Mechanical behaviour and deformation mechanisms of Zn-Al-Cu-Mg alloys

Von der Fakultät für Georessourcen und Materialtechnik der Rheinisch-Westfälischen Technischen Hochschule Aachen

zur Erlangung des akademischen Grades eines Doktors der Ingenieurwissenschaften

genehmigte Dissertation

vorgelegt von

M.Sc. Zhicheng Wu

aus Quyuan, Provinz Hunan, China

Berichter: Univ.-Prof. Dr. Sandra Korte-Kerzel Univ.-Prof. Dr.-Ing. Sebastian Münstermann

Tag der mündlichen Prüfung: 05.September 2018

Diese Dissertation ist auf den Internetseiten der Universitätsbibliothek online verfügbar

To my family

Acknowledgements

Acknowledgements

In the beginning, I would like to express my appreciation to all people who have helped me with my doctoral thesis.

Foremost, I would like to express my deepest thanks and sincere appreciation to my advisor Prof. Dr. Sandra Korte-Kerzel, Institute of Physical Metallurgy and Metal Physics (IMM), RWTH Aachen University, for her expert, creative and comprehensive guidance, and encouragement throughout my PhD. My greatest gratitude also goes to Prof. Dr. Günter Gottstein, Institute of Physical Metallurgy and Metal Physics (IMM), RWTH Aachen University, for giving me the opportunity to work at IMM, for suggesting the topic of this dissertation, and for his kind supervision in the first two years of my PhD. My special thanks to my second examiner Prof. Dr. Sebastian Münstermann, Steel Institute (IEHK), RWTH Aachen University, and the chair of my examination committee Prof. Dr. Dieter Georg Senk, Steel Institute (IEHK), RWTH Aachen University, for their insightful comments and discussions.

I would like to thank Dr. Stefanie Sandlöbes and Dr. Weiping Hu for the inspiring discussions, insightful corrections, valuable comments and continuous encouragement. Very special thanks to Dr. Stefanie Sandlöbes, for her unswerving support not only in my research, but also in my personal life.

I would like to thank my colleagues at IMM for the stimulating discussions, for the introductions and support to experimental equipment, and for all the great time we have had during my PhD period. It was fantastic to work together with them at IMM. In particular, I am grateful to David Beckers for the support in metallography, Thomas Burlet for setting up the creep testing and in-situ testing devices, Jann-Erik Brandenburg for the introduction to SEM, Dr. James Gibson for the technical support and inspiring discussions in nanoindentation, Sebastian Schröders for the introduction to FIB and AFM, Fengxin Mao for the help with MATLAB scripts, and Arndt Ziemons for the alloys casting.

I am also grateful to Prof. Dr. Rainer Schmid-Fezter and Dr. Song-mao Liang at TU Clausthal for the cooperation and suggestions for design in our joint projects, Prof. Dr. Christophe Tromas at CNRS-Université de Poitiers for the support with AFM experiments, Xiaoxiao Li at IEHK, RWTH Aachen University for the help with MATLAB scripts, Dr. Konda Gokuldoss Pradeep, Dr. Marcus Hans and Marshal Amalraj at Material Chemistry (MCh), RWTH Aachen University for the support with APT experiments.

I would like to extend my thanks to my student assistants Liang Wu, Jing Rao, Yufengnan Wang, An Zhang, Kuan Ding, Shuo Wang and Shanyu Chen for their great work in their theses and hiwi jobs.

i Acknowledgements

A very special gratitude goes to the German Research Foundation (DFG) for the financial support for my work within projects GO 335/47-1 and KO 4603/5-1. I am also grateful to Federation of German foundry industry e. V. (BDG) for the invitation to industry meet ups on Zn castings, and Dr. Didier Rollez at Grillo-Werke AG for his insightful comments from industry point of view.

Finally, I would like to explain my profound gratitude to my family and my friends, especially my parents, who have supported me along the way.

Thanks for all your encouragement and support!

ii Contents

Contents

Acknowledgements ...... i Contents ...... iii List of tables ...... vi List of figures ...... vii List of symbols...... xi List of abbreviations ...... xii 1. Introduction ...... 1 2. Theoretical background ...... 4 2.1. and Zinc alloy ...... 4 2.1.1. Alloying elements in Zn ...... 4 2.1.2. Application of Zn and Zn based alloys ...... 6 2.2. Deformation mechanisms ...... 7 2.2.1. Dislocation slip ...... 7 2.2.2. Kink bands ...... 10 2.2.3. Deformation twinning ...... 11 2.2.4. Grain and phase boundary sliding ...... 13 2.2.5. Creep ...... 14 2.2.6. Activation of deformation mechanisms ...... 21 2.2.6.1. Thermal activation of dislocations...... 21 2.2.6.2. Activation volume ...... 23 2.2.6.3. Activation energy ...... 24 2.2.7. Deformation mechanism maps ...... 25 2.3. Nanoindentation ...... 30 3. Experimental methods and materials ...... 34 3.1. Materials ...... 34 3.2. Sample preparation ...... 34 3.3. Macroscopic mechanical testing ...... 35 3.3.1. Constant strain rate tensile testing ...... 35 3.3.2. Strain rate jump test ...... 36 3.3.3. Creep test ...... 36 3.3.4. In-situ straining experiments ...... 37 3.4. Nano-indentation experiments ...... 38 3.4.1. Constant strain rate and strain rate jump tests ...... 38 3.4.2. Nanoindentation creep tests ...... 39 3.5. Microstructure characterisation ...... 40

iii Contents

3.6. Digital image correlation (DIC) measurements ...... 40 4. Macroscopic mechanical response and mechanisms ...... 42 4.1. Results and discussion ...... 42 4.1.1. As-cast microstructure ...... 42 4.1.2. Mechanical properties ...... 46 4.1.3. Fracture surfaces ...... 48 4.1.4. Deformation microstructure ...... 51 4.1.4.1. Room temperature and/or highest strain rate tested (5∙10−4 s-1) ...... 51 4.1.4.2. Elevated temperature and/or low strain rate ...... 55 4.2. Conclusions...... 59 5. Creep of ZnAl4Cu1 alloys ...... 60 5.1. Results ...... 60 5.1.1. Uniaxial tensile creep experiments ...... 60 5.1.2. Creep microstructure ...... 65 5.1.3. Creep behaviour of individual microstructural constituents ...... 68 5.2. Discussion ...... 70 5.2.1. Creep mechanisms ...... 70 5.2.1.1. Macroscopic uniaxial tensile creep ...... 70 5.2.1.2. Nanoindentation creep ...... 72 5.2.2. Linking micro- and macroscale deformation: geometrical constraints model....75 5.2.3. Influence of Mg on the creep properties of ZnAl4Cu1Mg alloys ...... 78 5.3. Conclusions...... 80 6. Micromechanical response and mechanisms ...... 82 6.1. Results ...... 82 6.1.1. Local mechanical properties ...... 82 6.1.1.1. Constant strain rate indentation ...... 82 6.1.1.2. Nanoindentation strain rate jump test ...... 83 6.1.2. Deformation microstructure ...... 86 6.1.2.1. Deformation microstructure in primary η-Zn phase ...... 86 6.1.2.2. Deformation microstructure in η-Zn + α-Al eutectoid structures ...... 88 6.1.2.3. Strain partitioning ...... 90 6.2. Discussion ...... 93 6.2.1. Local mechanical properties of individual microstructural constituents ...... 93 6.2.2. Deformation mechanisms and thermal activation ...... 94 6.2.2.1. Deformation of the primary η-Zn phase ...... 94 6.2.2.2. Deformatlion of η-Zn + α-Al eutectoid structures ...... 95

iv Contents

6.2.3. Local strain distribution and strain transfer ...... 96 6.2.4. Bulk deformation ...... 98 6.3. Conclusion ...... 99 7. Precipitation behaviour ...... 101 7.1. Results and discussion ...... 101 7.2. Conclusions...... 108 8. Summarising discussion and concluding remarks ...... 109 8.1. Temperature dependent activity of deformation mechanisms ...... 109 8.2. Micro- and macro-mechanical behaviour ...... 110 8.3. Fracture behaviour ...... 110 8.4. Creep behaviour ...... 111 8.5. Effect of Mg ...... 112 8.6. Precipitates and mechanical behaviour ...... 112 Bibliography ...... 113 Abstract ...... 126 Kurzzusammenfassung ...... 128

v List of tables

List of tables

Table 2.1. Important physical properties of zinc at room temperature [1, 33]...... 4 Table 2.2. Slip systems in zinc [72-74]...... 9 Table 2.3. Twinning elements in hexagonal structures [87]...... 11 Table 2.4: Creep parameters of Zn and Zn alloys reported in the literatures...... 18 Table 2.5. Activation energy of obstacles [83]...... 24 Table 2.6. Activation energies in Zn and Zn based alloys...... 24 Table 2.7. Deformation mechanisms in Zn and Zn alloys reported in literature. The corresponding references are included in the table...... 27 Table 3.1. Chemical composition of the zinc alloys investigated...... 35 Table 3.2. Applied parameters (temperature, stress) during uniaxial creep experiments*. ....37

Table 3.3. Detailed procedure of depositing SiO2 nanoparticles on the sample surface...... 41 Table 4.1. Grain size and phase fractions of investigated ZnAl4Cu1 alloys...... 45

Table 4.2. Mechanical properties of as-cast ZnAl4Cu1 alloys; 휎0.2 yield strength, 휀푓 elongation to fracture...... 48 Table 4.3. The area fraction (%) of ductile fracture under tensile deformation...... 51 Table 5.1. Time (in hours) to 1% creep strain of ZnAl4Cu1MgX (X = 0.04, 0.21, 0.31, in wt.%) alloys at 55°C, 85°C and 105°C and stresses of 0.6 – 0.8∙σ0.2...... 62 Table 5.2. The minimum creep rate 휀̇ (10-4h-1) of ZnAl4Cu1MgX (X = 0.04, 0.21, 0.31, in wt.%) alloys at 55 – 105°C and stresses of 0.6 – 0.8∙σ0.2...... 62 Table 5.3. Calculated shear moduli G of ZnAl4Cu1MgX (X = 0.04, 0.21, 0.31, in wt.%) alloys at different temperatures, [GPa]...... 63 Table 5.4. Creep stress exponents and activation energies of ZnAl4Cu1MgX (X = 0.04, 0.21, 0.31, in wt.%) alloys at 25 – 105°C determined from uniaxial tensile creep tests...... 64 Table 5.5. Nanoindentation creep stress exponents of different individual microstructural components at 25°C and 85°C at maximal loads of 15 – 35 mN...... 70 Table 5.6. Lattice diffusion coefficient of various elements in Zn-Al-Cu-Mg alloys [83]. The diffusion coefficients at 25°C and 85°C were calculated from the data in [83]...... 74 Table 5.7. Boundary diffusion coefficient of various elements in Zn-Al-Cu-Mg alloys [83]. The diffusion coefficients at 25°C and 85°C were calculated from the data in [83]...... 74 Table 6.1. Hardness, H, and elastic modulus, E, of individual microstructural constituents in ZnAl4Cu1Mg0.31 investigated at 25 ºC and 85°C, respectively. The relative values are given as a reduction of the mean value (e.g. 1.2 GPa for uncorrected Zn) with temperature to indicate the temperature dependence...... 83 Table 6.2. Nanoindentation strain rate sensitivity and activation volume of the primary η-Zn phase and η-Zn + α-Al eutectoid structures at RT and 85°C compared to bulk ZnAl4Cu1Mg0.31 at 85°C...... 86 Table 6.3. Average surface heights of Zn-Al phase boundaries in eutectoid structures around nano-indentation imprints. The indentations were performed at a maximum load of 25 mN at 25°C and 85 ºC at a strain rate of 0.1 s-1, and at 85°C at a strain rate of 0.01 s-1. For comparison the average height of Zn-Al phase boundaries in eutectoid structures in un-deformed material is also listed...... 90

vi List of figures

List of figures

Figure 1.1. Graphical abstract...... 3 Figure 2.1. Zn-Al equilibrium phase diagram [25, 38]...... 5 Figure 2.2. Atomic configuration of (a) an edge dislocation and (b) a screw dislocation [59]. . 8 Figure 2.3. Movement of an edge dislocation: the arrows indicate the applied shear stress [62]...... 8 Figure 2.4. Slip systems in the hexagonal structure of zinc: (a) basal slip; (b) prismatic slip; (c) pyramidal slip; and (d) pyramidal slip [69-71]...... 9 Figure 2.5. Schematic explanation of the formation of deformation kink bands in Zn. (a) Model of deformation kink band in Zn single crystal by [76] and [79]. (b) Schematic illustration on the crystal rotation angle in the deformation band [79]. (c) An example of kind bands in the Zn grains of ZnAlCuMg0.31 alloy after tension strain (current work)...... 10 Figure 2.6. Dependence of the twinning shear on the axial ratio in hexagonal metals [3]. ....12 Figure 2.7. Grain boundary sliding between two adjacent grains under tensile stress [93]. ...13 Figure 2.8. Schematic image of typical (a) creep curves and (b) creep rate curves [59]...... 15 Figure 2.9. Creep by diffusional transport through lattice or grain boundary [83]...... 16 Figure 2.10. Schematic illustration of the thermally activated glide process of a dislocation overcoming the Peierls barrier via the double kink mechanism. Reproduced from [61, 140]. 21 Figure 2.11. Deformation mechanism map of pure zinc with a grain size of 100 μm [83]...... 25 Figure 2.12. Deformation mechanism map of normalized grain size versus normalized stress for Zn–22% Al alloys tested at 200°C, reproduced from [160]...... 26 Figure 2.13. Geometries of commonly used indenters: (a) Spherical indenter; (b) Conical indenter; (c) Vickers indenter; (d) Berkovich indenter [165]...... 30 Figure 2.14. (a) Schematic illustration of the indenter and the specimen surface at full load and unloaded for a Berkovich indenter; (b) Schematic load versus displacement curve for elastic- plastic loading followed by elastic unloading. At the maximum load, Pmax, hmax is the depth from the original specimen surface to the tip of the indenter, hc is the contact depth of indenter and material and hs is the distance from the edge of the contact to the specimen surface. hf is the final depth of the residual impression after indentation. Upon elastic reloading, the eventual point of contact with the specimen surface moves through a distance hs [165, 167, 173]. ....31 Figure 3.1. Sketch of the samples for constant strain rate tensile tests...... 35 Figure 3.2. Geometry of specimens for tensile strain rate jump tests, in mm...... 36 Figure 3.3. Geometry of specimens for tensile creep tests, unit in mm...... 37 Figure 3.4. Geometry of specimens for in-situ straining experiments, unit in mm...... 38 Figure 3.5. Secondary electron (SE) micrographs of the region of interest prior to deformation. (a) In-lens SE image of the whole region of interest. (b) In-lens SE image of one individual image, enlarged from the red box in the top left corner in (a). (c) Enlarged in-lens SE micrograph of the region highlighted by the red box in (b), showing SiO2 particle speckle pattern on the sample surface...... 41 Figure 4.1. Microstructures of as-cast ZnAl4Cu1 alloys (BSE); (a) Z410, (b) ZnAl4Cu1Mg0.04, (c) ZnAl4Cu1Mg0.21, (d) ZnAl4Cu1Mg0.31...... 43 Figure 4.2. EDS analysis of alloys (a) ZnAl4Cu1Mg0.04 and (b) ZnAl4Cu1Mg0.21...... 44 Figure 4.3. True stress – true strain curves of as-cast laboratory alloys ZnAl4Cu1Mg0.04, ZnAl4Cu1Mg0.21, ZnAl4Cu1Mg0.31 and as-cast Z410 alloy (tested at lowest/highest rates and temperatures only). The legend given in (a) is valid for all diagrams, while testing conditions (strain rate, temperature) are given in the figures...... 47 Figure 4.4. Fracture surfaces of ZnAl4Cu1 alloys deformed at room temperature; (a) alloy

vii List of figures

ZnAl4Cu1Mg0.04 at strain rate of 5∙10-5 s-1, (b) alloy ZnAl4Cu1Mg0.21 at strain rate of 5∙10-4 s-1, (c) alloy ZnAl4Cu1Mg0.31 at strain rate of 6∙10-6 s-1...... 49 Figure 4.5. Fracture surfaces of alloy ZnAl4Cu1Mg0.04 deformed at elevated temperatures; (a) at 55°C and strain rate of 5∙10-4 s-1, (b) at 55°C and strain rate of 5∙10-5 s-1, (c) at 55°C and strain rate of 6∙10-6 s-1, (d) at 85°C and strain rate of 5∙10-4 s-1, (e) at 85°C and strain rate of -5 -1 -6 -1 5∙10 s , (f) at 85°C and strain rate of 6∙10 s ...... 50 Figure 4.6. Fracture surfaces of alloy ZnAl4Cu1Mg0.31 deformed at elevated temperatures; (a) at 55°C and strain rate of 5∙10-4 s-1, (b) at 55°C and strain rate of 5∙10-5 s-1, (c) at 55°C and strain rate of 6∙10-6 s-1, (d) at 85°C and strain rate of 5∙10-4 s-1, (e) at 85°C and strain rate of -5 -1 -6 -1 5∙10 s , (f) at 85°C and strain rate of 6∙10 s ...... 50 Figure 4.7. Microstructure of ZnAl4Cu1 alloys after deformation at low temperature and/or a strain rate of 5∙10−4 s-1. (a), (d), (g) alloy ZnAl4Cu1Mg0.04; (b), (e), (h) alloy ZnAl4Cu1Mg0.21; (c), (f), (i) alloy ZnAl4Cu1Mg0.31; (a-c) alloys at RT and strain rate of 5∙10-4 s-1; (d-f) alloys at RT and strain rate of 6∙10-6 s-1; (g-i) alloys at 55°C and strain rate of 5∙10-4 s-1; PB: phase boundary...... 52 Figure 4.8. (a), (b) SEM and EBSD images of deformation twins in alloy ZnAl4Cu1Mg0.31 after deformation at RT and strain rate of 6∙10-6 s-1, (c) SEM and EBSD images of cracks formed close to the fracture surface in alloy ZnAl4Cu1Mg0.31 after deformation at RT and strain rate of 6∙10-6 s-1...... 53 Figure 4.9. Microstructure evolution of alloy ZnAl4Cu1Mg0.31 during in-situ straining at RT in the SEM; a) initial microstructure, b) displacement of 34 µm, c) displacement of 36 µm, d) displacement of 62 µm, e) displacement of 63 µm, (f) failure at a displacement of 64 µm; TD: tensile direction...... 54 Figure 4.10. Microstructure of ZnAl4Cu1 alloys after deformation at elevated temperature and/or low strain rate; (a), (d), (g) alloy ZnAl4Cu1Mg0.04; (b), (e), (h) alloy ZnAl4Cu1Mg0.21; (c), (f), (i) alloy ZnAl4Cu1Mg0.31; (a-c) alloys at 55°C and strain rate of 6∙10-6 s-1; (d-f) alloys at 85°C and strain rate of 5∙10-5 s-1; (g-i) alloys at 85°C and strain rate of 6∙10-6 s-1; TD: tensile direction...... 56 Figure 4.11. Microstructure evolution of alloy ZnAl4Cu1Mg0.31 during in-situ straining at 85°C in the SEM; (a) initial microstructure, (b) displacement of 52 µm, (c) displacement of 67 µm, (d) displacement of 110 µm, (e) displacement of 126 µm, (f) displacement of 158 µm, (g) void formation at a displacement of 219 µm, (h) void coalescence at a displacement of 241 µm, (i) failure at a displacement of 259 µm. The red arrows in the enlarged area in the insets show the occurrence of grain boundary sliding in the eutectoid structures. TD: tensile direction. ...57 Figure 5.1. Creep curves of ZnAl4Cu1MgX (X = 0.04, 0.21, 0.31, in wt.%) alloys at 55°C, 85°C and 105°C and stresses of 0.6 – 0.8∙σ0.2...... 61 Figure 5.2. Relation between steady state creep rate 휀̇ vs. normalised stress 휎/퐺 (a, c ,e) and temperature dependence of the steady state creep rate 휀̇ (b, d, f) of ZnAl4Cu1MgX (X = 0.04, 0.21, 0.31, in wt.%) alloys...... 64 Figure 5.3. As-cast microstructure of (a) ZnAl4Cu1Mg0.04; (b) ZnAl4Cu1Mg0.21; and (c) ZnAl4Cu1Mg0.31...... 65 Figure 5.4. Post-deformation microstructures of alloy ZnAl4Cu1Mg0.31 after 3.7 % creep strain at 55°C and 102 MPa (a), and after 15.0% creep strain at 85°C and 93 MPa (b). The red boxes indicate the areas analysed by EBSD (Figure 5.5)...... 66 Figure 5.5. Inverse pole figure (IPF) (a, c, e) and corresponding grain reference orientation deviation (GROD) (b, d, f) maps of primary η-Zn grains in alloy ZnAl4Cu1Mg0.31. (a, b) show the as-cast undeformed material, (c, d) show material creep-deformed to a creep strain of 3.6% at 55°C and 102 MPa, and (e, f) show material creep-deformed to a creep strain of at 15.0%

viii List of figures at 85°C and 93 MPa. The corresponding primary η-Zn grains are marked by dashed rectangles in Figure 5.4 (e, f) are given with a higher magnification due to higher local misorientation angles which are not uniquely revealed at lower magnifications...... 67 Figure 5.6. Micrographs of individual microstructural constituents in ZnAl4Cu1Mg0.31 after indentation with a load of 25 mN at 25°C: (a) primary Zn, (b) η-Zn + α-Al eutectic structure, (c)

η-Zn + α-Al eutectoid structure, (d) Mg2Zn11 embedded in eutectoid structure ...... 68 Figure 5.7. Indentation depth increase (h) as a function of time (t) during the holding period of nanoindentation creep tests at different loads in primary η-Zn phase at (a) 25°C and (b) 85°C...... 68 Figure 5.8. Indentation depth as a function of time during the holding period of the individual microstructural constituents at 25 mN at (a) 25°C and (b) 85°C...... 69 Figure 5.9. Log-log plot of indentation strain rate as a function of the nominal contact pressure, pnom, for nanoindentation creep tests of the different individual microstructural components at 25 mN load at (a) 25°C and (b) 85°C...... 70 Figure 5.10. Typical microstructure of (a) ZnAl4Cu1Mg0.04; (d) ZnAl4Cu1Mg0.21; and (h) ZnAl4Cu1Mg0.31 with highlighted eutectic and eutectoid colony boundaries and corresponding schematic alignment of individual constituents in (b, c) ZnAl4Cu1Mg0.04; (e, f) ZnAl4Cu1Mg0.21; and (i, g) ZnAl4Cu1Mg0.31, where the black blocks represent primary η-Zn grains, blue lines represent interfaces in eutectic lamellar colonies and red lines represent interfaces in eutectoid colonies...... 77 Figure 5.11. Creep strain-time curves of ZnAl4Cu1MgX (X = 0.04, 0.21, 0.31, in wt.%) alloys at the same stress level at (a) 55°C, (b) 85°C and (c) 105°C...... 78 Figure 6.1. Dependence of the true stress on the strain rate of alloy ZnAl4Cu1Mg0.31 obtained by macroscopic tensile strain rate jump tests at 85°C (a). (b) shows how the corresponding stresses upon strain rate jumps from 5∙10-4 s-1 to 6∙10-6 s-1 were determined. The dependence of the hardness of the primary η-Zn phase and η-Zn + α-Al eutectoid structures on the strain rate obtained by indentation strain rate jump tests at RT (c) and 85°C (d)...... 85 Figure 6.2. Post-mortem micrographs of a typical indent in the primary η-Zn phase (RT, 0.1 s-1 strain rate, 25 mN peak load), (a) SE micrograph; (b) IPF map, twin boundaries are highlighted in yellow; (c) AFM micrograph, basal plane traces are highlighted in red; (d) height profile along line 1 in (c); (e) height profile along line 2 in (c) with corresponding AFM micrograph. The dashed white rectangle in (a) shows the region of the AFM map shown in (c)...... 87 Figure 6.3. Post-mortem micrographs of typical indents in the primary η-Zn phase (85°C, 0.1 s-1 strain rate, 25 mN peak load), (a) SE micrograph and (b) IPF map showing deformation twins highlighted by yellow boundaries, the IPF map has been rotated to guide the eyes; (c), (e) SE micrographs and (d), (f) IPF maps showing basal and non-basal slip traces; the insets in (c), (e) are enlarged micrographs of the slip traces...... 88 Figure 6.4. SE micrographs and AFM maps of typical indentation imprints in eutectoid structures and primary Zn phase, the indentations were performed at 25°C at a strain rate of 0.1 s-1 (a)-(c), 85 ºC at a strain rate of 0.1 s-1 (d)-(f), 85°C at a strain rate of 0.01 s-1 (g)-(h) and a maximum load of 25 mN. (a), (d), (g) are SEM images with imposed deformation traces; (b), (e), (h) are the corresponding AFM images with red arrows pointing to the surface traces; (c) and (f) are the corresponding IPF maps...... 89 Figure 6.5. SE micrographs of the region of interest (ROI) in alloy ZnAl4Cu1Mg0.31 at (a) 2% and (b) 5% tensile strain at 85°C; (c) BSE micrograph of the ROI at 0 strain, primary η-Zn grains are numbered by 1-15 (highlighted in red) and η-Zn + α-Al eutectic colonies labelled A- N (highlighted in blue); (d) IPF figure of the ROI at 0 strain...... 91 Figure 6.6. Local von Mises strain map of the region of interest in alloy ZnAl4Cu1Mg0.31 at (a, c, e) 2%, (b, d, f) 5% tensile strain at 85°C. (a, b) Strain maps calculated using the software

ix List of figures

GOM correlate (V8.1, GOM mbH).; (c, d) strain maps calculated using the non-rigid registration method; (e, f) DIC strain maps are overlaid with SE images of the microstructure...... 92 Figure 6.7. Enlarged DIC strain maps show strain transfer between primary η-Zn grains 1, 2 and eutectic colony A (a), and between grains 6, 7 and colonies O, K (b); (c) schematic displays the orientation of slip systems in Zn grain 9 (2nd order pyramidal plane traces observed) and eutectic colony O (basal traces observed) across the interface...... 98 Figure 7.1. True stress – true strain curves of ZnAl4Cu1Mg0.31 alloy at room temperature and at 85°C and at a strain rate of 5∙10−4 s-1(a) and the corresponding fracture surfaces of ZnAl4Cu1Mg0.31 alloy deformed at RT (b) and at 85°C (c) and strain rate of 5∙10-4 s-1...... 102 Figure 7.2. EDS analysis of alloy ZnAl4Cu1Mg0.31...... 103 Figure 7.3. TEM micrographs revealing precipitates in primary η-Zn grains of alloy ZnAl4Cu1Mg0.31 after deformation at RT (a, b) and 85°C (c, d); (a, c) TEM bright field micrographs show lenticular-shaped precipitates; (b, d) TEM dark field micrographs show the needle-shaped precipitates, the micrographs were taken from the spots marked by red circles; (e) diffraction patterns of the needle-shaped precipitates (b, d) and the primary η-Zn grain matrix with zone axis B=[011̅0], B=[0001], B=[112̅0]...... 104 Figure 7.4. 3-D elemental maps of alloy ZnAl4Cu1Mg0.31 deformed at RT showing the 3-D positions of (a, b) Al atoms, (h) Mg atoms and (i) Cu atoms in a primary η-Zn grain. The isoconcentration surfaces of 30 at.% Al (a, b), 0.7 at.% Mg (h) and 18 at.% Cu (i) are also shown. (c-g, j) Plots from different precipitates regions (marked by yellow) give the local chemical compositions...... 105 Figure 7.5. TEM bright field micrographs of alloy ZnAl4Cu1Mg0.31 deformed at RT: (a) area close to fracture surface; (b) enlarged images of the area marked by dashed yellow rectangle in (a), the insets show the diffraction patterns of the area marked by yellow circles...... 106 Figure 7.6. TEM dark field micrographs of alloy ZnAl4Cu1Mg0.31 after deformation at 85°C: (a) high strain sample (area close to fracture surface) with g=(112̅2); (b) low strain sample (area away from fracture surface) with g=(112̅2), (c) low strain sample with g=(0002), (d) low strain sample with g=(011̅0), (e) low strain sample with g=(011̅1); g: diffraction vector. Red arrows mark the dislocations on basal planes and orange arrows mark the dislocations on pyramidal planes...... 107

x List of symbols

List of symbols

1 Twinning shear direction Pmax Maximum indentation load

2 Conjugate twinning direction pnom Nominal contact pressure m' Luster-Morris parameter Q Activation energy R Universal gas constant UP Peierls potential S Stiffness, Plane of shear in twinning, A, ACO, ANH Material constant Ac Projected contact area Shear strain b Burgers vector T Temperature B Zone axis, Indentation pre-exponential T’ Al4Cu3Zn phase term Tm Melting point d Grain size, Lattice spacing V Activation volume * D Grain size V Apparent activation volume

DV Bulk self-diffusion coefficient νi Poisson's ratio of indenter material

D0 Pre-exponential factor for self-diffusion α Al-rich FCC phase

DGB Coefficient of grain boundary α' supersaturated α phase diffusion α1 Power law fitting constant E Young's modulus β Zn-rich FCC phase

Ei Young's modulus of indenter material β’ Zn-rich FCC phase

Er Residual modulus γ̇, ε̇ Strain rate g Geometrical factor δ Thickness of grain boundary G Shear modulus ε Strain, Intermetallic CuZn4 phase h Indentation depth εf Elongation to fracture H Hardness ϵ’ Geomtrical constant hc Contact depth ε’ Strain rate, creep rate hf Final indentation depth η Zn-rich hexagonal phase

HM Meyer hardness η’ supersaturated η phase hmax Depth from the original surface to the ν Poisson ratio

tip of the indenter ρm Mobile dislocation density hs Depth from the edge of the contact to σ, τ Stress the specimen surface τ0 Critical resolved shear stress k Boltzmann constant σ0.2 Yield strength K1 Twinning plane τP Perierls stress K2 Conjugate twinning plane φ Angle between two normal directions of l Dislocation length slip / twining planes m Strain rate sensitivity Ω Atomic volume 1 m Power law fitting constant ωA Frequency n Stress exponent ωD Debye frequency N Indentation stress exponent 휅 Angle between two slip / twinning P Load, Probability directions Ṗ Loading rate

xi List of abbreviations

List of abbreviations

AFM Atomic force microscopy APT Atom probe tomography BSE Backscattered electron microscopy CRSS Critical resolved shear stress DIC Digital image correlation DZM Electromechanical testing machine EBSD Electron backscatter diffraction EDS Energy dispersive X-ray spectroscopy FCC Face centered cubic structure HCP Hexagonal close packed structure IPF Inverse pole figure GBS Grain boundary sliding GROD Grain reference orientation deviation PB Phase boundary PBS Phase boundary sliding RT Room temperature ROI Region of interest SADP Slected area diffraction pattern SE Secondary electron SEM Scanning electron microscopy TD Tensile direction TEM Transmission electron microscopy UFG Ultra-fine grained

xii 1. Introduction

1. Introduction

Zinc is the most widely used metal in industrial production after steel, aluminium and copper and the 24th most abundant element in the earth’s crust [1]. However, Zn in its unalloyed form is relatively soft and brittle at room temperature. This is due to the low symmetry, high mechanical anisotropy of its hexagonal crystal structure and insufficient independent deformation systems for compatible polycrystalline deformation [2, 3]. Alloying with Al and Cu was shown to improve the mechanical properties [4-9] and Zn based alloys show attractive physical and mechanical properties such as high strength and wear resistance, excellent fluidity and castability, good surface quality as well as outstanding corrosion resistance [4, 10- 12]. Therefore, Zn-Al based alloys are widely used in the production of structural and decorative parts for automotive, architectural, electrical and electronic applications as well as machinery and equipment with a complex geometry or equipment needing a high manufacturing precision [9, 12-16]. In 2013 approximately 55,000 tons of zinc die castings products were produced in Germany, and about 453,000 tons were produced worldwide. However, wider application of Zn alloys is hindered by:

i. low creep resistance at moderately elevated temperatures [17, 18], ii. long-term mechanical and dimensional instability at ambient or slightly elevated temperatures [11, 19-21].

Efforts were therefore made to improve the creep resistance and mechanical stability of Zn alloys. To this end, an essential understanding of the mechanical properties and the corresponding plasticity mechanisms, creep mechanisms as well as precipitation behaviour in Zn alloys are necessary. Zn-Al based alloys solidify through a eutectic reaction followed by a eutectoid reaction upon cooling, forming a multi-phase microstructure composed of η-Zn and η-Zn + α-Al eutectic / eutectoid structures, where η-Zn is a Zn-rich hexagonal phase and α-Al is an Al-rich FCC phase. Therefore, to fully understand the mechanical behaviour and deformation mechanisms of Zn-Al alloys, it is also necessary to investigate both, the locally intrinsic properties of individual microstructural constituents as well as the mutual effect of these constituents as an aggregate.

Therefore, the thesis comprises the following four parts:

- Chapter 4 presents the macroscopic mechanical properties of three eutectic ZnAl4Cu1 alloys with different Mg content (0.04, 0.21 and 0.31 wt.% Mg) as well as the corresponding plastic deformation mechanisms and fracture mechanisms.

- Chapter 5 describes the creep properties of the three bulk ZnAl4Cu1 alloys as well as the local creep behaviour of individual microstructural constituents in alloy ZnAl4Cu1Mg0.31. A model bridging the macroscopic and microscopic creep

1 1. Introduction

properties is introduced.

- Chapter 6 focuses on the local mechanical response and corresponding deformation mechanisms of individual microstructural constituents as well as their specific role to the macroscopic deformation in alloy ZnAl4Cu1Mg0.31.

- Chapter 7 illustrates the local precipitation and decomposition behaviour in alloy ZnAl4Cu1Mg0.31 and discusses the influence of these precipitates on the mechanical properties of Zn based alloys.

Therefore, this dissertation bridges the gap between the mechanical properties as well as the plastic deformation mechanisms in ZnAl4Cu1 alloys from both, the macroscopic and microscale point of view.

2 1. Introduction

Figure 1.1. Graphical abstract1.

1 Part of this dissertation appeared as articles. The original citations are: Z. Wu, S. Sandlöbes, L. Wu, W. P. Hu, G. Gottstein and S. Korte-Kerzel (2016). "Mechanical behaviour of Zn-Al-Cu-Mg alloys: Deformation mechanisms of as-cast microstructures." Materials Science and Engineering: A 651: 675- 687., S. Sandlöbes, Z. Wu, K. Pradeep and S. Korte-Kerzel (2016). "Precipitation and decomposition phenomena in a Zn-Al-Cu-Mg alloy." Materials Letters 175: 27-31., Z. Wu, S. Sandlöbes, Y. Wang, J. S. K.-L. Gibson and S. Korte-Kerzel (2018). "Creep behaviour of eutectic Zn-Al-Cu-Mg alloys." Materials Science and Engineering: A 724: 80-94., Z. Wu, S. Sandlöbes, J. Rao, J. S. K.-L. Gibson, B. Berkels and S. Korte-Kerzel (2018). "Local mechanical properties and plasticity mechanisms in a Zn-Al eutectic alloy." Materials & Design 157: 337-350., and Z. Wu, S. Sandlöbes, J. Rao, J. S. K.-L. Gibson, B. Berkels and S. Korte-Kerzel (2018). "Data on measurement of the strain partitioning in a multiphase Zn-Al eutectic alloy." Data in Brief: https://doi.org/10.1016/j.dib.2018.09.010.

3

This page intentionally left blank.

2. Theoretical background

2. Theoretical background

2.1. Zinc and Zinc alloy

Pure zinc is a bluish-white metal which crystallizes in a hexagonal crystal structure [22]. Table 2.1 lists the detailed physical properties of Zn at room temperature. Pure Zn is relatively soft and brittle at room temperature. To improve its mechanical properties, several methods have been applied, such as addition of alloying elements [12, 20, 23-28], severe plastic deformation [11, 19, 29, 30], proper heat treatment [12, 20, 31] and non-metallic compound reinforcement [32].

Table 2.1. Important physical properties of zinc at room temperature [1, 33].

Crystallographic structure Hexagonal

Lattice parameter a (nm) 0.2665

Lattice parameter c (nm) 0.4947

c/a ratio 1.856

Density (g/cm3) 7.134

Melting point (ºC) 419.53

Shear modulus G (GPa) 43.4

Young’s modulus E (GPa) 108.4

2.1.1. Alloying elements in Zn

Alloying with Al, Cu and Mg [12, 20, 23-28] has been proven to be effective to improve the mechanical properties of Zn. Al is a major alloying element to increase the fluidity, strength, wear resistance, corrosion resistance and decrease the grain size of zinc alloys [4, 11, 34- 37]. For example, El-khair et al. [35] have investigated the mechanical properties of Zn-xAl (x=8,12 and 27 wt.%) alloys and reported an increase in strength and a decrease in ductility with increasing Al concentration at room temperature to 100 ºC. Similarly, Yan et al. [37] have observed a hardness increase in Zn-Al alloys with increasing Al content from 27 to 48 wt.%. Türk et al. [36] have also reported the same strengthening effect of Al in Zn alloys with Al content from 5 to 11 wt.%.

According to the Zn-Al phase diagram, Figure 2.1, Zn rich Zn-Al alloys solidify through a eutectic reaction at 382°C upon cooling, forming a dual-phase microstructure composed of Zn-rich hexagonal η phase and Zn-rich FCC β phase. The Zn-rich β phase decomposes through a eutectoid transformation at 275°C, generating η + α eutectoid structures, where α

4 2. Theoretical background is an Al-rich FCC phase.

Figure 2.1. Zn-Al equilibrium phase diagram [25, 38].

It has been reported that rapid cooling of Zn-Al alloys results in super-saturation of Al atoms in η-Zn which causes various phase transformation and decomposition reactions including a number of metastable phases upon aging [1, 21, 39], which has been reported to proceed for months or even years during natural aging [1, 21, 39] and considered to cause a loss of strength and also dimensional changes [40]. Specifically, Zhu and Opitz [41, 42] observed strong volume changes occurring during the decomposition of supersaturated Zn-rich FCC β’ phase into eutectoid Al-rich FCC α and Zn-rich hexagonal η’, while they observed no significant changes in the mechanical properties [1]. The metastable Zn-rich hexagonal η’ phase has been reported to continuously decompose into the thermodynamically stable η and α phases accompanied by a volume contraction of about 0.06% causing a decrease in tensile strength and an increase in elongation [1, 43, 44]. LeHuy et al., Zhu et al. and Mykura et al. [43, 45, 46] observed volume changes during decomposition of supersaturated α' solid solution following α’ → α + η. However, while LeHuy et al. [45] associated this transformation with a volume contraction, Mykura et al. [46] reported a volume expansion upon this decomposition reaction.

Cu is another main alloying element in Zn alloys that has been shown to increase the solubility of Al in Zn and improve the creep resistance, corrosion behaviour, mechanical as well as the tribological properties of Zn-Al based alloys [7, 20, 26, 47, 48]. However, Cu contents higher than 2 wt.% result in the formation of a copper rich metastable ε (CuZn4) phase in the grain boundaries and interdendritic regions, which has been reported to deteriorate the strengthening effect and corrosion resistance of Zn-Al alloys [20, 25, 26]. The addition of ≥ 2 wt.% Cu has also been assumed to be responsible for the mechanical and dimensional instability of Zn alloys due to the precipitation of metastable intermetallic ε (CuZn4) and Al-rich

5 2. Theoretical background ternary T’ phases [6, 13, 31]. LeHuy et al. and da Costa et al. [25, 45] reported that the formation of these Cu-rich intermetallic compounds leads to a decrease of the Cu content in primary η grains, causing a decrease in strength due to reduced solid solution strengthening. Further, it has been reported that decomposition of the metastable ε phase following α + ε →T’ + η causes a volume expansion of up to 4.5% during prolonged aging of Zn-AI-Cu alloys [8, 21, 25, 43, 44, 49].

Mg is an important alloying element in zinc alloys primarily to counteract the intergranular corrosion effects of impurities such as Pb, Sn, Cd, In and Tl [1]. Mg is also effective to improve the hardness and strength of Zn-Al based alloys [23, 25, 50, 51], which has been attributed to a grain refinement effect [23, 52] and formation of Mg rich intermetallic phases such as

MgZn2 and Mg2Zn11 at grain boundaries [51]. In addition, the presence of Mg has been reported to inhibit the η transformation in the Zn-Al alloys [1] and suppress the formation of metastable ε phase precipitates in Zn-Al-Cu ternary alloys [25]. However, Mg contents higher than 0.1% has been reported to reduce the ductility and impact strength in Zn-Al alloys [1, 50]. High Mg contents also lead to hot shortness due to the formation of a ternary Zn-Al-Mg eutectic structure [1].

The presence of other alloying elements such as Zr, Sn and Ti also has beneficial effects on the Zn alloys. For instance, the addition of 0.1 wt.% Zr has been reported to improve the strength of ZnAl4 alloys due to grain refinement [28]. The addition of Sn has been reported to be beneficial to the strength and creep resistance of Zn alloys [53]. It has been reported that the addition of Ti improves the creep resistant of Zn alloys especially in conjunction with Cu [54].The strength of Zn-Al alloys has also been reported to increase by the addition of small amounts of Li [27] and Na [24].

2.1.2. Application of Zn and Zn based alloys

Every year around 13 million tons of Zn are consumed worldwide [55]. End uses of zinc products mainly include galvanizing to protect steel from corrosion, zinc based alloys, brass and bronze (Cu-Zn alloy and Cu-Sn alloy with a small amount of Zn), rolled zinc in architectural applications and production of coins, chemicals such as zinc as an activator oxide for the tire manufacturing industry and zinc sulphate as a micronutrient additive in animal feed and fertilizers [55, 56].

Specifically, the earliest application of zinc dates back to the Romans with the appearance of brass. Depending on the Zn contents, different kinds of brasses can be produced, which together with bronze makes up to 17% of the zinc application. Higher Zn content leads to a lower melting point of brasses [57]. The most commonly used brass is called Muntz metal,

6 2. Theoretical background which contains 40% Zn [57]. Besides, around 50% of zinc finds its application as coating and galvanizing materials due to its unique corrosion resistance properties [55]. Due to the formation of a unique patina, an inert layer of Zn at the surface in ambient atmospheres, further reaction could be avoided [1]. Moreover, the stronger electronegativity of Zn than Fe makes Zn coatings an effective method to protect iron and steel products from corrosion (cathodic protection [1]). Around 7% of Zn products is used as rolled zinc alloys in the forms of sheet, wire, plate and rod [55]. These products have mainly been used in dry cell battery manufacturing process and as roofing, wall cladding and downpipes in the building industry [58].

Due to the combination of several impressive mechanical and physical properties such as unique corrosion resistance properties and excellent castability [4, 10-12, 16], Zn also finds its application in Zn casting alloys, which are widely used for various engineering applications, such as production of both structural and decorative parts for architectural, automotive, electrical and electronic industries as well as general purpose machinery and equipment which need a high manufacturing precision [9, 12-16]. The zinc casting alloys mainly refer to two groups of Zn-Al alloys, namely and ZA alloys [1]. The ZAMAK alloys have been widely used as pressure materials since the 1920s and contain around 4 wt.% Al and different contents of Cu and Mg in order to achieve optimum mechanical properties and castability [1, 12, 14]. ZA alloys are a new group of Zn-Al alloys with high Al contents which have been developed in the 1970s as high-performance gravity casting materials [1, 14]. However, wider application of both group of Zn alloys is restricted due to limited ductility, poor impact toughness and low strength at low temperature [5, 11, 19, 20, 28], low creep resistance at moderately elevated temperatures [17, 18] and long-term mechanical and geometrical instability at ambient or slightly elevated temperatures [11, 19-21]. In order to improve the creep resistance and long-term mechanical stability of Zn based alloys, a deeper and more quantitative knowledge of the creep mechanisms, deformation mechanisms and phase stabilities in Zn based alloys are essential.

2.2. Deformation mechanisms

2.2.1. Dislocation slip

As one-dimensional lattice defects within the crystal structure, dislocations are irregularities or perturbations of the perfect crystal along a line [59]. The spatial structure of dislocations can be visualized by imagining that an extra half-plane is inserted into a crystal, Figure 2.2 [59]. In the vicinity of the dislocations where the half plane ends, the crystal is locally distorted [60].

7 2. Theoretical background

Figure 2.2. Atomic configuration of (a) an edge dislocation and (b) a screw dislocation [59].

The movement of dislocations causes plastic deformation in crystalline solids [59, 61, 62]. Dislocations can move in two approaches, namely glide on the plane along which the dislocation line is displaced, i.e. slip plane [59], or leave the slip plane by absorption or emission of point defects, i.e. climb [59]. Dislocation slip results in glide steps or slip bands on the surfaces of crystals [63], it doesn’t change the volume of the crystalline, while dislocation climb results in a (very small) volume change of the crystal due to generation or emission of point defects [59, 61].

Due to the existence of dislocations, only small displacements of the adjacent atoms along dislocations rather than gliding of the whole slip plane is required during plastic deformation, Figure 2.3 [62]. In this case, the shape of the crystal is changed through the glide of dislocations, while the crystal structure remains unaffected according to Masing [64].

Figure 2.3. Movement of an edge dislocation: the arrows indicate the applied shear stress [62].

During movement, dislocations are also subjected to periodical lattice resistance, arising from the changes in energy from the atomic rearrangement and misfit of the atoms across the slip plane as dislocations move [61, 65]. The shear stress needed to overcome the periodical lattice resistance and move a single dislocation within a plane of atoms in a perfect lattice,

8 2. Theoretical background namely Peierls stress, or Peierls–Nabarro (P-N) stress, can be roughly estimated as [59, 61, 66-68]:

2퐺 2휋 푑 휏 = ∙ 푒푥푝⁡(− ∙ ) (2.1) 푃 1 − 휈 1 − 휈 푏 where 휏푃 is the Peierls stress, G the shear modulus, 휈 the Poisson ratio, d the lattice spacing and b the Burgers vector. The lowest shear resistance is therefore expected on slip systems where the ratio d/b is greatest, which generally corresponds to the most densely packed plane and closest-packed direction within the slip plane [59]. Figure 2.4 [69-71] and Table 2.2 [72- 74] list all possible slip systems in zinc. In Zn with an axis ratio c/a>1.63, the most densely packed planes and directions are the (0001) basal plane and the 〈12̅10〉 directions [59]. The basal slip system in Zn has the largest d/b ratio and therefore lowest Peierls stress 휏푃, and is conseuently easiest to be activated at room temperature.

Figure 2.4. Slip systems in the hexagonal structure of zinc: (a) basal slip; (b) prismatic slip; (c) pyramidal slip; and (d) pyramidal slip [69-71].

Table 2.2. Slip systems in zinc [72-74]. Number of Slip systems Slip planes Slip directions d/b independent slip systems

Basal (0001) 〈12̅10〉 0.93 2

Prismatic {101̅0} 〈12̅10〉 0.87 2

Pyramidal {101̅1} 〈12̅10〉 0.79 4

Pyramidal {112̅2} 〈1123̅̅̅̅〉 0.21 5

However, in addition to the intrinsic lattice resistance, dislocation motion is also subjected to other kinds of local energy barriers such as small obstacles as well as the long-range elastic interactions of mobile dislocations with other defects and microstructures such as forest

9 2. Theoretical background dislocations, grain boundaries, precipitates and dispersoids [59, 75]. The critical resolved shear stress (CRSS), 휏0, [59], is therefore comprised of both, the stress to overcome long- range elastic interactions of mobile dislocations with the microstructure as well as the stress necessary to overcome locally short-range energy barriers such as the elastic strain fields induced by point defects [75].

2.2.2. Kink bands

A kind band is a special form of a deformation band [76]. It usually occurs in strongly anisotropic materials in which one slip system is predominately operative and the c/a is large [77]. Kink bands were frequently observed in Zn and their formation was solely attributed to basal slip [76-79].

Figure 2.5. Schematic explanation of the formation of deformation kink bands in Zn. (a) Model of deformation kink band in Zn single crystal by [76] and [79]. (b) Schematic illustration on the crystal rotation angle in the deformation band [79]. (c) An example of kind bands in the Zn grains of ZnAlCuMg0.31 alloy after tension strain (current work).

Hess and Barrett [76] have proposed a model for the formation of kink bands in single crystal Zn, Figure 2.5, where basal dislocations glide synchronously on several parallel slip

10 2. Theoretical background planes and are arranged perpendicular to the slip planes, forming thereby the deformation kink band boundaries [76, 80]. Further, Hagihara et al. [79] have reported that the generation of kink bands is significantly affected by the loading direction during compressive tests in Zn single crystals. This kind of loading direction dependence of kink band formation has also been observed by Chen et al. in long period stacking ordered (LPSO) structure in an extruded

Mg97Y2Zn1 (at.%) alloy [81]. However, with increasing temperature, the loading orientation dependence of kink bands formation has been reported to be less important [79].

2.2.3. Deformation twinning

Deformation twinning is a simple shear deformation in which a portion of the crystal transforms into an orientation with mirror symmetry relative to the untwined lattice [59, 82]. It involves the motion of partial dislocations [83]. Twinning is an important deformation mode in hexagonal materials as slip alone does not offer a sufficient number of deformation systems for an arbitrary shape change at low temperatures [2, 3, 84]. High strain rates, low stacking-fault energy and low temperatures are found to facilitate deformation twinning [84-86], while in the plastic deformation of hexagonal metals twinning has also been reported to play a significant role at almost all temperatures [2].

The geometry and crystallography of mechanical twinning is characterised by the plane of symmetry, namely the twinning plane, K1, and the twinning shear direction, 1. Table 2.3 [87] lists the four most commonly observed twin systems in hexagonal materials, where K2 is the conjugate twinning plane (undistorted rotated plane), 2 is the conjugate twinning direction, and S is the plane of shear perpendicular to K1 and K2.

Table 2.3. Twinning elements in hexagonal structures [87].

K1 1 K2 2 S

{101̅2} <101̅1̅> {1̅012} <101̅1> {12̅10}

{101̅1} <101̅2̅> {101̅3̅} <303̅2> {12̅10}

{112̅2} <112̅-3> {112̅4̅} <224̅3> {11̅00}

{112̅1} <1̅1̅26> {0002} <112̅0> {11̅00}

The twinning shear in hexagonal metals is depending on the axial ratio [59]. Figure 2.6 [3] plots the variation of twinning shear for different twin systems with the axial ratio in hexagonal structures. Twinning systems showing a positive slope induce a compressive strain along the crystal c-axis, called compression twins. Equivalent, a negative slope induces a tensile strain

11 2. Theoretical background along the c-axis, these twins are called tension twins. In Zn, due to its high axial ratio of 1.856 (˃√3), 101̅2<101̅1̅> twin have the smallest twinning shear and are easily activated when loaded in compression parallel to the c-axis or when tension is applied perpendicular to the c-axis [3, 87]. Parisot et al. [88] have reported a critical resolved shear stress (CRSS) of 5 MPa for this twinning system in Zn at RT, which is much larger than that for basal slip and slightly lower than that for non-basal slip at RT. The {101̅2} 〈1011̅̅̅̅〉 twin in Zn induces a compressive strain along the c-axis of the crystal and is therefore a compression twins in Zn.

Figure 2.6. Dependence of the twinning shear on the axial ratio in hexagonal metals [3].

The amount of deformation that can be produced by twinning is small [89]. In hexagonal metals, the amount of shear strain, S, induced by one {101̅2} 〈1011̅̅̅̅〉 twin is [61]:

푐 2 √3 푆 = [( ) − 3] ∙ 푐⁡ (2.2) 푎 3 ∙ 푎

This gives a twinning shear strain of 0.138 in Zn. Because of the small shear, twinning is easy to occur in Zn since only few amount of atomic shuffling is necessary. The important role of mechanical twinning in plastic deformation therefore is not related merely to the strain produced by the twinning process but from the orientation changes induced by twinning. Twinning may place new slip systems in a favourable orientation with respect to the stress axis so that additional slip can take place [71].

In Zn, deformation twinning has been reported to play a significant role in the plastic deformation of Zn in a wide temperature range [2]. For instance, Dirras et al. [90] and Li et al. [91] have reported that {101̅2} 〈1011̅̅̅̅〉 twins are one of the predominant deformation mechanisms in polycrystalline zinc during compression test at room temperature. The {101̅2}

12 2. Theoretical background

〈1011̅̅̅̅〉 twinning has also been reported by Hughes et al. [92] in polycrystalline zinc during tensile test at -196.5°C to 150°C. On the other hand, Yoo et al. [3] and Partridge et al. [87] have reported that the frequency and importance of twinning decreases with increasing deformation temperature. Dirras et al. [90] have also concluded that the possibility of deformation twinning decreases with decreasing grain size and increasing strain rate.

2.2.4. Grain and phase boundary sliding

Grain boundary sliding (GBS) and phase boundary sliding (PBS) refer to the relative displacement of two neighbouring grains or phases against each other under an external loading stress and are thermally activated processes [93]. However, both processes are often referred to as GBS without differentiation if the gliding interface is formed by grains of the same crystalline phase or different crystalline phases. GBS is usually observed in small nano- sized grains and can accommodate a large amount of plastic deformation by inducing viscous sliding of grains without the formation of cracks or cavities [94]. Figure 2.7 shows a schematic example of grain boundary sliding, where a three dimensional displacement vector is formed [93].

Figure 2.7. Grain boundary sliding between two adjacent grains under tensile stress [93].

GBS was first reported by Pearson [95] in extruded lead-tin and bismuth-tin eutectic alloys after superplastic deformation. It has been reported that GBS occurs through the motion of intragranular dislocations, and to be controlled by grain boundary diffusion and accommodated by grain matrix deformation [96]. Therefore, the threshold stress for GBS is strongly temperature dependent [97]. The GBS has been reported to be greatly affected by the grain size [98], impurities [96, 99, 100] and type of grain boundaries [101]. It has been reported that GBS occurs easier in microstructures which contain a high fraction of high angle grain boundaries [101, 102].

13 2. Theoretical background

Due to the relatively low melting point of Zn, grain boundary sliding (GBS) is readily activated in Zn alloys at ambient temperature. Grain / phase boundary sliding has been frequently reported in ultra-fine grained Zn and Zn-Al alloys [2, 83, 94, 103, 104] or eutectoid Zn-22Al alloy with α-Al and η-Zn binary microstructure [6, 96, 97, 105-118].

In Zn and Zn-Al alloys with dilute Al contents, the sliding of Zn-Zn boundaries has been reported [2, 30, 83, 94, 103, 104, 119, 120]. For instance, GBS has been reported to be active in Zn with average grain size of around 0.24 µm during tensile deformation between 20°C to 60°C [94]. Ha et al. [119] and Demirtas et al. [30] have observed GBS in a Zn-0.3Al alloy after deformation at RT. However, it has been reported that the microstructure of low Al containing Zn alloys is not very stable and significant grain growth might occur at high temperatures [119]. GBS in these kind of Zn-Al alloys usually occurs at RT and might be deteriorated with increasing temperature due to grain growth [119, 121].

In Zn alloys with higher Al contents such as the eutectoid Zn-22Al alloy, grain / phase boundary sliding of Zn-Zn, Zn-Al and Al-Al boundaries have been reported [6, 96, 97, 105- 118]. Uesuqi et al. [96] have reported superplasticity in an eutectoid Zn-22Al alloy with an average grain size of 0.60 µm during tensile deformation at RT. Huang et al. [113] have also observed grain / phase boundary sliding in the same alloy with a grain size of 1.3 µm at RT. Tanaka et al. [122] have reported GBS in a Zn-22Al alloy with an average grain size of 1.3 µm after tensile tests in the temperature range of 20 - 200°C. Furthermore, Kawasaki et al. [101] have quantitatively measured the contribution of GBS of different types of boundaries by drawing marker lines on the surface of Zn-22Al alloy specimens during tension at 200 ºC. They have concluded that among the Zn-Zn, Zn-Al and Al-Al interfaces in Zn-22Al alloys the Zn-Zn interfaces contribute most to the total GBS due to the high homologous temperature of Zn at RT, whereas the Al-Al interfaces contribute the least to the total GBS [101, 123].

2.2.5. Creep

At elevated temperatures, creep deformation, which is the time dependent plastic deformation of materials when subjected to load [59], can occur. Figure 2.8 [59] shows a typical creep curve ε(t), namely the strain vs. time curve, and the corresponding creep rate curves 휀̇(t) at constant stress. Both curves are comprised of three stages. In the first stage, or primary creep stage, the creep strain increases rapidly while the strain rate decreases upon loading. At the end of the primary creep stage, the creep rate gradually reaches a minimum and becomes nearly constant, indicative of the second or stationary creep stage, where work hardening and thermal softening balances. The secondary creep stage dominates most of the elapsed time of the creep life [59]. In the range of tertiary creep, stage III, the creep rate increases again rapidly due to necking until creep fracture occurs.

14 2. Theoretical background

Figure 2.8. Schematic image of typical (a) creep curves and (b) creep rate curves [59].

The stationary creep rate, ε̇, strongly depends on the deformation conditions such as applied stress, temperature and material state such as the grain size. The stationary creep rate can be phenomenologically approached by the Mukherjee-Bird-Dorn equation [124, 125] through an Arrhenius-type equation:

퐺푏 푄 휎 푛 휀̇ = ⁡퐴 퐷 푒푥푝(− ) ( ) (2.3) 푘푇 0 푅푇 퐺 where A is a material constant, G the shear modulus, b the Burgers vector, T the temperature in Kelvin, k the Boltzmann constant, D0 the pre-exponential factor for self-diffusion, Q the creep activation energy, R the universal gas constant, σ the applied stress and n the stress exponent.

Depending on the testing conditions (temperature, stress) and material state, different creep mechanisms can be activated. Dislocation creep is a mechanism where edge dislocations overcome obstacles on their slip planes with the assistance of vacancy diffusion [59]. Since the vacancies are generated by self-diffusion, the activation energy for dislocation creep is found to be of the same magnitude as the activation energy for self-diffusion [59]. In this case, the stresses applied on the specimens are usually high. In a simplified approach, the stationary creep rate can be described as [53, 59, 126, 127]:

휎 3−7 푄 휀̇ ⁡ = ⁡퐴 ( ) 푒푥푝(− )⁡ (2.4) 퐺 푅푇 and yields a stress exponent value of 3 – 7 for dislocation creep [53, 59, 126, 127]. Due to this power law relationship the dislocation creep is also called power law creep [59].

At high temperatures, the vacancy transportation itself leads to creep deformation even if there is no dislocation movement. As Figure 2.9 [83] shows, under an external stress vacancies emerge at the grain boundaries normal to the stress, while those parallel to the stress develop a lower concentration of vacancies [128]. This concentration gradient induces a diffusive flux of vacancies towards the parallel grain boundaries and therefore the migration of atoms

15 2. Theoretical background through and around the boundaries of the grains to the perpendicular grain boundaries, leading to creep strain [59, 83, 128].

Figure 2.9. Creep by diffusional transport through lattice or grain boundary [83]. At high temperature and low stress, the resulting flow is controlled by lattice diffusion, this kind of diffusion creep is known as Nabarro-Herring creep [59, 83]. The steady creep rate during this diffusional creep is related to the applied stress, 휎, according to [59, 128]:

휎훺 퐷푉(푇) 휀̇ = 퐴푁퐻 ⁡ (2.5) 푘푇 푑2

3 Here, ⁡퐴푁퐻 is a material constant,⁡⁡훺⁡≈ b is the atomic volume, 퐷푉 is the bulk self-diffusion coefficient and d is the grain size. In this case the stress σ owns a linear relationship with the strain rate (n = 1).

At intermediate temperatures and in fine-grained materials, grain boundary diffusion dominates diffusional creep and the majority of diffusive flux takes place along the grain boundaries instead of through the lattice [128]. This kind of diffusional creep refers to Coble creep [59, 83]. Similar to Nabarro-Herring creep, the steady state flux of vacancies and the resulting steady state creep rate for Coble creep can be described as [59, 128]:

휎훺 훿퐷 (푇) 휀̇ = 퐴 퐺퐵 (2.6) 퐶푂 푘푇 푑3

Where 퐴퐶푂 is a material constant, DGB is the coefficient of grain boundary diffusion, 훿 is the thickness of a grain boundary.

Creep deformation is an important deformation mechanism in Zn and Zn-Al based alloys, as they are sensitive to creep already at room temperature due to their relatively low melting point (around 400°C). In Zn, dislocation creep has been generally reported to be the predominant mechanism during creep deformation. Specifically, the creep of Zn at temperatures up to 142°C has been reported to be controlled by the climb of edge dislocations [129-131], as the creep activation energy in this region (79 – 108 kJ/mol [129-131]) is close to the self-diffusion

16 2. Theoretical background energy of Zn (92 kJ/mol [83, 132]). At temperatures above 280°C, creep has been suggested to be dominated by prismatic dislocation glide [83, 133] due to a creep activation energy, Q, of 152 – 168 kJ/mol [130, 131, 133] which is similar to the activation energy of prismatic slip (152 kJ/mol [83]). Although grain boundary sliding (GBS) has been reported during creep at temperatures between 80°C and 200°C in Zn [126], GBS has been proposed not to be the rate-controlling creep mechanism [126]. At 300 K and a load of 19 MPa, Matsunaga et al. [134] have determined that GBS accommodates about 30% of the total creep strain in polycrystalline Zn with an average grain size of 100 µm.

In Zn-Al alloys, various creep mechanisms have been reported. Dislocation controlled creep has been reported to be the predominant creep mechanism in Zn-Al alloys at stress levels of 10 – 100 MPa and ambient temperatures [18, 40, 127, 135]. For instance, Murphy et al. [18] have investigated the creep properties of three pressure die-cast commercial zinc-based alloys: No.3 (4 wt.% Al, 0.05 wt.% Mg), ZA8 (8 wt.% Al, 1 wt.% Cu, 0.03 wt.% Mg) and ZA27 (27 wt.% Al, 2 wt.% Cu, 0.02 wt.% Mg) at stresses between 10 to 100 MPa and temperatures between 60°C and 150°C. They have obtained a stress exponent n = 3.5 and a creep activation energy Q = 106 kJ/mol for most of the tested alloys [18]. Roberti et al. [127] have investigated the creep behaviour of a commercial Z410 alloy (ZnAl4Cu1) using creep tests at 80°C under a constant load of 20 and 30 MPa, respectively, on samples with an average grain size of around 10 μm. The obtained stress exponent (n = 5) has been associated with dislocation creep mechanisms [127]. The same material has been studied by Kallien et al. [40] at RT with stresses between 40 MPa and 100 MPa and at 85°C with stresses between 12 MPa and 50 MPa, where a stress exponent n of 4.1 and an activation energy of 94 kJ/mol, which are indicative of dislocation creep, have been obtained.

Further, diffusional creep has been reported in Zn alloys at relatively low stresses and high temperatures [136, 137]. Specifically, Coble creep has been reported to be the predominant creep mechanism in an ultra-fine grained Zn-4.5Al alloy with an average grain size of 260 nm at temperatures between 22°C and 100°C [136], where a stress exponent close to 1 and activation energies close to that of grain boundary diffusion were obtained (60.5 kJ/mol [83], summarised in Table 2.6). Similarly, Prasad et al. [137] have reported Coble creep in a Zn- 22Al alloy with an average grain size around 1 µm at temperatures between 120 – 200°C and stresses lower than 1 MPa. Additionally, Nabarro-Herring creep has also been reported in a Zn-22Al (wt.%) alloy with larger grain size (1.3 – 3.7 µm) at lower stresses (0.1 – 0.3 MPa) and temperatures between 177°C and 252°C [138].

A comprehensive summary of creep parameters for zinc and zinc alloys reported in the literature are given in Table 2.4. Generally, zinc and zinc alloys possess different and distinct creep mechanisms which are sensitively determined by the temperature, stress state and material state such as composition and grain size.

17 2. Theoretical background

Table 2.4: Creep parameters of Zn and Zn alloys reported in the literatures2.

Grain Temperature Stress Stress Activation energy Proposed Material (in wt.%) Size Reference (°C) (MPa) exponent n Q (kJ/mol) mechanism (µm) Gilman 1956 single crystal zinc - 255 – 407 3.92 – 39.2 - 167.88 - [133] Tegart 1958 polycrystalline 100 70 – 150 1.4 – 9.7 4.7 87.9 – 97.1 dislocation creep [131], Flinn 1964 zinc [130] Tegart 1958 polycrystalline 100 280 – 400 1.4 – 9.7 4 151.9 – 159 prismatic climb [131], Flinn 1964 zinc [130] polycrystalline Murthy 1982 210 80 – 200 5 – 200 4.4 163.2 dislocation creep zinc [126] polycrystalline 200 330 – 360 30 – 60 - 211 – 252 dislocation creep Jenei 2014 [104] zinc grain boundary polycrystalline Matsunaga 2010 100 ambient 19 3 20 sliding contributes to zinc [134] 30% of creep strain polycrystalline grain boundary Matsunaga 2009 100 -70 – 70 24 – 36 3.4 12 zinc sliding [129] polycrystalline Matsunaga 2009 100 70 – 200 24 – 36 4.8 79 dislocation creep zinc [129]

2 This table was published in an article. The original citation is: Z. Wu, S. Sandlöbes, Y. Wang, J. S. K.-L. Gibson and S. Korte-Kerzel (2018). "Creep behaviour of eutectic Zn-Al-Cu-Mg alloys." Materials Science and Engineering: A 724:80-94.

18 2. Theoretical background

Grain Temperature Stress Stress Activation energy Proposed Material (in wt.%) Size Reference (°C) (MPa) exponent n Q (kJ/mol) mechanism (µm)

polycrystalline Godavarti 1987 - 150 – 225 20 – 50 3.4 – 4.4 82 ± 5 dislocation climb zinc [139] polycrystalline Godavarti 1987 - 125 – 150 20 – 50 3.4 – 4.4 58 ± 8 core diffusion zinc [139]

Zn-4Al - 60 – 150 10 – 100 3.5 106 - Murphy 1988 [18]

Zn-4Al-0.1Cu - 25 – 85 40 – 100 4.6 94.1 dislocation creep Kallien 2010 [40]

Roberti 2008 Zn-4Al-1Cu 10 80 20, 30 5 - dislocation creep [127]

Zn-4Al-1Cu - 25 – 85 40 – 100 4.15 94.1 dislocation creep Kallien 2010 [40]

Zn-4Al-3Cu - 25 – 85 40 – 100 4.5 94.1 dislocation creep Kallien 2010 [40]

Gobien 2010 Zn-4.5Al 0.26 22 – 100 25 – 130 1 69 coble creep [136] Zn-3Cu-xAl (x=4, Alibabaie 2012 20 72 – 132 70 – 250 7.3 – 7.9 52.5 – 56.1 pipe diffusion 5, 6) [53] Zn-3Cu-xAl (x=4, Alibabaie 2012 20 147 – 222 70 – 800 5.0 – 5.9 85.6 – 100.3 dislocation creep 5, 6) [53]

Zn-8Al-1Cu - 70 – 160 20 – 100 4.74 112 dislocation creep Anwar 2000 [17]

Zn-8Al-1Cu - 60 – 150 10 – 100 3.5 106 - Murphy 1988 [18]

Zn-12Al-1Cu - 70 – 160 20 – 100 4.71 109 dislocation creep Anwar 2000 [17]

19 2. Theoretical background

Grain Temperature Stress Stress Activation energy Proposed Material (in wt.%) Size Reference (°C) (MPa) exponent n Q (kJ/mol) mechanism (µm)

Zn-22Al 1.3 – 3.7 177 – 252 0.7 – 40 2.2 69.6 ± 2.1 Coble creep Arieli 1980 [138]

Zn-22Al 1.3 – 3.7 177 – 252 0.1 – 0.3 1 95.9 ± 2.1 Nabarro-Herring Arieli 1980 [138]

Prasad 1993 Zn-22Al 0.9 – 2.0 120 – 200 0.49 – 1 1 62 coble creep [137]

Zn-27Al-2Cu - 70 – 160 20 – 100 3.85 102 dislocation creep Anwar 2000 [17]

Zn-27Al-2Cu - 60 – 150 10 – 100 4.44 106 - Murphy 1988 [18]

20 2. Theoretical background

2.2.6. Activation of deformation mechanisms

2.2.6.1. Thermal activation of dislocations

At 0K, the CRSS of dislocation glide in a defect-free single crystal is equal to the Peierls stress,

휏푃 [61]. In this case, the Peierls potential, 푈푃, is the only obstacle to dislocation motion and the dislocations can move when an external stress larger than 휏푃 is applied. Figure 2.10 [61, 140] schematically describes the relationship between the position of a dislocation on its glide plane and the corresponding Peierls barrier,⁡푈푃.

At finite temperatures, under the action of thermal fluctuations, dislocations can move in such a way that first one part of the dislocation with length l overcomes the Peierls “hill”, generating two kinks, instead of overpassing the Peierls potential as a rigid straight line [59, 140]. Under an external stress, 휏, the kink pair can expand in opposite directions, enabling the whole dislocation line to propagate into the next Peierls valley [59, 140].

Figure 2.10. Schematic illustration of the thermally activated glide process of a dislocation overcoming the Peierls barrier via the double kink mechanism. Reproduced from [61, 140].

If 휔퐴 is the frequency of a dislocation segment attempting to overcome the obstacle, the probability, P, for this segment to surmount the energy barrier is [65, 75]:

훥퐺(휏) 푃 = 휔 푒푥푝⁡(− )⁡ (2.7) 퐴 푘푇 where k is the Boltzmann constant, T the absolute temperature,⁡훥퐺(휏) the energy necessary to initiate dislocation glide. 휔퐴 can be related with Debye frequency, ω퐷, the Burgers vector of the dislocation, b, and the length of dislocation segment, l, according to [75]:

휔퐴 = 푏휔퐷/푙⁡ (2.8)

The distance that the dislocation moves after one activation event corresponds to b [65]. The average velocity of the dislocation segment, 푣, can be therefore estimated by [65]:

21 2. Theoretical background

훥퐺(휏) 푣 = 푏푃 = 휔 푏 ∙ 푒푥푝⁡(− )⁡⁡⁡⁡ (2.9) 퐴 푘푇

If an external stress 휏 < 휏푃 is applied, the thermal activation barrier is reduced to the stress differential:

훥퐺(휏) = 훥퐺 − 휏푉 (2.10) where 훥퐺 is the total activation energy required to overcome the obstacle at 0 stress. V is the activation volume which is defined as [75]:

휕퐺 푉 = − ( ) (2.11) 휕휏 푇

The activation volume describes the critical atomic volume necessary to overcome the energy barrier during a thermal activation event (see cf. 2.2.6.2).

Consequently, the strain rate,⁡훾̇ , given by the Orowan equation [141]:

훾̇ = 휌푚푏푣 (2.12) can be expressed as:

훥퐺 − 휏푉 훾̇ = 휌 푏휔 푏 ∙ 푒푥푝⁡(− )⁡ (2.13) 푚 퐴 푘푇 where 휌푚⁡is the mobile dislocation density in the material.

By rearranging this equation the critical resolved shear stress at temperature T and a strain rate ⁡훾̇ can be obtained [65]:

훥퐺 푘푇 훾̇ 휏 = + 푙푛 2 ⁡ (2.14) 푉 푉 휌푚푏 휔퐴

For the kink pair mechanism V exhibits small values and the flow stress changes significantly with temperature [75]. If the stress to overcome a forest dislocation governs flow, i.e. forest mechanism, V shows significantly larger values and the flow stress is hardly affected by thermal activation [75].

In Zn, the CRSS values of the basal slip system has been reported to be much smaller than that of non-basal slip systems at ambient temperature [88, 142-144], arising from the lowest Peierls stress, 휏푃 , of basal slip (largest d/b ratio, as shown in Table 2.2).

Specifically, Bell and Cahn [142] have reported a 휏0⁡value of 0.3 MPa for basal slip in single crystal Zn and an up to 30 times higher⁡휏0 value for pyramidal slip at RT (4 – 10 MPa). Fundenberger et al. [143] have reported that the CRSS for prismatic slip is around 15 times larger than that for basal slip.

On the other hand, the CRSS for slip of dislocations on the basal plane in Zn has been reported to be only slightly temperature-dependent [145], while that for and

22 2. Theoretical background pyramidal slip decreases abruptly with increasing temperature [90, 104, 145]. Specifically, the flow stress for basal glide in hexagonal crystals has been reported to be determined by the long-range stresses of dislocation intersections [144, 146] and is therefore less thermally activated. On the other hand, non-basal slip such as the pyramidal dislocation slip in Zn has been reported to be controlled by overcoming the Peierls barrier due to its small activation energy (0.1 – 0.2 eV) and small activation volume (2 – 30 b3) [147]. Non-basal slip could be therefore easily activated at elevated temperatures. In fact, while Dirras et al. [90] have reported that basal slip is one of the dominant plastic deformation mechanisms of zinc at RT, the activation of non-basal slip systems in Zn have been frequently reported at elevated temperatures [90, 104, 148]. Specifically, Dirras et al. [90] have also found that increasing the temperature to around 150ºC leads to the activation of more type dislocations on pyramidal planes in polycrystalline Zn. Jenei et al. [104] have also observed increased activity of pyramidal dislocations during indentation creep tests in ultra-fine grained (UFG) zinc at 300 – 360ºC. Due to the thermal activation of pyramidal slip systems [2, 59, 90, 149], the ductility of zinc is greatly increased at elevated temperatures.

2.2.6.2. Activation volume

The experimental determination of V is of importance as it is directly related to the thermally activated glide of dislocations and can give information about the obstacles for dislocation motion [75, 150]. Experimentally, the value of V can be measured using strain rate jump experiments, stress relaxation tests or creep transient tests where the structure remains constant [75]. The apparent activation volume, V*, at a constant deformation temperature can be calculated as [150]:

휕푙푛훾̇ 푉∗ = 푘푇 ( ) (2.15) 휕휏 푇 where k is the Boltzmann constant, T is the absolute temperature, 훾̇ is the strain rate, 휏 is the shear stress. The obtained apparent activation volume V* is close to the physical activation volume, V, Eq. (2.11)[151].

Alternatively, the strain rate sensitivity, m, which reflects the dependency of the flow stress on the strain rate, can be also obtained by the same experiments [59, 152]:

휕푙푛휏 푚 = ( ) (2.16) 휕푙푛훾̇ 푇 m is also known as the reciprocal of the stress exponent n, Eq. (2.3), used in conventional creep tests [153], see section 2.2.5. The strain rate sensitivity, m, and activation volume, V*, are important mechanical parameters to evaluate the predominant thermally activated plastic deformation mechanisms of materials [150, 154, 155]. Typical values of V* are in the range of ~100b3 to ~1000b3 for dislocation glide in face-centered-cubic metals [156], ~10b3 for GBS

23 2. Theoretical background

[157, 158] and ~b3 for grain boundary or lattice diffusion processes [156]. Generally, plasticity mechanisms with higher V* values require higher stresses to be activated [158] and the stress is less sensitive to temperature changes [75], while plasticity mechanisms with a small V* correspond to a rapid decrease of stress with temperature [75]. In Zn-Al alloys, a strain rate sensitivity above 0.25 has been reported to indicate a grain boundary dominated deformation mechanism [101, 122, 159].

2.2.6.3. Activation energy

The activation energy, 훥퐺, which is defined as the total activation energy required to overcome obstacles thermally without the aid of external stress, characterises the strength of a single obstacle [65, 83]. Hence, the activation energy is directly related to thermally activated mechanisms. Generally, thermal activation is dominant for deformation mechanisms with a small activation energy, 훥퐺, see section 2.2.6.1. Frost and Ashby [83] have summarised the values of activation energies for different kinds of obstacles, Table 2.5.

Table 2.5. Activation energy of obstacles [83].

Obstacle strength Activation energy, 훥퐺 Example

Strong 2 Gb3 Dispersions; large or strong precipitates

Forest dislocations, radiation damage; Medium 0.2 – 1.0 Gb3 small or weak precipitates

Weak < 0.2 Gb3 Lattice resistance; solution hardening

Further, Table 2.6 summarises the reported activation energies for different kinds of deformation mechanisms in Zn and Zn alloys in literature.

Table 2.6. Activation energies in Zn and Zn based alloys.

Deformation mechanism Activation energy, 훥퐺 (kJ/mol)

Boundary diffusion 61 [83]

Lattice diffusion 92 [83]

Core diffusion 50 – 64 [53, 83, 139]

Low-temperature dislocation creep (T < 142°C) 79 – 112 [17, 40, 53, 129-131, 139]

High-temperature dislocation creep (T > 280°C) 152 – 168 [126, 130, 131, 133]

Coble creep 62 – 72 [136-138]

Nabarro-Herring creep 94 – 98 [138]

24 2. Theoretical background

2.2.7. Deformation mechanism maps

Frost and Ashby [83] have constructed a deformation mechanism map for pure Zn with a grain size of 100 µm from experimental data which were fitted to model-based rate-equations that describe the different deformation mechanisms, Figure 2.11, showing the predominant deformation mechanisms with regards to stress and temperature. It is apparent from the deformation mechanism map of pure Zn that dislocation-mediated mechanisms such as dislocation glide and deformation twinning dominate at high stress levels, while at elevated temperatures (T/Tm > 0.3, where Tm is the melting point) and intermediate stress levels power law creep can occur. At higher temperatures (T/Tm > 0.5) and very low stress levels, diffusional flow such as Nabarro-Herring creep and Coble creep dominates deformation.

Figure 2.11. Deformation mechanism map of pure zinc with a grain size of 100 μm [83].

Similar to Frost and Ashby’s work, Kawasaki and Langdon [160] have published a deformation mechanism map for eutectoid Zn-22Al alloys at 200°C in the form of D/b plotted against the normalised stress σ/G using experimental data obtained through tensile test over a wide range of strain rates [101, 123, 160], Figure 2.12 [160], where D is the grain size and b is the burger vector of Zn-22Al alloys (b = 2.7∙10−10 m [161]). The regions I, II and III in the map refer to the three regions of superplastic deformation, where a maximal amount of grain boundary sliding has been reported to occur in region II [100, 101, 109, 160-162]. According to this deformation mechanism map of Zn-22Al alloys at 200°C [160], superplastic deformation which is dominated

25 2. Theoretical background by grain boundary sliding occurs at relatively high stress levels, while diffusional flow dominates deformation at low stress levels. Further, Coble creep is predominant in alloys swith small grain size (finer than ~ 20 µm) while Nabarro-Herring creep is active in alloy with larger grains.

Figure 2.12. Deformation mechanism map of normalized grain size versus normalized stress for Zn–22% Al alloys tested at 200°C, reproduced from [160].

It is evident that the active deformation modes in Zn and Zn-Al based alloys are highly temperature and stress dependent. Table 2.7 gives an overview of the active deformation mechanisms of Zn and Zn-Al based alloys reported in the literature (creep mechanisms are shown in Table 2.4).

26 2. Theoretical background

Table 2.7. Deformation mechanisms in Zn and Zn alloys reported in literature. The corresponding references are included in the table3.

Strain rate Grain size Experimental Temperature Predominant Material Strain rate sensitivity, Reference (µm) method range (ºC) mechanism m Pyramidal + Single crystalline Compression 1.67∙10-4 - RT-200 - basal dislocation Hagihara 2016 [79] zinc test slip; kink bands a and c+a Compression Polycrystalline zinc 1∙10-4-10 0.2 - 0.25 RT - dislocation slip; Dirras 2013 [90] test GBS a and c+a Compression Polycrystalline zinc 1∙10-4-10 780 RT - dislocation slip; Dirras 2013 [90] test twinning Dislocation slip; Compression Polycrystalline zinc 5∙10-4 2000 RT - twinning; kink Li 2007 [91] test bands Dislocation slip; Polycrystalline zinc 1∙10-4 - 1∙10-2 70 Tensile test RT - twinning; kink Liu 2008 [103] bands Dislocation slip Tensile strain in large grains Polycrystalline zinc 10-4 - 10-2 ~ 0.24 20 - 40 0.15 Zhang 2002 [94] rate jump test and GBS in small nanograins

3 This table was published in an article. The original citation is: Z. Wu, S. Sandlöbes, J. Rao, J. S. K.-L. Gibson, B. Berkels and S. Korte-Kerzel (2018). "Local mechanical properties and plasticity mechanisms in a Zn-Al eutectic alloy." Materials & Design 157: 337-350.

27 2. Theoretical background

Strain rate Grain size Experimental Temperature Predominant Material Strain rate sensitivity, Reference (µm) method range (ºC) mechanism m a and c+a Indentation Polycrystalline zinc 1∙10-5 - 1∙10-3 0.2 330 - 360 0.13 dislocation slip; Jenei 2014 [104] creep test GBS

Zn-0.3Al 2∙10-4 - 5∙10-2 ~ 1 Tensile test RT 0.4 GBS Ha 2001 [163]

Zn-0.3Al 2∙10-4 - 5∙10-2 ~ 10 Tensile test RT 0.07 Dislocation slip Ha 2001 [163]

Zn-0.3Al 10-4 - 10-3 ~ 2 Tensile test RT 0.31 GBS Demirtas 2015 [30]

Zn-0.3Al 10-3 - 10-1 ~ 2 Tensile test RT 0.2 Dislocation slip Demirtas 2015 [30]

Dislocation Zn-0.4Al 10-1 ~ 0.6 Tensile test RT 0.05 Naziri 1974 [120] recovery

Zn-0.4Al 10-4 - 10-3 ~ 0.6 Tensile test RT 0.4 GBS Naziri 1974 [120]

10-4 - 10-2 ~ 0.54 Tensile test RT 0.25 GBS Demirtas 2015 [159] Zn-5Al 10-2 - 1 ~ 0.54 Tensile test RT 0.18 GBS Demirtas 2015 [159]

Zn-22Al 10-3 - 10-1 0.2 - 0.75 Tensile test RT 0.3 GBS Cetin 2016 [98]

Zn-22Al 10-4 - 10-2 0.6 Tensile test RT 0.24 GBS Uesugi 2015 [96]

28 2. Theoretical background

Strain rate Grain size Experimental Temperature Predominant Material Strain rate sensitivity, Reference (µm) method range (ºC) mechanism m

Zn-22Al 0.0125 - 0.1 ~ 1.4 Nanoinden- RT 0.124 GBS Choi 2014 [158] tation test

(Pmax=20 0.226 – Zn-22 Al 0.0125 - 0.1 ~ 0.35 RT GBS Choi 2014 [158] mN) 0.256

Zn-22Al 10-4 - 1 ~ 0.8 Tensile test RT 0.2 – 0.35 GBS Xia 2008 [164]

Zn-22Al 10-3 - 10-1 ~ 0.8 Tensile test 200 0.43 GBS Kawasaki 2008 [101]

Zn-22Al 10-2 - 10-1 0.35 Tensile test 200 0.5 GBS Kawasaki 2011 [160]

Zn-22Al 10-4 - 10-5 ~ 2.5 Tensile test 160-220 0.4 GBS Duong 1998 [100] (+ Cu and Fe) Dislocation glide Zn-22Al 10-2 ~ 1.3 Tensile test 30 0.2 Tanaka 2002 [122] and climb

Zn-22Al 10-5 ~ 1.3 Tensile test 45 0.3 GBS Tanaka 2002 [122]

Zn-22Al 2∙10-3 ~ 1.3 Tensile test 200 0.5 GBS Tanaka 2002 [122]

29 2. Theoretical background

2.3. Nanoindentation

Nanoindentation is an indirect method of measuring the hardness of a material as the dimensions of the residual impression are in the sub-micrometer range and difficult to be precisely detected upon removal of the load [165]. Nanoindentation enables to evaluate the relationship between the mechanical properties and the local crystallographic orientation or the microstructure with a high accuracy. Further, the internal plasticity mechanisms of nano- or micro-scale materials such as thin films, coatings or membranes can be investigated [165]. The plasticity behaviour of brittle materials such as ceramics, glasses and intermetallic phases can also be analysed by utilizing the size effect. Figure 2.13 [165] shows the different geometries of commonly used indenters. Among them, the three-sided pyramidal Berkovich indenter with a centreline-to-face angle of 65.3º is the most commonly used indenter type as it allows precise control of the indentation process [166].

Figure 2.13. Geometries of commonly used indenters: (a) Spherical indenter; (b) Conical indenter; (c) Vickers indenter; (d) Berkovich indenter [165]. Various approaches have been developed to analyse the mechanical data of indentation experiments, such as Oliver and Pharr [167], Field and Swain [168], Doerner and Nix [169] as well as Briscoe and Sebastian [170]. The most commonly used two systematic analysis methods for indentation are the “Oliver and Pharr” method, published in 1992, and the “Field and Swain” method. Both methods are based on the elastic equations of Hertz [171], the only difference between these two approaches is the determination of the plastic depth, which is the distance from the periphery of the indenter contact to the maximum penetration depth. The “Field and Swain” method is usually applied in experiments with a spherical indenter and the application range of the “Oliver and Pharr” method can be extended to all axis-symmetric indenters with a smooth profile [172] and is therefore applied in this study.

There are two critical parameters, the hardness, H, and the elastic modulus, E, which can be derived from indentation load-displacement data obtianed in one complete circle of indentation. The Meyer hardness, HM, has been frequently used to calculate the hardness from nanoindentation tests [165, 167]. The Meyer hardness is obtained after the material has reached the plastic deformation region, its relationship with the reduced modulus Er is [165, 167]:

30 2. Theoretical background

Pmax Pmax H = = (2.17) M A 2 c ghc

S π S π E = √ = (2.18) r 2 A 2 √ 2 c ghc

where Pmax is the maximum indentation load and Ac is the projected contact area of the hardness impression measured at the contact depth hc, as Figure 2.14 [165, 167, 173] presents. S is the stiffness, g is a geometrical factor which depends on the geometry of the indenter, for a Berkovich indenter g is 24.5 [165]. It has been reported [165] that Eq. (2.18) can be applied to any axi-symmetric indenter and describes also well the elastic-plastic contact although the derivation is based on the elastic contact.

Figure 2.14. (a) Schematic illustration of the indenter and the specimen surface at full load and unloaded for a Berkovich indenter; (b) Schematic load versus displacement curve for elastic-plastic loading followed by elastic unloading. At the maximum load, Pmax, hmax is the depth from the original specimen surface to the tip of the indenter, hc is the contact depth of indenter and material and hs is the distance from the edge of the contact to the specimen surface. hf is the final depth of the residual impression after indentation. Upon elastic reloading, the eventual point of contact with the specimen surface moves through a distance hs [165, 167, 173].

The elastic unloading stiffness in Eq. (2.18), is also known as contact stiffness and is defined as the slope of the upper portion of the curve during the initial stage of unloading from the maximum load Pmax [167]:

dP S= (2.19) dh

Oliver and Pharr [167] have fitted the unloading curve approximately with a power law relation

31 2. Theoretical background in order to get the S value:

1 푚1 P=훼 (h-h푓) (2.20)

1 1 where hf is the final depth of the residual impression after indentation, 훼 ⁡and 푚 are power law fitting constants. α1 depends on the indented material and m1 depends on the indenter behaviour during indentation.

According to Figure 2.14, the contact depth hc can be derived as [167]:

P h = h -h = h -ϵ’ max (2.21) c max s max S where hmax is the depth from the original specimen surface to the tip of the indenter, hs is the distance from the edge of the contact to the specimen surface, ϵ’ is a constant that depends on the geometry of the indenter. Durst and Maier [150] have recommended 푚1 = 1.5 and ϵ’ = 0.75 for indentation using a Berkovich or Vickers indenter that behaves like a paraboloid of revolution [173].

As Er is considered to be interfered by elastic displacements of both, the specimen and the indenter, the elastic modulus, E, of the specimen can be obtained by [167]:

1 1-휈2 1-휈2 = + i (2.22) Er E Ei where E and 휈 are Young’s modulus and Poisson’s ratio of the specimen and Ei and 휈푖 are Young’s modulus and Poisson’s ratio of the indenter, respectively.

The indentation can be set either in load or depth control. In the load-controlled mode, the maximum testing load and the incremental load intervals are recorded and the incremental evolution is set as square root or linear [165]. On the other hand, the maximum depth of penetration is set previously in the depth-controlled mode [165]. The load-controlled mode is more commonly used [165].

During a load-controlled indentation testing, two testing methods are commonly used to control the indentation. One way is to apply a constant loading rate upon time, which can be P interpreted as⁡Ṗ = [150]. Here, several different loads can be complied in one indentation to t study the mechanical properties of the material. The other way is to set the indentation strain rate constant, this method is employed more frequently in load-controlled indentation experiments [150]. When assuming that the hardness is independent of the indentation depth,

Ṗ the relation between the effective strain rate, ε̇, and the indentation strain rate, , can be P roughly described according to Lucas and Oliver [174]:

32 2. Theoretical background

ḣ 1 Ṗ Ḣ 1 Ṗ ε̇= = ( - ) ≈ (2.23) h 2 P H 2 P

The shape and bunt of the indenter, the compliance of the frame, pile-ups and sink-ins in the vicinity to the indenter as well as thermal drift etc. influence nanoindentation experiments. However, due to the experimental and time limitations, only a few of the critical influential factors like the compliance of the frame and thermal drift can be taken into consideration. Errors of the resultant hardness and elastic modulus cannot be avoided, but the approximate data can to some extend still be compared with uniaxial macroscopic mechanical results according to previous studies [150, 175].

33 3. Experimental methods and materials

3. Experimental methods and materials4

3.1. Materials

As a benchmark, plates of commercial zinc alloy Z410 with dimensions 3 mm x 30 mm x 150 mm were produced by hot chamber die casting at HTW Aalen (Beethovenstrasse 1, 73430 Aalen, Germany). On the basis of the composition of alloy Z410, three laboratory alloys ZnAl4Cu1Mg0.04, ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31 (in wt.%) were prepared in-house by gravity casting (IMM, Kopernikusstrasse 14, 52074 Aachen, Germany). The alloys were melted in air in a resistance furnace using Zn-3.8Al-1.4Cu-0.04Mg (wt.%), Zn-3.8Al-0.04Mg (wt.%) master alloys, 99.99% pure Al and 99.96% pure Mg. The melt was cast into a preheated copper mould and cooled down in air. The chemical compositions of the alloys were determined by both optical emission spectroscopy using a MAXx spectrometer and inductively coupled plasma atomic emission spectroscopy using a spectro Ciros Vision spectrometer. Optical emission spectroscopy was used to determine the composition of all elements and the accurate composition of Cu and Mg were determined by inductively coupled plasma atomic emission spectroscopy, Table 3.1. The impurities such as Fe and Si were controlled to not exceed 10 ppm Si (5 ppm ZnAl4Cu1Mg0.04; 6 ppm ZnAl4Cu1Mg0.21; 7 ppm ZnAl4Cu1Mg0.31) and 50 ppm Fe (48 ppm ZnAl4Cu1Mg0.04; 41 ppm ZnAl4Cu1Mg0.21; 43 ppm ZnAl4Cu1Mg0.31) since Mohamed et al. [99, 108, 176] and Uesugi et al. [96] have reported that Fe and Si impurities can significantly influence the mechanical properties of Zn alloys.

3.2. Sample preparation

Metallographic specimens were prepared by mechanical grinding with 1200#, 2400# and 4000# SiC grinding papers and polishing with 6 µm, 3 μm, 1 μm and 0.25 μm diamond suspension down to 50 nm gamma alumina powder finish (OPA particles). An intermediate etching step with 90% ethanol + 10% nitrate acid solution to remove the deformation layer was

4 Part of this chapter appeared as articles. The original citations are: Z. Wu, S. Sandlöbes, L. Wu, W. P. Hu, G. Gottstein and S. Korte-Kerzel (2016). "Mechanical behaviour of Zn-Al-Cu-Mg alloys: Deformation mechanisms of as-cast microstructures." Materials Science and Engineering: A 651: 675-687., S. Sandlöbes, Z. Wu, K. Pradeep and S. Korte-Kerzel (2016). "Precipitation and decomposition phenomena in a Zn-Al-Cu-Mg alloy." Materials Letters 175: 27-31., Z. Wu, S. Sandlöbes, Y. Wang, J. S. K.-L. Gibson and S. Korte-Kerzel (2018). "Creep behaviour of eutectic Zn-Al-Cu-Mg alloys." Materials Science and Engineering: A 724: 80-94., Z. Wu, S. Sandlöbes, J. Rao, J. S. K.-L. Gibson, B. Berkels and S. Korte-Kerzel (2018). "Local mechanical properties and plasticity mechanisms in a Zn-Al eutectic alloy." Materials & Design 157: 337-350., and Z. Wu, S. Sandlöbes, J. Rao, J. S. K.-L. Gibson, B. Berkels and S. Korte-Kerzel (2018). "Data on measurement of the strain partitioning in a multiphase Zn-Al eutectic alloy." Data in Brief: https://doi.org/10.1016/j.dib.2018.09.010.

34 3. Experimental methods and materials performed between 6 µm and 3 µm diamond polishing steps. The entire polishing procedure was performed water-free.

For transmission electron microscopy (TEM) observation, discs with a diameter of 3 mm and height of 0.8 mm were cut by electric discharge machining, mechanically ground to 100 µm thickness and electro-polished until perforation using 5 vol.% perchloric acid, 35 vol.% butanol, 60 vol.% ethanol at -30°C. APT samples were prepared using a focused ion beam (FEI Helios Nanolab 660) as described in [177].

Table 3.1. Chemical composition of the zinc alloys investigated.

Element content, wt.% Alloy Al Cu Mg Zn

Z410 4.32 0.55 0.04 Bal.

ZnAl4Cu1Mg0.04 4.30 0.59 0.04 Bal.

ZnAl4Cu1Mg0.21 4.30 0.58 0.21 Bal.

ZnAl4Cu1Mg0.31 4.30 0.59 0.31 Bal.

3.3. Macroscopic mechanical testing

3.3.1. Constant strain rate tensile testing

The mechanical properties of the as-cast zinc alloys ZnAl4Cu1Mg0.04, ZnAl4Cu1Mg0.21 and

ZnAl4Cu1Mg0.31 were determined by tensile tests at room temperature (RT) (0.45T/Tm), 55°C

−4 −5 −6 -1 (0.5T/Tm) and 85°C (0.55T/Tm) at strain rates (휀̇) of 5∙10 , 5∙10 and 6∙10 s . The alloys were additionally tested in tension at 105°C at a strain rate of 5∙10−4 s-1. For comparison, the

−4 -1 −6 -1 commercial alloy Z410 was tested at RT and 85°C (0.55T/Tm) at 휀̇ of 5∙10 s and 6∙10 s . An electromechanical testing machine (DZM) with an accuracy of 0.17 MPa equipped with an electric furnace was used for these experiments. Flat dog-bone shaped specimens with a cross section of 1.5 mm X 4.0 mm in the centre region and a gauge length of 26 mm were used for the tests, as shown in Figure 3.1.

Figure 3.1. Sketch of the samples for constant strain rate tensile tests.

35 3. Experimental methods and materials

3.3.2. Strain rate jump test

Tensile strain-rate jump tests were performed on bulk ZnAl4Cu1Mg0.04, ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31 alloys using an electromechanical testing machine at 55°C and 85°C with strain rates jumps between 5∙10-4 s-1 and 6∙10-6 s-1. During the tests, the strain rate was kept constant at 5∙10-5 s-1 until 4% true strain, after that the strain rate was changed among 5∙10-4 s-1, 5∙10-5 s-1 and 6∙10-6 s-1 at every 4% true strain. Figure 3.2 shows the sample geometry used for tensile strain rate jump tests, where the cross section of the centre region is 1.5 mm X 1.0 mm and the gauge length is 3.56 mm.

Figure 3.2. Geometry of specimens for tensile strain rate jump tests, in mm.

3.3.3. Creep test

Uniaxial tensile creep tests were performed in air at 25°C (only ZnAl4Cu1Mg0.04 and

ZnAl4Cu1Mg0.31), 55°C, 85°C and 105°C at stresses of 0.6∙σ0.2, 0.7∙σ0.2 and 0.8∙σ0.2, respectively, where σ0.2 is the yield strength of the alloy at the corresponding temperature as determined by tensile tests. The creep experiments were stopped when the stationary creep stage was reached. Table 3.2 lists the applied test parameters, where the normalised stress σ/G was calculated based on the corresponding shear moduli of the alloys (more details are given in Table 5.3). Specimens for tensile creep tests were dog-bone shaped tensile samples with a thickness of 1.50 mm, a width of 4 mm at the centre region and a gauge length of 5.33 mm, Figure 3.3.

36 3. Experimental methods and materials

Figure 3.3. Geometry of specimens for tensile creep tests, unit in mm.

Table 3.2. Applied parameters (temperature, stress) during uniaxial creep experiments*.

Alloy ZnAl4Cu1Mg0.04 ZnAl4Cu1Mg0.21 ZnAl4Cu1Mg0.31

-3 -3 -3 T, °C σ/σ0.2 σ, MPa σ/G, ∙10 σ, MPa σ/G, ∙10 σ, MPa σ/G, ∙10

0.6 87 1.89 98 2.14 87 1.90

55 0.7 101 2.20 114 2.49 102 2.23

0.8 115 2.50 130 2.84 117 2.56

0.6 76 1.69 93 2.07 80 1.79

85 0.7 89 1.98 108 2.41 93 2.08

0.8 101 2.25 124 2.77 106 2.37

0.6 61 1.38 65 1.47 63 1.43

105 0.7 71 1.60 75 1.70 73 1.66

0.8 81 1.83 85 1.92 83 1.88

* Additionally, alloy ZnAl4Cu1Mg0.21 was tested at 55°C and 87 MPa and alloy ZnAl4Cu1Mg0.04 was tested at 85°C and 93 MPa.

3.3.4. In-situ straining experiments

In-situ straining experiments were performed in a JEOL 820 SEM with a commercial tensile/compression stage (Kammrath & Weiss Co.) and an integrated resistance furnace.

37 3. Experimental methods and materials

Alloy ZnAl4Cu1Mg0.31 was tested at RT with a cross head velocity of 0.1 µm∙s-1, and at 85°C with a cross head velocity of 0.25 µm∙s-1. The specimens for in-situ straining were flat dog- bone shaped samples where the centre region of the gauge length was cut to 0.5 X 0.5 mm (width x length) for observation in the SEM, Figure 3.4. Due to this deviation from standard tensile specimen dimensions and the resulting inaccuracies in determining the strain within the gauge length, displacements and displacement rates rather than strains and strain rate are given throughout.

Figure 3.4. Geometry of specimens for in-situ straining experiments, unit in mm.

3.4. Nano-indentation experiments

3.4.1. Constant strain rate and strain rate jump tests

Load-controlled quasi-static nanoindentation testing of the individual microstructural constituents was conducted using a Nanomechanics iNano® nanoindenter at RT with maximum loads of 1 mN, 5 mN, 25 mN and 40 mN at a constant strain rate of 0.1 s-1 using a diamond Berkovich indenter (with a Poisson ratio of 0.07 and a tip radius of 50 nm as determined from the diamond area function). At least 120 independent indentations were performed at each test condition. Nanoindentation strain rate jump tests at RT were performed using the same machine setup. The indentation strain rate was changed from 5∙10-2 s-1 to 5∙10-4 s-1 and back with an jumping interval of 200 nm displacement. At least 10 indentations into each individual microstructural constituent were conducted.

Nanoindentation tests at 85°C were performed in vacuum (6 ∙ 10-5 mbar) using a MicroMaterials NanoTest® Platform3 (modified for vacuum operation) using a cubic boron nitride Berkovich indenter (with a Poisson ratio of 0.15 and a tip radius of 100 nm as determined from the diamond area function) at constant strain rates of 0.01 and 0.1 s-1 with a peak load of 25 mN and an unloading rate of 10 mN/s. As η-Zn and η-Zn+α-Al eutectoid

38 3. Experimental methods and materials structures have been reported to have a low creep resistance [178], their hardness and elastic modulus are affected by creep during the holding period [179, 180]. As the holding period only lasted one second, no significant reduction in hardness due to creep is expected [180]. However, the elastic modulus was observed to be affected by residual plasticity upon unloading, causing an under-estimation of the Young’s modulus, even with rapid unloading (~2 seconds). Therefore, additional tests that could utilise continuous stiffness measurements were performed. Nanoindentation strain rate jump tests at 85°C were conducted on a Nanomechanics InSEM-III® in vacuum at strain rates jumping between 5∙10-2 s-1 and 5∙10-3 s-1 with a jump interval of 200 nm displacement using a sapphire Berkovich indenter tip (with a Poisson ratio of 0.24 and a tip radius of 50 nm as determined from the diamond area function), the stiffness was continuously measured during indentation. To minimize the thermal drift during indentation, the indenter and specimen were heated independently. During indentation experiments, the thermal drift was below 0.4 nm∙s-1. 168 independent indentations were performed to ensure a sufficient number of indents into all microstructural constituents. Particularly at high temperature, pile-ups around the indents were observed. As an AFM scan of every indent is prohibitively time-intensive, for a representative, single indent in each phase, the actual contact area was calculated from AFM and used to calculate the true value of hardness and modulus [181]. All the subsequently reported high-temperature mechanical data are obtained from experiments performed using the InSEM-III nanoindenter, while the microstructural characterisation was performed on indents carried out using the MicroMaterials system. For calculation of the Young modulus, Poisson ratios of 0.24 for η-Zn and 0.30 for η-Zn+α-Al eutectoid eutectic / eutectoid structures were used [182].

Before indentation on all instruments, frame stiffness and tip shape calibrations were carried out on fused silica and verified on a polycrystalline tungsten sample according to the method of Oliver and Pharr [173]. The hardness and elastic moduli were continuously measured during indentation.

3.4.2. Nanoindentation creep tests

Nanoindentation creep experiments were conducted at RT and 85°C in vacuum (6∙10-5 mbar) on a MicroMaterials NanoTest Platform 3 (modified for vacuum operation) at maximum loads of 15 mN, 25 mN and 35 mN using a cubic boron nitride Berkovich indenter. Frame stiffness and tip shape calibrations were carried out on fused silica and verified using a polycrystalline tungsten sample according to the method of Oliver and Pharr [173] before indentation. Loading was carried out at a constant strain rate of 0.1 s-1, after which the load was held constant for 60 s and the change in displacement was continuously recorded. To minimize the thermal drift during indentation, the indenter and specimen were heated independently. During indentation

39 3. Experimental methods and materials experiments, the thermal drift was below 0.1 nm∙s-1. At least 45 independent indentations were performed at each test condition.

3.5. Microstructure characterisation

The samples were examined using a LEO1530, a JEOL JSM7000F and an FEI Helios Nanolab 600i scanning electron microscope (SEM) at an accelerating voltage of 20 kV using secondary electron microscopy (SE), backscattered electron microscopy (BSE), energy dispersive X-ray spectroscopy (EDS) and electron backscatter diffraction (EBSD). The EBSD data were analysed using the OIMTM data collection software.

The surface topology of specimens were characterised using an XE-70 (Park Systems) and a Dimension 3100 (Bruker) atomic force microscope (AFM), both operated in the non-contact mode. The AFM data were analysed using the software WSXM 5.0 [183].

TEM observations were performed using a Philips CM20 TEM operated at 200 kV. Identification of dislocation planes and precipitate habit planes was done using the software JEMS-SAAS. Pulsed laser APT was carried out in a local electrode atom probe (LEAP 4000X HR, Cameca Inc.) at a specimen temperature of 60 K.

3.6. Digital image correlation (DIC) measurements

The samples for DIC measurements were dog-bone shaped specimens with a gauge length of 3.56 mm and a cross-section of 1x1.5 mm2, Figure 3.3. Table 3.3 shows the detailed procedure of pattern deposition. A monolayer of SiO2 particles with an average particle size of 40 nm was dispersed on the specimen surface, cf. Figure 3.5 b. This kind of SiO2 particle pattern has been reported to be favourable for digital image correlation and to be transparent for EBSD measurements [184].

Quasi-insitu tensile tests were performed at 85°C and a constant strain rate of 5∙10-4 s-1 on an electromechanical testing machine. The tests were interrupted at 2% and 5% global elongation, respectively. SEM micrographs and EBSD maps were obtained prior to deformation and after 2% and 5% global elongation, respectively. Secondary electron images were achieved using an in-lens SE detector, a low beam voltage of 3 kV and a working distance of 3 mm. 7 x 9 images were acquired with 20% overlap between two adjacent images in both horizontal and longitudinal directions. The images were analysed using the software GOM Correlate (V8.1, GOM mbH).

40 3. Experimental methods and materials

Table 3.3. Detailed procedure of depositing SiO2 nanoparticles on the sample surface.

Step Procedure Purpose

Prepare the specimen using Obtaining a flat and deformation-free 1 standard metallographic procedures surface. until 50 nm gamma alumina finish.

Preparation of an OPS suspension Homogeneous distribution of SiO2 (SiO particles, Struers®) in ethanol particles in the suspension. 2 2 with a ratio of 1:3 (SiO2:ethanol) using an ultrasonic bath.

Dipping of the specimen in the Deposition of SiO2 particles on suspension for 30 s in ultrasonic specimen surface with a 3 bath. homogeneous distribution to avoid accumulation of particles. Rinsing the sample with ethanol Removal of floating particles,

4 spray, and subsequent drying of the formation of a monolayer of SiO2 sample with an air fan. particles on the sample surface.

The image correlation method was performed on the individual images separately. The final strain maps (shown in Figure 6.6) were stitched together from 7 x 9 images using the "Microsoft Image Composite Editor" software.

Figure 3.5. Secondary electron (SE) micrographs of the region of interest prior to deformation. (a) In-lens SE image of the whole region of interest. (b) In-lens SE image of one individual image, enlarged from the red box in the top left corner in (a). (c) Enlarged in-lens SE micrograph of the region highlighted by the red box in (b), showing SiO2 particle speckle pattern on the sample surface.

41 4. Macro mechanical response and mechanisms

4. Macroscopic mechanical response and mechanisms5

The commercial alloy Z410 (ZnAl4Cu1, in wt.%) is a widely used commercial zinc die casting alloy and provides a good combination of strength, ductility as well as excellent plating and finishing characteristics [12, 28]. However, most commercial Zn-Al alloys possess limited ductility, poor impact toughness and low strength at room temperature [5, 11, 19, 20, 28]. With only a slightly difference in temperature the ZnAl4Cu1 alloys displays obviously dissimilar mechanical properties: brittleness at room temperature and large ductility at slightly elevated temperature (e.g. 85°C). To understand why the alloys display such different mechanical properties at only slightly different temperatures, this thesis aims to establish a deeper understanding of the relation between microstructure and mechanical properties and the underlying deformation and work hardening mechanisms in Zn-Al-Cu alloys at different temperatures. Moreover, although the addition of Mg was shown to have beneficial effects in other Zn-Al alloys [1, 9, 14, 23, 25], the influence of Mg addition on the mechanical properties and microstructure of ZnAl4Cu1 have not yet been studied. Therefore, it is of great interest to explore the potential of Mg as a strengthening alloying element. To this end, this chapter mainly focuses on the microstructure, macro mechanical properties and deformation behaviour of ZnAl4Cu1 alloys with and without dilute Mg additions at different temperatures and strain rates.

4.1. Results and discussion

4.1.1. As-cast microstructure

Figure 4.1 shows the microstructures of the as-cast laboratory alloys ZnAl4Cu1Mg0.04, ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31 and the die-cast commercial alloy Z410 for comparison. It can be seen from Figure 4.1a that the microstructure of alloy Z410 is mainly comprised of bright globular grains with an average grain size of 5 – 6 µm, a fine lamellar and a coarse cellular structure. On the other hand, the microstructure of alloy ZnAl4Cu1Mg0.04 (Figure 4.1b), which has the same nominal chemical composition as alloy Z410 but was solidified with a slower cooling rate, is comprised of a much coarser dendritic phase with dendrite size of 27 – 70 µm, a lamellar eutectic structure and some darker areas at the boundaries of the dendrites exhibiting a fine lamellar structure, see the enlarged micrograph in Figure 4.1b. Similarly, the microstructures of alloys ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31 are also comprised of coarse dendrites, as well as coarse and fine lamellar eutectic structures, while an additional phase is observed in between the dendrites and lamellar areas, giving an

5 This part was published as an article. The original citation is: Z. Wu, S. Sandlöbes, L. Wu, W. P. Hu, G. Gottstein and S. Korte-Kerzel (2016). "Mechanical behaviour of Zn-Al-Cu-Mg alloys: Deformation mechanisms of as-cast microstructures." Materials Science and Engineering: A 651: 675-687.

42 4. Macro mechanical response and mechanisms additional, intermediate contrast in the BSE micrographs compared with the microstructures containing no additional magnesium.

Figure 4.1. Microstructures of as-cast ZnAl4Cu1 alloys (BSE); (a) Z410, (b) ZnAl4Cu1Mg0.04, (c) ZnAl4Cu1Mg0.21, (d) ZnAl4Cu1Mg0.31.

It is known from the Zn-Al phase diagram [185] that in hypoeutectic Zn-Al alloys the primary η- Zn phase forms first during solidification, below the eutectic temperature of 382°C the remaining liquid phase transforms to a η-Zn + β-Al eutectic structure. Below 275°C the β-Al phase subsequently decomposes into a η-Zn + α-Al eutectoid structure. Figure 4.2 shows the

43 4. Macro mechanical response and mechanisms

EDS results of the alloys ZnAl4Cu1Mg0.04 and ZnAl4Cu1Mg0.21 (the results for alloys ZnAl4Cu1Mg0.31 are not shown here because the phase identification is identical to alloy ZnAl4Cu1Mg0.21), revealing that the coarse dendrites in alloys ZnAl4Cu1Mg0.04 (spot 1 and selected area 2 in the left image of Figure 4.2a) and ZnAl4Cu1Mg0.21 (spot 4 in Figure 4.2b) are primary η-Zn phase. The lamellar structure in alloys ZnAl4Cu1Mg0.04, ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31 (area 1 in the left image of Figure 4.2a) is identified as η-Zn + α-Al eutectic, and the fine lamellar structure (dark areas in the lower magnification micrographs; selected area 1 in the right image of Figure 4.2a, spot 2 in Figure 4.2b) as η-Zn + α-Al eutectoid. It is evident from the enlarged micrograph of the eutectic area, Figure 4.2a, that the eutectic areas are comprised of secondary η-Zn (spot 1 in the right image of Figure 4.2a) and eutectoid η-Al + α-Zn lamellars (spot 2, 3 and 4 in the right image of Figure 4.2a).

The precipitates giving intermediate BSE-contrast in alloys ZnAl4Cu1Mg0.21 and

ZnAl4Cu1Mg0.31 are identified as Mg2Zn11 (spot 3 in Figure 4.2b). Figure 4.2b shows an EDS mapping of the area containing the Mg2Zn11 phase revealing that the Mg2Zn11 phase is embedded in between η-Zn and eutectoid η-Zn + α-Al.

Figure 4.2. EDS analysis of alloys (a) ZnAl4Cu1Mg0.04 and (b) ZnAl4Cu1Mg0.21. The phase fractions of each phase were determined by measuring the observed area fractions using the software Image-Pro© as listed in Table 4.1. As already mentioned, it is found that

44 4. Macro mechanical response and mechanisms

Mg alloying caused the formation of Mg2Zn11 phase precipitates in the alloys ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31. The formation of Mg2Zn11 phase precipitates has been reported before for a Zn-4Al-3Mg solder [186]. Additionally, with increasing Mg content the phase fraction of primary η-Zn phase is slightly increased. This is attributed to retarded diffusion of Zn and Al through trapping of vacancies by Mg atoms due to their slightly larger atomic size (125 pm Al, 145 pm Zn, 150 pm Mg) [187, 188]. Since the eutectic and eutectoid transformations are diffusion controlled processes, the addition of Mg is assumed to cause a significant retardation of the transformation kinetics of the eutectic and eutectoid transformation in Zn-Al alloys as has been similarly proposed by Ling et al. [189] for a Zn-Al eutectoid alloy and by da Costa et al. [25] for a Zn-4Al-3Cu alloy. Due to the delayed eutectic transformation the solidification of the primary η-Zn phase continues to lower temperatures and hence could promote nucleation and growth of primary η-Zn phase in the alloys ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31.

Table 4.1. Grain size and phase fractions of investigated ZnAl4Cu1 alloys.

Grain size, µm Phase fraction, % (area)

η-Zn + α- Alloy primary η-Zn eutectoid primary Al lamellar Mg Zn (diameter / short x η-Zn 2 11 eutectic / spacing long axis) eutectoid

Z410 5.56 ± 1.80 0.18 ± 0.03 31.7 ± 1.6 - bal.

48.3 ± 21.4 x 29.1 ZnAl4Cu1Mg0.04 0.27 ± 0.07 29.6 ± 5.0 - bal. ± 15.7 46.8 ± 20.4 x 28.7 ZnAl4Cu1Mg0.21 0.13 ± 0.03 35.2 ± 5.9 0.2 ± 0.1 bal. ± 15.9 44.8 ± 18.8 x 25.6 ZnAl4Cu1Mg0.31 0.17 ± 0.03 34.4 ± 5.3 0.5 ± 0.2 bal. ± 13.9

The grain size of the primary η-Zn is also shown in Table 4.1. For alloy Z410 the grain size was measured in terms of the diameter and in the alloys ZnAl4Cu1Mg0.04, ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31 in terms of the projected lengths along the long and short axes of the dendrites. The grain size of primary η-Zn phase in Z410 is much finer than in the laboratory alloy ZnAl4Cu1Mg0.04. This is assumed to result from the relatively low cooling rate in the gravity castings (ZnAl4Cu1Mg0.04) compared to the high cooling rate of die casting (Z410) consistent with observations by Krupinska et al. [9] who reported microstructural refinement with increasing cooling rate in a ZnAl4Cu1 alloy.

The secondary lamellar spacing, i.e. the eutectoid lamellar spacing, was also measured and the results are listed in Table 4.1. Here it is found that the lamellar structure in alloys

45 4. Macro mechanical response and mechanisms

ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31, which contain higher Mg contents, is much finer than that in alloy ZnAl4Cu1Mg0.04. Similarly, da Costa et al. [25] reported decreased eutectoid lamellar spacings for Zn-4Al-4Cu-XMg (X: 0.002; 0.5; 1 wt.%) alloys with increasing Mg content. Further, Li et al. [28] reported refinement of the eutectic structure by addition of 0.3 Zr in a Zn- 4Al alloy. We attribute the observed refinement of the η-Zn + α-Al eutectoid lamellar structures also to the retarded eutectic and eutectoid transformation kinetics through the addition of Mg. More specifically, due to the retarded eutectic transformation, it is assumed that directly after the eutectic reaction the transformation into the eutectoid structure starts, hence, causing a higher volume fraction of “isolated” eutectoid structure regions with a refined lamellar structure, as experimentally observed in alloys ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31 (Figure 4.1). Here, the smallest eutectoid lamellar spacing is observed in alloy ZnAl4Cu1Mg0.21 with a medium Mg concentration (0.21 wt.%). The observed coarser eutectoid lamellar spacing in alloy ZnAl4Cu1Mg0.31 (0.31 wt.% Mg concentration) than in alloy ZnAl4Cu1Mg0.21 might also be caused by retarded diffusion of Zn and Al through higher Mg addition leading to a higher growth rate of the eutectoid lamellae instead of the nucleation of new lamellae. However, this point is not yet fully understood and requires further investigation in the future.

4.1.2. Mechanical properties

Figure 4.3 shows the true stress - true strain curves of the as-cast alloys. The corresponding yield strength (휎0.2) and elongation to fracture (휀푓) values are listed in Table 4.2. At room temperature and/or highest strain rate tested (5∙10−4 s-1), the alloys show limited plastic deformation, with 휀푓 values less than 1.5%, and the flow curves show work hardening until fracture, Figure 4.3a, b, d, g. Compared with alloy Z410, the laboratory alloys ZnAl4Cu1Mg0.04,

ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31 are brittle and show much lower 휎0.2 at RT, as shown in Figure 4.3a, g. This is attributed to their much coarser microstructures caused by lower cooling rates in gravity casting. This is consistent with observations by Krupinska et al. [9] who observed microstructural refinement with increasing cooling rate causing increased hardness in a ZnAl4Cu1 alloy. With increasing temperature and / or decreasing strain rate, the strength of the alloys decreases below 200 MPa and their fracture elongation gradually increases to more than 25%. At 85°C, all alloys exhibit good ductility with 휀푓 over 20%. Similarly, Uesugi et al. [96] and Prasad et al. [190] reported that the yield stress of binary Zn-22Al (wt.%) and Zn- 27.5Al (wt.%) hypereutectic alloys decreases with increasing temperature or decreasing strain rate. Furthermore, at elevated temperature and / or low strain rate the flow curves exhibit a reduction of the true stress after reaching the maximum stress (Figure 4.3f, h, i), i.e. work softening. In comparison to alloy Z410, the laboratory alloys ZnAl4Cu1Mg0.04,

46 4. Macro mechanical response and mechanisms

ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31 show a higher strength at 85°C, as shown in Figure 4.3c, i.

Figure 4.3. True stress – true strain curves of as-cast laboratory alloys ZnAl4Cu1Mg0.04, ZnAl4Cu1Mg0.21, ZnAl4Cu1Mg0.31 and as-cast Z410 alloy (tested at lowest/highest rates and temperatures only). The legend given in (a) is valid for all diagrams, while testing conditions (strain rate, temperature) are given in the figures.

The effect of Mg addition on the mechanical properties of the alloys ZnAl4Cu1Mg0.04, ZnAl4Cu1Mg0.21, ZnAl4Cu1Mg0.31 is summarized in Table 4.2. Generally, the yield strength is increased with Mg addition. This is consistent with observations of da Costa et al. [25] who showed that Mg alloying can increase the hardness of a Zn-4Al-3Cu alloy. Further, among these three laboratory alloys, the alloy ZnAl4Cu1Mg0.21 with a medium Mg concentration of 0.21 wt.% exhibits the highest yield strength for all investigated temperatures and strain rates. This agrees well with the observation that alloy ZnAl4Cu1Mg0.21 possesses the finest eutectoid lamellar spacing. Furthermore, at elevated temperature and / or low strain rate an increase of the ductility concomitant with a reduction of the yield stress and a pronounced work

47 4. Macro mechanical response and mechanisms softening is observed. A possible explanation for the observed pronounced softening is the occurrence of grain boundary plasticity presumably in the fine eutectic and eutectoid structures, which has been associated with this type of true stress-strain curve [109, 159, 163].

Table 4.2. Mechanical properties of as-cast ZnAl4Cu1 alloys; 휎0.2 yield strength, 휀푓 elongation to fracture. ZnAl4Cu1 ZnAl4Cu1 ZnAl4Cu1 Alloy Z410 Mg0.04 Mg0.21 Mg0.31 σ , σ , σ , σ , 휀̇, s-1 T, °C 0.2 휀 , % 0.2 휀 , % 0.2 휀 , % 0.2 휀 , % MPa 푓 MPa 푓 MPa 푓 MPa 푓

RT 153.5 0.9 181.9 0.4 156.9 0.6 232.3 4.7

5∙10-4 55 144.1 3.0 162.9 2.1 146.1 2.3 - -

85 126.8 21.8 154.4 26.4 132.4 27.5 119.3 25.1

RT 154.3 0.8 171.0 0.6 151.3 0.7 - -

5∙10-5 55 138.9 4.8 158.1 7.1 140.1 17.1 - -

85 116.5 26.0 127.6 26.7 110.1 26.6 - -

RT 141 0.9 161.9 0.7 152 1.4 202.7 13.1

6∙10-6 55 126.9 17.5 133.9 22.5 119.1 27.3 - -

85 103.2 19.6 102.1 26.0 89.6 30.8 86.6 41.2

4.1.3. Fracture surfaces

Figure 4.4 shows the fracture surfaces of samples deformed at RT, where the alloys show very limited elongation to fracture (less than 1.5%). The fracture surfaces in these micrographs exhibit large and smooth cleavage facets, which is a typical feature of intragranular brittle fracture. Intragranular cleavage fracture have also been observed in Zn alloys deformed at temperatures between 77K and RT [92] and at RT at a strain rate of 10-2 s-1 [103]. The cleavage plane extends not only throughout the primary η-Zn grains, but also into the eutectic / eutectoid regions (e.g. Figure 4.4b). While most fracture surfaces after deformation at RT are dominated by cleavage facets, some small dimples are also observed, indicating locally ductile failure, probably in the eutectic and eutectoid regions.

48 4. Macro mechanical response and mechanisms

Figure 4.4. Fracture surfaces of ZnAl4Cu1 alloys deformed at room temperature; (a) alloy ZnAl4Cu1Mg0.04 at strain rate of 5∙10-5 s-1, (b) alloy ZnAl4Cu1Mg0.21 at strain rate of 5∙10-4 s-1, (c) alloy ZnAl4Cu1Mg0.31 at strain rate of 6∙10-6 s-1.

Figure 4.5 and Figure 4.6 show the fracture surfaces of alloys deformed at elevated temperatures. Here, the alloys ZnAl4Cu1Mg0.04 and ZnAl4Cu1Mg0.31 are given as representative for all alloys, since almost identical fracture surfaces are observed for alloys ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31. At 55°C and a strain rate of 5∙10−4 s-1 (Figure 4.5a and Figure 4.6a) the fracture surfaces are mainly comprised of cleavage facets as similarly observed after room temperature deformation. With decreasing strain rate the area fraction of cleavage facets decreases while that of dimples increases. At 55°C and a strain rate of 5∙10−5 s-1 cleavage is still the dominant feature of fracture in alloy ZnAl4Cu1Mg0.04 (Figure 4.5b), while the fracture surface of alloy ZnAl4Cu1Mg0.31 (Figure 4.6b) is mainly comprised of dimples, which probably nucleated from microvoid coalescence in regions of high local plastic deformation. At 55°C and a strain rate of 6∙10−6 s-1, cleavage planes are still visible, but dimples and tearing ridges are the main features in the case of alloy ZnAl4Cu1Mg0.21 (Figure 4.5c), while the fracture surface of alloy ZnAl4Cu1Mg0.31 (Figure 4.6c) shows only dimples. At 85°C (Figure 4.5d, e and Figure 4.6d, e), where the alloys underwent larger plastic deformation with elongations to fracture up to 30%, the fracture surfaces of both alloys show only dimples indicating ductile fracture. Hughes et al. [92] and Abou El-khair et al. [35] have observed a similar transition of brittle to ductile fracture with increasing temperature for pure Zn [92] and Zn-12Al [35] which has been assigned to increased activity of dislocation slip with increasing temperature in Zn by Hughes et al. [92].

49 4. Macro mechanical response and mechanisms

Figure 4.5. Fracture surfaces of alloy ZnAl4Cu1Mg0.04 deformed at elevated temperatures; (a) at 55°C and strain rate of 5∙10-4 s-1, (b) at 55°C and strain rate of 5∙10-5 s-1, (c) at 55°C and strain rate of 6∙10-6 s-1, (d) at 85°C and strain rate of 5∙10-4 s-1, (e) at 85°C and strain rate -5 -1 -6 -1 of 5∙10 s , (f) at 85°C and strain rate of 6∙10 s .

Figure 4.6. Fracture surfaces of alloy ZnAl4Cu1Mg0.31 deformed at elevated temperatures; (a) at 55°C and strain rate of 5∙10-4 s-1, (b) at 55°C and strain rate of 5∙10-5 s-1, (c) at 55°C and strain rate of 6∙10-6 s-1, (d) at 85°C and strain rate of 5∙10-4 s-1, (e) at 85°C and strain rate of -5 -1 -6 -1 5∙10 s , (f) at 85°C and strain rate of 6∙10 s .

50 4. Macro mechanical response and mechanisms

The relative fractions of brittle and ductile fracture were estimated by measuring the area fraction of cleavage planes and dimples from the projected images, see Table 4.3. With increasing temperature and / or decreasing strain rate the fraction of ductile fracture increases, here the alloys with higher Mg content showed better ductility and a higher fraction of ductile fracture. This is indeed consistent with all those sets of stress-strain curves exhibiting stable plastic deformation to very high strains, i.e. at elevated temperatures and / or low strain rates. Where catastrophic brittle fracture occurs early on, the elongation of fracture scatters between the three alloys as might be expected where failure depends on the largest fault to develop rather than stable flow.

Table 4.3. The area fraction (%) of ductile fracture under tensile deformation.

T, °C 휀̇, s-1 ZnAl4Cu1Mg0.04 ZnAl4Cu1Mg0.21 ZnAl4Cu1Mg0.31

5∙10-4 2.5 ± 0.1 6.1 ± 3.4 8.3 ± 5.9

RT 5∙10-5 3.9 ± 0.6 6.2 ± 1.1 14.8 ± 1.8

6∙10-6 3.9 ± 0.4 6.6 ± 2.2 23.8 ± 5.6

5∙10-4 7.8 ± 0.3 23.7 ± 6.4 35.7 ± 4.2

55 5∙10-5 25.7 ± 3.5 49.4 ± 2.8 60.0 ± 6.8

6∙10-6 93.4 ± 3.1 100 100

5∙10-4 100 100 100

85 5∙10-5 100 100 100

6∙10-6 100 100 100

4.1.4. Deformation microstructure

4.1.4.1. Room temperature and/or highest strain rate tested (5∙10−4 s-1)

Figure 4.7 shows the deformation microstructures of the alloys ZnAl4Cu1Mg0.04, ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31 after tensile deformation at room temperature and/or highest strain rate tested (5∙10−4 s-1). The deformation microstructures of all alloys are nearly identical to the microstructure of the as-cast condition, i.e. the grains maintain their initial morphologies. However, deformation twins are observed inside the primary η-Zn phase, as marked in the micrographs, and cracks can be found close to the fracture surfaces. Most of the cracks are observed to form and propagate along a straight path inside primary η-Zn grains, while some cracks are found to spread into the η-Zn + α-Al eutectic and eutectoid regions, as shown in Figure 4.7b and g. Moreover, it is observed that when a crack approaches Mg2Zn11

51 4. Macro mechanical response and mechanisms

precipitates, the crack changes its propagation direction and propagates along the Mg2Zn11 phase boundary, Figure 4.7e.

Figure 4.7. Microstructure of ZnAl4Cu1 alloys after deformation at low temperature and/or a strain rate of 5∙10−4 s-1. (a), (d), (g) alloy ZnAl4Cu1Mg0.04; (b), (e), (h) alloy ZnAl4Cu1Mg0.21; (c), (f), (i) alloy ZnAl4Cu1Mg0.31; (a-c) alloys at RT and strain rate of 5∙10-4 s-1; (d-f) alloys at RT and strain rate of 6∙10-6 s-1; (g-i) alloys at 55°C and strain rate of 5∙10-4 s-1; PB: phase boundary.

Figure 4.8a, b shows SEM images and inverse pole figure (IPF) maps of alloy ZnAl4Cu1Mg0.31 after tensile deformation at RT and a strain rate of 6∙10-6 s-1 with twin boundaries highlighted in red. The observed twins are indexed as {101̅2}[101̅1̅] compression

52 4. Macro mechanical response and mechanisms twins which is consistent with earlier studies that have reported {101̅2}[101̅1̅] as the dominant twinning system in Zn [2, 92].

Figure 4.8. (a), (b) SEM and EBSD images of deformation twins in alloy ZnAl4Cu1Mg0.31 after deformation at RT and strain rate of 6∙10-6 s-1, (c) SEM and EBSD images of cracks formed close to the fracture surface in alloy ZnAl4Cu1Mg0.31 after deformation at RT and strain rate of 6∙10-6 s-1.

The twins propagate not only through the primary η-Zn phase grains, but also penetrate into and cross through the eutectic and eutectoid regions. Easy twin propagation is assumed to be promoted by the orientation relationship of the primary η-Zn phase and the eutectic / eutectoid

53 4. Macro mechanical response and mechanisms

η-Zn + α-Al structures where {112̅0} η-Zn prim // {112̅0} η-Zn eut and {112̅0} η-Zn eut // {110} α-Al eut, since these interfaces do not act as strong obstacles for twin propagation.

Figure 4.8c shows an SE micrograph and IPF map of the sample area close to the fracture surface in alloy ZnAl4Cu1Mg0.31 after tensile deformation at RT and a strain rate of 6∙10-6s-1. Large cracks are propagating through primary η-Zn grains. EBSD-trace analysis revealed that all cracks formed along the basal plane trace, which agrees well with the observation of Hughes et al. and Curry et al. [92, 191]. This finding indicates localised planar basal dislocation slip at room temperature and/or a strain rate of 5∙10−4 s-1 being consistent with the observations of Hughes et al. [92] and Liu et al. [103]. Li et al. [91] concluded that the main deformation and damage mechanisms of commercially pure Zn in compression are twinning, dislocation slip, and that cracking occurs at grain boundaries and deformation twin boundaries. It is found that these cleavage cracks also penetrate through the eutectic and eutectoid structures adjacent to primary η-Zn grains (Figure 4.7d, e, g, Figure 4.8c), eventually causing brittle failure of the specimen.

Figure 4.9. Microstructure evolution of alloy ZnAl4Cu1Mg0.31 during in-situ straining at RT in the SEM; a) initial microstructure, b) displacement of 34 µm, c) displacement of 36 µm, d) displacement of 62 µm, e) displacement of 63 µm, (f) failure at a displacement of 64 µm; TD: tensile direction.

54 4. Macro mechanical response and mechanisms

To further analyse the deformation and fracture behaviour, in-situ straining experiments were carried out at room temperature in the SEM for alloy ZnAl4Cu1Mg0.31. Figure 4.9 shows selected in-situ test images of alloy ZnAl4Cu1Mg0.31 tested at RT. The initial microstructure of the sample is shown in Figure 4.9a. After 34 µm displacement (Figure 4.9b) deformation twins are observed to form inside the primary η-Zn phase (yellow arrows in Figure 4.9). With further displacement the twins propagate and grow in thickness and more twins are generated, as shown in Figure 4.9c and d. After displacement of 62 µm a cleavage crack nucleates inside a primary η-Zn grain (red arrow in Figure 4.9d). After displacement of 62 µm a cleavage crack nucleates inside a primary η-Zn grain (red arrow in Figure 4.9d). First, the crack grows and propagates slowly along its cleavage plane (Figure 4.9e), after reaching the critical size very sudden crack propagation and failure is observed (Figure 4.9f). These images also show that the twins are only observed in a relatively localized area close to the fracture surface. In contrast to Liu et al. [103] who observed cracking also along twin boundaries and Curry et al. [191] who assumed cracking to occur also along twin boundaries, we do not observe fracture in conjunction with deformation twins in the current study. In how far twins and cracks interact or cause the formation of the other might be investigated further in the future based on small- scale in-situ experiments in conjunction with 3D metallographic and focussed ion beam techniques.

These results confirm the dominant activity of dislocation-mediated deformation mechanisms, namely basal slip and {101̅2}[101̅1̅] twinning, during deformation at room temperature and/or a strain rate of 5∙10−4 s-1 in the η-Zn phase. Since basal dislocation slip and {101̅2}[101̅1̅] twinning alone do not offer enough independent slip and twinning systems to fulfill the von Mises criterion [2, 3], the observed limited plastic deformation at room temperature and / or highest strain rate tested is probably caused by the limitation of active deformation systems.

4.1.4.2. Elevated temperature and/or low strain rate

After tensile deformation at elevated temperatures and/or low strain rates (Figure 4.10), the microstructures of the deformed samples show a significant change when compared to their as-cast conditions (Figure 4.1). More specifically, the primary η-Zn phase grains and the eutectic and eutectoid structures are elongated along the tensile direction. Tearing cracks instead of cleavage cracks are visible near the fracture surface, as shown in Figure 4.10a, g and h.

55 4. Macro mechanical response and mechanisms

Figure 4.10. Microstructure of ZnAl4Cu1 alloys after deformation at elevated temperature and/or low strain rate; (a), (d), (g) alloy ZnAl4Cu1Mg0.04; (b), (e), (h) alloy ZnAl4Cu1Mg0.21; (c), (f), (i) alloy ZnAl4Cu1Mg0.31; (a-c) alloys at 55°C and strain rate of 6∙10-6 s-1; (d-f) alloys at 85°C and strain rate of 5∙10-5 s-1; (g-i) alloys at 85°C and strain rate of 6∙10-6 s-1; TD: tensile direction.

The coarse primary η-Zn grains show a large degree of plastic deformation, indicating the activation of non-basal dislocation motion by thermal activation. Since the activity of non-basal dislocation systems was reported before for pure Zn [2, 83, 92, 103, 104], it is assumed that also in the investigated Z410 alloys non-basal slip systems are activated at elevated temperatures and, thus, enable more compatible plastic deformation of primary η-Zn grains. This is in agreement to the observations of Hughes et al. [92] who observed decreased deformation twinning but increased dislocation slip with increasing temperature. Furthermore, both the flow curves exhibiting strain softening, and the deformation microstructures indicate the activity of grain boundary sliding (GBS) in the eutectic and eutectoid structures at elevated temperatures and / or low strain rates. Grain boundary sliding and superplasticity are well- known to occur in hypereutectic Zn-22Al alloys (e.g [6, 96, 97, 105-111, 113-118]) and has

56 4. Macro mechanical response and mechanisms also been reported to occur in severe plastically deformed hypoeutectic Zn-0.3Al and Zn-5Al alloys [159, 163] at RT. Zhang et al. [94] assumed also that dynamic recovery or recrystallisation might contribute to work softening in ultra-fine grained (UFG) Zn during deformation at a strain rate of 4∙10-3 s-1 in the temperature range of 20 to 60°C. They [94] further assumed plastic deformation to occur by dislocation creep in larger grains and grain boundary sliding in smaller grains in UFG Zn.

Figure 4.11. Microstructure evolution of alloy ZnAl4Cu1Mg0.31 during in-situ straining at 85°C in the SEM; (a) initial microstructure, (b) displacement of 52 µm, (c) displacement of 67 µm, (d) displacement of 110 µm, (e) displacement of 126 µm, (f) displacement of 158 µm, (g) void formation at a displacement of 219 µm, (h) void coalescence at a displacement of 241 µm, (i) failure at a displacement of 259 µm. The red arrows in the enlarged area in the insets show the occurrence of grain boundary sliding in the eutectoid structures. TD: tensile direction.

At elevated temperatures and / or low strain rate, the alloys show larger elongation to fracture and ductile failure, as shown in Figure 4.5c to f and Figure 4.6c to f.

57 4. Macro mechanical response and mechanisms

To further investigate the activation of the deformation mechanisms discussed above, in-situ straining experiments in the SEM were also performed for alloy ZnAl4Cu1Mg0.31 at 85°C, Figure 4.11. The elongation to fracture in these in-situ experiments was more than 4 times higher at 85°C than at RT, which is in excellent agreement with the ex-situ tensile tests (Figure 4.3).

Figure 4.11a shows the initial microstructure of the sample. After a displacement of 52 µm deformation twins are observed to form in primary η-Zn phase grains (yellow arrows in Figure 4.11). With increasing strain the deformation twins propagate and grow in thickness and first slip traces appear at the sample surface in the primary η-Zn phase (white arrows in Figure 4.11). With further straining the density and height of the slip traces in the primary η-Zn phase increases. From Figure 4.11 it is evident that the occurrence of the slip traces is not straight but wavy, indicating again the activity and interaction of more slip systems than only the basal slip system. The enlarged micrographs in the insets in Figure 4.11 show the activity of grain boundary sliding of the eutectoid structures; the arrows point to morphology changes and grain boundary displacements of the eutectoid structures. After a displacement of 219 µm, the onset of cracking by void coalescence is observed on the surface (red arrow in Figure 4.11g). More cracks are generated upon further straining (Figure 4.11h) until these cracks finally converge causing failure (Figure 4.11i).

Based on our experimental results we conclude that the higher elongation to fracture observed at elevated temperatures and / or low strain rates is caused by (i) a higher number of available deformation systems of the η-Zn phase at elevated temperatures and (ii) grain boundary sliding of the eutectic and eutectoid structures. Fracture is ductile and occurs due to void formation and coalescence.

Our observations show that dilute Mg alloying has a beneficial effect on both strength and ductility of the investigated gravity-cast Zn alloys at elevated temperatures and / or low strain rates. One of the underlying mechanisms is microstructural refinement of the eutectoid lamellar structure due to retarded diffusion through Mg alloying. Further experiments at higher cooling rates are required to confirm that dilute Mg is also beneficial in die cast Zn alloys. Furthermore, an increased activity of grain boundary sliding at elevated temperatures and / or low strain rates was observed with increasing Mg concentration. Hence, it is assumed that Mg decoration of the grain and phase boundaries in the eutectic and eutectoid structures might promote grain boundary sliding maybe by decreasing the grain boundary cohesion. However, this point is still speculative and requires further investigations in the future.

In summary, our study reveals that at slightly elevated temperatures and even at room temperature and low strain rates a mixture of classical low-temperature dislocation-mediated deformation mechanisms, i.e. dislocation glide and twinning, but also high-temperature

58 4. Macro mechanical response and mechanisms diffusion-driven deformation mechanisms, i.e. grain boundary sliding, can be activated. We showed that the activity and impact of the activity of these deformation systems does not only depend on temperature and strain rate but can also be effected by dilute alloying. Specifically, we observed that dilute Mg alloying essentially affects (i) the microstructure formation by refinement of the eutectic and eutectoid structures enhancing the material strength and (ii) increasing the activity of grain boundary sliding improving the ductility of Zn alloys. This brings increased strength and ductility at elevated temperatures with respect to the commercial alloy Z410. These findings are believed to contribute to a better understanding of the mechanisms causing the long-term mechanical instability of commercial Zn alloys and might open a direction for the design of Zn alloys with enhanced long-term mechanical stability.

4.2. Conclusions

The following conclusions are drawn from our experimental results on the mechanical properties and microstructures of Zn4Al1Cu alloys with different Mg concentrations:

1. Die casting (Z410) leads to a much finer microstructure and thus a higher strength and better ductility at room temperature compared to gravity casting (ZnAl4Cu1Mg0.04).

2. Dilute Mg addition leads to a refinement of the eutectoid structures and promotes the formation of Mg2Zn11 precipitates.

3. Dilute Mg alloying increases the strength and the ductility of Zn4Al1Cu alloys, particularly at elevated temperatures. It is assumed that the improved strength and ductility is caused by microstructural refinement of the eutectoid structures.

4. The investigated Zn alloys reveal two distinct deformation regimes depending on temperature and strain rate:

(a) At low temperature and/or highest strain rates tested (5∙10−4 s-1), the alloys are brittle and show work hardening. The dominant deformation mechanisms at room temperature and / or highest strain rate tested (5∙10−4 s-1) are basal slip and {101̅2}[101̅1̅] twinning in the primary η-Zn phase.

(b) At elevated temperature and / or low strain rate, the alloys are ductile and show pronounced working softening. The deformation mechanisms of the investigated zinc alloys are a mixture of twinning, dislocation motion, including non-basal dislocation slip within the primary η-Zn grains, and grain boundary sliding in the eutectic and eutectoid structures.

59 5. Creep of ZnAl4Cu1 alloys

5. Creep of ZnAl4Cu1 alloys6

Zn-Al based alloys are sensitive to creep due to their relatively low melting point (around 400°C). However, various creep mechanisms have been reported in Zn-Al based alloy and the reported results in literature are contradicting even for the same material tested at similar conditions, Table 2.4. Further, in order to increase the creep resistance of Zn-Al alloys, tailored alloys and microstructures should be used depending on the predominant creep mechanisms. It is therefore essential to fully understand the underlying creep mechanism of Zn-Al based alloys. Further, Chapter 4 has shown that Mg is effective in increasing the strength and ductility of Zn-Al alloys through the modification of eutectoid structures and the formation of Zn-Mg intermetallic compounds. However, the influence of Mg on the creep behaviour of Zn-Al-Cu alloys has not yet been comprehensively studied. Therefore, it is of great impact to explore the influence of Mg on the creep resistance of Zn-Al alloys. To this end, this chapter presents the creep properties and creep mechanisms alloys with and without dilute Mg additions. As the role of the individual microstructural components in Zn-Al alloys during creep still remains unknown, this chapter further explores the local creep properties of the individual microstructural constituents and their contributions to the global creep properties of bulk alloys.

5.1. Results

5.1.1. Uniaxial tensile creep experiments

Figure 5.1 shows the uniaxial tensile creep curves, namely the true strain vs. time curves, of the investigated ZnAl4Cu1Mg alloys obtained at 55°C, 85°C and 105°C, at stresses of 0.6 –

0.8∙σ0.2. In this temperature and stress range, all samples reveal typical creep curves with primary and stationary creep stages. During the primary creep stage, the creep strain increases rapidly while the strain rate decreases upon loading. At the end of the primary creep stage, the creep strain increases almost linearly with time and the creep rate reaches a minimum, indicative of the second or stationary creep stage. Depending on the testing temperature, the samples enter the stationary creep stage at a creep strain ε of 0.4% to 1.4%. At the same temperature, the primary stage appears shorter at higher stress (e.g. Figure 5.1a), which is consistent with the observations of Arieli et al. [138] in a Zn-22Al (wt.%) alloy. For some conditions (at high temperature and high stress level), the creep curves reach the tertiary stage within the testing time (e.g. Figure 5.1e), where the creep rate increases again rapidly until fracture.

6 This part was published as an article. The original citation is: Z. Wu, S. Sandlöbes, Y. Wang, J. S. K.- L. Gibson and S. Korte-Kerzel (2018). "Creep behaviour of eutectic Zn-Al-Cu-Mg alloys." Materials Science and Engineering: A 724: 80-94.

60 5. Creep of ZnAl4Cu1 alloys

Figure 5.1. Creep curves of ZnAl4Cu1MgX (X = 0.04, 0.21, 0.31, in wt.%) alloys at 55°C,

85°C and 105°C and stresses of 0.6 – 0.8∙σ0.2.

Table 5.1 shows the time required to reach 1% creep strain for the investigated ZnAl4Cu1Mg alloys at different experimental conditions. Generally, the time to 1% strain decreases with increasing stress and / or increasing temperature. It is noticed that the time to reach 1% creep strain at 105°C is longer than at 85°C, which is assumed to be due to the higher normalized stress σ/G at 85°C, Table 3.2. Additionally, under the same temperature and stress level, the time to reach 1% creep strain decreases with increasing Mg concentration in the alloy.

The creep rates of the secondary creep stage, i.e. the minimum or stationary creep rate, were determined by linear fitting, Table 5.2. In all alloys the minimum creep rate increases with increasing stress and temperature. The alloys with higher Mg content generally show higher creep rates under the same experimental conditions.

61 5. Creep of ZnAl4Cu1 alloys

Table 5.1. Time (in hours) to 1% creep strain of ZnAl4Cu1MgX (X = 0.04, 0.21, 0.31, in wt.%) alloys at 55°C, 85°C and 105°C and stresses of 0.6 – 0.8∙σ0.2.

Temperature, °C σ/σ0.2 ZnAl4Cu1Mg0.04 ZnAl4Cu1Mg0.21 ZnAl4Cu1Mg0.31

0.6 not reached 12.33 2.4

55 0.7 19.93 10.04 1.73

0.8 9.25 2.04 1.55

0.6 1.88 2.42 1.04

85 0.7 4.29 0.87 0.3

0.8 0.83 0.21 0.15

0.6 9.14 2.31 2.48

105 0.7 3.02 1.29 0.98

0.8 1.48 0.65 0.53

Table 5.2. The minimum creep rate 휀̇ (10-4h-1) of ZnAl4Cu1MgX (X = 0.04, 0.21, 0.31, in wt.%) alloys at 55 – 105°C and stresses of 0.6 – 0.8∙σ0.2.

Temperature, °C σ/σ0.2 ZnAl4Cu1Mg0.04 ZnAl4Cu1Mg0.21 ZnAl4Cu1Mg0.31

0.6 0.4 1.9 2.0

55 0.7 1.5 4.1 4.5

0.8 3.6 18.6 15.0

0.6 5.7 25.2 20.8

85 0.7 19.7 54.2 64.6

0.8 43.6 297.0 170.0

0.6 3.5 12.3 17.5

105 0.7 10.7 31.8 46.7

0.8 28.3 92.2 98.2

According to the Mukherjee-Bird-Dorn (MBD) equation [124, 125], the stationary creep rate, ε̇, and the applied stress, σ, can be related through an Arrhenius-type equation:

62 5. Creep of ZnAl4Cu1 alloys

퐺푏 푄 휎 푛 휀̇ = ⁡퐴 퐷 푒푥푝(− ) ( ) (5.1) 푘푇 0 푅푇 퐺 where A is a material constant, G the shear modulus, b the Burgers vector, T the temperature in Kelvin, k the Boltzmann constant, D0 the pre-exponential factor for self-diffusion, Q the creep activation energy, R the universal gas constant and n the stress exponent.

The shear moduli of the alloys were estimated using the rule of mixtures [138, 192]:

퐺퐴푙푙표푦 = 푓푍푛퐺푍푛 + 푓퐴푙퐺퐴푙 + 푓퐶푢퐺퐶푢 + 푓푀푔퐺푀푔 (5.2) where 푓푍푛, 푓퐴푙, 푓퐶푢 and 푓푀푔 are the atomic fractions of the corresponding elements in the alloy. The pure metal shear moduli at each temperature were estimated according to literature data [193-197]:

푇(퐾) − 300 퐺 (퐺푃푎) = 49.3 × (1 − 0.50 × ) (5.3) 푍푛 693

푇(퐾) − 300 (5.4) 퐺 (퐺푃푎) = 25.4 × (1 − 0.50 × ) 퐴푙 933

푇(퐾) − 300 (5.5) 퐺 (퐺푃푎) = 16.6 × (1 − 0.49 × ) 푀푔 924

푇(퐾) − 300 (5.6) 퐺 (퐺푃푎) = 42.1 × (1 − 0.50 × ) 퐶푢 1356

The calculated shear moduli, G, of the ZnAl4Cu1Mg alloys investigated at different temperatures are listed in Table 5.3.

Table 5.3. Calculated shear moduli G of ZnAl4Cu1MgX (X = 0.04, 0.21, 0.31, in wt.%) alloys at different temperatures, [GPa].

T [°C] Alloy 55 85 105

ZnAl4Cu1Mg0.04 45.96 44.96 44.29

ZnAl4Cu1Mg0.21 45.83 44.83 44.17

ZnAl4Cu1Mg0.31 45.75 44.75 44.09

The stress exponent, n, and the creep activation energy, Q, of the studied ZnAl4Cu1Mg alloys at 25 – 105°C and stresses of 0.6 – 0.8∙σ0.2 were quantitatively evaluated from the creep curves (Figure 5.1) and are presented in Figure 5.2. The determined stress exponents and creep activation energies of the three alloys are also listed in Table 5.4.

63 5. Creep of ZnAl4Cu1 alloys

휎 Figure 5.2. Relation between steady state creep rate 휀̇ vs. normalised stress ⁄퐺 (a, c ,e) and temperature dependence of the steady state creep rate 휀̇ (b, d, f) of ZnAl4Cu1MgX (X = 0.04, 0.21, 0.31, in wt.%) alloys. Table 5.4. Creep stress exponents and activation energies of ZnAl4Cu1MgX (X = 0.04, 0.21, 0.31, in wt.%) alloys at 25 – 105°C determined from uniaxial tensile creep tests.

Alloy n Q, kJ/mol

ZnAl4Cu1Mg0.04 7.4 ± 0.2 96 ± 3

ZnAl4Cu1Mg0.21 8.0 ± 0.9 104 ± 4

ZnAl4Cu1Mg0.31 6.9 ± 0.4 93 ± 2

64 5. Creep of ZnAl4Cu1 alloys

5.1.2. Creep microstructure

Figure 5.3 shows the microstructures of the as-cast ZnAl4Cu1Mg alloys. The microstructures are mainly comprised of coarse primary η-Zn phase dendrites with a typical dendrite size of 26 – 70 µm, η-Zn + α-Al lamellar eutectic structures and fine lamellar eutectoid structures with a lamellar spacing of 0.1 – 0.34 µm, as described in section 4.1.1. Additionally, precipitates of the intermetallic Mg2Zn11 phase with an average size of 0.4 to 0.9 µm are present in the η-Zn + α-Al eutectoid structure of ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31, see the enlarged micrograph in Figure 5.3c.

Figure 5.3. As-cast microstructure of (a) ZnAl4Cu1Mg0.04; (b) ZnAl4Cu1Mg0.21; and (c) ZnAl4Cu1Mg0.31.

Figure 5.4 shows SE micrographs of alloy ZnAl4Cu1Mg0.31 creep-deformed to a global creep strain of 3.6% (deformed at 55°C and 67 MPa; stationary creep stage, Figure 5.4.a) and 15.0 % (deformed at 85°C and 93 MPa, begin of the tertiary creep stage, Figure 5.4.b). No substantial microstructural changes are evident from the SE micrographs, when comparing the microstructures of the creep-deformed and as-cast conditions (Figure 5.3.c). To characterise

65 5. Creep of ZnAl4Cu1 alloys the in-grain misorientations in the primary Zn phase upon creep, EBSD measurements of as- cast undeformed (a, b), creep-deformed to a creep strain of 3.6% (c, d) and creep-deformed to a creep strain of 15.0% (e, f) ZnAl4Cu1Mg0.31 samples were performed, Figure 5.5. The left hand micrographs in Figure 5.5 (a, c, e) show inverse pole figure (IPF) maps and the right hand micrographs of Figure 5.5 (b, d, f) show grain reference orientation deviation (GROD) maps where the angular in-grain orientation deviations at each point within the grains are plotted. The corresponding primary Zn grains are marked by red dashed rectangles in Figure 5.4.

Figure 5.4. Post-deformation microstructures of alloy ZnAl4Cu1Mg0.31 after 3.7 % creep strain at 55°C and 102 MPa (a), and after 15.0% creep strain at 85°C and 93 MPa (b). The red boxes indicate the areas analysed by EBSD (Figure 5.5).

In the as-cast undeformed sample hardly any in-grain misorientations are present as evident from the IPF and GROD maps (Figure 5.5a, b), whereas the creep deformed samples show the formation of subgrains with orientations that deviate from the grain reference orientation. Specifically, slight in-grain misorientations, evident from faint colour differences within the grains, are visible in the inverse pole figure (IPF) maps of the creep-deformed samples, Figure 5.5c and e. Similarly, the corresponding GROD maps show sub-grains with angular orientation deviations of up to several degrees within the primary η-Zn grains, Figure 5.5d and f. These in-grain misorientations are more pronounced after a creep-deformation of 15.0%, Figure 5.5.f, than after a creep-deformation of 3.6%, Figure 5.5d. Quantitative measurement of the misorientation distributions reveals misorientations of 0.4 – 3.5° across the subgrain boundaries. In the sample subjected to 15.0% global creep strain, {101̅2} 〈1011̅̅̅̅〉⁡ twin orientations are visible in the primary η-Zn grain (appearing orange-red in the IPF map), Figure 5.5e. These twins are assumed to have formed during the primary or tertiary creep stage where work hardening is dominant [127, 128]. The twin boundaries are not sharp and the twinned areas contain subgrain structures, indicating that these twins have grown during creep due to the elevated testing temperatures.

66 5. Creep of ZnAl4Cu1 alloys

Similar subgrain structures are present in creep-deformed ZnAl4Cu1Mg0.04 and ZnAl4Cu1Mg0.21 alloys (not shown here).

Figure 5.5. Inverse pole figure (IPF) (a, c, e) and corresponding grain reference orientation deviation (GROD) (b, d, f) maps of primary η-Zn grains in alloy ZnAl4Cu1Mg0.31. (a, b) show the as-cast undeformed material, (c, d) show material creep-deformed to a creep strain of 3.6 % at 55°C and 102 MPa, and (e, f) show material creep-deformed to a creep strain of at 15.0% at 85°C and 93 MPa. The corresponding primary η-Zn grains are marked by dashed rectangles in Figure 5.4 (e, f) are given with a higher magnification due to higher local misorientation angles which are not uniquely revealed at lower magnifications.

67 5. Creep of ZnAl4Cu1 alloys

5.1.3. Creep behaviour of individual microstructural constituents

The creep properties of the individual microstructural constituents were further studied using nanoindentation. Figure 5.6 shows the individual microstructural constituents of alloy ZnAl4Cu1Mg0.31 after nanoindentation creep experiments with a peak load of 25 mN at room temperature (25°C). While the η-Zn phase and the η-Zn + α-Al eutectic and eutectoid structures were indented separately, the Mg2Zn11 phase were investigated together with the surrounding eutectoid structure due to its small size of less than 1 µm.

Figure 5.6. Micrographs of individual microstructural constituents in ZnAl4Cu1Mg0.31 after indentation with a load of 25 mN at 25°C: (a) primary Zn, (b) η-Zn + α-Al eutectic structure,

(c) η-Zn + α-Al eutectoid structure, (d) Mg2Zn11 embedded in eutectoid structure

Figure 5.7. Indentation depth increase (h) as a function of time (t) during the holding period of nanoindentation creep tests at different loads in primary η-Zn phase at (a) 25°C and (b) 85°C.

Figure 5.7 shows the indentation depth over time during the holding period at 25°C (Figure 5.7a) and at 85°C (Figure 5.7b) in the primary η-Zn phase. The displacement in this stage is caused by creep [165, 198]. The depth-time curves show two distinct regions: a region analogous to transient creep where the indentation depth increases rapidly (~ 0 – 10 s), and a region analogous to steady-state creep where the indentation depth increases almost linearly

68 5. Creep of ZnAl4Cu1 alloys with time (~ 20 – 60 s). The depth increase during the holding time is higher under higher loads and at elevated temperatures.

Figure 5.8 shows the indentation creep curves of the individual microstructural components at a load of 25 mN. It is apparent that the η-Zn phase shows the lowest increase in indentation depth among all microstructural components at RT (Figure 5.8a), while the Mg2Zn11 + eutectoid regions exhibit the highest indentation depth increase. The eutectic structures display slightly lower creep displacements than the eutectoid structures at RT (Figure 5.8a). This trend is more pronounced at 85°C (Figure 5.8b).

Figure 5.8. Indentation depth as a function of time during the holding period of the individual microstructural constituents at 25 mN at (a) 25°C and (b) 85°C.

According to Su et al. [199] as well as Phani and Oliver [200], the indentation strain rate, 휺̇ 풊, can be expressed in an equation similar to the uniaxial power law creep:

1 푑ℎ 휀̇ = ( ) ( ) = 퐵(푝 )푁 (5.7) 푖 ℎ 푑푡 푛표푚 where t is the time, h the indentation depth, which can be fitted according to ℎ(푡) − ℎ(0) =

푎푙푛(푏푡 + 1) (a and b are fitting parameters), B the indentation pre-exponential term, pnom the nominal contact pressure which is determined by the load and the projected contact area at the indentation depth and N the stress exponent during indentation [200]. In the present study the projected contact area was calibrated on fused silica at room temperature and determined to be 29.73ℎ2 + 9423.9ℎ.

The stress exponent N was calculated according to:

휕(푙표푔휀̇ ) 푁 = 푖 (5.8) 휕(푙표푔푝푛표푚)

Figure 5.9 shows the procedure of calculating the stress exponent during indentation at 25 mN. The obtained nanoindentation creep stress exponents are listed in Table 5.5.

69 5. Creep of ZnAl4Cu1 alloys

Figure 5.9. Log-log plot of indentation strain rate as a function of the nominal contact pressure, pnom, for nanoindentation creep tests of the different individual microstructural components at 25 mN load at (a) 25°C and (b) 85°C.

Table 5.5. Nanoindentation creep stress exponents of different individual microstructural components at 25°C and 85°C at maximal loads of 15 – 35 mN.

Microstructural RT 85°C

component 15 mN 25 mN 35 mN 15 mN 25 mN 35 mN

Zn 27.2 ± 4.2 22.4 ± 1.7 24.9 ± 2.5 16.6 ± 3.0 21.3 ± 2.8 21.3 ± 1.9

Eutectic 19.7 ± 3.8 20.2 ± 1.8 21.2 ± 3.1 10.0 ± 0.4 13.9 ± 1.8 14.1 ± 2.1

Eutectoid 16.7 ± 2.0 17.5 ± 1.7 17.8 ± 0.9 9.7 ± 3.2 11.9 ± 2.9 -*

Mg2Zn11 + 20.5 ± 1.9 16.2 ± 2.8 -* 8.5 ± 1.2 10.7 ± 0.3 10.3 ± 3.2 Eutectoid

* Data are excluded if fewer than 3 indents were measured for the corresponding microstructural constituents.

5.2. Discussion

5.2.1. Creep mechanisms

5.2.1.1. Macroscopic uniaxial tensile creep

The stress exponents of the ZnAl4Cu1Mg alloys investigated lie in the range 6.9 to 8, indicating a dislocation-controlled creep mechanism [201, 202]. The obtained activation energies amount to 93 to 104 kJ/mol which is close to the self-diffusion activation energy of pure zinc (92 kJ/mol) [83, 132]. This is consistent with the study by Kallien et al. [40], who investigated the creep properties of a ZnAl4Cu1 alloy (in wt.%) at 25 – 85°C and 40 – 100 MPa and obtained an

70 5. Creep of ZnAl4Cu1 alloys activation energy of 94.1 kJ/mol. Consistently, Murphy et al. [18] studied the creep behaviour of ZnAl4, ZnAl8Cu1, ZnAl27Cu2 alloys (in wt.%) at 60 – 150°C and 10 – 100 MPa and reported an activation energy of 106 kJ/mol. Similarly, Anwar et al. [17] have investigated the creep properties of ZnAl8, ZnAl12, ZnAl27 alloys (in wt.%) at 70 – 160°C and 20 – 100 MPa and determined an activation energy of 102 – 112 kJ/mol. These groups have all concluded that the predominant creep mechanism is dislocation creep in the η-Zn phase [17, 18, 40].

It has been further reported that the creep activation energy of dislocation creep in Zn is close to the self-diffusion activation energy of Zn and that the corresponding stress exponent amounts typically to n = 3 – 5 [203-207]. The stress exponents of the ZnAl4Cu1Mg alloys investigated lie in the range of 6.9 – 8. Such higher stress exponents in combination with activation energies around that for self-diffusion have been reported to be related to either dislocation climb in precipitation- or dispersion-strengthened materials in which the actual stress acting on dislocations for creep is lowered since extra stress is required due to the interaction between obstacles and dislocations [208-211], or dislocation climb at high stress level [212-215] where it has been assumed that the high applied stresses generate excess vacancies by dislocation interaction and thus enhance dislocation climb [216, 217].

The presence of ZnxAl1-x (x≥0.7) and pure Al precipitates in the primary η-Zn grains in ZnAl4Cu1Mg0.31 has been reported [148]. Further, a reduction of size and number density of these ZnxAl1-x precipitates after tensile deformation at 85°C has been observed [148], indicating the partial dissolution of the precipitation upon tensile deformation at 85°C. However, the stress exponents and activation energies obtained in the present study are consistent and constant in the temperature range 25°C to 105°C (Figure 5.2) and also the nanoindentation creep exponents of the primary η-Zn phase during indentation creep (Figure 5.9) are constant at RT and 85°C. This strongly indicates that the microstructure of the ZnAl4Cu1Mg alloys investigated remained stable during creep deformation in the investigated temperature range (25 – 105°C). It is therefore assumed that these precipitates could block dislocation movement and hereby leading to higher stress exponents in our alloys. However, further investigations are necessary to confirm to which extent these precipitates can hinder dislocation movement.

Stress exponents of 3 – 5 have been reported in alloys with similar composition such as ZnAl4, ZnAl4Cu1, ZnAl8Cu1, ZnAl27Cu2 alloys (in wt.%) for creep at the same temperature but lower stress levels (10 – 100 MPa) than in the present study [17, 18, 40]. Our creep tests were conducted at stresses of 61 – 130 MPa, which is slightly higher than those reported in literature [17, 18, 40], corresponding to intermediate to high stress creep ranges (휀̇/퐷 up to 1.57∙109, where D is the self-diffusion coefficient of Zn) [208]. It is therefore concluded that the slightly higher n values are also partially caused by the higher stress levels applied.

71 5. Creep of ZnAl4Cu1 alloys

It has been reported by Kassner et al. [218], that the dislocations form equiaxed subgrain structures in dislocation climb controlled creep processes. The formation of subgrains in primary η-Zn grains during creep (Figure 5.5) therefore further suggests that the rate controlling mechanism in the η-Zn phase of ZnAl4Cu1Mg alloys is dislocation climb. Further, although Zn- 22Al alloys (in wt.%) with fully eutectoid structures have been reported to deform by Coble creep [137, 138], the reported activation energies of 62 – 70 kJ/mol [137, 138] and stress exponents of 1 – 2.2 [137, 138] are much lower than the results of the present study. Therefore, it is assumed that the rate controlling mechanism in the ZnAl4Cu1Mg alloys investigated is stress assisted dislocation climb in the η-Zn phase between 25°C and 105°C and stresses between 61MPa and 130MPa.

5.2.1.2. Nanoindentation creep

The nanoindentation creep exponents (N) of the individual microstructural constituents amount to 16 – 22 at 25°C and 11 – 21 at 85°C. Ideally, the nanoindentation creep stress exponents should be equal to the ones obtained using uniaxial creep testing under effective steady state conditions [199, 200]. For instance, Phani and Oliver [200] have concluded that indentation creep could successfully predict uniaxial creep behavior by comparing the indentation creep data from a Berkovich tip and tensile test data at 10% uniaxial strain for pure Al. However, for ZnAl4Cu1Mg0.31 alloy, our results show that the creep exponents obtained from nanoindentation creep tests are two to three times higher than the creep exponents obtained from macroscopic uniaxial creep tests at 25°C to 105°C (n = 6.9 ± 0.4). The creep exponent obtained from constant strain rate tests data at 10% uniaxial strain at 85°C for the same ZnAl4Cu1Mg0.31 alloy (n = 7.0 ± 0.5, calculated from the data in [16]) are also much smaller than those obtained by nanoindentation. This is consistent with the study of Goodall et al. [219], who have comprehensively studied the nanoindentation creep behaviour of various metals and found that the nanoindentation creep exponents do not match with the macroscopic tensile creep exponents. In a material which undergoes work hardening, the nanoindentation hardness determined by a Berkovich indenter tip has been reported to correspond to the flow stress at 8% plastic strain rather than to the yield stress [175, 220]. Since the stresses applied in uniaxial creep experiments are lower than the yield strength, the stress level in nanoindentation creep experiments is physically much higher than that in uniaxial creep experiments. It has been reported that at high creep stresses, the stress exponent increases with increasing applied load [208, 212-215, 221]. Maier et al. [222] have suggested that indentation creep using a pyramidal indenter is causing a mixture of different creep stages at the same time due to a large amount of plastic deformation and strain gradients in the deformed volume. Similarly, Li et al. [223] have reported that indentation creep is actually dominated by dislocation glide dominated plasticity due to high local stresses.

72 5. Creep of ZnAl4Cu1 alloys

The measurement of the smaller microstructural components is further complicated by the size of the indentation stress field. Bull [224] has shown that the zone of plastic deformation is ~ 10 times the indenter penetration in diameter and the elastic stress field extends to about 100 times the indenter penetration. The η-Zn + α-Al eutectoid structure investigated in this study has a colony size of around 2.8 – 9.5 µm, while the maximum indenter penetration is around 0.9 µm at 25°C and 1.2 µm at 85°C. This indicates that not only the eutectic/eutectoid structures, but also the η-Zn phase underneath the eutectic/eutectoid structures contributes to deformation during indentation, which might lead to higher stress exponents in the eutectic / eutectoid structures. We also observed extensive pile-ups around the indent impressions. Although the pile-ups affect the contact area and thus change the apparent elastic modulus and hardness [225], the hardness gradient over the strain rate is not obviously affected in our case. This is readily seen in Figure 5.9 where the gradients of log-log plots of indentation strain rate over contact pressure remain constant and the linear fitting over entire plots show coefficients of determination (R2) > 0.995. Therefore, these pile-ups are not considered in the analysis of the indentation stress exponent.

Therefore, in our case direct comparison of the n values from macro-scale testing and N values from nanoindentation is not feasible. However, nanoindentation can be used to compare the relative rates of creep in individual microstructural constituents under similar conditions, thereby giving insight into the phases and mechanisms which are likely creep-rate controlling in the material.

The η-Zn phase has a higher resistance to creep than the eutectic and eutectoid structures, while the Mg2Zn11 + eutectoid regions show the highest creep strain compared to other microstructural constituents in the ZnAl4Cu1Mg0.31 alloy. This is concluded by comparing the indentation depth of individual microstructural constituents at the same load (Figure 5.8). The increase of indentation depth at the same load of individual microstructural constituents during

nanoindentation creep has the order: hMg2Zn11 + Eutectoid˃ hEutectoid˃⁡hEutectic ˃ hZn at both 25°C and 85°C. Higher creep displacements in the η-Zn + α-Al eutectoid structures and Mg2Zn11 + eutectoid regions than in the η-Zn phase are assumed to be caused by faster lattice diffusion of Al and Mg atoms than Zn atoms at 25°C and 85°C, see Table 5.6 and Table 5.7, as well as accelerated diffusion along interfaces due to a high fraction of grain / phase boundaries in the eutectoid structures. The eutectic structures show slightly lower indentation creep displacements than the eutectoid structures.

73 5. Creep of ZnAl4Cu1 alloys

Table 5.6. Lattice diffusion coefficient of various elements in Zn-Al-Cu-Mg alloys [83]. The diffusion coefficients at 25°C and 85°C were calculated from the data in [83].

Lattice diffusion

Diffusion coefficient, 퐷 Element 푣 Pre-exponential Activation energy, (m2/s) 2 factor, 퐷0푣 (m /s) 푄푣 (kJ/mol) 25°C 85°C

Zn 1.30∙10-05 91.7 1.25∙10-05 1.26∙10-05

Al 1.70∙10-04 142.0 1.61∙10-04 1.62∙10-04

Cu 2.00∙10-05 197.0 1.85∙10-05 1.87∙10-05

Mg 1.00∙10-04 135.0 9.47∙10-05 9.56∙10-05

Table 5.7. Boundary diffusion coefficient of various elements in Zn-Al-Cu-Mg alloys [83]. The diffusion coefficients at 25°C and 85°C were calculated from the data in [83].

Boundary diffusion

Diffusion coefficient, 훿퐷 Element 푏 Pre-exponential Activation energy, (m3/s) 3 factor, 훿퐷0푏 (m /s) 푄푏 (kJ/mol) 25°C 85°C

Zn 1.30∙10-14 60.5 1.27∙10-14 1.27∙10-14

Al 5.00∙10-14 84.0 4.83∙10-14 4.86∙10-14

Cu 5.00∙10-15 104.0 4.79∙10-15 4.83∙10-15

Mg 5.00∙10-12 92.0 4.82∙10-12 4.85∙10-12

When comparing indents at 25°C and 85°C at a constant load (25 mN), both eutectic and eutectoid structures already display a mean increase of around 120 nm in indentation depth (from 97 ± 8 nm to 220 ± 6 nm for eutectic structures and from 108 ± 6 nm to 230 ± 12 nm for eutectoid structures). The Mg2Zn11 + eutectoid regions show a mean increase of 155 nm in indentation depth (109 ± 7 nm to 264 ± 15 nm), while the η-Zn phase only shows a depth increase of 51 nm (93 ± 7 nm to 145 ± 10 nm). This strongly indicates that the eutectic/eutectoid structures and the Mg2Zn11 phase show higher thermal activation than the η-Zn phase, i.e. their hardness is more sensitive to a change in temperature.

74 5. Creep of ZnAl4Cu1 alloys

The nanoindentation stress exponent of the η-Zn phase is higher than the other microstructural components and that of Mg2Zn11 + eutectoid is lowest. It is noticed that at 25°C and a load of

15 mN, the stress exponent of Mg2Zn11 + eutectoid is slightly off this trend, which is presumably due to a large contribution of η-Zn phase underneath the indent. Specifically, the nanoindentation stress exponent of the η-Zn phase at 25°C amounts to 22.4 – 27.2, which is around 20% higher than that of the eutectic / eutectoid structures and Mg2Zn11 + eutectoid. At 85°C, the nanoindentation stress exponent of the η-Zn phase (16.6 – 21.3) is nearly two times higher than that of the other microstructural constituents. While the nanoindentation creep stress exponent of the η-Zn phase is less influenced by temperature in the temperature range investigated, there is a clear decrease of the nanoindentation stress exponents of the eutectic

/ eutectoid structures and the Mg2Zn11 + eutectoid with increasing temperature. This again indicates that deformation in the eutectic/eutectoid structures and the Mg2Zn11 phase is thermally activated to a greater extent in the temperature range 25°C to 85°C. It has been reported that η-Zn + α-Al eutectic / eutectoid structures deform plastically predominantly by grain / phase boundary sliding [6, 16, 96, 97, 105-109, 113-115, 117, 118, 158], while the η- Zn phase deforms predominantly by deformation twinning and dislocation glide [2, 3, 16, 90, 91, 148, 226]. When assuming that similar mechanisms are activated during creep, it is reasonable to assume that the predominant creep mechanism in the η-Zn phase is dislocation controlled while those in eutectic / eutectoid structures are interface related mechanisms, i.e. Coble creep or grain / phase boundary sliding. Indeed, it has been reported that creep in Zn is mainly controlled by the climb of edge dislocations [83, 129-131, 133]. Further, Coble creep has been observed to be the predominant creep mechanism at low stress levels in Zn-22Al (wt.%) with fully eutectoid structures [137, 138]. This highly indicates that the creep behaviour in η-Zn + α-Al eutectic / eutectoid structures might be interface controlled.

5.2.2. Linking micro- and macroscale deformation: geometrical constraints model

In summary, although η-Zn + α-Al eutectic / eutectoid structures and Mg2Zn11 + eutectoid exhibit higher creep displacements than the η-Zn phase in nanoindentation creep tests, the apparent stress exponents and activation energies calculated from macroscopic tensile creep tests indicate that the creep deformation of ZnAl4Cu1Mg alloys is controlled by the η-Zn phase. This is assumed to be caused by geometrical constraints of the η-Zn + α-Al eutectic and eutectoid lamellae colonies during deformation. In large lamellae colonies (Figure 5.10b), straining is restricted to the interface planes. In alloys with large eutectic / eutectoid colonies and large primary η-Zn grains, the ability to accommodate strain induced shape changes is therefore geometrically constrained. Since the primary η-Zn grains creep by dislocation climb,

75 5. Creep of ZnAl4Cu1 alloys they can deform more freely in all directions, while large eutectic and eutectoid colonies are geometrically constrained. Hence, during macroscopic tensile creep in ZnAl4Cu1Mg alloys, deformation in primary η-Zn grains is assumed to be the rate-controlling process, while the eutectic and eutectoid structures contribute only little to the global deformation. Mg addition to Zn-Al alloys causes a decrease of the size of eutectic colonies which have the same interface plane but a higher number density of smaller eutectic colonies with a more random distribution of interface planes, as well as a higher volume fraction of “isolated” eutectoid regions [16]. This is shown in the micrographs in Figure 5.10, where the η-Zn + α-Al eutectic colony boundaries are highlighted in purple and the eutectoid colony boundaries are highlighted in red. The geometrical constraints on the lamellae colonies are weaker in ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31 due to higher number density of smaller lamellae colonies which can deform together as an aggregate. This is assumed to allow strain accommodation in more directions, i.e. interface planes in more directions are available in colony aggregates than in large singular colonies, and therefore contribute more to the global creep deformation. On the other hand, during nanoindentation creep of eutectic and eutectoid colonies, the free surfaces allow deformation along the boundary planes out of the surface, hence, the geometrical constraints in indentation are minimal compared with the bulk material.

76 5. Creep of ZnAl4Cu1 alloys

Figure 5.10. Typical microstructure of (a) ZnAl4Cu1Mg0.04; (d) ZnAl4Cu1Mg0.21; and (h) ZnAl4Cu1Mg0.31 with highlighted eutectic and eutectoid colony boundaries and corresponding schematic alignment of individual constituents in (b, c) ZnAl4Cu1Mg0.04; (e, f) ZnAl4Cu1Mg0.21; and (i, g) ZnAl4Cu1Mg0.31, where the black blocks represent primary η- Zn grains, blue lines represent interfaces in eutectic lamellar colonies and red lines represent interfaces in eutectoid colonies.

77 5. Creep of ZnAl4Cu1 alloys

5.2.3. Influence of Mg on the creep properties of ZnAl4Cu1Mg alloys

We have investigated the creep behaviour of three ZnAl4Cu1MgX alloys (X=0.04, 0.21 and 0.31 wt.%) and their individual microstructural constituents using macroscopic tensile creep tests and nanoindentation creep experiments. Mg additions of 0.21 wt.% or more cause an accelerated macroscopic creep rate, Figure 5.11. The creep properties of the individual microstructural components, namely primary η-Zn phase, η-Zn + α-Al eutectic and eutectoid structures and Mg2Zn11 + eutectoid regions, revealed a higher resistance to creep of the primary η-Zn phase than the eutectic / eutectoid structures and a larger creep deformation in the Mg2Zn11 + eutectoid regions, Figure 5.8.

Figure 5.11. Creep strain-time curves of ZnAl4Cu1MgX (X = 0.04, 0.21, 0.31, in wt.%) alloys at the same stress level at (a) 55°C, (b) 85°C and (c) 105°C.

The effects of Mg on the microstructure and creep properties of the ZnAl4Cu1 alloys investigated are related to (i) solid solution Mg atoms (up to 0.11 % at 364°C according to equilibrium solubility [227]), (ii) formation of Mg2Zn11 precipitates, and (iii) defect segregation, particularly to interfaces.

Specifically, it has been reported that the addition of Mg promotes the formation of Mg2Zn11 precipitates and concurrently causes the formation of a higher volume fraction of isolated globular η-Zn + α-Al eutectoid structure regions with refined lamellar structures in Zn-Al alloys [16, 25, 186, 228]. During the solidification of ZnAl4Cu1Mg alloys with an Mg content of more than 0.11 wt.%, the primary η-Zn phase forms first, leaving excess Mg atoms in the liquid phase as the maximal solubility of Mg in Zn amounts 0.11 wt.% at 364°C [227]. Below the eutectic temperature (382°C [185]), the liquid phase gradually transforms to a η-Zn + β-Al eutectic structure. This eutectic transformation has been reported to be delayed by Mg atoms due to the retarded diffusion of Zn and Al as Mg atoms trap vacancies due to their slightly larger atomic size (125 pm Al, 145 pm Zn, 150 pm Mg) [187, 188, 229]. While a limited amount of Mg atoms (up to 0.11 wt.% [227]) dissolve in the η-Zn crystals and the solid solubility of Mg in Al

78 5. Creep of ZnAl4Cu1 alloys is high [230], the rest of the Mg atoms precipitate directly from the remaining liquid phase as

Mg2Zn11 phase at around 350°C. These Mg2Zn11 phase precipitates are assumed to nucleate preferentially inside the η-Zn and β-Al eutectic structures due to the pre-existing boundaries, as observed in ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.31 alloys, Figure 5.3. With subsequent cooling, the β-Al phase decomposes into a η-Zn + α-Al eutectoid structure at around 275°C

[185]. As most excess Mg atoms are already precipitated in the Mg2Zn11 phase, the amount of solute Mg atoms in the β-Al phase is substantially low and the influence of solute Mg atoms on the kinetics of the eutectoid transformation is therefore assumed to be small. Due to a delayed eutectic reaction and a less affected eutectoid reaction, higher Mg concentrations in Zn-Al alloys cause a decreased volume fraction of eutectic colonies and, hence, a larger number of eutectoid colonies with refined structures [16]. Similar changes in the microstructure have also been reported by da Costa et al. [25] for Zn-4Al-3Cu alloys with Mg contents up to 0.5 wt.%. Further, due to the decreased solubility of Mg in η-Zn and α-Al with decreasing temperature, the solute Mg atoms in η-Zn and α-Al grains are thought to gradually diffuse and segregate to grain and phase boundaries [59].

Consequently, the Mg atoms in ZnAl4Cu1Mg alloys are distributed as (i) solute solution atoms in η-Zn and α-Al, (ii) grain and phase boundaries segregation, and (iii) Mg2Zn11 phase precipitates which decorate the eutectoid structures.

Solute Mg atoms in the η-Zn grains are assumed to pin dislocations and drag the dislocation motion, leading to an improved creep resistance of the η-Zn phase [208] where dislocation climb is the predominant creep mechanism. This is consistent with our experimental results revealing that the primary Zn phase has the highest creep resistance of all microstructural components.

The segregation of Mg atoms to boundaries has been reported to retard the motion of grain and phase boundaries [59], leading to an increased creep resistance of the eutectic and eutectoid structures where coble creep and grain / phase sliding are assumed to be the predominant creep mechanism.

On the other hand, Mg2Zn11 precipitates which decorate η-Zn + α-Al eutectoid structures have been reported to be responsible for the hot shortness of Zn-Al alloys due to the low melting point of the Mg2Zn11 phase [1]. Our nanoindentation creep experiments showed that the eutectoid regions which contain Mg2Zn11 precipitates have the lowest creep resistance. The volume fraction of Mg2Zn11 precipitates increases with increasing Mg content in Zn-Al alloys

[16]. As a result, a higher amount of η-Zn + α-Al eutectoid structures containing Mg2Zn11 precipitates with low creep resistance would lead to higher creep rates of the bulk Zn-Al alloys.

As a higher Mg content causes reduced amounts of eutectic colonies and higher volume fractions of isolated eutectoid regions with refined lamellar structures, the geometrical

79 5. Creep of ZnAl4Cu1 alloys constraints on the eutectic / eutectoid lamellae colonies may be reduced, Figure 5.10, causing a higher contribution of the eutectic and eutectoid structures during creep. As the eutectic and eutectoid structures have a lower creep resistance than the primary η-Zn phase, the accelerated creep rate of ZnAl4Cu1Mg alloys with high Mg content (0.21 and 0.31 wt.%) may therefore also be due to reduced geometrical constraints.

The alloy with the lowest Mg content, ZnAl4Cu1Mg0.04, showed the highest creep resistance.

In this alloy no Mg2Zn11 precipitates are observed and the Mg atoms are in solid solution (maximum solubility of Mg in Zn is 0.11 wt.% at 364°C [227]). Alloys ZnAl4Cu1Mg0.21 and ZnAl4Cu1Mg0.21 with higher Mg contents showed higher creep rates. In these alloys (i) Mg segregation to interfaces, (ii) reduced volume fraction of eutectic colonies and increased volume fraction of eutectoid colonies, (iii) refined eutectoid colonies, and (iv) Mg2Zn11 precipitates in eutectoid regions are observed due to the addition of Mg. As discussed in detail above, only Mg segregation at interfaces causes increased creep resistance, while an increased volume fraction of eutectoid colonies, refined eutectoid colonies and Mg2Zn11 precipitates in eutectoid regions cause accelerated creep rates. Consequently, the macroscopic creep resistance of these alloys is lower than that of alloy ZnAl4Cu1Mg0.04.

Consistent with our results, Mulvania et al. [231] have shown that dilute addition of Mg up to 0.018 wt.% causes a significant improvement in creep resistance of ZnAl22 (in wt.%) alloy, while the addition of 0.06 wt.% Mg cause a higher creep rate than the addition of 0.018 wt.% Mg. We therefore conclude that only dilute solid solution addition of Mg enhances the creep resistance of Zn-Al alloys through solid solution strengthening of the primary Zn phase [59, 208], while higher Mg contents reduce the creep resistance of Zn-Al alloys.

5.3. Conclusions

The creep behaviour of three ZnAl4Cu1MgX alloys (X = 0.04, 0.21 and 0.31, in wt.%) were studied at stresses of 61 – 130 MPa and temperatures of 25 – 105°C. The following conclusions are drawn:

1. The ZnAl4Cu1Mg alloys investigated show stress exponents of 6.9 – 8 and activation energies of 93 – 104 kJ/mol. It is concluded that the rate controlling mechanism in ZnAl4Cu1Mg alloys is stress assisted dislocation climb in the primary η-Zn phase.

2. Primary η-Zn shows the highest creep resistance, and η-Zn + α-Al eutectoid

structures containing Mg2Zn11 precipitates show the lowest creep resistance in nanoindentation creep tests.

3. It is assumed that during macroscopic creep deformation, the eutectic / eutectoid structures creep-deform via Coble creep or grain / phase boundary sliding which

80 5. Creep of ZnAl4Cu1 alloys is not rate controlling due to geometrical constrains.

4. Dilute Mg alloying between 0.21 wt.% and 0.31 wt.% decreases the creep resistance of Zn-Al alloys due to increased grain / phase boundary activities and reduced geometrical constraints of eutectic and eutectoid colonies.

81 6. Micromechanical response and mechanisms

6. Micromechanical response and mechanisms7

In Chapter 4 it was shown that the ZnAl4Cu1 (in wt.%) alloys investigated have multi-phase microstructures composed of η-Zn and η-Zn + α-Al eutectic/eutectoid structures. Therefore, the plasticity mechanisms and mechanical properties of Zn-Al alloys are controlled by both, the intrinsic properties of the individual microstructural constituents and the joint effects of these constituents as an aggregate. However, the local mechanical behaviour and plastic deformation of the individual microstructural constituents, especially the Zn-Al eutectic / eutectoid structures, have not been investigated yet. Further, the distribution of strain within the different microstructural constituents and the co-deformation between η-Zn and η-Zn + α- Al eutectic/eutectoid structures have not been well understood yet. Therefore, to fully understand the local mechanical properties and their interplay in bulk deformation of ZnAl4Cu1 alloys, this chapter aims to explore the local mechanical properties and plastic deformation mechanisms of the individual microstructural constituents in ZnAl4Cu1 alloys as well as the role of the individual microstructural constituents to the global deformation of ZnAl4Cu1 alloys.

6.1. Results

6.1.1. Local mechanical properties

6.1.1.1. Constant strain rate indentation

The local mechanical properties of the individual microstructural constituents in the alloy ZnAl4Cu1Mg0.31 were studied using nanoindentation at room temperature (25°C) and 85°C.

The Mg2Zn11 phase was indented together with the surrounding eutectoid structures due to its small size of less than 1 µm. The averaged steady state indentation hardness, H, and elastic modulus, E, of the individual microstructural constituents are summarised in Table 6.1. Here, the steady state data at 25°C were obtained at an indentation depth of 800 to 1200 nm, a strain rate of 0.1 s-1 and a peak load of 25 mN. The steady state data at 85°C were obtained at an indentation depth of 500 to 700 nm, a strain rate of 0.05 s-1 and a peak load of 25 mN.

At RT, the η-Zn phase showed the lowest hardness of the individual microstructural constituents, while the Mg2Zn11 + eutectoid regions exhibited a significantly higher hardness than the other constituents. The hardness of the microstructural constituents displays the following trend: H ˃ H ˃ H ˃ H . Compared to RT, all microstructural Mg2Zn11+Eutectoid Eutectoid Eutectic Zn constituents showed a decrease in hardness at 85°C. While the Mg2Zn11 + eutectoid regions

7 This part was published as an article. The original citation is: Z. Wu, S. Sandlöbes, J. Rao, J. S. K.-L. Gibson, B. Berkels and S. Korte-Kerzel. (2018) "Local mechanical properties and plasticity mechanisms in a Zn-Al eutectic alloy." Materials & Design 157: 337-350.

82 6. Micromechanical response and mechanisms exhibited a hardness of 1.0 – 1.6 GPa, the η-Zn phase as well as the η-Zn+α-Al eutectic / eutectoid structures showed similar hardness values of around 0.6 – 1.1 GPa.

Table 6.1. Hardness, H, and elastic modulus, E, of individual microstructural constituents in ZnAl4Cu1Mg0.31 investigated at 25 ºC and 85°C, respectively. The relative values are given as a reduction of the mean value (e.g. 1.2 GPa for uncorrected Zn) with temperature to indicate the temperature dependence.

Microstructural Mg Zn + T (°C) Zn Eutectic Eutectoid 2 11 consitutent Eutectoid

25 1.2 ± 0.1 1.4 ± 0.2 1.5 ± 0.1 2.1 ± 0.2 Hardness, 85 1.1 ± 0.2 1.0 ± 0.1 1.0 ± 0.2 1.5 ± 0.3 H (GPa) Reduction 8% 29% 33% 29%

Hardness, H (GPa) 25 1.1 ± 0.1 1.2 ± 0.2 1.3 ± 0.1 1.7 ± 0.2

(Pile-Up corrected 85 0.9 ± 0.2 0.8 ± 0.1 0.8 ± 0.2 1.3 ± 0.3 from representative indent) Reduction 18% 33% 39% 24%

25 102 ± 13 104 ± 11 94 ± 10 103 ± 12 Elastic modulus, 85 82 ± 15 84 ± 11 80 ± 10 99 ± 6 E (GPa) Reduction 20% 19% 15% 4%

Elastic modulus, 25 96 ± 12 95 ± 11 85 ± 8 92 ± 11

E (GPa) 85 74 ± 14 74 ± 10 70 ± 9 89 ± 6 (Pile-Up corrected from representative Reduction 23% 22% 18% 3% indent)

6.1.1.2. Nanoindentation strain rate jump test

The strain rate sensitivity, m, which reflects the dependence of the flow stress, σ, on the strain rate, ε̇, is defined as [59, 152]:

83 6. Micromechanical response and mechanisms

∂ ln σ m= (6.1) ∂ ln ε̇

The apparent activation volume, 푉∗, is calculated by [232-235]:

휕퐺 휕푙푛휀̇ 푉∗ = − ( ) = 푘푇 ( ) (6.2) 휕휏 푇 휕휏 푇 where G is the Gibbs free energy of activation, 흉 is the externally applied shear stress, k is the Boltzmann constant, T is the temperature in K.

Similar to macroscopic strain rate jump tests, the strain rate sensitivity and activation volume in nanoindentation can be calculated according to [158, 236-238]:

∂ ln H m= (6.3) ∂ ln 휀푖̇

휕푙푛휀̇ 푉∗ = 3√3푘푇 ( ) (6.4) 휕퐻 푇 where H is the hardness obtained from indentation experiments and 휀푖̇ is the corresponding indentation strain rate. In a load controlled experiment, the indentation strain rate is given according to Lucas and Oliver [174] as:

ḣ 1 Ṗ Ḣ 1 Ṗ 휀̇ = = ( - ) ≈ (6.5) 푖 h 2 P H 2 P where h is the indentation depth, P is the load, Ṗ ⁡is the loading rate.

Figure 6.1a shows a typical true stress-strain curve obtained during macroscopic uniaxial tensile strain rate jump tests at 85°C, where the strain rate was changed between 5∙10-4 and 6∙10-6 s-1. The steady state stresses upon strain rate jumps were determined via linear fitting (Figure 6.1b), and used to calculate the values of m (0.108 ± 0.012) and V* (0.546 ± 0.053 nm3) at 85°C, Table 6.2.

Figure 6.1c and d show typical hardness vs indentation depth (H-h) curves and corresponding indentation strain rate vs. indentation depth curves of the primary η-Zn phase and η-Zn + α- Al eutectoid structures obtained by indentation strain rate jump tests at RT and 85°C. It is apparent that both the primary η-Zn phase and η-Zn + α-Al eutectoid structures exhibited a strain rate sensitivity, i.e. the hardness of both microstructural constituents showed considerable changes upon indentation strain rate jumps. The observed hardness changes upon strain rate jumps were higher at 85°C than at RT and the hardness changes of the eutectoid structure are obviously larger than those of the primary Zn phase at RT and 85°C, Figure 6.1c, d.

84 6. Micromechanical response and mechanisms

Figure 6.1. Dependence of the true stress on the strain rate of alloy ZnAl4Cu1Mg0.31 obtained by macroscopic tensile strain rate jump tests at 85°C (a). (b) shows how the corresponding stresses upon strain rate jumps from 5∙10-4 s-1 to 6∙10-6 s-1 were determined. The dependence of the hardness of the primary η-Zn phase and η-Zn + α-Al eutectoid structures on the strain rate obtained by indentation strain rate jump tests at RT (c) and 85°C (d).

Due to the apparent indentation size effect of the primary Zn phase and the eutectoid structures, linear fitting was applied in each segment of the H-h curves. The corresponding hardness values upon instantaneous strain rate changes were used to calculate the strain rate sensitivity and activation volume according to Maier et al. [239]. Table 6.2 shows the resulting indentation strain rate sensitivity, m, and activation volume, V*, of the primary Zn phase and the eutectoid structure in the strain rate range 5∙10-4 – 5∙10-2 s-1 at RT and 85°C. Both, the primary η-Zn phase and η-Zn + α-Al eutectoid structures showed an increase in strain rate sensitivity and a decrease in activation volume with increasing temperature. Generally, the eutectoid structures possess higher m and lower V* values than the primary η- Zn phase. The m value obtained from macroscopic tensile strain rate jump tests at 85°C lies between the m values of these two microstructural constituents, indicating a contribution of both, primary η-Zn phase and η-Zn + α-Al eutectoid structures to the global deformation at elevated temperatures.

85 6. Micromechanical response and mechanisms

Table 6.2. Nanoindentation strain rate sensitivity and activation volume of the primary η-Zn phase and η-Zn + α-Al eutectoid structures at RT and 85°C compared to bulk ZnAl4Cu1Mg0.31 at 85°C.

Strain rate sensitivity, m Activation volume, V* (nm3) T (°C) Zn Eutectoid Bulk Zn Eutectoid Bulk

0.026 ± 0.051 ± 0.781 ± 0.331 ± - - 25 0.004 0.009 0.124 0.097 0.051 ± 0.157 ± 0.108 ± 0.570 ± 0.209 ± 0.546 ± 85 0.016 0.044 0.012 0.246 0.083 0.053

6.1.2. Deformation microstructure

6.1.2.1. Deformation microstructure in primary η-Zn phase

Figure 6.2 shows an SE image, inverse pole figure (IPF) map and AFM images of a typical indent in the primary η-Zn phase (indentation was performed at RT with a strain rate of 0.1 s-1 and a peak load of 25 mN). Twins with lenticular shapes are visible near the corners of the residual imprint in the η-Zn grain, Figure 6.2b. The twins were identified as {101̅2} 〈1011̅̅̅̅〉 compression twins and the corresponding twin boundaries are highlighted in yellow in the IPF map.

The pile-ups at all three sides of the imprint were measured and considered in the pile-up correction of the hardness and modulus values reported in Table 6.1. Due to preferential orientations for dislocation slip in the surrounding grain, dislocation pile-ups mainly occurred on one side of the indent, Figure 6.2a. These are shown in more detail in the AFM images,

Figure 6.2c and e. The slip traces are parallel to the basal (0001)Zn plane trace, as highlighted in red in the AFM micrograph, indicating the activation of basal slip in the primary η-Zn phase. AFM height profile measurements across the deformation twin (line 1 in Figure 6.2c) and the slip traces (line 2 in Figure 6.2c) revealed that the surface step formed by the deformation twin has a height of 120 nm and the height of those formed by dislocation slip traces amount to 2 – 9 nm , Figure 6.2e. The height of the basal slip traces correspond to 8 – 33 Burgers vectors (b= 2.7∙10-10 m for dislocations in η-Zn [158, 161]).

86 6. Micromechanical response and mechanisms

Figure 6.2. Post-mortem micrographs of a typical indent in the primary η-Zn phase (RT, 0.1 s-1 strain rate, 25 mN peak load), (a) SE micrograph; (b) IPF map, twin boundaries are highlighted in yellow; (c) AFM micrograph, basal plane traces are highlighted in red; (d) height profile along line 1 in (c); (e) height profile along line 2 in (c) with corresponding AFM micrograph. The dashed white rectangle in (a) shows the region of the AFM map shown in (c).

Figure 6.3 shows SE images and IPF maps of typical indents in the primary η-Zn phase after indentation at 85°C, a strain rate of 0.1 s-1 and a peak load of 25 mN. Similar to indentation at RT, {101̅2} 〈1011̅̅̅̅〉 compression twins were formed near the corners of the imprint in primary η-Zn, as highlighted in yellow in Figure 6.3b. Additionally, slip traces are visible around indentation imprints in η-Zn, Figure 6.3c, e. These slip traces were indexed as basal

(0001)Zn, prismatic {101̅0}Zn and pyramidal {101̅1}Zn plane traces using EBSD.

87 6. Micromechanical response and mechanisms

Figure 6.3. Post-mortem micrographs of typical indents in the primary η-Zn phase (85°C, 0.1 s-1 strain rate, 25 mN peak load), (a) SE micrograph and (b) IPF map showing deformation twins highlighted by yellow boundaries, the IPF map has been rotated to guide the eyes; (c), (e) SE micrographs and (d), (f) IPF maps showing basal and non-basal slip traces; the insets in (c), (e) are enlarged micrographs of the slip traces.

6.1.2.2. Deformation microstructure in η-Zn + α-Al eutectoid structures

Figure 6.4 shows SE and corresponding AFM micrographs of η-Zn + α-Al eutectoid structures after nanoindentation at RT and 85°C. In the AFM micrographs, curved surface traces are visible around the indentation imprints in the eutectoid structures, Figure 6.4b, e, h. Comparison of the AFM observed surface traces with corresponding SE micrographs, Figure 6.4a, d, g, revealed that most of the formed surface traces are following η-Zn (bright) and α- Al (dark) phase boundaries, see the drawn-in dashed red lines.

88 6. Micromechanical response and mechanisms

Figure 6.4. SE micrographs and AFM maps of typical indentation imprints in eutectoid structures and primary Zn phase, the indentations were performed at 25°C at a strain rate of 0.1 s-1 (a)-(c), 85 ºC at a strain rate of 0.1 s-1 (d)-(f), 85°C at a strain rate of 0.01 s-1 (g)-(h) and a maximum load of 25 mN. (a), (d), (g) are SEM images with imposed deformation traces; (b), (e), (h) are the corresponding AFM images with red arrows pointing to the surface traces; (c) and (f) are the corresponding IPF maps.

The density of surface traces that follow Zn-Al phase boundaries in eutectoid structures increased with increasing temperature. Table 6.3 lists the average heights of the surface traces formed during nanoindentation in eutectoid structures measured by AFM, and for comparison the average height of Zn-Al phase boundaries in eutectoid structures of un- deformed material is also given. The height of Zn-Al interface traces induced during nanoindentation increases with increasing temperature and decreasing strain rate, Table 6.3.

89 6. Micromechanical response and mechanisms

Table 6.3. Average surface heights of Zn-Al phase boundaries in eutectoid structures around nano-indentation imprints. The indentations were performed at a maximum load of 25 mN at 25°C and 85 ºC at a strain rate of 0.1 s-1, and at 85°C at a strain rate of 0.01 s-1. For comparison the average height of Zn-Al phase boundaries in eutectoid structures in un- deformed material is also listed.

Temperature (ºC) 25 85

Strain rate (s-1) un-deformed 0.1 0.1 0.01 Height of Zn-Al phase boundaries in eutectoid 13 ± 3 23 ± 8 37 ± 14 69 ± 29 structures (nm)

6.1.2.3. Strain partitioning

Using quasi in-situ tension experiments we have deformed a bulk ZnAl4Cu1Mg0.31 alloy specimen to 2% and 5% tensile strain at 85°C, Figure 6.5. The local strain distribution in the region of interest (ROI) at each strain was calculated using DIC, Figure 6.6a, b, and the non rigid method, Figure 6.6c, d. The strain maps are scaled to 0 – 0.3 equivalent von Mises strain, allowing an optimum contrast of the local strain. We have named the primary η-Zn grains by numbers, 1-15, and η-Zn + α-Al eutectic / eutectoid colonies by letters, A-N, in the ROI, Figure 6.5c.

At 2% global strain, the microstructure in the ROI (Figure 6.5b) remained nearly identical to the one before deformation (Figure 3.5a), i.e. the grains mainly maintained their initial morphologies. On the other hand, the local strain map (Figure 6.6a) clearly revealed strain in the microstructure. Figure 6.6e shows the local von Mises strain map at 2% global strain overlaid with the corresponding SE micrograph. Here it is evident that plastic strain was accumulated in the η-Zn + α-Al eutectic and eutectoid colonies (area A, B, E, G), at the boundaries between primary η-Zn grains and the interfaces in eutectic colonies (see e.g. the boundaries between grain 10 and areas E, H). In contrast, most of the primary Zn grains carried only little or no strain (grains 3, 4, 11, 12, 15), Figure 6.6a, e. Further, straight localised lines are visible in several Zn grains (grains 1, 2, 13, 14) and in the eutectic / eutectoid colonies

(colonies A, K), Figure 6.6a. According to EBSD, these lines are parallel to basal (0001)Zn plane traces (highlighted in red in Figure 6.6a), indicating the activity of basal dislocation slip in the Zn grains of the primary η-Zn phase and η-Zn + α-Al eutectic / eutectoid colonies.

At 5% global strain, pronounced local deformation was observed at Zn-Al phase boundaries between primary η-Zn grains and eutectic colonies (grain 5, 10) and in the eutectic / eutectoid colonies (area H, L, J), see Figure 6.5b. Due to the large local strain, the DIC method was not able to accurately calculate the local strain in those regions, which appear white in the strain map (e.g. the boundaries between grain 10 and area E, H, Figure 6.6b, f). On the other hand,

90 6. Micromechanical response and mechanisms the non-rigid registration method showed strain concentrations in these regions. Unlike DIC, non-rigid registration does not need to assume that deformation is affine (or even just a translation) on small image patches. Thus, it is able to resolve the non-linear deformation in these areas in contrast to DIC. In addition to localised lines parallel to basal (0001)Zn plane traces, in grains 8 and 9 localised lines parallel to the (1011)Zn (highlighted in orange) and

(112 2)Zn plane traces (highlighted in black) were observed (Figure 6.6b), indicating the activation of non-basal slip in the primary Zn phase. Generally, the distribution of the local strain was heterogeneous and concentrated mainly at the interfaces of eutectic / eutectoid colonies.

Figure 6.5. SE micrographs of the region of interest (ROI) in alloy ZnAl4Cu1Mg0.31 at (a) 2% and (b) 5% tensile strain at 85°C; (c) BSE micrograph of the ROI at 0 strain, primary η-Zn grains are numbered by 1-15 (highlighted in red) and η-Zn + α-Al eutectic colonies labelled A-N (highlighted in blue); (d) IPF figure of the ROI at 0 strain.

91 6. Micromechanical response and mechanisms

Figure 6.6. Local von Mises strain map of the region of interest in alloy ZnAl4Cu1Mg0.31 at (a, c, e) 2%, (b, d, f) 5% tensile strain at 85°C. (a, b) Strain maps calculated using the software GOM correlate (V8.1, GOM mbH).; (c, d) strain maps calculated using the non-rigid registration method; (e, f) DIC strain maps are overlaid with SE images of the microstructure.

92 6. Micromechanical response and mechanisms

6.2. Discussion

6.2.1. Local mechanical properties of individual microstructural constituents

Due to the observed influence of pile-ups, the data were corrected using the true contact area of a representative indent as measured by AFM (Table 6.1), and these corrected values are thus used in the subsequent discussion. In all cases, a small reduction in hardness and modulus is found, with a slightly larger correction at 85˚C as would be expected from the softer material. Nevertheless, the relative trends between the phases remain the same before and after this contact area correction. On the other hand, as the contact area of the indent remains effectively constant upon one instantaneous change in strain rate during nanoindentation strain rate jump tests, the relative change in hardness is not obviously affected by the pile-ups. Therefore, the area correction for pile-ups was not considered in the analysis of strain rate sensitivity, m, and activation volume, V*.

The elastic modulus of 84 – 108 GPa of the primary η-Zn phase obtained using nanoindentation at RT is consistent to the study of Ledbetter [240] who has reported an elastic modulus of 83 – 121 GPa for polycrystalline zinc. The large scatter of E has been assumed to be caused by the high anisotropy of Zn [240]. The hardness of η-Zn at RT obtained in the present study amounts to 1.1 ± 0.1GPa, which is similar to the hardness reported by Tohid et. al. (1.1 GPa) [179] for pure Zn. Further, our nanoindentation results show that the hardness of the individual microstructural constituents follow the trend: H ˃ H ˃ Mg2Zn11+⁡Eutectoid Eutectoid

HEutectic˃ HZn at RT. The same trend has been reported by Li et al. [186] for ZnAl4Mg3 (in wt.%) using Vickers microhardness testing with a load of 250 mN, and by Mahmudi and Alibabaie [5] for Zn and eutectic structures in ZnAl6Cu3 (in wt.%) using Vickers microhardness testing with a load of 20 mN. Mahmudi and Alibabaie [5] have attributed the higher hardness of eutectic / eutectoid structures (with respect to primary Zn) to phase boundary strengthening [5]. Similarly, Gan et al [241] have reported strong dislocation retention by the interphase boundaries in eutectic Sn/Pb alloys when the sample size is significantly larger than the lamellar spacing. Due to the fine grains inside the eutectic / eutectoid structures exhibiting a secondary Zn lamellar spacing of 0.6 – 1.7 µm in eutectic structures and of 0.06 – 0.29 µm in eutectoid structures in the present study, it is assumed that the observed high hardness of eutectic / eutectoid structures is similarly caused by phase boundary strengthening.

At 85°C, all microstructural constituents show a lower hardness than at RT. Consistent to our results, Freeman et al. [242] have also reported a decrease of the Brinell hardness from 31 HB at 20°C to 19 HB at 110°C in pure Zn. Compared to the primary η-Zn phase, the eutectic

/ eutectoid structures and the Mg2Zn11 phase show a larger decrease in hardness with increasing temperature, indicating a higher thermal activation in the eutectic and eutectoid

93 6. Micromechanical response and mechanisms structures than in the primary η-Zn phase. Specifically, the decrease in hardness in eutectic and eutectoid structures has been assumed to be caused by a weaker phase boundary strengthening in eutectic and eutectoid structures at elevated temperatures [5], as phase boundary sliding becomes more active. Further, the higher strain rate sensitivity, m, and lower activation volume, V*, in eutectoid structures than in the primary η-Zn phase indicate higher thermal activation in eutectoid structures than in the primary η-Zn phase, Table 6.2. However, due to the relatively small colony size of η-Zn + α-Al eutectoid structures of around 2.8 – 9.5 µm, the η-Zn phase underneath the eutectoid structures is assumed to contribute to deformation during indentation into the eutectoid structures. Specifically, it has been reported by Bull [224] that the plastic deformation zone diameter is ~10 times the indenter penetration, in the present study the maximum indenter penetration was around 1.2 µm at 25°C and 0.8 µm at 85°C.

6.2.2. Deformation mechanisms and thermal activation

6.2.2.1. Deformation of the primary η-Zn phase

Using nanoindentation strain rate jump tests we have obtained a strain rate sensitivity, m, of 0.022 – 0.030 at RT and 0.035 – 0.071 at 85°C in the primary η-Zn phase. The activation volume, V*, of the primary η-Zn phase amounts to 0.657 – 0.905 nm3 at RT (corresponding to 39.7 ± 6.3b3) and 0.324 – 0.816 nm3 at 85°C. Both, m and V* values suggest thermally activated dislocation-dominated deformation in the primary η-Zn phase at RT and 85°C [120, 158, 163]. The strain rate sensitivity values obtained using nanoindentation strain rate jump tests in the present study are close to those of bulk Zn-0.3Al (m: 0.07 [163]) and Zn-0.4Al (m: 0.05 [120]) obtained using macroscopic tensile tests at RT. These alloys were reported to have microstructures mainly containing primary η-Zn grains and dislocation-dominant deformation has been proposed [120, 163]. The slightly lower m value of the primary η-Zn phase at RT in our study is assumed to be caused by a coarser grain size [98] (10 µm in Zn- 0.3Al alloy [163] and 0.6 µm in Zn-0.4Al alloy [120]) and a higher experimental strain rate [30, 120, 122, 159, 163].

Consistent with the m and V* values obtained using nanoindentation, we have observed {101̅2} 〈1011̅̅̅̅〉 compression twins and basal slip traces in primary η-Zn phase grains indented at RT, see Figure 6.2. Upon indentation at 85°C, slip traces indexed as basal and non-basal slip planes and deformation twins were observed in primary η-Zn phase grains, Figure 6.3. It is therefore concluded that the predominant deformation mechanisms in the primary η-Zn phase are {101̅2} 〈1011̅̅̅̅〉 compression twinning and basal slip at RT, and deformation twinning and basal and non-basal dislocation slip at 85°C.

94 6. Micromechanical response and mechanisms

The same deformation mechanisms have been observed before in pure Zn [2, 79, 90, 92, 104] and in the primary η-Zn phase in bulk Zn-Al alloys [16, 148]. Easy activation of basal slip at RT has been reported consistently for pure Zn [74, 79, 90-92, 103]. Further, it has been reported that the CRSS of non-basal dislocation slip decreases with increasing temperature [2, 59, 90, 104, 149]. The activation of non-basal slip systems has been reported in Zn and Zn alloys at elevated temperatures, such as the activation of pyramidal slip in Zn at 150°C [90], at 300 – 360 ºC [104] and in the η-Zn phase of a Zn-4Al-1Cu-0.3Mg alloy at 85°C [148].

6.2.2.2. Deformatlion of η-Zn + α-Al eutectoid structures

Nanoindentation of the η-Zn + α-Al eutectoid structures induced the formation of curved surface steps along Zn-Al boundaries at RT and 85°C, indicating the activation of grain / phase boundary sliding [158, 222, 238, 243]. Further, we observed an increased density and height of these surface steps with increasing temperature and decreasing strain rate, Table 6.3. This is consistent with the observations of Wheeler et al. [238] who have observed increasing activation of grain boundary sliding with increasing temperature during indentation of ultrafine-grained Al. It is therefore assumed that the predominant deformation mechanism in the η-Zn + α-Al eutectoid structures is grain / phase boundary sliding at RT and 85°C. Grain / phase boundary sliding has been widely reported for fully eutectoid Zn-22Al alloys at RT and at elevated temperatures [6, 96, 97, 105-118].

The strain rate sensitivity of the η-Zn + α-Al eutectoid structures obtained from nanoindentation strain rate jump tests amounts to 0.042 – 0.060 at RT and 0.113 – 0.201 at 85°C. In Zn-Al alloys, a strain rate sensitivity of more than 0.12 has been reported to be indicative for grain / phase boundary dominated deformation mechanisms [30, 94, 101, 122, 158, 159]. Specifically, in a Zn-22Al alloy with a fully η-Zn + α-Al eutectoid microstructure, Choi et al. [158] have reported an m value of 0.12 at RT associated with sliding of Zn-Zn and Zn-Al interfaces. During macroscopic tensile testing of a fully eutectoid Zn-22Al alloy at 45°C, Tanaka et al. [122] have observed grain and phase boundary sliding associated with an m value of 0.3. At 200°C, Kawasaki and Langdon [101] have observed grain boundary sliding in an eutectoid Zn-22Al alloy together with a strain rate sensitivity value of 0.43. Further, using nanoindentation strain rate jump tests we have obtained an activation volume, V*, of 0.234 - 0.428 nm3 at RT and 0.126 - 0.292 nm3 at 85°C for the eutectoid structures. An activation volume of ~ b3 has been related to grain boundary diffusion [75, 244], while grain / phase boundary sliding has been reported to have activation volumes of ~10 b3 [157, 158, 236]. The activation volumes obtained in the present study correspond to 11.7 – 21.4 b3 at RT and 6.3 3 1 ̅ – 14.6 b at 85°C (b: ⁄3 〈1120〉 dislocation). The m and V* values obtained for eutectoid structures at 85°C therefore strongly indicate grain / phase boundary sliding dominated deformation, which is consistent with post-mortem microstructure observations, Figure 6.4.

95 6. Micromechanical response and mechanisms

On the other hand, although the V* value obtained for the eutectoid structures at RT is slightly higher than commonly reported values associated with grain / phase boundary sliding, we have observed clear evidence of grain / phase boundary sliding in the eutectoid structures at RT. The slightly lower m and higher V* values of eutectoid structures at RT are assumed to be caused by a higher contribution of the η-Zn phase underneath the eutectoid structures to the deformation during indentation into the eutectoid structures. During nanoindentation strain rate jump tests, the maximum indenter depth was around 1.2 µm at 25°C and 0.8 µm at 85°C. Consequently, the η-Zn phase underneath the eutectic / eutectoid structures contributes to deformation during indentation into the eutectic / eutectoid structures as the colony size of eutectoid structures amounts to 2.8 – 9.5 µm while the plastic deformation zone under the indenter amounts to ~10 times the indenter penetration depth [224]. Due to the larger penetration depth at RT than at 85°C the contribution of the primary Zn phase is assumed to be larger at RT.

6.2.3. Local strain distribution and strain transfer

The local strain maps (Figure 6.6) show heterogeneous distribution of strain in the microstructure. Quantitative measurement of the local strain in the individual microstructural constituents shows that the η-Zn + α-Al eutectic and eutectoid structures carry higher strain than the primary η-Zn grains during tension at 85°C. Our nanoindentation tests at 85°C further reveal that the predominant deformation mechanisms of the primary η-Zn phase are dislocation slip and to a lesser extent deformation twinning, while in the eutectoid structures grain / phase boundary sliding is observed to be the predominant deformation mechanism at 85°C.

Further, it is observed that the local displacement of the surface pattern inside primary η-Zn grains is heterogeneously distributed and localises along straight lines parallel to slip planes in η-Zn (grain 10, 13, 14 in Figure 6.6b), presumably due to localised slip. It is well-known that the local stress state in polycrystalline aggregates differs from the globally applied stress, however, we had to consider the applied uniaxial global stress state (see red arrows in Figure 6.5a), as the actual local stress state is not available. With this simplified assumption, we have calculated the Schmid factors of each slip systems in the primary η-Zn phase grains and the η-Zn + α-Al eutectic and eutectoid colonies, Figure 6.5.

Comparison of the strain maps and the calculated Schmid factors of the primary η-Zn grains reveals that most of those primary η-Zn grains that carry high plastic strain have soft orientations for basal slip. For example, the Schmid factors for basal slip of the primary Zn grains 1, 2, 6, 7, 13, where relatively large local strains are observed, are higher than 0.35. Primary Zn grains with a low Schmid factor for basal slip are observed to carry only little

96 6. Micromechanical response and mechanisms strain, e.g. grains 3 and 4, both have Schmid factors for basal slip of 0.01. On the other hand, the primary Zn grains 11 and 12 exhibit only low local strain although they have large Schmid factors for basal slip (0.48). This is attributed to the fact that the actual stress state may differ from the global stress state.

Generally, higher strains are accummulated in η-Zn + α-Al eutectic and eutectoid colonies, mostly independent of the crystallographic orientation (and consequently the Schmid factor), where grain / phase boundary sliding is the predominant deformation mechanism.

Further, strain concentrations are observed at the boundaries between primary η-Zn grains and eutectic / eutectoid colonies (e.g. between grain 10 and colony E, H) and colony boundaries (e.g. between colonies A, B, C, F), Figure 6.6b, f.

It is visible from the strain maps that the lines of localised strain in grains 1, 2 and colony A as well as in grains 6, 7 and colonies O and K, retain their directions across the interfaces, Figure 6.6b, f and Figure 6.7a, b. We have therefore employed the Luster-Morris parameter, 푚′, to analyse the alignment of deformation modes across boundaries, i.e. the possibility of strain transfer across these boundaries [245-247]:

푚′ = cos(휑) cos⁡(휅) (6.6) where 휑 is the angle between two normal directions of slip / twining planes in two neighbouring grains and 휅 is the angle between the two slip / twining directions in the corresponding grains.

For grains 1, 2 and colony A, the Luster-Morris parameters for basal slip in the Zn phase in eutectic colony A and the primary η-Zn grains 1 and 2 are close to 1, indicating easy strain transfer. Further, easy slip transfer is assumed in this region due to the orientation relationship between Zn and Al grains in the η-Zn + α-Al eutectic and eutectoid structures where (0001) η-

Zn eut // (110) α-Al eut and [12̅10] η-Zn eut // [11̅0] α-Al eut, as these interfaces do not act as strong obstacles for dislocation slip [248]. The Luster-Morris parameter of basal slip in eutectic colony O (basal plane traces observed) and 2nd order pyramidal slip in the primary η-Zn grain 9 (2nd order pyramidal plane traces observed) amounts to 0.92, indicating the activation of 2nd order pyramidal slip in grain 9 due to strain transfer from colony O into grain 9, Figure 6.7b, c.

97 6. Micromechanical response and mechanisms

Figure 6.7. Enlarged DIC strain maps show strain transfer between primary η-Zn grains 1, 2 and eutectic colony A (a), and between grains 6, 7 and colonies O, K (b); (c) schematic displays the orientation of slip systems in Zn grain 9 (2nd order pyramidal plane traces observed) and eutectic colony O (basal traces observed) across the interface.

6.2.4. Bulk deformation

Using nanoindentation experiments into the individual microstructural constituents of a ZnAl4Cu1Mg0.31 alloy, we showed that the predominant deformation mechanisms in the primary η-Zn phase are {101̅2} 〈1011̅̅̅̅〉 deformation twinning and basal slip at RT. These deformation mechanisms alone only offer three independent deformation modes [72], which is not sufficient to fulfil the von Mises criterion requiring at least five independent deformation modes for homogeneous plastic deformation of polycrystalline aggregates, resulting in the intrinsic brittleness of Zn at RT. At 85°C, additionally non-basal dislocation slip was observed in the η-Zn phase. Together with the other active plastic deformation mechanisms non-basal slip offers sufficient independent deformation modes, hence, causing improved ductility of Zn

98 6. Micromechanical response and mechanisms at elevated temperatures. Grain and phase boundary sliding were observed to be the predominant deformation mechanism in the η-Zn + α-Al eutectoid structures at RT and 85°C.

At RT, the onset of plastic deformation is assumed to occur in the primary η-Zn phase due to its low hardness, Table 6.1. However, as Zn is intrinsically brittle at RT, fracture of the primary η-Zn phase grains occurs easily before the eutectic and eutectoid structures start to deform. This has been shown before by in-situ tensile tests of bulk ZnAl4Cu1Mg0.31 at RT [16], where the primary η-Zn phase fractured quickly, while the eutectic and eutectoid structures contributed only minorly to the bulk deformation at RT [16].

At 85°C, both, the primary η-Zn phase and the η-Zn + α-Al eutectic and eutectoid structures contribute to bulk plastic deformation of ZnAl4Cu1Mg0.31 [16]. This is further evident from the strain rate sensitivities of the individual microstructural constituents and of the bulk alloy obtained in the present study. Further, using DIC we showed that the eutectic and eutectoid structures carry higher strain than the primary η-Zn phase at 85°C. The observed ductility of alloy ZnAl4Cu1Mg0.31 at elevated temperatures [16] is therefore assumed to be caused by the increased ductility of the primary Zn phase due to a high number of available deformation modes on the one hand, and due to the contribution of η-Zn + α-Al eutectic and eutectoid structures to the bulk deformation on the other hand.

6.3. Conclusion

Due to the brittleness of bulk Zn-Al alloys at room temperature, the room temperature deformation mechanisms of Zn-Al eutectic / eutectoid structures are difficult or even impossible to study by macroscopic deformation experiments. Using instrumented indentation techniques, the local mechanical properties and deformation mechanisms of the individual microstructural constituents of a ZnAl4Cu1Mg0.31 alloy were studied at RT and 85°C. The combination of nanoindentation techniques and digital image correlation and non-rigid image registration allows the complete mechanical characterisation of complex, multi-phase alloys.

The following specific conclusions are also drawn:

1. The nanoindentation hardness follows the order: HMg2Zn11 + Eutectoid˃ HEutectoid˃ HEutectic˃

HZn at RT. At 85°C, the eutectic and eutectoid structures have a similar hardness as the primary η-Zn phase.

2. At RT, the predominant deformation mechanisms in the primary η-Zn phase are basal slip and {101̅2} 〈101̅1̅〉 deformation twinning. At 85°C, non-basal slip is additionally activated in the primary η-Zn phase.

3. The predominant deformation mechanism in η-Zn + α-Al eutectoid structures is grain

99 6. Micromechanical response and mechanisms

and phase boundary sliding at RT and 85°C.

4. At RT, the primary η-Zn phase is the main contributor to bulk deformation of the ZnAl4Cu1Mg0.31 alloy. At 85°C, both the primary η-Zn phase and the eutectic / eutectoid structures contribute to the deformation, and the eutectic / eutectoid structures carry higher strain than the primary η-Zn phase.

100

This page intentionally left blank.

7. Precipitation behaviour

7. Precipitation behaviour8

At present, the application of zinc alloys is hindered by their long-term mechanical and dimensional instability at ambient or slightly elevated temperatures [11, 19-21].

The long-term mechanical/dimensional instability of Zn alloys has been assumed to be mainly caused by aging, which is already activated at room temperature due to the high homologous temperature (T/Tm) of Zn alloys [11, 19, 21]. It has been reported that rapid cooling of Zn-Al and Zn-Al alloys results in super-saturation of Al and Cu in η-Zn which causes various phase transformation and decomposition reactions including a number of metastable phases upon aging [1, 6, 8, 13, 21, 25, 31, 39-46, 49] , namely:

. decomposition of the supersaturated Al-rich FCC α', Zn-rich FCC β' and Zn-rich hexagonal η' solid solution phases, . solid state phase transformation with the participation of different phases, α, ε, η and T’, . precipitation due to super-saturated solid solutions.

These phase transformations have been reported to proceed for months or even years during natural aging [1, 21, 39], and are assumed to cause a loss of strength and also dimensional changes [40]. However, only limited and partly contradicting studies focus on aging of Zn alloys and the related effects on the mechanical stability and the underlying mechanisms are not yet fully understood. In order to improve the long-term mechanical stability of Zn based alloys, a deeper and more quantitative knowledge of the aging phenomena and phase stabilities are essential. Therefore, this chapter aims at achieving a more comprehensive and quantitative description of the precipitation and decomposition phenomena in ZnAl4Cu1 alloys.

7.1. Results and discussion

Using uniaxial tensile tests the mechanical properties of ZnAl4Cu1 alloys were characterized at temperatures of 25 – 85°C and at strain rates of 5∙10-4 - 6∙10-6 s-1, Table 4.2. Figure 7.1 shows the true stress-strain curves of the ZnAl4Cu1Mg0.31 alloy at RT and at 85°C at a stain rate of 5∙10-4 s-1 and the corresponding fracture surfaces. The ZnAl4Cu1 alloys showed dissimilar mechanical properties at a narrow temperature range of 25 – 85°C, cf. 4.1.2. At RT, the alloys showed brittle behaviour with fracture surfaces mainly comprised of smooth cleavage facets, Figure 7.1b, while at 85°C the alloys exhibited good ductility with fracture surface mainly comprised of dimples, Figure 7.1c. The increase in ductility of ZnAl4Cu1 alloys at elevated temperatures were assumed to be caused by a higher number of active

8 This part was published as an article. The original citation is: S. Sandlöbes, Z. Wu, K. Pradeep and S. Korte-Kerzel (2016). "Precipitation and decomposition phenomena in a Zn-Al-Cu-Mg alloy." Materials Letters 175: 27-31.

101 7. Precipitation behaviour deformation mechanisms, such as the activation of non-basal dislocation slip in primary η-Zn phase and grain / phase boundary sliding in the η-Zn + α-Al eutectic / eutectoid structures, cf. 4.1.4, 6.2.2, 6.2.4.

Figure 7.1. True stress – true strain curves of ZnAl4Cu1Mg0.31 alloy at room temperature and at 85°C and at a strain rate of 5∙10−4 s-1(a) and the corresponding fracture surfaces of ZnAl4Cu1Mg0.31 alloy deformed at RT (b) and at 85°C (c) and strain rate of 5∙10-4 s-1.

The microstructure of the as-cast ZnAl4Cu1Mg0.31 alloy is mainly comprised of (i) dendritic primary η-Zn phase with dendrite sizes ranging from 26 to 70 µm, (ii) globular η-Zn + α-Al eutectoid structures in between the η-Zn dendrites with a lamellar spacing of 0.1 – 0.34 µm, (iii) a eutectic structure consisting of secondary η-Zn phase and a eutectoid η-Zn + α-Al lamellar structure, (iv) Mg2Zn11 precipitates with an average size of 0.40 – 0.89 µm, which are located in between primary η-Zn phase and eutectoid η-Zn +α-Al structures, cf. 4.1.1. Figure 7.2 shows the EDS results of the as-cast alloy ZnAl4Cu1Mg0.31. It is evident from the micrograph that elongated precipitates features which contain a high amount of Al (EDS spots 2 and 3 in Figure 7.2) formed in the primary η phase. However, it is difficult to identify these precipitates due to the limited resolution of SEM.

102 7. Precipitation behaviour

Figure 7.2. EDS analysis of alloy ZnAl4Cu1Mg0.31.

Therefore, the chemical composition and structure of these precipitates in the primary η-Zn phase were further quantitatively determined using TEM and APT analyses. Figure 7.3 shows the TEM micrographs of primary η-Zn grains in alloy ZnAl4Cu1Mg0.31 after tensile deformation at RT and 85°C (corresponding to the stress-strain curves in Figure 7.1). Two types of precipitates were observed in the primary η-Zn grains: (i) elongated lenticular precipitates with FCC crystal structure (Figure 7.3a, c) and (ii) needle-shaped precipitates with hexagonal crystal structure (Figure 7.3b, d). Compared to the sample deformation at RT (Figure 7.3a, b), both types of precipitates show a reduction in size after deformation at 85°C (Figure 7.3c, d). Specifically, the needle-shaped precipitates show a reduction of length from 200 nm (RT) to 50 – 100 nm (85°C) and an increase in spacing from 10 – 20 nm (RT) to 20 – 100 nm (85°C), indicating a dissolution of these needle-shaped precipitates induced by temperature and / or strain.

The lenticular-shaped precipitates form along the basal plane of Zn grains and exhibit a heterogeneous density and dispersion in different primary η-Zn grains. They have similar lattice constants than Al and are assumed to be the Al-rich α’ or β' phase, as reported in [6, 41, 44]. On the other hand, the needle-shaped precipitates are homogeneously distributed within the primary η-Zn grains. Figure 7.3e shows the selected area diffraction patterns (SADP) of the needle-shaped precipitates and the primary η-Zn grain from different zone axes. Identical diffraction pattern of the needle-shaped precipitates are observed after deformation at RT and at 85°C. SADP reveals that the precipitates form on (11̅05) planes of the η-Zn matrix. The lattice constants of the needle-shaped precipitates (a = 2.68-2.69 Å, c = 4.84-4.85 Å) are different from that of η-Zn (a = 2.67 Å, c = 4.95 Å [1, 33]) and not consistent with any reported phases in the Zn-Al-Cu(-Mg) alloy system [6, 41, 44].

103 7. Precipitation behaviour

Figure 7.3. TEM micrographs revealing precipitates in primary η-Zn grains of alloy ZnAl4Cu1Mg0.31 after deformation at RT (a, b) and 85°C (c, d); (a, c) TEM bright field micrographs show lenticular-shaped precipitates; (b, d) TEM dark field micrographs show the needle-shaped precipitates, the micrographs were taken from the spots marked by red circles; (e) diffraction patterns of the needle-shaped precipitates (b, d) and the primary η-Zn grain matrix with zone axis B=[011̅0], B=[0001], B=[112̅0].

Figure 7.4 shows the APT data of the primary η-Zn grains in alloy ZnAl4Cu1Mg0.31 deformed at RT (see the stress-strain curves in Figure 7.1). The overall concentrations of Al and Cu in the measured volume (cylinder along the specimen z-axis with a diameter of 30 nm and a length of 500 nm) amount to 3.29 wt.% (6.6 at.%) Al and 4.1 wt.% (4,05 at.%) Cu, which is far above their equilibrium solubility in Zn at RT. The 3-D elemental maps, Figure 7.4a, b, reveal a clear partitioning of Zn and Al atoms. Quantitative APT analysis of the chemical composition of the Al-enriched regions show the precipitation of pure Al (Figure 7.4d), and thin (~50 nm) elongated Al-enriched features with Al concentration of 25 – 50 at.% (Figure 7.4b, c, e-g) in the primary η-Zn grains. These pure Al precipitates are consistent with the lenticular-shaped precipitates that have similar lattice constants as Al as observed using TEM (Figure 7.3a, c), and the thin elongated Al-enriched features are consistent with the needle-shaped precipitates (Figure 7.3b, d). It is therefore concluded that the needle-shaped precipitates are a super- saturated ZnxAl1-x (x≥0.7) transition phase. In agreement to our study, several metastable

104 7. Precipitation behaviour phases have been reported in Zn-Al and Zn-Al-Cu alloy systems, such as the Zn-enriched α' and β' (FCC) [41, 44], Al-enriched η' (hexagonal) [41, 44] and Al-Zn-Cu ternary T’

(rhombohedral) [6, 41, 44] phases. The observed super-saturated ZnxAl1-x precipitates are assumed to be similar to the Zn rich η' phase [41, 44], where the observed differences in lattice constants are assumed to be caused by the high amounts of super-saturated Al atoms in the

ZnxAl1-x (x≥0.7) phase.

Figure 7.4. 3-D elemental maps of alloy ZnAl4Cu1Mg0.31 deformed at RT showing the 3-D positions of (a, b) Al atoms, (h) Mg atoms and (i) Cu atoms in a primary η-Zn grain. The isoconcentration surfaces of 30 at.% Al (a, b), 0.7 at.% Mg (h) and 18 at.% Cu (i) are also shown. (c-g, j) Plots from different precipitates regions (marked by yellow) give the local chemical compositions.

Further, Mg atoms partition into the Al-enriched regions where the local Mg concentration (0.26 wt.% or 0.7 at.%) reaches nearly the global Mg concentration in the alloy (0.31 wt.%). Cu atoms are observed as ε-CuZn4 precipitates (Figure 7.4i-j) in the primary η-Zn grains. As the global concentration of Cu is low (0.59 wt.%) and diffraction spots of ε phase have not been observed using TEM, it is assumed that the density of these precipitates is negligibly low. The ε-CuZn4

105 7. Precipitation behaviour precipitates have been reported to form in Z-Al alloys with Cu content higher than 2 wt.% and cause dimensional instability [6, 8, 13, 31, 249].

Figure 7.5. TEM bright field micrographs of alloy ZnAl4Cu1Mg0.31 deformed at RT: (a) area close to fracture surface; (b) enlarged images of the area marked by dashed yellow rectangle in (a), the insets show the diffraction patterns of the area marked by yellow circles.

In consistency with our previous investigation where {101̅2}[101̅1̅] deformation twinning and localised basal slip in primary η-Zn grains were observed as the predominant deformation mechanisms of Zn alloys at RT, while twinning, basal and non-basal slip in primary η-Zn grains and grain / phase boundary sliding in the eutectic and eutectoid structures are active in the Zn alloys at elevated temperatures (cf. 4.1.4, 6.2.2), deformation twins are frequently observed in the sample deformed at RT using TEM, Figure 7.5. Figure 7.5b show the diffraction patterns of the areas in the twinning and outside twining, where a misorientation of 86° across the twin boundary is measured, corresponding to the {101̅2}[101̅1̅] compression twining in Zn.

After deformation at 85°C, a high density of basal and 1st order pyramidal dislocations are observed in the sample in the area near fracture, Figure 7.6a. In the high strain sample (area near fracture), Figure 7.6a, it is difficult to differentiate individual dislocation segments and the needle-shaped precipitates due to the high dislocation density caused by high strain. Figure 7.6b-e show dark field images of a low strain sample (area away from the fracture surface) in the sample deformed at 85°C imaged under different two-beam-conditions. Due to the low strain in this area, only few dislocations are observed.

According to the g∙b invisibility criterion, only dislocations with a -component are visible under g=(0002), dislocations with an -component are visible under g=(XXX0), X≠0, and both, and components, are visible under g=(XXXX), X≠0. In consistency to the dislocation arrangements in the high strain region (Figure 7.6a), most dislocations in the low

106 7. Precipitation behaviour strain area are of type and lie on basal and pyramidal planes. Dislocations on basal planes are marked by red arrows and those on pyramidal planes by orange arrows. It is revealed from the dark fields images of the low strain sample, Figure 7.6b-e, that most dislocations are in conjunction with precipitates, i.e. pinned at dislocations. This highly indicates the ability of these metastable ZnxAl1-x (x≥0.7) precipitates to hinder the dislocation motion and, hence, strengthen the material.

Figure 7.6. TEM dark field micrographs of alloy ZnAl4Cu1Mg0.31 after deformation at 85°C: (a) high strain sample (area close to fracture surface) with g=(112̅2); (b) low strain sample (area away from fracture surface) with g=(112̅2), (c) low strain sample with g=(0002),(d) low strain sample with g=(011̅0), (e) low strain sample with g=(011̅1); g: diffraction vector. Red arrows mark the dislocations on basal planes and orange arrows mark the dislocations on pyramidal planes.

The ZnxAl1-x transition phase precipitates partially dissolve during deformation at elevated temperature, Figure 7.3. Further, the ZnxAl1-x precipitates are observed to pin dislocations and, hence, strengthen the material, Figure 7.6. The interaction of these precipitates with dislocations is assumed to partially contribute to the high stress exponents observed in the current ZnAl4Cu1Mg alloys, as discussed in cf. 5.2.1.1, where the actual stress acting on dislocations for creep is lowered as extra stress is required due to the interaction between obstacles and dislocations [208-211]. The decomposition of the non-equilibrium ZnxAl1-x precipitates in conjunction with the thermally activated non-basal dislocation slip and grain /

107 7. Precipitation behaviour phase boundary sliding leads to enhanced ductility of the ZnAl alloys at 85°C. On the other hand, although the kinetics of the decomposition process of these precipitates is not clear from our current study, it is reasonable to assume that the dissolution and decomposition of ZnxAl1-x precipitates would also proceed at RT due to the high homologous temperature of Zn (≈0.43 at RT), resulting in the reported long-term mechanical softening in ZnAl alloys.

7.2. Conclusions

Using TEM and APT the precipitates in a gravity cast ZnAl4Cu1Mg0.31 alloy were characterised in terms of both structure and chemistry. Two types of precipitates are determined in primary η-Zn grains: pure Al precipitates and non-equilibrium ZnxAl1-x (x≥0.7) transition phase precipitates. Deformation at slightly elevated temperatures (85°C) causes the dissolution and decomposition of the ZnxAl1-x precipitates. It is proposed that the dissolution of these transition phase precipitates contributes to the long-term mechanical softening observed in Zn-Al alloys.

108 8. Summarising discussion and concluding remarks

8. Summarising discussion and concluding remarks9

The mechanical properties, microstructures and deformation mechanisms of three eutectic ZnAl4Cu1 alloys with different Mg contents (0.04 wt.%, 0.21 wt.% and 0.31 wt.%) were comprehensively investigated from the macroscopic scale to the microscopic scale at temperatures ranging from room temperature (25°C) to 105°C. The following conclusions are drawn.

8.1. Temperature dependent activity of deformation mechanisms

The microstructure of eutectic ZnAl4Cu1Mg alloys is mainly comprised of dendritic primary η- Zn phase grains, coarse η-Zn + α-Al eutectic lamellar colonies and fine globular η-Zn + α-Al eutectoid lamellar colonies, as well as Mg2Zn11 precipitates that forms in eutectoid colonies, see cf. 4.1.1. The mechanical properties and deformation mechanisms of ZnAl4Cu1 alloys are therefore controlled by both the intrinsic properties of these individual microstructural components as well as the joint effect of these microstructural constituents as an aggregate.

Using nanoindentation in conjunction with post-mortem microstructural analysis into individual microstructural components we show that the predominant deformation mechanisms in the primary η-Zn phase are basal slip and {101̅2}〈101̅1̅〉 deformation twinning at room temperature. At 85°C, additionally non-basal slip is activated in the primary η-Zn phase due to thermal activation. The predominant deformation mechanism in η-Zn + α- Al eutectoid structures is grain and phase boundary sliding at RT and 85°C, cf. 6.2.2. This is due to the high homologous temperature of Zn (T/Tm = 0.43 at RT) as well as the fine grain size in the eutectic and eutectoid structures, where grain boundary sliding and phase boundary sliding as thermal activated processes can be readily activated at slightly elevated temperatures or even at RT.

Using DIC method and in-situ straining experiments in the SEM, the contribution of individual microstructural constituents during the bulk deformation of ZnAl4Cu1Mg alloys was determined: at RT, mainly primary η-Zn phase grains contribute to the global deformation; while both primary η-Zn phase grains and eutectic / eutectoid colonies contribute to the bulk

9 Part of this chapter appeared as articles. The original citations are: Z. Wu, S. Sandlöbes, L. Wu, W. P. Hu, G. Gottstein and S. Korte-Kerzel (2016). "Mechanical behaviour of Zn-Al-Cu-Mg alloys: Deformation mechanisms of as-cast microstructures." Materials Science and Engineering: A 651: 675-687., S. Sandlöbes, Z. Wu, K. Pradeep and S. Korte-Kerzel (2016). "Precipitation and decomposition phenomena in a Zn-Al-Cu-Mg alloy." Materials Letters 175: 27-31., Z. Wu, S. Sandlöbes, Y. Wang, J. S. K.-L. Gibson and S. Korte-Kerzel (2018). "Creep behaviour of eutectic Zn-Al-Cu-Mg alloys." Materials Science and Engineering: A 724: 80-94., Z. Wu, S. Sandlöbes, J. Rao, J. S. K.-L. Gibson, B. Berkels and S. Korte-Kerzel (2018). "Local mechanical properties and plasticity mechanisms in a Zn-Al eutectic alloy." Materials & Design 157: 337-350, and Z. Wu, S. Sandlöbes, J. Rao, J. S. K.-L. Gibson, B. Berkels and S. Korte-Kerzel (2018). "Data on measurement of the strain partitioning in a multiphase Zn-Al eutectic alloy." Data in Brief: https://doi.org/10.1016/j.dib.2018.09.010.

109 8. Summarising discussion and concluding remarks deformation at elevated temperatures (85°C), cf. 4.1.4, 6.2.2, 6.2.4. Therefore, the deformation mechanisms of bulk ZnAl4Cu1Mg alloys room temperature and / or highest strain rate tested (5∙10−4 s-1) are basal slip and {101̅2}[101̅1̅] twinning in the primary η-Zn phase, and at elevated temperatures and / or low strain rate, the deformation mechanisms of the investigated Zn alloys are a mixture of twinning, dislocation motion, including basal and non-basal dislocation slip within the primary η-Zn grains, and grain / phase boundary sliding in the eutectic and eutectoid structures, cf. 4.1.4.

8.2. Micro- and macro-mechanical behaviour

At room temperature, nanoindentation tests into individual microstructural components show

their nanoindentation hardness follows the order: HMg2Zn11 + Eutectoid˃ HEutectoid˃ HEutectic˃ HZn at RT. At 85°C, the eutectic and eutectoid structures have similar hardness values as the primary η-Zn phase.

The combination of these local mechanical properties and deformation mechanisms of individual microstructural constituents, cf. 6.1.1 and 6.2.2, as well as their joint effects, cf. 4.1.4 and 6.2.4, results in the observed two distinct deformation regimes of the ZnAl4Cu1 alloys investigated depending on temperature and strain rate, cf. 4.1.2:

 At low temperature and / or highest strain rate tested (5∙10−4 s-1), the alloys are brittle and show work hardening. The observed brittleness of these alloys is caused by the intrinsic brittleness of Zn grains as the number of available deformation modes are limited.  At elevated temperature and / or low strain rates, the alloys are ductile and show pronounced working softening. The observed enhanced ductility of ZnAl4Cu1 alloys is proposed to be caused by both the increased ductility of primary η-Zn grains from a higher amount of available deformation modes, as well as the contribution of η-Zn + α-Al eutectic / eutectoid structures.

8.3. Fracture behaviour

The bulk ZnAl4Cu1 alloys were observed to undergo a rather abrupt transition from brittle to ductile behaviour in a very short temperature range (from 25°C to 85°C).

At room temperature, the ZnAl4Cu1Mg alloys fractured in a transgranular brittle manner featured with large and smooth cleavage facets, with the cracks form either inside the primary η-Zn grains or at the cavities deriving from the casting process, cf.4.1.3. Most of the cracks have been observed to propagate along the basal (0001) plane inside primary η-Zn grains,

110 8. Summarising discussion and concluding remarks showing that the basal plane is the cleavage plane in Zn at RT. It has been reported that cleavage in Zn occurs preferentially along basal planes due to the relatively small critical cleavage stress in the basal plane, arising from local stress concentrations by basal dislocation pile-ups and tilt boundaries [250-252]. On the other hand, the presence of several dimples have been observed in the η-Zn + α-Al eutectic and eutectoid regions in the fracture surface of ZnAl4Cu1 alloys indicating that these microstructural components also contribute to the bulk deformation of ZnAl4Cu1 alloys at RT, however, to a very limited extent.

In contrast, at 55°C, mixed fracture features as cleavage facets, dimples and intergranular cracks as well as a clear transition from brittle fracture to ductile fracture was observed with decreasing strain rate in the bulk ZnAl4Cu1 alloys, cf. 4.1.3. At elevated temperatures (85°C and higher), the ZnAl4Cu1 alloys displayed a high level of ductility with ductile fracture featured by fracture surfaces comprising of only dimples, cf. 4.1.3. Instead of cleavage cracks which are responsible for room temperature fracture in ZnAl4Cu1 Mg alloys, tearing cracks and shearing cracks were observed at elevated temperatures. Most of the cracks were observed to initiate at macroscopic casting defects such as cavities where the strain is localised, or at colony boundaries between the primary η-Zn phase grains and η-Zn + α-Al eutectic / eutectoid colonies. It has been discussed in cf. 6.2.2 that plastic strain can either easily cross some primary η-Zn grains and eutectic / eutectoid colonies where the neighbouring regions across boundaries are well aligned for slip transfer (high Luster and Morris parameters), or accumulate at the boundaries between some primary η-Zn grains and eutectic / eutectoid colonies where the adjacent regions are oriented unfavourable for slip transfer (low Luster and Morris parameters). In the last case, the high strain localisation at the boundaries has been assumed to attribute to the crack formation in the ZnAl4Cu1 alloys.

8.4. Creep behaviour

The ZnAl4Cu1Mg alloys investigated show stress exponents of 6.9 – 8 and activation energies of 93 – 104 kJ/mol at temperatures of 25 – 105°C and stresses of 61 – 130 MPa. It is concluded that the rate controlling mechanism in ZnAl4Cu1Mg alloys is stress assisted dislocation climb in the primary η-Zn phase.

Further, nanoindentation creep tests into individual microstructural components show that the increase of indentation depth at the same load during nanoindentation creep has the order:

hMg2Zn11 + Eutectoid˃ hEutectoid˃⁡hEutectic ˃hZn at both 25°C and 85°C, where primary η-Zn shows the highest creep resistance, and η-Zn + α-Al eutectoid structures containing Mg2Zn11 precipitates show the lowest creep resistance in nanoindentation creep tests. It is assumed that during macroscopic creep deformation, the eutectic / eutectoid structures creep-deform via Coble creep or grain / phase boundary sliding which is not rate controlling due to

111 8. Summarising discussion and concluding remarks geometrical constrains.

8.5. Effect of Mg

In general, the effects of Mg on the microstructure and mechanical properties of the ZnAl4Cu1 alloys investigated are related to (i) solute solution atoms in η-Zn and α-Al, (ii) grain and phase boundaries segregation, and (iii) Mg2Zn11 phase precipitates which decorate the eutectoid structures, cf. 5.2.3. Further, the addition of Mg promotes the formation of Mg2Zn11 precipitates and concurrently causes the formation of a higher volume fraction of isolated globular η-Zn + α-Al eutectoid structure regions with refined lamellar structures in Zn-Al alloys, cf. 4.1.1.

Dilute Mg alloying increases the strength and the ductility of Zn4Al1Cu alloys, particularly at elevated temperatures, assumed to be caused by microstructural refinement of the eutectoid structures. Further, dilute Mg alloying between 0.21 wt.% and 0.31 wt.% decreases the creep resistance of Zn-Al alloys due to increased grain / phase boundary activities and reduced geometrical constraints of eutectic and eutectoid colonies.

8.6. Precipitates and mechanical behaviour

Using TEM and APT we structurally and chemically characterise non-equilibrium ZnxAl1-x (x≥0.7) transition phase precipitates and pure Al precipitates in the primary η-Zn grains in as- cast ZnAl4Cu1Mg0.31 alloy. The ZnxAl1-x transition phase precipitates are observed to pin dislocations and partially dissolve after tensile deformation at elevated temperature precipitates, cf. 7.1. It is therefore proposed that the dissolution of these transition phase precipitates contributes to the long-term mechanical softening of Zn-Al alloys that limits their wider application.

On the other hand, the constant nanoindentation creep exponents of the primary η-Zn phase during indentation creep at RT and 85°C (cf. 5.1.3) suggest that these precipitates remained stable during nanoindentation creep. Further, high stress exponents in the ZnAl4Cu1Mg alloys in the investigated temperature range (25 – 105°C) indicate that these precipitates have effectively hindered the dislocation movement.

However, the kinetics of the decomposition of these ZnxAl1-x (x≥0.7) transition phase precipitates are still not entirely clear from our current study, and a remaining question is that to which extent the precipitates could block the dislocation movement. Therefore, it is suggested that future studies could focus on the kinetics of the decomposition process and the effect of these metastable ZnxAl1-x precipitates, as well as to stabilise the metastable phases by alloying or heat treatment to improve the long-term stability of Zn-Al alloys.

112 Bibliography

Bibliography

[1] F.C. Porter, Zinc handbook: properties, processing, and use in design, 1 ed., CRC Press, New York, 1991. [2] M.H. Yoo, Slip, twinning, and fracture in hexagonal close-packed metals, Metallurgical Transactions A 12(3) (1981) 409-418. [3] M.H. Yoo, J.K. Lee, Deformation twinning in h.c.p. metals and alloys, Philosophical Magazine A 63(5) (1991) 987-1000. [4] M. Aydın, T. Savaşkan, Fatigue properties of zinc-aluminium alloys in 3.5% NaCl and 1% HCl solutions, International Journal of Fatigue 26(1) (2004) 103-110. [5] R. Mahmudi, S. Alibabaie, Elevated-temperature shear strength and hardness of Zn–3Cu– xAl ultra-high-temperature lead-free solders, Materials Science and Engineering: A 559 (2013) 421-426. [6] H.J. Dorantes-Rosales, V.M. Lopez-Hirata, J.L. Mendez-Velazquez, M.L. Saucedo-Munoz, D. Hernandez-Silva, Microstructure characterization of phase transformations in a Zn-22 wt%Al-2 wt%Cu alloy by XRD, SEM, TEM and FIM, Journal of Alloys and Compounds 313(1) (2000) 154-160. [7] S.-J. Kim, K.-S. Kim, S.-S. Kim, C.-Y. Kang, K. Suganuma, others, Characteristics of Zn- Al-Cu alloys for high temperature solder application, Materials Transactions 49(7) (2008) 1531- 1536. [8] R. Mojaver, H.R. Shahverdi, Relationship between cooling rate, microstructure features and wear behavior in end-chill cast Zn–27%Al alloys containing more than 2% Cu, Wear 271(11- 12) (2011) 2899-2908. [9] B. Krupinska, L.A. Dobrzanski, Z.M. Rdzawski, K. Labisz, Cooling rate influence on microstructure of the Zn-Al cast alloy, Archives of Materials Science 14 (2010) 13-20. [10] M. Gelfi, E. Bontempi, A. Pola, R. Roberti, D. Rollez, L.E. Depero, Microstructural and mechanical properties of zinc die casting alloys, Advanced Engineering Materials 6(10) (2004) 818-822. [11] M. Al-Maharbi, I. Karaman, G. Purcek, Flow response of a severe plastically deformed two-phase zinc–aluminum alloy, Materials Science and Engineering: A 527(3) (2010) 518-525. [12] T. Savaşkan, A.P. Hekimoğlu, Microstructure and mechanical properties of Zn–15Al- based ternary and quaternary alloys, Materials Science and Engineering: A 603 (2014) 52-57. [13] T. Savaşkan, G. Pürçek, S. Murphy, Sliding wear of cast zinc-based alloy bearings under static and dynamic loading conditions, Wear 252(9–10) (2002) 693-703. [14] A.A. Mir, The creep properties of a series of zinc-rich zinc-aluminium alloys. Doctoral Thesis, 1998, The University of Aston in Birmingham. [15] M. Durman, S. Murphy, Precipitation of metastable ϵ-phase in a hypereutectic zinc- aluminium alloy containing copper, Acta Metallurgica et Materialia 39(10) (1991) 2235-2242. [16] Z. Wu, S. Sandlobes, L. Wu, W.P. Hu, G. Gottstein, S. Korte-Kerzel, Mechanical behaviour of Zn-Al-Cu-Mg alloys: Deformation mechanisms of as-cast microstructures, Materials Science and Engineering: A 651 (2016) 675-687. [17] M. Anwar, S. Murphy, Creep kinetics in compression of sand cast commercial Zn–Al alloys, Materials Science and Technology 16(3) (2000) 321-327. [18] S. Murphy, M. Durman, J. Hill, Kinetics of creep in pressure diecast commercial zinc- aluminium alloys, Zeitschrift für Metallkunde 79(4) (1988) 243-247. [19] G. Pürçek, Improvement of mechanical properties for Zn–Al alloys using equal-channel angular pressing, Journal of Materials Processing Technology 169(2) (2005) 242-248.

113 Bibliography

[20] F. Wang, B. Xiong, Y. Zhang, H. Liu, Z. Li, X. Li, C. Qu, Effect of Cu addition on microstructure and corrosion behavior of spray-deposited Zn–30Al alloy, Materials Science and Engineering: A 532(0) (2012) 100-105. [21] T. Savaşkan, S. Murphy, Decomposition of Zn–Al alloys on quench–aging, Materials Science and Technology 6(8) (1990) 695-704. [22] W.M. Haynes, CRC Handbook of Chemistry and Physics, 96th ed., CRC Press, 2015. [23] T.J. Chen, Y. Hao, J. Sun, Y.D. Li, Effects of Mg and RE additions on the semi-solid microstructure of a zinc alloy ZA27, Science and Technology of Advanced Materials 4(6) (2003) 495-502. [24] T. Gancarz, G. Cempura, W. Skuza, Characterization of ZnAl cast alloys with Na addition, Materials Characterization 111 (2016) 147-153. [25] E.M. da Costa, C.E. da Costa, F.D. Vecchia, C. Rick, M. Scherer, C.A. dos Santos, B.A. Dedavid, Study of the influence of copper and magnesium additions on the microstructure formation of Zn–Al hypoeutectic alloys, Journal of Alloys and Compounds 488(1) (2009) 89- 99. [26] T. Savaşkan, A.P. Hekimoğlu, G. Pürçek, Effect of copper content on the mechanical and sliding wear properties of monotectoid-based zinc-aluminium-copper alloys, Tribology International 37(1) (2004) 45-50. [27] T. Gancarz, G. Cempura, Characterization of ZnAl cast alloys with Li addition, Materials & Design 104 (2016) 51-59. [28] H. Li, Z. Li, Y. Liu, H. Jiang, Effect of zirconium on the microstructure and mechanical properties of Zn–4%Al hypoeutectic alloy, Journal of Alloys and Compounds 592(0) (2014) 127-134. [29] G. Purcek, B.S. Altan, I. Miskioglu, P.H. Ooi, Processing of eutectic Zn–5% Al alloy by equal-channel angular pressing, Journal of Materials Processing Technology 148(3) (2004) 279-287. [30] M. Demirtas, G. Purcek, H. Yanar, Z.J. Zhang, Z.F. Zhang, Effect of equal-channel angular pressing on room temperature superplasticity of quasi-single phase Zn–0.3Al alloy, Materials Science and Engineering: A 644 (2015) 17-24. [31] B.K. Prasad, Microstructural alterations through heat treatment and its influence on wear response of a silicon containing zinc based alloy under different test conditions, Materials Science and Technology 19(3) (2003) 327-335. [32] Y. Xun, R. Rodriguez, E.J. Lavernia, F.A. Mohamed, Processing and microstructural evolution of powder metallurgy Zn-22 Pct Al eutectoid alloy containing nanoscale dispersion particles, Metallurgical and Materials Transactions A 36(10) (2005) 2849-2859. [33] L. Pugazhenthy, Zinc Handbook: Properties, Processing, and Use In Design, Taylor & Francis, 1991. [34] B.K. Prasad, A.K. Patwardhan, A.H. Yegneswaran, Microstructural modifications through compositional alterations and their influence on the mechanical and sliding wear properties of zinc-based alloys, Acta metallurgica 37(3) (1997) 323-328. [35] M.A. El-Khair, A. Daoud, A. Ismail, Effect of different Al contents on the microstructure, tensile and wear properties of Zn-based alloy, Materials Letters 58(11) (2004) 1754-1760. [36] A. Türk, M. Durman, E.S. Kayali, The effect of Cu and Al on the mechanical properties of gravity-cast hyper-eutectic Zn–Al-based alloys, Zeitschrift für Metallkunde 94(9) (2003) 1001- 1005. [37] S. Yan, J. Xie, Z. Liu, W. Wang, A. Wang, J. Li, Influence of different Al contents on microstructure, tensile and wear properties of Zn-based alloy, Journal of Materials Science & Technology 26(7) (2010) 648-652.

114 Bibliography

[38] H. Baker, H. Okamoto, ASM Handbook, Volume 03: Alloy phase diagrams, ASM International, 1992. [39] E. Gebhard, Über den Aufbau des Systems Zink-Aluminium-Kupfer und die Volumenänderung der Gusslegierung, Giesserei 29, Heft 24 (1942) 7. [40] L.H. Kallien, W. Leis, Ageing of zink alloys, Giessereiforschung including CD ROM 63(1) (2011) 2-2. [41] Y.h. Zhu, General rule of phase decomposition in Zn-Al based alloys (II)-On effects of external stresses on phase transformation, Materials Transactions 45(11) (2004) 3083-3097. [42] H. Opitz, Allgemeine Werkstoffkunde für Ingenieurschulen, 6 ed., VEB Fachbuchverlauf, Leipzig, 1971. [43] Y.H. Zhu, H.C. Man, H.J. Dorantes-Rosales, W.B. Lee, Ageing characteristics of furnace cooled eutectoid Zn-Al based alloy, Journal of Materials Science 38(13) (2003) 2925-2934. [44] Y.H. Zhu, J.J. Islas, Microstructures and dimensional stability of extruded eutectoid Zn-Al alloy, Journal of Materials Processing Technology 66(1–3) (1997) 244-248. [45] H. LeHuy, G. L'Espérance, Ageing characteristics of dendritic and non-dendritic (stir-cast) Zn-Al alloy (ZA-27), Journal of Materials Science 26(2) (1991) 559-568. [46] N. Mykura, Y.H. Zhu, S. Murphy, Solid-state reactions in Zn-Al based alloys, Canadian Metallurgical Quarterly 25(2) (1986) 151-159. [47] N. Kang, H.S. Na, S.J. Kim, C.Y. Kang, Alloy design of Zn–Al–Cu solder for ultra high temperatures, Journal of Alloys and Compounds 467(1–2) (2009) 246-250. [48] L.R. Xiao, X.M. Zhang, Y. Wang, W. Li, Q.S. Sun, Z.J. Geng, High Temperature Creep Behavior of Zn-1.0Cu-0.2Ti Alloy, Advanced Materials Research 287-290 (2011) 769-776. [49] K. Löhberg, The changements of the volumes of Zinc-Aluminium-Copper alloys connected with the transition equilibrium ε+α-η+Τ', Zeitschrift für Metallkunde 74 (1983) 2. [50] B. Wattiez, A.F. Gourgues, A. Deschamps, A. Roemer, Z. Zermout, Experimental investigation of microstructure and ageing behaviour of bulk Zn–(1–18)wt% Al–(0–0.06)wt% Mg alloys, Materials Science and Engineering: A 527(29-30) (2010) 7901-7911. [51] C. Yao, Z. Wang, S.L. Tay, T. Zhu, W. Gao, Effects of Mg on microstructure and corrosion properties of Zn–Mg alloy, Journal of Alloys and Compounds 602 (2014) 101-107. [52] Z. Liu, D. Qiu, F. Wang, J.A. Taylor, M. Zhang, Effect of grain refinement on tensile properties of cast zinc alloys, Metallurgical and Materials Transactions A 47(2) (2016) 830-841. [53] S. Alibabaie, R. Mahmudi, Microstructure and creep characteristics of Zn–3Cu–xAl ultra high-temperature lead-free solders, Materials & Design 39 (2012) 397-403. [54] C.H. Mathewson, Zinc: the science and technology of the metal, its alloys and compounds, Reinhold Publishing Corporation, New York, 1959. [55] A.c. Tolcin, 2014 Minerals Yearbook, in: U.S.G. Survey (Ed.) 2016. [56] American Galvanizers Association, Zinc Uses. , (accessed 22.08.2017). [57] S. Sohn, T. Kang, The effects of tin and nickel on the corrosion behavior of 60Cu–40Zn alloys, Journal of Alloys and Compounds 335(1–2) (2002) 281-289. [58] P. Chollet, Recent progress in the metallurgy of wrought-zinc alloys, Canadian Metallurgical Quarterly 7(3) (2013) 177-185. [59] G. Gottstein, Physical foundations of materials science, 1 ed., Springer-Verlag Berlin Heidelberg, 2004. [60] J. Rösler, H. Harders, M. Baeker, Mechanical behaviour of engineering materials: metals, ceramics, polymers, and composites, Springer Science & Business Media, 2007. [61] Y. Yu, Theory of metals, 2 ed., Metallurgy Industry Press, Beijing 2013. [62] D. Hull, D.J. Bacon, Chapter 3 - Movement of dislocations, introduction to dislocations (Fifth Edition), Butterworth-Heinemann, Oxford, 2011, pp. 43-62.

115 Bibliography

[63] G.I. Taylor, The mechanism of plastic deformation of crystals. Part I. Theoretical, Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences 145(855) (1934) 362-387. [64] G. Masing, Grundlagen der Metallkunde in anschaulicher Darstellung, Springer-Verlag Berlin Heidelberg, 1955. [65] S. Korte, Plasticity in ceramics-deformation in small volumes. Doctoral Thesis, 2010, University of Cambridge. [66] R. Peierls, The size of a dislocation, Proceedings of the Physical Society 52(1) (1940) 34. [67] F.R.N. Nabarro, Dislocations in a simple cubic lattice, Proceedings of the Physical Society 59(2) (1947) 256. [68] B. Joós, M.S. Duesbery, The Peierls stress of dislocations: An analytic formula, Physical Review Letters 78(2) (1997) 266-269. [69] B.D. Cullity, S.R. Cullity, S. Stock, Elements of X-ray diffraction, 2001. [70] G.E. Dieter, D.J. Bacon, Mechanical metallurgy, McGraw-Hill, New York 1988. [71] E. Meza García, Influence of alloying elements on the microstructure and mechanical properties of extruded Mg-Zn based alloys. Doctoral Thesis, 2010, Technische Universität Berlin. [72] G.W. Groves, A. Kelly, Independent slip systems in crystals, Philosophical Magazine 8(89) (1963) 877-887. [73] R. Parisot, S. Forest, A. Pineau, F. Grillon, X. Demonet, J.-M. Mataigne, Deformation and damage mechanisms of zinc coatings on hot-dip galvanized steel sheets: Part I. Deformation modes, Metallurgical and Materials Transactions A 35(3) (2004) 797-811. [74] M. Yoo, C. Wei, Slip modes of hexagonal‐close‐packed metals, Journal of Applied Physics 38(11) (1967) 4317-4322. [75] D. Caillard, J.-L. Martin, Thermally activated mechanisms in crystal plasticity, Elsevier, 2003. [76] J. Hess, C. Barrett, Structure and nature of kink bands in zinc, Tranctions of the American Institute of Mining and Metallurgical Engineers 185(9) (1949) 599-606. [77] K. Hagihara, M. Honnami, R. Matsumoto, Y. Fukusumi, H. Izuno, M. Yamasaki, T. Okamoto, T. Nakano, Y. Kawamura, In-situ observation on the formation behavior of the deformation kink bands in Zn single crystal and LPSO phase, Materials Transactions 56(7) (2015) 943-951. [78] J. Washburn, E. Parker, Kinking in zinc single-crystal tension specimens, J. Metals, 4, Tranctions of the American Institute of Mining and Metallurgical Engineers 194 (1952) 1076- 1078. [79] K. Hagihara, T. Mayama, M. Honnami, M. Yamasaki, H. Izuno, T. Okamoto, T. Ohashi, T. Nakano, Y. Kawamura, Orientation dependence of the deformation kink band formation behavior in Zn single crystal, International Journal of Plasticity 77 (2016) 174-191. [80] T.B. Anderson, Kink bands, structural geology and tectonics, Springer Berlin Heidelberg, Berlin, Heidelberg, 1987, pp. 373-377. [81] R. Chen, S. Sandlöbes, X. Zeng, D. Li, S. Korte-Kerzel, D. Raabe, Room temperature deformation of LPSO structures by non-basal slip, Materials Science and Engineering: A 682 (2017) 354-358. [82] B.A. Bilby, A.G. Crocker, The theory of the crystallography of deformation twinning, Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 288(1413) (1965) 240-255. [83] H.J. Frost, F. Ashby, Deformation-mechanism maps: the plasticity and creep of metals and ceramics, Pergamon Press, 1982.

116 Bibliography

[84] J.W. Christian, S. Mahajan, Deformation twinning, Progress in Materials Science 39(1) (1995) 1-157. [85] D.R. Steinmetz, T. Jäpel, B. Wietbrock, P. Eisenlohr, I. Gutierrez-Urrutia, A. Saeed–Akbari, T. Hickel, F. Roters, D. Raabe, Revealing the strain-hardening behavior of twinning-induced plasticity steels: Theory, simulations, experiments, Acta Materialia 61(2) (2013) 494-510. [86] M.A. Meyers, O. Vöhringer, V.A. Lubarda, The onset of twinning in metals: a constitutive description, Acta Materialia 49(19) (2001) 4025-4039. [87] P.G. Partridge, The crystallography and deformation modes of hexagonal close-packed metals, Metallurgical Reviews 12(1) (1967) 169-194. [88] R. Parisot, S. Forest, A.F. Gourgues, A. Pineau, D. Mareuse, Modeling the mechanical behavior of a multicrystalline zinc coating on a hot-dip galvanized steel sheet, Computational Materials Science 19(1) (2000) 189-204. [89] M.R. Barnett, Twinning and the ductility of magnesium alloys: Part I: “Tension” twins, Materials Science and Engineering: A 464(1) (2007) 1-7. [90] G. Dirras, J. Gubicza, H. Couque, A. Ouarem, P. Jenei, Mechanical behaviour and underlying deformation mechanisms in coarse- and ultrafine-grained Zn over a wide range of strain rates, Materials Science and Engineering: A 564 (2013) 273-283. [91] H. Li, Q.Q. Duan, X.W. Li, Z.F. Zhang, Compressive and fatigue damage behavior of commercially pure zinc, Materials Science and Engineering: A 466(1–2) (2007) 38-46. [92] G.M. Hughes, G.E. Smith, P.E.J. Flewitt, A.G. Crocker, The brittle fracture of polycrystalline zinc, Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences 463(2085) (2007) 2129-2151. [93] T.G. Langdon, An evaluation of the strain contributed by grain boundary sliding in superplasticity, Materials Science and Engineering A 174(2) (1994) 225-230. [94] X. Zhang, H. Wang, R.O. Scattergood, J. Narayan, C.C. Koch, A.V. Sergueeva, A.K. Mukherjee, Studies of deformation mechanisms in ultra-fine-grained and nanostructured Zn, Acta Materialia 50(19) (2002) 4823-4830. [95] C. Pearson, The viscous properties of extruded eutectic alloys of lead-tin and bismuth-tin, Journal of the Institute of Metals 54(1) (1934) 111-124. [96] T. Uesugi, M. Kawasaki, M. Ninomiya, Y. Kamiya, Y. Takigawa, K. Higashi, Significance of Si impurities on exceptional room-temperature superplasticity in a high-purity Zn-22%Al alloy, Materials Science and Engineering: A 645 (2015) 47-56. [97] F.A. Mohamed, Interpretation of superplastic flow in terms of a threshold stress, Journal of Materials Science 18(2) (1983) 582-592. [98] M. Cetin, M. Demirtas, H. Sofuoglu, Ö. Cora, G. Purcek, Effects of grain size on room temperature deformation behavior of Zn–22Al alloy under uniaxial and biaxial loading conditions, Materials Science and Engineering: A 672 (2016) 78-87. [99] P.K. Chaudhury, F.A. Mohamed, Effect of impurity content on superplastic flow in the Zn- 22% Al alloy, Acta Metallurgica 36(4) (1988) 1099-1110. [100] K. Duong, F.A. Mohamed, Effect of impurity content on boundary sliding behavior in the superplastic Zn–22% Al alloy, Acta Materialia 46(13) (1998) 4571-4586. [101] M. Kawasaki, T.G. Langdon, Grain boundary sliding in a superplastic zinc-aluminum alloy processed using severe plastic deformation, Materials Transactions 49(1) (2008) 84-89. [102] T. Watanabe, Grain boundary design and control for high temperature materials, Materials Science and Engineering A 166(1-2) (1993) 11-28. [103] J.H. Liu, C.X. Huang, S.D. Wu, Z.F. Zhang, Tensile deformation and fracture behaviors of high purity polycrystalline zinc, Materials Science and Engineering: A 490(1–2) (2008) 117- 125.

117 Bibliography

[104] P. Jenei, J. Gubicza, G. Dirras, J.L. Lábár, D. Tingaud, Indentation creep study on ultrafine-grained Zn processed by powder metallurgy, Materials Science and Engineering: A 596 (2014) 170-175. [105] Y. Xun, F.A. Mohamed, Microstructure and deformation in the spray deposited Zn–22% Al alloy, Materials Science and Engineering: A 432(1–2) (2006) 261-268. [106] M. Demirtas, G. Purcek, H. Yanar, Z.J. Zhang, Z.F. Zhang, Improvement of high strain rate and room temperature superplasticity in Zn–22Al alloy by two-step equal-channel angular pressing, Materials Science and Engineering: A 620(0) (2015) 233-240. [107] P. Kumar, C. Xu, T.G. Langdon, Mechanical characteristics of a Zn–22% Al alloy processed to very high strains by ECAP, Materials Science and Engineering: A 429(1–2) (2006) 324-328. [108] P. Chaudhury, K.-T. Park, F. Mohamed, Effect of Fe on the superplastic deformation of Zn-22 pct Al, Metallurgical and Materials Transactions A 25(11) (1994) 2391-2401. [109] P. Shariat, R.B. Vastava, T.G. Langdon, An evaluation of the roles of intercrystalline and interphase boundary sliding in two-phase superplastic alloys, Acta Metallurgica 30(1) (1982) 285-296. [110] A. Yousefiani, J.C. Earthman, F.A. Mohamed, Formation of cavity stringers during superplastic deformation, Acta Materialia 46(10) (1998) 3557-3570. [111] T. Tanaka, S.W. Chung, L.-F. Chaing, K. Makii, A. Kushibe, M. Kohzu, K. Higashi, On applying superplastic Zn-22 wt.% Al alloy with nanocrystalline grains to general residential seismic dampers, Materials Science and Engineering: A 410–411 (2005) 109-113. [112] T. Tanaka, M. Kohzu, Y. Takigawa, K. Higashi, Low cycle fatigue behavior of Zn–22mass% Al alloy exhibiting high-strain-rate superplasticity at room temperature, Scripta Materialia 52(3) (2005) 231-236. [113] Y. Huang, T.G. Langdon, Characterization of deformation processes in a Zn-22% Al alloy using atomic force microscopy, Journal of Materials Science 37(23) (2002) 4993-4998. [114] F.A. Mohamed, T.G. Langdon, Creep at low stress levels in the superplastic Zn-22% Al eutectoid, Acta Metallurgica 23(1) (1975) 117-124. [115] F.A. Mohamed, S. Shen-Ann, T.G. Langdon, The activation energies associated with superplastic flow, Acta Metallurgica 23(12) (1975) 1443-1450. [116] H. Ishikawa, F.A. Mohamed, T.G. Langdon, The influence of strain rate on ductility in the superplastic Zn–22% Al eutectoid, Philosophical Magazine 32(6) (1975) 1269-1271. [117] M.I. Ahmed, F. Mohamed, T. Langdon, Neck formation and cavitation in the superplastic Zn-22% Al eutectoid, Journal of Materials Science 14(12) (1979) 2913-2918. [118] F. Mohamed, M.I. Ahmed, T. Langdon, Factors influencing ductility in the superplastic Zn-22 Pct Al eutectoid, Metallurgical Transactions A 8(6) (1977) 933-938. [119] T.K. Ha, J.R. Son, W.B. Lee, C.G. Park, Y.W. Chang, Superplastic deformation of a fine- grained Zn–0.3 wt.% Al alloy at room temperature, Materials Science and Engineering: A 307(1) (2001) 98-106. [120] H. Naziri, R. Pearce, Superplasticity in a Zn-0.4% Al alloy, Acta metallurgica 22(11) (1974) 1321-1330. [121] T.K. Ha, H.W. Koo, Y.W. Chang, Temperature and grain size dependence of superplasticity in a Zn-0.3 wt.% Al alloy, Metals and Materials International 9(1) (2003) 29-33. [122] T. Tanaka, K. Makii, A. Kushibe, K. Higashi, Room temperature deformation behavior of Zn-22 mass%Al alloy with nanocrystalline structure, Materials Transactions 43(10) (2002) 2449-2454. [123] M. Kawasaki, H.-J. Lee, I.-C. Choi, J.-i. Jang, B. Ahn, T.G. Langdon, Evolution of hardness, microstructure, and strain rate sensitivity in a Zn-22% Al eutectoid alloy processed

118 Bibliography by high-pressure torsion, IOP Conference Series: Materials Science and Engineering, IOP Publishing, 2014, p. 012101. [124] A.K. Mukherjee, J.E. Bird, J.E. Dorn, Experimental correlations for high-temperature creep, Asm Transactions Quarterly 62(1) (1969) 155-179. [125] A.K. Mukherjee, An examination of the constitutive equation for elevated temperature plasticity, Materials Science and Engineering: A 322(1-2) (2002) 1-22. [126] G.S. Murthy, D.H. Sastry, Impression creep of zinc and the rate-controlling dislocation mechanism of plastic flow at high temperatures, Physica Status Solidi (A) 70(1) (1982) 63-71. [127] R. Roberti, A. Pola, M. Gilles, D. Rollez, Primary and steady state creep deformation in Zamak5 die-casting alloy at 80°C, Materials Characterization 59(12) (2008) 1747-1752. [128] J.M. Gobien, Creep properties of a zinc-aluminum die-casting alloy as a function of grain size. Doctoral Thesis, 2010, North Carolina State Universtity. [129] T. Matsunaga, T. Kameyama, K. Takahashi, E. Sato, Constitutive relation for ambient- temperature creep in hexagonal close-packed metals, Materials Transactions 50(12) (2009) 2858-2864. [130] J.E. Flinn, D.E. Munson, Stress dependence of the transition behaviour in multi- mechanism creep reactions: With special reference to zinc, Philosophical Magazine 10(107) (1964) 861-870. [131] W.J.M. Tegart, O.D. Sherby, Activation energies for high temperature creep of polycrstalline zinc, Philosophical Magazine 3(35) (1958) 1287-1296. [132] N.L. Peterson, S.J. Rothman, Isotope effect in self-diffusion in zinc, Physical Review 163(3) (1967) 645-649. [133] J.J. Gilman, Plastic anisotropy of zinc monocrystals, JOM 8(10) (1956) 1326-1336. [134] T. Matsunaga, T. Kameyama, S. Ueda, E. Sato, New creep region and mechanism in hexagonal close-packed metals, Journal of Physics: Conference Series 240(1) (2010). [135] M. Durman, The creep behaviour of pressure diecast zinc-aluminium based alloys. Doctoral Thesis, 1989, Aston University. [136] J.M. Gobien, K.L. Murty, R.O. Scattergood, F. Goodwin, C.C. Koch, Creep behavior of ultra-fine grained Zn–4.5Al, Materials Science and Engineering: A 527(27-28) (2010) 7382- 7386. [137] N. Prasad, G. Malakondaiah, D. Banerjee, P. Rama Rao, Low-stress creep behaviour of superplastic Zn-22% Al alloy, Journal of Materials Science 28(6) (1993) 1585-1594. [138] A. Arieli, A.K.S. Yu, A.K. Mukherjee, Low stress and superplastic creep behavior of Zn- 22 Pct Al eutectoid alloy, Metallurgical and Materials Transactions A 11(1) (1980) 181-191. [139] P.S. Godavarti, K.L. Murty, Creep anisotropy of zinc using impression tests, Journal of Materials Science Letters 6(4) (1987) 456-458.

[140] A. Kraych, P. Carrez, P. Cordier, On dislocation glide in MgSiO3 bridgmanite at high- pressure and high-temperature, Earth and Planetary Science Letters 452(Supplement C) (2016) 60-68. [141] E. Orowan, Problems of plastic gliding, Proceedings of the Physical Society 52(1) (1940) 8. [142] R.L. Bell, R.W. Cahn, The Dynamics of Twinning and the Interrelation of Slip and Twinning in Zinc Crystals, Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences 239(1219) (1957) 494-521. [143] J.J. Fundenberger, M.J. Philippe, F. Wagner, C. Esling, Modelling and prediction of mechanical properties for materials with hexagonal symmetry (zinc, titanium and zirconium alloys), Acta Materialia 45(10) (1997) 4041-4055. [144] K.H. Adams, Dislocation mobility and density in zinc single crystals. Doctoral Thesis, 1965, California Institute of Technology.

119 Bibliography

[145] Y.M. Soifer, V. Shteinberg, Mechanism of the low temperature plastic deformation in zinc single crystals, Physica Status Solidi (A) 10(2) (1972) K113-K115. [146] P. Rodriguez, Grain size dependence of the activation parameters for plastic deformation: Influence of crystal structure, slip system, and rate-controlling dislocation mechanism, Metallurgical and Materials Transactions A 35(9) (2004) 2697-2705. [147] I.V. Gektina, F.F. Lavrentev, N.D. Natsik, Thermally activated motion of pyramidal dislocations through Peierls barriers in a zinc single-crystal, Journal of Experimental and Theoretical Physics 79(5) (1980) 1927-1935. [148] S. Sandlobes, Z. Wu, K.G. Pradeep, S. Korte-Kerzel, Precipitation and decomposition phenomena in a Zn-Al-Cu-Mg alloy, Materials Letters 175 (2016) 27-31. [149] T.B. Britton, F.P.E. Dunne, A.J. Wilkinson, On the mechanistic basis of deformation at the microscale in hexagonal close-packed metals, Proceedings of the Royal Society A: Mathematical, Physical and Engineering Science 471(2178) (2015) 20140881. [150] K. Durst, V. Maier, Dynamic nanoindentation testing for studying thermally activated processes from single to nanocrystalline metals, Current Opinion in Solid State and Materials Science 19(6) (2015) 340-353. [151] D. Torre, E.V. Pereloma, C.H.J. Davies, Strain rate sensitivity and apparent activation volume measurements on equal channel angular extruded Cu processed by one to twelve passes, Scripta Materialia 51(5) (2004) 367-371. [152] E.W. Hart, Theory of the tensile test, Acta Metallurgica 15(2) (1967) 351-355. [153] N.Q. Chinh, T. Csanadi, T. Gyori, R.Z. Valiev, B.B. Straumal, M. Kawasaki, T.G. Langdon, Strain rate sensitivity studies in an ultrafine-grained Al-30 wt.% Zn alloy using micro- and nanoindentation, Materials Science and Engineering: A 543 (2012) 117-120. [154] J.C. Li, Mechanical properties of nanocrystalline materials, Pan Stanford Publishing Pte. Ltd., Singapore, 2011. [155] I.-C. Choi, Y.-J. Kim, M.-Y. Seok, B.-G. Yoo, J.-Y. Kim, Y. Wang, J.-i. Jang, Nanoscale room temperature creep of nanocrystalline nickel pillars at low stresses, International Journal of Plasticity 41 (2013) 53-64. [156] H. Conrad, Grain size dependence of the plastic deformation kinetics in Cu, Materials Science and Engineering: A 341(1) (2003) 216-228. [157] H. Conrad, Plastic deformation kinetics in nanocrystalline fcc metals based on the pile- up of dislocations, Nanotechnology 18(32) (2007) 325701. [158] I.-C. Choi, Y.-J. Kim, B. Ahn, M. Kawasaki, T.G. Langdon, J.-i. Jang, Evolution of plasticity, strain-rate sensitivity and the underlying deformation mechanism in Zn–22% Al during high-pressure torsion, Scripta Materialia 75 (2014) 102-105. [159] M. Demirtas, G. Purcek, H. Yanar, Z.J. Zhang, Z.F. Zhang, Achieving room temperature superplasticity in Zn–5Al alloy at high strain rates by equal-channel angular extrusion, Journal of Alloys and Compounds 623 (2015) 213-218. [160] M. Kawasaki, T.G. Langdon, Developing superplasticity and a deformation mechanism map for the Zn–Al eutectoid alloy processed by high-pressure torsion, Materials Science and Engineering: A 528(19-20) (2011) 6140-6145. [161] T.G. Langdon, F.A. Mohamed, The activation energies for superplasticity, Scripta Metallurgica 11(7) (1977) 575-579. [162] R.L. Bell, T.G. Langdon, An investigation of grain-boundary sliding during creep, Journal of Materials Science 2(4) (1967) 313-323. [163] T.K. Ha, J.R. Son, W.B. Lee, C.G. Park, Y.W. Chang, Superplastic deformation of a fine- grained Zn–0.3wt.%Al alloy at room temperature, Materials Science and Engineering: A 307(1–2) (2001) 98-106.

120 Bibliography

[164] S.H. Xia, J. Wang, J.T. Wang, J.Q. Liu, Improvement of room-temperature superplasticity in Zn–22wt.%Al alloy, Materials Science and Engineering: A 493(1–2) (2008) 111-115. [165] A.C. Fischer-Cripps, Nanoindentation, 3rd ed., Springer, New York, 2011. [166] K. Nogi, M. Naito, T. Yokoyama, Nanoparticle technology handbook, 2nd ed., Elsevier, 2012. [167] W.C. Oliver, G.M. Pharr, An improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments, Journal of Materials Research 7(06) (1992) 1564-1583. [168] J. Field, M. Swain, A simple predictive model for spherical indentation, Journal of Materials Research 8(02) (1993) 297-306. [169] M.F. Doerner, W.D. Nix, A method for interpreting the data from depth-sensing indentation instruments, Journal of Materials Research 1(04) (1986) 601-609. [170] B. Briscoe, K. Sebastian, S. Sinha, Application of the compliance method to microhardness measurements of organic polymers, Philosophical Magazine A 74(5) (1996) 1159-1169. [171] H. Hertz, On the contact of elastic solids, Journal für die reine und angewandte Mathematik 92(110) (1881) 156-171. [172] A. Fischer-Cripps, A review of analysis methods for sub-micron indentation testing, Vacuum 58(4) (2000) 569-585. [173] W.C. Oliver, G.M. Pharr, Measurement of hardness and elastic modulus by instrumented indentation: Advances in understanding and refinements to methodology, Journal of Materials Research 19(01) (2004) 3-20. [174] B.N. Lucas, W.C. Oliver, Indentation power-law creep of high-purity indium, Metallurgical and Materials Transactions a-Physical Metallurgy and Materials Science 30(3) (1999) 601-610. [175] H.N. Mathur, V. Maier-Kiener, S. Korte-Kerzel, Deformation in the γ-Mg17Al12 phase at 25–278 °C, Acta Materialia 113 (2016) 221-229. [176] X.-G. Jiang, S. Yang, J. Earthman, F. Mohamed, Effect of Fe on ductility and cavitation in the superplastic Zn-22 Pct Al eutectoid, Metallurgical and Materials Transactions A 27(4) (1996) 863-872. [177] H. Zhang, K.G. Pradeep, S. Mandal, D. Ponge, H. Springer, D. Raabe, Dynamic strain- induced transformation: An atomic scale investigation, Scripta Materialia 109(Supplement C) (2015) 23-27. [178] Z. Wu, S. Sandlöbes, Y. Wang, J.S.K.L. Gibson, S. Korte-Kerzel, Creep behaviour of eutectic Zn-Al-Cu-Mg alloys, Materials Science and Engineering: A 724 (2018) 80-94. [179] R.R. Tohid, S.J. Bull, Getting accurate nanoindentation data from time-dependent and microstructural effects of zinc, International Journal of Materials Research 98(5) (2007) 353- 359. [180] J.S.K.L. Gibson, S. Schröders, C. Zehnder, S. Korte-Kerzel, On extracting mechanical properties from nanoindentation at temperatures up to 1000 °C, Extreme Mechanics Letters 17 (2017) 43-49. [181] M. Hardiman, T.J. Vaughan, C.T. McCarthy, The effects of pile-up, viscoelasticity and hydrostatic stress on polymer matrix nanoindentation, Polymer Testing 52 (2016) 157-166. [182] H.M. Lee, W.-J. Lee, Poisson's ratio of pure metals and their non-metallic compounds, Scripta Metallurgica et Materialia 25(4) (1991) 965-968. [183] I. Horcas, R. Fernández, J. Gomez-Rodriguez, J. Colchero, J. Gómez-Herrero, A. Baro, WSXM: a software for scanning probe microscopy and a tool for nanotechnology, Review of Scientific Instruments 78(1) (2007) 013705.

121 Bibliography

[184] D. Yan, C.C. Tasan, D. Raabe, High resolution in situ mapping of microstrain and microstructure evolution reveals damage resistance criteria in dual phase steels, Acta Materialia 96 (2015) 399-409. [185] H. Baker, H. Okamoto, ASM Handbook, Volume 03: Alloy Phase Diagrams, ASM International, 1992. [186] L. Li, Y. Liu, H. Gao, Z. Gao, Phase formation sequence of high-temperature Zn-4Al-3Mg solder, Journal of Materials Science: Materials in Electronics 24(1) (2013) 336-344. [187] A.J. Perry, Solute-vacancy interaction energies and the effect of 0.009 at.% Mg on the ageing kinetics of an Al-4.01 at.% Zn alloy, Acta Metallurgica 14(10) (1966) 1143-1156. [188] F.-W. Ling, D.E. Laughlin, S.C. Chang, Room temperature decomposition of a superplastic Zn-Al eutectoid alloy, Materials Science and Engineering 51(1) (1981) 47-53. [189] F.-W. Ling, D. Laughlin, The kinetics of transformation in Zn-Al superplastic alloys, Metallurgical Transactions A 10(7) (1979) 921-928. [190] B.K. Prasad, Tensile properties of some zinc-based alloys comprising 27.5% Al: effects of alloy microstructure, composition and test conditions, Materials Science and Engineering: A 245(2) (1998) 257-266. [191] D.A. Curry, J.E. King, J.F. Knott, Effects of hydrostatic tension on cleavage fracture of pure polycrystalline zinc, Metal Science 12(5) (1978) 247-250. [192] L. Hu, G. Zhang, W. Hu, G. Gottstein, S. Bogner, A. Bührig-Polaczek, Tensile creep of directionally solidified NiAl–9Mo in situ composites, Acta Materialia 61(19) (2013) 7155-7165. [193] H.B. Huntington, The Elastic Constants of Crystals, Solid State Physics 7 (1958) 213- 351. [194] D. Lazarus, Diffusion in Metals, Solid State Physics 10 (1960) 71-126. [195] P.G. Shewmon, Self-Diffusion in Magnesium Single Crystals, Jom 8(8) (1956) 918-922. [196] Y.A. Chang, L. Himmel, Temperature dependence of the elastic constants of Cu, Ag, and Au above room temperature, Journal of Applied Physics 37(9) (1966) 3567-3572. [197] W.C. Overton, J. Gaffney, Temperature Variation of the Elastic Constants of Cubic Elements .1. Copper, Physical Review 98(4) (1955) 969-977. [198] N.Q. Chinh, P. Szommer, Mathematical description of indentation creep and its application for the determination of strain rate sensitivity, Materials Science and Engineering: A 611 (2014) 333-336. [199] C. Su, E.G. Herbert, S. Sohn, J.A. LaManna, W.C. Oliver, G.M. Pharr, Measurement of power-law creep parameters by instrumented indentation methods, Journal of the Mechanics and Physics of Solids 61(2) (2013) 517-536. [200] P.S. Phani, W.C. Oliver, A direct comparison of high temperature nanoindentation creep and uniaxial creep measurements for commercial purity aluminum, Acta Materialia 111 (2016) 31-38. [201] X.Y. Zhu, X.J. Liu, F. Zeng, F. Pan, Room temperature nanoindentation creep of nanoscale Ag/Fe multilayers, Materials Letters 64(1) (2010) 53-56. [202] N. Kaur, D. Kaur, Room temperature nanoindentation creep of nanograined NiTiW shape memory alloy thin films, Surface and Coatings Technology 260 (2014) 260-265. [203] F.A. Mohamed, T.G. Langdon, The transition from dislocation climb to viscous glide in creep of solid solution alloys, Acta Metallurgica 22(6) (1974) 779-788. [204] T.G. Langdon, Identifying creep mechanisms in plastic flow, Zeitschrift für Metallkunde 96(6) (2005) 522-531. [205] W. Blum, P. Eisenlohr, Dislocation mechanics of creep, Materials Science and Engineering: A 510-511 (2009) 7-13.

122 Bibliography

[206] R. Mahmudi, A. Karsaz, A. Akbari-Fakhrabadi, A.R. Geranmayeh, Impression creep study of a Cu–0.3Cr–0.1Ag alloy, Materials Science and Engineering: A 527(10) (2010) 2702- 2708. [207] P. Yavari, T.G. Langdon, An examination of the breakdown in creep by viscous glide in solid solution alloys at high stress levels, Acta Metallurgica 30(12) (1982) 2181-2196. [208] O.D. Sherby, P.M. Burke, Mechanical behavior of crystalline solids at elevated temperature, Progress in Materials Science 13 (1968) 323-390. [209] S.C. Tjong, Z.Y. Ma, Steady state creep deformation behaviour of SiC particle reinforced 2618 aluminium alloy based composites, Materials Science and Technology 15(4) (1999) 429- 436. [210] H. Dieringa, N. Hort, K.U. Kainer, Investigation of minimum creep rates and stress exponents calculated from tensile and compressive creep data of magnesium alloy AE42, Materials Science and Engineering: A 510 (2009) 382-386. [211] P. Jain, K.S. Kumar, Tensile creep of Mo–Si–B alloys, Acta Materialia 58(6) (2010) 2124- 2142. [212] W. Liu, X. Liu, F. Lu, X. Tang, H. Cui, Y. Gao, Creep behavior and microstructure evaluation of welded joint in dissimilar modified 9Cr–1Mo steels, Materials Science and Engineering: A 644 (2015) 337-346. [213] X. Nie, H. Liu, X. Zhou, D. Yi, B. Huang, Z. Hu, Y. Xu, Q. Yang, D. Wang, Q. Gao, Creep of Ti–5Al–5Mo–5V–1Fe–1Cr alloy with equiaxed and lamellar microstructures, Materials Science and Engineering: A 651 (2016) 37-44. [214] Z.F. Zhou, B.A. Parker, On the high-stress region stress exponent during creep, Scripta Metallurgica et Materialia 32(11) (1995) 1889-1893. [215] W.J. Evans, A.W. Beale, G.F. Harrison, Creep rate stress exponents, the friction stress (αo) and the Lagneborg particle by-passing model, Scripta Metallurgica 14(1) (1980) 149-154. [216] J. Weertman, Steady‐state creep through dislocation climb, Journal of Applied Physics 28(3) (1957) 362-364. [217] C.R. Barrett, W.D. Nix, A model for steady state creep based on the motion of jogged screw dislocations, Acta Metallurgica 13(12) (1965) 1247-1258. [218] M.E. Kassner, Fundamentals of Creep in Metals and Alloys, Fundamentals of Creep in Metals and Alloys, Butterworth-Heinemann, Boston, 2015. [219] R. Goodall, T.W. Clyne, A critical appraisal of the extraction of creep parameters from nanoindentation data obtained at room temperature, Acta Materialia 54(20) (2006) 5489-5499. [220] D. Tabor, The Hardness of Metals, OUP Oxford, 2000. [221] F. Wang, P. Huang, K. Xu, Strain rate sensitivity of nanoindentation creep in polycrystalline Al film on Silicon substrate, Surface and Coatings Technology 201(9) (2007) 5216-5218. [222] V. Maier, B. Merle, M. Göken, K. Durst, An improved long-term nanoindentation creep testing approach for studying the local deformation processes in nanocrystalline metals at room and elevated temperatures, Journal of Materials Research 28(09) (2013) 1177-1188. [223] W.B. Li, J.L. Henshall, R.M. Hooper, K.E. Easterling, The mechanisms of indentation creep, Acta Metallurgica et Materialia 39(12) (1991) 3099-3110. [224] S.J. Bull, Nanoindentation of coatings, Journal of Physics D: Applied Physics 38(24) (2005) 393-413. [225] C.E. Beck, F. Hofmann, J.K. Eliason, A.A. Maznev, K.A. Nelson, D.E.J. Armstrong, Correcting for contact area changes in nanoindentation using surface acoustic waves, Scripta Materialia 128 (2017) 83-86.

123 Bibliography

[226] A.G. Crocker, P.E.J. Flewitt, G.E. Smith, G.M. Hughes, The brittle fracture of polycrystalline zinc, Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 463(2085) (2007) 2129-2151. [227] S. Wasiur-Rahman, M. Medraj, Critical assessment and thermodynamic modeling of the binary Mg–Zn, Ca–Zn and ternary Mg–Ca–Zn systems, Intermetallics 17(10) (2009) 847-864. [228] C. Yao, S.L. Tay, T. Zhu, H. Shang, W. Gao, Effects of Mg content on microstructure and electrochemical properties of Zn–Al–Mg alloys, Journal of Alloys and Compounds 645 (2015) 131-136. [229] J. Elvins, J.A. Spittle, J.H. Sullivan, D.A. Worsley, The effect of magnesium additions on the microstructure and cut edge corrosion resistance of zinc aluminium alloy galvanised steel, Corrosion Science 50(6) (2008) 1650-1658. [230] Y. Zuo, Y.A. Chang, Thermodynamic calculation of the Al-Mg phase diagram, Calphad 17(2) (1993) 161-174. [231] L.E. Mulvania, R.D. Weltzin, P.A. Talcott, A superplastic Zn-Al alloy with creep resistance, Metals Engineering Quarterly 14 (1974) 55 - 60. [232] Q. Wei, S. Cheng, K.T. Ramesh, E. Ma, Effect of nanocrystalline and ultrafine grain sizes on the strain rate sensitivity and activation volume: fcc versus bcc metals, Materials Science and Engineering: A 381(1-2) (2004) 71-79. [233] E.W. Hart, A theory for flow of polycrystals, Acta Metallurgica 15(9) (1967) 1545-1549. [234] M.J. Mayo, W.D. Nix, A micro-indentation study of superplasticity in Pb, Sn, and Sn-38 wt-percent-Pb, Acta Metallurgica 36(8) (1988) 2183-2192. [235] G.B. Gibbs, The thermodynamics of thermally‐activated dislocation glide, Physica Status Solidi (B) 10(2) (1965) 507-512. [236] I.-C. Choi, D.-H. Lee, B. Ahn, K. Durst, M. Kawasaki, T.G. Langdon, J.-i. Jang, Enhancement of strain-rate sensitivity and shear yield strength of a magnesium alloy processed by high-pressure torsion, Scripta Materialia 94 (2015) 44-47. [237] V. Maier, K. Durst, J. Mueller, B. Backes, H.W. Höppel, M. Göken, Nanoindentation strain-rate jump tests for determining the local strain-rate sensitivity in nanocrystalline Ni and ultrafine-grained Al, Journal of Materials Research 26(11) (2011) 1421-1430. [238] J.M. Wheeler, V. Maier, K. Durst, M. Göken, J. Michler, Activation parameters for deformation of ultrafine-grained aluminium as determined by indentation strain rate jumps at elevated temperature, Materials Science and Engineering: A 585(Supplement C) (2013) 108- 113. [239] V. Maier, C. Schunk, M. Göken, K. Durst, Microstructure-dependent deformation behaviour of bcc-metals – indentation size effect and strain rate sensitivity, Philosophical Magazine 95(16-18) (2014) 1766-1779. [240] H. Ledbetter, Elastic properties of zinc: a compilation and a review, Journal of Physical and Chemical Reference Data 6(4) (1977) 1181-1203. [241] K.F. Gan, A.H.W. Ngan, The unusual size effect of eutectic Sn/Pb alloys in the micro regime: Experiments and modeling, Acta Materialia 151 (2018) 282-292. [242] J.R. Freeman, F. Sillers, P.F. Brandt, S. United, Pure zinc at normal and elevated temperatures. Part I., Some physical properties. Part II., Crystal structure, U.S. Dept. of Commerce, Bureau of Standards : U.S. Govt. Print. Off., Washington, D.C., 1926. [243] N.Q. Chinh, P. Szommer, Z. Horita, T.G. Langdon, Experimental evidence for grain- boundary sliding in ultrafine-grained aluminum processed by severe plastic deformation, Advanced Materials 18(1) (2006) 34-39. [244] H. Conrad, Grain-size dependence of the flow stress of Cu from millimeters to nanometers, Metallurgical and Materials Transactions A 35(9) (2004) 2681-2695.

124 Bibliography

[245] J. Luster, M.A. Morris, Compatibility of deformation in two-phase Ti-Al alloys: Dependence on microstructure and orientation relationships, Metallurgical and Materials Transactions A 26(7) (1995) 1745-1756. [246] F. Wang, S. Sandlöbes, M. Diehl, L. Sharma, F. Roters, D. Raabe, In situ observation of collective grain-scale mechanics in Mg and Mg–rare earth alloys, Acta Materialia 80 (2014) 77-93. [247] D. Mercier, C. Zambaldi, T.R. Bieler, A Matlab toolbox to analyze slip transfer through grain boundaries, IOP Conference Series: Materials Science and Engineering 82 (2015) 012090. [248] R. Elliott, Eutectic solidification processing: crystalline and glassy alloys, Elsevier, 1983. [249] M. Durman, S. Murphy, An electron metallographic study of pressure die-cast commercial zinc-aluminium-based alloy ZA27, Journal of Materials Science 32(6) (1997) 1603- 1611. [250] A. Deruyttere, G. Greenough, The criterion for the cleavage fracture of zinc single crystals, Journal of the Institute of Metals 84(9) (1956) 337-345. [251] A. Deruyttere, G. Greenough, Cleavage fracture of zinc single crystals, Nature (172) (1953) 170-171. [252] J. Gilman, Fracture of zinc-monocrystals and bicrystals, Tranctions of the American Institute of Mining and Metallurgical Engineers 212 (1958) 783-791.

125

This page intentionally left blank.

Abstract

Abstract

Zn-Al based alloys are widely used as structural and decorative parts as well as machinery and equipment with a complex geometry or equipment needing a high manufacturing precision, particular in the die casting industry. However, Zn-Al based alloys suffer from low creep resistance and long-term mechanical instability. This thesis therefore aims to investigate the underlying physical mechanisms, as well as to explore possible methods to overcome these drawbacks. To this end, three eutectic ZnAl4Cu1 alloys with different Mg contents (0.04 wt.%, 0.21 wt.% and 0.31 wt.%) were comprehensively investigated from the macroscopic scale to the microscopic scale in terms of their mechanical properties, microstructures and deformation mechanisms using macroscopic ex-situ / in-situ tensile tests and micromechanical test in conjunction with scanning electron microscopy (SEM), electron backscatter diffraction (EBSD), atomic force microscopy (AFM), transmission electron microscopy (TEM) and atom probe tomography (APT).

Chapter 4 introduces the macro mechanical response and mechanisms of the three eutectic ZnAl4Cu1 alloys investigated. Dilute Mg alloying caused an improvement of the yield strength and ductility of ZnAl4Cu1 base alloys. Uniaxial tensile tests revealed two distinct deformation regimes of the alloys: (i) work hardening and brittle fracture at low temperatures and / or high strain rates with basal slip and {101̅2}[101̅1̅ ] twinning as predominant deformation mechanisms of the primary η-Zn phase; (ii) work softening and ductile failure at elevated temperatures and / or low strain rates where the predominant deformation mechanisms are deformation twinning, dislocation motion in the primary η-Zn phase and grain / phase boundary sliding in the eutectic and eutectoid structures.

The creep behaviour of the bulk ZnAl4Cu1 alloys as well as the local creep behaviour of the individual microstructural constituents are presented in chapter 5. Tensile creep tests in the temperature range of 25 to 105°C and stress range of 61 to 130 MPa revealed stress exponents of 6.9 – 8.0 and creep activation energies of 93 – 104 kJ/mol in the bulk ZnAl4Cu1 alloys, suggesting that the creep behaviour of these alloys is controlled by dislocation movement. Primary η-Zn phase showed the highest creep resistance, and η-Zn + α-Al eutectoid structures containing Mg2Zn11 precipitates showed the lowest creep resistance of the microstructural constituents during nanoindentation creep tests. Moreover, dilute Mg alloying caused an accelerated creep rate of ZnAl4Cu1 alloys due to increased grain / phase boundary activities and reduced geometrical constraints of eutectic and eutectoid colonies.

Furthermore, the local mechanical properties and deformation mechanisms of the individual microstructural constituents in alloy ZnAl4Cu1Mg0.31 were systematically studied using nanoindentation tests at room temperature (25°C) and 85°C, presented in chapter 6. Estimation of the strain rate sensitivities and activation volumes from nanoindentation strain

126 Abstract rate jump tests suggested predominant deformation by dislocation-mediated mechanisms in the primary η-Zn phase and by grain / phase boundary sliding in the eutectoid structures. This was later verified using SEM-EBSD and AFM. Chapter 6 further provides information on the role of individual microstructural constituents in alloy ZnAl4Cu1Mg0.31 during bulk deformation obtained using quasi in-situ micro digital image correlation (µ-DIC) in the SEM during tensile deformation at 85°C. µ-DIC showed that eutectic / eutectoid colonies carried higher strain than the primary η-Zn phase grains, and confirmed strain transfer across Zn-Al phase boundaries and eutectic / eutectoid colony boundaries.

Chapter 7 is focussing on the understanding of the local precipitation behaviour in alloy ZnAl4Cu1Mg0.31 and its influence on the mechanical properties. The precipitation and decomposition phenomena in alloy ZnAl4Cu1Mg0.31 were investigated using TEM and APT.

A non-equilibrium ZnxAl1-x (x≥0.7) transition phase, which dissolved during deformation at 85°C, was identified and structurally as well as chemically characterised. The partial dissolution of these precipitates was proposed to contribute to the lack of long-term mechanical stability of Zn-Al alloys that currently poses one of the major drawbacks to their application.

Key words: Zn alloy, microstructure, mechanical property, plasticity, creep, nanoindentation, DIC.

127 Kurzzusammenfassung

Kurzzusammenfassung

Zn-Al Basis Legierungen werden für dekorative und strukturelle Bauteile eingesetzt sowie für Anwendungen mit komplexer Geometrie und hohen Anforderungen an die geometrische Fertigungspräzision, insbesondere als Druckgusslegierungen. Zn-Al Legierungen leiden jedoch unter einer geringen Kriechfestigkeit und einer langfristigen mechanischen Instabilität. Das Ziel dieser Arbeit ist es daher, die zugrundeliegenden physikalischen Mechanismen zu untersuchen und Möglichkeiten zur Überwindung dieser Nachteile zu erforschen. Zu diesem Zweck wurden drei eutektische ZnAl4Cu1-Legierungen mit unterschiedlichen Mg-Gehalten (0,04 Gew.-%, 0,21 Gew.-% und 0,31 Gew.-%) umfassend hinsichtlich ihrer mechanischen Eigenschaften, Mikrostrukturen und Verformungsmechanismen vom makroskopischen bis zum mikroskopischen Maßstab untersucht. Die verwendeten experimentellen Methoden reichten von makroskopischen Ex-situ / In-situ-Zugversuchen zu mikromechanischen Tests in Verbindung mit Rasterelektronenmikroskopie (REM), Elektronenrückstreubeugung (EBSD), Rasterkraftmikroskopie (AFM), Transmissionselektronenmikroskopie (TEM) und Atomsonden- Tomographie (APT).

Kapitel 4 stellt die makro-mechanischen Eigenschaften und die aktiven Verformungsmechanismen der drei untersuchten eutektischen ZnAl4Cu1-Legierungen vor. Zugabe einer geringen Menge Mg erhöht die Streckgrenze und Duktilität von ZnAl4Cu1- Basislegierungen. Uniaxiale Zugversuche zeigten zwei unterschiedliche Verformungsregime in den untersuchten Legierungen: (i) Kaltverfestigung und Sprödbruch bei niedrigen Temperaturen und / oder hohen Verformungsgeschwindigkeiten mit basalem Gleiten und {101̅2}[101̅1̅] Zwillingsbildung als vorherrschenden Verformungsmechanismen der primären η-Zn-Phase; (ii) Entfestigung und duktiles Versagen bei erhöhten Temperaturen und / oder niedrigen Verformungsgeschwindigkeiten mit Deformationszwillingsbildung und Versetzungsbewegung in der primären η-Zn-Phase und Korn / Phasengrenzengleitung in den eutektischen und eutektoiden Strukturen als vorherrschende Verformungsmechanismen.

Das makroskopische Kriechverhalten der ZnAl4Cu1 Legierungen sowie das lokale Kriechverhalten der einzelnen Gefügebestandteile sind in Kapitel 5 dargestellt. Zugkriechversuche im Temperaturbereich von 25 bis 105°C und Spannungsbereich von 61 bis 130 MPa ergaben Spannungsexponenten von 6.9 – 8.0 und Kriechaktivierungsenergien von 93 – 104 kJ / mol, was darauf hindeutet, dass das Kriechverhalten dieser Legierungen primär durch Versetzungsbewegung erfolgt. Die primäre η-Zn-Phase zeigte die höchste

Kriechbeständigkeit und η-Zn + α-Al eutektoide Strukturen mit Mg2Zn11 Ausscheidungen zeigten die geringste Kriechfestigkeit der mikrostrukturellen Bestandteile während der Nanoindentation-Kriechtests. Darüber hinaus verursachte Mg-Zugabe eine beschleunigte Kriechgeschwindigkeit von ZnAl4Cu1 Legierungen aufgrund von erhöhten Korn /

128 Kurzzusammenfassung

Phasengrenzen-Aktivitäten und reduzierten geometrischen Beschränkungen der eutektischen und eutektoiden Kolonien.

Darüber hinaus wurden die lokalen mechanischen Eigenschaften und Verformungsmechanismen der einzelnen Gefügebestandteile in der Legierung ZnAl4Cu1Mg0.31 unter Verwendung von Nanoindentationstests bei Raumtemperatur (25°C) und 85°C systematisch untersucht (siehe Kapitel 6). Die Aktivierungsvolumina wurden anhand von Nanoindentations-Dehnungsratenwechselversuchen ermittelt und weisen auf eine vorherrschende Verformung durch Versetzungs mechanismen in der primären η-Zn-Phase und durch Korn- / Phasengrenzengleiten in den eutektoiden Strukturen hin. Dies wurde später mittels SEM-EBSD und AFM Untersuchungen verifiziert. In Kapitel 6 wird darüberhinausgehend die Rolle der einzelnen mikrostrukturellen Bestandteile in der Legierung ZnAl4Cu1Mg0.31 während der Zugverformung bei 85 ° C mittels quasi-in-situ Mikro-Digital- Bildkorrelation (μ-DIC) im SEM untersucht. Diese Untersuchungen zeigten, dass die eutektischen / eutektoiden Kolonien eine höhere Dehnung aufweisen als die primären η-Zn- Körner und zeigten den Verformungstransfer über Zn-Al Phasengrenzen und eutektische / eutektoide Koloniegrenzen.

Kapitel 7 widmet sich dem Verständnis des lokalen Ausscheidungsverhaltens der Legierung ZnAl4Cu1Mg0.31 und deren Einfluss auf die mechanischen Eigenschaften. Die bei Verformung bei 85°C stattfindenen Dekompositionsphänomene in der Legierung ZnAl4Cu1Mg0.31 wurden mittels Transmissionselektronenmikroskopie und Atomsonden-

Tomographie untersucht. Eine Nichtgleichgewichts-Übergangsphase ZnxAl1-x (x≥0.7), welche sich während der Verformung bei 85°C auflöste, wurde identifiziert und strukturell und chemisch charakterisiert. Es wurde geschlussfolgert, dass die teilweise Auflösung dieser Ausscheidungen zu der langfristigen mechanischen Instabilität von Zn-Al-Legierungen beiträgt, die derzeit einen der Hauptnachteile ihrer Anwendung darstellt.

Schlüsselwörter: Zn-Legierung, Mikrostruktur, mechanische Eigenschaft, Plastizität, Kriechen, Nanoindentation, DIC

129