 [249]<br />

Ge(111) 183.4 5.905 461 [188]<br />

α-Al2O3(0001) 608.4 5.313 257 [86], [249]<br />

fused silica 86.5 0.52 162 [203]<br />

∗ at 25 ◦ C<br />

† at 40 ◦ C<br />

through the (111) pole figure <strong>and</strong> are sufficient to describe fiber textures, although a full<br />

ODF elim<strong>in</strong>ates the peak overlap sometimes present <strong>in</strong> fiber texture plots.<br />

The fiber texture plots were performed on the diffractometer shown <strong>in</strong> Figure 1.5. As<br />

discussed by Wcislak et al. [241], us<strong>in</strong>g a l<strong>in</strong>ear position sensitive detector to f<strong>in</strong>d the<br />

<strong>in</strong>tegrated <strong>in</strong>tensity <strong>of</strong> the peak as a function <strong>of</strong> χ elim<strong>in</strong>ates most <strong>of</strong> the corrections needed<br />

with a po<strong>in</strong>t detector [126, 135], but the th<strong>in</strong> film absorption correction (Appendix B) is still<br />

required. The volume fraction <strong>of</strong> 〈hkl〉-oriented gra<strong>in</strong>s is given by Baker et al. [10]:<br />

where I 111<br />

hkl<br />

fhkl =<br />

� χmax<br />

0<br />

� χmax<br />

0<br />

I 111<br />

hkl s<strong>in</strong> χ dχ<br />

I 111<br />

total<br />

s<strong>in</strong> χ dχ<br />

(2.5)<br />

is the scattered <strong>in</strong>tensity from the 〈hkl〉-oriented gra<strong>in</strong>s <strong>in</strong> the (111) texture plot.<br />

2.3.4 <strong>Cu</strong>rvature Measurements<br />

The wafer curvature apparatus used <strong>in</strong> these experiments was described <strong>in</strong> Chapter<br />

1, <strong>and</strong> the curvatures were converted to stresses us<strong>in</strong>g the Stoney equation (1.16). The<br />

thermomechanical properties <strong>of</strong> the various substrates used are listed <strong>in</strong> Table 2.3.<br />

The films were typically heated <strong>and</strong> cooled at approximately 5 ◦ C/m<strong>in</strong> <strong>in</strong> a form<strong>in</strong>g gas<br />

environment. For the elastic thermal cycles, the heat<strong>in</strong>g <strong>and</strong> cool<strong>in</strong>g rate was varied from


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 28<br />

1 ◦ C/m<strong>in</strong> to 6 ◦ C/m<strong>in</strong>. The temperature dependence <strong>of</strong> the modulus was studied dur<strong>in</strong>g the<br />

heat<strong>in</strong>g part <strong>of</strong> the thermal cycle (Figure 2.13).<br />

2.4 Results<br />

2.4.1 Microstructure<br />

Plan-view TEM micrographs are presented <strong>in</strong> Figure 2.2 (a) <strong>and</strong> (b) for the as-deposited<br />

<strong>and</strong> annealed 0.87 µm thick copper, respectively. The gra<strong>in</strong> size <strong>in</strong>creases from ∼ 40 nm<br />

to ∼ 500 nm, the latter be<strong>in</strong>g the average <strong>of</strong> a bimodal gra<strong>in</strong> size distribution. The cross-<br />

sectional micrographs <strong>of</strong> these films are presented <strong>in</strong> section 3.4.1. They show that even<br />

after anneal<strong>in</strong>g at 600 ◦ C for 0.5 hrs, the gra<strong>in</strong>s did not yet reach the stable microstructure<br />

consist<strong>in</strong>g <strong>of</strong> columnar gra<strong>in</strong>s with sizes on the order <strong>of</strong> the film thickness [219].<br />

A cross-sectional micrograph <strong>of</strong> <strong>Cu</strong> grown on Ge(111) is shown <strong>in</strong> Figure 2.3. The<br />

gra<strong>in</strong>s are columnar <strong>and</strong> extend from the substrate to the surface. The <strong>in</strong>-plane gra<strong>in</strong> size is<br />

∼ 50 nm.<br />

Both RBS <strong>and</strong> EDS show no impurities <strong>in</strong> the bulk <strong>of</strong> the copper. The argon level <strong>in</strong><br />

the annealed films was below the detection limit <strong>of</strong> EDS system.<br />

2.4.2 Film Texture<br />

Figure 2.4 (a) compares the powder scans for copper grown on Ge(111), α-Al2O3(0001),<br />

<strong>and</strong> Si(100). All three demonstrate 〈111〉 texture, although the copper grown on Ge(111)<br />

clearly has the strongest texture. This can be better seen <strong>in</strong> the χ-scan <strong>in</strong> Figure 2.4 (b)<br />

which shows the sharpness <strong>of</strong> the 〈111〉 texture pole.


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 29<br />

(a)<br />

(b)<br />

Figure 2.2: Plan-view TEM micrographs <strong>of</strong> 0.87 µm thick copper <strong>in</strong> the as-deposited state<br />

(a) <strong>and</strong> after anneal<strong>in</strong>g at 600 ◦ C for 0.5 hrs (b).


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 30<br />

Figure 2.3: A cross-sectional TEM micrograph <strong>of</strong> 1.90 µm thick copper grown on Ge(111).<br />

Anneal<strong>in</strong>g the films at 600 ◦ C <strong>in</strong>duces gra<strong>in</strong> growth, as seen <strong>in</strong> the micrographs, <strong>and</strong><br />

enhanced 〈111〉 texture, as seen <strong>in</strong> Figures 2.5 (a) (powder scan) <strong>and</strong> 2.5 (b) (χ-scan). The<br />

films annealed on α-Al2O3(0001) had much stronger 〈111〉 texture than those on Si(100)<br />

(Figure 2.6). The films grown on Ge(111) reacted with the Ge above 200 ◦ C, so the texture<br />

after anneal<strong>in</strong>g was not remeasured.<br />

Figure 2.7 shows that the 0.5 µm thick film had even stronger 〈111〉 texture <strong>in</strong> both<br />

the as-deposited <strong>and</strong> annealed states than the 1.25 µm thick film (cf. Figure 2.5 (b)). The<br />

as-deposited electron beam evaporated films had both 〈111〉 <strong>and</strong> 〈100〉 texture components<br />

(Figure 2.8).<br />

If we had the full ODF from several pole figures, we could calculate the biaxial mod-<br />

ulus <strong>of</strong> imperfectly textured films us<strong>in</strong>g equation (2.4). However, even with only χ-scans,<br />

we can still estimate the biaxial modulus as a function <strong>of</strong> the volume fraction <strong>of</strong> textured


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 31<br />

χ<br />

µ<br />

µ α<br />

µ µ<br />

χ<br />

〈100〉 〈111〉<br />

(a)<br />

(b)<br />

Figure 2.4: (a) Powder diffraction x-ray spectra for sputter-deposited copper films on<br />

Ge(111) (dashed), α-Al2O3(0001) (dotted), <strong>and</strong> Si(100) (solid). (b) (111) fiber texture<br />

plot <strong>of</strong> copper films on Ge(111) (dashed) <strong>and</strong> Si(100) (solid).


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 32<br />

χ<br />

χ<br />

µ µ µ<br />

〈100〉 〈111〉<br />

(a)<br />

(b)<br />

Figure 2.5: (a) Powder diffraction spectra, <strong>and</strong> (b) a (111) fiber texture plot for asdeposited<br />

<strong>and</strong> annealed copper films on Si(100).


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 33<br />

µ µ µ<br />

µ µ<br />

Figure 2.6: Powder diffraction spectra for annealed copper films on Si(100) (solid) <strong>and</strong><br />

α-Al2O3(0001) (dotted).<br />

χ<br />

χ<br />

〈100〉 〈111〉<br />

Figure 2.7: (111) fiber texture plot <strong>of</strong> as-deposited <strong>and</strong> annealed 0.50 µm thick copper<br />

films.


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 34<br />

χ<br />

χ<br />

〈100〉 〈111〉<br />

Figure 2.8: (111) fiber texture plot <strong>of</strong> an electron beam evaporated copper film.<br />

gra<strong>in</strong>s <strong>and</strong> the width <strong>of</strong> the texture pole. Figures 2.9 (a) <strong>and</strong> (b) show how the biaxial<br />

modulus depends on the pole width for both 〈111〉 <strong>and</strong> 〈100〉 texture, assum<strong>in</strong>g a Gaussian<br />

distribution <strong>of</strong> the orientations about the pole. We can then estimate the biaxial modulus as<br />

the volume-weighted average:<br />

Y ≈ fr<strong>and</strong>omYr<strong>and</strong>om + f100Y100(σ100) + f111Y111(σ111) (2.6)<br />

where f is the volume fraction <strong>of</strong> properly oriented gra<strong>in</strong>s, <strong>and</strong> σhkl is the st<strong>and</strong>ard de-<br />

viation <strong>of</strong> the texture pole. The results are presented <strong>in</strong> Table 2.4. All the calculations<br />

presented <strong>in</strong> this chapter are based on the Hill average, i.e., the arithmetic average <strong>of</strong> the<br />

Voigt <strong>and</strong> Reuss averages [100].


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 35<br />

Figure 2.9: The calculated biaxial modulus for 〈111〉 texture (a) <strong>and</strong> 〈100〉 texture (b) as a<br />

function <strong>of</strong> the texture pole st<strong>and</strong>ard deviation. A Gaussian texture pole is assumed.<br />

Table 2.4: Calculated values <strong>of</strong> the biaxial modulus for selected films. The st<strong>and</strong>ard deviation<br />

is ±22 GPa.<br />

Substrate Preparation Method Thickness (µm) Y (GPa) Y (GPa)<br />

(as-dep.) (ann.)<br />

Si(100) sputtered 1.25 226 256<br />

Si(100) sputtered 0.50 237 256<br />

Si(100) e-beam evaporated 0.94 203 -<br />

Ge(111) sputtered 1.87 253 -


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 36<br />

2.4.3 Biaxial Modulus<br />

Measur<strong>in</strong>g Both the Biaxial Modulus <strong>and</strong> the CTE<br />

As mentioned <strong>in</strong> section 2.2, by us<strong>in</strong>g two or more substrates with different thermo-<br />

mechanical constants, one can extract both the biaxial modulus <strong>and</strong> the CTE <strong>of</strong> the film.<br />

We did this us<strong>in</strong>g the as-deposited films grown on Si(100), α-Al2O3(0001), <strong>and</strong> fused sil-<br />

ica, choos<strong>in</strong>g these because the textures appeared to be similar. S<strong>in</strong>ce the sharpness <strong>of</strong><br />

the 〈111〉 texture <strong>in</strong>creases much more for copper on α-Al2O3(0001) than for Si(100) after<br />

anneal<strong>in</strong>g, the modulus after anneal<strong>in</strong>g could not be measured with this method. Us<strong>in</strong>g<br />

equations (2.2) <strong>and</strong> (2.3) for the data shown <strong>in</strong> Figure 2.10, we obta<strong>in</strong> Y = 156 ± 18 GPa<br />

<strong>and</strong> α = 20.2 ± 0.5 × 10 −6 K −1 .Ifwefit a l<strong>in</strong>e us<strong>in</strong>g only two substrates, Si(100) <strong>and</strong> α-<br />

Al2O3(0001), we arrive at Y = 179±32 GPa <strong>and</strong> α = 18.2±0.8×10 −6 K −1 . If we only use<br />

the Si(100) substrate <strong>and</strong> assume the bulk value <strong>of</strong> the CTE for <strong>Cu</strong>, α = 16.8 × 10 −6 K −1 ,<br />

we f<strong>in</strong>d Y = 197 ± 8 GPa. The latter value is judged to be the correct one s<strong>in</strong>ce the<br />

bulk CTE for a pure material should be <strong>in</strong>sensitive to microstructural defects. The ther-<br />

momechanical constants <strong>of</strong> fused silica <strong>and</strong> α-Al2O3(0001) are not as well established as<br />

that <strong>of</strong> Si (particularly the former; see Mull<strong>in</strong>’s thesis for problems associated with fused<br />

silica [158]). Most importantly, the textures <strong>of</strong> the copper films on fused silica <strong>and</strong> α-<br />

Al2O3(0001) are slightly different from that on Si(100). Us<strong>in</strong>g the bulk CTE for copper<br />

<strong>and</strong> the thermomechanical properties <strong>in</strong> Table 2.3, we obta<strong>in</strong> Y = 184 ± 8 GPa for copper<br />

on fused silica, Y = 197 ± 8 GPa for copper on Si(100), <strong>and</strong> Y = 201 ± 8 GPa for copper<br />

on α-Al2O3(0001), the variation <strong>of</strong> which can easily be accounted for by slight texture or<br />

microstructural differences caused by the substrate. Thus, all the biaxial moduli described<br />

hereafter were obta<strong>in</strong>ed us<strong>in</strong>g the bulk CTE <strong>of</strong> copper.


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 37<br />

∆σ ∆Τ<br />

α<br />

α × 10 −6 [Κ −1 ]<br />

Figure 2.10: Plot show<strong>in</strong>g how the biaxial modulus <strong>and</strong> CTE can be determ<strong>in</strong>ed for a set<br />

<strong>of</strong> different substrates.<br />

As-deposited <strong>and</strong> Annealed <strong>Films</strong> at Room Temperature<br />

Table 2.5 lists the average biaxial moduli for all the copper films with thicknesses<br />

greater than 1 µm. The st<strong>and</strong>ard deviation <strong>of</strong> ±8 GPa stems from the uncerta<strong>in</strong>ty <strong>in</strong> film<br />

thickness as measured by the pr<strong>of</strong>ilometer. The annealed films refer to samples heated to<br />

550 ◦ Cor600 ◦ C <strong>and</strong> held there for 0.5 hrs. When the sample is heated the first time, the<br />

room temperature modulus <strong>in</strong>creases monotonically with the maximum anneal<strong>in</strong>g temper-<br />

ature (Figure 2.11). After the first anneal <strong>and</strong> stabilization <strong>of</strong> the gra<strong>in</strong> size <strong>and</strong> texture,<br />

additional thermal cycles do not enhance the modulus further, as shown <strong>in</strong> Figure 2.12.<br />

The biaxial modulus <strong>of</strong> either the as-deposited or annealed film appears to be the same<br />

whether it is be<strong>in</strong>g stretched or compressed (true at any temperature). The e-beam evapo-<br />

rated films show similar behavior. This result suggests that the deficit is not due to gra<strong>in</strong><br />

boundary crack<strong>in</strong>g, which would predict a stiffer modulus <strong>in</strong> compression than <strong>in</strong> tension,


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 38<br />

Table 2.5: Measured values <strong>of</strong> the biaxial modulus for films <strong>of</strong> thickness 1 µm or greater.<br />

Substrate Preparation Method Y (GPa) (as-dep.) Y (GPa) (ann.)<br />

Si(100) sputtered 197 ± 8 217 ± 8<br />

Si(100) e-beam evaporated 155 ± 8 158 ± 8<br />

Ge(111) sputtered 262 ± 8 -<br />

α-Al2O3(0001) sputtered 201 ± 8 225 ± 8<br />

fused silica sputtered 184 ± 8 -<br />

Figure 2.11: Biaxial modulus as a function <strong>of</strong> the maximum anneal<strong>in</strong>g temperature exposed<br />

toa1.88 µm thick copper film on Si(100).


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 39<br />

Figure 2.12: Biaxial modulus as a function <strong>of</strong> the number <strong>of</strong> times thermally cycled to<br />

600 ◦ Cfora1.91 µm <strong>and</strong> 1.93 µm thick copper film.<br />

assum<strong>in</strong>g the crack closes completely <strong>in</strong> compression.<br />

Biaxial Modulus as a Function <strong>of</strong> Temperature<br />

Figure 2.13 shows an example <strong>of</strong> how the biaxial modulus can be determ<strong>in</strong>ed as a func-<br />

tion <strong>of</strong> temperature. As long as the elastic thermal cycle keeps a l<strong>in</strong>ear stress-temperature<br />

relationship, the biaxial modulus can be measured.<br />

Figures 2.14 <strong>and</strong> 2.15 depict the biaxial modulus as a function <strong>of</strong> temperature for the<br />

as-deposited films <strong>and</strong> annealed films, respectively. The moduli do not vary accord<strong>in</strong>g to<br />

the temperature dependence <strong>of</strong> the elastic constants found <strong>in</strong> the literature [207] (<strong>in</strong>dicated<br />

by the slope <strong>of</strong> the shaded areas), but drop <strong>of</strong>f faster at higher temperatures. The shaded<br />

areas <strong>in</strong> Figures 2.14 <strong>and</strong> 2.15 represent the texture-dependent biaxial moduli calculated<br />

us<strong>in</strong>g equation (2.6). The width <strong>of</strong> the shaded areas is equal to twice the st<strong>and</strong>ard devi-


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 40<br />

Figure 2.13: Stress as a function <strong>of</strong> temperature for a 1.19 µm thick <strong>Cu</strong> film on Si(100).<br />

The elastic thermal cycles for the measurements <strong>of</strong> biaxial modulus at five different temperatures<br />

are shown for the as-deposited film (a) <strong>and</strong> the thermally cycled film (b).<br />

ation <strong>of</strong> 22 GPa. As one can tell from Figure 2.13 (a), gra<strong>in</strong> growth causes substantial<br />

stress relaxation once the temperature exceeds 150 ◦ C. Thus, the data <strong>in</strong> Figure 2.14 are<br />

marred by a change <strong>in</strong> microstructure <strong>and</strong> texture, which would raise the modulus at higher<br />

temperatures <strong>in</strong> the absence <strong>of</strong> a modulus deficit.<br />

Biaxial Modulus as a Function <strong>of</strong> Oscillation Frequency<br />

The effect <strong>of</strong> the heat<strong>in</strong>g/cool<strong>in</strong>g rate (oscillation frequency) on the biaxial modulus was<br />

exam<strong>in</strong>ed. A typical result is shown <strong>in</strong> Figure 2.16. The modulus seems to be unaffected<br />

by the oscillation frequency, at least with<strong>in</strong> this frequency range.


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 41<br />

Figure 2.14: Biaxial modulus as a function <strong>of</strong> temperature for as-deposited copper films.<br />

The error bar denotes two times the st<strong>and</strong>ard deviation. The shaded area represents the<br />

calculated biaxial moduli after account<strong>in</strong>g for texture. It is based on a value <strong>of</strong> 226±22 GPa<br />

at room temperature (cf. Table 2.4) <strong>and</strong> has the temperature dependence found <strong>in</strong> Simmons<br />

<strong>and</strong> Wang [207].


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 42<br />

Figure 2.15: Biaxial modulus as a function <strong>of</strong> temperature for a copper film previously<br />

annealed at 600 ◦ C. The error bar denotes two times the st<strong>and</strong>ard deviation. The shaded<br />

area represents the calculated biaxial moduli after account<strong>in</strong>g for texture. It is based on<br />

a value <strong>of</strong> 256 ± 22 GPa at room temperature (cf. Table 2.4) <strong>and</strong> has the temperature<br />

dependence found <strong>in</strong> Simmons <strong>and</strong> Wang [207].


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 43<br />

×<br />

×<br />

Figure 2.16: Biaxial modulus as a function <strong>of</strong> temperature for copper films previously<br />

thermally cycled to 600 ◦ C for different stra<strong>in</strong> oscillation frequencies.<br />

Table 2.6: Comparison <strong>of</strong> the measured <strong>and</strong> calculated values <strong>of</strong> the biaxial modulus for<br />

the as-deposited films.<br />

Substrate Prep. Method Thickness Ymeas Ycalc Deficit<br />

(as-dep.) (as-dep.)<br />

Si(100) sputtered 1.25 µm 192 ± 8 GPa 226 ± 22 GPa 15%<br />

Si(100) sputtered 0.50 µm 175 ± 8 GPa 237 ± 22 GPa 26%<br />

Si(100) e-beam evap. 0.94 µm 155 ± 8 GPa 203 ± 22 GPa 23%<br />

Ge(111) sputtered 1.87 µm 262 ± 8 GPa 253 ± 22 GPa ∼ 0%<br />

2.5 Discussion<br />

2.5.1 Texture<br />

The measured biaxial moduli are compared to the calculated moduli <strong>in</strong> Tables 2.6 <strong>and</strong><br />

2.7. Exclud<strong>in</strong>g the Ge(111) sample, there still exists a modulus deficit, even when properly<br />

account<strong>in</strong>g for texture. The <strong>in</strong>crease <strong>in</strong> modulus with anneal<strong>in</strong>g appears to be nearly fully<br />

accounted for by a sharpen<strong>in</strong>g <strong>of</strong> the 〈111〉 texture.


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 44<br />

Table 2.7: Comparison <strong>of</strong> the measured <strong>and</strong> calculated values <strong>of</strong> the biaxial modulus for<br />

the annealed films.<br />

Substrate Prep. Method Thickness Ymeas Ycalc Deficit<br />

(ann.) (ann.)<br />

Si(100) sputtered 1.25 µm 224 ± 8 GPa 256 ± 22 GPa 12.5%<br />

Si(100) sputtered 0.50 µm 196 ± 8 GPa 256 ± 22 GPa 23%<br />

2.5.2 Cause <strong>of</strong> the Modulus Deficit<br />

The dependence <strong>of</strong> the biaxial modulus on temperature suggests an anelastic effect.<br />

Gra<strong>in</strong> boundary slid<strong>in</strong>g is a likely contributor. Bohn <strong>and</strong> Schill<strong>in</strong>g [97, 98, 17] have per-<br />

formed a series <strong>of</strong> <strong>in</strong>ternal friction experiments us<strong>in</strong>g both the vibrat<strong>in</strong>g reed <strong>and</strong> drumhead<br />

geometries to study gra<strong>in</strong> boundary slid<strong>in</strong>g <strong>in</strong> substrate-bonded <strong>and</strong> free-st<strong>and</strong><strong>in</strong>g Al th<strong>in</strong><br />

films approximately 4 µm thick. They have shown that the <strong>in</strong>ternal friction peak they stud-<br />

ied cannot be due to dislocation anelasticity, nor, for the substrate-bonded films, slid<strong>in</strong>g<br />

<strong>of</strong> the gra<strong>in</strong>s at the substrate <strong>in</strong>terface. V<strong>in</strong>ci et al. [237] have carried out frequency de-<br />

pendent tensile tests on free-st<strong>and</strong><strong>in</strong>g Al <strong>and</strong> partly attributed the modulus deficit to gra<strong>in</strong><br />

boundary slid<strong>in</strong>g. More recently, Kalkman et al. [122] have studied the Young modulus <strong>in</strong><br />

free-st<strong>and</strong><strong>in</strong>g Al, Au, <strong>and</strong> W films us<strong>in</strong>g a dynamic bulge test <strong>and</strong> found the modulus to<br />

decrease with oscillation frequency.<br />

However, the results <strong>in</strong> Tables 2.6 <strong>and</strong> 2.7 show that despite a ten-fold <strong>in</strong>crease <strong>in</strong> gra<strong>in</strong><br />

size, the deficit is reduced only marg<strong>in</strong>ally. This also does not support the gra<strong>in</strong> boundary<br />

crack<strong>in</strong>g theory, s<strong>in</strong>ce anneal<strong>in</strong>g would presumably heal the cracks. Furthermore, gra<strong>in</strong><br />

boundary slid<strong>in</strong>g cannot expla<strong>in</strong> the temperature dependence. The experimental frequen-<br />

cies are not with<strong>in</strong> the range required to detect a temperature dependent difference <strong>in</strong> the<br />

biaxial modulus.<br />

Accord<strong>in</strong>g to the gra<strong>in</strong> boundary slid<strong>in</strong>g mechanism, the characteristic relaxation time


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 45<br />

is dependent on the gra<strong>in</strong> size <strong>and</strong> the temperature [122]:<br />

τ ∝ d n exp(Eact/kBT ) (2.7)<br />

where 1 < n < 2 <strong>and</strong> Eact is the gra<strong>in</strong> boundary slid<strong>in</strong>g activation energy (equivalent to<br />

the activation energy for gra<strong>in</strong> boundary diffusion). The modulus <strong>in</strong> an <strong>in</strong>ternal friction<br />

experiment is given by [106]<br />

Y (ω, τ) ≈ YR<br />

�<br />

1 + δY<br />

YR<br />

τω<br />

π<br />

� �<br />

1 − exp − YU<br />

���<br />

π<br />

YR τω<br />

(2.8)<br />

where YU is the unrelaxed modulus, YR is the relaxed modulus, δY is the difference be-<br />

tween the relaxed <strong>and</strong> unrelaxed modulus, <strong>and</strong> ω is the oscillation frequency. If we <strong>in</strong>-<br />

sert reasonable values for the relaxed <strong>and</strong> unrelaxed moduli <strong>in</strong>to (2.8), for a frequency <strong>of</strong><br />

10 −3 rad sec −1 <strong>and</strong> Eact ≈ 1.0 eV, the biaxial modulus at room temperature <strong>and</strong> the biaxial<br />

modulus at 300 ◦ C are nearly equal. The only rema<strong>in</strong><strong>in</strong>g plausible anelasticity mechanism<br />

is dislocation anelasticity.<br />

A simple estimate <strong>of</strong> the reduction <strong>in</strong> modulus from dislocations is<br />

δY<br />

Y<br />

≈ 1<br />

20 ρℓ2<br />

(2.9)<br />

where ρ is the dislocation density <strong>and</strong> ℓ is the dislocation p<strong>in</strong>n<strong>in</strong>g length [68]. As shown<br />

by the schematic <strong>in</strong> Figure 2.17, it is possible for the total dislocation l<strong>in</strong>e length per unit<br />

volume, i.e., the dislocation density to rema<strong>in</strong> the same at high temperatures while the<br />

p<strong>in</strong>n<strong>in</strong>g length <strong>in</strong>creases by the thermal activation <strong>of</strong> dislocations past weak obstacles. This<br />

has some support <strong>in</strong> the literature [128, 48].<br />

Although the frequency dependence <strong>of</strong> the modulus was not observed (Figure 2.16),<br />

the tested frequency range (8.73 × 10 −4 to 5.24 × 10 −3 rad sec −1 ) was not wide enough to<br />

expect any substantial changes <strong>in</strong> modulus.


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 46<br />

l LT<br />

l HT<br />

Figure 2.17: Depiction <strong>of</strong> dislocation bow<strong>in</strong>g at high temperatures. Thermal fluctuations<br />

help the dislocation to bow past weak obstacles, but the radius <strong>of</strong> curvature rema<strong>in</strong>s below<br />

that required for slip.<br />

The electron beam evaporated film has a larger modulus deficit compared to the sput-<br />

tered films; its average gra<strong>in</strong> size is large, on the order <strong>of</strong> 1 µm [261], <strong>and</strong> the 〈111〉 texture<br />

is weak. Also, the “critical” temperature alluded to earlier is much lower for the evapo-<br />

rated film — about 165 ◦ C compared to about 300 ◦ C for the sputtered films. This may be<br />

related to the gra<strong>in</strong> boundary mobility: unlike the sputtered films with trapped Ar atoms,<br />

the e-beam films have no impurities to <strong>in</strong>hibit gra<strong>in</strong> growth.<br />

The e-beam microstructure should be contrasted with that <strong>of</strong> sputtered <strong>Cu</strong> on Ge(111);<br />

the gra<strong>in</strong>s extend from the substrate to the surface, but the <strong>in</strong>-plane gra<strong>in</strong> size is approxi-<br />

mately 50 nm. Recall there is very sharp 〈111〉 texture for this film <strong>and</strong> no modulus deficit.<br />

Indeed, there may be some relationship between the width <strong>of</strong> 〈111〉 texture pole <strong>and</strong> the<br />

deficit, but more evidence is required for confirmation. If there is a relationship, this <strong>in</strong>di-


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 47<br />

cates that the nature <strong>of</strong> the gra<strong>in</strong> boundary is different <strong>in</strong> the different films.<br />

2.6 Conclusions <strong>and</strong> Future Research<br />

There are three ma<strong>in</strong> observations concern<strong>in</strong>g these experiments. First, the biaxial mod-<br />

ulus <strong>of</strong> <strong>Cu</strong> films <strong>in</strong>creases after thermal cycl<strong>in</strong>g due primarily to enhanced 〈111〉 texture.<br />

The deficit only changes marg<strong>in</strong>ally despite significant gra<strong>in</strong> growth. Second, the modulus<br />

deficit varies from -23% to 0% depend<strong>in</strong>g on the deposition conditions <strong>and</strong> the substrate.<br />

Third, the magnitude <strong>of</strong> the deficit <strong>in</strong>creases at higher temperatures.<br />

From the temperature dependence <strong>of</strong> the modulus deficit, it appears that it can be par-<br />

tially attributed to dislocation anelasticity. This has been reported for free-st<strong>and</strong><strong>in</strong>g films<br />

<strong>in</strong> the literature [237], but not for substrate bonded films. It is unclear whether or not it<br />

is the dom<strong>in</strong>ant anelasticity mechanism, but it is anticipated that substrate-bonded films<br />

can experience more dislocation anelasticity than free-st<strong>and</strong><strong>in</strong>g films due to the dislocation<br />

storage capacity <strong>of</strong> the former.<br />

More work should be carried out on the frequency dependence <strong>of</strong> the modulus; if the<br />

material studied is copper, this needs to be performed <strong>in</strong> conjunction with χ-scans or even<br />

ODF measurements. Both sputtered <strong>and</strong> evaporated films should be studied.<br />

It also would be <strong>in</strong>terest<strong>in</strong>g to repeat the biaxial modulus as a function <strong>of</strong> tempera-<br />

ture experiments on the cool<strong>in</strong>g part <strong>of</strong> the thermal cycle rather than the heat<strong>in</strong>g part. As<br />

described <strong>in</strong> Chapter 3, the dislocation density <strong>and</strong> p<strong>in</strong>n<strong>in</strong>g length at a given temperature<br />

depend on whether or not the film is be<strong>in</strong>g heated or cooled.<br />

F<strong>in</strong>ally, the Young modulus <strong>and</strong> Poisson ratio should be checked us<strong>in</strong>g the Y -cut quartz<br />

substrates both before <strong>and</strong> after anneal<strong>in</strong>g [262]. An exam<strong>in</strong>ation <strong>of</strong> the changes <strong>in</strong> E <strong>and</strong> ν


Chapter 2: The Elastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 48<br />

due to anneal<strong>in</strong>g may further dist<strong>in</strong>guish between the gra<strong>in</strong> boundary crack<strong>in</strong>g <strong>and</strong> anelastic<br />

models.


Chapter 3<br />

The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong><br />

The experiment described <strong>in</strong> this chapter is the search for a dislocation “boundary<br />

layer”; pro<strong>of</strong> <strong>of</strong> the existence <strong>of</strong> such a layer would lend support to one <strong>of</strong> the major theo-<br />

ries on the strength <strong>of</strong> th<strong>in</strong> films: stra<strong>in</strong> gradient plasticity. This <strong>in</strong>troduction surveys some<br />

<strong>of</strong> the current theories regard<strong>in</strong>g th<strong>in</strong> film plasticity.<br />

3.1 The <strong>Cu</strong>rrent Underst<strong>and</strong><strong>in</strong>g <strong>of</strong> Plastic Properties <strong>of</strong><br />

Th<strong>in</strong> <strong>Films</strong><br />

Perhaps the first to carry out a systematic <strong>in</strong>vestigation <strong>of</strong> plastic properties <strong>in</strong> th<strong>in</strong> films<br />

was Murakami at IBM [166, 161, 160, 162, 165, 167], who studied thermal stra<strong>in</strong>s <strong>in</strong> Pb<br />

th<strong>in</strong> films for the expressed purpose <strong>of</strong> develop<strong>in</strong>g reliable Josephson junctions [168, 163,<br />

164]. His results stimulated two <strong>of</strong> his colleagues, Chaudhari <strong>and</strong> Ronay, to propose the<br />

first quantitative models <strong>of</strong> plastic flow <strong>in</strong> th<strong>in</strong> films <strong>in</strong> 1979 [34, 37, 35, 36, 191]. In the<br />

late 1980s, <strong>Ni</strong>x <strong>and</strong> his group at Stanford began study<strong>in</strong>g th<strong>in</strong> film stresses by perform<strong>in</strong>g<br />

49


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 50<br />

wafer curvature measurements dur<strong>in</strong>g thermal cycl<strong>in</strong>g [62]. Alum<strong>in</strong>um <strong>and</strong> alum<strong>in</strong>um alloy<br />

th<strong>in</strong> films were extensively studied at first because <strong>of</strong> their use as <strong>in</strong>terconnect materials<br />

[53, 233, 231, 232]. As it became apparent that copper would be more cost-effective than<br />

alum<strong>in</strong>um (hav<strong>in</strong>g a lower resistivity <strong>and</strong> a higher melt<strong>in</strong>g temperature), researchers at<br />

Stanford [263, 238, 236] <strong>and</strong> at the Max Planck Institute for Metals Research <strong>in</strong> Stuttgart<br />

[124, 125, 10] began to focus on copper th<strong>in</strong> films.<br />

The “classic” reviews <strong>in</strong> th<strong>in</strong> film mechanical properties are by <strong>Ni</strong>x <strong>and</strong> Arzt [54, 172,<br />

173, 6], although some <strong>of</strong> the more recent papers help to clarify several issues as well as<br />

raise more questions. In this <strong>in</strong>troduction, I will focus on the models currently used to<br />

describe low temperature flow, but first I will present a brief description <strong>of</strong> deformation<br />

mechanisms at high temperatures (above ∼ 200 ◦ C) s<strong>in</strong>ce they have some bear<strong>in</strong>g on the<br />

room temperature yield strength. The discussion is restricted to unpassivated th<strong>in</strong> films.<br />

3.1.1 High Temperature Deformation<br />

The earliest model <strong>of</strong> high temperature deformation allowed for gra<strong>in</strong> slid<strong>in</strong>g at the<br />

film/substrate <strong>in</strong>terface [221, 223, 222]; for films <strong>in</strong> which the gra<strong>in</strong> size is less than the<br />

film thickness, the data agreed with the model. If, however, the adhesion between the film<br />

<strong>and</strong> substrate is strong enough to prohibit diffusion along the <strong>in</strong>terface, only partial stress<br />

relaxation by gra<strong>in</strong> boundary diffusion can take place. This supposition was first made<br />

by Jackson <strong>and</strong> Li [110], who predicted the creation <strong>of</strong> stra<strong>in</strong> gradients <strong>in</strong> the film due<br />

to this phenomenon. More recently, Gao et al. [73] have made a mathematical model to<br />

predict the decay <strong>of</strong> normal tractions at the gra<strong>in</strong> boundaries with time. They also predicted<br />

that so-called gra<strong>in</strong> boundary diffusion wedges could serve as sources <strong>of</strong> dislocations. For


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 51<br />

copper films grown <strong>and</strong> annealed <strong>in</strong> high vacuum (HV), however, surface diffusion seems<br />

to be shut <strong>of</strong>f entirely (perhaps due to a metastable oxide), lead<strong>in</strong>g to negligible stress<br />

relaxation by gra<strong>in</strong> boundary diffusion, for which surface diffusion is a required pathway.<br />

The predom<strong>in</strong>ant stress relaxation mechanism <strong>in</strong> HV-grown <strong>Cu</strong> at high temperatures is<br />

power-law creep [125], or at least the mathematical formulation <strong>of</strong> power-law creep seems<br />

to fit the data. The actual mechanism by which power-law creep works <strong>in</strong> th<strong>in</strong> films is not<br />

yet understood. Incidentally, it is important to dist<strong>in</strong>guish between experiments performed<br />

<strong>in</strong> clean <strong>and</strong> “ultra-clean” conditions. For example, UHV-grown <strong>Cu</strong> films annealed <strong>in</strong> a<br />

UHV furnace do experience substantial surface diffusion [246].<br />

3.1.2 Low Temperature Deformation<br />

At low temperatures plasticity operates by dislocation motion. Most dislocation models<br />

have not been successful <strong>in</strong> simulat<strong>in</strong>g the low temperature portion <strong>of</strong> the stress-temperature<br />

pr<strong>of</strong>ile, although work on this is apparently <strong>in</strong> progress [246]. Instead, most models focus<br />

on a prediction <strong>of</strong> the room temperature yield strength.<br />

At room temperature, the yield strength hierarchy is as follows:<br />

σy, passivated >σy, unpassivated >σy, free−st<strong>and</strong><strong>in</strong>g >σy, bulk, (3.1)<br />

where σy is the yield strength. For free-st<strong>and</strong><strong>in</strong>g th<strong>in</strong> films <strong>and</strong> the bulk, the yield strength<br />

can be determ<strong>in</strong>ed by the usual 0.2% plastic stra<strong>in</strong> criterion. For unpassivated <strong>and</strong> pas-<br />

sivated th<strong>in</strong> films (the use <strong>of</strong> these terms imply the films are substrate-bonded), the yield<br />

strength is assumed to be the flow stress at room temperature as determ<strong>in</strong>ed by wafer curva-<br />

ture measurements [53], although this assumption is not always valid. Thus, from here on, I


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 52<br />

will call what others label the yield strength, σy, i.e., the biaxial stress at room temperature<br />

after a thermal cycle, the flow strength, σflow, <strong>in</strong>stead.<br />

The flow strength has several orig<strong>in</strong>s, each <strong>of</strong> which depends on the nature <strong>of</strong> the resis-<br />

tance to dislocation motion. There are three ma<strong>in</strong> contributions to the flow strength <strong>of</strong> th<strong>in</strong><br />

films: the constra<strong>in</strong>t <strong>of</strong> the film by the substrate, the small gra<strong>in</strong> size, <strong>and</strong> work harden<strong>in</strong>g.<br />

The mathematical form <strong>of</strong> these models can be written as<br />

σconstra<strong>in</strong>t = A<br />

, (3.2)<br />

σ I gra<strong>in</strong> size<br />

or σ II<br />

gra<strong>in</strong> size<br />

tf<br />

B1 = d , (3.3a)<br />

= B2<br />

√d , (3.3b)<br />

<strong>and</strong> σharden<strong>in</strong>g = C √ ρ (3.4)<br />

where tf is the film thickness, d is the gra<strong>in</strong> size, ρ is the dislocation density, <strong>and</strong> A, Bi,<br />

<strong>and</strong> C are model-dependent parameters that are discussed below. As we will see, these<br />

equations are only approximate. Researchers have usually taken the total flow strength to<br />

be equal to the superposition <strong>of</strong> all these strengthen<strong>in</strong>g mechanisms (dislocation obstacles)<br />

[124, 10]:<br />

σflow = σconstra<strong>in</strong>t + σgra<strong>in</strong> size + σharden<strong>in</strong>g, (3.5)<br />

although this is true only if every dislocation must overcome all the obstacles simultane-<br />

ously. Equations (3.3b) <strong>and</strong> (3.4) were previously derived for bulk materials but rema<strong>in</strong><br />

valid for th<strong>in</strong> films; the ma<strong>in</strong> difference is that the parameters B2 <strong>and</strong> C are much larger for<br />

th<strong>in</strong> films [124].


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 53<br />

There is some experimental verification for equation (3.5): Venkatraman <strong>and</strong> Bravman<br />

[233] have shown that for alum<strong>in</strong>um th<strong>in</strong> films,<br />

σflow = σconstra<strong>in</strong>t + σgra<strong>in</strong> size<br />

(3.6)<br />

where the gra<strong>in</strong> size refers to the <strong>in</strong>-plane size. However, the work harden<strong>in</strong>g contribution<br />

is ignored (perhaps for good reason, as discussed below), <strong>and</strong> there rema<strong>in</strong>s some debate<br />

on whether equation (3.3a) or (3.3b) is more appropriate [124, 56] for the gra<strong>in</strong> size rela-<br />

tionship.<br />

Many compet<strong>in</strong>g theories are developed for the empirical models described above, but<br />

<strong>of</strong>ten none <strong>of</strong> them can be dist<strong>in</strong>guished experimentally due to the difficulty <strong>in</strong> see<strong>in</strong>g how<br />

an ensemble <strong>of</strong> dislocations behaves <strong>in</strong> situ. The situation is naturally worse for poly-<br />

crystals than for s<strong>in</strong>gle crystals, yet there still rema<strong>in</strong>s contentious debate on the nature <strong>of</strong><br />

work harden<strong>in</strong>g <strong>in</strong> s<strong>in</strong>gle crystals [134]. A natural approach nowadays is to perform com-<br />

puter simulations <strong>of</strong> dislocation <strong>in</strong>teractions [171], but we rema<strong>in</strong> a long way away from<br />

simulat<strong>in</strong>g realistic structures. The field is too young for the development <strong>of</strong> a “statistical<br />

mechanics” <strong>of</strong> dislocations.<br />

In a recent review, Gao et al. [71] categorize the current theories accord<strong>in</strong>g to the length<br />

scale <strong>in</strong> which they are operative. For deformation length scales larger than 100 µm, clas-<br />

sical plasticity works. Stra<strong>in</strong> gradient plasticity is applicable <strong>in</strong> the 0.1 µm − 10 µm range,<br />

<strong>and</strong> discrete dislocation models work <strong>in</strong> the < 0.1 µm regime. Thus, stra<strong>in</strong> gradient plas-<br />

ticity can successfully fill the void left by the lack <strong>of</strong> a statistical mechanics formulation.<br />

I will beg<strong>in</strong> by discuss<strong>in</strong>g the discrete dislocation models <strong>and</strong> how they relate to the<br />

strength <strong>of</strong> th<strong>in</strong> films. Although Gao et al. [71] have stated stra<strong>in</strong> gradient plasticity to<br />

be the theory <strong>of</strong> choice <strong>in</strong> the 1 µm regime, nearly all <strong>of</strong> the models <strong>in</strong> the literature use


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 54<br />

discrete dislocation models to predict the flow strength for films <strong>of</strong> this thickness.<br />

Discrete Dislocation Models<br />

THE CONSTRAINT OF THE SUBSTRATE<br />

One <strong>of</strong> the more famous “criteria” <strong>in</strong> th<strong>in</strong> film physics is the Matthews-Blakeslee cri-<br />

terion, which gives the maximum film thickness <strong>of</strong> an epitaxial layer before nucleation <strong>of</strong><br />

a periodic array <strong>of</strong> misfit dislocations [152, 153, 151]. They arrived at their criterion by<br />

consider<strong>in</strong>g the forces necessary to propagate a thread<strong>in</strong>g dislocation across an epitaxial<br />

film. Freund later generalized <strong>and</strong> ref<strong>in</strong>ed their result us<strong>in</strong>g a dislocation dynamics ap-<br />

proach [65]. Here, I use <strong>Ni</strong>x’s <strong>in</strong>terpretation <strong>of</strong> Freund’s work [172], which <strong>in</strong>cludes many<br />

simplifications, render<strong>in</strong>g it nearly equivalent to the Matthews-Blakeslee result. The work<br />

per unit length performed by the coherency stress dur<strong>in</strong>g misfit dislocation formation is<br />

Wlayer =<br />

cos λ cos φ<br />

σfbtf, (3.7)<br />

s<strong>in</strong> φ<br />

where b is the Burgers vector, σf is the coherency stress, φ is the angle between the glide<br />

plane normal <strong>and</strong> the film normal, <strong>and</strong> λ is the angle between the Burgers vector <strong>and</strong> the<br />

film normal (see Figure 3.1); the geometric term cos λ cos φ is the Schmid factor. The work<br />

to make a unit length <strong>of</strong> edge dislocation at the <strong>in</strong>terface is (after the image forces at the<br />

<strong>in</strong>terface <strong>of</strong> a stiff substrate are considered [96] <strong>and</strong> ν = νs = νf is assumed)<br />

Wdis =<br />

b2 2GfGs<br />

4π (1 − ν) (Gf + Gs) ln<br />

� �<br />

βtf<br />

b<br />

(3.8)<br />

where Gi are the shear moduli <strong>of</strong> the film <strong>and</strong> substrate, β is a dislocation core radius cut<strong>of</strong>f<br />

(≈ 2), <strong>and</strong> elastic isotropy <strong>in</strong> the film <strong>and</strong> substrate is assumed. By equat<strong>in</strong>g (3.7) to (3.8),


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 55<br />

λ φ<br />

Figure 3.1: Diagram show<strong>in</strong>g the nucleation <strong>and</strong> propagation <strong>of</strong> dislocations along a slip<br />

plane from the surface to the gra<strong>in</strong> boundaries <strong>and</strong> film/substrate <strong>in</strong>terface.<br />

we arrive at<br />

tf,c =<br />

s<strong>in</strong> φ<br />

cos λ cos φ<br />

from which the critical thickness can be determ<strong>in</strong>ed.<br />

2GfGs b ln<br />

(Gf + Gs)<br />

� βtf,c/b �<br />

4π (1 − ν) σf<br />

(3.9)<br />

<strong>Ni</strong>x [172] recognized the importance <strong>of</strong> this equation. If the thickness <strong>of</strong> the film is<br />

larger than the critical thickness, σf is equal to the stress required to move a dislocation<br />

along the <strong>in</strong>terface. If this is the limit<strong>in</strong>g factor for dislocation motion,<br />

σconstra<strong>in</strong>t =<br />

s<strong>in</strong> φ b<br />

�<br />

GfGs<br />

cos φ cos λ 2π(1 − ν)tf<br />

Gf + Gs<br />

ln<br />

� ��<br />

βtf<br />

. (3.10)<br />

b<br />

As described below, Chaudhari was actually the first to derive this relationship [36]. This


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 56<br />

model predicts the experimentally observed 1/tf dependence <strong>and</strong> is <strong>of</strong>ten applied to poly-<br />

crystall<strong>in</strong>e films.<br />

Baker et al. [9, 181] have used this model to expla<strong>in</strong> the Bausch<strong>in</strong>ger effect (σy(tension)<br />

>σy(compression)) <strong>of</strong>ten found <strong>in</strong> stress-temperature pr<strong>of</strong>iles. They believe that when<br />

lower<strong>in</strong>g the stress from room temperature by heat<strong>in</strong>g, the dislocations run backwards eas-<br />

ily, remov<strong>in</strong>g the energetically expensive l<strong>in</strong>e segments from the film/substrate <strong>in</strong>terface.<br />

Thus, the plasticity found <strong>in</strong> compression is driven both by an applied compressive stress<br />

<strong>and</strong> the removal <strong>of</strong> misfit dislocation l<strong>in</strong>e length.<br />

Weihnacht <strong>and</strong> Brückner [242, 243, 245] have an alternate explanation for the Bausch<strong>in</strong>ger<br />

effect: pile-ups. They assume that dislocations nucleate at the surface <strong>and</strong> pile up at the<br />

substrate <strong>in</strong>terface when cool<strong>in</strong>g from the anneal<strong>in</strong>g temperature. The same dislocations<br />

then undergo “easy” back glid<strong>in</strong>g (aided by the mutually repulsive forces with<strong>in</strong> the pile-<br />

up) across the same slip planes when the film is heated (Figure 3.2). This back glid<strong>in</strong>g can<br />

occur even when the stress state is still positive, i.e., <strong>in</strong> tension.<br />

Weihnacht <strong>and</strong> Brückner’s model has some experimental support: us<strong>in</strong>g <strong>in</strong> situ thermal<br />

cycl<strong>in</strong>g <strong>of</strong> Al/Si(100) <strong>in</strong> the TEM, Allen et al. [1] showed that dislocations move <strong>in</strong> <strong>and</strong> out<br />

along the same slip planes as the temperature is cycled.<br />

THE GRAIN SIZE EFFECT<br />

As mentioned above, Chaudhari [36] was the first to derive (3.10). Rather than us<strong>in</strong>g<br />

a dislocation dynamics method, he relied on an energy m<strong>in</strong>imization approach, calculat<strong>in</strong>g<br />

the total stra<strong>in</strong> <strong>and</strong> dislocation energy <strong>in</strong> the film <strong>and</strong> sett<strong>in</strong>g the derivative with respect to<br />

stra<strong>in</strong> equal to zero. He obta<strong>in</strong>ed the same equation as above, but with ln(2tf/b) + 1 rather


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 57<br />

σ<br />

Figure 3.2: Schematic show<strong>in</strong>g the predicted dislocation behavior at different po<strong>in</strong>ts on the<br />

stress-temperature pr<strong>of</strong>ile. Based on the diagram by Weihnacht <strong>and</strong> Brückner [245].<br />

than ln βtf/b. In addition, Chaudhari considered the strengthen<strong>in</strong>g due to gra<strong>in</strong> boundaries:<br />

σ I gra<strong>in</strong> size =<br />

� �<br />

s<strong>in</strong> φ Gfb<br />

1.87 +<br />

cos φ cos λ 4πd<br />

1.29<br />

�� � � ��<br />

d<br />

ln + 1 , (3.11)<br />

1 − ν b<br />

as well as strengthen<strong>in</strong>g due to ledge formation at the film surface:<br />

σledge =<br />

s<strong>in</strong> φ γ<br />

cos φ cos λ tf<br />

T<br />

(3.12)<br />

where the gra<strong>in</strong> size d is the cut-<strong>of</strong>f radius for the dislocations near the gra<strong>in</strong> boundary, <strong>and</strong><br />

γ is the surface energy per unit area. The numerical values <strong>in</strong> (3.11) are due to some f<strong>in</strong>e<br />

details <strong>and</strong> are not important here. More recently, Thompson [218] <strong>in</strong>dependently derived<br />

a similar equation.<br />

In bulk materials, Hall [92] <strong>and</strong> Petch [185] found the flow strength to scale with the


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 58<br />

<strong>in</strong>verse square root <strong>of</strong> gra<strong>in</strong> size:<br />

σ II<br />

gra<strong>in</strong> size = σ0 + kHPd −1/2<br />

(3.13)<br />

where σ0 <strong>and</strong> kHP are material-dependent quantities. This empirical equation has been<br />

justified several ways [57, 143, 8]. The orig<strong>in</strong>al explanation was that the flow strength is<br />

equivalent to the critical stress (caused by dislocation pile-up) <strong>in</strong> a gra<strong>in</strong> needed to activate<br />

slip <strong>in</strong> the neighbor<strong>in</strong>g gra<strong>in</strong>. The value <strong>of</strong> kHP is larger <strong>in</strong> th<strong>in</strong> films than <strong>in</strong> the bulk [124].<br />

WORK HARDENING<br />

Taylor [216, 217] was the first to derive<br />

σwork harden<strong>in</strong>g = kTbGρ 1/2 , (3.14)<br />

<strong>and</strong> numerous subsequent derivations can be found [170]. Accord<strong>in</strong>g to the Taylor deriva-<br />

tion, the stress field a distance r away from a screw dislocation is given by bG/2πr. If<br />

we presume slip occurs at r<strong>and</strong>om po<strong>in</strong>ts <strong>in</strong> the crystal by the separation <strong>of</strong> positive <strong>and</strong><br />

negative dislocations, <strong>and</strong> on the average the dislocations move a distance ℓ before be-<br />

<strong>in</strong>g blocked, the plastic shear stra<strong>in</strong> equals ρbℓ, which leads to the Taylor relation with<br />

kT = 1/2π. Other derivations, <strong>in</strong>clud<strong>in</strong>g those that consider image forces on mov<strong>in</strong>g<br />

dislocations (long-range <strong>in</strong>teractions) <strong>and</strong> cross<strong>in</strong>g dislocations on different slip systems<br />

(short-range <strong>in</strong>teractions) can <strong>in</strong>crease the value <strong>of</strong> kT to 1 [191]. Another way <strong>of</strong> deriv<strong>in</strong>g<br />

(3.14) is simply to <strong>in</strong>sert the relationship ℓ = ρ −1/2 <strong>in</strong>to the Orowan equation (1.11). More<br />

recently, Weihnacht <strong>and</strong> Brückner [244] have developed a work harden<strong>in</strong>g model that con-<br />

siders the <strong>in</strong>teractions <strong>of</strong> a thread<strong>in</strong>g dislocation with a parallel array <strong>of</strong> misfit dislocations.<br />

Note that an implicit assumption <strong>in</strong> most dislocation models is that the flow is <strong>in</strong> a<br />

drag-controlled regime rather than an obstacle-controlled glide regime. New observations


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 59<br />

by Kobr<strong>in</strong>sky <strong>and</strong> Thompson [128] <strong>and</strong> Dehm et al. [48] have shown very short mean<br />

free paths for dislocation motion: 80 nm for annealed <strong>Ag</strong> <strong>and</strong> 70 nm for annealed <strong>Cu</strong>.<br />

These lengths were much smaller than the gra<strong>in</strong> size. The dislocations moved <strong>in</strong> a “jerky”<br />

fashion characteristic <strong>of</strong> dislocations requir<strong>in</strong>g thermal fluctuations to overcome obstacles<br />

(forest dislocations) [130]. They did not observe dislocation sweeps across the slip plane<br />

as depicted <strong>in</strong> Figure 3.1. These observations suggest more work needs to be done to<br />

underst<strong>and</strong> the k<strong>in</strong>etics <strong>of</strong> slip <strong>in</strong> th<strong>in</strong> films; also, the work harden<strong>in</strong>g contribution to the<br />

flow strength may be the most important one.<br />

Us<strong>in</strong>g these discrete dislocation models <strong>and</strong> the superposition equation (3.5) was the<br />

first step <strong>in</strong> try<strong>in</strong>g to underst<strong>and</strong> the flow strength <strong>in</strong> th<strong>in</strong> films. In general, equation (3.5)<br />

tends to underestimate the experimental th<strong>in</strong> film flow strength. Any good agreement be-<br />

tween the two should be considered fortuitous given the large number <strong>of</strong> assumptions <strong>in</strong>-<br />

volved. Moreover, equations (3.3a) (or (3.3b)) <strong>and</strong> (3.4) cannot be <strong>in</strong>dependent; Li has<br />

suggested the equations are <strong>in</strong> fact redundant s<strong>in</strong>ce ρ tends to scale with 1/d [142].<br />

Stra<strong>in</strong> Gradient Plasticity <strong>in</strong> Th<strong>in</strong> <strong>Films</strong><br />

The start<strong>in</strong>g po<strong>in</strong>t for any model predict<strong>in</strong>g flow strength <strong>in</strong> substrate-bonded films is<br />

the effect <strong>of</strong> the substrate on dislocation behavior s<strong>in</strong>ce the mechanical properties <strong>of</strong> free-<br />

st<strong>and</strong><strong>in</strong>g films can be predicted us<strong>in</strong>g the Hall-Petch relation. Accord<strong>in</strong>g to the theory <strong>of</strong><br />

Head [96], dislocations should be repelled by the rigid substrate if the shear modulus <strong>of</strong><br />

the substrate is larger than that <strong>of</strong> the film. Weihnacht <strong>and</strong> Brückner [242] have <strong>in</strong>direct<br />

TEM evidence <strong>of</strong> such pile-ups for <strong>Cu</strong> th<strong>in</strong> films on oxidized Si substrates. Dehm <strong>and</strong><br />

Arzt have supplied somewhat contradictory evidence <strong>in</strong> that they observe dislocations <strong>in</strong>


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 60<br />

<strong>Cu</strong> that are attracted <strong>and</strong> pulled <strong>in</strong>to an amorphous silicon nitride <strong>in</strong>terface [47]. This only<br />

happens after a some time under the electron beam, however, <strong>and</strong> may be due to core<br />

spread<strong>in</strong>g caused by irradiation-<strong>in</strong>duced diffusion [72]. For this discussion, I assume the<br />

rigid substrate is an effective barrier to dislocation propagation.<br />

The presence <strong>of</strong> a dislocation barrier creates gradients <strong>of</strong> plastic stra<strong>in</strong> s<strong>in</strong>ce dislocations<br />

pile-up at the <strong>in</strong>terface <strong>and</strong> are emitted from the surface. As discussed <strong>in</strong> Chapter 1, plastic<br />

stra<strong>in</strong> gradients imply the presence <strong>of</strong> geometrically necessary dislocations (GNDs) for<br />

compatibility.<br />

Ronay [191] was the first to use stra<strong>in</strong> gradient plasticity to expla<strong>in</strong> the large th<strong>in</strong> film<br />

flow strength. Her model will be discussed <strong>in</strong> some detail because it is easy to compare her<br />

SGP formulation to the discrete dislocation models; <strong>in</strong> fact, her formulation is ak<strong>in</strong> to us<strong>in</strong>g<br />

SGP to derive similar relationships. She assumed that the flow strength could be expressed<br />

as the superposition <strong>of</strong> statistically stored dislocations (SSDs), <strong>and</strong> long-range (LR) <strong>and</strong><br />

short-range (SR) <strong>in</strong>teraction GNDs:<br />

σflow = σSSD + σLR, GND + σSR, GND. (3.15)<br />

For the long-range <strong>in</strong>teractions, she based her results on Ashby’s treatment <strong>of</strong> rigid plate-<br />

like particles <strong>in</strong> a s<strong>of</strong>ter matrix [7]. If a shear stra<strong>in</strong> ɛsh is imposed on the film, the amount<br />

<strong>of</strong> shear at the surface is equal to ɛsh, while the shear at the <strong>in</strong>terface is 0. Thus, the<br />

average gradient on a s<strong>in</strong>gle slip system is ɛshζ/t if the gra<strong>in</strong> extends from the surface to<br />

the <strong>in</strong>terface. Here, ζ is a geometric factor that depends on the gra<strong>in</strong> orientation <strong>and</strong> the<br />

slip plane (= s<strong>in</strong> 70.5 ◦ for FCC materials). Thus, accord<strong>in</strong>g to equation (1.13) the density


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 61<br />

<strong>of</strong> GNDs is given by<br />

ρG = ɛsh<br />

. (3.16)<br />

bζ tf<br />

Us<strong>in</strong>g the Taylor relation (3.14), with kT = 1/π (the estimate for screw dislocations near<br />

an <strong>in</strong>terface),<br />

σLR, GND = 1<br />

π Gf<br />

� bɛsh<br />

Ronay then considered gra<strong>in</strong> boundary strengthen<strong>in</strong>g <strong>and</strong> found<br />

ζ tf<br />

� 1/2<br />

. (3.17)<br />

ρG = 2ɛsh<br />

. (3.18)<br />

bd<br />

This is the two-dimensional counterpart to Ashby’s three-dimensional result (4ɛsh/bd).<br />

S<strong>in</strong>ce the deformation <strong>of</strong> a gra<strong>in</strong> bonded to a substrate requires multiple slip for compati-<br />

bility, this is a short-range <strong>in</strong>teraction. As described above, the Taylor relation can be used<br />

with kT = 1:<br />

σSR, GND = Gf<br />

� 2bɛsh<br />

This is for a s<strong>in</strong>gle gra<strong>in</strong>. The stress for a polycrystal is<br />

σSR, GND = mGf<br />

where m is the Taylor orientation factor.<br />

Thus, the flow strength accord<strong>in</strong>g to Ronay is<br />

σRonay = σSSD + 1<br />

π Gf<br />

� bɛsh<br />

ζ tf<br />

d<br />

� 2bɛsh<br />

� 1/2<br />

d<br />

� 1/2<br />

. (3.19)<br />

� 1/2<br />

+ mGf<br />

� 2bɛsh<br />

d<br />

(3.20)<br />

� 1/2<br />

. (3.21)<br />

This differs from the above models <strong>in</strong> that it does not have the <strong>in</strong>verse film thickness depen-<br />

dence. As Ronay mentions, both the second <strong>and</strong> third terms on the RHS <strong>of</strong> this equation<br />

arise from the rigid substrate constra<strong>in</strong>t.


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 62<br />

Hutch<strong>in</strong>son [109] has recently used the cont<strong>in</strong>uum higher order stra<strong>in</strong> gradient plasticity<br />

theory (SGP) to derive the <strong>in</strong>verse thickness dependence <strong>of</strong> the flow stress. His derivation<br />

explicitly dist<strong>in</strong>guishes between the flow stress <strong>and</strong> the yield strength <strong>in</strong> a material. In the<br />

case <strong>of</strong> a th<strong>in</strong> film bonded to a rigid substrate there are only stretch gradients, with an as-<br />

sociated length scale, ℓSG, <strong>and</strong> no rotation gradients. Fleck <strong>and</strong> Hutch<strong>in</strong>son’s formulation<br />

[61] also considered SSDs as well as GNDs, which enabled Hutch<strong>in</strong>son to determ<strong>in</strong>e the<br />

strength as a function <strong>of</strong> plastic stra<strong>in</strong>, degree <strong>of</strong> plastic constra<strong>in</strong>t, <strong>and</strong> length parameter,<br />

ℓSG. He found that the flow stress can rise to over three times the yield strength at approxi-<br />

mately 1% stra<strong>in</strong> for a film 2 × ℓSG thick. A major dist<strong>in</strong>guish<strong>in</strong>g feature <strong>of</strong> this theory is<br />

the prediction <strong>of</strong> a dislocation boundary layer at the <strong>in</strong>terface, with a thickness on the order<br />

<strong>of</strong> the average distance <strong>of</strong> dislocation travel.<br />

There are several h<strong>in</strong>ts <strong>in</strong> the literature that stra<strong>in</strong> gradients exist <strong>in</strong> th<strong>in</strong> films, but no<br />

boundary layer has ever been observed. Murakami demonstrated the existence <strong>of</strong> stra<strong>in</strong><br />

gradients <strong>in</strong> planes parallel to the surface for 1 µm thick Pb/Si(111) films cooled to 77 K<br />

[160]. The stra<strong>in</strong> pr<strong>of</strong>ile was determ<strong>in</strong>ed by analyz<strong>in</strong>g the asymmetry <strong>of</strong> the (333) peak<br />

(taken <strong>in</strong> the powder diffraction geometry) with Houska’s method [103]. His measurement<br />

was performed <strong>in</strong> situ, with the sample mounted on a liquid helium cooled diffraction stage.<br />

The data were fit with the empirical equation<br />

ɛ⊥(z) = ɛ<strong>in</strong>t exp � −(tf − z)/z ∗�<br />

(3.22)<br />

where ɛ⊥ is the elastic stra<strong>in</strong> along the surface normal, ɛ<strong>in</strong>t is the stra<strong>in</strong> at the <strong>in</strong>terface, <strong>and</strong><br />

z ∗ is a characteristic length. Murakami found a stra<strong>in</strong> <strong>of</strong> 0.18% at the <strong>in</strong>terface <strong>and</strong> 0.04%<br />

at the surface. The ma<strong>in</strong> problem with this measurement is that the stra<strong>in</strong>s parallel to the<br />

<strong>in</strong>terface (the biaxial stra<strong>in</strong>s, ɛ�) cannot be measured directly. Doerner <strong>and</strong> Brennan [52]


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 63<br />

have also observed stra<strong>in</strong> gradients <strong>in</strong> annealed Al th<strong>in</strong> films (260 nm <strong>and</strong> 600 nm thick) on<br />

Si(100) us<strong>in</strong>g graz<strong>in</strong>g <strong>in</strong>cidence diffraction (GID), which does allow a direct measurement<br />

<strong>of</strong> ɛ�. The stra<strong>in</strong> rose from near zero at the surface, <strong>and</strong> reached a plateau 5 nm <strong>in</strong>to the<br />

260 nm thick film <strong>and</strong> 200 nm <strong>in</strong>to the 600 nm thick film.<br />

Most recently, Baker et al. [10] have found that for <strong>Cu</strong> th<strong>in</strong> films bonded to Si sub-<br />

strates, the dislocation densities are temperature-dependent <strong>and</strong> stra<strong>in</strong>-<strong>in</strong>dependent. This<br />

accurately describes the behavior <strong>of</strong> geometrically necessary dislocations [191].<br />

3.2 Determ<strong>in</strong><strong>in</strong>g the Stra<strong>in</strong> Pr<strong>of</strong>ile<br />

The complete plastic stra<strong>in</strong> pr<strong>of</strong>ile has never been determ<strong>in</strong>ed experimentally. If the<br />

exact distribution can be measured, <strong>and</strong> <strong>in</strong> particular, the presence <strong>of</strong> a boundary layer<br />

near the substrate <strong>in</strong>terface proven, this would allow for an estimate <strong>of</strong> ℓSG <strong>and</strong> a better<br />

underst<strong>and</strong><strong>in</strong>g <strong>of</strong> the configuration <strong>of</strong> GNDs. This is the goal <strong>of</strong> our experiment.<br />

For a thermally cycled film, the total biaxial stra<strong>in</strong>, ɛ total<br />

� , is approximately known s<strong>in</strong>ce<br />

it can be calculated from �α�T , where �α is the difference <strong>in</strong> CTE between the film <strong>and</strong><br />

substrate (discussed <strong>in</strong> Chapter 2). S<strong>in</strong>ce x-ray diffraction can measure the elastic stra<strong>in</strong> as<br />

a function <strong>of</strong> penetration depth, ɛ�(z), the plastic stra<strong>in</strong> pr<strong>of</strong>ile can be determ<strong>in</strong>ed us<strong>in</strong>g the<br />

relationship<br />

ɛ pl<br />

� (z) = ɛtotal � − ɛ�(z). (3.23)<br />

Although our goal is to determ<strong>in</strong>e ɛ pl<br />

� (z), I will more <strong>of</strong>ten refer to ɛ�(z) s<strong>in</strong>ce this is<br />

the measured parameter. Complicat<strong>in</strong>g matters is that our <strong>in</strong>terest is specifically <strong>in</strong> the<br />

plastic stra<strong>in</strong> from dislocation flow; where this beg<strong>in</strong>s on the stress-temperature pr<strong>of</strong>ile


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 64<br />

is unclear. Another complication is that it is possible for stra<strong>in</strong> gradients to arise from<br />

diffusive processes rather than dislocation processes [73] as well as from gra<strong>in</strong> boundary<br />

slid<strong>in</strong>g [17].<br />

One previous attempt at measur<strong>in</strong>g stra<strong>in</strong> gradients due to SGP was made by Leung<br />

et al. [141]. They studied thermal stra<strong>in</strong>s <strong>in</strong> gold th<strong>in</strong> films <strong>of</strong> various thickness us<strong>in</strong>g the<br />

scatter<strong>in</strong>g vector method described by Genzel [79, 80, 77]. An advantage <strong>of</strong> the scatter<strong>in</strong>g<br />

vector method is that one can measure the stra<strong>in</strong> as a function <strong>of</strong> penetration depth for a set<br />

<strong>of</strong> gra<strong>in</strong>s with the same orientation with respect to the surface normal (gra<strong>in</strong> population).<br />

Leung et al. did not f<strong>in</strong>d stra<strong>in</strong> gradients, but their data conta<strong>in</strong>ed large errors caused by<br />

reliance on a low diffraction angle; <strong>in</strong> addition, no error bars were <strong>in</strong>cluded, so small stra<strong>in</strong><br />

gradients could not be discerned. Furthermore, the stra<strong>in</strong>s measured were not parallel to<br />

the surface.<br />

We attempted to probe the elastic stra<strong>in</strong> distribution, ɛ�(z), us<strong>in</strong>g the graz<strong>in</strong>g <strong>in</strong>cidence<br />

diffraction method [52]. This allows an <strong>in</strong>vestigation <strong>of</strong> the surface stra<strong>in</strong>s when the x-rays<br />

enter the sample from the surface side <strong>and</strong> an <strong>in</strong>vestigation <strong>of</strong> the <strong>in</strong>terfacial stra<strong>in</strong>s when<br />

the enter the sample from the substrate side.<br />

3.3 Experiment<br />

3.3.1 Film Preparation<br />

Most <strong>of</strong> the <strong>Cu</strong> films were grown on oxidized Si(100) substrates approximately 400 µm<br />

thick. For better sensitivity with the curvature apparatus, the th<strong>in</strong>ner films were deposited<br />

onto 97 µm thick bare Si(100) substrates. Before deposition, the basel<strong>in</strong>e curvatures <strong>of</strong> the


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 65<br />

substrates were measured.<br />

Sputtered <strong>Cu</strong> films (from a 99.99 at.% <strong>Cu</strong> target) <strong>of</strong> thicknesses 0.22 µm, 0.5 µm,<br />

0.87 µm, 1.25 µm, <strong>and</strong> 1.73 µm were deposited with an Ar pressure <strong>of</strong> 1.1 × 10 −2 Pa<br />

after reach<strong>in</strong>g a base pressure <strong>of</strong> 1.3 × 10 −5 Pa. The sputter<strong>in</strong>g chamber is described by<br />

Spaepen et al. [213] <strong>and</strong> shown <strong>in</strong> Figure 1.7. The sputter<strong>in</strong>g parameters are given <strong>in</strong> Table<br />

1.2.<br />

For the films deposited onto bare Si(100) substrates, a th<strong>in</strong> amorphous silicon nitride<br />

(a-SiNx) layer, nom<strong>in</strong>ally 40 nm thick, was put down before deposit<strong>in</strong>g the <strong>Cu</strong> to act as<br />

a diffusion barrier between the film <strong>and</strong> the substrate. This layer was also deposited on<br />

the backside <strong>of</strong> the substrate after the copper deposition. The sputter<strong>in</strong>g parameters for<br />

the nitride deposition are <strong>in</strong> Table 2.2. Prior to deposition, the substrates were sputter-<br />

cleaned for 2-4 m<strong>in</strong>utes. The substrates were also water-cooled from the backside, lead<strong>in</strong>g<br />

to temperature rise <strong>of</strong> 10 ◦ C after three hours <strong>of</strong> deposition. A typical deposition rate was<br />

0.15 nm/s.<br />

3.3.2 Film Characterization<br />

After deposition, the thicknesses <strong>of</strong> the films were measured with a Tencor Alpha-Step<br />

200 Pr<strong>of</strong>ilometer. For the th<strong>in</strong>ner films (less than one micrometer thick), the thickness was<br />

also measured with Rutherford Backscatter<strong>in</strong>g Spectrometry (RBS).<br />

Microstructural characterization was performed by plan-view <strong>and</strong> cross sectional trans-<br />

mission electron microscopy (TEM). The film purity was checked by RBS <strong>and</strong> energy<br />

dispersive spectroscopy (EDS) <strong>in</strong> the TEM.


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 66<br />

3.3.3 <strong>Cu</strong>rvature Measurements<br />

The wafer curvature apparatus was described earlier <strong>in</strong> Chapter 1, <strong>and</strong> the curvatures<br />

were converted to stresses us<strong>in</strong>g the Stoney equation (1.16). The thermomechanical proper-<br />

ties <strong>of</strong> the Si(100) substrates are listed <strong>in</strong> Table 2.3. Thermal cycles to 660 ◦ C were carried<br />

out for one film <strong>of</strong> each thickness at rates <strong>of</strong> 5 ◦ C/m<strong>in</strong>. Another set <strong>of</strong> films <strong>of</strong> the same<br />

thicknesses was left <strong>in</strong> the as-deposited state.<br />

3.3.4 X-Ray Diffraction<br />

The GID measurements were carried out at Beaml<strong>in</strong>e X22C at the National Synchrotron<br />

Light Source at Brookhaven National Laboratory. The GID geometry is shown <strong>in</strong> Figure<br />

1.6 (c). A0.1 ◦ soller slit was placed before the detector to ensure a small diffraction peak<br />

width. The energy <strong>of</strong> the <strong>in</strong>com<strong>in</strong>g photons was tuned to 8960 eV, which is just below the<br />

<strong>Cu</strong> K edge. For studies <strong>in</strong> which the x-rays entered through the back <strong>of</strong> the substrate, the<br />

third order reflection <strong>of</strong> the Si(111) monochromator was used so that the <strong>in</strong>com<strong>in</strong>g energy<br />

was 26.9 keV. The energy w<strong>in</strong>dow <strong>of</strong> the sc<strong>in</strong>tillation detector was adjusted accord<strong>in</strong>gly.<br />

In a GID measurement, the diffraction angle 2θ does not lie <strong>in</strong> a plane parallel to the<br />

<strong>in</strong>com<strong>in</strong>g beam so a correction must be made to convert the diffractometer-specified 2θ ′ to<br />

the true 2θ. This geometry is displayed <strong>in</strong> Figure 3.3, <strong>and</strong> the conversion is given by<br />

cos 2θ = cos 2α cos 2θ ′<br />

(3.24)<br />

where α is the <strong>in</strong>com<strong>in</strong>g angle <strong>and</strong> 2θ ′ is the detector motor position. The correction was<br />

checked us<strong>in</strong>g a CeO2 powder st<strong>and</strong>ard from NIST (SRM 674a).


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 67<br />

2θ<br />

2θ<br />

2α<br />

Figure 3.3: The scatter<strong>in</strong>g geometry for the GID measurements. As shown <strong>in</strong> (a), the 2θ<br />

value is not the same as the motor position for the detector, 2θ ′ . The correction can be<br />

made us<strong>in</strong>g the well known formulas for the right-spherical triangle shown <strong>in</strong> (b).<br />

where<br />

The (1/e) penetration depth, τ, can be calculated us<strong>in</strong>g the follow<strong>in</strong>g equation [52]:<br />

2q 2 =<br />

τ = λ<br />

4πq<br />

2θ<br />

2α<br />

2θ<br />

(3.25)<br />

�<br />

2δ − α 2�<br />

�� + α 2 �2 − 2δ + 4β 2<br />

�1/2 . (3.26)<br />

Here, λ is the x-ray wavelength, q is the scatter<strong>in</strong>g vector, α is the <strong>in</strong>cidence angle, <strong>and</strong> δ<br />

<strong>and</strong> β are associated with the real <strong>and</strong> imag<strong>in</strong>ary parts <strong>of</strong> the <strong>in</strong>dex <strong>of</strong> refraction, n, i.e.,<br />

n = 1 − δ − iβ (3.27)<br />

<strong>and</strong> may be looked up <strong>in</strong> any x-ray reference table (e.g., www-cxro.lbl.gov).<br />

Follow<strong>in</strong>g Leung et al. [141], we only searched for stra<strong>in</strong> gradients <strong>in</strong> the three thickest<br />

films, (0.87 µm, 1.25 µm, <strong>and</strong> 1.73 µm) s<strong>in</strong>ce a stra<strong>in</strong> gradient from diffusional flow is ex-<br />

pected for the th<strong>in</strong>ner films. The stra<strong>in</strong> pr<strong>of</strong>iles <strong>in</strong> the thermally cycled films were compared


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 68<br />

to those <strong>in</strong> the as-deposited films. The measurements were made us<strong>in</strong>g the (220), (200),<br />

<strong>and</strong> (111) diffraction peaks with the <strong>in</strong>cidence angle vary<strong>in</strong>g from 0.3 ◦ to 4.3 ◦ . S<strong>in</strong>ce the<br />

texture poles shown <strong>in</strong> Chapter 2 are broad, a tilt <strong>of</strong> 4.3 ◦ does not significantly vary the<br />

sampled gra<strong>in</strong> population.<br />

For two <strong>of</strong> the samples (the as-deposited <strong>and</strong> annealed 1.25 µm thick <strong>Cu</strong>), the stra<strong>in</strong>s<br />

measured by the synchrotron were checked with s<strong>in</strong> 2 ψ measurements taken on the diffrac-<br />

tometer at <strong>Harvard</strong> University (Figure 1.5). The Kamm<strong>in</strong>ga method <strong>and</strong> the crystallite<br />

group method (Appendix A) were used for the as-deposited <strong>and</strong> annealed films, respec-<br />

tively.<br />

3.4 Results<br />

3.4.1 Microstructure<br />

Cross-sectional views <strong>of</strong> an as-deposited <strong>and</strong> an annealed 0.87 µm thick copper film are<br />

shown <strong>in</strong> Figure 3.4. The plan-view micrographs were presented <strong>in</strong> the results section <strong>of</strong><br />

Chapter 2. The as-deposited films appear to have a th<strong>in</strong> columnar morphology with an <strong>in</strong>-<br />

plane gra<strong>in</strong> size <strong>of</strong> about 50 nm, while the annealed films have graded morphology with a<br />

gra<strong>in</strong> size <strong>of</strong> about 500 nm near the surface <strong>of</strong> the film <strong>and</strong> a bimodal gra<strong>in</strong> distribution near<br />

the substrate-side <strong>of</strong> the film, with an average gra<strong>in</strong> size <strong>of</strong> 200 nm. Even after anneal<strong>in</strong>g,<br />

the gra<strong>in</strong>s do not extend from the surface to the substrate.<br />

copper.<br />

As discussed <strong>in</strong> Chapter 2, both RBS <strong>and</strong> EDS show no impurities <strong>in</strong> the bulk <strong>of</strong> the


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 69<br />

Figure 3.4: Cross-sectional views <strong>of</strong> an as-deposited (a) <strong>and</strong> annealed (b) 0.87 µm thick<br />

copper film.<br />

(a)<br />

(b)


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 70<br />

Figure 3.5: Deposition stress for ion beam sputtered <strong>Cu</strong> films as a function <strong>of</strong> total film<br />

thickness.<br />

3.4.2 <strong>Cu</strong>rvature Results<br />

The deposition stresses for films <strong>of</strong> various thickness are shown <strong>in</strong> Figure 3.5. A caveat<br />

should be mentioned here: the films<strong>of</strong>0.22 µm <strong>and</strong> 0.50 µm thickness have a silicon<br />

nitride layer on the backside <strong>of</strong> the substrate to cancel out the curvature change due to the<br />

silicon nitride layer <strong>in</strong> between the substrate <strong>and</strong> film. If the nitride thicknesses on either<br />

side <strong>of</strong> the Si substrate are not equal, this could result <strong>in</strong> an error <strong>in</strong> the deposition stress<br />

measurement. The thickness error is expected to be small, however.<br />

Figure 3.6 shows the room temperature flow strength after thermal cycl<strong>in</strong>g for the cop-<br />

per films. The <strong>in</strong>verse thickness relationship predicted by σconstra<strong>in</strong>t <strong>and</strong> the higher order<br />

SGP theory appears to hold.<br />

The three thermally cycled samples tested for stra<strong>in</strong> gradients have the stress-temperature<br />

pr<strong>of</strong>iles displayed <strong>in</strong> Figure 3.7.


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 71<br />

Figure 3.6: Room temperature flow strength as a function <strong>of</strong> <strong>in</strong>verse thickness for <strong>Cu</strong> samples<br />

thermally cycled to 660 ◦ C.<br />

Figure 3.7: The stress-temperature pr<strong>of</strong>iles for the <strong>Cu</strong> films used <strong>in</strong> the stra<strong>in</strong> gradient<br />

studies.


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 72<br />

d hkl (h 2 +k 2 +l 2 ) 1/2 [Å]<br />

5.420<br />

5.415<br />

5.410<br />

5.405<br />

5.400<br />

0<br />

1<br />

2<br />

α [degrees]<br />

Figure 3.8: Lattice parameter <strong>of</strong> CeO2 as a function <strong>of</strong> <strong>in</strong>cidence angle, α. The lattice<br />

parameters were measured with the (331) reflection.<br />

3.4.3 X-Ray Diffraction Results<br />

As mentioned earlier, a st<strong>and</strong>ard reference material, CeO2, was used to calibrate the<br />

<strong>in</strong>strument <strong>and</strong> check the 2θ correction (3.24). Figure 3.8 shows lattice parameter <strong>of</strong> CeO2<br />

as a function <strong>of</strong> <strong>in</strong>cidence angle after the correction is applied. The measurement was<br />

carried out us<strong>in</strong>g the (331) reflection. The large error bars stem from poor <strong>in</strong>tensities<br />

caused by nonuniform spread<strong>in</strong>g <strong>of</strong> the CeO2 powder on a substrate. That the powder<br />

shows no stra<strong>in</strong> gradients gives us confidence <strong>in</strong> our measurements.<br />

Measurements from the Surface <strong>of</strong> the Film<br />

The stra<strong>in</strong> distributions probed by x-rays enter<strong>in</strong>g the surface <strong>of</strong> the <strong>Cu</strong> are shown <strong>in</strong><br />

Figure 3.9. The solid <strong>and</strong> dotted curves refer to the thermally cycled films <strong>and</strong> as-deposited<br />

films, respectively, <strong>and</strong> the reflection plane used is <strong>in</strong>dicated <strong>in</strong> the legend. The vertical<br />

3<br />

4


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 73<br />

Figure 3.9: Elastic stra<strong>in</strong> gradients <strong>in</strong> <strong>Cu</strong> films. The solid <strong>and</strong> dotted curves refer to the<br />

thermally cycled films <strong>and</strong> as-deposited films, respectively. The reflection is denoted by<br />

different markers, as shown <strong>in</strong> the legend. The plastic stra<strong>in</strong> is given by 0.9% m<strong>in</strong>us the<br />

elastic stra<strong>in</strong>.<br />

dashed l<strong>in</strong>es <strong>in</strong>dicate the po<strong>in</strong>ts at which ∼ 63% <strong>of</strong> the diffracted <strong>in</strong>tensity comes from<br />

the total thickness <strong>of</strong> the film. To the right <strong>of</strong> that l<strong>in</strong>e, more <strong>of</strong> the diffracted <strong>in</strong>tensity<br />

orig<strong>in</strong>ates from the film/substrate <strong>in</strong>terface. If we take the maximum anneal<strong>in</strong>g temperature<br />

(660 ◦ C) for calculat<strong>in</strong>g the thermal stra<strong>in</strong>, the total stra<strong>in</strong> <strong>in</strong> all the films is ≈ 0.9%, <strong>and</strong><br />

the plastic stra<strong>in</strong>s can be determ<strong>in</strong>ed with equation (3.23).<br />

Measurements from the Film/Substrate Interface<br />

Despite the high brilliance <strong>of</strong> the synchrotron, we were unable to obta<strong>in</strong> a diffraction<br />

peak from the <strong>Cu</strong> when the x-rays entered through the back <strong>of</strong> the substrate. After scatter<strong>in</strong>g<br />

from the <strong>Cu</strong>, nearly all the photons were absorbed by the Si.


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 74<br />

Figure 3.10: s<strong>in</strong> 2 ψ-plot for an as-deposited <strong>Cu</strong> film.<br />

Comparison with s<strong>in</strong>2ψ 2ψ 2ψ Measurements<br />

Figures 3.10 <strong>and</strong> 3.11 show the s<strong>in</strong> 2 ψ plots for the as-deposited <strong>and</strong> thermally cycled<br />

1.25 µm thick films, respectively. The reference lattice parameter was taken to be the lit-<br />

erature value, aref = a0 = 0.3615 nm [115]. This agreed well with stra<strong>in</strong>-free lattice<br />

parameter determ<strong>in</strong>ed by s<strong>in</strong> 2 ψ analysis (Appendix A). The stra<strong>in</strong>s agree with those from<br />

the GID experiments, which gives us confidence <strong>in</strong> the absolute value <strong>of</strong> the lattice parame-<br />

ters. There is a significant difference between the stresses <strong>in</strong> the 〈111〉 gra<strong>in</strong>s <strong>and</strong> the 〈100〉<br />

gra<strong>in</strong>s for the thermally cycled film. The stresses <strong>in</strong> the r<strong>and</strong>omly oriented gra<strong>in</strong>s fell along<br />

the 〈111〉 s<strong>in</strong> 2 ψ l<strong>in</strong>e.<br />

ψ<br />

σ


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 75<br />

3.5 Discussion<br />

σ<br />

Figure 3.11: s<strong>in</strong> 2 ψ-plot for a thermally cycled <strong>Cu</strong> film.<br />

σ<br />

The cross-sectional TEM micrographs show gra<strong>in</strong> boundaries <strong>and</strong> tw<strong>in</strong>s parallel to the<br />

surface, which makes it difficult to analyze the stra<strong>in</strong> distribution <strong>in</strong> terms <strong>of</strong> an <strong>in</strong>homoge-<br />

neous dislocation distribution. Nevertheless, the films exhibit extremely high flow stresses,<br />

as depicted <strong>in</strong> Figure 3.7. If the large strength is primarily due to stra<strong>in</strong> gradient plasticity,<br />

the stra<strong>in</strong> gradients should be measurable even though the nature <strong>of</strong> the boundary layer may<br />

be different.<br />

Accord<strong>in</strong>g to the texture results <strong>in</strong> Chapter 2, the texture <strong>of</strong> the thermally cycled films is<br />

predom<strong>in</strong>antly 〈111〉; thus, the stress <strong>in</strong> the 〈111〉-oriented gra<strong>in</strong>s, 489 MPa (Figure 3.11),<br />

is taken to be the overall biaxial stress, <strong>and</strong> the stress with<strong>in</strong> the 〈100〉 gra<strong>in</strong>s is ignored.<br />

With this value for the annealed film <strong>and</strong> −124 MPa for the as-deposited film, the x-ray<br />

measurements underestimate the stress measured by substrate curvature <strong>in</strong> Figure 3.7 by<br />

ψ


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 76<br />

about 15-20%, which is <strong>in</strong>significant given the errors <strong>in</strong> both measurements. The s<strong>in</strong> 2 ψ<br />

results films are <strong>in</strong> agreement with the synchrotron measured stra<strong>in</strong>s, however. The large<br />

difference <strong>in</strong> stress between the 〈111〉 <strong>and</strong> 〈100〉 oriented gra<strong>in</strong>s is consistent with the re-<br />

sults <strong>of</strong> Baker et al. [10].<br />

Due to absorption, the biaxial stra<strong>in</strong> as a function <strong>of</strong> (1/e) penetration depth, ɛ�(τ), is<br />

related to the stra<strong>in</strong> as a function <strong>of</strong> position, ɛ�(z),byaf<strong>in</strong>ite thickness Laplace transform<br />

[13]:<br />

ɛ�(τ) =<br />

� tf<br />

0 dzɛ�(z) exp (−z/τ)<br />

. (3.28)<br />

dz exp (−z/τ)<br />

� tf<br />

0<br />

Figure 3.12 shows a hypothetical stra<strong>in</strong> pr<strong>of</strong>ile that demonstrates relaxation at the surface<br />

<strong>and</strong> a boundary layer near the <strong>in</strong>terface (the uppermost solid curve). The f<strong>in</strong>ite thickness<br />

Laplace transform <strong>of</strong> such a pr<strong>of</strong>ile is shown as the closed circles <strong>in</strong> the figure. The shape <strong>of</strong><br />

the f<strong>in</strong>ite thickness Laplace transform is <strong>in</strong>variably the same no matter what the magnitude<br />

<strong>of</strong> the boundary layer (e.g., from 0 to a full 0.9% stra<strong>in</strong> at the <strong>in</strong>terface). Only the surface<br />

relaxation can be detected from ɛ�(τ). Although <strong>in</strong> theory ɛ�(z) can be retrieved us<strong>in</strong>g the<br />

<strong>in</strong>verse Laplace transform, the <strong>in</strong>verse problem is nearly impossible given the error bars <strong>in</strong><br />

the data; even with <strong>in</strong>f<strong>in</strong>itesimal error, the unstable nature <strong>of</strong> the <strong>in</strong>verse problem would<br />

makeitdifficult to identify a boundary layer.<br />

This emphasizes the necessity <strong>of</strong> mak<strong>in</strong>g lattice parameter measurements from the back<br />

<strong>of</strong> the substrate. This is a difficult technique that has not even been attempted before accord-<br />

<strong>in</strong>g to the literature (<strong>and</strong> now we know why). The ma<strong>in</strong> problem is that the measurement<br />

is performed at graz<strong>in</strong>g angles; this requires the x-rays to enter <strong>and</strong> scattered x-rays to<br />

exit through the entire width <strong>of</strong> the substrate (≈ 0.25 <strong>in</strong>). The closest measurement to this<br />

that has been successful is a reflectivity measurement from an <strong>in</strong>terface, which researchers


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 77<br />

Biaxial Elastic Stra<strong>in</strong> [%]<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

-0.2<br />

0.0<br />

0.5<br />

ε<br />

ε τ<br />

1.0<br />

Depth [µm]<br />

Figure 3.12: Hypothetical stra<strong>in</strong> pr<strong>of</strong>iles <strong>and</strong> their normalized f<strong>in</strong>ite thickness Laplace<br />

transforms for a 1.8 µm thick <strong>Cu</strong> film.<br />

performed to determ<strong>in</strong>e the nature <strong>of</strong> the solid/liquid <strong>in</strong>terface for liquid Pb on Si [189].<br />

Thus, for now, I will focus only on the surface stra<strong>in</strong> behavior <strong>in</strong> Figure 3.9.<br />

3.5.1 The As-Deposited <strong>Films</strong><br />

Intr<strong>in</strong>sic stresses <strong>in</strong> as-deposited films have been the subject <strong>of</strong> several theses ([204,<br />

186, 49]) <strong>and</strong> reviews ([54, 254, 129, 212]). Only the relevant po<strong>in</strong>ts are discussed here.<br />

Tensile stresses <strong>in</strong> as-deposited films orig<strong>in</strong>ate from excess volume. For the microstruc-<br />

ture seen <strong>in</strong> Figure 3.4, one can see the ma<strong>in</strong> source <strong>of</strong> excess volume: gra<strong>in</strong> boundaries. In<br />

addition, there is a contribution from vacancy annihilation. Tensile stresses are very com-<br />

mon <strong>in</strong> high-vacuum deposited films, particularly those grown by evaporation. The stresses<br />

are <strong>of</strong>ten less tensile <strong>in</strong> UHV-grown films, however, presumably due to enhanced adatom<br />

mobility.<br />

ε τ<br />

ε<br />

1.5


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 78<br />

The usual explanation for compressive stresses <strong>in</strong> ion beam sputtered films is that neu-<br />

tral Ar atoms are back-reflected from the target toward the grow<strong>in</strong>g film [50, 51, 253]. The<br />

energies <strong>of</strong> these back-reflected atoms are <strong>in</strong> the hundreds <strong>of</strong> eV range, while the energies<br />

<strong>of</strong> the deposited atoms are <strong>in</strong> the 10 − 40 eV range. If the Ar gas pressure is low enough,<br />

the back-reflected atoms can arrive at the film with little energy loss, <strong>and</strong> the momentum<br />

transfer can produce a compressive component on top <strong>of</strong> the tensile component.<br />

The Stra<strong>in</strong> Pr<strong>of</strong>ile<br />

The surfaces <strong>of</strong> the as-deposited films are <strong>in</strong> tension (on the order <strong>of</strong> 200 MPa) although<br />

the average stress <strong>in</strong> the film is compressive (on the order <strong>of</strong> −150 MPa). Although the<br />

compressive stress component is not obvious from the ɛ�(τ) pr<strong>of</strong>ile <strong>in</strong> Figure 3.9, the hy-<br />

pothetical ɛ�(z) pr<strong>of</strong>ile (the lower solid l<strong>in</strong>e <strong>in</strong> Figure 3.12) shows how it can be obscured.<br />

It is not currently understood how this pr<strong>of</strong>ile develops, <strong>and</strong> further work is required.<br />

Anisotropy<br />

From our results <strong>in</strong> Chapter 2, we know the biaxial modulus varies with orientation <strong>in</strong><br />

the follow<strong>in</strong>g way:<br />

Y〈111〉 > Yr<strong>and</strong>om > Y〈100〉. (3.29)<br />

Thus, if the stress state is the same <strong>in</strong> all gra<strong>in</strong>s (the isostress limit [10]), then<br />

ɛ〈100〉 >ɛr<strong>and</strong>om >ɛ〈111〉. (3.30)<br />

Dur<strong>in</strong>g a GID measurement, the scatter<strong>in</strong>g vector is nearly parallel to the plane <strong>of</strong> the film.<br />

The (220) reflection conta<strong>in</strong>s the 〈111〉 directions, the (200) reflection conta<strong>in</strong>s the 〈100〉


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 79<br />

directions, <strong>and</strong> the (111) reflection conta<strong>in</strong>s the 〈110〉 directions. Accord<strong>in</strong>g to our texture<br />

results <strong>in</strong> Chapter 2, the as-deposited films are predom<strong>in</strong>antly 〈111〉 textured but also have<br />

〈100〉 <strong>and</strong> r<strong>and</strong>om components. Thus, the (220) stra<strong>in</strong>s represent the 〈111〉 textured gra<strong>in</strong>s,<br />

the (200) stra<strong>in</strong>s represent the 〈100〉 textured gra<strong>in</strong>s, <strong>and</strong> s<strong>in</strong>ce there is no 〈110〉 texture, the<br />

(111) stra<strong>in</strong>s may represent the r<strong>and</strong>omly oriented gra<strong>in</strong>s. Then, accord<strong>in</strong>g to Figure 3.9,<br />

equation (3.30) holds.<br />

3.5.2 The Thermally Cycled <strong>Films</strong><br />

The Stra<strong>in</strong> Pr<strong>of</strong>ile<br />

The thermally cycled films all show strong relaxation at the surface. (This qualitatively<br />

agrees with what happens if the thermally cycled film delam<strong>in</strong>ates: it immediately curls<br />

with negative curvature). The relaxed “layer” is approximately 300 nm thick for all the<br />

thermally cycled films. The cross-sectional micrograph (Figure 3.4 (b)) suggests that sur-<br />

face diffusion allows the top third <strong>of</strong> the films to undergo more 〈111〉 gra<strong>in</strong> growth than<br />

the bottom two-thirds. The surface gra<strong>in</strong>s are most likely 〈111〉-oriented because the 〈111〉<br />

texture <strong>in</strong>creases dur<strong>in</strong>g anneal<strong>in</strong>g while the 〈100〉 texture rema<strong>in</strong>s the same. However,<br />

the larger gra<strong>in</strong>s do not necessarily have less stra<strong>in</strong> than the other gra<strong>in</strong>s (other than that<br />

dictated by equation (3.30) if they are <strong>in</strong>deed 〈111〉 textured). The average film stress is ap-<br />

proximately 0 at 660 ◦ C. It is more likely that the stra<strong>in</strong> relaxation is caused by dislocation<br />

emission through the free surface.<br />

The (111) peak <strong>of</strong> the thermally cycled 1.73 µm thick films was also viewed <strong>in</strong> the pow-<br />

der diffraction geometry. As described above, strong gradients may cause an asymmetry <strong>in</strong><br />

this peak; however, no peak asymmetry <strong>in</strong> the (111) peak was discernible.


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 80<br />

Anisotropy<br />

Accord<strong>in</strong>g the results based on the Schmid factor <strong>in</strong> Appendix D, the 〈111〉 textured<br />

gra<strong>in</strong>s should yield before the 〈100〉 gra<strong>in</strong>s. If we imag<strong>in</strong>e that each gra<strong>in</strong> undergoes the<br />

same plastic stra<strong>in</strong> after the <strong>in</strong>itial yield, the 〈111〉 gra<strong>in</strong>s will have undergone more plastic<br />

stra<strong>in</strong> than 〈100〉 gra<strong>in</strong>s. Consequently, the elastic stra<strong>in</strong> left will be <strong>in</strong> the order<br />

ɛ〈100〉 >ɛ〈111〉. (3.31)<br />

The results <strong>in</strong> Figure 3.9 agree with this analysis. This may not be correct, however.<br />

Accord<strong>in</strong>g to a study by Owusu-Boahen <strong>and</strong> K<strong>in</strong>g [179], at the very onset <strong>of</strong> yield<strong>in</strong>g<br />

<strong>in</strong> th<strong>in</strong> films, dislocation nucleation at the surface takes place on apical slip planes (planes<br />

ly<strong>in</strong>g at the <strong>in</strong>tersection <strong>of</strong> gra<strong>in</strong> boundary triple junctions <strong>and</strong> the free surface) because<br />

a nearly <strong>in</strong>f<strong>in</strong>itesimally short l<strong>in</strong>e segment can be nucleated here. Furthermore, the gra<strong>in</strong><br />

boundaries at the junctions must conta<strong>in</strong> low energy gra<strong>in</strong> boundary dislocations. Conse-<br />

quently, the Schmid factor is not a good predictor <strong>of</strong> which gra<strong>in</strong>s will slip first <strong>in</strong> this very<br />

early yield regime.<br />

A better explanation is that if the large gra<strong>in</strong>s near the surface are <strong>in</strong>deed 〈111〉 textured,<br />

it would be easier for these gra<strong>in</strong>s to undergo plastic relaxation simply because <strong>of</strong> the<br />

proximity to the surface. The dislocations <strong>in</strong> 〈100〉 textured gra<strong>in</strong>s may be constra<strong>in</strong>ed by<br />

the gra<strong>in</strong> boundaries.<br />

The r<strong>and</strong>om texture component seems to have stra<strong>in</strong>s <strong>and</strong> stra<strong>in</strong> pr<strong>of</strong>iles very similar to<br />

those <strong>of</strong> the 〈111〉 gra<strong>in</strong>s. <strong>Cu</strong>rrently, it is not understood why this is.


Chapter 3: The Plastic Properties <strong>of</strong> <strong>Cu</strong> Th<strong>in</strong> <strong>Films</strong> 81<br />

3.6 Conclusions <strong>and</strong> Future Research<br />

Accord<strong>in</strong>g to the SGP theory, the stra<strong>in</strong> pr<strong>of</strong>ile <strong>in</strong> thermally cycled films should exhibit<br />

surface relaxation <strong>and</strong> a boundary layer at the substrate <strong>in</strong>terface. Our GID measurements<br />

show that the surface relaxation <strong>in</strong>deed exists: pro<strong>of</strong> <strong>of</strong> the boundary layer could not be ob-<br />

ta<strong>in</strong>ed. The as-deposited films exhibit large tensile stresses at the surface while the average<br />

stress is compressive.<br />

Stra<strong>in</strong> anisotropy <strong>in</strong> the as-deposited films is best expla<strong>in</strong>ed by the anisotropy <strong>in</strong> the<br />

biaxial modulus. The stra<strong>in</strong> anisotropy <strong>in</strong> the thermally cycled films may be due to a texture<br />

gradient along the surface normal.<br />

The best method for study<strong>in</strong>g the boundary layer would be to perform an <strong>in</strong> situ diffrac-<br />

tion experiment on tensile-tested free-st<strong>and</strong><strong>in</strong>g films with th<strong>in</strong> dislocation barrier layer (Ta<br />

or polyimide?) <strong>and</strong> very large plastic stra<strong>in</strong>s. A GID experiment may be performed through<br />

the th<strong>in</strong> barrier layer us<strong>in</strong>g very high x-ray energies <strong>and</strong> an energy sensitive detector.<br />

Stra<strong>in</strong> gradients should also dom<strong>in</strong>ate the plastic behavior <strong>in</strong> multilayers [234]. These<br />

gradients may be studied us<strong>in</strong>g the methods discussed <strong>in</strong> the next chapter.


Chapter 4<br />

The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress<br />

4.1 The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface<br />

Chapter 3 presented the Matthews-Blakeslee equation (3.9) for the maximum thickness<br />

<strong>of</strong> an epitaxial layer before the onset <strong>of</strong> dislocation nucleation. In much the same way,<br />

we can calculate a critical bilayer period for multilayers <strong>of</strong> equal phase thicknesses; the<br />

critical thickness <strong>of</strong> each layer is about four times greater than that for a film on an <strong>in</strong>f<strong>in</strong>ite<br />

substrate [65, 228]. This <strong>in</strong>troduction provides background on the structure <strong>and</strong> energy<br />

<strong>of</strong> the {111}<strong>Ag</strong>/{111}<strong>Ni</strong> <strong>in</strong>terface (hereafter called the {111} <strong>in</strong>terface), as well as some<br />

requisite thermodynamics for deriv<strong>in</strong>g a relationship between <strong>in</strong>terface energy <strong>and</strong> <strong>in</strong>terface<br />

stress. Subsequently, I will describe the experiment performed to measure the <strong>in</strong>terface<br />

stress <strong>and</strong> discuss its orig<strong>in</strong>.<br />

82


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 83<br />

4.1.1 The Structure <strong>and</strong> Energy <strong>of</strong> Semicoherent Interfaces<br />

Numerous reviews on the structure <strong>and</strong> energy <strong>of</strong> heterophase <strong>in</strong>terfaces exist; an ex-<br />

cellent start<strong>in</strong>g po<strong>in</strong>t is the book by Howe [104], <strong>and</strong> Sutton <strong>and</strong> Balluffi have recently<br />

completed a comprehensive treatise on solid <strong>in</strong>terfaces [214]. I beg<strong>in</strong> by first def<strong>in</strong><strong>in</strong>g the<br />

follow<strong>in</strong>g two parameters. The misfit between two materials α <strong>and</strong> β is given by<br />

<strong>and</strong> the spac<strong>in</strong>g between misfit dislocations is<br />

δ = 2 � �<br />

aα − aβ<br />

, (4.1)<br />

aα + aβ<br />

Dδ = b<br />

, (4.2)<br />

δ<br />

where a is the lattice parameter, <strong>and</strong> b is the Burgers vector <strong>of</strong> the misfit dislocation. In<br />

general, knowledge <strong>of</strong> the misfit, the elastic constants <strong>of</strong> the adjacent materials, <strong>and</strong> the<br />

<strong>in</strong>terface crystallography can provide an estimate <strong>of</strong> the <strong>in</strong>terface structure <strong>and</strong> energy <strong>in</strong><br />

an idealized system. For real systems, many other factors such as substrate surface defects<br />

<strong>and</strong> dislocation k<strong>in</strong>etics become important.<br />

Structure <strong>of</strong> the <strong>Ag</strong>/<strong>Ni</strong> {111} Interface<br />

Silver <strong>and</strong> nickel have negligible solid solubility at room temperature; the phase dia-<br />

gram is reproduced <strong>in</strong> Figure 4.1. They also have a large misfit <strong>of</strong> 14.8%, with a<strong>Ag</strong> =<br />

0.40862 nm <strong>and</strong> a<strong>Ni</strong> = 0.35238 nm [115]. For both these reasons, the structures <strong>of</strong> low-<br />

<strong>in</strong>dex <strong>Ag</strong>/<strong>Ni</strong> <strong>in</strong>terfaces have been widely studied by both cross-sectional TEM [75, 154]<br />

<strong>and</strong> simulation [55, 74, 76, 87]. Accord<strong>in</strong>g to these studies, the {111} <strong>in</strong>terface with 0 ◦<br />

twist has the lowest <strong>in</strong>terface energy. Figure 4.2 shows a view <strong>of</strong> this <strong>in</strong>terface along<br />

the [11¯1] direction. The <strong>Ag</strong> atoms are black <strong>and</strong> the <strong>Ni</strong> atoms are white. Accord<strong>in</strong>g to


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 84<br />

O-lattice theory [19], a {111} <strong>in</strong>terface produces a hexagonal network (three sets <strong>of</strong> dislo-<br />

cations <strong>in</strong>tersect<strong>in</strong>g at 60 ◦ )<strong>of</strong>1/2〈110〉-type Burgers vectors. If we take our top view <strong>of</strong><br />

the <strong>in</strong>terface <strong>and</strong> cut along a (¯110), we obta<strong>in</strong> the cross-sectional view <strong>in</strong> Figure 4.3, which<br />

looks along the [110]. Two types <strong>of</strong> dislocation nodes may be seen: one is associated with<br />

{...CABC}<strong>Ni</strong> / {BC AB ...}<strong>Ag</strong> stack<strong>in</strong>g <strong>and</strong> another with {...CABC}<strong>Ni</strong> / {CABC...}<strong>Ag</strong><br />

stack<strong>in</strong>g [76, 90]. The former is an <strong>in</strong>tr<strong>in</strong>sic stack<strong>in</strong>g fault (ISF), while the latter is a high<br />

energy fault. S<strong>in</strong>ce the energy <strong>of</strong> dislocations scales as b 2 (equation (1.9)), it is energet-<br />

ically favorable for these dislocations to dissociate <strong>in</strong>to 1/6〈211〉 Shockley partials [76].<br />

After dissociation, the misfit dislocations form the triangles seen <strong>in</strong> Figure 4.2; at the cen-<br />

ters <strong>of</strong> the white triangles lie planes with the correct stack<strong>in</strong>g while the <strong>in</strong>tr<strong>in</strong>sic stack<strong>in</strong>g<br />

faults lie at the centers <strong>of</strong> the shaded triangles. The high energy faults lie at the nodes <strong>of</strong><br />

the triangles.<br />

The cross-sectional TEM images showed that the misfit dislocation cores lie <strong>in</strong> the<br />

{111} <strong>in</strong>terface <strong>and</strong> are slightly delocalized [75]. Other <strong>Ag</strong>/<strong>Ni</strong> <strong>in</strong>terfaces, such as the<br />

{110}<strong>Ag</strong>/{110}<strong>Ni</strong> <strong>and</strong> the {100}<strong>Ag</strong>/{100}<strong>Ni</strong> are less closely packed than the {111} <strong>in</strong>terfaces<br />

<strong>and</strong> allow misfit dislocations to move out <strong>of</strong> the <strong>in</strong>terface <strong>and</strong> <strong>in</strong>to the <strong>Ni</strong> one plane deep<br />

[260]. For the {110}<strong>Ag</strong>/{110}<strong>Ni</strong> <strong>in</strong>terface, Gumbsch et al. [89] have found that this expulsion<br />

<strong>of</strong> misfit dislocations can lead to a collection <strong>of</strong> vacancies on the <strong>Ni</strong> side <strong>of</strong> the boundary,<br />

which lowers the <strong>in</strong>terfacial enthalpy. This could potentially happen at the high energy<br />

fault positions <strong>in</strong> the {111} <strong>in</strong>terface, but the energy decrease <strong>in</strong> the removal <strong>of</strong> the fault is<br />

less than the cost <strong>of</strong> vacancy formation [89].


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 85<br />

Figure 4.1: The <strong>Ag</strong>/<strong>Ni</strong> phase diagram [149].


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 86<br />

Figure 4.2: Top view <strong>of</strong> the <strong>Ag</strong>/<strong>Ni</strong> {111} <strong>in</strong>terface, with <strong>Ag</strong> <strong>and</strong> <strong>Ni</strong> denoted by closed <strong>and</strong><br />

open circles, respectively. After relaxation, misfit dislocations are formed, as del<strong>in</strong>eated by<br />

the dark l<strong>in</strong>es. The coherent areas <strong>and</strong> <strong>in</strong>tr<strong>in</strong>sic stack<strong>in</strong>g fault areas lie <strong>in</strong> the centers <strong>of</strong> the<br />

white <strong>and</strong> shaded triangles, respectively.


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 87<br />

Figure 4.3: An edge-on view <strong>of</strong> the <strong>Ag</strong>/<strong>Ni</strong> {111} <strong>in</strong>terface along the [110] direction, with<br />

the <strong>Ag</strong> <strong>and</strong> <strong>Ni</strong> atoms denoted by closed <strong>and</strong> open circles, respectively. The large circles are<br />

atoms <strong>in</strong> the plane <strong>of</strong> the diagram; the small circles are atoms <strong>in</strong> the planes immediately<br />

above <strong>and</strong> below the plane <strong>of</strong> the figure. Two dist<strong>in</strong>ct dislocation nodes can be seen <strong>and</strong><br />

associated with the different stack<strong>in</strong>g conditions. Adapted from Gao <strong>and</strong> Merkle [75].


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 88<br />

Energy <strong>of</strong> the <strong>Ag</strong>/<strong>Ni</strong> {111} Interface<br />

Turnbull suggested that the energy <strong>of</strong> a semicoherent <strong>in</strong>terface may be calculated as the<br />

sum <strong>of</strong> chemical <strong>and</strong> structural effects for small misfit [226]:<br />

γ = γc + γs<br />

(4.3)<br />

where γ is the energy <strong>of</strong> the solid-solid <strong>in</strong>terface, <strong>and</strong> the subscripts c <strong>and</strong> s refer to the<br />

chemical <strong>and</strong> structural components. The chemical component can be estimated with var-<br />

ious theories (e.g., Cahn-Hilliard [27, 28] or the discrete lattice plane model [139]), while<br />

the structural component has been described us<strong>in</strong>g the dislocation models <strong>of</strong> Frank-van der<br />

Merwe [63, 64] <strong>and</strong> Matthews [150]. A recent paper by Willis et al. [250] demonstrates the<br />

equivalence <strong>of</strong> these two dislocation models when the spac<strong>in</strong>g between the misfit disloca-<br />

tions goes to <strong>in</strong>f<strong>in</strong>ity.<br />

For simplicity, only the Matthews result is presented. The energy <strong>of</strong> two perpendicular<br />

<strong>and</strong> non<strong>in</strong>teract<strong>in</strong>g arrays <strong>of</strong> edge dislocations is<br />

� � � �<br />

(δ − εcoh) R<br />

GfGs<br />

γ = b ln + 1<br />

π b (Gf + Gs)(1 − ν)<br />

� �� � � �� �<br />

ξ<br />

Mcomp<br />

(4.4)<br />

where G is the shear modulus, ν is Poisson’s ratio <strong>of</strong> both the film <strong>and</strong> substrate (assumed<br />

equal), R is the cut-<strong>of</strong>f parameter, <strong>and</strong> εcoh is the coherency stra<strong>in</strong> between 0 <strong>and</strong> δ. The<br />

variable ξ is a dimensionless parameter represent<strong>in</strong>g the behavior <strong>of</strong> the stress field near the<br />

dislocation core, <strong>and</strong> Mcomp is the composite modulus. Ref<strong>in</strong>ements on this equation that<br />

consider anharmonic effects [155, 147, 146] as well as more complex dislocation arrays<br />

are available [20, 21].


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 89<br />

4.2 Interfacial Thermodynamics<br />

This section is based ma<strong>in</strong>ly on Sutton <strong>and</strong> Balluffi [214] who, <strong>in</strong> turn, base most <strong>of</strong><br />

their material on Cahn [26]. An <strong>in</strong>terest<strong>in</strong>g <strong>and</strong> rigorous treatment <strong>of</strong> the thermodynamics<br />

<strong>of</strong> solid surfaces is provided by Rusanov [193].<br />

4.2.1 Fluid-Fluid Interfaces<br />

For planar fluid-fluid <strong>in</strong>terfaces, a change <strong>in</strong> <strong>in</strong>ternal energy can be expressed as<br />

dU = TdS− PdV + µid<strong>Ni</strong> + γ dA (4.5)<br />

where U is the <strong>in</strong>ternal energy, T is the temperature, S is the entropy, P is the hydrostatic<br />

pressure, V is the volume, µi is the chemical potential per atom, <strong>Ni</strong> is the number <strong>of</strong><br />

atoms <strong>of</strong> species i, γ is the <strong>in</strong>terface free energy, A is the <strong>in</strong>terface area, <strong>and</strong> the E<strong>in</strong>ste<strong>in</strong><br />

summation notation is used. S<strong>in</strong>ce the <strong>in</strong>ternal energy is a homogeneous function <strong>of</strong> the<br />

first order,<br />

or<br />

Tak<strong>in</strong>g the total differential <strong>of</strong> equation (4.6) gives<br />

U = TS− PV + µi <strong>Ni</strong> + γ A (4.6)<br />

γ = 1<br />

A (U − TS+ PV − µi <strong>Ni</strong>) . (4.7)<br />

dU =TdS− PdV + µid<strong>Ni</strong> + γ dA+ (4.8)<br />

SdT − VdP + dµi <strong>Ni</strong> + Adγ.


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 90<br />

After subtract<strong>in</strong>g equation (4.5) from (4.8), we obta<strong>in</strong><br />

or<br />

0 = SdT − VdP + <strong>Ni</strong>dµ + Adγ (4.9)<br />

dγ =−[S]dT + [V ]dP − [<strong>Ni</strong>]dµ. (4.10)<br />

The bracketed variables are the so-called layer quantities [26] <strong>and</strong> are not uniquely def<strong>in</strong>ed.<br />

The Gibbs-Duhem equations for the two fluid phases α <strong>and</strong> β are<br />

0 = −S α dT + V α dP − N α<br />

i dµi<br />

0 = −S β dT + V β dP − N β<br />

i dµi.<br />

(4.11)<br />

Us<strong>in</strong>g Cramer’s rule to solve (4.10) <strong>and</strong> (4.11) for dγ , we obta<strong>in</strong> the Gibbs adsorption<br />

equation for fluid <strong>in</strong>terfaces<br />

dγ =−�SdT + �VdP − �<strong>Ni</strong>dµi<br />

(4.12)<br />

<strong>in</strong> which the excess quantities �Xi refer to the extra X per unit area compared to a system<br />

with the same composition α <strong>and</strong> β phases but no <strong>in</strong>terface. The excess quantities are<br />

well-def<strong>in</strong>ed for any choice <strong>of</strong> the divid<strong>in</strong>g surface.<br />

4.2.2 Solid-Fluid Interfaces<br />

S<strong>in</strong>ce surface atoms have different bond<strong>in</strong>g <strong>and</strong> average coord<strong>in</strong>ation from those <strong>in</strong> the<br />

bulk, the surface layer has a different lattice parameter than the bulk. However, the surface<br />

is coherent with the bulk (ignor<strong>in</strong>g surface dislocations <strong>and</strong> reconstructions), so a stress<br />

must be present <strong>in</strong> the layer.


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 91<br />

Here, I closely follow Sutton <strong>and</strong> Balluffi’s derivation <strong>of</strong> the relationship between sur-<br />

face stress <strong>and</strong> surface free energy [214]; Shuttleworth’s orig<strong>in</strong>al derivation can be found<br />

<strong>in</strong> his paper [206] or <strong>in</strong> Cammarata’s review [29].<br />

We can imag<strong>in</strong>e a free-st<strong>and</strong><strong>in</strong>g slab <strong>of</strong> thickness t, length L, <strong>and</strong> width L, where L ≫ t<br />

(Figure 4.4). In this schematic, the two surfaces are stretched to fit onto the bulk, much <strong>in</strong><br />

the same way as the film was stretched to fit onto the substrate <strong>in</strong> Figure 1.3. If the slab is<br />

thick enough, we can approximate the stress distribution <strong>in</strong> the bulk as uniform through its<br />

thickness result<strong>in</strong>g <strong>in</strong> the force balance relationship<br />

fij = −σijt<br />

2<br />

for i, j = 1, 2 (4.13)<br />

where fij is the force on the surface per unit width (surface stress), σij is the bulk stress,<br />

<strong>and</strong> I have already <strong>in</strong>serted the generalized tensor form for f <strong>and</strong> σ , which can easily be<br />

proven [159].<br />

Now a variational stra<strong>in</strong> δɛij is applied to the slab. The change <strong>in</strong> free energy is given<br />

by the change <strong>in</strong> surface free energy plus the change <strong>in</strong> bulk stra<strong>in</strong> energy:<br />

�<br />

∂(2γ L2 )<br />

δG =<br />

∂ɛij<br />

�<br />

2 ∂w<br />

+ tL δɛij, (4.14)<br />

∂ɛij<br />

where w is the stra<strong>in</strong> energy density <strong>of</strong> the bulk <strong>and</strong> related to the stresses by<br />

∂w<br />

∂ɛij<br />

= σij. (4.15)<br />

Under equilibrium conditions, δG = 0. After <strong>in</strong>sert<strong>in</strong>g <strong>in</strong> equation (4.13) <strong>and</strong> not<strong>in</strong>g that<br />

the δɛij are <strong>in</strong>dependent variations, we obta<strong>in</strong><br />

fijA = γ<br />

∂ A<br />

∂ɛij<br />

+ A ∂γ<br />

. (4.16)<br />

∂ɛij


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 92<br />

F<strong>in</strong>ally, s<strong>in</strong>ce<br />

∂ A<br />

∂ɛij<br />

we arrive at the Shuttleworth relationship:<br />

= A for i = j (4.17)<br />

= 0 for i �= j, (4.18)<br />

fij = δijγ + ∂γ<br />

∂ɛij<br />

. (4.19)<br />

We can now add the surface stress term to the Gibbs adsorption equation:<br />

dγ =−�SdT + �VdP − �<strong>Ni</strong>dµi + ( fij − δijγ)dɛij. (4.20)<br />

An <strong>in</strong>terest<strong>in</strong>g Maxwell relation between the surface stress <strong>and</strong> the excess volume can be<br />

obta<strong>in</strong>ed by tak<strong>in</strong>g second derivatives <strong>of</strong> γ <strong>and</strong> chang<strong>in</strong>g the order <strong>of</strong> differentiation:<br />

� �<br />

∂( fij − δijγ)<br />

∂ P<br />

� �<br />

∂�V<br />

=<br />

. (4.21)<br />

4.2.3 Solid-Solid Interfaces<br />

T,µk<br />

Solid-solid <strong>in</strong>terfaces can be treated <strong>in</strong> a similar way, although one needs to be care-<br />

ful about dist<strong>in</strong>guish<strong>in</strong>g between coherent <strong>and</strong> <strong>in</strong>coherent <strong>in</strong>terfaces. Figure 4.5 shows a<br />

∂ɛij<br />

T,µk<br />

similar schematic to Figure 4.4, but now applied to solid multilayers.<br />

In analogy to the previous section, the Shuttleworth relationship for solid-solid <strong>in</strong>ter-<br />

faces is<br />

fij = f α β<br />

ij + fij = δijγ + ∂γ<br />

. (4.22)<br />

∂ɛij<br />

This equation is for a coherent <strong>in</strong>terface. For an <strong>in</strong>coherent <strong>in</strong>terface, one phase may be<br />

stra<strong>in</strong>ed relative to the other one by an amount ɛ r ; we then require two <strong>in</strong>terface stresses


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 93<br />

t<br />

Figure 4.4: A schematic demonstrat<strong>in</strong>g how misfitt<strong>in</strong>g surface layers can lead to <strong>in</strong>ternal<br />

bulk stresses.<br />

t<br />

t<br />

β<br />

α<br />

β<br />

α<br />

Figure 4.5: A schematic show<strong>in</strong>g how <strong>in</strong>terface stresses can lead to <strong>in</strong>ternal bulk stresses.<br />

α<br />

f11 α<br />

f11 α<br />

f11 α<br />

f11 β<br />

σ11 f 11<br />

f 11<br />

α<br />

σ11 β<br />

σ11 α<br />

σ11 σ 11<br />

β<br />

f11 β<br />

f11 β<br />

f11 β<br />

f11


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 94<br />

[30],<br />

<strong>and</strong><br />

fij = δijγ + ∂γ<br />

∂ɛij<br />

gij = δijγ + ∂γ<br />

∂ɛ r ij<br />

(4.23)<br />

(4.24)<br />

where we can now connote fij with stra<strong>in</strong><strong>in</strong>g a unit area at constant <strong>in</strong>terface structure<br />

<strong>and</strong> gij with chang<strong>in</strong>g the <strong>in</strong>terface structure with different stra<strong>in</strong>s <strong>in</strong> the two phases but<br />

constant average stra<strong>in</strong> [247].<br />

The Maxwell relationship between the excess volume <strong>and</strong> <strong>in</strong>terface stress is the same<br />

as that <strong>in</strong> equation (4.21) if the <strong>in</strong>terfaces are <strong>in</strong>coherent; for coherent <strong>in</strong>terfaces, no simple<br />

relationship has yet been derived, although Johnson et al. [117, 118, 116] have made sig-<br />

nificant progress. The problem is that the Gibbs-Duhem equation needs to be reformulated<br />

when the system conta<strong>in</strong>s coherency stresses [137, 117]. Johnson found that for coher-<br />

ent <strong>in</strong>terfaces, the <strong>in</strong>terface stress <strong>and</strong> excess volume were physically mean<strong>in</strong>gful only for<br />

certa<strong>in</strong> multilayer compositions [116]. This result has not yet been corroborated.<br />

The general theory relat<strong>in</strong>g the <strong>in</strong>terface stress to the average volume stress was pre-<br />

sented by Weissmüller <strong>and</strong> Cahn [247]. Their result as applied to s<strong>in</strong>gle crystal multilayers<br />

is presented here:<br />

〈σij〉v = −2<br />

�<br />

⎡ ⎤<br />

⎢ f11<br />

⎢ f12 ⎢<br />

⎣<br />

f12<br />

f22<br />

0 ⎥<br />

0⎥<br />

,<br />

⎥<br />

⎦<br />

(4.25)<br />

0 0 0<br />

where 〈σij〉v is the average volume stress tensor <strong>and</strong> � is the bilayer thickness. For a biaxial


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 95<br />

stress state <strong>and</strong> <strong>in</strong>terfaces with three-fold symmetry or higher, this reduces to<br />

〈σ 〉v = −2<br />

�<br />

f (4.26)<br />

where f is now a scalar. Weissmüller <strong>and</strong> Cahn note that this equation does not assume<br />

<strong>in</strong>dependently relaxed stra<strong>in</strong>s <strong>in</strong> the <strong>in</strong>dividual layers. Also, the <strong>in</strong>terface stress gij does<br />

not contribute to to the average volume stress.<br />

Prior to Weissmüller <strong>and</strong> Cahn, Ruud et al. [196] derived a similar equation for use <strong>in</strong><br />

measur<strong>in</strong>g the <strong>Ag</strong>/<strong>Ni</strong> <strong>in</strong>terface stress. The <strong>Ag</strong> <strong>and</strong> <strong>Ni</strong> layers were not s<strong>in</strong>gle crystals so their<br />

value <strong>of</strong> f is a comb<strong>in</strong>ation <strong>of</strong> the <strong>Ag</strong>/<strong>Ni</strong> <strong>in</strong>terface stress <strong>and</strong> <strong>Ag</strong> <strong>and</strong> <strong>Ni</strong> gra<strong>in</strong> boundary<br />

stresses. Spaepen gives a correction <strong>in</strong> Ref. [212]. An <strong>in</strong>dependent determ<strong>in</strong>ation <strong>of</strong> the <strong>Ag</strong><br />

<strong>and</strong> <strong>Ni</strong> gra<strong>in</strong> boundary stresses has not yet been carried out.<br />

4.3 Pr<strong>in</strong>ciple <strong>of</strong> the Interface Stress Measurement<br />

In Chapter 1, we discussed several “types” <strong>of</strong> stresses <strong>in</strong> th<strong>in</strong> films <strong>and</strong> categorized<br />

them accord<strong>in</strong>g to their measurability by wafer curvature <strong>and</strong> x-ray methods (Table 1.1).<br />

If the thicknesses <strong>of</strong> the <strong>Ag</strong> <strong>and</strong> <strong>Ni</strong> layers are equal, the substrate curvature change due to<br />

coherency stresses is zero [101]; <strong>in</strong> this case, barr<strong>in</strong>g any <strong>in</strong>terface phase formation, the<br />

difference between the curvature measured stress <strong>and</strong> the x-ray measured stress orig<strong>in</strong>ates<br />

from the <strong>in</strong>terface stress.<br />

For a multilayer on a substrate, equation (4.26) becomes<br />

� � 2 Nf<br />

σsc −〈σ 〉x−ray = f =<br />

� tf<br />

(4.27)<br />

where σsc is the stress on the multilayer due to the substrate constra<strong>in</strong>t, 〈σ 〉x−ray is the<br />

average volume stress measured by x-rays, tf is the total film thickness, <strong>and</strong> N is the total


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 96<br />

Table 4.1: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Energy<br />

Chemical Structural Total Preparation Method Ref.<br />

- - 0.73 ± 0.12 N/m e-beam (725 ◦ C) [121]<br />

0.05 N/m ∗ 0.37 N/m 0.42 N/m EAM calculation [87]<br />

∗ Average <strong>of</strong> the coherent stack<strong>in</strong>g <strong>and</strong> <strong>in</strong>tr<strong>in</strong>sic stack<strong>in</strong>g fault areas.<br />

number <strong>of</strong> <strong>in</strong>terfaces [196]. By plott<strong>in</strong>g � �<br />

σsc −〈σ 〉x−ray versus �−1 <strong>and</strong> perform<strong>in</strong>g a<br />

l<strong>in</strong>ear fit, the <strong>in</strong>terface stress can be obta<strong>in</strong>ed from the slope.<br />

4.4 Previous Measurements <strong>of</strong> the <strong>Ag</strong>/<strong>Ni</strong> {111} Interface<br />

Energy <strong>and</strong> Stress<br />

There has been only one measurement <strong>of</strong> the <strong>Ag</strong>/<strong>Ni</strong> {111} <strong>in</strong>terface energy (Table 4.1)<br />

<strong>and</strong> four <strong>of</strong> the <strong>Ag</strong>/<strong>Ni</strong> {111} <strong>in</strong>terface stress (Table 4.2), <strong>in</strong>clud<strong>in</strong>g the work presented here.<br />

The measured <strong>in</strong>terface energy is larger than the embedded atom method (EAM) result,<br />

but this may be partially attributable to the temperature dependence <strong>of</strong> the <strong>in</strong>terface energy<br />

[211] s<strong>in</strong>ce the EAM calculation was performed at zero temperature. Accord<strong>in</strong>g to the<br />

EAM result, the structural component dom<strong>in</strong>ates the <strong>in</strong>terface energy.<br />

Three <strong>of</strong> the four <strong>Ag</strong>/<strong>Ni</strong> <strong>in</strong>terface stress measurements rely on equation (4.27) <strong>and</strong> one<br />

relies on equation (4.26) [120]. The multilayers <strong>in</strong> each <strong>of</strong> the four experiments vary <strong>in</strong><br />

deposition parameters as well as layer quality; however, these differences have little effect<br />

on the value <strong>of</strong> the <strong>in</strong>terface stress. The other FCC metallic multilayers listed <strong>in</strong> Table 4.2<br />

have similar misfits to <strong>Ag</strong>/<strong>Ni</strong>, <strong>and</strong> their <strong>in</strong>terface stresses are similar <strong>in</strong> sign <strong>and</strong> magnitude<br />

to that <strong>of</strong> <strong>Ag</strong>/<strong>Ni</strong>. All <strong>of</strong> these experimental results differ significantly from the EAM results,


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 97<br />

∗ Dynamic value<br />

Table 4.2: {111} Interface Stress Results for FCC Metallic <strong>Multilayers</strong><br />

System Value Preparation Method Ref.<br />

<strong>Ag</strong>/<strong>Ni</strong> −2.02 ± 0.26 N/m e-beam (∼ −196 ◦ C) [120]<br />

<strong>Ag</strong>/<strong>Ni</strong> −2.24 ± 0.21 N/m sputter<strong>in</strong>g (−2 ◦ C) [200]<br />

<strong>Ag</strong>/<strong>Ni</strong> −2.27 ± 0.67 N/m sputter<strong>in</strong>g (−156 ◦ C) [196]<br />

<strong>Ag</strong>/<strong>Ni</strong> −2.17 ± 0.15 N/m sputter<strong>in</strong>g (−190 ◦ C) this work<br />

<strong>Ag</strong>/<strong>Ni</strong> +0.32 N/m EAM calculation [88]<br />

<strong>Ag</strong>/<strong>Cu</strong> −3.19 ± 0.43 N/m sputter<strong>in</strong>g (35 ◦ C) [14]<br />

<strong>Ag</strong>/<strong>Cu</strong> −0.21 ± 0.10 N/m ∗ evaporation (40 ◦ C) [205]<br />

<strong>Ag</strong>/<strong>Cu</strong> +0.32 N/m EAM calculation [88]<br />

Au/<strong>Ni</strong> −2.69 ± 0.43 N/m † sputtered (−5 ◦ C) [198]<br />

Au/<strong>Ni</strong> −0.08 N/m EAM calculation [88]<br />

† After account<strong>in</strong>g for <strong>in</strong>termix<strong>in</strong>g<br />

which are smaller <strong>in</strong> magnitude <strong>and</strong> are generally tensile.<br />

To resolve the discrepancy between the EAM results <strong>and</strong> experiment, we consider the<br />

follow<strong>in</strong>g two scenarios. The first scenario is that the <strong>in</strong>terpretation <strong>of</strong> the experiments<br />

is flawed. Clemens et al. [39] have hypothesized that <strong>in</strong>termix<strong>in</strong>g between the multilayer<br />

components <strong>in</strong>terferes with the <strong>in</strong>terface stress measurement. The second scenario is that<br />

the <strong>in</strong>terpretation is correct, <strong>and</strong> the discrepancy with the EAM calculations arises from the<br />

<strong>in</strong>ability <strong>of</strong> EAM to consider stresses <strong>in</strong> a real system. For <strong>in</strong>stance, the EAM calculation<br />

was performed for bilayers with very small residual stra<strong>in</strong>s at a temperature <strong>of</strong> zero Kelv<strong>in</strong>,<br />

while the experiments measure f <strong>in</strong> multilayers with large stra<strong>in</strong>s at room temperature.<br />

Furthermore, the embedded atom method may only apply to the l<strong>in</strong>early elastic regime<br />

(only the second order elastic constants were used as <strong>in</strong>puts) [45].<br />

Two experiments were carried out to help shed light on the controversy. One remeasures<br />

the <strong>Ag</strong>/<strong>Ni</strong> {111} <strong>in</strong>terface stress. The other exam<strong>in</strong>es the stra<strong>in</strong> <strong>and</strong> compositional pr<strong>of</strong>iles<br />

<strong>of</strong> the multilayers. After present<strong>in</strong>g the results, I will return to the discrepancy <strong>in</strong> section


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 98<br />

4.7.<br />

4.5 Experiment<br />

4.5.1 Film Preparation<br />

<strong>Ag</strong>/<strong>Ni</strong> multilayered films were deposited by ion beam sputter<strong>in</strong>g 99.99 at.% <strong>Ag</strong> <strong>and</strong><br />

99.99 at.% <strong>Ni</strong> targets us<strong>in</strong>g a gettered Ar sputter<strong>in</strong>g gas. The chamber used for deposition<br />

is shown <strong>in</strong> Figure 1.7; it <strong>in</strong>cludes a rotat<strong>in</strong>g target block to switch between the <strong>Ag</strong> <strong>and</strong><br />

<strong>Ni</strong> targets [213], which elim<strong>in</strong>ates the need for precise shutter<strong>in</strong>g. The Si(100) substrates<br />

(381 µm thick) were kept near −190 ◦ C dur<strong>in</strong>g deposition to prevent ball<strong>in</strong>g up <strong>of</strong> <strong>Ag</strong> on <strong>Ni</strong><br />

<strong>and</strong> vice versa, as well as to m<strong>in</strong>imize <strong>in</strong>terdiffusion. The temperature was monitored with<br />

a thermocouple placed near the backside <strong>of</strong> the substrate (the side away from the deposition<br />

flux). Five different bilayer thicknesses were made: 3.2 nm, 5.9 nm, 11.3 nm, 17.3 nm, <strong>and</strong><br />

22.8 nm. All sputter<strong>in</strong>g targets <strong>and</strong> substrates were sputter cleaned prior to deposition. The<br />

sputter<strong>in</strong>g parameters are given <strong>in</strong> Table 1.2.<br />

Subsequent to the <strong>in</strong>terface stress measurement, another multilayer with � = 4.2nm<br />

was grown for specular reflection <strong>and</strong> diffuse scatter<strong>in</strong>g measurements. The results for this<br />

multilayer are given <strong>in</strong> section 4.7.1.<br />

4.5.2 Film Characterization<br />

Rutherford Backscatter<strong>in</strong>g Spectrometry (RBS) was used to determ<strong>in</strong>e the film compo-<br />

sition <strong>and</strong> the overall thickness <strong>of</strong> the multilayers. The thicknesses were also checked with<br />

a Tencor Alpha-Step 200 Pr<strong>of</strong>ilometer. The film quality was determ<strong>in</strong>ed by both low-angle


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 99<br />

x-ray reflectivity <strong>and</strong> cross-sectional transmission electron microscopy (TEM).<br />

4.5.3 <strong>Cu</strong>rvature Measurements<br />

The stress applied by the substrate on the multilayer was determ<strong>in</strong>ed by measur<strong>in</strong>g the<br />

change <strong>in</strong> curvature with the wafer curvature apparatus (Figure 1.4) described <strong>in</strong> Chapter 1;<br />

the Stoney equation (1.16) was used to convert the curvatures <strong>in</strong>to stresses. Measurements<br />

on a laser scann<strong>in</strong>g apparatus [257] were taken both before <strong>and</strong> after deposition. S<strong>in</strong>ce<br />

substantial stress relaxation took place after deposition, the x-ray scans were performed<br />

only after the stress stabilized.<br />

4.5.4 X-Ray Diffraction<br />

The multilayers were exam<strong>in</strong>ed at the National Synchrotron Light Source at beaml<strong>in</strong>e<br />

X22C with a wavelength <strong>of</strong> 0.154591 nm; some measurements were also performed with a<br />

shorter wavelength, 0.149020 nm. The wavelengths were chosen to be far from <strong>and</strong> close<br />

to the <strong>Ni</strong> K absorption edge to effect a large change <strong>in</strong> f ′ , the real part <strong>of</strong> the atomic<br />

scatter<strong>in</strong>g factor. Low-angle θ − 2θ scans were made to quantify the bilayer thickness <strong>and</strong><br />

<strong>in</strong>terface roughness, <strong>and</strong> high-angle scans were taken to measure the out-<strong>of</strong>-plane stra<strong>in</strong>s.<br />

The SUPREX s<strong>of</strong>tware package [70] <strong>and</strong> Bede Scientific’s REFS s<strong>of</strong>tware were used to<br />

analyze the high-angle <strong>and</strong> low-angle θ − 2θ results.<br />

The <strong>in</strong>-plane stra<strong>in</strong>s, ɛ�, were measured with graz<strong>in</strong>g <strong>in</strong>cidence diffraction (GID) [148]<br />

at <strong>in</strong>cidence angles <strong>of</strong> 1 ◦ <strong>and</strong> 2 ◦ . The {220} diffraction planes for silver <strong>and</strong> nickel were<br />

chosen for the measurement us<strong>in</strong>g the bulk lattice parameters <strong>of</strong> <strong>Ag</strong> <strong>and</strong> <strong>Ni</strong> for reference.


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 100<br />

The average volume stress was then calculated us<strong>in</strong>g<br />

〈σ 〉x−ray = Y<strong>Ag</strong> t<strong>Ag</strong> ɛ�,<strong>Ag</strong> + Y<strong>Ni</strong> t<strong>Ni</strong> ɛ�,<strong>Ni</strong><br />

t<strong>Ag</strong> + t<strong>Ni</strong><br />

(4.28)<br />

where Yi is the biaxial modulus, ti is the layer thickness, <strong>and</strong> ɛ�,i is the <strong>in</strong>-plane stra<strong>in</strong> <strong>of</strong><br />

element i. Due to the large stra<strong>in</strong>s <strong>in</strong>volved, Yi is stra<strong>in</strong> dependent, <strong>and</strong> higher order elastic<br />

constants are required. Assum<strong>in</strong>g perfect {111} texture, we use the formulas [196, 240]<br />

4.6 Results<br />

4.6.1 Microstructure<br />

Y<strong>Ag</strong>(ɛ) = 173.60 − 3764.9 ɛ − 16349 ɛ 2 (GPa) (4.29)<br />

<strong>and</strong> Y<strong>Ni</strong>(ɛ) = 389.32 − 4903.9 ɛ − 17006 ɛ 2 (GPa). (4.30)<br />

Figure 4.6 shows a cross-sectional TEM micrograph <strong>of</strong> a � = 4.2 nm <strong>Ag</strong>/<strong>Ni</strong> multilayer.<br />

The layers are <strong>of</strong> good quality <strong>and</strong> have well-def<strong>in</strong>ed <strong>in</strong>terfaces, as confirmed by the low-<br />

angle diffraction scans shown <strong>in</strong> Figure 4.7. Fitt<strong>in</strong>g <strong>of</strong> the low-angle data gave an average<br />

total <strong>in</strong>terface roughness <strong>of</strong> 0.4 nm for all <strong>of</strong> the multilayers used <strong>in</strong> the <strong>in</strong>terface stress<br />

measurement.<br />

The RBS data <strong>in</strong>dicated that the films were slightly <strong>Ag</strong> rich (51-53 at.%), <strong>and</strong> the total<br />

thicknesses varied from 1.14 to 1.29 µm. From these results, the deposition rates were<br />

calculated to be approximately 0.5nm/s for <strong>Ag</strong> <strong>and</strong> 0.1nm/s for <strong>Ni</strong>.


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 101<br />

Figure 4.6: A cross-sectional view <strong>of</strong> a <strong>Ag</strong>/<strong>Ni</strong> multilayer with � = 4.2nm.<br />

4.6.2 Substrate <strong>Cu</strong>rvature Stress<br />

The stress <strong>of</strong> the multilayer due to the substrate constra<strong>in</strong>t is shown as the open circles<br />

<strong>in</strong> Figure 4.8 (b). In general, the magnitude <strong>of</strong> σsc <strong>in</strong>creases slightly with bilayer thickness.<br />

4.6.3 Volume Stress<br />

FCC metallic multilayers are expected to have 〈111〉 texture; the high-angle θ − 2θ<br />

scans confirmed this <strong>and</strong> are shown later. S<strong>in</strong>ce most <strong>of</strong> the gra<strong>in</strong>s are 〈111〉-oriented,<br />

the macroscopic <strong>in</strong>-plane biaxial stra<strong>in</strong> can be determ<strong>in</strong>ed by the d-spac<strong>in</strong>gs <strong>of</strong> the {220}<br />

diffraction peaks.<br />

The <strong>Ag</strong> {220}, <strong>Ni</strong>{220}, <strong>Ag</strong>{311}, <strong>and</strong> <strong>Ag</strong> {222} GID peaks <strong>in</strong> Figure 4.9 were simul-<br />

taneously fit with four pseudo-Voigt functions [248]. The result<strong>in</strong>g stra<strong>in</strong>s are shown <strong>in</strong>


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 102<br />

Λ = 11.3 nm<br />

Λ = 5.9 nm<br />

Λ = 3.2 nm<br />

Figure 4.7: Offset low-angle θ − 2θ diffraction spectra for the three smallest bilayer thicknesses.<br />

The spectra were taken with an x-ray energy <strong>of</strong> 8020 eV.


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 103<br />

σ<br />

Λ<br />

Λ<br />

〈σ〉<br />

σ<br />

(a) (b)<br />

Figure 4.8: (a) Volume stresses <strong>in</strong> <strong>Ni</strong> (closed squares) <strong>and</strong> <strong>Ag</strong> (open squares) layers as<br />

a function <strong>of</strong> <strong>in</strong>verse bilayer thickness. (b) Substrate curvature stress (open circles) <strong>and</strong><br />

average volume stress (closed circles) as a function <strong>of</strong> <strong>in</strong>verse bilayer thickness. The error<br />

bars for all the data are with<strong>in</strong> the size <strong>of</strong> the symbols. The arrow drawn from (a) to (b)<br />

illustrates how the layer stresses are averaged to give 〈σ 〉x−ray.<br />

Λ<br />

Λ


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 104<br />

Λ<br />

θ [degrees]<br />

Figure 4.9: Offset GID spectra for <strong>Ag</strong>/<strong>Ni</strong> multilayers hav<strong>in</strong>g the <strong>in</strong>dicated bilayer thickness<br />

�. The vertical l<strong>in</strong>es <strong>in</strong>dicate the {220} peak positions for stra<strong>in</strong>-free bulk <strong>Ag</strong> <strong>and</strong> <strong>Ni</strong>. The<br />

x-ray energy was 8020 eV.<br />

Figure 4.10.<br />

The stra<strong>in</strong>s were converted to volume stresses with the use <strong>of</strong> higher order elastic con-<br />

stants, as discussed earlier. Figure 4.9 shows that the <strong>Ni</strong> {220} <strong>and</strong> <strong>Ag</strong> {311} peaks overlap<br />

significantly for the � = 17.3 nm <strong>and</strong> 22.8 nm samples, prevent<strong>in</strong>g an accurate measure <strong>of</strong><br />

〈σ 〉x−ray at these bilayer thicknesses. These samples were not used when calculat<strong>in</strong>g the<br />

<strong>in</strong>terface stress.<br />

The average volume stresses calculated us<strong>in</strong>g equation (4.28) are shown as the closed<br />

circles <strong>in</strong> Figure 4.8 (b). This figure also shows that σsc <strong>and</strong> 〈σ 〉x−ray converge as the<br />

bilayer thickness <strong>in</strong>creases. Without an <strong>in</strong>terface stress, these curves should co<strong>in</strong>cide for<br />

all bilayer thicknesses.


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 105<br />

ε<br />

Λ<br />

Λ<br />

Figure 4.10: In-plane stra<strong>in</strong>s <strong>in</strong> nickel <strong>and</strong> silver as a function <strong>of</strong> bilayer thickness.<br />

4.6.4 The Interface Stress<br />

The difference between the substrate curvature stress <strong>and</strong> the average volume stress is<br />

plotted as a function <strong>of</strong> <strong>in</strong>verse bilayer thickness <strong>in</strong> Figure 4.11. A weighted least squares<br />

fit for the three smallest bilayer thicknesses produces a l<strong>in</strong>e with slope −4.34, giv<strong>in</strong>g an<br />

<strong>in</strong>terface stress <strong>of</strong> −2.17 ± 0.15Nm −1 . The <strong>in</strong>terface stress appears to <strong>in</strong>crease sharply<br />

when �>11.3 nm for unknown reasons but is most likely due to a change <strong>in</strong> the degree<br />

<strong>of</strong> coherency <strong>of</strong> the <strong>in</strong>terface. Schweitz et al. [200] noted a similar trend <strong>in</strong> their data.<br />

This value agrees with the previous results listed <strong>in</strong> Table 4.2. As mentioned, the <strong>in</strong>ter-<br />

face stress is large <strong>and</strong> compressive while the EAM result is small <strong>and</strong> tensile.<br />

Before go<strong>in</strong>g on to discuss different models for the compressive <strong>in</strong>terface stress, I repeat


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 106<br />

σ 〈σ〉<br />

Λ<br />

Λ<br />

Figure 4.11: Difference between the substrate curvature stress <strong>and</strong> the average volume<br />

stress for <strong>Ag</strong>/<strong>Ni</strong> multilayers as a function <strong>of</strong> <strong>in</strong>verse bilayer thickness. The l<strong>in</strong>e is a<br />

weighted least-squares fit for the three smallest bilayer thicknesses.


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 107<br />

the Shuttleworth relation,<br />

f = γ + ∂γ<br />

. (4.31)<br />

∂ɛ<br />

The first term on the right side is the excess energy per unit deformed area, i.e., the area<br />

after the <strong>in</strong>terface has been deformed to fit between the bulk phases [174]. It is possible for<br />

γ to be negative, although not for a stable <strong>in</strong>terface; however, experiment <strong>and</strong> calculations<br />

agree that it is positive. Then for f to be negative, the derivative term must be negative,<br />

with the <strong>in</strong>terface energy decreas<strong>in</strong>g with <strong>in</strong>creas<strong>in</strong>g area. This result is not <strong>in</strong>tuitive <strong>and</strong> has<br />

led some to question the validity <strong>of</strong> the experiment. Note that the EAM simulation actually<br />

shows the same phenomenon, although the magnitude is much smaller; γ = 0.42 N/m<br />

with f = 0.32 N/m.<br />

4.7 Models for the Compressive Interface Stress<br />

Here, I return to the questions concern<strong>in</strong>g the disagreement between the EAM results<br />

<strong>and</strong> experiment <strong>and</strong> <strong>in</strong>vestigate the possibility that the <strong>in</strong>terpretation <strong>of</strong> the experiments is<br />

flawed by <strong>in</strong>termix<strong>in</strong>g. Afterwards, I will discuss another possibility: the <strong>in</strong>terpretation <strong>of</strong><br />

the experiments is correct <strong>and</strong> the EAM calculations do not consider an analogous system<br />

that merits comparison.<br />

4.7.1 Intermix<strong>in</strong>g<br />

<strong>Ni</strong>ckel appears to have a negative Poisson ratio <strong>in</strong> the plane <strong>of</strong> the <strong>Ag</strong>/<strong>Ni</strong> multilayers<br />

[190, 200]. Although the <strong>in</strong>-plane stra<strong>in</strong>s are strongly tensile, the out-<strong>of</strong>-plane stra<strong>in</strong>s are<br />

also tensile, as shown <strong>in</strong> Figure 4.12. This is very different from the s<strong>in</strong>gle crystal value.


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 108<br />

Λ<br />

Figure 4.12: The average {111} d-spac<strong>in</strong>g for nickel. S<strong>in</strong>ce the nickel is <strong>in</strong> tension parallel<br />

to the {111} planes, d111 is expected to be less than the stra<strong>in</strong>-free d-spac<strong>in</strong>g <strong>of</strong> 0.20345 nm<br />

for a positive Poisson ratio.<br />

Λ<br />

Gergaud et al. [83], us<strong>in</strong>g the s<strong>in</strong> 2 ψ-method (Appendix A), have claimed that the<br />

stra<strong>in</strong>-free lattice parameter <strong>of</strong> <strong>Ni</strong> <strong>in</strong>creases with smaller bilayer thicknesses, <strong>and</strong> attributes<br />

the <strong>in</strong>crease to <strong>in</strong>termix<strong>in</strong>g. Use <strong>of</strong> the s<strong>in</strong> 2 ψ-method, however, should be used with cau-<br />

tion when work<strong>in</strong>g with multilayer structures. Chocyk et al. [38] have demonstrated that<br />

the s<strong>in</strong> 2 ψ-method gives <strong>in</strong>correct stresses when applied to the Au/<strong>Ni</strong> multilayers. Further-<br />

more, as shown <strong>in</strong> Appendix A, this method <strong>of</strong> determ<strong>in</strong><strong>in</strong>g the stra<strong>in</strong>-free lattice parameter<br />

relies on the bulk Poisson ratio; hence the negative Poisson ratio effect <strong>and</strong> the stra<strong>in</strong>-free<br />

lattice parameter problem are related. The question is whether both can be expla<strong>in</strong>ed by a<br />

self-consistent model.<br />

In view <strong>of</strong> Gergaud’s results, Clemens et al. [39] reasoned that s<strong>in</strong>ce the surface energy<br />

<strong>of</strong> <strong>Ag</strong> is lower than that <strong>of</strong> <strong>Ni</strong>, <strong>Ag</strong> atoms prefer to lie on the film surface. If, dur<strong>in</strong>g


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 109<br />

<strong>Ni</strong> deposition, the deposition k<strong>in</strong>etics were such that a substantial amount <strong>of</strong> <strong>Ag</strong> could<br />

segregate <strong>in</strong>to the <strong>Ni</strong> <strong>and</strong> become k<strong>in</strong>etically trapped, this would <strong>in</strong>validate the use <strong>of</strong> the<br />

nickel bulk lattice parameter when calculat<strong>in</strong>g the stra<strong>in</strong>. Consequently, this would yield a<br />

spurious value for the <strong>in</strong>terface stress.<br />

Intermix<strong>in</strong>g, however, should be dependent on the deposition k<strong>in</strong>etics (Figure 4.13). If<br />

the segregation <strong>of</strong> <strong>Ag</strong> through <strong>Ni</strong> is too slow, this results <strong>in</strong> no mix<strong>in</strong>g. If the segregation<br />

is too fast, then <strong>Ag</strong> “floats” through the <strong>Ni</strong> <strong>and</strong> does not mix. If the segregation falls <strong>in</strong>to<br />

an <strong>in</strong>termediate regime, most likely a compositional gradient exists between the <strong>Ni</strong> <strong>and</strong> <strong>Ag</strong><br />

layers; it seems unlikely that <strong>Ag</strong> can homogenize itself <strong>in</strong> the <strong>Ni</strong> dur<strong>in</strong>g <strong>Ni</strong> deposition to<br />

achieve uniform mix<strong>in</strong>g. Recall that our experiments were performed at liquid nitrogen<br />

temperatures. If the gradient is too steep, there only exists <strong>in</strong>coherent scatter<strong>in</strong>g from the<br />

alloyed planes. The <strong>in</strong>tensity due to <strong>in</strong>coherent scatter<strong>in</strong>g is too weak to contribute to the<br />

<strong>Ni</strong> {220} diffraction peak <strong>and</strong> affect the <strong>in</strong>terface stress [196]. If the gradient is too shallow,<br />

this would result <strong>in</strong> an alloy peak between the <strong>Ag</strong> <strong>and</strong> <strong>Ni</strong> {220} peaks <strong>in</strong> Figure 4.9. This is<br />

not seen.<br />

In the case <strong>of</strong> uniform or nearly uniform <strong>in</strong>termix<strong>in</strong>g, Clemens et al. [39] found that<br />

equation (4.27) should be replaced by<br />

where<br />

δmix = 1<br />

�<br />

Y<strong>Ag</strong><br />

2<br />

� � 2<br />

σsc −〈σ 〉x−ray =<br />

� ( f + δmix) (4.32)<br />

(d<strong>Ag</strong>, 0 − d<strong>Ni</strong>, 0)<br />

(d<strong>Ag</strong>, 0 − d<strong>Ni</strong>, 0)<br />

t<strong>Ni</strong> <strong>in</strong>to <strong>Ag</strong> − Y<strong>Ni</strong><br />

t<strong>Ag</strong> <strong>in</strong>to <strong>Ni</strong><br />

d<strong>Ag</strong>, 0<br />

d<strong>Ni</strong>, 0<br />

�<br />

(4.33)<br />

assum<strong>in</strong>g a l<strong>in</strong>ear relationship between the degree <strong>of</strong> alloy<strong>in</strong>g <strong>and</strong> the change <strong>in</strong> lattice<br />

parameter (Vegard’s law). Here, di, 0 is the stra<strong>in</strong>-free lattice parameter <strong>of</strong> element i, <strong>and</strong>


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 110<br />

<strong>Ni</strong><br />

<strong>Ag</strong><br />

cold hot<br />

Figure 4.13: Depiction <strong>of</strong> the k<strong>in</strong>etics <strong>of</strong> segregation. At cold temperatures, the <strong>Ag</strong> lacks the<br />

mobility to segregate. At high temperatures, the <strong>Ag</strong> “floats” to the surface. At <strong>in</strong>termediate<br />

temperatures, a compositional gradient exists.<br />

tα <strong>in</strong>to β is the the equivalent thickness <strong>of</strong> α mixed with β. Accord<strong>in</strong>g to their argument,<br />

δmix is much larger than f <strong>and</strong> is consistently measured to be −2N/m.<br />

The composition gradient can be estimated us<strong>in</strong>g the follow<strong>in</strong>g method <strong>of</strong> determ<strong>in</strong><strong>in</strong>g<br />

the stra<strong>in</strong> <strong>and</strong> composition pr<strong>of</strong>iles.<br />

A Systematic Method <strong>of</strong> Extract<strong>in</strong>g Stra<strong>in</strong> <strong>and</strong> Composition Pr<strong>of</strong>iles <strong>of</strong> <strong>Multilayers</strong><br />

The stra<strong>in</strong> <strong>and</strong> composition pr<strong>of</strong>ile <strong>of</strong> a multilayer can be obta<strong>in</strong>ed from RBS <strong>and</strong> θ −2θ<br />

scans at both low <strong>and</strong> high angles. The low-angle data result from optical <strong>in</strong>terference <strong>of</strong><br />

the reflected rays from the multilayer <strong>and</strong> can be simulated us<strong>in</strong>g an optical multilayer<br />

formalism [85]. Hence, low-angle reflectivity is <strong>of</strong>ten performed on both amorphous <strong>and</strong><br />

crystall<strong>in</strong>e multilayers. The high-angle data result from Bragg diffraction <strong>of</strong>f a superlattice<br />

[70]. The data from each are complementary. For example, both give a bilayer thickness,<br />

but the high-angle data also provide d-spac<strong>in</strong>gs while the low-angle data provide mass


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 111<br />

N α<br />

N β<br />

Λ<br />

d α<br />

d β<br />

(111)<br />

(111)<br />

Figure 4.14: Depiction <strong>of</strong> a superlattice unit cell us<strong>in</strong>g the parameters <strong>in</strong> equation (4.34).<br />

densities. <strong>Ag</strong>reement between the two on the bilayer thickness allows a consistency check.<br />

However, it should be remembered that the high-angle data provide a bilayer thickness<br />

based on Bragg diffraction; for a s<strong>in</strong>gle crystal multilayer, the agreement should be perfect,<br />

butifthefilm is polycrystall<strong>in</strong>e, the bilayer thicknesses may not match exactly, especially<br />

if a texture gradient exists <strong>in</strong> the multilayer.<br />

The first step <strong>of</strong> the method is to use RBS to f<strong>in</strong>d the ratio Nα/Nβ, where Nα <strong>and</strong> Nβ<br />

are the number <strong>of</strong> α <strong>and</strong> β monolayers <strong>in</strong> one repeat unit, �. Now we turn to the high-angle<br />

θ − 2θ data. The superlattice unit cell is given by<br />

� = Nαdα<br />

� �� �<br />

tα<br />

+ Nβdβ<br />

� �� �<br />

tβ<br />

where the parameters are def<strong>in</strong>ed <strong>in</strong> Figure 4.14. Bragg’s law for a superlattice is<br />

2 s<strong>in</strong> θ<br />

λ<br />

= 1<br />

¯d<br />

± n<br />

�<br />

t α<br />

t β<br />

(4.34)<br />

(4.35)


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 112<br />

Figure 4.15: High-angle x-ray diffraction peaks for a <strong>Ag</strong>/<strong>Ni</strong> multilayer with � = 4.2 nm.<br />

The order <strong>of</strong> the satellite is given above its peak. The data taken at E = 8320 eV are <strong>of</strong>fset<br />

from the E = 8020 eV data.<br />

where n is the order <strong>of</strong> the satellite <strong>and</strong><br />

¯d =<br />

�<br />

Nα + Nβ<br />

. (4.36)<br />

By plott<strong>in</strong>g 2 s<strong>in</strong> θ/λ vs. n, we can determ<strong>in</strong>e ¯d <strong>and</strong> � by fitt<strong>in</strong>g a straight l<strong>in</strong>e. S<strong>in</strong>ce ¯d is<br />

given by (4.36) <strong>and</strong> we know Nα/Nβ from RBS, we can solve for Nα <strong>and</strong> Nβ.<br />

At this stage, we know �, Nα, <strong>and</strong> Nβ but have no <strong>in</strong>formation on dα <strong>and</strong> dβ. Aref<strong>in</strong>e-<br />

ment method can be used to determ<strong>in</strong>e these parameters. Figure 4.15 shows an example<br />

<strong>of</strong> a fit us<strong>in</strong>g SUPREX [70]. In addition to provid<strong>in</strong>g dα <strong>and</strong> dβ <strong>and</strong> therefore tα <strong>and</strong> tβ,<br />

SUPREX supplies additional <strong>in</strong>formation I will discuss later.<br />

We now use the low-angle θ −2θ data. Details on the theory beh<strong>in</strong>d specular reflectivity


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 113<br />

<strong>and</strong> diffuse scatter<strong>in</strong>g can be found <strong>in</strong> a recent text by Als-<strong>Ni</strong>elsen <strong>and</strong> McMorrow [2]. 1<br />

With tα <strong>and</strong> tβ used as consistency checks, true specular reflectivity pr<strong>of</strong>iles 2 are fit <strong>in</strong> order<br />

to determ<strong>in</strong>e the total <strong>in</strong>terface roughnesses, σtotal, <strong>of</strong> both the <strong>Ag</strong>/<strong>Ni</strong> <strong>and</strong> <strong>Ni</strong>/<strong>Ag</strong> <strong>in</strong>terfaces.<br />

As shown <strong>in</strong> Figure 4.16, specular reflectivity samples only along qz <strong>in</strong> reciprocal space <strong>and</strong><br />

cannot dist<strong>in</strong>guish between physical <strong>and</strong> chemical roughness. Longitud<strong>in</strong>al <strong>and</strong> transverse<br />

diffuse scatter<strong>in</strong>g scans are then recorded <strong>and</strong> simulated. The transverse scans sample<br />

qx <strong>in</strong>formation at fixed qz, where the x direction is parallel to the sample surface, <strong>and</strong><br />

longitud<strong>in</strong>al scans sample reciprocal space <strong>in</strong> both x <strong>and</strong> z directions just <strong>of</strong>f the specular<br />

“ridge” at qx = 0. By hav<strong>in</strong>g the qx <strong>in</strong>formation, a dist<strong>in</strong>ction between the physical <strong>and</strong><br />

chemical roughness can be made; they are related to the total <strong>in</strong>terface roughness by<br />

σ 2<br />

total = σ 2 physical + σ 2 chemical . (4.37)<br />

The diffuse scatter<strong>in</strong>g pr<strong>of</strong>ile simulations use the Distorted-Wave Born Approximation<br />

(DWBA) [209, 138]. An example <strong>of</strong> a specular reflectivity fit <strong>and</strong> a longitud<strong>in</strong>al diffuse<br />

scatter<strong>in</strong>g simulation are presented <strong>in</strong> Figure 4.17 (a) <strong>and</strong> (b), respectively.<br />

F<strong>in</strong>ally, the multilayer can be studied with different x-ray energies, as was done <strong>in</strong> this<br />

experiment. Two <strong>in</strong>dependent sets <strong>of</strong> data can be taken if the energies <strong>in</strong>duce scatter<strong>in</strong>g<br />

with different atomic scatter<strong>in</strong>g factors. This provides another consistency check for all the<br />

results.<br />

Results from Diffuse Scatter<strong>in</strong>g<br />

The systematic method described above was performed on the � = 4.2 nm sample.<br />

Only the diffuse scatter<strong>in</strong>g result is presented, as the other results are similar to those <strong>of</strong> the<br />

1 Goldman et al. [85] demonstrate how the specular data can be analyzed systematically.<br />

2 The rema<strong>in</strong><strong>in</strong>g pr<strong>of</strong>ile after the subtraction <strong>of</strong> a longitud<strong>in</strong>al diffuse scan from a specular reflectivity scan.


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 114<br />

q z<br />

0<br />

Specular scan<br />

Transverse scan<br />

0<br />

∆θ<br />

q x<br />

Longitud<strong>in</strong>al scan<br />

Figure 4.16: An illustration <strong>of</strong> a reciprocal space map along qx <strong>and</strong> qz. The specular ridge<br />

samples along qz, transverse scans sample along qx, <strong>and</strong> longitud<strong>in</strong>al scans sample both<br />

components.


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 115<br />

∆θ = 0°<br />

∆θ = 0.55°<br />

Figure 4.17: Specular reflectivity pr<strong>of</strong>iles (a) <strong>and</strong> longitud<strong>in</strong>al diffuse scatter<strong>in</strong>g pr<strong>of</strong>iles<br />

(b). The tilt <strong>of</strong> the sample from the specular condition is <strong>in</strong>dicated by �θ, <strong>and</strong> the x-ray<br />

spectra taken at E = 8020 eV <strong>and</strong> E = 8320 eV are <strong>of</strong>fset. The data <strong>and</strong> fitted/simulated<br />

pr<strong>of</strong>iles are given by the open circles <strong>and</strong> solid (or dashed) curves, respectively. The solid<br />

curves show the results <strong>of</strong> a simulation with a <strong>Ag</strong>/<strong>Ni</strong> (surface side / substrate side) chemical<br />

roughness <strong>of</strong> 0.24 nm <strong>and</strong> a <strong>Ni</strong>/<strong>Ag</strong> chemical roughness <strong>of</strong> 0.30 nm. The <strong>of</strong>fset dashed<br />

curves show the results <strong>of</strong> a simulation with a <strong>Ag</strong>/<strong>Ni</strong> chemical roughness <strong>of</strong> 0.30 nm <strong>and</strong> a<br />

<strong>Ni</strong>/<strong>Ag</strong> chemical roughness <strong>of</strong> 0.40 nm.


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 116<br />

other multilayers <strong>and</strong> are discussed <strong>in</strong> the next section.<br />

By simulat<strong>in</strong>g the diffuse scatter<strong>in</strong>g scans (Figure 4.17 (b)), it became evident that some<br />

chemical grad<strong>in</strong>g must be present at both <strong>in</strong>terfaces; very high order harmonics would exist<br />

otherwise, even with large physical roughness parameters. However, an upper bound can be<br />

determ<strong>in</strong>ed s<strong>in</strong>ce too much chemical roughness can elim<strong>in</strong>ate all the harmonics. The <strong>of</strong>fset<br />

dashed curves <strong>in</strong> Figure 4.17 (b) show the effect <strong>of</strong> extra chemical roughness added to both<br />

the <strong>Ag</strong>/<strong>Ni</strong> <strong>and</strong> <strong>Ni</strong>/<strong>Ag</strong> <strong>in</strong>terfaces. Specifically, the chemical roughness <strong>of</strong> <strong>Ag</strong>/<strong>Ni</strong> (surface<br />

side / substrate side) was changed from 0.24 nm to 0.30 nm, <strong>and</strong> the chemical roughness<br />

<strong>of</strong> <strong>Ni</strong>/<strong>Ag</strong> was changed from 0.30 nm to 0.40 nm. The additional chemical roughness <strong>in</strong> the<br />

latter simulation elim<strong>in</strong>ates the third harmonic <strong>in</strong> the 8020 eV scan. By perform<strong>in</strong>g several<br />

such simulations, we determ<strong>in</strong>ed an upper bound <strong>of</strong> σchemical ∼ 0.3 nm <strong>and</strong> a lower bound<br />

<strong>of</strong> σchemical ∼ 0.2 nm for both the <strong>Ag</strong>/<strong>Ni</strong> <strong>and</strong> <strong>Ni</strong>/<strong>Ag</strong> <strong>in</strong>terfaces.<br />

Ruud et al. [196] considered a situation <strong>in</strong> which two <strong>in</strong>terfacial {111} planes consist<br />

<strong>of</strong> <strong>Ni</strong> alloyed with <strong>Ag</strong>. With the use <strong>of</strong> Vegard’s law, the composition <strong>of</strong> these two planes<br />

produc<strong>in</strong>g the largest peak shift was calculated. This shift, however, corresponded to a<br />

change <strong>in</strong> the lattice parameter that was considerably smaller than the st<strong>and</strong>ard deviation <strong>of</strong><br />

the lattice parameter measurement. Our upper <strong>and</strong> lower bounds <strong>of</strong> σchemical <strong>in</strong>dicate that<br />

one to two <strong>in</strong>termixed {111} planes is reasonable; therefore, it is unlikely that <strong>in</strong>termix<strong>in</strong>g<br />

affects our measurement <strong>of</strong> the <strong>in</strong>terface stress.<br />

Accord<strong>in</strong>g to Josell et al. [120], even when the stra<strong>in</strong>-free lattice parameters extracted<br />

from the s<strong>in</strong> 2 ψ-method are used, the <strong>in</strong>terface stress rema<strong>in</strong>s compressive, with a value <strong>of</strong><br />

∼−0.75 N/m.<br />

The rema<strong>in</strong>der <strong>of</strong> this chapter is devoted to other rationales for a compressive <strong>in</strong>terface


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 117<br />

stress.<br />

4.7.2 Excess Volume / Misfit Dislocations<br />

We now return to some <strong>of</strong> the additional <strong>in</strong>formation supplied by SUPREX. The fits<br />

allow an extraction <strong>of</strong> the <strong>in</strong>terface width (Figure 4.18). We see that the width <strong>of</strong> the<br />

<strong>Ni</strong>/<strong>Ag</strong> <strong>in</strong>terface is close to the average {111} d-spac<strong>in</strong>g <strong>of</strong> <strong>Ni</strong> <strong>and</strong> <strong>Ag</strong>. This is expected<br />

for a coherent <strong>in</strong>terface. The <strong>Ag</strong>/<strong>Ni</strong> <strong>in</strong>terface width is larger, however, by approximately<br />

0.02 nm. This distance is related to the excess volume per unit area discussed earlier.<br />

The SUPREX simulations also show that stra<strong>in</strong> gradients are needed to fit the data.<br />

SUPREX allows the three planes adjacent to each side <strong>of</strong> the <strong>in</strong>terface to deviate from<br />

the “bulk” d-spac<strong>in</strong>g as shown schematically <strong>in</strong> Figure 4.19. The SUPREX fits for the<br />

� = 3.2 nm, � = 4.2 nm, <strong>and</strong> � = 5.9 nm multilayers all show the presence <strong>of</strong> steep<br />

d-spac<strong>in</strong>g gradients with<strong>in</strong> both the <strong>Ag</strong> <strong>and</strong> <strong>Ni</strong> layers. The multilayers with larger � have<br />

a significant overlap between the high-angle satellites <strong>and</strong> could not be fit. Figure 4.20<br />

shows the significance <strong>of</strong> the stra<strong>in</strong> gradients. A model disallow<strong>in</strong>g the gradients (the<br />

dotted curves) cannot reproduce the +3 satellite <strong>in</strong> the � = 3.2 nm sample or the +3<br />

<strong>and</strong> +4 satellite <strong>in</strong> the � = 5.9 nm sample while the model <strong>in</strong>clud<strong>in</strong>g the gradients fits the<br />

data well. S<strong>in</strong>ce our diffuse scatter<strong>in</strong>g results showed m<strong>in</strong>imal <strong>in</strong>terdiffusion, we assume<br />

the d-spac<strong>in</strong>g gradients to be stra<strong>in</strong> gradients rather than compositional gradients. This<br />

allows the construction <strong>of</strong> the d111 pr<strong>of</strong>iles <strong>in</strong> Figure 4.21.<br />

The d111 pr<strong>of</strong>iles resemble the results <strong>of</strong> Jaszczak et al., <strong>in</strong> which they simulated the<br />

stra<strong>in</strong> pr<strong>of</strong>iles <strong>of</strong> (100)-oriented FCC multilayers with as misfits <strong>of</strong> 10% [113] <strong>and</strong> 20%<br />

[114] us<strong>in</strong>g the Lennard-Jones potential (Figure 4.21 (c)). If we <strong>in</strong>terpret the <strong>Ag</strong>/<strong>Ni</strong> <strong>in</strong>ter-


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 118<br />

growth<br />

<strong>Ag</strong> (111)<br />

d 111, <strong>in</strong>terface<br />

<strong>Ni</strong> (111)<br />

Λ<br />

Λ<br />

(a)<br />

(b)<br />

Figure 4.18: (a) A schematic show<strong>in</strong>g the width between <strong>Ag</strong> <strong>and</strong> <strong>Ni</strong> (111) planes at<br />

the <strong>in</strong>terface. (b) The <strong>in</strong>terface width for <strong>Ag</strong>/<strong>Ni</strong> <strong>and</strong> <strong>Ni</strong>/<strong>Ag</strong> <strong>in</strong>terfaces. The mean (111)<br />

d-spac<strong>in</strong>g, 0.2197 nm, is <strong>in</strong>dicated by the horizontal l<strong>in</strong>e. (d111,<strong>Ag</strong> = 0.23592 nm, <strong>and</strong><br />

d111,<strong>Ni</strong> = 0.20345 nm).


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 119<br />

Nα<br />

∆d α2<br />

d α + ∆d α2e<br />

d α + ∆d α2e<br />

d α<br />

d α + ∆d α1e<br />

d α + ∆d α1e<br />

∆d α1<br />

Figure 4.19: Schematic show<strong>in</strong>g the method by which SUPREX treats gradients <strong>in</strong> the<br />

out-<strong>of</strong>-plane d-spac<strong>in</strong>g.<br />

Figure 4.20: High-angle x-ray diffraction peaks for the � = 3.2nm(a) <strong>and</strong> � = 5.9nm<br />

(b) <strong>Ag</strong>/<strong>Ni</strong> multilayers. The <strong>in</strong>tensity axis is on a logarithmic scale, <strong>and</strong> the data taken<br />

at E = 8320 eV is <strong>of</strong>fset above the E = 8020 eV data. The circles are the measured<br />

<strong>in</strong>tensities, the solid curve is the fit for the stra<strong>in</strong> gradient model, <strong>and</strong> the dotted curve is the<br />

best fit when stra<strong>in</strong> gradients are disallowed. The order <strong>of</strong> the reflection is <strong>in</strong>dicated above<br />

the peak.<br />

−ξ<br />

−2ξ<br />

−2ξ<br />

−ξ


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 120<br />

<strong>Ag</strong><br />

<strong>Ni</strong><br />

<strong>Ag</strong><br />

Si<br />

0.18<br />

0.20<br />

0.22<br />

d 111 [nm]<br />

0.24<br />

0.26<br />

<strong>Ag</strong><br />

<strong>Ni</strong><br />

<strong>Ag</strong><br />

0.18<br />

0.20<br />

0.22<br />

d 111 [nm]<br />

(a) (b)<br />

Si<br />

0.24<br />

0.26<br />

coherent<br />

0.14<br />

<strong>in</strong>coherent<br />

0.16<br />

0.18<br />

d 100 [nm]<br />

Figure 4.21: Calculated {111} d-spac<strong>in</strong>g pr<strong>of</strong>iles for the � = 3.2nm (a) <strong>and</strong> 5.9nm (b)<br />

<strong>Ag</strong>/<strong>Ni</strong> multilayers. The results <strong>of</strong> Jaszczak et al. for their theoretical multilayer are provided<br />

<strong>in</strong> (c) [113]. The vertical dashed l<strong>in</strong>es are the stra<strong>in</strong>-free d-spac<strong>in</strong>gs for the bulk phases.<br />

(c )<br />

0.20<br />

0.22


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 121<br />

faces us<strong>in</strong>g Jaszczak’s results, this suggests the <strong>Ni</strong>/<strong>Ag</strong> <strong>in</strong>terface is coherent while the <strong>Ag</strong>/<strong>Ni</strong><br />

is semicoherent.<br />

Accord<strong>in</strong>g to the orig<strong>in</strong>al Matthews-Blakeslee criterion, equation (3.9), there should<br />

exist no coherency between <strong>Ag</strong> <strong>and</strong> <strong>Ni</strong>; it predicts a coherent monolayer up to 9% mismatch<br />

[18]. However, Markov <strong>and</strong> Milchev [155, 147] have shown that add<strong>in</strong>g anharmonic terms<br />

to the potential allows coherency up to 13.5% mismatch <strong>in</strong> a one-dimensional system. This<br />

emphasizes the importance <strong>of</strong> the anharmonic terms.<br />

The asymmetry <strong>of</strong> the <strong>in</strong>terfaces most likely derives from the sign <strong>of</strong> the misfit [155,<br />

147]. <strong>Ni</strong>/<strong>Ag</strong> has a negative misfit with <strong>Ni</strong> hav<strong>in</strong>g to stretch to fit onto <strong>Ag</strong>, while <strong>Ag</strong>/<strong>Ni</strong> has<br />

a positive misfit with <strong>Ag</strong> hav<strong>in</strong>g to compress to fit onto the <strong>Ni</strong>. It should be easier to stretch<br />

than to compress because <strong>of</strong> the asymmetry <strong>of</strong> the potential well (Figure 1.1). Indeed,<br />

a room temperature growth experiment <strong>of</strong> <strong>Ni</strong> on <strong>Ag</strong>(111) probed by low energy electron<br />

diffraction (LEED) <strong>and</strong> Auger electron spectroscopy (AES) showed epitaxial growth up<br />

to six monolayers <strong>and</strong> no <strong>in</strong>termix<strong>in</strong>g [11]. On the other h<strong>and</strong>, a similar UHV-MBE ex-<br />

periment <strong>of</strong> <strong>Ag</strong> on <strong>Ni</strong>(100) have found negligible coherency [22] with the first layer <strong>of</strong><br />

<strong>Ag</strong> tend<strong>in</strong>g to close-pack<strong>in</strong>g. However, this may result from the competition between the<br />

surface energy <strong>and</strong> stra<strong>in</strong> energy; another study [157] found <strong>Ag</strong> could grow <strong>in</strong> an almost<br />

coherent fashion on <strong>Ni</strong>(111) up to six monolayers, but with the silver doma<strong>in</strong>s rotated with<br />

respect to the substrate by ±2 ◦ .


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 122<br />

Excess Volume<br />

There exists <strong>in</strong>direct evidence <strong>of</strong> excess volume 3 from elastic constant measurements.<br />

Carlotti et al. [31, 33, 32] found evidence <strong>of</strong> this <strong>in</strong> <strong>Ag</strong>/<strong>Ni</strong> multilayers, observ<strong>in</strong>g a reduc-<br />

tion <strong>of</strong> the effective shear elastic constant C44 by 40% when the bilayer thickness decreases<br />

from 18 to 2 nm us<strong>in</strong>g Brillou<strong>in</strong> light scatter<strong>in</strong>g. These results agree with the simulations <strong>of</strong><br />

Jaszczak et al. [113], who suggested that excess volume at the <strong>in</strong>terface can <strong>in</strong>duce a dra-<br />

matic s<strong>of</strong>ten<strong>in</strong>g <strong>of</strong> C44. This “s<strong>of</strong>ten<strong>in</strong>g” is not seen <strong>in</strong> multilayers <strong>in</strong> which the components<br />

are mutually soluble.<br />

Accord<strong>in</strong>g to Jaszczak et al. [113], the excess volume at semicoherent boundaries con-<br />

tracts the <strong>in</strong>-plane lattice parameter by the Poisson effect; this <strong>in</strong> turn leads to an expansion<br />

<strong>in</strong> the <strong>Ni</strong> <strong>and</strong> <strong>Ag</strong> planes adjacent to the <strong>in</strong>terfaces; these stra<strong>in</strong> gradients were seen <strong>in</strong> Fig-<br />

ure 4.21. Furthermore, this expansion can expla<strong>in</strong> the negative Poisson ratio found <strong>in</strong> these<br />

multilayers. <strong>Ni</strong> suffers a large expansion <strong>in</strong> the out-<strong>of</strong>-plane lattice parameter near the<br />

<strong>Ag</strong>/<strong>Ni</strong> <strong>in</strong>terface. This expansion dom<strong>in</strong>ates over the Poisson contraction from the <strong>in</strong>-plane<br />

tensile stress. 4<br />

Implicit <strong>in</strong> the argument <strong>of</strong> Clemens et al. [39] is that the excess volume, �V , is zero<br />

[119] or at least does not affect the out-<strong>of</strong>-plane d-spac<strong>in</strong>gs near the <strong>in</strong>terfaces. This clearly<br />

does not agree with our results.<br />

An excellent question is how the excess volume relates to the <strong>in</strong>terface energy <strong>and</strong><br />

3 Pelos<strong>in</strong> et al. [182, 183] have performed several anneal<strong>in</strong>g experiments that also <strong>in</strong>directly show the<br />

presence <strong>of</strong> excess volume <strong>in</strong> <strong>Ag</strong>/<strong>Ni</strong> multilayers. The measured thermal expansion coefficient <strong>of</strong> <strong>Ag</strong>/<strong>Ni</strong><br />

(19.1×10 −6 / ◦ C) decreases dramatically to zero when the sample is heated from room temperature to 375 ◦ C.<br />

This somewhat anomalous behavior can be expla<strong>in</strong>ed by densification at the <strong>in</strong>terfaces dur<strong>in</strong>g anneal<strong>in</strong>g. As a<br />

side note, they measured a biaxial modulus <strong>of</strong> 200 GPa [182], a Young modulus <strong>of</strong> 142 GPa [184], <strong>and</strong> noted<br />

very little dislocation plasticity <strong>in</strong> the multilayers [182].<br />

4 Other researchers have found negative Poisson ratios <strong>in</strong> <strong>Cu</strong> for Nb/<strong>Cu</strong> multilayers [59]. Nb/<strong>Cu</strong> also is<br />

similar to <strong>Ag</strong>/<strong>Ni</strong> <strong>in</strong> that it has large mismatch <strong>and</strong> low mutual solubilities. It is expected that a large excess<br />

volume also exists <strong>in</strong> Nb/<strong>Cu</strong> multilayers.


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 123<br />

stress. Through simulations, Wolf has shown that there exists an almost l<strong>in</strong>ear relationship<br />

between excess volume <strong>and</strong> the <strong>in</strong>terface energy [258, 259]. One may then expect that a<br />

compressive <strong>in</strong>terface stress can exp<strong>and</strong> the <strong>in</strong>terface plane, <strong>and</strong> by a Poisson effect, reduce<br />

the excess volume <strong>and</strong> hence energy. Equation (4.21) shows one relationship between<br />

excess volume <strong>and</strong> <strong>in</strong>terface stress, but further work along these l<strong>in</strong>es may able to show a<br />

more direct relationship for a biaxial stress state. However, there is the added complication<br />

<strong>of</strong> the more coherent <strong>Ni</strong>/<strong>Ag</strong> <strong>in</strong>terface affect<strong>in</strong>g the Gibbs-Duhem relationship.<br />

F<strong>in</strong>ally, an alternate approach to underst<strong>and</strong><strong>in</strong>g the relationship between excess volume<br />

<strong>and</strong> <strong>in</strong>terface stress can be made us<strong>in</strong>g a dislocation approach rather than a thermodynamic<br />

approach.<br />

Misfit Dislocations<br />

If we discount vacancy formation at the <strong>in</strong>terfaces, the most significant contribution to<br />

excess volume at the <strong>in</strong>terface is the misfit dislocation array. However, for a dislocation<br />

to have positive excess volume, anharmonic energy potentials are required [208]. We have<br />

already discussed the importance <strong>of</strong> anharmonic effects on the Matthews-Blakeslee crite-<br />

rion; this should be expected s<strong>in</strong>ce the dislocations represent a s<strong>in</strong>gularity <strong>in</strong> the stress field.<br />

Near the cores, the stra<strong>in</strong>s no longer obey Hookean elasticity.<br />

If we take the <strong>in</strong>terface energy to be positive, the problem <strong>of</strong> compressive <strong>in</strong>terface<br />

stress can be reduced to the consideration <strong>of</strong> why<br />

∂γc<br />

∂ɛ<br />

+ ∂γs<br />

∂ɛ<br />

< 0. (4.38)<br />

While the first term is not necessarily small, we shall assume the second term dom<strong>in</strong>ates.<br />

Further work needs to done to verify this assumption.


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 124<br />

Follow<strong>in</strong>g Cammarata et al. [30], we can estimate the contribution <strong>of</strong> the higher order<br />

terms <strong>in</strong> the potential energy function on the <strong>in</strong>terface stress, first by consider<strong>in</strong>g a one-<br />

dimensional array <strong>of</strong> dislocation cores, <strong>and</strong> then by us<strong>in</strong>g Matthews’ model <strong>of</strong> an edge<br />

dislocation array.<br />

Cammarata et al. [30] first assume a one term expansion <strong>of</strong> the harmonic potential:<br />

σ = Mcompɛ + M ′ comp ɛ2 . (4.39)<br />

The nonl<strong>in</strong>ear contribution to the stra<strong>in</strong> is equated to the volume expansion per unit length<br />

<strong>of</strong> a dislocation, which is approximately b 2 [208]. After determ<strong>in</strong><strong>in</strong>g the total stra<strong>in</strong> energy<br />

per unit volume <strong>of</strong> a dislocation l<strong>in</strong>e, they f<strong>in</strong>d<br />

� �<br />

δ − εcoh<br />

fcore ≈−b<br />

2<br />

or with ξ ≈ (δ − εcoh)/2 (cf. equation (4.4)),<br />

fcore ≈−<br />

Mcomp<br />

� Mcomp<br />

M ′ comp<br />

� 2<br />

� Mcomp<br />

M ′ comp<br />

� 2<br />

(4.40)<br />

γ. (4.41)<br />

The Matthews model <strong>of</strong> <strong>in</strong>terface energy is given by equation (4.4). We assume a<br />

completely relaxed <strong>in</strong>terface such that εcoh = 0. Us<strong>in</strong>g the Shuttleworth relationship (4.19),<br />

we take the derivative <strong>of</strong> γ with respect to the <strong>in</strong>terfacial stra<strong>in</strong>, ɛ11, <strong>and</strong> obta<strong>in</strong><br />

�<br />

f = bξ Mcomp 1 +<br />

ξ ′<br />

ξ + M′ comp<br />

+<br />

Mcomp<br />

b′<br />

�<br />

b<br />

(4.42)<br />

where ( ′ ) represents the derivative. After consider<strong>in</strong>g the major terms us<strong>in</strong>g a Born-Mayer<br />

potential, Cammarata et al. f<strong>in</strong>d that<br />

� �<br />

M ′<br />

comp<br />

f ≈ γ. (4.43)<br />

Mcomp


Chapter 4: The <strong>Ag</strong>/<strong>Ni</strong> {111} Interface Stress 125<br />

The higher order composite modulus, M ′ comp , is negative; hence the <strong>in</strong>terface stress is neg-<br />

ative for a positive γ . Although the core contribution from equation (4.41) can be small, it<br />

also results <strong>in</strong> a compressive <strong>in</strong>terface stress.<br />

4.8 Conclusions <strong>and</strong> Future Research<br />

The negative value <strong>of</strong> the <strong>Ag</strong>/<strong>Ni</strong> <strong>in</strong>terface stress has been confirmed us<strong>in</strong>g the bulk <strong>Ag</strong><br />

<strong>and</strong> <strong>Ni</strong> lattice parameters <strong>and</strong> is <strong>in</strong> agreement with previous results. Fits to low <strong>and</strong> high-<br />

angle x-ray scatter<strong>in</strong>g data show that <strong>in</strong>termix<strong>in</strong>g is unlikely to affect the measurement <strong>of</strong><br />

the <strong>Ag</strong>/<strong>Ni</strong> <strong>in</strong>terface stress. The high-angle results also demonstrate the existence <strong>of</strong> excess<br />

volume at the <strong>Ag</strong>/<strong>Ni</strong> <strong>in</strong>terface that causes an out-<strong>of</strong>-plane expansion <strong>in</strong> the adjacent layers<br />

<strong>and</strong> can result <strong>in</strong> a negative Poisson ratio. The excess volume orig<strong>in</strong>ates from the volume<br />

expansion <strong>of</strong> dislocation cores, thereby emphasiz<strong>in</strong>g the contribution <strong>of</strong> anharmonic effects<br />

to the <strong>in</strong>terface stress.<br />

Another way <strong>of</strong> exam<strong>in</strong><strong>in</strong>g asymmetric <strong>in</strong>terdiffusion was demonstrated by Huai et al.<br />

[105]. They analyzed the low <strong>and</strong> high-angle x-ray scans <strong>of</strong> Co/Re <strong>and</strong> Re/Co bilayers. Do-<br />

<strong>in</strong>g similar experiments with <strong>Ag</strong> <strong>and</strong> <strong>Ni</strong> may be <strong>of</strong> <strong>in</strong>terest. Also, experiments to measure<br />

the <strong>in</strong>terface stress after a low temperature heat treatment (to avoid multilayer destratifica-<br />

tion [84, 197, 201, 199]) may sharpen the <strong>in</strong>terfaces <strong>and</strong> ref<strong>in</strong>e the measurement.<br />

The added complications <strong>of</strong> <strong>in</strong>terface roughness effects [212, 180] <strong>and</strong> gra<strong>in</strong> boundary<br />

stresses [212, 16] on the <strong>in</strong>terface stress are beg<strong>in</strong>n<strong>in</strong>g to be studied, but the most important<br />

work that rema<strong>in</strong>s is a more rigorous calculation <strong>of</strong> the effect <strong>of</strong> an anharmonic potential on<br />

the <strong>in</strong>terface stress, possibly consider<strong>in</strong>g the effects <strong>of</strong> core delocalization <strong>and</strong> dislocation<br />

<strong>in</strong>teractions across the layer thickness.


Bibliography<br />

[1] C. W. Allen, H. Schroeder, <strong>and</strong> J. M. Hiller. In situ study <strong>of</strong> dislocation behavior<br />

<strong>in</strong> columnar Al th<strong>in</strong> films on Si substrate dur<strong>in</strong>g thermal cycl<strong>in</strong>g. In R. V<strong>in</strong>ci et al.,<br />

editors, Th<strong>in</strong>-<strong>Films</strong>: <strong>Stresses</strong> <strong>and</strong> Mechanical Properties VIII, volume 594 <strong>of</strong> Mater.<br />

Res. Soc. Symp. Proc., page 123, Warrendale, PA, 2000. Materials Research Society.<br />

[2] J. Als-<strong>Ni</strong>elsen <strong>and</strong> D. McMorrow. Elements <strong>of</strong> Modern X-Ray Physics. John Wiley<br />

& Sons, Ltd, New York, 2001.<br />

[3] A. S. Argon. Mechanical properties <strong>of</strong> s<strong>in</strong>gle-phase crystall<strong>in</strong>e media: Deformation<br />

at low temperature. Physical Metallurgy: Fourth, revised <strong>and</strong> enhanced edition,<br />

3:1877, 1996.<br />

[4] A. S. Argon. Mechanical properties <strong>of</strong> s<strong>in</strong>gle-phase crystall<strong>in</strong>e media: Deformation<br />

<strong>in</strong> the presence <strong>of</strong> diffusion. Physical Metallurgy: Fourth, revised <strong>and</strong> enhanced<br />

edition, 3:1957, 1996.<br />

[5] A. Arsenlis <strong>and</strong> D. M. Parks. Crystallographic aspects <strong>of</strong> geometrically-necessary<br />

<strong>and</strong> statistically-stored dislocation density. Acta Mater., 47:1597, 1999.<br />

[6] E. Arzt. Size effects <strong>in</strong> materials due to microstructural <strong>and</strong> dimensional constra<strong>in</strong>ts:<br />

A comparative review. Acta Mater., 46:5611, 1998.<br />

[7] M. F. Ashby. The deformation <strong>of</strong> plastically non-homogeneous materials. Philos.<br />

Mag., 21:399, 1970.<br />

[8] M. F. Ashby. The deformation <strong>of</strong> plastically non-homogeneous alloys. In A. Kelly<br />

<strong>and</strong> R. B. <strong>Ni</strong>cholson, editors, Strengthen<strong>in</strong>g Methods <strong>in</strong> Crystals, page 137, London,<br />

1971. Applied Science.<br />

[9] S. P. Baker, R.-M. Keller, <strong>and</strong> E. Arzt. Energy storage <strong>and</strong> recovery <strong>in</strong> th<strong>in</strong> metal<br />

films on substrates. In R. C. Cammarata, M. Nastasi, E. P. Busso, <strong>and</strong> W. C. Oliver,<br />

editors, Th<strong>in</strong>-<strong>Films</strong> — <strong>Stresses</strong> <strong>and</strong> Mechanical Properties VII, page 605, Warrendale,<br />

PA, 1998. Material Research Society.<br />

[10] S. P. Baker, A. Kretschmann, <strong>and</strong> E. Arzt. Thermomechanical behavior <strong>of</strong> different<br />

texture components <strong>in</strong> <strong>Cu</strong> th<strong>in</strong> films. Acta Mater., 49:2145, 2001.<br />

126


Bibliography 127<br />

[11] M.-G. Barthès. The first stages <strong>in</strong> the formation <strong>of</strong> ultrath<strong>in</strong> nickel layers on <strong>Ag</strong>(111):<br />

A low energy electron diffraction-Auger electron spectroscopy study. Th<strong>in</strong> Solid<br />

<strong>Films</strong>, 76:45, 1981.<br />

[12] R. Bechmann, R. F. S. Hearmon, <strong>and</strong> S. K. Kurtz. Elastic, piezoelectric, piezooptic,<br />

electrooptic constants, <strong>and</strong> nonl<strong>in</strong>ear dielectric susceptibilities <strong>of</strong> crystals, volume 2<br />

<strong>of</strong> L<strong>and</strong>olt-Börnste<strong>in</strong>: Numerical data <strong>and</strong> functional relationships <strong>in</strong> science <strong>and</strong><br />

technology. Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>, 1969.<br />

[13] S. Be<strong>in</strong>, C. Le Calvez, <strong>and</strong> J-L. Lebrun. Determ<strong>in</strong>ation <strong>of</strong> stress gradients by x-ray<br />

diffraction: Comparison <strong>of</strong> different methods <strong>and</strong> applications. Z. Metallkd., 89:289,<br />

1998.<br />

[14] S. Berger <strong>and</strong> F. Spaepen. The <strong>Ag</strong>/<strong>Cu</strong> <strong>in</strong>terface stress. Nanostruct. Mater., 6:201,<br />

1995.<br />

[15] B. A. Bilby, R. Bullough, <strong>and</strong> E. Smith. Cont<strong>in</strong>uous distributions <strong>of</strong> dislocations:<br />

a new application <strong>of</strong> the methods <strong>of</strong> non-Riemannian geometry. Proc. Roy. Soc. A,<br />

231:263, 1955.<br />

[16] R. Birr<strong>in</strong>ger <strong>and</strong> M. H<strong>of</strong>fmann. Interface stress <strong>in</strong> nanocrystall<strong>in</strong>e materials.<br />

[17] H. G. Bohn, M. Prieler, C. M. Su, H. Tr<strong>in</strong>kaus, <strong>and</strong> W. Schill<strong>in</strong>g. Internal friction<br />

effects due to gra<strong>in</strong> boundary slid<strong>in</strong>g <strong>in</strong> large- <strong>and</strong> small-gra<strong>in</strong>ed alum<strong>in</strong>ium. J. Phys.<br />

Chem. Solids, 55:1157, 1994.<br />

[18] B. C. Bold<strong>in</strong>g <strong>and</strong> E. A. Carter. Effect <strong>of</strong> stra<strong>in</strong> on th<strong>in</strong> film growth: Deposition <strong>of</strong><br />

<strong>Ni</strong> on <strong>Ag</strong>(100). Surf. Sci., 268:142, 1992.<br />

[19] W. Bollmann. Crystal Defects <strong>and</strong> Crystall<strong>in</strong>e Interfaces. Spr<strong>in</strong>ger–Verlag, New<br />

York, 1970.<br />

[20] R. Bonnet. Periodic elastic fields <strong>in</strong> anisotropic two-phase media. Application to<br />

<strong>in</strong>terfacial dislocations. Acta Metall., 29:437–445, 1981.<br />

[21] A. Bouzaher <strong>and</strong> R. Bonnet. Misfit dislocations arranged <strong>in</strong> a hexagonal network<br />

<strong>in</strong> anisotropic elasticity related displacement field <strong>and</strong> stored elastic energy. Acta<br />

Metall. Mater., 41:1595–1603, 1993.<br />

[22] A. Brodde, G. Wilhelmi, D. Badt, H. Wengelnik, <strong>and</strong> H. Neddermeyer. The growth<br />

<strong>of</strong> <strong>Ag</strong> films on <strong>Ni</strong>(100). J. Vac. Sci. Technol. B, 9:920, 1991.<br />

[23] K. Brugger. Thermodynamic def<strong>in</strong>ition <strong>of</strong> higher order elastic constants. Phys. Rev.,<br />

133:A1611, 1964.<br />

[24] H. J. Bunge. Texture Analysis <strong>in</strong> Materials Science. Butterworths, London, 1982.


Bibliography 128<br />

[25] W. K. Burton, N. Cabrera, <strong>and</strong> F. C. Frank. The growth <strong>of</strong> crystals <strong>and</strong> the equilibrium<br />

structure <strong>of</strong> their surfaces. Trans. Roy. Soc. A, 243:299, 1950.<br />

[26] J. W. Cahn. Thermodynamics <strong>of</strong> solid <strong>and</strong> fluid surfaces. In W. C. Johnson <strong>and</strong> J. M.<br />

Blakely, editors, Interfacial Segregation, page 3, Metals Park, OH, 1979. Am. Soc.<br />

Metals.<br />

[27] J. W. Cahn <strong>and</strong> J. E. Hilliard. Free energy <strong>of</strong> a nonuniform system. I. Interfacial free<br />

energy. J. Chem. Phys., 28:258, 1958.<br />

[28] J. W. Cahn <strong>and</strong> J. E. Hilliard. Free energy <strong>of</strong> a nonuniform system. III. Nucleation<br />

<strong>in</strong> a two-component <strong>in</strong>compressible fluid. J. Chem. Phys., 31:688, 1959.<br />

[29] R. C. Cammarata. Surface <strong>and</strong> <strong>in</strong>terface stress effects <strong>in</strong> th<strong>in</strong> films. Prog. Surf. Sci.,<br />

46:1, 1994.<br />

[30] R. C. Cammarata, K. Sieradzki, <strong>and</strong> F. Spaepen. Simple model for <strong>in</strong>terface stresses<br />

with application to misfit dislocation generation <strong>in</strong> epitaxial th<strong>in</strong> films. J. Appl.<br />

Phys., 87:1227, 2000.<br />

[31] G. Carlotti, D. Fioretto, G. Soc<strong>in</strong>o, B. Rodmacq, <strong>and</strong> V. Pelos<strong>in</strong>. Interface effects<br />

<strong>and</strong> elastic constants <strong>of</strong> <strong>Ag</strong>/<strong>Ni</strong> superlattices studied by Brillou<strong>in</strong> scatter<strong>in</strong>g. J. Appl.<br />

Phys., 71:4897, 1992.<br />

[32] G. Carlotti, D. Fioretto, G. Soc<strong>in</strong>o, L. Verd<strong>in</strong>i, <strong>and</strong> V. Pelos<strong>in</strong>. Effect <strong>of</strong> anneal<strong>in</strong>g<br />

on the elastic properties <strong>of</strong> <strong>Ag</strong>/<strong>Ni</strong> superlattices. J. Appl. Phys., 73:3028, 1993.<br />

[33] G. Carlotti, A. Montone, C. Petrillo, <strong>and</strong> M. Vittori Antisari. Relationship between<br />

structural properties <strong>and</strong> elastic anomalies <strong>in</strong> <strong>Ag</strong>/<strong>Ni</strong> superlattices. J. Phys.: Condens.<br />

Matter, 5:4611, 1993.<br />

[34] P. Chaudhari. Mechanisms <strong>of</strong> stress relief <strong>in</strong> polycrystall<strong>in</strong>e films. IBM J. Res.<br />

Develop., 13:197, 1969.<br />

[35] P. Chaudhari. Gra<strong>in</strong> growth <strong>and</strong> stress relief <strong>in</strong> th<strong>in</strong> films. J. Vac. Sci. Technol.,<br />

9:520, 1972.<br />

[36] P. Chaudhari. Plastic properties <strong>of</strong> polycrystall<strong>in</strong>e th<strong>in</strong> films on a substrate. Philos.<br />

Mag. A, 39:507, 1979.<br />

[37] P. Chaudhari, S. Mader, <strong>and</strong> J. F. Freedman. Mechanisms <strong>of</strong> stress relief <strong>in</strong> th<strong>in</strong><br />

films. J. Vac. Sci. Technol., 6:618, 1969.<br />

[38] D. Chocyk, G. Gladyszewski, T. Pienkos, <strong>and</strong> L. Gladyszewski. Stra<strong>in</strong> <strong>in</strong> multilayered<br />

systems <strong>and</strong> its <strong>in</strong>fluence on asymmetric x-ray diffraction pr<strong>of</strong>iles. Electron<br />

Tech., 33:329, 2000.


Bibliography 129<br />

[39] B. M. Clemens, W. D. <strong>Ni</strong>x, <strong>and</strong> V. Ramaswamy. Surface-energy-driven <strong>in</strong>termix<strong>in</strong>g<br />

<strong>and</strong> its effect on the measurement <strong>of</strong> <strong>in</strong>terface stress. J. Appl. Phys., 87:2816, 2000.<br />

[40] J. B. Cohen. Report on tungsten lattice parameter round rob<strong>in</strong>. Technical report, X-<br />

Ray Subcommittee <strong>of</strong> SAE Iron <strong>and</strong> Steel Technical Committee, Division 4, 1964.<br />

[41] F. Convert <strong>and</strong> B. Miege. The control <strong>of</strong> geometrical sources <strong>of</strong> error <strong>in</strong> x-ray diffraction<br />

applied to stress analysis. J. Appl. Cryst., 25:384, 1992.<br />

[42] A. H. Cottrell. Dislocations <strong>and</strong> Plastic Flow <strong>in</strong> Crystals. Clarendon Press, Oxford,<br />

1953.<br />

[43] A. H. Cottrell. The Mechanical Properties <strong>of</strong> Matter. John Wiley & Sons, Inc., New<br />

York, 1964.<br />

[44] B. D. <strong>Cu</strong>llity. Elements <strong>of</strong> X-Ray Diffraction. Addison-Wesley Pub., Read<strong>in</strong>g, MA,<br />

2d edition, 1978.<br />

[45] M. S. Daw <strong>and</strong> M. I. Baskes. Embedded-atom method: Derivation <strong>and</strong> application<br />

to impurities, surfaces, <strong>and</strong> other defects <strong>in</strong> metals. Phys. Rev. B, 29:6443, 1984.<br />

[46] M. M. de Lima, Jr., R. G. Lacerda, J. Vilcarromero, <strong>and</strong> F. C. Marques. Coefficient<br />

<strong>of</strong> thermal expansion <strong>and</strong> elastic modulus <strong>of</strong> th<strong>in</strong> films. J. Appl. Phys., 86:4936,<br />

1999.<br />

[47] G. Dehm <strong>and</strong> E. Arzt. In situ transmission electron microscopy study <strong>of</strong> dislocations<br />

<strong>in</strong> a polycrystall<strong>in</strong>e <strong>Cu</strong> th<strong>in</strong> film constra<strong>in</strong>ed by a substrate. Appl. Phys. Lett.,<br />

77:1126, 2000.<br />

[48] G. Dehm, D. Weiss, <strong>and</strong> E. Arzt. In situ transmission electron microscopy study <strong>of</strong><br />

thermal-stress-<strong>in</strong>duced dislocations <strong>in</strong> a th<strong>in</strong> <strong>Cu</strong> film constra<strong>in</strong>ed by a Si substrate.<br />

Mater. Sci. Eng. A, 309-310:468, 2001.<br />

[49] A. L. Del Vecchio. PhD thesis, <strong>Harvard</strong> University, 2003.<br />

[50] F. M. D’Heurle. Alum<strong>in</strong>um films deposited by RF sputter<strong>in</strong>g. Metall. Trans., 1:725,<br />

1970.<br />

[51] F. M. D’Heurle <strong>and</strong> J. M. E. Harper. Note on the orig<strong>in</strong>s <strong>of</strong> <strong>in</strong>tr<strong>in</strong>sic stresses <strong>in</strong> films<br />

deposited via evaporation <strong>and</strong> sputter<strong>in</strong>g. Th<strong>in</strong> Solid <strong>Films</strong>, 171:81, 1989.<br />

[52] M. F. Doerner <strong>and</strong> S. Brennan. Stra<strong>in</strong> distribution <strong>in</strong> th<strong>in</strong> alum<strong>in</strong>um films us<strong>in</strong>g x-ray<br />

depth pr<strong>of</strong>il<strong>in</strong>g. J. Appl. Phys., 63:126, 1987.<br />

[53] M. F. Doerner, D. S. Gardner, <strong>and</strong> W. D. <strong>Ni</strong>x. Plastic properties <strong>of</strong> th<strong>in</strong> films on<br />

substrates as measured by submicron <strong>in</strong>dentation hardness <strong>and</strong> substrate curvature<br />

techniques. J. Mater. Res., 1:845, 1987.


Bibliography 130<br />

[54] M. F. Doerner <strong>and</strong> W. D. <strong>Ni</strong>x. <strong>Stresses</strong> <strong>and</strong> deformation processes <strong>in</strong> th<strong>in</strong> films on<br />

substrates. CRC Crit. Rev. Solid St. Mater. Sci., 14:225, 1988.<br />

[55] S. A. Dregia. Equilibrium Segregation <strong>and</strong> Dislocation Structures at Epitaxial Interphase<br />

Boundaries. PhD thesis, Carnegie-Mellon University, 1989.<br />

[56] J. D. Embury <strong>and</strong> J. P. Hirth. On dislocation storage <strong>and</strong> the mechanical response <strong>of</strong><br />

f<strong>in</strong>e scale microstructures. Acta Metall. Mater., 42:2051, 1994.<br />

[57] J. D. Eshelby, F. C. Frank, <strong>and</strong> F. R. N. Nabarro. The equilibrium <strong>of</strong> l<strong>in</strong>ear arrays <strong>of</strong><br />

dislocations. Philos. Mag., 42:351, 1951.<br />

[58] Y. Estr<strong>in</strong>, G. Gottste<strong>in</strong>, E. Rabk<strong>in</strong>, <strong>and</strong> L. S. Shv<strong>in</strong>dlerman. Gra<strong>in</strong> growth <strong>in</strong> th<strong>in</strong><br />

metallic films. Acta Mater., 49:673, 2001.<br />

[59] A. Fartash, M. Grimsditch, E. E. Fullerton, <strong>and</strong> I. K. Schuller. Breakdown <strong>of</strong> Poisson’s<br />

effect <strong>in</strong> Nb/<strong>Cu</strong> superlattices. Phys. Rev. B, 47:12813, 1993.<br />

[60] P. F. Fewster <strong>and</strong> N. L. Andrew. Absolute lattice-parameter measurement. J. Appl.<br />

Cryst., 28:451, 1995.<br />

[61] N. A. Fleck <strong>and</strong> J. W. Hutch<strong>in</strong>son. Stra<strong>in</strong> gradient plasticity. Adv. Appl. Mech.,<br />

33:295, 1997.<br />

[62] P. A. Fl<strong>in</strong>n, D. S. Gardner, <strong>and</strong> W. D. <strong>Ni</strong>x. Measurement <strong>and</strong> <strong>in</strong>terpretation <strong>of</strong> stress<br />

<strong>in</strong> alum<strong>in</strong>um-based metallization as a function <strong>of</strong> thermal history. IEEE Trans. Electron.<br />

Dev., 34:689, 1987.<br />

[63] F. C. Frank <strong>and</strong> J. H. van der Merwe. One-dimensional dislocations. I. Static theory.<br />

Proc. Roy. Soc. London, A198:205, 1949.<br />

[64] F. C. Frank <strong>and</strong> J. H. van der Merwe. One-dimensional dislocations. II. Misfitt<strong>in</strong>g<br />

monolayers <strong>and</strong> oriented overgrowth. Proc. Roy. Soc. London, A198:216, 1949.<br />

[65] L. B. Freund. The stability <strong>of</strong> a dislocation thread<strong>in</strong>g <strong>of</strong> a stra<strong>in</strong>ed layer on a substrate.<br />

J. Appl. Mech., 54:553, 1987.<br />

[66] L. B. Freund. Some elemetary connections between curvature <strong>and</strong> mismatch stra<strong>in</strong><br />

<strong>in</strong> compositionally graded th<strong>in</strong> films. J. Mech. Phys. Solids, 44:723, 1996.<br />

[67] L. B. Freund, J. A. Floro, <strong>and</strong> E. Chason. Extensions <strong>of</strong> the Stoney formula for substrate<br />

curvature to configurations with th<strong>in</strong> substrates or large deformations. Appl.<br />

Phys. Lett., 74:1987, 1999.<br />

[68] J. Friedel, C. Boulanger, <strong>and</strong> C. Crussard. Constantes elastiques et frottement <strong>in</strong>terieur<br />

de l’alum<strong>in</strong>ium polygonise. Acta Metall., 3:380, 1955.


Bibliography 131<br />

[69] H. J. Frost. Deformation mechanisms <strong>in</strong> th<strong>in</strong> films. In C. V. Thompson <strong>and</strong> J. R.<br />

Lloyd, editors, Materials Reliability <strong>in</strong> Microelectronics II, volume 265, page 3,<br />

Pittsburgh, PA, 1992. Mater. Res. Soc. Symp.<br />

[70] E. E. Fullerton, I. K. Schuller, H. V<strong>and</strong>erstraeten, <strong>and</strong> Y. Bruynseraede. Structural<br />

ref<strong>in</strong>ement <strong>of</strong> superlattices from x-ray diffraction. Phys. Rev. B, 45:9292, 1991.<br />

[71] H. Gao, Y. Huang, <strong>and</strong> W. D. <strong>Ni</strong>x. Model<strong>in</strong>g plasticity at the micrometer scale.<br />

Naturwissenschaften, 86:507, 1999.<br />

[72] H. Gao, L. Zhang, <strong>and</strong> S. P. Baker. Dislocation core spread<strong>in</strong>g at <strong>in</strong>terfaces between<br />

crystall<strong>in</strong>e <strong>and</strong> amorphous solids. Mater. Res. Soc. Symp. Proc., 673:P6.6.1, 2001.<br />

[73] H. Gao, L. Zhang, W. D. <strong>Ni</strong>x, C. V. Thompson, <strong>and</strong> E. Arzt. Crack-like gra<strong>in</strong>boundary<br />

diffusion wedges <strong>in</strong> th<strong>in</strong> metal films. Acta Mater., 47(10):2865–2878,<br />

1999.<br />

[74] Y. Gao, S. A. Dregia, <strong>and</strong> P. G. Shewmon. Energy <strong>and</strong> structure <strong>of</strong> the (001) twist<br />

<strong>in</strong>terphase boundaries <strong>in</strong> the <strong>Ag</strong>/<strong>Ni</strong> system. Acta Metall., 37:1627, 1989.<br />

[75] Y. Gao <strong>and</strong> K. L. Merkle. High-resolution electron microscopy <strong>of</strong> metal/metal <strong>and</strong><br />

metal/metal-oxide <strong>in</strong>terfaces <strong>in</strong> the <strong>Ag</strong>/<strong>Ni</strong> <strong>and</strong> Au/<strong>Ni</strong> systems. J. Mater. Res., 5:1995,<br />

1990.<br />

[76] Y. Gao, P. G. Shewmon, <strong>and</strong> S. A. Dregia. Investigation <strong>of</strong> low energy <strong>in</strong>terphase<br />

boundaries <strong>in</strong> <strong>Ag</strong>/<strong>Ni</strong> by computer simulation <strong>and</strong> crystallite rotation. Acta Metall.,<br />

37:3165, 1989.<br />

[77] C. Genzel. A self-consistent method for x-ray diffraction analysis <strong>of</strong> multiaxial<br />

residual-stress fields <strong>in</strong> the near-surface region <strong>of</strong> polycrystall<strong>in</strong>e materials. I. Theoretical<br />

concept. J. Appl. Cryst., 32:770, 1999.<br />

[78] C. Genzel <strong>and</strong> W. Reimers. Some new aspects <strong>in</strong> x-ray stress analysis <strong>of</strong> th<strong>in</strong> layers.<br />

Surf. Coat. Technol., 116-119:404, 1999.<br />

[79] Ch. Genzel. Formalism for the evaluation <strong>of</strong> strongly non-l<strong>in</strong>ear surface stress fields<br />

by x-ray diffraction performed <strong>in</strong> the scatter<strong>in</strong>g vector mode. Phys. Stat. Sol. A,<br />

146:629, 1994.<br />

[80] Ch. Genzel. A study <strong>of</strong> x-ray residual stress gradient analysis <strong>in</strong> th<strong>in</strong> layers with<br />

strong fibre texture I. Evaluation <strong>of</strong> the stress factors Fij. Phys. Stat. Sol. A, 165:347,<br />

1998.<br />

[81] S. Geremia, H. Burzlaff, <strong>and</strong> W. Rothammel. A method <strong>of</strong> alignment <strong>of</strong> a four-circle<br />

diffractometer us<strong>in</strong>g a small collimator <strong>in</strong> the center <strong>of</strong> the Eulerian cradle. J. Appl.<br />

Cryst., 27(6):1061–3, 1994.


Bibliography 132<br />

[82] P. Gergaud, S. Labat, <strong>and</strong> O. Thomas. Limits <strong>of</strong> validity <strong>of</strong> the crystallite group<br />

method <strong>in</strong> stress determ<strong>in</strong>ation <strong>of</strong> th<strong>in</strong> film structures. Th<strong>in</strong> Solid <strong>Films</strong>, 319:9, 1998.<br />

[83] P. Gergaud, H. Yang, C. Pélissonnier-Grosjean, A. J. Bottger, P. S<strong>and</strong>ström, J. E.<br />

Sundgren, <strong>and</strong> O. Thomas. Internal stress <strong>in</strong> sputtered silver nickel th<strong>in</strong> films <strong>and</strong><br />

multilayers: Sputter<strong>in</strong>g pressure <strong>and</strong> thickness effects. In D. E. Laughl<strong>in</strong>, K. P.<br />

Rodbell, O. Thomas, <strong>and</strong> B. Zhang, editors, Polycrystall<strong>in</strong>e Metal <strong>and</strong> Magnetic<br />

Th<strong>in</strong> <strong>Films</strong>, page 123, Warrendale, PA, 2000. Materials Research Society.<br />

[84] V. Ghetta, J. Senateur, J. La Manna, G. Delabouglise, I. Matko, B. Chenevier,<br />

P. S<strong>and</strong>strom, <strong>and</strong> J.-E. Sundgren. Temperature dependence <strong>of</strong> <strong>Ag</strong>/<strong>Ni</strong> multilayer<br />

microstructure studied by <strong>in</strong> situ x-ray diffraction. In D. E. Laughl<strong>in</strong>, K. P. Rodbell,<br />

O. Thomas, <strong>and</strong> B. Zhang, editors, Polycrystall<strong>in</strong>e Metal <strong>and</strong> Magnetic Th<strong>in</strong> <strong>Films</strong>,<br />

page 195, Warrendale, PA, 2000. Materials Research Society.<br />

[85] L. M. Goldman, H. A. Atwater, <strong>and</strong> F. Spaepen. A systematic method for extract<strong>in</strong>g<br />

structural parameters from low angle x-ray reflectivity measurements on multilayers.<br />

In B. W. Dodson, L. J. Schowalter, J. E. <strong>Cu</strong>nn<strong>in</strong>gham, <strong>and</strong> F. H. Pollak, editors, Layered<br />

Structures — Heteroepitaxy, Superlattices, Stra<strong>in</strong>, <strong>and</strong> Metastability, volume<br />

160, page 577, Pittsburgh, PA, 1990. Mater. Res. Soc. Symp. Proc.<br />

[86] T. Goto, O. L. Anderson, I. Ohno, <strong>and</strong> S. Yamamoto. Elastic constants <strong>of</strong> corundum<br />

up to 1825 K. J. Geophys. Res., 94:7588, 1989.<br />

[87] P. Gumbsch. Atomistic study <strong>of</strong> misfit accommodation <strong>in</strong> cube-on-cube oriented<br />

<strong>Ag</strong>/<strong>Ni</strong> heterophase boundaries. Z. Metallkd., 83:500, 1992.<br />

[88] P. Gumbsch <strong>and</strong> M. S. Daw. Interface stresses <strong>and</strong> their effects on the elastic moduli<br />

<strong>of</strong> metallic multilayers. Phys. Rev. B, 44:3934, 1991.<br />

[89] P. Gumbsch, M. S. Daw, S. M. Foiles, <strong>and</strong> H. F. Fischmeister. Accomodation <strong>of</strong><br />

lattice mismatch <strong>in</strong> a <strong>Ag</strong>/<strong>Ni</strong> heterophase boundary. Phys. Rev. B, 43:13833, 1991.<br />

[90] P. Gumbsch <strong>and</strong> H. F. Fischmeister. The accomodation <strong>of</strong> lattice mismatch on the<br />

(111) <strong>in</strong>terphase boundary plane <strong>in</strong> FCC metals. Mater. Res. Symp. Proc., 209:59,<br />

1991.<br />

[91] M. E. Gurt<strong>in</strong>. An Introduction to Cont<strong>in</strong>uum Mechanics, volume 158 <strong>of</strong> Mathematics<br />

<strong>in</strong> Science <strong>and</strong> Eng<strong>in</strong>eer<strong>in</strong>g. Academic Press, Inc., New York, 1981.<br />

[92] E. O. Hall. The deformation <strong>and</strong> age<strong>in</strong>g <strong>of</strong> mild steel: III Discussion <strong>of</strong> results. Proc.<br />

Phys. Soc. B, 64:747, 1951.<br />

[93] Z. Hash<strong>in</strong> <strong>and</strong> S. Shtrikman. On some variational pr<strong>in</strong>ciples <strong>in</strong> anisotropic <strong>and</strong> nonhomogeneous<br />

elasticity. J. Mech. Phys. Solids, 10:335, 1962.


Bibliography 133<br />

[94] Z. Hash<strong>in</strong> <strong>and</strong> S. Shtrikman. A variational approach to the theory <strong>of</strong> the elastic<br />

behaviour <strong>of</strong> polycrystals. J. Mech. Phys. Solids, 10:343, 1962.<br />

[95] V. Hauk. Structural <strong>and</strong> Residual Stress Analysis by Nondestructive Methods. Elsevier,<br />

Amsterdam, 1997.<br />

[96] A. K. Head. The <strong>in</strong>teraction <strong>of</strong> dislocations <strong>and</strong> boundaries. Philos. Mag., 44:92,<br />

1953.<br />

[97] D. He<strong>in</strong>en, H. G. Bohn, <strong>and</strong> W. Schill<strong>in</strong>g. Internal friction <strong>in</strong> free-st<strong>and</strong><strong>in</strong>g th<strong>in</strong> Al<br />

films. J. Appl. Phys., 78:893, 1995.<br />

[98] D. He<strong>in</strong>en, H. G. Bohn, <strong>and</strong> W. Schill<strong>in</strong>g. On the mechanical strength <strong>of</strong> freest<strong>and</strong><strong>in</strong>g<br />

<strong>and</strong> substrate-bonded Al th<strong>in</strong> films. J. Appl. Phys., 77:3742, 1995.<br />

[99] R. Hill. The Mathematical Theory <strong>of</strong> Plasticity. Oxford University Press, Oxford,<br />

1950.<br />

[100] R. Hill. The elastic behavior <strong>of</strong> a crystall<strong>in</strong>e aggregate. Proc. Phys. Soc., 65A:350,<br />

1952.<br />

[101] J. P. Hirth <strong>and</strong> A. G. Evans. Damage <strong>of</strong> coherent multilayer structures by <strong>in</strong>jection<br />

<strong>of</strong> dislocations or cracks. J. Appl. Phys., 60:2372, 1986.<br />

[102] J. Hornstra <strong>and</strong> W. J. Bartels. Determ<strong>in</strong>ation <strong>of</strong> the lattice constant <strong>of</strong> epitaxial layers<br />

<strong>of</strong> III-V compounds. J. Cryst. Growth, 44:513, 1978.<br />

[103] C. R. Houska. The <strong>in</strong>vestigation <strong>of</strong> composition variations by diffraction, volume<br />

19A <strong>of</strong> Treatise on Materials Science <strong>and</strong> Technology, chapter 2. Academic Press,<br />

Inc., 1980.<br />

[104] J. M. Howe. Interfaces <strong>in</strong> Materials. John Wiley & Sons, Inc., New York, NY, 1997.<br />

[105] Y. Huai, R. W. Cochrane, <strong>and</strong> M. Sutton. X-ray diffraction studies <strong>of</strong> Co/Re superlattices.<br />

Phys. Rev. B, 48:2568, 1993.<br />

[106] H. Huang. Mechanical Properties <strong>of</strong> Free-st<strong>and</strong><strong>in</strong>g Polycrystall<strong>in</strong>e Metallic Th<strong>in</strong><br />

<strong>Films</strong> <strong>and</strong> <strong>Multilayers</strong>. PhD thesis, <strong>Harvard</strong> University, 1998.<br />

[107] H. Huang <strong>and</strong> F. Spaepen. Tensile test<strong>in</strong>g <strong>of</strong> free-st<strong>and</strong><strong>in</strong>g <strong>Cu</strong>, <strong>Ag</strong> <strong>and</strong> Al th<strong>in</strong> films<br />

<strong>and</strong> <strong>Ag</strong>/<strong>Cu</strong> multilayers. Acta Mater., 48:3261, 2000.<br />

[108] H. B. Hunt<strong>in</strong>gton. The elastic constants <strong>of</strong> crystals. In F. Seitz <strong>and</strong> D. Turnbull,<br />

editors, Solid State Physics, volume 7, page 213. Academic Press, New York, 1958.<br />

[109] J. W. Hutch<strong>in</strong>son. Thickness dependence <strong>of</strong> stress <strong>in</strong> th<strong>in</strong> films from stra<strong>in</strong> gradient<br />

plasticity. unpublished.


Bibliography 134<br />

[110] M. S. Jackson <strong>and</strong> C.-Y. Li. Stress relaxation <strong>and</strong> hillock growth <strong>in</strong> th<strong>in</strong> films. Acta<br />

Metall., 30:1993, 1982.<br />

[111] M. R. James. The use <strong>of</strong> oscillation on PSD-based <strong>in</strong>struments for x-ray measurement<br />

<strong>of</strong> residual stress. Expt. Mech., 27:164, 1987.<br />

[112] M. R. James <strong>and</strong> J. B. Cohen. Geometrical problems with a position-sensitive detector<br />

employed on a diffractometer, <strong>in</strong>clud<strong>in</strong>g its use <strong>in</strong> the measurement <strong>of</strong> stress.<br />

J. Appl. Cryst., 12:339, 1979.<br />

[113] J. A. Jaszczak, S. R. Phillpot, <strong>and</strong> D. Wolf. Role <strong>of</strong> coherency <strong>in</strong> the elastic behavior<br />

<strong>of</strong> composition-modulated superlattices. J. Appl. Phys., 68:4573, 1990.<br />

[114] J. A. Jaszczak <strong>and</strong> D. Wolf. On the elastic behavior <strong>of</strong> composition-modulated superlattices.<br />

J. Mater. Res., 6:1207, 1991.<br />

[115] JCPDS – International Center for Diffraction Data. Card Nos. 04-0850 (<strong>Ni</strong>), 04-0783<br />

(<strong>Ag</strong>), 04-0836 (<strong>Cu</strong>), 1995.<br />

[116] W. C. Johnson. Superficial stress <strong>and</strong> stra<strong>in</strong> at coherent <strong>in</strong>terfaces. Acta Mater.,<br />

48:433, 2000.<br />

[117] W. C. Johnson <strong>and</strong> H. Schmalzried. Gibbs-Duhem <strong>and</strong> Clausius-Clapeyron type<br />

equations for elastically stressed crystals. Acta Metall. Mater., 40:2337–2342, 1992.<br />

[118] W. C. Johnson <strong>and</strong> P. W. Voorhees. Interfacial stress, <strong>in</strong>terfacial energy, <strong>and</strong> phase<br />

equilibria <strong>in</strong> b<strong>in</strong>ary alloys. J. Stat. Phys., 95:1281, 1999.<br />

[119] D. Josell. The impact <strong>of</strong> <strong>in</strong>termix<strong>in</strong>g on <strong>in</strong>terface stress determ<strong>in</strong>ation from studies<br />

<strong>of</strong> multilayer materials.<br />

[120] D. Josell, J. E. Bonevich, I. Shao, <strong>and</strong> R. C. Cammarata. Measur<strong>in</strong>g the <strong>in</strong>terface<br />

stress: Silver/nickel <strong>in</strong>terfaces. J. Mater. Res., 14:4358, 1999.<br />

[121] D. Josell <strong>and</strong> F. Spaepen. Determ<strong>in</strong>ation <strong>of</strong> the <strong>in</strong>terfacial tension by zero creep<br />

experiments on multilayers — II. Experiment. Acta Metall. Mater., 41:3017, 1993.<br />

[122] A. J. Kalkman, A. H. Verbruggen, <strong>and</strong> G. C. A. M. Janssen. Young’s modulus<br />

measurements <strong>and</strong> gra<strong>in</strong> boundary slid<strong>in</strong>g <strong>in</strong> free-st<strong>and</strong><strong>in</strong>g th<strong>in</strong> metal films. Appl.<br />

Phys. Lett., 78:2673, 2001.<br />

[123] J.-D. Kamm<strong>in</strong>ga, Th. H. de Keijser, E. J. Mittemeijer, <strong>and</strong> R. Delhez. New methods<br />

for diffraction stress measurement: A critical evaluation <strong>of</strong> new <strong>and</strong> exist<strong>in</strong>g<br />

methods. J. Appl. Cryst., 33:1059, 2000.


Bibliography 135<br />

[124] R.-M. Keller, S. P. Baker, <strong>and</strong> E. Arzt. Quantitative analysis <strong>of</strong> strengthen<strong>in</strong>g mechanisms<br />

<strong>in</strong> th<strong>in</strong> <strong>Cu</strong> films: Effects <strong>of</strong> film thickness, gra<strong>in</strong> size, <strong>and</strong> passivation. J.<br />

Mater. Res., 13:1307, 1998.<br />

[125] R.-M. Keller, S. P. Baker, <strong>and</strong> E. Arzt. Stress-temperature behavior <strong>of</strong> unpassivated<br />

th<strong>in</strong> copper films. Acta Mater., 47:415, 1999.<br />

[126] D. B. Knorr <strong>and</strong> D. P. Tracy. A review <strong>of</strong> microstructure <strong>in</strong> vapor deposited copper<br />

th<strong>in</strong> films. Mater. Chem. Phys., 41:206, 1995.<br />

[127] G. Knuyt, C. Quaeyhaegens, J. D’Haen, <strong>and</strong> L. M. Stals. A model for texture evolution<br />

<strong>in</strong> a grow<strong>in</strong>g film. In D. J. Srolovitz, C. A. Volkert, M. J. Fluss, <strong>and</strong> R. J. Kee,<br />

editors, Model<strong>in</strong>g <strong>and</strong> Simulation <strong>of</strong> Th<strong>in</strong>-Film Process<strong>in</strong>g, volume 389, page 137,<br />

Pittsburgh, PA, 1995. Mater. Res. Soc.<br />

[128] M. J. Kobr<strong>in</strong>sky <strong>and</strong> C. V. Thompson. Activation volume for <strong>in</strong>elastic deformation<br />

<strong>in</strong> polycrystall<strong>in</strong>e <strong>Ag</strong> th<strong>in</strong> films. Acta Mater., 48:625, 2000.<br />

[129] R. Koch. The <strong>in</strong>tr<strong>in</strong>sic stress <strong>of</strong> polycrystall<strong>in</strong>e <strong>and</strong> epitaxial th<strong>in</strong> metal films. J.<br />

Phys. Condens. Matter, 6:9519, 1994.<br />

[130] U. F. Kocks, A. S. Argon, <strong>and</strong> M. F. Ashby. Thermodynamics <strong>and</strong> k<strong>in</strong>etics <strong>of</strong> slip.<br />

Prog. Mater. Sci., 19:1, 1975.<br />

[131] U.F. Kocks, C.N. Tome, <strong>and</strong> H.-R. Wenk. Texture <strong>and</strong> Anisotropy : Preferred Orientations<br />

<strong>in</strong> Polycrystals <strong>and</strong> Their Effect on Materials Properties. Cambridge University<br />

Press, New York, 1998.<br />

[132] E. Kröner. The <strong>in</strong>ternal mechanical state <strong>of</strong> solids with defects. Int. J. Solids Structures,<br />

29:1849, 1992.<br />

[133] E. Kröner. Dislocation theory as a physical field theory. Meccanica, 31:577, 1996.<br />

[134] D. Kuhlmann-Wilsdorf. Questions you always wanted (or should have wanted) to<br />

ask about workharden<strong>in</strong>g. Mater. Res. Innovat., 1:265, 1998.<br />

[135] W.-M. Kuschke, A. Kretschmann, R.-M. Keller, R. P. V<strong>in</strong>ci, C. Kaufmann, <strong>and</strong><br />

E. Arzt. Textures <strong>of</strong> th<strong>in</strong> copper films. J. Mater. Res., 13:2962, 1998.<br />

[136] L. D. L<strong>and</strong>au <strong>and</strong> E. M. Lifshitz. Theory <strong>of</strong> Elasticity, volume 7 <strong>of</strong> Course <strong>of</strong> Theoretical<br />

Physics. Pergamon Press, 1959.<br />

[137] F. C. Larché <strong>and</strong> J. W. Cahn. The <strong>in</strong>teractions <strong>of</strong> composition <strong>and</strong> stress <strong>in</strong> crystall<strong>in</strong>e<br />

solids. Acta Metall., 33:331, 1985.


Bibliography 136<br />

[138] D. R. Lee, Y. J. Park, D. Kim, Y. H. Jeong, <strong>and</strong> K.-B. Lee. Nonspecular x-ray<br />

reflectivity study <strong>of</strong> partially correlated <strong>in</strong>terface roughness <strong>of</strong> a Mo/Si multilayer.<br />

Phys. Rev. B, 57:8786, 1998.<br />

[139] Y. W. Lee <strong>and</strong> H. I. Aaronson. Anisotropy <strong>of</strong> coherent <strong>in</strong>terphase boundary energy.<br />

Acta Metall., 28:539, 1980.<br />

[140] J. E. Lennard-Jones. Proc. R. Soc. London, Ser. A, 106:463, 1924.<br />

[141] O. S. Leung, A. Munkholm, S. Brennan, <strong>and</strong> W. D. <strong>Ni</strong>x. A search for stra<strong>in</strong> gradients<br />

<strong>in</strong> gold films on substrates us<strong>in</strong>g x-ray diffraction. J. Appl. Phys., 88:1389, 2000.<br />

[142] J. C. M. Li. Petch relation <strong>and</strong> gra<strong>in</strong> boundary sources. Trans. Metall. Soc. AIME,<br />

227:239, 1963.<br />

[143] J. C. M. Li <strong>and</strong> Y. T. Chou. The role <strong>of</strong> dislocations <strong>in</strong> the flow stress gra<strong>in</strong> size<br />

relationships. Metall. Trans., 1:1145, 1970.<br />

[144] E. S. Machl<strong>in</strong>. A critique <strong>of</strong> a recent zone model for the orig<strong>in</strong> <strong>of</strong> gra<strong>in</strong> structure <strong>in</strong><br />

th<strong>in</strong> films. Scripta Metall. Mater., 32:1233, 1995.<br />

[145] E. S. Machl<strong>in</strong>. Materials Science <strong>in</strong> Microelectronics. Giro Press, Croton-on-<br />

Hudson, NY, 1995.<br />

[146] I. Markov. Recent theoretical developments <strong>in</strong> epitaxy. Mater. Chem. Phys., 36:1–<br />

30, 1993.<br />

[147] I. Markov <strong>and</strong> A. Milchev. The effect <strong>of</strong> anharmonicity <strong>in</strong> epitaxial <strong>in</strong>terfaces: II.<br />

Equilibrium structure <strong>of</strong> th<strong>in</strong> epitaxial films. Surf. Sci., 136:519–531, 1984.<br />

[148] W. C. Marra, P. Eisenberger, <strong>and</strong> A. Y. Cho. X-ray total-external-reflection-Bragg<br />

diffraction: A structural study <strong>of</strong> the GaAs-Al <strong>in</strong>terface. J. Appl. Phys., 50:6927,<br />

1979.<br />

[149] T. B. Massalski. B<strong>in</strong>ary Alloy Phase Diagrams. American Society for Metals, Metals<br />

Park, OH, 1986.<br />

[150] J. W. Matthews. Coherent <strong>in</strong>terfaces <strong>and</strong> misfit dislocations. In J. W. Matthews,<br />

editor, Epitaxial Growth: Part B, chapter 8, page 560. Academic Press, Inc., New<br />

York, NY, 1975.<br />

[151] J. W. Matthews. Defects associated with the accommodation <strong>of</strong> misfit between crystals.<br />

J. Vac. Sci. Technol., 12:126, 1975.<br />

[152] J. W. Matthews <strong>and</strong> A. E. Blakeslee. Defects <strong>in</strong> epitaxial multilayers. I. Misfit dislocations.<br />

J. Cryst. Growth, 27:118, 1974.


Bibliography 137<br />

[153] J. W. Matthews <strong>and</strong> A. E. Blakeslee. Defects <strong>in</strong> epitaxial multilayers. II. Dislocation<br />

pile-ups, thread<strong>in</strong>g dislocations, slip l<strong>in</strong>es, <strong>and</strong> cracks. J. Cryst. Growth, 29:273,<br />

1975.<br />

[154] K. L. Merkle, M. I. Buckett, <strong>and</strong> Y. Gao. Atomic structure <strong>of</strong> heterophase <strong>in</strong>terfaces.<br />

Acta Metall. Mater., 40:S249, 1992.<br />

[155] A. Milchev <strong>and</strong> I. Markov. The effect <strong>of</strong> anharmonicity <strong>in</strong> epitaxial <strong>in</strong>terfaces: I.<br />

Substrate-<strong>in</strong>duced dissociation <strong>of</strong> f<strong>in</strong>ite epitaxial isl<strong>and</strong>s. Surf. Sci., 136:503–518,<br />

1984.<br />

[156] B. A. Movchan <strong>and</strong> A. V. Demchish<strong>in</strong>. Investigation <strong>of</strong> the structure <strong>and</strong> properties<br />

<strong>of</strong> thick vacuum-deposited films <strong>of</strong> nickel, titanium, tungsten, alum<strong>in</strong>a <strong>and</strong> zirconium<br />

dioxide. Phys. Met. Metallogr., 28:83, 1969.<br />

[157] S. Mróz, Z. Jankowski, <strong>and</strong> M. Nowicki. Growth <strong>and</strong> isothermal desorption <strong>of</strong> ultrath<strong>in</strong><br />

silver layers on the <strong>Ni</strong>(111) face at the substrate temperature from 180 to 900<br />

K. Surf. Sci., 454-456:702, 2000.<br />

[158] J. A. Mull<strong>in</strong>. Viscous Flow <strong>and</strong> Structural Relaxation <strong>in</strong> Amorphous Silicon <strong>and</strong><br />

Amorphous Selenium Th<strong>in</strong> <strong>Films</strong>. PhD thesis, <strong>Harvard</strong> University, 2000.<br />

[159] W. W. Mull<strong>in</strong>s. Solid surface morphologies governed by capillarity. In Metal Surfaces:<br />

Structure, K<strong>in</strong>etics, <strong>and</strong> Energetics, page 17, Metals Park, OH, 1963. Am.<br />

Soc. Metals.<br />

[160] M. Murakami. Residual stra<strong>in</strong>s <strong>of</strong> Pb th<strong>in</strong> films deposited onto Si substrates. Acta<br />

Metall., 26:175, 1978.<br />

[161] M. Murakami. Thermal stra<strong>in</strong> <strong>in</strong> lead th<strong>in</strong> films. II: Stra<strong>in</strong> relaxation mechanisms.<br />

Th<strong>in</strong> Solid <strong>Films</strong>, 55:101, 1978.<br />

[162] M. Murakami. Thermal stra<strong>in</strong> <strong>in</strong> lead th<strong>in</strong> films. III: Dependences <strong>of</strong> the stra<strong>in</strong> on<br />

film thickness <strong>and</strong> on gra<strong>in</strong> size. Th<strong>in</strong> Solid <strong>Films</strong>, 59:105, 1979.<br />

[163] M. Murakami. Stra<strong>in</strong> relaxation mechanisms <strong>of</strong> th<strong>in</strong> deposited films. CRC Crit. Rev.<br />

Solid State Mater. Sci., 11:317, 1984.<br />

[164] M. Murakami. Deformation <strong>in</strong> th<strong>in</strong> films by thermal stra<strong>in</strong>. J. Vac. Sci. Technol. A,<br />

9:2469, 1991.<br />

[165] M. Murakami, J. Angelillo, H.-C.W. Huang, A. Segmuller, <strong>and</strong> C. J. Kircher. Thermal<br />

stra<strong>in</strong> <strong>in</strong> lead th<strong>in</strong> films. IV: Effects <strong>of</strong> multiple cycl<strong>in</strong>g to 4.2 K. Th<strong>in</strong> Solid<br />

<strong>Films</strong>, 60:1, 1979.


Bibliography 138<br />

[166] M. Murakami <strong>and</strong> P. Chaudhari. Thermal stra<strong>in</strong> <strong>in</strong> lead th<strong>in</strong> films. I: Dependence <strong>of</strong><br />

the stra<strong>in</strong> on crystal orientation. Th<strong>in</strong> Solid <strong>Films</strong>, 46:109, 1977.<br />

[167] M. Murakami <strong>and</strong> T. S. Kuan. Thermal stra<strong>in</strong> <strong>in</strong> lead th<strong>in</strong> films. V: Stra<strong>in</strong> relaxation<br />

above room temperature. Th<strong>in</strong> Solid <strong>Films</strong>, 66:381, 1980.<br />

[168] M. Murakami, T.-S. Kuan, <strong>and</strong> I. A. Blech. Mechanical properties <strong>of</strong> th<strong>in</strong> films on<br />

substrates. In K. N. Tu <strong>and</strong> R. Rosenberg, editors, Preparation <strong>and</strong> Properties <strong>of</strong><br />

Th<strong>in</strong> <strong>Films</strong>, volume 24 <strong>of</strong> Treatise on Materials Science <strong>and</strong> Technology, page 164.<br />

Academic Press, Inc., New York, 1982.<br />

[169] F. D. Murnaghan. F<strong>in</strong>ite Deformation <strong>of</strong> an Elastic Solid. John Wiley & Sons, Inc.,<br />

New York, 1951.<br />

[170] F. R. N. Nabarro, D. B. Holt, <strong>and</strong> Z. S. Bas<strong>in</strong>ski. Plasticity <strong>of</strong> pure s<strong>in</strong>gle crystals.<br />

Adv. Phys., 13:193, 1964.<br />

[171] A. Needleman <strong>and</strong> E. Van der Giessen. Discrete dislocation <strong>and</strong> cont<strong>in</strong>uum descriptions<br />

<strong>of</strong> plastic flow. Mater. Sci. Eng. A, 309-310:1, 2001.<br />

[172] W. D. <strong>Ni</strong>x. Mechanical properties <strong>of</strong> th<strong>in</strong> films. Metall. Trans. A, 20:2217, 1989.<br />

[173] W. D. <strong>Ni</strong>x. Yield<strong>in</strong>g <strong>and</strong> stra<strong>in</strong> harden<strong>in</strong>g <strong>of</strong> th<strong>in</strong> metal films on substrates. Scripta<br />

Mater., 39(4-5):545–54, 1998.<br />

[174] W. D. <strong>Ni</strong>x <strong>and</strong> H. Gao. An atomistic <strong>in</strong>terpretation <strong>of</strong> <strong>in</strong>terface stress. Scripta Mater.,<br />

39:1653, 1998.<br />

[175] A. S. Nowick <strong>and</strong> B. S. Berry. Anelastic Relaxation <strong>in</strong> Crystall<strong>in</strong>e Solids. Academic<br />

Press, New York, NY, 1972.<br />

[176] I. C. Noyan <strong>and</strong> J. B. Cohen. Residual Stress: Measurement By Diffraction And<br />

Interpretation. Spr<strong>in</strong>ger-Verlag, New York, NY, 1987.<br />

[177] I. C. Noyan, T. C. Huang, <strong>and</strong> B. R. York. Residual stress/stra<strong>in</strong> analysis <strong>in</strong> th<strong>in</strong><br />

films by x-ray diffraction. CRC Crit. Rev. Solid St. Mater. Sci., 20:125, 1995.<br />

[178] J. F. Nye. Some geometrical relations <strong>in</strong> dislocated crystals. Acta Metall., 1:153,<br />

1953.<br />

[179] K. Owusu-Boahen <strong>and</strong> A. H. K<strong>in</strong>g. The early stages <strong>of</strong> plastic yield<strong>in</strong>g <strong>in</strong> polycrystall<strong>in</strong>e<br />

gold th<strong>in</strong> films. Acta Mat., 49:237, 2001.<br />

[180] G. Palasantzas <strong>and</strong> J. Th. M. De Hosson. Roughness effect on the measurement <strong>of</strong><br />

<strong>in</strong>terface stress. Acta Mater., 48:3641, 2000.


Bibliography 139<br />

[181] P. Pant, K. W. Schwarz, <strong>and</strong> S. P. Baker. Dislocation dynamics simulations <strong>of</strong> dislocation<br />

<strong>in</strong>teractions <strong>in</strong> th<strong>in</strong> fcc metal films. Mater. Res. Soc. Symp. Proc., 673:P2.2.1,<br />

2001.<br />

[182] V. Pelos<strong>in</strong> <strong>and</strong> J. Hillairet. Stress development <strong>and</strong> relief <strong>in</strong> <strong>Ag</strong> or <strong>Ni</strong> films <strong>and</strong> <strong>Ag</strong><strong>Ni</strong><br />

multilayers <strong>in</strong> relation to structural evolution. Nanostruct. Mater., 4:229, 1994.<br />

[183] V. Pelos<strong>in</strong>, J. Hillairet, <strong>and</strong> B. Rodmacq. The dynamics <strong>of</strong> <strong>in</strong>terface relaxation <strong>and</strong><br />

destratification <strong>in</strong> <strong>Ag</strong><strong>Ni</strong> multilayers. J. Phys.: Condens. Matter, 6:1099, 1994.<br />

[184] V. Pelos<strong>in</strong>, B. Rodmacq, J. Hillairet, G. Carlotti, D. Fioretto, <strong>and</strong> G. Soc<strong>in</strong>o. Elastic<br />

characterization <strong>of</strong> <strong>Ag</strong><strong>Ni</strong> superlattices. Mater. Sci. Forum, 119-121:359–364, 1993.<br />

[185] N. J. Petch. The cleavage strength <strong>of</strong> polycrystals. J. Iron Steel Inst., 174:25, 1953.<br />

[186] V. Ramaswamy. Stress Evolution <strong>in</strong> Sputtered FCC Metal <strong>Multilayers</strong>. PhD thesis,<br />

Stanford University, 2000.<br />

[187] D. T. Read. Young’s modulus <strong>of</strong> th<strong>in</strong> films by speckle <strong>in</strong>terferometry. Meas. Sci.<br />

Technol., 9:676, 1998.<br />

[188] R. R. Reeber <strong>and</strong> K. Wang. Thermal expansion <strong>and</strong> lattice parameters <strong>of</strong> group IV<br />

semiconductors. Mater. Chem. Phys., 46:259, 1996.<br />

[189] H. Reichert, O. Kle<strong>in</strong>, H. Dosch, M. Denk, V. Honkimäki, T. Lippmann, <strong>and</strong> G. Reiter.<br />

Observation <strong>of</strong> five-fold local symmetry <strong>in</strong> liquid lead. Nature, 408:839, 2000.<br />

[190] B. Rodmacq. X-ray diffraction study <strong>of</strong> silver-nickel superlattices. J. Appl. Phys.,<br />

70:4194, 1991.<br />

[191] M. Ronay. Yield stress <strong>of</strong> th<strong>in</strong> fcc polycrystall<strong>in</strong>e metal films bonded to rigid substrates.<br />

Philos. Mag. A, 40:145, 1979.<br />

[192] C. T. Rosenmayer, F. R. Brotzen, <strong>and</strong> R. J. Gale. Mechanical test<strong>in</strong>g <strong>of</strong> th<strong>in</strong> films. In<br />

J. C. Bravman, W. D. <strong>Ni</strong>x, D. M. Barnett, <strong>and</strong> D. A. Smith, editors, Th<strong>in</strong> <strong>Films</strong>:<br />

<strong>Stresses</strong> <strong>and</strong> Mechnical Properties Symposium, volume 130 <strong>of</strong> Mater. Res. Soc.<br />

Symp. Proc., page 77, Pittsburgh, PA, 1989. Materials Research Society.<br />

[193] A. I. Rusanov. Thermodynamics <strong>of</strong> solid surfaces. Surf. Sci. Reports, 23:173–247,<br />

1996.<br />

[194] T. P. Russell. X-ray <strong>and</strong> neutron reflectivity for the <strong>in</strong>vestigation <strong>of</strong> polymers. Mater.<br />

Sci. Rep., 5:171, 1990.<br />

[195] J. A. Ruud. Interface Stress <strong>and</strong> Mechanical Properties <strong>of</strong> Multilayered Th<strong>in</strong> <strong>Films</strong>.<br />

PhD thesis, <strong>Harvard</strong> University, 1992.


Bibliography 140<br />

[196] J. A. Ruud, A. Witvrouw, <strong>and</strong> F. Spaepen. Bulk <strong>and</strong> <strong>in</strong>terface stresses <strong>in</strong> silver-nickel<br />

multilayered th<strong>in</strong> films. J. Appl. Phys., 74:2517, 1993.<br />

[197] P. S<strong>and</strong>ström, E. B. Svedberg, M. P. Johansson, J. Birch, <strong>and</strong> J.-E. Sundgren. Structure<br />

evolution <strong>in</strong> <strong>Ag</strong>/<strong>Ni</strong> multilayers grown by ultra high vacuum DC magnetron sputter<strong>in</strong>g.<br />

Th<strong>in</strong> Solid <strong>Films</strong>, 353:166, 1999.<br />

[198] K. O. Schweitz, J. Bøttiger, J. Chevallier, R. Feidenhans’l, M. M. <strong>Ni</strong>elsen, <strong>and</strong> F. B.<br />

Rasmussen. Interface stress <strong>in</strong> Au/<strong>Ni</strong> multilayers. J. Appl. Phys., 88:1401, 2000.<br />

[199] K. O. Schweitz, J. Bøttiger, A. L. Greer, P. J. Thomas, <strong>and</strong> D. T. Foord. The mechanism<br />

<strong>of</strong> degradation <strong>of</strong> <strong>Ag</strong>/<strong>Ni</strong> multilayers deposited at different temperatures. Philos.<br />

Mag. A, 81:1, 2001.<br />

[200] K. O. Schweitz, H. Geisler, J. Chevallier, J. Bøttiger, <strong>and</strong> R. Feidenhans’l. Interface<br />

stress <strong>and</strong> an apparent negative Poisson’s ratio <strong>in</strong> <strong>Ag</strong>/<strong>Ni</strong> multilayers. In R. C. Cammarata,<br />

M. Nastasi, E. P. Busso, <strong>and</strong> W. C. Oliver, editors, Th<strong>in</strong>-<strong>Films</strong> — <strong>Stresses</strong> <strong>and</strong><br />

Mechanical Properties VII, volume 505, page 559, Warrendale, PA, 1998. Materials<br />

Research Society.<br />

[201] K. O. Schweitz, K. Rätzke, D. Foord, P. J. Thomas, A. L. Greer, H. Geisler,<br />

J. Chevallier, R. Feidenhans’l, <strong>and</strong> J. Bøttiger. The microstructural development<br />

<strong>of</strong> <strong>Ag</strong>/<strong>Ni</strong> multilayers dur<strong>in</strong>g anneal<strong>in</strong>g. Philos. Mag. A, 80:1867, 2000.<br />

[202] A. Segmüller <strong>and</strong> M. Murakami. X-ray diffraction analysis <strong>of</strong> stra<strong>in</strong>s <strong>and</strong> stresses<br />

<strong>in</strong> th<strong>in</strong> films. In K. N. Tu <strong>and</strong> R. Rosenberg, editors, Analytical Techniques for<br />

Th<strong>in</strong> <strong>Films</strong>, volume 27 <strong>of</strong> Treatise on Materials Science <strong>and</strong> Technology, page 143.<br />

Academic Press, Inc., Boston, 1988.<br />

[203] Corn<strong>in</strong>g Semiconductor. Corn<strong>in</strong>g Semiconductor Materials HPFS Data Sheet.<br />

http://www.corn<strong>in</strong>g.com/semiconductormaterials/.<br />

[204] A. L. Shull. Measurements <strong>of</strong> Stress Dur<strong>in</strong>g Deposition <strong>in</strong> Copper <strong>and</strong> Silver Th<strong>in</strong><br />

<strong>Films</strong> <strong>and</strong> <strong>Multilayers</strong>. PhD thesis, <strong>Harvard</strong> University, 1996.<br />

[205] A. L. Shull <strong>and</strong> F. Spaepen. Measurements <strong>of</strong> stress dur<strong>in</strong>g vapor deposition <strong>of</strong><br />

copper <strong>and</strong> silver th<strong>in</strong> films <strong>and</strong> multilayers. J. Appl. Phys., 80:6243, 1996.<br />

[206] R. Shuttleworth. The surface tension <strong>of</strong> solids. Proc. Phys. Soc. A, 63:445, 1950.<br />

[207] G. Simmons <strong>and</strong> H. Wang. S<strong>in</strong>gle Crystal Elastic Constants <strong>and</strong> Calculated <strong>Ag</strong>gregate<br />

Properties: A H<strong>and</strong>book. MIT Press, Cambridge, MA, 1971.<br />

[208] J. E. S<strong>in</strong>clair, P. C. Gehlen, R. G. Hoagl<strong>and</strong>, <strong>and</strong> J. P. Hirth. Flexible boundary<br />

conditions <strong>and</strong> nonl<strong>in</strong>ear geometric effects <strong>in</strong> atomic dislocation model<strong>in</strong>g. J. Appl.<br />

Phys., 49:3890, 1978.


Bibliography 141<br />

[209] S. K. S<strong>in</strong>ha, E. B. Sirota, S. Gar<strong>of</strong>f, <strong>and</strong> H. B. Stanley. X-ray <strong>and</strong> neutron scatter<strong>in</strong>g<br />

from rough surfaces. Phys. Rev. B, 38:2297, 1988.<br />

[210] I. S. Sokolnik<strong>of</strong>f. Mathematical Theory <strong>of</strong> Elasticity. R.E. Krieger Pub. Co., Malabar,<br />

FL, 1956.<br />

[211] F. Spaepen. Homogenous nucleation <strong>and</strong> the temperature dependence <strong>of</strong> the crystalmelt<br />

<strong>in</strong>terfacial tension. volume 47 <strong>of</strong> Solid State Physics, page 1. Academic Press,<br />

New York, 1994.<br />

[212] F. Spaepen. Interfaces <strong>and</strong> stresses <strong>in</strong> th<strong>in</strong> films. Acta Mater., 48:31, 2000.<br />

[213] F. Spaepen, A. L. Greer, K. F. Kelton, <strong>and</strong> J. L. Bell. Ion beam sputter<strong>in</strong>g apparatus<br />

for fabrication <strong>of</strong> compositionally modulated materials. Rev. Sci. Instrum., 56:1340,<br />

1985.<br />

[214] A. P. Sutton <strong>and</strong> R. W. Balluffi. Interfaces <strong>in</strong> Crystall<strong>in</strong>e Materials, volume 51 <strong>of</strong><br />

Monographs on the Physics <strong>and</strong> Chemistry <strong>of</strong> Materials. Oxford: Clarendon Press,<br />

Oxford University, 1995.<br />

[215] W. Suwito, M. L. Dunn, S. J. <strong>Cu</strong>nn<strong>in</strong>gham, <strong>and</strong> D. T. Read. Elastic moduli, strength,<br />

<strong>and</strong> fracture <strong>in</strong>itiation at sharp notches <strong>in</strong> etched s<strong>in</strong>gle crystal silicon microstructures.<br />

J. Appl. Phys., 85:7, 1999.<br />

[216] G. I. Taylor. The mechanism <strong>of</strong> plastic deformation <strong>of</strong> crystals. Proc. R. Soc. A,<br />

145:362, 1934.<br />

[217] G. I. Taylor. Plastic stra<strong>in</strong> <strong>in</strong> metals. J. Inst. Met., 62:307, 1938.<br />

[218] C. V. Thompson. The yield stress <strong>of</strong> polycrystall<strong>in</strong>e th<strong>in</strong> films. J. Mater. Res., 8:237,<br />

1993.<br />

[219] C. V. Thompson. Structure evolution dur<strong>in</strong>g process<strong>in</strong>g <strong>of</strong> polycrystall<strong>in</strong>e films.<br />

Annu. Rev. Mater. Sci., 30:159, 2000.<br />

[220] C. V. Thompson <strong>and</strong> R. Carel. Stress <strong>and</strong> gra<strong>in</strong> growth <strong>in</strong> th<strong>in</strong> films. J. Mech. Phys.<br />

Solids, 44:657, 1996.<br />

[221] M. D. Thouless. Effect <strong>of</strong> surface diffusion on the creep <strong>of</strong> th<strong>in</strong> films <strong>and</strong> s<strong>in</strong>tered<br />

arrays <strong>of</strong> particles. Acta Metall. Mater., 41(4):1057, 1993.<br />

[222] M. D. Thouless. Model<strong>in</strong>g the developoment <strong>and</strong> relaxation <strong>of</strong> stresses <strong>in</strong> films.<br />

Annu. Rev. Mater. Sci., 25:69, 1995.<br />

[223] M. D. Thouless, J. Gupta, <strong>and</strong> J. M. E. Harper. Stress development <strong>and</strong> relaxation <strong>in</strong><br />

copper films dur<strong>in</strong>g thermal cycl<strong>in</strong>g. J. Mater. Res., 8:1845, 1993.


Bibliography 142<br />

[224] S. P. Timoshenko <strong>and</strong> J. N. Goodier. Theory <strong>of</strong> Elasticity. McGraw-Hill, New York,<br />

3rd edition, 1987.<br />

[225] M. F. Toney <strong>and</strong> S. Brennan. Observation <strong>of</strong> the effect <strong>of</strong> refraction on xrays<br />

diffracted <strong>in</strong> a graz<strong>in</strong>g-<strong>in</strong>cidence asymmetric Bragg geometry. Phys. Rev. B,<br />

39:7963, 1989.<br />

[226] D. Turnbull. Role <strong>of</strong> structural impurities <strong>in</strong> phase transformations. In Impurities<br />

<strong>and</strong> Imperfections, page 121, Metals Park, OH, 1955. American Society for Metals.<br />

[227] K. Van Acker, L. De Buyser, J. P. Celis, <strong>and</strong> P. Van Houtte. Characterization <strong>of</strong> th<strong>in</strong><br />

nickel electrocoat<strong>in</strong>gs by the low-<strong>in</strong>cident-beam-angle diffraction method. J. Appl.<br />

Cryst., 27:56, 1994.<br />

[228] J. H. van der Merwe <strong>and</strong> W. A. Jesser. An exactly solvable model for calculat<strong>in</strong>g critical<br />

misfit <strong>and</strong> thickness <strong>in</strong> epitaxial superlattices: Layers <strong>of</strong> equal elastic constants<br />

<strong>and</strong> thicknesses. J. Appl. Phys., 63:1509, 1988.<br />

[229] M. van Leeuwen, J.-D. Kamm<strong>in</strong>ga, <strong>and</strong> E. J. Mittemeijer. Diffraction stress analysis<br />

<strong>of</strong> th<strong>in</strong> films: Model<strong>in</strong>g <strong>and</strong> experimental evaluation <strong>of</strong> elastic constants <strong>and</strong> gra<strong>in</strong><br />

<strong>in</strong>teraction. J. Appl. Phys., 86:1904, 1999.<br />

[230] J. A. Venables, G. D. T. Spiller, <strong>and</strong> M. Hanbücken. Nucleation <strong>and</strong> growth <strong>of</strong> th<strong>in</strong><br />

films. Rep. Prog. Phys., 47:399, 1984.<br />

[231] R. Venkatraman. Plasticity, microstructure <strong>and</strong> the thermal dependence <strong>of</strong> flow<br />

stresses <strong>in</strong> alum<strong>in</strong>um th<strong>in</strong> film <strong>in</strong>terconnects. In P. Børgesen, W. Filter, J. E. Sanchez,<br />

Jr., K. P. Rodbell, <strong>and</strong> J. C. Coburn, editors, Materials Reliability <strong>in</strong> Microelectronics<br />

IV, volume 338, page 215, Pittsburgh, PA, 1994. Mater. Res. Soc.<br />

[232] R. Venkatraman, P. R. Besser, J. C. Bravman, <strong>and</strong> S. Brennan. Elastic stra<strong>in</strong> gradients<br />

<strong>and</strong> x-ray l<strong>in</strong>e broaden<strong>in</strong>g effects as a function <strong>of</strong> temperature <strong>in</strong> alum<strong>in</strong>um th<strong>in</strong> films<br />

on silicon. J. Mater. Res., 9:328, 1994.<br />

[233] R. Venkatraman <strong>and</strong> J. C. Bravman. Separation <strong>of</strong> film thickness <strong>and</strong> gra<strong>in</strong> boundary<br />

strengthen<strong>in</strong>g effects <strong>in</strong> al th<strong>in</strong> films on si. J. Mater. Res., 7:2040, 1992.<br />

[234] M. Verdier, H. Huang, F. Spaepen, J. D. Embury, M. Hawley, <strong>and</strong> H. Kung. Microstructure,<br />

nano<strong>in</strong>dentation <strong>and</strong> work harden<strong>in</strong>g <strong>of</strong> <strong>Cu</strong>/<strong>Ag</strong> multilayers.<br />

[235] A. C. Vermeulen <strong>and</strong> E. Houtman. Determ<strong>in</strong>ation <strong>of</strong> alignment errors <strong>in</strong> classical<br />

XRD residual stress methods. Mater. Sci. Forum, 347-349:17, 2000.<br />

[236] R. P. V<strong>in</strong>ci <strong>and</strong> J. J. Vlassak. Mechanical behavior <strong>of</strong> th<strong>in</strong> films. volume 26 <strong>of</strong> Annu.<br />

Rev. Mater. Sci., page 431. Annual Reviews, Palo Alto, CA, 1996.


Bibliography 143<br />

[237] R.P. V<strong>in</strong>ci, G. Cornella, <strong>and</strong> J.C. Bravman. Anelastic contributions to the behavior<br />

<strong>of</strong> freest<strong>and</strong><strong>in</strong>g Al th<strong>in</strong> films. In Stress Induced Phenomena <strong>in</strong> Metallization: Fifth<br />

International Workshop, volume 491 <strong>of</strong> AIP Conference Proceed<strong>in</strong>gs, pages 240–<br />

248, 1999.<br />

[238] R.P. V<strong>in</strong>ci, E.M. Ziel<strong>in</strong>ski, <strong>and</strong> J.C. Bravman. Thermal stra<strong>in</strong> <strong>and</strong> stress <strong>in</strong> copper<br />

th<strong>in</strong> films. Th<strong>in</strong> Solid <strong>Films</strong>, 262(1-2):142–153, 1995.<br />

[239] R. W. Vook <strong>and</strong> F. Witt. Thermally <strong>in</strong>duced stra<strong>in</strong>s <strong>in</strong> evaporated films. J. Appl.<br />

Phys., 36:2169, 1965.<br />

[240] D. C. Wallace. Thermoelasticity <strong>of</strong> stressed materials <strong>and</strong> comparison <strong>of</strong> various<br />

elastic constants. Phys. Rev., 162:776, 1967.<br />

[241] L. Wcislak, H. J. Bunge, <strong>and</strong> C. Nauer-Gerhardt. XRD texture <strong>in</strong>vestigations with<br />

the employment <strong>of</strong> location sensitive measur<strong>in</strong>g technique. In R. Delhez <strong>and</strong> E. J.<br />

Mittemeijer, editors, Proceed<strong>in</strong>gs <strong>of</strong> the Second European Powder Diffraction Conference,<br />

volume 133-136 <strong>of</strong> Mater. Sci. Forum, page 157, Switzerl<strong>and</strong>, 1993. Trans<br />

Tech.<br />

[242] V. Weihnacht <strong>and</strong> W. Brückner. Stress <strong>and</strong> plasticity <strong>in</strong> <strong>Cu</strong> th<strong>in</strong> films. In O. Kraft<br />

et al., editors, Stress Induced Phenomena <strong>in</strong> Metallization: Fifth International Workshop,<br />

AIP Conference Proceed<strong>in</strong>gs, page 283, USA, 1999. AIP Conf. Proc.<br />

[243] V. Weihnacht <strong>and</strong> W. Brückner. Plasticity <strong>in</strong> copper th<strong>in</strong> films. In R. V<strong>in</strong>ci et al.,<br />

editors, Th<strong>in</strong>-films: <strong>Stresses</strong> <strong>and</strong> mechanical properties VIII, volume 594 <strong>of</strong> Mater.<br />

Res. Soc. Symp. Proc., page 87, Warrendale, PA, 2000. Materials Research Society.<br />

[244] V. Weihnacht <strong>and</strong> W. Brückner. Dislocation accumulation <strong>and</strong> strengthen<strong>in</strong>g <strong>in</strong> <strong>Cu</strong><br />

th<strong>in</strong> films. Acta Mater., 49:2365, 2001.<br />

[245] V. Weihnacht <strong>and</strong> W. Brückner. Mechanical behavior <strong>of</strong> th<strong>in</strong> <strong>Cu</strong> films studied by a<br />

four-po<strong>in</strong>t bend<strong>in</strong>g technique. Mater. Res. Soc. Symp. Proc., 673:P1.10.1, 2001.<br />

[246] D. Weiss, H. Gao, <strong>and</strong> E. Arzt. Constra<strong>in</strong>ed diffusional creep <strong>in</strong> UHV-produced<br />

copper th<strong>in</strong> films. Acta Mater., 49:2395, 2001.<br />

[247] J. Weissmüller <strong>and</strong> J. W. Cahn. Mean stresses <strong>in</strong> microstructures due to <strong>in</strong>terface<br />

stresses: A generalization <strong>of</strong> a capillary equation for solids. Acta Mater., 45:1899,<br />

1997.<br />

[248] G. K. Wertheim, M. A. Butler, K. W. West, <strong>and</strong> D. N. E. Buchanan. Determ<strong>in</strong>ation <strong>of</strong><br />

the Gaussian <strong>and</strong> Lorentzian content <strong>of</strong> experimental l<strong>in</strong>e shapes. Rev. Sci. Instrum.,<br />

45:1369, 1974.


Bibliography 144<br />

[249] G. K. White. Reference materials for thermal expansion: certified or not? Thermochimica<br />

acta, 218:83, 1993.<br />

[250] J. R. Willis, S. C. Ja<strong>in</strong>, <strong>and</strong> R. Bullough. The energy <strong>of</strong> an array <strong>of</strong> dislocations:<br />

implications for stra<strong>in</strong> relaxation <strong>in</strong> semiconductor heterostructures. Philos. Mag. A,<br />

62:115, 1990.<br />

[251] A. J. C. Wilson. The location <strong>of</strong> peaks. Brit. J. Appl. Phys., 16:665, 1965.<br />

[252] A. J. C. Wilson. Statistical variance <strong>of</strong> l<strong>in</strong>e-pr<strong>of</strong>ile parameters. Measures <strong>of</strong> <strong>in</strong>tensity,<br />

location <strong>and</strong> dispersion. Acta Cryst., 23:888, 1967.<br />

[253] H. W<strong>in</strong>dischmann. Intr<strong>in</strong>sic stress <strong>in</strong> sputtered th<strong>in</strong> films. J. Vac. Sci. Tehnol. A,<br />

9:2431–2436, 1991.<br />

[254] H. W<strong>in</strong>dischmann. Intr<strong>in</strong>sic stress <strong>in</strong> sputter-deposited th<strong>in</strong> films. CRC Crit. Rev.<br />

Solid St. Mater. Sci., 17:547, 1992.<br />

[255] R. A. W<strong>in</strong>holtz. Corrections to equations for estimat<strong>in</strong>g the count<strong>in</strong>g-statistical errors<br />

<strong>in</strong> diffraction stress measurements. J. Appl. Cryst., 28:590, 1995.<br />

[256] A. Witvrouw. Viscosity <strong>and</strong> Elastic Constants <strong>of</strong> Amorphous Th<strong>in</strong> <strong>Films</strong>. PhD thesis,<br />

<strong>Harvard</strong> University, 1992.<br />

[257] A. Witvrouw <strong>and</strong> F. Spaepen. Determ<strong>in</strong>ation <strong>of</strong> the plane stress elastic constants<br />

<strong>of</strong> th<strong>in</strong> films from substrate curvature measurements: Applications to amorphous<br />

metals. J. Appl. Phys., 73:7344, 1993.<br />

[258] D. Wolf. Correlation between energy <strong>and</strong> volume expansion for gra<strong>in</strong> boundaries <strong>in</strong><br />

fcc metals. Scripta Metall., 23:1913, 1989.<br />

[259] D. Wolf. Correlation between energy, surface tension <strong>and</strong> structure <strong>of</strong> free surfaces<br />

<strong>in</strong> fcc metals. Surf. Sci., 226:389, 1990.<br />

[260] U. Wolf, S. M. Foiles, <strong>and</strong> H. F. Fischmeister. Study <strong>of</strong> misfit dislocations at the<br />

<strong>in</strong>teface <strong>of</strong> weakly bonded metal/metal systems. Acta Metall. Mater., 39:373, 1991.<br />

[261] D. Y. W. Yu. private communication.<br />

[262] J.-H. Zhao, Y. Du, M. Morgan, <strong>and</strong> P. S. Ho. Simultaneous measurement <strong>of</strong> Young’s<br />

modulus, Poisson ratio, <strong>and</strong> coefficient <strong>of</strong> thermal expansion <strong>of</strong> th<strong>in</strong> films on substrates.<br />

J. Appl. Phys., 87:1575, 2000.<br />

[263] E. M. Ziel<strong>in</strong>ski, R. P. V<strong>in</strong>ci, <strong>and</strong> J. C. Bravman. Effects <strong>of</strong> barrier layer <strong>and</strong> anneal<strong>in</strong>g<br />

on abnormal gra<strong>in</strong> growth <strong>in</strong> copper th<strong>in</strong> films. J. Appl. Phys., 76:4516, 1994.


Appendix A<br />

X-Ray Stress Analysis<br />

Although the pr<strong>in</strong>ciple beh<strong>in</strong>d x-ray stress analysis (XSA) is simple, the actual exper-<br />

iment <strong>and</strong> analysis can be difficult. This section discusses some <strong>of</strong> the methods used to<br />

convert x-ray data <strong>in</strong>to a stress tensor.<br />

By x-ray diffraction we directly measure<br />

dhkl = nλx−ray<br />

2 s<strong>in</strong> θhkl<br />

(A.1)<br />

where 2θhkl is the peak position <strong>of</strong> the (hkl)-reflection. We calculate the stra<strong>in</strong> perpendic-<br />

ular to a particular set <strong>of</strong> (hkl) lattice planes by<br />

ɛ L 33 (hkl) = dhkl − dhkl, ref<br />

. (A.2)<br />

dhkl, 0<br />

The superscript L refers to the laboratory reference frame; the relationship between this<br />

coord<strong>in</strong>ate system <strong>and</strong> the sample coord<strong>in</strong>ate system, S, is shown <strong>in</strong> Figure A.1. The<br />

<strong>in</strong>dividual crystals have their own coord<strong>in</strong>ate system denoted by C. Naturally, there are<br />

several ways to tilt by ψ; the most common is to rotate the sample normal <strong>in</strong> the plane <strong>of</strong><br />

the diffractometer, i.e., the plane def<strong>in</strong>ed by the <strong>in</strong>com<strong>in</strong>g <strong>and</strong> outgo<strong>in</strong>g x-rays; s<strong>in</strong>ce this<br />

145


Appendix A: X-Ray Stress Analysis 146<br />

S 1<br />

S 3<br />

ψ<br />

φ<br />

q, L3 Figure A.1: Diagram show<strong>in</strong>g the relationship between the sample coord<strong>in</strong>ate system, S,<br />

<strong>and</strong> the laboratory coord<strong>in</strong>ate system, L.<br />

motor is <strong>of</strong>ten labeled ω, this is called the ω-diffractometer. Some diffractometers can tilt<br />

the sample normal <strong>in</strong> a plane orthogonal to the diffractometer plane; this is angle is usually<br />

labeled ψ <strong>in</strong> the literature, <strong>and</strong> so this is called the ψ-diffractometer. Some diffractometers,<br />

such as four-circle diffractometers, can tilt <strong>in</strong> both planes. In an attempt to avoid confusion,<br />

I will call angle <strong>in</strong> the plane orthogonal to the diffractometer plane χ <strong>and</strong> leave ψ as simply<br />

a general tilt angle. The angle φ is always called φ <strong>and</strong> refers to a rotation around the sample<br />

normal.<br />

It is best to analyze (hkl) reflections at high 2θ angles for the best sensitivity. Differen-<br />

tiat<strong>in</strong>g Bragg’s law gives<br />

S 2<br />

�θhkl = ɛ L 33 (hkl) tan θhkl<br />

(A.3)<br />

so for a given stra<strong>in</strong>, �2θhkl is larger for higher 2θhkl.<br />

The measured stra<strong>in</strong> is related to the stra<strong>in</strong>s <strong>in</strong> the sample reference frame by the rota-


Appendix A: X-Ray Stress Analysis 147<br />

tion matrix a(ψ, φ):<br />

ɛ L 33 (φ, ψ) = aLS 3k (ψ)aLS 3l (φ)ɛS kl<br />

= ɛ S 11 cos2 φ s<strong>in</strong> 2 ψ + ɛ S 22 s<strong>in</strong>2 φ s<strong>in</strong> 2 ψ + ɛ S 33 cos2 ψ +<br />

ɛ S 12 s<strong>in</strong> 2φ s<strong>in</strong>2 ψ + ɛ S 13 cos φ s<strong>in</strong> 2ψ + ɛS 23 s<strong>in</strong> φ s<strong>in</strong> 2ψ. (A.4)<br />

Therefore, by mak<strong>in</strong>g six <strong>in</strong>dependent measurements <strong>of</strong> ɛ L 33 (φ, ψ), we can solve for<br />

all six components <strong>of</strong> ɛ S ij . Typically, however, we hope a plot <strong>of</strong> dhkl vs. s<strong>in</strong> 2 ψ yields a<br />

straight l<strong>in</strong>e; <strong>in</strong> such a case, ɛ S 13 = ɛS 23<br />

= 0.<br />

The stress tensor can then be determ<strong>in</strong>ed us<strong>in</strong>g Hooke’s law,<br />

ɛ S ij = s S ijkl<br />

A.1 Untextured Th<strong>in</strong> <strong>Films</strong><br />

S<br />

σkl . (A.5)<br />

Comb<strong>in</strong><strong>in</strong>g (A.4) <strong>and</strong> (A.5), we obta<strong>in</strong> for an elastically isotropic sample<br />

ɛ L 1<br />

33 (φ, ψ) =<br />

where S hkl<br />

i<br />

2 Shkl<br />

2<br />

+ 1<br />

2 Shkl<br />

2<br />

�<br />

σ S 11 cos2 φ s<strong>in</strong> 2 ψ + σ S 22 s<strong>in</strong>2 φ s<strong>in</strong> 2 ψ + σ S 33 cos2 �<br />

ψ +<br />

�<br />

σ S 12 s<strong>in</strong> 2φ s<strong>in</strong>2 ψ + σ S 13 cos φ s<strong>in</strong> 2ψ + σ S 23 s<strong>in</strong> φ s<strong>in</strong> 2ψ<br />

�<br />

+<br />

S hkl<br />

�<br />

1 σ S 11 + σ S 22 + σ S �<br />

33<br />

are the x-ray elastic constants (XECs) for elastically isotropic materials.<br />

(A.6)


Appendix A: X-Ray Stress Analysis 148<br />

A.1.1 Biaxial Stress Case<br />

If we assume the stress tensor is biaxial, i.e., σi3 = 0, we can simplify this further <strong>and</strong><br />

arrive at<br />

where<br />

Equibiaxial Stress Case<br />

ɛ L 33 (φ, ψ, hkl) = Shkl<br />

1 (σ S 11 + σ S 1<br />

22 ) +<br />

2 Shkl<br />

2 σ S φ s<strong>in</strong>2 ψ (A.7)<br />

σ S φ = σ S 11 cos2 φ + σ S 22 s<strong>in</strong>2 φ + σ S 12 s<strong>in</strong> 2φ. (A.8)<br />

Furthermore, <strong>in</strong> many cases σ S 11 = σ S 22 = σ S � <strong>and</strong> σ S 12<br />

stress state), <strong>and</strong> we can reduce (A.7) to<br />

ɛ L 33 (ψ, hkl) = 2Shkl<br />

1 σ S �<br />

Therefore, if S hkl<br />

1 <strong>and</strong> S hkl<br />

2 are known, the stress σ S �<br />

= 0 (result<strong>in</strong>g <strong>in</strong> an equibiaxial<br />

+ 1<br />

2 Shkl<br />

2 σ S � s<strong>in</strong>2 ψ. (A.9)<br />

can be determ<strong>in</strong>ed from the slope <strong>of</strong><br />

the dhkl(ψ) vs. s<strong>in</strong> 2 ψ l<strong>in</strong>e. Equation (A.9) is the commonly used equation for r<strong>and</strong>omly<br />

oriented polycrystall<strong>in</strong>e th<strong>in</strong> films.<br />

For an isotropic solid, we can also def<strong>in</strong>e S mech<br />

1<br />

<strong>and</strong> Smech 2 , which are <strong>in</strong>dependent<br />

<strong>of</strong> (hkl) <strong>and</strong> depend only on the macroscopic elastic properties <strong>of</strong> the material (<strong>and</strong> thus<br />

labelled the mechanical elastic constants):<br />

S mech<br />

1<br />

1<br />

2 Smech 2 = 1 + ν<br />

= −ν<br />

E<br />

(A.10)<br />

. (A.11)<br />

E


Appendix A: X-Ray Stress Analysis 149<br />

Another advantage <strong>of</strong> us<strong>in</strong>g S mech<br />

1<br />

<strong>and</strong> S mech<br />

2<br />

rather than S hkl<br />

1 <strong>and</strong> S hkl<br />

2 is that the XECs<br />

have no specific bounds while the mechanical elastic constants are bounded by the Voigt<br />

<strong>and</strong> Reuss limits or the Hash<strong>in</strong>-Shtrikman bounds [93, 94]. Kamm<strong>in</strong>ga et al. [123] have<br />

used formulas relat<strong>in</strong>g S hkl<br />

i<br />

σ S �<br />

to S mech<br />

i<br />

<strong>in</strong> a sample with known Smech<br />

1<br />

ɛ L 33 (ψ, hkl) = dhkl(ψ)(h 2 + k 2 + l 2 ) 1/2 − a0<br />

where<br />

=<br />

�<br />

2S mech<br />

1 + 2K2<br />

The values a0, K2, <strong>and</strong> σ S �<br />

aref<br />

�<br />

Ɣ − 1<br />

5<br />

for elastically isotropic samples so one can solve for<br />

<strong>and</strong> Smech 2 . Their equation is<br />

�<br />

+<br />

�<br />

1<br />

2 Smech<br />

�<br />

2 − 3K2 Ɣ − 1<br />

��<br />

s<strong>in</strong><br />

5<br />

2 �<br />

ψ σ S � (A.12)<br />

Ɣ = h2k 2 + k2l 2 + l2h 2<br />

�<br />

h2 + k2 + l2 � . (A.13)<br />

2<br />

can then be determ<strong>in</strong>ed by at least three measurements <strong>of</strong><br />

dhkl(ψ)(h 2 + k 2 + l 2 ) 1/2 <strong>and</strong> perform<strong>in</strong>g a multivariate fit. This is called the direction<br />

solution method (or the Kamm<strong>in</strong>ga method) <strong>and</strong> the method I use for untextured th<strong>in</strong> films.<br />

A.2 Textured Th<strong>in</strong> <strong>Films</strong> <strong>in</strong> Equibiaxial Stress<br />

A major problem <strong>in</strong> stress analysis <strong>of</strong> textured th<strong>in</strong> films is that sufficient diffracted<br />

<strong>in</strong>tensity can be found only at certa<strong>in</strong> tilt angles, ψ. We can label these peaks as texture<br />

poles. For <strong>in</strong>stance, if our sample has pure 〈111〉 fiber texture <strong>and</strong> we exam<strong>in</strong>e a (420)<br />

peak, we should tilt to<br />

�<br />

[111] · [420]<br />

ψ = arccos √<br />

12 + 12 + 12 √ 42 + 22 + 02 �<br />

(A.14)<br />

to analyze stra<strong>in</strong>s <strong>in</strong> the 〈111〉-textured gra<strong>in</strong>s. Therefore, some have proposed only tak<strong>in</strong>g<br />

diffraction data at these texture poles (this is called the crystallite group method or CGM).


Appendix A: X-Ray Stress Analysis 150<br />

1 While this is useful for exam<strong>in</strong><strong>in</strong>g stresses <strong>in</strong> different gra<strong>in</strong> populations (as Baker et al.<br />

[10] have done), extract<strong>in</strong>g the macroscopic stress value <strong>in</strong> this manner can be dangerous<br />

if the degree <strong>of</strong> preferred orientation is not well known. For the situation <strong>of</strong> perfect 〈111〉<br />

fiber texture [123], the CGM gives<br />

where s C 0 = sC 11 − sC 12<br />

ɛ L 33 (ψ) =<br />

− 1<br />

2 sC 44<br />

, <strong>and</strong><br />

ɛ L 33 (ψ) =<br />

for perfect 〈100〉 fiber texture [82].<br />

�<br />

2s C 2<br />

12 +<br />

3 sC 0<br />

�<br />

2s C 12 +<br />

1<br />

+<br />

2 sC 44 s<strong>in</strong>2 �<br />

ψ σ S �<br />

�<br />

1<br />

2 sC 44 + sC �<br />

0 s<strong>in</strong> 2 �<br />

ψ σ S �<br />

(A.15)<br />

(A.16)<br />

Therefore, after plott<strong>in</strong>g d vs. s<strong>in</strong> 2 ψ for a sample with perfect 〈111〉 texture, the biaxial<br />

stress can be determ<strong>in</strong>ed by<br />

∂ɛ L 33 (ψ)<br />

∂ s<strong>in</strong>2 1<br />

=<br />

ψ d0<br />

Additionally, the stra<strong>in</strong> at ψ = 0isgivenby<br />

ɛ L 33 (ψ = 0) =<br />

∂d<br />

∂ s<strong>in</strong>2 ψ = sC 44<br />

2 σ S � . (A.17)<br />

�<br />

2s C 2<br />

12 +<br />

3 sC 0<br />

�<br />

σ S � . (A.18)<br />

We can use either equation to arrive at the stress or both to provide a check.<br />

A.3 The Stra<strong>in</strong>-Free Lattice Parameter <strong>in</strong> Biaxially Stressed<br />

Samples<br />

For untextured films, the best way <strong>of</strong> obta<strong>in</strong><strong>in</strong>g a0 is by the direct solution method as<br />

mentioned earlier. The most common method <strong>of</strong> f<strong>in</strong>d<strong>in</strong>g the stra<strong>in</strong>-free lattice spac<strong>in</strong>g for a<br />

1 Another way around the th<strong>in</strong> film <strong>in</strong>tensity problem is to keep a large “footpr<strong>in</strong>t” on the sample by fix<strong>in</strong>g<br />

the <strong>in</strong>com<strong>in</strong>g beam at a low angle, such as 3 ◦ <strong>and</strong> to vary the tilt by tak<strong>in</strong>g measurements at different 2θhkls.<br />

This is known as the LIBAD method [227]. Of course <strong>in</strong> such a case, d is no longer l<strong>in</strong>ear <strong>in</strong> s<strong>in</strong> 2 ψ.


Appendix A: X-Ray Stress Analysis 151<br />

sample with known elastic constants is to solve (A.9) for zero stra<strong>in</strong>, which gives<br />

s<strong>in</strong> 2 ψ0 = −4Shkl<br />

1<br />

S hkl<br />

2<br />

≈ −4Smech<br />

1<br />

Smech 2<br />

= 2ν<br />

. (A.19)<br />

1 + ν<br />

For a perfectly textured 〈111〉 sample, we can solve (A.15) for zero stra<strong>in</strong>:<br />

s<strong>in</strong> 2 ψ0 = 2sC 44 − 4sC 11 − 8sC 12<br />

3sC . (A.20)<br />

44<br />

For more on x-ray stress analysis, <strong>in</strong>clud<strong>in</strong>g problems associated with triaxial stress<br />

states, see Noyan <strong>and</strong> Cohen [176] <strong>and</strong> Hauk [95].


Appendix B<br />

X-Ray Errors<br />

In this section, I assume the diffractometer is well aligned. Even after good alignment,<br />

however, we <strong>of</strong>ten have to contend with defocus<strong>in</strong>g <strong>and</strong> geometric errors. In general, these<br />

errors can be accounted for by spread<strong>in</strong>g a th<strong>in</strong> layer <strong>of</strong> stra<strong>in</strong>-free powder onto the sample<br />

<strong>and</strong> measur<strong>in</strong>g the peak shifts <strong>in</strong> the powder <strong>in</strong> addition to the peak shifts <strong>in</strong> the film. This<br />

may lead to dim<strong>in</strong>ished <strong>in</strong>tensity from the sample <strong>and</strong> peak overlap, however. The second<br />

best option is to mount a stra<strong>in</strong>-free powder sample <strong>and</strong> carry out the same measurements<br />

as for the sample <strong>of</strong> <strong>in</strong>terest. The only rema<strong>in</strong><strong>in</strong>g error here is the sample missett<strong>in</strong>g; we<br />

will learn how to m<strong>in</strong>imize this error.<br />

If either <strong>of</strong> these procedures are performed, the data should be reliable, <strong>and</strong> none <strong>of</strong> the<br />

corrections outl<strong>in</strong>ed <strong>in</strong> this appendix are required. However, I would advise those <strong>in</strong>terested<br />

<strong>in</strong> XSA at least to skim this section to underst<strong>and</strong> what the major sources <strong>of</strong> error are <strong>and</strong><br />

how best to avoid them. The ma<strong>in</strong> lessons from this section are to m<strong>in</strong>imize �X (the<br />

sample missett<strong>in</strong>g from the 2θ-axis), to work at high 2θ, <strong>and</strong> to work at +ψ tilt angles<br />

when us<strong>in</strong>g the ω-diffractometer.<br />

152


Appendix B: X-Ray Errors 153<br />

We must def<strong>in</strong>e our angles carefully. ψ is, <strong>in</strong> general, the tilt <strong>of</strong> the sample normal with<br />

respect to the scatter<strong>in</strong>g vector. When work<strong>in</strong>g with an ω-diffractometer, i.e., work<strong>in</strong>g <strong>in</strong><br />

ω-mode, ψ = �ω. When work<strong>in</strong>g <strong>in</strong> ψ-mode, ψ = χ. (�χ is unnecessary here because<br />

χ = 0 implies that the sample normal lies <strong>in</strong> the diffractometer plane). Also, I occasionally<br />

refer to η, the angle <strong>of</strong> rotation <strong>of</strong> the sample normal around the scatter<strong>in</strong>g vector [78]. η<br />

can be described <strong>in</strong> terms <strong>of</strong> �ω <strong>and</strong> χ but will be used only as a “switch” between ω-mode<br />

<strong>and</strong> ψ-mode: η = π for the ω-mode <strong>and</strong> π/2 for the ψ-mode.<br />

B.1 M<strong>and</strong>atory Corrections<br />

B.1.1 Lorenz-Polarization-Absorption Corrections to the Intensity<br />

The LPA corrections are angle-dependent modifications to the <strong>in</strong>tensity. One should<br />

divide the measured <strong>in</strong>tensity by the follow<strong>in</strong>g expressions:<br />

PR<br />

� �� �<br />

1 + cos 2 2θ| cos 2θm|<br />

s<strong>in</strong> 2 θ<br />

� �� �<br />

LR<br />

for an imperfect monochromator or<br />

PR<br />

� �� �<br />

1 + cos 2 2θ cos 2 2θm<br />

s<strong>in</strong> 2 θ<br />

� �� �<br />

LR<br />

(1 + tan ψ cot θ cos η)<br />

� �� �<br />

AR<br />

(1 + tan ψ cot θ cos η)<br />

� �� �<br />

AR<br />

(B.1)<br />

(B.2)<br />

for a perfect monochromator. Both these expressions refer to a “reflection” diffraction<br />

geometry rather than transmission.<br />

For a specimen <strong>of</strong> non-zero thickness (such as a th<strong>in</strong> film), the absorption correction is


Appendix B: X-Ray Errors 154<br />

different. In such <strong>in</strong>stances, we multiply the above AR by<br />

� �<br />

1 − exp −µt<br />

1<br />

s<strong>in</strong> θ cos ψ − cos θ s<strong>in</strong> ψ cos η +<br />

��<br />

1<br />

s<strong>in</strong> θ cos ψ + cos θ s<strong>in</strong> ψ cos η<br />

(B.3)<br />

where t is the film thickness <strong>and</strong> µ is the l<strong>in</strong>ear absorption coefficient. This can be sig-<br />

nificant for very th<strong>in</strong> films. For th<strong>in</strong> films <strong>in</strong> transmission, the total absorption correction<br />

is<br />

AT =<br />

exp(−µt/ cos αout)<br />

µ cos α<strong>in</strong><br />

�<br />

1 1<br />

cos − α<strong>in</strong> cos αout<br />

where α<strong>in</strong> <strong>and</strong> αout are the <strong>in</strong>com<strong>in</strong>g <strong>and</strong> outgo<strong>in</strong>g angles.<br />

B.1.2 Refraction Corrections<br />

�<br />

�<br />

1 − e −µt<br />

�<br />

1<br />

cos α −<br />

<strong>in</strong> 1<br />

��<br />

cos αout<br />

(B.4)<br />

Due to refraction, there is a correction to the 2θ position when work<strong>in</strong>g at low <strong>in</strong>cidence<br />

angles 1 [225]:<br />

2θcorrect = (2θmeasured − α<strong>in</strong>) + 1 √ 2<br />

��� α 2 <strong>in</strong> − θ 2 �2 c + 4β 2<br />

�1/2 − θ 2 c + α2 �1/2 <strong>in</strong><br />

(B.5)<br />

where β = (λx−rayµ)/(4π) is the imag<strong>in</strong>ary part <strong>of</strong> the <strong>in</strong>dex <strong>of</strong> refraction, <strong>and</strong> θc is<br />

the critical angle. The values <strong>of</strong> β <strong>and</strong> θc can be found on the CXRO homepage, www-<br />

cxro.lbl.gov. There is a similar correction for graz<strong>in</strong>g exit mode: 2<br />

2θcorrect = (2θmeasured − αout) + 1 √ 2<br />

where αout is the exit angle.<br />

��� α 2 out + θ 2 �2 c + 4β 2<br />

�1/2 + θ 2 c + α2 �1/2 out<br />

1 This correction is not needed for graz<strong>in</strong>g <strong>in</strong>cidence diffraction [148].<br />

2 Caveat lector: I derived this formula myself.<br />

(B.6)


Appendix B: X-Ray Errors 155<br />

B.2 Geometric Errors<br />

B.2.1 Receiv<strong>in</strong>g Position Errors<br />

Follow<strong>in</strong>g Wilson [251, 252], we def<strong>in</strong>e the error �2θ by<br />

2θcorrect = 2θ + �2θ, (B.7)<br />

where for defocus<strong>in</strong>g errors associated with the receiv<strong>in</strong>g position <strong>of</strong> the detector,<br />

�2θdefocus = δr + δfr + δrs<br />

. (B.8)<br />

s<strong>in</strong> 2ϕ<br />

The terms δr + δfr + δrs have been made explicit by James <strong>and</strong> Cohen [112]. They refer to<br />

errors associated with the receiv<strong>in</strong>g position, correlated errors from the receiv<strong>in</strong>g position<br />

<strong>and</strong> x-ray focal po<strong>in</strong>t, <strong>and</strong> correlated errors from the receiv<strong>in</strong>g position <strong>and</strong> sample position,<br />

respectively. The average error due to these is given by ≈ 1/3 <strong>of</strong><br />

�2θdefocus = R −2<br />

sd (Rsd 〈yr〉−〈xr〉〈yr〉+ 〈y2 r 〉<br />

cot 2θ (B.9)<br />

2<br />

−〈xr xs〉 cos(2θ − ψw)<br />

−〈yr ys〉 cos(2θ − ψw)<br />

−〈xr ys〉 s<strong>in</strong>(2θ − ψw)<br />

+ 〈yr xs〉<br />

s<strong>in</strong> 2θ<br />

�<br />

cos(4θ − ψw) + Rsd<br />

+ Rsd<br />

〈yr yf〉 cot 2θ)<br />

Rxs<br />

where I have assumed 〈xr〉〈yr〉 =〈xryr〉,<br />

Rxs is the x-ray focus to sample distance,<br />

Rsd is the sample to detector distance,<br />

Rxs<br />

�<br />

cos ψw cos 2θ


Appendix B: X-Ray Errors 156<br />

ψ = ∆ω<br />

O<br />

ψ = 0°<br />

Rxs<br />

Rsd<br />

Figure B.1: The focus<strong>in</strong>g circle.<br />

xf <strong>and</strong> yf are the deviations <strong>of</strong> the x-ray focus from po<strong>in</strong>t A,<br />

xs <strong>and</strong> ys are the deviations <strong>of</strong> the sample surface from O,<br />

xr <strong>and</strong> yr are the deviations <strong>of</strong> the PSD from B<br />

<strong>and</strong><br />

ψw = θ + �ω.<br />

See Figure B.1 for A, B, <strong>and</strong> O. It may be obvious from the formulation that a good<br />

estimate <strong>of</strong> this error is difficult to obta<strong>in</strong>. The formulation is good from a heuristic po<strong>in</strong>t<br />

<strong>of</strong> view, however; maximiz<strong>in</strong>g Rsd <strong>and</strong> the 2θ angle reduces defocus<strong>in</strong>g errors.<br />

B<br />

q<br />

A


Appendix B: X-Ray Errors 157<br />

B.2.2 Horizontal Divergence Errors<br />

S<strong>in</strong>ce the <strong>in</strong>com<strong>in</strong>g beam has f<strong>in</strong>ite divergence, the maximum <strong>in</strong> diffracted <strong>in</strong>tensity<br />

varies with the sample-detector geometry. The peak shift <strong>in</strong> degrees (for a flat plate sample)<br />

is given by ≈ 1/3<strong>of</strong><br />

�2θdiv = 360<br />

π<br />

Rxs<br />

Rsd<br />

tan<br />

� δ<br />

2<br />

⎛<br />

�<br />

⎜<br />

s<strong>in</strong>(θ − �ω + δ/2) s<strong>in</strong>(θ − �ω − δ/2) ⎟<br />

⎝<br />

− ⎟<br />

s<strong>in</strong>(θ + �ω − δ/2) s<strong>in</strong>(θ + �ω + δ/2) ⎠ .<br />

� �� � � �� �<br />

(B.10)<br />

BO<br />

AO<br />

If the reader is us<strong>in</strong>g Noyan <strong>and</strong> Cohen [176], �ω + θ − δ/2 = �NC <strong>and</strong> π − �ω −<br />

θ − δ/2 = χNC, <strong>and</strong> the two terms <strong>in</strong> the parentheses are their BO <strong>and</strong> AO, respectively.<br />

Note that their error is much smaller s<strong>in</strong>ce they move the detector to an optimized position.<br />

Horizontal errors can be significant if us<strong>in</strong>g small <strong>in</strong>cidence angles; aga<strong>in</strong>, a stra<strong>in</strong>-free<br />

powder should be used to correct for this error.<br />

B.2.3 Alignment Errors<br />

Specimen <strong>and</strong> Beam Displacement<br />

For the ω-mode, the error <strong>in</strong> 2θ due to specimen <strong>and</strong> beam displacement is given by<br />

�2θω−disp ≈ 180 s<strong>in</strong> θ(2�Xω cos θ − �eω s<strong>in</strong>(θ − �ω))<br />

π<br />

Rsd s<strong>in</strong>(θ + �ω)<br />

⎞<br />

(B.11)<br />

where �Xω <strong>and</strong> �eω are the missett<strong>in</strong>gs <strong>of</strong> the specimen surface <strong>and</strong> <strong>in</strong>cident beam with<br />

respect to the 2θ-rotation axis, respectively. This effect is largest for low Bragg angles. The<br />

specimen displacement, �Xω, also affects the calibration constant � (described <strong>in</strong> section<br />

B.3.3), <strong>and</strong> is extremely important to m<strong>in</strong>imize. This error is negligible for systems with<br />

both <strong>in</strong>com<strong>in</strong>g <strong>and</strong> outgo<strong>in</strong>g parallel x-rays [176].


Appendix B: X-Ray Errors 158<br />

The result<strong>in</strong>g 2θ shift (�2θ�ω − �2θ0) due to these errors is given by [41]<br />

δ(�2θ) = 180<br />

�<br />

2 cos θ s<strong>in</strong> θ s<strong>in</strong> �ω<br />

π Rsd s<strong>in</strong>(�ω + θ) �eω<br />

�<br />

� �<br />

s<strong>in</strong> θ<br />

− 1 −<br />

�Xω . (B.12)<br />

s<strong>in</strong>(θ + �ω)<br />

Assum<strong>in</strong>g the beam displacement is negligible, Cohen [40] developed an easy way to<br />

check the sample displacement for cubic materials us<strong>in</strong>g the equation:<br />

ahkl − a0<br />

a0<br />

=− �Xω<br />

Rsd<br />

cos2 θ<br />

. (B.13)<br />

s<strong>in</strong> θ<br />

By determ<strong>in</strong><strong>in</strong>g the lattice parameter ahkl for various reflections at �ω = 0 <strong>and</strong> plott<strong>in</strong>g<br />

ahkl vs. cos 2 θ/s<strong>in</strong> θ, the fitted slope is equal to �Xωa0/Rsd. a0 is given by the <strong>in</strong>tercept.<br />

A positive slope <strong>in</strong>dicates that the sample is displaced to a position beh<strong>in</strong>d the center <strong>of</strong><br />

the diffractometer. I <strong>of</strong>ten use this method for FCC th<strong>in</strong> films with the (111) <strong>and</strong> (222)<br />

reflections.<br />

For materials with low stack<strong>in</strong>g fault energies, the Cohen method <strong>of</strong> elim<strong>in</strong>at<strong>in</strong>g sample<br />

displacement may not work. For such cases, see Noyan <strong>and</strong> Cohen [176].<br />

For the ψ-mode, the error due to sample <strong>and</strong> beam displacement <strong>of</strong>f the ψ-rotation axis<br />

is given by:<br />

�2θψ−disp = 180<br />

π<br />

2<br />

Rsd<br />

cos θ � �<br />

�Xψ − s<strong>in</strong> χ�eψ<br />

cos χ<br />

(B.14)<br />

where �eψ is the missett<strong>in</strong>g <strong>of</strong> the <strong>in</strong>cident beam with the ψ-rotation axis <strong>and</strong> �Xψ is the<br />

missett<strong>in</strong>g <strong>of</strong> the sample surface with respect to the ψ-rotation axis.<br />

The shift �2θχ − �2θ0 is given by [41]:<br />

δ(�2θ) = 180<br />

π<br />

2<br />

Rsd<br />

cos θ<br />

��<br />

1 − 1<br />

�<br />

�<br />

�Xψ + �eψ tan χ<br />

cos χ<br />

(B.15)<br />

Generally, we should use low θ <strong>and</strong> high χ to measure �eψ <strong>and</strong> either elim<strong>in</strong>ate or account<br />

for it.


Appendix B: X-Ray Errors 159<br />

Effect <strong>of</strong> the ω-Axis Not Correspond<strong>in</strong>g to the 2θ-Axis<br />

For the ω-diffractometer, this <strong>of</strong>fset gives an error<br />

where X ′ is the <strong>of</strong>fset.<br />

�2θω−axis = 180 (�X<br />

π<br />

′ ) s<strong>in</strong> 2θ(1 − cos �ω)<br />

Rsd s<strong>in</strong>(θ + �ω)<br />

δ(�2θ) = 180 (�X<br />

π<br />

′ ) s<strong>in</strong> 2θ(1 − cos �ω)<br />

Rsd s<strong>in</strong>(θ + �ω)<br />

(B.16)<br />

(B.17)<br />

In general, this error should be negligible if the Eulerian cradle is mounted properly<br />

onto the two-circle base. This mount<strong>in</strong>g should be performed at a mach<strong>in</strong>e shop us<strong>in</strong>g a<br />

dial counter. This was not done for the <strong>Harvard</strong> four-circle, so �X ′ ≈ 25 µm.<br />

B.3 Corrections for a Position Sensitive Detector<br />

B.3.1 Parallax Corrections for a PSD<br />

or<br />

Accord<strong>in</strong>g to James <strong>and</strong> Cohen [112], the true value <strong>of</strong> 2θ is given by<br />

2θ = 2θPSD + arctan(n tan �) (B.18)<br />

� �<br />

nz<br />

2θ = 2θPSD + arctan<br />

Rsd<br />

(B.19)<br />

where 2θPSD is the 2θ value at which lies the center channel <strong>of</strong> the PSD, n is the number<br />

<strong>of</strong> pixels away from 2θPSD, z is the length per pixel <strong>and</strong> � is the number <strong>of</strong> degrees 2θ per<br />

pixel for the PSD.


Appendix B: X-Ray Errors 160<br />

B.3.2 Small Beam Spot Corrections<br />

When us<strong>in</strong>g a PSD, one needs to oscillate the sample (usually around ω) by a couple <strong>of</strong><br />

degrees <strong>in</strong> order to sample enough gra<strong>in</strong>s for good statistics [111].<br />

B.3.3 PSD Calibration Constant Errors<br />

The calibration constants the position along the PSD <strong>in</strong>to degrees 2θ <strong>and</strong> depends on<br />

the sample to detector distance; if there is non-zero �Xω, the calibration will be <strong>of</strong>f. The<br />

calibration constant is given by<br />

� =<br />

second peak − peak at center channel (<strong>in</strong> degrees)<br />

channel difference<br />

(B.20)<br />

where �(Rsd) is determ<strong>in</strong>ed for a fixed Rsd us<strong>in</strong>g a calibration sample. If the next sample<br />

is misset by �Xω, the calculated peak shift will be error due to an <strong>in</strong>correct �(Rsd) given<br />

by<br />

σcal = �2θ �(Rsd) − �(Rsd + �Xω)<br />

. (B.21)<br />

�(Rsd)<br />

In general, I ignore this correction by ensur<strong>in</strong>g that �Xω is small.<br />

B.4 Count<strong>in</strong>g <strong>and</strong> Fitt<strong>in</strong>g Errors<br />

The variance <strong>in</strong> d as a function <strong>of</strong> variance <strong>in</strong> the 2θ peak position is given by [255]<br />

σ 2 (d) =<br />

where σ(2θ) is given <strong>in</strong> degrees.<br />

� �2 λ cos θp σ 2 (2θp)<br />

2 s<strong>in</strong> 2 θp<br />

4<br />

�<br />

π<br />

�2 180<br />

(B.22)


Appendix C<br />

Alignment Procedure for the <strong>Harvard</strong><br />

Four-Circle Diffractometer<br />

In general, alignment consists <strong>of</strong> three major steps:<br />

• alignment <strong>of</strong> the diffractometer with itself,<br />

• alignment <strong>of</strong> the x-rays with the diffractometer (2θ)-axis, <strong>and</strong><br />

• alignment <strong>of</strong> the sample surface with the 2θ-axis.<br />

The first step refers to ensur<strong>in</strong>g the Eulerian cradle is properly mounted onto the two-circle<br />

base. We focus on the latter two steps.<br />

C.1 Coarse Alignment<br />

We can beg<strong>in</strong> with Geremia’s method [81]. This <strong>in</strong>volves first align<strong>in</strong>g a p<strong>in</strong>hole colli-<br />

mator <strong>in</strong> place <strong>of</strong> the sample <strong>and</strong> then maximiz<strong>in</strong>g the <strong>in</strong>tensity <strong>of</strong> the beam com<strong>in</strong>g through<br />

161


Appendix C: Alignment Procedure for the <strong>Harvard</strong> Four-Circle Diffractometer 162<br />

the p<strong>in</strong>hole.<br />

C.2 F<strong>in</strong>e Alignment<br />

Good studies on the elim<strong>in</strong>ation <strong>of</strong> both sample <strong>and</strong> x-ray tube missett<strong>in</strong>gs were com-<br />

pleted by Convert <strong>and</strong> Miege [41] <strong>and</strong> Vermeulen <strong>and</strong> Houtman [235]. This section is based<br />

primarily on their work.<br />

Align<strong>in</strong>g the x-rays with the diffractometer axis can be performed by mount<strong>in</strong>g a stan-<br />

dard <strong>in</strong>to the diffractometer (such as NIST’s Si powder 640) <strong>and</strong> perform<strong>in</strong>g the follow<strong>in</strong>g<br />

procedure.<br />

We refer back to an equation <strong>in</strong> the previous section:<br />

δ(�2θ) = 180 2 cos θ<br />

π Rsd<br />

� s<strong>in</strong> θ s<strong>in</strong> �ω<br />

s<strong>in</strong>(�ω + θ) �eω −<br />

�<br />

1 −<br />

� �<br />

s<strong>in</strong> θ<br />

�Xω<br />

s<strong>in</strong>(θ + �ω)<br />

(C.1)<br />

where �eω is the displacement <strong>of</strong> the x-ray tube from the diffractometer axis <strong>and</strong> �Xω is<br />

the displacement <strong>of</strong> the sample surface from the diffractometer axis. The goal here is to<br />

move the x-ray tube <strong>in</strong> the horizontal plane until �eω = 0. One can do this even when<br />

there exists a f<strong>in</strong>ite �Xω by tilt<strong>in</strong>g the sample by an amount �ωc where<br />

1 −<br />

the solution is simply (besides �ω = 0)<br />

s<strong>in</strong> θ<br />

s<strong>in</strong>(θ + �ωc)<br />

= 0; (C.2)<br />

�ωc = π − 2θ. (C.3)<br />

When �ω = �ωc, the shift <strong>in</strong> 2θ is zero when �eω = 0. �Xω can be zeroed afterwards.<br />

Unfortunately, there is no similar method with �eψ <strong>and</strong> the equation<br />

δ(�2θ) = 180<br />

π<br />

2<br />

��<br />

cos θ 1 − 1<br />

�<br />

�<br />

�Xψ + �eψ tan χ .<br />

cos χ<br />

(C.4)<br />

Rsd


Appendix C: Alignment Procedure for the <strong>Harvard</strong> Four-Circle Diffractometer 163<br />

We have to determ<strong>in</strong>e �eψ <strong>and</strong> �Xψ by fitt<strong>in</strong>g to the diffraction results for an unstra<strong>in</strong>ed<br />

sample to this equation.<br />

C.3 Align<strong>in</strong>g Samples<br />

The first th<strong>in</strong>g to do after mount<strong>in</strong>g a sample is to ensure that the sample surface normal<br />

lies <strong>in</strong> the diffraction plane by us<strong>in</strong>g the laser. Reflect the laser <strong>of</strong>f the surface <strong>and</strong> mark its<br />

position on a backdrop. Rotate the sample around its normal by 180 ◦ <strong>and</strong> aga<strong>in</strong> mark the<br />

reflected spot. The midpo<strong>in</strong>t between the two spots is the ideal location. Adjust the sample<br />

goniometer accord<strong>in</strong>gly.<br />

Once �eω <strong>and</strong> �eψ are zero, the most important adjustment is to m<strong>in</strong>imize �Xω. This<br />

can be done with Cohen’s method, as discussed <strong>in</strong> the previous section. Many th<strong>in</strong> films<br />

are textured, however, <strong>and</strong> only one or two peaks are available (or the two peaks differ<br />

<strong>in</strong> stra<strong>in</strong>). As stated <strong>in</strong> Appendix B, usually the (111) <strong>and</strong> (222) reflections can be used.<br />

Also, we can take advantage <strong>of</strong> films with smooth surfaces <strong>and</strong> align the sample by x-ray<br />

reflectivity. Move the sample <strong>in</strong> the diffraction plane until the surface splits the ma<strong>in</strong> beam<br />

<strong>in</strong> half. Then rotate the detector to 1 ◦ <strong>and</strong> adjust ω (or whatever rotates the sample normal <strong>in</strong><br />

the diffraction plane) such that you obta<strong>in</strong> a reflection peak at 1 ◦ . Then set ω to 0.5 ◦ . Br<strong>in</strong>g<br />

the detector back to 0 ◦ <strong>and</strong> split the beam <strong>in</strong> half aga<strong>in</strong>. Keep reiterat<strong>in</strong>g this procedure<br />

until everyth<strong>in</strong>g agrees.


Appendix C: Alignment Procedure for the <strong>Harvard</strong> Four-Circle Diffractometer 164<br />

C.4 Calibrat<strong>in</strong>g the Position Sensitive Detector<br />

First, make sure the x-ray tube is aligned well. The PSD is then calibrated by the<br />

follow<strong>in</strong>g rout<strong>in</strong>e.<br />

1. Guess the start<strong>in</strong>g parameters for the PSD. Right now, the center channel <strong>of</strong> the PSD<br />

is ≈ 868, tan � = 0.00027, z = 0.04003 mm/chn, Rxs = 147 mm, <strong>and</strong> Rsd =<br />

147.702 mm.<br />

2. Mount <strong>and</strong> align the powder st<strong>and</strong>ard so that �X = 0.<br />

3. Choose a detector position featur<strong>in</strong>g two high 2θhkl peaks, one at the center channel<br />

<strong>and</strong> another <strong>of</strong>f to the left or right side <strong>of</strong> the PSD. Adjust the detector so that one<br />

<strong>of</strong> the diffraction peaks lies at the center channel; record this scan. Then pull the<br />

detector away from the sample <strong>and</strong> perform a distant scan. F<strong>in</strong>ally, push the detector<br />

back <strong>in</strong>to the previous position for a scan; the peak should stay at the same channel<br />

for both the far <strong>and</strong> the near scans! If it doesn’t, adjust the slide on which the de-<br />

tector sits until the detector slide is aligned with path <strong>of</strong> the diffracted x-rays. This<br />

determ<strong>in</strong>es the true “center channel”.<br />

4. Now that we know the sample <strong>and</strong> x-ray tube are perfectly adjusted, <strong>and</strong> now that<br />

we know the precise value <strong>of</strong> the center channel, we can use this to obta<strong>in</strong> the exact<br />

values <strong>of</strong> z <strong>and</strong> Rsd. Wefirst ensure the same diffraction peak as before lies on the<br />

center channel. Then we pull the detector out by a known amount <strong>and</strong> take a scan;<br />

this is followed by push<strong>in</strong>g the detector <strong>in</strong>, <strong>and</strong> tak<strong>in</strong>g a scan. We can use this data<br />

<strong>and</strong> value <strong>of</strong> ast<strong>and</strong>ard to obta<strong>in</strong> z <strong>and</strong> Rsd.


Appendix D<br />

Elastic Constants<br />

Throughout this appendix, we will use the prime ( ′ ) to denote quantities <strong>in</strong> the sample<br />

coord<strong>in</strong>ate system <strong>and</strong> no prime to denote quantities <strong>in</strong> the crystal coord<strong>in</strong>ate system.<br />

D.1 Important Elastic Quantities<br />

For cubic s<strong>in</strong>gle crystals, the Young modulus (E), Poisson ratio (ν), <strong>and</strong> biaxial modu-<br />

lus (Y ) are def<strong>in</strong>ed by:<br />

E = (c′ 11 − c′ 12 )(c′ 11 + 2c′ 12 )<br />

c ′ 11 + c′ 12<br />

ν =<br />

c ′ 12<br />

c ′ 11 + c′ 12<br />

Y = c ′ 11 + c′ 12 − 2c2 13<br />

c ′ 33<br />

165<br />

= −s′ 12<br />

s ′ 11<br />

=<br />

= 1<br />

s ′ 11<br />

1<br />

s ′ 11 + s′ 12<br />

(D.1)<br />

(D.2)<br />

. (D.3)


Appendix D: Elastic Constants 166<br />

D.2 The Rotation Matrix<br />

When transform<strong>in</strong>g from the crystal coord<strong>in</strong>ate system to the sample coord<strong>in</strong>ate system,<br />

we use the Euler angles (ϕ1,�,ϕ2) where ϕ1 is the rotation around the [001] axis, � is the<br />

rotation around the new [100] ′ axis, <strong>and</strong> ϕ2 is the rotation around the new [001] ′′ axis. Such<br />

a rotation matrix is given by<br />

�<br />

a =<br />

cos ϕ1 cos ϕ2 − cos � s<strong>in</strong> ϕ1 s<strong>in</strong> ϕ2 cos ϕ2 s<strong>in</strong> ϕ1 + cos � cos ϕ1 s<strong>in</strong> ϕ2<br />

− cos � cos ϕ2 s<strong>in</strong> ϕ1 − cos ϕ1 s<strong>in</strong> ϕ2 cos � cos ϕ1 cos ϕ2 − s<strong>in</strong> ϕ1 s<strong>in</strong> ϕ2<br />

s<strong>in</strong> � s<strong>in</strong> ϕ1<br />

<strong>and</strong> any tensor quantity <strong>of</strong> order n can be transformed by<br />

− cos ϕ1 s<strong>in</strong> �<br />

s<strong>in</strong> � s<strong>in</strong> ϕ2<br />

�<br />

cos ϕ2 s<strong>in</strong> �<br />

cos �<br />

(D.4)<br />

M ′ = a n M. (D.5)<br />

One can then easily determ<strong>in</strong>e the follow<strong>in</strong>g relationships for (111)-oriented crystals<br />

(ϕ1 =−π/4,�=−arccos(1/ √ 3), ϕ2 = 0)<br />

c ′ 11<br />

c ′ 33<br />

c ′ 12<br />

c ′ 13<br />

= 1<br />

2 (c11 + c12 + 2c44) (D.6)<br />

= 1<br />

3 (c11 + 2c12 + 4c44) (D.7)<br />

= 1<br />

6 (c11 + 5c12 − 2c44) (D.8)<br />

= 1<br />

3 (c11 + 2c12 − 2c44) (D.9)


Appendix D: Elastic Constants 167<br />

s ′ 12<br />

s ′ 11<br />

= 1<br />

4 (2s11 + 2s12 + s44) (D.10)<br />

= 1<br />

12 (2s11 + 10s12 − s44) . (D.11)<br />

D.3 Average Elastic Quantities For Polycrystall<strong>in</strong>e <strong>Ag</strong>gre-<br />

gates<br />

In general, a tensor property can be averaged by <strong>in</strong>tegrat<strong>in</strong>g over all space us<strong>in</strong>g the<br />

Euler angles <strong>and</strong> normaliz<strong>in</strong>g. For example, a fourth order tensor Mijkl can be averaged<br />

us<strong>in</strong>g the equation<br />

〈Mijkl〉= 1<br />

8π 2<br />

� ϕ1=2π � �=π � ϕ2=2π<br />

Mijkl(ϕ1,�,ϕ2) s<strong>in</strong> � dϕ1 d� dϕ2. (D.12)<br />

ϕ1=0 �=0 ϕ2=0<br />

With cubic symmetry, this simplifies to<br />

〈Mijkl〉= 4<br />

π 2<br />

� ϕ1=π/2 � �=π/2 � ϕ2=π/2<br />

ϕ1=0<br />

�=0<br />

ϕ2=0<br />

Mijkl(ϕ1,�,ϕ2) s<strong>in</strong> � dϕ1 d� dϕ2 (D.13)<br />

s<strong>in</strong>ce all the <strong>in</strong>formation is conta<strong>in</strong>ed with<strong>in</strong> an octant <strong>in</strong> space. If the material is textured,<br />

an orientation distribution function (ODF) f (ϕ1,�,ϕ2) (<strong>and</strong> its associated normalization<br />

constant) may be used as a weight<strong>in</strong>g factor.<br />

D.4 The Vook-Witt Model <strong>of</strong> Gra<strong>in</strong> Interaction<br />

If we take every gra<strong>in</strong> <strong>in</strong> the film to have equal stra<strong>in</strong> tensors (thus provid<strong>in</strong>g stress<br />

discont<strong>in</strong>uities at the gra<strong>in</strong> boundaries) we can average over the cmn <strong>and</strong> obta<strong>in</strong> a Voigt


Appendix D: Elastic Constants 168<br />

average. If we assume equal stress tensors <strong>in</strong> each gra<strong>in</strong> (thus violat<strong>in</strong>g the stra<strong>in</strong> com-<br />

patibility relations), we can average over the smn <strong>and</strong> obta<strong>in</strong> a Reuss average. These are<br />

called “gra<strong>in</strong> <strong>in</strong>teraction” models. Hill showed that the Voigt average is an upper bound<br />

while the Reuss average is a lower bound [100]. He thus proposed the Hill average for the<br />

elastic constants, 〈Mijkl〉=(1/2)(〈M Voigt<br />

ijkl 〉+〈MReuss<br />

ijkl 〉). Vook <strong>and</strong> Witt suggested a gra<strong>in</strong><br />

<strong>in</strong>teraction model particularly suited to th<strong>in</strong> films with a columnar morphology, where each<br />

gra<strong>in</strong> has the follow<strong>in</strong>g stra<strong>in</strong> elements [239]:<br />

ɛ ′ 11 = ɛ′ 22<br />

ɛ ′ 12 = ɛ′ 21<br />

σ ′<br />

i3 = σ ′ 3i<br />

= e (D.14)<br />

= 0 (D.15)<br />

= 0. (D.16)<br />

The rema<strong>in</strong><strong>in</strong>g stra<strong>in</strong> <strong>and</strong> stress tensor elements can be solved for us<strong>in</strong>g Hooke’s law. They<br />

are given by<br />

where<br />

σ ′ 1 =<br />

σ ′ 2 =<br />

�<br />

e s ′ 2<br />

26 − s ′ 16s′ 26 + s′ �<br />

66 s ′<br />

12 − s ′ �<br />

22<br />

�<br />

ζ<br />

�<br />

e s ′ 2<br />

16 − s ′ 16s′ 26 + s′ 66<br />

σ ′ 6 = e � s ′ 11s′ 26 + s′ 22s′ 16 − s′ 12<br />

2ζ<br />

ζ<br />

�<br />

s ′<br />

12 − s ′ �<br />

11<br />

�<br />

�<br />

s ′<br />

16 + s ′ ��<br />

26<br />

ζ = s ′ 2<br />

16 s ′ 22 − 2s′ 12 s′ 16 s′ 26 + s′ 11 s′ 2<br />

26 + s ′ 2<br />

12 s ′ 66 − s′ 11 s′ 22 s′ 66<br />

(D.17)<br />

(D.18)<br />

(D.19)<br />

(D.20)


Appendix D: Elastic Constants 169<br />

<strong>and</strong> (us<strong>in</strong>g the above stresses)<br />

ɛ ′ i<br />

�<br />

= e 2s ′ 2<br />

16s ′ 2i − 2s′ 16s′ 26 (s′ 1i + s′ 2i ) − s′ 16s′ i6 (s′ 12 − s′ 22 ) +<br />

where i = 3, 4, 5.<br />

�<br />

/<br />

�<br />

2(s ′ 2<br />

16s ′ 22 − 2s′ 12s′ 16s′ 26 + s′ 11s′ 2<br />

26 + s ′ 2<br />

12s ′ 66 − s′ 11s′ 22s′ 66 )<br />

�<br />

2s ′ 1i (s′ 2<br />

26 + s ′ 12 s′ 66 − s′ 22 s′ 66 ) + (s′ 11 − s′ 12 )(s′ 26 s′ i6 − 2s′ 2i s′ 66 )<br />

(D.21)<br />

The Vook-Witt model may have important consequences for x-ray diffraction work as<br />

well as for elastic parameters [229], but this rema<strong>in</strong>s to be corroborated.<br />

D.4.1 The Hornstra-Bartels Calculation Technique<br />

For the follow<strong>in</strong>g elastic calculations, the method <strong>of</strong> Hornstra <strong>and</strong> Bartels [102, 202]<br />

has been used. This method speeds up the calculations considerably. Accord<strong>in</strong>g to this<br />

formulation (which relies on the Vook-Witt model), the stra<strong>in</strong> <strong>in</strong> each gra<strong>in</strong> is given by<br />

ɛij = eδij + aiℓj<br />

(D.22)<br />

where δij is the Krönecker delta, ℓ j is the <strong>in</strong>terface normal (unit vector), <strong>and</strong> ai is a vector<br />

to be solved for.<br />

The stra<strong>in</strong> elements are found by us<strong>in</strong>g equation (D.22). We can simplify the <strong>in</strong>dex


Appendix D: Elastic Constants 170<br />

notation us<strong>in</strong>g the def<strong>in</strong>itions<br />

ɛ1 = ɛ11 (D.23)<br />

ɛ2 = ɛ22 (D.24)<br />

ɛ3 = ɛ33 (D.25)<br />

ɛ4 = ɛ23 + ɛ32 (D.26)<br />

ɛ5 = ɛ13 + ɛ31 (D.27)<br />

ɛ6 = ɛ12 + ɛ21. (D.28)<br />

We can then write the stress elements as (aga<strong>in</strong> for cubic crystals)<br />

σ1 = c11ɛ1 + c12(ɛ2 + ɛ3) (D.29)<br />

σ2 = c11ɛ2 + c12(ɛ3 + ɛ1) (D.30)<br />

σ3 = c11ɛ3 + c12(ɛ1 + ɛ2) (D.31)<br />

σ4 = c44ɛ4 (D.32)<br />

σ5 = c44ɛ5 (D.33)<br />

σ6 = c44ɛ6. (D.34)<br />

The solution for ai can be obta<strong>in</strong>ed by us<strong>in</strong>g the above equations <strong>and</strong> aga<strong>in</strong> not<strong>in</strong>g that the<br />

components <strong>of</strong> force normal to the sample are zero:<br />

σ1ℓ1 + σ6ℓ2 + σ5ℓ3 = 0 (D.35)<br />

σ6ℓ1 + σ2ℓ2 + σ4ℓ3 = 0 (D.36)<br />

σ5ℓ1 + σ4ℓ2 + σ3ℓ3 = 0. (D.37)<br />

This gives ai <strong>in</strong> terms <strong>of</strong> ℓ j, e, c11, c12, <strong>and</strong> c44. The stress <strong>and</strong> stra<strong>in</strong> elements <strong>in</strong> the


Appendix D: Elastic Constants 171<br />

sample coord<strong>in</strong>ates are then given by<br />

<strong>and</strong><br />

σ ′<br />

ij = aika jlσkl<br />

D.4.2 Results with the Vook-Witt Model<br />

(D.38)<br />

ɛ ′ ij = aika jlɛkl. (D.39)<br />

Us<strong>in</strong>g the Vook-Witt model <strong>and</strong> the Hornstra-Bartels calculation method, we can deter-<br />

m<strong>in</strong>e some useful elastic properties for anisotropic materials.<br />

Follow<strong>in</strong>g the example <strong>of</strong> Murakami <strong>and</strong> Chaudhari [166], I calculated the ratio <strong>of</strong><br />

out-<strong>of</strong>-plane stra<strong>in</strong> to <strong>in</strong>-plane stra<strong>in</strong> (Figure D.1) <strong>and</strong> the biaxial modulus (Figure D.2)<br />

<strong>and</strong> mapped them onto a (111) stereographic projection (the (111) plane is parallel to the<br />

figure).<br />

For FCC materials such as <strong>Ni</strong> <strong>and</strong> <strong>Cu</strong>, slip preferentially occurs on the {111} planes <strong>in</strong><br />

the 〈110〉 directions. The maximum resolved shear stress <strong>in</strong> a crystal with a surface normal<br />

[hkl] is shown <strong>in</strong> Figure D.3 for an arbitrary value <strong>of</strong> ɛ ′ 11<br />

= 0.704% (orig<strong>in</strong>ally used to<br />

compare my results with those <strong>of</strong> Murakami <strong>and</strong> Chaudhari). This shows that orientations<br />

with the {331} <strong>and</strong> {110} family <strong>of</strong> planes parallel to the substrate yield before the {111}<br />

family, which <strong>in</strong> turn yield before the {100} family.


Appendix D: Elastic Constants 172<br />

<strong>Ni</strong><br />

112<br />

<strong>Cu</strong><br />

112<br />

101 313<br />

113<br />

101 313<br />

113<br />

211<br />

001<br />

211<br />

001<br />

111<br />

113<br />

111<br />

113<br />

123<br />

123<br />

110<br />

111<br />

110<br />

111<br />

010<br />

110<br />

010<br />

110<br />

131<br />

121<br />

131<br />

131<br />

121<br />

131<br />

111<br />

111<br />

Figure D.1: The ratio <strong>of</strong> ɛ ′ 33 /ɛ′ 11 <strong>in</strong> the Vook-Witt model.<br />

133<br />

133<br />

011<br />

011<br />

112<br />

112


Appendix D: Elastic Constants 173<br />

112<br />

112<br />

<strong>Ni</strong><br />

101 313<br />

<strong>Cu</strong><br />

113<br />

101 313<br />

113<br />

211<br />

001<br />

211<br />

001<br />

111<br />

113<br />

111<br />

113<br />

123<br />

123<br />

Figure D.2: The biaxial modulus <strong>in</strong> units <strong>of</strong> 100 GPa us<strong>in</strong>g the Vook-Witt model.<br />

110<br />

111<br />

110<br />

111<br />

010<br />

110<br />

010<br />

110<br />

131<br />

121<br />

131<br />

131<br />

121<br />

131<br />

111<br />

111<br />

133<br />

133<br />

011<br />

011<br />

112<br />

112


Appendix D: Elastic Constants 174<br />

<strong>Ni</strong><br />

112<br />

<strong>Cu</strong><br />

112<br />

101 313<br />

113<br />

101 313<br />

113<br />

211<br />

001<br />

211<br />

001<br />

111<br />

113<br />

111<br />

113<br />

133<br />

133<br />

011<br />

011<br />

Figure D.3: The maximum resolved shear stress on the (111) slip plane given <strong>in</strong> GPa us<strong>in</strong>g<br />

ɛ ′ 11 = 0.704%.<br />

110<br />

111<br />

110<br />

111<br />

010<br />

110<br />

010<br />

110<br />

131<br />

131<br />

121<br />

131<br />

121<br />

131<br />

111<br />

111<br />

133<br />

133<br />

011<br />

011<br />

112<br />

112


Appendix E<br />

The Stoney Equation<br />

We consider a th<strong>in</strong> film on a substrate, with the bottom layer <strong>of</strong> the substrate as the<br />

z = 0 reference plane [66]. The lattice parameter <strong>of</strong> the composite is allowed to vary such<br />

that the misfit stra<strong>in</strong> ɛm 11 is given by<br />

From the stra<strong>in</strong> compatibility equations,<br />

ɛ m 11 = d0 − d(z)<br />

. (E.1)<br />

d(z)<br />

ɛij(z) = ɛ m ij (z) − zκij(t) + ɛ r ij (t) (E.2)<br />

where κij(t) is the curvature <strong>of</strong> the reference plane at a total composite thickness t <strong>and</strong><br />

ɛr ij (t) is the stra<strong>in</strong> <strong>of</strong> the reference plane at thickness t.<br />

In what follows, we will focus on ɛ11(z) <strong>and</strong> therefore use the notation<br />

ɛ m 11 = ɛm 22<br />

ɛ r 11 = ɛr 22<br />

= ɛm<br />

= ɛr<br />

κ11 = κ22 = κ.<br />

175


Appendix E: The Stoney Equation 176<br />

If we assume the film has a constant misfit stra<strong>in</strong> ɛm <strong>and</strong> the substrate has none,<br />

<strong>and</strong><br />

Then we use force <strong>and</strong> moment balance:<br />

<strong>and</strong><br />

The force balance term written out is<br />

Ys<br />

� ts<br />

<strong>and</strong> the moment balance term is<br />

Ys<br />

0<br />

� ts<br />

0<br />

ɛ s 11 (z) =−zκ + ɛr<br />

(E.3)<br />

ɛ f 11 (z) = ɛm − zκ + ɛr. (E.4)<br />

� tf+ts<br />

dzσ11(z) = 0 (E.5)<br />

0<br />

� tf+ts<br />

dzσ11(z)z = 0. (E.6)<br />

0<br />

dzɛ s 11 (z) + Yf<br />

dzɛ s 11 (z)z + Yf<br />

� ts+hf<br />

ts<br />

� ts+hf<br />

ts<br />

dzɛ f 11 (z) = 0 (E.7)<br />

dzɛ f 11 (z)z = 0. (E.8)<br />

Substitut<strong>in</strong>g <strong>in</strong> the above equations for ɛs 11 (z) <strong>and</strong> ɛf 11 (z) <strong>and</strong> <strong>in</strong>tegrat<strong>in</strong>g gives two equations<br />

with the two unknowns κ <strong>and</strong> ɛr. Solv<strong>in</strong>g for these gives<br />

<strong>and</strong><br />

ɛr = ξYftf<br />

κ = 6ξYfYstfts(tf + ts)ɛm<br />

�<br />

Yst 2 s (3tf + 2ts) − Yft 3 �<br />

f<br />

(E.9)<br />

(E.10)


Appendix E: The Stoney Equation 177<br />

where<br />

ξ =<br />

When we use σ f 11 = Yfɛm we see that<br />

�<br />

Y 2<br />

f t4 f + Y 2 s t4 s + 2YfYstfts(2t 2 f + 3tfts + 2t 2 s )<br />

�−1 . (E.11)<br />

σ f 11 =<br />

κ<br />

. (E.12)<br />

6ξYstfts(tf + ts)<br />

Freund et al. [67] have a more attractive expression for equation (E.9):<br />

κ = 6ɛm<br />

ts<br />

where t = tf/ts <strong>and</strong> Y = Yf/Ys.<br />

�<br />

1 + t<br />

1 + tY(4 + 6t + 4t2 ) + t4Y 2<br />

�<br />

(E.13)<br />

Stoney’s approximation is used when tf ≪ ts (such that ξ → (Y 2 s t4 s )−1 ) so that this<br />

simplifies to<br />

<strong>and</strong><br />

σ f 11 = Yst 2 s<br />

6tf<br />

κ (E.14)<br />

ɛr = 2 Yftf<br />

. (E.15)<br />

Ysts<br />

Note that there is an implicit assumption <strong>in</strong> Stoney’s equation that the biaxial modulus<br />

<strong>of</strong> the film <strong>in</strong> near or less than that <strong>of</strong> the substrate; strictly speak<strong>in</strong>g, ξ really goes to<br />

(Y 2 s t4 s + 4YfYstft 3 s )−1 . If Yf is much greater than Ys (near (tsYs)/(4tf) — which may be<br />

possible for rigid films on polymer substrates), then one should use the approximations<br />

σ f 11 =<br />

�<br />

Yst 2 s<br />

+<br />

6tf<br />

2Yfts<br />

�<br />

κ (E.16)<br />

3<br />

<strong>and</strong><br />

ɛr =<br />

2Yftf<br />

. (E.17)<br />

Ysts + 4Yftf


Appendix E: The Stoney Equation 178<br />

<strong>and</strong><br />

When the film misfit stra<strong>in</strong> ɛm is not constant through the thickness <strong>of</strong> the film, we take<br />

ɛ s 11 (z) =−zκ + ɛr<br />

(E.18)<br />

ɛ f 11 (z) = ɛm(z) − zκ + ɛr. (E.19)<br />

This gives for κ <strong>and</strong> ɛr,<br />

κ =−6Yfξ(Yst 2 s + Yftf(tf<br />

� ts+tf<br />

+ 2ts)) dzɛm(z) +<br />

ts<br />

� ts+tf<br />

12Yfξ(Yftf + Ysts) dzɛm(z)z (E.20)<br />

ts<br />

<strong>and</strong><br />

ɛr =−4Yfξ(Yst 3 s + Yftf(t 2 f + 3tfts + 3t 2 s ))<br />

� ts+tf<br />

dzɛm(z) +<br />

ts<br />

6Yfξ(Yst 2 s + Yftf(tf<br />

� ts+tf<br />

+ 2ts)) dzɛm(z)z. (E.21)<br />

In the th<strong>in</strong> film approximation,<br />

κ = 6Yf<br />

Yst 2 � � ts+tf<br />

� ts+tf<br />

2<br />

dzɛm(z)z −<br />

s ts ts<br />

ts<br />

or<br />

or <strong>in</strong> terms <strong>of</strong> force per unit width,<br />

ts<br />

�<br />

dzɛm(z)<br />

(E.22)<br />

ɛm(tf) = Yst 2 s<br />

κ<br />

6Yf<br />

′ (tf) (E.23)<br />

ɛm(tf) = 1<br />

Yf<br />

� F ′ (tf)/w �<br />

(E.24)<br />

where κ ′ (tf) <strong>and</strong> F ′ (tf)w −1 are the slopes <strong>of</strong> the curvature or force per unit width versus<br />

thickness curves, respectively. This allows an extraction <strong>of</strong> the stra<strong>in</strong> pr<strong>of</strong>ile from curvature<br />

data.


Appendix F<br />

List <strong>of</strong> Symbols<br />

179


Appendix F: List <strong>of</strong> Symbols 180<br />

A area<br />

a lattice spac<strong>in</strong>g<br />

a0 equilibrium lattice spac<strong>in</strong>g<br />

aref reference lattice spac<strong>in</strong>g<br />

b Burgers vector<br />

cijkl elastic stiffness<br />

cos λ cos φ Schmid factor<br />

Dδ misfit dislocation spac<strong>in</strong>g<br />

d gra<strong>in</strong> size<br />

dα, dβ out-<strong>of</strong>-plane d-spac<strong>in</strong>gs <strong>of</strong> the α or β-phase<br />

E Young’s modulus<br />

Eact activation energy<br />

F force<br />

fij <strong>in</strong>terface stress<br />

fhkl volume fraction <strong>of</strong> hkl-oriented gra<strong>in</strong>s<br />

G shear modulus or Gibbs free energy<br />

GID graz<strong>in</strong>g <strong>in</strong>cidence diffraction<br />

GND geometrically necessary dislocation<br />

gij relative <strong>in</strong>terface stress<br />

Mcomp composite modulus <strong>of</strong> the film <strong>and</strong> substrate<br />

Nα, Nβ number <strong>of</strong> monolayers <strong>of</strong> the α or β-phase<br />

P hydrostatic pressure<br />

q scatter<strong>in</strong>g vector<br />

RBS Rutherford Backscatter<strong>in</strong>g Spectrometry<br />

SSD statistically stored dislocation<br />

sijkl elastic compliance<br />

Tij thermodynamic tensions<br />

T temperature or dislocation l<strong>in</strong>e tension<br />

t film thickness<br />

U energy<br />

ui displacement<br />

w work<br />

Y biaxial modulus<br />

YU unrelaxed biaxial modulus<br />

relaxed biaxial modulus<br />

YR


Appendix F: List <strong>of</strong> Symbols 181<br />

α coefficient <strong>of</strong> thermal expansion or <strong>in</strong>cidence angle for x-ray scatter<strong>in</strong>g<br />

β dislocation cut-<strong>of</strong>f parameter<br />

γ <strong>in</strong>terface energy<br />

γc, γs chemical or structural components to the solid-solid <strong>in</strong>terface energy<br />

�V excess volume<br />

δ misfit or the real part <strong>of</strong> the <strong>in</strong>dex <strong>of</strong> refraction<br />

ɛij stra<strong>in</strong><br />

ɛsh,ɛ�,ɛ⊥ shear stra<strong>in</strong>, <strong>in</strong>-plane stra<strong>in</strong>, <strong>and</strong> out-<strong>of</strong>-plane stra<strong>in</strong><br />

εcoh coherency stra<strong>in</strong><br />

ζ Ronay’s geometric factor<br />

ηijk<br />

Lagrangian stra<strong>in</strong>s<br />

θ diffraction angle/2<br />

κ curvature<br />

� bilayer thickness<br />

λ x-ray wavelength<br />

µ chemical potential<br />

ν Poisson’s ratio<br />

ξ dimensionless parameter for the stress field near a dislocation core<br />

ρ dislocation density<br />

σij stress<br />

σ shear stress, <strong>in</strong>terface roughness, or st<strong>and</strong>ard deviation<br />

σy yield strength<br />

σflow flow strength<br />

σsc substrate curvature stress<br />

volume stress<br />

〈σ 〉v<br />

〈σ 〉x−ray<br />

x-ray measured volume stress<br />

τ characteristic relaxation time or 1/e penetration depth<br />

ϕ1,�,ϕ2 Euler angles<br />

χ sample tilt angle out <strong>of</strong> the diffractometer plane<br />

ψ general sample tilt angle used <strong>in</strong> the s<strong>in</strong> 2 ψ-method<br />

ω stra<strong>in</strong> oscillation frequency<br />

ℓ characteristic length scale for plasticity

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

Saved successfully!

Ooh no, something went wrong!